url
stringlengths
31
38
title
stringlengths
7
229
abstract
stringlengths
39
2.87k
text
stringlengths
1
3.74M
meta
dict
https://arxiv.org/abs/1805.11660
Notes on hyperbolic dynamics
These expository notes present a proof of the Stable/Unstable Manifold Theorem (also known as the Hadamard--Perron Theorem). They also give examples of hyperbolic dynamics: geodesic flows on surfaces of negative curvature and dispersing billiards.
\section{Stable/unstable manifolds in a simple setting} \label{s:base} In this section we present the Stable/Unstable Manifold Theorem (broken into two parts, Theorem~\ref{t:stun-1} in~\S\ref{s:base-1} and Theorem~\ref{t:stun-2} in~\S\ref{s:base-4}) under several simplifying assumptions: \begin{itemize} \item we study iterates of a single map $\varphi$, defined on a neighborhood of $0$ in $\mathbb R^2$; \item $\varphi(0)=0$ and $d\varphi(0)$ is a hyperbolic matrix, with eigenvalues $2$ and~$1\over 2$; \item the map $\varphi$ is close to the linearized map $x\mapsto d\varphi(0)\cdot x$ in the $C^{N+1}$ norm. \end{itemize} These assumptions are made to make the notation below simpler, however they do not impact the substance of the proof. As explained below in~\S\S\ref{s:general}--\ref{s:maps-and-flows}, the arguments of this section can be adapted to the setting of general hyperbolic maps and flows. \subsection{Existence of stable/unstable manifolds} \label{s:base-1} Throughout this section we use the following notation for $\ell_\infty$ balls in~$\mathbb R^2$: $$ \overline B_\infty(0,r):=\{(x_1,x_2)\in\mathbb R^2\colon \max(|x_1|,|x_2|)\leq r\}. $$ We assume that $U_\varphi,V_\varphi\subset \mathbb R^2$ are open sets with $\overline B_\infty(0,1)\subset U_\varphi\cap V_\varphi$ and $$ \varphi:U_\varphi\to V_\varphi $$ is a $C^{N+1}$ map (here $N\geq 1$ is fixed) which satisfies the following assumptions: \begin{enumerate} \item $\varphi(0)=0$; \item The differential $d\varphi(0)$ is equal to \begin{equation} \label{e:hypas-1} d\varphi(0)=\begin{pmatrix} 2 & 0 \\ 0 & 1/2 \end{pmatrix}; \end{equation} \item for a small constant $\delta>0$ (chosen later in Theorems~\ref{t:stun-1} and~\ref{t:stun-2}) and all multiindices $\alpha$ with $2\leq |\alpha|\leq N+1$, we have \begin{equation} \label{e:hypas-2} \sup_{U_\varphi}|\partial^\alpha \varphi|\leq \delta; \end{equation} \item $\varphi$ is a diffeomorphism onto its image. \end{enumerate} We remark that assumptions~(3) and~(4) above can be arranged to hold locally by zooming in to a small neighborhood of 0, see~\S\ref{s:reduce-1} below. It follows from~\eqref{e:hypas-1} that the space $E_u(0):=\mathbb R\partial_{x_1}$ is preserved by the linearized map $x\mapsto d\varphi(0)\cdot x$ and vectors in this space are expanded exponentially by the powers of $d\varphi(0)$. Similarly the space $E_s(0):=\mathbb R\partial_{x_2}$ is invariant and contracted exponentially by the powers of $d\varphi(0)$. We call $E_u(0)$ the \emph{unstable space} and $E_s(0)$ the \emph{stable space} of $\varphi$ at~0. The main results of this section are the nonlinear versions of the above observations: namely there exist one-dimensional \emph{unstable/stable submanifolds} $W_u,W_s\subset\mathbb R^2$ which are (locally) invariant under the map $\varphi$; the iterates of $\varphi$ are exponentially expanding on the unstable manifold and exponentially contracting on the stable one. We construct the unstable/stable manifolds as graphs of $C^N$ functions. For a function $F:[-1,1]\to\mathbb R$ we define its unstable/stable graphs \begin{equation} \label{e:stun-graphs} \mathcal G_u(F):=\{x_2=F(x_1),\ |x_1|\leq 1\},\quad \mathcal G_s(F):=\{x_1=F(x_2),\ |x_2|\leq 1\} \end{equation} which are subsets of $\mathbb R^2$. Theorem~\ref{t:stun-1} below asserts existence of unstable/stable manifolds. The fact that $\varphi$ is expanding on $W_u$ and contracting on $W_s$ is proved later in Theorem~\ref{t:stun-2}. \begin{theo} \label{t:stun-1} Assume that $\delta$ is small enough (depending only on $N$) and assumptions~(1)--(4) above hold. Then there exist $C^N$ functions $$ F_u,F_s:[-1,1]\to [-1,1],\quad F_u(0)=F_s(0)=0,\quad \partial_{x_1}F_u(0)=\partial_{x_2}F_s(0)=0 $$ such that, denoting the graphs (see Figure~\ref{f:stun-base}) \begin{equation} \label{e:W-u-def} W_u:=\mathcal G_u(F_u),\quad W_s:=\mathcal G_s(F_s), \end{equation} we have \begin{equation} \label{e:graph-inv} \varphi(W_u)\cap \overline B_\infty(0,1)=W_u,\quad \varphi^{-1}(W_s)\cap \overline B_\infty(0,1)=W_s. \end{equation} Moreover, $W_u\cap W_s=\{0\}$. \end{theo} \begin{figure} \includegraphics{stunnote.1} \caption{The manifolds $W_u,W_s$. The square is $\overline B_\infty(0,1)$, the horizontal direction is $x_1$, and the vertical direction is $x_2$.} \label{f:stun-base} \end{figure} The proof of the theorem, given in~\S\S\ref{s:base-2}--\ref{s:base-3} below, is partially based on the proof of the more general Hadamard--Perron theorem in~\cite[Theorem~6.2.8]{KaHa}. The main idea is to show that the action of $\varphi$ on unstable graphs is a contraction mapping with respect to an appropriately chosen metric (and same with the action of $\varphi^{-1}$ on stable graphs). There are however two points in which our proof differs from the one in~\cite{KaHa}: \begin{itemize} \item We run the contraction mapping argument on the metric space of $C^N$ functions whose $N$-th derivative has Lipschitz norm bounded by 1, with the $C^N$ metric. This is slightly different from the space used in~\cite[\S6.2.d, Step 3]{KaHa} and it requires having $N+1$ derivatives of the map $\varphi$ to obtain $C^N$ regularity for invariant graphs (rather than $N$ derivatives as in~\cite{KaHa}). The upshot is that we do not need separate arguments for establishing regularity of the manifolds~$W_u,W_s$ \cite[\S6.2.d, Steps 1--2, 4--5]{KaHa}. \item We only consider the action of $\varphi$ on the ball $\overline B_\infty(0,1)$ rather than extending it to the entire $\mathbb R^2$ as in~\cite[Lemma~6.2.7]{KaHa}. Because of this parts~(3)--(4) of Theorem~\ref{t:stun-2} below have a somewhat different proof than the corresponding statement~\cite[Theorem~6.2.8(iii)]{KaHa}. \end{itemize} \noindent\textbf{Notation:} In the remainder of this section we denote by $C$ constants which depend only on $N$ (in particular they do not depend on $\delta$), and write $R=\mathcal O(\delta)$ if $|R|\leq C\delta$. We assume that $\delta>0$ is chosen small (depending only on $N$). \subsection{Action on graphs and derivative bounds} \label{s:base-2} We start the proof of Theorem~\ref{t:stun-1} by considering the action of $\varphi$ on unstable graphs. (The action of $\varphi^{-1}$ on stable graphs is handled similarly.) This action, denoted by $\Phi_u$ and called the \emph{graph transform}, is defined by \begin{lemm} \label{l:gmap} Let $F:[-1,1]\to \mathbb R$ satisfy \begin{equation} \label{e:gmap-1} F(0)=0,\quad \sup |\partial_{x_1} F|\leq 1. \end{equation} Then there exists a function $$ \Phi_uF:[-1,1]\to\mathbb R,\quad \Phi_uF(0)=0 $$ such that (see Figure~\ref{f:gmap}) \begin{equation} \label{e:gmap-def} \varphi(\mathcal G_u(F))\cap \{|x_1|\leq 1\}=\mathcal G_u(\Phi_u F). \end{equation} \end{lemm} \begin{figure} \includegraphics[scale=0.35]{matfig/bstun1.jpg} \includegraphics[scale=0.35]{matfig/bstun2.jpg} \caption{Left: the graph $\mathcal G_u(F)$ for some function $F$ satisfying~\eqref{e:gmap-1}. Right: the image of $\mathcal G_u(F)$ under $\varphi$. The solid blue part is the graph $\mathcal G_u(\Phi_u F)$. Figures~\ref{f:gmap}--\ref{f:exc-2} are plotted numerically using the map $\varphi(x_1,x_2)=(2x_1+{1\over 2}x_2^2,{1\over 2}x_2+{1\over 2}x_1^2)$.} \label{f:gmap} \end{figure} \begin{proof} Define the components $\varphi_1,\varphi_2:\overline B_\infty(0,1)\to\mathbb R$ of $\varphi$ by \begin{equation} \label{e:phi-12-def} \varphi(x)=(\varphi_1(x),\varphi_2(x)),\quad x\in \overline B_\infty(0,1). \end{equation} Next, define the functions $G_1,G_2:[-1,1]\to\mathbb R$ by \begin{equation} \label{e:G-12-def} G_1(x_1):=\varphi_1(x_1,F(x_1)),\quad G_2(x_1):=\varphi_2(x_1,F(x_1)), \end{equation} so that the manifold $\varphi(\mathcal G_u(F))$ has the form $$ \varphi(\mathcal G_u(F))=\{(G_1(x_1),G_2(x_1))\colon |x_1|\leq 1\}. $$ To write $\varphi(\mathcal G_u(F))$ as a graph, we need to show that $G_1$ is invertible. Note that $G_1(0)=0$. We next compute $$ \partial_{x_1}G_1(x_1)=\partial_{x_1}\varphi_1(x_1,F(x_1))+\partial_{x_2}\varphi_1(x_1,F(x_1))\cdot\partial_{x_1}F(x_1). $$ Together~\eqref{e:hypas-1} and~\eqref{e:hypas-2} imply for all $(x_1,x_2)\in \overline B_\infty(0,1)$ $$ \partial_{x_1}\varphi_1(x_1,x_2)=2+\mathcal O(\delta),\quad \partial_{x_2}\varphi_1(x_1,x_2)=\mathcal O(\delta), $$ so for $\delta$ small enough \begin{equation} \label{e:G-1-der} \partial_{x_1}G_1(x_1)=2+\mathcal O(\delta)\geq {3\over 2}\quad\text{for all } x_1\in [-1,1]. \end{equation} Therefore $G_1$ is a diffeomorphism and its image contains $[-1,1]$. It follows that $\varphi(\mathcal G_u(F))\cap \{|x_1|\leq 1\}=\mathcal G_u(\Phi_u F)$ for the function $\Phi_u F$ defined by \begin{equation} \label{e:Phi-u-def} \Phi_uF(y_1)=G_2(G_1^{-1}(y_1)),\quad y_1\in [-1,1] \end{equation} where \begin{equation} \label{e:G-1-inv} G_1^{-1}:[-1,1]\to [-1,1] \end{equation} is the inverse of $G_1$. \end{proof} We now want to estimate the function $\Phi_u F$ in terms of $F$, ultimately showing that $\Phi_u$ is a contraction with respect to a certain norm. For that we use the following formula for the derivatives of $\Phi_u F$: \begin{lemm} \label{l:derform} Let $1\leq k\leq N$. Assume that $F\in C^k([-1,1];\mathbb R)$ satisfies \begin{equation} \label{e:gmap-2} F(0)=0,\quad \sup |\partial_{x_1}^j F|\leq 1\quad\text{for all }j=1,\dots,k. \end{equation} Then we have for all $y_1\in [-1,1]$ and $G_1^{-1}$ defined in~\eqref{e:G-1-inv} \begin{equation} \label{e:derform} \partial^k_{x_1}(\Phi_u F)(y_1)=L_k(x_1,F(x_1),\partial_{x_1}F(x_1),\dots,\partial_{x_1}^kF(x_1)),\quad x_1:=G_1^{-1}(y_1) \end{equation} where the function $L_k(x_1,\tau_0,\dots,\tau_k)$, depending on $\varphi$ but not on $F$, is continuous on the cube $Q_k:=[-1,1]^{k+2}$. Moreover $L_k(x_1,\tau_0,\dots,\tau_k)=2^{-k-1}\tau_k+\mathcal O(\delta)$, with the remainder satisfying the derivative bounds \begin{equation} \label{e:derform-2} \begin{gathered} \sup_{Q_k}\big|\partial^\alpha_{x_1}\partial^{\beta_0}_{\tau_0}\dots\partial^{\beta_k}_{\tau_k} \big(L_k(x_1,\tau_0,\dots,\tau_k)-2^{-k-1}\tau_k\big)\big|\leq C_{\alpha\beta}\,\delta\\ \text{for all}\quad\alpha,\beta_0,\dots,\beta_k\quad\text{such that}\quad \alpha+\beta_0+k\leq N+1. \end{gathered} \end{equation} \end{lemm} \Remark In the linear case $\varphi(x)=d\varphi(0)\cdot x$ we have $G_1(x_1)=2x_1$, $G_2(x_1)={1\over 2}F(x_1)$, therefore $\Phi_u F(y_1)={1\over 2}F({1\over 2}y_1)$. The meaning of~\eqref{e:derform-2} is that the action of $\Phi_u$ on the derivatives of~$F$ for nonlinear $\varphi$ is $\mathcal O(\delta)$-close to the linear case. \begin{proof} For $x\in \overline B_\infty(0,1)$ define the matrix $$ A(x):=d\varphi(x)=(A_{jk}(x)),\quad A_{jk}(x):=\partial_{x_k}\varphi_j(x). $$ By~\eqref{e:hypas-1} and~\eqref{e:hypas-2} we have \begin{equation} \label{e:A-prop} A(x)=\begin{pmatrix} 2& 0 \\ 0 & 1/2\end{pmatrix}+\mathcal O(\delta), \end{equation} and the remainder in~\eqref{e:A-prop} is $\mathcal O(\delta)$ with derivatives of order $\leq N$. We argue by induction on $k$. For $k=1$, from the definition~\eqref{e:Phi-u-def} of $\Phi_u F$ we have $$ \partial_{x_1}(\Phi_u F)(y_1)={\partial_{x_1}G_2(x_1)\over \partial_{x_1}G_1(x_1)},\quad x_1:=G_1^{-1}(y_1) $$ so~\eqref{e:derform} holds with \begin{equation} \label{e:lulla} L_1(x_1,\tau_0,\tau_1)={A_{21}(x_1,\tau_0)+A_{22}(x_1,\tau_0) \tau_1\over A_{11}(x_1,\tau_0)+A_{12}(x_1,\tau_0) \tau_1}. \end{equation} From~\eqref{e:A-prop} we see that $L_1(x_1,\tau_0,\tau_1)={1\over 4}\tau_1+\mathcal O(\delta)$ and the stronger remainder estimate~\eqref{e:derform-2} holds. Now assume that $2\leq k\leq N$ and~\eqref{e:derform}, \eqref{e:derform-2} hold for $k-1$. Then by the chain rule~\eqref{e:derform} holds for~$k$ with $$ L_{k}(x_1,\tau_0,\dots,\tau_{k}):={\partial_{x_1} L_{k-1}(x_1,\tau_0,\dots,\tau_{k-1}) +\sum_{j=0}^{k-1} \partial_{\tau_j}L_{k-1}(x_1,\tau_0,\dots,\tau_{k-1})\tau_{j+1} \over A_{11}(x_1,\tau_0)+A_{12}(x_1,\tau_0)\tau_1}. $$ It is straightforward to check that $L_k(x_1,\tau_0,\dots,\tau_k)=2^{-k-1}\tau_k+\mathcal O(\delta)$ and the stronger remainder estimate~\eqref{e:derform-2} holds. \end{proof} Armed with Lemma~\ref{l:derform} we estimate the derivatives of $\Phi_u F$ in terms of the derivatives of $F$. Let $1\leq k\leq N$. We use the following seminorm on $C^k([-1,1];\mathbb R)$: \begin{equation} \label{e:C-k-norm} \|F\|_{C^k}:=\max_{1\leq j\leq k}\sup|\partial^j_{x_1}F|. \end{equation} We will work with functions satisfying $F(0)=0$, on which $\|\bullet\|_{C^k}$ is a norm: \begin{equation} \label{e:F-sup-bound} F(0)=0\quad\Longrightarrow\quad \sup|F|\leq \|F\|_{C^1}. \end{equation} To establish the contraction property (see the remark following the proof of Lemma~\ref{l:derest-2}) we also need the space of functions $C^{k,1}([-1,1];\mathbb R)$ with Lipschitz continuous $k$-th derivative, endowed with the seminorm \begin{equation} \label{e:C-k-1-norm} \|F\|_{C^{k,1}}:=\max\bigg(\|F\|_{C^k},\sup_{x_1\neq\tilde x_1} {|\partial^k_{x_1}F(x_1)-\partial^k_{x_1}F(\tilde x_1)|\over |x_1-\tilde x_1|}\bigg). \end{equation} Note that $\|F\|_{C^k}\leq \|F\|_{C^{k,1}}\leq \|F\|_{C^{k+1}}$. Our first estimate implies that $\Phi_u$ maps the unit balls in $C^k$, $C^{k,1}$ into themselves: \begin{lemm} \label{l:derest-1} Let $1\leq k\leq N$ and assume that $F(0)=0$ and $\|F\|_{C^k}\leq 1$. Then \begin{equation} \label{e:derest-1.1} \|\Phi_u F\|_{C^k}\leq {1\over 4}\|F\|_{C^k}+C\delta. \end{equation} If additionally $\|F\|_{C^{k,1}}\leq 1$ then \begin{equation} \label{e:derest-1.2} \|\Phi_u F\|_{C^{k,1}}\leq {1\over 4}\|F\|_{C^{k,1}}+C\delta. \end{equation} \end{lemm} \begin{proof} Let $y_1\in [-1,1]$ and $x_1:=G_1^{-1}(y_1)\in [-1,1]$. By Lemma~\ref{l:derform} we have for all $j=1,\dots,k$ $$ |\partial_{x_1}^j (\Phi_uF)(y_1)|\leq 2^{-j-1}|\partial_{x_1}^jF(x_1)|+C\delta \leq {1\over 4}\|F\|_{C^k}+C\delta $$ which implies~\eqref{e:derest-1.1}. Next, assume that $\|F\|_{C^{k,1}}\leq 1$. Take $y_1,\tilde y_1\in [-1,1]$ such that $y_1\neq \tilde y_1$. Put $x_1:=G_1^{-1}(y_1)$, $\tilde x_1:=G_1^{-1}(\tilde y_1)$. Then by Lemma~\ref{l:derform} $$ \begin{aligned} |\partial_{x_1}^k (\Phi_u F)(y_1)-\partial_{x_1}^k(\Phi_u F)(\tilde y_1)| &\leq 2^{-k-1}|\partial_{x_1}^k F(x_1)-\partial_{x_1}^k F(\tilde x_1)| +C\delta|x_1-\tilde x_1|\\ &\quad\ +C\delta\max_{0\leq j\leq k}|\partial_{x_1}^j F(x_1)-\partial_{x_1}^j F(\tilde x_1)| \\ &\leq \Big({1\over 4}\|F\|_{C^{k,1}}+C\delta+C\delta \|F\|_{C^{k,1}}\Big)|x_1-\tilde x_1|. \end{aligned} $$ Since $|x_1-\tilde x_1|\leq |y_1-\tilde y_1|$ by~\eqref{e:G-1-der}, this implies~\eqref{e:derest-1.2}. \end{proof} The next estimate gives the contraction property of $\Phi_u$ in $C^k$ norm: \begin{lemm} \label{l:derest-2} Let $1\leq k\leq N$. Assume that $F,\widetilde F\in C^{k,1}$ satisfy $F(0)=\widetilde F(0)=0$ and $\|F\|_{C^{k,1}},\|\widetilde F\|_{C^{k,1}}\leq 1$. Then \begin{equation} \label{e:derest-2} \|\Phi_u F-\Phi_u \widetilde F\|_{C^k}\leq \Big({1\over 4}+C\delta\Big)\|F-\widetilde F\|_{C^k}. \end{equation} \end{lemm} \begin{proof} Let $G_1$ and~$\widetilde G_1$ be defined by~\eqref{e:G-12-def} using the functions $F$ and~$\widetilde F$ respectively. Take $y_1\in [-1,1]$ and put $x_1:=G_1^{-1}(y_1)$, $\tilde x_1:=\widetilde G_1^{-1}(y_1)$, see Figure~\ref{f:contractor}. We first estimate the difference between the two inverses $x_1$, $\tilde x_1$: \begin{equation} \label{e:anor-1} |x_1-\tilde x_1|\leq C\delta\|F-\widetilde F\|_{C^1}. \end{equation} \begin{figure} \includegraphics[scale=0.35]{matfig/bcon1.jpg} \includegraphics[scale=0.35]{matfig/bcon2.jpg} \caption{Left: the points $(x_1,F(x_1))$, $(\tilde x_1,\widetilde F(\tilde x_1))$. The blue curve is the graph of $F$ and the red curve is the graph of $\widetilde F$. Right: the image of the picture on the left by $\varphi$.} \label{f:contractor} \end{figure} To show~\eqref{e:anor-1}, we write $$ \begin{aligned} |x_1-\tilde x_1|&\leq |\widetilde G_1(x_1)-\widetilde G_1(\tilde x_1)|= |\widetilde G_1(x_1)-G_1(x_1)| \\&=|\varphi_1(x_1,\widetilde F(x_1))-\varphi_1(x_1,F(x_1))|\leq C\delta \|F-\widetilde F\|_{C^1} \end{aligned} $$ where the first inequality follows from~\eqref{e:G-1-der} and the last inequality uses that $\partial_{x_2}\varphi_1=\mathcal O(\delta)$ by~\eqref{e:A-prop} and $|\widetilde F(x_1)-F(x_1)|\leq \|F-\widetilde F\|_{C^1}$ by~\eqref{e:F-sup-bound}. Next we have for all $j=0,\dots,k$ \begin{equation} \label{e:anor-2} |\partial_{x_1}^j F(x_1)-\partial_{x_1}^j \widetilde F(\tilde x_1)|\leq (1+C\delta) \|F-\widetilde F\|_{C^k}. \end{equation} Indeed, $$ \begin{aligned} |\partial_{x_1}^j F(x_1)-\partial_{x_1}^j \widetilde F(\tilde x_1)|&\leq |\partial_{x_1}^j F(x_1)-\partial_{x_1}^j F(\tilde x_1)| +|\partial_{x_1}^j F(\tilde x_1)-\partial_{x_1}^j\widetilde F(\tilde x_1)|\\ &\leq |x_1-\tilde x_1|+\|F-\widetilde F\|_{C^k}\leq (1+C\delta)\|F-\widetilde F\|_{C^k} \end{aligned} $$ where the second inequality uses the fact that $\|F\|_{C^{k,1}}\leq 1$ and the last inequality used~\eqref{e:anor-1}. Finally, by Lemma~\ref{l:derform} we estimate for all $j=1,\dots,k$ $$ \begin{aligned} |\partial^j_{x_1}(\Phi_uF)(y_1)-\partial^j_{x_1}(\Phi_u\widetilde F)(y_1)|&\leq 2^{-j-1}|\partial_{x_1}^j F(x_1)-\partial_{x_1}^j \widetilde F(\tilde x_1)| \\&\quad +C\delta |x_1-\tilde x_1|+C\delta\max_{0\leq \ell\leq j}|\partial_{x_1}^\ell F(x_1)-\partial_{x_1}^\ell\widetilde F(\tilde x_1)| \\&\leq \Big({1\over 4}+C\delta\Big)\|F-\widetilde F\|_{C^k} \end{aligned} $$ where the second inequality uses~\eqref{e:anor-1} and~\eqref{e:anor-2}. This implies~\eqref{e:derest-2}. \end{proof} \Remark The a priori bound $\|F\|_{C^{k,1}}\leq 1$ was used in the proof of~\eqref{e:anor-2}. Without it we would not be able to estimate the difference $|\partial^k_{x_1} F(x_1)-\partial^k_{x_1}\widetilde F(\tilde x_1)|$ since the functions $\partial^k_{x_1}F$, $\partial^k_{x_1}\widetilde F$ are evaluated at two different values of $x_1$. We do not use the stronger a priori bound $\|F\|_{C^{k+1}}\leq 1$ because it would make it harder to set up a complete metric space for the contraction mapping argument in~\eqref{e:X-N-def} below. \subsection{Contraction mapping argument} \label{s:base-3} \begin{figure} \includegraphics[scale=0.225]{matfig/bstuniter0.jpg} \includegraphics[scale=0.225]{matfig/bstuniter1.jpg} \includegraphics[scale=0.225]{matfig/bstuniter2.jpg} \hbox to\hsize{\quad\hss $n=0$\hss\hss $n=1$\hss\hss $n=2$\hss\quad} \caption{The iterations $\varphi^n(\mathcal G_{F_0})$ for two different choices of $F_0$. Both converge to $W_u$, illustrating~\eqref{e:contramap-u}.} \label{f:contramap} \end{figure} We now give the proof of Theorem~\ref{t:stun-1} using the estimates from the previous section. We show existence of the function $F_u$; the function $F_s$ is constructed similarly, replacing $\varphi$ by $\varphi^{-1}$ and switching the roles of $x_1$ and $x_2$. Consider the metric space $(\mathcal X_N,d_N)$ defined using the seminorms~\eqref{e:C-k-norm}, \eqref{e:C-k-1-norm}: \begin{equation} \label{e:X-N-def} \begin{aligned} \mathcal X_N&:=\{F\in C^{N,1}([-1,1];\mathbb R)\colon F(0)=0,\ \|F\|_{C^{N,1}}\leq 1\},\\ d_N(F,\widetilde F)&:=\|F-\widetilde F\|_{C^N}. \end{aligned} \end{equation} Then $(\mathcal X_N,d_N)$ is a complete metric space. Indeed, it is the subset of the closed unit ball in $C^N$ defined using the closed conditions $$ F(0)=0,\quad |\partial^N_{x_1}F(x_1)-\partial^N_{x_1} F(\tilde x_1)|\leq |x_1-\tilde x_1|\quad\text{for all } x_1,\tilde x_1\in [-1,1]. $$ By Lemmas~\ref{l:gmap} and~\ref{l:derest-1}, for $\delta$ small enough the graph transform defines a map $$ \Phi_u:\mathcal X_N\to\mathcal X_N. $$ By Lemma~\ref{l:derest-2}, for $\delta$ small enough this map is a contraction, specifically \begin{equation} \label{e:contraction-achieved} d_N(\Phi_u F,\Phi_u\widetilde F)\leq {1\over 3}d_N(F,\widetilde F)\quad\text{for all } F,\widetilde F\in \mathcal X_N. \end{equation} Therefore by the Contraction Mapping Principle the map $\Phi_u$ has a unique fixed point $$ F_u\in\mathcal X_N,\quad \Phi_uF_u=F_u. $$ In fact for each fixed $F_0\in \mathcal X_N$ we have (see Figure~\ref{f:contramap}) \begin{equation} \label{e:contramap-u} (\Phi_u)^n F_0\to F_u\quad\text{in }C^N\quad\text{as }n\to\infty. \end{equation} Let $W_u:=\mathcal G_u(F_u)\subset \overline B_\infty(0,1)$ be the unstable graph of $F_u$. Recalling the definition~\eqref{e:gmap-def} of $\Phi_uF$, we have $$ \varphi(W_u)\cap \{|x_1|\leq 1\}=\mathcal G_u(\Phi_u F_u)=W_u. $$ It follows that $\varphi(W_u)\cap \overline B_\infty(0,1)=W_u$, giving~\eqref{e:graph-inv}. Next, we see from~\eqref{e:lulla} that $F(0)=\partial_{x_1}F(0)=0$ implies $\partial_{x_1}(\Phi_u F)(0)=0$. Using~\eqref{e:contramap-u} with $F_0\equiv 0$, we get $\partial_{x_1}F_u(0)=0$. By Lemma~\ref{l:derest-1} we have the following derivative bounds on $F_u$, $F_s$: \begin{equation} \label{e:F-u-derb} \|F_u\|_{C^{N,1}}\leq C\delta,\quad \|F_s\|_{C^{N,1}}\leq C\delta. \end{equation} This implies that $W_u\cap W_s=\{0\}$. Indeed, if $(x_1,x_2)\in W_u\cap W_s$, then $x_2=F_u(x_1)$ and $x_1=F_s(x_2)$. Therefore, $x_1=F_s(F_u(x_1))$. However~\eqref{e:F-u-derb} implies that \begin{equation} \label{e:F-comp-contra} \sup_{[-1,1]}|\partial_{x_1}(F_s\circ F_u)|\leq C\delta. \end{equation} Therefore $F_s\circ F_u:[-1,1]\to[-1,1]$ is a contraction and the equation $x_1=F_s(F_u(x_1))$ has only one solution, $x_1=0$. \Remark The above proof shows that $F_u\in C^{N,1}$. One can prove in fact that $F_u\in C^{N+1}$, see~\cite[\S6.2.d, Step~5]{KaHa}. \subsection{Further properties} \label{s:base-4} In this section we prove the following theorem which relates the manifolds $W_u,W_s$ to the behavior of large iterates $\varphi^n$ of the map $\varphi$: \begin{theo} \label{t:stun-2} Let $\varphi$ be as in Theorem~\ref{t:stun-1} and $\delta$ be small enough depending only on $N$. Let $W_u,W_s$ be defined in~\eqref{e:W-u-def}. Then: \begin{enumerate} \item If $w\in W_u$ then $\varphi^{-n}(w)\to 0$ as $n\to\infty$, more precisely \begin{equation} \label{e:stun-2-1} |\varphi^{-n}(w)|\leq \Big({1\over 2}+C\delta\Big)^n|w|\quad\text{ for all }n\geq 0. \end{equation} \item If $w\in W_s$ then $\varphi^{n}(w)\to 0$ as $n\to\infty$, more precisely \begin{equation} \label{e:stun-2-2} |\varphi^{n}(w)|\leq \Big({1\over 2}+C\delta\Big)^n|w|\quad\text{ for all }n\geq 0. \end{equation} \item If $w\in \overline B_\infty(0,1)$ satisfies $\varphi^{-n}(w)\in \overline B_\infty(0,1)$ for all $n\geq 0$, then $w\in W_u$. \item If $w\in \overline B_\infty(0,1)$ satisfies $\varphi^{n}(w)\in \overline B_\infty(0,1)$ for all $n\geq 0$, then $w\in W_s$. \end{enumerate} \end{theo} \Remark Theorem~\ref{t:stun-2} implies the following dynamical characterization of the unstable/stable manifolds $W_u,W_s$: \begin{equation} \label{e:stun-char} \begin{aligned} w\in W_u&\quad\Longleftrightarrow\quad \varphi^{-n}(w)\in \overline B_\infty(0,1)\quad\text{for all }n\geq 0;\\ w\in W_s&\quad\Longleftrightarrow\quad \varphi^{n}(w)\in \overline B_\infty(0,1)\quad\text{for all }n\geq 0. \end{aligned} \end{equation} See Figure~\ref{f:dynch}. We only give the proof of parts~(1) and~(3) in Theorem~\ref{t:stun-2}. Parts~(2) and~(4), characterizing the stable manifold, are proved similarly, replacing $\varphi$ by $\varphi^{-1}$ and switching the roles of $x_1$ and~$x_2$. \begin{figure} \includegraphics[scale=0.115]{matfig/bfill0.jpg} \includegraphics[scale=0.115]{matfig/bfill1.jpg} \includegraphics[scale=0.115]{matfig/bfill2.jpg} \includegraphics[scale=0.115]{matfig/bfill3.jpg} \includegraphics[scale=0.115]{matfig/bfill4.jpg} \includegraphics[scale=0.115]{matfig/bfill5.jpg} \hbox to\hsize{\hss $n=0$ \hss\hss $n=1$ \hss\hss $n=2$ \hss\hss $n=3$ \hss\hss $n=4$ \hss\hss $n=5$ \hss} \caption{The sets of points $w$ such that $w,\varphi^{-1}(w),\dots,\varphi^{(-n)}(w)\in\overline B_\infty(0,1)$. By~\eqref{e:stun-char}, in the limit $n\to\infty$ we obtain $W_u$.} \label{f:dynch} \end{figure} Part~(1) of Theorem~\ref{t:stun-2} follows by iteration (putting $y:=w$, $\tilde y:=0$) from \begin{lemm} \label{l:dyndyn-1} Let $y,\tilde y\in W_u$ and put $x:=\varphi^{-1}(y)$, $\tilde x:=\varphi^{-1}(\tilde y)$. Then (see Figure~\ref{f:exc-1}) \begin{equation} \label{e:dyndyn-1} |x-\tilde x|\leq \Big({1\over 2}+C\delta\Big)|y-\tilde y|. \end{equation} \end{lemm} \begin{proof} By~\eqref{e:graph-inv} we have $x,\tilde x\in W_u\subset \overline B_\infty(0,1)$. We write $$ x=(x_1,F_u(x_1)),\quad \tilde x=(\tilde x_1,F_u(\tilde x_1)),\quad y=(y_1,F_u(y_1)),\quad \tilde y=(\tilde y_1,F_u(\tilde y_1)). $$ Since $\|F_u\|_{C^1}\leq C\delta$ by~\eqref{e:F-u-derb}, we have \begin{align} \label{e:dyndyn-1.1} |x-\tilde x|&\leq (1+C\delta)|x_1-\tilde x_1|,\\ \label{e:dyndyn-1.2} |y-\tilde y|&\geq (1-C\delta)|y_1-\tilde y_1|. \end{align} We have $y_1=\varphi_1(x_1,F_u(x_1))$ and $\tilde y_1=\varphi_1(\tilde x_1,F_u(\tilde x_1))$. Thus by~\eqref{e:A-prop} \begin{equation} \label{e:dyndyn-1.3} y_1-\tilde y_1=2(x_1-\tilde x_1)+\mathcal O(\delta)|x_1-\tilde x_1|. \end{equation} Together~\eqref{e:dyndyn-1.1}--\eqref{e:dyndyn-1.3} give~\eqref{e:dyndyn-1}. \end{proof} \begin{figure} \includegraphics[scale=0.35]{matfig/bstab1.jpg} \includegraphics[scale=0.35]{matfig/bstab2.jpg} \caption{The points $x,\tilde x,y,\tilde y$ from Lemma~\ref{l:dyndyn-1}. The curve is $W_u$.} \label{f:exc-1} \end{figure} It remains to show part~(3) of Theorem~\ref{t:stun-2}. For a point $w=(w_1,w_2)\in \overline B_\infty(0,1)$, define the distance from it to the unstable manifold by \begin{equation} \label{e:unstable-distance} d(w,W_u):=|w_2-F_u(w_1)|. \end{equation} The key component of the proof is \begin{lemm} \label{l:dyndyn-2} Assume that $w\in \overline B_\infty(0,1)$ and $\varphi(w)\in\overline B_\infty(0,1)$. Then (see Figure~\ref{f:exc-2}) \begin{equation} \label{e:dyndyn-2} d(\varphi(w),W_u)\leq \Big({1\over 2}+C\delta\Big)d(w,W_u). \end{equation} \end{lemm} \begin{proof} We write $$ w=(w_1,w_2),\quad z:=\varphi(w)=(z_1,z_2). $$ Define $$ x:=(w_1,F_u(w_1)),\quad y:=\varphi(x)=(y_1,F_u(y_1)). $$ (Since it might happen that $y_1\notin [-1,1]$, strictly speaking we extend the function $F_u$ to a larger interval by making $\varphi(W_u)$ the graph of the extended $F_u$. The resulting function still satisfies the bound~\eqref{e:F-u-derb}.) By~\eqref{e:A-prop} we have $$ z-y=\begin{pmatrix}2 & 0 \\ 0 & 1/2\end{pmatrix}(w-x)+\mathcal O(\delta)|w-x|. $$ Since $|w-x|=d(w,W_u)$ this implies \begin{align} \label{e:frap-1} z_1-y_1&=\mathcal O(\delta)d(w,W_u),\\ \label{e:frap-2} z_2-F_u(y_1)&={1\over 2}(w_2-F_u(w_1))+\mathcal O(\delta)d(w,W_u). \end{align} It follows from~\eqref{e:F-u-derb} and~\eqref{e:frap-1} that $$ F_u(z_1)-F_u(y_1)=\mathcal O(\delta)d(w,W_u). $$ From here and~\eqref{e:frap-2} we obtain $$ z_2-F_u(z_1)={1\over 2}(w_2-F_u(w_1))+\mathcal O(\delta)d(w,W_u). $$ Since $d(w,W_u)=|w_2-F_u(w_1)|$ and $d(z,W_u)=|z_2-F_u(z_1)|$ this implies~\eqref{e:dyndyn-2}. \end{proof} \begin{figure} \includegraphics[scale=0.35]{matfig/bstab3.jpg} \includegraphics[scale=0.35]{matfig/bstab4.jpg} \caption{An illustration of Lemma~\ref{l:dyndyn-2}. The curve is $W_u$ and the picture on the right is the image of the picture on the left under $\varphi$. The blue segment on the left has length $d(w,W_u)$ and the red segment on the right has length $d(z,W_u)$.} \label{f:exc-2} \end{figure} We now finish the proof of part~(3) of Theorem~\ref{t:stun-2}. Assume that $w\in \overline B_\infty(0,1)$ and $$ w^{(n)}:=\varphi^{-n}(w)\in \overline B_\infty(0,1)\quad\text{for all }n\geq 0. $$ Since $w^{(n-1)}=\varphi(w^{(n)})$, by Lemma~\ref{l:dyndyn-2} for small enough $\delta$ we have \begin{equation} \label{e:dd3-1} d(w^{(n-1)},W_u)\leq {2\over 3}d(w^{(n)},W_u). \end{equation} Since $d(w^{(n)},W_u)\leq 2$ for all $n\geq 0$, we iterate this to get \begin{equation} \label{e:dd3-2} d(w,W_u)\leq 2\cdot \Big({2\over 3}\Big)^nd(w^{(n)},W_u)\leq \Big({2\over 3}\Big)^n\quad\text{for all }n\geq 0 \end{equation} which implies $d(w,W_u)=0$ and thus $w\in W_u$. \Remark The above proof in fact gives stronger versions of parts~(3) and~(4) of Theorem~\ref{t:stun-2}: if $0\leq \sigma\leq 1$ and $n\geq 0$, then \begin{align} \label{e:redoer-1} w,\varphi^{-1}(w),\dots,\varphi^{-n}(w)\in \overline B_\infty(0,\sigma)\quad\Longrightarrow\quad d(w,W_u)\leq (2/3)^n\cdot 2\sigma,\\ \label{e:redoer-2} w,\varphi(w),\dots,\varphi^{n}(w)\in \overline B_\infty(0,\sigma)\quad\Longrightarrow\quad d(w,W_s)\leq (2/3)^n\cdot 2\sigma. \end{align} Here we define $d(w,W_s):=|w_1-F_s(w_2)|$ similarly to~\eqref{e:unstable-distance}. Another version of~\eqref{e:redoer-1}, \eqref{e:redoer-2} is available using the following estimate: \begin{equation} \label{e:redoer-3} \begin{aligned} |w|\leq 4\big(d(w,W_u)+d(w,W_s)\big)\quad\text{for all}\quad w\in \overline B_\infty(0,1). \end{aligned} \end{equation} To prove~\eqref{e:redoer-3} we use~\eqref{e:F-u-derb} and~\eqref{e:F-comp-contra}: $$ \begin{aligned} {1\over 2}|w_1|&\leq |w_1-F_s(F_u(w_1))|\leq |w_1-F_s(w_2)|+|F_s(w_2)-F_s(F_u(w_1))| \\&\leq |w_1-F_s(w_2)|+|w_2-F_u(w_1)|=d(w,W_s)+d(w,W_u) \end{aligned} $$ and $|w_2|$ is estimated similarly. Combining~\eqref{e:redoer-1}--\eqref{e:redoer-3} we get the following bound: for $n,r\geq 0$ and $0\leq\sigma\leq 1$ \begin{equation} \label{e:redoer-4} \begin{gathered} \varphi^{-n}(w),\dots,\varphi^{-1}(w),w,\varphi(w),\dots,\varphi^r(w)\in \overline B_\infty(0,\sigma)\\\Longrightarrow\quad |w|\leq \big((2/3)^n+(2/3)^r\big)\cdot 8\sigma. \end{gathered} \end{equation} The bound~\eqref{e:redoer-4} can be interpreted as `long time strict convexity': if the trajectory of a point $w$ stays in the ball $\overline B_\infty(0,1)$ for long positive and negative times, then $w$ is close to 0. \section{The general setting} \label{s:general} In this section we explain how to extend the proofs of Theorems~\ref{t:stun-1} and~\ref{t:stun-2} to general families of hyperbolic transformations, yielding the general Theorem~\ref{t:stun-adv} stated in~\S\ref{s:reduce-5}. Rather than give a complete formal proof we explain below several generalizations of Theorems~\ref{t:stun-1} and~\ref{t:stun-2} which together give Theorem~\ref{t:stun-adv}. We refer the reader to~\cite[Theorem~6.2.8]{KaHa} for a detailed proof. \subsection{Making $\delta$ small by rescaling} \label{s:reduce-1} We first show how to arrange for the assumption~(3) in~\S\ref{s:base-1} (that is, $\varphi$ being close to its linearization $d\varphi(0)$) to hold by a rescaling argument. Assume that $\varphi:U_\varphi\to V_\varphi$ is a $C^{N+1}$ map and it satisfies assumptions~(1)--(2) in~\S\ref{s:base-1}, namely \begin{equation} \label{e:stay-put} \varphi(0)=0,\quad d\varphi(0)=\begin{pmatrix} 2 & 0 \\ 0 & 1/2\end{pmatrix}. \end{equation} Fix small $\delta_1>0$ and consider the rescaling map $$ T:\overline B_\infty(0,1)\to \overline B_\infty(0,\delta_1),\quad T(x)=\delta_1 x. $$ We conjugate $\varphi$ by $T$ to get the map $$ \widetilde\varphi:=T^{-1}\circ\varphi\circ T:\overline B_\infty (0,1)\to \mathbb R^2. $$ The map $\widetilde\varphi$ still satisfies~\eqref{e:stay-put}, and its higher derivatives are given by $$ \partial^\alpha\widetilde\varphi(x)=\delta_1^{|\alpha|-1}\partial^\alpha\varphi(\delta_1 x). $$ Therefore, $\widetilde\varphi$ satisfies the assumption~(3) with $\delta=C\delta_1$, where $C$ depends on $\varphi$. By the Inverse Mapping Theorem, $\widetilde\varphi$ also satisfies the assumption~(4); that is, it is a diffeomorphism $\widetilde U_\varphi\to \widetilde V_\varphi$ for some open sets $\widetilde U_\varphi,\widetilde V_\varphi$ containing $\overline B_\infty(0,1)$. It follows that for $\delta_1$ small enough (depending on $\varphi$) Theorems~\ref{t:stun-1} and~\ref{t:stun-2} apply to $\widetilde\varphi$, giving the unstable/stable manifolds $\widetilde W_u,\widetilde W_s$. The manifolds $$ W_{u,\delta_1}:=T(\widetilde W_u),\quad W_{s,\delta_1}:=T(\widetilde W_s) $$ satisfy the conclusions of Theorems~\ref{t:stun-1} and~\ref{t:stun-2} for $\varphi$ with the ball $\overline B_\infty(0,1)$ replaced by $\overline B_\infty(0,\delta_1)$. We call these the ($\delta_1$-)\emph{local unstable/stable manifolds} of $\varphi$ at~$0$. \subsection{General expansion/contraction rates} \label{s:reduce-2} We next explain why Theorems~\ref{t:stun-1}--\ref{t:stun-2} hold if the condition~\eqref{e:stay-put} is replaced by \begin{equation} \label{e:stay-put-2} \varphi(0)=0,\quad d\varphi(0)=\begin{pmatrix} \mu & 0 \\ 0 & \lambda\end{pmatrix}\quad\text{where } 0<\lambda<1<\mu\text{ are fixed.} \end{equation} Note that the value of $\delta$ for which Theorems~\ref{t:stun-1} and~\ref{t:stun-2} apply will depend on $\lambda,\mu$, in particular it will go to~0 if $\lambda\to 1$ or $\mu\to 1$. The proofs in~\S\ref{s:base} apply with the following changes: \begin{itemize} \item in the proof of Lemma~\ref{l:gmap}, \eqref{e:G-1-der} is replaced by $$ \partial_{x_1}G_1(x_1)=\mu+\mathcal O(\delta)>1\quad\text{for all }x_1\in [-1,1]; $$ \item in Lemma~\ref{l:derform}, we have $$ L_k(x_1,\tau_0,\dots,\tau_k)=\lambda\mu^{-k}\tau_k+\mathcal O(\delta) $$ and the estimate~\eqref{e:derform-2} is changed accordingly; \item in the estimates~\eqref{e:derest-1.1}, \eqref{e:derest-1.2}, and~\eqref{e:derest-2} the constant ${1\over 4}$ is replaced by $\lambda\over\mu$; \item in the contraction property~\eqref{e:contraction-achieved} the constant $1\over 3$ is replaced by any fixed number in the interval $({\lambda\over\mu},1)$; \item in the estimate~\eqref{e:stun-2-1} in Theorem~\ref{t:stun-2}, as well as in Lemma~\ref{l:dyndyn-1}, the constant $1\over 2$ is replaced by~$\mu^{-1}$; \item in the estimate~\eqref{e:stun-2-2} in Theorem~\ref{t:stun-2} the constant ${1\over 2}$ is replaced by~$\lambda$; \item in Lemma~\ref{l:dyndyn-2} the constant ${1\over 2}$ is replaced by~$\lambda$; \item in the estimates~\eqref{e:dd3-1} and~\eqref{e:dd3-2}, as well as in~\eqref{e:redoer-1}, the constant ${2\over 3}$ is replaced by any fixed number in the interval $(\lambda,1)$; \item in~\eqref{e:redoer-2}, the constant $2\over 3$ is replaced by any fixed number in the interval $(\mu^{-1},1)$; \item in~\eqref{e:redoer-4}, the conclusion becomes $|w|\leq (\tilde\lambda^n+\tilde\mu^{-r})\cdot 8\sigma$ where $\tilde\lambda\in (\lambda,1)$ and $\tilde\mu\in (1,\mu)$ are fixed. \end{itemize} \subsection{Higher dimensions} \label{s:reduce-3} We now generalize Theorems~\ref{t:stun-1} and~\ref{t:stun-2} to the case of higher dimensions. More precisely, consider a diffeomorphism $$ \varphi:U_\varphi\to V_\varphi,\quad U_\varphi,V_\varphi\subset \mathbb R^d,\quad \overline B_\infty(0,1)\subset U_\varphi\cap V_\varphi, $$ where $d=d_u+d_s$, we write elements of $\mathbb R^d$ as $(x_1,x_2)$ with $x_1\in \mathbb R^{d_u}$, $x_2\in\mathbb R^{d_s}$, and (with $|\bullet|$ denoting the Euclidean norm) \begin{equation} \label{e:B-infty-new} \overline B_\infty(0,r):=\{(x_1,x_2)\in\mathbb R^d\colon \max(|x_1|,|x_2|)\leq r\}. \end{equation} The condition~\eqref{e:stay-put} is replaced by \begin{equation} \label{e:stay-put-3} \varphi(0)=0,\quad d\varphi(0)=\begin{pmatrix} A_1 & 0 \\ 0 & A_2 \end{pmatrix}, \end{equation} where $A_1:\mathbb R^{d_u}\to\mathbb R^{d_u}$, $A_2:\mathbb R^{d_s}\to\mathbb R^{d_s}$ are linear isomorphisms satisfying \begin{equation} \label{e:A-j-ineq} \|A_1^{-1}\|\leq \mu^{-1},\quad \|A_2\|\leq \lambda,\quad \max(\|A_1\|,\|A_2^{-1}\|)\leq C_0 \end{equation} for some fixed constants $\lambda,\mu,C_0$ such that $0<\lambda<1<\mu$. We assume that the bounds~\eqref{e:hypas-2} on higher derivatives still hold for some small $\delta>0$. The definitions~\eqref{e:stun-graphs} of the unstable/stable graphs still apply with the following adjustments. Define the unstable/stable balls \begin{equation} \label{e:stab-the-ball} \overline B_u(0,1):=\{x_1\in\mathbb R^{d_u}\colon |x_1|\leq 1\},\quad \overline B_s(0,1):=\{x_2\in\mathbb R^{d_s}\colon |x_2|\leq 1\}. \end{equation} Then for a $C^N$ function $F:\overline B_u(0,1)\to\mathbb R^{d_s}$, its unstable graph $\mathcal G_u(F)$ is a $d_u$-dimensional submanifold (with boundary) of $\mathbb R^d$. If instead $F:\overline B_s(0,1)\to\mathbb R^{d_u}$, then the stable graph $\mathcal G_s(F)$ is a $d_s$-dimensional submanifold of $\mathbb R^d$. Theorems~\ref{t:stun-1} and~\ref{t:stun-2} still hold for the map $\varphi$, with the constant $\delta$ now depending on $d,\lambda,\mu,C_0$ and the constants in~\eqref{e:stun-2-1} and~\eqref{e:stun-2-2} modified as in~\S\ref{s:reduce-2}. The proofs in~\S\ref{s:base} need to be modified as follows (in addition to the changes described in~\S\ref{s:reduce-2}): \begin{itemize} \item in Lemma~\ref{l:gmap}, the invertibility of $G_1$ and the fact that the image of $G_1$ contains $\overline B_u(0,1)$ follow from the standard (contraction mapping principle) proof of the Inverse Mapping Theorem, with the estimate~\eqref{e:G-1-der} replaced by $$ \|\partial_{x_1}G_1(x_1)-A_1\|=\mathcal O(\delta)\quad\text{for all }x_1\in \overline B_u(0,1); $$ \item we use the notation $\mathbf D^k_{x_1} F(x_1)=(\partial^\alpha_{x_1}F(x_1))_{|\alpha|=k}$, giving a vector in a finite-dimensional space $\mathcal V^k$ which is identified with the space of homogeneous polynomials in $d_u$ variables with values in $\mathbb R^{d_s}$ using the operation $$ \mathbf D^k_{x_1} F(x_1)\cdot v:=\sum_{|\alpha|=k}\binom{k}{\alpha}\partial^\alpha_{x_1}F(x_1)v^\alpha =\partial_t|_{t=0}(F(x_1+tv)),\quad v\in\mathbb R^{d_u}; $$ \item we use the following norms on $\mathcal V^k$: \begin{equation} \label{e:new-norms} \|\mathbf D^k_{x_1} F(x_1)\|:=\sup\big\{|\mathbf D^k_{x_1}F(x_1)\cdot v|\colon v\in\mathbb R^{d_u},\ |v|=1\big\}; \end{equation} \item in Lemma~\ref{l:derform}, the derivative bounds in~\eqref{e:gmap-2} are now on $\|\mathbf D^j_{x_1} F(x_1)\|$ for $j=1,\dots,k$; \item in Lemma~\ref{l:derform}, the derivative formula~\eqref{e:derform} is replaced by $$ \mathbf D^k_{x_1}(\Phi_u F)(y_1)=L_k(x_1,F(x_1),\mathbf D^1_{x_1}F(x_1),\dots,\mathbf D^k_{x_1}F(x_1)),\quad x_1:=G_1^{-1}(y_1) $$ where $L_k(x_1,\tau_0,\tau_1,\dots,\tau_k)\in\mathcal V^k$ is defined for $x_1\in \overline B_{\mathbb R^{d_u}}(0,1)$, $\tau_0\in\overline B_{\mathbb R^{d_s}}(0,1)$ and $\tau_j\in \mathcal V^j$, $\|\tau_j\|\leq 1$ for $j=1,\dots, k$; \item in Lemma~\ref{l:derform}, the approximation for $L_k$ is changed to the following: $$ L_k(x_1,\tau_0,\tau_1,\dots,\tau_k)\cdot v =A_2\big(\tau_k\cdot(A_1^{-1}v)\big)+\mathcal O(\delta)\quad\text{for all } v\in\mathbb R^{d_u},\ |v|=1 $$ and the derivative estimates~\eqref{e:derform-2} generalize naturally to the case of higher dimensions; \item in the proof of Lemma~\ref{l:derform}, the formula~\eqref{e:lulla} is replaced by $$ L_1(x_1,\tau_0,\tau_1)=(A_{21}(x_1,\tau_0)+A_{22}(x_1,\tau_0)\tau_1)(A_{11}(x_1,\tau_0)+A_{12}(x_1,\tau_0)\tau_1)^{-1} $$ where $A_{jk}(x_1,\tau_0)$ and $\tau_1$ are now matrices; \item in the definitions~\eqref{e:C-k-norm} and~\eqref{e:C-k-1-norm} of $\|F\|_{C^k}$ and~$\|F\|_{C^{k,1}}$ we use the norms~\eqref{e:new-norms} for the derivatives $\mathbf D^j_{x_1}F$, and we have the revised inequality $\|F\|_{C^{k,1}}\leq C\|F\|_{C^{k+1}}$; \item in the proof of Lemma~\ref{l:dyndyn-1}, the equation~\eqref{e:dyndyn-1.3} is replaced by $$ y_1-\tilde y_1=A_1(x_1-\tilde x_1)+\mathcal O(\delta)|x_1-\tilde x_1|; $$ \item in the proof of Lemma~\ref{l:dyndyn-2}, the equation~\eqref{e:frap-2} is replaced by $$ z_2-F_u(y_1)=A_2(w_2-F_u(w_1))+\mathcal O(\delta)d(w,W_u). $$ \end{itemize} \subsection{Iterating different transformations} \label{s:reduce-4} We next discuss a generalization of Theorems~\ref{t:stun-1} and~\ref{t:stun-2} from the case of a single map $\varphi$ to a $\mathbb Z$-indexed family of maps. More precisely, we assume that $$ \varphi_m:U_\varphi\to V_\varphi,\quad m\in\mathbb Z $$ is a family of maps each of which satisfies the assumptions in~\S\ref{s:reduce-3} uniformly in $m$, that is the constants $\lambda,\mu,C_0,\delta$ are independent of~$m$. The linear maps $A_1,A_2$ in~\eqref{e:stay-put-3} are allowed to depend on~$m$. We explain how the construction of the unstable manifold in Theorem~\ref{t:stun-1} generalizes to the case of a family of transformations. The case of stable manifolds is handled similarly, and Theorem~\ref{t:stun-2} generalizes naturally to this setting, see~\S\ref{s:reduce-5} below. Instead of a single function $F_u$ we construct a family of functions $$ F^u_m:\overline B_u(0,1)\to\overline B_s(0,1),\quad m\in\mathbb Z;\quad F^u_m(0)=0,\quad dF^u_m(0)=0. $$ Denote the graphs $W^u_m:=\mathcal G_u(F^u_m)$. Then the invariance property~\eqref{e:graph-inv} generalizes to \begin{equation} \label{e:graph-inv-2} \varphi_m(W^u_m)\cap \overline B_\infty(0,1)=W^u_{m+1}. \end{equation} To construct $F^u_m$ we use the graph transform $\Phi^u_m$ of the map $\varphi_m$ defined in Lemma~\ref{l:gmap}. This transform satisfies the derivative bounds of Lemmas~\ref{l:derest-1}--\ref{l:derest-2} uniformly in $m$. As in~\S\ref{s:base-3}, to show~\eqref{e:graph-inv-2} it suffices to construct $F^u_m$ such that for all $m\in\mathbb Z$, \begin{equation} \label{e:graph-inv-2.1} F^u_m\in\mathcal X_N,\quad \Phi^u_mF^u_m=F^u_{m+1}. \end{equation} To do this we modify the argument of~\S\ref{s:base-3} as follows: consider the space $$ \mathcal X_N^{\mathbb Z}:=\{(F_m)_{m\in\mathbb Z}\mid F_m\in\mathcal X_N\text{ for all }m\in\mathbb Z\} $$ with the metric $$ d_N^{\mathbb Z}\big((F_m),(\widetilde F_m)\big):=\sup_{m\in\mathbb Z}d_N(F_m,\widetilde F_m). $$ Then $(\mathcal X_N^{\mathbb Z},d_N^{\mathbb Z})$ is a complete metric space. Consider the map on $\mathcal X_N^{\mathbb Z}$ $$ \Phi_u^{\mathbb Z}:(F_m)\mapsto (\widehat F_m),\quad \widehat F_{m+1}:=\Phi^u_mF_m. $$ It follows from Lemma~\ref{l:derest-2} that $\Phi_u^{\mathbb Z}$ is a contracting map on $(\mathcal X_N^{\mathbb Z},d_N^{\mathbb Z})$. Applying the Contraction Mapping Principle, we obtain a fixed point $$ (F^u_m)\in\mathcal X_N^{\mathbb Z},\quad \Phi_u^{\mathbb Z}(F^u_m)=(F^u_m) $$ which satisfies~\eqref{e:graph-inv-2.1}, finishing the proof. \subsection{The general Stable/Unstable Manifold Theorem} \label{s:reduce-5} We finally combine the generalizations in~\S\S\ref{s:reduce-2}--\ref{s:reduce-4} and state the general version of the Stable/Unstable Manifold Theorem. We assume that: \begin{enumerate} \item $d=d_u+d_s$, $d_u,d_s\geq 0$, elements of $\mathbb R^d$ are written as $(x_1,x_2)$ where $x_1\in\mathbb R^{d_u}$, $x_2\in\mathbb R^{d_s}$, we use the Euclidean norm on $\mathbb R^d$, and $\overline B_\infty(0,1)$, $\overline B_u(0,1)$, $\overline B_s(0,1)$ are defined by~\eqref{e:B-infty-new}, \eqref{e:stab-the-ball}; \item we are given a family of $C^{N+1}$ diffeomorphisms (here $N\geq 1$ is fixed) \begin{equation} \label{e:gst-1} \varphi_m:U_\varphi\to V_\varphi,\quad m\in\mathbb Z;\quad U_\varphi,V_\varphi\subset\mathbb R^d,\quad \overline B_\infty(0,1)\subset U_\varphi\cap V_\varphi; \end{equation} \item we have for all $m\in\mathbb Z$ \begin{equation} \label{e:gst-2} \varphi_m(0)=0,\quad d\varphi_m(0)(x_1,x_2)=(A_{1,m}x_1,A_{2,m}x_2) \end{equation} where the linear maps $A_{1,m}:\mathbb R^{d_u}\to\mathbb R^{d_u}$, $A_{2,m}:\mathbb R^{d_s}\to\mathbb R^{d_s}$ satisfy \begin{equation} \label{e:gst-3} \|A_{1,m}^{-1}\|\leq \mu^{-1},\quad \|A_{2,m}\|\leq \lambda,\quad \max(\|A_{1,m}\|,\|A_{2,m}^{-1}\|)\leq C_0 \end{equation} for some constants $0<\lambda<1<\mu$, $C_0> 0$; \item we have the derivative bounds for some $\delta>0$ \begin{equation} \label{e:gst-4} \sup_{U_\varphi}|\partial^\alpha \varphi_m|\leq \delta\quad\text{for all}\quad m\in\mathbb Z,\quad 2\leq|\alpha|\leq N+1. \end{equation} \end{enumerate} Note that the stable/unstable spaces at~$0$ are now given by $$ E_u(0):=\{(x_1,0)\mid x_1\in\mathbb R^{d_u}\},\quad E_s(0):=\{(0,x_2)\mid x_2\in\mathbb R^{d_s}\}. $$ The general form of Theorems~\ref{t:stun-1} and~\ref{t:stun-2} is then \begin{theo} \label{t:stun-adv} There exists $\delta>0$ depending only on $d,N,\lambda,\mu,C_0$ such that the following holds. Assume that $\varphi_m$ satisfy assumptions~(1)--(4) of the present section. Then there exist families of $C^N$ functions \begin{equation} \label{e:fufsy} \begin{aligned} F^u_m:\overline B_u(0,1)\to\overline B_s(0,1),&\quad F^s_m:\overline B_s(0,1)\to\overline B_u(0,1),\\ F^u_m(0)=0,\quad dF^u_m(0)=0,&\quad F^s_m(0)=0,\quad dF^s_m(0)=0 \end{aligned} \end{equation} bounded in $C^N$ uniformly in $m$ and such that the graphs \begin{equation} \label{e:wuwsy} \begin{aligned} W^u_m&:=\{(x_1,x_2)\colon |x_1|\leq 1,\ x_2=F^u_m(x_1)\},\\ W^s_m&:=\{(x_1,x_2)\colon |x_2|\leq 1,\ x_1=F^s_m(x_2)\} \end{aligned} \end{equation} have the following properties for all $m\in\mathbb Z$: \begin{enumerate} \item $\varphi_m^{-1}(W^u_{m+1})\subset W^u_m$ and $\varphi_m(W^s_m)\subset W^s_{m+1}$, more precisely \begin{equation} \label{e:stuna-1} \varphi_m(W^u_m)\cap \overline B_\infty(0,1)=W^u_{m+1},\quad \varphi_m^{-1}(W^s_{m+1})\cap \overline B_\infty(0,1)=W^s_m; \end{equation} \item $W^u_m\cap W^s_m=\{0\}$; \item for each $x\in W^u_m$ we have $\varphi_{m-n}^{-1}\cdots\varphi_{m-1}^{-1}(x)\to 0$ as $n\to\infty$; \item for each $x\in W^s_m$ we have $\varphi_{m+n-1}\cdots\varphi_m(x)\to 0$ as $n\to\infty$; \item if $\varphi_{m-n}^{-1}\cdots\varphi_{m-1}^{-1}(x)\in \overline B_\infty(0,1)$ for all $n\geq 0$, then $x\in W^u_m$; \item if $\varphi_{m+n-1}\cdots\varphi_m(x)\in\overline B_\infty(0,1)$ for all $n\geq 0$, then $x\in W^s_m$. \end{enumerate} \end{theo} \Remarks 1. Similarly to~\eqref{e:stun-char}, we obtain the following dynamical definition of the manifolds $W^u_m,W^s_m$: \begin{equation} \label{e:stun-char-2} \begin{aligned} x\in W^u_m&\quad\Longleftrightarrow\quad \varphi_{m-n}^{-1}\cdots\varphi_{m-1}^{-1}(x)\in \overline B_\infty(0,1)\quad\text{for all }n\geq 0;\\ x\in W^s_m&\quad\Longleftrightarrow\quad \varphi_{m+n-1}\cdots\varphi_{m}(x)\in \overline B_\infty(0,1)\quad\text{for all }n\geq 0. \end{aligned} \end{equation} \noindent 2. Similarly to Theorem~\ref{t:stun-2} and Lemma~\ref{l:dyndyn-1}, parts~(3) and~(4) of Theorem~\ref{t:stun-adv} can be made quantitative as follows. Fix $\tilde\lambda,\tilde\mu$ such that \begin{equation} \label{e:tildes} 0<\lambda<\tilde\lambda<1<\tilde\mu<\mu. \end{equation} Then for $\delta$ small enough depending on $d,\lambda,\tilde\lambda,\mu,\tilde\mu,C_0$ and all $m\in\mathbb Z$, $n\geq 0$ we have \begin{equation} \label{e:stuna-2} \begin{aligned} x,\tilde x\in W^u_m&\quad\Longrightarrow\quad |\varphi_{m-n}^{-1}\cdots\varphi_{m-1}^{-1}(x)- \varphi_{m-n}^{-1}\cdots\varphi_{m-1}^{-1}(\tilde x)|\leq \tilde\mu^{-n}|x-\tilde x|,\\ x,\tilde x\in W^s_m&\quad\Longrightarrow\quad |\varphi_{m+n-1}\cdots\varphi_{m}(x)- \varphi_{m+n-1}\cdots\varphi_{m}(\tilde x)|\leq \tilde\lambda^{n}|x-\tilde x|. \end{aligned} \end{equation} \noindent 3. Similarly to the remark at the end of~\S\ref{s:base-4}, there is a quantitative version of parts~(5) and~(6) of Theorem~\ref{t:stun-adv} as well. Namely, if $\tilde\lambda,\tilde\mu$ satisfy~\eqref{e:tildes} and $\delta$ is small enough depending on $d,\lambda,\tilde\lambda,\mu,\tilde\mu,C_0$, then for all $0\leq\sigma\leq 1$ and $n\geq 0$ \begin{equation} \label{e:stuna-3} \begin{aligned} \varphi_{m-\ell}^{-1}\cdots\varphi_{m-1}^{-1}(x)\in \overline B_\infty(0,\sigma),\quad\ell=0,1,\dots,n &\quad\Longrightarrow\quad d(x,W_u)\leq \tilde\lambda^n\cdot 2\sigma;\\ \varphi_{m+\ell-1}\cdots\varphi_{m}(x)\in \overline B_\infty(0,\sigma),\quad\ell=0,1,\dots,n &\quad\Longrightarrow\quad d(x,W_s)\leq \tilde\mu^{-n}\cdot 2\sigma. \end{aligned} \end{equation} The following analog of the estimate~\eqref{e:redoer-4} holds: for all $n,r\geq 0$ and $0\leq\sigma\leq 1$ \begin{equation} \label{e:stuna-4} \begin{gathered} \text{if}\quad \varphi_{m-\ell}^{-1}\cdots\varphi_{m-1}^{-1}(x)\in \overline B_\infty(0,\sigma),\quad \ell=0,1,\dots,n\\ \text{and}\quad \varphi_{m+\ell-1}\cdots\varphi_{m}(x)\in \overline B_\infty(0,\sigma),\quad\ell=0,1,\dots,r\\ \text{then}\quad |x|\leq \big(\tilde\lambda^n+\tilde\mu^{-r}\big)\cdot 8\sigma. \end{gathered} \end{equation} \noindent 4. The rescaling argument of~\S\ref{s:reduce-1} applies to the setting of Theorem~\ref{t:stun-adv}. That is, if $\varphi_m$ satisfy~\eqref{e:gst-1}--\eqref{e:gst-3}, then one can make~\eqref{e:gst-4} hold by zooming in to a sufficiently small neighborhood of the origin. \subsection{The case of expansion/contraction rate~1} \label{s:reduce-6} For applications to hyperbolic flows (which have a neutral direction) we also need to discuss which parts of Theorem~\ref{t:stun-adv} still hold when either $\lambda$ or $\mu$ is equal to~1. Specifically, we replace the condition $\lambda<1<\mu$ with \begin{equation} \label{e:degrade-1} \lambda=1<\mu. \end{equation} That is, there is still expansion in the unstable directions but there does not have to be (strict) contraction in the stable directions. The construction of the unstable manifolds $W^u_m$ applies to the case~\eqref{e:degrade-1} without any changes, and the resulting manifolds $W^u_m$ satisfy conclusions~(1) and~(3) of Theorem~\ref{t:stun-adv}. (In fact, this would work under an even weaker condition $\mu>\max(1,\lambda)$.) The estimate~\eqref{e:stuna-2} still holds for $W^u_m$, assuming that $\tilde\mu$ satisfies $1<\tilde\mu<\mu$. However, we cannot construct the stable manifolds $W^s_m$ under the assumption~\eqref{e:degrade-1}. The problem is in the proof of Lemma~\ref{l:gmap}: the projection of $\varphi^{-1}(\mathcal G_s(F))$ onto the $x_2$ variable might no longer cover the unit ball, so the function $\Phi_sF$ cannot be defined. Moreover, conclusion~(5) of Theorem~\ref{t:stun-adv} (as well as~\eqref{e:stuna-3}) no longer holds, so the dynamical characterization~\eqref{e:stun-char-2} of the unstable manifolds $W^u_m$ is no longer valid. The `strict convexity' property~\eqref{e:stuna-4} no longer holds. Similarly, one can still construct the stable manifolds $W^s_m$ and establish conclusions~(1) and~(4) of Theorem~\ref{t:stun-adv} for them if we replace the condition $\lambda<1<\mu$ with \begin{equation} \label{e:degrade-2} \lambda<1=\mu. \end{equation} \section{Hyperbolic maps and flows} \label{s:maps-and-flows} In this section we apply Theorem~\ref{t:stun-adv} to obtain the Stable/Unstable Manifold Theorem for hyperbolic maps (Theorem~\ref{t:stun-maps} in~\S\ref{s:hyp-maps}) and for hyperbolic flows (Theorem~\ref{t:stun-flows} in~\S\ref{s:hyp-flows}). This involves constructing an adapted metric (Lemma~\ref{l:adapted-metric}) and taking adapted coordinates to bring the map/flow into the model case handled by Theorem~\ref{t:stun-adv}. \subsection{Hyperbolic maps} \label{s:hyp-maps} Assume that $M$ is a $d$-dimensional manifold without boundary and $d=d_u+d_s$ where $d_u,d_s\geq 0$. Let $\varphi:M\to M$ be a $C^{N+1}$ diffeomorphism (here $N\geq 1$ is fixed) and assume that $\varphi$ is hyperbolic on some compact $\varphi$-invariant set $K\subset M$ in the following sense: \begin{defi} \label{d:hyp-map} Let $K\subset M$ be a compact set such that $\varphi(K)=K$. We say that $\varphi$ is \textbf{hyperbolic} on~$K$ if there exists a splitting \begin{equation} \label{e:hyp-map-1} T_x M=E_u(x)\oplus E_s(x),\quad x\in K \end{equation} where $E_u(x),E_s(x)\subset T_x M$ are subspaces of dimensions $d_u,d_s$ and: \begin{itemize} \item $E_u,E_s$ are invariant under $d\varphi$, namely \begin{equation} \label{e:hyp-map-2} d\varphi(x)E_u(x)=E_u(\varphi(x)),\quad d\varphi(x)E_s(x)=E_s(\varphi(x))\quad\text{for all }x\in K; \end{equation} \item large negative iterates of $\varphi$ are contracting on $E_u$, namely there exist constants $C>0$, $0<\lambda<1$ such that for some Riemannian metric $|\bullet|$ on $M$ we have \begin{equation} \label{e:hyp-map-u} |d\varphi^{-n}(x)v|\leq C\lambda^n |v|\quad\text{for all}\quad v\in E_u(x),\ x\in K,\ n\geq 0; \end{equation} \item large positive iterates of $\varphi$ are contracting on $E_s$, namely \begin{equation} \label{e:hyp-map-s} |d\varphi^{n}(x)v|\leq C\lambda^n |v|\quad\text{for all}\quad v\in E_s(x),\ x\in K,\ n\geq 0. \end{equation} \end{itemize} \end{defi} \Remarks 1. The contraction properties~\eqref{e:hyp-map-u}, \eqref{e:hyp-map-s} do not depend on the choice of the metric on $M$, though the constant $C$ (but not $\lambda$) will depend on the metric. Later in Lemma~\ref{l:adapted-metric} we construct metrics which give~\eqref{e:hyp-map-u}, \eqref{e:hyp-map-s} with $C=1$. \noindent 2. We do not assume a priori that the maps $x\mapsto E_u(x),E_s(x)$ are continuous. However we show in~\S\ref{s:continuity} below that Definition~\ref{d:hyp-map} implies that these maps are in fact H\"older continuous. \noindent 3. The basic example of a hyperbolic set is $K=\{x_0\}$ where $x_0\in M$ is a fixed point of~$\varphi$ which is hyperbolic, namely $d\varphi(x_0)$ has no eigenvalues on the unit circle. More generally one can take as $K$ a hyperbolic closed trajectory of $\varphi$. The opposite situation is when $K=M$; in this case $\varphi$ is called an \emph{Anosov diffeomorphism}. \noindent 4. Similarly to~\S\ref{s:reduce-5} we could take two different constants in~\eqref{e:hyp-map-u}, \eqref{e:hyp-map-s}, corresponding to different minimal contraction rates in the unstable and the stable directions. We do not do this here to simplify notation, and since the examples in this note have time-reversal symmetry and thus equal stable/unstable contraction rates. Fix a Riemannian metric on $M$ which induces a distance function $d(\bullet,\bullet)$; denote for $x\in M$ and $r\geq 0$ \begin{equation} \label{e:B-d-def} \overline B_d(x,r):=\{y\in M\mid d(x,y)\leq r\}. \end{equation} We now state the Stable/Unstable Manifold Theorem for hypebolic maps: \begin{theo} \label{t:stun-maps} Assume that $\varphi$ is hyperbolic on $K\subset M$. Then for each $x\in K$ there exist \textbf{local unstable/stable manifolds} $$ W_u(x),W_s(x)\subset M $$ which have the following properties for some $\varepsilon_0>0$ depending only on $\varphi,K$: \begin{enumerate} \item $W_u(x),W_s(x)$ are $C^N$ embedded disks of dimensions $d_u,d_s$, that is images of closed balls in $\mathbb R^{d_u}$, $\mathbb R^{d_s}$ under $C^N$ embeddings, and the $C^N$ norms of these embeddings are bounded uniformly in $x$; \item $W_u(x)\cap W_s(x)=\{x\}$ and $T_xW_u(x)=E_u(x)$, $T_x W_s(x)=E_s(x)$; \item the boundaries of $W_u(x),W_s(x)$ do not intersect $\overline B_d(x,\varepsilon_0)$; \item $\varphi^{-1}(W_u(x))\subset W_u(\varphi^{-1}(x))$ and $\varphi(W_s(x))\subset W_s(\varphi(x))$; \item for each $y\in W_u(x)$, we have $d(\varphi^{-n}(y),\varphi^{-n}(x))\to 0$ as $n\to\infty$; \item for each $y\in W_s(x)$, we have $d(\varphi^n(y),\varphi^n(x))\to 0$ as $n\to\infty$; \item if $y\in M$ and $d(\varphi^{-n}(y),\varphi^{-n}(x))\leq \varepsilon_0$ for all $n\geq 0$, then $y\in W_u(x)$; \item if $y\in M$ and $d(\varphi^n(y),\varphi^n(x))\leq \varepsilon_0$ for all $n\geq 0$, then $y\in W_s(x)$; \item if $x,y\in K$ and $d(x,y)\leq\varepsilon_0$ then $W_s(x)\cap W_u(y)$ consists of exactly one point. \end{enumerate} \end{theo} \Remarks 1. Similarly to~\eqref{e:stuna-2} there are quantitative versions of the statements~(5) and~(6): if we fix $\tilde\lambda$ such that $\lambda<\tilde\lambda<1$ then for all $n\geq 0$ and $x\in K$ \begin{equation} \label{e:cream-1} \begin{aligned} y,\tilde y\in W_u(x)&\quad\Longrightarrow\quad d(\varphi^{-n}(y),\varphi^{-n}(\tilde y))\leq C\tilde\lambda^n d(y,\tilde y);\\ y,\tilde y\in W_s(x)&\quad\Longrightarrow\quad d(\varphi^{n}(y),\varphi^{n}(\tilde y))\leq C\tilde\lambda^n d(y,\tilde y). \end{aligned} \end{equation} where $C$ is a constant depending only on $\varphi,K,\tilde\lambda$. Here and in Remark~2 below the manifolds $W_u,W_s$ and the constant $\varepsilon_0$ depend on $\tilde\lambda$, in particular when $\tilde\lambda\to \lambda$ the stable/unstable manifolds might degenerate to a point and $\varepsilon_0$ might go to 0. (One can get rid of this dependence, but it is rather tedious and typically unnecessary.) \noindent 2. Similarly to~\eqref{e:stuna-3} there are quantitative versions of the statements~(7) and~(8) as well: if we fix $\tilde\lambda$ as before then for all $n\geq 0$, $0\leq\sigma\leq\varepsilon_0$, $x\in K$, and $y\in M$ \begin{equation} \label{e:cream-2} \begin{aligned} d(\varphi^{-\ell}(y),\varphi^{-\ell}(x))\leq \sigma\text{ for all }\ell=0,\dots,n &\quad\Longrightarrow\quad d(y,W_u(x))\leq C\tilde\lambda^n\sigma,\\ d(\varphi^{\ell}(y),\varphi^{\ell}(x))\leq \sigma\text{ for all }\ell=0,\dots,n &\quad\Longrightarrow\quad d(y,W_s(x))\leq C\tilde\lambda^n\sigma \end{aligned} \end{equation} where $C$ is a constant depending only on $\varphi,K,\tilde\lambda$. We also have the following analog of the `strict convexity' property~\eqref{e:stuna-4}: for all $n\geq 0$, $0\leq\sigma\leq\varepsilon_0$, $x\in K$, and $y\in M$ \begin{equation} \label{e:cream-3} d(\varphi^\ell(y),\varphi^\ell(x))\leq\sigma\quad\text{for all}\quad|\ell|\leq n \quad\Longrightarrow\quad d(y,x)\leq C\tilde\lambda^n\sigma. \end{equation} \begin{figure} \includegraphics[scale=0.2875]{matfig/catstun1.jpg} \includegraphics[scale=0.2875]{matfig/catstun2.jpg} \caption{Numerically computed unstable (blue) and stable (red) manifolds for a perturbed Arnold cat map on the torus $\mathbb R^2/\mathbb Z^2$. On the left are the local stable/unstable manifolds $W_u(x),W_s(x)$ for several choices of $x$. On the right are the manifolds $W_u^{(k)}(x),W_s^{(k)}(x)$ for $x=(0.5,0.5)$ and $k=4$.} \label{f:stun-global} \end{figure} \noindent 3. The manifolds $W_u(x)$ are not defined canonically since they have a somewhat arbitrarily determined boundary. For the same reason, if $W_u(x)\cap W_u(y)\neq \emptyset$, this does not imply that $W_u(x)=W_u(y)$. However, we can show that in this case $W_u(x)$ and $W_u(y)$ are subsets of the same $d_u$-dimensional manifold, see~\eqref{e:intersector} below. For $k\geq 0$ and $x\in K$ define \begin{equation} \label{e:iterated-unstable} W_u^{(k)}(x):=\varphi^k(W_u(\varphi^{-k}(x))). \end{equation} This is still a $d_u$-dimensional embedded disk in $M$. Moreover, by statement~(4) in Theorem~\ref{t:stun-maps} we have $$ W_u(x)=W_u^{(0)}(x)\subset W_u^{(1)}(x)\subset\dots\subset W_u^{(k)}(x)\subset W_u^{(k+1)}(x)\subset\dots $$ There exists $k_0\geq 0$ such that for all $x,y\in K$ \begin{equation} \label{e:intersector} W_u(x)\cap W_u(y)\neq\emptyset\quad\Longrightarrow\quad W_u(x)\cup W_u(y)\subset W_u^{(k_0)}(x). \end{equation} Indeed, assume that $z\in W_u(x)\cap W_u(y)$. By~\eqref{e:cream-1} if $k_0$ is large enough then $$ d(\varphi^{-k}(x),\varphi^{-k}(y)) \leq d(\varphi^{-k}(z),\varphi^{-k}(x)) +d(\varphi^{-k}(z),\varphi^{-k}(y)) \leq{\varepsilon_0\over 2}\quad\text{for all }k\geq k_0. $$ Let $w\in W_u(x)\cup W_u(y)$. Then for $k_0$ large enough we get from~\eqref{e:cream-1} \begin{equation} \label{e:prada} d(\varphi^{-k}(w),\varphi^{-k}(x))\leq \varepsilon_0\quad\text{for all }k\geq k_0. \end{equation} It follows from statement~(7) in Theorem~\ref{t:stun-maps} that $\varphi^{-k_0}(w)\in W_u(\varphi^{-k_0}(x))$ and thus $w\in W_u^{(k_0)}(x)$, proving~\eqref{e:intersector}. Note that~\eqref{e:intersector} implies that the tangent spaces to $W_u(x)$ are the unstable spaces: \begin{equation} \label{e:tangentor} y\in W_u(x)\cap K\quad\Longrightarrow\quad T_y(W_u(x))=E_u(y). \end{equation} The above discussion applies to stable manifolds where we define \begin{equation} \label{e:iterated-stable} W_s^{(k)}(x):=\varphi^{-k}(W_s(\varphi^k(x))). \end{equation} \noindent 4. Here is another version of `local uniqueness' of stable/unstable manifolds: there exists $\varepsilon_1>0$ such that for all $x,y\in K$ we have \begin{align} \label{e:locally-unique-1} W_u(x)\cap W_u(y)\neq\emptyset&\quad\Longrightarrow\quad W_u(y)\cap \overline B_d(x,\varepsilon_1)\subset W_u(x),\\ \label{e:locally-unique-2} W_s(x)\cap W_s(y)\neq\emptyset&\quad\Longrightarrow\quad W_s(y)\cap \overline B_d(x,\varepsilon_1)\subset W_s(x). \end{align} We show~\eqref{e:locally-unique-1}, with~\eqref{e:locally-unique-2} proved similarly. Let $w\in W_u(y)\cap \overline B_d(x,\varepsilon_1)$. By~\eqref{e:prada} we have $d(\varphi^{-k}(w),\varphi^{-k}(x))\leq\varepsilon_0$ for all $k\geq k_0$. On the other hand $$ d(\varphi^{-k}(w),\varphi^{-k}(x))\leq C d(w,x)\leq C\varepsilon_1\quad \text{for}\quad 0\leq k <k_0. $$ Choosing $\varepsilon_1$ small enough we get $d(\varphi^{-k}(w),\varphi^{-k}(x))\leq\varepsilon_0$ for all $k\geq 0$, which by statement~(7) in Theorem~\ref{t:stun-maps} gives $w\in W_u(x)$ as needed. \noindent 5. One can take the unions of the manifolds~\eqref{e:iterated-unstable}, \eqref{e:iterated-stable} to obtain \emph{global unstable/stable manifolds}: for $x\in K$, \begin{equation} \label{e:global} W_u^{(\infty)}(x):=\bigcup_{k\geq 0}W_u^{(k)}(x),\quad W_s^{(\infty)}(x):=\bigcup_{k\geq 0}W_s^{(k)}(x). \end{equation} By statements (5)--(8) in Theorem~\ref{t:stun-maps} we can characterize these dynamically as follows: \begin{equation} \label{e:global-char} \begin{aligned} y\in W_u^{(\infty)}(x)&\quad\Longleftrightarrow\quad d(\varphi^{-n}(y),\varphi^{-n}(x))\to 0\quad\text{as }n\to\infty;\\ y\in W_s^{(\infty)}(x)&\quad\Longleftrightarrow\quad d(\varphi^{n}(y),\varphi^{n}(x))\to 0\quad\text{as }n\to\infty. \end{aligned} \end{equation} Therefore the global stable/unstable manifolds are disjoint: if $W_u^{(\infty)}(x)\cap W_u^{(\infty)}(y)\neq\emptyset$, then $W_u^{(\infty)}(x)=W_u^{(\infty)}(y)$, and same is true for $W_s^{(\infty)}$. The sets $W_u^{(\infty)}(x)$ and~$W_s^{(\infty)}(x)$ are $d_u$ and $d_s$-dimensional immersed submanifolds without boundary in $M$, however they are typically not embedded. In fact, in many cases these submanifolds are dense in $M$. See Figure~\ref{f:stun-global}. \subsection{Continuity of the stable/unstable spaces} \label{s:continuity} In this section we study the regularity of the maps $x\mapsto E_u(x),E_s(x)$. To talk about these, it is convenient to introduce the Grassmanians $$ \begin{aligned} \mathscr G_u&:=\{(x,E)\colon x\in M,\ E\subset T_xM\text{ is a $d_u$-dimensional subspace}\},\\ \mathscr G_s&:=\{(x,E)\colon x\in M,\ E\subset T_xM\text{ is a $d_s$-dimensional subspace}\}. \end{aligned} $$ These are smooth manifolds fibering over $M$. If $\varphi$ is hyperbolic on $K$, then we have the maps \begin{equation} \label{e:stun-mapps} E_u:K\to\mathscr G_u,\quad E_s:K\to\mathscr G_s. \end{equation} We first show that $E_u(x),E_s(x)$ depend continuously on $x$: \begin{lemm} \label{l:contin} Assume that $\varphi$ is hyperbolic on $K$. Then the maps~\eqref{e:stun-mapps} are continuous. \end{lemm} \begin{proof} We show continuity of $E_s$, the continuity of $E_u$ is proved similarly. It suffices to show that if $$ x_k\in K,\quad x_k\to x_\infty,\quad v_k\in E_s(x_k),\quad |v_k|=1,\quad v_k\to v_\infty\in T_{x_\infty}M $$ then $v_\infty\in E_s(x_\infty)$. By~\eqref{e:hyp-map-s} we have for all $k$ and all $n\geq 0$ $$ |d\varphi^n(x_k)v_k|\leq C\lambda^n. $$ Passing to the limit $k\to\infty$, we get for all $n\geq 0$ $$ |d\varphi^n(x_\infty)v_\infty|\leq C\lambda^n. $$ Given~\eqref{e:hyp-map-1} and~\eqref{e:hyp-map-u} this implies that $v_\infty\in E_s(x_\infty)$, finishing the proof. \end{proof} Lemma~\ref{l:contin} and the fact that $E_u(x),E_s(x)$ are transverse for each $x$ implies that $E_u(x)$, $E_s(x)$ are uniformly transverse, namely there exists a constant $C$ such that \begin{equation} \label{e:transverse} \max\big(|v_u|,|v_s|\big)\leq C|v_u+v_s|\quad\text{for all}\quad v_u\in E_u(x),\ v_s\in E_s(x),\ x\in K. \end{equation} A quantitative version of the proof of Lemma~\ref{l:contin} shows that in fact $E_u(x),E_s(x)$ are H\"older continuous in $x$. (This statement is not used in the rest of these notes.) \begin{lemm} \label{l:holder} Fix some smooth metrics on $\mathscr G_u,\mathscr G_s$. Assume that $\varphi$ is hyperbolic on $K$. Then there exists $\gamma>0$ such that the maps~\eqref{e:stun-mapps} have $C^\gamma$ regularity. \end{lemm} \begin{proof} In this proof we denote by $C$ constants which only depend on $\varphi,K$. We show H\"older continuity of $E_s$, with the case of $E_u$ handled similarly. Let $\overline TM$ be the fiber-radial compactification of $TM$, which is a manifold with interior $TM$ and boundary diffeomorphic to the sphere bundle $SM$, with the boundary defining function $|v|^{-1}$. Denote by $\Phi$ the action of $d\varphi$ on $TM$: $$ \Phi(x,v):=(\varphi(x),d\varphi(x)v),\quad (x,v)\in TM. $$ Note that $\Phi^n(x,v)=(\varphi^n(x),d\varphi^n(x)v)$. Moreover, since $\Phi$ is homogeneous with respect to dilations in the fibers of $TM$, it extends to a smooth map on $\overline TM$. Fix a Riemannian metric on $\overline TM$ (smooth up to the boundary) and denote by $d_{\overline TM}$ the corresponding distance function. Then there exists a constant $\Lambda\geq 1$ depending only on $\varphi,K$ such that \begin{equation} \label{e:holder-1} d_{\overline TM}\big(\Phi(x,v),\Phi(y,w)\big)\leq \Lambda\cdot d_{\overline TM}\big((x,v),(y,w)\big)\quad\text{for}\quad x,y\in K. \end{equation} We will show that the map $E_s$ is H\"older continuous with exponent $$ \gamma:=-{2\log\lambda\over\log(\Lambda/\lambda)}>0. $$ To do this, assume that $$ x,y\in K,\quad v\in E_s(x),\quad w\in T_yM,\quad |v|=1,\quad d_{\overline TM}\big((x,v),(y,w)\big)\leq Cd(x,y). $$ We write $$ w=w_u+w_s,\quad w_u\in E_u(y),\quad w_s\in E_s(y). $$ Then it suffices to prove that \begin{equation} \label{e:holder-goal} |w_u|\leq Cd(x,y)^\gamma. \end{equation} To show~\eqref{e:holder-goal}, we use the following bound true for all $n\geq 0$: \begin{equation} \label{e:holder-2} \begin{aligned} d_{\overline TM}\big(\Phi^n(y,w),(\varphi^n(x),0)\big) &\leq C\Lambda^n d(x,y)+d_{\overline TM}\big(\Phi^n(x,v),(\varphi^n(x),0)\big)\\ &\leq C\Lambda^n d(x,y)+C\lambda^n \end{aligned} \end{equation} where the first inequality uses~\eqref{e:holder-1} iterated $n$ times and the second one uses \eqref{e:hyp-map-s}. Assume that $d(x,y)$ is small and choose $$ n:=\Big\lfloor -{\log d(x,y)\over\log(\Lambda/\lambda)}\Big\rfloor\geq 0. $$ Then $\Lambda^n d(x,y)\leq \lambda^n$, so~\eqref{e:holder-2} implies $$ d_{\overline TM}\big(\Phi^n(y,w),(\varphi^n(x),0)\big) \leq C\lambda^n. $$ If $d(x,y)$ is small, then $n$ is large, thus $\Phi^n(y,w)$ is close to the zero section. Therefore \begin{equation} \label{e:holder-3} |d\varphi^n(y)w|\leq C\lambda^n. \end{equation} Recalling the decomposition $w=w_u+w_s$ and using the bounds~\eqref{e:hyp-map-u}, \eqref{e:hyp-map-s}, and~\eqref{e:transverse}, we see that~\eqref{e:holder-3} implies $$ |w_u|\leq C\lambda^n |d\varphi^n(y)w_u|\leq C\lambda^n|d\varphi^n(y)w|+C\lambda^{2n} \leq C\lambda^{2n}\leq Cd(x,y)^\gamma. $$ This gives~\eqref{e:holder-goal}, finishing the proof. \end{proof} \subsection{Adapted metrics} \label{s:adapted} In preparation for the proof of Theorem~\ref{t:stun-maps}, we show that there exist Riemannian metrics on $M$ which are adapted to the map $\varphi$: \begin{lemm} \label{l:adapted-metric} Assume that $\varphi$ is hyperbolic on $K$. Fix $\tilde\lambda$ such that $\lambda<\tilde\lambda<1$ where $\lambda$ is given in Definition~\ref{d:hyp-map}. Then there exist $C^N$ Riemannian metrics $|\bullet|_u,|\bullet|_s$ on~$M$ such that \begin{align} \label{e:hyp-map-u-1} |d\varphi^{-1}(x)v|_u\leq \tilde\lambda |v|_u&\quad\text{for all}\quad v\in E_u(x),\ x\in K;\\ \label{e:hyp-map-s-1} |d\varphi(x)v|_s\leq \tilde\lambda |v|_s&\quad\text{for all}\quad v\in E_s(x),\ x\in K. \end{align} \end{lemm} \Remark Iterating~\eqref{e:hyp-map-u-1}, \eqref{e:hyp-map-s-1} we get analogs of~\eqref{e:hyp-map-u}, \eqref{e:hyp-map-s} with $C=1$. \begin{proof} We first construct the metric $|\bullet|_s$. Let $|\bullet|$ be a Riemannian metric on~$M$. Take large fixed $m$ to be chosen later and define the Riemannian metric $|\bullet|_s$ by $$ |v|_s^2:=\sum_{n=0}^{m-1} \tilde\lambda^{-2n}|d\varphi^n(x)v|^2,\quad x\in M,\quad v\in T_xM. $$ Assume that $x\in K$ and $v\in E_s(x)$. Then $$ \begin{aligned} |d\varphi(x)v|_s^2&=\sum_{n=0}^{m-1}\tilde\lambda^{-2n}|d\varphi^{n+1}(x)v|^2 =\sum_{n=1}^m \tilde\lambda^{2-2n}|d\varphi^n(x)v|^2\\ &=\tilde\lambda^2\big(|v|_s^2-|v|^2+\tilde\lambda^{-2m}|d\varphi^m(x)v|^2\big)\\ &\leq \tilde\lambda^2\big(|v|_s^2-|v|^2+C\lambda^{2m}\tilde\lambda^{-2m}|v|^2\big) \end{aligned} $$ where in the last inequality we used~\eqref{e:hyp-map-s} and $C$ is a constant depending only on $\varphi,K$. Since $\tilde\lambda>\lambda$, choosing $m$ large enough we can guarantee that $C\lambda^{2m}\tilde\lambda^{-2m}\leq 1$, thus~\eqref{e:hyp-map-s-1} holds. The inequality~\eqref{e:hyp-map-u-1} is proved similarly, using the metric $$ |v|_u^2:=\sum_{n=0}^{m-1} \tilde\lambda^{-2n}|d\varphi^{-n}(x)v|^2,\quad x\in M,\quad v\in T_xM.\qedhere $$ \end{proof} \subsection{Adapted charts} \label{s:adapted-coord} To reduce Theorem~\ref{t:stun-maps} to Theorem~\ref{t:stun-adv}, we introduce charts on~$M$ which are adapted to the map $\varphi$. We assume that $\varphi$ is hyperbolic on $K\subset M$, fix $\tilde\lambda\in (\lambda,1)$ and let $|\bullet|_u$, $|\bullet|_s$ be the Riemannian metrics on $M$ constructed in Lemma~\ref{l:adapted-metric}. We write the elements of $\mathbb R^d$ as $(x_1,x_2)$ where $x_1\in \mathbb R^{d_u}$, $x_2\in\mathbb R^{d_s}$. We use the canonical stable/unstable subspaces of $\mathbb R^d$ \begin{equation} \label{e:e-su-0} E_u(0):=\{(v_1,0)\mid v_1\in \mathbb R^{d_u}\},\quad E_s(0):=\{(0,v_2)\mid v_2\in\mathbb R^{d_s}\}. \end{equation} Recall from~\eqref{e:B-infty-new} the notation $$ \overline B_\infty(0,r)=\{(x_1,x_2)\in\mathbb R^d\colon \max(|x_1|,|x_2|)\leq r\}. $$ \begin{defi} \label{d:adapted-chart} Let $x_0\in K$. A diffeomorphism $$ \varkappa:U_\varkappa\to V_\varkappa,\quad x_0\in U_\varkappa\subset M,\quad 0\in V_\varkappa\subset \mathbb R^d $$ is called an \textbf{adapted chart} for $\varphi$ centered at $x_0$, if: \begin{enumerate} \item $\varkappa(x_0)=0$; \item $d\varkappa(x_0)E_u(x_0)=E_u(0)$ and the restriction of $d\varkappa(x_0)$ to $E_u(x_0)$ is an isometry from the metric $|\bullet|_u$ to the Euclidean metric; \item $d\varkappa(x_0)E_s(x_0)=E_s(0)$ and the restriction of $d\varkappa(x_0)$ to $E_s(x_0)$ is an isometry from the metric $|\bullet|_s$ to the Euclidean metric. \end{enumerate} \end{defi} For each $x_0\in K$, there exists an adapted chart for $\varphi$ centered at $x_0$. Moreover, it follows from uniform transversality~\eqref{e:transverse} of $E_u$, $E_s$ that we can select for each $x_0\in K$ an adapted chart for $\varphi$ centered at $x_0$ \begin{equation} \label{e:good-charts} \varkappa_{x_0}:U_{x_0}\to V_{x_0},\quad x_0\in U_{x_0}\subset M,\quad 0\in V_{x_0}\subset\mathbb R^d \end{equation} such that the set $\{\varkappa_{x_0}\mid x_0\in K\}$ is bounded in the class of $C^{N+1}$ charts, more precisely: \begin{enumerate} \item there exists $\delta_0>0$ such that $\overline B_\infty(0,\delta_0)\subset V_{x_0}$ for all $x_0\in K$; \item all order $\leq N+1$ derivatives of $\varkappa_{x_0}$ and $\varkappa_{x_0}^{-1}$ are bounded uniformly in $x_0$. \end{enumerate} Note that we do not require continuous dependence of $\varkappa_{x_0}$ on $x_0$, in fact in many cases such dependence is impossible because the bundles $E_u,E_s$ are not topologically trivial. We now study the action of the map $\varphi$ in adapted charts. For each $x_0\in K$, define the diffeomorphism $\psi_{x_0}$ of neighborhoods of $0$ in $\mathbb R^d$ by \begin{equation} \label{e:phi-adapted} \psi_{x_0}:=\varkappa_{\varphi(x_0)}\circ\varphi\circ \varkappa_{x_0}^{-1}. \end{equation} Note that the set $\{\psi_{x_0}\mid x_0\in K\}$ is bounded in the class of $C^{N+1}$ diffeomorphisms. From the definition of adapted charts and the fact that $E_u(x),E_s(x)$ are $\varphi$-invariant we see that \begin{equation} \label{e:butter-1} \psi_{x_0}(0)=0,\quad d\psi_{x_0}(0)=\begin{pmatrix} A_{1,x_0} & 0 \\ 0 & A_{2,x_0}\end{pmatrix} \end{equation} where $A_{1,x_0}:\mathbb R^{d_u}\to \mathbb R^{d_u}$, $A_{2,x_0}:\mathbb R^{d_s}\to\mathbb R^{d_s}$ are linear isomorphisms. Moreover, from the properties~\eqref{e:hyp-map-u-1}, \eqref{e:hyp-map-s-1} of the adapted metrics $|\bullet|_u$, $|\bullet|_s$ we get \begin{equation} \label{e:butter-3} \|A_{1,x_0}^{-1}\|\leq \tilde\lambda,\quad \|A_{2,x_0}\|\leq \tilde\lambda,\quad \max(\|A_{1,x_0}\|,\|A_{2,x_0}^{-1}\|)\leq C_0\quad\text{for all }x_0\in K \end{equation} for some fixed $C_0$, where $\|\bullet\|$ is the operator norm with respect to the Euclidean norm. \begin{figure} \includegraphics{stunnote.2} \caption{An illustration of the commutative diagram~\eqref{e:commdiag}. The blue/red lines are the unstable/stable subspaces of the tangent spaces.} \label{f:charts} \end{figure} To make the maps $\psi_{x_0}$ close to their linearizations $d\psi_{x_0}(0)$, we use the rescaling procedure introduced in~\S\ref{s:reduce-1}. Fix small $\delta_1>0$ to be chosen later (when we apply Theorem~\ref{t:stun-adv}) and consider the rescaling map $$ T:\mathbb R^d\to\mathbb R^d,\quad T(x)=\delta_1x. $$ Define the rescaled charts \begin{equation} \label{e:rescaled-chart} \widetilde\varkappa_{x_0}:=T^{-1}\circ\varkappa_{x_0},\quad x_0\in K \end{equation} and the corresponding maps \begin{equation} \label{e:phi-adapted-2} \widetilde\psi_{x_0}:=\widetilde\varkappa_{\varphi(x_0)}\circ\varphi\circ \widetilde\varkappa_{x_0}^{-1} =T^{-1}\circ\psi_{x_0}\circ T. \end{equation} That is, we have the commutative diagram (see also Figure~\ref{f:charts}) \begin{equation} \label{e:commdiag} \begin{tikzcd}[column sep=large, row sep=large] M \arrow[r, "\varphi"] \arrow[d, "\varkappa_{x_0}"] \arrow[dd, bend right,swap,"\widetilde\varkappa_{x_0}"] & M \arrow[d, swap, "\varkappa_{\varphi(x_0)}"] \arrow[dd, bend left, "\widetilde\varkappa_{\varphi(x_0)}"]\\ \mathbb R^d \arrow[r, "\psi_{x_0}"] \arrow[d, "T^{-1}"] & \mathbb R^d \arrow[d, swap, "T^{-1}"] \\ \mathbb R^d \arrow[r, "\widetilde\psi_{x_0}"] & \mathbb R^d \end{tikzcd}. \end{equation} We choose $\delta_1$ small enough so that $\overline B_{\infty}(0,1)$ is contained in both the domain and the range of each $\widetilde\psi_{x_0}$. The map $\widetilde\psi_{x_0}$ still satisfies~\eqref{e:butter-1}: \begin{equation} \label{e:butter-2} \widetilde\psi_{x_0}(0)=0,\quad d\widetilde\psi_{x_0}(0)=\begin{pmatrix} A_{1,x_0}&0\\ 0&A_{2,x_0}\end{pmatrix} \end{equation} where $A_{j,x_0}$ are the same transformations as in~\eqref{e:butter-1}. Moreover, as explained in~\S\ref{s:reduce-1}, for any given $\delta>0$, if we choose $\delta_1$ small enough depending on $\delta,\varphi,K$ (but not on $x_0$) then we have the derivative bounds \begin{equation} \label{e:delta-1-chosen} \sup |\partial^\alpha\widetilde\psi_{x_0}|\leq \delta\quad\text{for all}\quad x_0\in K,\quad 2\leq|\alpha|\leq N+1. \end{equation} In fact, one can take $\delta_1:=\delta/C$ where $C$ is some constant depending on $\varphi,K$. \subsection{Proof of Theorem~\ref{t:stun-maps}} \label{s:maps-proof} We now give the proof of the Stable/Unstable Manifold Theorem for hyperbolic maps. We fix $\tilde\lambda$ such that $\lambda<\tilde\lambda<1$. Let $\delta>0$ be the constant in Theorem~\ref{t:stun-adv} where we use $\tilde\lambda,\tilde\lambda^{-1}$ in place of $\lambda,\mu$ and $C_0$ is the constant from~\eqref{e:butter-3}. Choose the rescaling parameter $\delta_1>0$ in~\S\ref{s:adapted-coord} small enough so that~\eqref{e:delta-1-chosen} holds. We use the rescaled adapted charts $\widetilde\varkappa_{x_0}$, $x_0\in K$, defined in~\eqref{e:rescaled-chart}, and the maps $\widetilde\psi_{x_0}$ giving the action of $\varphi$ in these charts, defined in~\eqref{e:phi-adapted-2}. For every $x\in K$ and $m\in\mathbb Z$ denote \begin{equation} \label{e:maps-defi} \psi_{x,m}:=\widetilde\psi_{\varphi^m(x)}=\widetilde\varkappa_{\varphi^{m+1}(x)}\circ\varphi \circ\widetilde\varkappa^{-1}_{\varphi^m(x)}. \end{equation} The dynamics of the iterates of $\varphi$ near the trajectory $(\varphi^m(x))$ is conjugated by the charts $\widetilde\varkappa_{\varphi^m(x)}$ to the dynamics of the compositions of the maps $\psi_{x,m}$. We will prove Theorem~\ref{t:stun-maps} by applying Theorem~\ref{t:stun-adv} to the maps $\psi_{x,m}$ and pulling back the resulting stable/unstable manifolds by $\widetilde\varkappa_{\varphi^m(x)}$ to get the stable/unstable manifolds for $\varphi$. As shown in~\S\ref{s:adapted-coord}, for each $x\in K$ the sequence of maps $(\psi_{x,m})_{m\in\mathbb Z}$ satisfies the assumptions in~\S\ref{s:reduce-5}, where we use $\tilde\lambda,\tilde\lambda^{-1}$ in place of $\lambda,\mu$. We apply Theorem~\ref{t:stun-adv} to get the stable/unstable manifolds for this sequence, which we denote $$ W^u_{x,m},W^s_{x,m}\subset \overline B_\infty(0,1)\subset\mathbb R^d. $$ We define the unstable/stable manifolds for $\varphi$ at $x$ as follows: \begin{equation} \label{e:stun-maps-defined} W_u(x):=\widetilde\varkappa_{x}^{-1}(W^u_{x,0}),\quad W_s(x):=\widetilde\varkappa_{x}^{-1}(W^s_{x,0}). \end{equation} The unstable/stable manifolds at the iterates $\varphi^n(x)$ are given by \begin{equation} \label{e:works-well} W_u(\varphi^n(x))=\widetilde\varkappa_{\varphi^n(x)}^{-1}(W^u_{x,n}),\quad W_s(\varphi^n(x))=\widetilde\varkappa_{\varphi^n(x)}^{-1}(W^s_{x,n}),\quad n\in\mathbb Z. \end{equation} The statement~\eqref{e:works-well} is not a tautology since the left-hand sides were obtained by applying Theorem~\ref{t:stun-adv} to the sequence of maps $(\psi_{\varphi^n(x),m})_{m\in\mathbb Z}$ while the right-hand sides were obtained using the maps $(\psi_{x,m})_{m\in\mathbb Z}$. To prove~\eqref{e:works-well} we note that \eqref{e:maps-defi} implies $$ \psi_{\varphi^n(x),m}=\psi_{x,m+n}. $$ Therefore, the sequence $(\psi_{\varphi^n(x),m})_{m\in\mathbb Z}$ is just a shift of the sequence $(\psi_{x,m})_{m\in\mathbb Z}$. From the construction of the manifolds $W^u_{x,m}$, $W^s_{x,m}$ in~\S\ref{s:reduce-4} we see that $$ W^u_{\varphi^n(x),m}=W^u_{x,m+n},\quad W^s_{\varphi^n(x),m}=W^s_{x,m+n}. $$ Putting $m=0$ and recalling~\eqref{e:stun-maps-defined}, we get~\eqref{e:works-well}. We now show that the manifolds $W_u(x)$, $W_s(x)$ defined in~\eqref{e:stun-maps-defined} satisfy the statements~(1)--(8) in Theorem~\ref{t:stun-maps}. This is straightforward since the above construction effectively reduced Theorem~\ref{t:stun-maps} to Theorem~\ref{t:stun-adv}. \begin{itemize} \item[(1):] This follows from the definitions~\eqref{e:wuwsy} of $W^u_{x,m},W^s_{x,m}$. The uniform boundedness of the embeddings in $C^N$ follows from the fact that the functions $F^u_{x,m}$, $F^s_{x,m}$ used to define $W^u_{x,m}$, $W^s_{x,m}$ are bounded by~1 in $C^N$ norm, see~\S\ref{s:reduce-4} and~\eqref{e:X-N-def}. \smallskip \item[(2):] We have $W^u_{x,m}\cap W^s_{x,m}=\{0\}$ by statement~(2) in Theorem~\ref{t:stun-adv}. By~\eqref{e:fufsy} we have also $T_0 W^u_{x,m}=E_u(0)$, $T_0 W^s_{x,m}=E_s(0)$ where $E_u(0),E_s(0)\subset\mathbb R^d$ are defined in~\eqref{e:e-su-0}. It remains to use~\eqref{e:stun-maps-defined} and the fact that $d\widetilde\varkappa_x(x)$ maps $E_u(x),E_s(x)$ to $E_u(0),E_s(0)$ by Definition~\ref{d:adapted-chart} and~\eqref{e:rescaled-chart}. \smallskip \item[(3):] Every $w=(w_1,w_2)\in\partial W^u_{x,m}$ satisfies $|w_1|=1$, thus $|w|\geq 1$. The latter is also true for all $w\in \partial W^s_{x,m}$. It then suffices to choose $\varepsilon_0$ small enough so that \begin{equation} \label{e:eps-0-chosen} x\in K,\ d(x,y)\leq \varepsilon_0 \quad\Longrightarrow\quad |\widetilde\varkappa_x(y)|<1 \end{equation} which is possible since $\widetilde\varkappa_x(x)=0$ and $\widetilde\varkappa_x$ are bounded in $C^N$ uniformly in $x$. \smallskip \item[(4):] By statement~(1) in Theorem~\ref{t:stun-adv} we have \begin{equation} \label{e:yoga} \psi_{x,-1}^{-1}(W^u_{x,0})\subset W^u_{x,-1},\quad \psi_{x,0}(W^s_{x,0})\subset W^s_{x,1}. \end{equation} From~\eqref{e:maps-defi} we have $$ \psi_{x,-1}^{-1}=\widetilde\varkappa_{\varphi^{-1}(x)}\circ\varphi^{-1} \circ\widetilde\varkappa^{-1}_x,\quad \psi_{x,0}=\widetilde\varkappa_{\varphi(x)}\circ\varphi\circ\widetilde\varkappa^{-1}_x. $$ Applying $\widetilde\varkappa_{\varphi^{-1}(x)}^{-1}$ to the first statement in~\eqref{e:yoga} and $\widetilde\varkappa_{\varphi(x)}^{-1}$ to the second one and using~\eqref{e:works-well} we get $\varphi^{-1}(W_u(x))\subset W_u(\varphi^{-1}(x))$ and $\varphi(W_s(x))\subset W_s(\varphi(x))$ as needed. \smallskip \item[(5)--(6):] Define the closed neighborhoods of $x$ $$ \overline B_x:=\widetilde\varkappa_x^{-1}(\overline B_\infty(0,1)),\quad x\in K. $$ Note that $W_u(x)\cup W_s(x)\subset\overline B_x$. By~\eqref{e:maps-defi}, if $n\geq 1$ and $\varphi^{-\ell}(y)\in \overline B_{\varphi^{-\ell}(x)}$ for all $\ell=0,\dots,n-1$, then \begin{equation} \label{e:oid-1} \widetilde\varkappa_{\varphi^{-n}(x)}\big(\varphi^{-n}(y)\big) =\psi_{x,-n}^{-1}\cdots\psi_{x,-1}^{-1}(\widetilde\varkappa_x(y)). \end{equation} Similarly if $n\geq 1$ and $\varphi^\ell(y)\in \overline B_{\varphi^\ell(x)}$ for all $\ell=0,\dots,n-1$, then \begin{equation} \label{e:oid-2} \widetilde\varkappa_{\varphi^n(x)}\big(\varphi^n(y)\big) =\psi_{x,n-1}\cdots\psi_{x,0}(\widetilde\varkappa_x(y)). \end{equation} If $y\in W_u(x)$, then for all $\ell\geq 0$ we have $\varphi^{-\ell}(y)\in W_u(\varphi^{-\ell}(x))\subset \overline B_{\varphi^{-\ell}(x)}$. Moreover, $\widetilde\varkappa_x(y)\in W^u_{x,0}$ by~\eqref{e:stun-maps-defined}. Applying the statement~(3) in Theorem~\ref{t:stun-adv} with $m:=0$, we get $\psi^{-1}_{x,-n}\cdots\psi_{x,-1}^{-1}(\widetilde\varkappa_x(y))\to 0$ as $n\to\infty$, thus $d(\varphi^{-n}(y),\varphi^{-n}(x))\to 0$ by~\eqref{e:oid-1}. The case $y\in W_s(x)$ is handled similarly. \smallskip \item[(7)--(8):] We show (7), with~(8) proved similarly. Assume that $d(\varphi^{-n}(y),\varphi^{-n}(x))\leq\varepsilon_0$ for all $n\geq 0$. By~\eqref{e:eps-0-chosen} this implies that $\varphi^{-n}(y)\in \overline B_{\varphi^{-n}(x)}$ for all $n\geq 0$. Then by~\eqref{e:oid-1} we have $$ \psi_{x,-n}^{-1}\cdots\psi_{x,-1}^{-1}(\widetilde\varkappa_x(y))\in\overline B_\infty(0,1)\quad\text{for all }n\geq 0. $$ By statement~(5) in Theorem~\ref{t:stun-adv} with $m:=0$ we have $\widetilde\varkappa_x(y)\in W^u_{x,0}$ and thus $y\in W_u(x)$ by~\eqref{e:stun-maps-defined}. \end{itemize} The quantitative statements~\eqref{e:cream-1}--\eqref{e:cream-3} follow from~\eqref{e:stuna-2}--\eqref{e:stuna-4} similarly to the proofs of statements~(5)--(8) above. Finally, statement~(9) in Theorem~\ref{t:stun-maps} essentially follows from the continuous dependence of $E_u(x),E_s(x)$ on $x$ (Lemma~\ref{l:contin}) and the fact that $E_u(x)$, $E_s(x)$ are transversal to each other. We give a more detailed (straightforward but slightly tedious) explanation below. Recall from~\eqref{e:stun-maps-defined} that $W_u(x)=\widetilde\varkappa_x^{-1}(W_{x,0}^u)$ and $W_s(x)=\widetilde\varkappa_x^{-1}(W_{x,0}^s)$ where $\widetilde\varkappa_x$ is the rescaled adapted chart defined in~\eqref{e:rescaled-chart} and $W_{x,0}^u,W_{x,0}^s\subset\overline B_\infty(0,1)$ are the unstable/stable graphs constructed in Theorem~\ref{t:stun-adv}. Writing elements of $\mathbb R^d$ as $(w_1,w_2)$ where $w_1\in\mathbb R^{d_u}$, $w_2\in\mathbb R^{d_s}$ we have \begin{align} \label{e:yappy-1} \widetilde\varkappa_x(W_u(x))=W_{x,0}^u&=\{(w_1,w_2)\colon |w_1|\leq 1,\ w_2=F_{x,u}(w_1)\},\\ \label{e:yappy-2} \widetilde\varkappa_x(W_s(x))=W_{x,0}^s&=\{(w_1,w_2)\colon |w_2|\leq 1,\ w_1=F_{x,s}(w_2)\} \end{align} for some functions $F_{x,u}:\overline B_u(0,1)\to\overline B_s(0,1)$, $F_{x,s}:\overline B_s(0,1)\to\overline B_u(0,1)$, where $\overline B_u(0,1),\overline B_s(0,1)$ are defined in~\eqref{e:stab-the-ball}, such that (recalling~\eqref{e:fufsy}, \eqref{e:F-u-derb}, and~\eqref{e:delta-1-chosen}) \begin{equation} \label{e:fluffy} F_{x,u}(0)=0,\quad F_{x,s}(0)=0,\quad \max\big(\|F_{x,u}\|_{C^1},\|F_{x,s}\|_{C^1}\big)\leq C\delta_1; \end{equation} here $\delta_1>0$ is the rescaling parameter used in the definition~\eqref{e:rescaled-chart} of the chart $\widetilde\varkappa_x$. Now, we assume that $x,y\in K$ and $d(x,y)\leq \varepsilon_0$ where $\varepsilon_0>0$ is small, in particular $\varepsilon_0\ll \delta_1$. Then $W_u(y)$ is contained in the domain of the chart $\widetilde\varkappa_x$. The points in $W_s(x)\cap W_u(y)$ have the form $\widetilde\varkappa_x^{-1}(F_{x,s}(w_2),w_2)=\widetilde\varkappa_y^{-1}(w_1,F_{y,u}(w_1))$ where $w=(w_1,w_2)\in\overline B_\infty(0,1)$ solves the equation \begin{equation} \label{e:equator} G_{x,y}(w)=0,\quad G_{x,y}(w_1,w_2):= \widetilde\varkappa_{x,y}(w_1,F_{y,u}(w_1))-(F_{x,s}(w_2),w_2) \end{equation} where we put $$ \widetilde\varkappa_{x,y}:=\widetilde\varkappa_x\circ\widetilde\varkappa_y^{-1} =T^{-1}\circ\varkappa_{x,y}\circ T,\quad \varkappa_{x,y}:=\varkappa_x\circ\varkappa_y^{-1},\quad T:w\mapsto\delta_1 w, $$ and $\varkappa_x,\varkappa_y$ are the unrescaled charts defined in~\eqref{e:good-charts}. Since $\varkappa_x,\varkappa_y$ are adapted charts (see Definition~\ref{d:adapted-chart}), $d(x,y)\leq\varepsilon_0$, and $E_u(x),E_s(x)$ depend continuously on $x$, for small enough $\varepsilon_0$ we have $$ d\varkappa_{x,y}(0)=A_{x,y}+\mathcal O(\delta_1),\quad A_{x,y}:=\begin{pmatrix}A_{1,x,y} & 0 \\ 0 & A_{2,x,y} \end{pmatrix} $$ where $A_{1,x,y}:\mathbb R^{d_u}\to\mathbb R^{d_u}$, $A_{2,x,y}:\mathbb R^{d_s}\to\mathbb R^{d_s}$ are isometries. The rescaled change of coordinates map $\widetilde\varkappa_{x,y}$ then satisfies (assuming $\varepsilon_0\leq\delta_1^2$) $$ |\widetilde\varkappa_{x,y}(0)|\leq C\delta_1,\quad \sup_{w\in \overline B_\infty(0,1)} \|d\widetilde\varkappa_{x,y}(w)-A_{x,y}\|\leq C\delta_1. $$ Combining this with~\eqref{e:fluffy} we get $$ |G_{x,y}(0)|\leq C\delta_1,\quad \sup_{w\in\overline B_\infty(0,1)}\bigg\|dG_{x,y}(w)-\begin{pmatrix} A_{1,x,y}&0 \\ 0 & -I\end{pmatrix}\bigg\|\leq C\delta_1. $$ If $\delta_1$ is small enough, then by the Contraction Mapping Principle the equation~\eqref{e:equator} has a unique solution in $\overline B_\infty(0,1)$. Therefore, $W_s(x)\cap W_u(y)$ consists of a single point as required. \subsection{Hyperbolic flows} \label{s:hyp-flows} We finally consider the setting of hyperbolic flows. Let $M$ be a $d$-dimensional manifold without boundary and $d=d_u+d_s+1$ where $d_u,d_s\geq 0$. Let $$ \varphi^t=\exp(tX):M\to M,\quad t\in\mathbb R $$ be the flow generated by a $C^{N+1}$ vector field $X$ on $M$. For simplicity we assume that $\varphi^t$ is globally well-defined for all $t$, though in practice it is enough to require this in the neighborhood of the set $K$ below. We assume that $K\subset M$ is a $\varphi^t$-invariant hyperbolic set in the following sense: \begin{defi} \label{d:hyp-flows} Let $K\subset M$ be a compact set such that $\varphi^t(K)=K$ for all $t\in\mathbb R$. We say that the flow $\varphi^t$ is \textbf{hyperbolic} on $K$ if the generating vector field $X$ does not vanish on $K$ and there exists a splitting \begin{equation} \label{e:hypflow-split} T_xM=E_0(x)\oplus E_u(x)\oplus E_s(x),\quad x\in K,\quad E_0(x):=\mathbb R X(x), \end{equation} where $E_u(x),E_s(x)\subset T_x M$ are subspaces of dimensions $d_u,d_s$ and: \begin{itemize} \item $E_u,E_s$ are invariant under the flow, namely for all $x\in K$ and $t\in\mathbb R$ \begin{equation} \label{e:hypflow-inv} d\varphi^t(x)E_u(x)=E_u(\varphi^t(x)),\quad d\varphi^t(x)E_s(x)=E_s(\varphi^t(x)); \end{equation} \item $d\varphi^t$ is expanding on $E_u$ and contracting on $E_s$, namely there exist constants $C>0$, $\nu>0$ such that for some Riemannian metric $|\bullet|$ on $M$ and all $x\in K$ \begin{equation} \label{e:hypflow-exp} |d\varphi^t(x)v|\leq Ce^{-\nu|t|}\cdot|v|,\quad \begin{cases} v\in E_u(x),& t\leq 0;\\ v\in E_s(x),& t\geq 0. \end{cases} \end{equation} \end{itemize} \end{defi} \Remarks 1. The time-$t$ map $\varphi^t$ of a hyperbolic flow is \textbf{not} a hyperbolic map in the sense of Definition~\ref{d:hyp-map} because of the flow direction $E_0$. \noindent 2. The property~\eqref{e:hypflow-exp} does not depend on the choice of the metric on $M$, though the constant $C$ (but not $\nu$) will depend on the metric. \noindent 3. The basic example of a hyperbolic set is a closed trajectory $$ K=\{\varphi^t(x_0)\colon 0\leq t\leq T\}\quad\text{for some}\quad x_0\in M,\ T>0\quad\text{such that}\quad \varphi^T(x_0)=x_0 $$ which is hyperbolic, namely $d\varphi^T(x_0)$ has a simple eigenvalue~$1$ and no other eigenvalues on the unit circle. The opposite situation is when $K=M$; in this case $\varphi^t$ is called an \emph{Anosov flow}. An important class of Anosov flows are geodesic flows on negatively curved manifolds, discussed in~\S\ref{s:surf-neg} below. As in~\S\ref{s:hyp-maps}, fix a distance function $d(\bullet,\bullet)$ on $M$ and define the balls $\overline B_d(x,r)$ by~\eqref{e:B-d-def}. We now state the Stable/Unstable Manifold Theorem for hyperbolic flows, which is similar to the case of maps (Theorem~\ref{t:stun-maps}): \begin{theo} \label{t:stun-flows} Assume that the flow $\varphi^t$ is hyperbolic on $K\subset M$. Then for each $x\in K$ there exist \textbf{local unstable/stable manifolds} $$ W_u(x),W_s(x)\subset M $$ which have the following properties for some $\varepsilon_0>0$ depending only on $\varphi^t,K$: \begin{enumerate} \item $W_u(x),W_s(x)$ are $C^N$ embedded disks of dimensions $d_u,d_s$, and the $C^N$ norms of the embeddings are bounded uniformly in~$x$; \item $W_u(x)\cap W_s(x)=\{x\}$ and $T_x W_u(x)=E_u(x)$, $T_xW_s(x)=E_s(x)$; \item the boundaries of $W_u(x),W_s(x)$ do not intersect $\overline B_d(x,\varepsilon_0)$; \item $\varphi^{-1}(W_u(x))\subset W_u(\varphi^{-1}(x))$ and $\varphi^1(W_s(x))\subset W_s(\varphi^1(x))$; \item for each $y\in W_u(x)$, we have $d(\varphi^t(y),\varphi^t(x))\to 0$ as $t\to -\infty$; \item for each $y\in W_s(x)$, we have $d(\varphi^t(y),\varphi^t(x))\to 0$ as $t\to \infty$; \item if $y\in M$, $d(\varphi^t(y),\varphi^t(x))\leq\varepsilon_0$ for all $t\leq 0$, and $d(\varphi^{t}(y),\varphi^{t}(x))\to 0$ as $t\to-\infty$, then $y\in W_u(x)$; \item if $y\in M$, $d(\varphi^t(y),\varphi^t(x))\leq\varepsilon_0$ for all $t\geq 0$, and $d(\varphi^{t}(y),\varphi^{t}(x))\to 0$ as $t\to\infty$, then $y\in W_s(x)$. \end{enumerate} \end{theo} \Remarks 1. The statement~(4) is somewhat artificial since it involves the time-one map $\varphi^1$ and its inverse $\varphi^{-1}$. By rescaling the flow we can construct local stable/unstable manifolds such that the statement~(4) holds with $\varphi^{t_0},\varphi^{-t_0}$ instead, where $t_0>0$ is any fixed number. A more natural statement would be that $\varphi^{-t}(W_u(x))\subset W_u(\varphi^{-t}(x))$ and $\varphi^t(W_s(x))\subset W_s(\varphi^t(x))$ for all $t\geq 0$, but this would be more difficult to arrange since our method of proof is tailored to discrete time evolution. However, below in~\S\ref{s:flow-invariance} we explain that the \emph{global} stable/unstable manifolds are invariant under $\varphi^t$ for all $t$. \noindent 2. Compared to Theorem~\ref{t:stun-maps}, properties~(7) and~(8) impose the additional condition that $d(\varphi^t(y),\varphi^t(x))\to 0$. This is due to the presence of the flow direction: for instance, if $y=\varphi^s(x)$ where $s\neq 0$ is small, then $d(\varphi^t(y),\varphi^t(x))\leq\varepsilon_0$ for all $t$ but $y\notin W_u(x)\cup W_s(x)$. Without this additional condition we can only assert that $y$ lies in the \emph{weak} stable/unstable manifold of $x$, see~\S\ref{s:flows-weak}. \noindent 3. The analog of statement~(9) of Theorem~\ref{t:stun-maps} is given in~\eqref{e:weak-transverse}. The analogs of the quantitative statements~\eqref{e:cream-1}--\eqref{e:cream-3} are discussed in~\S\ref{s:flows-quant}. We now sketch the proof of Theorem~\ref{t:stun-flows}. We follow the proof of Theorem~\ref{t:stun-maps} in~\S\S\ref{s:continuity}--\ref{s:maps-proof}, indicating the changes needed along the way. First of all, the proof of Lemma~\ref{l:contin} applies without change, so the spaces $E_u(x),E_s(x)$ depend continuously on $x$. (The H\"older continuity Lemma~\ref{l:holder} holds as well.) This implies the following version of uniform transversality~\eqref{e:transverse}: there exists a constant $C$ such that for all $x\in K$ \begin{equation} \label{e:transverse-flows} \max\big(|v_0|,|v_u|,|v_s|\big)\leq C|v_0+v_u+v_s|\quad\text{if}\quad v_0\in E_0(x),\ v_u\in E_u(x),\ v_s\in E_s(x). \end{equation} Next, existence of adapted metrics is given by the following analog of Lemma~\ref{l:adapted-metric}: \begin{lemm} \label{l:adapted-metric-flow} Assume that $\varphi^t$ is hyperbolic on~$K$. Fix $\tilde\nu$ such that $0<\tilde\nu<\nu$ where $\nu$ is given in Definition~\ref{d:hyp-flows}. Then there exist $C^N$ Riemannian metrics $|\bullet|_u$, $|\bullet|_s$ on $M$ such that for all $x\in K$ \begin{equation} \label{e:adapted-metric-flow} \begin{aligned} |d\varphi^t(x)v|_u&\leq e^{-\tilde\nu |t|}\cdot|v|_u\quad\text{for all}\quad v\in E_u(x),\ t\leq 0;\\ |d\varphi^t(x)v|_s&\leq e^{-\tilde\nu |t|}\cdot|v|_s\quad\text{for all}\quad v\in E_s(x),\ t\geq 0. \end{aligned} \end{equation} \end{lemm} \begin{proof} This follows by a similar argument to the proof of Lemma~\ref{l:adapted-metric}, fixing a metric $|\bullet|$ on $M$ and defining the adapted metrics as follows: for $v\in T_xM$, $$ |v|_u^2:=\int_0^T e^{2\tilde\nu s}|d\varphi^{-s}(x)v|^2\,ds,\quad |v|_s^2:=\int_0^T e^{2\tilde\nu s}|d\varphi^s(x)v|^2\,ds $$ where $T>0$ is a sufficiently large constant. \end{proof} We next define adapted charts, similarly to~\S\ref{s:adapted-coord}. Fix $\tilde\nu\in (\nu,1)$ and let $|\bullet|_u$, $|\bullet|_s$ be the adapted metrics constructed in Lemma~\ref{l:adapted-metric-flow}. We write elements of $\mathbb R^d$ as $(x_0,x_1,x_2)$ where $x_0\in\mathbb R$, $x_1\in\mathbb R^{d_u}$, and $x_2\in \mathbb R^{d_s}$. Consider the subspaces of $\mathbb R^d$ $$ E_u(0):=\{(0,v_1,0)\mid v_1\in\mathbb R^{d_u}\},\quad E_s(0):=\{(0,0,v_2)\mid v_2\in\mathbb R^{d_s}\}. $$ \begin{defi} \label{d:adapted-chart-flows} Let $x\in K$. A $C^{N+1}$ diffeomorphism $$ \varkappa:U_\varkappa\to V_\varkappa,\quad x\in U_\varkappa\subset M,\quad 0\in V_\varkappa\subset\mathbb R^d $$ is called an \textbf{adapted chart} for $\varphi^t$ centered at $x$, if: \begin{enumerate} \item $\varkappa(x)=0$; \item for each $y\in U_\varkappa$, $d\varkappa(y)$ sends the generator of the flow $X(y)$ to $\partial_{x_0}$; \item $d\varkappa(x)E_u(x)=E_u(0)$ and the restriction of $d\varkappa(x)$ to $E_u(x)$ is an isometry from the metric $|\bullet|_u$ to the Euclidean metric; \item $d\varkappa(x)E_s(x)=E_s(0)$ and the restriction of $d\varkappa(x)$ to $E_s(x)$ is an isometry from the metric $|\bullet|_s$ to the Euclidean metric. \end{enumerate} \end{defi} Similarly to~\S\ref{s:adapted-coord}, it follows from the uniform transversality property~\eqref{e:transverse-flows} that we can select for each $x\in K$ an adapted chart for $\varphi^t$ centered at $x$ $$ \varkappa_x:U_x\to V_x,\quad x\in K $$ such that the set $\{\varkappa_x\mid x\in K\}$ is bounded in the class of $C^{N+1}$ charts. Similarly to~\eqref{e:rescaled-chart} we next define rescaled charts \begin{equation} \label{e:risky-charts} \widetilde\varkappa_x:=T^{-1}\circ \varkappa_x,\quad x\in K;\quad T(w)=\delta_1 w \end{equation} where $\delta_1>0$ is chosen small depending only on $\varphi^t,K$ when we apply Theorem~\ref{t:stun-adv}. The action of the time-one map $\varphi^1$ in the charts $\widetilde\varkappa_x$ is given by the maps \begin{equation} \label{e:flow-acting} \widetilde\psi_x:=\widetilde\varkappa_{\varphi^1(x)}\circ\varphi^1\circ\widetilde\varkappa_x^{-1}. \end{equation} Arguing as in~\S\ref{s:adapted-coord}, we see that $\widetilde\psi_x$ has the following properties: \begin{itemize} \item the domain and the range of $\widetilde\psi_x$ contain the closed ball in $\mathbb R^d$; \item we have \begin{equation} \label{e:flow-differ} \widetilde\psi_x(0)=0,\quad d\widetilde\psi_x(0)=\begin{pmatrix} 1 & 0 & 0\\ 0 & A_{1,x} & 0\\ 0 & 0 & A_{2,x} \end{pmatrix} \end{equation} where the linear maps $A_{1,x}:\mathbb R^{d_u}\to\mathbb R^{d_u}$, $A_{2,x}:\mathbb R^{d_s}\to\mathbb R^{d_s}$ satisfy for all $x\in K$ $$ \|A_{1,x}^{-1}\|\leq e^{-\tilde\nu},\quad \|A_{2,x}\|\leq e^{-\tilde\nu},\quad \max(\|A_{1,x}\|,\|A_{2,x}\|^{-1})\leq C_0 $$ where $C_0$ is some constant depending only on $\varphi^t,K$; \item for any given $\delta>0$, if we choose $\delta_1$ small enough depending on $\delta$, then \begin{equation} \label{e:flow-derbounds} \sup|\partial^\alpha\widetilde\psi_x|\leq\delta\quad\text{for all}\quad x\in K,\quad 2\leq |\alpha|\leq N+1. \end{equation} \end{itemize} The matrix $d\widetilde\psi_x(0)$ in~\eqref{e:flow-differ} has eigenvalue 1 coming from the flow direction, which makes some parts of the proof problematic because Theorem~\ref{t:stun-adv} only partly applies when either the expansion or the contraction rate is equal to~1~-- see~\S\ref{s:reduce-6}. To deal with this problem we introduce the \emph{reduced} maps $\omega_x$, which act on subsets of $\mathbb R^{d-1}$ and correspond to the action of the flow on Poincar\'e sections. Define the projection map $$ \pi_{us}:(w_0,w_1,w_2)\mapsto (w_1,w_2) $$ where $w_0\in\mathbb R$, $w_1\in \mathbb R^{d_u}$, $w_2\in\mathbb R^{d_s}$. From the definition of adapted charts and~\eqref{e:flow-acting} we see that \begin{equation} \label{e:flow-flowing} d\widetilde\psi_x(w)\partial_{x_0}=\partial_{x_0}\quad\text{for all }w. \end{equation} Therefore (shrinking the domain of $\widetilde\psi_x$ if necessary) there exists a diffeomorphism $\omega_x$ of open neighborhoods of $\{(x_1,x_2)\in\mathbb R^{d-1}\colon \max(|x_1|,|x_2|)\leq 1\}$ such that \begin{equation} \label{e:omega-x-def} \pi_{us}(\widetilde\psi_x(w))=\omega_x(\pi_{us}(w))\quad\text{for all }w. \end{equation} The maps $\omega_x$ satisfy $$ \omega_x(0)=0,\quad d\omega_x(0)=\begin{pmatrix} A_{1,x} & 0 \\ 0 & A_{2,x}\end{pmatrix} $$ where $A_{1,x},A_{2,x}$ are the same matrices as in~\eqref{e:flow-differ}. They also satisfy the derivative bounds~\eqref{e:flow-derbounds}. We can finally give the \begin{proof}[Proof of Theorem~\ref{t:stun-flows}] We argue similarly to~\S\ref{s:maps-proof}. For each $x\in K$ and $m\in\mathbb Z$ define the maps $$ \psi_{x,m}:=\widetilde\psi_{\varphi^m(x)}=\widetilde\varkappa_{\varphi^{m+1}(x)}\circ\varphi^1 \circ\widetilde\varkappa_{\varphi^m(x)}^{-1}. $$ The sequence of maps $(\psi_{x,m})_{m\in\mathbb Z}$ satisfies the assumptions in~\S\ref{s:reduce-5} where we absorb the $\partial_{x_0}$ direction into the stable space and put $$ \lambda:=1,\quad \mu:=e^{\tilde\nu}>1. $$ As explained in~\S\ref{s:reduce-6}, Theorem~\ref{t:stun-adv} still partially applies to the maps $(\psi_{x,m})_{m\in\mathbb Z}$ even though $\lambda=1$, yielding the unstable manifolds $W^u_{x,m}\subset\mathbb R^d$. We then define the unstable manifold for $\varphi^t$ at $x$ by \begin{equation} \label{e:flow-u-def} W_u(x):=\widetilde\varkappa_x^{-1}(W^u_{x,0}). \end{equation} If we instead absorb the $\partial_{x_0}$ direction into the unstable space, then the sequence $(\psi_{x,m})_{m\in\mathbb Z}$ satisfies the assumptions in~\S\ref{s:reduce-5} where $$ \lambda:=e^{-\tilde\nu}<1,\quad \mu:=1. $$ As explained in~\S\ref{s:reduce-6}, Theorem~\ref{t:stun-adv} gives the stable manifolds $W^s_{x,m}\subset\mathbb R^d$, and we define the stable manifold for $\varphi^t$ at $x$ by \begin{equation} \label{e:flow-s-def} W_s(x):=\widetilde\varkappa_x^{-1}(W^s_{x,0}). \end{equation} Statements~(1)--(4) of Theorem~\ref{t:stun-flows} are then proved in the same way as for Theorem~\ref{t:stun-maps}. The proof of Theorem~\ref{t:stun-maps} also gives the convergence statements~(5)--(6) for integer~$t$, which imply these statements for all $t$. To show statements~(7)--(8) we use the reduced maps $\omega_x$. The sequence $$ \omega_{x,m}:=\omega_{\varphi^m(x)} $$ satisfies the assumptions in~\S\ref{s:reduce-5} with $$ \lambda:=e^{-\tilde\nu}<1<\mu:=e^{\tilde\nu}. $$ Therefore Theorem~\ref{t:stun-adv} applies to give unstable/stable manifolds for the sequence $(\omega_{x,m})_{m\in\mathbb Z}$. Recalling the construction of these manifolds in~\S\ref{s:reduce-4} it is straightforward to see that the unstable/stable manifolds for $(\omega_{x,m})_{m\in\mathbb Z}$ are equal to $\pi_{us}(W^u_{x,m})$, $\pi_{us}(W^s_{x,m})$ where $W^u_{x,m}$, $W^s_{x,m}$ are the unstable/stable manifolds for the sequence $(\psi_{x,m})_{m\in\mathbb Z}$. We now show statement~(7), with the statement~(8) proved similarly. Assume that $y\in M$ and $d(\varphi^t(y),\varphi^t(x))\leq\varepsilon_0$ for all $t\leq 0$. Arguing similarly to the proof of Theorem~\ref{t:stun-maps} we see that $$ \psi_{x,-n}^{-1}\cdots\psi_{x,-1}^{-1}(\widetilde\varkappa_x(y))\in \{(w_0,w_1,w_2)\colon \max(|w_1|,|(w_0,w_2)|)\leq 1\}\quad\text{for all}\quad n\geq 0. $$ It follows from~\eqref{e:omega-x-def} that $$ \omega_{x,-n}^{-1}\cdots\omega_{x,-1}^{-1}(\pi_{us}(\widetilde\varkappa_x(y))) \in\{(w_1,w_2)\colon \max(|w_1|,|w_2|)\leq 1\}\quad\text{for all}\quad n\geq 0. $$ Applying statement~(5) in Theorem~\ref{t:stun-adv} for the maps $(\omega_{x,m})_{m\in\mathbb Z}$, we see that $\pi_{us}(\widetilde\varkappa_x(y))\in \pi_{us}(W^u_{x,0})$ and thus \begin{equation} \label{e:fiji} \widetilde\varkappa_x(y)=w+(s,0,0)\quad\text{for some}\quad w\in W^u_{x,0},\quad s\in [-2,2]. \end{equation} Now, assume additionally that $d(\varphi^t(y),\varphi^t(x))\to 0$ as $t\to -\infty$. Then \begin{equation} \label{e:fiji2} \psi_{x,-n}^{-1}\cdots\psi_{x,-1}^{-1}(\widetilde\varkappa_x(y))\to 0\quad\text{as}\quad n\to\infty. \end{equation} Using~\eqref{e:flow-flowing} and~\eqref{e:fiji}, we see that $$ \psi_{x,-n}^{-1}\cdots\psi_{x,-1}^{-1}(\widetilde\varkappa_x(y))= \psi_{x,-n}^{-1}\cdots\psi_{x,-1}^{-1}(w)+(s,0,0). $$ Since $w\in W^u_{x,0}$, by part~(3) of Theorem~\ref{t:stun-adv} we have $\psi_{x,-n}^{-1}\cdots\psi_{x,-1}^{-1}(w)\to 0$ as $n\to\infty$. Together with~\eqref{e:fiji2} this shows that $s=0$. Therefore $\widetilde\varkappa_x(y)=w\in W^u_{x,0}$ which implies that $y\in W_u(x)$ as needed. \end{proof} \subsection{Further properties of hyperbolic flows} \label{s:flows-more} We now discuss some further properties of the stable/unstable manifolds constructed in Theorem~\ref{t:stun-flows}. Throughout this section we assume that the conditions of Theorem~\ref{t:stun-flows} hold. \subsubsection{Weak stable/unstable manifolds} \label{s:flows-weak} Let $\delta_1>0$ be the rescaling parameter used in the proof of Theorem~\ref{t:stun-flows}, see~\eqref{e:risky-charts}. Recall that $\delta_1$ is chosen small depending only on $\varphi^t,K$. For each $x\in K$ define the \emph{weak unstable/stable manifolds} $$ W_{u0}(x):=\bigcup_{|s|\leq 2\delta_1} \varphi^s(W_u(x)),\quad W_{s0}(x):=\bigcup_{|s|\leq 2\delta_1} \varphi^s(W_s(x)). $$ In the adapted chart $\widetilde\varkappa_x$ from~\eqref{e:risky-charts}, we have (using~\eqref{e:flow-u-def}--\eqref{e:flow-s-def} and the fact that $d\widetilde\varkappa_x$ maps the generator $X$ of the flow to $\delta_1^{-1}\partial_{x_0}$) \begin{equation} \label{e:weak-in-chart} \begin{aligned} W_{u0}(x)=\widetilde\varkappa_x^{-1}(W^{u0}_{x,0}),&\quad W^{u0}_{x,0}:=\{w+(s,0,0)\colon w\in W^u_{x,0},\ |s|\leq 2\};\\ W_{s0}(x)=\widetilde\varkappa_x^{-1}(W^{s0}_{x,0}),&\quad W^{s0}_{x,0}:=\{w+(s,0,0)\colon w\in W^s_{x,0},\ |s|\leq 2\} \end{aligned} \end{equation} where $W^u_{x,0}$, $W^s_{x,0}$ are the unstable/stable manifolds for the maps $\widetilde \psi_x$, see the proof of Theorem~\ref{t:stun-flows}. Since $W^u_{x,0}$, $W^s_{x,0}$ are graphs of functions of $x_1$, $x_2$ (see~\eqref{e:wuwsy}), we see that $W^{u0}_{x,0},W^{s0}_{x,0}$, and thus $W_{u0}(x),W_{s0}(x)$, are embedded $C^N$ submanifolds of dimensions $d_u+1,d_s+1$. It is also clear that \begin{equation} \label{e:weak-tangent} T_x W_{u0}(x)=E_u(x)\oplus E_0(x),\quad T_x W_{s0}(x)=E_s(x)\oplus E_0(x). \end{equation} It follows from statement~(4) in Theorem~\ref{t:stun-flows} that the weak stable/unstable manifolds are invariant under the positive/negative integer time maps of the flow $\varphi^t$: \begin{equation} \label{e:weak-invariant} \varphi^{-1}(W_{u0}(x))\subset W_{u0}(\varphi^{-1}(x)),\quad \varphi^1(W_{s0}(x))\subset W_{s0}(\varphi^1(x)). \end{equation} We next have the following versions of the statements~(7)--(8) in Theorem~\ref{t:stun-flows}: for all $y\in M$, \begin{equation} \label{e:weak-characterized} \begin{aligned} d(\varphi^t(y),\varphi^t(x))\leq\varepsilon_0\quad\text{for all }t\leq 0 &\quad\Longrightarrow\quad y\in W_{u0}(x);\\ d(\varphi^t(y),\varphi^t(x))\leq\varepsilon_0\quad\text{for all }t\geq 0 &\quad\Longrightarrow\quad y\in W_{s0}(x). \end{aligned} \end{equation} The properties~\eqref{e:weak-characterized} follow immediately from~\eqref{e:fiji} (and its analog for stable manifolds) and~\eqref{e:weak-in-chart}. Similarly to the proof of statement~(9) of Theorem~\ref{t:stun-maps} one can show the following transversality properties: if $x,y\in K$ and $d(x,y)\leq\varepsilon_0$ then \begin{equation} \label{e:weak-transverse} W_{s0}(x)\cap W_u(y),W_s(x)\cap W_{u0}(y)\quad\text{have exactly one point each}. \end{equation} \subsubsection{Quantitative statements} \label{s:flows-quant} We now discuss quantitative versions of the statements of Theorem~\ref{t:stun-flows}, which are the analogs of~\eqref{e:cream-1}--\eqref{e:cream-3}. We fix $\tilde\nu$ such that $$ 0<\tilde\nu<\nu $$ and allow the manifolds $W_u,W_s$ and the constant $\varepsilon_0$ to depend on $\tilde\nu$. The analog of~\eqref{e:cream-1} is given by the following: there exists a constant $C$ such that for all $x\in K$ and $t\geq 0$ \begin{equation} \label{e:milk-1} \begin{aligned} y,\tilde y\in W_u(x)&\quad\Longrightarrow\quad d(\varphi^{-t}(y),\varphi^{-t}(\tilde y))\leq Ce^{-\tilde\nu t}d(y,\tilde y);\\ y,\tilde y\in W_s(x)&\quad\Longrightarrow\quad d(\varphi^{t}(y),\varphi^{t}(\tilde y))\leq Ce^{-\tilde\nu t}d(y,\tilde y). \end{aligned} \end{equation} To show~\eqref{e:milk-1}, it suffices to consider the case of integer~$t$. The latter case is proved similarly to~\eqref{e:cream-1} (see~\S\ref{s:maps-proof}), since~\eqref{e:stuna-2} still applies to the maps $\psi_{x,m}$. The analog of~\eqref{e:cream-2} is given by the following: for all $x\in K$, $y\in M$, $t\geq 0$, and $0\leq\sigma\leq\varepsilon_0$ \begin{equation} \label{e:milk-2} \begin{aligned} d(\varphi^{-s}(y),\varphi^{-s}(x))\leq\sigma\text{ for all }s\in [0,t] &\quad\Longrightarrow\quad d(y,W_{u0}(x))\leq Ce^{-\tilde\nu t}\sigma,\\ d(\varphi^{s}(y),\varphi^{s}(x))\leq\sigma\text{ for all }s\in [0,t] &\quad\Longrightarrow\quad d(y,W_{s0}(x))\leq Ce^{-\tilde\nu t}\sigma. \end{aligned} \end{equation} This is proved by applying~\eqref{e:stuna-3} for the reduced maps $\omega_{x,m}$ defined in~\eqref{e:omega-x-def}, see the proof of~\eqref{e:fiji}. Finally the analog of~\eqref{e:cream-3} is given by the following: for all $x\in K$, $y\in M$, $t\geq 0$, and $0\leq \sigma\leq \varepsilon_0$, \begin{equation} \label{e:milk-3} \begin{gathered} d(\varphi^s(y),\varphi^s(x))\leq\sigma\quad\text{for all}\quad s\in [-t,t]\\ \Longrightarrow\quad d(y,\varphi^r(x))\leq Ce^{-\tilde\nu t}\sigma\quad\text{for some }r\in [-2\delta_1,2\delta_1]. \end{gathered} \end{equation} This is proved by applying~\eqref{e:stuna-4} to the reduced maps $\omega_{x,m}$, arguing similarly to the proof of~\eqref{e:fiji}. \subsubsection{Local invariance and global stable/unstable manifolds} \label{s:flow-invariance} In this section we discuss local invariance of the stable/unstable manifolds. We first discuss invariance under the flow $\varphi^t$. Theorem~\ref{t:stun-flows} does not imply that $\varphi^{-t}(W_u(x))\subset W_u(x)$, $\varphi^t(W_s(x))\subset W_s(x)$ for non-integer $t\geq 0$, however local versions of this statement are established in~\eqref{e:victor-1u}--\eqref{e:victor-2s} below. Similarly to~\eqref{e:iterated-unstable}, \eqref{e:iterated-stable} for each $x\in K$ and integer $k\geq 0$ define the iterated unstable/stable manifolds \begin{equation} \label{e:iterated-stun-flows} W^{(k)}_u(x):=\varphi^k(W_u(\varphi^{-k}(x))),\quad W^{(k)}_s(x):=\varphi^{-k}(W_s(\varphi^k(x))). \end{equation} It follows from statement~(4) in Theorem~\ref{t:stun-flows} that for all $k\geq 0$ $$ W^{(k)}_u(x)\subset W^{(k+1)}_u(x),\quad W^{(k)}_s(x)\subset W^{(k+1)}_s(x). $$ Local invariance of the unstable/stable manifolds under the flow is given by the following statements: there exist $k_0\geq 0$ and $\varepsilon_1>0$ depending only on $\varphi^t,K$ such that for all $x\in K$ and $s\in [-1,1]$ \begin{align} \label{e:victor-1u} \varphi^s(W_u(x))&\subset W_u^{(k_0)}(\varphi^s(x)),\\ \label{e:victor-1s} \varphi^s(W_s(x))&\subset W_s^{(k_0)}(\varphi^s(x)),\\ \label{e:victor-2u} \varphi^s(W_u(x))\cap \overline B_d(\varphi^s(x),\varepsilon_1)&\subset W_u(\varphi^s(x)),\\ \label{e:victor-2s} \varphi^s(W_s(x))\cap \overline B_d(\varphi^s(x),\varepsilon_1)&\subset W_s(\varphi^s(x)). \end{align} Note that~\eqref{e:victor-1u}, \eqref{e:victor-2u} can be extended to all $s\leq 0$ and~\eqref{e:victor-1s}, \eqref{e:victor-2s} to all $s\geq 0$ using statement~(4) in Theorem~\ref{t:stun-flows}. We show~\eqref{e:victor-1s}, \eqref{e:victor-2s}, with~\eqref{e:victor-1u}, \eqref{e:victor-2u} proved similarly. We start with~\eqref{e:victor-1s}. Assume that $y\in W_s(x)$. Then by~\eqref{e:milk-1} we have for all $t\geq 0$ \begin{equation} \label{e:milky-way-1} d(\varphi^{t+s}(y),\varphi^{t+s}(x))\leq Cd(\varphi^t(y),\varphi^t(x))\leq Ce^{-\tilde\nu t}. \end{equation} Therefore, $d(\varphi^{t+s}(y),\varphi^{t+s}(x))\to 0$ as $t\to\infty$ and there exists $k_0\geq 0$ such that \begin{equation} \label{e:milky-way-2} d(\varphi^{t+s+k_0}(y),\varphi^{t+s+k_0}(x))\leq \varepsilon_0\quad\text{for all}\quad t\geq 0. \end{equation} It follows from statement~(8) in Theorem~\ref{t:stun-flows} that $\varphi^{s+k_0}(y)\in W_s(\varphi^{s+k_0}(x))$ and thus $\varphi^s(y)\in W_s^{(k_0)}(\varphi^s(x))$ as needed. To show~\eqref{e:victor-2s}, assume that $y\in W_s(x)$ and $d(\varphi^s(y),\varphi^s(x))\leq\varepsilon_1$. Then~\eqref{e:milky-way-2} holds. If $\varepsilon_1$ is small enough, then~\eqref{e:milky-way-2} holds also for all $t\in [-k_0,0]$, and thus for all $t\geq -k_0$. It then follows from statement~(8) in Theorem~\ref{t:stun-flows} that $\varphi^s(y)\in W_s(\varphi^s(x))$ as needed. Next, we note that the statements~\eqref{e:intersector}, \eqref{e:tangentor}, \eqref{e:locally-unique-1}, and~\eqref{e:locally-unique-2} relating the stable/unstable manifolds at different points are still valid in the case of flows, with very similar proofs. Finally, similarly to~\eqref{e:global} we can define the \emph{global unstable/stable manifolds}: for $x\in K$, \begin{equation} \label{e:global-flows} W^{(\infty)}_u(x):=\bigcup_{k\geq 0}W_u^{(k)}(x),\quad W^{(\infty)}_s(x):=\bigcup_{k\geq 0}W_s^{(k)}(x). \end{equation} By statements~(5)--(8) in Theorem~\ref{t:stun-flows}, these can be characterized as follows: \begin{equation} \label{e:global-flows-char} \begin{aligned} y\in W^{(\infty)}_u(x)&\quad\Longleftrightarrow\quad d(\varphi^t(y),\varphi^t(x))\to 0\quad\text{as }t\to -\infty;\\ y\in W^{(\infty)}_s(x)&\quad\Longleftrightarrow\quad d(\varphi^t(y),\varphi^t(x))\to 0\quad\text{as }t\to \infty. \end{aligned} \end{equation} The manifolds $W^{(\infty)}_u(x)$, $W^{(\infty)}_s(x)$ are $d_u$ and $d_s$-dimensional immersed submanifolds without boundary in $M$. They are invariant under the flow $\varphi^t$ and disjoint from each other. \section{Examples} \label{s:examples} In this section we provide two examples of hyperbolic systems: geodesic flows on negatively curved surfaces (\S\ref{s:surf-neg}) and billiard ball maps on Euclidean domains with concave boundary (\S\ref{s:billiard}). \subsection{Surfaces of negative curvature} \label{s:surf-neg} Let $(M,g)$ be a closed (compact without boundary) oriented surface. To simplify the computations below, we will often use (positively oriented) \emph{isothermal coordinates} $(x,y)$ in which the metric is conformally flat: \begin{equation} \label{e:flatty} g=e^{2G(x,y)}(dx^2+dy^2). \end{equation} Such coordinates exist locally near each point of $M$. In isothermal coordinates, the Gauss curvature is given by \begin{equation} \label{e:kurvy} K(x,y)=-e^{-2G(x,y)}(\partial_x^2+\partial_y^2)G(x,y). \end{equation} We show below in Theorem~\ref{t:geodesic-hyperbolic} that if $K<0$ then the geodesic flow on $(M,g)$ is hyperbolic. We define the geodesic flow as the Hamiltonian flow \begin{equation} \label{e:gerry} \begin{aligned} \varphi^t:=\exp(tX):S^*M\to S^*M,&\qquad X:=H_p,\\ S^*M:=\{(x,\xi)\in T^*M\colon p(x,\xi)=1\},&\qquad p(x,\xi):=|\xi|_g \end{aligned} \end{equation} where $T^*M$ is the cotangent bundle of $M$ and $S^*M$ is the unit cotangent bundle (with respect to the metric $g$). In local coordinates $(x,y)$, if $(\xi,\eta)$ are the corresponding momentum variables (i.e. coordinates on the fibers of $T^*M$) then \begin{equation} \label{e:hammy} X=H_p=(\partial_\xi p)\partial_x+(\partial_\eta p)\partial_y -(\partial_x p)\partial_\xi-(\partial_y p)\partial_\eta. \end{equation} In isothermal coordinates~\eqref{e:flatty} we have \begin{equation} \label{e:poppy} p(x,y,\xi,\eta)=e^{-G(x,y)}\sqrt{\xi^2+\eta^2}. \end{equation} \begin{theo} \label{t:geodesic-hyperbolic} Let $(M,g)$ be a closed oriented surface with geodesic flow $\varphi^t:S^*M\to S^*M$ defined in~\eqref{e:gerry}. Assume that the Gauss curvature $K$ is negative everywhere. Then $\varphi^t$ is an Anosov flow, that is $\varphi^t$ is hyperbolic on the entire $S^*M$ in the sense of Definition~\ref{d:hyp-flows}. \end{theo} \Remark Theorem~\ref{t:geodesic-hyperbolic} extends to higher dimensional manifolds of \emph{negative sectional curvature}~-- see for instance~\cite[Theorem~17.6.2]{KaHa}. The orientability hypothesis is made only for convenience of the proof, one can remove it for instance by passing to a double cover of $M$. In the remainder of this section we prove Theorem~\ref{t:geodesic-hyperbolic}. We first define a convenient frame on $S^*M$. Let $V$ be the vector field on $S^*M$ generating rotations on the circle fibers (counterclockwise with respect to the fixed orientation). If $(x,y)$ are isothermal coordinates~\eqref{e:flatty}, we use local coordinates $(x,y,\theta)$ on $S^*M$ where $\theta$ is defined by $$ \xi=e^{G(x,y)}\cos\theta,\quad \eta=e^{G(x,y)}\sin\theta, $$ and we have (here $X$ is the generator of the geodesic flow) $$ V=\partial_\theta,\quad X=e^{-G(x,y)}\big(\cos\theta\,\partial_x+\sin\theta\,\partial_y +(\partial_y G(x,y)\cos\theta-\partial_x G(x,y)\sin\theta)\partial_\theta\big). $$ Define the vector field $$ X_\perp:=[X,V], $$ in isothermal coordinates $$ X_\perp=e^{-G(x,y)}\big(\sin\theta\,\partial_x-\cos\theta\,\partial_y +(\partial_x G(x,y)\cos\theta+\partial_y G(x,y)\sin\theta)\partial_\theta\big). $$ The vector fields $X,V,X_\perp$ form a global frame on $S^*M$ and we have (using~\eqref{e:kurvy}) \begin{equation} \label{e:commy} [X,V]=X_\perp,\quad [X_\perp,V]=-X,\quad [X,X_\perp]=-KV. \end{equation} For any vector field $W$ on $S^*M$, we have (by the standard properties of Lie bracket) \begin{equation} \label{e:lie} \partial_t(\varphi^{-t}_*W)=\varphi^{-t}_*[X,W]\quad\text{where}\quad \varphi^{-t}_*W(\rho):=d\varphi^{-t}(\rho)W(\varphi^t(\rho)),\ \rho\in S^*M. \end{equation} It then follows from~\eqref{e:commy} that the following two-dimensional subbundle of $T(S^*M)$ is invariant under the flow $\varphi^t$: $$ E_{us}:=\Span(V,X_\perp). $$ The space $E_{us}$ will be the direct sum of the stable and the unstable subspaces for the flow $\varphi^t$. For a vector $v\in E_{us}(\rho)$, $\rho\in S^*M$, we define its coordinates $(a,b)$ with respect to the frame $V,X_\perp$: \begin{equation} \label{e:cordial} v=aV(\rho)+bX_\perp(\rho). \end{equation} We have the following differential equations for the action of $d\varphi^t$ on $E_{us}$ (which are a special case of Jacobi's equations): \begin{lemm} \label{l:jacobi} Let $\rho\in S^*M$, $v\in E_{us}(\rho)$, and denote $$ \rho(t)=(x(t),\xi(t)):=\varphi^t(\rho)\in S^*M,\quad v(t):=d\varphi^t(\rho)v\in E_{us}(\rho(t)). $$ Let $a(t),b(t)$ be the coordinates of~$v(t)$ defined in~\eqref{e:cordial}. Put $K(t):=K(x(t))$. Then, denoting by dots derivatives with respect to~$t$, the functions $a(t),b(t)$ satisfy the ordinary differential equation \begin{equation} \label{e:diffy} \dot a=K(t)b,\quad \dot b=-a. \end{equation} \end{lemm} \begin{proof} Define the vector field $W_t:=a(t)V+b(t)X_\perp$. Then $W_t(\varphi^t(\rho))=v(t)=d\varphi^t(\rho)v$ and thus $v=\varphi^{-t}_*W_t(\rho)$. By~\eqref{e:lie} we have $$ 0=\partial_t(\varphi^{-t}_*W_t)(\rho)=\varphi^{-t}_*\big(\partial_t W_t+[X,W_t]\big)(\rho). $$ Using~\eqref{e:commy}, we get $$ 0=(\partial_t W_t+[X,W_t])(\rho(t))=\big(\dot a(t)-K(x(t))b(t)\big)V(\rho(t))+\big(\dot b(t)+a(t)\big)X_\perp(\rho(t)\big). $$ This implies~\eqref{e:diffy}. \end{proof} Lemma~\ref{l:jacobi} immediately implies Theorem~\ref{t:geodesic-hyperbolic} in the special case of \emph{constant curvature} $K\equiv -1$, with expansion rate $\nu=1$ in~\eqref{e:hypflow-exp} and $E_u,E_s$ given by $$ E_u=\Span(V-X_\perp),\quad E_s=\Span(V+X_\perp). $$ To handle the case of variable curvature we first construct cones in $E_{us}$ which are invariant under the flow $\varphi^t$ for positive and negative times (see Figure~\ref{f:cones}): \begin{lemm} \label{l:cones-1} For each $\rho\in S^*M$, define the closed cones $\mathcal C^u_0(\rho),\mathcal C^s_0(\rho)\subset E_{us}(\rho)$: $$ \mathcal C^u_0(\rho):=\{aV(\rho)+bX_\perp(\rho)\mid ab\leq 0\},\quad \mathcal C^s_0(\rho):=\{aV(\rho)+bX_\perp(\rho)\mid ab\geq 0\}. $$ Assume that $K\leq 0$ everywhere on $M$. Then for all $t\geq 0$ \begin{equation} \label{e:cones-1} d\varphi^t(\mathcal C^u_0(\rho))\subset \mathcal C^u_0(\varphi^t(\rho)),\quad d\varphi^{-t}(\mathcal C^s_0(\rho))\subset \mathcal C^s_0(\varphi^{-t}(\rho)). \end{equation} \end{lemm} \begin{proof} In the notation of Lemma~\ref{l:jacobi} we have $$ \partial_t(a(t)b(t))=-a(t)^2+K(x(t))b(t)^2\leq 0 $$ and~\eqref{e:cones-1} follows immediately. \end{proof} We now prove an upgraded version of Lemma~\ref{l:cones-1}, constructing invariant cones on which the differentials $d\varphi^t$, $d\varphi^{-t}$ are expanding. Fix small constants $\zeta>0$, $\gamma>0$ to be chosen later in Lemma~\ref{l:cones-2}. Define the norm $|\bullet|$ on the fibers of $E_{us}$ as follows: $$ |aV(\rho)+bX_\perp(\rho)|:=\sqrt{\zeta a^2+b^2}. $$ Define also the following dilation invariant function $\Theta$ on the fibers of $E_{us}\setminus 0$: $$ \Theta(aV(\rho)+bX_\perp(\rho)):={ab\over \zeta a^2+b^2}. $$ The upgraded invariant cones are constructed in \begin{figure} \includegraphics[scale=0.25]{matfig/jacobi1.jpg}\qquad \includegraphics[scale=0.25]{matfig/jacobi2.jpg} \caption{Direction fields for the equation~\eqref{e:diffy} for $K=1$ (left) and $K=-1$ (right) in the $(a,b)$ plane. The cones on the right are $\mathcal C^u_\gamma$ (blue) and $\mathcal C^s_\gamma$ (red).} \label{f:cones} \end{figure} \begin{lemm} \label{l:cones-2} Assume that $K<0$ everywhere on $M$. Then there exist $\zeta>0$, $\gamma>0$, $\nu>0$ such that the closed cones $\mathcal C^u_\gamma(\rho),\mathcal C^s_\gamma(\rho)\subset E_{us}(\rho)$ defined by (see Figure~\ref{f:cones}) $$ \mathcal C^u_\gamma(\rho):=\{v\in E_{us}(\rho)\colon \Theta(v)\leq -\gamma\}\cup\{0\},\quad \mathcal C^s_\gamma(\rho):=\{v\in E_{us}(\rho)\colon \Theta(v)\geq \gamma\}\cup\{0\} $$ have the following properties for all $\rho\in S^*M$ and $t\geq 0$ \begin{align} \label{e:coney-1} d\varphi^t(\rho)\mathcal C^u_\gamma(\rho)&\subset \mathcal C^u_\gamma(\varphi^t(\rho));\\ \label{e:coney-2} d\varphi^{-t}(\rho)\mathcal C^s_\gamma(\rho)&\subset \mathcal C^s_\gamma(\varphi^{-t}(\rho));\\ \label{e:coney-3} |d\varphi^t(\rho)v|&\geq e^{\nu t}|v|\quad\text{for all}\quad v\in \mathcal C^u_\gamma(\rho);\\ \label{e:coney-4} |d\varphi^{-t}(\rho)v|&\geq e^{\nu t}|v|\quad\text{for all}\quad v\in \mathcal C^s_\gamma(\rho). \end{align} \end{lemm} \begin{proof} We fix constants $K_0,K_1>0$ such that $$ 0<K_0\leq -K(x)\leq K_1\quad\text{for all}\quad x\in M $$ and put \begin{equation} \label{e:zeta-gamma-def} \zeta:={1\over K_1},\quad \gamma:={\sqrt K_0\over 3},\quad \nu:=\gamma(1+\zeta K_0). \end{equation} Let $\rho\in S^*M$, $v\in E_{us}(\rho)$, and put $\rho(t):=\varphi^t(\rho)$, $v(t)=d\varphi^t(\rho)v$. We write $v(t)=a(t)V(\rho(t))+b(t)X_\perp(\rho(t))$ and recall that $a(t),b(t)$ satisfy the differential equations~\eqref{e:diffy}. Denote $K(t):=K(x(t))$ where $\rho(t)=(x(t),\xi(t))$ and $$ R(t):=|v(t)|^2=\zeta a(t)^2+b(t)^2,\quad \Theta(t):=\Theta(v(t))={a(t)b(t)\over \zeta a(t)^2+b(t)^2}. $$ Then it follows from~\eqref{e:diffy} that (denoting by dots derivatives with respect to~$t$) \begin{equation} \label{e:dotty} \dot R=-2(1-\zeta K(t))\Theta R,\quad \dot\Theta=-{a^2-Kb^2\over R}+2(1-\zeta K(t))\Theta^2. \end{equation} Therefore $$ \dot\Theta\leq -{a^2+K_0b^2\over R}+2(1+\zeta K_1)\Theta^2 \leq -K_0+4\Theta^2. $$ Thus if $|\Theta(t)|=\gamma$ for some $t$, then $\dot\Theta(t)<0$. In particular, if $\Theta(0)\leq -\gamma$, then $\Theta(t)\leq -\gamma$ for all $t\geq 0$, which implies~\eqref{e:coney-1}. Similarly if $\Theta(0)\geq \gamma$, then $\Theta(t)\geq\gamma$ for all $t\leq 0$, which implies~\eqref{e:coney-2}. We next prove~\eqref{e:coney-3}. Assume that $v\in \mathcal C^u_\gamma(\rho)\setminus 0$, then $\Theta(t)\leq -\gamma$ for all $t\geq 0$. From~\eqref{e:dotty} we have $$ \dot R(t)\geq 2\gamma(1+\zeta K_0)R(t)= 2\nu R(t)\quad\text{for all}\quad t\geq 0. $$ Therefore $R(t)\geq e^{2\nu t}R(0)$ for all $t\geq 0$, which gives~\eqref{e:coney-3}. The bound~\eqref{e:coney-4} is proved similarly. \end{proof} We finally use the cones from Lemma~\ref{l:cones-2} to construct the stable/unstable spaces, finishing the proof of Theorem~\ref{t:geodesic-hyperbolic}: \begin{figure} \includegraphics[scale=0.35]{matfig/jacobis.jpg} \caption{The cones $\mathcal C^u_{\gamma,t}(\rho)$ (blue) and $\mathcal C^s_{\gamma,t}(\rho)$ (red) for several values of~$t$, with the darker colors representing larger values of~$t$. The solid lines are the spaces $E_u(\rho),E_s(\rho)$.} \label{f:manycones} \end{figure} \begin{lemm} \label{l:geodfinal} Let $\gamma$ be chosen in Lemma~\ref{l:cones-2}. For each $\rho\in S^*M$ and $t\geq 0$ define the subsets of $E_{us}(\rho)$ $$ \mathcal C^u_{\gamma,t}(\rho):=d\varphi^t(\varphi^{-t}(\rho))\mathcal C^u_\gamma(\varphi^{-t}(\rho)),\quad \mathcal C^s_{\gamma,t}(\rho):=d\varphi^{-t}(\varphi^{t}(\rho))\mathcal C^s_\gamma(\varphi^{t}(\rho)), $$ and consider their intersections (see Figure~\ref{f:manycones}) \begin{equation} \label{e:geodfinal-2} E_u(\rho):=\bigcap_{t\geq 0}\mathcal C^u_{\gamma,t}(\rho),\quad E_s(\rho):=\bigcap_{t\geq 0}\mathcal C^s_{\gamma,t}(\rho). \end{equation} Then $E_u(\rho),E_s(\rho)$ are one-dimensional subspaces of $E_{us}(\rho)$ which satisfy the conditions of Definition~\ref{d:hyp-flows}. \end{lemm} \begin{proof} We show the properties of $E_u(\rho)$. The properties of $E_s(\rho)$ are proved similarly and the transversality of $E_u(\rho)$ and $E_s(\rho)$ follows from the fact that $E_u(\rho)\subset \mathcal C^u_\gamma(\rho)$, $E_s(\rho)\subset \mathcal C^s_\gamma(\rho)$, and $\mathcal C^u_\gamma(\rho)\cap \mathcal C^s_\gamma(\rho)=\{0\}$. It follows from~\eqref{e:coney-1} that $$ \mathcal C^u_{\gamma,s}(\rho)\subset\mathcal C^u_{\gamma,t}(\rho)\quad\text{when}\quad s\geq t\geq 0. $$ We first claim that $E_u(\rho)$ contains a one-dimensional subspace of $E_{us}(\rho)$. Indeed, let $\mathscr G$ be the Grassmanian of all one-dimensional subspaces of $E_{us}(\rho)$ and $\mathscr V_t\subset\mathscr G$ consist of the subspaces which are contained in $\mathcal C^u_{\gamma,t}(\rho)$. Then $\mathscr V_s\subset\mathscr V_t$ for $s\geq t$ and all the sets $\mathscr V_t$ are compact. Moreover, each $\mathscr V_t$ is nonempty since (recalling~\eqref{e:zeta-gamma-def}) \begin{equation} \label{e:coller0} V^0_u(\rho):=\{ aV(\rho)+bX_\perp(\rho)\mid b=-a\sqrt{\zeta}\} \subset \mathcal C^u_\gamma(\rho). \end{equation} Therefore the intersection $\bigcap_{t\geq 0}\mathscr V_t\subset \mathscr G$ is nonempty. Take an element $V_u(\rho)$ of this intersection, then $V_u(\rho)$ is a one-dimensional subspace of $E_{us}(\rho)$ and $V_u(\rho)\subset E_u(\rho)$. We now claim that \begin{equation} \label{e:coller1} E_u(\rho)=V_u(\rho). \end{equation} For that, it suffices to show that every $v\in E_u(\rho)$ lies in $V_u(\rho)$. Define the one-dimensional space $V^0_s(\rho)\subset \mathcal C^s_\gamma(\rho)$ similarly to~\eqref{e:coller0} but putting $b=a\sqrt{\zeta}$. Since $V_u(\rho)\subset \mathcal C^u_\gamma(\rho)$, the spaces $V_u(\rho)$ and $V^0_s(\rho)$ are transverse to each other. Thus we can write $$ v=v_1+v_2\quad\text{for some}\quad v_1\in V_u(\rho),\quad v_2\in V^0_s(\rho). $$ Denote $$ v(t):=d\varphi^{-t}(\rho)v,\quad v_1(t):=d\varphi^{-t}(\rho)v_1,\quad v_2(t):=d\varphi^{-t}(\rho) v_2. $$ Since $v,v_1\in E_u(\rho)$, we have $v(t),v_1(t)\in\mathcal C^u_\gamma(\varphi^{-t}(\rho))$ for all $t\geq 0$. It follows from~\eqref{e:coney-3} applied to $v(t),v_1(t)$ that \begin{equation} \label{e:viper-1} |v_2(t)|\leq |v(t)|+|v_1(t)|\leq e^{-\nu t}(|v|+|v_1|)\quad\text{for all}\quad t\geq 0. \end{equation} On the other hand, since $v_2\in \mathcal C^s_\gamma(\rho)$ we have by~\eqref{e:coney-4} \begin{equation} \label{e:viper-2} |v_2|\leq e^{-\nu t}|v_2(t)|\quad\text{for all}\quad t\geq 0. \end{equation} Combining~\eqref{e:viper-1} and~\eqref{e:viper-2} and letting $t\to\infty$ we get $v_2=0$, thus $v=v_1\in V_u(\rho)$ which gives~\eqref{e:coller1}. Now, \eqref{e:coller1} implies immediately that $E_u(\rho)$ is a one-dimensional subspace of $E_{us}(\rho)$. We have $d\varphi^{-t}(\rho)E_u(\rho)\subset E_u(\varphi^{-t}(\rho))$ for all $t\geq 0$, which gives the invariance property~\eqref{e:hypflow-inv}. The expansion property~\eqref{e:hypflow-exp} follows from~\eqref{e:coney-3} and the inclusion $E_u(\rho)\subset\mathcal C^u_\gamma(\rho)$. \end{proof} \Remark The proof of Lemma~\ref{l:geodfinal} can be used to show the stability of Anosov maps/flows under perturbations, namely a small $C^N$ perturbation of an Anosov map/flow is still Anosov. This uses the fact that (a slightly modified version of) the properties~\eqref{e:coney-1}--\eqref{e:coney-4} is stable under perturbations; note it is enough to require these properties for $0\leq t\leq 1$. See~\cite[Corollary~6.4.7 and Proposition~17.4.4]{KaHa} for details. \subsection{A simple case of the billiard ball map} \label{s:billiard} We finally briefly discuss two-dimensional billiard ball maps. Let $\Omega\subset\mathbb R^2$ be a domain with smooth boundary, referring the reader to~\cite{ChernovBook} for a comprehensive treatment and history of the subject. We do not require $\Omega$ to be bounded but we require that the boundary $\partial\Omega$ be compact; this implies that either $\Omega$ is compact (\emph{interior case}) or $\mathbb R^2\setminus\Omega^\circ$ is compact (\emph{exterior case}). The boundary $\partial\Omega$ is diffeomorphic to the union of finitely many circles. We parametrize it locally by a real number variable $\theta$, denoting the parametrization $$ \mathbf x:\partial\Omega\to \mathbb R^2. $$ We assume that $\mathbf x$ is a unit speed parametrization: $$ |\mathbf v|\equiv 1\quad\text{where}\quad \mathbf v(\theta):=\partial_\theta \mathbf x(\theta). $$ We choose the inward pointing unit normal vector field $$ \mathbf n:\partial\Omega\to\mathbb R^2. $$ We assume that $(\mathbf v,\mathbf n)$ is a positively oriented frame on $\mathbb R^2$ at every point of $\partial\Omega$. If $\partial\Omega$ consists of a single circle, then the direction of increasing $\theta$ is counterclockwise in the interior case and clockwise in the exterior case. For each $\theta\in\partial\Omega$ we define the \emph{curvature} $K(\theta)$ of the boundary at $\theta$ by the following identity: $$ \partial_\theta\mathbf v(\theta)=K(\theta)\mathbf n(\theta). $$ We say that the boundary is (strictly) \emph{convex} at some point $\theta$ if $K(\theta)>0$ and (strictly) \emph{concave} if $K(\theta)<0$. See Figure~\ref{f:obstacles}. \begin{figure} \includegraphics{stunnote.3}\qquad \includegraphics{stunnote.4} \caption{The interior (left) and exterior (right) case when the boundary is a circle of radius~$r$. The domain $\Omega$ is shaded. The curvature is equal to $1/r$ in the interior case and to $-1/r$ in the exterior case.} \label{f:obstacles} \end{figure} The billiard ball map acts on the phase space $M$ which consists of inward pointing unit vectors at the boundary: $$ M=\{(\theta,\mathbf w)\mid \theta\in\partial\Omega,\ \mathbf w\in\mathbb R^2,\ |\mathbf w|=1,\ \langle \mathbf w,\mathbf n(\theta)\rangle\geq 0\}. $$ The boundary of the phase space $\partial M$ consists of \emph{glancing} directions, where $\mathbf w$ is tangent to $\partial\Omega$. These directions are what makes billiards much more difficult to handle than closed manifolds; in these notes we ignore entirely the complications resulting from glancing, restricting to the interior of $M$. The interior $M^\circ$ can be parametrized by two numbers $\theta\in\partial\Omega$ and $\sigma\in (-1,1)$ defined by $$ \sigma:=\langle \mathbf w,\mathbf v(\theta)\rangle. $$ In particular $\sigma=0$ corresponds to vectors which are orthogonal to the boundary. We now define the billiard ball map $$ \varphi:M^\circ\to M^\circ,\quad \varphi(\theta_1,\sigma_1)=(\theta_2,\sigma_2) $$ where $\mathbf x(\theta_2)$ is the first intersection of $\partial\Omega$ with the ray $\{\mathbf x(\theta_1)+t\mathbf w_1\mid t>0\}$ and $\mathbf w_2$ is obtained by the law of reflection~-- see Figure~\ref{f:billiards}. The map $\varphi$ is in fact only defined on an open subset of $M^\circ$ since the ray might intersect $\partial\Omega$ in a glancing direction or (in the exterior case) may escape to infinity without intersecting $\partial\Omega$, we ignore here the issues arising from this fact. \begin{figure} \includegraphics{stunnote.5} \caption{The billiard ball map for the disk (interior case), given in case of the unit disk by the formulas $\theta_2=\theta_1+2\arccos\sigma_1$, $\sigma_2=\sigma_1$.} \label{f:billiards} \end{figure} Define the function on $\partial\Omega\times\partial\Omega$ $$ \Phi(\theta_1,\theta_2):=|\mathbf x(\theta_1)-\mathbf x(\theta_2)|. $$ Then $\Phi$ is the generating function of $\varphi$, namely \begin{equation} \label{e:billimap-1} \varphi(\theta_1,\sigma_1)=(\theta_2,\sigma_2)\quad\Longleftrightarrow\quad \sigma_1=-\partial_{\theta_1}\Phi(\theta_1,\theta_2),\ \sigma_2=\partial_{\theta_2}\Phi(\theta_1,\theta_2). \end{equation} (Strictly speaking, we should restrict the right-hand side above to $\theta_1\neq\theta_2$ such that the interior of the line segment between $\mathbf x(\theta_1)$ and $\mathbf x(\theta_2)$ does not intersect $\partial\Omega$.) We have $$ \begin{gathered} \partial_{\theta_1}^2\Phi={1-\sigma_1^2\over\Phi}-K(\theta_1)\sqrt{1-\sigma_1^2},\quad \partial_{\theta_2}^2\Phi={1-\sigma_2^2\over\Phi}-K(\theta_2)\sqrt{1-\sigma_2^2},\\ \partial_{\theta_1\theta_2}\Phi={\sqrt{(1-\sigma_1^2)(1-\sigma_2^2)}\over\Phi}. \end{gathered} $$ Therefore, if $\varphi(\theta_1,\sigma_1)=(\theta_2,\sigma_2)$ and we denote $\ell:=\Phi(\theta_1,\theta_2)$ (the distance traveled between the bounces) and $K_1:=K(\theta_1)$, $K_2:=K(\theta_2)$, then \begin{equation} \label{e:huge-matrix} d\varphi(\theta_1,\sigma_1)=\begin{pmatrix}\displaystyle {\ell K_1-\sqrt{1-\sigma_1^2}\over \sqrt{1-\sigma_2^2}} & \displaystyle -{\ell\over \sqrt{(1-\sigma_1^2)(1-\sigma_2^2)}}\\ \noalign{\bigskip} \displaystyle K_1\sqrt{1-\sigma_2^2}+K_2\sqrt{1-\sigma_1^2}-\ell K_1K_2 & \displaystyle {\ell K_2-\sqrt{1-\sigma_2^2}\over \sqrt{1-\sigma_1^2}} \end{pmatrix}. \end{equation} Note that $\det d\varphi\equiv 1$. \begin{figure} \includegraphics{stunnote.6} \caption{A hyperbolic trajectory on a Bunimovich stadium} \label{f:billimore} \end{figure} We now discuss under which conditions $\varphi$ is hyperbolic. We start with the simplest case of a closed trajectory of period 2 (which is necessarily orthogonal to the boundary): \begin{prop} \label{l:billiard-per2} Assume that $(\theta_1,0)$ is a periodic point for $\varphi$ with period~2, namely $\varphi(\theta_1,0)=(\theta_2,0)$ and $\varphi(\theta_2,0)=(\theta_1,0)$ for some $\theta_2\in\partial\Omega$. Let $\ell:=|\mathbf x(\theta_1)-\mathbf x(\theta_2)|$, $K_1:=K(\theta_1)$, $K_2:=K(\theta_2)$. Then $\varphi$ is hyperbolic on the closed trajectory $\{(\theta_j,0)\}$, in the sense of Definition~\ref{d:hyp-map}, if and only if the following condition holds: \begin{equation} \label{e:billiard-per2} (1-\ell K_1)(1-\ell K_2)\notin [0, 1]. \end{equation} \end{prop} \Remark Condition~\eqref{e:billiard-per2} always holds in the concave case (when $K_1,K_2<0$). However it sometimes also holds in the convex case, see Figure~\ref{f:billimore}. \begin{proof} Put $$ A:=d\varphi^2(\theta_1,0)=d\varphi(\theta_2,0)d\varphi(\theta_1,0). $$ Then $\varphi$ is hyperbolic on $\{(\theta_j,0)\}$ if and only if $A$ has no eigenvalues on the unit circle. Since $\det A=1$, this is equivalent to $|\tr A|>2$. Computing $$ \tr A=2+4\ell(\ell K_1K_2-K_1-K_2) $$ we arrive to the condition~\eqref{e:billiard-per2}. \end{proof} For general sets we restrict to the concave case (the corresponding billiards are sometimes called \emph{dispersing}): \begin{prop} \label{l:billiard-gen} Assume that $\mathcal K\subset M^\circ$ is a $\varphi$-invariant compact set and the curvature $K$ satisfies $K(\theta)<0$ for all $(\theta,\sigma)\in\mathcal K$. Then the billiard ball map $\varphi$ is hyperbolic on~$\mathcal K$ in the sense of Definition~\ref{d:hyp-map}. \end{prop} \Remark One example of a compact $\varphi$-invariant set is a closed (non-glancing) trajectory. Another example is when $\Omega$ is the complement of several strictly convex obstacles (that is, the exterior concave case), we impose the \emph{no-eclipse condition} that no obstacle intersects the convex hull of the union of any two other obstacles, and $\mathcal K$ is the reduction to boundary of the \emph{trapped set} which consists of all billiard ball trajectories which stay in a bounded set for all times. The no-eclipse condition ensures that trapped trajectories cannot be glancing. See Figure~\ref{f:several-obstacles}. \begin{figure} \includegraphics[width=5.5cm]{matfig/billiard1.jpg}\quad \includegraphics[width=8.25cm]{matfig/billiard2.jpg} \caption{Left: numerically computed set of trapped trajectories in the exterior of 3 disks. Right: the corresponding reduction to the boundary $\mathcal K$, that is the trapped set of the billiard ball map, restricted to the bottom circle; the horizontal direction is $\theta$ and the vertical direction is $\sigma$. Both sets exhibit fractal structure.} \label{f:several-obstacles} \end{figure} \begin{proof}[Sketch of the proof] We argue similarly to~\S\ref{s:surf-neg}. Consider the cones in $\mathbb R^2$ $$ \mathcal C^u_0:=\{(v_\theta,v_\sigma)\mid v_\theta\cdot v_\sigma\geq 0\},\quad \mathcal C^s_0:=\{(v_\theta,v_\sigma)\mid v_\theta\cdot v_\sigma\leq 0\}. $$ By the concavity condition, for all $\rho:=(\theta_1,\sigma_1)\in\mathcal K$ we have $K_1,K_2<0$ in~\eqref{e:huge-matrix}. Thus all entries of the matrix $d\varphi(\rho)$ are negative. It follows that \begin{equation} \label{e:conifer-1} d\varphi(\rho)\mathcal C^u_0\subset \mathcal C^u_0,\quad d\varphi(\rho)^{-1}\mathcal C^s_0\subset\mathcal C^s_0. \end{equation} Next, at each $\rho=(\theta,\sigma)\in \mathcal K$ define the norm $|\bullet|_{\rho}$ by $$ |(v_\theta,v_\sigma)|^2_{\rho}:=(1-\sigma^2) v_\theta^2+{v_\sigma^2\over 1-\sigma^2}. $$ Then there exists $\lambda>1$ such that for all $\rho\in \mathcal K$ and $v\in\mathbb R^2$ \begin{align} \label{e:conifer-2} v\in \mathcal C^u_0&\quad\Longrightarrow\quad |d\varphi(\rho)v|_{\varphi(\rho)}\geq\lambda |v|_\rho,\\ \label{e:conifer-3} v\in \mathcal C^s_0&\quad\Longrightarrow\quad |d\varphi^{-1}(\rho)v|_{\varphi(\rho)}\geq\lambda |v|_\rho. \end{align} The hyperbolicity of $\varphi$ on $\mathcal K$ now follows by adapting the proof of Lemma~\ref{l:geodfinal}, using~\eqref{e:conifer-1}--\eqref{e:conifer-3} in place of~\eqref{e:coney-1}--\eqref{e:coney-4}. \end{proof}
{ "timestamp": "2018-05-31T02:01:37", "yymm": "1805", "arxiv_id": "1805.11660", "language": "en", "url": "https://arxiv.org/abs/1805.11660", "abstract": "These expository notes present a proof of the Stable/Unstable Manifold Theorem (also known as the Hadamard--Perron Theorem). They also give examples of hyperbolic dynamics: geodesic flows on surfaces of negative curvature and dispersing billiards.", "subjects": "Dynamical Systems (math.DS)", "title": "Notes on hyperbolic dynamics", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9935117315560086, "lm_q2_score": 0.805632181981183, "lm_q1q2_score": 0.8004050241173706 }
https://arxiv.org/abs/2204.04729
On dually-CPT and strong-CPT posets
A poset is a containment of paths in a tree (CPT) if it admits a representation by containment where each element of the poset is represented by a path in a tree and two elements are comparable in the poset if and only if the corresponding paths are related by the inclusion relation. Recently Alcón, Gudiño and Gutierrez introduced proper subclasses of CPT posets, namely dually-CPT, and strongly-CPT. A poset $\mathbf{P}$ is dually-CPT, if and only if $\mathbf{P}$ and its dual $\mathbf{P}^{d}$ both admit a CPT representation. A poset $\mathbf{P}$ is strongly-CPT, if and only if $\mathbf{P}$ and all the posets that share the same underlying comparability graph admit a CPT representation. Where as the inclusion between Dually-CPT and CPT was known to be strict. It was raised as an open question by Alcón, Gudiño and Gutierrez whether strongly-CPT was a strict subclass of dually-CPT. We provide a proof that both classes actually coincide.
\section{Introduction} A poset is called a containment order of paths in a tree (CPT for short) if it admits a representation by containment where each element of the poset corresponds to a path in a tree and for two elements $x$ and $y$, we have $x < y$ in the poset if and only if the path corresponding to $x$ is properly contained in the path corresponding to $y$. Several classes of posets are known to admit specific containment models, for example, containment orders of circular arcs on a circle~\cite{NiMaNa88, RoUr82}, containment orders of axis-parallel boxes in $\mathbb{R}^d$ \cite{GoSc89}, or containment orders of disks in the plane \cite{BrWi89, Fi88, Fi89} to cite just a few. All the aforementioned classes, as well as CPT, generalize the class {CI} of containment orders of intervals on a line~\cite{DU-MI-41}. It is well known that this class coincides with the class of $2$-dimensional posets and are also equivalent to the transitive orientations of permutation graphs~\cite{Go2004}. In 1984, Corneil and Golumbic observed that a graph $G$ may be the comparability graph of a CPT poset, yet a different transitive orientation of $G$ may not necessarily have a CPT representation, (see Golumbic~\cite{GO-84}). This stands in contrast to poset dimension, interval orders, unit interval orders, box containment orders, tolerance orders and others which are comparability invariant. Golumbic and Scheinerman~\cite{GoSc89} called such classes \emph{strong containment poset classes}. Recently, interest in CPT posets has been revived and several groups of researchers have considered various aspects of this class \cite{ALC-GUD-GUT-2, AGG-k-tree, GolumbicL21, MajumderMR21}. Since the CPT posets are not a strong containment class, Alc\'on, Gudi\~{n}o and Gutierrez~\cite{ALC-GUD-GUT-2} introduced the study of the subclasses dually-CPT and strongly-CPT posets. A poset ${\mathbf P}$ is called \emph{dually-CPT} if ${\mathbf P}$ and its dual ${\mathbf P}^{d}$ admit a CPT representation. A poset ${\mathbf P}$ is called \emph{strongly-CPT} if ${\mathbf P}$ and all the posets that share the same underlying comparability graph admit CPT representations. From the definition it is clear that the class of strongly-CPT posets is included in the class of dually-CPT posets. Many families of separating examples are now known between the class of dually-CPT and general CPT posets, however, concerning the strongly and dually-CPT, it was left as an open problem for many years to determine whether the inclusion is strict or if the two classes coincide. We present in this paper a solution to this question with the following main theorem. \begin{thm} \label{thm:Main} A poset ${\mathbf P}$ is strongly-CPT if and only if it is dually-CPT. \end{thm} To prove our main result we rely on the link between modular decomposition of the underlying comparability graph and its transitive orientations. Our strategy consists of considering a dually-CPT poset and proving that any poset with the same comparability graph also admits a CPT representation. At first we consider the representation and perform some modifications to obtain a representation with particular properties. Once this is done, we rely on the specific structure of modules in dually-CPT posets, and we provide a method to obtain the representation of any poset with the same comparability graph. ~\\ The paper is organized as follows: In Section \ref{sec:Def}, we present the definitions related to posets, CPT and modular decomposition and recall some fundamental results that we will use throughout the paper. In Section \ref{sec:TrivPathsModules}, we prove that for dually-CPT posets it is possible to obtain a representation where no element of a strong module is represented by a trivial path. Then, in Section \ref{sec:ModuleVsTrivPath}, we show how to modify a CPT representation of a dually-CPT poset so that either the paths of a strong module do not end on a trivial path or the considered module admits very specific properties. Finally, in Section~\ref{sec:Substition}, we show how to use an operation called substitution to prove our main result. \section{Definitions and notations} \label{sec:Def} A \emph{partially ordered set} or \emph{poset} is a pair ${\mathbf P}=(X,P)$ where $X$ is a finite non-empty set and $P$ is a reflexive, antisymmetric and transitive binary relation on $X$. The elements of $X$ are also called \emph{vertices} of the poset. As usual, we write $x \leq y$ in ${\mathbf P}$ for $(x,y)\in P$; and $x<y$ in ${\mathbf P}$ when $(x,y)\in P$ and $x\neq y$. If $x<y$ or $y<x$, we say that $x$ and $y$ are \emph{comparable} in ${\mathbf P}$ and write $x\perp y$. When there is no relationship between $x$ and $y$ we say that they are \emph{incomparable} and write $x\parallel y$. An element $x$ is \emph{covered} by $y$ in \textbf{P}, denoted by $x<:y$ in \textbf{P}, when $x<y$ and there is no element $z\in X$ for which $x<z$ and $z<y$. The \emph{down-set} $\{x\in X: x< z\}$ and the \emph{up-set} $\{x\in X : z< x\}$ of an element $z$ are denoted by $D(z)$ and $U(z)$, respectively. We let $D[z]=D(z) \cup \{z\}$ and $U[z]=U(z) \cup \{z\}$. The \emph{dual} of ${\mathbf P}=(X,P)$ is the poset ${\mathbf P}^d=(X,P^d)$ where $x\leq y$ in ${\mathbf P}^d$ if and only if $y \leq x$ in ${\mathbf P}$. A \emph{containment representation} $\Model{{\mathbf P}}$ or \emph{model} of a poset ${\mathbf P}=(X,P)$ maps each element $x$ of $X$ into a set $\Path{x}$ in such a way that $x < y$ in ${\mathbf P}$ if and only if $W_x$ is a proper subset of $W_y$. We identify the containment representation $\Model{{\mathbf P}}$ with the set family $\{\Path{x}\}_{x\in X}$. A poset ${\mathbf P}=(X,P)$ is a \emph{containment order of paths in a tree}, or $CPT$ poset for brevity, if it admits a containment representation $\Model{{\mathbf P}} = \{W_x\}$ where every $W_x$ is a path of a tree $T$, which is called the \emph{host tree} of the model. When $T$ is a path, ${\mathbf P}$ is said to be a \emph{containment order of intervals} or $CI$ poset for short. (We generally consider a path as the set of vertices that induces it.) The comparability graph $G_{{\mathbf P}}$ of a poset ${\mathbf P}=(X,P)$ is the simple graph with vertex set $V(G_{{\mathbf P}})=X$ and edge set $E(G_{{\mathbf P}})=\{xy: x\perp y\}$. In what follows, a poset ${\mathbf P}$, such that $G_{{\mathbf P}}$ is complete (resp. without edges), is called a \emph{total order} (resp. an \emph{empty order}). We say that two posets are \emph{associated} if their comparability graphs are isomorphic. A graph $G$ is a \emph{comparability graph} if there exists some poset ${\mathbf P}$ such that $G=G_{{\mathbf P}}$. A \emph{transitive orientation} $\overrightarrow{E}$ of a graph $G$ is an assignment of one of the two possible directions, $\overrightarrow{xy}$ or $\overrightarrow{yx}$, to each edge $xy\in E(G)$ in such a way that if $\overrightarrow{xy}\in \overrightarrow{E}$ and $\overrightarrow{yz}\in \overrightarrow{E}$ then $\overrightarrow{xz}\in \overrightarrow{E}$. The graphs whose edges can be transitively oriented are exactly the comparability graphs \cite{GA-67, GH-HO-62, Go2004}. Furthermore, given a transitive orientation $\overrightarrow{E}$ of a graph $G$, we let ${\mathbf P}_{\overrightarrow{E}}$ denote the poset $(V(G),P_{\overrightarrow{E}})$ where $u<v$ in ${\mathbf P}_{\overrightarrow{E}}$ if and only if $\overrightarrow{uv}\in \overrightarrow{E}$. The comparability graph of ${\mathbf P}_{\overrightarrow{E}}$ is $G$. Thereby, the transitive orientations of $G$ are put in one-to-one correspondence with the posets whose comparability graphs are $G$. Let ${\mathbf P}=(X, P)$ be a poset. A set $M\subseteq X$ is a \emph{module} (\emph{homogeneous set} \cite{GA-67}) if for every $y\in X-M$, either $y\perp x$ for all $x\in M$, or $y\parallel x$ for all $x \in M$. The whole set $X $ and the singleton sets $\left\{x\right\}$, for any $x\in X$, are modules of ${\mathbf P}$. These modules are called \emph{trivial modules}. A poset ${\mathbf P}$ is \emph{prime} or \emph{indecomposable} if all its modules are trivial. Otherwise ${\mathbf P}$ is \emph{decomposable} or \emph{degenerate}. A module $M$ is \emph{strong} if for all modules $M'$ either $M\cap M'={\varnothing}$ or $M\subseteq M'$ or $M'\subseteq M$. A module (respectively, strong module) $M\neq X$ is called \emph{maximal} if there exists no module (respectively, strong module) $Y$ such that $M\subset Y \subset X$. \begin{thm}(Modular decomposition theorem) \cite{GA-67} Let ${\mathbf P}=(X, P)$ be a poset with at least two vertices. Then exactly one of the following three conditions is satisfied: \begin{enumerate} \item [(i)] $G_{{\mathbf P}}$ is not connected and the maximal strong modules of ${\mathbf P}$ are the connected components of $G_{{\mathbf P}}$. \item [(ii)] $\overline{G_{{\mathbf P}}}$ is not connected and the maximal strong modules of ${\mathbf P}$ are the connected components of $\overline{G_{{\mathbf P}}}$. \item [(iii)] $G_{{\mathbf P}}$ and $\overline{G_{{\mathbf P}}}$ are connected. There is some $Y\subseteq X$ and a unique partition $\mathcal{S}$ of $X$ such that \begin{enumerate} \item [(a)] $|Y|\geq 4$, \item [(b)] ${\mathbf P}\left[Y\right]$ is the biggest prime subposet of ${\mathbf P}$ (in the sense that it is not included in any other prime subposet), \item [(c)] for every part $S$ of the partition $\mathcal{S}$, $S$ is a module of ${\mathbf P}$ and $|S\cap Y|=1$. \end{enumerate} \end{enumerate} \end{thm} The previous theorem defines a partition $\mathcal{M}({\mathbf P}) = \{M_1,...,M_k\}$ of $X$, which is called the \emph{canonical partition} or \emph{maximal modular partition} of ${\mathbf P}$. In the first case, $G_{{\mathbf P}}$ is said to be \textit{parallel} or \emph{stable} and the partition is formed by the vertices of the connected components of $G_{{\mathbf P}}$. In the second case, $G_{{\mathbf P}}$ is \emph{series} or \emph{clique} and the partition is formed by the vertices of each connected component of $\overline{G_{{\mathbf P}}}$. And, in the last case, $G_{{\mathbf P}}$ is \emph{neighborhood} or \emph{prime}, and the partition is $\mathcal{S}$. The \emph{quotient poset} of ${\mathbf P}$, denoted by ${\mathbf P}/\mathcal{M}({\mathbf P})$, has a vertex $v_i$ for each part $M_i$ of $\mathcal{M}({\mathbf P})$; and two vertices $v_i$ and $v_j$ of ${\mathbf P}/\mathcal{M}({\mathbf P})$ are comparable if and only if for all $x\in M_i$ and for all $y\in M_j$, $x \perp y$ in ${\mathbf P}$. The quotient poset is \emph{empty} (iff $G_{{\mathbf P}}$ is parallel), a \emph{total order} (iff $G_{{\mathbf P}}$ is series) or \emph{indecomposable} (iff $G_{{\mathbf P}}$ is neighborhood). On some occasions, when referring to a module, we will mean the subposet induced by it. For instance, we will say that a module $M$ of ${\mathbf P}$ is $CI$ or that it is prime, meaning that ${\mathbf P}(M)$ is. This will be clear from the context and will cause no confusion. \begin{thm}\cite{GA-67} \label{t:indecommposable-poset} Given posets ${\mathbf P}$ and ${\mathbf P}'$, if $G_{{\mathbf P}}=G_{{\mathbf P}'}$ and ${\mathbf P}$ is indecomposable, then ${\mathbf P}'={\mathbf P}$ or ${\mathbf P}'={\mathbf P}^d$. \end{thm} \begin{prop}\cite{GA-67} \label{p:associated-posets} Given posets ${\mathbf P}$ and ${\mathbf P}'$, if $G_{{\mathbf P}}=G_{{\mathbf P}'}$, then ${\mathbf P}$ and ${\mathbf P}'$ have the same strong modules and, consequently, $\mathcal{M}({\mathbf P})=\mathcal{M}({\mathbf P}')$. \end{prop} Given a vertex $v$ of a poset ${\mathbf P}=(X, P)$ and a poset ${\mathbf H}=(X_1, H)$, \emph{substituting or replacing} $v$ by ${\mathbf H}$ in ${\mathbf P}$ results in the poset ${\mathbf P}_{{\mathbf H}\rightarrow v}=\left(X-\{v\}\cup X_1, P_{{\mathbf H}\rightarrow v}\right)$ such that $P_{{\mathbf H}\rightarrow v}=P-\{(x,y) : x=v \vee y=v\} \cup H \cup \{(x,y):x\in X_1 \wedge y\in U(v)\}$ $\cup \{(x,y):y\in X_1 \wedge x\in D(v)\}$. \begin{thm} \label{t:orientations} Let $\mathcal{M}({\mathbf P}) = \{M_1, . . . , M_k\}$ be the maximal modular partition of a connected poset ${\mathbf P}=(X, P)$ whose quotient is prime, and call ${\mathbf H}$ the quotient poset ${\mathbf P}/\mathcal{M}({\mathbf P})$. A poset ${\mathbf Q}$ is associated to ${\mathbf P}$ if and only if there exist posets ${\mathbf Q}_i$ for $1\leq i \leq k$ such that ${\mathbf Q}_i$ is associated to ${\mathbf P}_i={\mathbf P}(M_i)$ for each $i$, and ${\mathbf Q}$ is obtained by replacing each vertex $v_i$ of ${\mathbf H}$ by the poset ${\mathbf Q}_i$ or replacing each vertex $v_i$ of ${\mathbf H}^d$ by the poset ${\mathbf Q}_i$. \end{thm} \begin{thm}\cite{GA-67} \label{t:strong-property} A poset ${\mathbf P}$ is $CI$ if and only if the quotient poset and all the maximal strong modules of ${\mathbf P}$ are $CI$. \end{thm} \begin{lem} \cite{ALC-GUD-GUT-2} \label{l:neces} If $z$ is a vertex of a $CPT$ poset ${\mathbf P}$ then the subposet induced by the closed down-set of $z$ is $CI$. In particular, if ${\mathbf P}$ is dually-$CPT$, then also the subposet induced by the closed up-set of $z$ is $CI$.\end{lem} \begin{rem}\cite{GA-67} \label{r:strong} Let ${\mathbf P}$ and ${\mathbf P}'$ be associated posets. Then, ${\mathbf P}$ is a $CI$ poset if and only if ${\mathbf P}'$ is a $CI$ poset. In particular, ${\mathbf P}$ is a $CI$ poset if and only if ${\mathbf P}^d$ is a $CI$ poset. \end{rem} \begin{thm} \label{t:dually-nprime} Let ${\mathbf P}=(X, P)$ be a connected dually-$CPT$ poset. Then the quotient poset of ${\mathbf P}$ is dually-$CPT$ and every maximal strong module of ${\mathbf P}$ is $CI$. In particular, if the quotient poset is $CI$, then ${\mathbf P}$ is $CI$. \end{thm} \begin{proof} Let $\mathcal{M}({\mathbf P}) = \{M_1, . . . , M_k\}$ be the maximal modular partition of ${\mathbf P}$. The quotient poset ${\mathbf H}={\mathbf P}/\mathcal{M}({\mathbf P})$ is a subposet of ${\mathbf P}$, so ${\mathbf H}$ is dually-$CPT$. We can assume that ${\mathbf P}$ is not empty, and since ${\mathbf P}$ is connected we have that ${\mathbf H}$ is connected, and so every vertex $v_i$ of ${\mathbf H}$ is in the down-set or in the up-set of some other vertex. Which implies that in ${\mathbf P}$ the whole module $M_i$ is in the up-set or in the down-set of some other vertex. It follows from Lemma \ref{l:neces} that each ${\mathbf P}_i={\mathbf P}(M_i)$ is $CI$. Therefore, by Theorem \ref{t:strong-property}, if ${\mathbf H}$ is $CI$, then ${\mathbf P}$ is $CI$.\end{proof} The converse of Theorem \ref{t:dually-nprime} is not true in general. For instance, if in the quotient poset ${\mathbf H}$ there exists a vertex $v_i$ such that in any $CPT$ representation of ${\mathbf H}$ the corresponding path $W_{v_i}$ is reduced to a vertex, then for ${\mathbf P}$ to be $CPT$ the module $M_i$ has to be a singleton. In a representation $\Model{{\mathbf P}}$ of a CPT poset ${\mathbf P}$, a subset $X$ of paths of \Model{{\mathbf P}} is called \emph{one-sided} if all the paths that represent $X$ arrive at a vertex $a$ of the host tree and all paths of $X$, except possibly one trivial path, pass through a vertex $b$ of $T$ neighbor of $a$. If all the paths of $X$ arrive at a vertex $a$ and $X$ is not one-sided, then it is called \emph{two-sided}. Addressing that issue in the proof of the main theorem will requires the following lemmas and properties. \begin{property}\cite{DU-MI-41, GO-84} \label{p:compresing-CI-model} Every $CI$ poset admits a $CI$ representation where the intersection of all the intervals used in the representation is a non-trivial interval. \end{property} \section{Trivial paths into modules} \label{sec:TrivPathsModules} The goal of this section is to prove that for any dually-CPT poset ${\mathbf P}$, there exists a representation $\Model{{\mathbf P}}$ where all the elements contained in strong modules are represented by non-trivial paths. At first we prove that if an element of module is represented by a trivial path, it does mean that the module (all its elements) are not greater than any other element not in the module. \begin{lem} Let ${\mathbf P}$ be a poset and let $M$ be a strong module of ${\mathbf P}$. If there exists a representation $\Model{{\mathbf P}}$ where an element $x$ of $M$ is represented by a trivial path, then all the elements of $M$ are not greater than any element of ${\mathbf P}$ not in $M$. \label{lem:TrivialPathModule} \end{lem} \begin{proof} Let us proceed by contradiction and let us assume that there exists an element $z \notin M$ such that $z < x$. Then in any representation \Model{{\mathbf P}} we have $\Path{z} \subset \Path{x}$ but since $\Path{x}$ is already a trivial path it cannot properly contain some other object. \end{proof} Hence from the previous lemma, if in a representation \Model{{\mathbf P}} an element of a module is represented by a trivial path, the module is a minimal subset of ${\mathbf P}$. \begin{lem} \label{lem:ModuleContained} Let $M$ be a strong module of a CPT poset ${\mathbf P}$, if in a representation $\Model{{\mathbf P}}$ one of its elements is represented as a trivial path, then there exists an element $x$ not in $M$ such that the path $\Path{x}$ contains all the paths representing the elements of $M$. \end{lem} \begin{proof} Since the poset is connected, and by the previous lemma, we know that the module cannot contain any other element, to ensure the connection outside the module, there might be at least one element $x$ that is greater than every element of $M$. \end{proof} \begin{lem} \label{lem:PassingThroughA} Let $M$ be a strong module of a CPT poset ${\mathbf P}$. If in a representation $\Model{{\mathbf P}}$ one of its elements $z$ is represented as a trivial path, then this path is hosted on some vertex $a$ of $T$. If for an element $x$ not in $M$ its path $\Path{x}$ passes through $a$, then $\Path{x}$ has to contain all the paths corresponding to the elements of $M$. \end{lem} \begin{proof} From the definition of a module, every element not in the module is either completely disconnected from $M$ or completely connected to $M$. In that case, if for an element $x$, in a representation $\Model{{\mathbf P}}$ its path $\Path{x}$ passes through $a$, then it is connected to the element $z$. Hence it has to be connected to every element of $M$. In addition, in a transitive orientation of a graph, the containment relation between $x$ and the elements of $M$ is the same for every element of $M$. Thus if $\Path{x}$ contains $\Path{z}$ it contains all the paths of the elements of $M$. \end{proof} \begin{lem} \label{lem:non-triv} Let $M$ be a strong module of a CPT poset ${\mathbf P}$. If in a representation $\Model{{\mathbf P}}$ one of its elements is represented as a trivial path and $M$ is a clique or prime module, then there exists at least one element of $M$ represented as a non-trivial path. \end{lem} In the case of dually-CPT posets, the next three lemmas consider the presence of trivial paths in a representation of strong modules and show how to obtain an equivalent representation where all the elements of the module are non-trivial paths. For these lemmas, we consider each strong module to be a CI poset and the element of the module represented by a trivial path is denoted by $z$. \begin{lem} \label{lem:CliqueModuleTrivialPath} Let $M$ be a strong CI clique module of a dually-CPT poset ${\mathbf P}$. If an element $z$ of $M$ is represented by a trivial path in a representation $\Model{{\mathbf P}}$, then there exists a representation $\Model[']{{\mathbf P}}$ where $z$ is represented as non-trivial path. \end{lem} \begin{proof} By Lemma \ref{lem:ModuleContained} we know that there exists an element $x$ such that all the paths of $M$ are contained in $\Path{x}$ in all CPT representations. Let us consider three cases. ~ \\(1) Suppose the trivial path of $z$ is not an extremity of any path that represents the elements of $M$. Let $a$ be the vertex of $T$ that hosts the trivial path of $z$. Since $\Path{z}$ is not an extremity of any path of $M$, $a$ admits at least one neighbor $b$ in $T$ such that all the paths of $M$ (except for $z$) pass through $b$ (see Figure \ref{fig:CliqueModule}$(i)$). Let us subdivide the edge $a,b$ by adding a vertex $c$. Then it suffices to replace the trivial path of $z$ by a non-trivial path that goes from $c$ to $a$ in $T$. The containment relations among $M$ are preserved and no new containment relation is added nor deleted with respect to the elements not in $M$. ~ \\(2) Suppose now, the trivial path of $z$ is a common extremity for all the elements of $M$ and $M$ is one-sided (see Figure \ref{fig:CliqueModule}$(ii)$). We proceed as in the previous case; we consider a vertex $b$ of $T$ that is a neighbor of $a$ and such that all the paths of $M$ except for $z$ pass through $b$. Since $M$ is a clique, it only admits at most one element represented by a trivial path, such a vertex $b$ exists, then we subdivide the edge by adding a vertex $c$ and the path of $z$ goes from $a$ to $b$. Note that the technique still works if some paths of $M$ continue after $a$. ~ \\(3) Suppose now, the trivial path of $z$ is the common extremity for some paths of the module in a 2-sided manner (see Figure \ref{fig:CliqueModule}$(iii)$). Let $b$ and $c$ be two vertices of $T$ that are neighbors of $a$, such that $b$ and $c$ lie on the path of $x$, $x$ being an element not in $M$ that contains all elements of $M$. We can partition the elements of $M$ into three sets: $B$ the elements for which paths arrive at $a$ and pass through $b$, $C$ defined in a similar way but \emph{w.r.t.} $c$ instead of $b$, and $A$, the paths of $M$ that go through $b$ and $c$. This time we need to subdivide the edges $a,b$ and $a,c$ of $T$. We add a vertex $i$ between $a$ and $b$ and a vertex $j$ between $a$ and $c$. Then it suffices to extend the paths of $B$ until $j$ and the paths of $C$ until $i$. The path of $z$ now goes from $i$ to $j$. By subdividing several times the edges $a,b$ and $a,c$, we can make sure that all the extremities are distinct. \end{proof} \begin{figure} \begin{center} \includegraphics[width=\textwidth]{CliqueModule} \end{center} \caption{Representation of cliques modules with trivial paths.} \label{fig:CliqueModule} \end{figure} \begin{lem} \label{lem:StableModuleTrivialPath} Let $M$ be a strong CI stable module of a dually-CPT poset ${\mathbf P}$. If an element $z$ of $M$ is represented by a trivial path in a representation $\Model{{\mathbf P}}$, then there exists a representation $\Model[']{{\mathbf P}}$ where $z$ is represented as non-trivial path. \end{lem} \begin{proof} Let us first remark that in a strong stable module, several elements can be represented as trivial paths in a representation $\Model[']{{\mathbf P}}$. In addition, if an element of $M$ is represented by a trivial path, the trivial path is disjoint from all the other paths representing the elements of $M$. Let $z$ be such an element. We will transform $\Model{{\mathbf P}}$ such that all the elements of $M$ represented by trivial paths in $\Model{{\mathbf P}}$ will be represented by non-trivial paths. Let $a$ be the vertex of $T$ that hosts the path of $z$. Thanks to Lemma \ref{lem:ModuleContained}, we know that there exists an element $x$ of ${\mathbf P}$ such that in $\Model{{\mathbf P}}$ the paths of the elements of $M$ are contained in the path of $x$. Since $M$ is a non-trivial module it contains at least two elements, hence in $\Model{{\mathbf P}}$ there exists a vertex $b$ of $T$ that is adjacent to $a$, and $b$ is contained in all the paths of the elements not in $M$ that contain $M$, since such a path has to contain $\Path{z}$ and all the other elements of $M$. Let us denote by $U=\{u_1,u_2,\ldots,u_k\}$ the elements of $M$ that are represented by trivial paths in $\Model{{\mathbf P}}$. To obtain an equivalent representation $\Model[']{{\mathbf P}}$, we subdivide $2k-1$ times the edge $a,b$. We then rename $a$ as $a_1$, and we number the newly created vertices $a_2,a_3,\ldots,a_{2k}$ (the transformation is presented in Figure \ref{fig:StableModule}). In this new representation each element $u_i$ of $U$ is replaced by a path that goes from $a_{i}$ to $a_{k+i}$ in $T$. It remains to prove that this representation is equivalent. First observe that for any element $x$ connected to $M$, its path in $\Model{{\mathbf P}}$ contains all the elements of $M$. By the choice of vertex $b$ to perform the transformation, we can guarantee that any path of such an element $x$ will pass through $a,b$ in $\Model{{\mathbf P}}$. Since we subdivided this edge to obtain $\Model[']{{\mathbf P}}$ , this path will still pass through $a$ and $b$ and all the vertices introduced by the transformation. Now for any element $y$ not connected to $M$, we know by Lemma \ref{lem:PassingThroughA} that no path of such an element will pass through $a$. \end{proof} \begin{figure} \begin{center} \includegraphics{StableModule} \end{center} \caption{$(i)$ Representation of a stable module with elements represented by trivial paths; $(ii)$ transformation to eliminate trivial paths from the representation.} \label{fig:StableModule} \end{figure} \begin{lem} \label{lem:PrimeModuleTrivialPath} Let $M$ be a strong CI prime module of a dually-CPT poset ${\mathbf P}$. If an element $z$ of $M$ is represented by a trivial path in a representation $\Model{{\mathbf P}}$, then there exists a representation $\Model[']{{\mathbf P}}$ where $z$ is represented as non-trivial path. \end{lem} \begin{proof} For this proof, we consider three cases: (1) either $\Path{z}$ the trivial path of $z$ is properly contained (\emph{i.e.} $\Path{z}$ is not an extremity of any path of the element of $M$) in all the paths of the elements of the module $M$, or (2) there exists at least two elements $q$ and $r$ of $M$ such that $\Path{z}$ is the right bound of $\Path{q}$ and the left bound of $\Path{r}$, or (3) the path $\Path{z}$ is the right (respectively left) bound for some paths representing elements of $M$, and is not the left (respectively right) bound of any elements of $M$. These three cases are illustrated in Figure \ref{fig:PrimeModuleTrivialPath}$(i)-(iii)$. ~ \\ \noindent { (1)} Let $a$ be the vertex of $T$ that hosts $\Path{z}$, the trivial path representing $z$. By hypothesis, all the paths that represent the elements of $M$ properly contain $\Path{z}$ and thus pass through vertex $a$. Since it is a proper containment, no path of elements of $M$ (other than $z$) starts or finishes at $a$. Thus $a$ admits at least one neighbor $b$ in $T$ such that all the paths that represent elements of $M$, except for $z$, pass through $b$. To obtain a new representation $\Model[']{{\mathbf P}}$ we subdivide the edge $a,b$ by a adding a vertex $d$. Then $\Path{z}$ in $\Model[']{{\mathbf P}}$ is replaced by the path $a,d$. (See Figure \ref{fig:PrimeModuleTrivialPath}$(iv)$). Since the representation of $\Path{z}$ is the only modification of the representation, by the previous discussion all the paths that represent the elements of $M$ pass through $a$ and $b$ and as a consequence pass through $a$ and $d$ since $d$ is in between $a$ and $b$. By Lemma \ref{lem:PassingThroughA} we know that all the paths of the elements not in $M$ that pass through $a$ will also contain all the paths of $M$. Hence the modification of $\Path{z}$ preserves the containment relation of $\Model{{\mathbf P}}$. ~ \\ \noindent {(2)} Let us now consider that there exist at least two elements $q$ and $r$ of $M$ such that in $\Model{{\mathbf P}}$, the vertex $a$ is the right bound of the path $\Path{q}$ and the left bound of the path $\Path{r}$ (see Figure \ref{fig:PrimeModuleTrivialPath}$(ii)$). Let us denote by $L$ the set of elements of $M$ for which $a$ is the right bound in the representation $\Model{{\mathbf P}}$ and similarly let us denote by $R$ the set of elements of $M$ for which $a$ is the left bound in $\Model{{\mathbf P}}$. Let us remark that $L\cap R = {\varnothing}$ and some elements of $M \setminus (L\cup R)$ might not be empty. Let $b$ be the neighbor of $a$ in $T$ such that the paths of the elements of $L$ pass through $b$. And similarly let $c$ be the neighbor of $a$ in $T$ such that the paths of the elements of $R$ pass through $c$. To obtain a new representation $R'(P)$ we subdivide the edge $a,b$ $|R|+1$ times, and the edge $a,c$ $|L|+1$ times. The added vertices are called $ab_{i}$ for the vertices between $a$ and $b$ and $ac_{j}$ for the vertices between $a$ and $c$. Let $ab_{1}$ and $ac_{1}$ be the neighbors of $a$ in $\Model[']{{\mathbf P}}$. The path $\Path{z}$ now goes from $ab_{1}$ to $ac_{1}$. The left bound of the paths of the elements of $R$ are moved on the $ab_{i}$ vertices. The coordinates are chosen to preserve the containment relation. We proceed symmetrically for the paths of the elements in $L$. It remains to prove that the obtained representation still corresponds to ${\mathbf P}$. Again we know by Lemma \ref{lem:PassingThroughA} that no path of an element not connected to $M$ passes through $a$, by construction it remains valid for $a$ and for all the newly introduced vertices. For any other path their relation to $\Path{z}$ and the paths of the elements of $L$ and $R$ are unchanged. If the path $\Path{s}$ of an element $s$ was containing the path $\Path{l}$ of an element $l$ of $L$ in $\Model{{\mathbf P}}$, it is still the case in $\Model[']{{\mathbf P}}$. In that case the left bound of $\Path{l}$ is contained in $\Path{s}$ and the right bound of $\Path{s}$ will be at the right of $c$ in $\Model{{\mathbf P}}$. This property will be preserved in $\Model[']{{\mathbf P}}$. Similarly if both paths $\Path{s}$ and $\Path{l}$ were overlapping in $\Model{{\mathbf P}}$, they are still overlapping in $\Model[']{{\mathbf P}}$. ~ \\ \noindent {(3)} Since $\Path{z}$ is the right (\emph{resp.} left) bound of some paths representing some elements of $M$, and is not the left (\emph{resp.} right) bound of any other elements of $M$, there exists a vertex $b$ in $T$ adjacent to $a$ and such that all the paths representing elements of $M$ that end at $a$ pass through $b$. To obtain the new representation $\Model[']{{\mathbf P}}$ it suffices to subdivide this edge one time. Let $c$ the newly introduced vertex. Then the trivial path $\Path{z}$ in $\Model{{\mathbf P}}$ is replaced by a path going from $a$ to $c$. By the transformation, we can observe that all the paths that were containing $\Path{z}$ in $\Model{{\mathbf P}}$ still contain $\Path{z}$ in $\Model[']{{\mathbf P}}$. Let $s$ be an element of ${\mathbf P}$ such that $\Path{z} \subset \Path{s}$ in $\Model{{\mathbf P}}$. If $\Path{s}$ was containing $\Path{z}$ it had to pass through $a$ and $b$, thus by subdividing $a,b$ we can also conclude that this paths will pass through $a$ and $c$, the added vertex, in $\Model[']{{\mathbf P}}$. \end{proof} \begin{figure} \begin{center} \includegraphics{PrimeModule} \end{center} \caption{Representation of prime modules with the element $z$ represented as trivial path.} \label{fig:PrimeModuleTrivialPath} \end{figure} \begin{thm} If ${\mathbf P}$ is dually-CPT and in a representation $\Model{{\mathbf P}}$ some elements of strong modules are represented by trivial paths, then there exists an equivalent representation $\Model[']{{\mathbf P}}$ where all the paths representing elements of strong modules are non-trivial paths. \end{thm} \begin{proof} It is a direct consequence of Lemmas \ref{lem:CliqueModuleTrivialPath}, \ref{lem:StableModuleTrivialPath} and \ref{lem:PrimeModuleTrivialPath} and the fact that each time a trivial path is replaced by a non-trivial one, no trivial path is created in $\Model[']{{\mathbf P}}$. \end{proof} From the preceding theorem, we know how to obtain a representation a dually-CPT poset where all the elements contained in non-trivial strong modules are represented by non-trivial paths. Hence, in this representation some elements that do not belong to strong modules might be represented by trivial paths. \section{Ending of modules onto trivial paths} \label{sec:ModuleVsTrivPath} In the previous section we proved that for a dually-CPT poset, one can always obtain a representation where no element of a strong module is represented by a trivial path. It therefore remains to consider how the paths that represent a strong module $M$ can connect to an element $z$, not contained in a strong module, where $z$ is represented by a trivial path in the representation $\Model{{\mathbf P}}$. Since we need to reconfigure the containment relation inside the module, this operation could be prevented or constrained if the trivial path is misplaced. In the case where the trivial path is in the middle of the paths of the module, it will be easy to reconfigure the containment relation. In the opposite case, if all the paths representing elements of a module arrive at a trivial path, we cannot perform the intended operation as planned. In this section, we will identify the problematic situations, and we will show how to overcome these problems. As in the previous section, we will perform local changes to the representation to suppress problematic cases. When the paths that represent elements of a module are connected to a trivial path in a representation, several configurations could arise. The most favorable one, is when the trivial path is properly contained in the paths of the module (\emph{i.e.} the trivial path does not lie on any extremity of the path of the module). Actually this is a configuration we aim at obtaining. The other two configurations is when all the path have their extremities that end at a trivial path, or just some of them end at this trivial path. In most cases we will be able to reconfigure our representation to obtain a representation that is favorable to our purpose. \subsection{Complete ending of a module on a trivial path} Let us assume that all paths in $\Model{{\mathbf P}}$ corresponding to elements of $M$ have all their extremities end at a vertex $a$ of the host tree $T$. In that case, there are several possibilities: either all the paths that represent $M$ will arrive at $a$ by passing by a vertex $b$ of $T$ and such that $a,b$ is an edge of $T$, or there is another vertex $c$ that is a neighbor of $a$ in $T$ different from $b$ and such that some paths of the module pass through $c$. In this section, even if it is not explicitly stated, the representation of the module $M$ will contain the trivial path of $z$ located at the vertex $a$ in $T$. \begin{rem} If a strong module of a dually-CPT poset ${\mathbf P}$ is two-sided in a representation $\Model{{\mathbf P}}$, then the induced graph is not connected. Hence the strong module is a stable module. \end{rem} \begin{lem} Let $M$ be a strong module of a dually-CPT poset ${\mathbf P}$. If $M$ is one-sided in a representation \Model{{\mathbf P}} and the poset induced by $M$ is connected, then $M$ is a clique module. \end{lem} \begin{proof} If the graph induced by $M$ is connected, $M$ is either a clique module or a prime module. If $M$ is a clique module, then there is nothing to prove. If $M$ is a prime module, then the graph induced by $M$ necessarily contains an induced $P_4$. Let us show that it is not possible to represent a $P_4$ as a CPT representation where all the paths end up at a same vertex $a$ of the host tree $T$. Consider the representation of the $P_4$ as presented in Figure \ref{fig:P4}$(i)$ with the containment relation represented in Figure \ref{fig:P4}$(ii)$. For a contradiction, let us assume that such a representation exists. Since the paths of $2$ and $4$ have to contain the path of $3$ and all these paths have to arrive at vertex $a$ of the host tree, we have a configuration similar to the one depicted in Figure \ref{fig:P4}$(iii)$ and a part of the host tree is depicted in Figure \ref{fig:P4}($iv$). Since $2$ and $4$ are not connected, their paths have to diverge in $T$. Call $x$ the vertex of $T$ where these paths diverge. It remains to represent the path of $1$. Since $1$ is connected to $2$ but not to $4$, call $y$ the vertex of $T$ where the path of $1$ begins. The vertex $y$ has to lie in the proper part of the path of $2$ (see Figure \ref{fig:P4}($iv$)), and this path, by hypothesis, has to go all the way to $a$. But in that case it has to contain the path of $3$, hence there is a contradiction. \end{proof} \begin{figure} \begin{center} \includegraphics{P4} \end{center} \caption{$(i)$ a $P_4$, $(ii)$ a CI representation of $P_4$, $(iii)$ tentative representation with all the paths arriving at a vertex, $(iv)$ the host tree of the tentative representation.\label{fig:P4}} \end{figure} We have proven that if in the representation of a strong module all its paths arrive at a same vertex of the host tree, then the module is either a clique or a stable module. We now consider in which cases can we obtain an alternative representation where all the paths do not arrive at a same vertex of the host tree. When the modification is possible, we will show how by starting from $\Model{{\mathbf P}}$ one can obtain an equivalent representation $\Model[']{{\mathbf P}}$, that is, a containment representation that still corresponds to ${\mathbf P}$. \begin{lem} \label{lem:ModNotIncluded} Let $M$ be a strong module of a dually-CPT poset ${\mathbf P}$ and $\Model{{\mathbf P}}$ a representation of ${\mathbf P}$ where all the paths of $M$ arrive at a same vertex. If there is no element of ${\mathbf P}$ that contains the elements of $M$, then there exists an alternative representation $\Model[']{{\mathbf P}}$ of ${\mathbf P}$ where all the paths of $M$ will have different endpoints. \end{lem} \begin{proof} Let us assume that all the paths of a strong module $M$ arrive at a vertex $a$ in the representation $\Model{{\mathbf P}}$. If there is no element of ${\mathbf P}\setminus M$ that contains all the paths of $M$, then we can add a new branch to the host tree starting at $a$ and ending at $b$ (see Figure \ref{fig:ModNotIncluded}). Let us denote by $k$ the cardinality of $M$. In order to guarantee that all paths end on a dedicated vertex, the new branch needs to have at least $k$ new vertices. It is easy to make sure that the containment relation inside the module is not altered in this new representation. It is simple to notice that the previous containments of ${\mathbf P}$ are preserved by this modification and no new containment is added since the branch only contains paths of $M$. \end{proof} \begin{figure} \begin{center} \includegraphics{ModNotIncluded} \end{center} \caption{Example of modification on a representation of a poset ${\mathbf P}$} \label{fig:ModNotIncluded} \end{figure} We now consider the case when there is at least one element $x$ not in $M$ that is greater than all the elements of $M$. In that case $M$ is either one-sided or two-sided. Let us start with this second case. \begin{lem} \label{lem:Stable-Path} Let $M$ be a strong stable module of a dually-CPT poset ${\mathbf P}$ and let $x$ be an element of ${\mathbf P}\setminus M$ that contains all the elements of $M$. Let us assume that in a representation $\Model{{\mathbf P}}$ all the paths of $M$ arrive at a vertex $a$ of $T$. Then there exists an equivalent representation $\Model[']{{\mathbf P}}$ where all the endpoints of the paths of $M$ near $a$ are distinct. \end{lem} \begin{proof} By hypothesis, since the elements of $M$ are all contained in an element $x$ of ${\mathbf P}$, it means that in any representation $\Model{{\mathbf P}}$ of ${\mathbf P}$ the union of the paths of $M$ is a path. If in a representation $\Model{{\mathbf P}}$ of ${\mathbf P}$ all the paths of $M$ arrive at $a$, let $b$ and $c$ be the immediate neighbors of $a$ on $T$ along the path that hosts all the paths of $M$. % Since the strong module considered is stable and in the representation every element lies under the path of $x$, the module is two-sided at $a$. Since $M$ is two-sided in the representation, its elements can be partitioned into two sets $B$ and $C$ as follows: An element $r$ is in $B$ if its path in $\Model{{\mathbf P}}$ passes by the vertex $b$. Similarly, an element $s$ is in $C$ if its path in $\Model{{\mathbf P}}$ passes by $c$ (see Figure \ref{fig:Stable2-sided}). % To obtain $\Model[']{{\mathbf P}}$ it suffices to subdivide the edges $a,b$ and $a,c$ of $T$. All the paths of the elements of $B$ that previously ended at $a$ will now end between $a$ and $c$. Hence it is necessary to add $|B|$ new vertices between $a$ and $c$. In a symmetric manner, the paths of the elements of $C$ will be elongated to end on a new vertex between $a$ and $b$; thus it is necessary to add $|C|$ new vertices between $a$ and $b$. It is simple to see that the introduced modification does not alter the containment relationship. Any path that contained all the elements of $M$ will still contain all the elements of $M$. And any path that crossed the section of tree spanned by the elements of $M$ but did not contain them, will still not contain them. \end{proof} \begin{figure} \begin{center} \includegraphics{Stable2-sided} \end{center} \caption{Modification of the representation of a two-sided stable module.} \label{fig:Stable2-sided} \end{figure} \begin{lem} \label{lem:Clique-Proper} Let $M$ be a strong clique module of a dually-CPT poset ${\mathbf P}$ and let $x$ be an element of ${\mathbf P}\setminus M$ that contains all the elements of $M$. If in a representation $\Model{{\mathbf P}}$ all the paths of $M$ arrive at a vertex $a$, then $M$ does not contain any other strong module. \end{lem} \begin{proof} Because of the element $x$, the union of all the paths of the elements of $M$ in $\Model{{\mathbf P}}$ is included in the path of $x$ and hence itself forms a path. Since all these paths are bounded at $a$, then for any pair of elements $p$ and $q$ of $M$ either the path of $p$ is contained in the path of $q$ or the converse. There is no pair of non-adjacent vertices. As a consequence it does not contain any other module. \end{proof} \begin{figure} \begin{center} \includegraphics[width=\textwidth]{Clique-Path} \end{center} \caption{Configuration of a clique path. The set of paths $C$ represents the paths of a strong clique module $M$} \label{fig:Clique-Path} \end{figure} Let $M$ be a strong clique module with representation $\Model{{\mathbf P}}$ where all the paths of the elements of $M$ stop at a vertex $a$. We say that $M$ is \emph{free} in $\Model{{\mathbf P}}$ if there is at least one vertex $b$ of $T$ such that $a,b$ is an edge of $T$, no path of $M$ passes through $b$ and all the paths that contain the paths of $M$ pass through $b$. (See Figure \ref{fig:Clique-Path} $(i)$ and $(ii)$.) \begin{lem} \label{lem:Clique-Path} Let $M$ be a strong clique module of a dually-CPT poset ${\mathbf P}$, and let $x$ be an element of ${\mathbf P}\setminus M$ that contains all the elements of $M$. If $M$ is free in a representation $\Model{{\mathbf P}}$ where all the paths of $M$ arrive at a vertex $a$, then we can find a representation $\Model[']{{\mathbf P}}$ where all the endpoints of $M$ arrive on different vertices of $T$. \end{lem} \begin{proof} Since $M$ is free in $\Model{{\mathbf P}}$ we can re-use the technique used in Lemma \ref{lem:Stable-Path} by subdividing the edge $a,b$ of $T$. \end{proof} Thanks to Lemmas \ref{lem:ModNotIncluded}, \ref{lem:Stable-Path}, and \ref{lem:Clique-Path}, we know how modify a representation in almost all the cases. However, one case is not covered, namely, when the module is a clique and it is blocked. We say that a strong clique module bounded at a vertex $a$ in a representation $\Model{{\mathbf P}}$ is \emph{blocked} if it is not free. There are two reasons why $M$ may be blocked: (1) It might be because a path that contains the elements of $M$ also stops at $a$, or (2) because there are two elements $x$ and $y$ that contain all the elements of $M$ and in $\Model{{\mathbf P}}$ their corresponding paths diverge at $a$. (See Figure \ref{fig:Clique-Path}$(iii)$.) \begin{rem} \label{rem:CliqueBlocked} Let $M$ be a strong clique blocked module of a dually-CPT. From Lemma \ref{lem:Clique-Proper}, we know that it does not contain any other strong module. Hence, a reconfiguration of this subposet is just a matter of relabelling the elements. \end{rem} \subsection{Partial ending of a module on a trivial path} In the previous section, we proved that whenever a module is connected to an element $z$ of ${\mathbf P}$ represented by a trivial path in a representation $\Model{{\mathbf P}}$ and all paths that represent the element of $M$ end at this path, we can either alter the representation to ensure that all the paths do not end on that trivial path or, the module is a clique and does not contain any other modules. Hence it is possible to alter the containment relation. If, in the completely opposite direction, a module $M$ is connected to an element $z$ represented by a trivial path, but no path that represents an element of $M$ ends at this trivial path, it does not create any problem to change the containment relation of the module. The last case to consider is when $M$ is connected to a trivial path, but only some paths of $M$ (not all) end at this trivial path. We will prove that in that case an equivalent representation, where no path of $M$ ends at this trivial path, can be obtained. \begin{lem} \label{lem:PartialEndingModule} Let $M$ be a strong module of a dually-CPT poset ${\mathbf P}$ connected to an element $z$ ($z\notin M$). If in a representation $\Model{{\mathbf P}}$ the element $z$ is represented by a trivial path $\Path{z}$ and the paths of some elements of $M$ end at the path of $\Path{z}$ and some other paths of elements of $M$ properly contain $\Path{z}$, then there exists an equivalent representation $\Model[']{{\mathbf P}}$ where no element of $M$ ends at a trivial path. \end{lem} \begin{proof} Let $I$ denote the set of elements not in $M$ such that the paths of the elements of $I$ are contained in the paths of the elements of $M$. In the representation $\Model{{\mathbf P}}$ all the paths that represent the element of $I$ are all contained in $\cap_{m\in M} P_m$. Since by hypothesis not all the paths of $M$ end at a trivial path, if there are some elements of $M$ that end paths represented by trivial paths, there are at most two trivial paths in that situation. Call these trivial paths $y$ and $z$. Let us assume that the part common to all the paths of $M$ in $\Model{{\mathbf P}}$ is on a horizontal line, and that \emph{w.l.o.g.}~that $\Path{y}$ is the leftmost and $\Path{z}$ is the rightmost of this common part. We assume further, in the representation $\Model{{\mathbf P}}$, that $\Path{z}$ lies on vertex $a$ of $T$ and $\Path{y}$ lies on vertex $b$ of $T$. We denote by $L$ (resp.~$R$) the set of all elements of $M$ whose paths in $\Model{{\mathbf P}}$ end at $b$ (resp.~at $a$.) Note that there is at most one element of $M$ that belongs to both $L$ and $R$, since the containment relation is proper. There are two cases to consider: (1) either there is no element $x$ such that all the paths of $M$ are contained in the path of $x$, or (2) such an element $x$ exists. ~\\(1) For the first case, let us assume that such an element does not exists. Hence there is no path in $\Model{{\mathbf P}}$ that contains any path of the elements of $M$. In that case, to obtain an equivalent representation, in $T$ we can add one path with $|M|$ new vertices connected to $a$ and another path with $|M|$ new vertices connected to $b$. Since the poset induced by $M$ is CI, it suffices to represent this module as a containment of intervals using these new branches for the endpoints. The transformation process is presented in Figure \ref{fig:PartialEndingCase1}. The containment relation between elements of $R$ (resp. $L$) and $I$ remain unchanged. Moreover, for any element $q$ not connected to $M$, since the endpoints of the paths of the elements of $M$ have been relocated in the two new branches, there is no containment relation between $\Path{q}$ and the paths of the elements of $M$, since $\Path{q}$ does not contain any of the new branches in $\Model[']{{\mathbf P}}$. ~\\(2) Let us now consider the case when there is an element $x$ not in $M$ such that in $\Model{{\mathbf P}}$ the path of $x$ contains all the paths of the elements of $M$. In the host tree $T$ we denote by $c$ the neighbor of $a$ such that no path of $R$ passes through $c$ but some paths of elements of $M$ do (by our initial hypothesis). Let $d$ be the neighbor of $b$ in $T$ such that paths of some elements of $M$ pass through but no element of $L$ does. To obtain an alternative representation $\Model[']{{\mathbf P}}$ we subdivide the edge $a,c$ $|R|$ times and subdivide the edge $b,d$ $|L|$ times. (This transformation is presented in Figure \ref{fig:PartialEndingCase2}). Then it is just a matter of extending the paths of the elements in $R$ such that they end on a vertex located between $a$ and $c$. For each element of $R$, its new ending vertex is determined according to the containment relation in $R$. For the elements of $L$, we proceed in a similar manner. It remains to prove that the new representation still represents the poset ${\mathbf P}$. The only paths that are transformed are the paths that correspond to elements of $R$ and $L$. Without loss of generality, let $l$ be an element of $R$ and let $\Path{l}$ be its path in $\Model[']{{\mathbf P}}$. Since $\Path{l}$ has been extended, it is clear that all the paths in $\Model{{\mathbf P}}$ that were contained in $\Path{l}$ remain contained in $\Model[']{{\mathbf P}}$. In addition, since the extension occurred between $a,c$ or $b,d$. Equivalently let $k$ be an element of ${\mathbf P}$. If $\Path{l} \subset \Path{k}$ in $\Model{{\mathbf P}}$ then $\Path{l} \subset \Path{k}$ in $\Model[']{{\mathbf P}}$. If $k$ is an element of $R$, by the transformation we ensure that the containment relation is preserved. If $k$ is not an element of $R$, then in $\Model{{\mathbf P}}$, the path $\Path{k}$ passed by vertex $c$ of $T$, hence by extending $\Path{l}$, it will not reach $c$, then it is still contained in $\Path{k}$ in $\Model[']{{\mathbf P}}$. Let us now consider an element $q$ such that $\Path{q}$ intersects $\Path{l}$ but there is no containment relation in $\Model{{\mathbf P}}$. If $\Path{l} \cup \Path{q}$ is not a path in $\Model{{\mathbf P}}$ then it contains a claw pattern and this pattern will be preserved in $\Model[']{{\mathbf P}}$. Let us now consider the case when $\Path{l} \cup \Path{q}$ forms a path in $\Model{{\mathbf P}}$. If $\Path{q}$ passes through $a$ in $\Model{{\mathbf P}}$ it has one endpoint contained between the endpoint of $\Path{l}$. Thus the first endpoint of $\Path{q}$ is at the left of $a$ in $\Model{{\mathbf P}}$ and the endpoint at the right of $c$ (possibly $c$). Since $\Path{l}$ does not reach $c$ in $\Model[']{{\mathbf P}}$, the overlap relation is preserved in the new representation. If both paths were disjoint, they remain disjoint in $\Model[']{{\mathbf P}}$. \end{proof} \begin{figure}[h!] \begin{center} \includegraphics{PartialEndingCase1} \end{center} \caption{Illustration of case (1) of Lemma \ref{lem:PartialEndingModule}. Elements $1,2,3$ and $4$ are parts of the modules. The module is connected to the elements represented by the paths in the box $B$. Elements $1$ and $3$ belong to $L$ and elements $2$ and $3$ belong to $R$.} \label{fig:PartialEndingCase1} \end{figure} \begin{figure}[h!] \begin{center} \includegraphics{PartialEndingCase2} \end{center} \caption{The same example as in Figure \ref{fig:PartialEndingCase1}, but this time there is an element $x$ not in $M$ that contains all the elements of $M$ and that prevents performing the modification of case (1).} \label{fig:PartialEndingCase2} \end{figure} From Lemmas \ref{lem:ModNotIncluded}, \ref{lem:Stable-Path}, \ref{lem:Clique-Path} and Remark \ref{rem:CliqueBlocked}, we can summarize the results of this section with the following theorem: \begin{thm} Let ${\mathbf P}$ be a dually-CPT poset. Either for each strong module $M$ of ${\mathbf P}$ there exists a representation $\Model{{\mathbf P}}$ such that all the paths of $M$ do not end on a trivial path, or $M$ is a clique blocked module. \label{thm:StandardizedRepresentation} \end{thm} We call a representation that fulfills the condition of the previous theorem a \emph{normalized representation}. \section{Substitution} \label{sec:Substition} The last step to obtain our main result is to prove that for any dually-CPT poset ${\mathbf P}$ all the posets $\mathcal{Q}=\{{\mathbf Q}_1,\ldots,{\mathbf Q}_l\}$ that are associated to ${\mathbf P}$ admit a CPT representation. Let us consider one particular poset ${\mathbf Q}$ of this set. If ${\mathbf Q}$ is associated to ${\mathbf P}$ it means by definition that their underlying comparability graphs are identical. We assume that ${\mathbf P}$ is not CI, otherwise the results already stand from Theorem \ref{t:strong-property}. Thus we deduce that the quotient poset of ${\mathbf P}$ is not CI, by Theorem \ref{t:dually-nprime}, and thus is prime. Since ${\mathbf P}$ and ${\mathbf Q}$ are associated, by Property \ref{p:associated-posets} they admit the same set of strong modules. The quotient ${\mathbf H}$ of ${\mathbf P}$ is obtained by keeping one element of each strong maximal module and the quotient ${\mathbf K}$ of ${\mathbf Q}$ is either equal to ${\mathbf H}$ or to its dual ${\mathbf H}^d$. Let us consider that ${\mathbf H}$ is equal to ${\mathbf K}$. To obtain a representation for ${\mathbf Q}$, we will use the normalized representation $\Model{{\mathbf P}}$ obtained for ${\mathbf P}$. From $\Model{{\mathbf P}}$ it is immediate to obtain a representation $\Model{{\mathbf H}}$ for ${\mathbf H}$ as it suffices to keep one path for each strong module of ${\mathbf P}$. In addition, since it is obtained by removing paths from a normalized representation, we can consider that all the paths that correspond to strong modules which are not clique blocked modules, do not end on trivial paths of other elements. Then we will show that for such elements, we can replace this path by an arbitrary CI poset. Finally, to obtain a CPT representation for ${\mathbf Q}$ it suffices to replace each path that is a representative of a strong module, by the corresponding CI poset in ${\mathbf Q}$. For the clique blocked modules, as they do not contain other strong modules, they correspond to total orders, hence the representation can be preserved, but the labelling has to be changed to suit the total order in ${\mathbf Q}$. Let ${v_0}$ be an element of ${\mathbf H}$ that is a representative of some maximal strong module of ${\mathbf P}$ that is not a clique blocked module. Let $\Path{v_0}=(x_1,x_2,\ldots,x_k)$ be its path in $\Model{H}$. We will assume that $k$ is at least $4$. We will show how to replace $\Path{v_0}$ by a CI poset ${\mathbf N}$. Let $\Model{{\mathbf N}}=\{I_i\}_{1\leq i\leq n}$ be a $CI$ representation of a poset ${\mathbf N}$ whose vertices are $u_1,u_2,\ldots, u_m$. Assume that the intervals $I_i$ (subpaths of a path $I$) are non-trivial, no two of them share an end vertex and there is an edge $cd$ of $I$ contained in the total intersection of the intervals $I_i$ -- this assumption is guaranteed by Proposition \ref{p:compresing-CI-model}. Name $a$ and $b$ the end vertices of the interval union of the intervals $I_i$. Clearly $[c,d]\subset [a,b]$. We also assume that $a$, $b$, $c$ and $d$ are distinct, and that neither $c$ nor $d$ are end vertices of an interval $I_i$. \begin{process} \label{p:substitution} The process of replacing in the representation $\Model{{\mathbf H}}$ the path $\Path{v_0}$ by the intervals $\{I_i\}_{1\leq i \leq n}$ of the representation $\Model{{\mathbf N}}$ consists of: \begin{itemize} \item[(i)] subdividing the edges $x_1 x_2$ and $x_{k-1} x_k$ of $T$ by adding in each one $n-1$ vertices. \item[(ii)]subdividing the edge $c d$ of $I$ by adding as many vertices as there are in $T$ between $x_2$ and $x_{k-1}$. \item[(iii)] removing from $\Model{{\mathbf H}}$ the path $W_{v_0}$ and embedding in its place the intervals of $S$ in such a way that the vertices $a$, $c$, $d$, $b$ and all others between them match with the vertices $x_1$, $x_2$, $x_{k-1}$, $x_k$ and all others between them, respectively, as it is shown in Figure \ref{f:grande}. \end{itemize} \end{process} \begin{figure} \centering \includegraphics[width=15cm]{figura_grande.pdf} \caption{Description of the Replacement process}\label{f:grande} \end{figure} \begin{lem}\label{l:substitution-1} If in $\Model{{\mathbf H}}$ the path $\Path{v_0}$ that represents a module of a dually-CPT poset ${\mathbf P}$ does not end on trivial paths, then we can obtain the representation $\Model{{\mathbf H}_{{\mathbf N}\rightarrow v_0}}$ by replacing $\Path{v_0}$ by the intervals $\{I_i\}_{1\leq i \leq n}$ of the representation $\Model{{\mathbf N}}$ in $\Model{{\mathbf H}_{{\mathbf N}\rightarrow v_0}}$. If any of the paths $I_{i}$ contains (resp.~is contained in) a path $\Path{v}$, then all the paths $I_i$ contain (resp.~are contained in) $\Path{v}$. Moreover, a path $\Path{v}$ of $\Model{{\mathbf H}}$ contains (is contained in) $\Path{v_0}$ if and only if $\Path{v}$ contains (is contained in) every one of the intervals $I_i$ in $\Model{{\mathbf H}_{{\mathbf N}\rightarrow v_0}}$. \end{lem} \begin{proof} This result is a direct consequence of two facts: first, that in $\Model{{\mathbf N}}$ no interval $\Path{v}$ of $\Model{{\mathbf H}}$ has an end-vertex between $x_1$ and $x_2$, nor between $x_{k-1}$ and $x_{k+1}$, and second, that in $\Model{{\mathbf N}}$, all the intervals $I_i$ contain the interval $x_2x_{k-1}$. See Figure \ref{f:grande}. \end{proof} \begin{lem}\label{l:substitution-3} If in $\Model{{\mathbf H}}$ the path $\Path{v_0}$, that represents a blocked clique module of a dually-CPT poset ${\mathbf P}$, ends on a trivial path, then we can obtain the representation $\Model{{\mathbf H}_{{\mathbf N}\rightarrow v_0}}$ by replacing $\Path{v_0}$ by the a collection of paths that represent a clique. \end{lem} \begin{proof} Let us assume that $\Path{z}$ is the trivial path that $\Path{v_0}$ ends on in $\Model{{\mathbf H}}$. Let us denote by $a$ the vertex of the host tree that hosts $\Path{z}$. Since the containment relation is proper, we can assume that $\Path{v_0}$ passes through at least two vertices of the host tree. One of the extremities of $\Path{v_0}$ is $a$. Let us call the other extremity $b$. Since the length of $\Path{v_0}$ is at least two, we know there exists in the host tree a vertex $c$ that is the immediate neighbor of $b$ on the path going to $a$. The vertex $c$ is possibly equal to $a$. By subdividing an appropiate number of times the edge $bc$ of the host tree, we can add as many paths as we need to place a clique module. From the transformation, it is easy to see that the containment relation is preserved with respect to the module. \end{proof} We restate here our main theorem: \begin{thm} \label{thm:Main2} A poset ${\mathbf P}$ is strongly-CPT if and only if it is dually-CPT. \end{thm} \begin{proof} Let ${\mathbf H}={\mathbf P}/\mathcal{M}({\mathbf P})$ be the quotient poset, where $\mathcal{M}({\mathbf P})=\left\{M_1, \ldots, M_k\right\}$ is the maximal modular partition of ${\mathbf P}$. Since ${\mathbf P}$ is a dually-$CPT$ poset and ${\mathbf H}$ is a subposet of ${\mathbf P}$, then ${\mathbf H}$ and ${\mathbf H}^d$ admit a normalized $CPT$-representation. If ${\mathbf H}$ is $CI$, by Remark \ref{r:strong} and Theorem \ref{t:dually-nprime}, ${\mathbf P}$ is $CI$ and so strongly-$CPT$. Thus let us assume that ${\mathbf H}$ is a prime dually-$CPT$ poset. Let ${\mathbf Q}$ be an associated poset of ${\mathbf P}$ and let ${\mathbf K}$ be its quotient poset. Since ${\mathbf P}$ and ${\mathbf Q}$ are associated, an immediate consequence is that ${\mathbf H}$ and ${\mathbf K}$ are associated; in addition by hypothesis they are both prime, hence by Theorem \ref{t:indecommposable-poset}, ${\mathbf K}$ is either equal to ${\mathbf H}$ or to ${\mathbf H}^{d}$. Let us assume, \emph{w.l.o.g.}, that ${\mathbf H}={\mathbf K}$. We will prove that ${\mathbf Q}$ admits a $CPT$ representation. By Theorem \ref{t:orientations} and \emph{w.l.o.g}, we assume that ${\mathbf Q}$ is obtained by replacing in ${\mathbf H}$ each vertex $v_i$ of ${\mathbf H}$ for ${\mathbf Q}_i={\mathbf Q}(M_i)$. By Proposition \ref{p:associated-posets}, ${\mathbf P}$ and ${\mathbf Q}$ possess the same strong modules and by Theorem \ref{t:dually-nprime} since ${\mathbf P}$ is dually-CPT, all the strong modules of ${\mathbf P}$ and ${\mathbf Q}$ are CI. For each ${\mathbf Q}_i$ we have a $CI$ representation. Let $\Model{{\mathbf H}}$ be a $CPT$ representation of ${\mathbf H}$, obtained from a normalized representation of ${\mathbf P}$. The representation is obtained by only keeping one path for each strong module of ${\mathbf P}$. For each path $\Path{v_i}$ that corresponds to a module $M_i$ of ${\mathbf Q}$, if $\Path{v_i}$ does not end on a trivial path of $\Model{{\mathbf H}}$ then it corresponds to a module which is not a blocked clique module, hence by Lemma \ref{l:substitution-1}, we can replace $\Path{v_i}$ by a CI representation of ${\mathbf Q}_{i}$. The only remaining case is if $\Path{v_i}$ ends on a trivial path in $\Model{{\mathbf H}}$. In that case, it means that it corresponds to a blocked clique module of ${\mathbf P}$ in the representation $\Model{{\mathbf P}}$. Hence by Lemma \ref{l:substitution-3}, we can replace $\Path{v_i}$ by a CI representation of the maximal strong clique module ${\mathbf Q}_i$. By proceeding in that way for each maximal strong module, we are able to obtain a CPT representation $\Model{{\mathbf Q}}$ of ${\mathbf Q}$. \end{proof}
{ "timestamp": "2022-04-12T02:25:31", "yymm": "2204", "arxiv_id": "2204.04729", "language": "en", "url": "https://arxiv.org/abs/2204.04729", "abstract": "A poset is a containment of paths in a tree (CPT) if it admits a representation by containment where each element of the poset is represented by a path in a tree and two elements are comparable in the poset if and only if the corresponding paths are related by the inclusion relation. Recently Alcón, Gudiño and Gutierrez introduced proper subclasses of CPT posets, namely dually-CPT, and strongly-CPT. A poset $\\mathbf{P}$ is dually-CPT, if and only if $\\mathbf{P}$ and its dual $\\mathbf{P}^{d}$ both admit a CPT representation. A poset $\\mathbf{P}$ is strongly-CPT, if and only if $\\mathbf{P}$ and all the posets that share the same underlying comparability graph admit a CPT representation. Where as the inclusion between Dually-CPT and CPT was known to be strict. It was raised as an open question by Alcón, Gudiño and Gutierrez whether strongly-CPT was a strict subclass of dually-CPT. We provide a proof that both classes actually coincide.", "subjects": "Discrete Mathematics (cs.DM)", "title": "On dually-CPT and strong-CPT posets", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9875683484150417, "lm_q2_score": 0.8104789132480439, "lm_q1q2_score": 0.8004033217815886 }
https://arxiv.org/abs/2111.13342
On connected components with many edges
We prove that if $H$ is a subgraph of a complete multipartite graph $G$, then $H$ contains a connected component $H'$ satisfying $|E(H')||E(G)|\geq |E(H)|^2$. We use this to prove that every three-coloring of the edges of a complete graph contains a monochromatic connected subgraph with at least $1/6$ of the edges. We further show that such a coloring has a monochromatic circuit with a fraction $1/6-o(1)$ of the edges. This verifies a conjecture of Conlon and Tyomkyn. Moreover, for general $k$, we show that every $k$-coloring of the edges of $K_n$ contains a monochromatic connected subgraph with at least $\frac{1}{k^2-k+\frac{5}{4}}\binom{n}{2}$ edges.
\section{Introduction} A classical observation of Erd\H{o}s and Rado is that in any two-coloring of the edges of the complete graph $K_n$, one of the color classes forms a connected graph. In \cite{gyarfas1977partition}, Gyárfás proves the following generalization of this observation: For any $k\geq 2$, in every $k$-coloring of the edges of $K_n$, there is a monochromatic connected component with at least $n/(k-1)$ vertices. This observation has since been extended in numerous ways, such as by replacing $K_n$ with a graph of high minimum degree \cite{gyarfas2017mindeg} or with a nearly-complete bipartite graph \cite{deblasio2020bipartite}, or adding a constraint on the diameter of the large monochromatic component \cite{Ruszinko2012diameter}. See \cite{Gyarfas2011Survey} for a survey of earlier work on the subject. The arguments used in this subject tend to focus on sparse spanning structures like double stars. As such, there is a surprising lack of progress on the corresponding question about edges in a monochromatic connected component. That is, what is the largest value of $M=M(n,k)$ such that every $k$-edge-coloring of $K_n$ has a monochromatic connected component with at least $M$ edges? This question was raised by Conlon and Tyomkyn in \cite{conlon2021ramsey}, in the context of determining the multicolor Ramsey numbers for trails (see Section~\ref{subsec:trails} for the relevant definitions). After showing that $M(n,2)=\frac{2}{9}n^2+o(n^2)$, they sketch a simple argument that shows $M(n,k)\geq \frac{1}{16k^2}n^2+O(n)$ for all $k$; with a slightly more careful analysis, their argument in fact yields $M(n,k)\geq \frac{1}{4k^2}n^2+O(n)$. In the other direction, they examine a construction of Gyárfás in \cite{Gyarfas2011Survey} to show that $M(n,k)\leq \frac{1}{2k(k-1)}n^2+O(n)$ for infinitely many values of $k$ (specifically, when $k-1$ is a prime power), conjecturing that this upper bound is tight in the case $k=3$. In this note, we improve the general lower bound on $M(n,k)$, as well as a corresponding lower bound for the trail Ramsey problem, bringing it to asymptotically within a factor $1-O\left(\frac{1}{k^2}\right)$ of the upper bound for infinitely many values of $k$. \begin{theorem} \label{thm:colorlb} For any $k\geq 2$, in every $k$-coloring of the edges of $K_n$, there is a monochromatic connected component with at least $M(n,k)\geq \frac{1}{k^2-k+\frac{5}{4}}\binom{n}{2}$ edges. \end{theorem} By building on the overarching ideas in \cite{Gyarfas2011Survey} and introducing some key new insights, we manage to strengthen this lower bound in the case $k=3$ to prove the tight lower bound conjectured by Conlon and Tyomkyn. \begin{theorem} \label{thm:color3} In every $3$-coloring of the edges of $K_n$, there is a monochromatic component containing at least a sixth of the edges. That is, $M(n,3)\geq \lceil\frac{1}{6}\binom{n}{2}\rceil$. Moreover, for $n$ sufficiently large, this bound is sharp. \end{theorem} While equality holds in this bound for sufficiently large $n$, there are, as we will see from the proof, small values of $n$ for which $M(n,3)>\lceil\frac{1}{6}\binom{n}{2}\rceil$. In particular, equality holds for all $n\geq 18$, but not for $n=17$. The key result we use is a new inequality that may be of independent interest. Given a subgraph $H$ of a complete multipartite graph $G$, it relates the edge counts of $G$ and $H$ to the largest edge count of a connected component $H'$ of $H$. \begin{theorem} \label{thm:main} Let $G$ be a complete $r$-partite graph for some $r\geq 2$, and let $H$ be a subgraph of $G$. Then $H$ contains a connected component $H'$ satisfying \[ |E(H')|\geq \frac{|E(H)|^2}{|E(G)|}. \] \end{theorem} In effect, this result settles the density analogue of the coloring question of determining $M(n,k)$. Instead of partitioning the edges of a graph $G$ into color classes, we are fixing a subgraph $H$ of $G$ with a given fraction $\delta=\frac{|E(H)|}{|E(G)|}$ of its edges, and asking about the component of $H$ with the most edges. Theorem~\ref{thm:main} can then be restated as: if $|E(H)| = \delta |E(G)|$, then $H$ contains a connected component $H'$ with $|E(H')|\geq \delta^2 |E(G)|$. Equality can be attained asymptotically when $\delta=\frac{1}{k}$ for any positive integer $k\geq 2$: Let $V(G)=V_1\cup\cdots\cup V_r$ be an $r$-partition of $V(G)$, split each $V_i$ into $k$ (roughly) equally sized vertex sets $\{V_{i,j}\}_{j=1}^k$, and for $1\leq j\leq k$, let $H_j$ be the subgraph of $G$ induced on $\bigcup_{i=1}^r V_{i,j}$. Then indeed, $H=\bigcup_{j=1}^k H_j$ is a subgraph of $G$ whose components have edge counts \[ |E(H_j)|=\frac{1}{k^2}|E(G)|+O(n) = \frac{|E(H)|^2}{|E(G)|} + O(n). \] The simplest case of Theorem~\ref{thm:main}, when $G=K_n$, already immediately implies an improvement over previously known lower bounds on $M(n,k)$. \begin{corollary}\label{cor:weaklb} In every $k$-coloring of the edges of $K_n$, there is a monochromatic connected component with at least $\frac{1}{k^2}\binom{n}{2}$ edges. \end{corollary} \begin{proof} In a $k$-coloring of the edges of $K_n$, one of the color classes has at least $\frac{1}{k}\binom{n}{2}$ edges. Taking $H$ to be this color class in Theorem~\ref{thm:main} with $G=K_n$ yields Corollary~\ref{cor:weaklb}. \end{proof} The proof of Theorem~\ref{thm:color3} requires a more detailed argument, but similarly follows from this one case of Theorem~\ref{thm:main}, while the proof of Theorem~\ref{thm:colorlb} requires the use of Theorem~\ref{thm:main} in full generality. In the next section, we give an elementary proof of Theorem~\ref{thm:main}. We then study the coloring version of the problem, as well as the corresponding trail Ramsey problem, in Section~\ref{sec:color}. \section{\texorpdfstring{Proof of Theorem~\ref{thm:main}}{Proof of Theorem 3}} \label{sec:mainpf} In the case $G=K_n$, the proof of Theorem~\ref{thm:main} is very simple, but nevertheless includes a key insight: Instead of taking a component that maximizes the number of edges right away, we consider a component $H'$ with maximum \textbf{average degree} $\bar{d}(H')=\frac{|E(H')|}{\frac{1}{2}|V(H')|}$. Two observations are crucial here: First, by the so-called \emph{generalized mediant inequality}, this highest average degree must be at least the average degree of the whole graph $H$, since \[ \frac{|E(H)|}{\frac{1}{2}|V(H)|}=\frac{\sum|E(H_i)|}{\frac{1}{2}\sum|V(H_i)|}, \] where the sums are over connected components $H_i$ of $H$, and the right hand side is a generalized mediant of the average degrees of the individual components. Second, the number of vertices of $H'$ is at least one more than its maximum degree, so $|V(H')|\geq \bar{d}(H')+1$. Letting $\delta = \frac{|E(H)|}{|E(G)|}$, we then obtain the bound \begin{align*} |E(H')|&\geq \frac{1}{2}(\bar{d}(H'))(\bar{d}(H')+1) = \binom{\bar{d}(H')+1}{2}\\ &\geq \binom{\bar{d}(H)+1}{2} = \binom{\frac{2|E(H)|}{n}+1}{2} \\ &= \binom{\delta(n-1)+1}{2} \geq \delta^2 \binom{n}{2} \\ &= \delta^2 |E(G)| = \frac{|E(H)|^2}{|E(G)|}, \end{align*} as desired. The general case will use both of these observations in a modified setting. Instead of the average degree, we will work with a slightly different quantity whose denominator is a weighted vertex count; we will then, perhaps counterintuitively, lower bound this weighted vertex count by the modified analogue of the average degree, in order to obtain the bound we seek. The core of our proof is the following general inequality. \begin{lemma} \label{lem:weightcs} If $a_1,\dots,a_r$ and $b_1,\dots,b_r$ are nonnegative real numbers, then \[ \left(\left(\sum_{i=1}^r a_i\right)\left(\sum_{i=1}^r b_i\right) - \sum_{i=1}^r a_i b_i \right)^2 \geq \left(\left(\sum_{i=1}^r a_i\right)^2 - \sum_{i=1}^r a_i^2 \right)\left(\left(\sum_{i=1}^r b_i\right)^2 - \sum_{i=1}^r b_i^2 \right). \] \end{lemma} \begin{proof} The Cauchy-Schwarz inequality yields \begin{align*} \left(\sum_{i=1}^r a_i\right)\left(\sum_{i=1}^r b_i\right) - \sum_{i=1}^r a_i b_i \geq \left(\sum_{i=1}^r a_i\right)\left(\sum_{i=1}^r b_i\right) - \sqrt{\left(\sum_{i=1}^r a_i^2\right)\left(\sum_{i=1}^r b_i^2\right)}. \end{align*} This last quantity is clearly nonnegative, since $\left(\sum_{i=1}^r a_i\right)^2\geq \sum_{i=1}^r a_i^2$ and $\left(\sum_{i=1}^r b_i\right)^2\geq \sum_{i=1}^r b_i^2$. So, \begin{align*} &\left(\left(\sum_{i=1}^r a_i\right)\left(\sum_{i=1}^r b_i\right) - \sum_{i=1}^r a_i b_i \right)^2 \geq \left(\left(\sum_{i=1}^r a_i\right)\left(\sum_{i=1}^r b_i\right) - \sqrt{\left(\sum_{i=1}^r a_i^2\right)\left(\sum_{i=1}^r b_i^2\right)}\right)^2 \\ &= \left(\sum_{i=1}^r a_i\right)^2\left(\sum_{i=1}^r b_i\right)^2 + \left(\sum_{i=1}^r a_i^2\right)\left(\sum_{i=1}^r b_i^2\right) - 2 \left(\sum_{i=1}^r a_i\right)\left(\sum_{i=1}^r b_i\right)\sqrt{\left(\sum_{i=1}^r a_i^2\right)\left(\sum_{i=1}^r b_i^2\right)} \\ &\geq \left(\sum_{i=1}^r a_i\right)^2\left(\sum_{i=1}^r b_i\right)^2 + \left(\sum_{i=1}^r a_i^2\right)\left(\sum_{i=1}^r b_i^2\right) - \left(\sum_{i=1}^r a_i\right)^2\left(\sum_{i=1}^r b_i^2\right) - \left(\sum_{i=1}^r b_i\right)^2\left(\sum_{i=1}^r a_i^2\right) \\ &= \left(\left(\sum_{i=1}^r a_i\right)^2 - \sum_{i=1}^r a_i^2 \right)\left(\left(\sum_{i=1}^r b_i\right)^2 - \sum_{i=1}^r b_i^2 \right), \end{align*} where the last inequality is an application of the AM-GM inequality. \end{proof} Given a graph $G$ and vertex sets $S,T\subseteq V(G)$, let \[ e_G(S,T)=\#\{(s,t)\in S\times T:\: s\sim_G t\}. \] We write $e(S,T)$ for $e_G(S,T)$ when the graph in question is unambiguous. Let $G[S]$ denote the induced subgraph of $G$ on the vertex set $S$. Lemma~\ref{lem:weightcs} immediately implies the following. \begin{corollary} \label{cor:weightcs} Let $G$ be a complete multipartite graph. For any $S, T\subseteq V(G)$, we have \[ e(S,T)^2\geq 4|E(G[S])||E(G[T])|. \] \end{corollary} \begin{proof} Suppose that $G$ is $r$-partite. Let $V(G)=V_1\cup \cdots \cup V_r$ be a partition of the vertices of $G$ into $r$ independent sets. Let $a_i=|V_i\cap S|$ and $b_i=|V_i\cap T|$, so \[ e(S,T)=\sum_{\substack{1\leq i,j\leq r\\ i\neq j}} a_{i}b_{j}=\left(\sum_{i=1}^r a_i\right)\left(\sum_{i=1}^r b_i\right) - \sum_{i=1}^r a_i b_i, \] while \[ |E(G[S])|=\frac{1}{2}\left(\left(\sum_{i=1}^r a_i\right)^2 - \sum_{i=1}^r a_i^2 \right), \qquad |E(G[T])|=\frac{1}{2}\left(\left(\sum_{i=1}^r b_i\right)^2 - \sum_{i=1}^r b_i^2 \right). \] Then Lemma~\ref{lem:weightcs} indeed yields $e(S,T)^2\geq 4|E(G[S])||E(G[T])|$, as desired. \end{proof} \begin{proof}[Proof of Theorem~\ref{thm:main}] Let $V=V(G)=V_1\cup\cdots\cup V_r$ be a partition of the vertices of $G$ into $r$ independent sets, let $H_1,\dots, H_k$ be the connected components of $H$, and let $V_{i,\ell}=V_i\cap H_\ell$. We have \[ |E(H)|=\sum_{j=1}^k |E(H_\ell)|, \qquad |V_i|=\sum_{\ell=1}^k |V_{i,\ell}|\: \text{ for all } i. \] For any subset $S\subseteq V$, define $f(S)=\frac{1}{2}e_G(S,V)$. Note that \[ f(S)=\frac{1}{2}\sum_{\substack{1\leq i,j\leq r\\i\neq j}} |V_{i}\cap S||V_{j}|, \] so $f(S)$ can be viewed as a weighted vertex count for $S$, with the property that \[ \sum_{\ell=1}^k f(V(H_\ell))=\frac{1}{2}\sum_{\ell=1}^k e_G(V(H_\ell),V)=\frac{1}{2}e_G(V,V)=|E(G)|. \] Then by the generalized mediant inequality, for some $H'=H_\ell$ we have \[ \frac{|E(H')|}{f(V(H'))}\geq \frac{\sum_{\ell=1}^k |E(H_\ell)|}{\sum_{\ell=1}^k f(V(H_\ell))} = \frac{|E(H)|}{|E(G)|}. \] Now, Corollary~\ref{cor:weightcs} applied with $S=V(H')$, $T=V$ yields $f(V(H'))^2\geq |E(G[V(H')])||E(G)|\geq |E(H')||E(G)|$, so that \[ \frac{|E(H)|}{|E(G)|}\leq \frac{|E(H')|}{f(V(H'))}\leq \sqrt{\frac{|E(H')|}{|E(G)|}}, \] which rearranges to the desired inequality. \end{proof} \section{Coloring problems} \label{sec:color} We can now apply Theorem~\ref{thm:main} to the corresponding coloring problems. We have already seen that applying Theorem~\ref{thm:main} with $G=K_n$ readily yields the simple lower bound on $M(n,k)$ given in Corollary~\ref{cor:weaklb}. We now discuss how to improve this lower bound, before turning to a closely related problem on monochromatic trails and circuits. \subsection{Lower bound on $M(n,k)$ for general $k$} Our strategy for improving the lower bound on $M(n,k)$ is as follows: First, assuming that no monochromatic component has too many edges, we show that in the color with the highest density (say, red), we can upper bound the number of vertices covered by any set of components (or else we can finish with an average degree argument on the rest of the red components). Through a smoothing argument, this yields an upper bound on the sum of squares of the number of vertices in each red component, and thus a lower bound on the number of edges in the complete multipartite graph $G$ formed by deleting from $K_n$ all edges within the vertex sets of the red components. We finish by applying Theorem~\ref{thm:main} to $G$ to find a non-red component with many edges. \begin{proof}[Proof of Theorem~\ref{thm:colorlb}] Fix a $k$-coloring $\chi$ of the edges of $K_n$, and suppose that the largest number of edges in a monochromatic connected component in this coloring is $z\binom{n}{2}$. Without loss of generality, let red be the color with the most edges, so there are at least $\frac{1}{k}\binom{n}{2}$ red edges in our coloring. Let $\mathcal{C}_1=\{R_1,R_2,\dots,R_m\}$ be the set of red components, with \begin{equation} \label{eqn:totalvtx} |V(R_1)|\geq |V(R_2)|\geq \cdots \geq |V(R_m)|, \qquad \sum_{i=1}^m |V(R_i)|=n. \end{equation} Let $x=\frac{1}{\binom{n}{2}}\sum_{i=1}^m |E(R_i)|$, so $x\geq \frac{1}{k}$. By assumption, $|E(R_i)|\leq z\binom{n}{2}$ for $1\leq i\leq m$. As in the proof of Corollary~\ref{cor:weaklb}, applying Theorem~\ref{thm:main} with $K_n$ as $G$ and its red color class (i.e. the spanning subgraph formed by its red edges) as $H$ yields a red component with at least $x^2 \binom{n}{2}$ edges, so by assumption we have $z\geq x^2\geq \frac{1}{k^2}$. Define $\delta=k-\frac{1}{\sqrt{z}}\geq 0$, so $z=\frac{1}{(k-\delta)^2}$. For any $j\in [1,m-1]$, let $G_j$ be the complete graph on $\bigcup_{i=j+1}^m V(R_i)$, with its coloring induced by $\chi$. Applying Theorem~\ref{thm:main} with $G_j$ as $G$ and the red color class of $G_j$ as $H$ yields a red connected component $H'=R_i$ with $|E(H')|\geq \frac{|E(H)|^2}{|E(G_j)|}$. Since \[ |E(H)|=\sum_{i=j+1}^m |E(R_i)|=x\binom{n}{2}-\sum_{i=1}^{j}|E(R_i)| \geq x\binom{n}{2}-jz\binom{n}{2}, \] while $|E(G_j)|=\binom{|V(G_j)|}{2}=\binom{n-\sum_{i=1}^{j}|V(R_i)|}{2}$, we have \[ z\binom{n}{2}\geq |E(H')|\geq \frac{\max(x-jz,0)^2 \binom{n}{2}^2}{\binom{n-\sum_{i=1}^{j}|V(R_i)|}{2}}. \] Since $\frac{\max(x-jz,0)}{\sqrt{z}}\leq \frac{x}{\sqrt{z}}\leq 1$, this implies \[ \binom{n-\sum_{i=1}^{j}|V(R_i)|}{2}\geq \frac{\max(x-jz,0)^2}{z}\binom{n}{2}\geq \binom{\frac{\max(x-jz,0)}{\sqrt{z}}n}{2}. \] Since $n-\sum_{i=1}^{j}|V(R_i)|\geq 1$, and the function $f(X)=\binom{X}{2}$ is increasing for $X\geq 1$, this then implies \begin{equation} \label{eqn:vtxbounds} \sum_{i=1}^{j}|V(R_i)|\leq (1-x/\sqrt{z} + j\sqrt{z})n \qquad \text{for all }j\in [1,m-1]. \end{equation} We can now give an upper bound on $\sum_{i=1}^m \binom{|V(R_i)|}{2}$ by solving the corresponding convex optimization problem. The proof of the following technical lemma will be deferred until the end of the section. \begin{lemma} \label{lem:smoothing} Let $x,z>0$. Subject to the constraints $v_1\geq \cdots \geq v_m \geq 0$, $\sum_{i=1}^m v_i= 1$, and \[ \sum_{i=1}^j v_i \leq 1-x/\sqrt{z}+j\sqrt{z} \qquad \text{for all }j\in [1,m-1], \] the quantity $\sum_{i=1}^m v_i^2$ is maximized when $v_1=1-x/\sqrt{z}+\sqrt{z}$, $v_i=\sqrt{z}$ for $2\leq i\leq \lfloor \frac{x}{z}\rfloor$, $v_{\lfloor \frac{x}{z}\rfloor+1}=x/\sqrt{z}-\lfloor \frac{x}{z}\rfloor \sqrt{z}$, and $v_i=0$ for $i>\lfloor \frac{x}{z}\rfloor+1$. \end{lemma} Let $v_i=\frac{|V(R_i)|}{n}$. Since \eqref{eqn:totalvtx} and \eqref{eqn:vtxbounds} hold, we can apply Lemma~\ref{lem:smoothing} to obtain \[ \sum_{i=1}^m v_i^2\leq (1-x/\sqrt{z}+\sqrt{z})^2 + \left(\left\lfloor \frac{x}{z}\right\rfloor-1\right) z +\left(\left(x/z-\left\lfloor \frac{x}{z}\right\rfloor\right) \sqrt{z}\right)^2 \leq (1-x/\sqrt{z}+\sqrt{z})^2 + (x/z-1)z, \] so that we have \begin{align*} \sum_{i=1}^m \binom{|V(R_i)|}{2} &= \frac{n^2 \sum_{i=1}^m v_i^2 -n }{2} \leq \frac{n^2 ((1-x/\sqrt{z}+\sqrt{z})^2 + (x/z-1)z) - n}{2}. \end{align*} Finally, let $G$ be the complete $m$-partite graph obtained from $K_n$ by removing all edges within each $V(R_i)$, with its coloring induced by $\chi$. There are no red edges in $G$, so by the pigeonhole principle, one of the $k-1$ remaining colors has at least $\frac{1}{k-1}|E(G)|$ edges in $G$. Let $H$ be the spanning subgraph of $G$ induced by the edges in that color. Applying Theorem~\ref{thm:main} then yields a monochromatic connected component $H'$ with at least $\frac{1}{(k-1)^2}|E(G)|$ edges. Then by assumption we have \begin{align*} z &\geq \frac{1}{(k-1)^2}\frac{|E(G)|}{\binom{n}{2}} = \frac{1}{(k-1)^2} \left(1- \frac{\sum_{i=1}^m \binom{|V(R_i)|}{2}}{\binom{n}{2}} \right) \\ &\geq \frac{1}{(k-1)^2} \left(1- \frac{n^2((1-x/\sqrt{z}+\sqrt{z})^2 + (x/z-1)z) - n}{n^2-n} \right)\\ &\geq \frac{1}{(k-1)^2} \left(1- (1-x/\sqrt{z}+\sqrt{z})^2 - (x/z-1)z \right), \end{align*} which rearranges to give \[ 1-(k-1)^2 z \leq (1-x/\sqrt{z}+\sqrt{z})^2 + (x/z-1)z=\frac{1}{z}x^2 - (1+2/\sqrt{z})x+(1+2\sqrt{z}). \] The right hand side is a quadratic in $x$ that is decreasing for $x\leq \frac{z}{2}+\sqrt{z}$. Since by assumption $\sqrt{z}\geq x\geq \frac{1}{k}$, we then have \[ 1-(k-1)^2 z \leq \frac{1}{z k^2} - (1+2/\sqrt{z})\frac{1}{k}+(1+2\sqrt{z}). \] Substituting in $z=\frac{1}{(k-\delta)^2}$ yields \[ 1-\frac{(k-1)^2}{(k-\delta)^2}\leq \frac{(k-\delta)^2}{k^2} - \frac{1+2(k-\delta)}{k}+1+\frac{2}{k-\delta}, \] which upon rearrangement becomes \begin{align*} 0 &\leq (k-1)^2 k^2 + (k-\delta)^4 - k(k-\delta)^2 - 2(k-\delta)^3 k + 2(k-\delta) k^2 \\ &= -k^3+k^2-k\delta^2+2k^3\delta -2k\delta^3+\delta^4\\ &=(k-\delta^2)^2 - k(k^2-\delta^2)(1-2\delta). \end{align*} This implies \[ 1-2\delta \leq \frac{(k-\delta^2)^2}{k(k^2-\delta^2)}<\frac{1}{k}, \] so that $\delta>\frac{k-1}{2k}$. Thus, the coloring $\chi$ contains a monochromatic connected component with at least $z\binom{n}{2}$ edges, where \[ z=\frac{1}{(k-\delta)^2}>\frac{1}{(k-\frac{1}{2}+\frac{1}{2k})^2}\geq \frac{1}{k^2-k+\frac{1}{4}+1-\frac{1}{4k}+\frac{1}{4k^2}} \geq \frac{1}{k^2-k+\frac{5}{4}}, \] as desired. \end{proof} \begin{proof}[Proof of Lemma~\ref{lem:smoothing}] Since the feasible region is compact, there exists a choice of the $v_i$ such that the desired maximum is attained. Fix such a maximizing choice of the $v_i$. For $j\in [m]$, let $A_j$ denote the given constraint on $\sum_{i=1}^j v_i$, and let $S$ be the set of $j$ for which equality holds in $A_j$. Let $m'=\lfloor \frac{x}{z}\rfloor$. Since $\sum_{i=1}^m v_i=1$, the constraints $A_j$ for $j>m'$ cannot be tight, so $S\subseteq [m']$. For any $i<j$ and any $\varepsilon\in (0,v_j]$, replacing $(v_i,v_j)$ with $(v_i+\varepsilon, v_j-\varepsilon)$ increases the value of $\sum_{i=1}^m v_i^2$, so by maximality, no such ``smoothing'' operation is possible. That is, we can assume the following \textbf{equality condition} for all $(i,j)$ with $i<j$: Either $v_i=v_{i-1}$, or $v_j=v_{j+1}$, or $S\cap [i,j-1]\neq \emptyset$. If $S=[m']$, then we are in exactly the maximizing case described (since by the equality conditions for $(m'+1,i)$ we have $v_i=0$ for all $i\geq m'+2$), so assume otherwise. First, suppose there is some $i_0\in [m']$ such that $v_{i_0}<\sqrt{z}$, and pick the smallest such index $i_0$. Let $i_1\in [m]$ be the largest index such that $v_{i_1}>0$, and note that $i_1>i_0$ since $\sum_{i=1}^{i_0} v_i<1$. Then we have $v_{i_0}<v_{i_0-1}$ and $v_{i_1}>v_{i_1+1}$. But as $\sum_{i=1}^j v_i\leq (1-x/\sqrt{z}+(j-1)\sqrt{z}) + v_j < (1-x/\sqrt{z}+(j-1)\sqrt{z})+ \sqrt{z}$ for any $j\geq i_0$, we have $S\cap [i_0,i_1]=\emptyset$, contradicting the equality condition for $(i_0,i_1)$. So, we can assume $v_i\geq \sqrt{z}$ for all $i\leq m'$. In particular, if $j\in S$ for some $j\in [m']$, then we recursively obtain $v_i=\sqrt{z}$ for all $i\in [j+1,m']$, so $[j,m']\subseteq S$. Thus, we can assume $1\notin S$, i.e. $v_1<1-x/\sqrt{z}+\sqrt{z}$. By the equality condition for $(1,2)$, we must then have $v_2=v_3$. Let $j_0\geq 3$ be the smallest index such that $v_{j_0+1}\neq v_{j_0}$. By the equality condition for $(1,j_0)$, there is some $j_1\in S\cap [2,j_0-1]$. Then \[ 1-x/\sqrt{z}+j_1 \sqrt{z} = \sum_{i=1}^{j_1} v_i = v_1+(j_1-1) v_2 < 1-x/\sqrt{z}+\sqrt{z}+(j_1-1)v_2, \] which implies $v_2>\sqrt{z}$. But then $\sum_{i=1}^{j_1+1} v_i=1-x/\sqrt{z}+\sqrt{z}+j_1 v_2 > 1-x/\sqrt{z}+(j_1+1) \sqrt{z}$, violating $A_{j_1+1}$. This is a contradiction, so in fact $S=[m']$, and we are in the desired maximizing case. \end{proof} We remark that in Gyárfás's construction, after removing all edges contained in the vertex set of each red component, each non-red component is left with approximately $\frac{1}{(k-1)^2} (1-\frac{1}{k-1}) \binom{n}{2} = \frac{k-2}{(k-1)^3} \binom{n}{2}$ edges. Since $\frac{k-2}{(k-1)^3}=\frac{1}{k^2-k+1+\frac{1}{k-2}}$, this suggests that the method used to prove Theorem~\ref{thm:colorlb} cannot show a lower bound on $M(n,k)$ better than $\frac{1}{k^2-k+1}\binom{n}{2}$ without introducing additional ideas. \subsection{Multicolor Ramsey numbers of trails and circuits} \label{subsec:trails} A \emph{trail} is a walk without repeated edges, and a \emph{circuit} is a trail with the same first and last vertex. The ($k$-color) Ramsey problem for trails is the question of finding the largest $m$ such that every $k$-coloring of the edges of $K_n$ contains a monochromatic trail of length $m$. Answering a question of Osumi \cite{osumi2021ramsey}, Conlon and Tyomkyn \cite{conlon2021ramsey} show that every $2$-coloring of the edges of $K_n$ contains a monochromatic circuit with at least $\frac{2}{9}n^2+O(n^{3/2})$ edges, and this is asymptotically tight. For the case of general $k$, they observe that by deleting a forest in each color class of a $k$-coloring of $K_n$ to make each color class Eulerian (i.e. ensuring every vertex has even degree in each color), one can reduce this Ramsey problem to a variant of the problem of determining $M(n,k)$. Where previously we colored the edges of $K_n$ and found a large monochromatic component, we now apply the same procedure to the graph obtained by deleting at most $kn$ edges from $K_n$. We now prove a lower bound for the general case of this problem, analogous to the bound on $M(n,k)$ given in Theorem~\ref{thm:colorlb}. \begin{corollary} \label{thm:trailsweak} Every $k$-coloring of the edges of $K_n$ contains a monochromatic circuit (and hence a monochromatic trail) of length at least $\frac{1}{k^2-k+\frac{5}{4}}\binom{n}{2}+O_k(n)$. \end{corollary} \begin{proof} Fix a $k$-coloring of $E(K_n)$. As in \cite{conlon2021ramsey}, we can remove from each color class a forest that meets all odd degree vertices, leaving a coloring where every color class, and hence every monochromatic connected component, is Eulerian. Let $\chi$ be the resulting partial $k$-coloring of $E(K_n)$, where the (at most $kn$) removed edges are left uncolored, and let $z\binom{n}{2}$ be the largest number of edges in a monochromatic component in this coloring. It suffices to show that $z\geq \frac{1}{k^2-k+\frac{5}{4}}+O_k(n^{-1})$, since every monochromatic component is Eulerian, and thus contains an Eulerian circuit. Fix a color (say, red) with $x\binom{n}{2}$ edges, where $x\geq \frac{1}{k} \frac{\binom{n}{2}-nk}{\binom{n}{2}}=\frac{1}{k}-\frac{2}{n-1}$. As before, applying Theorem~\ref{thm:main} with $G=K_n$ immediately yields $z\geq x^2\geq \frac{1}{k^2}-O(n^{-1})$. Note that this means there is a monochromatic circuit with at least $\left(\frac{1}{k^2}-O(n^{-1})\right)\binom{n}{2}=\frac{1}{k^2}\binom{n}{2}+O(n)$ edges. To improve this lower bound further, we proceed as in the proof of Theorem~\ref{thm:colorlb}. Letting $R_1,\dots,R_m$ be the red components, such that \eqref{eqn:totalvtx} holds, we obtain \eqref{eqn:vtxbounds} in the same manner as before, so we can once again apply Lemma~\ref{lem:smoothing} to obtain the same upper bound on $\sum_{i=1}^m \binom{|V(R_i)|}{2}$ in terms of $x$ and $z$. Let $G$ be the complete $m$-partite graph with parts $V(R_1),\dots,V(R_m)$, with its partial coloring induced by $\chi$. Since $G$ has no red edges, and at most $kn$ edges are uncolored, one of its color classes $H$ has at least $\frac{1}{k-1}(|E(G)|-kn)=\left(\frac{1}{k-1}-O_k(n^{-1})\right)|E(G)|$ edges. Applying Theorem~\ref{thm:main} as before yields \begin{align*} z &\geq \left(\frac{1}{k-1}-O_k(n^{-1})\right)^2\frac{|E(G)|}{\binom{n}{2}} \geq \left(\frac{1}{(k-1)^2}-O_k(n^{-1})\right) \left(1- \frac{\sum_{i=1}^m \binom{|V(R_i)|}{2}}{\binom{n}{2}} \right) \\ &\geq \left(\frac{1}{(k-1)^2+O_k(n^{-1})}\right) \left(1- (1-x/\sqrt{z}+\sqrt{z})^2 - (x/z-1)z \right). \end{align*} Letting $z=\frac{1}{(k-\delta)^2}$ and noting that $x\geq \frac{1}{k}-O(n^{-1})$, we can perform the same rearrangements and substitutions as in the proof of Theorem~\ref{thm:colorlb}, simply separating out the $O_k(n^{-1})$ terms at each step, to derive the inequality $1-2\delta < \frac{1}{k} + O_k(n^{-1})$, and thus \[ z\geq \frac{1}{k^2-k+\frac{5}{4}-O_k(n^{-1})}= \frac{1}{k^2-k+\frac{5}{4}}+O_k(n^{-1}). \] Then the coloring $\chi$ contains a monochromatic connected component, and hence a monochromatic Eulerian circuit, with at least $z\binom{n}{2}\geq \frac{1}{k^2-k+\frac{5}{4}}\binom{n}{2}+O_k(n)$ edges, as claimed. \end{proof} \subsection{Three colors} In this section, we prove the lower and upper bounds on $M(n,3)$ in separate lemmas in order to establish Theorem~\ref{thm:color3}. We then discuss the behavior of $M(n,3)$ for small values of $n$, and conclude by describing how to adapt our proofs to obtain asymptotically tight bounds for the Ramsey numbers of trails and circuits in three colors. \begin{lemma} \label{lem:3lb} Every $3$-coloring of the edges of $K_n$ contains a monochromatic component with at least $\lceil \frac{1}{6}\binom{n}{2}\rceil$ edges. \end{lemma} \begin{proof} Let $G=K_n$, and call the three colors red, green, and blue. First, suppose one of the color classes (say, red) is connected. If there are at least $\frac{1}{6}|E(G)|$ red edges, we are done. Otherwise, the other two colors together have at least $\frac{5}{6}|E(G)|$ edges, so one of them has at least $ \frac{5}{12}|E(G)|$ edges. Applying Theorem~\ref{thm:main} with $G=K_n$ to the graph in that color then gives a monochromatic component with at least $(\frac{5}{12})^2 |E(G)| >\frac{1}{6}|E(G)|$ edges, so we are again done. Thus, we can assume every color has at least two components. Without loss of generality, let red be the color with the most edges, so the red graph $H$ has at least $\frac{1}{3}|E(G)|$ edges. Let $H'$ be the component of $H$ with the highest average degree, so $|V(H')|\geq \bar{d}(H')+1\geq \frac{1}{3}n$. We can assume $|V(H')|\leq \frac{1}{2}n$; otherwise, we would have $|E(H')|\geq \frac{1}{2}|V(H')|\bar{d}(H') > \frac{1}{6}|E(G)|$. Let $V_1=V(H')$ and $V_2=V(G)\setminus V_1$, and let $G'$ be the bipartite graph induced by $G$ between $V_1$ and $V_2$, so $G'$ has at least $|V_1||V_2|\geq \frac{2}{9}n^2>\frac{1}{3}|E(G)|$ edges, all of which must be green or blue. Fix an edge in $G'$ and consider the monochromatic component $C_1$ of $G'$ containing this edge. Without loss of generality, assume $C_1$ is green. Suppose $C_1$ covers all of $V_1$. Since every green edge in $G'$ intersects $V_1$, this means there is exactly one green component of $G'$ with a nonzero number of edges. Since there are at least two green components in $G$, there is a vertex $v\in V_2$ not in $C_1$. Then all edges between $v$ and $V_1$ must be blue, so all vertices of $V_1$ are in the same blue component in $G'$. This implies that there is also exactly one blue component of $G'$ with nonzeroly many edges. Thus, all edges of $G'$ are in one of two monochromatic components, one of which then has at least $\frac{1}{2}|E(G')|>\frac{1}{6}|E(G)|$ edges, as desired. \begin{figure}[h] \centering \includegraphics[scale=0.8]{3colorwithkey2.png} \caption{Lower bound: Exactly two components of each color} \label{fig:3color} \end{figure} We are likewise done if $C_1$ covers all of $V_2$, so we can assume $V_1\setminus C_1$ and $V_2\setminus C_1$ are both nonempty. Let $A_1=V_1\cap C_1$, $B_1=V_2\cap C_1$, $A_2=V_1\setminus A_1$, $B_2=V_2\setminus B_1$. Then all edges between $A_1$ and $B_2$, or between $A_2$ and $B_1$, can only be blue. Then there are at most two blue components, and thus by assumption exactly two. This in turn implies that all edges between $A_1$ and $B_1$, or between $A_2$ and $B_2$, can only be green, so there are exactly two green components $C_1$ and $C_2$. Finally, all edges between $B_1$ and $B_2$ can only be red, so we conclude that there are exactly two red components, $V_1$ and $V_2$; see Figure~\ref{fig:3color}. Since each of the three colors has exactly two components, one of the components has at least $\frac{1}{6}|E(G)|$ edges, as desired. \end{proof} \begin{lemma} \label{lem:3ub} For sufficiently large $n$, there exists a $3$-coloring of the edges of $K_n$ such that every monochromatic component contains at most $\lceil \frac{1}{6} \binom{n}{2}\rceil$ edges. \end{lemma} \begin{proof} \begin{figure}[h] \centering \includegraphics[scale=0.8]{3ubnokey.png} \caption{Initial construction for the $M(n,3)$ upper bound} \label{fig:3ub} \end{figure} We consider the following modification of a construction by Gyárfás: Let $V=V_1\cup V_2\cup V_3\cup V_4$ be a partition of the vertices of $G=K_n$ into four parts, with \[ \left\lceil \frac{n}{4}\right \rceil = |V_1| \geq \cdots \geq |V_4| = \left\lfloor \frac{n}{4}\right \rfloor. \] Letting $E(U,W)$ denote the set of edges between vertex sets $U$ and $W$, color $E(V_1,V_2)\cup E(V_3,V_4)$ red, $E(V_1,V_3)\cup E(V_2,V_4)$ green, and $E(V_1,V_4)\cup E(V_2,V_3)$ blue. For large enough $n$, $e(V_i,V_j)\leq \lceil \frac{1}{6}\binom{n}{2}\rceil$ for $1\leq i<j\leq 4$, so it remains to extend this partial coloring by coloring the edges within each of the $V_i$ such that each monochromatic component has at most $\lceil \frac{1}{6}\binom{n}{2}\rceil$ edges at the end. It is natural to attempt the simple approach of distributing the edges within each $V_i$ as evenly as possible among the three colors. However, the possibility of a slight difference in size among the $V_i$ can yield a $\Theta(n)$ difference among the numbers of edges in each component if we are not sufficiently careful; a different coloring strategy will turn out to be simpler to analyze. We call an extension of the above partial coloring \emph{nice} if each green or blue component contains exactly $\lceil\frac{1}{6}\binom{n}{2}\rceil$ edges. At least one nice coloring exists: we can color exactly $\lceil \frac{1}{6}\binom{n}{2}\rceil - e(V_1,V_3)$ of the edges within $V_1$ and $\lceil \frac{1}{6}\binom{n}{2}\rceil - e(V_2,V_4)$ of the edges within $V_2$ green, and exactly $\lceil \frac{1}{6}\binom{n}{2}\rceil - e(V_2,V_3)$ of the edges within $V_3$ and $\lceil \frac{1}{6}\binom{n}{2}\rceil - e(V_1,V_4)$ of the edges within $V_4$ blue (there are enough edges within each $V_i$ to do this when $n$ is sufficiently large), and color all remaining edges red. See Figure~\ref{fig:3ub} for a diagram of this coloring. Fix a nice coloring where the larger of the two red components contains as few edges as possible. Suppose one of the red components in this coloring, without loss of generality the one on $V_1\cup V_2$, has more than $\lceil\frac{1}{6}\binom{n}{2}\rceil$ edges. Then $V_3\cup V_4$ must have less than $\lceil\frac{1}{6}\binom{n}{2}\rceil$ red edges. Without loss of generality let $V_1$ contain a red edge $e$. If either $V_3$ contains a green edge, or $V_4$ contains a blue edge, we can switch the color of that edge with edge $e$, preserving the sizes of the green and blue components while decreasing the size of the larger red component by one. Otherwise, $V_3$ is entirely blue and red, and $V_4$ is entirely green and red. Since the red component on $V_3\cup V_4$ has less than $\lceil\frac{1}{6}\binom{n}{2}\rceil$ edges, neither $V_3$ nor $V_4$ can be entirely red (for sufficiently large $n$, $\lfloor \frac{n}{4}\rfloor^2 + \binom{\lfloor \frac{n}{4}\rfloor}{2}>\lceil \frac{1}{6}\binom{n}{2}\rceil$). Then if $V_2$ contains a red edge, we can similarly switch two edges to reduce the size of the larger red component by one, so we can assume $V_2$ is entirely green and blue. But then there is one component of each color (including the larger red component) that does not have any edges within $V_2\cup V_3\cup V_4$, which means there are at least $3\lceil\frac{1}{6}\binom{n}{2}\rceil$ edges incident to $V_1$, a contradiction since \[ e(V_1,V)\leq \binom{\lceil \frac{n}{4}\rceil}{2}+\left\lceil \frac{n}{4}\right\rceil\left(n-\left\lceil \frac{n}{4}\right\rceil\right) < 3\left\lceil\frac{1}{6}\binom{n}{2}\right\rceil, \] for sufficiently large $n$. Thus indeed there is a construction of this form where every monochromatic component has at most $\lceil\frac{1}{6} \binom{n}{2}\rceil$ edges. \end{proof} Combining Lemmas~\ref{lem:3lb} and~\ref{lem:3ub} yields Theorem~\ref{thm:color3} as desired. The construction in the proof of Lemma~\ref{lem:3ub} is well-defined for all $n\geq 46$. A more careful analysis, splitting into cases based on the value of $n$ modulo $4$ and then using a more explicit construction in each case, shows that in fact $M(n,3)=\left\lceil\frac{1}{6}\binom{n}{2}\right\rceil$ for all $n\geq 11$ except $n=13,17$. However, the lower bound from Lemma~\ref{lem:3lb} is not sharp for some of these small values of $n$. Closely inspecting each step of our proof for $n=17$, for example, we can deduce that $M(17,3)=24$, instead of the expected $23$; indeed, all of the bounds before the final step in the proof of Lemma~\ref{lem:3lb} are loose enough to yield a component with at least $24$ edges unless we are in the case depicted in Figure~\ref{fig:3color}, where one of the vertex sets $A_i$ or $B_i$ has size $5$, and the other three sets have size $4$. The set of size $5$ then contains $10$ internal edges, some $4$ of which are in the same color, yielding a component with $4+20=24$ edges as claimed. This shows a genuine difference in the behavior of $M(n,3)$ for these small values of $n$ due to integer-related constraints in the extremal configurations. We can adjust the proof of Lemma~\ref{lem:3lb} to give a lower bound on the size of the largest monochromatic circuit in a $3$-coloring of the edges of $K_n$, as follows. First, as before, we can remove a forest in each color and leave each color class Eulerian. The resulting graph $G$ has at least $\binom{n}{2}-3n$ edges, so all but at most $3\sqrt{n}$ of the vertices have degree at least $n-1-2\sqrt{n}$. We then pass to the induced subgraph $G'$ on these $n'\geq n-3\sqrt{n}$ vertices; the minimum degree of $G'$ is at least $n'-1-2\sqrt{n}\geq n'-3\sqrt{n'}$ when $n$ is sufficiently large. The argument then proceeds largely as in the proof of Lemma~\ref{lem:3lb}, except that the condition of every green or blue component intersecting both $V_1$ and $V_2$ is strengthened by requiring every green or blue component to intersect each of $V_1$ and $V_2$ in more than $6\sqrt{n'}$ vertices. The proof then concludes as before, reducing to the case where there are exactly two components in each color. When the edges between $V(G')$ and $V(G)\setminus V(G')$ are added back in, it remains true that there are at most two components in each color with a positive number of edges. So, at least one monochromatic component in $G$ has at least $\frac{1}{6}(|E(G)|-(3\sqrt{n})^2) \geq \frac{1}{12}n^2 - O(n)$ edges. Since this component of $G$ is Eulerian, we have a circuit, and hence a trail, of the desired length, for all sufficiently large $n$. The upper bound from Lemma~\ref{lem:3ub} likewise applies to the size of the longest circuit, showing that the lower bound is asymptotically tight. \vspace{3mm} \noindent {\bf Acknowledgments.} I would like to thank my advisor Jacob Fox for introducing me to this problem and for helpful conversations along the way, as well as David Conlon and Mykhaylo Tyomkyn for the insights from our later joint work that served as inspiration for some of the strengthened arguments in the revised version of this paper. In addition, I would like to thank the anonymous referees for their careful reading and helpful comments, including a specific suggestion that led to a significant strengthening in the bound for general $k$ in Theorem~\ref{thm:colorlb}. \bibliographystyle{acm}
{ "timestamp": "2022-08-30T02:11:27", "yymm": "2111", "arxiv_id": "2111.13342", "language": "en", "url": "https://arxiv.org/abs/2111.13342", "abstract": "We prove that if $H$ is a subgraph of a complete multipartite graph $G$, then $H$ contains a connected component $H'$ satisfying $|E(H')||E(G)|\\geq |E(H)|^2$. We use this to prove that every three-coloring of the edges of a complete graph contains a monochromatic connected subgraph with at least $1/6$ of the edges. We further show that such a coloring has a monochromatic circuit with a fraction $1/6-o(1)$ of the edges. This verifies a conjecture of Conlon and Tyomkyn. Moreover, for general $k$, we show that every $k$-coloring of the edges of $K_n$ contains a monochromatic connected subgraph with at least $\\frac{1}{k^2-k+\\frac{5}{4}}\\binom{n}{2}$ edges.", "subjects": "Combinatorics (math.CO)", "title": "On connected components with many edges", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9875683476832692, "lm_q2_score": 0.810478913248044, "lm_q1q2_score": 0.8004033211885024 }
https://arxiv.org/abs/1010.5005
Error Estimates for Generalized Barycentric Interpolation
We prove the optimal convergence estimate for first order interpolants used in finite element methods based on three major approaches for generalizing barycentric interpolation functions to convex planar polygonal domains. The Wachspress approach explicitly constructs rational functions, the Sibson approach uses Voronoi diagrams on the vertices of the polygon to define the functions, and the Harmonic approach defines the functions as the solution of a PDE. We show that given certain conditions on the geometry of the polygon, each of these constructions can obtain the optimal convergence estimate. In particular, we show that the well-known maximum interior angle condition required for interpolants over triangles is still required for Wachspress functions but not for Sibson functions.
\section{Introduction} While a rich theory of finite element error estimation exists for meshes made of triangular or quadrilateral elements, relatively little attention has been paid to meshes constructed from arbitrary polygonal elements. Many quality-controlled domain meshing schemes could be simplified if polygonal elements were permitted for dealing with problematic areas of a mesh and finite element methods have been applied to such meshes~\cite{TS06,WBG07}. Moreover, the theory of Discrete Exterior Calculus has identified the need for and potential usefulness of finite element methods using interpolation methods over polygonal domain meshes (e.g. Voronoi meshes associated to a Delaunay domain mesh)~\cite{GB2010}. Therefore, we seek to develop error estimates for functions interpolated from data at the vertices of a polygon $\Omega$. Techniques for interpolation over polygons focus on generalizing barycentric coordinates to arbitrary $n$-gons; this keeps the degrees of freedom associated to the vertices of the polygon which is exploited in nodal finite element methods. The seminal work of Wachspress~\cite{W1975} explored this exact idea and has since spawned a field of research on rational finite element bases over polygons. Many alternatives to these `Wachspress coordinates' have been defined as well, including the Harmonic and Sibson interpolants. To our knowledge, however, no careful analysis has been made as to which, if any, of these interpolation functions provide the correct error estimates required for finite element schemes. We consider first-order interpolation operators from some generalization of barycentric coordinates to arbitrary convex polygons. A set of barycentric coordinates $\{\lambda_i\}$ for $\Omega$ associated with the interpolation operator $I: H^2(\Omega) \rightarrow \text{span} \{\lambda_i\} \subset H^1(\Omega)$ is given by \begin{equation} \label{eq:genintop} Iu:=\sum_iu(\textbf{v}_i)\lambda_i. \end{equation} Since barycentric coordinates are unique on triangles (described in Section~\ref{ssec:triangulation}) this is merely the standard linear Lagrange interpolation operator when $\Omega$ is a triangle. Before stating any error estimates, we fix some notation. For multi-index ${\bf \alpha} = (\alpha_1, \alpha_2)$ and point $\textbf{x} = (x,y)$, define $\textbf{x}^{\bf \alpha} := x^{\alpha_1} y^{\alpha_2}$, $\alpha ! := \alpha_1 \alpha_2$, $|{\bf \alpha}| := \alpha_1 + \alpha_2$, and $D^{\bf \alpha} u := \partial^{|{\bf \alpha}|} u/\partial x^{\alpha_1}\partial y^{\alpha_2}$. The Sobolev semi-norms and norms over an open set $\Omega$ are defined by \begin{align*} \hpsn{u}{m}{\Omega}^2 &:= \int_\Omega \sum_{|\alpha| = m} |D^\alpha u(\textbf{x})|^2 \,{\rm d} \textbf{x} &{\rm and} & & \hpn{u}{m}{\Omega}^2 &:= \sum_{0\leq k\leq m}\hpsn{u}{m}{\Omega}^2. \end{align*} The $H^{0}$-norm is the $L^2$-norm and will be denoted $\lpn{\cdot}{2}{\Omega}$. Analysis of the finite element method often yields bounds on the solution error in terms of the best possible approximation in the finite-dimensional solution space. Thus the challenge of bounding the solution error is reduced to a problem of finding a good interpolant. In many cases Lagrange interpolation can provide a suitable estimate which is asymptotically optimal. For first-order interpolants that we consider, this \textbf{optimal convergence estimate} has the form \begin{equation} \label{eq:hconv} \hpn{u - I u }{1}{\Omega} \leq C\,\diam(\Omega)\hpsn{u}{2}{\Omega},\quad\forall u\in H^2(\Omega). \end{equation} To prove estimate (\ref{eq:hconv}) in our setting, it is sufficient (see Section~\ref{sec:intsobsp}) to restrict the analysis to a class of domains with diameter one and show that $I$ is a bounded operator from $H^2(\Omega)$ into $H^1(\Omega)$, that is \begin{equation} \label{eq:iuh1uh2} \hpn{Iu}{1}{\Omega} \leq C_I \hpn{u}{2}{\Omega},\quad \forall u\in H^1(\Omega). \end{equation} We call equation (\ref{eq:iuh1uh2}) the \textbf{$H^1$-interpolant estimate} associated to the barycentric coordinates $\lambda_i$ used to define $I$. The optimal convergence estimate (\ref{eq:hconv}) does not hold uniformly over all possible domains; a suitable geometric restriction must be selected to produce a uniform bound. Even in the simplest case (Lagrange interpolation on triangles), there is a gap between geometric criteria which are simple to analyze (e.g. the minimum angle condition) and those that encompass the largest possible set of domains (e.g. the maximum angle condition). This paper is devoted to finding geometric criteria under which the optimal convergence estimate (\ref{eq:hconv}) holds for several types of generalized barycentric coordinates on arbitrary convex polygons. We begin by establishing some notation (shown in Figure \ref{fig:notation}) to describe the specific geometric criteria. \begin{figure}[ht] \begin{center} \psfrag{vi}{$\textbf{v}_i$} \psfrag{Bi}{$\beta_i$} \psfrag{c}{\textcolor{red}{$\textbf{c}$}} \psfrag{diam}{\textcolor{blue}{$\diam(\Omega)$}} \psfrag{Om}{$\Omega$} \psfrag{r}{\textcolor{red}{$\rho(\Omega)$}} \includegraphics[width=.3\linewidth]{img/notation.eps} \end{center} \caption{Notation used throughout paper. } \label{fig:notation} \end{figure} Let $\Omega$ be a convex polygon with $n$ vertices. Denote the vertices of $\Omega$ by $\textbf{v}_i$ and the interior angle at $\textbf{v}_i$ by $\beta_i$. The largest distance between two points in $\Omega$ (the diameter of $\Omega$) is denoted $\diam(\Omega)$ and the radius of the largest inscribed circle is denoted $\rho(\Omega)$. The center of this circle is denoted $\textbf{c}$ and is selected arbitrarily when no unique circle exists. The \textbf{aspect ratio} (or chunkiness parameter) $\gamma$ is the ratio of the diameter to the radius of the largest inscribed circle, i.e. \[\gamma := \frac{\diam(\Omega)}{\rho(\Omega)}.\] We will consider domains satisfying one or more of the following geometric conditions. \renewcommand{\labelenumi}{G\arabic{enumi}.} \begin{enumerate} \item \textbf{Bounded aspect ratio:} There exists $\gamma^*\in\R$ such that $\gamma < \gamma^*$. \label{g:ratio} \item \textbf{Minimum edge length: } There exists $d_*\in\R$ such that $|\textbf{v}_i - \textbf{v}_j| > d_* > 0$ for all $i\neq j$. \label{g:minedge} \item \textbf{Maximum interior angle:} There exists $\beta^*\in\R$ such that $\beta_i < \beta^* < \pi$ for all $i$.\label{g:maxangle} \end{enumerate} Using several definitions of generalized barycentric functions from the literature, we show which geometric constraints on $\Omega$ are either necessary or sufficient to ensure the estimate for each definition. The main results of this paper are summarized by the following theorem and Table \ref{tab:conditions}. Primary attention is called to the difference between Wachspress and Sibson coordinates: while G\ref{g:maxangle} is a necessary requirement for Wachspress coordinates, it is demonstrated to be unnecessary for the Sibson coordinates. \begin{theorem} In Table \ref{tab:conditions}, any necessary geometric criteria to achieve the $H^1$ interpolant estimate (\ref{eq:iuh1uh2}) are denoted by N. The set of geometric criteria denoted by S in each row are sufficient to guarantee the $H^1$ interpolant estimate (\ref{eq:iuh1uh2}). \begin{table}[ht] \centering \sbox{\strutbox}{\rule{0pt}{0pt}} \begin{tabular}[.75\textwidth]{@{\extracolsep{\fill}} ccccc} && \begin{tabular}{c} G\ref{g:ratio} \\ (aspect \\ ratio)\end{tabular} & \begin{tabular}{c} G\ref{g:minedge} \\ (min edge \\ length)\end{tabular} & \begin{tabular}{c} G\ref{g:maxangle} \\ (max interior \\ angle)\end{tabular} \\[2mm] \hline\hline\\[2mm] Triangulated & $\displaystyle\lambda^{\rm Tri}$ & - & - & S,N \\[2mm] \hline\\[2mm] Harmonic & $\displaystyle\lambda^{\rm Har}$ & S & - & - \\[2mm] \hline\\[2mm] Wachspress & $\displaystyle\lambda^{\rm Wach}$ & S & S & S,N \\[2mm] \hline\\[2mm] Sibson & $\displaystyle\lambda^{\rm Sibs}$ & S & S & - \\[2mm] \hline \end{tabular} \caption{`N' indicates a necessary geometric criterion for achieving the $H^1$ interpolant estimate (\ref{eq:iuh1uh2}). The set of criteria denoted `S' in each row, taken together, are sufficient to ensure the $H^1$ interpolant estimate (\ref{eq:iuh1uh2}).} \label{tab:conditions} \end{table} \end{theorem} In Section \ref{sec:genbary}, we define the various types of generalized barycentric coordinates, compare their properties, and mention prior applications. In Section \ref{sec:geomcond}, we review some general geometric results needed for subsequent proofs. In Section \ref{sec:intsobsp}, we give the relevant background on interpolation theory for Sobolev spaces and state some classical results used to motivate our approach. In Section \ref{sec:trimeth}, we show that the simplest technique of triangulating the polygon achieves the estimate if and only if G\ref{g:maxangle} holds. In Section \ref{sec:optimal}, we show that if harmonic coordinates are used, G\ref{g:ratio} alone is sufficient. In Section \ref{sec:wachpress}, we show that Wachspress coordinates require G\ref{g:maxangle} to achieve the estimate but all three criteria are sufficient. In Section \ref{sec:sibson}, we show that Sibson coordinates achieve the estimate with G\ref{g:ratio} and G\ref{g:minedge} alone. We discuss the implications of these results and future directions in Section \ref{sec:conc}. \section{Generalized Barycentric Coordinate Types} \label{sec:genbary} Barycentric coordinates on general polygons are any set of functions satisfying certain key properties of the regular barycentric functions for triangles. \begin{definition}\label{def:barcoor} Functions $\lambda_i:\Omega\rightarrow\R$, $i=1,\ldots, n$ are \textbf{barycentric coordinates} on $\Omega$ if they satisfy two properties. \renewcommand{\labelenumi}{B\arabic{enumi}.} \begin{enumerate} \item \textbf{Non-negative}: $\lambda_i\geq 0$ on $\Omega$.\label{b:nonneg} \item \textbf{Linear Completeness}: For any linear function $L:\Omega\rightarrow\R$, $\displaystyle L=\sum_{i=1}^{n} L(\textbf{v}_i)\lambda_i$.\label{b:lincomp} \end{enumerate} \end{definition} \begin{remark} Property B\ref{b:lincomp} is the key requirement needed for our interpolation estimates. It ensures that the interpolation operation preserves linear functions, i.e. $IL = L$. \end{remark} We will restrict our attention to barycentric coordinates satisfying the following invariance property. Let $T:\R^2 \rightarrow \R^2$ be a composition of rotation, translation, and uniform scaling transformations and let $\{\lambda^T_i\}$ denote a set of barycentric coordinates on $T\Omega$. \renewcommand{\labelenumi}{B\arabic{enumi}.} \begin{enumerate} \setcounter{enumi}{2} \item \textbf{Invariance:} $\displaystyle\lambda_i(\textbf{x})=\lambda_i^T(T(\textbf{x}))$.\label{b:invariance} \end{enumerate} This assumption will allow estimates over the class of convex sets with diameter one to be immediately extended to generic sizes since translation, rotation and uniform scaling operations can be easily passed through Sobolev norms (see Section~\ref{sec:intsobsp}). At the expense of requiring uniform bounds over a class of diameter-one domains rather than a single reference element, complications associated with handling non-affine mappings between reference and physical elements are avoided \cite{ABF02}. A set of barycentric coordinates $\{\lambda_i\}$ also satisfies these additional familiar properties: \renewcommand{\labelenumi}{B\arabic{enumi}.} \begin{enumerate} \setcounter{enumi}{3} \item \textbf{Partition of unity:} $\displaystyle\sum_{i=1}^{n}\lambda_i\equiv 1$. \label{b:partition} \item \textbf{Linear precision:} $\displaystyle\sum_{i=1}^{n}\textbf{v}_i\lambda_i(\textbf{x})=\textbf{x}$. \label{b:linprec} \item \textbf{Interpolation:} $\displaystyle\lambda_i(\textbf{v}_j) = \delta_{ij}$. \label{b:interpolation} \end{enumerate} The precise relationship between these properties and those defining the barycentric coordinates is given in the following proposition. \begin{proposition} The properties B\ref{b:nonneg}-B\ref{b:interpolation} are related as follows: \label{prop:Brelation} \vspace{-.07in} \begin{enumerate}[(i)] \item B\ref{b:lincomp} $\Leftrightarrow$ (B\ref{b:partition} and B\ref{b:linprec}) \item (B\ref{b:nonneg} and B\ref{b:lincomp}) $\Rightarrow$ B\ref{b:interpolation} \end{enumerate} \end{proposition} \begin{proof} Given B\ref{b:lincomp}, setting $L\equiv 1$ implies B\ref{b:partition} and setting $L(\textbf{x})=\textbf{x}$ yields B\ref{b:linprec}. Conversely, assuming B\ref{b:partition} and B\ref{b:linprec}, let $L(x,y)=ax+by+c$ where $a,b,c\in\R$ are constants. Let $\textbf{v}_i$ have coordinates $(\textbf{v}_i^x,\textbf{v}_i^y)$. Then \begin{align*} \sum_{i=1}^{n}L(\textbf{v}_i)\lambda_i(x,y) & =\sum_{i=1}^{n} (a\textbf{v}_i^x+b\textbf{v}_i^y+c)\lambda_i(\textbf{x})\\ & =a\left(\sum_{i=1}^{n}\textbf{v}_i^x\lambda_i(\textbf{x})\right)+b\left(\sum_{i=1}^{n}\textbf{v}_i^y\lambda_i(\textbf{x})\right)+c\left(\sum_{i=1}^{n}\lambda_i(\textbf{x})\right)\\ & = ax+by+c = L(x,y). \end{align*} A proof that B\ref{b:nonneg} and B\ref{b:lincomp} imply B\ref{b:interpolation} can be found in \cite[Corollary 2.2]{FHK2006}. \qed \end{proof} Thus, while other definitions of barycentric coordinates appear in the literature, requiring only properties B\ref{b:nonneg} and B\ref{b:lincomp} is a minimal definition still achieving all the desired properties. In the following subsections, we define common barycentric coordinate functions from the literature. Additional comparisons of barycentric functions can be found in the survey papers of Cueto et al. \cite{CSCMCD2003} and Sukumar and Tabarraei \cite{ST2004}. \subsection{Triangulation Coordinates} \label{ssec:triangulation} The simplest method for constructing barycentric coordinates on a polygon is to triangulate the polygon and use the standard barycentric coordinate functions of these triangles. Interpolation properties of this scheme are well known from the standard analysis of the finite element method over triangular meshes, but this construction serves as an important point of comparison with the alternative barycentric coordinates discussed later. Let ${\mathcal{T}}$ be a triangulation of $\Omega$ formed by adding edges between the $\textbf{v}_j$ in some fashion. Define \[\lambda^{\rm Tri}_{i,{\mathcal{T}}}:\Omega\rightarrow \R\] to be the barycentric function associated to $\textbf{v}_i$ on triangles in ${\mathcal{T}}$ containing $\textbf{v}_i$ and identically 0 otherwise. Trivially, these functions define a set of barycentric coordinates on $\Omega$. Two particular triangulations are of interest. For a fixed $i$, let ${\mathcal{T}}_m$ denote any triangulation with an edge between $\textbf{v}_{i-1}$ and $\textbf{v}_{i+1}$. Let ${\mathcal{T}}_M$ denote the triangulation formed by connecting $\textbf{v}_i$ to all the other $\textbf{v}_j$. Examples are shown in Figure \ref{fig:optimal}. \begin{figure}[ht] \begin{center} \psfrag{vi}{$\textbf{v}_i$} \psfrag{tmin}{${\mathcal{T}}_m$} \psfrag{tmax}{${\mathcal{T}}_M$} \[\begin{array}{ccc} \includegraphics[width=.3\linewidth]{img/triang-min.eps} & \quad\text{ }\quad & \includegraphics[width=.3\linewidth]{img/triang-max.eps} \end{array}\] \end{center} \caption{Triangulations ${\mathcal{T}}_m$ and ${\mathcal{T}}_M$ are used to produce the minimum and maximum barycentric functions associated with $\textbf{v}_i$, respectively.} \label{fig:optimal} \end{figure} \begin{proposition}(Floater et al.~\cite{FHK2006}) \label{prop:fhk} Any barycentric coordinate function $\lambda_i$ according to Definition \ref{def:barcoor} satisfies the bounds \begin{equation} \label{eq:fhk} 0\leq \lambda^{\rm Tri}_{i,{\mathcal{T}}_m}(\textbf{x})\leq \lambda_i(\textbf{x}) \leq \lambda^{\rm Tri}_{i,{\mathcal{T}}_M}(\textbf{x})\leq 1,\quad\forall\textbf{x}\in\Omega. \end{equation} \end{proposition} Proposition \ref{prop:fhk} tells us that the triangulation coordinates are, in some sense, the extremal definitions of generalized barycentric coordinates. In any triangulation of $\Omega$, at least one triangle will be of the form $(\textbf{v}_{i-1},\textbf{v}_i,\textbf{v}_{i+1})$, and hence the lower bound in (\ref{eq:fhk}) is always realized by some $\lambda^{\rm Tri}_i$. Thus, the examination of alternative barycentric coordinates can be motivated as an attempt to find non-extremal generalized barycentric coordinates. \subsection{Harmonic Coordinates} A particularly well-behaved set of barycentric coordinates, harmonic coordinates, can be defined as the solutions to certain boundary value problems. Let $g_i:\partial\Omega\rightarrow\R$ be the piecewise linear function satisfying \[g_i(\textbf{v}_j)=\delta_{ij},\quad g_i \text{ linear on each edge of $\Omega$}.\] The harmonic coordinate function $\lambda^{\rm Har}_i$ is defined to be the solution of Laplace's equations with $g_i$ as boundary data, \begin{equation} \label{eq:optpde} \displaystyle\left\{\begin{array}{rcll} \Delta\left(\lambda^{\rm Har}_i\right) & = & 0, & \text{on $\Omega$}, \\ \lambda^{\rm Har}_i & = & g_i. & \text{on $\partial\Omega$}. \end{array}\right. \end{equation} Existence and uniqueness of the solution are well known results~\cite{Ev98,RR04}. Properties B\ref{b:nonneg} and B\ref{b:lincomp} are a consequence of the maximum principle and linearity of Laplace's equation. These coordinates are optimal in the sense that they minimize the norm of the gradient over all functions satisfying the boundary conditions, \[ \lambda^{\rm Har}_i = \text{argmin} \left\{\hpsn{\lambda}{1}{\Omega} \, : \, \lambda = g_i \,\text{on $\partial\Omega$} \right\}. \] This natural construction extends nicely to polytopes, as well as to a similar definition for barycentric-like (Whitney) vector elements on polygons. Christensen~\cite{C2008} has explored theoretical results along these lines. Numerical approximations of the $\lambda^{\rm Har}_i$ functions have been used to solve Maxwell's equations over polyhedral grids~\cite{E2007} and for finite element simulations for computer graphics~\cite{MKBWG2008,JMRGS07}. While it may seem excessive to solve a PDE just to derive the basis functions for a larger PDE solver, the relatively limited geometric requirements required for their use (see Section \ref{sec:optimal}) make these functions a useful reference point for comparison with simpler constructions and a suitable choice in contexts where mesh element quality is hard to control. \subsection{Wachspress Coordinates} \begin{figure}[ht] \begin{center} \psfrag{vi}{$\textbf{v}_i$} \psfrag{vip1}{$\textbf{v}_{i+1}$} \psfrag{vim1}{$\textbf{v}_{i-1}$} \psfrag{x}{$\textbf{x}$} \psfrag{Aix}{$A_i(\textbf{x})$} \psfrag{Aip1x}{$A_{i+1}(\textbf{x})$} \psfrag{Aim1x}{$A_{i-1}(\textbf{x})$} \psfrag{wBi}{$B_i$} \psfrag{dots}{$\cdots$} \[\begin{array}{ccc} \includegraphics[width=.3\linewidth]{img/triang-wach1.eps} & \includegraphics[width=.3\linewidth]{img/triang-wach2.eps} \end{array}\] \end{center} \caption{Left: Notation for $A_i(\textbf{x})$. Right: Notation for $B_i$.} \label{fig:wach-ntn} \end{figure} One of the earliest generalizations of barycentric coordinates was provided by Wachspress \cite{W1975}. Definition of these coordinates is defined based on some notation shown in Figure \ref{fig:wach-ntn}. Let $\textbf{x}$ denote an interior point of $\Omega$ and let $A_i(\textbf{x})$ denote the area of the triangle with vertices $\textbf{x}$, $\textbf{v}_i$, and $\textbf{v}_{i+1}$ where, by convention, $\textbf{v}_{0} := \textbf{v}_{n}$ and $\textbf{v}_{n+1} := \textbf{v}_{1}$. Let $B_i$ denote the area of the triangle with vertices $\textbf{v}_{i-1}$, $\textbf{v}_i$, and $\textbf{v}_{i+1}$. Define the Wachspress weight function \[w^{\rm Wach}_i(\textbf{x}) = B_i \prod_{j\not=i,i-1}A_j(\textbf{x}).\] The Wachspress coordinates are then given by \begin{equation} \label{eq:wach} \lambda^{\rm Wach}_i(\textbf{x})=\frac{w^{\rm Wach}_i(\textbf{x})}{\sum_{j=1}^{n} w^{\rm Wach}_j(\textbf{x})} \end{equation} These coordinates have received extensive attention in the literature since they can be represented as rational functions in Cartesian coordinates. Their use in finite element schemes has been numerically tested in specific application contexts but to our knowledge has not been evaluated in the general Sobolev error estimate context considered here. We note that $\lambda^{\rm Wach}_i\in H^1(\Omega)$ since it is a rational function with strictly positive denominator on $\Omega$. \begin{remark} Since $B_i$ does not depend on $\textbf{x}$ and $A_i(\textbf{x})$ is linear in $\textbf{x}$, the Wachspress functions are degree $n-2$. By a result from Warren~\cite{W2003}, the Wachspress functions are the unique, lowest degree rational barycentric functions over polygons. For finite element applications, however, the $\lambda_i$ need not be rational. \end{remark} \subsection{Sibson (Natural Neighbor) Coordinates} \begin{figure}[ht] \psfrag{vi}{$\textbf{v}_i$} \psfrag{Ci}{$C_i$} \psfrag{x}{$\textbf{x}$} \psfrag{Dix}{$D(\textbf{x})$} \psfrag{DixCi}{$D(\textbf{x})\cap C_i$} \[\begin{array}{ccc} \includegraphics[width=.3\linewidth]{img/NECexampleA.eps} & \includegraphics[width=.3\linewidth]{img/NECexampleB.eps} & \includegraphics[width=.3\linewidth]{img/NECexampleC.eps} \end{array}\] \caption{Geometric calculation of a Sibson coordinate. $C_i$ is the area of the Voronoi region associated to vertex $\textbf{v}_i$ inside $\Omega$. $D(\textbf{x})$ is the area of the Voronoi region associated to $\textbf{x}$ if it is added to the vertex list. The quantity $D(\textbf{x})\cap C_i$ is exactly $D(\textbf{x})$ if $\textbf{x}=\textbf{v}_i$ and decays to zero as $\textbf{x}$ moves away from $\textbf{v}_i$, with value identically zero at all vertices besides $\textbf{v}_i$.} \label{fig:sibson} \end{figure} The Sibson coordinates \cite{S1980}, also called the natural neighbor or natural element coordinates, make use of Voronoi diagrams on the vertices $\textbf{v}_i$ of $\Omega$. Let $\textbf{x}$ be a point inside $\Omega$. Let $P$ denote the set of vertices $\{\textbf{v}_i\}$ and define \[P'=P\cup \{\textbf{x}\} = \{\textbf{v}_1,\ldots,\textbf{v}_{n},\textbf{x}\}.\] We denote the \textbf{Voronoi cell} associated to a point $\textbf{p}$ in a pointset $Q$ by \[V_Q(\textbf{p}):= \left\{\textbf{y}\in \Omega \, : \, \vsn{\textbf{y}-\textbf{p}} < \vsn{\textbf{y}-\textbf{q}} \, , \, \forall\textbf{q}\in Q\setminus\{\textbf{p}\} \right\}.\] Note that these Voronoi cells have been restricted to $\Omega$ and are thus always of finite size. We fix the notation \[\begin{array}{lcccl} C_i &:=& |V_{P}(\textbf{v}_i)| &=& \left|\{\textbf{y}\in\Omega \, : \, \vsn{\textbf{y}-\textbf{v}_i} < \vsn{\textbf{y}-\textbf{v}_j}\, , \, \forall j\not=i\}\right| \\ && &=& \text{area of cell for $\textbf{v}_i$ in Voronoi diagram on the points of $P$,} \\ \\ D(\textbf{x}) &:=& |V_{P'}(\textbf{x})| &=& \left|\{\textbf{y}\in\Omega \, : \, \vsn{\textbf{y}-\textbf{x}} < \vsn{\textbf{y}-\textbf{v}_i}\, , \, \forall i\}\right| \\ && &=& \text{area of cell for $\textbf{x}$ in Voronoi diagram on the points of $P'$}. \end{array} \] By a slight abuse of notation, we also define \[ D(\textbf{x})\cap C_i := |V_{P'}(\textbf{x})\cap V_{P}(\textbf{v}_i)|. \] The notation is shown in Figure \ref{fig:sibson}. The Sibson coordinates are defined to be \begin{align*} \lambda^{\rm Sibs}_i(\textbf{x}) &:= \frac{D(\textbf{x})\cap C_i}{D(\textbf{x})} & \textnormal{or, equivalently,} & & \lambda^{\rm Sibs}_i(\textbf{x}) &= \frac{D(\textbf{x})\cap C_i}{\sum_{j=1}^{n}D_j(\textbf{x})\cap C_j}. \end{align*} It has been shown that the $\lambda^{\rm Sibs}_i$ are $C^\infty$ on $\Omega$ except at the vertices $\textbf{v}_i$ where they are $C^0$ and on circumcircles of Delaunay triangles where they are $C^1$ \cite{S1980,F1990}. Since the finite set of vertices are the only points at which the function is not $C^1$, we conclude that $\lambda^{\rm Sibs}_i\in H^1(\Omega)$. To close this section, we compare the intra-element smoothness properties of the coordinate types on the interior of $\Omega$. The triangulation coordinates are $C^0$, the Sibson coordinates are $C^1$, and the Wachspress functions and the harmonic coordinates are both $C^\infty$. \section{Generalized Shape Regularity Conditions} \label{sec:geomcond} The invariance property B\ref{b:invariance} allows estimates on diameter-one polygons to be scaled to polygons of arbitrary size. Several well-known properties of planar convex sets to be used throughout the analysis are given in Proposition~\ref{prop:convexfacts}. Let $|\Omega|$ denote the area of convex polygon $\Omega$ and let $|\partial \Omega|$ denote the perimeter of $\Omega$. \begin{proposition}\label{prop:convexfacts} If $\Omega$ is a convex polygon with $\diam(\Omega) = 1$, then \vspace{-.07in} \begin{enumerate}[(i)] \item $|\Omega| < \pi/4$,\label{cf:maxarea} \item $|\partial \Omega| \leq \pi$, \label{cf:maxperimeter} \item $\Omega$ is contained in a ball of radius no larger than $1/\sqrt{2}$, and \label{cf:jungs} \item If convex polygon $\Upsilon$ is contained in $\Omega$, then $|\partial \Upsilon| \leq |\partial \Omega|$. \label{cf:subset} \end{enumerate} \end{proposition} The first three statements are the isodiametric inequality, a corollary to Barbier's theorem, and Jung's theorem, respectively. The last statement is a technical result along the same lines. See \cite{Eg58,YB61,SA00} for more details. Certain combinations of the geometric restrictions (G1-G3) imply additional useful properties for the analysis. These resulting conditions are listed below. \renewcommand{\labelenumi}{G\arabic{enumi}.} \begin{enumerate} \setcounter{enumi}{3} \item \textbf{Minimum interior angle:} There exists $\beta_*\in\R$ such that $\beta_i > \beta_* > 0$ for all $i$.\label{g:minangle} \item \textbf{Maximum vertex count:} There exists $n^*\in\R$ such that $n < n^*$.\label{g:maxdegree} \end{enumerate} For triangles, G\ref{g:minangle} and G\ref{g:maxangle} are the only two important geometric restrictions since G\ref{g:maxdegree} holds trivially and G\ref{g:ratio}$\Leftrightarrow$G\ref{g:minangle}$\Rightarrow$G\ref{g:minedge}. For general polygons, the relationships between these conditions are more complicated; for example, a polygon satisfying G\ref{g:ratio} may have vertices which are arbitrarily close to each other and thus might not satisfy G\ref{g:maxdegree}. Proposition~\ref{prop:Grelation} below specifies when the original geometric assumptions (G1-G3) imply G\ref{g:minangle} or G\ref{g:maxdegree}. \begin{proposition} The following implications hold. \label{prop:Grelation} \vspace{-.07in} \begin{enumerate}[(i)] \item G\ref{g:ratio} $\Rightarrow$ G\ref{g:minangle} \label{gr:getminangle} \item (G\ref{g:minedge} or G\ref{g:maxangle}) $\Rightarrow$ G\ref{g:maxdegree}\label{gr:getmaxdegree} \end{enumerate} \end{proposition} \begin{proof} G\ref{g:ratio} $\Rightarrow$ G\ref{g:minangle}: If $\beta_i$ is an interior angle, then $\rho(\Omega) \leq \sin(\beta_i/2)$ (see Figure~\ref{fig:minIntAngProof}). Thus $\gamma > \frac{1}{\sin(\beta_i/2)}$. We conclude that $\beta_i > 2 \arcsin \frac{1}{\gamma^*}$. Note that $\gamma^*\geq 2$ so this is well-defined. \begin{figure}[ht] \begin{center} \psfrag{vi}{$\textbf{v}_i$} \psfrag{Bi}{$\beta_i$} \psfrag{c}{\textcolor{red}{$\textbf{c}$}} \psfrag{r}{\textcolor{red}{$\rho(\Omega)$}} \includegraphics[width=.3\linewidth]{img/minIntAngProof.eps} \end{center} \caption{Proof that G1 $\Rightarrow$ G4. The upper angle in the triangle is $\leq\beta_i/2\leq \pi/2$ and the hypoteneuse is $\leq\diam(\Omega)=1$. Thus $\rho(\Omega)\leq \sin(\beta_i/2)$.} \label{fig:minIntAngProof} \end{figure} G\ref{g:minedge} $\Rightarrow$ G\ref{g:maxdegree}: By Jung's theorem (Proposition~\ref{prop:convexfacts}(\ref{cf:jungs})), there exists $\textbf{x}\in\Omega$ such that $\Omega \subset B(\textbf{x},1/\sqrt{2})$. By G\ref{g:minedge}, $\{B(\textbf{v}_i,d_*/2)\}_{i=1}^{n}$ is a set of disjoint balls. Thus $B(\textbf{x}, 1/\sqrt{2}+d_*/2)$ contains all of these balls. Comparing the areas of $\bigcup_{i=1}^{n} B(\textbf{v}_i,d_*/2)$ and $B(\textbf{x}, 1/\sqrt{2}+d_*/2)$ gives $n\frac{\pi d_*^2}{4} < \pi (\frac 1{\sqrt 2}+d_*/2)^2$, so $n < \frac{(\sqrt{2}+d_*)^2}{d_*^2}$. G\ref{g:maxangle} $\Rightarrow$ G\ref{g:maxdegree}: Since $\Omega$ is convex, $\sum_{i=1}^{n} \beta_i = \pi(n-2)$. So $n \beta^* \geq \pi(n-2)$. Thus $n \leq \frac{2\pi}{\pi-\beta^*}$. \qed \end{proof} \section{Interpolation in Sobolev Spaces} \label{sec:intsobsp} Interpolation error estimates are typically derived from the Bramble-Hilbert lemma which says that Sobolev functions over a certain domain or class of domains can be approximated well by polynomials. The original lemma \cite{BH70} applied to a fixed domain (typically the ``reference'' element) and did not indicate how the estimate was impacted by domain geometry. Later, a constructive proof based on the averaged Taylor polynomial gave a uniform estimate under the geometric restriction G\ref{g:ratio} \cite{DS80,BS08}. Recent improvements to this construction have demonstrated that even the condition G\ref{g:ratio} is unnecessary \cite{Ve99,DL04}. This modern version of the Bramble-Hilbert lemma is stated below and has been specialized to our setting, namely, the $H^1$ estimate for diameter $1$, convex domains. \begin{lemma}[\cite{Ve99,DL04}]\label{lem:bramblehilbert} Let $\Omega$ be a convex polygon with diameter $1$. For all $u\in H^2(\Omega)$, there exists a first order polynomial $p_u$ such that $\hpn{u-p_u}{1}{\Omega} \leq C_{BH} \, \hpsn{u}{2}{\Omega}$. \end{lemma} We emphasize that the constant $C_{BH}$ is uniform over all convex sets of diameter $1$. The $H^1$-interpolant estimate (\ref{eq:iuh1uh2}) and Lemma~\ref{lem:bramblehilbert} together ensure the desired optimal convergence estimate (\ref{eq:hconv}). \begin{theorem} Let $\Omega$ be a convex polygon with diameter $1$. If the $H^1$-interpolant estimate (\ref{eq:iuh1uh2}) holds, then for all $u\in H^2(\Omega)$, \[ \hpn{u-Iu}{1}{\Omega} \leq (1 + C_I)\, \sqrt{1+C_{BH}^2} \, \hpsn{u}{2}{\Omega}. \] \end{theorem} \begin{proof} Let $p_u$ be the polynomial given in Lemma~\ref{lem:bramblehilbert} which closely approximates $u$. By property B\ref{b:lincomp}, $Ip_u = p_u$ yielding the estimate \begin{align*} \hpn{u-Iu}{1}{\Omega} & \leq \hpn{u-p_u}{1}{\Omega} + \hpn{I(u-p_u)}{1}{\Omega}\\ & \leq (1 + C_I)\hpn{u-p_u}{2}{\Omega} \leq (1 + C_I)\sqrt{1+C_{BH}^2} \hpsn{u}{2}{\Omega}.\qed \end{align*} \end{proof} \begin{corollary} Let $\diam(\Omega) \leq 1$. If the $H^1$-interpolant estimate (\ref{eq:iuh1uh2}) holds, then for all $u\in H^2(\Omega)$, \[ \hpn{u-Iu}{1}{\Omega} \leq (1 + C_I)\, \sqrt{1+C_{BH}^2}\, \diam(\Omega) \, \hpsn{u}{2}{\Omega}. \] \end{corollary} \begin{proof} This follows from the standard scaling properties of Sobolev norms since property B\ref{b:invariance} allows for a change of variables to a unit diameter domain. Note: the $L^2$-component of the $H^1$-norm satisfies a stronger estimate containing an extra power of $\diam(\Omega)$. \qed \end{proof} Section~\ref{sec:errorest} is an investigation of the geometric conditions under which the $H^1$-interpolant estimate (\ref{eq:iuh1uh2}) holds for the barycentric functions discussed in Section~\ref{sec:genbary}. Under the geometric restrictions G\ref{g:ratio} and G\ref{g:maxdegree}, one method for verifying (\ref{eq:iuh1uh2}) (utilized in \cite{BS08} for simplicial interpolation) is to bound the $H^1$-norm of the basis functions. In several cases we will utilize this criteria which is justified by the following lemma. \begin{lemma} \label{lem:basisbd} Under G\ref{g:ratio} and G\ref{g:maxdegree}, the $H^1$-interpolant estimate (\ref{eq:iuh1uh2}) holds whenever there exists a constant $C_\lambda$ such that \begin{equation}\label{eq:basisbound} \hpn{\lambda_i}{1}{\Omega} \leq C_\lambda. \end{equation} \end{lemma} \begin{proof} This follows almost immediately from the Sobolev embedding theorem; see \cite{Ad03,Le09}: \[ \hpn{Iu}{1}{\Omega} \leq \sum_{i=1}^{n} |u(\textbf{v}_i)| \hpn{\lambda_i}{1}{\Omega} \leq n^* C_\lambda\, \vn{u}_{C^0(\overline{\Omega})} \leq n^* \,C_\lambda\, C_s\, \hpn{u}{2}{\Omega}, \] where $C_s$ is the Sobolev embedding; i.e., $\vn{u}_{C^0(\overline{\Omega})} \leq C_s\, \hpn{u}{2}{\Omega}$ for all $u\in H^2(\Omega)$. The constant $C_s$ is independent of the domain $\Omega$ since the boundaries of all polygons satisfying G\ref{g:ratio} are uniformly Lipschitz~\cite{Le09}. \qed \end{proof} \section{Error Estimate Requirements}\label{sec:errorest} \subsection{Estimate Requirements for Triangulation Coordinates} \label{sec:trimeth} Interpolation error estimates on triangles are well understood: the optimal convergence estimate (\ref{eq:hconv}) holds as long as the triangle satisfies a maximum angle condition \cite{BA76,Ja76}. In fact, it has been shown that the triangle circumradius controls the error independent of any other geometric criteria \cite{Kr91}. This result can be directly applied to $I^{\rm Tri}$, the interpolation operator associated to coordinates $\lambda^{\rm Tri}_i$. This convention will also be used to define $I^{\rm Opt}$, $I^{\rm Wach}$, and $I^{\rm Sibs}$ as the interpolation operators associated with with harmonic, Wachspress, and Sibson coordinates, respectively. \begin{lemma} \label{lem:triangulated} Under G\ref{g:maxangle}, the $H^1$ interpolant estimate (\ref{eq:iuh1uh2}) holds for $I^{\rm Tri}$. Conversely, G\ref{g:maxangle} is a necessary assumption to achieve (\ref{eq:iuh1uh2}) with $I^{\rm Tri}$. \end{lemma} \begin{proof} All angles of all triangles of any triangulation ${\mathcal{T}}$ of $\Omega$ satisfying G\ref{g:maxangle} are less than $\beta^*$. Thus, the sufficiency of G\ref{g:maxangle} follows immediately from the maximum angle condition on simplices~\cite{BA76}. An example from the same paper involving the interpolation of a quadratic function over a triangle also establishes the necessity of the condition.\qed \end{proof} \subsection{Estimate Requirements for Harmonic Coordinates} \label{sec:optimal} \begin{figure}[ht] \begin{center} \psfrag{vi}{$\textbf{v}_i$} \psfrag{vip1}{$\textbf{v}_{i+1}$} \psfrag{vim1}{$\textbf{v}_{i-1}$} \psfrag{c}{\textcolor{red}{$\textbf{c}$}} \psfrag{y}{\textcolor{red}{$\textbf{y}$}} \psfrag{x}{$\textbf{x}$} \psfrag{Aix}{$A_i(\textbf{x})$} \psfrag{wBi}{$B_i$} \[\begin{array}{cccc} \includegraphics[height=.2\linewidth]{img/triang-opt.eps} & \includegraphics[height=.35\linewidth]{img/triang-opt-est1.eps} & \text{ }\quad\text{ } & \includegraphics[height=.35\linewidth]{img/triang-opt-est2.eps} \\ \\ \textbf{(a)} & \textbf{(b)} && \textbf{(c)} \end{array}\] \end{center} \caption{\textbf{(a)} Triangulation used in the analysis of Harmonic coordinates. \textbf{(b)} Notation for proof of the bound for $\angle \textbf{c}\textbf{v}_i\textbf{v}_{i+1}$ in a case where it is $>\pi/2$. \textbf{(c)} Notation for proof of the bound for $\angle \textbf{v}_i\textbf{c}\textbf{v}_{i+1}$ in a case where it is $>\pi/2$.} \label{fig:opt-tri} \end{figure} Recalling the notation from Figure~\ref{fig:notation}, let $T$ be the triangulation of $\Omega$ formed by connecting $\textbf{c}$ to each of the $\textbf{v}_i$; see Figure~\ref{fig:opt-tri}a. \begin{proposition}\label{prop:nlatriangulation} Under G\ref{g:ratio} all angles of all triangles of $T$ are less than $\pi - \arcsin (1/\gamma^*)$. \end{proposition} \begin{proof} Consider the triangle with vertices $\textbf{c}$, $\textbf{v}_i$ and $\textbf{v}_{i+1}$. Without loss of generality, assume that $|\textbf{c} - \textbf{v}_i| < |\textbf{c}- \textbf{v}_{i+1}|$. First we bound $\angle\textbf{c} \textbf{v}_{i+1} \textbf{v}_i$. By the law of sines, \begin{equation} \label{eq:lawsinarg} \frac{\sin(\angle\textbf{c} \textbf{v}_{i+1} \textbf{v}_i)}{\sin(\angle\textbf{c} \textbf{v}_i \textbf{v}_{i+1})}=\frac{|\textbf{c} - \textbf{v}_i|}{|\textbf{c} - \textbf{v}_{i+1}|}<1. \end{equation} If $\angle\textbf{c} \textbf{v}_i \textbf{v}_{i+1}>\pi/2$ then $\angle\textbf{c} \textbf{v}_{i+1} \textbf{v}_i<\pi/2$. Otherwise, (\ref{eq:lawsinarg}) implies $\angle\textbf{c}\textbf{v}_{i+1}\textbf{v}_i<\pi/2$. To bound angle $\angle \textbf{c}\textbf{v}_i\textbf{v}_{i+1}$, it suffices to consider the case when $\angle \textbf{c}\textbf{v}_i\textbf{v}_{i+1}>\pi/2$, as shown in Figure~\ref{fig:opt-tri}b. Define $\textbf{y}$ to be the point on the line through $\textbf{v}_i$ and $\textbf{v}_{i+1}$ which forms a right triangle with $\textbf{v}_i$ and $\textbf{c}$. Since $\angle \textbf{c}\textbf{v}_i\textbf{v}_{i+1}>\pi/2$, $\textbf{y}$ is exterior to $\Omega$, as shown. Observe that \[\frac{|\textbf{c}-\textbf{v}_i|}{|\textbf{c}-\textbf{y}|}<\frac{|\textbf{c}-\textbf{v}_{i+1}|}{|\textbf{c}-\textbf{y}|}<\frac{\diam(\Omega)}{\rho(\Omega)}=\gamma<\gamma^*.\] Since $\sin(\pi-\angle \textbf{c}\textbf{v}_i\textbf{v}_{i+1})=\frac {|\textbf{c}-\textbf{y}|}{|\textbf{c}-\textbf{v}_i|}$, the result follows. For the final case, it suffices to assume $\angle \textbf{v}_i\textbf{c}\textbf{v}_{i+1} > \pi/2$, as shown in Figure~\ref{fig:opt-tri}c. Define $\textbf{y}$ in the same way, but note that in this case $\textbf{y}$ is between $\textbf{v}_i$ and $\textbf{v}_{i+1}$, as shown. Similarly, $\frac{|\textbf{c}-\textbf{v}_{i+1}|}{|\textbf{c}-\textbf{y}|}<\gamma^*$, implying $\angle\textbf{v}_i\textbf{v}_{i+1}\textbf{c} >\arcsin(1/\gamma^*)$. Since $\angle\textbf{v}_i\textbf{c}\textbf{v}_{i+1}<\pi-\angle\textbf{v}_i\textbf{v}_{i+1}\textbf{c}$, the result follows.\qed \end{proof} \begin{lemma}\label{lem:optimal} Under G\ref{g:ratio} the operator $I^{\rm Opt}$ satisfies the $H^1$ interpolant estimate (\ref{eq:iuh1uh2}). \end{lemma} \begin{proof} Since the differential equation (\ref{eq:optpde}) is linear, $I^{\rm Opt} u$ is the solution to the differential equation, \begin{equation} \label{eq:ioptpde} \displaystyle\left\{\begin{array}{rcll} \Delta\left(I^{\rm Opt} u \right) & = & 0, & \text{on $\Omega$} \\ I^{\rm Opt}_i u & = & g_u, & \text{on $\partial\Omega$} \end{array}\right. \end{equation} where $g_u$ is the piecewise linear function which equals $u$ at the vertices of $\Omega$. Following the standard approach for handling nonhomogeneous boundary data we divide: $I^{\rm Opt} u = u_{\rm hom} + u_{\rm non}$ where $u_{\rm non}\in H^1(\Omega)$ is some function satisfying the boundary condition (i.e., $u_{\rm non} = g_u$ on $\partial\Omega$) and $u_{\rm hom}$ solves, \begin{equation} \label{eq:uhom} \displaystyle\left\{\begin{array}{rcll} \Delta u_{\rm hom} & = & -\Delta u_{\rm non}, & \text{on $\Omega$} \\ u_{\rm hom} & = & 0, & \text{on $\partial\Omega$}. \end{array}\right. \end{equation} Specifically we select $u_{\rm non}$ to be the standard Lagrange interpolant of $u$ over triangulation $T$ (described earlier). Since Proposition~\ref{prop:nlatriangulation} guarantees that no large angles exist in the triangulation, the standard interpolation error estimate holds, \begin{equation} \label{eq:nlainterpolation} \hpn{u - u_{\rm non}}{1}{\Omega} \leq C_{BA} \, \hpsn{u}{2}{\Omega} \end{equation} where $C_{BA}$ only depends upon the aspect ratio bound $\gamma^*$ since $\diam(\Omega) = 1$. The triangle inequality then implies that $\hpn{u_{\rm non}}{1}{\Omega} \leq \max(1,C_{BA}) \hpn{u}{2}{\Omega}$. Next a common energy estimate (see \cite{Ev98}) for (\ref{eq:uhom}) implies that $\hpsn{u_{\rm hom}}{1}{\Omega} \leq \hpsn{u_{\rm non}}{1}{\Omega}$. The Poincar\'e inequality (see \cite{Le09}) ensures that $\lpn{u_{\rm hom}}{2}{\Omega} \leq C_P \hpsn{u}{1}{\Omega}$ where $C_P$ only depends on the diameter of $\Omega$ which we have fixed to be $1$. The argument is completed by combining the previous estimates: \begin{align*} \hpn{I^{\rm Opt} u}{1}{\Omega} & \leq \hpn{u_{\rm hom}}{1}{\Omega} + \hpn{u_{\rm non}}{1}{\Omega} \\ & \leq (1 + C_P) \hpsn{u_{\rm hom}}{1}{\Omega} + \hpn{u_{\rm non}}{1}{\Omega}\\ & \leq (1 + C_P)\hpn{u_{\rm non}}{1}{\Omega} \leq (1 + C_P)\max(1,C_{BA}) \hpn{u}{2}{\Omega}. \qed \end{align*} \end{proof} \subsection{Estimate Requirements for Wachspress Coordinates} \label{sec:wachpress} \begin{figure}[ht] \begin{center} \psfrag{v1}{$\textbf{v}_1$} \[\begin{array}{ccc} \includegraphics[width=.3\linewidth]{img/wach-cex-eps-pt10.eps} & \includegraphics[width=.3\linewidth]{img/wach-cex-eps-pt05.eps} & \includegraphics[width=.3\linewidth]{img/wach-cex-eps-pt025.eps} \end{array}\] \end{center} \caption{Example showing the necessity of condition G\ref{g:maxangle} for attaining the optimal convergence estimate (\ref{eq:hconv}) with the Wachspress coordinates. As the shape approaches a square, the level sets of $\lambda^{\rm Wach}_1$ collect at the top edge, causing a steep gradient and thus preventing a bound on the $H^1$ norm of the error. The figures from left to right correspond to $\epsilon$ values of 0.1, 0.05, and 0.025. } \label{fig:wach-cex} \end{figure} Unlike the harmonic coordinate functions, the Wachspress coordinates can produce unsatisfactory interpolants unless additional geometric conditions are imposed. We present a simple counterexample (observed qualitatively in \cite{FHK2006} and in Figure~\ref{fig:wach-cex}) to show what can go wrong. Let ${\Omega_\epsilon}$ be the pentagon defined by the vertices \[ \textbf{v}_1=(0,1+\epsilon),\quad \textbf{v}_2=(1,1),\quad \textbf{v}_3=(1,-1),\quad \textbf{v}_4=(-1,-1),\quad \textbf{v}_5=(-1,1), \] with $\epsilon>0$. As $\epsilon\rightarrow 0$, ${\Omega_\epsilon}$ approaches a square so G1 is not violated. Consider the interpolant of $u(\textbf{x})=1-x_1^2$ where $\textbf{x} =(x_1,x_2)$. Observe that $u$ has value 1 at $\textbf{v}_1$ and value $0$ at the other vertices of ${\Omega_\epsilon}$. Hence \[I^{\rm Wach} u=\sum_{i=1}^5 u(\textbf{v}_i)\lambda^{\rm Wach}_i=\lambda^{\rm Wach}_1\] Using the fact that $\partial u/\partial y = 0$, we write \begin{align*} \hpn{u - I^{\rm Wach} u}{1}{{\Omega_\epsilon}}^2 &= \hpn{u - \lambda^{\rm Wach}_1}{1}{{\Omega_\epsilon}}^2 \\ & = \int_{\Omega_\epsilon} |u - \lambda^{\rm Wach}_1|^2+\left|\frac{\partial (u - \lambda^{\rm Wach}_1)}{\partial x}\right|^2+\left|\frac{\partial \lambda^{\rm Wach}_1}{\partial y}\right|^2. \end{align*} The last term in this sum blows up as $\epsilon \rightarrow 0$. \begin{lemma} \label{lem:wachcex} $\displaystyle\lim_{\epsilon\rightarrow 0}\int_{\Omega_\epsilon}\left|\frac{\partial \lambda^{\rm Wach}_1}{\partial y}\right|^2=\infty$. \end{lemma} The proof of the lemma is given in the appendix. As a corollary, we observe that $\hpn{u - Iu}{1}{{\Omega_\epsilon}}$ cannot be bounded independently of $\epsilon$. Since $\hpn{u}{2}{{\Omega_\epsilon}}$ is finite, this means the optimal convergence estimate (\ref{eq:hconv}) cannot hold without additional geometric criteria on the domain $\Omega$. This establishes the necessity of a maximum interior angle bound on vertices if Wachspress coordinates are used. Under the three geometric restrictions G\ref{g:ratio}, G\ref{g:minedge}, and G\ref{g:maxangle}, (\ref{eq:hconv}) does hold which will be shown in Lemma~\ref{lm:wach}. We begin with some preliminary estimates. \begin{proposition}\label{pr:gradarea} For all $\textbf{x} \in \Omega$, $\vsn{\nabla A_i(\textbf{x})} \leq \frac{1}{2}$. \end{proposition} \begin{proof} In \cite[Equation (17)]{FK10} it is shown that the $\vsn{\nabla A_i(\textbf{x})}=\frac{1}{2}\vsn{\textbf{v}_i - \textbf{v}_{i+1}}$. Since $\diam (\Omega) = 1$ the result follows. \qed \end{proof} Next we show that the triangular areas $B_i$ are uniformly bounded from below given our geometric assumptions. \begin{proposition}\label{pr:carealower} Under G\ref{g:ratio}, G\ref{g:minedge}, and G\ref{g:maxangle}, there exists $B_*$ such that $B_i >B_*$. \end{proposition} \begin{proof} By G\ref{g:minedge}, the area of the isosceles triangle with equal sides of length $d_*$ meeting with angle $\beta_i$ at $\textbf{v}_i$ is a lower bound for $B_i$, as shown in Figure~\ref{fig:ajxlower} (left). More precisely, $B_i > (d_*)^2 \sin (\beta_i/2) \cos(\beta_i/2)$. G\ref{g:maxangle} implies that $\cos(\beta_i/2) > \cos(\beta^*/2)$. G\ref{g:minangle} (which follows from G\ref{g:ratio} by Proposition \ref{prop:Grelation}) implies that $\sin (\beta_i/2) > \sin(\beta_*/2)$. Thus $B_i > B_* := (d_*)^2 \sin (\beta_*/2) \cos(\beta^*/2)$. \qed \end{proof} Proposition~\ref{pr:carealower} can be extended to guarantee a uniform lower bound on the sum of the Wachspress weight functions. \begin{proposition}\label{pr:wachweightlower} Under G\ref{g:ratio}, G\ref{g:minedge}, and G\ref{g:maxangle}, there exists $w_*$ such that for all $\textbf{x}\in \Omega$, $$\sum_k w^{\rm Wach}_k(\textbf{x}) > w_*.$$ \end{proposition} \begin{proof} Let $\textbf{v}_i$ be the nearest vertex to $\textbf{x}$, breaking any tie arbitrarily. We will produce a lower bound on $w_i(\textbf{x})$. Let $j \notin \{i-1,i\}$. G\ref{g:minedge} implies that $\vsn{\textbf{x}-\textbf{v}_j} > d_*/2$ and $\vsn{\textbf{x}-\textbf{v}_{j+1}} > d_*/2$. G\ref{g:maxangle} implies that $\angle \textbf{x}\textbf{v}_j\textbf{v}_{j+1} < \beta^*$ and $\angle \textbf{x}\textbf{v}_{j+1}\textbf{v}_j < \beta^*$. It follows that $A_j(\textbf{x}) > (d_*)^2\sin(\pi-\beta^*)/4$ (see Figure~\ref{fig:ajxlower} (right)). We now use Proposition~\ref{pr:carealower} and property G\ref{g:maxdegree} (which follows from either G\ref{g:minedge} or G\ref{g:maxangle} by Proposition~\ref{prop:Grelation}) to conclude that \[ \sum_k w^{\rm Wach}_k(\textbf{x}) > w^{\rm Wach}_i(\textbf{x}) = B_i \prod_{j\not=i,i-1}A_j(\textbf{x}) \geq B_* \left[(d_*)^2\sin(\pi-\beta^*)/4\right]^{n^*-2}. \] \qed \end{proof} \begin{figure}[ht] \begin{center} \psfrag{vi}{$\textbf{v}_i$} \psfrag{vj}{$\textbf{v}_j$} \psfrag{vjp1}{$\textbf{v}_{j+1}$} \psfrag{vip1}{$\textbf{v}_{i+1}$} \psfrag{vim1}{$\textbf{v}_{i-1}$} \psfrag{dstar}{$d_*$} \psfrag{x}{$\textbf{x}$} \psfrag{Ajx}{$A_j(\textbf{x})$} \[ \includegraphics[width=1.8in]{img/bilower.eps}\quad\text{ }\quad \includegraphics[width=1.5in]{img/ajxlower.eps} \] \end{center} \caption{\textbf{Left:} Justification of claim that $B_i > (d_*)^2 \sin (\beta_i/2) \cos(\beta_i/2)$ in the proof of Proposition~\ref{pr:carealower}. The shaded triangle is isosceles with angle $\beta_i$ and two side lengths equal to $d_*$ as indicated. Computing the area of this triangle using the dashed edge as the base yields the estimate. \textbf{Right:} Justification of claim that $A_j(\textbf{x}) > (d_*)^2\sin(\pi-\beta^*)/4$ in the proof of Proposition~\ref{pr:wachweightlower}. The indicated angle is at least $\pi-\beta^*$ by G\ref{g:maxangle} and $\vsn{\textbf{v}_{j}-\textbf{v}_{j+1}}>d_*$. Computing the area of the triangle using edge $\textbf{v}_{j}\textbf{v}_{j+1}$ as the base yields the estimate.} \label{fig:ajxlower} \end{figure} \begin{lemma}\label{lm:wach} Under G1, G2, and G3, (\ref{eq:basisbound}) holds for the Wachspress coordinates. \end{lemma} \begin{proof} The gradient of $\lambda^{\rm Wach}_i(\textbf{x})$ can be bounded using Proposition \ref{pr:wachweightlower}: \begin{align} \vsn{\nabla \lambda^{\rm Wach}_i(\textbf{x})} & \leq \frac{\vsn{\nabla w^{\rm Wach}_i(\textbf{x})}}{\sum_j w^{\rm Wach}_j(\textbf{x})} + \frac{w^{\rm Wach}_i(\textbf{x})\sum_k \vsn{\nabla w^{\rm Wach}_k(\textbf{x})}}{\left(\sum_j w^{\rm Wach}_j(\textbf{x})\right)^2}\notag\\ & \leq \frac{\vsn{\nabla w^{\rm Wach}_i(\textbf{x})} + \sum_k \vsn{\nabla w^{\rm Wach}_k(\textbf{x})}}{\sum_j w^{\rm Wach}_j(\textbf{x})} \leq \frac{2 \sum_k \vsn{\nabla w^{\rm Wach}_k(\textbf{x})}}{w_*}.\label{eq:wachgradbound} \end{align} Recalling Proposition~\ref{prop:convexfacts}, $\sum_{j=1}^{n} A_j(\textbf{x}) < \pi/4$ and $B_i<\pi/4$. Using Proposition~\ref{pr:gradarea} and the arithmetic mean-geometric mean inequality, we derive \begin{align} \vsn{\nabla w^{\rm Wach}_i(\textbf{x})} &= \vsn{\sum_{j\neq i-1,i} B_i \nabla A_j(\textbf{x}) \prod_{k\neq i-1,i,j} A_k(\textbf{x})}\notag \\ & \leq \sum_{j\neq i-1,i}\left[ B_i \vsn{\nabla A_j(\textbf{x})} \prod_{k\neq i-1,i,j} A_k(\textbf{x})\right]\notag \\ & \leq \sum_{j\neq i-1,i}\frac{\pi}{8}\left[ \frac{\sum_{k\neq i-1,i,j} A_k(\textbf{x})}{n-3} \right]^{n-3} \leq \sum_{j\neq i-1,i}\frac{\pi}{8}\left[ \frac{\pi}{4(n-3)} \right]^{n-3} \notag \\ & = \frac{\pi}{8}(n-2)\left[ \frac{\pi}{4(n-3)}\right]^{n-3}. \label{eq:wachgradweight} \end{align} By induction, one can show that $n(n-2)\left[ \frac{\pi}{4(n-3)}\right]^{n-3} \leq 2\pi$ for $n\geq 4$. Using this, we substitute (\ref{eq:wachgradweight}) into (\ref{eq:wachgradbound}) to get \[\vsn{\nabla \lambda^{\rm Wach}_i(\textbf{x})}\leq \frac 2{w_*}\sum_k|\nablaw^{\rm Wach}_k(\textbf{x})|\leq\frac 2{w_*}n\frac{\pi}{8}(n-2)\left[ \frac{\pi}{4(n-3)}\right]^{n-3}\leq \frac{\pi}{4w_*}2\pi =\frac{\pi^2}{2w_*}. \] Since $|\Omega|<\pi/4$ by Proposition \ref{prop:convexfacts}, we thus have a uniform bound \[\hpn{\lambda^{\rm Wach}_i}{1}{\Omega} \leq \sqrt{\left(1 + \frac{\pi^4}{4w_*^2}\right)\frac{\pi}{4}}.\qed \] \end{proof} \subsection{Estimate Requirements for Sibson Coordinates} \label{sec:sibson} The interpolation estimate for Sibson coordinates is computed using a very similar approach to that of the previous section on Wachspress coordinates. However in this case the geometric condition G\ref{g:maxangle} is not necessary. We begin with a technical property of domains satisfying conditions G\ref{g:ratio} and G\ref{g:minedge}. \begin{proposition}\label{prop:ballinvoronoi} Under G\ref{g:ratio} and G\ref{g:minedge}, there exists $h_* > 0$ such that for all $\textbf{x}\in \Omega$, $B(\textbf{x},h_*)$ does not intersect any three edges or any two non-adjacent edges of $\Omega$. \end{proposition} \begin{proof} Let $\textbf{x} \in \Omega$, $h\in (0,d_*/2)$, and suppose that two disjoint edges of $\Omega$, $e_i$ and $e_j$, intersect $B(\textbf{x}, h)$. Let $L_i$ and $L_j$ be the lines containing $e_i$ and $e_j$ and let $\theta$ be the angle between these lines; see Figure~\ref{fig:ballinset}. We first consider the case where $L_i$ and $L_j$ are not parallel and define $\textbf{z} = L_i \cap L_j$. \begin{figure} \centering \psfrag{x}{$\textbf{x}$} \psfrag{z}{$\textbf{z}$} \psfrag{ei}{$e_i$} \psfrag{ej}{$e_j$} \psfrag{vi}{$\textbf{v}_i$} \psfrag{vj}{$\textbf{v}_j$} \psfrag{Li}{$L_i$} \psfrag{Lj}{$L_j$} \psfrag{h}{$h$} \psfrag{W}{$W$} \psfrag{theta}{$\theta$} \psfrag{omega}{$\Omega$} \includegraphics[width=.75\linewidth]{img/ballinset.eps} \caption{Notation for proof of Proposition~\ref{prop:ballinvoronoi}.} \label{fig:ballinset} \end{figure} Let $\textbf{v}_i$ and $\textbf{v}_j$ be the endpoints of $e_i$ and $e_j$ nearest to $\textbf{z}$. Since $h < d_*/2$ both $\textbf{v}_i$ and $\textbf{v}_j$ cannot live in $B(\textbf{x}, h)$; without loss of generality assume that $\textbf{v}_i\notin B(\textbf{x}, h)$. Since $\dist(\textbf{v}_j,L_i) < 2h$, \begin{equation}\label{eq:thetabound} \sin\theta < 2h/\vsn{\textbf{z} - \textbf{v}_j}. \end{equation} Let $W$ be the sector between $L_i$ and $L_j$ containing $x$. Now $\Omega \subset B(\textbf{v}_j,1)\cap W \subset B(\textbf{z}, 1 + \vsn{\textbf{z}-\textbf{v}_j})\cap W$. It follows that $\rho(\Omega) \leq (1 + \vsn{\textbf{v}_j - \textbf{z}})\sin\theta $. Using (\ref{eq:thetabound}) and G\ref{g:ratio}, \[ \frac{1}{\gamma^*} \leq \frac{2h}{\vsn{\textbf{z} - \textbf{v}_j}}(1 + \vsn{\textbf{z} - \textbf{v}_j}) \leq 2h\left(\frac{1}{d_*} + 1 \right) \] where the final inequality holds because by G\ref{g:minedge} $\vsn{\textbf{z} - \textbf{v}_j} \geq \vsn{\textbf{v}_i - \textbf{v}_j} \geq d_*$. Thus \begin{equation}\label{eq:hlwrbd} h > \frac{d_*}{2\gamma^*(1+d_*)}. \end{equation} Estimate (\ref{eq:hlwrbd}) holds in the limiting case: when $L_i$ and $L_j$ are parallel. In this case $\Omega$ must be contained in a strip of width $2h$ which for small $h$ violates the aspect ratio condition. The triangle is the only polygon with three or more pairwise non-adjacent edges. So it remains to find a suitable $h_*$ so that $B(\textbf{x},h_*)$ does not intersect all three edges of the triangle. For a triangle, $\rho(\Omega)$ is the radius of the smallest circle touching all three edges. Since under G\ref{g:ratio} $\rho(\Omega) \geq 1/\gamma^*$, $B(\textbf{x},\frac{1}{2\gamma^*})$ intersects at most two edges. Thus $h_* = \frac{d_*}{2\gamma^*(1+d_*)}$ is sufficiently small to satisfy the proposition in all cases. \qed \end{proof} Proposition~\ref{prop:ballinvoronoi} is a useful tool for proving a lower bound on $D(\textbf{x})$, the area of the Voronoi cell of $\textbf{x}$ intersected with $\Omega$. \begin{proposition}\label{prop:minvoronoiarea} Under G\ref{g:ratio} and G\ref{g:minedge}, there exists $D_* >0$ such that $D(\textbf{x}) > D_*$. \end{proposition} \begin{proof} Let $h_*$ be the constant in Proposition~\ref{prop:ballinvoronoi}. We consider two cases, based on whether the point $\textbf{x}$ is near any vertex of $\Omega$, as shown in Figure~\ref{fig:sibsproof} (left). \noindent \underline{Case 1}: There exists $\textbf{v}_i$ such that $\textbf{x} \in B(\textbf{v}_i,h_*/2)$. \begin{figure} \centering \psfrag{x}{$\textbf{x}$} \psfrag{vi}{$\textbf{v}_i$} \psfrag{hstar}{$h_*$} \psfrag{Bi}{$\beta_i$} \[\begin{array}{ccc} \includegraphics[width=.3\linewidth]{img/cutCorners.eps} & \quad\text{ }\quad & \includegraphics[width=.23\linewidth]{img/nearvertexcase.eps} \end{array}\] \caption{The proof of Proposition~\ref{prop:minvoronoiarea} has two cases based on whether $\textbf{x}$ is within $h_*/2$ of some $\textbf{v}_i$ or not. When $\textbf{x}$ is within $h_*/2$ of $\textbf{v}_i$, the shaded sector shown on the right is contained in $V_{P'}(\textbf{x})\cap\Omega$.} \label{fig:sibsproof} \end{figure} Consider the sector of $B(\textbf{x},h_*/2)$ specified by segments which are parallel to the edges of $\Omega$ containing $\textbf{v}_i$, as shown in Figure~\ref{fig:sibsproof} (right). This sector must be contained in $\Omega$ by Proposition~\ref{prop:ballinvoronoi} and in the Voronoi cell of $\textbf{x}$ by choice of $h_*<d_{*}$. Thus by G\ref{g:minangle} (using Proposition~\ref{prop:Grelation}(\ref{gr:getminangle})) $D(\textbf{x}) \geq \beta_* h_*^2/8$. \noindent \underline{Case 2}: For all $\textbf{v}_i$, $\textbf{x} \notin B(\textbf{v}_i,h_*/2)$. In this case, $B(\textbf{x}, h_*/4) \cap \Omega \subset V_{P'}(\textbf{x})$. If $B(\textbf{x}, h_*/4)$ intersects zero or one boundary edge of $\Omega$, then $D(\textbf{x}) \geq \pi h_*^2/32$. Otherwise $B(\textbf{x}, h_*/4)$ intersects two adjacent boundary edges. By G\ref{g:minangle}, $D(\textbf{x}) \geq \beta_* h_*^2/32$. \qed \end{proof} General formulas for the gradient of the area of a Voronoi cell are well-known and can be used to bound the gradients of $D(\textbf{x})$ and $D(\textbf{x})\cap C_i$. \begin{proposition}\label{prop:voronoigrad} $|\nabla D(\textbf{x})| \leq \pi$ and $|\nabla (D(\textbf{x})\cap C_i)| \leq 1$. \end{proposition} \begin{proof} The gradient of the area of a Voronoi region is known to be \[ \nabla D(\textbf{x}) = \sum_{j=1}^{n} \frac{\textbf{v}_j - \textbf{x}}{\vsn{\textbf{v}_j - \textbf{x}}} F_j, \] where $F_j$ is the length of the segment separating the Voronoi cells of $\textbf{x}$ and $\textbf{v}_j$~\cite{OA91,OBSC00}. Then applying Proposition~\ref{prop:convexfacts} gives \[ \vsn{\nabla D(\textbf{x})} \leq \sum_{i=1}^{n} F_i \leq |\partial\Omega|\leq \pi. \] Similarly, \[ \nabla (D(\textbf{x})\cap C_i) = \frac{\textbf{v}_i - \textbf{x}}{\vsn{\textbf{v}_i - \textbf{x}}} F_i, \] and since $F_i \leq \diam(\Omega)$, $|\nabla (D(\textbf{x})\cap C_i)| \leq 1$. \qed \end{proof} Propositions~\ref{prop:minvoronoiarea} and \ref{prop:voronoigrad} give estimates for the key terms needed in proving (\ref{eq:basisbound}) for the Sibson coordinates $\lambda^{\rm Sibs}$. \begin{lemma}\label{lm:sibs} Under G1 and G2, (\ref{eq:basisbound}) holds for the Sibson coordinates. \end{lemma} \begin{proof} $\vsn{\nabla \lambda^{\rm Sibs}_i}$ is estimated by applying Propositions~\ref{prop:minvoronoiarea} and \ref{prop:voronoigrad}: \begin{align*} \vsn{\nabla \lambda^{\rm Sibs}_i} &\leq \frac{|\nabla (D(\textbf{x})\cap C_i)|}{D(\textbf{x})} + \frac{(D(\textbf{x})\cap C_i) \vsn{\nabla D(\textbf{x})} }{D(\textbf{x})^2} \leq \frac{|\nabla (D(\textbf{x})\cap C_i)| + |\nabla D(\textbf{x})|}{D(\textbf{x})} \\ &\leq \frac{1 + \pi}{D_*}. \end{align*} Integrating this estimate completes the result. \qed \end{proof} \begin{corollary} By Lemma \ref{lem:basisbd}, the $H^1$ interpolant estimate (\ref{eq:iuh1uh2}) holds for the Sibson coordinates. \end{corollary} \section{Final Remarks} \label{sec:conc} Geometric requirements needed to ensure optimal interpolation error estimates are necessary for guaranteeing the compatibility of polygonal meshes with generalized barycentric interpolation schemes in finite element methods. Moreover, the identification of necessary and unnecessary geometric restrictions provides a tool for comparing various approaches to barycentric interpolation. Specifically we have demonstrated the necessity of a maximum interior angle restriction for Wachspress coordinates, which was empirically observed in \cite{FHK2006}, and shown that this restriction is unneeded when using Sibson coordinates. Table \ref{tab:conditions} provides a guideline for how to choose barycentric basis functions given geometric criteria or, conversely, which geometric criteria should be guaranteed given a choice of basis functions. While utilized throughout our analysis, the aspect ratio requirement G\ref{g:ratio} can likely be substantially weakened. Due to a dependence on specific affine transformations, such techniques on triangular domains \cite{BA76,Ja76} (i.e., methods for proving error estimates under the maximum angle condition rather than the minimum angle condition) cannot be naturally extended to polygonal domains. Although aimed at a slightly different setting that we have analyzed, challenges in identifying sharp geometric restrictions are apparent from the numerous studies on quadrilateral elements, e.g., \cite{Ja77,ZV95,AD01,MNS08}. A satisfactory generalization of the maximum angle condition to arbitrary polygons is a subject of further investigation. This paper emphasizes three specific barycentric coordinates (harmonic, Wachspress, and Sibson) but several others have been introduced in the literature. Maximum entropy \cite{Su04}, metric \cite{MLBD02}, and discrete harmonic \cite{PP93} coordinates can all be studied either by specific analysis or generalizing the arguments given here to wider classes of functions. The mean value coordinates defined by Floater~\cite{F2003} are of particular interest in this regard as they are defined by an explicit formula and appear to not require a maximum angle condition. The formal analysis of these functions, however, is not trivial. Additional generalizations could be considered by dropping certain restrictions on the coordinates, such as non-negativity, or the mesh elements, such as convexity. Working with non-convex elements, however, would require some non-obvious generalization of the geometric restrictions G1-G5. \section{Proof of Lemma \ref{lem:wachcex}} \noindent \begin{proof} The explicit formula for the Wachspress weight associated to $\textbf{v}_1$ is \[w^{\rm Wach}_1(\textbf{x})= B_1A_2A_3A_4 =\epsilon(1-x)(1+x)(1+y) \] where $\textbf{x}=(x,y)$ is an arbitrary point inside ${\Omega_\epsilon}$. The other weights can be computed similarly, yielding the coordinate function \[\lambda^{\rm Wach}_1= \frac{w_{i}(\textbf{v})}{\sum_{j=1}^5w_{j}(\textbf{v})} = \frac{\epsilon(1-x)(1+x)(1+y)}{\epsilon^2 (1 - x^2) + 4\epsilon + 2(1 - y)}.\] The $y$ partial derivative term is computed to be \[\frac{\partial \lambda^{\rm Wach}_1}{\partial y} = \frac{4\epsilon(1-x^2) + 4\epsilon^2(1-x^2) + \epsilon^3(1 - x^2)^2}{(\epsilon^2 (1 - x^2) + 4\epsilon + 2(1 - y))^2}\] Define the subregion ${\Omega'_P}\subset{\Omega_\epsilon}$ by \[{\Omega'_P}=\left\{(x,y)\in{\Omega_\epsilon}:\frac 14\leq x\leq\frac 34,\quad 1\leq y\leq 1+\epsilon\right\}\] Observe that $\frac{7}{16}\leq 1 - x^2 \leq \frac{15}{16}$ on ${\Omega'_P}$. Fix $0<\epsilon<1$. We bound the numerator by \[ 4\epsilon(1-x^2) + 4\epsilon^2(1-x^2) + \epsilon^3(1 - x^2)^2 > 4\epsilon\cdot \frac{7}{16} + 4 \epsilon^2 \cdot \frac{7}{16} + \epsilon^3\cdot \frac{49}{256} > \frac{7}{4}\epsilon. \] Since $|y-1|<\epsilon$ on ${\Omega'_P}$, we can bound the denominator by \[ |\epsilon^2 (1 - x^2) + 4\epsilon + 2(1 - y)| \leq |\epsilon^2 (1 - x^2)| + |4\epsilon| + |2(1 - y)| \leq \epsilon^2 + 4\epsilon + 2\epsilon \leq 7\epsilon. \] Putting these results together, we have that \[\left|\frac{\partial \lambda^{\rm Wach}_1}{\partial y}\right|>\frac{\frac{7}{4}\epsilon}{49\epsilon^2} = \frac{1}{28\epsilon }>0.\] Let $C=\frac {1}{28}$ for ease of notation. Since $|{\Omega'_P}|>\frac 18 \epsilon$, \[\lim_{\epsilon\rightarrow 0}\int_{\Omega_\epsilon}\left|\frac{\partial \lambda^{\rm Wach}_1}{\partial y}\right|^2\geq \lim_{\epsilon\rightarrow 0}\int_{{\Omega'_P}}\left|\frac{\partial \lambda^{\rm Wach}_1}{\partial y}\right|^2>\lim_{\epsilon\rightarrow 0}\int_{{\Omega'_P}}\frac{C^2}{\epsilon^2}=C^2\lim_{\epsilon\rightarrow 0}\frac{|{\Omega'_P}|}{\epsilon^2}>\frac{C^2}{8}\lim_{\epsilon\rightarrow 0}\frac 1\epsilon=\infty,\] thereby proving the lemma. \qed \end{proof} \cancel{ The explicit formula for the Wachspress weight associated to $\textbf{v}_1$ is \[w^{\rm Wach}_1(\textbf{x})= B_1A_2A_3A_4 =\epsilon xy(1-x)\] where $\textbf{x}=(x,y)$ is an arbitrary point inside ${\Omega_\epsilon}$. The other weights can be computed similarly, yielding the coordinate function \[\lambda^{\rm Wach}_1= \frac{w_{i}(\textbf{v})}{\sum_{j=1}^5w_{j}(\textbf{v})} = \frac{4\epsilon x (x-1)}{\epsilon^2 x(x-3)(x-1)^2-2\epsilon(x-1)(1-3x+x^2y)+(y-1)(x^2y+3xy+x^2-x-2)}.\] Then we compute the $y$ partial derivative term and fix the following notation, \begin{eqnarray*} \frac{\partial \lambda^{\rm Wach}_1}{\partial y} &=& \frac {\epsilon^2(8x^4(x-1)-8x^3(x-1))+\epsilon(8x(x-1)+x^2(x-1)(2-3y))} {(\epsilon^2 x(x-3) (x-1)^2- 2 \epsilon (x-1)(1 - 3 x + x^2y) + (y-1) (-2 + x^2 (1 + y) + x (3y -1)))^2} \\ &=& \frac{\epsilon^2 f_1+\epsilon f_2}{(\epsilon^2f_3+\epsilon f_4 + (y-1)f_5)^2} \end{eqnarray*} where the $f_i$ are polynomials in $x$ and $y$ with the natural correspondence to the expression for the partial derivative. Define the subregion ${\Omega'_P}\subset{\Omega_\epsilon}$ by \[{\Omega'_P}=\left\{(x,y)\in{\Omega_\epsilon}:\frac14\leq x\leq\frac 34,\quad 1\leq y\leq 1+\epsilon\right\}\] Observe that $|f_1|$ is bounded above on ${\Omega'_P}$ by 8. Further, it is easily confirmed that $|f_2|>1$ on ${\Omega'_P}$ since the $8x(x-1)$ term dominates on this region and the range of $x$ values does not include 0 or 1. Thus for $\epsilon< 1/9$ we have $\epsilon|f_1|<8\epsilon<\frac 89 <|f_2|$, meaning we have the uniform bound \[|f_2|-\epsilon|f_1|\geq 1-8\epsilon>\frac 19 > 0.\] Now, since the $f_i$ are polynomials, there exists $M\in\R$ an upper bound for $|f_3|$, $|f_4|$ and $|f_5|$ on ${\Omega'_P}$. Since $|y-1|<\epsilon$ on ${\Omega'_P}$, we have the estimate \[|\epsilon^2f_3+\epsilon f_4 + (y-1)f_5|^2\leq |\epsilon^2f_3|^2+|\epsilon f_4|^2+|\epsilon f_5|^2\leq \epsilon^2(3M^2).\] Putting these results together, we have that \[\left|\frac{\partial \lambda^{\rm Wach}_1}{\partial y}\right|\geq \frac{\epsilon(|f_2|-\epsilon|f_1|)}{|\epsilon^2f_3+\epsilon f_4 + (y-1)f_5|^2}>\frac{\epsilon(1/9)}{\epsilon^2(3M^2)}= \frac{1}{\epsilon 27M^2}>0.\] Let $C=1/27M^2$ for ease of notation. Observe that the area $|{\Omega'_P}|>\frac 18 \epsilon$. Thus we have that \[\lim_{\epsilon\rightarrow 0}\int_{\Omega_\epsilon}\left|\frac{\partial \lambda^{\rm Wach}_1}{\partial y}\right|^2\geq \lim_{\epsilon\rightarrow 0}\int_{{\Omega'_P}}\left|\frac{\partial \lambda^{\rm Wach}_1}{\partial y}\right|^2>\lim_{\epsilon\rightarrow 0}\int_{{\Omega'_P}}\frac{C^2}{\epsilon^2}=C^2\lim_{\epsilon\rightarrow 0}\frac{|{\Omega'_P}|}{\epsilon^2}>\frac{C^2}{8}\lim_{\epsilon\rightarrow 0}\frac 1\epsilon=\infty,\] thereby proving the lemma. } \section{Additional references}
{ "timestamp": "2011-04-19T02:00:24", "yymm": "1010", "arxiv_id": "1010.5005", "language": "en", "url": "https://arxiv.org/abs/1010.5005", "abstract": "We prove the optimal convergence estimate for first order interpolants used in finite element methods based on three major approaches for generalizing barycentric interpolation functions to convex planar polygonal domains. The Wachspress approach explicitly constructs rational functions, the Sibson approach uses Voronoi diagrams on the vertices of the polygon to define the functions, and the Harmonic approach defines the functions as the solution of a PDE. We show that given certain conditions on the geometry of the polygon, each of these constructions can obtain the optimal convergence estimate. In particular, we show that the well-known maximum interior angle condition required for interpolants over triangles is still required for Wachspress functions but not for Sibson functions.", "subjects": "Numerical Analysis (math.NA)", "title": "Error Estimates for Generalized Barycentric Interpolation", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9875683476832691, "lm_q2_score": 0.8104789109591832, "lm_q1q2_score": 0.8004033189280959 }
https://arxiv.org/abs/2105.00247
Extension of tetration to real and complex heights
The continuous tetrational function ${^x}r=\tau(r,x)$, the unique solution of equation $\tau(r,x)=r^{\tau(r,x-1)}$ and its differential equation $\tau'(r,x) =q \tau(r,x) \tau'(r,x-1)$, is given explicitly as ${^x}r=\exp_{r}^{\lfloor x \rfloor+1}[\{x\}]_q$, where $x$ is a real variable called height, $r$ is a real constant called base, $\{x\}=x-\lfloor x \rfloor$ is the sawtooth function, $\lfloor x \rfloor$ is the floor function of $x$, and $[\{x\}]_q=(q^{\{x\}}-1)/(q-1)$ is a q-analog of $\{x\}$ with $q=\ln r$, respectively. Though ${^x}r$ is continuous at every point in the real $r-x$ plane, extensions to complex heights and bases have limited domains. The base $r$ can be extended to the complex plane if and only if $x\in \mathbb{Z}$. On the other hand, the height $x$ can be extended to the complex plane at $\Re(x)\notin \mathbb{Z}$. Therefore $r$ and $x$ in ${^x}r$ cannot be complex values simultaneously. Tetrational laws are derived based on the explicit formula of ${^x}r$.
\section{Introduction} Tetration, i.e., iterated exponentiation, is the fourth hyperoperation after addition, multiplication, and exponentiation \cite{Goodstein1947}. Tetration is defined as \[^{n}r\colonequals \underset{n}{\underbrace{r^{r^{\udots^{r}}}}},\] meaning that copies of $r$ are exponentiated \(n\) times in a right-to-left direction, or recursively defined as \[^{n}r\colonequals \begin{cases} 1 & (n=0)\\ r^{(^{n-1}r)} & (n \geq 1). \end{cases}\] Obviously \(^{n-1}r=\log_r (^{n}r)\). The parameter \(r\) and \(n\) are referred to as the base and height, respectively. In this paper, we use the notations to express tetration as \[^{n}r=\tau(r,n)=\exp_{r}^{n}1,\] and we write \(\tau(n)\) for simplicity if \(r\) is a constant.\vspace{0.2in} \noindent Tetration has been applied in fundamental physics \cite{Scot2006}, degree of connection in air transportation system \cite{Sun2017}, signal interpolation \cite{Khetkeeree2019}, data compression \cite{Furuya2019}, etc. \vspace{0.2in} \noindent Tetration is known as a rapid growing function that may be useful for expressing superexponential growth or huge numbers, but it also shows saturation and oscillation. \begin{figure} \includegraphics[scale=0.43]{fig_discrete.pdf} \centering \caption{Behaviors of discrete tetration for different \(r\) (left), and cobweb plots (right).} \label{fig:discrete tetration} \end{figure} As in Fig. \ref{fig:discrete tetration} the protean dynamics can be illustrated by cobweb plots. \vspace{0.2in} \noindent Euler \cite{Euler1783} proved that the limit of tetration \(^{n}r\), as \(n\) goes to infinity, converges if \(e^{(-e)}\leq r \leq e^{1/e}\). Eisenstein \cite{Eisenstein1844} gave the closed form of \(^{\infty}r\) for complex \(r\), and Corless et al. \cite{Corless1996} expressed it as \(^{\infty}r=W(-\ln r)/(-\ln r)\) by using Lambert's W function. \vspace{0.2in} \noindent Different from established extension of base \(r\) to real or complex values under integers or infinite height, extension of height \(n\) to real or complex values is still under debate. The object of this paper is to derive rigorously the explicit formula of continuous tetration. \vspace{0.2in} \begin{dfn} \label{dfn_tau} Let \(r>0\) be real constant and \(x>-2\) be real variable. Tetrational function \(\tau (x)\) is a continuous function defined as: \[\tau(0)\colonequals 1,\hspace{15pt}\tau(x)\colonequals r^{\tau(x-1)}.\] \end{dfn} \begin{cly} Independent of \(r\), every tetrational function \(\tau(x)\) goes through \(\tau(-1)=0\) and \(\tau(0)=1\). \end{cly} \begin{cly} \label{cly_tau} If a curve segment connecting between \(\tau(-1)=0\) and \(\tau(0)=1\) is defined, and is extended to other region by exponentiation or logarithm, then the whole curve satisfies Definition\ref{dfn_tau}. \end{cly} \vspace{0.2in} \noindent Because of Colloray \ref{cly_tau}, it is taken for granted that \(\tau(x)\) cannot be uniquely determined without an extra requirement. It follows that there exist many unique solutions for different extra conditions. Hooshmand \cite{Hooshmand2006} applied a linear segment between \(\tau(-1)=0\) and \(\tau(0)=1\), and proved uniqueness of the function \(\tau(x)=\exp_{r}^{\lfloor x \rfloor+1}\{x\}\), where \(\lfloor x \rfloor\) is the floor function of \(x\) and \(\{x\}=x-\lfloor x \rfloor\) is the sawtooth function. He found that the first derivative exists at the connecting points of \(x=n\in\mathbb{N}\) only for the case of \(r=e\), and the second derivative does not exist at the connecting points. Kouznetsov \cite{Kouznetsov2009} gave a numerical solution for \(r>e^{1/e}\) in the complex plane under the requirement that \(\tau(x+iy)\) is holomorphic, and Paulsen and Cowgill \cite{Paulsen2017} proved that such a holomorphic solution is unique.\vspace{0.2in} \noindent In this paper, we analytically derive an unique solution only from the minimum requirement of differentiability: the delay-differential equation in Lemma \ref{lem_delaydiff}. \begin{lem} \label{lem_delaydiff} Let \(q=\ln{r}\) be a real constant.\\ If a tetrational function is differentiable, the following delay-differential equation holds at every point in \(x>-2\): \[\frac{d\tau(x)}{dx}=q\tau(x)\frac{d\tau(x-1)}{dx}\] \end{lem} \vspace{0.2in} \begin{proof} The derivative of \(\tau(x)\) in Definition \ref{dfn_tau}. \end{proof}\vspace{0.2in} \noindent This paper consists of eight sections. In Section \ref{sec_multiple}, a new function, named multi-tetrational function, is defined to express the restriction placed in the delay-differential equation. In Section \ref{sec_Taylor}, the tetrational function and multi-tetrational function are then expressed in Taylor series with coefficients characteristic to these functions. In Section \ref{sec_trans}, based on the general forms of coefficients, the Taylor series are transformed into new expressions. In section \ref{sec_explicit} explicit formulae of tetrational functions and multi-tetrational functions, the main results of this paper, are derived. In Section \ref{sec_complex}, Analytical properties as well as the extension to the complex heights and bases are discussed. In Section \ref{sec_rule}, calculation rules of tetration are presented. Concluding remarks are given in section \ref{sec_summary}. \section{Multi-tetrational function} \label{sec_multiple} In this section we define a key function, and by using this, we transform the delay-differential function into an ordinary differential equation.\vspace{0.2in} \noindent Let us consider a continuous function \(\mu(x)\) which goes through the following discrete values: \[\mu(0)=\tau(0)=1\] \[\mu(1)=\tau(1)\tau(0)=r\] \[\mu(2)=\tau(2)\tau(1)\tau(0)=r^{r}r\] \[\vdots\] \[\mu(n)=\tau(n)\mu(n-1)\] It is reasonable to define \(\mu(x)\) by replacing \(n\) with \(x\).\vspace{0.2in} \noindent \begin{dfn} \label{dfn_mu} A continuous function \(\mu(x)\) is called a multi-tetrational function if it satisfies the following relations: \[\mu(0)\colonequals\tau(0),\hspace{15pt}\mu(x)\colonequals\tau(x)\mu(x-1)\] \end{dfn} \begin{lem} \label{lem_mu_value} The equation \(\mu(-1)=\mu(0)=1\) holds independent of \(r\). \end{lem} \begin{proof} From Definition \ref{dfn_tau}, we get \(\mu(0)=\tau(0)=1\). Assigning \(x=0\) to second equation in Definition\ref{dfn_mu} gives \(\mu(-1)=\mu(0)/\tau(0)=1\). \end{proof}\vspace{0.2in} \noindent The delay-differential equation in Lemma \ref{lem_delaydiff} is transformed in to an ordinary differential equation including \(\mu(x)\) as in Theorem \ref{thm_dif_tau_mu}, which places strong restriction on both \(\tau(x)\) and \(\mu(x)\) .\vspace{0.2in} \begin{thm} \label{thm_dif_tau_mu} Let \(q=\ln{r}\) be a real constant.\\ Let \(\omega\in\mathbb{C}\) be q-dependent constant. \\ The delay-defferentail equation \ref{lem_delaydiff} is equivalent to the differential equation: \[\frac{d\tau(x)}{dx}=\omega q^{x}\mu(x)\] \end{thm} \begin{proof} By assigning integer heights \(x=n\in\mathbb{N}\) to the equation \ref{lem_delaydiff}, we get: \[\tau'(0)=q\tau(0)\tau'(-1)\] \[\tau'(1)=q\tau(1)\tau'(0)=q^{2}\tau(1)\tau(0)\tau'(-1)\] \[\tau'(2)=q\tau(2)\tau'(1)=q^{3}\tau(2)\tau(1)\tau(0)\tau'(-1)\] \[\vdots\] \[\tau'(n)=q^{n+1}\mu(n)\tau'(-1).\] By defining \(\omega\colonequals{q}\tau'(-1)\), we get \[\tau'(n)={\omega}q^{n}\mu(n).\] By definitions \ref{dfn_tau} and \ref{dfn_mu}, the continuous functions \(\tau(x)\), \(\mu(x)\) and \(q^{x}\) go through all of the discrete points \(\tau(n)\), \(\mu(n)\) and \(q^{n}\), respectively. Therefore equation above is extended to real values: \[\tau'(x)={\omega}q^{x}\mu(x).\] Conversely, if \(\tau'(x)={\omega}q^{x}\mu(x)\) and \(\mu(x)=\tau(x)\mu(x-1)\) are given, then \[\frac{\tau'(x)}{\tau'(x-1)}=\frac{{\omega}q^{x}\mu(x)}{{\omega}q^{x-1}\mu(x-1)}=\frac{q\mu(x)}{\mu(x-1)}=q\tau(x).\] Therefore the delay-defferential equation \(\tau'(x)=q\tau(x)\tau'(x-1)\) is reproduced.\\ \end{proof} \section{Taylor series} \label{sec_Taylor} In this section, we will find general forms of the coefficients in Taylor series of the tetrational function and multi-tetrational function. \vspace{0.2in} \noindent First we express \(\tau(x)\) and \(\mu(x)\) as Taylor series. \begin{dfn} \label{dfn_taylor} Let \(a_n, b_n, c_n\) and \(d_n\) be real Taylor coefficients. Taylor series of tetrational and multi-tetrational functions near the points \(x=0\) or \(x=-1\) are expressed as: \[\tau(x)\colonequals\sum_{n=0}^{\infty}\frac{1}{n!}{a_n}{x^{n}},\hspace{15pt}\tau(x-1)\colonequals\sum_{n=0}^{\infty}\frac{1}{n!}{b_n}{x^{n}},\] \[\mu(x)\colonequals\sum_{n=0}^{\infty}\frac{1}{n!}{c_n}{x^{n}},\hspace{15pt}\mu(x-1)\colonequals\sum_{n=0}^{\infty}\frac{1}{n!}{d_n}{x^{n}},\] \end{dfn} \begin{cly} The following coefficients are given as follows independent of \(r\): \[a_0=1,\hspace{15pt}b_0=0,\hspace{15pt}c_0=1,\hspace{15pt}d_0=1\] \end{cly} \begin{proof} Corollary \ref{cly_tau} gives \(a_0=\tau(0)=1\) and \(b_0=\tau(-1)=0\). Lemma \ref{lem_mu_value} gives \(c_0=\mu(0)=1\), and \(d_0=\mu(-1)=1\). \end{proof}\vspace{0.2in} \noindent Next we find the correlations of coefficients \(a_n, b_n, c_n\) and \( d_n\). \begin{lem} \label{lem_coefficient_abcd} Let \(s=\ln{q}\) be the real or complex constant dependent on \(q\), where \(s\) is the principal value for \(q<0\).\\ Let \(\binom{n}{k}\) be the binomial coefficient.\\ Taylor coefficients \(a_n, b_n, c_n\),and \( d_n\) for \(n\geq1\) are related to each other in the following ways: \begin{eqnarray} a_n&=&\omega\sum_{k=1}^{n}\binom{n-1}{k-1}c_{k-1}s^{n-k},\\ qb_n&=&\omega\sum_{k=1}^{n}\binom{n-1}{k-1}d_{k-1}s^{n-k},\\ c_n&=&\sum_{k=0}^{n}\binom{n}{k}a_{n-k}d_{k}. \end{eqnarray} \end{lem} \begin{proof} Repeated differentiation of the equation in Theorem \ref {thm_dif_tau_mu} and assigning \(x=0\) and \(x=-1\) to it give relations (1) and (2). The relation (3) is given by assigning \(x=0\) to the equation in Definition \ref{dfn_mu} after repeated differentiation of it. \end{proof} \noindent From Lemma \ref{lem_delaydiff}, we can derive another relationship between \(a_{n}\) and \(b_{n}\) similar to Lemma \ref{lem_coefficient_abcd}. The relation, however, can be derived from equations in Lemma \ref{lem_coefficient_abcd}.\vspace{0.2in} \noindent Finally we show that general formulae of \(a_n, b_n, c_n\) and \( d_n\) have the common structure with the Striling numbers of the second kind \(\stirling{m}{n}\) and constants \(A_{n}\) or \(B_{n}\) e.g., \[a_{0}=A_{0},\] \[a_{1}=A_{0}{\omega},\] \[a_{2}=A_{0}s{\omega}+A_{1}{\omega}^{2},\] \[a_{3}=A_{0}s^{2}{\omega}+3A_{1}s{\omega}^{2}+A_{2}{\omega}^{3},\] \[a_{4}=A_{0}s^{3}{\omega}+7A_{1}s^{2}{\omega}^{2}+6A_{2}s{\omega}^{3}+A_{3}{\omega}^{4}.\] Generally, \[a_{0}=A_{0},\hspace{10pt}a_{n}=\sum_{k=1}^{n}\stirling{n}{k}A_{k}s^{n-k}{\omega}^{k}\hspace{10pt}(n\geq{1}).\] The proof is given in Theorem \ref{thm_coefficient_AB}. \begin{thm} \label{thm_coefficient_AB} Let \(A_{n}\) and \(B_{n}\) be a real constants having the relation: \[A_{0}\colonequals 1,\hspace{10pt}A_{1}\colonequals 1,\hspace{10pt}B_{0}\colonequals 0,\hspace{10pt}B_{1}\colonequals 1,\] \[A_{n}\colonequals\sum_{k=1}^{n}\binom{n}{k-1}A_{n-k}B_{k}.\hspace{15pt}(n\geq 2)\] Let \(\stirling{n}{m}\) be the Stirling numbers of the second kind.\\ Then the coefficients \(a_n, b_n, c_n\),and \( d_n\) for \(n\geq 1\) with the following formulae satisfy the relations given in Lemma \ref{lem_coefficient_abcd}: \[a_{n}=\sum_{k=1}^{n}\stirling{n}{k}A_{k}s^{n-k}\omega^{k},\hspace{15pt}qb_{n}=\sum_{k=1}^{n}\stirling{n}{k}B_{k}s^{n-k}\omega^{k},\] \[c_{n}=\sum_{k=1}^{n}\stirling{n}{k}A_{k+1}s^{n-k}\omega^{k},\hspace{10pt}d_{n}=\sum_{k=1}^{n}\stirling{n}{k}B_{k+1}s^{n-k}\omega^{k}.\] \end{thm} \begin{proof} We shall prove this by induction.\\ For \(n=1\), relations in Lemma \ref{lem_coefficient_abcd} (1), (2) and (3) are \(a_{1}=\omega\), \(qb_{1}=\omega\) and \(c_{1}=a_{1}d_{0}+a_{0}d_{1}\), respectively. Clearly, the formulae of \(a_{1}=A_{1}\omega=\omega\) and \(qb_{1}=B_{1}\omega=\omega\) satisfy the first and second relations above, respectively. The relation \ref{lem_coefficient_abcd} (3) also holds since the left hand side is \(c_{1}=A_{2}\omega\), and the right hand side is \(a_{1}d_{0}+a_{0}d_{1}=(A_{1}B_{1}+A_{0}B_{2})\omega=A_{2}\omega\).\vspace{0.1in} \noindent Then suppose the given formulae hold for \(n=k\). \\ By using \(c_{n}\) \((n\leq{k})\) and the following relationship \cite{Abramowitz1964}\cite{Graham1994}: \[\stirling{n+1}{k+1}=\sum_{j=k}^{n}\binom{n}{j}\stirling{j}{k},\] the relation of \ref{lem_coefficient_abcd} (1) for \(n=k+1\) is expressed as \[a_{k+1}=\omega\sum_{j=1}^{k+1}\binom{k}{j-1}c_{j-1}s^{k+1-j}\] \[={\omega}\left[\binom{k}{0}c_{0}s^{k}+\binom{k}{1}c_{1}s^{k-1}+\cdots+\binom{k}{k}c_{k}\right]\] \[=\binom{k}{0}\stirling{0}{0}A_{1}s^{k}{\omega}+\binom{k}{1}\stirling{1}{1}A_{2}{\omega}^{2}s^{k-1}+\cdots+\binom{k}{k}\left(\sum_{j=1}^{k}\stirling{k}{j}A_{j+1}s^{k-j}\omega^{k+1}\right)\] \[=\binom{k}{0}\stirling{0}{0}A_{1}s^{k}{\omega}+\left[\binom{k}{1}\stirling{1}{1}+\binom{k}{2}\stirling{2}{1}\right]A_{2}s^{k-1}{\omega}^{2}+\cdots+\binom{k}{k}\stirling{k}{k}A_{k+1}{\omega}^{k+1}\] \[=\stirling{k+1}{1}A_{1}s^{k}{\omega}+\stirling{k+1}{2}A_{2}s^{k-1}{\omega}^2+\cdots+\stirling{k+1}{k+1}A_{k+1}{\omega}^{k+1}\] Therefore, if the relation holds for \(n=k\), it holds for \(n=k+1\). It follows that the first relation is generally satisfied by the given formulae. \\ Similarly, proof for the relation of \ref{lem_coefficient_abcd} (2) is given by replacing \(a_{n}\) and \(c_{n}\) with \(qb_{n}\) and \(d_{n}\), respectively. Again, if the relation holds for \(n=k\), it holds for \(n=k+1\). \\ Therefore relation \ref{lem_coefficient_abcd} (1) and (2) are generally satisfied by given expressions. Next we check the given formulae of \(a_{n}\), \(c_{n}\) and \(d_{n}\) generally satisfy the relation of \ref{lem_coefficient_abcd} (3) at any \(k\). \[c_{k}=\sum_{j=0}^{k}\binom{k}{j}a_{k-j}d_{j}\] \[=\binom{k}{0}a_{k}d_{0}+\binom{k}{1}a_{k-1}d_{1}+\cdots+\binom{k}{k}a_{0}d_{k}\] By replacing each \(a_{n}\) and \(d_{n}\) by the given formulae, we have \[c_{k}=\binom{k}{0}\left[\stirling{k}{1}A_{1}s^{k-1}{\omega}+\cdots+\stirling{k}{k}A_{k}s^{0}{\omega}^{k}\right]B_{1}\] \[+\binom{k}{1}\left[\stirling{k-1}{1}A_{1}s^{k-2}{\omega}+\cdots+\stirling{k-1}{k-1}A_{k-1}s^{0}{\omega}^{k-1}\right]B_{2}\omega+\cdots\] \[+\binom{k}{k}A_{0}\left[\stirling{k}{1}B_{2}s^{k-1}{\omega}+\cdots+\stirling{k}{k}B_{k+1}s^{0}{\omega}^{k}\right].\] By rearrangement, we have \[c_{k}=\left[\binom{k}{0}\stirling{k}{1}A_{1}B_{1}+\binom{k}{k}\stirling{k}{1}A_{0}B_{2}\right]s^{k-1}\omega\] \[+\left[\binom{k}{0}\stirling{k}{2}A_{2}B_{1}+\sum_{h=0}^{k-2}\binom{k}{j+h}\stirling{k-j-h}{2-j}\stirling{j+h}{j}A_{1}B_{2}+\binom{k}{k}\stirling{k}{2}A_{0}B_{3}\right]s^{k-1}{\omega}^{2}\] \[+\cdots\] \[+\left[\binom{k}{0}\stirling{k}{k}A_{k}B_{1}{\omega}^{k}+\binom{k}{1}\stirling{k-1}{k-1}A_{k-1}B_{2}+\cdots+\binom{k}{k}\stirling{k}{k}A_{0}B_{k+1}\right]{\omega}^{k}.\] By using the following relation for \(k<m\) \cite{Abramowitz1964}\cite{Graham1994}: \[\stirling{n}{m}\binom{m}{j}=\sum_{k=0}^{n-m}\binom{n}{j+k}\stirling{n-j-k}{m-j}\stirling{j+k}{j},\] We have \[c_{k}=\stirling{k}{1}\left[\binom{1}{0}A_{1}B_{1}+\binom{1}{1}A_{0}B_{2}\right]s^{k}\omega\] \[+\stirling{k}{2}\left[\binom{2}{0}A_{2}B_{1}+\binom{2}{1}A_{1}B_{2}+\binom{2}{2}A_{0}B_{3}\right]s^{k-1}{\omega}^{2}\] \[+\cdots\] \[+\stirling{k}{k}\left[\binom{k-1}{0}A_{k-1}B_{1}+\binom{k-1}{1}A_{k-2}B_{2}+\cdots+\binom{k-1}{k-1}A_{0}B_{k-1}\right]\] By using the relationship \(A_{n}=\sum_{k=1}^{n}\binom{n}{k-1}A_{n-k}B_{k}\), finally we get \[c_{k}=\stirling{k}{1}A_{2}s^{k}{\omega}+\stirling{k}{2}A_{3}s^{k-1}{\omega}^{2}+\cdots+\stirling{k}{k-1}A_{k}s{\omega}^{k-1}+\stirling{k}{k}A_{k+1}{\omega}^{k},\] which is consistent with the expression of \(c_{k}\) given above.\\ Therefore the relation of \ref{lem_coefficient_abcd} (3) is generally satisfied by the given formulae.\\ \end{proof}\vspace{0.2in} \noindent Now we have the general form of Taylor series of the tetrational and muti-tetrational functions near the points \(x=0\) or \(x=-1\): \[\tau(x)=A_{0}+A_{1}{\omega}x+\frac{1}{2}(A_{1}s{\omega}+A_{2}{\omega}^{2})x^{2}+\frac{1}{3!}(A_{1}s^{2}{\omega}+3A_{2}s{\omega}^2+A_{3}{\omega}^{3})x^{3}+\cdots,\] \[\mu(x)=A_{1}+A_{2}{\omega}x+\frac{1}{2}(A_{2}s{\omega}+A_{3}{\omega}^{2})x^{2}+\frac{1}{3!}(A_{2}s^{2}{\omega}+3A_{3}s{\omega}^2+A_{4}{\omega}^{3})x^{3}+\cdots,\] \[\tau(x-1)=\frac{B_{0}}{q}+\frac{B_{1}{\omega}}{q}x+\frac{1}{2}\frac{B_{1}s{\omega}+B_{2}{\omega}^{2}}{q}x^{2}+\frac{1}{3!}\frac{B_{1}s^{2}{\omega}+3B_{2}s{\omega}^2+B_{3}{\omega}^{3}}{q}x^{3}+\cdots,\] \[\mu(x-1)=B_{1}+B_{2}{\omega}x+\frac{1}{2}(B_{2}s{\omega}+B_{3}{\omega}^{2})x^{2}+\frac{1}{3!}(B_{2}s^{2}{\omega}+3B_{3}s{\omega}^2+B_{4}{\omega}^{3})x^{3}+\cdots.\] \section{Tetrational series} \label{sec_trans} In this section, the Taylor series of the tetrational functions and multi-tetrational functions are transformed into new expressions based on the property of Stirling numbers of the second kind.\vspace{0.2in} \noindent First we define a key function to express new series. \begin{dfn} \label{dfn_phi} Let \(\varphi(x)\) be a function defined as \[\varphi(x)\colonequals\frac{{\omega}(q^{x}-1)}{s}.\] \end{dfn} \begin{cly} \label{cly_diff_Phi} The following equation holds: \[\frac{d\varphi(x)}{dx}={\omega}q^{x}.\] \end{cly}\vspace{0.2in} \noindent By using \(\varphi(x)\), the Taylor series are then simplified as in Theorem \ref{thm_tetrational_series}. \begin{thm} \label{thm_tetrational_series} Let Taylor series of tetrational functions and mutiple tetrational functions near the points \(x=0\) or \(x=-1\) be absolutely convergent, then they are expressed in power series of \(\varphi(x)\), named tetrational series: \[\tau(x)=\sum_{n=0}^{\infty}\frac{1}{n!}A_{n}\varphi^{n}(x),\hspace{15pt}\tau(x-1)=\frac{1}{q}\sum_{n=0}^{\infty}\frac{1}{n!}B_{n}\varphi^{n}(x),\] \[\mu(x)=\sum_{n=0}^{\infty}\frac{1}{n!}A_{n+1}\varphi^{n}(x),\hspace{15pt}\mu(x-1)=\sum_{n=0}^{\infty}\frac{1}{n!}B_{n+1}\varphi^{n}(x).\] \end{thm} \begin{proof} Since the Taylor series of tetrational function is absolutely convergent, it can be rearranged as follows: \[\tau(x)=A_{0}+A_{1}{\omega}x+\frac{1}{2}(A_{1}s{\omega}+A_{2}{\omega}^{2})x^{2}+\frac{1}{3!}(A_{1}s^{2}{\omega}+3A_{2}s{\omega}^2+A_{3}{\omega}^{3})x^{3}+\cdots\] \[=A_{0}+A_{1}{\omega}\left(x+\frac{1}{2}sx^{2}+\frac{1}{3!}s^{2}x^{3}+\cdots\right)+A_{2}{\omega}^{2}\left(\frac{1}{2}x^2+\frac{1}{3!}3sx^{3}+\cdots\right)+\cdots\] \[=A_{0}+A_{1}\frac{\omega}{s}\left(sx+\frac{1}{2}s^{2}x^{2}+\frac{1}{3!}s^{2}x^{3}+\cdots\right)+A_{2}\frac{\omega^{2}}{s^{2}}\left(\frac{1}{2}s^{2}x^2+\frac{1}{3!}3s^{3}x^{3}+\cdots\right)+\cdots.\] By using the exponential generating function of the Stiring numbers of the second kind \cite{Abramowitz1964}\cite{Graham1994}: \[\frac{1}{k!}(e^{x}-1)^{k}=\sum_{n=k}^{\infty}\stirling{n}{k}\frac{x^{n}}{n!},\] we have \[\tau(x)=A_{0}+A_{1}\left[\frac{{\omega}(e^{sx}-1)}{s}\right]+\frac{1}{2}A_{2}\left[\frac{{\omega}(e^{sx}-1)}{s}\right]^{2}+\frac{1}{3!}A_{3}\left[\frac{{\omega}(e^{sx}-1)}{s}\right]^{3}+\cdots.\] Since \(s=\ln{q}\), we can replace \(e^{sx}\) with \(q^{x}\). The Taylor series are expressed as the power series of \(\varphi(x).\) Similarly, we can transform \(\tau(x-1)\), \(\mu(x)\) and \(\mu(x-1)\) into the tetrational series: \begin{align*} \tau(x)&=A_{0}+A_{1}\varphi(x)+\frac{1}{2}A_{2}\varphi^{2}(x)+\frac{1}{3!}A_{3}\varphi^{3}(x)+\cdots,\\ \mu(x)&=A_{1}+A_{2}\varphi(x)+\frac{1}{2}A_{3}\varphi^{2}(x)+\frac{1}{3!}A_{4}\varphi^{3}(x)+\cdots,\\ \tau(x-1)&=\frac{B_{0}}{q}+\frac{B_{1}}{q}\varphi(x)+\frac{1}{2}\frac{B_{2}}{q}\varphi^{2}(x)+\frac{1}{3!}\frac{B_{3}}{q}\varphi^{3}(x)+\cdots,\\ \mu(x-1)&=B_{1}+B_{2}\varphi(x)+\frac{1}{2}B_{3}\varphi^{2}(x)+\frac{1}{3!}B_{4}\varphi^{3}(x)+\cdots. \end{align*} \end{proof}\vspace{0.2in} \begin{cly} \label{cly_dtau_dphi_mu} The following equations hold near \(x=0\): \[\frac{\partial\tau(x)}{\partial\varphi}=\mu(x),\hspace{10pt}\frac{\partial\tau(x-1)}{\partial\varphi}=q^{-1}\mu(x-1).\] \end{cly}\vspace{0.2in} \noindent We check the consistency of the formulae as shown in Lemma \ref{lem_consistency}. \begin{lem} \label{lem_consistency} Formulae \ref{thm_tetrational_series} satisfy the differential equation \ref{thm_dif_tau_mu}: \[\frac{d\tau(x)}{dx}={\omega}q^{x}\mu(x).\] \end{lem} \begin{proof} Since \ref{cly_diff_Phi} and \ref{cly_dtau_dphi_mu}, with chain rule, \[\frac{d\tau(x)}{dx}=\frac{\partial\tau(x)}{\partial\varphi}\frac{\partial\varphi(x)}{\partial{x}}={\omega}q^{x}\mu(x),\] \[\frac{d\tau(x-1)}{dx}=\frac{\partial\tau(x-1)}{\partial\varphi}\frac{\partial\varphi(x)}{\partial{x}}={\omega}q^{x-1}\mu(x-1).\] \end{proof}\vspace{0.2in} \noindent We further describe the general case of tetrational series with an integer shift \(n\) as in Lemma \ref{lem_integer_shift}. The formulae above are special case of \(n=0\). \begin{lem} \label{lem_integer_shift} Let \(0\leq{x}<1\) be the real heights.\\ Let \(n\) and \(m\) be integers and natural numbers, respectively.\\ Let \(A_{m}^{[n]}\) and \(B_{m}^{[n]}\) be n-dependent real constants.\\ The tetrational and multi-tetrational functions, having horizontal translation by integers \(n\) or \(n-1\), are expressed as: \begin{align*} \tau(x+n)&=q^{n}\sum_{k=0}^{\infty}\frac{x^{k}}{k!}\sum_{j=1}^k\stirling{k}{j}A_{j}^{[n]}s^{k-j}{\omega}^{j}=q^{n}\sum_{k=0}^{\infty}\frac{1}{k!}A_{k}^{[n]}{\varphi}^{k}(x),\\ \mu(x+n)&=\sum_{k=0}^{\infty}\frac{x^{k}}{k!}\sum_{j=1}^k\stirling{k}{j}A_{j+1}^{[n]}s^{k-j}{\omega}^{j}=\sum_{k=0}^{\infty}\frac{1}{k!}A_{k+1}^{[n]}{\varphi}^{k}(x),\\ \tau(x+n-1)&=q^{n-1}\sum_{k=0}^{\infty}\frac{x^{k}}{k!}\sum_{j=1}^k\stirling{k}{j}B_{j}^{[n]}s^{k-j}{\omega}^{j}=q^{n-1}\sum_{k=0}^{\infty}\frac{1}{k!}B_{k}^{[n]}{\varphi}^{k}(x),\\ \mu(x+n-1)&=\sum_{k=0}^{\infty}\frac{x^{k}}{k!}\sum_{j=1}^k\stirling{k}{j}B_{j+1}^{[n]}s^{k-j}{\omega}^{j}=\sum_{k=0}^{\infty}\frac{1}{k!}B_{k+1}^{[n]}{\varphi}^{k}(x). \end{align*} Coefficients have the relationships of: \[A_{0}^{[n]}=\tau(n),\hspace{15pt}B_{0}^{[n]}=\tau(n-1),\] \[A_m^{[n]}=q^{n}\sum_{k=1}^{m}\binom{m-1}{k-1}A_{m-k}^{[n]}B_{k}^{[n]}\hspace{10pt}(m\geq{1}).\] Generally, \(\tau(x)\) and \(\mu(x)\) having horizontal translation by \(n\) have the relationship of: \[\frac{\partial \tau(x+n)}{\partial\varphi}=q^{n}\mu(x+n).\] \end{lem} \begin{proof} Obviously the tetrational series satisfy \(\partial\tau(x+n)/\partial\varphi=q^{n}\mu(x+n)\). It follows that these series satisfy the differential equation \(\tau'(x+n)={\omega}q^{x+n}\mu(x+n)\), as similarly in \ref{lem_consistency}. Tetrational series and the Taylor series are converted into each other with the exponential generating function of the Stirling numbers of the second kind as in Theorem \ref{thm_tetrational_series}.\\ \noindent The relationships of coefficients are given by assigning \(x=0\) to the repeated partial derivatives by \(\varphi\) of \(\mu(x+n)=\tau(x+n)\mu(x+n-1)\). \end{proof}\vspace{0.2in} \section{Explicit formulae} \label{sec_explicit} In this section, we derive the explicit formulae of tetrational and muti-tetrational function by determining the coefficients \(A_{n}\) and \(B_{n}\).\vspace{0.2in} \begin{lem} \label{lem_mu_piecewise} Let \(0\leq{x}<1\). The following equations hold: \[\mu(x-1)=1,\] \[\mu(x)=\tau(x),\] \[\mu(x+n)=\prod_{k=0}^{n}\tau(x+k).\] \end{lem} \begin{proof} We defined \(\mu(x)\) in Definition \ref{dfn_mu} as a continuous function satisfying \[\mu(x)=\tau(x)\mu(x-1),\] which goes through the following discrete points: \[\mu(0)=\tau(0)\] \[\mu(1)=\tau(1)\tau(0)\] \[\mu(2)=\tau(2)\tau(1)\tau(0)\] \[\vdots\] \[\mu(n)=\prod_{k=0}^{n}\tau(k)=\tau(n)\mu(n-1).\] Similarly, \(\mu(x)\) also goes through the discrete points with changes of \(0\leq{x}<1\) as follows, since continuous functions \(\mu(x)\) and \(\tau(x)\) go through \(\mu(n)\) and \(\tau(n)\) respectively: \[\mu(x)=\tau(x)\] \[\mu(x+1)=\tau(x+1)\tau(x)\] \[\mu(x+2)=\tau(x+2)\tau(x+1)\tau(x)\] \[\vdots\] \[\mu(x+n)=\prod_{k=0}^{n}\tau(x+k)=\tau(x+n)\mu(x+n-1)\] The relationship \(\mu(x+n)=\tau(x+n)\mu(x+n-1)\) is consistent with the definition \ref{dfn_mu}.\\ \noindent Since from the definition \(\mu(x-1)=\mu(x)/\tau(x)\) and \(\mu(x)=\tau(x)\) as given above, we have \[\mu(x-1)=\frac{\tau(x)}{\tau(x)}=1.\] \end{proof} \begin{cly} Tetrational function \(\tau(x)\) and \(\mu(x)\) are piecewise connected functions. \end{cly} \begin{proof} Obviously from Lemma \ref{lem_mu_piecewise}, multi-tetrational function \(\mu(x)\) is a piecewise connected function. Since tetrational function have the relationship of \(\tau'(x)=\omega{q}^x\mu(x)\), tetrational function \(\tau(x)\) is also a piecewise connected function. \end{proof} \begin{lem} \label{lem_AB_value} The coefficients \(A_{n}\) and \(B_{n}\) are determined as \[A_{n}=1,\] \[B_{n\neq{1}}=0,\hspace{15pt}B_{1}=1.\] \end{lem} \begin{proof} From Theorem \ref{thm_tetrational_series}, \[\tau(x)=\sum_{n=0}^{\infty}\frac{1}{n!}A_{n}\varphi^{n}(x),\hspace{10pt}\mu(x)=\sum_{n=0}^{\infty}\frac{1}{n!}A_{n+1}\varphi^{n}(x).\] Since \(\partial{\tau(x)}/\partial{\varphi}=\mu(x)\) from \ref{cly_dtau_dphi_mu} and \(\tau(x)=\mu(x)\) from \ref{lem_mu_piecewise}, we have \[1=A_{0}=A_{1}=A_{2}=\cdots.\] From the relationship of coefficients \(A_{n}=\sum_{k=1}^{n}\binom{n}{k-1}A_{n-k}B_{k}\), we get \(B_{n\neq{1}}=0\) and \(B_{1}=1\).\\ \end{proof} \noindent Similarly, we can get the values of \(A_{m}^{[n]}\) and \(B_{m}^{[n]}\) given in Lemma \ref{lem_integer_shift}.\vspace{0.2in} \begin{lem} \label{lem_omega_value} The constant \(\omega\) is determined as \[\omega=\frac{sq}{q-1}.\] \end{lem} \begin{proof} Since we determined the values of \(B_{n}\) as in Lemma \ref{lem_AB_value}, and with the expression of \(\varphi(x)\) in Definition \ref{dfn_phi}, we can express \(\tau(x-1)\) for \(0\leq{x}<1\) as \[\tau(x-1)=\frac{1}{q}\varphi(x)=\frac{\omega(q^{x}-1)}{sq}\] Then by assigning \(x=1\), we get the equation \(\omega(q-1)/sq=1\) to derive \(\omega\). \end{proof}\vspace{0.2in} \noindent Now in Theorem \ref{thm_explicit} we can derive the explicit solution for the delay-differential equation \ref{lem_delaydiff}. The domains of the tetrational function as well as the multi-tetrational function are naturally extended by applying the sawtooth function. \begin{thm} \label{thm_explicit} Let \(r\) be real constant.\\ Let \(x\) be real variable heights in \(x\in\mathbb{R}\setminus\{n:n\in\mathbb{Z}, n\leq{-2}\}.\)\\ Let \(\lfloor{x}\rfloor\) and \(\{x\}\colonequals{x}-\lfloor{x}\rfloor\) be the floor function and the sawtooth function of \(x\), respectively.\\ Let \([x]_q\) be the q-analog of \(x\) defined as \[[x]_{q}\colonequals\frac{q^{x}-1}{q-1},\] where \(q=\ln{r}\) is the principal branch if \(r<0\).\\ Let \(\exp_{r}^{n}=\log_{r}^{-n}\) be the extended iterated exponential operator defined as \[\exp_{r}^{n}{f}\colonequals \begin{cases} \underset{n}{\underbrace{\exp_r\exp_r\cdots\exp_r}} {f} & \hspace{15pt}(n\geq{0})\\ \hspace{3pt}\underset{n}{\underbrace{\log_r\log_r\cdots\log_r}} {f} & \hspace{15pt}(n<{0}), \end{cases} \] where \(f\) is the operand.\\ Let \(\prod_{k=m}^{n}\) be the extended product operator defined as \[\prod_{k=m}^{n}f_{k}\colonequals \begin{dcases} f_{m}f_{m+1}{\cdots}f_{n-1}f_{n} & (0\leq{m}\leq{n})\\ \frac{1}{f_{m+1}f_{m+2}{\cdots}f_{n+1}} & (n\leq{m}<0), \end{dcases} \] where \(f_{k}\) are the operands.\\ Then the tetrational function and the multi-tetratinal functions are expressed as \[{^x}r=\tau(x)=\exp_{r}^{\lfloor{x}\rfloor+1}[\{x\}]_q\] \[\mu(x)=\prod_{k=0}^{\lfloor{x}\rfloor}\exp_{r}^{k+1}[\{x\}]_q=q^{-\lfloor{x}\rfloor-1}\frac{\partial\tau(x)}{\partial[\{x\}]_q}\] \end{thm} \begin{proof} Let us consider the case of \(0\leq x<1\).\\ From Definition \ref{dfn_phi} and Lemma \ref{lem_omega_value}, we have \[\varphi(x)=\frac{q(q^{x}-1)}{q-1}.\] By using the notation of q-analog of the variable, we can express it as \[\varphi(x)=q[x]_q.\] Then the tetrational and multi-tetrational functions for \(n\geq{0}\) are given as follows since \(\tau(x)=\sum_{n=0}^{\infty}{(q[x]_q)}^{n}/n!=\exp_{r}[x]_{q}\) in Theorem \ref{thm_tetrational_series} and \(\tau(x+n+1)=\exp_{r}{\tau(x+n)}\) as well as \(\mu(x+n)=\tau(x+n)\mu(x+n-1)\): \begin{align*} \tau(x)=\exp_{r}[x]_q, & \hspace{15pt}\mu(x)=\exp_{r}[x]_q\\ \tau(x+1)=\exp_{r}^{2}[x]_q, & \hspace{15pt}\mu(x+1)=\exp_{r}^{2}[x]_q\exp_{r}[x]_q\\ \tau(x+2)=\exp_{r}^{3}[x]_q, & \hspace{15pt}\mu(x+2)=\exp_{r}^{3}[x]_q\exp_{r}^{2}[x]_q\exp_{r}[x]_q\\ & \hspace{5pt}\vdots\\ \tau(x+n)=\exp_{r}^{n+1}[x]_q, & \hspace{15pt}\mu(x+n)=\prod_{k=0}^{n}\exp_{r}^{k+1}[x]_q \end{align*} In the case for \(n<0\). The relation \(\tau(x-1)=[x]_q\) and \(\tau(x-n-1)=\log_{r}\tau(x-n)\), as well as \(\mu(x-1)=1\) and \(\mu(x-n)=\mu(x-n+1)/\tau(x-n+1)\), give the following series: \begin{align*} \tau(x-1)=[x]_q, & \hspace{15pt}\mu(x-1)=\frac{\tau(x)}{\tau(x)}=1,\\ \tau(x-2)=\log_{r}[x]_q, & \hspace{15pt}\mu(x-2)=\frac{\tau(x)}{\tau(x)\tau(x-1)}=\frac{1}{[x]_q},\\ \tau(x-3)=\log_{r}^{2}[x]_q, & \hspace{15pt}\mu(x-3)=\frac{\tau(x)}{\tau(x)\tau(x-1)\tau(x-2)}=\frac{1}{[x]_q\log_{r}[x]_q},\\ & \hspace{5pt}\vdots\\ \tau(x-n)=\log_{r}^{n+1}[x]_q, & \hspace{15pt}\mu(x-n)=\frac{\prod_{k=0}^{0}\tau(x+k)}{\prod_{k=1}^{n}\tau(x-k+1)}=\prod_{k=0}^{-n}\tau(x+k). \end{align*} In the last line, we applied the extended operator.\\ Therefore we can also give the same expression for \(n<0\) as \[\tau(x+n)=\exp_{r}^{n+1}[x]_q,\hspace{15pt}\mu(x+n)=\prod_{k=0}^{n}\exp_{r}^{k+1}[x]_q.\] Obviously, the following equation holds for both cases of \(n\geq{0}\) and \(n<0\): \[\mu(x+n)=q^{-n-1}\frac{\partial\tau(x+n)}{\partial[x]_q}.\] By replacing \(x\) and \(n\) above with \(\{x\}\) and \(\lfloor{x}\rfloor\), respectively, we have the expressions of tetrational and multi-tetrational functions for \(x\in(-\infty, \infty)\): \[\tau(x)=\exp_{r}^{\lfloor{x}\rfloor+1}[\{x\}]_q,\hspace{10pt}\mu(x)=\prod_{k=0}^{\lfloor{x}\rfloor}\exp_{r}^{k+1}[\{x\}]_q=q^{-\lfloor{x}\rfloor-1}\frac{\partial\tau(x)}{\partial[\{x\}]_q}.\] \end{proof} \begin{cly} \label{cly_core} The following equation holds: \[ ^{\{x\}}r=r^{[\{x\}]_q}. \] \end{cly} \begin{proof} Since \(0\leq{\{x\}}<1\) and \(\lfloor\{x\}\rfloor=0\), the equation is given from Theorem \ref{thm_explicit}. \end{proof} \noindent Let us consider the special case of \(r=e\). If \(r\rightarrow{e}\), hence \(q=\ln{r}\rightarrow{1}\), then q-analog becomes \([\{x\}]_q\rightarrow{\{x\}}\). So we have \(\tau(x)=\exp_{r}^{\lfloor{x}\rfloor+1}\{x\}.\) This formula is the same as Hooshmand's ultra exponential function for \(r=e\) \cite{Hooshmand2006}.\vspace{0.2in} \noindent The curves of continuous tetrational functions \({^x}r=\tau(x)\) for different bases \(r\) are shown in Fig. \ref{fig:continuous tetration}. It is obvious that the tetrational functions for \(r<1\) have complex values. \begin{figure} \includegraphics[scale=0.4]{fig_continuous.pdf} \centering \caption{Behaviours of the continuous tetrational functions \({^x}r\) for different bases \(r\), where only the real component are expressed for \(r=1/e\) and \(r=e^{-e}\) (left). The tetrational functions of \(r=1/e\) and \(r=e^{-e}\) are expressed in the complex plane (right).} \label{fig:continuous tetration} \end{figure} \vspace{0.2in} \noindent Figure \ref{fig:multiple} shows the behaviours of the mutiple tetrational functions \(\mu(x)\) for different bases \(r\). In general the functions are not differentiable at \(x=n\in\mathbb{Z}\) \begin{figure} \includegraphics[scale=0.4]{fig_multiple.pdf} \centering \caption{Behaviors of the multi-tetrational functions \(\mu(x)\) for different bases \(r\), where only the real component are expressed for \(r=1/e\) and \(r=e^{-e}\) (left). The multi-tetrational functions of \(r=1/e\) and \(r=e^{-e}\) are expressed in the complex plane (right).} \label{fig:multiple} \end{figure} \vspace{0.2in} \section{Analytical properties} \label{sec_complex} First, let us study the properties of \({^y}x=\tau(x,y)\) as the multivariable function of base \(x\) and height \(y\). From this point, we shall freely choose the letters to express the bases and integers. \begin{thm} \label{thm_total_differential} Let \(x>0\) be real bases and let \(q=\ln{x}\).\\ Let \(y\) be real heights in \(y\in\mathbb{R}\setminus\{n:n\in\mathbb{Z}, n\leq{-2}\}.\)\\ Then the tetrational function \({^{y}}x=\tau(x,y)=\exp_{x}^{\lfloor{y}\rfloor+1}[\{y\}]_{q}\) is totally differentiable: \[d\tau(x,y)=\frac{\partial\tau(x,y)}{\partial{x}}dx+\frac{\partial\tau(x,y)}{\partial{y}}dy.\] \end{thm} \begin{proof} We shall prove this by induction. Let \(-1\leq{y}<0\), then the tetrational function is \[\tau(x,y)=\frac{q^{y+1}-1}{q-1}=\frac{e^{(\ln{\ln{x}})(y+1)}-1}{\ln{x}-1}.\] Then the partial derivatives for \(-1\leq{y}<0\) are: \[\frac{\partial\tau(x,y)}{\partial{x}}=\frac{ye^{(\ln{\ln{x}})(y+1)}}{x\ln{x}(\ln{x}-1)}-\frac{e^{(\ln{\ln{x}})(y+1)}-1}{x(\ln{x}-1)^{2}}\] \[\frac{\partial\tau(x,y)}{\partial{y}}=\frac{(\ln{\ln{x}})e^{(\ln{\ln{x}})(y+1)}}{\ln{x}-1}\] The following equation holds for \(-1\leq{y}<0\). \[\frac{\partial^{2}\tau(x,y)}{\partial{x}\partial{y}}=\frac{\partial^{2}\tau(x,y)}{\partial{y}\partial{x}}=\frac{(1+(y+1)\ln{\ln{x}})e^{(\ln{\ln{x}})(y+1)}}{x\ln{x}(\ln{x}-1)}-\frac{(\ln{\ln{x}})e^{(\ln{\ln{x}})(y+1)}}{x(\ln{x}-1)^{2}}\] Therefore \[d\tau(x,y)=\frac{\partial\tau(x,y)}{\partial{x}}dx+\frac{\partial\tau(x,y)}{\partial{y}}dy.\] Next, let us consider the general case.\\ From the Definition \ref{dfn_tau}: \[{^{y+1}}x=\tau(x,y+1)=e^{(\ln{x})\tau(x,y)},\] the partial derivatives of this general equation are \[\frac{\partial\tau(x,y+1)}{\partial{x}}=\tau(x,y+1)\left[\frac{\tau(x,y)}{x}+(\ln{x})\frac{\partial\tau(x,y)}{\partial{x}}\right],\] \[\frac{\partial\tau(x,y+1)}{\partial{y}}=(\ln{x})\tau(x,y+1)\frac{\partial\tau(x,y)}{\partial{y}}.\] \noindent Let \(\tau(x,y)\) be defined in \(n\leq{y}<n+1\), and \(\tau(x,y+1)\) be defined in \(n+1\leq{y+1}<n+2\).\\ \noindent Suppose the following equation holds for \(n\leq{y}<n+1\): \[\frac{\partial^{2}\tau(x,y)}{\partial{x}\partial{y}}=\frac{\partial^{2}\tau(x,y)}{\partial{y}\partial{x}}\] Since \(\tau(x,y)\) in Theorem \ref{thm_explicit} is determined so as to be continuous at integer heights \(y\), and differentiation along base \(x\) is not affected by the existence of boundary, obviously the connection at \(y=n\) is continuous as follows: \[\underset{y\rightarrow{n+1}}\lim\frac{\partial\tau(x,y)}{\partial{x}}=\underset{y\rightarrow{n}}\lim\frac{\partial\tau(x,y+1)}{\partial{x}}=\frac{\partial\tau(x,n+1)}{\partial{x}}\] \[\underset{y\rightarrow{n+1}}\lim\frac{\partial\tau(x,y)}{\partial{y}}=\underset{y\rightarrow{n}}\lim\frac{\partial\tau(x,y+1)}{\partial{y}}=\frac{\partial\tau(x,n+1)}{\partial{y}}\] \noindent The following equation holds: \[\frac{\partial^{2}\tau(x,y+1)}{\partial{x}\partial{y}}=\frac{\partial^{2}\tau(x,y+1)}{\partial{y}\partial{x}}\] \[=(\ln{x})\tau(x,y+1)\frac{\partial\tau(x,y)}{\partial{y}}\left[\frac{\tau(x,y)}{x}+(\ln{x})\frac{\partial\tau(x,y)}{\partial{x}}\right]\] \[+\tau(x,y+1)\left[\frac{1}{x}\frac{\partial\tau(x,y)}{\partial{y}}+(\ln{x})\frac{\partial^{2}\tau(x,y)}{\partial{y}\partial{x}}\right]\] Hence: \[d\tau(x,y+1)=\frac{\partial\tau(x,y+1)}{\partial{x}}dx+\frac{\partial\tau(x,y+1)}{\partial{y}}dy.\] In general, therefore, tetrational function is totally differentiable.\\ \end{proof}\vspace{0.2in} \noindent The property above allow us to define the inverse operations of tetration, super-logarithm or super-root, in monotonic changing range: base \(r\geq{1}\) and height \(h>-2\).\vspace{0.2in} \noindent Next we study the behaviours of the tetrational function \({^{h}x}=\tau(x,h)\) with a variable base \(x\) and a constant height \(h\).\vspace{0.2in} \begin{lem} \label{lem_(^n)x:x=0} let \(n\) be the integer constant.\\ Let \(x>0\) be the real variable. \[ \underset{x\rightarrow{0}}\lim(^{n}x)= \begin{cases} 0 & (n: \mathrm{even})\\ 1 & (n: \mathrm{odd}) \end{cases} \] \end{lem} \begin{proof} We shall prove this by induction. From Definition \ref{dfn_tau}, the relations \(^{-1}x=0\) and \(^{0}x=1\) hold for \(x\rightarrow{0}\).\\ Obviously we have \(\underset{x\rightarrow{0}}{\lim}(\ln{x})=-\infty\).\\ If \(\underset{x\rightarrow{0}}{\lim}({^{2n}}x)=1\), then \[\underset{x\rightarrow{0}}{\lim}[(^{2n}x)\ln{x}]=-\infty.\] It follows \[\underset{x\rightarrow{0}}{\lim}({^{2n+1}}x)=\underset{x\rightarrow{0}}{\lim}\left[e^{(^{2n}x)\ln{x}}\right]=0,\] If \(\underset{x\rightarrow{0}}{\lim}({^{2n+1}}x)=0\), then \[\underset{x\rightarrow{0}}{\lim}[(^{2n+1}x)\ln{x}]=\underset{x\rightarrow{0}}{\lim}[\ln[x^{(^{2n+1}x)}]]=\underset{x\rightarrow{0}}{\lim}[\ln{(^{2n}x)}]=0.\] It follows \[\underset{x\rightarrow{0}}{\lim}({^{2n+2}}x)=\underset{x\rightarrow{0}}{\lim}\left[e^{(^{2n+1}x)\ln{x}}\right]=1.\] Therefore the relation generally holds.\\ \end{proof}\vspace{0.2in} \noindent The relation in Lemma \ref{lem_(^n)x:x=0} can be confirmed by the curves shown in Figure \ref{fig:(^n)x_real} (left). The tetrational functions \({^n}x\) for integers \(n\) have real values and go to 0 or 1 for even \(n\) or odd \(n\) respectively under \(x\rightarrow{0}\). As shown in the right graph of Figure \ref{fig:(^n)x_real}, in the general case including non-integers \(h\), the functions \({^h}x\) have complex values and non-monotonic behaviours for \(x<1\) while the functions have real values and monotonic changes for \(x\geq{1}\). \begin{figure} [ht] \includegraphics[scale=0.35]{fig_n_x_real.pdf} \centering \caption{Behaviors of \({^n}x\) for integer heights \(-1\leq{n}\leq{4}\) (left). The tetrational function \({^h}x\) for various heights \(-1\leq{h}\leq{3}\) with an interval of 0.2 (right). In general, the functions have complex values for non-integer heights if \(x<1\), and only the real components are shown in the right figures.} \label{fig:(^n)x_real} \end{figure} \begin{figure}[ht] \includegraphics[scale=0.5]{fig_n_x_complex.pdf} \centering \caption{The trajectories of tetrational functions \({^h}x\) in complex plane for non-integer heights \(h\) with an interval of 0.1. (a) \(-1<h<0\), (b) \(0<h<1\), (c) \(1<h<2\), (d) \(2<h<3\). } \label{fig:(^n)x_complex} \end{figure} \noindent In the right graph of Figure \ref{fig:(^n)x_real}, only the real components are shown, while trajectories for \(x<1\) in complex plane are given in Figures \ref{fig:(^n)x_complex}. It is obvious that \({^h}x\) do not always go to 0 or 1 for \(x\rightarrow{0}\), e.g., the trajectory for \(h=0.5\) circulates around 0 and hence never reach 0 or 1.\vspace{0.2in} \noindent It is also shown in Figure \ref{fig:(^n)x_complex} (a) that the trajectories for \(h=-0.5\) is a semicircle and the distorted semicircles for \(h\) and \(-(1+h)\) are axially symmetric each other, e.g., -0.2 and -0.8. The relation between \({^h}x\) and \(^{-(1+h)}y\) is given as Lemma \ref{lem_(^n)x_symmetry}, which explains not only the symmetry but also existence of a pair function convertible each other if \(-1\leq{h}\leq{0}\). \begin{lem} \label{lem_(^n)x_symmetry} Let \(-1\leq{h}\leq{0}\).\\ Let \(x>0\) and \(y>0\).\\ Let \(({^h}x)^{*}\) be the complex conjecture of \({^h}x\).\\ Then the following equation holds: \[({^h}x)^{*}=1-({^{-(1+h)}}y),\] or with the q-analog notation: \[[h+1]_{\ln{x}}=1-[-h]_{\ln{y}},\] where \[\ln{y}=\frac{1}{\ln{x}}.\] \end{lem} \begin{proof} Let \(q=\ln{x}\) and \(k>0\).\\ For \(0<x<1\), obviously \(q<0\), and \(q\) can be written as \(q=-k=ke^{i\pi}\). Hence we have \[q^{h+1}=k^{h+1}[\cos\pi(h+1)+i\sin\pi(h+1)]\] The tetrational function \({^h}x\) for \(-1\leq{h}\leq{0}\) is then expressed as \[ {^h}x=[h+1]_q=\frac{q^{h+1}-1}{q-1}=\frac{h^{k+1}\cos\pi(h+1)}{-k-1}+i\frac{k^{h+1}\sin\pi(h+1)}{-k-1}. \] Therefore \[ ({^h}x)^{*}=-\frac{h^{k+1}\cos\pi(h+1)}{k+1}+i\frac{k^{h+1}\sin\pi(h+1)}{k+1}. \] On the other hand, \[\ln{y}=\frac{1}{q}=k^{-1}e^{i\pi}.\] Hence we get \[ \left(\frac{1}{q}\right)^{-(h+1)}=k^{h+1}[\cos\pi(h+1)-i\sin\pi(h+1)]. \] Then \[ 1-(^{-(1+h)}y)=1-[-h]_{1/q}=1-\frac{\left(\frac{1}{q}\right)^{-h}-1}{\frac{1}{q}-1}=1-\frac{\left(\frac{1}{q}\right)^{-(h+1)}-q}{1-q} \] \[ =1-\frac{k^{h+1}\cos\pi(h+1)+k}{1+k}+i\frac{k^{h+1}\sin\pi(h+1)}{1+k} \] \[ =-\frac{h^{k+1}\cos\pi(h+1)}{k+1}+i\frac{k^{h+1}\sin\pi(h+1)}{k+1}. \] Therefore \(({^h}x)^{*}=1-({^{-(1+h)}}y)\) holds.\vspace{0.2in}\\ \noindent For \(x\geq{1}\), obviously \(q\geq{0}\), and we have \[({^h}x)^{*}={^h}x=\frac{q^{h+1}-1}{q-1}.\] On the other hand, we have \[1-(^{-(1+h)}y)=1-\frac{\left(\frac{1}{q}\right)^{-h}-1}{\frac{1}{q}-1}=\frac{q^{h+1}-1}{q-1}.\] Therefore generally \(({^h}x)^{*}=1-({^{-(1+h)}}y)\) holds.\\ \end{proof}\vspace{0.2in} \noindent Now let us consider the extension of the tetration to complex bases and heights. \begin{thm} \label{thm_complex_domains} Let \(r\) be real constant.\\ Let \(z\) be a complex variable. \\ Then the following statements hold: \begin{itemize} \item \({^z}r\) is holomorphic if and only if \(\Re(z)\notin\mathbb{Z}\). \item \({^r}z\) is holomorphic if and only if \(r\in\mathbb{Z}\). \end{itemize} \end{thm} \begin{proof} Let \(z=x+iy\).\\ (1) Proof of the first statement.\\ We shall prove this by induction.\\ Let \(s=\ln{q}=\ln{r}\).\\ Then \(q^{z+1}=e^{s(x+1+iy)}=e^{s(x+1)}[\cos(sy)+i\sin(sy)]\).\\ For \(-1\leq{x}<{0}\) we have \[ {^x}r=\frac{q^{x+1+iy}-1}{q-1}=\frac{e^{s(x+1)}\cos(sy)-1}{q-1}+i\frac{e^{s(x+1)}\sin(sy)}{q-1} =u_{0}+iv_{0},\] where \(u_{0}\) and \(v_{0}\) are real and imaginary component of \({^x}r\), respectively.\\ Obviously the Cauchy-Riemann equations holds: \[ \frac{\partial{u_{0}}}{\partial{x}} =\frac{\partial{v_{0}}}{\partial{y}} =\frac{se^{s(x+1)}\cos(sy)}{q-1}, \] \[ \frac{\partial{u_{0}}}{\partial{y}} =-\frac{\partial{v_{0}}}{\partial{v}} =-\frac{se^{s(x+1)}\sin(sy)}{q-1}. \] Therefore \({^x}r\) is holomorphic in \(-1\leq{x}<{0}\).\\ Let \(n\in\mathbb{Z}\). Suppose \(^{(x+iy)}r=u_{n}+iv_{n}\) for \((n-1)\leq{x}<{n}\) is holomorphic.\\ Then \[ ^{(x+1+iy)}r=r^{(u_{n}+iv_{n})}=e^{qu_{n}}\cos(qv_{n})+ie^{qu_{n}}\sin(qv_{n})=u_{n+1}+iv_{n+1}. \] By using the chain rule, e.g., \[ \frac{\partial{u_{n+1}}}{\partial{x}} =\frac{\partial{u_{n+1}}}{\partial{u_{n}}} \frac{\partial{u_{n}}}{\partial{x}} +\frac{\partial{u_{n+1}}}{\partial{v_{n}}} \frac{\partial{v_{n}}}{\partial{x}}, \] we can prove that Cauchy-Riemann equations holds: \[ \frac{\partial{u_{n+1}}}{\partial{x}} =\frac{\partial{v_{n+1}}}{\partial{y}} =qe^{qu_{n}}\cos(qv_{n})\frac{\partial{u_{n}}}{\partial{x}}-qe^{qu_{n}}\sin(qv_{n})\frac{\partial{v_{n}}}{\partial{x}}, \] \[ \frac{\partial{v_{n+1}}}{\partial{x}} =-\frac{\partial{u_{n+1}}}{\partial{y}} =qe^{qu_{n}}\sin(qv_{n})\frac{\partial{u_{n}}}{\partial{x}}-qe^{qu_{n}}\cos(qv_{n})\frac{\partial{v_{n}}}{\partial{x}}. \] Therefore \({^z}r\) is generally holomorphic in inside of each segment \((n-1)<x<n\).\\ However the Cauchy-Riemann equations do not hold at boundaries of segments \(x=n\) as follows:\\ The real and imaginary components of \({^z}r\) in segment \(-1\leq{x}<{0}\) approach to the following values as \(x\rightarrow{0}\): \[ \underset{x\rightarrow{0}}{\lim}{u_{0}}=\frac{q\cos(sy)-1}{q-1},\hspace{20pt}\underset{x\rightarrow{0}}{\lim}{v_{0}}=\frac{q\sin(sy)}{q-1}. \] On the other hand, \({^z}r\) approaches the following values as \(x\rightarrow{0}\) since \(u_{1}+iv_{1}=e^{qu_{0}}\cos(qv_{0})+ie^{qu_{0}}\sin(qv_{0})\): \[ \underset{x\rightarrow{0}}{\lim}{u_{1}}=\exp\left[q\frac{q\cos(sy)-1}{q-1}\right]\cos\left(q\frac{q\sin(sy)}{q-1}\right) \] \[ \underset{x\rightarrow{0}}{\lim}{v_{1}}=\exp\left[q\frac{q\cos(sy)-1}{q-1}\right]\sin\left(q\frac{q\sin(sy)}{q-1}\right) \] Obviously, in general \({^r}z\) is not continuous at the segment boundary. It is continuous if and only if the heights are real, \(y=0\) hence \(v_{1}=v_{0}=0\).\\ Similarly, \[ \underset{x\rightarrow{0}}{\lim}{u_{n+1}}=e^{qu_{n}}\cos(qv_{n}),\hspace{20pt}\underset{x\rightarrow{0}}{\lim}{v_{n+1}}=e^{qu_{n}}\sin(qv_{n}). \] Then \(\underset{x\rightarrow{n}}{\lim}(u_{n+1}+iv_{n+1})=\underset{x\rightarrow{n}}{\lim}(u_{n}+iv_{n})\) holds if and only if \(v_{n+1}=v_{n}=0\).\\ Therefore, the first statement holds.\vspace{0.2in} \noindent (2) Proof of the second statement\\ Let \(h,k\in\mathbb{R}\) and let \(\ln{z}=e^{h+ik}=e^{h}\cos{k}+ie^{h}\sin{k}\).\\ Since \(z=x+iy=e^{e^{h}\cos{k}}\cos{(e^{h}\cos{k})}+ie^{e^{h}\cos{k}}\sin{(e^{h}\cos{k})}\), we can confirm that the Cauchy-Riemann equations hold and expressed as follows: \[ \frac{\partial{h}}{\partial{x}}=\frac{\partial{k}}{\partial{y}}=\frac{1}{e^{(e^{h}\cos{k}+h)}\cos{(e^{h}\sin{k}+k)}}, \] \[ \frac{\partial{k}}{\partial{x}}=-\frac{\partial{h}}{\partial{y}}=-\frac{1}{e^{(e^{h}\cos{k}+h)}\sin{(e^{h}\sin{k}+k)}}. \] Let \(-1\leq{r}<0\). Then we have \[ {^r}z=\frac{(\ln{z})^{r+1}-1}{\ln{z}-1}=\frac{e^{(r+1)h}\cos{k(r+1)}+ie^{(r+1)h}\sin{k(r+1)}-1}{e^{h}\cos{k}+ie^{h}\sin{k}-1} \] \[ =\frac{[e^{(r+1)h}\cos{k(r+1)}+ie^{(r+1)h}\sin{k(r+1)}-1][(e^{h}\cos{k}-1)-ie^{h}\sin{k}]}{(e^{h}\cos{k}-1)^{2}+e^{2h}\sin^{2}{k}} \] \[ =\frac{e^{(r+2)h}\cos{rk}-e^{(r+1)h}\cos{k(r+1)}-e^{h}\cos{k}+1}{e^{2h}-2e^{h}\cos{k}+1} \] \[ +i\frac{e^{(r+2)h}\sin{rk}-e^{(r+1)h}\sin{k(r+1)}+e^{h}\sin{k}}{e^{2h}-2e^{h}\cos{k}+1} \] \[=u+iv,\] where \(u\) and \(v\) are real and imaginary component of \({^r}z\), respectively.\\ The partial derivatives of \(u\) and \(v\) by \(h\) or \(k\) are: \begin{align*} \frac{\partial{u}}{\partial{h}}&=\frac{(r+2)e^{(r+2)h}\cos{kr}-(r+1)e^{(r+1)h}\cos{k(r+1)}-e^{h}\cos{k}}{e^{2h}-2e^{h}\cos{k}+1}\\ &-\frac{\left(e^{(r+2)h}\cos{rk}-e^{(r+1)h}\cos{k(r+1)}-e^{h}\cos{k}+1\right)(2e^{2h}-2e^{h}\cos{k})}{(e^{2h}-2e^{h}\cos{k}+1)^{2}},\\ \frac{\partial{u}}{\partial{k}}&=\frac{-re^{(r+2)h}\sin{kr}+(r+1)^{2}e^{(r+1)h}\sin{k(r+1)}+e^{h}\sin{k}}{e^{2h}-2e^{h}\cos{k}+1}\\ &-\frac{\left(e^{(r+2)h}\cos{rk}-e^{(r+1)h}\cos{k(r+1)}-e^{h}\cos{k}+1\right)(2e^{h}\sin{k})}{(e^{2h}-2e^{h}\cos{k}+1)^{2}},\\ \end{align*} \begin{align*} \frac{\partial{v}}{\partial{h}}&=\frac{(r+2)e^{(r+2)h}\sin{kr}-(r+1)e^{(r+1)h}\sin{k(r+1)}+e^{h}\sin{k}}{e^{2h}-2e^{h}\cos{k}+1}\\ &-\frac{\left(e^{(r+2)h}\sin{rk}-e^{(r+1)h}\sin{k(r+1)}+e^{h}\sin{k}\right)(2e^{2h}-2e^{h}\cos{k})}{(e^{2h}-2e^{h}\cos{k}+1)^{2}},\\ \frac{\partial{v}}{\partial{k}}&=\frac{re^{(r+2)h}\cos{kr}-(r+1)^{2}e^{(r+1)h}\cos{k(r+1)}+e^{h}\cos{k}}{e^{2h}-2e^{h}\cos{k}+1}\\ &-\frac{\left(e^{(r+2)h}\sin{rk}-e^{(r+1)h}\sin{k(r+1)}+e^{h}\sin{k}\right)(2e^{h}\sin{k})}{(e^{2h}-2e^{h}\cos{k}+1)^{2}}. \end{align*} It is obvious that the Cauchy-Riemann equations do not hold. \[ \frac{\partial{u}}{\partial{x}}=\frac{\partial{u}}{\partial{h}}\frac{\partial{h}}{\partial{x}}+\frac{\partial{u}}{\partial{k}}\frac{\partial{k}}{\partial{x}} \neq \frac{\partial{v}}{\partial{h}}\frac{\partial{h}}{\partial{y}}+\frac{\partial{v}}{\partial{k}}\frac{\partial{k}}{\partial{y}}=\frac{\partial{v}}{\partial{y}}, \] \[ \frac{\partial{u}}{\partial{y}}=\frac{\partial{u}}{\partial{h}}\frac{\partial{h}}{\partial{y}}+\frac{\partial{u}}{\partial{k}}\frac{\partial{k}}{\partial{y}} \neq -\frac{\partial{v}}{\partial{h}}\frac{\partial{h}}{\partial{x}}-\frac{\partial{v}}{\partial{k}}\frac{\partial{k}}{\partial{x}}=-\frac{\partial{v}}{\partial{x}}. \] Similarly the tetrational function in other segments \({^{(r+n)}}z=\exp_{z}^{n}({^r}z)\) are generally not holomorphic inside the segment \((n-1)<x<n\) \\ However, at the boundaries of the segment \(r=n\), the function is expressed as \({^n}z=\exp_{z}^{n}[1]\) and is holomorphic proved by induction as follows .\\ Let \(z=e^{x+iy}\) and \({^{(n+1)}}z=x_{n}+iy_{n}\).\\ For \(n=1\), \[ {^1}z=e^{x+iy}=e^{x}\cos{y}+ie^{x}\sin{y}=u_{1}+iv_{1} \] The Cauchy-Riemann equations hold: \[\frac{\partial{u}}{\partial{x}}=\frac{\partial{v}}{\partial{y}}=e^{x}\cos{y},\hspace{20pt}\frac{\partial{u}}{\partial{y}}=-\frac{\partial{v}}{\partial{x}}=-e^{x}\sin{y}.\] Hence \({^1}z\) is holomorphic.\\ Suppose \({^n}z=u_{n-1}+iv_{n-1}\) is holomorphic: \[\frac{\partial{u_{n-1}}}{\partial{x}}=\frac{\partial{v_{n-1}}}{\partial{y}},\hspace{20pt}\frac{\partial{u_{n-1}}}{\partial{y}}=-\frac{\partial{v_{n-1}}}{\partial{x}}.\] Then \(^{(n+1)}z\) is expressed as: \[^{(n+1)}z=z^{(^{n}z)}=e^{(x+iy)(u_{n-1}+iv_{n-1})}=e^{(xu_{n-1}-yv_{n-1})+i(xv_{n-1}+yu_{n-1})}\] \[=e^{(xu_{n-1}-yv_{n-1})}\cos{(xv_{n-1}+yu_{n-1})}+ie^{(xu_{n-1}-yv_{n-1})}\sin{(xv_{n-1}+yu_{n-1})}\] \[=u_{n}+v_{n}.\] By using the relations above, it is shown that the Cauchy-Riemann equations hold: \[ \frac{\partial{u_{n}}}{\partial{x}}=\frac{\partial{v_{n}}}{\partial{y}} \] \[ =\left(u_{n-1}+x\frac{\partial{u_{n-1}}}{\partial{x}}-y\frac{\partial{v_{n-1}}}{\partial{x}}\right)e^{(xu_{n-1}-yv_{n-1})}\cos{(xv_{n-1}+yu_{n-1})} \] \[ -\left(v_{n-1}+x\frac{\partial{v_{n-1}}}{\partial{x}}+y\frac{\partial{u_{n-1}}}{\partial{x}}\right)e^{(xu_{n-1}-yv_{n-1})}\sin{(xv_{n-1}+yu_{n-1})}, \] \[ \frac{\partial{u_{n}}}{\partial{y}}=-\frac{\partial{v_{n}}}{\partial{x}} \] \[ =-\left(v_{n-1}+x\frac{\partial{v_{n-1}}}{\partial{x}}+y\frac{\partial{u_{n-1}}}{\partial{x}}\right)e^{(xu_{n-1}-yv_{n-1})}\cos{(xv_{n-1}+yu_{n-1})} \] \[ -\left(u_{n-1}+x\frac{\partial{u_{n-1}}}{\partial{x}}-y\frac{\partial{v_{n-1}}}{\partial{x}}\right)e^{(xu_{n-1}-yv_{n-1})}\sin{(xv_{n-1}+yu_{n-1})}. \] Hence \(^{(n+1)}z\) is holomorphic.\\ Therefore the second statement holds. \end{proof} \begin{cly} If both the base and height are complex values without integer components, \(x,y\in\mathbb{C}\) and \(x,y\notin\mathbb{Z}\), then the tetrational function \({^y}x\) is not holomorphic. \end{cly} \section{Calculation rules of tetration} \label{sec_rule} In this section, we study calculation rules of tetration based on the explicit formulae and properties of the functions.\vspace{0.2in} \noindent First let us extend the exponentiation operator. \begin{dfn} \label{dfn_extended_exp} Let \(r,x,y\in\mathbb{R}\).\\ Extended exponentiation operator \(\exp_{r}^{x}\) is defined as: \[\exp_{r}^{x}1\colonequals\exp_{r}^{\lfloor{x}\rfloor}\left(r^{[\{x\}]_q}\right)={^x}r,\] where the notations in Theorem \ref{thm_explicit} are used. \end{dfn} \begin{lem} \label{lem_e^(x-a)(^a)x} Let \(r,x,y\in\mathbb{R}\). The following equation holds: \[{^x}r=\exp_r^{x-y}\left({{^y}r}\right)\] \end{lem} \begin{proof} Since \(x=\lfloor{x}\rfloor+\{x\}\), by using the relation \ref{cly_core}, we have the following relations: \[ {^x}r=\exp_{r}^{x}1=\exp_{r}^{\lfloor{x}\rfloor+\{x\}}1=\exp_{r}^{\lfloor{x}\rfloor}\left({^{\{x\}}r}\right) .\] Then for any \(w\in\mathbb{R}\), \[ \exp_{r}^{x}1=\exp_{r}^{\lfloor{x}\rfloor-w+\{x\}+w}1=\exp_{r}^{\lfloor{x}\rfloor-w}\left({^{\{x\}+w}r}\right) .\] By using \(y=\{x\}+w\) and hence \(\lfloor{x}\rfloor-w=\lfloor{x}\rfloor+\{x\}-y=x-y\), we get the relation. \end{proof} \begin{lem} \label{lem_exp_summation} Let \(r,x,y\in\mathbb{R}\). The following equation holds: \[\exp_{r}^{x}\exp_{r}^{y}1=\exp_{r}^{x+y}1={^{x+y}}r.\] \end{lem} \begin{proof} From Lemma\ref{lem_e^(x-a)(^a)x}, \[{^x}r=\exp_{r}^{x-a}\left({^a}r\right)=\exp_{r}^{x-a}\exp_{r}^{a}1\] By rewriting \(y=x-a\), we get the relation. \end{proof}\vspace{0.2in} \noindent Next, we distinguish the exponentiation (right associative) from the power (left associative) for clarity.\vspace{0.2in} \begin{dfn} \label{dfn_vee_wedge} Let \(\vee\) and \(\wedge\) be the symbols of the exponentiation (right associative) and the power (left associative), respectively. \begin{align*} {^x}r\vee{f} & \colonequals{\exp_r^{x}{f}},\\ f\wedge{{^x}r} & \colonequals{f}^{({^x}r)}, \end{align*} where \(f\) is the operand.\\ \noindent The left-hand side of \(\vee\) is not a number but the part of the exponentiation operator. The iterated tetration is given as: \[{^x}({^y}r)\vee{f}\colonequals(\exp_{r}^{y})^{x}f=\exp_{r}^{xy}f.\] \end{dfn}\vspace{0.2in} \noindent The notation \({^x}r\vee{^y}r\) refers to \(\exp_{r}^x({^y}x)\), where the operand \({^y}r\) is a number while \({^x}r\vee\) is not a number but the extended exponent operator, not referring to \(\exp_{({^x}r)}\). On the other hand, the notation \({^x}r\wedge{^y}r\) refers to \(({^x}r)^{({^y}r)}=\exp_{({^x}r)}({^y}r)\), where both \({^x}r\) and \({^y}r\) are numbers. Similarly, \({^x}({^y}r)\vee{1}\) is not \({^x}t\) with \(t={^y}r\vee{1}\).\vspace{0.2in} \noindent Some calculation rules are given as in Theorem \ref{thm_tet_rules}. \begin{thm} \label{thm_tet_rules} Let \(r,t,x,y,z\in\mathbb{R}\). The following relations hold: \begin{eqnarray} \setcounter{equation}{1} {^x}r\vee{1}&=&\hspace{3pt}{^x}r.\\ -{^x}r&=&{^{-1}}r\vee\left(\frac{1}{^{x+1}r}\right)\\ \frac{1}{{^x}r}&=&r\vee({-^{x-1}r})\\ {^x}r+{^y}r&=&{^{-1}}r\vee({^{x+1}}r\cdot{^{y+1}}r).\\ {^x}r-{^y}r&=&{^{-1}}r\vee\left(\frac{{^{x+1}}r}{{^{y+1}}r}\right).\\ {^{x}}r\wedge{^y}r&=&{^{y+1}}r\wedge{^{x-1}r}.\\ {^x}r\cdot{^y}r&=&{^{-1}r\vee\left({^x}r\wedge{^y}r\right)}=r\vee\left({^{x-1}}r+{^{y-1}r}\right)\\ \frac{{^x}r}{{^y}r}&=&{^{-1}r}\vee({^{x+1}}r\wedge\frac{1}{{^y}r})=r\vee({^{x-1}}r-{^{y-1}}r). \end{eqnarray} \begin{eqnarray} {^x}r\vee{^y}r&=&{^y}r\vee{^x}r\hspace{3pt}=\hspace{3pt}{^{x+y}}r.\\ {^y}({^x}r)&=&{^x}({^y}r)\hspace{3pt}=\hspace{3pt}{^{xy}}r.\\ {^z}({^x}r\vee{^y}r)&=&{^{z(z+y)}}r.\\ {^{\{x\}}}r&=&r\vee\frac{(\ln{r})^{\{x\}}-1}{\ln{r-1}}\\ {^x}r&=&{^{\lfloor{x}\rfloor}}r\vee{^{\{x\}}}r\\ {^{\{x\}+\{y\}-1}}r&=&{^{\{x\}-1}}r+(\ln{r})^{\{x\}}\cdot\left({^{\{y\}-1}}r\right),\hspace{10pt}\{x\}+\{y\}<1.\\ {^x}r&=&\left({^{\lfloor{x}\rfloor+1}}r\right)\vee\left(1-{^{-\{x\}}}t\right)^{*},\hspace{10pt}\ln{r}\cdot\ln{t}=1.\\ \frac{\partial({^x}r)}{\partial({^{\{x\}-1}r})}&=&(\ln{r})^{\lfloor{x}\rfloor+1}\prod_{k=0}^{\lfloor{x}\rfloor}({^k}}r\vee{^{\{x\}}r). \end{eqnarray} \end{thm} \begin{proof} The proofs are given as fellows respectively:\\ (1) Directly given from Definition \ref{dfn_extended_exp}. This is the relation between operator \({^x}r\vee\) and number \({^x}r\): \[{^x}r=\exp_{r}^{x}1={^x}r\vee{1}.\] (2) By Definition \ref{dfn_vee_wedge}, we have \(\log_{r}f={^{-1}}r\vee{f}\). \[-{^x}r=-\log_{r}[{^{x+1}}r]=\log_{r}\left[\frac{1}{{^{x+1}}r}\right]={^{-1}}r\vee\left[\frac{1}{^{x+1}r}\right].\] (3) By the exponent rule, we have \[\frac{1}{{^x}r}=({^x}r)^{-1}=(r^{^{x-1}r})^{(-1)}=r^{-(^{x-1}r)}=r\vee(-^{x-1}r).\] (4) By the logarithm rule, \[{^x}r+{^y}r=\log_{r}({^{x+1}}r)+\log_{r}({^{y+1}}r)=\log_{r}({^{x+1}}r\cdot{^{y+1}}r)={^{-1}}r\vee({^{x+1}}r\cdot{^{y+1}}r).\] (5) By the logarithm rule, \[{^x}r-{^y}r=\log_{r}({^{x+1}}r)-\log_{r}({^{y+1}}r)=\log_{r}\left(\frac{{^{x+1}}r}{{^{y+1}}r}\right)={^{-1}}r\vee\left(\frac{{^{x+1}}r}{{^{y+1}}r}\right).\] (6) By the exponent rule, \[{^x}r\wedge{^y}r=({^x}r)^{({^y}r)}=r^{(^{x-1}r)({^y}r)}=r^{({^y}r)(^{x-1}r)}={^{y+1}}r\wedge{^{x-1}r}.\] (7) By the logarithm rule, \[{^x}r\cdot{^y}r={^x}r\cdot\log_{r}({^{y+1}}r)=\log_{r}\left[({^{y+1}}r)^{({^x}r)}\right]={^{-1}r}\vee({^{y+1}r}\wedge{^{x}r})={^{-1}r}\vee({^{x}r}\wedge{^{y}r}),\] where the (6) is used at the last step. By the exponent rule, \[{^x}r\cdot{^y}r=r^{({^{x-1}}r)}\cdot{r}^{({^{y-1}}r)}=r^{({^{x-1}}r+{^{y-1}}r)}=r\vee({^{x-1}}r+{^{y-1}}r).\] (8) By the logarithm rule, \[\frac{{^x}r}{{^y}r}=\frac{1}{{^y}r}\log_{r}({^{x+1}}r)=\log_{r}\left[({^{x+1}}r)^{\frac{1}{({^y}r)}}\right]={^{-1}r}\vee({^{x+1}}r\wedge\frac{1}{{^y}r}).\] By the exponent rule, \[\frac{{^x}r}{{^y}r}=r^{({^{x-1}}r)}\cdot{r}^{(-{^{y-1}}r)}=r^{({^{x-1}}r-{^{y-1}}r)}=r\vee({^{x-1}}r-{^{y-1}}r).\] (9) Directly give from Lemma \ref{lem_e^(x-a)(^a)x} and \ref{lem_exp_summation}.\\ (10) From definition \ref{dfn_vee_wedge}, we have the relation by assining \(f=1\).\\ (11) \({^z}({^x}r\vee{^y}r)={^z}({^x}r\vee{^y}r)\vee{1}=(\exp_{r}^{x+y})^{z}1=\exp_{r}^{z(x+y)}1={^{z(x+y)}}r\vee{1}={^{z(x+y)}}r\).\\ (12) Corollary \ref{cly_core} directly give the relation.\\ (13) We have the relation from Definition \ref{dfn_extended_exp} and (12).\\ (14) This relation is the property of q-numbers: \[ {^{\{x\}+\{y\}-1}}r=\frac{(\ln{r})^{\{x\}+\{y\}}-1}{\ln{r}-1}=\frac{(\ln{r})^{\{x\}}-1}{\ln{r}-1}+(\ln{x})^{\{x\}}\frac{(\ln{r})^{\{y\}}-1}{\ln{r}-1} \] \[={^{\{x\}-1}}r+(\ln{r})^{\{x\}}\cdot\left({^{\{y\}-1}}r\right).\] (15) From Lemma \ref{lem_(^n)x_symmetry}, we have the following relation. \[\left({^{\{x\}-1}}r\right)^{*}=1-{^{-\{x\}}}t\] Then \({^{\{x\}-1}}r=\left(1-{^{-\{x\}}}t\right)^{*}\).\\ By applying the operation of \(\left({^{\lfloor{x}\rfloor+1}}r\vee\right)\), we have \[{^{\lfloor{x}\rfloor}}r\vee{^{\{x\}}}r=\left({^{\lfloor{x}\rfloor+1}}r\right)\vee\left(1-{^{-\{x\}}}t\right)^{*}.\] Then we use (13) and get the relation. \\ (16) We get the relation from Theorem \ref{thm_explicit} and Corollary \ref{cly_core}.\\ \end{proof} \section{Concluding remarks} \label{sec_summary} We derived the simple explicit formulae of the continuous tetrational function as well as multi-tetrational funcion based only on the delay-differential equation. The definition of the multi-tetrational function and finding the Stirling numbers of the second kind in Taylor coefficients are the key steps of our approach. The solutions is piecewise connected function and is class \(C^{1}\) at connecting point for real heights, and the height is extended to complex heights except the segment boundaries. Our solution has advantages in simple expression as well as in continuity in real r-x plane. The series for each segment converges absolutely and has infinite radius of convergence . The solution should be compared with those of numerical approaches of class \(C^{\infty}\) \cite{Kouznetsov2009}\cite{Paulsen2017} with limited domain of base \(r>e^{1/e}\) and finite radius of convergence for Taylor expansion. The calculation rules of tetration are given by distinguishing exponentiation from power operation. \begin{bibdiv} \begin{biblist} \bib{Goodstein1947}{article}{ title={Transfinite ordinals in recursive number theory}, volume={12}, DOI={10.2307/2266486}, journal={J. Symb. Log.}, author={Goodstein, R. L.}, year={1947}, pages={123--129} } \bib{Scot2006}{article}{ title={General relativity and quantum mechanics: toward a generalization of the Lambert W function}, volume={17}, DOI={10.1007/s00200-006-0196-1}, journal={Appl. Algebra Eng. Commun. Comput.}, author={Scot, T. C.}, author={Mann, R.}, author={Martinez II, R. E.}, year={2006}, pages={41--47} } \bib{Sun2017}{article}{ title={On the topology of air navigation route systems}, volume={170}, DOI={10.1680/jtran.15.00106}, journal={P I CIVIL ENG-ENERGY}, author={Sun, X.}, author={Wandelt, S.}, author={Linke, F.}, year={2017}, pages={46--59} } \bib{Khetkeeree2019}{article}{ title={Signal Reconstruction using Second Order Tetration Polynomial}, DOI={10.1109/ITC-CSCC.2019.8793435}, journal={2019 34th International Technical Conference on Circuits/Systems, Computers and Communications (ITC-CSCC), JeJu, Korea (South)}, author={Khetkeeree, S.}, author={Chansamorn, C.}, year={2019}, pages={1--4} } \bib{Furuya2019}{article}{ title={Compaction of Church Numerals}, VOLUME = {12}, DOI={10.3390/a12080159}, journal={Algorithms (Basel)}, author={Furuya, I.}, author={Kida, T.}, year={2019}, pages={159} } \bib{Euler1783}{article}{ title={De serie Lambertina Plurimisque eius insignibus proprietatibus,}, volume={2}, journal={Acta Acad. Scient. Petropol.}, author={Euler, L.}, year={1783}, pages={29--51} } \bib{Eisenstein1844}{article}{ title={Entwicklung von \(\alpha^{\alpha^{\alpha^{\udots}}}\),}, volume={28}, doi = {10.1515/crll.1844.28.49}, journal={J. Reine angew. Math.}, author={Eisenstein, G.}, year={1844}, pages={49--52} } \bib{Corless1996}{article}{ title={On the Lambert W function}, volume={5}, doi = {10.1007/BF02124750}, journal={Adv. Comput. Math.}, author={Corless, R. M.}, author={Gonnet, G. H.}, author={Hare, D. E. G.}, author={Jeffrey, D. J.}, author={Knuth, D. E.}, year={1996}, pages={329--359} } \bib{Hooshmand2006}{article}{ title={Ultra power and ultra exponential functions}, volume={17}, doi = {10.1080/10652460500422247}, journal={Integral Transforms Spec. Funct.}, author={Hooshmand, M. H.}, year={2006}, pages={549--558} } \bib{Kouznetsov2009}{article}{ title={Solution of \(F(z+1) = \exp(F(z))\) in Complex \(z\)-Plane}, volume={78}, doi = {10.1090/S0025-5718-09-02188-7}, journal={Math. Comp.}, author={Kouznetsov, D.}, year={2009}, pages={1647--1670} } \bib{Paulsen2017}{article}{ title={Solving \(F(z+1)=b^{F(z)}\) in the complex plane,}, volume={43}, doi = {10.1007/s10444-017-9524-1}, journal={Adv. Comput. Math.}, author={Paulsen, W.}, author={Cowgill, S.}, year={2009}, pages={1261--1282} } \bib{Abramowitz1964}{book}{ title={Handbook of Mathematical Functions with Formulas, Graphs, and Mathematical Tables}, publisher={NationalBureau of Standards}, isbn = {0-486-61272-4}, author={Abramowitz, M.}, author={Stegun, A.}, year={1964}, } \bib{Graham1994}{book}{ title={Concrete Mathematics: A Foundation for Computer Science}, publisher={Addison-Wesley Longman Publishing Co., Inc.}, isbn={0201558025}, edition={2nd}, author={Graham, Ronald L.}, author={Knuth, Donald E.}, year={1994}, } \end{biblist} \end{bibdiv} \end{document}
{ "timestamp": "2021-09-01T02:13:58", "yymm": "2105", "arxiv_id": "2105.00247", "language": "en", "url": "https://arxiv.org/abs/2105.00247", "abstract": "The continuous tetrational function ${^x}r=\\tau(r,x)$, the unique solution of equation $\\tau(r,x)=r^{\\tau(r,x-1)}$ and its differential equation $\\tau'(r,x) =q \\tau(r,x) \\tau'(r,x-1)$, is given explicitly as ${^x}r=\\exp_{r}^{\\lfloor x \\rfloor+1}[\\{x\\}]_q$, where $x$ is a real variable called height, $r$ is a real constant called base, $\\{x\\}=x-\\lfloor x \\rfloor$ is the sawtooth function, $\\lfloor x \\rfloor$ is the floor function of $x$, and $[\\{x\\}]_q=(q^{\\{x\\}}-1)/(q-1)$ is a q-analog of $\\{x\\}$ with $q=\\ln r$, respectively. Though ${^x}r$ is continuous at every point in the real $r-x$ plane, extensions to complex heights and bases have limited domains. The base $r$ can be extended to the complex plane if and only if $x\\in \\mathbb{Z}$. On the other hand, the height $x$ can be extended to the complex plane at $\\Re(x)\\notin \\mathbb{Z}$. Therefore $r$ and $x$ in ${^x}r$ cannot be complex values simultaneously. Tetrational laws are derived based on the explicit formula of ${^x}r$.", "subjects": "Classical Analysis and ODEs (math.CA)", "title": "Extension of tetration to real and complex heights", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9875683484150417, "lm_q2_score": 0.8104789063814617, "lm_q1q2_score": 0.8004033150003693 }
https://arxiv.org/abs/1812.00610
Weak discrete maximum principle and $L^\infty$ analysis of the DG method for the Poisson equation on a polygonal domain
We derive several $L^\infty$ error estimates for the symmetric interior penalty (SIP) discontinuous Galerkin (DG) method applied to the Poisson equation in a two-dimensional polygonal domain. Both local and global estimates are examined. The weak maximum principle (WMP) for the discrete harmonic function is also established. We prove our $L^\infty$ estimates using this WMP and several $W^{2,p}$ and $W^{1,1}$ estimates for the Poisson equation. Numerical examples to validate our results are also presented.
\section{Introduction} The discontinuous Galerkin (DG) method, which was proposed originally by Reed and Hill \cite{osti_4491151} in 1973, is a powerful method for solving numerically a wide range of partial differential equations (PDEs). We use a discontinuous function which is a polynomial on each element and introduce the numerical flux on each element boundary. The DG scheme is then derived by controlling the numerical flux to ensure the local conservation law. For linear elliptic problems, the study of stability and convergence developed well in the early 2000s; see the standard references \cite{MR1885715}, \cite{riv08} and \cite{MR2485457} for the detail. However, most of those works are based on the $L^2$ and DG energy norms, and a little is done using other norms. An exception is \cite{MR2113680}, where the optimal order error estimate in the $L^\infty$ norm was proved using the discrete Green function. However, the DG scheme in \cite{MR2113680} is defined in the exactly-fitted triangulation of a smooth domain; this restriction is somewhat unrealistic for practical applications. From the view point of applications, the $L^p$ theory, especially the $L^\infty$ theory, plays an important role in the analysis of nonlinear problems. Therefore, the development of the $L^p$ theory for the DG method is a subject of great importance. Another important subject for confirming the validity of numerical methods is the discrete maximum principle (DMP). Nevertheless, only a few works has been devoted to DMP for DG method. Horv\'{a}th and Mincsovics \cite{MR3015392} proved DMP for DG method applied to the one-dimensional Poisson equation. Badia, Bonilla and Hierro (\cite{MR3315069}, \cite{MR3646366}) proposed nonlinear DG schemes satisfying DMP for linear convection-diffusion equations in the one and two space dimensions. However, to the best of our knowledge, no results are known regarding DMP for the standard DG method in the higher-dimensional space domain. In contrast, the $L^p$ theory and DMP have been actively studied regarding the standard finite element method (FEM). The pioneering work by Ciarlet and Raviart \cite{MR0375802} studied $L^\infty$ and $W^{1,p}$ error estimates together with DMP; in particular, those error estimates were proved as a consequence of DMP. Then, the $L^\infty$ error estimates were proved using several methods; Scott \cite{MR0436617} applied the discrete Green function and Nitsche \cite{MR568857} utilized the weighted norm technique. Rannacher and Scott succeeded in deriving the optimal $W^{1,p}$ error estimate for $1\le p\le \infty$ in \cite{MR645661}. Detailed local and global pointwise estimates have been studied by many researchers, such as \cite{MR1464148}. Recently, the optimal order $W^{1,\infty}$ and $L^\infty$ stability and error estimates were established for the Poisson equation defined in a smooth domain; the effect of polyhedral approximations of a smooth domain was precisely examined. See \cite{2018arXiv180400390K} for the detail. As is well known, the non-negativity assumption (non-obtuse assumption) on the triangulation is necessary for DMP to hold in the standard FEM. In this connection, Schatz \cite{MR551291} is noteworthy in this area for deriving the \emph{weak maximum principle} (WMP) without the non-negativity assumption and applying it to the proof of the stability, local and global error estimates in the $L^\infty$ norm. In this paper, we are motivated by \cite{MR551291}, and extend the results of that study to the symmetric interior penalty (SIP) DG method which is one of the popular DG method for the Poisson equation. Our results are summarized as follows. Let $u$ be the solution of the Dirichlet boundary value problem for the Poisson equation \eqref{eq:poisson} defined in a polygonal domain $\Omega\subset\mathbb{R}^2$, and let $u_h$ be the solution of the SIPDG method \eqref{eq:dg} in the finite dimensional space $V_h$ defined as \eqref{eq:vh}. Then, we have the $L^\infty$ interior error estimate (see Theorem \ref{thm1}) \[ \norm{u-u_h}_{L^\infty(\Omega_0)} \le C \left(\inf_{\chi \in V_h}\norm{u-\chi}_{\alpha(h),\Omega_1}+\norm{u-u_h}_{L^2(\Omega_1)}\right), \] where $\Omega_0$ and $\Omega_1$ are open subsets such that $\Omega_0 \subset \Omega_1 \subset \Omega$, and $C$ is independent of $h$ and the choice of $\Omega_0$ and $\Omega_1$. This interior estimate is valid under Assumption \ref{asm1} below, where $\alpha(h)$ and $\|\cdot\|_{\alpha(h),\Omega_1}$ are defined. Using this interior error estimate, we prove the WMP (see Theorem \ref{thm2}) \[ \norm{u_h}_{L^\infty(\Omega)} \le C \norm{u_h}_{ L^\infty(\partial \Omega)} \] for the discrete harmonic function $u_h\in V_h$. Finally, under some assumptions on the triangulation (see Assumption \ref{asm2}), we prove the $L^\infty$ error estimate (see Theorem \ref{thm3}) \[ \norm{u-u_h}_{L^\infty(\Omega)} \le C \left(\inf_{\chi \in V_h}\norm{u-\chi}_{\alpha(h),\Omega} + \norm{u-u_h}_{L^\infty(\partial \Omega)}\right) \] for the solution $u$ of the Poisson equation \eqref{eq:poisson} and its DG approximation $u_h$. As a matter of fact, the WMP is a key point of the proof of this error estimate. Moreover, we obtain the $L^\infty$ error estimate of the form (see Corollary \ref{cor2}) \[ \norm{u-u_h}_{L^\infty(\Omega)} \le C h^{r}\norm{u}_{W^{1+r,\infty}(\Omega)}, \] where $r$ denotes the degree of approximate polynomials. Unfortunately, this error estimate is only sub-optimal even for the piecewise linear element ($r=1$). This is because the Dirichlet boundary condition is imposed ``weakly'' by the variational formulation (Nitsche' method) in the DG method. On the other hand, it is imposed ``strongly'' by the nodal interpolation in the FEM. This implies that we further need to more deeply consider the imposition of the Dirichlet boundary condition in the DG method. In particular, study of better, more precise estimates of $\alpha(h)$ and $\norm{u-u_h}_{L^\infty(\partial \Omega)}$ are necessary, and will be a focus of our future works. Our SIPDG scheme and main results, Theorems \ref{thm1}, \ref{thm2} and \ref{thm3}, are presented in Section \ref{result}. The main tool of our analysis is several $W^{2,p}$ and $W^{1,1}$ estimates for solutions of the Poisson equation. Several local error estimates developed in previous works (see \cite{MR0520174}, \cite{MR2113680} and \cite{MR0431753}) are also used. After having presented those preliminary results in Section \ref{sect2}, we state the proofs of Theorems \ref{thm1}, \ref{thm2} and \ref{thm3} in Sections \ref{s:I}, \ref{s:II} and \ref{s:III}, respectively. Finally, we report the results of numerical experiments to confirm the validity of our theoretical results in Section \ref{s:ne}. Before concluding this Introduction, here, we list the notation used in this paper. \paragraph{Notation.} We follow the standard notation, for example, of \cite{MR2424078} as for function spaces and their norms. In particular, for $1\le p \le \infty$ and a positive integer $j$, we use the standard Lebesgue space $L^{p}(\mathcal{O})$ and the Sobolev space $W^{j,p}(\mathcal{O})$. Here and hereinafter, $\mathcal{O}$ denotes a bounded domain in $\mathbb{R}^2$. The semi-norms and norm of $W^{j,p}(\mathcal{O})$ are denoted, respectively, by \[ \abs{v}_{W^{i,p}(\mathcal{O})} = \left(\sum_{\abs{\alpha} = i}\norm{\pderiv[\alpha]{v}{x}}_{L^p(\mathcal{O})}^p\right)^{{1}/{p}},\quad \norm{v}_{W^{j,p}(\mathcal{O})} = \left(\sum_{i = 0}^{j} \abs{v}_{W^{i,p}(\mathcal{O})}^p\right)^{{1}/{p}}. \] Letting $C_0^\infty(\mathcal{O})$ be the set of all infinity differentiable functions with compact support in $\mathcal{O}$, $W^{j,p}_0(\mathcal{O})$ denotes the closure of $C^\infty_0(\mathcal{O})$ in the $W^{j,p}(\mathcal{O})$ norm. The space $W^{1,p}_0(\mathcal{O})$ is characterized as $W^{j,p}_0(\mathcal{O})=\{v\in W^{1,p}(\mathcal{O})\colon v|_{\partial\mathcal{O}=0}\}$ if $\partial\mathcal{O}$ is a Lipschitz boundary. Let $p'$ be the H\"older conjugate exponent of $p$; $1/p+1/p'=1$. The inner product of $L^2(\mathcal{O})$ is denoted by $(\cdot,\cdot)_{\mathcal{O}}$. The $\mathbb{R}^d$-Lebesgue measure of $\mathcal{O}$ is denoted by $\abs{\mathcal{O}}_d$. We also use the fractional order Sobolev space $W^{s,p}(\mathcal{O})$ for $s>0$. As usual, we write as $H^s(\mathcal{O}) = W^{s,2}(\mathcal{O})$. For $\Gamma \subset \partial \mathcal{O}$, we define $W^{j,p}(\Gamma)$ and $H^s(\Gamma)$ using a surface measure $ds=ds_\Gamma$. For $\mathcal{O}_1,\mathcal{O}_2\subset\mathbb{R}^2$, we write $\mathcal{O}_1\Subset\mathcal{O}_2$ and $\mathcal{O}_2\Supset\mathcal{O}_1$ to express $\mathcal{O}_1\subset\mathcal{O}_2$ and $\overline{\mathcal{O}_1}\subset\mathcal{O}_2$. Finally, the letter $C$ denotes a generic positive constant depending only on $\Omega$, $\partial\Omega$, the criterion $\sigma_0$ of the penalty parameter and the shape-regularity constant $C_*$ defined in Section \ref{result}. \section{DG scheme and results} \label{result} Throughout this paper, letting $\Omega$ be a bounded polygonal domain in $\mathbb{R}^2$, we consider the Dirichlet boundary value problem for the Poisson equation \begin{equation} \left\{ \begin{array}{rcc} -\Delta u = f & \text{in} & \Omega \\ u = g & \text{on} & \partial \Omega, \end{array}\right.\label{eq:poisson} \end{equation} where $f \in L^2(\Omega)$ and $g \in H^{1/2}(\partial\Omega)$. There exists an extension $\tilde{g}\in H^1(\Omega)$ such that $\tilde{g}=g$ on $\partial\Omega$ and $\|\tilde{g}\|_{H^1(\Omega)}\le C\|g\|_{H^{1/2}(\partial\Omega)}$. The following discussion does not depend on the way of extension. Then, the weak formulation of \eqref{eq:poisson} is stated as follows. \smallskip \textup{(BVP;$f,g$)} Find $u=u_0+\tilde{g}\in H^1(\Omega)$ and $u_0\in H^1_0(\Omega)$ such that \begin{equation} \int_\Omega \nabla u_0 \cdot \nabla v ~dx = (f,v)_\Omega-\int_\Omega \nabla \tilde{g} \cdot \nabla v ~dx \quad {}^\forall v \in H^1_0(\Omega). \label{eq:weakhomodiriclet} \end{equation} The Lax--Milgram theory guarantees that \textup{(BVP;$f,g$)} admits a unique solution $u\in H^1(\Omega)$. The regularity of $u$ is well studied. See \cite[Theorem 1]{MR0466912} and \cite[Lemma 1.2]{MR551291} for the detail of the following results. \begin{prop}\label{prop:poisson} Let $0 < \alpha < 2\pi$ be the maximum (interior) angle of $\partial\Omega$, and set $\beta = \pi/\alpha$. Letting $g=0$, we suppose that $u \in H^1_0(\Omega)$ is the solution of \textup{(BVP;$f,0$)}. \begin{enumerate} \item[\textup{(i)}] If $\Omega$ is convex ($\beta > 1$), then $u$ belongs to $H^2(\Omega) \cap H^1_0(\Omega)$, and we have \begin{equation} \abs{u}_{H^2(\Omega)} \le \norm{f}_{L^2(\Omega)}. \label{eq:convexregulality} \end{equation} \item[\textup{(ii)}] If $1/2 < \beta < 1$ and $f \in L^p(\Omega)$ for some $1<p<2/(2-\beta)$, then $u$ belongs to $ W^{2,p}(\Omega) \cap H^1_0(\Omega)$, and there exists a positive constant $C$ depending only on $\Omega$ and $p$ such that \begin{equation} \norm{u}_{W^{2,p}(\Omega)} \le C \norm{f}_{L^p(\Omega)}. \label{eq:nonconvexregulality} \end{equation} \end{enumerate} \end{prop} \begin{rem} Because $\beta > 1/2$, we have $2/(2-\beta) > 4/3$ and, consequently, \eqref{eq:nonconvexregulality} holds for some $4/3 < p \le 2$. \end{rem} Let $\{\mathcal{T}_h\}_h$ be a family of shape-regular and quasi-uniform triangulations of $\Omega$ (see \cite[(4.4.15), (4.4.16)]{bs08}). That is, there exists a positive constant $C_*$ satisfying \begin{equation} \frac{h_K}{\rho_K} \le C_* ,\, \,\frac{h}{h_K} \le C_*\qquad (K\in\mathcal{T}_h\in\{\mathcal{T}_h\}_h). \label{eq:triangulation} \end{equation} Therein, $h_K$ and $\rho_K$ denote the diameters of the circumscribed and inscribed circles of $K$, respectively. Moreover, the granularity parameter $h$ is defined as $h = \displaystyle \max_{K \in \mathcal{T}_h}h_K$. We set, for $1\le p\le \infty$, \begin{equation} V^p = V^p(\Omega) \coloneqq \{ v \in L^p(\Omega) \colon v|_K \in W^{1,p}(K) , (\nabla v)|_{\partial K} \in L^{p}(\partial K) {}^\forall K \in \mathcal{T}_h\}. \end{equation} It is noteworthy that $v$ is a continuous function on $K \in \mathcal{T}_h$ if $v \in V^\infty$. For a positive integer $r$, we define the finite element spaces $V_h$ and $\mathring V_h$ as \begin{align} V_h &= V_h^r(\Omega) \coloneqq\{ v_h \in L^2(\Omega) \colon v_h|_K \in\mathcal{P}^r(K) \text{ for each }K \in \mathcal{T}_h\}, \label{eq:vh}\\ \mathring V_h & = \mathring V_h(\Omega) \coloneqq \{v_h \in V_h \colon \operatorname{supp}{v_h} \subset \Omega\},\label{eq:vho} \end{align} where $\mathcal{P}^r(K)$ denotes the set of all polynomials of degree $\le r$. For $\mathcal{O}\subset\Omega$, we set $V^p(\mathcal{O})=\{v|_{\mathcal{O}}\colon v\in V^p\}$, $V_h(\mathcal{O})=\{v|_{\mathcal{O}}\colon v\in V_h\}$ and $\mathring V_h(\mathcal{O}) \coloneqq \{v_h \in V_h(\mathcal{O}) \colon \operatorname{supp}{v_h} \subset \mathcal{O}\}$. We let $\mathcal{E}_h$ be the set of all edges of $K \in \mathcal{T}_h$, and set \[ \E^{\partial}_h \coloneqq \{e \in \mathcal{E}_h \colon e \subset \partial \Omega \},\quad \E^{\circ}_h \coloneqq \mathcal{E}_h \setminus \E^{\partial}_h. \] For $v \in V^p$ and $e \in \mathcal{E}_h$, we define $\mean{\cdot}$ and $\jump{\cdot}$ as follows. If $e \in \E^{\circ}_h$, we set \begin{align} \mean{v} \coloneqq \frac{1}{2}(v_1 +v_2)\,,\,\,& \jump{v} \coloneqq v_1n_1 + v_2n_2 \,,\,\\ \mean{\nabla v} \coloneqq \frac{1}{2}(\nabla v_1 + \nabla v_2)\,,\,& \jump{\nabla v} \coloneqq \nabla v_1 \cdot n_1 + \nabla v_2 \cdot n_2 \,. \end{align} If $e \in \E^{\partial}_h$, we set \begin{align} \mean{v} \coloneqq v \,,\,\jump{v} \coloneqq v n\,,\,\mean{\nabla v} \coloneqq \nabla v \,,\,\jump{\nabla v} \coloneqq \nabla v \cdot n \,. \end{align} Therein, for $e \in \E^{\circ}_h$, there exist distinct $K_1,\,K_2 \in \mathcal{T}_h$ satisfying $e \subset \partial K_1 \cap \partial K_2$ and $v_i = v|_{K_i}$, where $n_i$ denotes the outward unit normal vector to $e$ of $K_i$, and $n$ denotes the outward unit normal vector on $\partial \Omega$. We define norm $\norm{v}_{V^p(\mathcal{O})}$ on $V^p(\mathcal{O})$, where $\mathcal{O}=\Omega$ or $\mathcal{O}\subset\Omega$, as \[ \norm{v}_{V^p(\mathcal{O})}^p \coloneqq \sum_{K \in \mathcal{T}_h} \norm{v}_{W^{1,p}(K \cap \mathcal{O})}^p + \sum_{e \in \mathcal{E}_h} h_e^{1-p}\norm{\jump{v}}_{L^p(e \cap \overline{\mathcal{O}})}^p + \sum_{e \in \mathcal{E}_h} h_e \norm{\mean{\nabla v}}_{L^p(e \cap \overline{\mathcal{O}})}^p \] and \[ \norm{v}_{V^\infty(\mathcal{O})} \coloneqq \max_{K \in \mathcal{T}_h} \norm{v}_{W^{1,\infty}(K \cap \mathcal{O})} + \max_{e \in \mathcal{E}_h}h_e^{-1}\norm{\jump{v}}_{L^\infty(e\cap\overline{\mathcal{O}})}+ \max_{e \in \mathcal{E}_h} \norm{\mean{\nabla v}}_{L^\infty(e\cap\overline{\mathcal{O}})}, \] where $h_e = (h_{K_1} + h_{K_2})/2$ if $e \in \E^{\circ}_h$ and $h_e = h_K$ if $e \in \E^{\partial}_h$. Letting $1 \le p ,p'\le \infty$ and $1/p+1/p'=1$, we introduce the DG bilinear form on $V^p\times V^{p'}$ as \begin{multline} a(u,v) \coloneqq \sum_{K \in \mathcal{T}_h} \int_K \nabla u \cdot \nabla v ~dx \\ -\sum_{e \in \mathcal{E}_h} \int_e(\mean{\nabla u}\jump{v} + \mean{\nabla v}\jump{u})~ ds + \sum_{e \in \mathcal{E}_h}\frac{\sigma}{h_e}\int_e \jump{u}\jump{v} ~ ds \label{eq:bilinear} \end{multline} for $u \in V^p$ and $v \in V^{p'}$. Herein, $\sigma$ is a sufficiently large constant. The linear form $F$ on $V^2$ is defined by \begin{equation} F(v) \coloneqq \int_\Omega fv ~ dx + \sum_{e \in \E^{\partial}_h}\int_e g\left( \frac{\sigma}{h_e}v - \nabla v \cdot n\right)~ds. \end{equation} Now we can state the DG scheme to be addressed in this paper: \begin{equation} \textup{(DG;$f,g$)}\quad \text{Find}\quad u_h \in V_h \quad \text{ s.t. }\quad a(u_h,\chi) = F(\chi) \quad {}^\forall \chi \in V_h. \label{eq:dg} \end{equation} This scheme is usually called the symmetric interior penalty DG (SIPDG) method, and the $L^2$ theory is well-developed at present (see \cite{MR1885715}). For example, the DG bilinear form $a$ is continuous in the sense that, for any $1 \le p \le \infty$, there exists $C>0$ satisfying \begin{equation} a(u,v) \le C \norm{u}_{V^p}\norm{v}_{V^{p'}} \quad {}^\forall u \in V^p, {}^\forall v \in V^{p'}. \label{eq:dgconti} \end{equation} Moreover, there exists $\sigma_0>0$ such that, if $\sigma\ge \sigma_0$, then we have \begin{equation} a(\chi,\chi) \ge C \norm{\chi}_{V^2}^2 \quad {}^\forall \chi \in V_h. \label{eq:dgcoercive} \end{equation} Consequently, \textup{(DG;$f,g$)} with $\sigma\ge \sigma_0$ admits a unique solution $u_h \in V_h$ and, it satisfies \[ \|u_h\|_{V^2} \le \sup_{\chi\in V_h}\frac{F(\chi)}{\|\chi\|_{V^2}}. \] If the solution $u$ of \textup{(BVP;$f,g$)} belongs to $u \in H^s(\Omega)$ for some $s > \frac{3}{2}$, we have \begin{equation} a(u,v) = F(v) \quad {}^\forall v \in V^2. \label{eq:dgconsistency} \end{equation} As a result, we have the Galerkin orthogonality (consistency) \begin{equation} a(u-u_h,\chi) = 0 \quad {}^\forall \chi \in V_h. \label{eq:galerkin} \end{equation} Our main theorem below will be formulated using the pair of functions $u\in V^\infty$ and $u_h\in V_h$ satisfying \eqref{eq:galerkin}. More generally, we consider $u\in V^\infty$ and $u_h\in V_h$ satisfying \begin{equation} a(u-u_h,\chi) = 0 \quad {}^\forall \chi \in \mathring V_h. \label{eq:go} \end{equation} Below, we always assume that $\sigma\ge \sigma_0$. We are now in a position to state the main results of this paper, but to do so, we need additional notations. Suppose that we are given an open disk $D \Subset \Omega$ with center $x_0$ and radius $R$. In the disk $D$, we consider an auxiliary Neumann problem: \begin{equation} \left\{ \begin{array}{rcc} -\Delta w + w = \varphi & \text{in} & D \\ \partial_n w = 0 & \text{on} & \partial D, \end{array}\right. \label{eq:circle} \end{equation} where $\partial_n=n\cdot \nabla$ denotes the outward normal derivative to $\partial D$. Because $\partial D$ is smooth, for a given $\varphi\in L^2(D)$, there exists a unique solution $w\in H^2(D)$ of \eqref{eq:circle}; this correspondence is denoted by $w=\mathcal{G}_D\varphi$. We recall the $W^{2,p}$ and $W^{1,1}$ regularity results in Section \ref{sect2}. The DG bilinear form $a_D^1$ corresponding to \eqref{eq:circle} is introduced as \begin{multline} a_D^1(u,v) \coloneqq \sum_{K \in \mathcal{T}_h} \int_{K \cap D}( \nabla u \cdot \nabla v + uv)~dx \\ -\sum_{e \in \mathcal{E}_h} \int_{e\cap \overline D}(\mean{\nabla u}\jump{v} + \mean{\nabla v}\jump{u})~ds + \sum_{e \in \mathcal{E}_h}\frac{\sigma}{h_e}\int_{e\cap \overline D} \jump{u}\jump{v} ~ds. \label{eq:bilinearoncircle} \end{multline} We introduce the operator $\Pi^1_h$ of $L^2(D)\to V_h(D)$ as \begin{equation} a_D^1(\mathcal{G}_D\varphi - \Pi_h^1 \varphi , \chi) = 0 \quad {}^\forall \chi \in V_h(D) \label{eq:circlegalerkin} \end{equation} and make the assumption below: \begin{asm}\label{asm1} There exist a function $\alpha$ of $\mathbb{R}_+=(0,\infty)\to \mathbb{R}_+$ and constant $C>0$ which are independent of $h$ such that \begin{gather} \mbox{$\alpha$ is bounded in a neighborhood of $0$;}\label{eq:asm1a}\\ a_D^1(\mathcal{G}_D^1 \varphi -\Pi_h^1 \varphi,v) \le C h\alpha(h) \norm{\varphi}_{L^2(D)}\norm{v}_{V^{\infty}(D)} \quad {}^\forall \varphi \in L^2(D). {}^\forall v \in V^{\infty}(D) \label{eq:asm1} \end{gather} for a sufficiently small $h$. \end{asm} \begin{rem} In view of \eqref{eq:circleconti} and \eqref{eq:circlel1} of Proposition \ref{prop:D}, we can take at least $\alpha(h) = 1$. \end{rem} Using this $\alpha(h)$, we define $\norm{\cdot}_{\alpha(h),\Omega_0}$ as \begin{equation} \norm{v}_{\alpha(h),\mathcal{O}} \coloneqq \norm{v}_{L^\infty(\mathcal{O})} + \alpha(h)\norm{v}_{V^\infty(\mathcal{O})} \end{equation} for $\mathcal{O}\subset \Omega$ or $\mathcal{O}=\Omega$. Our first result is the following interior error estimate in the $L^\infty$ norm. \begin{thm}[$L^\infty$ interior error estimate]\label{thm1} Letting $u \in V^\infty,u_h \in V_h$ satisfy \eqref{eq:go} and supposing that $\kappa>0$ and open sets $\Omega_0 \subset \Omega_1 \subset \Omega$ satisfy $\operatorname{dist}(\Omega_0,\partial\Omega_1) \ge \kappa h$, then we have, under Assumption \ref{asm1}, \begin{equation} \norm{u-u_h}_{L^\infty(\Omega_0)} \le C \left(\inf_{\chi \in V_h}\norm{u-\chi}_{\alpha(h),\Omega_1}+\norm{u-u_h}_{L^2(\Omega_1)}\right) \label{eq:localerror} \end{equation} for a sufficiently small $h$. \end{thm} The next result is the weak discrete maximum principle. \begin{thm}[Weak discrete maximum principle]\label{thm2} Supposing that Assumption \ref{asm1} is satisfied and letting $u_h \in V_h$ be the discrete harmonic function, i.e., \begin{equation} a(u_h,\chi) = 0 \quad{}^\forall \chi \in \mathring V_h, \label{eq:discreteharmonic} \end{equation} then we have \begin{equation} \norm{u_h}_{L^\infty(\Omega)} \le C \norm{u_h}_{ L^\infty(\partial \Omega)} \label{eq:dmp} \end{equation} for a sufficiently small $h$. \end{thm} To state the final $L^\infty$ error estimate, we make the following assumption on triangulations. \begin{asm}\label{asm2} There exist a convex polygonal domain $\widetilde \Omega \Supset \Omega$ and its triangulation $\widetilde \mathcal{T}_h$ such that $\mathcal{T}_h$ is the restriction of $\widetilde \mathcal{T}_h$ to $\Omega$, and that \eqref{eq:triangulation} holds for any $\widetilde K \in \widetilde \mathcal{T}_h\in \{\widetilde \mathcal{T}_h\}_h$ with the same constant $C_*$. \end{asm} We define $\widetilde \mathcal{E}_h$, $V^p(\widetilde \Omega)$, $V_h(\widetilde \Omega)$ similarly, using $\widetilde \mathcal{T}_h$. \begin{thm}[$L^\infty$ error estimate]\label{thm3} Letting $u \in V^\infty,u_h \in V_h$ satisfy \eqref{eq:go}, and supposing that Assumptions \ref{asm1} and \ref{asm2} are satisfied, then we have \begin{equation} \norm{u-u_h}_{L^\infty(\Omega)} \le C \left(\inf_{\chi \in V_h}\norm{u-\chi}_{\alpha(h),\Omega} + \norm{u-u_h}_{L^\infty(\partial \Omega)}\right) \label{eq:thm3} \end{equation} for a sufficiently small $h$. \end{thm} \section{Preliminaries} \label{sect2} In this section, we collect some preliminary results. \subsection{Some local estimates} \label{sec:l2} For $x_0 \in \mathbb{R}^2$ and $d>0$, we denote $B_d(x_0)$ as an open disk with center $x_0$ and radius $d$. We define $N_d(\Omega_0) \coloneqq \{ x \in \overline\Omega \colon \operatorname{dist}{(x,\Omega_0)} < d \}$ and $S_d(x_0) \coloneqq N_d(\{x_0\}) = \overline\Omega \cap B_d(x_0)$. For $\Omega_0 \subset \Omega_1 \subset \Omega$, we define \[ d(\Omega_0,\Omega_1) \coloneqq \operatorname{dist}{(\Omega_0,\partial \Omega_1)},\quad d_{\Omega}(\Omega_0,\Omega_1) \coloneqq \operatorname{dist}{(\Omega_0,\partial \Omega_1 \setminus \partial\Omega}). \] If $d_{\Omega}(\Omega_0,\Omega_1) \ge d$, then we have $N_d(\Omega_0) \subset \Omega_1$. First, we recall further regularity results for the solution of \textup{(BVP;$f,0$)}; see \cite[Lemma 1.2, Lemma 1.3]{MR551291} for more detail. \begin{prop} \label{prop:poisson2} Under the same settings of Proposition \ref{prop:poisson}, we have the following. \begin{enumerate} \item[\textup{(iii)}] \label{prop:poisson3} Assume that $f \in L^2(\Omega)$ and $\operatorname{supp}{f} \subset S_d(x_0)$ for some $d > 0$ and $x_0 \in \overline \Omega$ with $\operatorname{dist}(x_0,\partial \Omega) \le d$. Then, we have \begin{equation} \abs{u}_{H^1(\Omega)} \le C d \norm{f}_{L^2(S_d(x_0))}. \label{eq:localstability} \end{equation} \item[\textup{(iv)}] \label{prop:poisson4} Assume that $\Omega_0 \subset \Omega_1 \subset \Omega$ and $d>0$ with $d_{\Omega}(\Omega_0,\Omega_1) \ge d$. Then, there exists a positive constant $C$ such that \begin{equation} \abs{u}_{W^{2,p}(\Omega_0)} \le C(\abs{f}_{L^p(\Omega_1)}+d^{-1}\abs{u}_{W^{1,p}(\Omega_1)}+d^{-2}\norm{u}_{L^p(\Omega_1)}) \label{interiorstability} \end{equation} under the same assumption of \textup{(i)} or \textup{(ii)} in Proposition \ref{prop:poisson}. \end{enumerate} \end{prop} A version of the Poincar\'e inequality is available (see \cite[Lemma 1.1]{MR551291}). \begin{prop}\label{prop:poincare} Let $\Omega$ be a simply connected polygonal domain. Then, there exists a positive constant $C$ satisfying \begin{equation} \norm{v}_{L^2(S_d(x_0))} \le C d\abs{v}_{H^1(S_d(x_0))} \quad {}^\forall v \in H^1_0(\Omega)\label{eq:poincare} \end{equation} for all $x_0 \in \partial \Omega$ and $d>0$. \end{prop} \begin{prop}\label{prop:holder} For a domain $\Omega_0 \subset \Omega$ and $1 \le p < q \le \infty$, we have $V^q(\Omega_0) \subset V^p(\Omega_0)$. In particular, there exists a positive constant $C$ independent of $h$ and $\Omega_0$ satisfying \begin{equation} \norm{v}_{V^p(\Omega_0)} \le C \abs{\Omega_0}_2^{\frac{1}{p}-\frac{1}{q}} \norm{v}_{V^q(\Omega_0)} \label{eq:holder} \end{equation} for a sufficiently small $h$. \end{prop} \begin{proof} Using H\"older's inequality, we have \begin{align*} \norm{v}_{V^p(\Omega_0)}^p &\le C \abs{\Omega_0}_2^{\frac{q-p}{q}}\sum_{K \in \mathcal{T}_h} \norm{v}_{W^{1,q}(K \cap \Omega_0)}^p \\ &+ \left(\sum_{e \in \mathcal{E}_h} h_e \abs{e\cap\overline\Omega_0}_1\right)^{\frac{q-p}{q}}\left(\sum_{e \in \mathcal{E}_h} h_e^{1-q}\norm{\jump{v}}_{L^q(e \cap \overline \Omega_0 )}^p + \sum_{e \in \mathcal{E}_h} h_e \norm{\mean{\nabla v}}_{L^q(e \cap \overline \Omega_0)}^p\right). \end{align*} Therefore, \begin{align*} \sum_{e \in \mathcal{E}_h} h_e \abs{e\cap\overline\Omega_0}_1 \le \sum_{K \in \mathcal{T}_h(\Omega_0)}\abs{K}_2 \le C\abs{\Omega_0}_2, \end{align*} where $\mathcal{T}_h(\Omega_0) \coloneqq \{K \in \mathcal{T}_h \colon \overline K \cap \overline \Omega_0 \ne \emptyset\}$. Consequently, \eqref{eq:holder} follows. \end{proof} For $\mathcal{O}\subset\Omega$, we denote broken Sobolev space $W^{j,p}_h(\mathcal{O})$ as \[ W^{j,p}_h(\mathcal{O})=W^{j,p}_h(\mathcal{O},\mathcal{T}_h) \coloneqq\{v \in L^p(\mathcal{O}) \colon v|_{K\cap \mathcal{O}} \in W^{j,p}(K\cap \mathcal{O})\} \] equipped with the norm \[ \norm{v}_{W^{j,p}_h(\mathcal{O})} = \left(\sum_{K \in \mathcal{T}_h} \norm{v}_{W^{j,p}( K \cap \mathcal{O})}^p\right)^{{1}/{p}}. \] The following results are available; see \cite[Chapter 3]{MR0520174}, \cite[Propositions 2.1 and 2.2]{MR2113680} and \cite[Proposition 2.2]{MR0431753}. \begin{prop}\label{prop:bestapprx} Let $1 \le p \le \infty$, and $0 \le i \le 1 \le j \le 1 + r$. Assume that $\kappa>0$ and open sets $\Omega_0 \subset \Omega_1 \subset \Omega$ satisfy $d_{\Omega}(\Omega_0,\Omega_1)\ge\kappa h$. Then, there exists positive constant $C$ independent of $h$ such that for $v \in W^{j,p}_h(\Omega_1)$, $\chi \in V_h(\Omega_1)$ exists and satisfies \begin{equation} \norm{v-\chi}_{W^{i,p}_h(\Omega_0)} \le C h^{j-i}\norm{v}_{W^{j,p}_h(\Omega_1)}. \label{eq:bestapprx} \end{equation} \end{prop} \begin{prop}\label{prop:inverseineq} Let $1 \le p \le \infty$, Assume that $\kappa>0$ and open sets $\Omega_0 \subset \Omega_1 \subset \Omega$ satisfy $d_{\Omega}(\Omega_0,\Omega_1)\ge\kappa h$. Then, there exists a positive constant $C$ independent of $h$ satisfying \begin{equation} \norm{v_h}_{V^p(\Omega_0)} \le C h^{-1} \norm{v_h}_{L^p(\Omega_1)} \label{eq:inverseineq} \end{equation} for $v_h \in V_h(\Omega_1)$. \end{prop} \begin{prop}\label{prop:superapprx} Let open sets $\Omega_1 \Subset \Omega_2 \Subset \Omega_3 \Subset\Omega_4 \Subset \Omega$. Then, there exists a positive constant $C$ independent of $h$, and the following property holds for a sufficiently small $h$. For each $\eta_h \in V_h(\Omega_4)$, there exists $\chi \in \mathring V_h(\Omega_3)$ satisfying $\chi \equiv \eta_h$ on $\Omega_2$ and \begin{equation} \norm{\eta_h-\chi}_{V^2(\Omega_3)} \le C \norm{\eta_h}_{V^2(\Omega_4\setminus \Omega_1)}. \end{equation} \end{prop} \subsection{$L^2$ theory for DG method} \label{sec:l2dg} The following results, Propositions \ref{prop:energyerror} and \ref{prop:interiorl2}, are well-known (see \cite[\S 4]{MR1885715} and \cite[\S 3 and \S 4]{MR2113680}). \begin{prop}\label{prop:energyerror} For $f\in L^2(\Omega)$ and $g\in H^{1/2}(\Gamma)$, there exists a unique solution $u_h \in V_h$ of DG scheme \textup{(DG;$f,g$)}. In addition, if the solution $u$ of \eqref{eq:poisson} belongs to $H^s(\Omega)$ with $s>\frac{3}{2}$, then, we have \begin{equation} \norm{u-u_h}_{V_2(\Omega)} \le C \inf_{\chi \in V_h}\norm{u-\chi}_{V_2(\Omega)}. \label{eq:energyestimate} \end{equation} Moreover, if $u \in H^{r+1}(\Omega)$, we have \begin{equation} \|u-u_h\|_{L^2(\Omega)}+h\norm{u-u_h}_{V_2(\Omega)} \le C h^{r+1}|u|_{H^2(\Omega)}. \label{eq:energyestimate1} \end{equation} \end{prop} \begin{prop}\label{prop:interiorl2} Assume that $\kappa>0$ and open sets $\Omega_0 \subset \Omega_1 \subset \Omega$ satisfy $d = d_{\Omega}(\Omega_0,\Omega_1) \ge \kappa h$. We set $S_h = V_h$ or $\mathring V_h$. If $S_h = V_h$, we also assume that $d(\Omega_0,\Omega_1) >0$. If $u \in H^1(\Omega)$ and $u_h \in S_h$ satisfy \[ a(u-u_h,\chi) = 0 \quad {}^\forall \chi \in \mathring V_h \] Then, there exists a positive constant $C_1$ independent of $h$, $u$, and $u_h$ satisfying \begin{equation} \norm{u-u_h}_{V^2(\Omega_0)} \le C_1\left( \inf_{\chi \in S_h} \norm{u-\chi}_{V^2(\Omega_1)} + \norm{u-u_h}_{L^2(\Omega_1)}\right). \label{dg:interiorerror1} \end{equation} Moreover, if $u \in H^{1+r}(\Omega)$, there exists a positive constant $C_2$ independent of $h$, $u$, $u_h$, and $d$ satisfying \begin{equation} \norm{u-u_h}_{V^2(\Omega_0)} \le C_2\left(h^r \norm{u}_{H^{1+r}(\Omega_1)}+d^{-1}\norm{u-u_h}_{L^2(\Omega_1)}\right). \label{eq:interiorerror2} \end{equation} \end{prop} \subsection{Estimates on the disk} \label{sec:disk} We here state some estimates for functions defined on the open disk $D$ with center $x_0$ and radius $R$. Recall that we consider the Neumann boundary value problem \eqref{eq:circle} and the corresponding bilinear form $a_D^1$ defined as \eqref{eq:bilinearoncircle}. For the positive constant $c$, we denote by $cD$ an open disk with center $x_0$ and radius $cR$. The following property (i) is well-known (see \cite{MR0188615} for example). However, we can find no explicit reference to (ii), and we prove it by essentially the same way as \cite{MR0336050}. \begin{prop}\label{prop:circleregulality} \begin{enumerate} \item[\textup{(i)}] If $f \in L^p(D)$ for some $1 < p < \infty$, then the solution $u \in W^{2,p}(D)$ of \eqref{eq:circle} exists and satisfies \begin{equation} \norm{u}_{W^{2,p}(D)} \le C \norm{f}_{L^p(D)}. \label{eq:circleregulality} \end{equation} \item[\textup{(ii)}] If $f \in L^1(D)$, then the weak solution $u \in W^{1,1}(D)$ of \eqref{eq:circle} exists and satisfies \begin{equation} \norm{u}_{W^{1,1}(D)} \le C \norm{f}_{L^1(D)}. \label{eq:circlel1} \end{equation} \end{enumerate} \end{prop} \begin{prop} \label{prop:D} \textup{(i)} Consistency. If the solution $u$ of \eqref{eq:circle} belongs to $u \in H^s(D)$ with $s > \frac{3}{2}$, then we have \begin{equation} a_D^1(u,v) = (f,v)_D \quad {}^\forall v \in V^2(D). \label{eq:circleconsistency} \end{equation} \textup{(ii)} Continuity. For $1 \le p \le \infty$, we have \begin{equation} a_D^1(u,v) \le C \norm{u}_{V^p(D)}\norm{v}_{V^{p'}(D)} \quad {}^\forall u \in V^p(D), {}^\forall v \in V^{p'}(D). \label{eq:circleconti} \end{equation} \textup{(iii)} Coercivity. \begin{equation} a_D^1(\chi,\chi) \ge C \norm{\chi}_{V^2(D)}^2 \quad {}^\forall \chi \in V_h(D). \label{eq:circlecoercive} \end{equation} \end{prop} \begin{lem}\label{lem1} Assume that $\tilde u \in V^\infty(D)$ satisfies $\operatorname{supp} \tilde u \subset \frac{1}{2}D$. Let $\tilde u_h \in V_h(D)$ satisfy \[ a_D^1(\tilde u - \tilde u_h ,\chi) = 0 \quad {}^\forall \chi \in V_h(D). \] Then, under Assumption \ref{asm1}, there exists a positive constant $C$ independent of $h$ and $\tilde u$ satisfying \begin{equation} \norm{\tilde u - \tilde u_h}_{L^\infty(\frac{1}{4}D)} \le C \norm{\tilde u}_{\alpha(h),D} \label{eq:lem1} \end{equation} for a sufficiently small $h$. \end{lem} \begin{proof} We take $x_1 \in \frac{1}{4}D$ such that $\abs{\tilde u(x_1)-\tilde u_h(x_1)} = \norm{\tilde{u}-\tilde{u}_h}_{L^\infty(\frac{1}{4}D)}$. For a sufficiently large $M$, we denote by $D_h \subset D$ the open disk with center at $x_1$ and radius $Mh$. Then, we have \begin{align*} \abs{\tilde u(x_1)-\tilde u_h (x_1)} & \le \abs{\tilde u(x_1)} + \abs{\tilde u_h(x_1)} \\ & \le \abs{\tilde u (x_1)} + Ch^{-1}\abs{\tilde u_h}_{L^2(D_h)} \\ & \le C\norm{\tilde u}_{L^\infty(D_h)} + Ch^{-1}\norm{\tilde u - \tilde u_h}_{L^2(D_h)}. \end{align*} For $\phi \in C_0^\infty(D_h)$, we set $v = \mathcal{G}_D \phi$ and $v_h = \Pi_h^1 \phi$. Then, \begin{align*} (\tilde u - \tilde u_h,\phi)_{D_h} &= a_D^1(\tilde u -\tilde u_h , v) = a_D^1(\tilde u - \tilde u_h,v-v_h) =a_D^1(\tilde u ,v-v_h)\\ & \le C h \alpha(h)\norm{\tilde u}_{V^\infty(D)}\norm{\phi}_{L^2(D_h)}. \end{align*} Therefore, we deduce $\norm{\tilde u -\tilde u_h}_{L^2(D_h)} \le Ch\alpha(h) \norm{\tilde u}_{V^\infty(D)}$, which implies \eqref{eq:lem1}. \end{proof} \begin{lem}\label{lem2} Assume that $w_h \in V_h(D)$ satisfies \[ a_D^1(w_h,\chi) = 0 \quad {}^\forall \chi \in \mathring V_h(D). \] Then, there exists a positive constant $C$ independent of $h$ and $w_h$ satisfying \begin{equation} \abs{w_h(x_0)} \le C \norm{w_h}_{L^2(D)} \label{eq:lem2} \end{equation} for a sufficiently small $h$. \end{lem} \begin{proof} In view of Proposition \ref{prop:superapprx}, there exists $\eta \in \mathring V_h(\frac{3}{4}D)$ satisfying $\eta_h \equiv w_h$ on $\frac{1}{2}D$ and $\norm{\eta_h}_{V^2(\frac{3}{4}D)} \le C \norm{w_h}_{V^2(D)}$. For a sufficiently large $M$, let $D_h \subset D$ be the disk with center at $x_0$ and radius $Mh$. Then, \begin{align} \abs{w_h(x_0)} = \abs{\eta_h(x_0)} \le C h^{-1} \norm{\eta_h}_{L^2(D_h)}. \label{eq:lem2first} \end{align} For $\phi \in C_0^\infty(D_h)$, we set $v = \mathcal{G}_D \phi$ and $v_h = \Pi_h^1 \phi$. According to Proposition \ref{prop:superapprx}, there exists $\chi_h \in \mathring V_h(\frac{1}{2}D)$ satisfying $\chi_h \equiv v_h$ on $\frac{1}{4}D$ and $\norm{v_h-\chi_h}_{V^2(\frac{3}{4}D)} \le C \norm{v_h}_{V^2(\frac{3}{4}D\setminus\frac{1}{8}D)}$. We can estimate this as \begin{align} (\eta_h,\phi)_{D_h} &= a_D^1(\eta_h,v) = a_D^1(\eta_h,v_h) = a_D^1(\eta_h,v_h-\chi_h) \nonumber \\ & \le C \norm{\eta_h}_{V^2(\frac{3}{4}D)} \norm{v_h-\chi_h}_{V^2(\frac{3}{4}D)} \nonumber \\ & \le C \norm{w_h}_{V^2(D)}\norm{v_h}_{V^2(\frac{3}{4}D\setminus\frac{1}{8}D)}. \end{align} Because $a_D^1(v_h,\chi) = (\phi,\chi)_D = 0$ for $\chi \in \mathring V_h(D\setminus D_h)$, using Proposition \ref{prop:interiorl2} and the Sobolev inequality, we obtain \begin{align} \norm{v_h}_{V^2(\frac{3}{4}D \setminus \frac{1}{8}D)} &\le C \norm{v_h}_{L^2(D)} \nonumber\\ & \le C\left(\norm{v-v_h}_{W^{1,1}_h(D)} + \norm{v}_{W^{1,1}(D)}\right), \end{align} where $\norm{v}_{W^{1,1}_h(D)} \coloneqq \sum_{K \in \mathcal{T}_h} \norm{v}_{W^{1,1}(K \cap D)}$. By Propositions \ref{prop:energyerror} and \eqref{eq:circleregulality}, we have \begin{align} \norm{v-v_h}_{W^{1,1}_h(D)} &\le C\left(\sum_{K \in \mathcal{T}_h}\norm{v-v_h}_{H^1(K\cap D)}^2\right)^{1/2} \nonumber \\ & \le C h \norm{v}_{H^2(D)} \nonumber \\ & \le C h \norm{\phi}_{L^2(D_h)}. \end{align} Using Proposition \ref{prop:circleregulality}, we have \begin{equation} \norm{v}_{W^{1,1}(D)} \le C \norm{\phi}_{L^1(D_h)} \le Ch\norm{\phi}_{L^2(D_h)}. \label{eq:lem2last} \end{equation} From \eqref{eq:lem2first}--\eqref{eq:lem2last}, we obtain \eqref{eq:lem2}. \end{proof} \section{Interior error estimates (Proof of Theorem \ref{thm1})} \label{s:I} We first consider the homogeneous Neumann boundary value problem in $\Omega$. Set $a^1(u,v) \coloneqq a(u,v) + (u,v)_\Omega$. \begin{lem}\label{thm1:coercive} Assume that $\kappa>0$ and open sets $\Omega_0 \subset \Omega_1 \subset \Omega$ satisfy $d = d(\Omega_0,\Omega_1) \ge \kappa h$. Let $u \in V^\infty$ and $u_h \in V_h$ satisfy \[ a^1(u-u_h,\chi) = 0 \quad {}^\forall \chi \in \mathring V_h. \] Then, under Assumption \ref{asm1}, there exists a positive constant $C$ independent of $h$, $u$, and $u_h$ satisfying \begin{equation} \norm{u-u_h}_{L^\infty(\Omega_0)} \le C \left(\inf_{\chi \in V_h}\norm{u-\chi}_{\alpha(h),\Omega_1}+\norm{u-u_h}_{L^2(\Omega_1)}\right) \label{eq:coercivelocalerror} \end{equation} for a sufficiently small $h$. \end{lem} \begin{proof} Letting $\chi\in V_h$ be arbitrary, we set $v=u-\chi$ and $v_h=u_h-\chi$. We take $x_0 \in \Omega_0$ such that $\abs{v(x_0)-v_h(x_0)}= \norm{v-v_h}_{L^\infty(\Omega_0)}= \norm{u-u_h}_{L^\infty(\Omega_0)}$. Let $D\Subset \Omega_1$ be an open disk with center at $x_0$ and radius $R < d$. Because $d \ge \kappa h$, we can take $R$ independent of $x_0$. Letting $\omega \in C^\infty_0(\frac{1}{2}D)$ satisfy $0 \le \omega \le 1 $ and $\omega \equiv 1$ on $\frac{1}{4}D$, we set $\tilde v = \omega v \in V^\infty(D)$. We define $\tilde v_h \in V_h(D)$ as \[ a_D^1(\tilde v - \tilde v_h ,\xi) = 0 \quad {}^\forall \xi \in V_h(D). \] Then, $a_D^1(\tilde v_h - v_h,\eta_h) = a^1(v-v_h,\eta_h) = 0$ for $\eta_h \in \mathring V_h(\frac{1}{4}D)$, because $v=v_h$ in $\operatorname{supp} \eta_h\subset \frac14 D$. Using Lemmas \ref{lem2} and \ref{lem1}, we have \begin{align*} \abs{\tilde v_h(x_0) - v_h(x_0)} &\le C \norm{\tilde v_h -v_h}_{L^2(\frac{1}{4}D)} \\ & \le C \norm{\tilde v -\tilde v_h}_{L^\infty(\frac{1}{4}D)} + C \norm{v-v_h}_{L^2(\frac{1}{4}D)} \\ & \le C \norm{\tilde v}_{\alpha(h),\frac{1}{2}D} + C \norm{v-v_h}_{L^2(\frac14 D)}\\ & \le C \norm{v}_{\alpha(h),D} + C \norm{v-v_h}_{L^2(\frac14 D)}. \end{align*} Therefore, \begin{align*} \norm{u-u_h}_{L^\infty(\Omega_0)}&= \abs{v(x_0)-v_h(x_0)} \\ & \le \abs{\tilde v(x_0)-\tilde v_h(x_0)} + \abs{\tilde v_h(x_0) - v_h(x_0)}\\ & \le C \norm{v}_{\alpha(h),D} + C \norm{v-v_h}_{L^2(\frac14 D)}\\ &\le C\norm{u-\chi}_{\alpha(h),\Omega_1} + C\norm{u-u_h}_{L^2(\Omega_1)}. \end{align*} \end{proof} \begin{lem}\label{lem3} Assume that $D' \Subset D \subset \Omega$ are open disks with the same center. Let $u \in V^\infty(D)$ and $u_h \in V_h(D)$ satisfy \[ a_D(u-u_h,\chi)=0\quad{}^\forall \chi \in \mathring V_h(D) \] where $a_D(v,w) = a_D^1(v,w)-(v,w)_D$. Then, letting Assumption \ref{asm1} be satisfied, we have, for $p$ and $q$ which satisfy $2 \le q < p \le \infty$ and $\frac{1}{q} < \frac{1}{p} + \frac{1}{2} \displaystyle$ \begin{equation} \norm{u-u_h}_{L^p(D')} \le C \left(\norm{u}_{\alpha(h),D} + \norm{u-u_h}_{L^q(D)}\right) \label{eq:lem3} \end{equation} for a sufficiently small $h$. \end{lem} \begin{proof} Setting $\psi = \mathcal{G}_D(u-u_h)$, $\psi_h = \Pi_h^1(u-u_h)$, we have \[ a_D^1(u-u_h-\psi_h,\chi) = 0 \quad{}^\forall \chi \in \mathring V_h(D). \] We apply Lemma \ref{thm1:coercive} to obtain \begin{equation*} \norm{u-u_h-\psi_h}_{L^\infty(D')} \le C\norm{u}_{\alpha(h),D} + C\norm{u-u_h}_{L^2(D)} + C\norm{\psi_h}_{L^2(D)}. \end{equation*} On the other hand, using \eqref{eq:circlecoercive}, we have \begin{equation*} \norm{\psi_h}_{L^2(D)}^2 \le C a_D(\psi_h,\psi_h)= C (u-u_h,\psi_h)_{D} \le C\norm{u-u_h}_{L^2(D)}\norm{\psi_h}_{L^2(D)} \end{equation*} and, therefore, \begin{align} \norm{u-u_h}_{L^p(D')}& \le \norm{u-u_h-\psi_h}_{L^p(D')} + \norm{\psi_h}_{L^p(D')} \nonumber \\ & \le C \norm{u}_{\alpha(h),D} + C\norm{u-u_h}_{L^2(D)} + \norm{\psi_h}_{L^p(D')}. \label{thm1coercive:second} \end{align} Because $\psi$ is smooth, we again apply Lemma \ref{thm1:coercive} to obtain \begin{align*} \norm{\psi_h}_{L^\infty(D')} & \le \norm{\psi-\psi_h}_{L^\infty(D')} + \norm{\psi}_{L^\infty(D')} \\ & \le \norm{\psi}_{\alpha(h),D} + C\norm{\psi-\psi_h}_{L^2(D)} \\ &\le C \norm{\psi}_{\alpha(h),D} \\ &\le C\left(\norm{\psi}_{W^{1,\infty}(D)} + \max_{e \in \mathcal{E}_h,e\cap \overline D \ne \emptyset}\norm{\nabla \psi}_{L^\infty(e\cap\overline D)}\right). \end{align*} The Sobolev inequality and elliptic regularity give \[ \norm{\psi}_{W^{1,\infty}(D)} \le C \norm{\psi}_{W^{2,s}(D)} \le C \norm{u-u_h}_{L^s(D)} \] for $s >2$. Because \[ \abs{\nabla \psi (x)} \le C \int_D \frac{\abs{u(y)-u_h(y)}}{\abs{x-y}} dy \le C \norm{u-u_h}_{L^s(D)} \] for $x \in D$, we have \begin{equation} \norm{\psi_h}_{L^\infty(D')} \le C \norm{u-u_h}_{L^s(D)}. \label{thm1:linfty} \end{equation} Similarly, we deduce \begin{align*} \norm{\psi_h}_{L^2(D')} &\le C \norm{\psi}_{V^2(D)} \\ & \le C\norm{\psi}_{H^1(D)} + C \left( \sum_{e \in \mathcal{E}_h,e\cap \overline D \ne \emptyset} h_e\norm{\nabla \psi}_{L^2(e\cap\overline D)}^2\right)^{1/2}. \end{align*} Applying the Young inequality for convolution, we have \begin{align*} \abs{\nabla \psi(x)} &\le C \int_D\frac{\abs{u(y)-u_h(y)}}{\abs{x-y}} dy \\ &\le C \norm{u-u_h}_{L^t(D)} \norm{\abs{x-y}^{-1}}_{L^{2t/(3t-2)}(D)} \\ & \le C \norm{u-u_h}_{L^t(D)} . \end{align*} for $x \in D$ and $1<t<2$. Therefore, \begin{align} \norm{\psi_h}_{L^2(D')} &\le C\norm{\psi}_{W^{2,t}(D)} + C \left( \sum_{e \in \mathcal{E}_h,e\cap \overline D \ne \emptyset}h_e^2\norm{u-u_h}_{L^t(D)}^2\right)^{1/2} \nonumber \\ & \le C \norm{u-u_h}_{L^t(D)}. \label{thm1:l2} \end{align} In view of the Riesz--Thorin interpolation theorem, we have \begin{align} \norm{\psi_h}_{L^p(D')} \le C \norm{u-u_h}_{L^q(D)} \end{align} for $\theta = \frac{2}{p}$, and $\frac{1}{q} = \frac{1-\theta}{s} + \frac{\theta}{t} < \frac{1}{2}+\frac{1}{p}$. This, together with \eqref{thm1coercive:second}, implies \eqref{eq:lem3}. \end{proof} We can now state the following proof. \begin{proof}[Proof of Theorem \ref{thm1}.] Letting $\chi\in V_h$ be arbitrary, we set $v=u-\chi$ and $v_h=u_h-\chi$. Similar to the proof of Lemma \ref{thm1:coercive}, we take $x_0 \in \Omega_0$, $D \Subset \Omega_1$, and $\omega \in C^\infty_0(\frac{1}{2}D)$. Setting $\tilde v = \omega v \in V_h(D)$, we define $\tilde v_h \in V_h(D)$ as \[ a_D^1(\tilde v - \tilde v_h ,\xi) = 0 \quad {}^\forall \xi \in V_h(D). \] Then, Lemma \ref{lem1} gives \[ \norm{\tilde v -\tilde v_h}_{L^\infty(\frac{1}{4}D)} \le C\norm{\tilde v}_{\alpha(h),D}. \] Setting $D' = \varepsilon D$ for $\varepsilon < \frac{1}{4}$, we let $\psi$ be the unique solution of \[ \left\{ \begin{array}{ccc} -\Delta \psi = -(\tilde v_h - \tilde v) & \text{in} & D' \\ \psi = 0 & \text{on} & \partial D' \end{array} \right.. \] Then, we have \[ a_D(\psi,w) = -(\tilde v_h - \tilde v,w)_{D'} \quad {}^\forall w \in \{w \in V^2(D') \colon w|_{\partial D'} = 0 \} \] and \[ a_D(\psi-(v_h-\tilde v_h),\xi) = 0 \quad {}^\forall \xi \in \mathring V_h(D'). \] Applying Lemma \ref{lem3} several times, we obtain \begin{align*} \norm{\psi-(v_h-\tilde v_h)}_{L^\infty(\frac{1}{4}D')} & \le C \norm{\psi}_{\alpha(h),D'} + C \norm{\psi-(v_h-\tilde v_h)}_{L^2(D')} \\ & \le C \norm{\psi}_{\alpha(h),D'} + C \norm{v-v_h}_{L^2(D')} + C \norm{\tilde v - \tilde v_h}_{L^2(D')} \\ & \le C \norm{\psi}_{\alpha(h),D'} + C \norm{v-v_h}_{L^2(D')} + C\norm{\tilde v}_{\alpha(h),D}. \end{align*} Then, we deduce $\norm{\psi}_{\alpha(h),D'} \le C \norm{\tilde v -\tilde v_h}_{L^\infty(D')} \le C\norm{v}_{\alpha(h),D}$ in the similar way as the proof of Lemma \ref{lem3}. Using the triangle inequality, we have \begin{align*} \norm{v_h-\tilde v_h}_{L^\infty(\frac14D')}\le C\norm{v}_{\alpha(h),D} + C\norm{v-v_h}_{L^2(D)}. \end{align*} Therefore, \begin{align*} \norm{u-u_h}_{L^\infty(\Omega_0)} &\le \abs{\tilde v(x_0)-\tilde v_h(x_0)} + \abs{\tilde v_h(x_0) - v_h(x_0)} \\ & \le C\norm{v}_{\alpha(h),D} + C\norm{v-v_h}_{L^2(D)} \\ &\le C\norm{u-\chi}_{\alpha(h),\Omega_1} + C\norm{u-u_h}_{L^2(\Omega_1)}. \end{align*} \end{proof} \begin{cor}\label{cor1} Under the same assumption of Theorem \ref{thm1}, we further assume that $d(\Omega_1,\Omega) \ge \kappa h$ and $u \in W^{1+r,\infty}(\Omega)$. Then, there exists a positive constant $C$ independent of $h$, $u$, $u_h$, and $D$ satisfying \begin{equation} \norm{u-u_h}_{L^\infty(\Omega_0)} \le C h^r\left[h+\alpha\left(\frac{h}{d}\right)\right]\norm{u}_{W^{1+r,\infty}(\Omega_1)} + Cd^{-1}\norm{u-u_h}_{L^2(\Omega_1)} \label{eq:cor1} \end{equation} for a sufficiently small $h$. \end{cor} \section{Weak discrete maximum principle (Proof of Theorem \ref{thm2})} \label{s:II} We follow the same method as the proof of Theorem 1 of \cite{MR551291} to prove Theorem \ref{thm2} described below. \begin{proof}[Proof of Theorem \ref{thm2}.] Let $x_0 \in \Omega$ satisfy $\abs{u_h(x_0)} = \norm{u_h}_{L^\infty(\Omega)}$. Set $d = \operatorname{dist}(x,\partial \Omega)$. First, we consider the case $d \ge 2\kappa h$ for some $\kappa>1$. Then, applying Corollary \ref{cor1} to an open disk with center $x_0$ and $u\equiv 0$, we have \[ \abs{u_h(x_0)} \le C d^{-1}\norm{u_h}_{L^2(S_{\frac{1}{2}d}(x_0))} \le C d^{-1}\norm{u_h}_{L^2(S_{d}(x_0))}. \] Now we assume that $d < 2\kappa h$. Using the inverse inequality, we have \[ \abs{u_h(x_0)} \le Ch^{-1}\norm{u_h}_{L^2(S_{h}(x_0))}. \] Therefore, \begin{equation} \norm{u_h}_{L^\infty(\Omega)} \le C \rho^{-1}\norm{u_h}_{L^2(S_\rho(x_0))} \label{eq:thm2step1} \end{equation} where $\rho = \max\{d,h\}$. Let $\phi \in C^\infty_0(S_\rho(x_0))$ satisfy $\norm{\phi}_{L^2(S_\rho(x_0))}=1$. Let $v \in H^1_0(\Omega)$ be the solution of \eqref{eq:poisson} with $f=\phi$ and $g=0$. Then, $v \in W^{2,p}(\Omega)$ for some $4/3< p\le 2 $, and $a(v,w) = (\phi,w)_\Omega$ for all $w \in V^2$. Let $v_h \in \mathring V_h$ be satisfy \[ a(v_h,\chi) = (\phi,\chi) \quad {}^\forall \chi \in \mathring V_h. \] In view of the assumption of $u_h$, we get \begin{align} \abs{(u_h,\phi)_\Omega} &= \abs{a(u_h,v)} = \abs{a(u_h,v-v_h)} \nonumber \\ & = \abs{a(u_h-\chi,v-v_h)} \label{eq:thm2step2first} \end{align} for $\chi \in \mathring V_h$. We define $\hat u_h \in \mathring V_h$ such that $\hat u_h = 0$ at nodal points on $\partial \Omega$ and $\hat u_h = u_h$ at interior nodal points. Then, we have \begin{align*} \operatorname{supp}(u_h - \hat u_h) & \subset \Lambda_h = \{x \in \overline \Omega \colon \operatorname{dist}(x,\partial \Omega) \le h \},\\ \norm{u_h - \hat u_h}_{L^\infty(\Omega)} & \le C \norm{u_h}_{L^\infty(\partial \Omega)}. \end{align*} Substituting \eqref{eq:thm2step2first} for $\chi = \hat u_h$, and using the inverse inequality, we have \begin{align} \abs{(u_h,\phi)_\Omega} &\le C \norm{u_h-\hat u_h}_{V^\infty} \norm{v-v_h}_{V^1(\Lambda_h)} \nonumber \\ & \le Ch^{-1}\norm{u_h - \hat u_h}_{L^\infty(\Omega)}\norm{v-v_h}_{V^1(\Lambda_h)} \nonumber \\ & \le Ch^{-1}\norm{u_h}_{L^\infty(\partial \Omega)}\norm{v-v_h}_{V^1(\Lambda_h)}. \label{eq:thm2step2last} \end{align} Set $R_0 = \operatorname{diam}\Omega$ and $d_j = R_02^{-j}$ for non-negative integer $j$. We define $A_j$ as \[ A_j \coloneqq \{x \in \overline \Omega \colon d_{j+1} \le \abs{x-x_0} \le d_j \}. \] Then, $\abs{A_j \cap \Lambda_h}_2 \le C d_jh$. Set $A_j^l =\displaystyle \bigcup_{k=j-l}^{j+l}A_k$ and $J \coloneqq \min \{j \in {\mathbb Z} \colon d_{j+1} \le 8 \rho\}$. Then, we have \begin{align} \norm{v-v_h}_{V^1(\Lambda_h)} & \le \sum_{j=0}^J \norm{v-v_h}_{V^1(\Lambda_h \cap A_j)} + \norm{v-v_h}_{V^1(\Lambda_h \cap S_{8\rho})} \nonumber \\ & \le C\sum_{j=0}^J h^{1/2}d_j^{1/2}\norm{v-v_h}_{V^2(\Lambda_h\cap A_j)} + C\rho^{1/2}h^{1/2} \norm{v-v_h}_{V^2(\Lambda_h \cap S_{8\rho}(x_0))}. \label{eq:thm2step3first} \end{align} To estimate the second term of \eqref{eq:thm2step3first}, we apply Propositions \ref{prop:energyerror} and \ref{prop:bestapprx} and get \begin{align} \norm{v-v_h}_{V^2(\Lambda_h \cap S_{8\rho}(x_0))} &\le Ch^{2-\frac{2}{p}}\norm{v}_{W^{2,p}(\Omega)} \nonumber \\ & \le Ch^{2-\frac{2}{p}}\norm{\phi}_{L^p(S_\rho(x_0))} \nonumber \\ & \le Ch^{2-\frac{2}{p}}\rho^{\frac{2}{p}-1} \label{eq:thm2step3circle}. \end{align} Meanwhile, using Proposition \ref{prop:interiorl2} for $j$ satisfying $\Lambda_h \cap A_j \ne \emptyset$, we have \begin{align*} \norm{v-v_h}_{V^2(\Lambda_h \cap A_j)} &\le C\left( \norm{v-v_h}_{V^2(A_j^1)} + d_j^{-1}\norm{v-v_h}_{L^2(A_j^1)} \right) \nonumber \\ & \le Ch^{2-\frac{2}{p}}\norm{v}_{W^{2,p}(A_j^2)} + C d_j^{-1} \norm{v-v_h}_{L^2(A_j^1)}. \end{align*} In view of Proposition \ref{prop:poisson4}, \begin{align*} \norm{v}_{W^{2,p}(A_j^2)} &\le C\left( \norm{\phi}_{L^p(A_j^3)} + d_j^{-1}\abs{v}_{W^{1,p}(A_j^3)} + d_j^{-2}\norm{v}_{L^p(A_j^3)} \right) \\ & \le C d_j^{\frac{2}{p}-1}\left(1 + d_j^{-1}\abs{v}_{H^1(A_j^4)} + d_j^{-2}\norm{v}_{L^2(A_j^4)} \right). \end{align*} Because $\operatorname{diam} A_j^4 \le 32 d_j$ and $\operatorname{dist}(A_j^4,\partial \Omega) \le h$, there exists $\overline x_j \in \partial \Omega$ satisfying \[ A_j^4 \subset S_{64d_j}(\overline x_j)\,,\,S_\rho(x_0) \subset S_{64d_j}(\overline x_j). \] Therefore, we have \begin{align} \norm{v}_{W^{2,p}(A_j^2)} \le C d_j^{\frac{2}{p}-1} \label{eq:thm2step3w2p} \end{align} by Propositions \ref{prop:poisson3} and \ref{prop:poincare}. Let $\eta \in C^\infty_0(S_{64d_j}(\overline x_j))$ and let $w \in H^1_0(\Omega)$ be the solution of \eqref{eq:poisson} with $f=\eta$ and $g=0$. Let $w_h \in \mathring V_h$ satisfy \[ a(w_h,\chi) = (\eta,\chi)_{\Omega} \quad {}^\forall \chi \in \mathring V_h. \] Similar to the above, we deduce \begin{align*} \abs{(v-v_h,\eta)_{S_{64d_j}(\overline x_j)}} & = \abs{a(v-v_h,w-w_h)} \nonumber \\ & \le C \norm{v-v_h}_{V^2}\norm{w-w_h}_{V^2} \nonumber \\ & \le Ch^{4-\frac{4}{p}}\norm{\phi}_{L^p(S_\rho(x_0))}\norm{\eta}_{L^p(S_{64d_j}(\overline x_j))} \nonumber \\ & \le Ch^{4-\frac{4}{p}}(\rho d_j)^{\frac{2}{p}-1}\norm{\eta}_{L^2(S_{64d_j}(\overline x_j))}. \end{align*} Therefore, \begin{align} \norm{v-v_h}_{L^2(A_j^1)} \le Ch^{4-\frac{4}{p}}(\rho d_j)^{\frac{2}{p}-1}. \label{eq:thm2step3l2} \end{align} Summing up \eqref{eq:thm2step3first}--\eqref{eq:thm2step3l2} and using $h \le \rho \le C d_j \le R_0$ and $p>\frac{3}{4}$, we obtain \begin{align} \norm{v-v_h}_{V^1(\Lambda_h)} & \le \sum_{j=0}^J h^{1/2}d_j^{1/2}\left(h^{2-\frac{2}{p}}d_j^{\frac{2}{p}-1} + h^{4-\frac{4}{p}}\rho^{\frac{2}{p}-1}d_j^{\frac{2}{p}-1} \right) + h\rho \left( \frac{h}{\rho} \right)^{\frac{3}{2}-\frac{2}{p}} \nonumber \\ & \le Ch\rho. \label{eq;thm2step3last} \end{align} The desired \eqref{eq:dmp} now follows \eqref{eq:thm2step2last} and \eqref{eq:thm2step1}. \end{proof} \section{$L^\infty$ error estimate (Proof of Theorem \ref{thm3})} \label{s:III} We finally state the following proof. \begin{proof}[Proof of Theorem \ref{thm3}.] Let $\tilde u \in V^\infty(\widetilde \Omega)$ be the extension of $u$ satisfying $\norm{\tilde u}_{\alpha(h), \widetilde \Omega} \le C \norm{u}_{\alpha(h),\Omega}$ and $\tilde u = 0$ on $\partial \widetilde \Omega$. Moreover, let $\tilde u_h \in \mathring V_h(\widetilde \Omega)$ solve \[ \tilde a(\tilde u - \tilde u_h ,\xi) = 0 \quad {}^\forall \xi \in \mathring V_h(\widetilde \Omega), \] where $\tilde a$ is the bilinear form \eqref{eq:bilinear} with replacement of $\mathcal{T}_h$, $\mathcal{E}_h$ by $\widetilde \mathcal{T}_h$, $\widetilde \mathcal{E}_h$, respectively. For arbitrary $\chi \in V_h$, we define $\tilde \chi \in \mathring V_h(\widetilde \Omega)$ as a zero extension. Then, in view of Theorem \ref{thm1}, we have \begin{equation} \norm{\tilde u -\tilde u_h}_{L^\infty(\Omega)} \le C \norm{\tilde u- \tilde \chi}_{\alpha(h),\widetilde \Omega} + C \norm{\tilde u - \tilde u_h}_{L^2(\widetilde \Omega)}. \label{eq:thm3step1} \end{equation} Let $\psi \in H^1(\widetilde \Omega)$ be the solution of \begin{equation} \left\{ \begin{array}{ccc} -\Delta \psi = \tilde u - \tilde u_h & \text{in} & \widetilde \Omega \\ \psi = 0 & \text{on} & \partial \widetilde \Omega . \end{array}\right. \end{equation} Then, $\psi \in H^2(\widetilde \Omega)$ and $\tilde a (\psi,\eta) = (\psi,\eta)_{\widetilde \Omega}$ for $\eta \in V^2(\widetilde \Omega)$. Let $\psi_h \in \mathring V_h(\widetilde \Omega)$ solve \[ \tilde a(\psi-\psi_h,\xi) = 0 \quad {}^\forall \xi \in \mathring V_h(\widetilde \Omega). \] Then, by the continuity of $\tilde a$ and elliptic regularity, we have \begin{align} \norm{\tilde u -\tilde u_h}_{L^2(\widetilde \Omega)}^2 &= \tilde a(\tilde u-\tilde u_h ,\psi) = \tilde a(\tilde u-\tilde \chi,\psi-\psi_h) \nonumber \\ & \le C\norm{\tilde u-\tilde \chi}_{V^\infty(\widetilde \Omega)}\norm{\psi - \psi_h}_{V^1(\widetilde \Omega)} \nonumber \\ & \le Ch \norm{\tilde u-\tilde \chi}_{V^\infty(\widetilde \Omega)}\norm{\tilde u -\tilde u_h}_{L^2(\widetilde \Omega)}. \label{eq:thm3step2first} \end{align} Because $a(u_h - \tilde u_h,\xi) = 0$ for $\xi \in \mathring V_h$, using Theorems \ref{thm2} and \ref{thm1} and \eqref{eq:thm3step2first}, we deduce \begin{align} \norm{u_h - \tilde u_h}_{L^\infty(\Omega)} & \le C\norm{u_h-\tilde u_h}_{L^\infty(\partial \Omega)} \nonumber \\ & \le C\norm{\tilde u - \tilde u_h}_{L^\infty(\partial \Omega)} + C\norm{u- u_h}_{L^\infty(\partial \Omega)} \nonumber \\ & \le C\norm{\tilde u-\tilde \chi}_{\alpha(h),\widetilde \Omega} + C\norm{\tilde u -\tilde u_h}_{L^2(\widetilde \Omega)} + C\norm{u- u_h}_{L^\infty(\partial \Omega)} \nonumber \\ & \le C\norm{\tilde u-\tilde \chi}_{\alpha(h),\widetilde \Omega} + C\norm{u- u_h}_{L^\infty(\partial \Omega)}. \label{eq:thm3step3} \end{align} Therefore, using triangle inequality, we obtain \[ \norm{u-u_h}_{L^\infty(\Omega)} \le C \norm{u-\chi}_{\alpha(h),\Omega} + C\norm{u-u_h}_{L^\infty(\partial \Omega)}. \] \end{proof} \begin{cor}\label{cor2} In addition to the assumption of Theorem \ref{thm3}, we assume $u \in W^{1+r,\infty}(\Omega)$. Then, we have \begin{equation} \norm{u-u_h}_{L^\infty(\Omega)} \le C h^{r}\norm{u}_{W^{1+r,\infty}(\Omega)}.\label{eq:cor2} \end{equation} \end{cor} \begin{proof} First, in view of the standard interpolation error estimate, we have \[ \inf_{\chi \in V_h}\norm{u-\chi}_{\alpha(h),\Omega} \le C h^r(h+\alpha(h))\norm{u}_{W^{1+r,\infty}(\Omega)}. \] To perform the estimation for $\|u-u_h\|_{L^\infty(\partial\Omega)}$, we let $e \in \E^{\partial}_h$ and $K \in \mathcal{T}_h$ such that $e \subset \overline K$. Moreover, let $\chi\in V_h$ be arbitrary. By the inverse inequality, we have \begin{align*} \norm{u-u_h}_{L^\infty(e)} & \le \norm{u-\chi}_{L^\infty(e)} + \norm{u_h-\chi}_{L^\infty(e)} \\ &\le \norm{u-\chi}_{L^\infty(K)} + Ch_e^{-1/2}\norm{u_h-\chi}_{L^2(e)} . \end{align*} Using \eqref{eq:dgconti}, \eqref{eq:dgcoercive}, and \eqref{eq:galerkin}, we have $\|\chi-u_h\|_{V^2}\le C\|\chi-u\|_{V^2}$ and, consequently, \[ h_e^{-1/2} \norm{\chi-u_h}_{L^2(e)}\le C\|\chi-u\|_{V^2}. \] Therefore, \[ \norm{u-u_h}_{L^\infty(e)}\le \norm{u-\chi}_{L^\infty(K)}+C\|\chi-u\|_{V^2}. \] Choosing $\chi$ as the Lagrange interpolation of $u$, we deduce \[ \norm{u-u_h}_{L^\infty(e)}\le Ch^{r}(h+\alpha(h))|u|_{W^{r+1,\infty}(K)}+Ch^r|u|_{W^{r+1,\infty}(\Omega)}. \] Summing up those estimate, we obtain the desired \eqref{eq:cor2}. \end{proof} \section{Numerical examples} \label{s:ne} In this section, we examine the weak discrete maximum principle (Theorem \ref{thm2}) and the $L^\infty$ error estimate (Corollary \ref{cor2}) using numerical examples. We consider the square domain $\Omega$ (see Fig. \ref{fig:squaremesh}) and the L-shape domain $\Omega$ (see Fig. \ref{fig:lshapemesh}). \begin{figure}[bt] \centering \subfloat[][Square domain]{\includegraphics[scale=0.7]{Fig1a.pdf}\label{fig:squaremesh}} \subfloat[][L-shape domain]{\includegraphics[scale=0.7]{Fig1b.pdf}\label{fig:lshapemesh}} \caption{Domain $\Omega$} \end{figure} First, we solve (DG;$f,g$) with $f = 0$ and $g = \cos(\pi x) \cos (\pi y)$; the solution $u_h$ satisfies \eqref{eq:discreteharmonic}. The minimum and maximum values of $u_h$ on $\Omega$ and $\partial \Omega$ are reported in Tab.~\ref{tb:wdmp}. We see from Tab.~\ref{tb:wdmp} that the minimum and maximum values on $\Omega$ agree with those on $\partial \Omega$. We infer that the discrete maximum principle \eqref{eq:dmp} actually holds with $C=1$. \begin{table}[bt] \caption{Minimum and Maximum on $\Omega$ or $\partial \Omega$} \centering \begin{tabular}{lc|cc|cc}\hline Domain & $h$ & $\displaystyle \min_{\Omega}u_h$ & $\displaystyle \min_{\partial \Omega}u_h$ & $\displaystyle \max_{\Omega}u_h$ & $\displaystyle \max_{\partial \Omega}u_h$ \\ \hline\hline \multirow{2}{*}{Square} & $0.152069063$ & $-1.01415829$ & $-1.01415829$ & $1.01407799$ & $1.01407799$ \\ \cline{2-6} & $0.0762297934$ & $-1.00438815$&$-1.00438815$&$1.00437510$&$1.00437510$ \\ \hline \multirow{2}{*}{L-shape} & $0.152069063$ & $-1.01406865$ & $-1.01406865$ & $1.01414407$ & $1.01414407$ \\ \cline{2-6} & $0.0790226728$ & $-1.00437424$&$-1.00437424$&$1.00437503$&$1.00437503$ \\ \hline \end{tabular} \label{tb:wdmp} \end{table} Finally, we consider (BVP;$f,g$) with $f(x,y)=2\pi^2\sin(\pi x)\sin(\pi y)$ and $g(x,y)=\sin(\pi x)\sin(\pi y)$. The exact solution is given as $u(x,y) = \sin(\pi x)\sin(\pi y)$. We examine errors $\|u-u_h\|_{L^\infty(\Omega)}$ with $r=1$ ($\mathcal{P}^1$ element) and $r=2$ ($\mathcal{P}^2$ element). Results are shown in Fig. \ref{fig:p1error} and Fig. \ref{fig:p2error}. We observe that the order is almost $O(h^{1+r})$: the optimal convergence rate is actually observed. This implies that our $L^\infty$ error estimate, Corollary \ref{cor2}, has room for improvement. \begin{figure}[bt] \centering \subfloat[][$r=1$]{\includegraphics[scale=0.8]{Fig2a.pdf}\label{fig:p1error}} \subfloat[][$r=2$]{\includegraphics[scale=0.8]{Fig2b.pdf}\label{fig:p2error}} \caption{$L^\infty$ errors $\|u-u_h\|_{L^\infty(\Omega)}$} \end{figure} \section{Conclusion} We have shown the interior error estimate and discrete weak maximum principle of the DG method for the Poisson equation. Those results are extensions of the standard FEM \cite{MR551291} to the DG method. Moreover, we have derived the $L^\infty$ error estimate as an application of the discrete weak maximum principle. Unfortunately, our $L^\infty$ error estimate is only sub-optimal. The optimal rate has been proved in \cite{MR2113680} by another method. This implies that we need to deep consider the imposition of the Dirichlet boundary condition in the DG method. In particular, we will study more precise estimates of $\alpha(h)$ and $\norm{u-u_h}_{L^\infty(\partial \Omega)}$ in the future works. \section*{Acknowledgment} The first author was supported by Program for Leading Graduate Schools, MEXT, Japan. The second author was supported by JST CREST Grant Number JPMJCR15D1, Japan, and JSPS KAKENHI Grant Number 15H03635, Japan. \bibliographystyle{spmpsci}
{ "timestamp": "2018-12-04T02:25:23", "yymm": "1812", "arxiv_id": "1812.00610", "language": "en", "url": "https://arxiv.org/abs/1812.00610", "abstract": "We derive several $L^\\infty$ error estimates for the symmetric interior penalty (SIP) discontinuous Galerkin (DG) method applied to the Poisson equation in a two-dimensional polygonal domain. Both local and global estimates are examined. The weak maximum principle (WMP) for the discrete harmonic function is also established. We prove our $L^\\infty$ estimates using this WMP and several $W^{2,p}$ and $W^{1,1}$ estimates for the Poisson equation. Numerical examples to validate our results are also presented.", "subjects": "Numerical Analysis (math.NA)", "title": "Weak discrete maximum principle and $L^\\infty$ analysis of the DG method for the Poisson equation on a polygonal domain", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9817357211137666, "lm_q2_score": 0.8152324826183822, "lm_q1q2_score": 0.8003428491987237 }
https://arxiv.org/abs/1306.5872
Geometric properties of inverse polynomial images
Given a polynomial $\T_n$ of degree $n$, consider the inverse image of $\R$ and $[-1,1]$, denoted by $\T_n^{-1}(\R)$ and $\T_n^{-1}([-1,1])$, respectively. It is well known that $\T_n^{-1}(\R)$ consists of $n$ analytic Jordan arcs moving from $\infty$ to $\infty$. In this paper, we give a necessary and sufficient condition such that (1) $\T_n^{-1}([-1,1])$ consists of $\nu$ analytic Jordan arcs and (2) $\T_n^{-1}([-1,1])$ is connected, respectively.
\section{Introduction} Let $\PP_n$ be the set of all polynomials of degree $n$ with complex coefficients. For a polynomial $\T_n\in\PP_n$, consider the inverse images $\T_n^{-1}(\R)$ and $\T_n^{-1}([-1,1])$, defined by \begin{equation} \T_n^{-1}(\R):=\bigl\{z\in\C:\T_n(z)\in\R\bigr\} \end{equation} and \begin{equation} \T_n^{-1}([-1,1]):=\bigl\{z\in\C:\T_n(z)\in[-1,1]\bigr\}, \end{equation} respectively. It is well known that $\T_n^{-1}(\R)$ consists of $n$ analytic Jordan arcs moving from $\infty$ to $\infty$ which cross each other at points which are zeros of the derivative $\T_n'$. In \cite{Peh-2003}, Peherstorfer proved that $\T_n^{-1}(\R)$ may be split up into $n$ Jordan arcs (not necessarily analytic) moving from $\infty$ to $\infty$ with the additional property that $\T_n$ is strictly monotone decreasing from $+\infty$ to $-\infty$ on each of the $n$ Jordan arcs. Thus, $\T_n^{-1}([-1,1])$ is the union of $n$ (analytic) Jordan arcs and is obtained from $\T_n^{-1}(\R)$ by cutting off the $n$ arcs of $\T_n^{-1}(\R)$. In \cite[Thm.\,3]{PehSch-2004}, we gave a necessary and sufficient condition such that $\T_n^{-1}([-1,1])$ consists of $2$ Jordan arcs, compare also \cite{Pakovich-1995}, where the proof can easily be extended to the case of $\ell$ arcs, see also \cite[Remark after Corollary\,2.2]{Peh-2003}. In the present paper, we will give a necessary and sufficient condition such that (1)~$\T_n^{-1}([-1,1])$ consists of $\nu$ (but not less than $\nu$) \emph{analytic} Jordan arcs (in Section\,2) and (2)~$\T_n^{-1}([-1,1])$ is connected (in Section\,3), respectively. From a different point of view as in this paper, inverse polynomial images are considered, e.g., in \cite{PehSt}, \cite{Pakovich-1998}, \cite{Pakovich-2007}, and \cite{Pakovich-2008}. Inverse polynomial images are interesting for instance in approximation theory, since each polynomial (suitable normed) of degree $n$ is the minimal polynomial with respect to the maximum norm on its inverse image, see \cite{KamoBorodin}, \cite{Peh-1996}, \cite{FischerPeherstorfer}, and \cite{Fischer-1992}. \section{The Number of (Analytic) Jordan Arcs of an Inverse Polynomial Image} Let us start with a collection of important properties of the inverse images $\T_n^{-1}(\R)$ and $\T_n^{-1}([-1,1])$. Most of them are due to Peherstorfer\,\cite{Peh-2003} or classical well known results. Let us point out that $\T_n^{-1}(\R)$ (and also $\T_n^{-1}([-1,1])$), on the one hand side, may be characterized by $n$ analytic Jordan arcs and, on the other side, by $n$ (not necessarily analytic) Jordan arcs, on which $\T_n$ is strictly monotone. Let $C:=\{\gamma(t):t\in[0,1]\}$ be an analytic Jordan arc in $\C$ and let $\T_n\in\PP_n$ be a polynomial such that $\T_n(\gamma(t))\in\R$ for all $t\in[0,1]$. We call a point $z_0=\gamma(t_0)$ a \emph{saddle point} of $\T_n$ on $C$ if $\T_n'(z_0)=0$ and $z_0$ is no extremum of $\T_n$ on $C$. \begin{lemma}\label{Lem-1} Let $\T_n\in\PP_n$ be a polynomial of degree $n$. \begin{enumerate} \item $\T_n^{-1}(\R)$ consists of $n$ analytic Jordan arcs, denoted by $\tilde{C}_1,\tilde{C}_2,\dots,\tilde{C}_n$, in the complex plane running from $\infty$ to $\infty$. \item $\T_n^{-1}(\R)$ consists of $n$ Jordan arcs, denoted by $\tilde{\Gamma}_1,\tilde{\Gamma}_2,\dots,\tilde{\Gamma}_n$, in the complex plane running from $\infty$ to $\infty$, where on each $\tilde{\Gamma}_j$, $j=1,2,\ldots,n$, $\T_n(z)$ is strictly monotone decreasing from $+\infty$ to $-\infty$. \item A point $z_0\in\T_n^{-1}(\R)$ is a crossing point of exactly $m$, $m\geq2$, analytic Jordan arcs $\tilde{C}_{i_1},\tilde{C}_{i_2},\ldots,\tilde{C}_{i_m}$, $1\leq{i}_1<i_2<\ldots<i_m\leq{n}$, if and only if $z_0$ is a zero of $\T'_n$ with multiplicity $m-1$. In this case, the $m$ arcs are cutting each other at $z_0$ in successive angles of $\pi/m$. If $m$ is odd then $z_0$ is a saddle point of $\re\{\T_n(z)\}$ on each of the $m$ arcs. If $m$ is even then, on $m/2$ arcs, $z_0$ is a minimum of $\re\{\T_n(z)\}$ and on the other $m/2$ arcs, $z_0$ is a maximum of $\re\{\T_n(z)\}$. \item A point $z_0\in\T_n^{-1}(\R)$ is a crossing point of exactly $m$, $m\geq2$, Jordan arcs\\ $\tilde{\Gamma}_{i_1},\tilde{\Gamma}_{i_2},\ldots,\tilde{\Gamma}_{i_m}$, $1\leq{i}_1<i_2<\ldots<i_m\leq{n}$, if and only if $z_0$ is a zero of $\T'_n$ with multiplicity $m-1$. \item $\T_n^{-1}([-1,1])$ consists of $n$ analytic Jordan arcs, denoted by $C_1,C_2,\dots,C_n$, where the $2n$ zeros of $\T_n^2-1$ are the endpoints of the $n$ arcs. If $z_0\in\C$ is a zero of $\T_n^2-1$ of multiplicity $m$ then exactly $m$ analytic Jordan arcs $C_{i_1},C_{i_2},\ldots,C_{i_m}$ of $\T_n^{-1}([-1,1])$, $1\leq{i}_1<i_2<\ldots<i_m\leq{n}$, have $z_0$ as common endpoint. \item $\T_n^{-1}([-1,1])$ consists of $n$ Jordan arcs, denoted by $\Gamma_1,\Gamma_2,\dots,\Gamma_n$, with $\Gamma_j\subset\tilde{\Gamma}_j$, $j=1,2,\dots,n$, where on each $\Gamma_j$, $\T_n(z)$ is strictly monotone decreasing from $+1$ to $-1$. If $z_0\in\C$ is a zero of $\T_n^2-1$ of multiplicity $m$, then exactly $m$ Jordan arcs $\Gamma_{i_1},\ldots,\Gamma_{i_m}$ of $\T_n^{-1}([-1,1])$, $1\leq{i}_1<i_2<\ldots<i_m\leq{n}$, have $z_0$ as common endpoint. \item Two arcs $C_j,C_k$, $j\neq k$, cross each other at most once (the same holds for $\Gamma_j,\Gamma_k$). \item Let $S:=\T_n^{-1}([-1,1])$, then the complement $\C\setminus{S}$ is connected. \item Let $S:=\T_n^{-1}([-1,1])$, then, for $P_n(z):=\T_n((z-b)/a)$, $a,b\in\C$, $a\neq0$, the inverse image is $P_n^{-1}([-1,1])=aS+b$. \item $\T_n^{-1}([-1,1])\subseteq\R$ if and only if the coefficients of $\T_n$ are real, $\T_n$ has $n$ simple real zeros and $\min\bigl\{|\T_n(z)|:\T'_n(z)=0\bigr\}\geq1$. \item $\T_n^{-1}(\R)$ is symmetric with respect to the real line if and only if $\T_{n}(z)$ or $\ii\T_{n}(z)$ has real coefficients only. \end{enumerate} \end{lemma} \begin{proof} (i), (iii), (iv), and (xi) are well known.\\ For (ii), see \cite[Thm.\,2.2]{Peh-2003}.\\ Concerning the connection between (iii),(iv) and (v),(vi) note that each zero $z_0$ of $Q_{2n}(z)=\T_n^2(z)-1\in\PP_{2n}$ with multiplicity $m$ is a zero of $Q_{2n}'(z)=2\T_n(z)\,\T_n'(z)$ with multiplicity $m-1$, hence a zero of $\T_n'(z)$ with multiplicity $m-1$. Thus, (v) and (vi) follow immediately from (i)\&(iii) and (ii)\&(iv), respectively.\\ (vii) follows immediately from (viii).\\ Concerning (viii), suppose that there exists a simple connected domain $B$, which is surrounded by a subset of $\T_n^{-1}([-1,1])$. Then the harmonic function $v(x,y):=\im\{\T_n(x+\ii{y})\}$ is zero on $\partial{B}$ thus, by the maximum principle, $v(x,y)$ is zero on $B$, which is a contradiction.\\ (ix) follows from the definition of $\T_n^{-1}([-1,1])$.\\ For (x), see \cite[Cor.\,2.3]{Peh-2003}. \end{proof} \begin{example}\label{Ex} Consider the polynomial $\T_n(z):=1+z^2(z-1)^3(z-2)^4$ of degree $n=9$. Fig.\,\ref{Fig_InverseImageGeneral} shows the inverse images $\T_n^{-1}([-1,1])$ (solid line) and $\T_n^{-1}(\R)$ (dotted and solid line). The zeros of $\T_n+1$ and $\T_n-1$ are marked with a circle and a disk, respectively. One can easily identitfy the $n=9$ analytic Jordan arcs $\tilde{C}_1,\tilde{C}_2,\ldots,\tilde{C}_n$ which $\T_n^{-1}(\R)$ consists of, compare Lemma\,\ref{Lem-1}\,(i), and the $n=9$ analytic Jordan arcs $C_1,C_2,\ldots,C_n$ which $\T_n^{-1}([-1,1])$ consists of, compare Lemma\,\ref{Lem-1}\,(v), where the endpoints of the arcs are exactly the circles and disks, i.e., the zeros of $\T_n^2-1$. Note that $\tilde{C}_1=\R$, $C_1=[-0.215\ldots,0]$ and $C_2=[0,1]$. \end{example} \begin{figure}[ht] \begin{center} \includegraphics[scale=1.6]{Fig_InverseImageGeneral} \caption{\label{Fig_InverseImageGeneral} Inverse images $\T_9^{-1}([-1,1])$ (solid line) and $\T_9^{-1}(\R)$ (dotted and solid line) for the polynomial $\T_9(z):=1+z^2(z-1)^3(z-2)^4$} \end{center} \end{figure} Before we state the result concerning the minimal number of analytic Jordan arcs $\T_n^{-1}([-1,1])$ consists of, let us do some preparations. Let $\T_n\in\PP_n$ and consider the zeros of the polynomial $\T_n^2-1\in\PP_{2n}$. Let $\{a_1,a_2,\ldots,a_{2\ell}\}$ be the set of all zeros of $\T_n^2-1$ with \emph{odd} multiplicity, where $a_1,a_2,\ldots,a_{2\ell}$ are pairwise distinct and each $a_j$ has multiplicity $2\beta_j-1$, $j=1,\ldots,2\ell$. Further, let \begin{equation} (b_1,b_2,\ldots,b_{2\nu}):=(\underbrace{a_1,\ldots,a_1}_{(2\beta_1-1)-\text{times}}, \underbrace{a_2,\ldots,a_2}_{(2\beta_2-1)-\text{times}},\ldots,\underbrace{a_{2\ell},\ldots,a_{2\ell}}_{(2\beta_{2\ell}-1)-\text{times}}), \end{equation} thus \begin{equation} 2\nu=\sum_{j=1}^{2\ell}(2\beta_j-1), \end{equation} i.e., $b_1,b_2,\ldots,b_{2\nu}$ are the zeros of odd multiplicity \emph{written according to their multiplicity}. \begin{theorem}\label{Thm-NuArcs} Let $\T_n\in\PP_n$ be any polynomial of degree $n$. Then, $\T_n^{-1}([-1,1])$ consists of $\nu$ (but not less than $\nu$) analytic Jordan arcs with endpoints $b_1,b_2,\ldots,b_{2\nu}$ if and only if $\T_n^2-1$ has exactly $2\nu$ zeros $b_1,b_2,\ldots,b_{2\nu}$ (written according to their multiplicity) of odd multiplicity. \end{theorem} \begin{proof} By Lemma\,\ref{Lem-1}\,(v), $\T_n^{-1}([-1,1])$ consists of $n$ analytic Jordan arcs $C_1,C_2,\ldots,C_n$, which can be combined into $\nu$ analytic Jordan arcs in the following way. Clearly, two analytic Jordan arcs $C_{i_1}$ and $C_{i_2}$ can be joined together into one analytic Jordan arc if they have the same endpoint, which is a zero of $\T_n^2-1$, and if they lie on the same analytic Jordan arc $\tilde{C}_{i_3}$ of Lemma\,\ref{Lem-1}\,(i). By Lemma\,\ref{Lem-1}\,(iii) and (v), such combinations are possible only at the zeros of $\T_n^2-1$ of \emph{even} multiplicity. More precisely, let $d_1,d_2,\ldots,d_k$ be the zeros of $\T_n^2-1$ with even multiplicities $2\alpha_1,2\alpha_2,\ldots,2\alpha_k$, where, by assumption, \[ 2\alpha_1+2\alpha_2+\ldots+2\alpha_k=2n-2\nu. \] By Lemma\,\ref{Lem-1}\,(iii) and (v), at each point $d_j$, the $2\alpha_j$ analytic Jordan arcs of $\T_n^{-1}([-1,1])$ can be combined into $\alpha_j$ analytic arcs, $j=1,2,\ldots,k$. Altogether, the number of such combinations is $\alpha_1+\alpha_2+\ldots+\alpha_k=n-\nu$, thus the the total number of $n$ analytic Jordan arcs is reduced by $n-\nu$, hence $\nu$ analytic Jordan arcs remain and the sufficiency part is proved. Since, for each polynomial $\T_n\in\PP_n$, there is a unique $\nu\in\{1,2,\ldots,n\}$ such that $\T_n^2-1$ has exactly $2\nu$ zeros of odd multiplicity (counted with multiplicity), the necessity part follows. \end{proof} \begin{example} For a better understanding of the combination of two analytic Jordan arcs into one analytic Jordan arc, as done in the proof of Theorem\,\ref{Thm-NuArcs}, let us again consider the inverse image of the polynomial of Example\,\ref{Ex}. \begin{itemize} \item The point $d_1=0$ is a zero of $\T_n-1$ with multiplicity $2\alpha_1=2$, thus $2$ analytic Jordan arcs, here $C_1$ and $C_2$, have $d_1$ as endpoint, compare Lemma\,\ref{Lem-1}\,(v). Along the arc $\tilde{C}_1$, $d_1$ is a maximum, along the arc $\tilde{C}_2$, $d_1$ is a minimum, compare Lemma\,\ref{Lem-1}\,(iii), thus the $2$ analytic Jordan arcs $C_1$ and $C_2$ can be joined together into one analytic Jordan arc $C_1\cup{C}_2$. \item The point $d_2=2$ is a zero of $\T_n-1$ with multiplicity $2\alpha_2=4$, thus $4$ analytic Jordan arcs, here $C_6$, $C_7$, $C_8$ and $C_9$, have $d_2$ as endpoint. Along the arc $\tilde{C}_7$ or $\tilde{C}_9$, $d_3$ is a maximum, along the arc $\tilde{C}_8$ or $\tilde{C}_1$, $d_3$ is a minimum, compare Lemma\,\ref{Lem-1}\,(iii). Hence, the analytic Jordan arcs $C_6$ and $C_9$ can be combined into one analytic Jordan arc $C_6\cup{C}_9$, analogously $C_7$ and $C_8$ can be combined into $C_7\cup{C}_8$. \item The point $a_1=1$ is a zero of $\T_n-1$ with multiplicity $3$, thus $3$ analytic Jordan arcs, here $C_2$, $C_4$ and $C_5$, have $a_1$ as endpoint. Since $a_1$ is a saddle point along each of the three analytic Jordan arcs $\tilde{C}_1,\tilde{C}_4,\tilde{C}_5$, compare Lemma\,\ref{Lem-1}\,(iii), no combination of arcs can be done. \end{itemize} Altogether, we get $\alpha_1+\alpha_2=3=n-\nu$ combinations and therefore $\T_n^{-1}([-1,1])$ consists of $\nu=6$ analytic Jordan arcs, which are given by $C_1\cup{C}_2$, $C_3$, $C_4$, $C_5$, $C_6\cup{C}_9$ and $C_7\cup{C}_8$. \end{example} \begin{lemma}\label{Lemma-PolEq} For any polynomial $\T_n(z)=c_nz^n+\ldots\in\PP_n$, $c_n\in\C\setminus\{0\}$, there exists a unique $\ell\in\{1,2,\ldots,n\}$, a unique monic polynomial $\HH_{2\ell}(z)=z^{2\ell}+\ldots\in\PP_{2\ell}$ with pairwise distinct zeros $a_1,a_2,\ldots,a_{2\ell}$, i.e., \begin{equation}\label{H} \HH_{2\ell}(z)=\prod_{j=1}^{2\ell}(z-a_j), \end{equation} and a unique polynomial $\U_{n-\ell}(z)=c_nz^{n-\ell}+\ldots\in\PP_{n-\ell}$ with the same leading coefficient $c_n$ such that the polynomial equation \begin{equation}\label{TU} \T_n^2(z)-1=\HH_{2\ell}(z)\,\U_{n-\ell}^2(z) \end{equation} holds. Note that the points $a_1,a_2,\ldots,a_{2\ell}$ are exactly those zeros of $\T_n^2-1$ which have odd multiplicity. \end{lemma} \begin{proof} The assertion follows immediately by the fundamental theorem of algebra for the polynomial $Q_{2n}(z):=\T_n^2(z)-1=c_n^2z^{2n}+\ldots\in\PP_{2n}$, where $2\ell$ is the number of distinct zeros of $Q_{2n}$ with odd multiplicity. It only remains to show that the case $\ell=0$ is not possible. If $\ell=0$, then all zeros of $Q_{2n}$ are of even multiplicity. Thus there are at least $n$ zeros (counted with multiplicity) of $Q_{2n}'$ which are also zeros of $Q_{2n}$ but not zeros of $\T_n$. Since $Q_{2n}'(z)=2\,\T_n(z)\,\T_n'(z)$, there are at least $n$ zeros (counted with multiplicity) of $\T_n'$, which is a contradiction. \end{proof} Let us point out that the polynomial equation \eqref{TU} (sometimes called Pell equation) is the starting point for investigations concerning minimal or orthogonal polynomials on several intervals, see, e.g., \cite{Bogatyrev}, \cite{Peh-1993}, \cite{Peh-1996}, \cite{Peh-2001}, \cite{PehSch-1999}, \cite{SoYu-1995}, and \cite{Totik-2001}.\\ \indent In \cite[Theorem\,3]{PehSch-2004}, we proved that the polynomial equation \eqref{TU} (for $\ell=2$) is equivalent to the fact that $\T_n^{-1}([-1,1])$ consists of $2$ Jordan arcs (not necessarily analytic), compare also \cite{Pakovich-1995}. The condition and the proof can be easily extended to the general case of $\ell$ arcs, compare also \cite[Remark after Corollary\,2.2]{Peh-2003}. In addition, we give an alternative proof similar to that of Theorem\,\ref{Thm-NuArcs}. \begin{theorem}\label{Thm-EllArcs} Let $\T_n\in\PP_n$ be any polynomial of degree $n$. Then $\T_n^{-1}([-1,1])$ consists of $\ell$ (but not less than $\ell$) Jordan arcs with endpoints $a_1,a_2,\ldots,a_{2\ell}$ if and only if $\T_n^2-1$ has exactly $2\ell$ pairwise distinct zeros $a_1,a_2,\ldots,a_{2\ell}$, $1\leq\ell\leq{n}$, of odd multiplicity, i.e., if and only if $\T_n$ satisfies a polynomial equation of the form \eqref{TU} with $\HH_{2\ell}$ given in \eqref{H}. \end{theorem} \begin{proof} By Lemma\,\ref{Lem-1}\,(vi), $\T_n^{-1}([-1,1])$ consists of $n$ Jordan arcs $\Gamma_1,\Gamma_2,\dots,\Gamma_n$, which can be combined into $\ell$ Jordan arcs in the following way: Let $d_1,d_2,\ldots,d_k$ be those zeros of $\T_n^2-1$ with \emph{even} multiplicities $2\alpha_1,2\alpha_2,\ldots,2\alpha_k$ and let, as assumed in the Theorem, $a_1,a_2,\ldots,a_{2\ell}$ be those zeros of $\T_n^2-1$ with \emph{odd} multiplicities $2\beta_1-1,2\beta_2-1,\ldots,2\beta_{2\ell}-1$, where \begin{equation}\label{SumMult} 2\alpha_1+2\alpha_2+\ldots+2\alpha_k+(2\beta_1-1)+(2\beta_2-1)+\ldots+(2\beta_{2\ell}-1)=2n \end{equation} holds. By Lemma\,\ref{Lem-1}\,(vi), at each point $d_j$, the $2\alpha_j$ Jordan arcs can be combined into $\alpha_j$ Jordan arcs, $j=1,2,\ldots,\nu$, and at each point $a_j$, the $2\beta_j-1$ Jordan arcs can be combined into $\beta_j$ Jordan arcs, $j=1,2,\ldots,2\ell$. Altogether, the number of such combinations, using \eqref{SumMult}, is \[ \alpha_1+\alpha_2+\ldots+\alpha_{\nu}+(\beta_1-1)+(\beta_2-1)+\ldots+(\beta_{2\ell}-1)=(n+\ell)-2\ell=n-\ell, \] i.e., the total number $n$ of Jordan arcs is reduced by $n-\ell$, thus $\ell$ Jordan arcs remain and the sufficiency part is proved. Since, by Lemma\,\ref{Lemma-PolEq}, for each polynomial $\T_n\in\PP_n$ there is a unique $\ell\in\{1,2,\ldots,n\}$ such that $\T_n^2-1$ has exactly $2\ell$ distinct zeros of odd multiplicity, the necessity part is clear. \end{proof} \begin{example} Similar as after the proof of Theorem\,\ref{Thm-NuArcs}, let us illustrate the combination of Jordan arcs by the polynomial of Example\,\ref{Ex}. Taking a look at Fig.\,\ref{Fig_InverseImageGeneral}, one can easily identitfy the $n=9$ Jordan arcs $\Gamma_1,\Gamma_2,\ldots,\Gamma_n\in\T_n^{-1}([-1,1])$, where each arc $\Gamma_j$ runs from a disk to a circle. Note that the two arcs, which cross at $z\approx0.3$, may be chosen in two different ways. Now, $\T_n^2-1$ has the zero $d_1=0$ with multiplicity $2\alpha_1=2$, the zero $d_2=2$ with multiplicity $2\alpha_2=4$, and a zero $a_1=1$ with multiplicity $2\beta_1-1=3$, all other zeros $a_j$ have multiplicity $2\beta_j-1=1$, $j=2,3,\ldots,2\ell$. Thus, it is possible to have one combination at $d_1=0$, two combinations at $d_2=2$ and one combination of Jordan arcs at $a_1=1$. Altogether, we obtain $\alpha_1+\alpha_2+(\beta_1-1)=4=n-\ell$ combinations and the number of Jordan arcs is $\ell=5$. \end{example} For the sake of completeness, let us mention two simple special cases, first the case $\ell=1$, see, e.g., \cite[Remark\,4]{PehSch-2004}, and second, the case when all endpoints $a_1,a_2,\ldots,a_{2\ell}$ of the arcs are real, see \cite{Peh-1993}. \begin{corollary} Let $\T_n\in\PP_n$. \begin{enumerate} \item $\T_n^{-1}([-1,1])$ consists of $\ell=1$ Jordan arc with endpoints $a_1,a_2\in\C$, $a_1\neq{a}_2$, if and only if $\T_n$ is the classical Chebyshev polynomial of the first kind (suitable normed), i.e., $\T_n(z)=T_n((2z-a_1-a_2)/(a_2-a_1))$, where $T_n(z):=\cos(n\arccos{z})$. In this case, $\T_n^{-1}([-1,1])$ is the complex interval $[a_1,a_2]$. \item $\T_n^{-1}([-1,1])=[a_1,a_2]\cup[a_3,a_4]\cup\ldots\cup[a_{2\ell-1},a_{2\ell}]$, $a_1,a_2,\ldots,a_{2\ell}\in\R$, $a_1<a_2<\ldots<a_{2\ell}$, if and only if $\T_n$ satisfies the polynomial equation \eqref{TU} with $\HH_{2\ell}$ as in \eqref{H} and $a_1,a_2,\ldots,a_{2\ell}\in\R$, $a_1<a_2<\ldots<a_{2\ell}$. \end{enumerate} \end{corollary} Let us consider the case of $\ell=2$ Jordan arcs in more detail. Given four pairwise distinct points $a_1,a_2,a_3,a_4\in\C$ in the complex plane, define \begin{equation}\label{H2} \HH_4(z):=(z-a_1)(z-a_2)(z-a_3)(z-a_4), \end{equation} and suppose that $\T_n(z)=c_nz^n+\ldots\in\PP_n$ satisfies a polynomial equation of the form \begin{equation}\label{TU2} \T_n^2(z)-1=\HH_4(z)\,\U_{n-2}^2(z) \end{equation} with $\U_{n-2}(z)=c_nz^{n-2}+\ldots\in\PP_{n-2}$. Then, by \eqref{TU2}, there exists a $z^*\in\C$ such that the derivative of $\T_n$ is given by \begin{equation}\label{z*} \T_n'(z)=n(z-z^*)\,\U_{n-2}(z). \end{equation} By Theorem\,\ref{Thm-EllArcs}, $\T_n^{-1}([-1,1])$ consists of two Jordan arcs. Moreover, it is proved in \cite[Theorem\,3]{PehSch-2004} that the two Jordan arcs are crossing each other if and only if $z^*\in\T_n^{-1}([-1,1])$ (compare also Theorem\,\ref{Theorem-Connectedness}). In this case, $z^*$ is the only crossing point. Interestingly, the minimum number of analytic Jordan arcs is not always two, as the next theorem says. In order to prove this result, we need the following lemma\,\cite[Lemma\,1]{PehSch-2004}. \begin{lemma}\label{Lemma-PehSch} Suppose that $\T_n\in\PP_n$ satisfies a polynomial equation of the form \eqref{TU2}, where $\HH_4$ is given by \eqref{H2}, and let $z^*$ be given by \eqref{z*}. \begin{enumerate} \item If $z^*$ is a zero of $\U_{n-2}$ then it is either a double zero of $\U_{n-2}$ or a zero of $\HH$. \item If $z^*$ is a zero of $\HH$ then $z^*$ is a simple zero of $\U_{n-2}$. \item The point $z^*$ is the only possible common zero of $\HH$ and $\U_{n-2}$. \item If $\U_{n-2}$ has a zero $y^*$ of order greater than one then $y^*=z^*$ and $z^*$ is a double zero of $\U_{n-2}$. \end{enumerate} \end{lemma} \begin{theorem} Suppose that $\T_n\in\PP_n$ satisfies a polynomial equation of the form \eqref{TU2}, where $\HH_4$ is given by \eqref{H2}, and let $z^*$ be given by \eqref{z*}. If $z^*\notin\{a_1,a_2,a_3,a_4\}$ then $\T_n^{-1}([-1,1])$ consists of two analytic Jordan arcs. If $z^*\in\{a_1,a_2,a_3,a_4\}$ then $\T_n^{-1}([-1,1])$ consists of three analytic Jordan arcs, all with one endpoint at $z^*$, and an angle of $2\pi/3$ between two arcs at $z^*$. \end{theorem} \begin{proof} We distinguish two cases: \begin{enumerate} \item[1.] $\T_n(z^*)\notin\{-1,1\}$: By Lemma\,\ref{Lemma-PehSch}, $\T_n^2-1$ has 4 simple zeros $\{a_1,a_2,a_3,a_4\}$ and $n-2$ double zeros. Thus, by Theorem\,\ref{Thm-NuArcs}, $\T_n^{-1}([-1,1])$ consists of two analytic Jordan arcs. \item[2.] $\T_n(z^*)\in\{-1,1\}$: \begin{enumerate} \item[2.1] If $z^*\in\{a_1,a_2,a_3,a_4\}$ then, by Lemma\,\ref{Lemma-PehSch}, $\T_n^2-1$ has 3 simple zeros given by $\{a_1,a_2,a_3,a_4\}\setminus\{z^*\}$, $n-3$ double zeros and one zero of multiplicity 3 (that is $z^*$). Thus, by Theorem\,\ref{Thm-NuArcs}, $\T_n^{-1}([-1,1])$ consists of three analytic Jordan arcs. \item[2.2] If $z^*\notin\{a_1,a_2,a_3,a_4\}$ then, by Lemma\,\ref{Lemma-PehSch}, $z^*$ is a double zero of $\U_{n-2}$. Thus $\T_n^2-1$ has 4 simple zeros $\{a_1,a_2,a_3,a_4\}$, $n-4$ double zeros and one zero of multiplicity 4 (that is $z^*$). Thus, by Theorem\,\ref{Thm-NuArcs}, $\T_n^{-1}([-1,1])$ consists of two analytic Jordan arcs. \end{enumerate} \end{enumerate} The very last statement of the theorem follows immediately by Lemma\,\ref{Lem-1}\,(iii). \end{proof} Let us mention that in \cite{PehSch-2004}, see also \cite{Sch-2007a} and \cite{Sch-2009}, necessary and sufficient conditions for four points $a_1,a_2,a_3,a_4\in\C$ are given with the help of Jacobian elliptic functions such that there exists a polynomial of degree $n$ whose inverse image consists of two Jordan arcs with the four points as endpoints. Concluding this section, let us give two simple examples of inverse polynomial images. \begin{example}\hfill{} \begin{enumerate} \item Let $a_1=-1$, $a_2=-a$, $a_3=a$ and $a_4=1$ with $0<a<1$ and \[ \HH_4(z)=(z-a_1)(z-a_2)(z-a_3)(z-a_4)=(z^2-1)(z^2-a^2). \] If \[ \T_2(z):=\frac{2z^2-a^2-1}{1-a^2},\quad\U_0(z):=\frac{2}{1-a^2}, \] then \[ \T_2^2(z)-\HH_4(z)\U_0^2(z)=1. \] Thus, by Theorem\,\ref{Thm-EllArcs}, $\T_2^{-1}([-1,1])$ consists of two Jordan arcs with endpoints $a_1$, $a_2$, $a_3$, $a_4$, more precisely $\T_2^{-1}([-1,1])=[-1,-a]\cup[a,1]$. \item Let $a_1=\ii$, $a_2=-\ii$, $a_3=a-\ii$ and $a_4=a+\ii$ with $a>0$ and \[ \HH_4(z)=(z-a_1)(z-a_2)(z-a_3)(z-a_4)=(z^2+1)((z-a)^2+1). \] If \[ \T_2(z):=\frac{\ii}{a}\bigl(z^2-az+1\bigr),\quad\U_0(z):=\frac{\ii}{a}, \] then \[ \T_2^2(z)-\HH_4(z)\U_0^2(z)=1. \] Thus, by Theorem\,\ref{Thm-EllArcs}, $\T_2^{-1}([-1,1])$ consists of two Jordan arcs with endpoints $a_1$, $a_2$, $a_3$, $a_4$. More precisely, if $0<a<2$, \[ \T_2^{-1}([-1,1])=\bigl\{x+\ii{y}\in\C:-\frac{(x-a/2)^2}{1-a^2/4}+\frac{y^2}{1-a^2/4}=1\bigr\}, \] i.e., $\T_2^{-1}([-1,1])$ is an equilateral hyperbola (not crossing the real line) with center at $z_0=a/2$ and asymptotes $y=\pm(x-a/2)$.\\ If $a=2$, $\T_2^{-1}([-1,1])=[\ii,a-\ii]\cup[-\ii,a+\ii]$, i.e., the union of two complex intervals.\\ If $2<a<\infty$, \[ \T_2^{-1}([-1,1])=\bigl\{x+\ii{y}\in\C:\frac{(x-a/2)^2}{a^2/4-1}-\frac{y^2}{a^2/4-1}=1\bigr\}, \] i.e., $\T_2^{-1}([-1,1])$ is an equilateral hyperbola with center at $z_0=a/2$, crossing the real line at $a/2\pm\sqrt{a^2/4-1}$ and asymptotes $y=\pm(x-a/2)$.\\ In Fig.\,\ref{Fig_Rectangle}, the sets $\T_2^{-1}([-1,1])$ including the asymptotes are plotted for the three cases discussed above. \end{enumerate} \end{example} \begin{figure}[ht] \begin{center} \includegraphics[scale=0.7]{Fig_Rectangle} \caption{\label{Fig_Rectangle} The inverse image $\T_2^{-1}([-1,1])$ for $0<a<2$ (left plot), for $a=2$ (middle plot) and for $a>2$ (right plot)} \end{center} \end{figure} \section{The Connectedness of an Inverse Polynomial Image} In the next theorem, we give a necessary and sufficient condition such that the inverse image is connected. \begin{theorem}\label{Theorem-Connectedness} Let $\T_n\in\PP_n$. The inverse image $\T_n^{-1}([-1,1])$ is connected if and only if all zeros of the derivative $\T_n'$ lie in $\T_n^{-1}([-1,1])$. \end{theorem} \begin{proof} Let $\Gamma:=\bigl\{\Gamma_1,\Gamma_2,\dots,\Gamma_n\bigr\}$ denote the set of arcs of $\T_n^{-1}([-1,1])$ as in Lemma\,\ref{Lem-1}\,(vi). ``$\Leftarrow$'': Suppose that all zeros of $\T_n'$ lie in $\Gamma$. Let $A_1\in\Gamma$ be such that it contains at least one zero $z_1$ of $\T'_n$ with multiplicity $m_1\geq1$. By Lemma\,\ref{Lem-1}\,(ii), (iv) and (vi), there are $m_1$ additional arcs $A_2,A_3,\dots,A_{m_1+1}\in\Gamma$ containing $z_1$. By Lemma\,\ref{Lem-1}\,(vii), \[ A_j\cap{A}_k=\{z_1\}~\text{for}~j,k\in\{1,2,\dots,m_1+1\},~j\neq{k}. \] Now assume that there is another zero $z_2$ of $\T'_n$, $z_2\neq{z}_1$, with multiplicity $m_2$, on $A_{j^*}$, $j^*\in\{1,2,\dots,m_1+1\}$. Since no arc $A_j$, $j\in\{1,2,\dots,m_1+1\}\setminus\{j^*\}$ contains $z_2$, there are $m_2$ curves $A_{m_1+1+j}\in\Gamma$, $j=1,2,\dots,m_2$, which cross each other at $z_2$ and for which, by Lemma\,\ref{Lem-1}\,(vii), \[ \begin{aligned} A_j\cap{A}_k&=\{z_2\} \quad &&\text{for } &&j,k\in\{m_1+2,\dots,m_1+m_2+1\}, j\neq k,\\ A_j\cap{A}_k&=\emptyset &&\text{for } &&j\in\{1,2,\dots,m_1+1\}\setminus\{j^*\},\\ & && &&k\in\{m_1+2,\dots,m_1+m_2+1\}\\ A_{j^*}\cap{A}_k&=\{z_2\} &&\text{for } &&k\in\{m_1+2,\dots,m_1+m_2+1\}. \end{aligned} \] If there is another zero $z_3$ of $\T_n'$, $z_3\notin\{z_1,z_2\}$, on $A_{j^{**}}$, $j^{**}\in\{1,2,\dots,m_1+m_2+1\}$, of multiplicity $m_3$, we proceed as before.\\ We proceed like this until we have considered all zeros of $\T'_n$ lying on the constructed set of arcs. Thus, we get a connected set of $k^*+1$ curves \[ A^*:=A_1\cup{A}_2\cup\ldots\cup{A}_{k^*+1} \] with $k^*$ zeros of $\T'_n$, counted with multiplicity, on $A^*$.\\ Next, we claim that $k^*=n-1$. Assume that $k^*<n-1$, then, by assumption, there exists a curve $A_{k^*+2}\in\Gamma$, for which \[ A_{k^*+2}\cap{A}^*=\{\} \] and on which there is another zero of $\T'_n$. By the same procedure as before, we get a set $A^{**}$ of $k^{**}+1$ arcs of $\Gamma$ for which $A^*\cap{A}^{**}=\{\}$ and $k^{**}$ zeros of $\T'_n$, counted with multiplicity. If $k^*+k^{**}=n-1$, then we would get a set of $k^*+k^{**}+2=n+1$ arcs, which is a contradiction to Lemma\,\ref{Lem-1}\,(i). If $k^*+k^{**}<n-1$, we proceed analogously and again, we get too many arcs, i.e., a contradiction to Lemma\,\ref{Lem-1}\,(vi). Thus, $k^*=n-1$ must hold and thus $\Gamma$ is connected. ``$\Rightarrow$'': Suppose that $\Gamma$ is connected. Thus, it is possible to reorder $\Gamma_1,\Gamma_2,\ldots,\Gamma_n$ into $\Gamma_{k_1},\Gamma_{k_2},\ldots,\Gamma_{k_n}$ such that $\Gamma_{k_1}\cup\ldots\cup\Gamma_{k_j}$ is connected for each $j\in\{2,\ldots,n\}$. Now we will count the crossing points (common points) of the arcs in the following way: If there are $m+1$ arcs $A_1,A_2,\ldots,A_{m+1}\in\Gamma$ such that $z_0\in{A}_j$, $j=1,2,\ldots,A_{m+1}$, then we will count the crossing point $z_0$ $m$-times, i.e., we say $A_1,\ldots,A_{m+1}$ has $m$ crossing points. Hence, $\Gamma_{k_1}\cup\Gamma_{k_2}$ has one crossing point, $\Gamma_{k_1}\cup\Gamma_{k_2}\cup\Gamma_{k_3}$ has two crossing points, $\Gamma_{k_1}\cup\Gamma_{k_2}\cup\Gamma_{k_3}\cup\Gamma_{k_4}$ has 3 crossing points, and so on. Summing up, we arrive at $n-1$ crossing points which are, by Lemma\,\ref{Lem-1}\,(iv) the zeros of $\T_n'$. \end{proof} Theorem\,\ref{Theorem-Connectedness} may be generalized to the question how many connected sets $\T_n^{-1}([-1,1])$ consists of. The proof runs along the same lines as that of Theorem\,\ref{Theorem-Connectedness}. \begin{theorem} Let $\T_n\in\PP_n$. The inverse image $\T_n^{-1}([-1,1])$ consists of $k$, $k\in\{1,2,\ldots,n\}$, connected components $B_1,B_2,\ldots,B_k$ with $B_1\cup{B}_2\cup\ldots\cup{B}_k=\T_n^{-1}([-1,1])$ and $B_i\cap{B}_j=\{\}$, $i\neq{j}$, if and only if $n-k$ zeros of the derivative $\T_n'$ lie in $\T_n^{-1}([-1,1])$. \end{theorem} \bibliographystyle{amsplain}
{ "timestamp": "2013-06-26T02:01:40", "yymm": "1306", "arxiv_id": "1306.5872", "language": "en", "url": "https://arxiv.org/abs/1306.5872", "abstract": "Given a polynomial $\\T_n$ of degree $n$, consider the inverse image of $\\R$ and $[-1,1]$, denoted by $\\T_n^{-1}(\\R)$ and $\\T_n^{-1}([-1,1])$, respectively. It is well known that $\\T_n^{-1}(\\R)$ consists of $n$ analytic Jordan arcs moving from $\\infty$ to $\\infty$. In this paper, we give a necessary and sufficient condition such that (1) $\\T_n^{-1}([-1,1])$ consists of $\\nu$ analytic Jordan arcs and (2) $\\T_n^{-1}([-1,1])$ is connected, respectively.", "subjects": "Complex Variables (math.CV)", "title": "Geometric properties of inverse polynomial images", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9904405986777259, "lm_q2_score": 0.8080672066194946, "lm_q1q2_score": 0.8003425678960499 }
https://arxiv.org/abs/1203.5829
Ensemble estimators for multivariate entropy estimation
The problem of estimation of density functionals like entropy and mutual information has received much attention in the statistics and information theory communities. A large class of estimators of functionals of the probability density suffer from the curse of dimensionality, wherein the mean squared error (MSE) decays increasingly slowly as a function of the sample size $T$ as the dimension $d$ of the samples increases. In particular, the rate is often glacially slow of order $O(T^{-{\gamma}/{d}})$, where $\gamma>0$ is a rate parameter. Examples of such estimators include kernel density estimators, $k$-nearest neighbor ($k$-NN) density estimators, $k$-NN entropy estimators, intrinsic dimension estimators and other examples. In this paper, we propose a weighted affine combination of an ensemble of such estimators, where optimal weights can be chosen such that the weighted estimator converges at a much faster dimension invariant rate of $O(T^{-1})$. Furthermore, we show that these optimal weights can be determined by solving a convex optimization problem which can be performed offline and does not require training data. We illustrate the superior performance of our weighted estimator for two important applications: (i) estimating the Panter-Dite distortion-rate factor and (ii) estimating the Shannon entropy for testing the probability distribution of a random sample.
\section{Introduction} Non-linear functionals of probability densities $f$ of the form $G(f) = \int g(f(x),x) f(x) dx$ arise in applications of information theory, machine learning, signal processing and statistical estimation. Important examples of such functionals include Shannon $g(f,x)=-\log(f)$ and R\'enyi $g(f,x) = f^{\alpha-1}$ entropy, and the quadratic functional $g(f,x) = f^{2}$. In these applications, the functional of interest often must be estimated empirically from sample realizations of the underlying densities. Functional estimation has received significant attention in the mathematical statistics community. However, estimators of functionals of multivariate probability densities $f$ suffer from mean square error (MSE) rates which typically decrease with dimension $d$ of the sample as $O(T^{-{\gamma}/{d}})$, where $T$ is the number of samples and $\gamma$ is a positive rate parameter. Examples of such estimators include kernel density estimators~\cite{kde}, $k$-nearest neighbor ($k$-NN) density estimators~\cite{fuk2}, $k$-NN entropy functional estimators~\cite{hero,kks,litt}, intrinsic dimension estimators~\cite{kks}, divergence estimators~\cite{wang}, and mutual information estimators. This slow convergence is due to the curse of dimensionality. In this paper, we introduce a simple affine combination of an ensemble of such slowly convergent estimators and show that the weights in this combination can be chosen to significantly improve the rate of MSE convergence of the weighted estimator. In fact our ensemble averaging method can improve MSE convergence to the parametric rate $O(T^{-1})$. Specifically, for $d$-dimensional data, it has been observed that the variance of estimators of functionals $G(f)$ decays as $O(T^{-1})$ while the bias decays as $O(T^{-1/(1+d)})$. To accelerate the slow rate of convergence of the bias in high dimensions, we propose a weighted ensemble estimator for ensembles of estimators that satisfy conditions ${\mathscr C}.1$(\ref{BE}) and ${\mathscr C}.2$(\ref{VE}) defined in Sec. II below. Optimal weights, which serve to lower the bias of the ensemble estimator to $O(T^{-1/2})$, can be determined by solving a convex optimization problem. Remarkably, this optimization problem does not involve any density-dependent parameters and can therefore be performed offline. This then ensures MSE convergence of the weighted estimator at the parametric rate of $O(T^{-1})$. \subsection{Related work} When the density $f$ is $s > d/4$ times differentiable, certain estimators of functionals of the form $\int g(f(x),x) f(x) dx$, proposed by Birge and Massart~\cite{birge}, Laurent~\cite{laurent} and Gin\'e and Mason~\cite{gini}, can achieve the parametric MSE convergence rate of $O(T^{-1})$. The key ideas in ~\cite{birge,laurent,gini} are: (i) estimation of quadratic functionals $\int f^2(x) dx$ with MSE convergence rate $O(T^{-1})$; (ii) use of kernel density estimators with kernels that satisfy the following symmetry constraints: \begin{equation} \int K(x) dx =1, \hspace{0.5in} \int x^r K(x) dx = 0, \label{eq:symmetrickernels} \end{equation} for $r=1,..,s$; and finally (iii) truncating the kernel density estimate so that it is bounded away from $0$. By using these ideas, the estimators proposed by ~\cite{birge,laurent,gini} are able to achieve parametric convergence rates. In contrast, the estimators proposed in this paper require additional higher order smoothness conditions on the density, i.~e.~ the density must be $s>d$ times differentiable. However, our estimators are much simpler to implement in contrast to the estimators proposed in ~\cite{birge,laurent,gini}. In particular, the estimators in ~\cite{birge,laurent,gini} require separately estimating quadratic functionals of the form $\int f^2(x) dx$, and using truncated kernel density estimators with symmetric kernels (\ref{eq:symmetrickernels}), conditions that are not required in this paper. Our estimator is a simple affine combination of an ensemble of estimators, where the ensemble satisfies conditions ${\mathscr C}.1$ and ${\mathscr C}.2$. Such an ensemble can be trivial to implement. For instance, in this paper we show that simple uniform kernel plug-in estimators (\ref{eq:plugin}) satisfy conditions ${\mathscr C}.1$ and ${\mathscr C}.2$. Ensemble based methods have been previously proposed in the context of classification. For example, in both boosting~\cite{boosting} and multiple kernel learning~\cite{multipleK} algorithms, lower complexity weak learners are combined to produce classifiers with higher accuracy. Our work differs from these methods in several ways. First and foremost, our proposed method performs estimation rather than classification. An important consequence of this is that the weights we use are {\emph {data independent}}, while the weights in boosting and multiple kernel learning must be estimated from training data since they depend on the unknown distribution. \subsection{Organization} The remainder of the paper is organized as follows. We formally describe the weighted ensemble estimator for a general ensemble of estimators in Section~\ref{sec:genmeth}, and specify conditions ${\mathscr C}.1$ and ${\mathscr C}.2$ on the ensemble that ensure that the ensemble estimator has a faster rate of MSE convergence. Under the assumption that conditions ${\mathscr C}.1$ and ${\mathscr C}.2$ are satisfied, we provide an MSE optimal set of weights as the solution to a convex optimization(\ref{convexsoltheory}). Next, we shift the focus to entropy estimation in Section~\ref{sec:entropyest}, propose an ensemble of simple uniform kernel plug-in entropy estimators, and show that this ensemble satisfies conditions ${\mathscr C}.1$ and ${\mathscr C}.2$. Subsequently, we apply the ensemble estimator theory in Section~\ref{sec:genmeth} to the problem of entropy estimation using this ensemble of kernel plug-in estimators. We present simulation results in Section~\ref{sec:exp} that illustrate the superior performance of this ensemble entropy estimator in the context of (i) estimation of the Panter-Dite distortion-rate factor~\cite{riten} and (ii) testing the probability distribution of a random sample. We conclude the paper in Section~\ref{sec:con}. \subsection*{Notation} We will use bold face type to indicate random variables and random vectors and regular type face for constants. We denote the statistical expectation operator by the symbol ${{\mathbb{E}}}$ and the conditional expectation given random variable $\mathbf{Z}$ using the notation ${{\mathbb{E}}}_{\mathbf{Z}}$. We also define the variance operator as ${{\mathbb{V}}}[\mathbf{X}] = {{\mathbb{E}}}[(\mathbf{X}-{{\mathbb{E}}}[\mathbf{X}])^2] $ and the covariance operator as $\mathrm{Cov}[\mathbf{X},\mathbf{Y}] = {{\mathbb{E}}}[(\mathbf{X}-{{\mathbb{E}}}[\mathbf{X}])(\mathbf{Y}-{{\mathbb{E}}}[\mathbf{Y}])] $. We denote the bias of an estimator by $\mathbb{B}$. \section{Ensemble estimators} \label{sec:problem} \label{sec:genmeth} Let $\bar{l} = \{l_1,..,l_L\}$ denote a set of parameter values. For a parameterized ensemble of estimators $\{\hat{\mathbf{E}}_l\}_{l \in \bar{l}}$ of $E$, define the weighted ensemble estimator with respect to weights $w = \{w(l_1),\ldots,w(l_L)\}$ as $$\hat{\mathbf{E}}_w = \sum_{l \in \bar{l}} w(l)\hat{\mathbf{E}}_l$$ where the weights satisfy $\sum_{l \in \bar{l}} w(l) = 1$. This latter sum-to-one condition guarantees that $\hat{\mathbf{E}}_w$ is asymptotically unbiased if the component estimators $\{\hat{\mathbf{E}}_l\}_{l \in \bar{l}}$ are asymptotically unbiased. Let this ensemble of estimators $\{\hat{\mathbf{E}}_l\}_{l \in \bar{l}}$ satisfy the following two conditions: \begin{itemize} \item ${\mathscr C}.1$ The bias is given by \begin{eqnarray} \label{Bias_Ensemble} \label{BE} \mathbb{B}({\hat{\mathbf{E}}}_{l}) &=& \sum_{i \in {\cal I}} c_{i}\psi_i(l)T^{-{i/2d}} + O(1/\sqrt{T}), \end{eqnarray} where $c_{i}$ are constants that depend on the underlying density, ${\cal I} = \{i_1,..,i_I\}$ is a finite index set with cardinality $I<L$, $\min({\cal I}) = i_0 > 0$ and $\max({\cal I}) = i_d \leq d$, and $\psi_i(l)$ are basis functions that depend only on the estimator parameter $l$. \item ${\mathscr C}.2$ The variance is given by \begin{eqnarray} \label{Variance_Ensemble} \label{VE} \mathbb{V}({\hat{\mathbf{E}}}_{l}) &=& c_{v}\left({\frac{1}{T}}\right) + o\left(\frac{1}{T}\right). \end{eqnarray} \end{itemize} \begin{theorem} \label{lemma:weightedensemble} For an ensemble of estimators $\{\hat{\mathbf{E}}_l\}_{l \in \bar{l}}$, assume that the conditions ${\mathscr C}.1$ and ${\mathscr C}.2$ hold. Then, there exists a weight vector $w_o$ such that $${{\mathbb{E}}}[(\hat{\mathbf{E}}_{w_o} - E)^2] = O(1/T).$$ This weight vector can be found by solving the following convex optimization problem: \begin{equation} \begin{aligned} & \underset{w}{\text{minimize}} & & ||w||_2 \\ & \text{subject to} & & \sum_{l \in \bar{l}} w(l) = 1, \\ &&& \gamma_w(i) = \sum_{l \in \bar{l}} w(l)\psi_i(l) = 0, \; i \in {\cal I}, \end{aligned} \label{convexsoltheory} \end{equation} where $\psi_i(l)$ is the basis defined in (2.1). \end{theorem} \begin{proof} The bias of the ensemble estimator is given by \begin{eqnarray} \label{Bias_EnsembleEstimate} \mathbb{B}({\hat{\mathbf{E}}}_{w}) &=& \sum_{i \in {\cal I}} c_{i}\gamma_w(i)T^{-{i/2d}} + O\left(\frac{||w||_1}{\sqrt{T}}\right) \nonumber \\ &=& \sum_{i \in {\cal I}} c_{i}\gamma_w(i)T^{-{i/2d}} + O\left(\frac{\sqrt{L}||w||_2}{\sqrt{T}}\right) \end{eqnarray} Denote the covariance matrix of $\{\hat{\mathbf{E}}_l; l \in \bar{l}\}$ by $\Sigma_L$. Let $\bar{\Sigma}_L = \Sigma_LT$. Observe that by (\ref{Variance_Ensemble}) and the Cauchy-Schwarz inequality, the entries of $\bar{\Sigma}_L$ are $O(1)$. The variance of the weighted estimator $\hat{\mathbf{E}}_w$ can then be bounded as follows: \begin{eqnarray} \label{eq:var} {{\mathbb{V}}}(\hat{\mathbf{E}}_w) &=& {{\mathbb{V}}}(\sum_{l \in \bar{l}} w_l\hat{\mathbf{E}}_l) = w' \Sigma_L w = \frac{w' \bar{\Sigma}_L w}{T}\nonumber \\ &\leq& \frac{\lambda_{\max}(\bar{\Sigma}_L)||w||^2_2}{T} \leq \frac{trace(\bar{\Sigma}_L)||w||^2_2}{T} \leq \frac{L||w||^2_2}{T} \end{eqnarray} We seek a weight vector $w$ that (i) ensures that the bias of the weighted estimator is $O(T^{-1/2})$ and (ii) has low $\ell_2$ norm $||w||_2$ in order to limit the contribution of the variance, and the higher order bias terms of the weighted estimator. To this end, let $w_{o}$ be the solution to the convex optimization problem defined in (2.3). The solution $w_o$ is the solution of \begin{equation} \begin{aligned} & \underset{w}{\text{minimize}} & & ||w||^2_2 \\ & \text{subject to} & & A_0w = b, \end{aligned} \nonumber \label{convexsol2} \end{equation} where $A_0$ and $b$ are defined below. Let $a_0$ be the vector of ones: $[1,1. .. ,1]_{1 \times L}$; and let $a_{i}$, for each $i \in \cal{I}$ be given by $a_{i} = [\psi_i(l_1), .. ,\psi_i(l_L)]$. Define $A_0 = [a'_0, a'_{i_1}, ... , a'_{i_I}]'$, $A_1 = [a'_{i_1}, ... , a'_{i_I}]'$ and $b = [1;0;0;..;0]_{(I+1) \times 1}$. Since $L > I$, the system of equations $A_0 w = b$ is guaranteed to have at least one solution (assuming linear independence of the rows $a_i$). The minimum squared norm $\eta_L(d) := ||w_0||_2^2$ is then given by $$\eta_L(d) = {\frac{\text{det}(A_1A'_1)}{\text{det}(A_0A'_0)}}.$$ Consequently, by (\ref{Bias_EnsembleEstimate}), the bias $\mathbb{B}[\hat{\mathbf{E}}_{w_o}] = O(\sqrt{L\eta_L(d)}/\sqrt{T})$. By (\ref{eq:var}), the estimator variance ${{\mathbb{V}}}[\hat{\mathbf{E}}_{w_0}] = O(L\eta_L(d)/T)$. The overall MSE is also therefore of order $O(L\eta_L(d)/T)$. For any fixed dimension $d$ and fixed number of estimators $L>I$ in the ensemble independent of sample size $T$, the value of $\eta_L(d)$ is also independent of $T$. Stated mathematically, $L\eta_L(d) = \Theta(1)$ for any fixed dimension $d$ and fixed number of estimators $L>I$ independent of sample size $T$. This concludes the proof. \end{proof} In the next section, we will verify conditions $\mathscr C.1$(\ref{BE}) and $\mathscr C.2$(\ref{VE}) for plug-in estimators $\hat{\mathbf{G}}_{k}(f)$ of entropy-like functionals $G(f) = \int g(f(x),x) f(x) dx$. \section{Application to estimation of functionals of a density} \label{sec:entropyest} Our focus is the estimation of general non-linear functionals $G(f)$ of $d$-dimensional multivariate densities $f$ with known finite support ${\cal S} = [a,b]^d$, where $G(f)$ has the form \begin{equation} \label{eq:oracle} G(f) = \int g(f(x),x) f(x) dx, \end{equation} for some smooth function $g(f,x)$. Let ${\cal B}$ denote the boundary of ${\cal S}$. Assume that $T=N+M$ i.i.d realizations $\{\mathbf{X}_1, \ldots, \mathbf{X}_N, \mathbf{X}_{N+1}, \ldots, \mathbf{X}_{N+M}\}$ are available from the density $f$. \subsection{Plug-in estimators of entropy} \label{sec:weightedpluginest} The truncated \emph{uniform kernel} density estimator is defined below. For any positive real number $k \leq M$, define the distance $d_k$ to be: $d_k = (k/M)^{1/d}$. Define the truncated kernel region for each $X \in {\cal S}$ to be $S_k(X) = \{Y \in {\cal S} : ||X-Y||_\infty \leq d_k/2\}$, and the volume of the truncated uniform kernel to be $V_k(X) = \int_{S_k(X)} dz$. Note that when the smallest distance from $X$ to ${\cal B}$ is greater than $d_k/2$, $V_k(X) = d_k^d = k/M$. Let $\mathbf{l}_k(X)$ denote the number of samples falling in $S_k(X)$: $\mathbf{l}_k(X) = \sum_{i=1}^{M} 1_{\{\mathbf{X}_i \in S_k(X)\}}$. The truncated {uniform kernel} density estimator is defined as \begin{equation} \hat{\mathbf{f}}_{k}(X) = \frac{\mathbf{l}_k(X)}{MV_k(X)}. \end{equation} The plug-in estimator of the density functional is constructed using a data splitting approach as follows. The data is randomly subdivided into two parts $\{\mathbf{X}_1, \ldots, \mathbf{X}_N\}$ and $\{\mathbf{X}_{N+1}, \ldots, \mathbf{X}_{N+M}\}$ of $N$ and $M$ points respectively. In the first stage, we form the kernel density estimate ${\hat{\mathbf{f}}_k}$ at the $N$ points $\{\mathbf{X}_1, \ldots, \mathbf{X}_N\}$ using the $M$ realizations $\{\mathbf{X}_{N+1}, \ldots, \mathbf{X}_{N+M}\}$. Subsequently, we use the $N$ samples $\{\mathbf{X}_1, \ldots, \mathbf{X}_N\}$ to approximate the functional $G(f)$ and obtain the plug-in estimator: \begin{eqnarray} \label{eq:plugin} \hat{\mathbf{G}}_k &=& \frac{1}{N}\sum_{i=1}^N g({\hat{\mathbf{f}}{_k}(\mathbf{X}_i)},\mathbf{X}_i). \end{eqnarray} Also define a standard kernel density estimator $\tilde{\mathbf{f}}_k$, which is identical to $\hat{\mathbf{f}}_k$ except that the volume $V_k(X)$ is always set to the untruncated value $V_k(X) = k/M$. Define \begin{eqnarray} \label{eq:plugin2} \tilde{\mathbf{G}}_k &=& \frac{1}{N}\sum_{i=1}^N g({\tilde{\mathbf{f}}{_k}(\mathbf{X}_i)},\mathbf{X}_i). \end{eqnarray} The estimator $\tilde{\mathbf{G}}_k$ is identical to the estimator of Gy{\"{o}}rfi and van der Meulen~\cite{gyrofi}. Observe that the implementation of $\tilde{\mathbf{G}}_k$, unlike $\hat{\mathbf{G}}_k$, does not require knowledge about the support of the density. \subsubsection{Assumptions} \label{sec:assump} We make a number of technical assumptions that will allow us to obtain tight MSE convergence rates for the kernel density estimators defined above. $({\cal {A}}.0)$ : Assume that $k = k_0M^\beta$ for some rate constant $0<\beta<1$, and assume that $M$, $N$ and $T$ are linearly related through the proportionality constant $\alpha_{frac}$ with: $0 < \alpha_{frac} < 1$, $M = \alpha_{frac}T$ and $N = (1-\alpha_{frac})T$. $({\cal {A}}.1)$ : Let the density $f$ be uniformly bounded away from $0$ and upper bounded on the set ${\cal S}$, i.e., there exist constants $\epsilon_0$, $\epsilon_\infty$ such that $0 < \epsilon_0 \leq f(x) \leq \epsilon_\infty < \infty$ $\forall x \in {\cal S}$. $({\cal {A}}.2)$: Assume that the density $f$ has continuous partial derivatives of order $d$ in the interior of the set ${\cal S}$, and that these derivatives are upper bounded. $({\cal {A}}.3)$: Assume that the function $g(f,x)$ has $\max\{\lambda,d\}$ partial derivatives w.r.t. the argument $f$, where $\lambda$ satisfies the condition $\lambda \beta>1$. Denote the $n$-th partial derivative of $g(f,x)$ wrt $x$ by $g^{(n)}(f,x)$. $({\cal {A}}.4)$: Assume that the absolute value of the functional $g(f,x)$ and its partial derivatives are strictly upper bounded in the range $\epsilon_0 \leq f \leq \epsilon_\infty$ for all $x$. $({\cal {A}}.5)$: Let $\epsilon \in (0,1)$ and $\delta \in (2/3,1)$. Let ${\cal C}(M)$ be a positive function satisfying the condition $ {\cal C}(M) = \Theta(\exp(-M^{\beta(1-\delta)}))$. For some fixed $0 < \epsilon< 1$, define $p_l = (1-\epsilon)\epsilon_0$ and $p_u = (1+\epsilon)\epsilon_\infty $. Assume that the conditions $$ (i) \sup_{x}|h(0,x)| < G_1 < \infty, $$ $$ (ii) \sup_{f \in (p_l,p_u),x}|h(f,x)| < G_2 < \infty, $$ $$(iii) \sup_{f \in (1/k,p_u),x }|h(f,x)|{\cal C}(M) < G_3 < \infty \hspace{0.15in} \forall M,$$ $$(iv)\sup_{f \in (p_l,2^dM/k),x} |h(f,x)|{\cal C}(M) < G_4 < \infty \hspace{0.15in} \forall M,$$ are satisfied by $h(f,x) = g(f,x), g^{(3)}(f,x)$ and $g^{(\lambda)}(f,x)$, for some constants $G_1$, $G_2$, $G_3$ and $G_4$. These assumptions are comparable to other rigorous treatments of entropy estimation. The assumption $({\cal {A}}.0)$ is equivalent to choosing the bandwidth of the kernel to be a fractional power of the sample size~\cite{raykar}. The rest of the above assumptions can be divided into two categories: (i) assumptions on the density $f$, and (ii) assumptions on the functional $g$. The assumptions on the smoothness, boundedness away from $0$ and $\infty$ of the density $f$ are similar to the assumptions made by other estimators of entropy as listed in Section II,~\cite{beir}. The assumptions on the functional $g$ ensure that $g$ is sufficiently smooth and that the estimator is bounded. These assumptions on the functional are readily satisfied by the common functionals that are of interest in literature: Shannon $g(f,x) = - \log(f)I(f>0) + I(f=0)$ and R\'enyi $g(f,x) = f^{\alpha-1}I(f>0) + I(f=0)$ entropy, where $I(.)$ is the indicator function, and the quadratic functional $g(f,x) = f^{2}$. \subsubsection{Analysis of MSE} \label{sec:mseanal} Under the assumptions stated above, we have shown the following in the Appendix: \begin{theorem} \label{knnbiasH} The biases of the plug-in estimators $\hat{\mathbf{G}}_{k}, \tilde{\mathbf{G}}_{k}$ are given by \begin{eqnarray} \label{Biasu} \mathbb{B}(\hat{\mathbf{G}}_{k}) &=& \sum_{i=1}^{d} c_{1,i}\left({\frac{k}{M}}\right)^{i/d} + \frac{c_{2}}{k} \nonumber + o\left(\frac{1}{k} + \frac{k}{M}\right) \nonumber \\ \mathbb{B}(\tilde{\mathbf{G}}_{k}) &=& c_{1}\left({\frac{k}{M}}\right)^{1/d} + \frac{c_{2}}{k} \nonumber + o\left(\frac{1}{k} + \frac{k}{M}\right), \nonumber \end{eqnarray} where $c_{1,i}$, $c_1$ and $c_2$ are constants that depend on $g$ and $f$ \end{theorem} \begin{theorem} \label{knnvarH} The variances of the plug-in estimators $\hat{\mathbf{G}}_{k}, \tilde{\mathbf{G}}_{k}$ are identical up to leading terms, and are given by \begin{eqnarray} \label{Variance} {{\mathbb{V}}}(\hat{\mathbf{G}}_{k}) &=& c_4\left(\frac{1}{N}\right)+ c_5\left(\frac{1}{M}\right) + o\left(\frac{1}{M} + \frac{1}{N}\right) \nonumber \\ {{\mathbb{V}}}(\tilde{\mathbf{G}}_{k}) &=& c_4\left(\frac{1}{N}\right)+ c_5\left(\frac{1}{M}\right) + o\left(\frac{1}{M} + \frac{1}{N}\right), \nonumber \end{eqnarray} where $c_4$ and $c_5$ are constants that depend on $g$ and $f$. \end{theorem} \subsubsection{Optimal MSE rate} From Theorem \ref{knnbiasH}, observe that the conditions $k \to \infty$ and $k/M \to 0$ are necessary for the estimators $\hat{\mathbf{G}}_{k}$ and $\tilde{\mathbf{G}}_{k}$ to be unbiased. Likewise from Theorem \ref{knnvarH}, the conditions $N \to \infty$ and $M \to \infty$ are necessary for the variance of the estimator to converge to $0$. Below, we optimize the choice of bandwidth $k$ for minimum MSE, and also show that the optimal MSE rate is invariant to the choice of $\alpha_{frac}$. \paragraph{Optimal choice of $k$} Minimizing the MSE over $k$ is equivalent to minimizing the square of the bias over $k$. The optimal choice of $k$ is given by \begin{eqnarray} \label{kopt} k_{opt} &=& \Theta({M^{{1}/{1+d}}}), \end{eqnarray} and the bias evaluated at $k_{opt}$ is $\Theta({M^{{-1}/{1+d}}})$. \paragraph{Choice of $\alpha_{frac}$} Observe that the MSE of $\hat{\mathbf{G}}_{k}$ and $\tilde{\mathbf{G}}_{k}$ are dominated by the squared bias $(\Theta(M^{-2/(1+d)}))$ as contrasted to the variance $(\Theta(1/N+1/M))$. This implies that the asymptotic MSE rate of convergence is invariant to the selected proportionality constant $\alpha_{frac}$. In view of (a) and (b) above, the optimal MSE for the estimators $\hat{\mathbf{G}}{_k}$ and $\tilde{\mathbf{G}}{_k}$ is therefore achieved for the choice of $k=\Theta(M^{1/(1+d)})$, and is given by $\Theta(T^{-2/(1+d)})$. Our goal is to reduce the estimator MSE to $O(T^{-1})$. We do so by applying the method of weighted ensembles described in Section~\ref{sec:genmeth}. \subsection{Weighted ensemble entropy estimator} For a positive integer $L > I = d-1$, choose $\bar{l} = \{l_1, \ldots, l_L\}$ to be positive real numbers. Define the mapping $k(l) = l\sqrt{M}$ and let $\bar{k} = \{k(l); l \in \bar{l}\}$. Define the weighted ensemble estimator \begin{equation} \hat{\mathbf{G}}_w = \sum_{l \in \bar{l}} w(l)\hat{\mathbf{G}}_{k(l)}. \label{eq:weightedensemble} \end{equation} From Theorems \ref{knnbiasH} and \ref{knnvarH}, we see that the biases of the ensemble of estimators $\{\hat{\mathbf{G}}_{k(l)}; l \in \bar{l}\}$ satisfy ${\mathscr C}.1$(\ref{BE}) when we set $\psi_i(l) = l^{i/d}$ and ${\cal I} = \{1,..,d-1\}$. Furthermore, the general form of the variance of $\hat{\mathbf{G}}_{k(l)}$ follows ${\mathscr C}.2$(\ref{VE}) because $N, M = \Theta(T)$. This implies that we can use the weighted ensemble estimator $\hat{\mathbf{G}}_w$ to estimate entropy at $O(L\eta_L(d)/T)$ convergence rate by setting $w$ equal to the optimal weight $w_o$ given by (\ref{convexsoltheory}). \section{Experiments} \label{sec:exp} We illustrate the superior performance of the proposed weighted ensemble estimator for two applications: (i) estimation of the Panter-Dite rate distortion factor, and (ii) estimation of entropy to test for randomness of a random sample. For finite $T$ direct use of Theorem 1 can lead to excessively high variance. This is because forcing the condition (2.3) that $\gamma_w(i)=0$ is too strong and, in fact, not necessary. The careful reader may notice that to obtain $O(T^{-1}) $ MSE convergence rate in Theorem 1 it is sufficient that $\gamma_w(i)$ be of order $O(T^{-1/2+i/2d})$. Therefore, in practice we determine the optimal weights according to the optimization: \begin{equation} \begin{split} \min_w \quad &\epsilon\\ \text{subject to} \quad &\gamma_w(0) = 1,\\ &\lvert \gamma_w(i) T^{1/2 - i/2d} \rvert \leq \epsilon, \quad i \in \mathcal{I},\\ &\lVert w \rVert_2^2 \leq \eta. \end{split} \label{convexsol} \end{equation} The optimization \eqref{convexsol} is also convex. Note that, as contrasted to \eqref{convexsoltheory}, the norm of the weight vector $w$ is bounded instead of being minimized. By relaxing the constraints $\gamma_w(i)=0$ in \eqref{convexsoltheory} to the softer constraints in \eqref{convexsol}, the upper bound $\eta$ on $\lVert w \rVert_2^2$ can be reduced from the value $\eta_L(d)$ obtained by solving \eqref{convexsoltheory}. This results in a more favorable trade-off between bias and variance for moderate sample sizes. In our experiments, we find that setting $\eta = 3d$ yields good MSE performance. Note that as $T \to \infty$, we must have $\gamma_w(i) \to 0$ for $i \in \mathcal{I}$ in order to keep $\epsilon$ finite, thus recovering the strict constraints in \eqref{convexsoltheory}. For fixed sample size $T$ and dimension $d$, observe that increasing $L$ increases the number of degrees of freedom in the convex problem \eqref{convexsol}, and therefore will result in a smaller value of $\epsilon$ and in turn improved estimator performance. In our simulations, we choose $\bar{l}$ to be $L=50$ equally spaced values between $0.3$ and $3$, ie the $l_i$ are uniformly spaced as $$l_i = \frac{x}{a} + \frac{(a-1)ix}{aL}; i=1,..,L,$$ with scale and range parameters $a=10$ and $x=3$ respectively. We limit $L$ to 50 because we find that the gains beyond $L=50$ are negligible. The reason for this diminishing return is a direct result of the increasing similarity among the entries in $\bar{l}$, which translates to increasingly similar basis functions $\psi_i(l) = l^{i/d}$. \subsection{Panter-Dite factor estimation} \begin{figure}[!t] \centering \subfigure[\small{Variation of MSE of Panter-Dite factor estimates as a function of sample size $T$. From the figure, we see that the proposed weighted estimator has the fastest MSE rate of convergence wrt sample size $T$ ($d=6$).}]{ \includegraphics[scale=.25]{MSEvsT.png} \label{a-compare} } \subfigure[\small{Variation of MSE of Panter-Dite factor estimates as a function of dimension $d$. From the figure, we see that the MSE of the proposed weighted estimator has the slowest rate of growth with increasing dimension $d$ ($T=3000$).}]{ \includegraphics[scale=.25]{MSEvsd.png} \label{ad-compare} } \caption{Variation of MSE of Panter-Dite factor estimates using standard kernel plug-in estimator~[14], truncated kernel plug-in estimator~(3.3), histogram plug-in estimator[17], $k$-NN estimator~[20], entropic graph estimator~[18] and the weighted ensemble estimator~(3.6). } \end{figure} For a $d$-dimensional source with underlying density $f$, the Panter-Dite distortion-rate function~\cite{riten} for a $q$-dimensional vector quantizer with $n$ levels of quantization is given by $ \delta(n)= n^{-2/q} \int f^{q/(q+2)}(x) dx. $ The Panter-Dite factor corresponds to the functional $G(f)$ with $g(f,x) = n^{-2/q} f^{-2/(q+2)}I(f>0) + I(f=0)$. The Panter-Dite factor is directly related to the R\'{e}nyi $\alpha$-entropy, for which several other estimators have been proposed~\cite{gyrofi0,heroJ,pal,leo2}. In our simulations we compare six different choices of functional estimators - the three estimators previously introduced: (i) the standard kernel plug-in estimator $\tilde{\mathbf{G}}_k$, (ii) the boundary truncated plug-in estimator $\hat{\mathbf{G}}_k$ and (iii) the weighted estimator $\hat{\mathbf{G}}_w$ with optimal weight $w = w^*$ given by (\ref{convexsol}), and in addition the following popular entropy estimators: (iv) histogram plug-in estimator~\cite{gyrofi0}, (v) $k$-nearest neighbor ($k$-NN) entropy estimator~\cite{leo2} and (vi) entropic $k$-NN graph estimator~\cite{heroJ,pal}. For both $\tilde{\mathbf{G}}_k$ and $\hat{\mathbf{G}}_k$, we select the bandwidth parameter $k$ as a function of $M$ according to the optimal proportionality $k=M^{1/(1+d)}$ and $N=M=T/2$. We choose $f$ to be the $d$ dimensional mixture density $f(a,b,p,d) = pf_\beta(a,b,d) + (1-p)f_u(d)$; where $d=6$, $f_\beta(a,b,d)$ is a $d$-dimensional Beta density with parameters $a=6,b=6$, $f_u(d)$ is a $d$-dimensional uniform density and the mixing ratio $p$ is $0.8$. The reason we choose the beta-uniform mixture for our experiments is because it trivially satisfies all the assumptions on the density $f$ listed in Section 3.1, including the assumptions of finite support and strict boundedness away from 0 on the support. The true value of the Panter-Dite factor $\delta(n)$ for the beta-uniform mixture is calculated using numerical integration methods via the 'Mathematica' software (http://www.wolfram.com/mathematica/). Numerical integration is used because evaluating the entropy in closed form for the beta-uniform mixture is not tractable. The MSE values for each of the six estimators are calculated by averaging the squared error $[\hat{\delta}_i(n) - \delta(n)]^2$, $i=1,..,m$ over $m = 1000$ Monte-Carlo trials, where each $\hat{\delta}_i(n)$ corresponds to an independent instance of the estimator. \subsubsection{Variation of MSE with sample size $T$} The MSE results of the different estimators are shown in Fig.~\ref{a-compare} as a function of sample size $T$, for fixed dimension $d=6$. It is clear from the figure that the proposed ensemble estimator $\hat{\mathbf{G}}_w$ has significantly faster rate of convergence while the MSE of the rest of the estimators, including the truncated kernel plug-in estimator, have similar, slow rates of convergence. It is therefore clear that the proposed optimal ensemble averaging significantly accelerates the MSE convergence rate. \subsubsection{Variation of MSE with dimension $d$} For fixed sample size $T$ and fixed number of estimators $L$, it can be seen that $\epsilon$ increases monotonically with $d$. This follows from the fact that the number of constraints in the convex problem \ref{convexsol} is equal to $d+1$ and each of the basis functions $\psi_i(l) = l^{i/d}$ monotonically approaches $1$ as $d$ grows, . This in turn implies that for a fixed sample size $T$ and number of estimators $L$, the overall MSE of the ensemble estimator should increase monotonically with the dimension $d$. The MSE results of the different estimators are shown in Fig.~\ref{ad-compare} as a function of dimension $d$, for fixed sample size $T=3000$. For the standard kernel plug-in estimator and truncated kernel plug-in estimator, the MSE increases rapidly with $d$ as expected. The MSE of the histogram and $k$-NN estimators increase at a similar rate, indicating that these estimators suffer from the curse of dimensionality as well. On the other hand, the MSE of the weighted estimator also increases with the dimension as predicted, but at a slower rate. Also observe that the MSE of the weighted estimator is smaller than the MSE of the other estimators for all dimensions $d>3$. \subsection{Distribution testing} \begin{figure}[!t] \centering \subfigure[\small{Entropy estimates for random samples corresponding to hypothesis $H_0$ (experiments 1-500) and $H_1$ (experiments 501-1000).}]{ \includegraphics[scale=.30]{responsemod.png} \label{c-compare} } \subfigure[\small{Histogram envelopes of entropy estimates for random samples corresponding to hypothesis $H_0$ (blue) and $H_1$ (red).}]{ \includegraphics[scale=.30]{histmod.png} \label{d-compare} } \caption{Entropy estimates using standard kernel plug-in estimator, truncated kernel plug-in estimator and the weighted estimator, for random samples corresponding to hypothesis $H_0$ and $H_1$. The weighted estimator provides better discrimination ability by suppressing the bias, at the cost of some additional variance.} \end{figure} In this section, we illustrate the weighted ensemble estimator for non-parametric estimation of Shannon differential entropy. The Shannon differential entropy is given by $G(f)$ where $g(f,x) = - \log(f)I(f>0) + I(f=0)$. The improved accuracy of the weighted ensemble estimator is demonstrated in the context of hypothesis testing using estimated entropy as a statistic to test for the underlying probability distribution of a random sample. Specifically, the samples under the null and alternate hypotheses $H_0$ and $H_1$ are drawn from the probability distribution $f(a,b,p,d)$, described in Section IV.A, with fixed $d=6$, $p=0.75$ and two sets of values of $a,b$ under the null and alternate hypothesis, $H_0: a=a_0,b=b_0$ versus $H_1: a=a_1,b=b_1$. First, we fix $a_0=b_0=6$ and $a_1=b_1=5$. The density under the null hypothesis $f(6,6,0.75,6)$ has greater curvature relative to $f(5,5,0.75,6)$ and therefore has smaller entropy. Five hundred (500) experiments are performed under each hypothesis with each experiment consisting of 1000 samples drawn from the corresponding distribution. The true entropy and estimates $\tilde{\mathbf{G}}_k$, $\hat{\mathbf{G}}_k$ and $\hat{\mathbf{G}}_w$ obtained from each instance of $10^3$ samples are shown in Fig.~\ref{c-compare} for the 1000 experiments. This figure suggests that the ensemble weighted estimator provides better discrimination ability by suppressing the bias, at the cost of some additional variance. To demonstrate that the weighted estimator provides better discrimination, we plot the histogram envelope of the entropy estimates using standard kernel plug-in estimator, truncated kernel plug-in estimator and the weighted estimator for the cases corresponding to the hypothesis $H_0$ (color coded {blue}) and $H_1$ (color coded {red}) in Fig.~\ref{d-compare}. Furthermore, we quantitatively measure the discriminative ability of the different estimators using the deflection statistic $ds = {|\mu_1-\mu_0|}/{\sqrt{\sigma_0^2+\sigma_1^2}},$ where $\mu_0$ and $\sigma_0$ (respectively $\mu_1$ and $\sigma_1$) are the sample mean and standard deviation of the entropy estimates. The deflection statistic was found to be $1.49$, $1.60$ and $1.89$ for the standard kernel plug-in estimator, truncated kernel plug-in estimator and the weighted estimator respectively. The receiver operating curves (ROC) for this entropy-based test using the three different estimators are shown in Fig.~\ref{b-compare}. The corresponding areas under the ROC curves (AUC) are given by $0.9271$, $0.9459$ and $0.9619$. \begin{figure}[!t] \centering \subfigure[\small{ROC curves corresponding to entropy estimates obtained using standard and truncated kernel plug-in estimators and the weighted estimator. The corresponding AUC are given by $0.9271$, $0.9459$ and $0.9619$.}]{ \includegraphics[scale=.30]{ROC1000.pdf} \label{b-compare} } \subfigure[\small{Variation of AUC curves vs $\delta (= a_0-a_1, b_0-b_1)$ corresponding to Neyman-Pearson omniscient test, entropy estimates using the standard and truncated kernel plug-in estimators and the weighted estimator.}]{ \includegraphics[scale=.30]{auc10000.pdf} \label{e-compare} } \caption{Comparison of performance in terms of ROC for the distribution testing problem. The weighted estimator uniformly outperforms the individual plug-in estimators.} \end{figure} In our final experiment, we fix $a_0=b_0=10$ and set $a_1=b_1=10-\delta$, perform 500 experiments each under the null and alternate hypotheses with samples of size 5000, and plot the AUC as $\delta$ varies from $0$ to $1$ in Fig.~\ref{e-compare}. For comparison, we also plot the AUC for the Neyman-Pearson likelihood ratio test. The Neyman-Pearson likelihood ratio test, unlike the Shannon entropy based tests, is an omniscient test that assumes knowledge of both the underlying beta-uniform mixture parametric model of the density and the parameter values $a_0$, $b_0$ and $a_1$, $b_1$ under the null and alternate hypothesis respectively. Figure 4 shows that the weighted estimator {\emph {uniformly and significantly}} outperforms the individual plug-in estimators and comes closest to the performance of the omniscient Neyman-Pearson likelihood test. The relatively superior performance of the Neyman-Pearson likelihood test is due to the fact that the weighted estimator is a nonparametric estimator that has marginally higher variance (proportional to $||w^*||_2^2$) as compared to the underlying parametric model for which the Neyman-Pearson test statistic provides the most powerful test. \section{Conclusions} \label{sec:con} We have proposed a new estimator of functionals of a multivariate density based on weighted ensembles of kernel density estimators. For ensembles of estimators that satisfy general conditions on bias and variance as specified by ${\mathscr C}.1$(\ref{BE}) and ${\mathscr C}.2$(\ref{VE}) respectively, the weight optimized ensemble estimator has parametric $O(T^{-1})$ MSE convergence rate that can be much faster than the rate of convergence of any of the individual estimators in the ensemble. The optimal weights are determined as a solution to a convex optimization problem that can be performed offline and does not require training data. We illustrated this estimator for uniform kernel plug-in estimators and demonstrated the superior performance of the weighted ensemble entropy estimator for (i) estimation of the Panter-Dite factor and (ii) non-parametric hypothesis testing. Several extensions of the framework of this paper are being pursued: (i) using $k$-nearest neighbor ($k$-NN) estimators in place of kernel estimators; (ii) extending the framework to the case where support ${\cal S}$ {\emph {is not known}}, but for which conditions ${\mathscr C}.1$ and ${\mathscr C}.2$ hold; (iii) using ensemble estimators for estimation of other functionals of probability densities including divergence, mutual information and intrinsic dimension; and (iv) using an $l_1$ norm $\|w\|_1$in place of the $l_2$ norm $\|w\|_2$ in the weight optimization algorithm (2.3) so as to introduce sparsity into the weighted ensemble. \section*{Acknowledgement} This work was partially supported by (i) ARO grant W911NF-12-1-0443 and (ii) NIH grant 2P01CA087634-06A2. \begin{figure}[h] \centering \includegraphics[width=6.2in]{sss.pdf} \caption{Illustration for the proof of Lemma~\ref{biaslemma}.} \label{i-compare} \end{figure}
{ "timestamp": "2013-03-05T02:01:58", "yymm": "1203", "arxiv_id": "1203.5829", "language": "en", "url": "https://arxiv.org/abs/1203.5829", "abstract": "The problem of estimation of density functionals like entropy and mutual information has received much attention in the statistics and information theory communities. A large class of estimators of functionals of the probability density suffer from the curse of dimensionality, wherein the mean squared error (MSE) decays increasingly slowly as a function of the sample size $T$ as the dimension $d$ of the samples increases. In particular, the rate is often glacially slow of order $O(T^{-{\\gamma}/{d}})$, where $\\gamma>0$ is a rate parameter. Examples of such estimators include kernel density estimators, $k$-nearest neighbor ($k$-NN) density estimators, $k$-NN entropy estimators, intrinsic dimension estimators and other examples. In this paper, we propose a weighted affine combination of an ensemble of such estimators, where optimal weights can be chosen such that the weighted estimator converges at a much faster dimension invariant rate of $O(T^{-1})$. Furthermore, we show that these optimal weights can be determined by solving a convex optimization problem which can be performed offline and does not require training data. We illustrate the superior performance of our weighted estimator for two important applications: (i) estimating the Panter-Dite distortion-rate factor and (ii) estimating the Shannon entropy for testing the probability distribution of a random sample.", "subjects": "Statistics Theory (math.ST); Methodology (stat.ME)", "title": "Ensemble estimators for multivariate entropy estimation", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9845754501811438, "lm_q2_score": 0.8128673178375734, "lm_q1q2_score": 0.8003292053974677 }
https://arxiv.org/abs/1302.7252
Stability results for some fully nonlinear eigenvalue estimates
In this paper, we give some stability estimates for the Faber-Krahn inequality relative to the eigenvalues of Hessian operators
\section{Introduction} In this paper we prove some stability estimates for the eigenvalue $\lambda_k(\Omega)$ of the $k$-Hessian operator, that has the variational characterization \begin{equation*} \lambda_k(\Omega) = \min\left\{\int_\Omega (-u) S_k(D^2u)\,dx,\; u\in \Phi_k^2(\Omega) \text{ and } \int_\Omega (-u)^{k+1}\,dx=1 \right\}. \end{equation*} Here $\Omega$ is a bounded, strictly convex, open set of $\mathbb R^n$, $n\ge 2$, with $C^2$ boundary, $\Sk_k(D^2u)$, with $1\le k\le n$, is the $k$-th elementary symmetric function of the eigenvalues of $D^2 u$ with $u\in C^2(\Omega)$, and $\Phi_k^2(\Omega)$ denotes the class of the admissible functions for $\Sk_k$, the so-called $k$-convex functions (see Section 2 for the precise definitions). Notice that $S_1(D^2u)=\Delta u$, the Laplacian operator, while $\Sk_n(D^2u)=\det D^2 u$, the Monge-Amp\`ere operator. It is known that, under suitable assumptions on $\Omega$, for this kind of operators a Faber-Krahn inequality holds, that is the eigenvalue $\lambda_k(\Omega)$ attains its minimum value on the ball $\Omega_{k-1}^*$, which preserves an appropriate curvature measure of $\Omega$, the $(k-1)$-th quermassintegral: \begin{equation} \label{introfksk} \lambda_k(\Omega) \ge \lambda_k(\Omega^*_{k-1}), \quad 1\le k\le n \end{equation} (see \cite{bt07,ga09}). Moreover, for $k=n$, $\lambda_n(\Omega)$ is also bounded from above by $\lambda_n(\Omega^*_0)$, with $|\Omega^*_0|=|\Omega|$ (see \cite{bntpoin}). For sake of completeness, we recall that in the case of Neumann boundary condition, for $k=1$, the reverse inequality in \eqref{introfksk} holds (see \cite{w56,sz54}, \cite{chdb12} and \cite{brchtr09,dpg2} for related results). In \cite{dpg5} we give some stability estimates of \eqref{introfksk}, proving that \begin{equation*} \frac{\lambda_k(\Omega) - \lambda_k(\Omega^*_{k-1})}{\lambda_k(\Omega)} \le C_{n,k} \frac{|\Omega^*_k|-|\Omega|}{|\Omega^*_k|}, \quad 1\le k\le n-1, \end{equation*} for some constant $C_{n,k}$ depending only on $n$ and $k$, while for $k=n$, \begin{equation*} \frac{\lambda_n(\Omega) - \lambda_n(\Omega^*_{n-1})}{\lambda_n(\Omega)} \le C_n \frac{|\Omega^*_{n-1}|-|\Omega|}{|\Omega^*_{n-1}|}, \end{equation*} where $C_n$ which depends only on $n$. Roughly speaking, such inequalities state that if $\Omega$ is close to a ball with respect the $L^1$ norm, then their corresponding eigenvalues are near. Such result is in the spirit of a well-known result due to Payne and Weinberger for the Laplace operator (see \cite{pw61}), and given in \cite{bnt10} for the $p$-Laplace operator (see also \cite{nitsch12} for the best constant in the case $p=2$, and \cite{dpgtors} for the anisotropic case). Viceversa, the aim of this paper is to prove some stability results which ensure that if $\lambda_k(\Omega)$ is near to $\lambda_k(\Omega^*_{k-1})$, then, in an appropriate sense, $\Omega$ is close to a suitable ball of $\mathbb R^n$ (see Section 2 for the precise statements). There are several contributions in this direction, for the first eigenvalue of the Laplacian operator (see \cite{melas92}, \cite{hana94}) or, more generally, for the $p$-Laplacian (see \cite{bhatta01}, \cite{fmp09}). In such papers, depending on the assumptions on $\Omega$, suitable notions of the distance between the set $\Omega$ and a ball are considered. In particular, under the convexity assumption on the domain, it seems natural to take into account the Hausdorff distance (see \cite{melas92}), while, in a more general setting, such notion is replaced by the so-called Fraenkel asymmetry (see \cite{bhatta01}, \cite{fmp09}). Both arguments are considered in \cite{hana94}. Dealing with convex sets, our aim is to prove some stability result for Hessian operators in the spirit of the results given in \cite{melas92,hana94}. In particular, we prove that for a strictly convex, smooth domain $\Omega$, such that \[ \lambda_k(\Omega) \le \lambda_k(\Omega^*_{k-1})(1+\varepsilon), \] for some $\varepsilon>0$ sufficiently small, then there exist two balls $B_{r_\Omega}$ and $B_{R_\Omega}$ such that $B_{r_\Omega}\subseteq \Omega \subseteq B_{R_\Omega}$ and two suitable asymmetry coefficients of $\Omega$ with respect to $\Omega^*_{k-1}$ vanish when $\varepsilon$ goes to zero. This will imply that the Hausdorff asymmetry of $\Omega$ is close to zero (see Section 2.4 for the precise definitions). The paper is organized as follows. In Section 2, we recall some basic definitions of convex geometry, and the properties of symmetrization for quermassintegrals. Moreover, we summarize some useful results on the eigenvalue problem for Hessian operators. In sections 3 and 4 we state and prove the main results. We distinguish the case of the Monge-Amp\`ere operator (see Section 3) from the case of $\Sk_k$, $1\le k\le n-1$. Our approach makes use of a quantitative version of a suitable isoperimetric inequality and a symmetrization for quermassintegral technique. \section{Notation and preliminaries} Throughout the paper, we will denote with $\Omega$ a set of $\mathbb R^n$, $n\ge 2$ such that \begin{equation} \label{ipomega} \Omega\text{ is a bounded, strictly convex, open set with }C^2 \text{ boundary}. \end{equation} By strict convexity of $\Omega$ we mean that the Gauss curvature is strictly positive at every point of $\partial \Omega$. Given a function $u \in C^2(\Omega)$, we denote by $\lambda(D^2u)= (\lambda_1,\lambda_2, \ldots,\lambda_n)$ the vector of the eigenvalues of $D^2u$. The $k$-Hessian operator $\Sk_k(D^2u)$, with $k=1,2,\ldots,n$, is \begin{equation*} \Sk_k(D^2u)=\sum_{i_1<i_2<\cdots<i_k} \lambda_{i_1} \cdot \lambda_{i_2} \cdots \lambda_{i_k}. \end{equation*} Hence $\Sk_k(D^2u)$ is the sum of all $k \times k$ principal minors of the matrix $D^2u$. The $k$-Hessian operator can be written also in divergence form, that is \begin{equation*} \label{div} \Sk_k(D^2u)=\frac{1}{k}\sum_{i,j=1}^n (\Sk_k^{ij}u_i)_j, \end{equation*} where $S_k^{ij} = \frac{\partial S_k(D^2u)}{\partial u_{ij}}$ (see for instance \cite{trudi1}, \cite{trudi2}, \cite{wangeigen}). Well known examples ok $k$-Hessian operators are $\Sk_1(D^2u)=\Delta u$, the Laplace operator, and $\Sk_n(D^2u)=\det(D^2u)$, the Monge-Amp\`ere operator. It is well-known that $\Sk_1(D^2u)$ is elliptic. This property is not true in general for $k>1$. As matter of fact, the $k$-Hessian operator is elliptic when it acts on the class of the so-called $k$-convex function, defined below. \begin{definiz} Let $\Omega$ be as in \eqref{ipomega}. A function $u \in C^2(\Omega)$ is called a $k$-convex function (strictly $k$-convex) in $\Omega$ if \begin{equation*} \Sk_j(D^2u)\geq 0 \text{ }(>0) \quad \text{for }j=1, \ldots, k. \end{equation*} We denote the class of $k$-convex functions in $\Omega$ such that $u \in C^2(\Omega)\cap C(\bar{\Omega})$ and $u=0$ on $\partial \Omega$ by $\Phi^2_k(\Omega)$. \end{definiz} Clearly, the set $\Phi_n^2(\Omega)$ coincides with the set of the convex and $C^2(\Omega)$ functions vanishing on $\partial \Omega$. If we define with $\Gamma_k$ the following convex open cone \begin{equation*} \Gamma_k=\{ \lambda \in \mathbb R^n : S_1(\lambda)>0, S_2(\lambda)>0, \ldots, S_k(\lambda)>0\}, \end{equation*} in \cite{ivo} it is proven that $\Gamma_k$ is the cone of ellipticity of $\Sk_k$. Hence the $k$-Hessian operator is elliptic with respect to the $k$-convex functions. By definition, it follows that the $k$-convex functions are subharmonic in $\Omega$ and then negative in $\Omega$ if zero on $\partial\Omega$. We go on by recalling some definitions of convex geometry which will be largely used in next sections. Standard references for this topic are \cite{bz}, \cite{schn}. \subsection {Quermassintegrals and the Aleksandrov-Fenchel inequalities} Let $K$ be a convex body, and let be $\rho>0$. We denote by $|K|$ the Lebesgue measure of $K$, by $P(K)$ the perimeter of $K$ and by $\omega_n$ the measure of the unit ball in $\mathbb R^n$. The well-known Steiner formula for the Minkowski sum is \[ |K+\rho B_1| =\sum_{i=0}^{n} \binom{n}{i} W_i(K) \rho^i. \] The coefficient $W_i(K)$, $i=0,\ldots,n$, is known as the $i$-th quermassintegral of $K$. Some special cases are $W_0(K)=|K|$, $nW_1(K)=P(K)$, $W_n(K)=\omega_n$. If $K$ has $C^2$ boundary, with nonvanishing Gaussian curvature, the quermassintegrals can be related to the principal curvatures of $\partial K$. Indeed, in such a case \[ W_i(K)=\frac 1 n \int_{\partial K} H_{i-1} d \mathcal H^{n-1}, \quad i={1,\ldots n}. \] Here $H_j$ denotes the $j$-th normalized elementary symmetric function of the principal curvatures $\kappa_1,\ldots,\kappa_{n-1}$ of $\partial K$, that is $H_0=1$ and \[ H_j= \binom{n-1}{j}^{-1} \sum_{1\le i_1\le \ldots \le i_j\le n-1} \kappa_{i_1}\cdots \kappa_{i_j},\quad j={1,\ldots,n-1}. \] An immediate computation shows that if $B_R$ is a ball of radius $R$, then \begin{equation} \label{querball} W_i(B_R)= \omega_n R^{n-i}, \quad i=0,\ldots,n. \end{equation} Moreover, the $i$-th quermassintegral, $0\le i \le n$, rescales as \[ W_{i}(tK)=t^{n-i}W_i(K), \quad t>0. \] The Aleksandrov-Fenchel inequalities state that \begin{equation} \label{afineq} \left( \frac{W_j(K)}{\omega_n} \right)^{\frac{1}{n-j}} \ge \left( \frac{W_i(K)}{\omega_n} \right)^{\frac{1}{n-i}}, \quad 0\le i < j \le n-1, \end{equation} where the inequality is replaced by an equality if and only if $K$ is a ball. In what follows, we use the Aleksandrov-Fenchel inequalities for particular values of $i$ and $j$. If $i=1$, and $j=k-1$, we have that \begin{equation}\label{af-2} W_{k-1}(K) \ge \omega_n^{\frac{k-2}{n-1}} n^{-\frac{n-k+1}{n-1}} P(K)^{\frac{n-k+1}{n-1}}, \quad 3\le k \le n. \end{equation} When $i=0$ and $j=1$, we have the classical isoperimetric inequality: \[ P(K) \ge n \omega_n^{\frac 1 n} |K|^{1-\frac 1 n}. \] Moreover, if $i=k-1$, and $j=k$, we have \[ W_k(K) \ge \omega_n^{\frac{1}{n-k+1}} W_{k-1}(K)^{\frac{n-k}{n-k+1}}. \] It can be also shown a derivation formula for quermassintegral of level sets of a function $u \in \Phi_k^2(\Omega)$ (see \cite{reilly}): \begin{equation} \label{reillyder} - \frac{d}{dt} W_{k}(\Omega_t)= \frac{n-k}{n} \int_{\Sigma_t}H_{k}(\Sigma_t) |Du|^{-1} d \mathcal H^{n-1}, \end{equation} where $\Sigma_t$ is the boundary of $\Omega_t=\{ -u > t \}$. Moreover, we recall the following equality (see \cite{reilly} again): \begin{equation} \label{introintcurv} \int_{\Omega_t}\Sk_k (D^2u)dx = \frac{1}{k} \int_{\Sigma_t} H_{k-1} |Du|^k d\mathcal H^{n-1}. \end{equation} \subsection{Rearrangements for quermassintegrals} Now we recall some basic facts on rearrangements for quermassintegrals. For an exhaustive treatment of the properties of such rearrangements we refer the reader, for example, to \cite{ta81}, \cite{tso}, \cite{trudiso}. Let $1\le k\le n$, and denote by $\Omega^*_{k-1}$ the ball centered at the origin and with the same $W_{k-1}$-measure than $\Omega$, that is $W_{k-1}(\Omega^*_{k-1})=W_{k-1}(\Omega)$. The $(k-1)$-symmetrand of a function $u \in \Phi_{k}(\Omega)$, $k=1,\ldots,n$, is the radially symmetric increasing function $u^*_{k-1}$, defined in the ball $\Omega^*_{k-1}$, which preserves the $W_{k-1}$-measure of the level sets of $u$. More precisely, we have that, for $x\in\Omega^*_{k-1}$, \begin{equation*} u_{k-1}^{*}(x)=-\inf \left\{ t \ge 0 \colon W_{k-1}(\Omega_t) \le \omega_n\left|x\right|^{n-k+1},\; Du \ne 0 \text{ on } \Sigma_t\right\}, \end{equation*} where $\Sigma_{t} = \partial \Omega_t=\left\{x \in \Omega\colon -u(x)>t\right\}$, with $t\ge 0$. We stress that, for $k=1$, $u^{*}_{0}(x)$ coincides with the classical Schwarz symmetrand of $u$, while for $k=2$, $u^*_1(x)$ is the rearrangement of $u$ which preserves the perimeter of the level sets of $u$. Denoting with $R$ the radius of $\Omega^*_{k-1}$, the following statements hold true (see \cite{tso,trudiso}): \begin{enumerate} \item writing $u_{k-1}^{*}(x) = \rho(r)$ for $r=\left|x\right|$, we have $\rho(0)=\min_{\Omega}(u)$ and $\rho(R)=0$, \item $\rho(r)$ is a negative and increasing function on $\left[0,R\right]$, \item $\rho(r) \in C^{0,1}(\left[0,R\right])$ and moreover $0 \le \rho'(r)\leq \sup_{\Omega}\left|Du\right|$ almost everywhere. \end{enumerate} If the function $u$ has convex level sets, the Aleksandrov-Fenchel inequalities \eqref{afineq} imply that $|\{ -u>t \}|\le |\{-u^*_{k-1}>t\}|$ and then, for any $p\ge 1$, \begin{equation*} \label{norme} \left\|u\right\|_{L^p(\Omega)} \le \left\|u_{k-1}^{*}\right\|_{L^p(\Omega^*_{k-1})}, \end{equation*} while, by property $(1)$, \[ \|u\|_{{L^\infty(\Omega)}} =\|u^*_{k-1}\|_{{L^\infty(\Omega^*_{k-1})}}. \] Now, it is possible to define the following functional associated to the $k$-Hessian operator, known as $k$-Hessian integral: \begin{equation*} I_k\left[u,\Omega \right]=\int_{\Omega} (-u)\Sk_k(D^2u) \, dx. \end{equation*} In the radial case the Hessian integrals can be defined as follows: \begin{equation*} I_{k}\left[ u^{*}_{k-1}, \Omega^*_{k-1}\right]= \binom{n}{k} \omega_n \int_0^R f^{k+1}\big(\omega_n r^{n-k+1}\big) r^{n-k} \, dr \end{equation*} where $f\big(\omega_n \left|x\right|^{n-k+1}\big) = \left|\nabla u^{*}_{k-1}(x)\right|$. Finally we recall that for the Hessian integrals the following extension of P\'olya-Szeg\"o principle holds (see \cite{trudiso,tso}): \begin{equation} \label{pol} I_{k}\left[u, \Omega\right]\geq I_{k}\left[u^{*}_{k-1}, \Omega^*_{k-1}\right], \quad p \geq 1. \end{equation} \subsection{Eigenvalue problems for $\Sk_k$} In this subsection we give a quick review on the main properties of eigenvalues and eigenfunctions of the $k$-Hessian operators, namely the couples $(\lambda,u)$ which solve \begin{equation} \label{autsk} \left\{ \begin{array}{ll} S_k(D^2u)=\lambda (-u)^k &\text{in } \Omega,\\ u=0 &\text{on } \partial\Omega. \end{array} \right. \end{equation} The following existence result holds (see \cite{lionsma} for $k=n$, and \cite{wangeigen}, \cite{geng} in the general case): \begin{theo} Let $\Omega$ as in \eqref{ipomega}. Then, there exists a positive constant $\lambda_k(\Omega)$ depending only on $n,k$, and $\Omega$, such that problem (\ref{autsk}) admits a solution $u \in C^2(\Omega)\cap C^{1,1}(\overline{\Omega})$, negative in $\Omega$, for $\lambda=\lambda_k(\Omega)$ and $u$ is unique up to positive scalar multiplication. Moreover, $\lambda_k(\Omega)$ has the following variational characterization: \begin{equation*} \lambda_k(\Omega)=\min_{\substack{u \in \Phi_k^2(\Omega) \\ u \ne 0 }} \displaystyle \frac{\int_{\Omega}(-u) S_k(D^2u)\,dx}{\int_{\Omega}(-u)^{k+1}\,dx}. \end{equation*} \end{theo} As matter of fact, if $k<n$ the above theorem holds under a more general assumption on $\Omega$, namely requiring that $\Omega$ is strictly $k$-convex (see \cite{wangeigen}, \cite{geng}). As matter of fact, we observe that if $k=1$, or $k=n$, $\lambda_k(\Omega)$ coincides respectively with the first eigenvalue of the Laplacian operator, or with the eigenvalue of the Monge-Amp\`ere operator. A simple but useful property of the eigenvalue $\lambda_k(\Omega)$ is that it rescales as \begin{equation} \label{risc_palla} \lambda_k(t\Omega)=t^{-2k}\lambda_k(\Omega),\quad t>0. \end{equation} If $k=1$, the well-known Faber-Krahn inequality states that \[ \lambda_1(\Omega) \ge \lambda_1(\Omega^\#), \] where $\Omega^\#$ is the ball centered at the origin with the same Lebesgue measure of $\Omega$. Moreover, the equality holds if $\Omega=\Omega^\#$. In \cite{bt07}, \cite{ga09} it is proved that if $k=n$ and $\Omega$ is a bounded strictly convex open set, then \begin{equation*} \lambda_n(\Omega) \ge \lambda_n(\Omega_{n-1}^*). \end{equation*} In general, in \cite{ga09} it is proven that if $\Omega$ is a strictly convex set such that the eigenfunctions have convex level sets, then, for $2\le k \le n$, \begin{equation} \label{fksk} \lambda_k(\Omega) \ge \lambda_k(\Omega^*_{k-1}). \end{equation} \subsection{Asymmetry measures and isoperimetric deficit} A purpose of this paper is to prove that the difference between the two sides in \eqref{fksk} controls the exterior and interior deficiencies, defined as follows (see also \cite{hana94} for $k=1$). Given $\Omega$ bounded nonempty domain of $\mathbb R^n$, denoted by $R$ the radius of the ball $\Omega^*_{k-1}$, then the exterior and interior $k$-deficiency of $\Omega$ are, respectively, the nonnegative numbers \begin{equation} \label{defe} D_{k}(\Omega)= \frac{R_\Omega}{R}-1,\quad d_k(\Omega)= 1-\frac{r_\Omega}{R}, \end{equation} where $r_\Omega$ and $R_\Omega$ are the inradius and the circumradius of $\Omega$. Such numbers give a measure of the asymmetry of $\Omega$ with respect to the ball with the same $(k-1)$-quermassintegral than $\Omega$. Furthermore, the deficiency of $\Omega$ is \begin{equation*} \Delta(\Omega)=\frac{R_\Omega}{r_\Omega}-1. \end{equation*} In order to have a measure of the asymmetry of $\Omega$ in terms of the Hausdorff distance $d$, we define the following coefficient: \begin{equation} \label{haus} \delta_H(\Omega) = \inf \{d(\Omega, \Omega^*_{n-1}+x_0),\; x_0\in \mathbb R^n\}. \end{equation} We refer to $\delta_H$ as the Hausdorff asymmetry of $\Omega$. In the class of convex sets, it is possible to obtain some stability estimates for the Aleksandrov-Fenchel inequalities \eqref{afineq}. More precisely, if $s$ denotes the Steiner point of $\Omega$, then in \cite{grsc} it has been proved that \begin{equation} \label{gshaus} \begin{array}{l} d(\Omega,\Omega^*_{n-1}+s)^{(n+3)/2} \le \bar C \frac{P(\Omega)^{(n^2-3)/2}}{|\Omega|^{(n+3)(n-2)/2}} \left[ \left(\frac{P(\Omega)}{n\omega_n}\right)^n - \left(\frac{|\Omega|}{\omega_n}\right)^{n-1} \right],\\ d(\Omega,\Omega^*_{n-1}+s)^{(n+3)/2} \le \bar C_1 \frac{W_{n-2}(\Omega)W_{n-1}(\Omega)^{\frac{n-1}{2}}}{W_{k-1}(\Omega)^{n-k}} \left( \frac{W_k(\Omega)^{n-k+1}}{\omega_n} -W_{k-1}(\Omega)^{n-k} \right), \end{array} \end{equation} where $\bar C,\bar C_1$ are two constants depending only on $n$, which can be explicitly determined. These estimates justify the definition of $\delta_H$. As matter of fact, in \cite{grsc} it is observed that, the inequalities \eqref{gshaus} can be rewritten as a Bonnesen-type inequality in terms of the inradius $r_\Omega$ and the circumradius $R_\Omega$ of $\Omega$: \begin{equation} \label{gs} \begin{array}{l} (R_\Omega-r_\Omega)^{(n+3)/2} \le \tilde C \frac{P(\Omega)^{(n^2-3)/2}}{|\Omega|^{(n+3)(n-2)/2}}\left[ \left(\frac{P(\Omega)}{n\omega_n}\right)^n - \left(\frac{|\Omega|}{\omega_n}\right)^{n-1} \right],\\ (R_\Omega-r_\Omega)^{(n+3)/2} \le \tilde C_1 \frac{W_{n-2}(\Omega)W_{n-1}(\Omega)^{\frac{n-1}{2}}}{W_{k-1}(\Omega)^{n-k}} \left( \frac{W_k(\Omega)^{n-k+1}}{\omega_n} - W_{k-1}(\Omega)^{n-k} \right). \end{array} \end{equation} \section{The case of the Monge-Amp\`ere operator} In this section we consider the eigenvalue problem for the Monge-Amp\`ere operator, \begin{equation}\label{eig_eq} \left\{ \begin{array}{ll} -\det D^2 u= \lambda (-u)^{n} & \text{in }\Omega, \\ [.2cm] u = 0 & \text{on }\partial\Omega, \end{array} \right. \end{equation} and we prove the first stability result, stated below. \begin{theo} \label{main} Let $\Omega \subset \mathbb R^n$ be as in \eqref{ipomega} such that \begin{equation} \label{small} \lambda_n(\Omega) \le (1+\varepsilon)\lambda_n(\Omega_{n-1}^*), \end{equation} where $\varepsilon>0$ is sufficiently small and $\Omega_{n-1}^*$ is the ball such that $W_{n-1}(\Omega)=W_{n-1}(\Omega_{n-1}^*)$. Then, if $\delta_H(\Omega)$ is the Hausdorff asymmetry \eqref{haus}, it holds that \begin{equation*} \label{maindelta} \delta_H(\Omega)\le C_n \varepsilon^{\frac{1}{(n+1)(n+3)}}, \end{equation*} where $C_n$ is a constant which depend only on $n$. Moreover, denoting by $d_n(\Omega)$ and $D_n(\Omega)$, respectively, the interior and exterior $n$-deficiency of $\Omega$ as in \eqref{defe}, we have the following: \begin{enumerate} \item if $n=2$, then \begin{equation} \label{defdet2} d_2(\Omega) \le C_2 \sqrt[6]\varepsilon, \quad D_2(\Omega) \le C_2 \sqrt[12]\varepsilon, \end{equation} where $C_2$ denotes a positive constant which depends only on the dimension $n=2$. \item If $n\ge 3$, then \begin{equation} \label{defdetn} d_n(\Omega) \le C_n \varepsilon^{\frac{1}{2n+2}}, \quad D_n(\Omega)\le C_n \varepsilon^{\frac{1}{(n+1)(n+3)}}, \end{equation} where $C_n$ depends only on $n$. \end{enumerate} \end{theo} \begin{rem} \label{remeq} The estimates \eqref{defdet2} and \eqref{defdetn} can be read as \[ \frac{P(\Omega)-P(B_{r_\Omega})}{P(\Omega)} \le C_2 \sqrt[6]\varepsilon, \qquad\quad \frac{P(B_{R_\Omega})-P(\Omega)}{P(\Omega)} \le C_2 \sqrt[12]\varepsilon, \] and \begin{equation*} \frac{W_{n-1}(\Omega) - W_{n-1}(B_{r_\Omega})}{W_{n-1}(\Omega)} \le C_n \varepsilon^{\frac{1}{2n+2}}, \qquad \frac{W_{n-1}(B_{R_\Omega}) - W_{n-1}(\Omega)}{W_{n-1}(\Omega)} \le C_n \varepsilon^{\frac{1}{(n+1)(n+3)}}, \end{equation*} in the spirit of the stability result contained in \cite{melas92}. \end{rem} To prove the Theorem, we need some preliminary lemmas. For $\delta\ge 0$, we denote \[ \Omega_\delta = \{x\in \Omega \colon -u > \delta \}. \] In the following result we estimate $W_{n-1}(\Omega_{\delta})$ in term of $W_{n-1}(\Omega)$. \begin{lemma}\label{lemma1} Under the hypotheses of Theorem \ref{main}, if $u$ is the eigenfunction of \eqref{eig_eq} such that $\|u\|_{L^{n+1}(\Omega)}=1$, then for any $\delta$ such that $0<\delta<\frac 1 2 |\Omega|^{-\frac{1}{n+1}}$, we have \[ W_{n-1}(\Omega_\delta)\ge W_{n-1}(\Omega) (1-\max\{\varepsilon, 2\delta |\Omega|^{\frac{1}{n+1}}\}). \] \begin{proof} For $\delta>0$, we compute the Rayleigh quotient of the function $\phi=u+\delta$ in $\Omega_\delta$. Then, \begin{equation} \label{udelta} \lambda_n(\Omega_\delta) \le \dfrac{\int_{\Omega_\delta} (-\phi)\det D^2 \phi\,dx }{ \int_{\Omega_\delta} (-\phi)^{n+1}\,dx} = \dfrac{\int_{\Omega_\delta} (-u-\delta)\det D^2 u \,dx }{ \int_{\Omega_\delta} (-u-\delta)^{n+1}\,dx}. \end{equation} Moreover, being $u$ a solution of \eqref{eig_eq} with $\lambda=\lambda_n(\Omega)$, we get, by H\"older inequality, and recalling that $\int_\Omega (-u)^{n+1} dx =1$, that \[ \begin{split} \int_{\Omega_\delta}(-u-\delta)\det D^2 u \,dx &= \lambda_n(\Omega) \int_{\Omega_\delta} (-u-\delta)(-u)^n\,dx \\ &\le \lambda_n(\Omega) \left(\int_{\Omega_\delta} (-u-\delta)^{n+1}\,dx\right)^{\frac{1}{n+1}} \left(\int_{\Omega_\delta} (-u)^{n+1}\,dx\right)^{\frac{n}{n+1}} \\ &\le \lambda_n(\Omega) \left(\int_{\Omega_\delta} (-u-\delta)^{n+1}\,dx\right)^{\frac{1}{n+1}}. \end{split} \] Hence, combining the above estimate with \eqref{udelta} it follows that \begin{equation} \label{udelta2} \lambda_n(\Omega_\delta) \le \lambda_n(\Omega) \left(\int_{\Omega_\delta} (-u-\delta)^{n+1}\,dx \right)^{-\frac{n}{n+1}} \end{equation} On the other hand, by Minkowski inequality and choosing $\delta<\frac 1 2 |\Omega|^{-\frac{1}{n+1}}$, we obtain that \[ \begin{split} \left(\int_{\Omega_\delta} (-u-\delta)^{n+1}\,dx\right)^{\frac{1}{n+1}} &\ge \left( \int_{\Omega_\delta} (-u)^{n+1} \,dx \right)^{\frac {1}{n+1}} - \left( \int_{\Omega_\delta} \delta^{n+1} \,dx \right)^{\frac {1}{n+1}} \\ &\ge \left(1- \int_{\Omega\setminus\Omega_\delta} \delta^{n+1} \,dx \right)^{\frac{1}{n+1}}- \delta |\Omega_\delta|^{\frac{1}{n+1}} \\ &= \left(1- \delta^{n+1} \big(|\Omega|-|\Omega_\delta|\big) \right)^{\frac{1}{n+1}}- \delta |\Omega_\delta|^{\frac{1}{n+1}} \\ &\ge 1-\delta\big( |\Omega| - |\Omega_\delta| \big)^{\frac{1}{n+1}} - \delta |\Omega_\delta|^{\frac{1}{n+1}} \ge 1-2\delta |\Omega|^{\frac {1}{n+1}}. \end{split} \] Hence, from \eqref{udelta2}, \eqref{small} and Faber-Krahn inequality it follows that \[ \lambda_n((\Omega_\delta)^*_{n-1}) \le \lambda_n(\Omega_\delta) \le (1+\varepsilon) \lambda_n(\Omega^*_{n-1}) \big( 1-2\delta |\Omega|^{\frac {1}{n+1}} \big)^{-n} \] which implies, by \ref{risc_palla}, that \begin{equation} \label{udelta3} \left(\frac{W_{n-1}(\Omega_\delta)}{W_{n-1}(\Omega)}\right)^{2n} = \frac{\lambda_n(\Omega^*_{n-1})}{\lambda_n((\Omega_\delta)^*_{n-1})} \ge \frac{\big( 1-2\delta |\Omega|^{\frac {1}{n+1}} \big)^n}{1+\varepsilon}, \end{equation} where we used that the balls $\Omega^*_{n-1}$ and $(\Omega_\delta)^*_{n-1}$ preserve, respectively, the $(n-1)$-th quermassintegral of $\Omega$ and $\Omega_\delta$. Hence, by \eqref{udelta3} we get that \[ \frac{W_{n-1}(\Omega_\delta)}{W_{n-1}(\Omega)} \ge \left( 1 - \frac{\varepsilon+ 2 \delta |\Omega|^{\frac {1}{n+1}}}{1+\varepsilon} \right)^{1/2} \ge 1-\max\{\varepsilon, 2\delta |\Omega|^{\frac{1}{n+1}}\}, \] obtaining the thesis. \end{proof} \end{lemma} The second lemma we need is the following. \begin{lemma} Under the hypotheses of Theorem \ref{main}, if $\Omega_t=\{-u>t\}$, and $u$ is the eigenfunction of \eqref{eig_eq} in $\Omega$ such that $\|u\|_{L^{n+1}(\Omega)}=1$, then \begin{equation} \label{eq:3} \int_0^{+\infty} t^n \big( W_{n-1}(\Omega_t)^n - \omega_n^{n-1}|\Omega_t|\big) dt \le \frac{\omega_n^{n-1}}{n+1} \varepsilon. \end{equation} \end{lemma} \begin{proof} We consider the difference of the eigenvalues related to the sets $\Omega$ and $\Omega^*_{n-1}$. Choosing $u$ as the normalized eigenfunction of \eqref{eig_eq} in $\Omega$, using the P\'olya-Szeg\"o principle \eqref{pol} \begin{equation} \label{eq:2} \begin{split} \lambda_n(\Omega) - \lambda_n(\Omega^*_{n-1}) & \ge \int_\Omega (-u)\det D^2 u\,dx - \frac{\int_{\Omega^*_{n-1}}(-u_{n-1}^* )\det D^2 u_{n-1}^*\,dx }{\int_{\Omega^*_{n-1}}(-u_{n-1}^* )^{n+1} \,dx} \\[.2cm] & \ge \frac{\int_{\Omega^*_{n-1}}(-u_{n-1}^* )\det D^2 u_{n-1}^*\,dx }{\int_{\Omega^*_{n-1}}(-u_{n-1}^* )^{n+1} \,dx} \left(\int_{\Omega^*_{n-1}}(-u_{n-1}^* )^{n+1} \,dx - 1 \right) \\[.2cm] &\ge \lambda_n(\Omega^*_{n-1})\left(\int_{\Omega^*_{n-1}}(-u_{n-1}^* )^{n+1} \,dx -1 \right). \end{split} \end{equation} On the other hand, recalling that $u$ has normalized $L^{n+1}$ norm, the coarea formula and an integration by parts give that \[ \begin{split} \int_{\Omega^*_{n-1}}(-u_{n-1}^*)^{n+1} \,dx -1 &= (n+1)\int_0^{+\infty} t^n \big( |\{-u^*_{n-1}>t\}|-|\{-u>t\}|\big) dt \\ &= (n+1) \omega_n^{1-n} \int_0^{+\infty} t^n \big( W_{n-1}(\Omega_t)^n - \omega_n^{n-1}|\Omega_t|\big). \end{split} \] Hence, joining \eqref{eq:2} with the above equality, and using \eqref{small} we obtain that \[ \int_0^{+\infty} t^n \big( W_{n-1}(\Omega_t)^n - \omega_n^{n-1}|\Omega_t|\big) dt \le \frac{\omega_n^{n-1}}{n+1} \varepsilon, \] that is the thesis. \end{proof} Last lemma plays a key role in order to obtain that the constant $C_n$ involved in (1) and (2) in Theorem \ref{main} is independent on $\Omega$. \begin{lemma} \label{boundmis} Under the hypotheses of Theorem \ref{main}, it holds that \begin{equation*} |\Omega| \ge \tilde C_n \left[W_{n-1}(\Omega)\right]^n, \end{equation*} where $\tilde C_n$ denotes a positive constant depending only on $n$. \end{lemma} \begin{proof} Let $u$ be an eigenfunction of \eqref{eig_eq} corresponding to the eigenvalue $\lambda=\lambda_n(\Omega)$. Then \begin{equation} \label{eq} \det D^2u = \lambda (-u)^n \quad \text{ in }\Omega. \end{equation} Integrating both sides in \eqref{eq} on the level set $\Omega_t=\{-u>t\}$, and denoting by $\Sigma_t=\partial \Omega_t=\{-u=t\}$, and using the H\"older inequality we have \begin{multline} \label{sinistra} \int_{\Omega_t}\det D^2u\,dx = \frac{1}{n} \int_{\Sigma_t} H_{n-1} |Du|^n d\mathcal H^{n-1} \ge \\ \ge \frac{1}{n} \frac{\left(\int_{\Sigma_t} H_{n-1} d\mathcal H^{n-1}\right)^{n+1}}{ \left(\int_{\Sigma_t} H_{n-1} |Du|^{-1} d\mathcal H^{n-1}\right)^{n}} = \frac{1}{n} \frac{(n\omega_n)^{n+1}}{\left(-\frac{d}{dt}W_{n-1}(\Omega_t)\right)^n}. \end{multline} Last inequality follows by the H\"older inequality and the properties of quermassintegrals. Moreover, being $|\Omega_t|\le |\Omega|$, we have \begin{equation} \label{destra} \left(\int_{\Omega_t}(-u)^n\,dx \right)^\frac{1}{n} \le |\Omega|^{\frac 1 n} \|u\|_{L^{\infty}(\Omega)}. \end{equation} Putting togheter \eqref{sinistra} and \eqref{destra}, by \eqref{eq} we get \[ -\frac{d}{dt}W_{n-1}(\Omega_t) \ge n \omega_n^{1+\frac{1}{n}} \lambda^{-\frac{1}{n}}|\Omega|^{-\frac 1 n} \|u\|_{L^\infty(\Omega)}^{-1} \] and, integrating between $0$ and $\|u\|_{L^{\infty}(\Omega)}$, \[ W_{n-1}(\Omega) \ge n \omega_n^{1+\frac{1}{n}} \lambda^{-\frac{1}{n}}|\Omega|^{-\frac{1}{n}}, \] that is, being $\lambda=\lambda_n(\Omega)\le (1+\varepsilon)\lambda_n(\Omega^*_{n-1})$, \begin{equation} \label{misbas} |\Omega|^{\frac 1 n} \ge n \omega_n^{1+\frac 1 n} W_{n-1}(\Omega)^{-1} \lambda_{n}(\Omega^*_{n-1})^{-\frac{1}{n}} (1+\varepsilon)^{-\frac 1 n}. \end{equation} As matter of fact, being $W_{n-1}(\Omega)=W_{n-1}(\Omega^*_{n-1})$, properties \eqref{risc_palla} and \eqref{querball} give that \[ \lambda_n(\Omega^*_{n-1})= \left( \frac{W_{n-1}(\Omega)}{\omega_n}\right)^{-2n} \lambda_n(B), \] where $B=\{|x|<1\}$. Then \eqref{misbas} becomes \[ |\Omega|^{\frac 1 n} \ge n \omega_n^{\frac 1 n - 1} W_{n-1}(\Omega) \lambda_n(B)^{-\frac 1 n} (1+\varepsilon)^{-\frac 1 n}, \] and this concludes the proof. \end{proof} Now we are in position to prove the main theorem of this section. \begin{proof}[Proof of the Theorem \ref{main}] First of all, we observe that the quotient \[ \frac{W_{n-1}(K)-W_{n-1}(L)}{W_{n-1}(K)} \] is rescaling invariant, hence we suppose that $W_{n-1}(\Omega)=1$. Consequently, by Lemma \ref{boundmis} and the Aleksandrov-Fenchel inequality, we have that there exists two positive constants $c_1(n)$ and $c_2(n)$, which depend only on the dimension, such that \begin{equation} \label{bound} c_1(n) \le |\Omega| \le c_2(n). \end{equation} For $\delta$ as in Lemma \ref{lemma1}, by \eqref{eq:3} we get that \[ \begin{split} \inf_{0\le t \le \delta}\big\{ W_{n-1}(\Omega_t)^n - \omega_n^{n-1}|\Omega_t|\big\} &\le \frac {n+1} {\delta^{n+1}} \int_0^{\delta} t^n \big( W_{n-1}(\Omega_t)^n - \omega_n^{n-1}|\Omega_t|\big) dt \\ &\le \omega_{n}^{n-1} \frac {\varepsilon}{\delta^{n+1}} = \omega_{n}^{n-1} \sqrt\varepsilon, \end{split} \] where we finally choose $\delta^{n+1}=\sqrt{\varepsilon}$. Hence, this gives that there exists $0\le \tau \le \delta$ such that \begin{equation} \label{eq:4} W_{n-1}(\Omega_\tau)^n \le \omega_n^{n-1}|\Omega_\tau| + \omega_{n}^{n-1} \sqrt\varepsilon. \end{equation} \fbox{\bf Case $n=2$.} In such a case, \eqref{eq:4} becomes \begin{equation*} \label{eq:7} \frac{P(\Omega_\tau)^2}{4\pi}- |\Omega_\tau| \le \sqrt \varepsilon. \end{equation*} Then, denoting by $r_\tau$ and $R_\tau$ the inradius and the circumradius of $\Omega_\tau$ respectively, and by $\rho_\tau$ the radius of $(\Omega_\tau)^*_1$, using the Bonnesen inequality (see for example \cite{oss79}, and \cite{afn09,afn11} for some related questions) we have \[ (\rho_\tau-r_\tau)^2 \le (R_\tau-r_\tau)^2 \le \sqrt\varepsilon. \] Being $2\pi \rho_\tau = P(\Omega_\tau) $, we have by Lemma \ref{lemma1}, for $\varepsilon$ sufficiently small, that \[ r_\tau\ge \frac{ P(\Omega_\tau) }{2\pi} -\sqrt[4]{\varepsilon}\ge \frac{P(\Omega)}{2\pi}\left( 1- 2\sqrt[6]{\varepsilon}|\Omega|^{\frac 1 3} \right)-\sqrt[4]{\varepsilon}\ge R(1-C_{2}|\Omega|^{\frac 1 3} \sqrt[6]{\varepsilon}), \] where $R=\frac{P(\Omega)}{2\pi}$ is the radius of $\Omega^*_1$ and $C_{2}$ denotes a constant which depends only on the dimension $n=2$. Being $r_\tau < r_\Omega$, by \eqref{bound} we have that \begin{equation} \label{eq:6} d_2(\Omega) \le 1- \frac{r_\tau}{R} \le C_{2}|\Omega|^{\frac 1 3} \sqrt[6] \varepsilon\le C_{2} \sqrt[6]{\varepsilon} \end{equation} where $B_{r_\tau}$ is a ball of radius $r_\tau$ contained in $\Omega$. Then, by \eqref{eq:6} and being $P(\Omega)=2$, we have that \begin{equation} \label{eq:8} \left( \frac{P^2(\Omega)}{4\pi} -|\Omega| \right)^{\frac 1 2}\le \left( \frac{\big(P(B_{r_\Omega})+C_2\sqrt[6]{\varepsilon}\big)^2}{4\pi} - |B_{r_\Omega}| \right)^{\frac 1 2} \le C_2 \sqrt[12]\varepsilon, \end{equation} where last inequality follows being $P(B_{r_\Omega})\le P(\Omega)=2$. Then, \eqref{eq:8}, \eqref{bound} and \eqref{gshaus} give \[ \delta_H(\Omega) \le C_2 \sqrt[15] \varepsilon. \] On the other hand, applying to \eqref{eq:8} the well-known Bonnesen inequality, we get that \[ D_2(\Omega) \le 2\pi(R_\Omega-r_\Omega)\le C_2 \sqrt[12] \varepsilon. \] \fbox{\bf Case $n> 2$.} From \eqref{eq:4} and the Aleksandrov-Fenchel inequalities \eqref{af-2} with $k=n$, we have that \[ \frac{P(\Omega_\tau)^{\frac{n}{n-1}}}{(n^n\omega_n)^{\frac{1}{n-1}}}\le |\Omega_\tau|+\sqrt \varepsilon. \] Hence, by \eqref{bound} and for $\varepsilon$ sufficiently small, an elementary inequality gives that \[ \frac{P(\Omega_\tau)^{{n}}}{n^n\omega_n}\le |\Omega_\tau|^{n-1}+C_n\sqrt \varepsilon. \] Then, applying the stability result \eqref{gs}, and using again \eqref{bound}, it follows that \begin{equation} \label{gs2} (R_\tau-r_\tau)^{(n+3)/2} \le C_n \frac{P(\Omega)^{(n^2-3)/2}}{|\Omega|^{(n+3)(n-2)/2}} \sqrt\varepsilon\le C_n\sqrt \varepsilon, \end{equation} where, as before, $R_\tau$ and $r_\tau$ are, respectively, the circumradius and the inradius of $\Omega_\tau$. Now, let $\rho_\tau$ be the radius of the ball $(\Omega_\tau)^*_{n-1}$, having the same $W_{n-1}$ measure of $\Omega_\tau$. Similarly as before, being $\rho_\tau <R_\tau$, by \eqref{gs2}, Lemma \ref{lemma1} and \eqref{bound}, for $\varepsilon$ sufficiently small we have \begin{multline} \label{catenella} r_\tau \ge \rho_\tau - C_n \varepsilon^{\frac{1}{n+3}} = \frac{W_{n-1}(\Omega_\tau)}{\omega_n} - C_n \varepsilon^{\frac{1}{n+3}} \ge \\ \ge \omega_n^{-1}W_{n-1}(\Omega)\left(1-2\varepsilon^{1/(2n+2)}|\Omega|^{\frac{1}{n+1}} \right) - C_n \varepsilon^{\frac{1}{n+3}} \ge \\ \ge R\left(1- C_n \varepsilon^{1/(2n+2)} \right), \end{multline} where $R=\omega_n^{-1}W_{n-1}(\Omega)$ is the radius of the ball $\Omega^*_{n-1}$. Denoting again with $r_\Omega$ the inradius of $\Omega$, we have that $r_\tau \le r_\Omega$ and \begin{equation*} d_n(\Omega) \le 1- \frac{r_\tau}{R} \le C_n \varepsilon^{1/(2n+2)}. \end{equation*} As matter of fact, by the Aleksandrov-Fenchel inequalities, \eqref{catenella} and being $W_{n-1}(\Omega)=1$, it follows that \begin{equation}\label{catena} \begin{split} \left[\left( \frac{P(\Omega)}{n\omega_n} \right)^n - \left( \frac{ |\Omega| }{\omega_n}\right)^{n-1}\right]^{\frac{2}{n+3}} &\le \left[ \left( \frac{W_{n-1}(\Omega)^n}{\omega_n^n} \right)^{n-1} - \left( \frac{ |\Omega| }{\omega_n}\right)^{n-1}\right]^{\frac{2}{n+3}} \le \\ & \le \left[ \left( \frac{W_{n-1}(B_{r_\Omega})+C_n \varepsilon^{\frac{1}{2n+2}}}{\omega_n} \right)^{n(n-1)}- \left( \frac{ |B_{r_\Omega}| }{\omega_n}\right)^{n-1}\right]^{\frac{2}{n+3}} \le \\ & \le \left[ \left( \frac{W_{n-1}(B_{r_\Omega})}{\omega_n} \right)^{n(n-1)} + C_n \varepsilon^{\frac{1}{2n+2}} - \left( \frac{ |B_{r_\Omega}| }{\omega_n}\right)^{n-1}\right]^{\frac{2}{n+3}} = \\ &= C_n \varepsilon^{\frac{1}{(n+1)(n+3)}}. \end{split} \end{equation} Hence, applying \eqref{gshaus}, from \eqref{catena}, \eqref{bound} and being $W_{n-1}(\Omega)=\omega_n R=1$, we get that \begin{equation*} \delta_H(\Omega)\le C_n \varepsilon^{\frac{1}{(n+1)(n+3)}}, \end{equation*} while applying \eqref{gs}, we get \begin{equation*} D_n(\Omega) \le \omega_n(R_\Omega-r_\Omega) \le C_n \varepsilon^{\frac{1}{(n+1)(n+3)}}, \end{equation*} and this concludes the proof. \end{proof} \begin{rem} \label{remdef} Under the assumption of Theorem \ref{main}, from \eqref{catena} and \eqref{catenella} an estimate for the deficiency of $\Omega$ holds, that is \[ \Delta(\Omega) \le C_{n} \varepsilon^{\frac{1}{(n+1)(n+3)}}. \] \end{rem} \section{The case of the $k$-Hessian operator, $1 \le k\le n-1$} In this section we consider the eigenvalue problem related to the $k$-Hessian operators, $1 \le k\le n-1$, namely \begin{equation*} \left\{ \begin{array}{ll} \Sk_k(D^2u)=\lambda (-u)^k &\text{in } \Omega,\\ u=0 &\text{on } \partial\Omega, \end{array} \right. \end{equation*} obtaining the stability result as follows. \begin{theo} \label{main2} Let $1\le k \le n-1$, and $\Omega \subset \mathbb R^n$ be as in \eqref{ipomega} such that \begin{equation} \label{small2} \lambda_k(\Omega) \le (1+\varepsilon)\lambda_k(\Omega_{k-1}^*), \end{equation} where $\varepsilon>0$ is sufficiently small and $\Omega_{k-1}^*$ is the ball such that $W_{k-1}(\Omega)=W_{k-1}(\Omega_{k-1}^*)$. Moreover we suppose that the eigenfunctions related to $\lambda_k(\Omega)$ have convex level sets. Then, \[ \delta_H(\Omega) \le C_{n,k}\varepsilon^{\frac{2\alpha}{n+3}}, \] and \begin{equation} \label{tesik} d_k(\Omega)\le C_{n,k} \varepsilon^{\alpha},\quad D_k(\Omega) \le C_{n,k} \varepsilon^{\frac{2\alpha}{n+3}}, \end{equation} where $\alpha=\max\left\{\frac{1}{k+1}, \frac {2k}{(k+1)(n+3)}\right\}$, $C_{n,k}$ is a positive constant which depends only on $n$ and $k$, and $d_k(\Omega)$ and $D_k(\Omega)$ are, respectively, the interior and exterior $k$-deficiency of $\Omega$ as in \eqref{defe}. \end{theo} \begin{rem} As observed in Section 2.3, the additional hypothesis on the convexity of the level sets of the eigenfunctions corresponding to $\lambda_k(\Omega)$ is necessary to have that a Faber-Krahn inequality holds. On the other hand this assumption seems to be natural. Indeed, for $k=1$ this is due to the Korevaar concavity maximum principle (see\cite{kor}), while it is trivial for $k=n$. For the $k$-Hessian operators, at least in the case $n=3$ and $k=2$, it in \cite{lmx10} and \cite{sa12} is proved that if $\Omega$ is sufficiently smooth, the eigenfunctions of $\Sk_2$ have convex level sets. Up to our knowledge, the general case is an open problem. \end{rem} \begin{rem} Similarly as observed in Remark \ref{remeq}, by the estimates \eqref{tesik} we can obtain that \begin{equation*} \frac{W_{k-1}(\Omega) - W_{k-1}(B_{r_\Omega})}{W_{k-1}(\Omega)} \le C_{n,k} \varepsilon^{\alpha},\quad \frac{W_{k-1}(B_{R_\Omega}) - W_{k-1}(\Omega)}{W_{k-1}(\Omega)} \le C_{n,k} \varepsilon^{\frac{2\alpha}{n+3}}. \end{equation*} \end{rem} Similarly to the case of the Monge-Amp\`ere operator, to give the proof of Theorem \ref{main2} we first consider some preliminary results. Using the same notations of section 3, for $\delta\ge 0$, we denote \[ \Omega_\delta = \{x\in \Omega \colon -u > \delta \}. \] \begin{lemma} \label{lemma1kkk} Under the hypotheses of Theorem \ref{main2}, if $u$ is the eigenfunction of $\Sk_k$ in $\Omega$ such that $\|u\|_{L^{k+1}(\Omega)}=1$, then for any $\delta$ such that $0<\delta<\frac 1 2 |\Omega|^{-\frac{1}{k+1}}$, we have \begin{equation} \label{lemma1k} W_{k-1}(\Omega_\delta) \ge W_{k-1}(\Omega) \left[1-(n-k+1)\max\{\varepsilon, 2\delta |\Omega|^{\frac{1}{k+1}}\}\right]. \end{equation} \end{lemma} \begin{proof} For $\delta>0$, we compute the Rayleigh quotient of the function $\phi=u+\delta$ in $\Omega_\delta$. Then, \begin{equation} \label{udeltak} \lambda_k(\Omega_\delta) \le \dfrac{\int_{\Omega_\delta} (-\phi)\Sk_k (D^2 \phi)\,dx }{ \int_{\Omega_\delta} (-\phi)^{k+1}\,dx} = \dfrac{\int_{\Omega_\delta} (-u-\delta)\Sk_k (D^2 u) \,dx }{ \int_{\Omega_\delta} (-u-\delta)^{k+1}\,dx}. \end{equation} \[ \begin{split} \int_{\Omega_\delta}(-u-\delta)\Sk_k (D^2u) \,dx &= \lambda_k(\Omega) \int_{\Omega_\delta} (-u-\delta)(-u)^k\,dx \\ &\le \lambda_k(\Omega) \left(\int_{\Omega_\delta} (-u-\delta)^{k+1}\,dx\right)^{\frac{1}{k+1}} \left(\int_{\Omega_\delta} (-u)^{k+1}\,dx\right)^{\frac{k}{k+1}} \\ &\le \lambda_k(\Omega) \left(\int_{\Omega_\delta} (-u-\delta)^{k+1}\,dx\right)^{\frac{1}{k+1}}. \end{split} \] Hence, combining the above estimate with \eqref{udeltak} it follows that \begin{equation} \label{udelta2k} \lambda_k(\Omega_\delta) \le \lambda_k(\Omega) \left(\int_{\Omega_\delta} (-u-\delta)^{k+1}\,dx \right)^{-\frac{k}{k+1}} \end{equation} On the other hand, by Minkowski inequality and choosing $\delta<\frac 1 2 |\Omega|^{-\frac{1}{k+1}}$, we obtain that \[ \begin{split} \left(\int_{\Omega_\delta} (-u-\delta)^{k+1}\,dx\right)^{\frac{1}{k+1}} &\ge \left( \int_{\Omega_\delta} (-u)^{k+1} \,dx \right)^{\frac {1}{k+1}} - \left( \int_{\Omega_\delta} \delta^{k+1} \,dx \right)^{\frac {1}{k+1}} \\ &\ge \left(1- \int_{\Omega\setminus\Omega_\delta} \delta^{k+1} \,dx \right)^{\frac{1}{k+1}}- \delta |\Omega_\delta|^{\frac{1}{k+1}} \\ &= \left(1- \delta^{k+1} \big(|\Omega|-|\Omega_\delta|\big) \right)^{\frac{1}{k+1}}- \delta |\Omega_\delta|^{\frac{1}{k+1}} \\ &\ge 1-\delta\big( |\Omega| - |\Omega_\delta| \big)^{\frac{1}{k+1}} - \delta |\Omega_\delta|^{\frac{1}{k+1}} \ge 1-2\delta |\Omega|^{\frac {1}{k+1}}. \end{split} \] Hence, from \eqref{udelta2k}, \eqref{small2} and Faber-Krahn inequality it follows that \[ \lambda_k((\Omega_\delta)^*_{k-1}) \le \lambda_k(\Omega_\delta) \le (1+\varepsilon) \lambda_k(\Omega^*_{k-1}) \big( 1-2\delta |\Omega|^{\frac {1}{k+1}} \big)^{-k} \] which implies, by \eqref{risc_palla}, that \begin{equation} \label{udelta3k} \left(\frac{W_{k-1}(\Omega_\delta)}{W_{k-1}(\Omega)} \right)^{\frac{2k}{n-k+1}} = \frac{\lambda_k(\Omega^*_{k-1})}{\lambda_k((\Omega_\delta)^*_{k-1})} \ge \frac{\big( 1-2\delta |\Omega|^{\frac{1}{k+1}} \big)^k}{1+\varepsilon}, \end{equation} where we used that the balls $\Omega^*_{k-1}$ and $(\Omega_\delta)^*_{k-1}$ preserve, respectively, the $(k-1)$-th quermassintegral of $\Omega$ and $\Omega_\delta$. Hence, by \eqref{udelta3k} we get that \[ \frac{W_{k-1}(\Omega_\delta)}{W_{k-1}(\Omega)} \ge \left( 1 - \frac{\varepsilon+ 2 \delta |\Omega|^{\frac {1}{k+1}}}{1+\varepsilon} \right)^{\frac{n-k+1}{2}} \ge 1-(n-k+1)\max\{\varepsilon, 2\delta |\Omega|^{\frac{1}{k+1}}\}, \] obtaining the thesis. \end{proof} \begin{lemma} Under the hypotheses of Theorem \ref{main2}, if $u$ is the eigenfunction of $\Sk_k$ in $\Omega$ such that $\|u\|_{L^{k+1}(\Omega)}=1$, we have that \begin{equation} \label{eq:10} \frac {n(n-k+1)^{k}}{k} \int_0^{\max|u|} \frac{W_k(\Omega_t)^{k+1}-W_k((\Omega_t)^*_{k-1})^{k+1}}{\big[-\frac {d}{dt} W_{k-1}(\Omega_t)\big]^k} dt \le \varepsilon \lambda_k(\Omega_{k-1}^*). \end{equation} \end{lemma} \begin{proof} The divergence form of $\Sk_k$ and the coarea formula give that (see also \cite{trudiso}) \[ \lambda_k(\Omega) = \frac 1 k \int_0^{\max(-u)} dt \int_{\Sigma_t} H_{k-1}(\Sigma_t) |Du|^k d\mathcal H^{n-1}, \] where $\Sigma_t=\partial \Omega_t$. Then, the H\"older inequality and the Reilly formula \eqref{reillyder} give that \begin{multline} \label{lemmak} \lambda_k(\Omega) \ge \frac 1 k \int_0^{\max(-u)} \dfrac{ \left( \int_{\Sigma_t} H_{k-1}(\Sigma_t) d\mathcal H^{n-1}\right)^{k+1}} {\left( \int_{\Sigma_t} \frac{H_{k-1}(\Sigma_t)}{|Du|}d\mathcal H^{n-1} \right)^{k} } dt = \\ = \frac {n(n-k+1)^{k}}{k} \int_0^{\max(-u)} \frac{ W_k(\Omega_t)^{k+1} } {\big[-\frac{d}{dt} W_{k-1}(\Omega_t)\big]^{k} } dt. \end{multline} Moreover, being $\|u^*_{k-1}\|_{k+1} \ge \|u\|_{k+1}=1$, we have \begin{multline} \label{lemmakball} \lambda_k(\Omega^*_{k-1}) \le \frac{ \int_{\Omega^*_{k-1}} (-u^*_{k-1})\Sk_k(D^2 u^*_{k-1}) dx } {\int_{\Omega^*_{k-1}} (-u^*_{k-1})^{k+1} dx } \le \int_{\Omega^*_{k-1}} (-u^*_{k-1})\Sk_k(D^2 u^*_{k-1}) dx = \\ = \frac {n(n-k+1)^{k}}{k} \int_0^{\max(-u)} \frac{ W_k((\Omega_t)^*_{k-1})^{k+1}} {\big[-\frac{d}{dt} W_{k-1}(\Omega_t)\big]^{k} } dt. \end{multline} Last equality follows from the symmetry of $u^*_{k-1}$ and being $W_{k-1}(\Omega_t)= W_{k-1}\big((\Omega_t)^*_{k-1}\big)$. Hence, taking \eqref{lemmak} and \eqref{lemmakball} and subtracting, from we have that \begin{multline*} \varepsilon \lambda_k(\Omega^*_{k-1}) \ge \\ \ge \lambda_k(\Omega) -\lambda_k(\Omega^*_{k-1}) \ge \frac {n(n-k+1)^{k}}{k} \int_0^{\max(-u)} \frac{ W_k(\Omega_t)^{k+1}- W_k((\Omega_t)^*_{k-1})^{k+1}} {\big[-\frac{d}{dt} W_{k-1}(\Omega_t)\big]^{k} } dt, \end{multline*} that gives the thesis. \end{proof} In the next result we prove a lower bound for $|\Omega|$ in term of $W_{k-1}(\Omega)$. \begin{lemma} \label{boundmisk} Under the hypotheses of Theorem \ref{main2}, it holds that \begin{equation*} |\Omega| \ge C_{n,k} W_{k-1}^{\frac{n}{n-k+1}}(\Omega), \end{equation*} where $C_{n,k}$ denotes a positive constant depending only on $n$ and $k$. \end{lemma} \begin{proof} Let be $u$ an eigenfunction corresponding to the eigenvalue $\lambda=\lambda_k(\Omega)$ and such that $\|u\|_{L^{k+1}}=1$. Then, \begin{equation} \label{eqk} \Sk_k (D^2u) = \lambda (-u)^k \quad\text{ in }\Omega. \end{equation} Arguing as in Lemma \ref{boundmis}, by \eqref{introintcurv} \eqref{reillyder} and the H\"older inequality we have \begin{equation} \label{sinistrak} \int_{\Omega_t}\Sk_k (D^2u)dx = \frac{1}{k} \int_{\Sigma_t} H_{k-1} |Du|^k d\mathcal H^{n-1} \ge C_{n,k} \frac{\left(W_k(\Omega_t)\right)^{k+1}} {\left(-\frac{d}{dt}W_{k-1}(\Omega_t)\right)^k}. \end{equation} We divide the proof in three cases. Case $k> \frac n 2$. By H\"older inequality we have: \begin{equation} \label{destrak} \left(\int_{\Omega_t}(-u)^kdx\right)^{\frac 1 k} \le |\Omega|^{\frac 1 k} \|u\|_{L^{\infty}(\Omega)}. \end{equation} Putting togheter \eqref{sinistrak} and \eqref{destrak}, by \eqref{eqk} we get that \[ W_k(\Omega_t)^{-\frac{k+1}{k}} \left( -\frac{d}{dt}W_{k-1}(\Omega_t) \right) \ge C_{n,k} |\Omega|^{-\frac 1 k} \|u\|_{\infty}^{-1}\lambda^{-\frac 1 k}. \] Using the Aleksandrov-Fenchel inequalities \eqref{afineq} with $j=k$ and $i=k-1$, and integrating between $0$ and $\|u\|_{L^{\infty}(\Omega)}$, being $k>\frac n 2$ we get \[ W_{k-1}(\Omega)^{\frac{2k-n}{k(n-k+1)}}\ge C_{n,k} \lambda^{-\frac{1}{k}}|\Omega|^{-\frac{1}{k}}. \] Being $\lambda_k(\Omega) \le (1+\varepsilon)\lambda_k(\Omega^*_{k-1})$, and recalling the properties \eqref{risc_palla} and \eqref{querball}, we have: \begin{gather*} \begin{split} |\Omega|^{\frac{1}{k}} & \ge C_{n,k} W_{k-1}(\Omega)^{-\frac{2k-n}{k(n-k+1)}} {\lambda_{k}(\Omega^*_{k-1})^{-\frac{1}{k}}} {(1+\varepsilon)^{-\frac{1}{k}}} = \\ &= C_{n,k} W_{k-1}(\Omega)^{-\frac{2k-n}{k(n-k+1)}} W_{k-1}(\Omega^*_{k-1})^{\frac {2}{n-k+1}} {\lambda_{k}(B_1)^{-\frac{1}{k}}} {(1+\varepsilon)^{-\frac{1}{k}}} = \\ &= C_{n,k} {(1+\varepsilon)^{-\frac{1}{k}}}W_{k-1}(\Omega)^{\frac{n}{k(n-k+1)}}, \end{split} \end{gather*} and the first case is completed. Case $k< \frac n 2$. By H\"older inequality, and being $\|u\|_{k+1}=1$ we have: \begin{equation} \label{destrak2} \left(\int_{\Omega_t}(-u)^kdx\right)^{\frac 1 k} \le |\Omega_t|^{\frac {1}{k(k+1)}} \left( \int_{\Omega_t} u^{k+1}dx \right)^{\frac {1}{k+1}}\le |\Omega|^{\frac {1}{k(k+1)}}. \end{equation} Then, joining \eqref{sinistrak} and \eqref{destrak2}, and using the Aleksandrov-Fenchel inequalities we get \[ W_{k-1}(\Omega_t)^{-\frac{(k+1)(n-k)}{k(n-k+1)}} \left( -\frac{d}{dt}W_{k-1}(\Omega_t) \right) \ge C_{n,k} |\Omega|^{-\frac {1} {k(k+1)}}\lambda^{-\frac 1 k}. \] Integrating between $0$ and $\delta$ sufficiently small, we get that \[ W_{k-1}(\Omega_\delta)^{-\frac{n-2k}{k(n-k+1)}}- W_{k-1}(\Omega)^{-\frac{n-2k}{k(n-k+1)}} \ge C_{n,k} |\Omega|^{-\frac {1} {k(k+1)}}\lambda^{-\frac 1 k} \delta. \] Now we apply Lemma \ref{lemma1kkk}. Let $\varepsilon$ and $\delta$ such that $\varepsilon<2\delta|\Omega|^{\frac{1}{k+1}}<(n-k+1)^{-1}$. Hence, writing $\alpha=-\frac{n-2k}{k(n-k+1)}<0$, we get \begin{equation} \label{boh1} W_{k-1}(\Omega)^{\alpha}\left[ (1-2\delta (n-k+1) |\Omega|^{\frac{1}{k+1}} )^\alpha-1 \right] \ge C_{n,k} |\Omega|^{-\frac{1} {k(k+1)}}\lambda^{-\frac 1 k}\delta. \end{equation} Moreover, if $\delta$ is such that $2\delta |\Omega|^{\frac{1}{k+1}}(n-k+1)\le 1-{2^{-\frac{1}{1-\alpha}}}$, from \eqref{boh1} we get \[ W_{k-1}(\Omega)^{\alpha}\left[ -4\alpha (n-k+1) |\Omega|^{\frac{1}{k+1}} \right] \delta \ge C_{n,k} |\Omega|^{-\frac{1} {k(k+1)}}\lambda^{-\frac 1 k}\delta, \] that is \[ |\Omega|^{\frac{1}{k}} \ge C_{n,k} W_{k-1}(\Omega)^{\frac{n-2k}{k(n-k+1)}} \lambda^{-\frac 1 k}= C_{n,k} (1+\varepsilon)^{-\frac{1}{k}} \lambda_k(B_1)^{-\frac 1 k} W_{k-1}(\Omega)^{\frac{n}{k(n-k+1)}}, \] that is the thesis. Case $k= \frac n 2$. Arguing as before, we get \[ \log\left(\frac{W_{\frac n 2-1}(\Omega)} {W_{\frac n 2-1}(\Omega_\delta)}\right) \ge C_{n} |\Omega|^{-\frac {4} {n (n+2)}}\lambda^{-\frac 2 n} \delta. \] By Lemma \ref{lemma1kkk}, it follows that if $\varepsilon<2\delta|\Omega|^{\frac{2}{n+2}}<\frac{2}{n+2}$, \[ -\log\left( 1-\delta\left(n +2\right)|\Omega|^{\frac{2}{n+2}} \right) \ge C_{n} |\Omega|^{-\frac {4} {n (n+2)}}\lambda^{-\frac 2 n} \delta. \] Then, for $\delta$ such that $\delta (n+2) |\Omega|^{\frac{2}{n+2}}<\frac {1} {2(n+2)}$, \[ 2(n+2)|\Omega|^{\frac{2}{n+2}}\delta \ge C_{n}|\Omega|^{-\frac{4}{n(n+2)}}\lambda^{-\frac 2 n} \delta. \] Then, similarly as before, \[ |\Omega|^{\frac{2}{n}} \ge C_{n} W_{\frac n 2 -1}(\Omega)^{\frac{4}{(n+2)}}, \] and the proof of the Lemma is completed. \end{proof} Now we can prove the main theorem of this section. \begin{proof}[Proof of the Theorem \ref{main2}] Without loss of generality, we may suppose that $W_{k-1}(\Omega)=1$. Indeed, the quotient \[ \frac{W_{k-1}(K)-W_{k-1}(L)}{W_{k-1}(K)} \] is rescaling invariant. Consequently, by Lemma \ref{boundmisk} and the Aleksandrov-Fenchel inequality, we have that there exist two positive constants $c_1(n,k)$ and $c_2(n,k)$, which depend only on $n$ and $k$, such that \begin{equation} \label{boundk} c_1(n,k) \le |\Omega| \le c_2(n,k). \end{equation} The H\"older inequality gives that \begin{multline} \label{eq:holder} \varepsilon= \delta^{k+1} = \left( \int_0^\delta dt \right)^{k+1} \le \left(\int_0^\delta \big[-\frac{d}{dt} W_{k-1}(\Omega_t)\big] dt\right)^k \int_0^\delta \frac{1}{\big[-\frac{d}{dt} W_{k-1}(\Omega_t)\big]^k}dt = \\ = \left[W_{k-1}(\Omega)-W_{k-1}(\Omega_\delta)\right]^k \int_0^\delta \frac{1}{\big[-\frac{d}{dt} W_{k-1}(\Omega_t)\big]^{k}}dt. \end{multline} Hence, for $\varepsilon>0$ sufficiently small, $\delta$ verifies the hypothesis in Lemma \ref{lemma1kkk}, and the inequalities \eqref{eq:holder}, \eqref{lemma1k}, \eqref{eq:10} and \eqref{boundk} imply that \begin{multline*} \inf_{t\in [0,\delta]} \left( W_k(\Omega_t)^{k+1}- W_k((\Omega_t)^*_{k-1})^{k+1}\right) \le \\ \le \frac 1 \varepsilon \left[W_{k-1}(\Omega)-W_{k-1}(\Omega_\delta)\right]^k \int_0^{\max(-u)} \frac{ W_k(\Omega_t)^{k+1}- W_k((\Omega_t)^*_{k-1})^{k+1}} {\big[-\frac{d}{dt} W_{k-1}(\Omega_t)\big]^{k} } dt \le \\ \le C_{n,k} \varepsilon^{\frac {k}{k+1}} \lambda_k(\Omega^*_{k-1}). \end{multline*} Hence, for some $\tau \in [0,\delta]$, we have that \[ W_k(\Omega_\tau)^{k+1} \le W_k((\Omega_\tau)^*_{k-1})^{k+1} + C_{n,k} \varepsilon^{\frac{k}{k+1}}, \] being $W_{k-1}(\Omega)=1$. Moreover, an algebraic inequality and \eqref{boundk} give that \[ {\omega_n}^{-1} W_{k}(\Omega_\tau)^{n-k+1} \le W_{k-1}(\Omega_\tau)^{n-k} + C_{n,k} \varepsilon^{\frac{k}{k+1}}, \] Applying the estimates \eqref{gs} and \eqref{boundk}, and using the same notation of the proof of Theorem \ref{main}, we have \begin{equation*} (R_\tau-r_\tau)^{(n+3)/2} \le C_{n,k}\varepsilon^{\frac{k}{k+1}}. \end{equation*} Moreover, by \eqref{lemma1k} it follows that \begin{multline} \label{katenella} r_\tau \ge \rho_\tau - C_{n,k} \varepsilon^{\frac{2k}{(k+1)(n+3)}} = \left(\frac{W_{k-1}(\Omega_\tau)}{\omega_n}\right)^{\frac{1}{n-k+1}} - C_{n,k} \varepsilon^{\frac{2k}{(k+1)(n+3)}} \ge \\ \ge \left[ \frac{ W_{k-1}(\Omega) }{\omega_n} \left(1-C_{n,k}\varepsilon^{\frac{1}{k+1}}|\Omega|^{\frac{1}{k+1}} \right) \right]^{\frac{1}{n-k+1}} - C_{n,k} \varepsilon^{\frac{2k}{(k+1)(n+3)}} \ge \\ \ge \left[\frac{ W_{k-1}(\Omega) }{\omega_n}\right]^{\frac{1}{n-k+1}} \left( 1-\tilde C_{n,k}\varepsilon^{\frac{1}{k+1}} -C_{n,k} \varepsilon^{\frac{2k}{(k+1)(n+3)}} \right)\ge R \left( 1- C_{n,k}\varepsilon^\alpha \right), \end{multline} where $R=\left[ \omega_n^{-1}W_{k-1}(\Omega)\right]^{\frac{1}{n-k+1}}$ is the radius of $\Omega^*_{k-1}$, and $\alpha=\max\left\{\frac{1}{k+1}, \frac {2k}{(k+1)(n+3)}\right\}$. Hence, recalling \eqref{querball}, and being $r_\tau\le r_\Omega$ and $W_{k-1}(\Omega)=1$, then \begin{equation*} \label{kkkk} d_k(\Omega) \le \omega_n^{\frac{1}{n-k+1}}(R-r_\tau)\le C_{n,k} \varepsilon^{\alpha}, \end{equation*} that is the first estimate in \eqref{tesik}. In order to obtain the remaining estimates of the theorem, using the Aleksandrov-Fenchel inequalities, \eqref{katenella} and being $W_{k-1}(\Omega)=1$, we have that \begin{equation}\label{catenak} \begin{split} \left[\left( \frac{P(\Omega)}{n\omega_n} \right)^n - \left( \frac{ |\Omega| }{\omega_n}\right)^{n-1}\right]^{\frac{2}{n+3}} &\le \left[ \left( \frac{W_{k-1}(\Omega)^n}{\omega_n^n} \right)^{\frac{n-1}{n-k+1}} - \left( \frac{ |\Omega| }{\omega_n}\right)^{n-1}\right]^{\frac{2}{n+3}} \le \\ & \le \left[ \left( \frac{W_{k-1}(B_{r_\Omega})+C_{n,k} \varepsilon^{\alpha}}{\omega_n} \right)^{\frac{n(n-1)}{n-k+1}} - \left( \frac{ |B_{r_\Omega}| }{\omega_n}\right)^{n-1}\right]^{\frac{2}{n+3}} \le \\ & \le \left[ \left( \frac{W_{k-1}(B_{r_\Omega})}{\omega_n} \right)^{\frac{n(n-1)}{n-k+1}} + C_{n,k} \varepsilon^{\alpha} - \left( \frac{ |B_{r_\Omega}| }{\omega_n}\right)^{n-1}\right]^{\frac{2}{n+3}} = \\ & = C_{n,k} \varepsilon^{\frac{2\alpha}{n+3}}. \end{split} \end{equation} Hence, \eqref{catenak} and \eqref{gshaus} imply \[ \delta_H(\Omega)\le C_{n,k} \varepsilon^{\frac{2\alpha}{n+3}}, \] while, from \eqref{gs} we get \begin{equation*} D_k(\Omega) \le \omega_n^{\frac{1}{n-k+1}}(R_\Omega-r_\Omega) \le C_{n,k} \varepsilon^{\frac{2\alpha}{n+3}}, \end{equation*} and this concludes the proof. \end{proof} \begin{rem} Similarly as observed in Remark \ref{remdef}, from the proof of Theorem \ref{main2} it is possible to obtain that \[ \Delta(\Omega) \le C_{n,k} \varepsilon^{\frac{2\alpha}{n+3}}. \] \end{rem}
{ "timestamp": "2013-03-27T01:01:43", "yymm": "1302", "arxiv_id": "1302.7252", "language": "en", "url": "https://arxiv.org/abs/1302.7252", "abstract": "In this paper, we give some stability estimates for the Faber-Krahn inequality relative to the eigenvalues of Hessian operators", "subjects": "Analysis of PDEs (math.AP)", "title": "Stability results for some fully nonlinear eigenvalue estimates", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9845754461077707, "lm_q2_score": 0.8128673087708699, "lm_q1q2_score": 0.8003291931595022 }
https://arxiv.org/abs/1709.07290
Comparing the Switch and Curveball Markov Chains for Sampling Binary Matrices with Fixed Marginals
The Curveball algorithm is a variation on well-known switch-based Markov chain approaches for uniformly sampling binary matrices with fixed row and column sums. Instead of a switch, the Curveball algorithm performs a so-called binomial trade in every iteration of the algorithm. Intuitively, this could lead to a better convergence rate for reaching the stationary (uniform) distribution in certain cases. Some experimental evidence for this has been given in the literature. In this note we give a spectral gap comparison between two switch-based chains and the Curveball chain. In particular, this comparison allows us to conclude that the Curveball Markov chain is rapidly mixing whenever one of the two switch chains is rapidly mixing. Our analysis directly extends to the case of sampling binary matrices with forbidden entries (under the assumption of irreducibility). This in particular captures the case of sampling simple directed graphs with given degrees. As a by-product of our analysis, we show that the switch Markov chain of the Kannan-Tetali-Vempala conjecture only has non-negative eigenvalues if the sampled binary matrices have at least three columns. This shows that the Markov chain does not have to be made lazy, which is of independent interest. We also obtain an improved bound on the smallest eigenvalue for the switch Markov chain studied by Greenhill for uniformly sampling simple directed regular graphs.
\section{Introduction} The problem of uniformly sampling binary matrices with fixed row and column sums (marginals) has received a lot of attention, see, e.g., \cite{Rao1996,Kannan1999,Erdos2013,Erdos2015,Erdos2016}. Equivalent formulations for this problem are the uniform sampling of undirected bipartite graphs, or the uniform sampling of directed graphs with possible a self-loop at every node (but no parallel edges). One approach is to define a Markov chain on the space of all binary matrices for given fixed row and column sums, and study a random walk on this space induced by making small changes to a matrix using a given probabilisitic procedure (that defines the transition matrix). The idea, roughly speaking, is that after a sufficient amount of time, the so-called \emph{mixing time}, the resulting matrix almost corresponds to a sample from the uniform distribution over all binary matrices with given row and column sums. The most well-known probabilistic procedures for making these small changes use so-called switches, see, e.g., \cite{Rao1996}. More recently the Curveball algorithm was introduced in some experimental papers \cite{Verhelst2008,Strona2014}, which is a procedure that intuitively speeds up the mixing time of switch-based chains in many settings. The goal of this paper is to confirm this intuition by giving a spectral gap comparison for the Markov chains of the classical switch algorithm of Kannan, Tetali and Vempala \cite{Kannan1999} and the Curveball algorithm as formulated by Verhelst \cite{Verhelst2008}. We will start with an informal description of both algorithms. For a given initial binary matrix $A$, in every step of the switch algorithm we randomly choose two distinct rows and two distinct columns uniformly at random. If the $2 \times 2$ submatrix corresponding to these rows and columns is a \emph{checkerboard} $C_i$ for $i = 1,2$, where, $$ C_1 = \left(\begin{matrix} 1 & 0 \\ 0 & 1 \end{matrix} \right) \ \ \ \ \text{ and } \ \ \ \ C_2 = \left(\begin{matrix} 0 & 1 \\ 1 & 0 \end{matrix} \right), $$ then the $2 \times 2$ submatrix is replaced by $C_{i+1}$ for $i$ modulo $2$. That is, if the checkerboard is $C_1$, it is replaced by $C_2$, and vice versa. If the submatrix does not correspond to a checkerboard, nothing is changed. Such an operation is called a \emph{switch}. The Curveball algorithm intuitively speeds up the switch algorithm. In every step of the algorithm, first two rows are chosen uniformly at random from $A$ as in the switch algorithm. Then, a so-called \emph{binomial trade} is performed. In such a trade, we first look at all the columns in the $2 \times n$ submatrix given by the chosen rows, and we identify all the columns for which the column sum, in this submatrix, is one. That is, the column consist of precisely one $1$ and one $0$. For example if the $2 \times 6$ submatrix (i.e., $n = 6$) is given by $$ \begin{pmatrix} 1 & \mathbf{1} & 0 & \mathbf{0} & \mathbf{0} & \mathbf{1} \\ 1 & \mathbf{0} & 0 & \mathbf{1} & \mathbf{1} & \mathbf{0} \end{pmatrix} , $$ then we consider the (auxiliary) submatrix $$ \left(\begin{matrix} 1 & 0 & 0 & 1 \\ 0 & 1 & 1 & 0 \end{matrix} \right) $$ given by the second, fourth, fifth and sixth column. Let $u$ and $l$ respectively be the number of columns where the $1$ appears on the upper row and the lower row ($u = l = 2$ here) . We now uniformly at random draw a $2 \times (u+l)$ matrix with columns sums equal to $1$, and row sums equal to $u$ and $l$. Note that there are $\binom{u+l}{u}$ possible choices, hence the name binomial trade. We then replace the (auxiliary) submatrix with this new submatrix in $A$. Note that such a drawing can be obtained by uniformly choosing $u$ out of $u+l$ column indices Both these algorithms define a Markov chain on the set of all $m \times n$ binary matrices satisfying given row and columns sums $r$ and $c$. The main result of this work is a comparison of their relaxation times, or, equivalently, spectral gaps (see next section for definitions). \begin{theorem}[Relaxation time comparison]\label{thm:relax_comparison} Let $(1 - \lambda_*^c)^{-1}$ and $(1 - \lambda_*^s)^{-1}$, be the relaxation times of the Curveball and switch Markov chains respectively. Then, with $r_{\max}$ the maximum row sum, $$ \frac{2}{n(n-1)}\cdot (1 - \lambda_*^s)^{-1} \ \leq \ (1 - \lambda_*^c)^{-1} \ \leq \ \min\left\{1, \frac{(2r_{\max} + 1)^2}{2n(n-1)}\right\} \cdot (1 - \lambda_*^s)^{-1}. $$ \end{theorem} \noindent We present a more general comparison framework inspired by, and based on, the notion of a heat-bath Markov chain as introduced by Dyer, Greenhill and Ullrich \cite{Dyer2014}. We prove Theorem \ref{thm:relax_comparison} as an application of this framework in the the more general setting where the binary matrices can also have \emph{forbidden entries} that must be zero. This allows us to also compare the chains for the sampling of a simple directed graph with given degree sequence, as its adjacency matrix can be modeled by a square binary matrix with zeros on the diagonal. \subsection{Related work} Before going into related work, we would also like to refer the reader to \cite{Erdos2016} for a nice exposition on related work concerning the switch Markov chain. Kannan, Tetali and Vempala \cite{Kannan1999} conjectured that the KTV-switch chain is rapidly mixing for all fixed row and column sums. Mikl\' os, Erd\H{o}s and Soukup \cite{Erdos2013} proved the conjecture for half-regular binary matrices, in which all the row sums are equal (or all column sums), and Erd\H{o}s, Kiss, Mikl\' os and Soukup \cite{Erdos2015} extended this result to almost half-regular marginals. The authors prove this in a slightly more general context where there might be certain forbidden edge sets. The Curveball algorithm was first described by Verhelst \cite{Verhelst2008} and a slightly different version was later independently formulated by Strona, Nappo, Boccacci, Fattorini and San-Miguel-Ayanz \cite{Strona2014}. The name Curveball algorithm was introduced in \cite{Strona2014}. Theorem \ref{thm:relax_comparison} directly implies that the Curveball Markov chain is also rapidly mixing for (almost) half-regular marginals. For the uniform sampling of simple directed graphs with a given degree sequence, the most used switch algorithm is the \emph{edge-switch} version,\footnote{We will address this version as well.} see Greenhill \cite{Greenhill2011}, who gives a polynomial upper bound on the mixing time for the case of $d$-regular directed graphs, and Greenhill and Sfragara \cite{Greenhill2017} for some recent results on certain irregular degree sequences. The latter paper \cite{Greenhill2017} only considers degree sequences for which the edge-switch Markov chain is irreducible for a given degree sequence. The Curveball chain has also been formulated for (un)directed graphs, see Carstens, Berger and Strona \cite{Carstens2016}. A theoretical analysis for the mixing time of the Curveball Markov chain was raised as an open problem there. All the results regarding rapid mixing mentioned above rely on the multi-commodity flow method developed by Sinclair \cite{Sinclair1992}. In this work we omit multi-commodity flow techniques in order to compare the switch and Curveball Markov chains, but rather take a more elementary approach based on comparing eigenvalues of transition matrices. One seeming advantage of the eigenvalue comparison is that it allows us to compare the switch and Curveball chains for arbitrary fixed row and column sums. Our spectral gap comparisons are special cases of the classical comparison framework developed largely by Diaconis and Saloff-Coste and is based on so-called Dirichlet form comparisons of Markov chains, see, e.g., \cite{Diaconis1993b,Diaconis1993a}, and also Quastel \cite{Quastel1992}. See also the expository paper by Dyer, Goldberg, Jerrum and Martin \cite{Dyer2006}. As the stationary distributions are the same for all our Markov chains, we use a more direct, but equivalent, framework based on positive semidefiniteness. We briefly elaborate on this in Appendix \ref{app:dirichlet} for the interested reader. The transition matrix of the Curveball Markov chain is a special case of a heat-bath Markov chain, as introduced by Dyer, Greenhill and Ullrich \cite{Dyer2014}. Our work partially builds on \cite{Dyer2014} in the sense that we compare a Markov chain, with a similar decomposition property as in the definition of a heat-bath chain, to its heat-bath variant. We explain these ideas in the next section. \section{General framework}\label{sec:general} We consider an ergodic Markov chain $\mathcal{M} = (\Omega,P)$ with stationary distribution $\pi$, being strictly positive for all $x \in \Omega$, that is of the form\footnote{This description is almost the same as that of a heat-bath chain \cite{Dyer2014}, and is introduced to illustrate the conceptual idea.} \begin{equation}\label{eq:general_chain} P = \sum_{a \in \mathcal{L}} \rho(a) \sum_{R \in \mathcal{R}_a} P_{R} \end{equation} which is given by a \begin{enumerate}[i)] \item finite index set $\mathcal{L}$, and probability distribution $\rho$ over $\mathcal{L}$, \item partition $\mathcal{R}_a = \cup R_{k,a}$ of $\Omega$ for $a \in \mathcal{L}$. \end{enumerate} Moreover, the restriction of a matrix $P_R$ to the rows and columns of $R = R_{k,a}$ defines the transition matrix of an ergodic, time-reversible Markov chain on $R$ (and is zero elsewhere), with stationary distribution $$ \tilde{\pi}_R(x) = \frac{\pi(x)}{\pi(R)} $$ for $x \in R$. We use $1 = \lambda_0^R \geq \lambda_1^R \geq \dots \geq \lambda_{|R|-1}^R$ to denotes its eigenvalues. Note that these are also eigenvalues of $P_R$ and that all other eigenvalues of $P_R$ are zeros (as all rows and columns not corresponding to elements in $R$ only contain zeros). We use $\mathcal{R}$ to denote the multi-set $\cup_a \mathcal{R}_a$ indexed by pairs $(k,a)$. Note that the chain $\mathcal{M}$ proceeds by drawing an index $a$ from the set $\mathcal{L}$, and then performs a transition in the Markov chain on the set $R$ that the current state is in. The \emph{heat-bath variant} $\mathcal{M}_{heat}$ of the chain $\mathcal{M}$ is given by the transition matrix \begin{equation}\label{eq:heat_chain} P_{heat} = \sum_{a \in \mathcal{L}} \rho(a) \sum_{R \in \mathcal{R}_a} \mathbf{1}\cdot \sigma_R \end{equation} with $\sigma_R$ is a row-vector given by $\sigma_R(x) = \tilde{\pi}_R(x)$ if $x \in R$ and zero otherwise, and $\mathbf{1}$ the all-ones column vector. It can be shown that $\mathcal{M}_{heat}$ is an ergodic Markov chain as well. It is reversible by construction \cite{Dyer2014}.\footnote{The Curveball chain is the heat-bath variant of the KTV-switch chain as we will later prove.} \begin{theorem}\label{thm:heat_comparison} Let $\mathcal{M}$ be a Markov chain as in (\ref{eq:general_chain}), and $\mathcal{M}_{heat}$ its heat-bath variant as in (\ref{eq:heat_chain}). If $\alpha$ and $\beta$ are non-zero constants, with $\alpha\cdot \beta > 0$, such that \begin{equation} \label{eq:assumption} \min_{R \in \mathcal{R}} \ \ \min_{i = 1,\dots,R-1} \{ \lambda_i^R, \alpha - \beta(1 - \lambda_i^R) \} \geq 0, \end{equation} then \begin{equation} \label{eq:alpha_beta} \frac{1}{\alpha}\frac{1}{1 - \lambda_*^{heat}} \leq \frac{1}{\beta}\frac{1}{1 - \lambda_*}, \end{equation} where $\lambda_*^{(heat)}$ is the second largest eigenvalue of $P_{(heat)}$. In particular, if $\lambda_{R-1}^R \geq 0$ for every $R \in \mathcal{R}$, then $ (1 - \lambda_*^{heat})^{-1} \leq (1 - \lambda_*)^{-1} $. \end{theorem} The intuition behind Theorem \ref{thm:heat_comparison} is that in order to compare the relaxation times of a Markov chain and its heat-bath variant, it suffices to compare them locally on the sets $R$. Note that $\alpha$ and $\beta$ can both be negative, so that this statement can be used to lower bound the relaxation time of the heat-bath variant in terms of the original relaxation time as well. We will use the following propositions in the proof of Theorem \ref{thm:heat_comparison}. For $S \subseteq \Omega$, the matrix $I_S$ is defined by $I_S(x,x) = 1$ if $x \in S$ and zero otherwise. Also, a symmetric real-valued matrix $A$ is positive semidefinite if all its eigenvalues are non-negative, and this is denoted by $A \succeq 0$. \begin{proposition}[\cite{Zhang1999}]\label{prop:eigenvalue_comparison} Let $X,Y$ be symmetric $l \times l$ matrices. If $X - Y \succeq 0$, then $\lambda_i(X) \geq \lambda_i(Y)$ for $i = 1,\dots, l$, where $\lambda_i(C)$ is the $i$-th largest eigenvalue of $C = X,Y$. \end{proposition} \begin{proposition} \label{prop:eigen_difference} Let $X$ be the $k \times k$ transition matrix of an ergodic reversible Markov chain with stationary distribution $\pi$, and eigenvalues $1 = \lambda_0 > \lambda_1 \geq \dots \geq \lambda_{k-1}$. Let $X^* = \lim_{t \rightarrow \infty} X^t$ be the matrix containing the row vector $\pi$ on every row. Then the eigenvalues of $\alpha(I - X^*) - \beta(I - X)$ are $$ \{0\} \cup \{ \alpha - \beta(1 - \lambda_i) \ \big| \ i = 1,\dots,k-1\} . $$ for given constants $\alpha$ and $\beta$. \end{proposition} \begin{proof} As $X$ is the transition matrix of a reversible Markov chain, it holds that the matrix $V X V^{-1}$ is symmetric, where $V = \text{diag}(\pi_1^{1/2},\pi_2^{1/2},\dots,\pi_k^{1/2}) = \text{diag}(\sqrt{\pi})$.\footnote{This is the same argument for showing that a reversible Markov chain only has real eigenvalues.} Note that the eigenvalues of $\alpha(I - X^*) - \beta(I - X)$ are the same as those of $$ V(\alpha(I - X^*) - \beta(I - X))V^{-1} = \alpha(I - \sqrt{\pi}^T\sqrt{\pi}) - \beta(I - VXV^{-1}). $$ Moreover, with $1 = (1,1,1,\dots,1)^T$ the all-ones vector, we have $$ VXV^{-1}\sqrt{\pi}^T = V X \mathbf{1} = V \mathbf{1} = \sqrt{\pi}^T, $$ so that $\sqrt{\pi}^T$ is an eigenvector of $VXV^{-1}$ with eigenvalue $1$. It then follows that $\sqrt{\pi}^T$ is an eigenvector of $\alpha(I - \sqrt{\pi}^T\sqrt{\pi}) - \beta(I - VXV^{-1})$ with eigenvalue $0$. Let $\sqrt{\pi}^T = w_0, w_1,\dots,w_{k-1}$ be a basis of orthogonal eigenvectors for $VXV^{-1}$ corresponding to eigenvalues $\lambda_1,\dots,\lambda_{k-1}$ (note that $X$ and $VXV^{-1}$ have the same eigenvalues). It then follows that $$ [\alpha(I - \sqrt{\pi}^T\sqrt{\pi}) - \beta(I - VXV^{-1})] w_i = \alpha - \beta(1 - \lambda_i) $$ because of orthogonality. This completes the proof. \end{proof} \begin{proof}[Proof of Theorem \ref{thm:heat_comparison}] Let $D$ be the $|\Omega| \times |\Omega|$ diagonal matrix with $(D)_{xx} = \sqrt{\pi(x)}$. As the matrices $\mathbf{1} \cdot \sigma_R$ and $P_R$ define reversible Markov chains on $R$, the matrix $$ Y_R = D^{-1}[\alpha(I_R - \mathbf{1}\cdot \sigma_R) - \beta(I_R - P_{R})]D $$ is symmetric. Moreover, from the assumption in (\ref{eq:assumption}), together with Proposition \ref{prop:eigen_difference} and the fact that similar\footnote{Two square matrices $A$ and $B$ are \emph{similar} if there exists an invertible matrix $T$ such that $A = T^{-1}BT$.} matrices have the same set of eigenvalues, it follows that $Y_R$ is positive semidefinite. Since any non-negative linear combination of positive semidefinite matrices is again positive semidefinite, the matrix $$ D^{-1}[\alpha(I - P_{heat}) - \beta(I - P)]D = \sum_{a \in \mathcal{L}} \rho(a) \sum_{R \in \mathcal{R}_a} D^{-1}[\alpha(I_R - \mathbf{1}\cdot \sigma_R) - \beta(I_R - P_{R})]D $$ is also positive semidefinite. Using Proposition \ref{prop:eigenvalue_comparison}, and again the fact that similar matrices $A,B$ have the same set of eigenvalues, it follows that $$ \alpha(1 - \lambda_i^{heat}) \geq \beta (1 - \lambda_i) $$ where $\lambda_i^{(heat)}$ is the $i$-th largest eigenvalue of $P_{(heat)}$. Note that $P$ has non-negative eigenvalues as $D^{-1} P D$ is a non-negative linear combination of positive semidefinite matrices. A similar argument holds for $P_{heat}$ and was shown in \cite{Dyer2014}. In particular, it follows that $\lambda_1^{(heat)}$ is the second-largest eigenvalue of $P_{(heat)}$. This proves (\ref{eq:alpha_beta}). \end{proof} \subsection{Markov chain definitions} \label{sec:background} Let $\mathcal{M} = (\Omega,P)$ be an ergodic, time-reversible Markov chain over state space $\Omega$ with transition matrix $P$ and stationary distribution $\pi$. We write $P_x^t = P^t(x,\cdot)$ for the distribution over $\Omega$ at time step $t$ given that the initial state is $x \in \Omega$. It is well-known that the matrix $P$ only has real eigenvalues $1 = \lambda_0 > \lambda_1 \geq \lambda_2 \geq \dots \geq \lambda_{N-1} > -1$, where $N = |\Omega|$. Moreover, we define $\lambda_{*} = \max\{\lambda_1,|\lambda_{N-1}|\}$ is the second-largest eigenvalue of $P$. The \emph{variation distance} at time $t$ with initial state $x$ is $$ \Delta_x(t) = \max_{S \subseteq \Omega} \big| P^t(x,S) - \pi(s)\big| = \frac{1}{2}\sum_{y \in \Omega} \big| P^t(x,y) - \pi(y)\big| $$ and the mixing time $\tau(\epsilon)$ is defined as $$ \tau(\epsilon) = \max_{x \in \Omega}\left\{ \min\{ t : \Delta_x(t') \leq \epsilon \text{ for all } t' \geq t\}\right\}. $$ A Markov chain is said to be \emph{rapidly mixing} if the mixing time can be upper bounded by a function polynomial in $\ln(|\Omega|/\epsilon)$. It is well-known, e.g., following directly from Proposition 1 \cite{Sinclair1992}, that \begin{equation}\label{eq:mixing_time} \frac{1}{2}\frac{\lambda_*}{1 - \lambda_*} \ln(1/2\epsilon) \ \leq \ \tau(\epsilon) \ \leq \ \frac{1}{1 - \lambda_*}\cdot (\ln(1/\pi_*) + \ln(1/\epsilon)) \end{equation} where $\pi_* = \min_{x \in \Omega} \pi(x)$. This roughly implies that the mixing time is determined by the \emph{spectral gap} $(1 - \lambda_*)$, or its inverse, the \emph{relaxation time} $(1 - \lambda_*)^{-1}$. We also introduce some additional notation. We let $G_{\Omega} = (\Omega,A)$ be the state space graph, with an arc $(a,b) \in A$ if and only if $P(a,b) > 0$ for $a,b \in \Omega$ with $a \neq b$. If $P$ is symmetric, we define $H_{\Omega} = (\Omega,E)$ as the undirected counterpart of $G_{\Omega}$ with $\{a,b\} \in E$ if and only if $(a,b),(b,a) \in A$ with $a \neq b$. Moreover, the $\delta$-lazy version of $\mathcal{M}$ is the Markov chain defined by transition matrix $(1 - \delta)I + \delta P$ for $0 < \delta < 1$. Note that this chain is also ergodic, and time-reversible with stationary distribution $\pi$. \begin{proposition}\label{prop:lazy} If $0 < \delta < 1$ is such the transition matrix $(1- \delta)I + \delta P$ of the $\delta$-lazy version of $\mathcal{M}$ only has non-negative eigenvalues. Then $$ \frac{1}{1 - \lambda_{*,\delta}} \leq \frac{1}{\delta} \frac{1}{1 - \lambda_{*}} $$ where $\lambda_{*,\delta} = \lambda_{1,\delta} = (1 - \delta) + \delta \lambda_{1}$ is the second-largest eigenvalue of $(1- \delta) + \delta P$. \end{proposition} \begin{proof} If $\lambda_i$ is an eigenvalue of $P$ then $\lambda_{i,\delta} := (1 - \delta) + \delta \lambda_i$ is an eigenvalue of $(1 - \delta)I + \delta P$. Note that $\lambda_i \leq \lambda_j$ if and only if $\delta \lambda_i \leq \delta \lambda_j$, which is true if and only if $$ \lambda_{i,\delta} = (1 - \delta) + \delta \lambda_i \leq (1 - \delta) + \delta \lambda_j = \lambda_{j,\delta}. $$ This in particular shows that $\lambda_{1,\delta} = (1 - \delta) + \delta \lambda_{1}$ is indeed the second-largest eigenvalue of $(1 - \delta)I + \delta P$. Moreover, $\lambda_{i,\delta} = (1 - \delta) + \delta \lambda_i$ is equivalent to $$ \frac{1}{1 - \lambda_{i,\delta}} = \frac{1}{\delta} \frac{1}{1 - \lambda_{i}} $$ for $i > 0$. As the eigenvalues of $(1- \delta) + \delta P$ are all non-negative, we have $$ \frac{1}{1 - \lambda_{*,\delta}} = \frac{1}{1 - \lambda_{1,\delta}} = \frac{1}{\delta} \frac{1}{1 - \lambda_{1}} \leq \frac{1}{\delta} \frac{1}{1 - \lambda_{*}} $$ and this completes the proof. Note that the final inequality is true independent of the sign of $\lambda_1$. \end{proof} \subsection{Johnson graphs.}\label{sec:johnson} One class of graphs that are of particular interest in this work, are the so-called Johnson graphs. For given integers $1 \leq q \leq p$, the undirected Johnson graph $J(p,q)$ contains as nodes all subsets of size $q$ of $\{1,\dots,p\}$, and two subsets $u,v \subseteq \{1,\dots,p\}$ are adjacent if and only if $|u \cap v| = q - 1$. We refer the reader to \cite{Holton1993,Brouwer2011} for the following facts. The Johnson graph $J(p,q)$ is a $q(p-q)$-regular graph and the eigenvalues of its adjacency matrix are given by $$ (q - i)(p - q -i) - i \ \ \ \ \text{ with multiplicity }\ \ \ \ \binom{p}{i} - \binom{p}{i-1} $$ for $i = 0,\dots,q$, with the convention that $\binom{p}{-1} = 0$. The following observation is included for ease of reference. It will often be used to lower bound the smallest eigenvalue of a Johnson graph. \begin{proposition}\label{prop:johnson} Let $p,q \in \mathbb{N}$ be given. The continuous function $f : \mathbb{R} \rightarrow \mathbb{R}$ defined by $$ f(x) = [(q - x)(p - q - x) - x] - q(p - q) = x(x - (p+1)) $$ is minimized for $x^* = (p+1)/2$ and $f(x^*) = -(p+1)^2/4$. \end{proposition} \section{Binary matrices and the switch chain} We are given $n,m \in \mathbb{N}$, fixed row sums $r = (r_1,\dots, r_m)$, column sums $c = (c_1,\dots,c_n)$, and a set of forbidden entries $\mathcal{F} \subseteq \{1,\dots,m\} \times \{1,\dots,n\}$. The state space $\Omega = \Omega(r,c,\mathcal{F})$ is the set of all binary $m \times n$-matrices $A$ satisfying these row and column sums, and for which $A(a,b) = 0$ if $(a,b) \in \mathcal{F}$. For $A \in \Omega$, we let $A_{ij}$ be the $2 \times n$-submatrix formed by rows $i$ and $j$, for $1 \leq i < j \leq m$. We define \begin{equation}\label{eq:forbidden_T} U_{ij}(A) = \{k \in \{1,\dots,n\} : A(i,k) = 1,\ A(j,k) = 0 \text{ and } (j,k) \notin \mathcal{F}\}, \end{equation} with $u_{ij}(A) = |U_{ij}(A)|$, and similarly \begin{equation}\label{eq:forbidden_B} L_{ij}(A) = \{k \in \{1,\dots,n\} : A(i,k) = 0,\ A(j,k) = 1 \text{ and } (i,k) \notin \mathcal{F}\}, \end{equation} with $l_{ij}(A) = |L_{ij}(A)|$. Note that $L_{ij} \cup U_{ij}$ are precisely the columns $k$ for which $A_{ij}$ has different values on its rows and for which neither $(i,k)$ or $(j,k)$ is forbidden. Matrices $A, B \in \Omega$ are \emph{switch-adjacent for row $i$ and $j$} if $A = B$ or if $A - B$ contains exactly four non-zero elements that occur on rows $i$ and $j$, and the columns $k$ and $l$ containing these non-zero elements do not have forbidden entries in $A_{ij}$. Two matrices are switch-adjacent if they are switch-adjacent for some rows $i$ and $j$. \emph{$\gamma$-Switch chain.} We next introduce the notion of a $\gamma$-switch Markov chain which is done for notational convenience as there are multiple switch-based chains available in the literature. For feasible $\gamma > 0$, the transition matrix of such a chain on state space $\Omega = \Omega(r,c,\mathcal{F})$ is given by $$ P_{\gamma}(A,B) = \left\{ \begin{array}{ll} \binom{m}{2}^{-1}\cdot \gamma & \ \ \ \ \ \ \text{if } A \neq B \text{ are switch-adjacent}, \\ \binom{m}{2}^{-1} \sum_{1 \leq i < j \leq m} 1 - u_{ij}l_{ij} \cdot \gamma & \ \ \ \ \ \ \text{if } A = B, \\ 0 &\ \ \ \ \ \ \text{otherwise,} \end{array}\right. $$ provided $\gamma$ satisfies the following assumption. \begin{assumption}\label{assump:gamma} For given $n,m,r,c$ and $\mathcal{F}$, we assume that $\gamma$ is such that $$ 1 - u_{ij}(A)l_{ij}(A)\cdot \gamma > 0 $$ for all $A \in \Omega$ and $1 \leq i < j \leq m$. \end{assumption} Note that the transition probability for switch-adjacent matrices is the same everywhere in the state space, and does not depend on the matrices. In particular, the transition matrix $P_{\gamma}$ is symmetric and hence the chain is reversible with respect to the uniform distribution. The factor $2/(m(m-1))$ is included for notational convenience. The chain can roughly be interpreted as follows. We first choose two distinct rows $i$ and $j$ uniformly at random, and then transition to a different matrix switch-adjacent for rows $i$ and $j$, of which there are $u_{ij}l_{ij}$ possibilities, where every matrix has probability $\gamma$ of being chosen; and with probability $1 - u_{ij}l_{ij}\gamma$ we do nothing. Taking $\gamma = 2/(n(n-1))$ we get back the KTV-switch chain \cite{Kannan1999}. We will later show that (a lazy version of) the edge-switch chain in \cite{Greenhill2011,Greenhill2017} also falls within this definition. \medskip \begin{remark} We always assume that the set $\Omega(r,c,\mathcal{F})$ is non-empty, and that the $\gamma$-switch chain is irreducible (it is clearly always aperiodic and finite). Irreducibility is in particular guaranteed in the case there are no forbidden entries \cite{Rao1996}; or in case $n = m \geq 4$, with $\mathcal{F}$ is the set of diagonal entries, and regular marginals $c_i = r_i = d$ for some given $d \geq 1$ \cite{Greenhill2011}. Note that the condition of irreducibility is independent of the value of $\gamma$. \end{remark}\medskip We next explain that the $\gamma$-switch chain is of the form (\ref{eq:general_chain}). The index set $$ \mathcal{L} = \{(i,j) : 1 \leq i < j \leq m\} $$ is the set of all pairs of distinct rows, and $\rho$ is the uniform distribution over $\mathcal{L}$, that is, $\rho(a) = \binom{m}{2}^{-1}$ for all $a \in \mathcal{L}$. The partitions $\mathcal{R}_a$ for $a \in \mathcal{L}$ rely on the notion of a binomial neighborhood, that is also defined in \cite{Verhelst2008} to describe the Curveball Markov chain (the decomposition idea given here is novel). \begin{definition}[Binomial neighborhood] For a fixed binary matrix $A$ and row-pair $(i,j)$, the $(i,j)$-binomial neighborhood $\mathcal{N}_{ij}(A)$ of $A$ is the set of matrices that can be reached by only applying switches on rows $i$ and $j$. More formally, $N_{ij}(A)$ contains all binary matrices $B \in \Omega$ for which $A(k,l) = B(k,l)$ whenever $(k,l) \notin \{i,j\} \times U_{ij}(A) \cup L_{ij}(A)$, which in particular implies that $U_{ij}(A) \cup L_{ij}(A) = U_{ij}(B) \cup L_{ij}(B)$.\footnote{Said differently, $\mathcal{N}_{ij}(A)$ contains all matrices that can be reached by one trade on rows $i$ and $j$ in the Curveball algorithm, as described in the introduction.} \end{definition} It should be clear that two matrices $A, B \in \Omega$ can be part of \emph{at most} one common binomial neighborhood, see also \cite{Verhelst2008}. This follows directly from the observation that if $B \in \mathcal{N}_{ij}(A) \setminus \{A\}$, then $A$ and $B$ differ on precisely rows $i$ and $j$, so switches using any other pair of rows $\{k,l\} \neq \{i,j\}$ can never transform $A$ into $B$. Moreover, we have $A \in \mathcal{N}_{ij}(A)$; if $B \in \mathcal{N}_{ij}(A)$, then $A \in \mathcal{N}_{ij}(B)$ \cite{Verhelst2008}; and, if $A \in \mathcal{N}_{ij}(B)$, $B \in \mathcal{N}_{ij}(C)$, then $A \in \mathcal{N}_{ij}(C)$. That is, the relation $\sim_{ij}$ defined by $a \sim b$ if and only if $a \in \mathcal{N}_{ij}(b)$, is an equivalence relation on $\Omega$. The equivalence classes of $\sim_{ij}$ define the set $\mathcal{R}_{(i,j)}$. Finally, note that $u_{ij}(A) = u_{ij}(B)$ and $l_{ij}(A) = l_{ij}(B)$ if $A$ and $B$ are part of the same binomial neighborhood $\mathcal{N}$. Therefore, these numbers are only neighborhood-dependent, and not element-dependent within a fixed neighborhood. Observe that $$ |\mathcal{N}| = \binom{u_{ij} + l_{ij}}{u_{ij}}. $$ Moreover, another important observation is that the undirected state space graph (see Section \ref{sec:background}) $H$ of the $\gamma$-switch chain (which is the same for all $\gamma$) induced on a binomial neighborhood is isomorphic to a Johnson graph $J(u+l,u)$ whenever $u,l \geq 1$ (see Section \ref{sec:johnson} for notation and definition). If either $u = 0$ or $l = 0$ it consists of a single binary matrix. To see this, note that every element in the $(i,j)$-binomial neighborhood $\mathcal{N}_{ij}(A)$ can be represented by the set of indices of the columns $k$ for which $A(i,k) = 1, A(j,k) = 0$ and $(j,k) \notin \mathcal{F}$, which we denote by $Z(A_{ij})$. The set $\{1,\dots,l_{ij} + u_{ij}\}$ here is then the set indices of \emph{all} columns with precisely one $1$ and one $0$ on rows $i,j$ and that do not contain forbidden entries. Indeed, matrices $A \neq B$ are switch-adjacent for rows $i$ and $j$ if $Z(A_{ij}) \cap Z(B_{ij}) = u_{ij} - 1$. Informally, the Markov chain resulting from always deterministically choosing rows $i$ and $j$ in the switch algorithm, is the disjoint union of smaller Markov chains each with a state space graph isomorphic to some Johnson graph. \begin{example}\label{exmp:binom_neighborhood} Consider the binary matrix $$ A = \left(\begin{matrix} 0 & 1 & 1 & 0 & 1 & 0 & 1\\ 1 & 0 & 0 & 1 & 1 & 0 & 1 \\ 0 & 1 & 0 & 0 & 0 & 1 & 1 \end{matrix} \right) $$ and the $2 \times 7$-submatrix formed by rows $1$ and $2$, which is $$ A_{12} = \left(\begin{matrix} 0 & 1 & 1 & 0 & 1 & 0 & 1\\ 1 & 0 & 0 & 1 & 1 & 0 & 1 \end{matrix} \right). $$ For sake of simplicity, we (uniquely) describe every element of the $(1,2)$-binomial neighborhood $\mathcal{N}_{12}(A)$ by the first four columns (precisely those with column sums equal to one in the submatrix). For the switch chain, the induced subgraph of the undirected state space graph $H$ on the $(1,2)$-binomial neighborhood of $A$, the Johnson graph $J(4,2)$ is given in Figure \ref{fig:johnson}. \begin{figure}[h!] \centering \begin{tikzpicture}[scale=3.75] \coordinate (A1) at (-0.25,0); \coordinate (A2) at (0.25,0); \coordinate (M1) at (-0.5,0.375); \coordinate (M2) at (0.5,0.375); \coordinate (T1) at (-0.25,0.75); \coordinate (T2) at (0.25,0.75); \node at (A1) [circle,scale=0.7,fill=black] {}; \node (a1) [below=0.1cm of A1] {$\left(\begin{matrix} 1 & 0 & 0 & 1 \\ 0 & 1 & 1 & 0 \end{matrix} \right)$}; \node at (A2) [circle,scale=0.7,fill=black] {}; \node (a2) [below=0.1cm of A2] {$\left(\begin{matrix} 1 & 0 & 1 & 0 \\ 0 & 1 & 0 & 1 \end{matrix} \right)$}; \node at (M1) [circle,scale=0.7,fill=black] {}; \node (m1) [left=0.1cm of M1] {$\left(\begin{matrix} 0 & 1 & 1 & 0 \\ 1 & 0 & 0 & 1 \end{matrix} \right)$}; \node at (M2) [circle,scale=0.7,fill=black] {}; \node (m2) [right=0.1cm of M2] {$\left(\begin{matrix} 1 & 1 & 0 & 0 \\ 0 & 0 & 1 & 1 \end{matrix} \right)$}; \node at (T1) [circle,scale=0.7,fill=black] {}; \node (t1) [above=0.1cm of T1] {$\left(\begin{matrix} 0 & 1 & 0 & 1 \\ 1 & 0 & 1 & 0 \end{matrix} \right)$}; \node at (T2) [circle,scale=0.7,fill=black] {}; \node (t2) [above=0.1cm of T2] {$\left(\begin{matrix} 0 & 0 & 1 & 1 \\ 1 & 1 & 0 & 0 \end{matrix} \right)$}; \path[every node/.style={sloped,anchor=south,auto=false}] (T1) edge[-,very thick] node {} (T2) (T1) edge[-,very thick] node {} (M2) (T1) edge[-,very thick] node {} (A1) (T1) edge[-,very thick] node {} (M1) (T2) edge[-,very thick] node {} (M1) (T2) edge[-,very thick] node {} (A1) (T2) edge[-,very thick] node {} (A2) (M1) edge[-,very thick] node {} (A2) (M1) edge[-,very thick] node {} (M2) (A1) edge[-,very thick] node {} (M2) (A1) edge[-,very thick] node {} (A2) (A2) edge[-,very thick] node {} (M2); \end{tikzpicture} \quad \begin{tikzpicture}[scale=3] \coordinate (A1) at (-0.25,0); \coordinate (A2) at (0.25,0); \coordinate (M1) at (-0.5,0.375); \coordinate (M2) at (0.5,0.375); \coordinate (T1) at (-0.25,0.75); \coordinate (T2) at (0.25,0.75); \node at (A1) [circle,scale=0.7,fill=black] {}; \node (a1) [below=0.1cm of A1] {$\{1,4\}$}; \node at (A2) [circle,scale=0.7,fill=black] {}; \node (a2) [below=0.1cm of A2] {$\{1,3\}$}; \node at (M1) [circle,scale=0.7,fill=black] {}; \node (m1) [left=0.1cm of M1] {$\{2,3\}$}; \node at (M2) [circle,scale=0.7,fill=black] {}; \node (m2) [right=0.1cm of M2] {$\{1,2\}$}; \node at (T1) [circle,scale=0.7,fill=black] {}; \node (t1) [above=0.1cm of T1] {$\{2,4\}$}; \node at (T2) [circle,scale=0.7,fill=black] {}; \node (t2) [above=0.1cm of T2] {$\{3,4\}$}; \path[every node/.style={sloped,anchor=south,auto=false}] (T1) edge[-,very thick] node {} (T2) (T1) edge[-,very thick] node {} (M2) (T1) edge[-,very thick] node {} (A1) (T1) edge[-,very thick] node {} (M1) (T2) edge[-,very thick] node {} (M1) (T2) edge[-,very thick] node {} (A1) (T2) edge[-,very thick] node {} (A2) (M1) edge[-,very thick] node {} (A2) (M1) edge[-,very thick] node {} (M2) (A1) edge[-,very thick] node {} (M2) (A1) edge[-,very thick] node {} (A2) (A2) edge[-,very thick] node {} (M2); \end{tikzpicture} \caption{The induced subgraph $H$ for the switch chain on the $(1,2)$-binomial neighborhood of $A$. On the left we have indexed the nodes by the submatrices of the first four columns, and on the right by label sets, indicating the positions of the $1$'s on the top row (i.e., row $1$).} \label{fig:johnson} \end{figure} \end{example} \begin{remark} A fixed binomial neighborhood is reminiscient of the Bernoulli-Laplace Diffusion model, see, e.g., \cite{Diaconis1987,Donnelly1994} for an analysis of this model. Here, there are two bins with resp. $k$ and $n- k$ balls, and in every transition two randomly chosen balls, one from each bin, are interchanged between the bins. Indeed, the state space graph is then a Johnson graph \cite{Donnelly1994}. The transition probabilities are different, due to the non-zero holding probabilities in the switch algorithm, but the eigenvalues of this Markov chain are related to the eigenvalues of the switch Markov chain on a fixed binomial neighborhood, see also \cite{Diaconis1987,Donnelly1994} \end{remark} \medskip For a binomial neighborhood $\mathcal{N} = \mathcal{N}_{ij}(A)$ for given $i < j$ and $A \in \Omega$, the undirected graph $H_{\mathcal{N}} = (\Omega, E_{\mathcal{N}})$ is the graph where $E_{\mathcal{N}}$ forms the edge-set of the Johnson graph $J(u_{ij} + l_{ij},u_{ij})$ on $\mathcal{N} \subseteq \Omega$, and where all binary matrices $B \in \Omega \setminus \mathcal{N}$ are isolated nodes. We use $M(H_{\mathcal{N}})$ do denote its adjacency matrix. The discussion above leads to the following result summarizing that the $\gamma$-switch chain is of the form (\ref{eq:general_chain}), and that its heat-bath variant is precisely the Curveball Markov chain as in \cite{Verhelst2008} defined by transition matrix $$ P_c(A,B) = \left\{ \begin{array}{ll} \binom{m}{2}^{-1}\cdot \binom{u_{ij} + l_{ij}}{u_{ij}}^{-1} & \ \ \ \ \ \ \text{if } B \in \mathcal{N}_{ij}(A) \setminus \{A\}, \\ \binom{m}{2}^{-1}\sum_{1 \leq i < j \leq m} \binom{u_{ij} + l_{ij}}{u_{ij}}^{-1} & \ \ \ \ \ \ \text{if } A = B, \\ 0 &\ \ \ \ \ \ \text{otherwise.} \end{array}\right. $$ Roughly speaking, the Curveball chain is precisely the chain sampling uniform within a fixed binomial neighborhood. For $S \subseteq \Omega$, the identity matrix $I_{S}$ on $S$ is defined by $I_{S}(x,x) = 1$ if $x \in S$ and zero elsewhere, and the all-ones matrix $J_{S}$ on $S$ is defined by $J_S(x,y) = 1$ if $x,y \in S$ and zero elsewhere. \begin{theorem}\label{thm:switch_decomposition} The transition matrix $P_\gamma$ of the $\gamma$-switch chain is of the form (\ref{eq:general_chain}) namely \begin{equation}\label{eq:switch_decomposition} P_\gamma = \sum_{1 \leq i < j \leq m} \binom{m}{2}^{-1} \sum_{\mathcal{N} \in \mathcal{R}_{(i,j)}} (1 - u_{ij}l_{ij}\cdot\gamma)\cdot I_{\mathcal{N}} + \gamma \cdot M(H_\mathcal{N}). \end{equation} The heat-bath variant of the $\gamma$-switch chain is given by the Curveball chain, and can be written as \begin{equation}\label{eq:curveball_decomposition} P_c = \sum_{1 \leq i < j \leq m} \binom{m}{2}^{-1} \sum_{\mathcal{N} \in \mathcal{R}_{(i,j)}} \binom{u_{ij} + l_{ij}}{u_{ij}}^{-1} J_{\mathcal{N}}. \end{equation} \end{theorem} \begin{proof} The decomposition in (\ref{eq:switch_decomposition}) follows from the discussion above, and Assumption \ref{assump:gamma} guarantees that the matrix $$ (1 - u_{ij}l_{ij}\cdot\gamma)\cdot I_{\mathcal{N}} + \gamma \cdot M(H_\mathcal{N}) $$ indeed defines the transition matrix of a Markov chain for every $\mathcal{N}$. Moreover, remember that the $\gamma$-switch chain has uniform stationary distribution $\pi$ over $\Omega$. Indeed, for a binomial neighborhood $\mathcal{N} = \mathcal{N}_{ij}(A)$ for given $i < j$ and $A \in \Omega$, the vector $\sigma_{\mathcal{N}}$ as used in (\ref{eq:heat_chain}) is then given by $$ \sigma_{\mathcal{N}}(x) = \frac{\pi(x)}{\pi(\mathcal{N})} = \frac{1}{|\Omega|} \cdot \frac{|\Omega|}{|\mathcal{N}|} = \frac{1}{|\mathcal{N}|} = \binom{u_{ij} + l_{ij}}{u_{ij}}^{-1} $$ if $x \in \mathcal{N}$, and zero otherwise. This implies that $$ \mathbf{1}\cdot \sigma_{\mathcal{N}} = \binom{u_{ij} + l_{ij}}{u_{ij}}^{-1} J_{\mathcal{N}}. $$ as desired. \end{proof} This completes our description of the $\gamma$-switch chain as a Markov chain of the form (\ref{eq:general_chain}) with heat-bath variant the Curveball chain. We next study two explicit $\gamma$-switch chains. \subsection{KTV-switch chain} The switch chain of the Kannan-Tetali-Vempala conjecture, as described in the introduction, can be obtained by setting $\gamma = 2/(n(n-1))$. As the product $u_{ij}(A)l_{ij}(A)$ can be at most $n^2/4$ for any $A \in \Omega$ and $1 \leq i < j \leq m$, we see that $\gamma$ satisfies Assumption \ref{assump:gamma}. \begin{theorem}\label{thm:curveball_KTV} Let $P_c$ and $P_{KTV}$ be the transition matrices of resp. the Curveball and KTV-switch Markov chains with $n \geq 3$. Then $$ \frac{2}{n(n-1)}\cdot (1 - \lambda_*^{KTV})^{-1} \ \leq \ (1 - \lambda_*^c)^{-1} \ \leq \ \min\left\{1, \frac{(2r_{\max} + 1)^2}{2n(n-1)}\right\} \cdot (1 - \lambda_*^{KTV})^{-1}, $$ where $\lambda_*^{(KTV,c)} = \lambda_1^{(KTV,c)} $ is the second largest eigenvalue of $P_{(KTV,c)}$. In particular, $P_{KTV}$ only has non-negative eigenvalues. \end{theorem} \begin{proof} Let $\mathcal{N} = \mathcal{N}_{ij}(A)$ for given $i < j$ and $A \in \Omega$. We apply Theorem \ref{thm:heat_comparison} for various pairs $(\alpha,\beta)$. \emph{Case 1: $\alpha = \beta = 1$.} From (\ref{eq:assumption}) it follows that it suffices to show that for any binomial neighborhood $\mathcal{N}$ the submatrix of $$ Y_{\mathcal{N}} = \left[1 - u_{ij}l_{ij}\cdot\binom{n}{2}^{-1}\right] I_{\mathcal{N}} + \binom{n}{2}^{-1}M(H_{\mathcal{N}}) $$ formed by the rows and columns of $\mathcal{N}$ only has non-negative eigenvalues. For any eigenvalue $\lambda$ of this submatrix, we have $$ \lambda = 1 + (\mu - u_{ij}l_{ij})\binom{n}{2}^{-1} $$ where $\mu = \mu(\lambda)$ is an eigenvalue of the Johnson graph $J(u_{ij} +l_{ij}, u_{ij})$ on $\mathcal{N}$. In particular, using Proposition \ref{prop:johnson} with $p = u_{ij} + l_{ij}$ and $q = u_{ij}$, we get $ (\mu - u_{ij}l_{ij}) \geq -\frac{1}{4}(u_{ij} + l_{ij} + 1)^2 \geq -\frac{1}{4}(n + 1)^2 $ using that $0 \leq u_{ij} + l_{ij} \leq n$. Therefore, when $n \geq 5$, we have $$ \lambda \geq 1 - \frac{1}{2}\frac{(n+1)^2}{n(n-1)} \geq 0. $$ The cases $n = 3,4$ can be checked with some elementary arguments. This is left to the reader. Note that, in particular, this implies that $P_{KTV}$ only has non-negative eigenvalues when $n \geq 3$. \emph{Case 2: $\alpha = 1$ and $\beta = (2n(n-1))/((2r_{\max} + 1)^2)$.} Using similar notation as in the previous case, we show that $$ \lambda = 1 - \beta \left(1 - \left(1 + \left(\mu - u_{ij} \cdot l_{ij}\right)\binom{n}{2}^{-1}\right)\right) = 1 + \beta(\mu - u_{ij} \cdot l_{ij})\binom{n}{2}^{-1} \geq 0 $$ for any $\mu = \mu(\lambda)$ that is an eigenvalue of the Johnson graph $J(u_{ij} + l_{ij}, u_{ij})$. Again, using Proposition \ref{prop:johnson} in order to lower bound the quantity $(\mu -u_{ij} \cdot l_{ij})$, we find $$ 1 + \beta \cdot \left(\mu - u_{ij} \cdot l_{ij}\right)\binom{n}{2}^{-1} \geq 1 - \frac{\beta}{4}(u_{ij} + l_{ij} +1)^2\binom{n}{2}^{-1} \geq 1 - \frac{\beta}{4}(2r_{\max}+1)^2\binom{n}{2}^{-1} \geq 0, $$ using the fact that $0 \leq u_{ij} + l_{ij} \leq 2r_{\max}$ and the choice of $\beta$. \emph{Case 3: $\alpha = -1$ and $\beta = -2/(n(n-1))$.} We have to show that $$ \lambda = \binom{n}{2} \left(1 - \left(1 + \left(\mu - u_{ij} \cdot l_{ij}\right)\binom{n}{2}^{-1} \right)\right) - 1 = u_{ij} \cdot l_{ij} - \mu - 1 \geq 0 $$ for all $$ \mu = \mu(k) = (u - k)(l - k) - k $$ where $k = 1,\dots,u$. Note that the eigenvalue $u_{ij} \cdot l_{ij}$ for the case $k = 0$ yields the largest eigenvalue $1 = \lambda_0^{\mathcal{N}}$ of $Y_{\mathcal{N}}$, and does not have to be considered here. The maximum over $k = 1,\dots,u$ is then attained for $k = 1$, and we have $ u_{ij} \cdot l_{ij}- \mu - 1 \geq u_{ij} \cdot l_{ij} - ((u_{ij} - 1)(l_{ij} - 1) - 1) - 1 = u_{ij} + l_{ij} - 1 \geq 0, $ since $u_{ij}, l_{ij} \geq 1$. \end{proof} \subsection{Edge-switch chain} In every step of the edge-switch algorithm, two matrix-entries $(i,a)$ and $(j,b)$ from the set $\{(c,d) : A(c,d) = 1\}$ are chosen uniformly at random. We refer to it as the edge-switch algorithm, as for the interpretation of uniformly sampling directed graphs (where every node can have at most one self-loop), it corresponds to choosing two distinct edges uniformly at random. If the $2 \times 2$ submatrix corresponding to rows $i,j$ and columns $a,b$ forms a checkerboard, and if $(i,b)$ and $(j,b)$ are not forbidden entries, the checkerboard is adjusted (similar as for the KTV-switch algorithm as described in the introduction). Note that $$ P_{edge}(A,B) = \binom{\rho}{2}^{-1} $$ if $A$ and $B$ are switch-adjacent, where $\rho = \sum_{i} r_i$ is the total number of ones in every binary matrix in $\Omega$. Note that $$ \gamma = \binom{m}{2} \binom{\rho}{2}^{-1} $$ in this case. The analysis in the main part of this section implies that we can write \begin{equation}\label{eq:switch_reformulated} P_{edge} = \sum_{1 \leq i < j \leq m} \binom{m}{2}^{-1} \sum_{\mathcal{N} \in \mathcal{R}_{(i,j)}} \left[1 - u_{ij}l_{ij}\cdot\binom{m}{2}\binom{\rho}{2}^{-1}\right] I_{\mathcal{N}} + \binom{m}{2}\binom{\rho}{2}^{-1}M(H_{\mathcal{N}}) \end{equation} where $M(H_{\mathcal{N}})$ is the adjacency matrix of a Johnson graph for every $\mathcal{N}$. However, the matrix \begin{equation}\label{eq:switch_S} S_{\mathcal{N}} = \left[1 - u_{ij}l_{ij}\cdot\binom{m}{2}\binom{\rho}{2}^{-1}\right] I_{\mathcal{N}} + \binom{m}{2}\binom{\rho}{2}^{-1}M(H_{\mathcal{N}}) \end{equation} does not necessarily define the transition matrix of a Markov chain on $\mathcal{N}$, as the holding probabilities might be negative.\footnote{In versions (v1,v2) we wrongfully claim that these matrices are stochastic, from which we conclude that $(1 - \lambda_*^c)^{-1} \leq 2 (1 - \lambda_*^{edge})^{-1}$. We fix this claim in Theorem \ref{thm:comparison_edge} at the cost of a polynomial factor. We can therefore still conclude that the Curveball chain is rapidly mixing whenever the edge-switch chain is rapidly mixing. The results on regular instances, given later on, remain unchanged.} We circumvent this problem by making the edge-switch chain $\delta$-lazy for $\delta$ sufficiently small. This procedure can be carried out for any $\gamma$ that does not satisfy Assumption \ref{assump:gamma}, provided $\gamma$ is polynomially bounded. \begin{theorem}\label{thm:comparison_edge} There exists a non-negative $\delta = \text{poly}(n,m,\rho)^{-1}$ such that $$ \frac{1}{1 - \lambda_*^c} \leq \frac{1}{\delta}\cdot \frac{1}{1 - \lambda_*^{edge}} $$ where $\lambda_*^{c, (edge)}$ is the second largest eigenvalue of $P_{c, (edge)}$. \end{theorem} \begin{proof} Note that \begin{eqnarray} (1 - \delta)I + \delta P_{edge} &=& \sum_{i < j } \binom{m}{2}^{-1} \sum_{\mathcal{N} \in \mathcal{R}_{(i,j)}} (1 - \delta) + \delta \cdot S_{\mathcal{N}} \nonumber \\ & = & \sum_{i < j } \binom{m}{2}^{-1} \sum_{\mathcal{N} \in \mathcal{R}_{(i,j)}} \left[1 - \delta \cdot u_{ij}l_{ij}\binom{m}{2}\binom{\rho}{2}^{-1}\right] I_{\mathcal{N}} + \delta\cdot \binom{m}{2}\binom{\rho}{2}^{-1}M(H_{\mathcal{N}}) \nonumber \end{eqnarray} so by taking, e.g., $$ \delta = \frac{1}{2}\left[ \frac{n^2}{4}\binom{m}{2}\binom{\rho}{2}^{-1}\right]^{-1} $$ we see that the matrices $(1 - \delta) + \delta \cdot S_{\mathcal{N}}$ are stochastic matrices with non-negative eigenvalues, as all holding probabilities are at least $1/2$. Here we also use the fact that $u_{ij}(A)l_{ij}(A) \leq n^2/4$ for all $A \in \Omega$ and $1 \leq i < j \leq m$. We may conclude that $(1 - \lambda_{*}^c)^{-1} \leq (1 - \lambda_{*,\delta}^{edge})^{-1} \leq (1 - \lambda_{*}^{edge})^{-1}/\delta$ where we use Proposition \ref{prop:lazy} in the last inequality. \end{proof}\medskip For certain instances we can do better than the $\delta$ in the proof of the previous theorem. \begin{theorem}\label{thm:regular} Let $\Omega = \Omega(n,d,\mathcal{F})$ be the set of square $n \times n$ binary matrices with row and column sums equal to $d \in \mathbb{N}$, so that $\rho = nd$, and forbidden entries $\mathcal{F}$. Then $$ (1 - \lambda_*^{c})^{-1} \leq \left(\frac{2d+1}{2d}\right)^2 (1 - \lambda_*^{edge})^{-1} $$ \end{theorem} \begin{proof}With $S_{\mathcal{N}}$ as in (\ref{eq:switch_S}) we have that any eigenvalue $\lambda$ of $S_{\mathcal{N}}$ is of the form $$ \lambda = 1 + (\mu - u_{ij}l_{ij})\binom{n}{2}\binom{nd}{2}^{-1} $$ where $\mu = \mu(\lambda)$ is an eigenvalue of the Johnson graph $J(u_{ij} + l_{ij},u_{ij})$. Proposition \ref{prop:johnson} shows that $ (\mu - u_{ij}l_{ij}) \geq -(u_{ij}+ l_{ij} + 1)^2/4 \geq - (2d + 1)^2/4, $ using $0 \leq u_{ij} + l_{ij} \leq 2d$ in the last inequality. It then follows that $$ 1 + (\mu - u_{ij}l_{ij})\binom{n}{2}\binom{nd}{2}^{-1} = 1 - \frac{1}{4}\frac{(2d+1)^2n(n-1)}{nd(nd-1)} = 1 - \frac{1}{4}\frac{4d^2(n-1)}{d(nd-1)} - \frac{1}{4}\frac{(4d+1)(n-1)}{d(nd-1)}. $$ Note that $d(n-1) \leq nd - 1$ for all $n, d \geq 1$, from which it follows that $$ 1 + (\mu - u_{ij}l_{ij})\binom{n}{2}\binom{m}{2}^{-1} \geq - \frac{1}{4}\frac{(4d+1)(n-1)}{d(nd-1)} \geq -\frac{1}{d} - \frac{1}{4d^2} $$ for all $n \in \mathbb{N}$. This implies that $(1/d+1/(4d^2))I_{\mathcal{N}} + S_{\mathcal{N}}$ is positive semidefinite. Rescaling, and rewriting, gives that $$ \left[1 - \left(\frac{2d}{2d+1}\right)^2\right]I_{\mathcal{N}} + \left(\frac{2d}{2d+1}\right)^2 S_{\mathcal{N}} $$ is a symmetric stochastic transition matrix with only non-negative eigenvalues, i.e., we can take $$ \delta = \left(\frac{2d}{2d+1}\right)^2. $$ \end{proof} \begin{corollary}\label{cor:regular} Let $\Omega = \Omega(n,d,\mathcal{F})$ be the set of square $n \times n$ binary matrices with row and column sums equal to $d \in \mathbb{N}$, so that $\rho = nd$, and forbidden entries $\mathcal{F}$, and let $\lambda_{|\Omega|-1}^{edge}$ be the smallest eigenvalue of $P_{edge}$. Then $$ (1 + \lambda_{|\Omega|-1}^{edge})^{-1} \leq \frac{4d^2}{4d^2 - 4d - 1} \leq \frac{5}{2} $$ if $d \geq 2$. \end{corollary} \begin{proof} In the proof of Theorem \ref{thm:regular} it was shown that $$ \left( \frac{1}{d} + \frac{1}{4d^2}\right) I + P_{edge} = \sum_{1 \leq i < j \leq m} \binom{m}{2}^{-1} \sum_{\mathcal{N} \in \mathcal{R}_{(i,j)}} \left( \frac{1}{d} + \frac{1}{4d^2}\right)I_{\mathcal{N}} + S_{\mathcal{N}} \succeq 0, $$ and hence $ \lambda_{|\Omega|-1}^{edge} \geq -\left( \frac{1}{d} + \frac{1}{4d^2}\right). $ Rewriting this gives the result. \end{proof} With $\mathcal{F}$ the set of diagonal entries, this improves a bound of $(1 + \lambda_{|\Omega|-1}^{edge})^{-1} \leq n^2d^2/4$ of Greenhill \cite{Greenhill2011} for the edge-switch chain for the sampling of simple directed regular graphs. \section{Parallelism in the Curveball chain} As a binary matrix is only adjusted on two rows at the time in the Curveball algorithm, one might perform multiple binomial trades in parallel on distinct pairs of rows \cite{Carstens2016}. To be precise, in every step of the so-called \emph{$k$-Curveball algorithm}, we choose a set of $k \leq \lfloor m/2 \rfloor$ disjoint pairs of rows uniformly at random and perform a binomial trade on every pair (see introduction). For $k = \lfloor m/2 \rfloor$ this corresponds to the Global Curveball algorithm described in \cite{Carstens2016}. We show that the induced $k$-Curveball chain is of the form (\ref{eq:general_chain}). The index set $\mathcal{L} = \mathcal{L}_k$ is the collection of all sets containing $k$ pairwise disjoint sets of two rows, i.e., $$ \left\{\{(1_a,1_b),(2_a,2_b),\dots,(k_a,k_b)\} \ : \ 1_a,1_b,\dots,k_a,k_b \in [m], \ |\{1_a,1_b,2_a,2_b,\dots,k_a,k_b\}| = 2k \right\}, $$ and $\rho$ is the uniform distribution over $\mathcal{L}$. For a fixed collection $\kappa \in \mathcal{L}_k$, we define the $\kappa$-neighborhood $\mathcal{N}_{\kappa}(A)$ of binary matrix $A \in \Omega$ as the set of binary matrices $B \in \Omega$ that can be obtained from $A$ by binomial trade-operations (see introduction) only involving the row-pairs in $\kappa$. Formally speaking, we have $B \in \mathcal{N}_{\kappa}(A)$ if and only if there exist binary matrices $A_l$ for $l = 0,\dots,k-1$, so that $$ A_{l+1} \in \mathcal{N}_{(l+1)_a,(l+1)_b}(A_l) $$ where $A = A_0$ and $B = A_{k}$. Note that the matrices $A_l$ might not all be pairwise distinct, as $A$ and $B$ could already coincide on certain pairs of rows in $\kappa$. Also note that $u_{i_ai_b}(A) = u_{i_ai_b}(B)$ and $l_{i_ai_b}(A) = l_{i_ai_b}(B)$ if $B \in \mathcal{N}_{\kappa}(A)$ for $i = 1,\dots,k$. It is not hard to see that such a neighborhood is isomorphic to a Cartesian product $W_1 \times W_2 \times \dots \times W_k$ of finite sets $W_1,\dots,W_k$ with $$ |W_i| = \binom{u_{i_ai_b} + l_{i_ai_b}}{u_{i_ai_b}}.\footnote{That is, the elements of $W_i$ describe a matrix on row-pair $(i_a,i_b)$.} $$ Moreover, the relation $\sim_{\kappa}$ defined by $a \sim_{\kappa} b$ if and only if $b \in \mathcal{N}_{\kappa}(a)$ defines an equivalence relation, and its equivalence classes give the set $\mathcal{R}_{\kappa}$. We now consider the following artificial formulation of the original Curveball chain: we first select $k$ pairs of distinct rows uniformly at random, and then we choose one of those pairs uniformly at random and apply a binomial trade on that pair. It should be clear that this generates the same Markov chain as when we directly select a pair of distinct rows uniformly at random. For $\mathcal{N}_{\kappa} \in \mathcal{R}_{\kappa}$ the matrix $P_{\mathcal{N}_{\kappa}}$ restricted to the rows and columns in $\mathcal{N}_{\kappa}$ is then the transition matrix of a Markov chain over $W_1 \times \dots \times W_k$, where in every step we choose an index $i \in [k]$ uniformly at random and make a transition in $W_i$ based on the (uniform) transition matrix $$ Q_{i} = \binom{u_{i_ai_b} + l_{i_ai_b}}{u_{i_ai_b}} ^{-1} J $$ where $J$ is the all-ones matrix of approriate size. More formally, the matrix $P_{\mathcal{N}_{\kappa}}$ restricted to the columns and rows in $\mathcal{N}_{\kappa}$ is given by \begin{equation}\label{eq:trans_product} \frac{\sum_{i = 1}^k \left[ \mathbf{\otimes}_{j = 1}^{i-1} \mathcal{I}_j\right] \otimes Q_{i} \otimes \left[\otimes_{j = i+1}^k \mathcal{I}_j\right]}{k}, \end{equation} forming a transition matrix on $\mathcal{N}_{\kappa}$, and is zero elsewhere. Here $\mathcal{I}_j$ is the identity matrix with the same size as $Q_j$ and $\otimes$ the usual tensor product. The eigenvalues of the matrix in (\ref{eq:trans_product}) are given by \begin{equation}\label{eq:eigen_product} \lambda_{\mathcal{N}_{\kappa}} = \left\{\frac{1}{k} \sum_{i = 1}^k \lambda_{{j_i},i} : 0 \leq j_i \leq |W_i| - 1\right\} \end{equation} where $1= \lambda_{0,i} \geq \lambda_{1,i} \geq \dots \geq \lambda_{|W_i| - 1,i}$ are the eigenvalues of $Q_i$ for $i = 1,\dots,k$.\footnote{See, e.g., \cite{Erdos2015decomposition} for a similar argument regarding the transition matrix, and eigenvalues, of a Markov chain of this form. These statements follow directly from elementary arguments involving tensor products.} It then follows that $$ P_{c} = \sum_{\kappa \in \mathcal{L}_k} \frac{1}{|\mathcal{L}_k|} \sum_{\mathcal{N}_{\kappa} \in \mathcal{R}_{\kappa}} P_{\mathcal{N}_{\kappa}} $$ which is of the form (\ref{eq:general_chain}). For $k = 1$, we get back the description of the previous section. Now, its heat-bath variant is precisely the $k$-Curveball Markov chain $$ P_{k-Curveball} = \sum_{\kappa \in \mathcal{L}_k} \frac{1}{|\mathcal{L}_k|} \sum_{\mathcal{N}_{\kappa} \in \mathcal{R}_{\kappa}} \frac{1}{|\mathcal{N}_{\kappa}|} J_{\mathcal{N}_{\kappa}}, $$ where $$ |\mathcal{N}_{\kappa}| = \prod_{i = 1}^k \binom{u_{i_ai_b} + l_{i_ai_b}}{u_{i_ai_b}}^{-1} $$ as, roughly speaking, for a fixed neighborhood $\mathcal{N}_{\kappa}$, the $k$-Curveball chain is precisely the uniform sampler over such a neighborhood. \begin{theorem}\label{thm:global_curveball} We have $$ \frac{(1 - \lambda_*^{c})^{-1}}{k} \ \leq \ (1 - \lambda_*^{k,c})^{-1} \ \leq \ (1 - \lambda_*^{c})^{-1} $$ where $\lambda_*^{k,c}$ is the second-largest eigenvalue of the $k$-Curveball chain, and $\lambda_*^c$ the second-largest eigenvalue of the $1$-Curveball chain. \end{theorem} \begin{proof} The upper bound follows from Theorem \ref{thm:heat_comparison}, with $\alpha = \beta = 1$, as the eigenvalues of all the $Q_i$ are non-negative, and therefore (\ref{eq:eigen_product}) implies that the eigenvalues of the matrix in (\ref{eq:trans_product}) are also non-negative. For the lower bound, we take $\alpha = -1$ and $\beta = -k$. That is, we have to show that $$ -1 + k(1 - \mu) $$ with $\mu \in \lambda_{\mathcal{N}_{\kappa}} \setminus \{1\}$ as in (\ref{eq:eigen_product}). It is not hard to see that the second-largest eigenvalue in $\lambda_{\mathcal{N}_{\kappa}}$ is $(k-1)/k$, as the eigenvalues of every fixed $Q_i$ are $1 = \lambda_{0,i} > \lambda_{1,i} = \dots = \lambda_{|W_i| - 1} = 0$. This implies that $$ -1 + k(1 - \mu) \geq -1 + k(1 - (k-1)/k) \geq 0 $$ for all $\mu \in \lambda_{\mathcal{N}_{\kappa}} \setminus \{1\}$. \end{proof} In general, the upper bound is tight for certain (degenerate) cases, that is, parallelism in the Curveball chain does not necessarily guarantee an improvement in its relaxation time. E.g., take column marginals $c_i = 1$ for $i = 1,\dots,n$, and row-marginals $r_1 = r_2 = n/2$ and $r_3 = r_4 = 0$, and consider $k = 2$. \section{Conclusion} We believe similar ideas as in this work can be used to prove that the Curveball chain is rapidly mixing for the sampling of undirected graphs with given degree sequences \cite{Carstens2016}, whenever one of the switch chains is rapidly mixing for those marginals. We leave this for future work, as the proof we have in mind is a bit more involved, but of a very similar nature as the ideas described here. An interesting direction for future work is to give a better comparison between the edge-switch chain and Curveball chain. It would also be interesting to see if there exist classes of marginals for which one can give a strict improvement over the result in Theorem \ref{thm:global_curveball}. \subsection*{Acknowledgements} Pieter Kleer is grateful to Annabell Berger and Catherine Greenhill for some useful discussions and comments regarding this work. \bibliographystyle{plain}
{ "timestamp": "2017-10-19T02:07:27", "yymm": "1709", "arxiv_id": "1709.07290", "language": "en", "url": "https://arxiv.org/abs/1709.07290", "abstract": "The Curveball algorithm is a variation on well-known switch-based Markov chain approaches for uniformly sampling binary matrices with fixed row and column sums. Instead of a switch, the Curveball algorithm performs a so-called binomial trade in every iteration of the algorithm. Intuitively, this could lead to a better convergence rate for reaching the stationary (uniform) distribution in certain cases. Some experimental evidence for this has been given in the literature. In this note we give a spectral gap comparison between two switch-based chains and the Curveball chain. In particular, this comparison allows us to conclude that the Curveball Markov chain is rapidly mixing whenever one of the two switch chains is rapidly mixing. Our analysis directly extends to the case of sampling binary matrices with forbidden entries (under the assumption of irreducibility). This in particular captures the case of sampling simple directed graphs with given degrees. As a by-product of our analysis, we show that the switch Markov chain of the Kannan-Tetali-Vempala conjecture only has non-negative eigenvalues if the sampled binary matrices have at least three columns. This shows that the Markov chain does not have to be made lazy, which is of independent interest. We also obtain an improved bound on the smallest eigenvalue for the switch Markov chain studied by Greenhill for uniformly sampling simple directed regular graphs.", "subjects": "Discrete Mathematics (cs.DM)", "title": "Comparing the Switch and Curveball Markov Chains for Sampling Binary Matrices with Fixed Marginals", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9873750518329434, "lm_q2_score": 0.8104789155369047, "lm_q1q2_score": 0.800246661237759 }
https://arxiv.org/abs/math/9805084
Coloring Distance Graphs on the Integers
Given a set D of positive integers, the associated distance graph on the integers is the graph with the integers as vertices and an edge between distinct vertices if their difference lies in D. We investigate the chromatic numbers of distance graphs. We show that, if $D = {d_1,d_2,d_3,...}$, with $d_n | d_{n+1}$ for all n, then the distance graph has a proper 4-coloring. We further find the exact chromatic numbers of all such distance graphs. Next, we characterize those distance graphs that have periodic proper colorings and show a relationship between the chromatic number and the existence of periodic proper colorings.
\section{Introduction} \label{S:intro} What is the least number of classes into which the integers can be partitioned, so that no two members of the same class differ by a square? What if ``square'' is replaced by ``factorial''? Questions like these can be formulated as graph coloring problems. Given a set $D$ of positive integers, the \emph{distance graph} $\gz{D}$ is the graph with the integers as vertices and an edge between distinct vertices if their difference lies in $D$; we call $D$ the \emph{distance set} of this graph. A \emph{proper coloring} of a graph is an assignment of colors to the vertices so that no two vertices joined by an edge receive the same color. The \emph{chromatic number} of a graph $G$, denoted by $\chi(G)$, is the least number of colors in a proper coloring. We abbreviate $\chi\big(\gz{D}\big)$ by $\ensuremath{\chi(D)}$. We refer to~\cite{BoMu76,WesD96} for graph-theoretic terminology not defined here. When $D$ is the set of all positive squares, we call $\gz{D}$ the \emph{square distance graph}. When $D$ is the set of all factorials, we obtain the \emph{factorial distance graph}. The questions at the beginning of this section ask for the chromatic numbers of these two graphs. We will study the chromatic numbers of these and other distance graphs on the integers. Distance graphs on the integers were introduced by Eggleton, Erd{\H o}s, and Skilton in~\cite{EES85}. In \cite{EES85,EES86}, the problem was posed of characterizing those distance sets $D$, containing only primes, such that $\ensuremath{\chi(D)}=4$. This problem was studied in \cite{EggR88,EES90,VoWa91,VoWa94}; see also \cite{EES88}. More recently, \cite{CCH97,DeZh97} have discussed the chromatic numbers of more general distance graphs with distance sets having 3 or 4 elements. In this paper, we are primarily interested in distance graphs for which the distance set is infinite, although our results apply to finite distance sets as well. We begin in Section~\ref{S:easy} with some easy lemmas on connectedness and bounds on the chromatic number. In Section~\ref{S:dc}, we consider distance graphs for which the distance set is totally ordered by the divisibility relation. We determine the chromatic numbers of all such graphs; in particular, we prove that they are all $4$-colorable. In Section~\ref{S:pc}, we study periodic proper colorings of distance graphs and their relationship to the chromatic number. Throughout this paper we will use standard notation for intervals to denote sets of consecutive integers. For example, $\iv{2}{6}$ denotes the set $\left\{2,3,4,5,6\right\}$. \section{Basic Results} \label{S:easy} In this section, we establish some basic facts about the connectedness and chromatic number of distance graphs. The results of this section have all been at least partially stated in earlier works. Our first result characterizes those distance sets for which the distance graph is connected. This result has been partially stated or implicitly assumed in a number of earlier works; see~\cite[p.~95]{EES85}. For $D$ a set of positive integers, we note that $\ensuremath{\gcd(D)}$ is well defined when $D$ is infinite. Given a real number $k$ and a set $D$, we denote by $k\cdot D$ the set $\left\{\,kd:d\in D\,\right\}$. \begin{lemma} \label{L:conn} Let $D$ be a nonempty set of positive integers. The graph $\gz{D}$ is connected if and only if $\ensuremath{\gcd(D)}=1$. Further, each component of $\gz{D}$ is isomorphic to $\gzmatch{\frac{1}{\ensuremath{\gcd(D)}}\cdot D}$. \end{lemma} \begin{proof} There is a path between vertices $k$ and $k+1$ if and only if there exist $d_1,\dotsc,d_a,e_1,\dotsc,e_b\in D$ such that $d_1+\dotsb+d_a-e_1-\dotsb-e_b=1$. This happens precisely when $\ensuremath{\gcd(D)}=1$, and so the first statement of the lemma is true. For the second statement, one isomorphism is the function\break $\varphi\colon\ensuremath{\gcd(D)}\cdot\ensuremath{\mathbb{Z}}\to\ensuremath{\mathbb{Z}}$ defined by $\varphi(k)=\frac{k}{\ensuremath{\gcd(D)}}$.\ggcqed \end{proof} When we determine the chromatic numbers of distance graphs,\break Lemma~\ref{L:conn} will often allow us to assume that the GCD of the distance set is $1$. Next, we prove a useful upper bound on the chromatic number. This result is a slight generalization of a result of Chen, Chang, and Huang \cite[Lemma~2]{CCH97}. \begin{lemma} \label{L:nomult} Let $D$ be a nonempty set of positive integers, and let $k$ be a positive integer. If $\frac{1}{\ensuremath{\gcd(D)}}\cdot D$ contains no multiple of $k$, then $\ensuremath{\chi(D)}\le k$. \end{lemma} \begin{proof} Let $D$ and $k$ be as stated. By Lemma~\ref{L:conn} we may assume that $\ensuremath{\gcd(D)}=1$. Thus, we assume that $D$ contains no multiple of $k$. We color the integers with colors $\iv{0}{k-1}$, assigning to each integer $i$ the color corresponding to the residue class of $i$ modulo $k$. Two integers will be assigned the same color precisely when they differ by a multiple of $k$. Since no multiple of $k$ occurs in $D$, this is a proper $k$-coloring of $\gz{D}$.\ggcqed\end{proof} The converse of Lemma~\ref{L:nomult} holds when $k=2$. This gives us a characterization of bipartite distance graphs: $\gz{D}$ is bipartite precisely when $\frac{1}{\ensuremath{\gcd(D)}}\cdot D$ contains no multiple of $2$, that is, when all elements of $D$ have the same power of $2$ in their prime factorizations. This result has been partially stated in earlier works; see~\cite[Thms.~8 \&~10]{EES85} and~\cite[Thms.~3 \&~4]{CCH97}. \begin{proposition} \label{P:chi2} Let $D$ be a set of positive integers. The graph $\gz{D}$ is bipartite if and only if there exists a non-negative integer $k$ so that $\frac{1}{2^k}\cdot D$ contains only odd integers. \end{proposition} \begin{proof} We may assume $D\ne\emptyset$. Since a graph is bipartite if and only if each component is bipartite, we may also assume, by Lemma~\ref{L:conn}, that $\ensuremath{\gcd(D)}=1$. For such $D$ we show that $\gz{D}$ is bipartite if and only if each element of $D$ is odd. $(\Longrightarrow)$ Since $\ensuremath{\gcd(D)}=1$, $D$ must have an odd element $d$. Suppose that $D$ has an even element $e$. If we begin at $0$, take $e$ steps in the positive direction, each of length $d$, ending at $de$, and then take $d$ steps in the negative direction, each of length $e$, ending at $0$, then we have followed a closed walk of odd length. Formally, the set \[\left\{0,d,d\cdot2,\dotsc,d(e-1),de,(d-1)e,(d-2)e,\dotsc,2e,e\right\}\] is the vertex set of an odd circuit, and so $\gz{D}$ is not bipartite. $(\Longleftarrow)$ If every element of $D$ is odd, then $\ensuremath{\chi(D)}\le2$, by Lemma~\ref{L:nomult}.\ggcqed\end{proof} The converse of Lemma~\ref{L:nomult} does not hold when $k>2$. For example, let $k>2$, and let $D=\left\{1,k\right\}$. Then $\frac{1}{\ensuremath{\gcd(D)}}\cdot D$ contains a multiple of $k$, and yet $\ensuremath{\chi(D)}\le3\le k$ (this is not hard to show; it will also follow from Lemma~\ref{L:dc3}). As with general graphs, it appears to be quite difficult to determine when a distance graph has a proper $k$-coloring, for $k\ge3$. However, when $D$ is finite, there does exist an algorithm to determine $\ensuremath{\chi(D)}$. This was proven for $D$ a finite set of primes by Eggleton, Erd\H os, and Skilton \cite[Corollary to Thm.~2]{EES90}; essentially the same proof works for more general sets. \begin{theorem} \label{T:existsalg} There exists an algorithm to determine $\ensuremath{\chi(D)}$ for $D$ a finite set of positive integers.\end{theorem} \begin{proof} (Outline---see~\cite[Thm.~2]{EES90}) Let $q=\max(D)$. Then $\ensuremath{\chi(D)}\le q+1$, by Lemma~\ref{L:nomult}. We consider the colorings of the subgraph of $\gz{D}$ induced by $S=\iv{1}{q^q+q}$. We show that, for $k\le q$, if $S$ has a proper $k$-coloring, then $\ensuremath{\chi(D)}\le k$; thus, $\ensuremath{\chi(D)}$ can be determined by a bounded search. Let $k\le q$, and suppose that $S$ has a proper $k$-coloring. The number of $k$-colorings of a block of $q$ consecutive integers is at most $q^q$. Since $S$ contains $q^q+1$ such blocks, two such blocks contained in $S$ (say $\iv{a}{a+q-1}$ and $\iv{b}{b+q-1}$, with $a<b$) receive the same pattern of colors. We extend the coloring of $\iv{a}{b+q-1}$ to a coloring $f$ of $\ensuremath{\mathbb{Z}}$ using the rule $f(i+a-b)=f(i)$, for all $i$. We can show that this is a proper coloring if $\gz{D}$, and so $\ensuremath{\chi(D)}\le k$.\ggcqed\end{proof} While an algorithm exists to determine $\ensuremath{\chi(D)}$ for finite $D$, we do not know whether there is an efficient algorithm. For finite graphs, determining whether the chromatic number is at most $k$ is NP-complete~\cite{GaJo79}. We conjecture that this is also true for distance graphs with finite distance sets. \begin{conjecture} \label{J:3npc} Let $k\ge3$. Determining whether $\ensuremath{\chi(D)}\le k$ for finite sets $D$ is NP-complete.\ggcqed\end{conjecture} \section{Divisibility Chains} \label{S:dc} We now focus on a particular class of distance graphs: those in which the distance set is totally ordered by divisibility. We show that all such graphs are $4$-colorable, and we determine their chromatic numbers. A \emph{divisibility chain} is a set of positive integers that is totally ordered by the divisibility relation. When $D$ is a (finite or infinite) divisibility chain we denote the elements of $D$ by $d_1,d_2,\dotsc$, where $d_1\mid d_2\mid\dotsb$. The \emph{ratios} of $D$ are the numbers $r_i=\frac{d_{i+1}}{d_i}$, for each $i$. When determining $\ensuremath{\chi(D)}$, we may, by Lemma~\ref{L:conn}, assume that $\ensuremath{\gcd(D)}=d_1=1$. Thus, $\ensuremath{\chi(D)}$ depends only on the ratios. We may also assume that all the $d_i$'s are distinct, that is, that none of the ratios is equal to $1$. A \emph{string} over $\{1,2\}$ is a finite sequence of $1$'s and $2$'s, written without spaces or separators. For example, $\alpha=1211$ is a string of length 4 with $\alpha_1=1$, $\alpha_2=2$, etc. For $k$ a positive integer, a string $\alpha$ is \emph{$k$-compatible} with a distance set $D$ if there is a proper $k$-coloring of $\gz{D}$ with colors $\iv{0}{k-1}$ such that the differences, modulo $k$, between colors of consecutive vertices form repeated copies of $\alpha$. Below is part of such a coloring with $k=4$ and $\alpha=1211$. \newcommand{{}}{{}} \newcommand{\phantom{0}}{\phantom{0}} \newcommand{{\,}}{{\,}} \newcommand{\qp\qp\qs}{\phantom{0}\qp{}} \newcommand{\qx\qp\qs}{{\,}\phantom{0}{}} \[ \begin{array}{rl} \text{vertex}& \qx0{} \qp\qp\qs \qp1{} \qp\qp\qs \qp2{} \qp\qp\qs \qp3{} \qp\qp\qs \qp4{} \qp\qp\qs \qp5{} \qp\qp\qs \qp6{} \qp\qp\qs \qp7{} \qp\qp\qs \qp8{} \qp\qp\qs \qp9{} \qp\qp\qs 10{} \qp\qp\qs 11{} \qp\qp\qs 12{} \cr \text{color}& \qx0{} \qp\qp\qs \qp1{} \qp\qp\qs \qp3{} \qp\qp\qs \qp0{} \qp\qp\qs \qp1{} \qp\qp\qs \qp2{} \qp\qp\qs \qp0{} \qp\qp\qs \qp1{} \qp\qp\qs \qp2{} \qp\qp\qs \qp3{} \qp\qp\qs \qp1{} \qp\qp\qs \qp2{} \qp\qp\qs \qp3{} \cr \text{difference}& \qx\qp\qs \qp1{} \qp\qp\qs \qp2{} \qp\qp\qs \qp1{} \qp\qp\qs \qp1{} \qp\qp\qs \qp1{} \qp\qp\qs \qp2{} \qp\qp\qs \qp1{} \qp\qp\qs \qp1{} \qp\qp\qs \qp1{} \qp\qp\qs \qp2{} \qp\qp\qs \qp1{} \qp\qp\qs \qp1{} \qp\qp\qs \cr \end{array} \] We see that $1211$ is not $4$-compatible with $\{3\}$, since, for example, $1$ and $4$ receive the same color; this is because the sum of three consecutive entries of the repeated copies of $\alpha$ is divisible by $4$ (i.e., $2+1+1=4$). Generally, a string $\alpha$ is $k$-compatible with $\{d\}$ if the concatenation of repeated copies of $\alpha$ contains no $d$ consecutive entries whose sum is a multiple of $k$. \begin{theorem} \label{T:dc4} If $D$ is a divisibility chain, then $\ensuremath{\chi(D)}\le4$. \end{theorem} \begin{proof} We may assume that the $d_i$'s are all distinct, and that $d_1=1$. We use notation such as $\alpha^n$ to denote a string; the superscript does not denote exponentiation or concatenation. \smallskip\noindent\emph{Claim.} For $n=1,2,3,\dotsc$, there exist strings $\alpha^n$, $\beta^n$ of length $d_n$ over $\left\{1,2\right\}$ such that \begin{enumerate} \item $\alpha^n$, $\beta^n$ differ only in the first entry, with $\alpha^n_1=1$, and $\beta^n_1=2$, and \label{I:dc4pf-differ} \item if $\gamma$ is a string resulting from the concatenation of any number of copies of $\alpha^n$ and/or $\beta^n$, in any order, then $\gamma$ is $4$-compatible with $\left\{d_1,d_2,d_3,\dotsc,d_n\right\}$. \label{I:dc4pf-compat} \end{enumerate} Before we prove the claim, we show that the theorem follows from it. If the claim holds, then, for each $n$, $\alpha^n$ is $4$-compatible with $\left\{d_1,d_2,\dotsc,d_n\right\}$, and so $\gzmatch{\left\{d_1,d_2,\dotsc,d_n\right\}}$ has a proper 4-coloring. Since every finite subgraph of $\gz{D}$ is isomorphic to a finite subgraph of $\gzmatch{\left\{d_1,\dotsc,d_n\right\}}$ for some $n$, every finite subgraph of $\gz{D}$ is 4-colorable, and we may conclude that $\ensuremath{\chi(D)}\le4$, by a compactness argument. Hence, it suffices to prove the claim. \smallskip\noindent\emph{Proof of Claim.} We proceed by induction on $n$. For $n=1$, we assumed that $d_n=1$. Let $\alpha^n=1$, and let $\beta^n=2$; these satisfy the claim for $n=1$. Now suppose that $n\ge1$, and that the claim holds for $n$. Define $s$ and $t$ as follows. \[ s:=\sum_{i=1}^{d_n}\alpha^n_i;\qquad t:=\sum_{i=1}^{d_n}\beta^n_i=s+1. \] We show first that $\iv{r_n\cdot s}{r_n\cdot t}$ contains integers $w$, $w+1$, neither a multiple of $4$. If $r_n>2$, then this is true since there are at least 4 consecutive integers in $\iv{r_n\cdot s}{r_n\cdot t}$. On the other hand, if $r_n=2$, then $r_n\cdot s$ and $r_n\cdot t$ are both even. Exactly one of the two is divisible by four. If $4\mid\left(r_n\cdot s\right)$, then let $w=r_n\cdot s+1$; otherwise, let $w=r_n\cdot s$. Now we choose $a\ge1$, $b\ge0$ so that $a+b=r_n$ and $as+bt=w$: let $b=w-r_n\cdot s$, and let $a=r_n-b$. We define $\alpha^{n+1}$ to be the concatenation of $a$ copies of $\alpha^n$ followed by $b$ copies of $\beta^n$. We let $\beta^{n+1}$ be the concatenation of $\beta^n$ followed by $a-1$ copies of $\alpha^n$ followed by $b$ copies of $\beta^n$; equivalently, $\beta^{n+1}$ is $\alpha^{n+1}$ with its first entry replaced by $2$. Now, $\alpha^{n+1}$ and $\beta^{n+1}$ both have length $d_{n+1}$, since $a+b=r_n$, and $\alpha^{n+1}$ and $\beta^{n+1}$ differ only in the first entry. Let $\gamma$ be a concatenation of copies of $\alpha^{n+1}$, $\beta^{n+1}$. Then $\gamma$ is a concatenation of copies of $\alpha^n$ and $\beta^n$, and so, by the induction hypothesis, $\gamma$ is $4$-compatible with $\left\{d_1,d_2,\dotsc,d_n\right\}$. In order to prove that $\alpha^{n+1}$, $\beta^{n+1}$ satisfy the claim, it remains only to show that $\gamma$ is $4$-compatible with $\left\{d_{n+1}\right\}$. This is true if the concatenation of repeated copies of $\gamma$ has no $d_{n+1}$ consecutive entries whose sum is a multiple of $4$. Since $\alpha^{n+1}$ and $\beta^{n+1}$ differ in only one entry, the sum of $d_{n+1}$ consecutive entries of repeated copies of $\gamma$ is equal either to the sum of the entries of $\alpha^{n+1}$ or to the sum of the entries of $\beta^{n+1}$; that is, it is equal \begin{align*} \text{either to}\quad\sum_{i=1}^{d_{n+1}}\alpha^{n+1}_i&=as+bt=w,\\ \text{or to}\quad\sum_{i=1}^{d_{n+1}}\beta^{n+1}_i&=t+(a-1)s+bt=w+1. \end{align*} Neither of these is a multiple of $4$. Thus, the claim is proven.\ggcqed\end{proof} The bound in Theorem~\ref{T:dc4} is sharp: if $D=\left\{1,2,6\right\}$, then the subgraph of $\gz{D}$ induced by $\iv{1}{7}$ has no proper 3-coloring. On the other hand, graphs satisfying the hypotheses of the theorem need not have chromatic number 4, even if $D$ is infinite. For example, if $D$ is the set of all powers of $3$, then every element of $D$ is odd, and so $\ensuremath{\chi(D)}=2$, by Proposition~\ref{P:chi2}. \begin{example} \label{G:usedc4} Let $D=\left\{d_1,d_2,d_3,\dotsc\right\}$, where $d_i=i!$ for each $i$. We use the technique of the above proof to produce part of a proper 4-coloring of $\gz{D}$, the factorial distance graph. Let $\alpha^1=1$ and $\beta^1=2$. We find consecutive nonmultiples of 4 in $\iv{2\cdot1}{2\cdot2}=\{2,3,4\}$: let $w=2$, so that $w+1=3$. So, $a=2$, and $b=0$. The string $\alpha^2$ is $2$ copies of $\alpha^1$ followed by $0$ copies of $\beta^1$. That is, $\alpha^2=11$, and so $\beta^2=21$. Continuing, we find consecutive nonmultiples of 4 in $\iv{3\cdot2}{3\cdot3}=\{6,7,8,9\}$: let $w=6$, so that $w+1=7$. So, $a=3$, and $b=0$. The string $\alpha^3$ is $3$ copies of $\alpha^2$ followed by $0$ copies of $\beta^2$. That is, $\alpha^3=111111$, and so $\beta ^3=211111$. Once again, we find consecutive nonmultiples of 4 in $\iv{4\cdot6}{4\cdot7}=\iv{24}{28}$: let $w=25$, so that $w+1=26$. So, $a=3$, and $b=1$. The string $\alpha^4$ is $3$ copies of $\alpha^3$ followed by $1$ copy of $\beta^3$. That is, $\alpha^4=111111111111111111211111$, and so $\beta^4=211111111111111111211111$. The coloring of $\iv{1}{24}$ obtained from $\alpha^4$ is the following. \[0,1,2,3,0,1,2,3,0,1,2,3,0,1,2,3,0,1,2,0,1,2,3,0.\ggcenddef\] \end{example} In \emph{almost} the entire proof of Theorem~\ref{T:dc4}, ``4'' can be replaced by ``3''; that is, we use $3$-compatibility instead of $4$-compatibility, we find a $3$-coloring instead of a $4$-coloring, and we find consecutive nonmultiples of $3$ instead of $4$. The one place where $4$ is required is the argument in the proof showing the existence of two consecutive nonmultiples when $r_n=2$. Thus, if we require that $r_n\ne 2$ for each $n$, then we can replace $4$ by $3$ in the proof, and we have the following result. \begin{lemma} \label{L:dc3} Let $D$ be a divisibility chain, with ratios $r_1,r_2,\dotsc$. If $r_i\ne2$ for all $i$, then $\ensuremath{\chi(D)}\le3$.\ggcqed\end{lemma} Again, the bound in this result is sharp: if $D=\left\{1,4\right\}$, then the subgraph of $\gz{D}$ induced by $\iv{1}{5}$ has no proper 2-coloring. We now find $\ensuremath{\chi(D)}$ for every divisibility chain $D$. \begin{theorem} \label{T:dcchi} Let $D$ be a divisibility chain, with ratios $r_1,r_2,\dotsc$. All of the following hold. \begin{enumerate} \item $\ensuremath{\chi(D)}\le4$. \label{I:dcchi4} \item $\ensuremath{\chi(D)}\le3$ if and only if there do not exist $i$, $j$ with $i<j$, $r_i=2$, and $3\mid r_j$. \label{I:dcchi3} \item $\ensuremath{\chi(D)}\le2$ if and only if $r_i$ is odd, for each $i$. \label{I:dcchi2} \item $\ensuremath{\chi(D)}=1$ if and only if $D=\emptyset$. \label{I:dcchi1} \end{enumerate} \end{theorem} \begin{proof} Statement~\ref{I:dcchi4} follows from Theorem~\ref{T:dc4}, statement~\ref{I:dcchi2} follows from Proposition~\ref{P:chi2}, and statement~\ref{I:dcchi1} holds because a graph is 1-colorable precisely when it has no edges. It remains to prove statement~\ref{I:dcchi3}. We may assume that $d_1=1$. $(\Longrightarrow)$ Suppose that there exist $i$ and $j$ with $i<j$, $r_i=2$, and $3\mid r_j$. Then $d_{i+1}=2d_i$, and $d_{j+1}$ is divisible by $3d_i$. Suppose that $\gz{D}$ has a proper $3$-coloring. Consider the colors assigned to the multiples of $d_i$. Since $d_i,2d_i\in D$, vertices $0$, $d_i$, and $2d_i$ induce a complete subgraph and so must be assigned 3 different colors. Similarly, $d_i$, $2d_i$, and $3d_i$ must receive 3 different colors, and so $0$ and $3d_i$ have the same color. Continuing this argument, all multiples of $3d_i$ must receive the same color, including $0$ and $d_{j+1}$, which is impossible. $(\Longleftarrow)$ Suppose there do not exist $i$ and $j$ with the properties specified in statement~\ref{I:dcchi3}; that is, every ratio divisible by $3$ precedes every ratio equal to $2$ in the list $\left\{r_1,r_2,\dotsc\right\}$. If there exist infinitely many ratios that are divisible by $3$, then, by our assumption, there exists no ratio equal to $2$, and so $\ensuremath{\chi(D)}\le3$, by Lemma~\ref{L:dc3}. Thus, we may assume that there are only finitely many ratios that are divisible by $3$. Let $c$ be the least positive integer such that $3\nmid r_i$, for all $i\ge c$. Then none of $r_1,r_2,\dotsc,r_{c-1}$ is equal to $2$. Thus, by Lemma~\ref{L:dc3}, the graph $\gzmatch{\left\{d_1,d_2,\dotsc,d_c\right\}}$ has a proper $3$-coloring. By the proof of Lemma~\ref{L:dc3}---that is, the proof of Theorem~\ref{T:dc4}, as modified to prove Lemma~\ref{L:dc3}---there is a string $\alpha^c$ of length $d_c$ over $\left\{1,2\right\}$ such that $\alpha^c$ is $3$-compatible with $\left\{d_1,d_2,\dotsc,d_c\right\}$. We claim that $\alpha^c$ is $3$-compatible with $D$. To see this, first note that \[ \sum_{i=1}^{d_c}\alpha^c_i \] is not a multiple of $3$, since $\alpha^c$ is $3$-compatible with $\left\{d_c\right\}$. Thus, if integers $x$ and $y$ differ by a multiple of $d_c$, then, in a $3$-coloring whose differences, modulo $3$, form repeated copies of $\alpha^c$, $x$ and $y$ receive the same color precisely when their difference is a multiple of $3d_c$. Now, no $r_i$ with $i\ge c$ is divisible by $3$; thus, no $d_i$ with $i\ge c$ is divisible by $3d_c$. We conclude that, for each $i\ge c$, no two integers with difference $d_i$ receive the same color, and so $\alpha^c$ is $3$-compatible with $\left\{d_c,d_{c+1},d_{c+2},\dotsc\right\}$. Thus, $\alpha^c$ is $3$-compatible with $D$, and we have $\ensuremath{\chi(D)}\le3$.\ggcqed\end{proof} By Theorem~\ref{T:dcchi}, the chromatic number of the factorial distance graph is $4$. We will have more to say about this graph in the next section. \section{Periodic Colorings} \label{S:pc} In this section, we consider periodic proper colorings of distance graphs. We characterize those distance graphs that have no periodic proper coloring, and we find a relationship between the chromatic number and the nonexistence of periodic proper colorings. Periodic colorings have been previously studied in~\cite{EES90}. \begin{lemma} \label{L:nomultper} Let $D$ be a set of positive integers, and let $k$ be a positive integer. If $D$ contains no multiple of $k$, then $\gz{D}$ has a periodic proper $k$-coloring. \end{lemma} \begin{proof} We may assume $D\ne\emptyset$. The proof of Lemma~\ref{L:nomult} gives a periodic proper $k$-coloring of each component of $\gz{D}$; this results in a periodic proper $k$-coloring of the graph.\ggcqed\end{proof} We can use Lemma~\ref{L:nomultper} to characterize those distance graphs that have no periodic proper coloring. The following result generalizes an observation of Eggleton~\cite{EggR97} that the square distance graph has no periodic proper coloring. \begin{proposition} \label{P:noper} Let $D$ be a set of positive integers. The graph $\gz{D}$ has no periodic proper coloring if and only if $D$ contains a multiple of every positive integer. \end{proposition} \begin{proof} $(\Longrightarrow)$ If there is some positive integer $k$ such that $D$ contains no multiple of $k$, then, by Lemma~\ref{L:nomultper}, $\gz{D}$ has a periodic proper coloring. $(\Longleftarrow)$ Let $D$ contain a multiple of every positive integer. Let $\gz{D}$ be colored in a periodic manner; say this coloring has period $k$. Every pair of vertices whose difference is a multiple of $k$ will have the same color. Since $D$ contains some multiple of $k$, this cannot be a proper coloring.\ggcqed\end{proof} \begin{remark} \label{R:fact} It follows from Theorem~\ref{T:dcchi} that the chromatic number of the factorial distance graph is $4$. However, by Proposition~\ref{P:noper}, the factorial distance graph has no periodic proper coloring.\ggcenddef\end{remark} Now we examine the effect of the existence of uniquely colorable subgraphs on proper colorings of distance graphs. We prove a useful lower bound on the chromatic number based on uniquely colorable subgraphs and periodic colorings. \begin{proposition} \label{P:uniqueper} Let $D$ be a set of positive integers, and let $k$ be a positive integer. If $\gz{D}$ has a finite, uniquely $k$-colorable subgraph, then every proper $k$-coloring of $\gz{D}$ is periodic. \end{proposition} \begin{proof} Suppose that $H$ is a uniquely $k$-colorable subgraph of $\gz{D}$. We may assume that the least integer that is a vertex of $H$ is $1$. Let $n$ be the greatest-numbered vertex of $H$. Since $H$ is uniquely $k$-colorable, every $k$-coloring of $\iv{1}{n-1}$ that can be extended to a proper $k$-coloring of $\gz{D}$ has a unique extension to a proper $k$-coloring of $\iv{1}{n}$. In short, once we have $k$-colored $\iv{1}{n-1}$, the color of vertex $n$ is forced. But, $\iv{2}{n+1}$ also contains a copy of $H$, and so once we have colored $\iv{2}{n}$, the color of vertex $n+1$ is forced. By an inductive argument, we can see that $k$-coloring $\iv{1}{n-1}$ completely determines the coloring of $\left[1,\infty\right)$. Essentially the same argument works in the opposite direction: $k$-coloring $\iv{2}{n}$ forces a certain color to occur at vertex $1$. Hence, $k$-coloring any set of $n-1$ consecutive vertices determines the coloring of all of $\ensuremath{\mathbb{Z}}$. Now, there are only a finite number of $k$-colorings of $n-1$ consecutive integers. Since the colorings of blocks of $n-1$ consecutive integers must eventually repeat, every proper $k$-coloring of the distance graph is periodic.\ggcqed\end{proof} Proposition~\ref{P:noper} and Proposition~\ref{P:uniqueper} have nearly opposite conclusions; the former concludes that the graph has no periodic proper coloring, while the latter concludes that every proper $k$-coloring of the distance graph is periodic. Suppose that a distance graph satisfies the hypothesis of both propositions, that is, the distance set contains a multiple of every positive integer, and the graph has a finite, uniquely $k$-colorable subgraph. Then the conclusions of both propositions must be true: there is no periodic proper coloring, and yet every proper $k$-coloring is periodic. We can only conclude that the distance graph must have no proper $k$-coloring at all, and so we have the following result. \begin{theorem} \label{T:omega+1} Let $D$ be a set of positive integers, and let $k$ be a positive integer. If $D$ contains a multiple of every positive integer, and $\gz{D}$ has a finite, uniquely $k$-colorable subgraph, then $\ensuremath{\chi(D)}\ge k+1$.\ggcqed\end{theorem} We can use Theorem~\ref{T:omega+1} to place a lower bound on the chromatic number of the square distance graph. Let $D$ be the set of all positive squares. Any Pythagorean triple gives a $K_3$ in the square distance graph. For example, the vertices $0$, $3^2$, $5^2$ induce a $K_3$, since $3,4,5$ is a Pythagorean triple. Since $\gz{D}$ has a $K_3$ subgraph, $\ensuremath{\chi(D)}\ge3$. Furthermore, $K_3$ is uniquely $3$-colorable, and $D$ contains a multiple of every positive integer. Thus, $\ensuremath{\chi(D)}\ge4$, by Theorem~\ref{T:omega+1}. Eggleton~\cite{EggR97} has found a $K_4$ in the square distance graph: the vertices are $0$, $672^2$, $680^2$, and $697^2$. We have $680^2-672^2=104^2$, $697^2-680^2=153^2$, and $697^2-672^2=185^2$. Since the square distance graph has a uniquely $4$-colorable subgraph, we have the following result. \begin{corollary} \label{C:sqchige5} The chromatic number of the square distance graph is at least $5$.\ggcqed\end{corollary} We do not know whether the square distance graph contains a $K_5$ or whether its chromatic number is greater than $5$. \begin{problem} \label{O:sqchi} What is the chromatic number of the square distance graph? Equivalently, what is the least number of classes into which the integers can be partitioned, so that no two members of the same class differ by a square?\ggcqed\end{problem} It seems likely that no finite number of colors suffices. We can ask similar questions about the distance graph resulting when $D$ is the set of all positive $n$th powers, for $n\ge3$. We know that these graphs contain no $K_3$ (this is equivalent to ``Fermat's Last Theorem'', proven by Wiles \cite{WilA95}), that they do not have periodic proper colorings, by Proposition~\ref{P:noper}, and that their chromatic numbers are all at least $3$, by Theorem~\ref{T:omega+1} (or Proposition~\ref{P:chi2}). It seems likely that these graphs have infinite chromatic number as well. As noted in Section~\ref{S:intro}, determining which distance graphs have chromatic number at most $k$, for a given $k\ge3$, appears to be difficult. A similar problem, whose difficultly we cannot estimate at this time, is the following. \begin{problem} \label{O:charinf} Characterize those sets $D$ such that $\ensuremath{\chi(D)}$ is\break infinite.\ggcqed \end{problem} No coloring requiring an infinite number of colors is periodic. Thus, by Proposition~\ref{P:noper}, a necessary condition for such sets $D$ is that they contain a multiple of every integer. However, this condition is not sufficient, by Remark~\ref{R:fact}. \section*{Acknowledgments} \label{S:ack} The author is grateful to Professor Roger Eggleton for bringing this topic to his attention and for helpful discussions.
{ "timestamp": "1998-05-19T22:08:10", "yymm": "9805", "arxiv_id": "math/9805084", "language": "en", "url": "https://arxiv.org/abs/math/9805084", "abstract": "Given a set D of positive integers, the associated distance graph on the integers is the graph with the integers as vertices and an edge between distinct vertices if their difference lies in D. We investigate the chromatic numbers of distance graphs. We show that, if $D = {d_1,d_2,d_3,...}$, with $d_n | d_{n+1}$ for all n, then the distance graph has a proper 4-coloring. We further find the exact chromatic numbers of all such distance graphs. Next, we characterize those distance graphs that have periodic proper colorings and show a relationship between the chromatic number and the existence of periodic proper colorings.", "subjects": "Combinatorics (math.CO)", "title": "Coloring Distance Graphs on the Integers", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.987375050346933, "lm_q2_score": 0.8104789155369047, "lm_q1q2_score": 0.8002466600333789 }
https://arxiv.org/abs/1810.02281
A Convergence Analysis of Gradient Descent for Deep Linear Neural Networks
We analyze speed of convergence to global optimum for gradient descent training a deep linear neural network (parameterized as $x \mapsto W_N W_{N-1} \cdots W_1 x$) by minimizing the $\ell_2$ loss over whitened data. Convergence at a linear rate is guaranteed when the following hold: (i) dimensions of hidden layers are at least the minimum of the input and output dimensions; (ii) weight matrices at initialization are approximately balanced; and (iii) the initial loss is smaller than the loss of any rank-deficient solution. The assumptions on initialization (conditions (ii) and (iii)) are necessary, in the sense that violating any one of them may lead to convergence failure. Moreover, in the important case of output dimension 1, i.e. scalar regression, they are met, and thus convergence to global optimum holds, with constant probability under a random initialization scheme. Our results significantly extend previous analyses, e.g., of deep linear residual networks (Bartlett et al., 2018).
\section{Approximate Balancedness and Deficiency Margin Under Customary Initialization} \label{app:balance_margin_init} Two assumptions concerning initialization facilitate our main convergence result (Theorem~\ref{theorem:converge}): \emph{(i)}~the initial weights $W_1(0),\ldots,W_N(0)$ are approximately balanced (see Definition~\ref{def:balance}); and \emph{(ii)}~the initial end-to-end matrix~$W_{1:N}(0)$ has positive deficiency margin with respect to the target~$\Phi$ (see Definition~\ref{def:margin}). The current appendix studies the likelihood of these assumptions being met under customary initialization of random (layer-wise independent) Gaussian perturbations centered at zero. For approximate balancedness we have the following claim, which shows that it becomes more and more likely the smaller the standard deviation of initialization is: \vspace{1mm} \begin{claim} \label{claim:balance_init} Assume all entries in the matrices $W_j\in\R^{d_j\times{d}_{j-1}}$, $j=1,\ldots,N$, are drawn independently at random from a Gaussian distribution with mean zero and standard deviation~$s>0$. Then, for any~$\delta>0$, the probability of $W_1,\ldots,W_N$ being $\delta$-balanced is at least~$\max\{0,1-10\delta^{-2}Ns^4d_{max}^3\}$, where $d_{max}:=\max\{d_0,\ldots,d_N\}$. \end{claim} \vspace{-3mm} \begin{proof} See Appendix~\ref{app:proofs:balance_init}. \end{proof} In terms of deficiency margin, the claim below treats the case of a single output model (scalar regression), and shows that if the standard deviation of initialization is sufficiently small, with probability close to~$0.5$, a deficiency margin will be met. However, for this deficiency margin to meet a chosen threshold~$c$, the standard deviation need be sufficiently large. \vspace{1mm} \begin{claim} \label{claim:margin_init} There is a constant $C_1 > 0$ such that the following holds. Consider the case where~$d_N=1$, $d_0\geq20$,\note{ The requirement $d_0\geq20$ is purely technical, designed to simplify expressions in the claim. } and suppose all entries in the matrices $W_j\in\R^{d_j\times{d}_{j-1}}$, $j=1,\ldots,N$, are drawn independently at random from a Gaussian distribution with mean zero, whose standard deviation~$s>0$ is small with respect to the target, \ie~$s\leq \| \Phi\|_F^{1/N} \big/ (10^5 d_0^3d_1 \cdots d_{N-1} C_1)^{1/(2N)}$. Then, for any~$c$ with $0<c\leq \norm{\Phi}_{F} \big/ \big(10^{5}d_0^{3}C_1(C_{1}N)^{2N}\big)$, the probability of the end-to-end matrix~$W_{1:N}$ having deficiency margin~$c$ with respect to~$\Phi$ is at least~$0.49$ if:\,\footnote{ The probability $0.49$ can be increased to any $p < 1/2$ by increasing the constant $10^5$ in the upper bounds for~$s$ and~$c$. } \footnote{ It is not difficult to see that the latter threshold is never greater than the upper bound for~$s$, thus sought-after standard deviations always exist. } $$s\geq{c}^{1/(2N)}\cdot\big(C_{1}N\norm{\Phi}_F^{1/(2N)}/(d_1\cdots{d}_{N-1})^{1/(2N)}\big) \text{~.}$$ \end{claim} \vspace{-3mm} \begin{proof} See Appendix~\ref{app:proofs:margin_init}. \end{proof} \section{Convergence Failures} \label{app:fail} In this appendix we show that the assumptions on initialization facilitating our main convergence result (Theorem~\ref{theorem:converge})~---~approximate balancedness and deficiency margin~---~are both necessary, by demonstrating cases where violating each of them leads to convergence failure. This accords with widely observed empirical phenomena, by which successful optimization in deep learning crucially depends on careful initialization (\cf~\citet{sutskever2013importance}). Claim~\ref{claim:diverge_balance} below shows\note{ For simplicity of presentation, the claim treats the case of even depth and uniform dimension across all layers. It can easily be extended to account for arbitrary depth and input/output/hidden dimensions. } that if one omits from Theorem~\ref{theorem:converge} the assumption of approximate balancedness at initialization, no choice of learning rate can guarantee convergence: \vspace{1mm} \begin{claim} \label{claim:diverge_balance} Assume gradient descent with some learning rate~$\eta>0$ is a applied to a network whose depth~$N$ is even, and whose input, output and hidden dimensions $d_0,\ldots,d_N$ are all equal to some $d\in\N$. Then, there exist target matrices~$\Phi$ such that the following holds. For any~$c$ with $0<c< \sigma_{min}(\Phi)$, there are initializations for which the end-to-end matrix~$W_{1:N}(0)$ has deficiency margin~$c$ with respect to~$\Phi$, and yet convergence will fail~---~objective will never go beneath a positive constant. \end{claim} \vspace{-3mm} \begin{proof} See Appendix~\ref{app:proofs:diverge_balance}. \end{proof} In terms of deficiency margin, we provide (by adapting Theorem~4 in~\citet{bartlett2018gradient}) a different, somewhat stronger result~---~there exist settings where initialization violates the assumption of deficiency margin, and despite being perfectly balanced, leads to convergence failure, for any choice of learning rate:\note{ This statement becomes trivial if one allows initialization at a suboptimal stationary point, \eg~$W_j(0)=0,~j=1,\ldots,N$. Claim~\ref{claim:diverge_margin} rules out such trivialities by considering only non-stationary initializations. } \vspace{1mm} \begin{claim} \label{claim:diverge_margin} Consider a network whose depth~$N$ is even, and whose input, output and hidden dimensions $d_0,\ldots,d_N$ are all equal to some $d\in\N$. Then, there exist target matrices~$\Phi$ for which there are non-stationary initializations $W_1(0),\ldots,W_N(0)$ that are $0$-balanced, and yet lead gradient descent, under any learning rate, to fail~---~objective will never go beneath a positive constant. \end{claim} \vspace{-3mm} \begin{proof} See Appendix~\ref{app:proofs:diverge_margin}. \end{proof} \section{Implementation Details} \label{app:impl} Below we provide implementation details omitted from our experimental report (Section~\ref{sec:exper}). The platform used for running the experiments is PyTorch \citep{paszke2017automatic}. For compliance with our analysis, we applied PCA whitening to the numeric regression dataset from UCI Machine Learning Repository. That is, all instances in the dataset were preprocessed by an affine operator that ensured zero mean and identity covariance matrix. Subsequently, we rescaled labels such that the uncentered cross-covariance matrix~$\Lambda_{yx}$ (see Section~\ref{sec:prelim}) has unit Frobenius norm (this has no effect on optimization other than calibrating learning rate and standard deviation of initialization to their conventional ranges). With the training objective taking the form of Equation~\eqref{eq:loss_whitened}, we then computed~$c$~---~the global optimum~---~in accordance with the formula derived in Appendix~\ref{app:whitened}. In our experiments with linear neural networks, balanced initialization was implemented with the assignment written in step~\emph{(iii)} of Procedure~\ref{proc:balance_init}. In the non-linear network experiment, we added, for each $j\in\{1,\ldots,N-1\}$, a random orthogonal matrix to the right of~$W_j$, and its transpose to the left of~$W_{j+1}$~---~this assignment maintains the properties required from balanced initialization (see Footnote~\ref{note:balance_init_props}). During all experiments, whenever we applied grid search over learning rate, values between $10^{-4}$ and~$1$ (in regular logarithmic intervals) were tried. \section{Deferred Proofs} \label{app:proofs} We introduce some additional notation here in addition to the notation specified in Section~\ref{sec:prelim}. We use~$\norm{A}_\sigma$ to denote the spectral norm (largest singular value) of a matrix~$A$, and sometimes~$\norm{\vv}_2$ as an alternative to~$\norm{\vv}$~---~the Euclidean norm of a vector~$\vv$. Recall that for a matrix $A$, $vec(A)$ is its vectorization in column-first order. We let $F(\cdot)$ denote the cumulative distribution function of the standard normal distribution, \ie~$F(x) = \int_{-\infty}^x\frac{1}{\sqrt{2\pi}}e^{-\frac12u^2}du$ ($x\in\R$). To simplify the presentation we will oftentimes use~$W$ as an alternative (shortened) notation for~$W_{1:N}$~---~the end-to-end matrix of a linear neural network. We will also use~$L(\cdot)$ as shorthand for~$L^1(\cdot)$~---~the loss associated with a (directly parameterized) linear model, \ie~$L(W) := \frac12\norm{W-\Phi}_F^2$. Therefore, in the context of gradient descent training a linear neural network, the following expressions all represent the loss at iteration $t$: \[ \ell(t)=L^N(W_1(t),\ldots,W_N(t))=L^1(W_{1:N}(t)) = L^1(W(t)) = L(W(t)) = \frac12\norm{W(t)-\Phi}_F^2\,. \] Also, for weights $W_j\in\R^{d_j\times{d}_{j-1}}$, $j=1,\ldots,{N}$ of a linear neural network, we generalize the notation~$W_{1:N}$, and define $W_{j:j'}:=W_{j'}W_{j'-1}\cdots{W}_j$ for every $1\leq{j}\leq{j'}\leq{N}$. Note that $W_{j:j'}^{\top}=W_{j}^{\top}W_{j+1}^{\top}\cdots{W}_{j'}^{\top}$. Then, by a simple gradient calculation, the gradient descent updates (\ref{eq:gd}) can be written as \begin{equation} \label{eq:wjupdate} W_j(t+1) = W_j(t) - \eta W_{j+1:N}^\top(t) \cdot \frac{dL}{dW}(W(t)) \cdot W_{1:j-1}^\top(t) \quad , 1\le j \le N\,, \end{equation} where we define $W_{1:0}(t) := I_{d_0}$ and $W_{N+1:N}(t) := I_{d_N}$ for completeness. Finally, recall the standard definition of the tensor product of two matrices (also known as the Kronecker product): for matrices $A \in \BR^{m_A \times n_A}, B \in \BR^{m_B \times n_B}$, their tensor product $A \otimes B \in \BR^{m_A m_B \times n_A n_B}$ is defined as $$ A \otimes B = \left ( \begin{matrix} a_{1,1} B& \cdots &a_{1,n_A} B \\ \vdots & \ddots & \vdots \\ a_{m_A,1} B & \cdots & a_{m_A,n_A} B \end{matrix} \right), $$ where $a_{i,j}$ is the element in the $i$-th row and $j$-th column of $A$. \subsection{Proof of Claim~\ref{claim:margin_interp}} \label{app:proofs:margin_interp} \begin{proof} Recall that for any matrices $A$ and~$B$ of compatible sizes $\sigma_{min}(A+B)\geq\sigma_{min}(A)-\sigma_{max}(B)$, and that the Frobenius norm of a matrix is always lower bounded by its largest singular value (\citet{horn1990matrix}). Using these facts, we have: \beas &&\sigma_{min}(W')=\sigma_{min}\big(\Phi+(W'-\Phi)\big)\geq\sigma_{min}(\Phi)-\sigma_{max}(W'-\Phi) \\[0.5mm] &&~\qquad\qquad\geq\sigma_{min}(\Phi)-\norm{W'-\Phi}_F\geq\sigma_{min}(\Phi)-\norm{W-\Phi}_F \\[1mm] &&~\qquad\qquad\geq\sigma_{min}(\Phi)-(\sigma_{min}(\Phi)-c)=c\,. \eeas \end{proof} \subsection{Proof of Lemma~\ref{lemma:descent}} \label{app:proofs:descent} To prove Lemma \ref{lemma:descent}, we will in fact prove a stronger result, Lemma \ref{lem:remainsmooth} below, which states that for each iteration $t$, in addition to (\ref{eq:descent}) being satisfied, certain other properties are also satisfied, namely: \emph{(i)} the weight matrices $W_1(t), \ldots, W_N(t)$ are $2\delta$-balanced, and \emph{(ii)} $W_1(t), \ldots, W_N(t)$ have bounded spectral norms. \begin{lemma} \label{lem:remainsmooth} Suppose the conditions of Theorem \ref{theorem:converge} are satisfied. Then for all $t \in \BN \cup \{0\}$, \begin{enumerate} \item[($\MA(t)$)] For $1 \leq j \leq N-1$, $\| W_{j+1}^\top(t) W_{j+1}(t) - W_j(t) W_j^\top(t) \|_F \leq 2\delta$. \item[($\MA'(t)$)] If $t \geq 1$, then for $1 \leq j \leq N-1$, \begin{eqnarray} \label{eq:sumbalancet} &&\| W_{j+1}^\top(t) W_{j+1}(t) - W_j(t) W_j^\top(t) \|_F \nonumber\\ &\leq& \| W_{j+1}^\top(t-1) W_{j+1}(t-1) - W_j(t-1) W_j^\top(t-1) \|_F \nonumber \\ && + \eta^2 \left\| \frac{dL^1}{dW} W(t-1)\right\|_F \cdot \left\| \frac{dL^1}{dW} W(t-1) \right\|_\sigma \cdot 4 \cdot (2 \| \Phi\|_F)^{2(N-1)/N}\nonumber. \end{eqnarray} \item[($\MB(t)$)] If $t = 0$, then $\ell(t) \leq \frac 12 \| \Phi\|_F^2$. If $t \geq 1$, then \begin{equation*} \ell(t) \leq \ell(t-1) -\frac \eta 2 \sigma_{min}(W(t-1))^{\frac{2(N-1)}{N}} \left\| \frac{dL^1}{dW} (W(t-1)) \right\|_F^2.\quad \end{equation*} \item[($\MC(t)$)] For $1 \leq j \leq N$, $\| W_j(t) \|_\sigma \leq (4 \| \Phi\|_F)^{1/N}$. \end{enumerate} \end{lemma} First we observe that Lemma \ref{lemma:descent} is an immediate consequence of Lemma \ref{lem:remainsmooth}. \begin{proof}[Proof of Lemma \ref{lemma:descent}] Notice that condition $\MB(t)$ of Lemma \ref{lem:remainsmooth} for each $t \geq 1$ immediately establishes the conclusion of Lemma \ref{lemma:descent} at time step $t-1$. \end{proof} \subsubsection{Preliminary lemmas} We next prove some preliminary lemmas which will aid us in the proof of Lemma \ref{lem:remainsmooth}. The first is a matrix inequality that follows from Lidskii's theorem. For a matrix $A$, let $\Sing(A)$ denote the rectangular diagonal matrix {of the same size,} whose diagonal elements are the singular values of $A$ arranged in non-increasing order (starting from the $(1,1)$ position). \begin{lemma}[\cite{bhatia_matrix_1997}, Exercise IV.3.5] \label{lemma:bhatiaexc} For any two matrices $A,B$ of the same size, $\| \Sing(A) - \Sing(B) \|_\sigma \leq \| A-B \|_\sigma$ and $\| \Sing(A) - \Sing(B) \|_F \leq \| A-B\|_F$. \end{lemma} Using Lemma \ref{lemma:bhatiaexc}, we get: \begin{lemma} \label{lem:rearrange} Suppose $D_1, D_2 \in \BR^{d \times d}$ are non-negative diagonal matrices with non-increasing values along the diagonal and $O \in \BR^{d \times d}$ is an orthogonal matrix. Suppose that $\| D_1 - OD_2O^\top\|_F \leq \ep$, for some $\ep > 0$. Then: \begin{enumerate} \item $\| D_1 - OD_1O^\top \|_F \leq 2\ep$. \item $\| D_1 - D_2 \|_F \leq \ep$. \end{enumerate} \end{lemma} \begin{proof} Since $D_1$ and $OD_2O^T$ are both symmetric positive semi-definite matrices, their singular values are equal to their eigenvalues. Moreover, the singular values of $D_1$ are simply its diagonal elements and the singular values of $OD_2O^T$ are simply the diagonal elements of $D_2$. Thus by Lemma \ref{lemma:bhatiaexc} we get that $\|D_1 - D_2 \|_F \leq \| D_1 - OD_2O^T\|_F \leq \ep$. Since the Frobenius norm is unitarily invariant, $\| D_1 - D_2 \|_F = \| OD_1O^T - OD_2O^T \|_F$, and by the triangle inequality it follows that \begin{align*} \| D_1 - OD_1O^T\|_F \leq \| OD_1O^T - OD_2O^T\|_F + \| D_1 - OD_2O^T\|_F \leq 2\ep. \end{align*} \end{proof} Lemma \ref{lem:boundcommute} below states that if $W_1, \ldots, W_N$ are approximately balanced matrices, \ie~$W_{j+1}^\top W_{j+1} - W_j W_j^\top$ has small Frobenius norm for $1 \leq j \leq N-1$, then we can bound the Frobenius distance between $W_{1:j}^\top W_{1:j}$ and $(W_1^\top W_1)^j$ (as well as between $W_{j:N}W_{j:N}^\top$ and $(W_NW_N^\top)^{N-j+1}$). \begin{lemma} \label{lem:boundcommute} Suppose that $d_N \leq d_{N-1}, d_0 \leq d_1$, and that for some $\nu > 0,M>0$, the matrices $W_j \in \BR^{d_j \times d_{j-1}}$, $1 \leq j \leq N$ satisfy, for $1 \leq j \leq N-1$, \begin{equation} \label{eq:pcndiff} \| W_{j+1}^\top W_{j+1} - W_j W_j^\top \|_F \leq \nu, \end{equation} and for $1 \leq j \leq N$, $\| W_j\|_\sigma \leq M$. Then, for $1 \leq j \leq N$, \begin{equation} \label{eq:wj1w} \| W_{1:j}^\top W_{1:j} - (W_1^\top W_1)^{j} \|_F \leq \frac 32 \nu \cdot M^{2(j-1)} j^2, \end{equation} and \begin{equation} \label{eq:wjnw} \| W_{j:N} W_{j:N}^\top - (W_NW_N^\top)^{N-j+1}\|_F \leq \frac 32 \nu \cdot M^{2(N-j)} (N-j+1)^2. \end{equation} Moreover, if $ \sigma_{min}$ denotes the minimum singular value of $W_{1:N}$, $\sigma_{1,min}$ denotes the minimum singular value of $W_1$ and $\sigma_{N,min}$ denotes the minimum singular value of $W_N$, then \begin{equation} \label{eq:singvallb} \sigma_{min}^2 - \frac 32 \nu M^{2(N-1)} N^2 \leq \begin{cases} \sigma_{N,min}^{2N} \quad : \quad d_N \geq d_0.\\ \sigma_{1,min}^{2N} \quad : \quad d_N \leq d_0. \end{cases} \end{equation} \end{lemma} \begin{proof} For $1 \leq j \leq N$, let us write the singular value decomposition of $W_j$ as $W_j = U_j \Sigma_j V_j^\top$, where $U_j\in \BR^{d_j\times d_j}$ and $V_j\in \BR^{d_{j-1}\times d_{j-1}}$ are orthogonal matrices and $\Sigma_j\in \BR^{d_j\times d_{j-1}}$ is diagonal. We may assume without loss of generality that the singular values of $W_j$ are non-increasing along the diagonal of $\Sigma_j$. Then we can write (\ref{eq:pcndiff}) as $$ \| V_{j+1}\Sigma_{j+1}^\top \Sigma_{j+1} V_{j+1}^\top - U_j \Sigma_j \Sigma_j^\top U_j^\top \|_F \leq \nu. $$ Since the Frobenius norm is invariant to orthogonal transformations, we get that $$ \| \Sigma_{j+1}^\top\Sigma_{j+1} - V_{j+1}^\top U_j \Sigma_j \Sigma_j^\top U_j^\top V_{j+1} \|_F \leq \nu. $$ By Lemma \ref{lem:rearrange}, we have that $\| \Sigma_{j+1}^\top \Sigma_{j+1} - \Sigma_j \Sigma_j^\top \|_F \leq \nu$ and $\| \Sigma_j \Sigma_j^\top - V_{j+1}^\top U_j \Sigma_j \Sigma_j^\top U_j^\top V_{j+1} \|_F \leq 2\nu$. We may rewrite the latter of these two inequalities as $$ \| [\Sigma_j \Sigma_j^\top ,V_{j+1}^\top U_j] \|_F = \| [\Sigma_j \Sigma_j^\top, V_{j+1}^\top U_j]U_j^\top V_{j+1} \|_F =\| \Sigma_j \Sigma_j^\top- V_{j+1}^\top U_j \Sigma_j \Sigma_j^\top U_j^\top V_{j+1} \|_F \leq 2\nu. $$ Note that $$ W_{j:N} W_{j:N}^\top = W_{j+1:N} U_j\Sigma_j\Sigma_j^\top U_j^\top W_{j+1:N}^\top. $$ For matrices $A,B$, we have that $\| AB\|_F \leq \| A \|_\sigma \cdot \| B \|_F$. Therefore, for $j+1 \leq i \leq N$, we have that \begin{eqnarray} && \| W_{i:N} U_{i-1} (\Sigma_{i-1} \Sigma_{i-1}^\top)^{i-j} U_{i-1}^\top W_{i:N}^\top - W_{i+1:N} U_{i} (\Sigma_{i}\Sigma_{i}^\top)^{i-j+1} U_{i}^\top W_{i+1:N}^\top\|_F \nonumber\\ & =& \| W_{i+1:N}U_{i} \left(\Sigma_{i}V_{i}^\top U_{i-1} (\Sigma_{i-1} \Sigma_{i-1}^\top)^{i-j} U_{i-1}^\top V_{i}\Sigma_{i}^\top - (\Sigma_{i}\Sigma_{i}^\top)^{i-j+1} \right) U_{i}^\top W_{i+1:N}^\top \|_F \nonumber\\ &\leq& \| W_{i+1:N} U_{i}\Sigma_{i}\|_\sigma^2 \cdot \| (\Sigma_{i-1} \Sigma_{i-1}^\top)^{i-j} + [V_{i}^\top U_{i-1}, (\Sigma_{i-1}\Sigma_{i-1}^\top)^{i-j}]U_{i-1}^\top V_{i} - (\Sigma_{i}^\top\Sigma_{i})^{i-j}\|_F\nonumber\\ & \leq & \| W_{i:N}\|_\sigma^2 \left( \| [V_{i}^\top U_{i-1}, (\Sigma_{i-1} \Sigma_{i-1}^\top)^{i-j}]\|_F + \| (\Sigma_{i-1}\Sigma_{i-1}^\top)^{i-j} - (\Sigma_{i}^\top \Sigma_{i})^{i-j}\|_F\right)\nonumber. \end{eqnarray} Next, we have that \begin{eqnarray} \| [V_{i}^\top U_{i-1}, (\Sigma_{i-1} \Sigma_{i-1}^\top)^{i-j}]\|_F & \leq & \sum_{k=0}^{i-j-1} \| (\Sigma_{i-1} \Sigma_{i-1}^\top)^k [V_i^\top U_{i-1}, \Sigma_{i-1} \Sigma_{i-1}^\top]( \Sigma_{i-1} \Sigma_{i-1}^\top)^{i-j-1-k}\|_F \nonumber\\ & \leq & \sum_{k=0}^{i-j-1} \| (\Sigma_{i-1} \Sigma_{i-1}^\top)^{i-j-1}\|_\sigma \cdot \| [V_i^\top U_{i-1} ,\Sigma_{i-1} \Sigma_{i-1}^\top]\|_F\nonumber\\ & \leq & (i-j) \| W_{i-1} \|_\sigma^{2(i-j-1)} \cdot 2\nu\nonumber. \end{eqnarray} We now argue that $\| (\Sigma_{i-1}\Sigma_{i-1}^\top)^k - (\Sigma_i^\top \Sigma_i)^k\|_F \leq \nu \cdot kM^{2(k-1)}$. Note that $\| \Sigma_{i-1} \Sigma_{i-1}^\top - \Sigma_i^\top \Sigma_i\|_F \leq \nu$, verifying the case $k=1$. To see the general case, since square diagonal matrices commute, we have that \begin{eqnarray} \| (\Sigma_{i-1}\Sigma_{i-1}^\top)^k - (\Sigma_i^\top \Sigma_i)^k\|_F&=& \left\| (\Sigma_{i-1}\Sigma_{i-1}^\top - \Sigma_i^\top \Sigma_i)\cdot \left( \sum_{\ell=0}^{k-1} (\Sigma_{i-1}\Sigma_{i-1}^\top)^\ell (\Sigma_i^\top \Sigma_i)^{k-1-\ell}\right)\right\|_F\nonumber\\ & \leq & \nu \cdot \sum_{\ell=0}^{k-1} \| W_{i-1}\|_\sigma^{2\ell} \cdot \| W_i\|_\sigma^{2(k-\ell-1)}\nonumber\\ & \leq & \nu k M^{2(k-1)}\nonumber. \end{eqnarray} It then follows that \begin{eqnarray} && \| W_{i:N} U_{i-1} (\Sigma_{i-1} \Sigma_{i-1}^\top)^{i-j} U_{i-1}^\top W_{i:N}^\top - W_{i+1:N} U_{i} (\Sigma_{i}\Sigma_{i}^\top)^{i-j+1} U_{i}^\top W_{i+1:N}^\top\|_F \nonumber\\ & \leq & \| W_{i:N}\|_\sigma^2 \cdot \left((i-j) M^{2(i-j-1)} \cdot 2\nu + \nu(i-j) M^{2(i-j-1)}\right)\nonumber\\ & = & \| W_{i:N}\|_\sigma^2 \cdot 3\nu (i-j) M^{2(i-j-1)}\nonumber. \end{eqnarray} By the triangle inequality, we then have that \begin{eqnarray} && \| W_{j:N}W_{j:N}^\top - U_N (\Sigma_N \Sigma_N^\top)^{N-j+1} U_N^\top\|_F\nonumber\\ & \leq & \nu \sum_{i=j+1}^N \| W_{i:N}\|_\sigma^2 \cdot 3(i-j) M^{2(i-j-1)}\nonumber\\ & \leq & 3\nu \sum_{i=j+1}^N (i-j) M^{2(N-i+1)} M^{2(i-j-1)}\nonumber\\ \label{eq:diffN} &=& 3\nu M^{2(N-j)} \sum_{i=j+1}^N (i-j) \leq \frac 32 \nu \cdot M^{2(N-j)} \cdot (N-j+1)^2. \end{eqnarray} By an identical argument (formally, by replacing $W_j$ with $W_{N-j+1}^\top$), we get that \begin{equation} \label{eq:diff1} || W_{1:j}^\top W_{1:j} - V_1 (\Sigma_1^\top \Sigma_1)^{j} V_1^\top\|_F \leq \frac 32\nu \cdot M^{2(j-1)} \cdot j^2. \end{equation} (\ref{eq:diffN}) and (\ref{eq:diff1}) verify (\ref{eq:wjnw}) and (\ref{eq:wj1w}), respectively, so it only remains to verify (\ref{eq:singvallb}). Letting $j=1$ in (\ref{eq:diffN}), we get \begin{equation} \label{eq:w1nundiff} \| W_{1:N}W_{1:N}^\top - U_N(\Sigma_N \Sigma_N^\top)^{N} U_N^\top\|_F \leq \frac 32\nu \cdot M^{2(N-1)} \cdot N^2. \end{equation} Let us write the eigendecomposition of $W_{1:N}W_{1:N}^\top$ with an orthogonal eigenbasis as $W_{1:N}W_{1:N}^\top = U \Sigma U^\top$, where $\Sigma$ is diagonal with its (non-negative) elements arranged in non-increasing order and $U$ is orthogonal. We can write the left hand side of (\ref{eq:w1nundiff}) as $\| U \Sigma U^\top - U_N(\Sigma_N \Sigma_N^\top)^N U_N^\top\|_F = \| \Sigma - U^\top U_N (\Sigma_N \Sigma_N^\top)^N U_N^\top U\|_F$. By Lemma \ref{lem:rearrange}, we have that \begin{equation} \label{eq:ntildediff} \| \Sigma - (\Sigma_N\Sigma_N^\top)^N \|_F \leq \frac 32 \nu M^{2(N-1)} N^2. \end{equation} Recall that $W \in \BR^{d_N \times d_0}$. Suppose first that $d_N \leq d_0$. Let $\sigma_{min}$ denote the minimum singular value of $W_{1:N}$ (so that $\sigma_{min}^2$ is the element in the $(d_N, d_N)$ position of $ \Sigma \in \BR^{d_N \times d_N}$), and $\sigma_{N,min}$ denote the minimum singular value (i.e. diagonal element) of $\Sigma_N$, which lies in the $(d_N, d_N)$ position of $\Sigma_N$. (Note that the $(d_N, d_N)$ position of $\Sigma_N \in \BR^{d_N \times d_{N-1}}$ exists since $d_{N-1} \geq d_N$ by assumption.) Then $$ (\sigma_{N,min}^{2N} - \sigma_{min}^2)^2 \leq \left(\frac 32 \nu M^{2(N-1)} N^2\right)^2, $$ so $$ \sigma_{N,min}^{2N} \geq \sigma_{min}^2 - {\frac 32 \nu M^{2(N-1)} N^2}. $$ By an identical argument using (\ref{eq:diff1}), we get that, in the case that $d_0 \leq d_N$, if $\sigma_{1,min}$ denotes the minimum singular value of $\Sigma_1$, then $$ \sigma_{1,min}^{2N} \geq \sigma_{min}^2 - \frac{3}{2} \nu M^{2(N-1)} N^2. $$ (Notice that we have used the fact that the nonzero eigenvalues of $W_{1:N}W_{1:N}^\top$ are the same as the nonzero eigenvalues of $W_{1:N}^\top W_{1:N}$.) This completes the proof of (\ref{eq:singvallb}). \end{proof} Using Lemma \ref{lem:boundcommute}, we next show in Lemma \ref{lem:boundindiv} that if {$W_1,\ldots,W_N$ are approximately balanced,} then an upper bound on $\| W_N \cdots W_1\|_\sigma$ implies an upper bound on $\| W_j\|_\sigma$ for $1 \leq j \leq N$. \begin{lemma} \label{lem:boundindiv} Suppose $\nu, C$ are real numbers satisfying $C > 0$ and $0 < \nu \leq \frac{ C^{2/N}}{30 N^2}$. Moreover suppose that the matrices $W_1, \ldots, W_N$ satisfy the following: \begin{enumerate} \item For $1 \leq j \leq N-1$, $\| W_{j+1}^\top W_{j+1} - W_j W_j^\top \|_F \leq \nu$. \item $\| W_N \cdots W_1 \|_\sigma \leq C$. \end{enumerate} Then for $1 \leq j \leq N$, $\| W_j\|_\sigma \leq C^{1/N} \cdot 2^{1/(2N)}$. \end{lemma} \begin{proof} For $1 \leq j \leq N$, let us write the singular value decomposition of $W_j$ as $W_j = U_j \Sigma_j V_j^\top$, where the singular values of $W_j$ are decreasing along the main diagonal of $\Sigma_j$. By Lemma \ref{lem:rearrange}, we have that for $1 \leq j \leq N-1$, $\| \Sigma_{j+1}^\top \Sigma_{j+1} - \Sigma_j \Sigma_j^\top \|_F \leq \nu$, which implies that $\left|\| \Sigma_{j+1}^\top \Sigma_{j+1}\|_\sigma -\| \Sigma_j \Sigma_j^\top \|_\sigma \right| \leq \nu$. Write $M = \max_{1 \leq j \leq N} \| W_j\|_\sigma = \max_{1 \leq j \leq N} \| \Sigma_j \|_\sigma$. By the above we have that $\| \Sigma_j \Sigma_j^\top\|_\sigma \geq M^2 - N\nu$ for $1 \leq j \leq N$. Let the singular value decomposition of $W_{1:N}$ be denoted by $W_{1:N} = U\Sigma V^\top$, so that $\| \Sigma \|_\sigma \leq C$. Then by (\ref{eq:wjnw}) of Lemma \ref{lem:boundcommute} and Lemma \ref{lem:rearrange} (see also (\ref{eq:ntildediff}), where the same argument was used), we have that $$ \| \Sigma\Sigma^\top - (\Sigma_N \Sigma_N^\top)^N \|_F\leq \frac 32 \nu M^{2(N-1)}N^2. $$ Then \begin{equation} \label{eq:sigman1} \| (\Sigma_N \Sigma_N^\top)^N\|_\sigma \leq \|\Sigma\Sigma^\top\|_\sigma + \frac 32 \nu M^{(2(N-1))}N^2 \leq \| \Sigma \Sigma^\top \|_\sigma + \frac 32 \nu \left( \| \Sigma_N \Sigma_N^\top\|_\sigma + \nu N\right)^{N-1} N^2. \end{equation} Now recall that $\nu$ is chosen so that $\nu \leq \frac{ C^{2/N}}{30 \cdot N^2}.$ Suppose for the purpose of contradiction that there is some $j$ such that $\| W_jW_j^\top \|_\sigma > 2^{1/N} C^{2/N}$. Then it must be the case that \begin{equation} \label{eq:2nabove} \| \Sigma_N \Sigma_N^\top\|_\sigma > 2^{1/N} C^{2/N} - \nu \cdot N \geq (5/4)^{1/N} C^{2/N} > \nu \cdot 30N^2, \end{equation} where we have used that $$ 2^{1/N} - (5/4)^{1/N} \geq \frac{1}{30N} $$ for all $N \geq 2$, which follows by considering the Laurent series $\exp(1/z) = \sum_{i=1}^\infty \frac{1}{i! z^i}$, which converges in $|z| > 0$ for $z \in \BC$. We now rewrite inequality (\ref{eq:2nabove}) as \begin{equation} \label{eq:boundepcontr} \nu \leq \frac{\| \Sigma_N \Sigma_N^\top\|_\sigma}{30N^2}. \end{equation} Next, using (\ref{eq:boundepcontr}) and $(1+1/x)^x \leq e$ for all $x > 0$, \begin{equation} \label{eq:sigman2} \frac 32 \nu \left( \| \Sigma_N \Sigma_N^\top\|_\sigma + \nu N\right)^{N-1} N^2 \leq \frac{e^{1/30}}{20} \cdot \| \Sigma_N \Sigma_N^\top \|_\sigma^{N} < \frac{e}{20} \cdot \| \Sigma_N \Sigma_N^\top \|_\sigma^{N}. \end{equation} Since $\| (\Sigma_N \Sigma_N^\top)^N \|_\sigma = \| \Sigma_N \Sigma_N^\top\|_\sigma^N$, we get by combining (\ref{eq:sigman1}) and (\ref{eq:sigman2}) that $$ \| \Sigma_N \Sigma_N^\top \|_\sigma < (1-e/20)^{-1/N} \cdot \| \Sigma \Sigma^\top\|_\sigma^{1/N} \leq (1-e/20)^{-1/N} \cdot C^{2/N}, $$ and since $1-e/20 > 1/(5/4)$, it follows that $\| \Sigma_N \Sigma_N^\top\|_\sigma < (5/4)^{1/N} C^{2/N}$, which contradicts (\ref{eq:2nabove}). It follows that for all $1 \leq j \leq N$, $\| W_j W_j^\top \|_\sigma \leq 2^{1/N} C^{2/N}$. The conclusion of the lemma then follows from the fact that $\| W_j W_j^\top\|_\sigma = \| W_j\|_\sigma^2$. \end{proof} \subsubsection{Single-Step Descent} Lemma~\ref{lem:boundgdupdate} below states that if certain conditions on $W_1(t),\ldots,W_N(t)$ are met, the sought-after descent~---~Equation~\eqref{eq:descent}~---~will take place at iteration~$t$. We will later show (by induction) that the required conditions indeed hold for every~$t$, thus the descent persists throughout optimization. The proof of Lemma~\ref{lem:boundgdupdate} is essentially a discrete, single-step analogue of the continuous proof for Lemma~\ref{lemma:descent} (covering the case of gradient flow) given in Section~\ref{sec:converge}. \begin{lemma} \label{lem:boundgdupdate} Assume the conditions of Theorem~\ref{theorem:converge}. Moreover, suppose that for some $t$, the matrices $W_1(t), \ldots, W_N(t)$ and the end-to-end matrix $W(t) := W_{1:N}(t)$ satisfy the following properties: \begin{enumerate} \item $\| W_j(t)\|_\sigma \leq (4 \| \Phi\|_F)^{1/N}$ for $1 \leq j \leq N$. \item $\| W(t) - \Phi\|_\sigma \leq \| \Phi\|_F$. \item $\| W_{j+1}^\top(t) W_{j+1}(t) - W_j(t)W_j^\top(t)\|_F \leq 2\delta$ for $1 \leq j \leq N-1$. \item ${\sigma}_{min}:={\sigma}_{min}(W(t)) \geq c$. \end{enumerate} Then, after applying a gradient descent update (\ref{eq:gd}) we have that $$ L(W(t+1)) - L(W(t)) \leq - \frac \eta 2 \sigma_{min}^{2(N-1)/N} \left\| \frac{dL}{dW}(W(t)) \right\|_F^2. $$ \end{lemma} \begin{proof} For simplicity write $M = (4 \| \Phi\|_F)^{1/N}$ and $B = \| \Phi\|_F$. We first claim that {\small\begin{equation} \label{eq:assumeeta} \eta \leq \min \left\{ \frac{1}{2M^{N-2}BN}, \frac{\sigma_{min}^{2(N-1)/N}}{24 \cdot 2 M^{3N-4} N^2B}, \frac{\sigma_{min}^{2(N-1)/N}}{24N^2 M^{4(N-1)}}, \frac{\sigma_{min}^{2(N-1)/(3N)}}{ \left(24 \cdot 4M^{6N-8}N^4B^2 \right)^{1/3}}\right\}. \end{equation}} Since $c \leq \sigma_{min}$, for (\ref{eq:assumeeta}) to hold it suffices to have \begin{eqnarray} \eta &\leq& \min \left\{ \frac{1}{8 \| \Phi\|_F^{(2N-2)/N} N}, \frac{c^{2(N-1)/N}}{3 \cdot 2^{11} \| \Phi\|_F^{4(N-1)/N} N^2},\frac{c^{2(N-1)/(3N)}}{3 \cdot 2^6 \left(\| \Phi\|_F^{(8N-8)/N}\right)^{1/3} N^{4/3}}\right\}\nonumber. \end{eqnarray} As the minimum singular value of $\Phi$ must be at least $c$, we must have $c \leq \| \Phi\|_\sigma$. Since then $\frac{c}{\| \Phi\|_F} \leq \frac{c}{\| \Phi\|_\sigma} \leq 1$, it holds that $$ \frac{c^{2(N-1)/N}}{\| \Phi\|_F^{4(N-1)/N}} \leq \min \left\{ \frac{1}{\| \Phi\|_F^{2(N-1)/N}}, \frac{c^{2(N-1)/(3N)}}{\| \Phi\|_F^{(8N-8)/(3N)}} \right\}, $$ meaning that it suffices to have $$ \eta \leq \frac{ c^{2(N-1)/N}}{3 \cdot 2^{11} N^2 \| \Phi\|_F^{4(N-1)/N}}, $$ which is guaranteed by (\ref{eq:descent_eta}). Next, we claim that \begin{eqnarray} \label{eq:assumeep} 2\delta &\leq &\min \left\{ \frac{ c^{2(N-1)/N}}{8 \cdot 2^4N^3 \| \Phi\|_F^{2(N-2)/N}}, \frac{c^{2}}{6 \cdot 2^4 N^2 \| \Phi\|_F^{2(N-1)/N}}\right\}\\ & \leq & \min \left\{ \frac{ \sigma_{min}^{2(N-1)/N}}{8N^3M^{2(N-2)}}, \frac{\sigma_{min}^{2}}{6N^2 M^{2(N-1)}}\right\} .\nonumber \end{eqnarray} The second inequality above is trivial, and for the first to hold, since $c \le \| \Phi\|_F$, it suffices to take $$ 2\delta \leq \frac{c^2}{128 \cdot N^3 \cdot \| \Phi\|_F^{2(N-1)/N}}, $$ which is guaranteed by the definition of $\delta$ in Theorem \ref{theorem:converge}. Next we continue with the rest of the proof. It follows from (\ref{eq:wjupdate}) that\footnote{Here, for matrices $A_1, \ldots, A_K$ such that $A_K A_{K-1} \cdots A_1$ is defined, we write $\prod_{1}^{j=K} A_j := A_K A_{K-1} \cdots A_1$. } \begin{eqnarray} && W(t+1) - W(t) \nonumber\\ &=& \prod_{1}^{j=N} \left( W_j(t) - \eta W_{j+1:N}^\top(t) \frac{dL}{dW}(W(t)) W_{1:j-1}^\top(t) \right) - W_{1:N}(t)\nonumber\\ \label{eq:wt1wt} &=& -\eta \left(\sum_{j=1}^N W_{j+1:N}W_{j+1:N}^\top(t) \frac{dL}{dW}(W(t)) W_{1:j-1}^\top(t) W_{1:j-1}(t)\right) + (\star), \end{eqnarray} where $(\star)$ denotes higher order terms in $\eta$. We now bound the Frobenius norm of $(\star)$. To do this, note that since $L(W) = \frac 12 \| W - \Phi\|_F^2$, $\frac{dL}{dW}(W(t)) = W(t) - \Phi$. Then \begin{eqnarray} \| (\star) \|_F & \leq & \sum_{k=2}^N \eta^k \cdot M^{k(N-1) + N-k} \cdot \left\| \frac{dL}{dW}(W(t)) \right\|_F \cdot \left\| \frac{dL}{dW}(W(t)) \right\|_\sigma^{k-1} \cdot {N \choose k}\nonumber\\ & \leq & \eta M^{2N-2}N \left\| \frac{dL}{dW}(W(t)) \right\|_F \sum_{k=2}^N \left(\eta M^{N-2} BN\right)^{k-1}\nonumber\\ \label{eq:starnorm} & \leq & \eta \cdot (2\eta M^{3N-4} N^2B) \cdot \left\| \frac{dL}{dW}(W(t)) \right\|_F, \end{eqnarray} where the last inequality uses $\eta M^{N-2} BN \leq 1/2$, which is a consequence of (\ref{eq:assumeeta}). Next, by Lemma \ref{lem:boundcommute} with $\nu = 2\delta$, {\small\begin{eqnarray} && \left\|\sum_{j=1}^N W_{j+1:N}W_{j+1:N}^\top(t) \frac{dL}{dW}(W(t)) W_{1:j-1}^\top(t) W_{1:j-1}(t) \right.\nonumber\\ &&- \left. \sum_{j=1}^N (W_NW_N^\top)^{N-j} \frac{dL}{dW}(W(t)) (W_1^\top W_1)^{j-1} \right\|_F\nonumber\\ & \leq & \left\|\sum_{j=1}^N (W_{j+1:N}W_{j+1:N}^\top(t) - (W_NW_N^\top)^{N-j}) \frac{dL}{dW}(W(t)) W_{1:j-1}^\top(t) W_{1:j-1}(t) \right\|_F \nonumber\\ & + & \left\| \sum_{j=1}^N (W_NW_N^\top)^{N-j} \frac{dL}{dW}(W(t)) (W_{1:j-1}^\top W_{1:j-1} - (W_1^\top W_1)^{j-1})\right\|_F\nonumber\\ & \leq & \left\| \frac{dL}{dW}(W(t)) \right\|_F \cdot \left(\sum_{j=1}^{N-1} \frac 32 2\delta \cdot M^{2(N-j)}(N-j)^2 M^{2(j-1)} + \sum_{j=2}^N \frac 32 2\delta \cdot M^{2(j-2)}(j-1)^2 M^{2(N-j)}\right)\nonumber\\ & \leq & \left\| \frac{dL}{dW}(W(t)) \right\|_F \cdot 2\delta N^3 M^{2(N-2)}\nonumber. \end{eqnarray}} Next, by standard properties of tensor product, we have that \begin{eqnarray} && vec\left(\sum_{j=1}^N (W_NW_N^\top)^{N-j} \frac{dL}{dW}(W(t)) (W_1^\top W_1)^{j-1} \right)\nonumber\\ &=& \sum_{j=1}^N \left( (W_1^\top W_1)^{j-1} \otimes (W_NW_N^\top)^{N-j} \right)vec \left( \frac{dL}{dW}(W(t))\right)\nonumber. \end{eqnarray} Let us write eigenvalue decompositions $W_1^\top W_1 = UDU^\top, W_NW_N^\top = VEV^\top$. Then \begin{eqnarray} && \sum_{j=1}^N \left( (W_1^\top W_1)^{j-1} \otimes (W_NW_N^\top)^{N-j} \right)\nonumber\\ &=& \sum_{j=1}^N \left( UD^{j-1}U^\top \otimes VE^{N-j} V^\top \right)\nonumber\\ &=& (U \otimes V) \left(\sum_{j=1}^N D^{j-1} \otimes E^{N-j}\right) (U \otimes V)^\top\nonumber\\ &=& O\Lambda O^\top\nonumber, \end{eqnarray} with $O = U \otimes V$, and $\Lambda = \sum_{j=1}^N D^{j-1} \otimes E^{N-j}$. As $W_1 \in \BR^{d_1 \times d_0}$, and $W_N \in \BR^{d_N \times d_{N-1}}$, then $D \in \BR^{d_0 \times d_0}, E \in \BR^{d_N \times d_N}$, so $\Lambda \in \BR^{d_0d_N \times d_0d_N}$. Moreover note that $\Lambda \succeq D^0 \otimes E^{N-1} + D^{N-1} \otimes E^0 = I_{d_0} \otimes E^{N-1} + D^{N-1} \otimes I_{d_N}$. If $\lambda_D$ denotes the minimum diagonal element of $D$ and $\lambda_E$ denotes the minimum diagonal element of $E$, then the minimum diagonal element of $\Lambda$ is therefore at least $\lambda_D^{N-1} + \lambda_E^{N-1}$. But, it follows from Lemma \ref{lem:boundcommute} (with $\nu = 2\delta$) that $$ \max\{ \lambda_D^{N}, \lambda_E^{N} \} \geq \sigma_{min}^2 - \frac 32 2\delta M^{2(N-1)} N^2 \geq 3\sigma_{min}^2/4, $$ where the second inequality follows from (\ref{eq:assumeep}). Hence the minimum diagonal element of $\Lambda$ is at least $( \sigma_{min}^2/(4/3))^{(N-1)/N} \geq \sigma_{min}^{2(N-1)/N}/(4/3)$. It follows as a result of the above inequalities that if we write $E(t) = vec(W(t+1)) - vec(W(t)) + \eta (O\Lambda O^\top) vec\left( \frac{dL}{dW}(W(t)) \right)$, then \begin{eqnarray} \| E(t) \|_2 & = &\left\| vec(W(t+1)) - vec(W(t)) + \eta (O\Lambda O^\top) vec\left( \frac{dL}{dW}(W(t)) \right) \right\|_2\nonumber\\ &\leq & \eta \left\| \frac{dL}{dW}(W(t)) \right\|_F \cdot (2\eta M^{3N-4} N^2B + 2\delta N^3 M^{2(N-2)}) \nonumber. \end{eqnarray} Then we have {\small\begin{eqnarray} && L(W(t+1)) - L(W(t)) \nonumber\\ & \leq & vec\left( \frac{d}{dW} L(W(t)) \right)^\top vec \left( W(t+1) - W(t) \right) + \frac 12 \| W(t+1) - W(t) \|_F^2\nonumber\\ & = & \eta \left( - vec\left(\frac{d}{dW} L(W(t))\right)^\top (O\Lambda O^\top) vec \left(\frac{d}{dW} L(W(t)) \right) + \frac{1}{\eta}vec\left(\frac{d}{dW} L(W(t))\right)^\top E(t) \right) \nonumber\\ && + \frac 12 \| W(t+1) - W(t) \|_F^2\nonumber\\ & \leq & \eta \left(-\left\| \frac{d}{dW} L(W(t)) \right\|_F^2 \cdot \frac{ \sigma_{min}^{2(N-1)/N}}{4/3} + \left\| \frac{d}{dW} L(W(t)) \right\|_F^2 \cdot \left(2\eta M^{3N-4} N^2B + 2\delta N^3 M^{2(N-2)}\right) \right) \nonumber\\ && + \frac 12 \| W(t+1) - W(t) \|_F^2\nonumber, \end{eqnarray}} where the first inequality follows since $L(W) = \frac 12 \| W - \Phi\|_F^2$ is $1$-smooth as a function of $W$. Next, by (\ref{eq:wt1wt}) and (\ref{eq:starnorm}), \begin{eqnarray} && \| W(t+1) - W(t) \|_F^2 \nonumber\\ & \leq & 2\eta^2 \cdot \left(N M^{2(N-1)} \cdot \left\| \frac{dL}{dW}(W(t)) \right\|_F\right)^2 + 2\eta^2 \cdot (2\eta M^{3N-4} N^2B)^2 \cdot \left\| \frac{dL}{dW}(W(t)) \right\|_F^2\nonumber\\ \label{eq:boundwtwt1} & = &2 \eta^2 \left\| \frac{dL}{dW}(W(t)) \right\|_F^2 \cdot \left( N^2M^{4(N-1)} + (4\eta^2 M^{6N-8} N^4B^2)\right). \end{eqnarray} Thus \begin{eqnarray} && L(W(t+1)) - L(W(t))\nonumber\\ & \leq & \eta \cdot \left\| \frac{dL}{dW}(W(t)) \right\|_F^2 \cdot \left( -\frac{\sigma_{min}^{2(N-1)/N}}{4/3} + 2\eta M^{3N-4} N^2B + 2\delta N^3 M^{2(N-2)} \right.\nonumber\\ && \left. + \eta \cdot (N^2M^{4(N-1)} + 4\eta^2 M^{6N-8}N^4B^2) \right)\nonumber. \end{eqnarray} By (\ref{eq:assumeeta}, \ref{eq:assumeep}), which bound $\eta, 2\delta$, respectively, we have that {\small\begin{eqnarray} \label{eq:descentl} &&L(W(t+1)) - L(W(t))\nonumber\\ & \leq& \eta \cdot \left\| \frac{dL}{dW}(W(t)) \right\|_F^2 \cdot \left( -\frac{\sigma_{min}^{2(N-1)/N}}{4/3} + \frac{\sigma_{min}^{2(N-1)/N}}{24} + \frac{\sigma_{min}^{2(N-1)/N}}{8} + \frac{\sigma_{min}^{2(N-1)/N}}{24} + \frac{\sigma_{min}^{2(N-1)/N}}{24} \right)\nonumber\\ &=& - \frac 12 \sigma_{min}^{2(N-1)/N} \eta \left\| \frac{dL}{dW}(W(t)) \right\|_F^2. \end{eqnarray}} \end{proof} \subsubsection{Proof of Lemma \ref{lem:remainsmooth}} \begin{proof}[Proof of Lemma \ref{lem:remainsmooth}] We use induction on $t$, beginning with the base case $t=0$. Since the weights $W_1(0), \ldots, W_N(0)$ are $\delta$-balanced, we get that $\MA(0)$ holds automatically. To establish $\MB(0)$, note that since $W_{1:N}(0)$ has deficiency margin $c > 0$ with respect to $\Phi$, we must have $\| W_{1:N}(0) - \Phi \|_F \leq \sigma_{min}(\Phi) \leq \|\Phi\|_F$, meaning that $L^1(W_{1:N}(0)) \leq \frac 12 \| \Phi\|_F^2$. Finally, by $\MB(0)$, which gives $\| W(0) - \Phi\|_F \leq \| \Phi\|_F$, we have that \begin{equation} \label{eq:boundprod0} \| W(0) \|_\sigma \leq \| W(0) \|_F \leq \| W(0) - \Phi\|_F + \| \Phi\|_F \leq 2 \| \Phi\|_F. \end{equation} To show that the above implies $\MC(0)$, we use condition $\MA(0)$ and Lemma \ref{lem:boundindiv} with $C = 2 \| \Phi\|_F$ and $\nu = 2\delta$. By the definition of $\delta$ in Theorem \ref{theorem:converge} and since $c \leq \| \Phi\|_F$, we have that \begin{equation} \label{eq:epltc} 2\delta \leq \frac{c^2}{128 \cdot N^3 \cdot \| \Phi\|_F^{2(N-1)/N}} = \frac{\| \Phi\|_F^{2/N}}{128N^3} \cdot \frac{c^2}{\| \Phi\|_F^2}< \frac{\| \Phi\|_F^{2/N}}{30 N^2}, \end{equation} as required by Lemma \ref{lem:boundindiv}. As $\MA(0)$ and (\ref{eq:boundprod0}) verify the preconditions 1.~and 2., respectively, of Lemma \ref{lem:boundindiv}, it follows that for $1 \leq j \leq N$, $\| W_j(t) \|_\sigma \leq (2 \| \Phi\|_F)^{1/N} \cdot 2^{1/(2N)} < (4 \| \Phi\|_F)^{1/N}$, verifying $\MC(0)$ and completing the proof of the base case. The proof of Lemma \ref{lem:remainsmooth} follows directly from the following inductive claims. \begin{enumerate} \item $\MA(t), \MB(t), \MC(t) \Rightarrow \MB(t+1)$. To prove this, we use Lemma \ref{lem:boundgdupdate}. We verify first that the preconditions hold. First, $\MC(t)$ immediately gives condition 1.~of Lemma \ref{lem:boundgdupdate}. By $\MB(t)$, we have that $\| W(t) - \Phi\|_\sigma \leq \| W(t) - \Phi\|_F \leq \| \Phi\|_F$, giving condition 2.~of Lemma \ref{lem:boundgdupdate}. $\MA(t)$ immediately gives condition 3.~of Lemma \ref{lem:boundgdupdate}. Finally, by $\MB(t)$, we have that $L^N(W_1(t), \ldots, W_N(t)) \leq L^N(W_1(0), \ldots, W_N(0))$, so $\sigma_{min}(W_{1:N}(t)) \geq c$ by Claim \ref{claim:margin_interp}. This verifies condition 4.~of Lemma \ref{lem:boundgdupdate}. Then Lemma \ref{lem:boundgdupdate} gives that $L^N(W_{1}(t+1), \ldots, W_N(t+1)) \leq L^N(W_1(t), \ldots, W_N(t)) - \frac 12 \sigma_{min}(W(t))^{2(N-1)/N} \eta \left\| \frac{dL}{dW}(W(t)) \right\|_F^2$, establishing $\MB(t+1)$. \item $\MA(0), \MA'(1), \ldots, \MA'(t), \MA(t), \MB(0), \ldots, \MB(t), \MC(t) \Rightarrow \MA(t+1), \MA'(t+1)$. To prove this, note that for $1 \leq j \leq N-1$, \begin{eqnarray} && W_{j+1}^\top(t+1) W_{j+1}(t+1) - W_j(t+1) W_j^\top(t+1)\nonumber\\ &=& \left( W_{j+1}^\top(t) - \eta W_{1:j}(t) \frac{dL}{dW}(W(t))^\top W_{j+2:N}(t) \right) \nonumber\\ && \cdot \left( W_{j+1}(t) - \eta W_{j+2:N}^\top(t) \frac{dL}{dW}(W(t)) W_{1:j}^\top(t) \right)\nonumber\\ &&- \left( W_j(t) - \eta W_{j+1:N}^\top(t) \frac{dL}{dW}(W(t)) W_{1:j-1}^\top(t) \right)\nonumber\\ && \cdot \left( W_j^\top(t) - \eta W_{1:j-1}(t) \frac{dL}{dW}(W(t))^\top W_{j+1:N}(t) \right)\nonumber. \end{eqnarray} By $\MB(0), \ldots, \MB(t)$, $\| W_{1:N}(t) - \Phi \|_F \leq \| \Phi\|_F$. By the triangle inequality it then follows that $\| W_{1:N}(t)\|_\sigma \leq 2 \| \Phi\|_F$. Also $\MA(t)$ gives that for $1 \leq j \leq N-1$, $\| W_{j}(t) W_j^\top(t) - W_{j+1}^\top(t) W_{j+1}(t) \|_F \leq 2\delta$. By Lemma \ref{lem:boundindiv} with $C = 2 \| \Phi\|_F, \nu = 2\delta$ (so that (\ref{eq:epltc}) is satisfied), \begin{eqnarray} && \left\| W_{j+1}^\top(t+1) W_{j+1}(t+1) - W_j(t+1) W_j^\top(t+1) \right\|_F\nonumber\\ & \leq & \| W_{j+1}^\top(t) W_{j+1}(t) - W_j(t) W_j^\top(t) \|_F + \eta^2 \left\| \frac{dL}{dW}(W(t)) \right\|_F \cdot \left\| \frac{dL}{dW}(W(t)) \right\|_\sigma \nonumber\\ && \cdot \left( \| W_{j+2:N}(t)\|_\sigma^2 \| W_{1:j}(t) \|_\sigma^2 + \| W_{1:j-1}\|_\sigma^2 \|W_{j+1:N} \|_\sigma^2\right)\nonumber\\ \label{eq:wtwt1diverge} & \leq & \| W_{j+1}^\top(t) W_{j+1}(t) - W_j(t) W_j^\top(t) \|_F \nonumber\\ && + 4\eta^2 \left\| \frac{dL}{dW}(W(t)) \right\|_F \left\| \frac{dL}{dW}(W(t)) \right\|_\sigma (2 \| \Phi\|_F)^{2(N-1)/N}. \end{eqnarray} In the first inequality above, we have also used the fact that for matrices $A, B$ such that $AB$ is defined, $\| AB\|_F \leq \| A \|_\sigma \| B \|_F$. (\ref{eq:wtwt1diverge}) gives us $\MA'(t+1)$. We next establish $\MA(t+1)$. By $\MB(i)$ for $0 \leq i \leq t$, we have that $\left\| \frac{dL}{dW}(W(i)) \right\|_F = \left\| W - \Phi \right\|_F \leq \| \Phi\|_F$. Using $\MA'(i)$ for $0 \leq i \leq t$ and summing over $i$ gives \begin{eqnarray} && \| W_{j+1}^\top(t+1) W_{j+1}(t+1) - W_j(t+1) W_j^\top(t+1) \|_F\nonumber\\ \label{eq:hybridep} & \leq & \| W_{j+1}^\top(0) W_{j+1}(0) - W_j(0) W_j^\top(0) \|_F \nonumber\\ && + 4 (2 \| \Phi\|_F)^{2(N-1)/N} \cdot \eta^2 \sum_{i=0}^{t} \left\| \frac{dL}{dW}(W(i)) \right\|_F^2. \end{eqnarray} Next, by $\MB(0), \ldots, \MB(t)$, we have that $L(W(i)) \leq L(W(0))$ for $i \leq t$. Since $W(0)$ has deficiency margin of $c$ and by Claim \ref{claim:margin_interp}, it then follows that $\sigma_{min}(W(i)) \geq c$ for all $i \leq t$. Therefore, by summing $\MB(0), \ldots, \MB(t)$, \begin{eqnarray} && \frac{1}{2} c^{2(N-1)/N} \eta\sum_{i=0}^t \left\| \frac{dL}{dW} W(i) \right\|_F^2\nonumber\\ & \leq & \frac 12 \eta \sum_{i=0}^t \sigma_{min}(W(i))^{2(N-1)/N} \left\| \frac{dL}{dW} (W(i)) \right\|_F^2\nonumber\\ & \leq & L(W(0)) - L(W(t))\nonumber\\ & \leq & L(W(0)) \leq \frac 12 \| \Phi\|_F^2\nonumber. \end{eqnarray} Therefore, \begin{eqnarray} &&4 \left( 2 \| \Phi\|_F\right)^{2(N-1)/N} \eta^2 \sum_{i=0}^t \left\| \frac{dL}{dW} W(i)\right\|_F^2 \nonumber\\ & \leq & 16 \| \Phi\|_F^{2(N-1)/N} \eta \frac{\| \Phi\|_F^2}{c^{2(N-1)/N}}\nonumber\\ \label{eq:sumetagrad} & \leq & 16 \| \Phi\|_F^{2(N-1)/N} \cdot \frac{1}{3 \cdot 2^{11} \cdot N^3} \cdot \frac{c^{(4N-2)/N}}{\| \Phi\|_F^{(6N-4)/N}} \cdot \frac{\| \Phi\|_F^2}{c^{2(N-1)/N}}\\ & \leq & \frac{c^2}{256 N^3 \| \Phi\|_F^{2(N-1)/N}}\nonumber\\ & = & \delta\nonumber, \end{eqnarray} where \eqref{eq:sumetagrad} follows from the definition of $\eta$ in (\ref{eq:descent_eta}), and the last equality follows from definition of $\delta$ in Theorem \ref{theorem:converge}. By (\ref{eq:hybridep}), it follows that $$ \| W_{j+1}^\top(t+1) W_{j+1}(t+1) - W_j(t+1) W_j^\top(t+1) \|_F \leq 2\delta, $$ verifying $\MA(t+1)$. \item $ \MA(t), \MB(t) \Rightarrow \MC(t)$. We apply Lemma \ref{lem:boundindiv} with $\nu = 2\delta$ and $C = 2 \| \Phi\|_F$. First, the triangle inequality and $\MB(t)$ give $$ \| W_{1:N}(t) \|_\sigma \leq \| \Phi\|_\sigma + \| \Phi - W_{1:N}(t) \|_\sigma \leq \| \Phi\|_F + \sqrt{2 \cdot L(W_{1:N}(t))} \leq 2 \| \Phi\|_F, $$ verifying precondition 2.~of Lemma \ref{lem:boundindiv}. $\MA(t)$ verifies condition 1.~of Lemma \ref{lem:boundindiv}, so for $1 \leq j \leq N$, $\| W_j(t) \|_\sigma \leq (4 \| \Phi\|_F)^{1/N}$, giving $\MC(t)$. \end{enumerate} The proof of Lemma \ref{lem:remainsmooth} then follows by induction on $t$. \end{proof} \subsection{Proof of Theorem~\ref{theorem:converge_balance_init}} \label{app:proofs:converge_balance_init} Theorem \ref{theorem:converge_balance_init} is proven by combining Lemma \ref{lem:integrate_rad} below, which implies that the balanced initialization is likely to lead to an end-to-end matrix $W_{1:N}(0)$ with sufficiently large deficiency margin, with Theorem~\ref{theorem:converge}, which establishes convergence. \begin{lemma} \label{lem:integrate_rad} Let $d \in \BN, d \geq 20$; $b_2 > b_1 \geq 1$ be real numbers (possibly depending on $d$); and $\Phi \in \BR^d$ be a vector. Suppose that $\mu$ is a rotation-invariant distribution\footnote{Recall that a distribution on vectors $V \in \BR^d$ is {\it rotation-invariant} if the distribution of $V$ is the same as the distribution of $OV$, for any orthogonal $d \times d$ matrix $O$. If $V$ has a well-defined density, this is equivalent to the statement that for any $r > 0$, the distribution of $V$ conditioned on $\| V\|_2 = r$ is uniform over the sphere centered at the origin with radius $r$.} over $\BR^d$ with a well-defined density, such that, for some $0 < \ep < 1$, $$ \pr_{V \sim \mu} \left[ \frac{\| \Phi\|_2}{\sqrt{b_2d}} \leq \| V \|_2 \leq \frac{\| \Phi\|_2}{\sqrt{b_1d}}\right] \geq 1-\ep. $$ Then, with probability at least $(1-\ep) \cdot \frac{3 - 4F(2/\sqrt{b_1})}{2}$, $V$ will have deficiency margin $\| \Phi\|_2 / (b_2d)$ with respect to $\Phi$. \end{lemma} The proof of Lemma~\ref{lem:integrate_rad} is postponed to Appendix~\ref{app:proofs:margin_init}, where Lemma~\ref{lem:integrate_rad} will be restated as Lemma~\ref{lem:integrate_rad_restate}. One additional technique is used in the proof of Theorem \ref{theorem:converge_balance_init}, which leads to an improvement in the guaranteed convergence rate. Because the deficiency margin of $W_{1:N}(0)$ is very small, namely $\OO(\| \Phi\|_2/d_0)$ (which is necessary for the theorem to maintain constant probability), at the beginning of optimization, $\ell(t)$ will decrease very slowly. However, after a certain amount of time, the deficiency margin of $W_{1:N}(t)$ will increase to a constant, at which point the decrease of $\ell(t)$ will be much faster. To capture this acceleration, we apply Theorem~\ref{theorem:converge} a second time, using the larger deficiency margin at the new ``initialization.'' From a geometric perspective, we note that the matrices $W_1(0), \ldots, W_N(0)$ are very close to 0, and the point at which $W_j(0) = 0$ for all $j$ is a saddle. Thus, the increase in $\ell(t) - \ell(t+1)$ over time captures the fact that the iterates $(W_1(t), \ldots, W_N(t))$ escape a saddle point. \begin{proof}[Proof of Theorem \ref{theorem:converge_balance_init}] Choose some $a \geq 2$, to be specified later. By assumption, all entries of the end-to-end matrix at time 0, $W_{1:N}(0)$, are distributed as independent Gaussians of mean 0 and standard deviation $s \leq \|\Phi\|_2/\sqrt{ad_0^2}$. We will apply Lemma~\ref{lem:integrate_rad} to the vector $W_{1:N}(0) \in \BR^{d_0}$. Since its distribution is obviously rotation-invariant, in remains to show that the distribution of the norm $\| W_{1:N}(0) \|_2$ is not too spread out. The following lemma~---~a direct consequence of the Chernoff bound applied to the $\chi^2$ distribution with $d_0$ degrees of freedom~---~will give us the desired result: \begin{lemma}[\cite{laurent_adaptive_2000}, Lemma 1] \label{lem:laurent} Suppose that $d \in \BN$ and $V \in \BR^d$ is a vector whose entries are i.i.d.~Gaussians with mean 0 and standard deviation $s$. Then, for any $k > 0$, \begin{eqnarray} \pr\left[\| V \|_2^2 \geq s^2 \left(d + 2k + 2\sqrt{kd} \right) \right] &\leq& \exp(-k)\nonumber\\ \pr\left[\| V \|_2^2 \leq s^2 \left( d - 2\sqrt{kd} \right) \right] & \leq & \exp(-k)\nonumber. \end{eqnarray} \end{lemma} By Lemma \ref{lem:laurent} with $k = d_0/16$, we have that $$ \pr\left[ \frac{s^2 d_0}{2} \leq \| V \|_2^2 \leq 2s^2 d_0\right] \geq 1 - 2 \exp(-d_0/16). $$ We next use Lemma \ref{lem:integrate_rad}, with $b_1 = \| \Phi\|_2^2/(2s^2d_0^2), b_2 = 2\| \Phi\|_2^2/(s^2d_0^2)$; note that since $a \geq 2$, $b_1 \geq 1$, as required by the lemma. Lemma \ref{lem:integrate_rad} then implies that with probability at least \be \label{eq:expF} \left(1 - 2\exp(-d_0/16) \right) \frac{3-4F\left(2/\sqrt{a/2}\right)}{2}, \ee $W_{1:N}(0)$ will have deficiency margin ${s^2d_0}/{2\| \Phi\|_2}$ with respect to $\Phi$. By the definition of balanced initialization (Procedure~\ref{proc:balance_init}) $W_1(0), \ldots, W_N(0)$ are $0$-balanced. Since $2^4 \cdot 6144 < 10^5$, our assumption on $\eta$ gives \begin{equation} \label{eq:eta_single_output} \eta \leq \frac{(s^2d_0)^{4-2/N}}{2^4 \cdot 6144N^3 \| \Phi\|_2^{10-6/N}}, \end{equation} so that Equation~(\ref{eq:descent_eta}) holds with $c = \frac{s^2d_0}{2\|\Phi\|_2}$. The conditions of Theorem \ref{theorem:converge} thus hold with probability at least that given in Equation~(\ref{eq:expF}). In such a constant probability event, by Theorem \ref{theorem:converge} (and the fact that a positive deficiency margin implies $L^1(W_{1:N}(0))\leq\frac{1}{2}\| \Phi\|_2^2$), if we choose \begin{equation} \label{eq:t0lb} t_0 \geq \eta^{-1} \left( \frac{ 2\| \Phi\|_2}{s^2d_0} \right)^{2-2/N} \ln(4), \end{equation} then $L^1(W_{1:N}(t_0)) \leq \frac 18 \| \Phi\|_2^2 $, meaning that $\| W_{1:N}(t_0) - \Phi\|_2 \leq \frac 12 \| \Phi\|_2 = \| \Phi\|_2 - \frac 12 \sigma_{min}(\Phi)$. Moreover, by condition $\MA(t_0)$ of Lemma \ref{lem:remainsmooth} and the definition of $\delta$ in Theorem \ref{theorem:converge}, we have, for $1 \leq j \leq N-1$, \begin{equation} \label{eq:balanced_single_output} \| W_{j+1}^T(t_0) W_{j+1}(t_0) - W_j(t_0) W_j^T(t_0) \|_F \leq \frac{2s^4d_0^2}{(2\| \Phi\|_2)^2 \cdot 256 N^3 \| \Phi\|_2^{2-2/N}} = \frac{s^4 d_0^2}{512 N^3 \| \Phi\|_2^{4-2/N}}. \end{equation} We now apply Theorem \ref{theorem:converge} again, verifying its conditions again, this time with the initialization $(W_1(t_0), \ldots, W_N(t_0))$. First note that the end-to-end matrix $W_{1:N}(t_0)$ has deficiency margin $c = \| \Phi\|_2/2$ as shown above. The learning rate $\eta$, by Equation~(\ref{eq:eta_single_output}), satisfies Equation~(\ref{eq:descent_eta}) with $c = \| \Phi\|_2/2$. Finally, since $$ \frac{s^4 d_0^2}{512 N^3 \| \Phi\|_2^{4-2/N}} \leq \frac{\|\Phi\|^{2/N}}{(a^2d_0^2) \cdot 512 N^3} \leq \frac{\| \Phi\|^{2/N}(1/2)^2}{256 N^3} $$ for $d_0 \geq 2$, by Equation~(\ref{eq:balanced_single_output}), the matrices $W_1(t_0), \ldots, W_N(t_0)$ are $\delta$-balanced with $\delta = \frac{\| \Phi\|^{2/N}(1/2)^{2}}{256 N^3}$. Iteration~$t_0$ thus satisfies the conditions of Theorem~\ref{theorem:converge} with deficiency margin $\| \Phi\|_2/2$, meaning that for \be \label{eq:tmt0lb} T - t_0 \geq \eta^{-1} \cdot 2^{2-2/N} \cdot \| \Phi\|^{2/N-2} \ln\left( \frac{\| \Phi\|_2^2}{8\ep} \right), \ee we will have $\ell(T) \leq \ep$. Therefore, by Equations~(\ref{eq:t0lb}) and~(\ref{eq:tmt0lb}), to ensure that $\ell(T) \leq \ep$, we may take $$ T \geq 4\eta^{-1} \left( \ln(4) \left(\frac{\|\Phi\|_2}{s^2d_0}\right)^{2-2/N} + \| \Phi\|_2^{2/N-2}\ln(\| \Phi\|_2^2/(8\ep)) \right). $$ Recall that this entire analysis holds only with the probability given in Equation~(\ref{eq:expF}). As $\lim_{d \ra \infty} (1-2\exp(-d/16)) = 1$ and $\lim_{a \ra \infty} (3 - 4F(2\sqrt{2/a}))/2 = 1/2$, for any $0 < p < 1/2$, there exist $a, d_0' > 0$ such that for $d_0 \geq d_0'$, the probability given in Equation~(\ref{eq:expF}) is at least $p$. This completes the proof. \end{proof} In the context of the above proof, we remark that the expressions $1- 2\exp(-d_0/16)$ and $(3 - 4F(2\sqrt{2/a}))/2$ converge to their limits of $1$ and $1/2$, respectively, as $d_0,a \ra \infty$ quite quickly. For instance, to obtain a probability of greater than $0.25$ of the initialization conditions being met, we may take $d_0 \geq 100, a \geq 100$. \subsection{Proof of Claim~\ref{claim:balance_init}} \label{app:proofs:balance_init} We first consider the probability of $\delta$-balancedness holding between any two layers: \begin{lemma} \label{lem:abvarbound} Suppose $a,b,d \in \BN$ and $A \in \BR^{a \times d}, B \in \BR^{d \times b}$ are matrices whose entries are distributed as i.i.d.~Gaussians with mean 0 and standard deviation $s$. Then for $k \geq 1$, \begin{equation} \label{eq:abd} \pr\left[ \left\|A^TA - BB^T\right\|_F \geq ks^2 \sqrt{2d(a+b)^2 + d^2(a+b)} \right] \leq 1/k^2. \end{equation} \end{lemma} \begin{proof} Note that for $1 \leq i,j \leq d$, let $X_{ij}$ be the random variable $(A^TA - BB^T)_{ij}$, so that $$X_{ij} = (A^TA - BB^T)_{ij} = \sum_{1 \leq \ell \leq a} A_{\ell i} A_{\ell j} - \sum_{1 \leq r \leq b} B_{ir} B_{jr}.$$ If $i \neq j$, then $$\ex[X^2] = \sum_{1 \leq \ell \leq a} \ex[A_{\ell i}^2 A_{\ell j}^2] + \sum_{1 \leq r \leq b} \ex[B_{ir}^2 B_{jr}^2]=(a+b)s^4.$$ We next note that for a normal random variable $Y$ of variance $s^2$ and mean 0, $\ex[Y^4] = 3s^4$. Then if $i = j$, $$ \ex[X^2] = s^4 \cdot (3(a+b) + a(a-1) + b(b-1) {-} ab) \leq s^4((a+b)^2 + 2(a+b)). $$ Thus \begin{eqnarray} \ex[\| A^TA - BB^T\|_F^2] &{\leq}& s^4(d((a+b)^2 + 2(a+b)) + d(d-1)(a+b)) \nonumber\\ &\leq & s^4(2d(a+b)^2 + d^2(a+b))\nonumber. \end{eqnarray} Then (\ref{eq:abd}) follows from Markov's inequality. \end{proof} Now the proof of Claim \ref{claim:balance_init} follows from a simple union bound: \begin{proof}[Proof of Claim \ref{claim:balance_init}] By (\ref{eq:abd}) of Lemma \ref{lem:abvarbound}, for each $1 \leq j \leq N-1$, $k \geq 1$, $$ \pr\left[ \| W_{j+1}^T W_{j+1} - W_jW_j^T \|_F {\geq} ks^2 \sqrt{10d_{max}^3} \right] \leq 1/k^2. $$ By the union bound, $$ \pr \left[ \forall 1 \leq j \leq N-1, \ \ \| W_{j+1}^T W_{j+1} - W_jW_j^T \|_F \leq ks^2 \sqrt{10d_{max}^3} \right] \geq 1-N/k^2, $$ and the claim follows with $\delta = ks^2 \sqrt{10d_{max}^3}$. \end{proof} \subsection{Proof of Claim~\ref{claim:margin_init}} \label{app:proofs:margin_init} We begin by introducing some notation. Given $d \in \BN$ and $r > 0$, we let $B^d(r)$ denote the open ball of radius $r$ centered at the origin in $\BR^d$. For an open subset $U \subset \BR^d$, let $\partial U := \bar U \backslash U$ be its boundary, where $\bar U$ denotes the closure of $U$. For the special case of $U = B^d(r)$, we will denote by $S^d(r)$ the boundary of such a ball, i.e.~the sphere of radius $r$ centered at the origin in $\BR^d$. Let $S^d := S^d(1)$ and $B^d := B^d(1)$. There is a well-defined uniform (Haar) measure on $S^d(r)$ for all $d, r$, which we denote by $\sigma^{d,r}$; we assume $\sigma^{d,r}$ is normalized so that $\sigma^{d,r}(S^d(r)) = 1$. Finally, since in the context of this claim we have~$d_N=1$, we allow ourselves to regard the end-to-end matrix $W_{1:N}\in\R^{1\times{d}_0}$ as both a matrix and a vector. \medskip To establish Claim \ref{claim:margin_init}, we will use the following low-degree anti-concentration result of \cite{carbery_distributional_2001} (see also \cite{lovett_elementary_2010,meka_anti-concentration_2016}): \begin{lemma}[\cite{carbery_distributional_2001}] \label{lem:carbery} There is an absolute constant $C_{{0}}$ such that the following holds. Suppose that ${h}$ is a multilinear polynomial of $K$ variables $X_1, \ldots, X_K$ and of degree $N$. Suppose that $X_1, \ldots, X_K$ are i.i.d.~Gaussian. Then, for any $\epsilon>0$: $$ \mathbb{P} \left[|{h}(X_1, \ldots, X_K)| \leq \ep \cdot \sqrt{\Var[{h}(X_1, \ldots, X_K)]}\right] \leq C_{{0}} N\ep^{1/N}. $$ \end{lemma} The below lemma characterizes the norm of the end-to-end matrix~$W_{1:N}$ following zero-centered Gaussian initialization: \begin{lemma} \label{lem:norm_concentrate} For any constant $0 < C_2 < 1$, there is an absolute constant $C_1 > 0$ such that the following holds. Let $N, d_0, \ldots, d_{N-1} \in \BN$. Set $d_N = 1$. Suppose that for $1 \leq j \leq N$, $W_j \in \BR^{d_j \times d_{j-1}}$ are matrices whose entries are i.i.d.~Gaussians of standard deviation $s$ and mean 0. Then $$ \pr\left[ s^{2N} d_1 \cdots d_{N-1} \left( \frac{1}{C_1N} \right)^{2N} \leq \| W_{1:N}\|_2^2 \leq C_1d_0^2 d_1 \cdots d_{N-1} s^{2N} \right] \geq C_2. $$ \end{lemma} \begin{proof} Let $f(W_1, \ldots, W_N) = \| W_{1:N}\|_2^2$, so that $f$ is a polynomial of degree $2N$ in the entries of $W_1, \ldots, W_N$. Notice that $$ f(W_1, \ldots, W_N) = \sum_{i_0 = 1}^{d_0} \left( \sum_{i_1=1}^{d_1} \cdots \sum_{i_{N-1} = 1}^{d_{N-1}} (W_N)_{1,i_{N-1}} (W_{N-1})_{i_{N-1}, i_{N-2}} \cdots (W_1)_{i_1, i_0} \right)^2. $$ For $1 \leq i_0 \leq d_0$, set $$ g_{i_0}(W_1, \ldots, W_N) = \sum_{i_1=1}^{d_1} \cdots \sum_{i_{N-1} = 1}^{d_{N-1}} (W_N)_{1,i_{N-1}} (W_{N-1})_{i_{N-1}, i_{N-2}} \cdots (W_1)_{i_1, i_0}, $$ so that $f = \sum_{i_0 = 1}^{d_0} g_{i_0}^2$. Since each $g_{i_0}$ is a multilinear polynomial in $W_1,\ldots,W_N$, we have that $\ex[g_{i_0}(W_1, \ldots, W_N)] = 0$ for all $1 \leq i_0 \leq d_0$. Also \begin{eqnarray} \Var[g_{i_0}(W_1, \ldots, W_N)] &=& \ex[g_{i_0}(W_1, \ldots, W_N)^2]\nonumber\\ &=& \sum_{i_1=1}^{d_1} \cdots \sum_{i_{N-1} = 1}^{d_{N-1}} \ex\left[ (W_N)_{1,i_{N-1}}^2 (W_{N-1})_{i_{N-1}, i_{N-2}}^2 \cdots (W_1)_{i_1, i_0}^2 \right]\nonumber\\ &=& d_1 d_2 \cdots d_{N-1} s^{2N} \nonumber. \end{eqnarray} It then follows by Markov's inequality that for any $k \geq 1$, $\pr[g_{i_0}^2 \geq k s^{2N} d_1 \cdots d_{N-1}] \leq 1/k$. For any constant $B_1$ (whose exact value will be specified below), it follows that \begin{eqnarray} &&\pr[f(W_1, \ldots, W_N) {\geq} B_1 d_0^2 d_1 d_2 \cdots d_{N-1} s^{2N}]\nonumber\\ &=& \pr\left[ \sum_{i_0=1}^{d_0} g_{i_0}(W_1, \ldots, W_N)^2 {\geq} B_1 d_0^2 d_1 d_2 \cdots d_{N-2} s^{2N} \right]\nonumber\\ & \leq & d_0 \cdot \pr[g_1(W_1, \ldots, W_N)^2 {\geq} B_1 d_0d_1 \cdots d_{N-1} s^{2N}]\nonumber\\ \label{eq:poly_ub} & \leq & 1/B_1. \end{eqnarray} Next, by Lemma~\ref{lem:carbery}, there is an absolute constant $C_0 > 0$ such that for any $\ep > 0$, and any $1 \leq i_0 \leq d_0$, $$ \pr\left[ |g_{i_0}(W_1, \ldots, W_N)| \leq \ep^N \sqrt{s^{2N} d_1 \cdots d_{N-1}} \right] \leq C_0 N \ep. $$ Since $f^2 \geq g_{i_0}^2$ for each $i_0$, it follows that \begin{equation} \label{eq:poly_lb} \pr[ f(W_1, \ldots, W_N) \geq \ep^{2N} s^{2N} d_1 \cdots d_{N-1} ] \geq 1 - C_0 N \ep. \end{equation} Next, given $0 < C_2 < 1$, choose $\ep = (1-C_2)/(2C_0N)$, and $B_1 = 2/(1-C_2)$. Then by (\ref{eq:poly_ub}) and (\ref{eq:poly_lb}) and a union bound, we have that $$ \pr\left[\left( \frac{1-C_2}{2C_0N}\right)^{2N} s^{2N} d_1 \cdots d_{N-1} \leq f(W_1, \ldots, W_N) \leq \frac{2}{1-C_2} {s^{2N}}d_0^2 d_1 \cdots d_{N-1} \right] \geq C_2. $$ The result of the lemma then follows by taking $C_1 = \max\left\{ \frac{2}{1-C_2}, \frac{2C_0}{1-C_2} \right\}$. \end{proof} \begin{lemma} \label{lem:rotation_inv} Let $N, d_0, \ldots, d_{N-1} \in \BN$, and set $d_N = 1$. Suppose $W_j \in \BR^{d_j \times d_{j-1}}$ for $1 \leq j \leq N$, are matrices whose entries are i.i.d.~Gaussians with mean~$0$ and standard deviation~$s$. Then, the distribution of $W_{1:N}$ is rotation-invariant. \end{lemma} \begin{proof} First we remark that for any orthogonal matrix $O \in \BR^{d_0 \times d_0}$, the distribution of $W_1$ is the same as that of $W_1O$. To see this, let us denote the rows of $W_1$ by $(W_1)_1, \ldots, (W_1)_{d_1}$, and the columns of $O$ by $O^1, \ldots, O^{d_0}$. Then the $(i_1, i_0)$ entry of $W_1O$, for $1 \leq i_1 \leq d_1, 1 \leq i_0 \leq d_0$ is $\langle (W_1)_{i_1}, O^{i_0} \rangle$, which is a Gaussian with mean~$0$ and standard deviation~$s$, since $\| O^{i_0}\|_2 = 1$. Since $\langle O^{i_0}, O^{i_0'} \rangle = 0$ for $i_0 \neq i_0'$, the covariance between any two distinct entries of $W_1O$ is 0. Therefore, the entries of $W_1O$ are independent Gaussians with mean~$0$ and standard deviation~$s$, just as are the entries of $W_1$. But now for any matrix $O \in \BR^{d_0 \times d_0}$, the distribution of $W_{1:N}O$ is the distribution of $W_N W_{N-1} \cdots W_2 (W_1O)$, which is the same as the distribution of $W_N W_{N-1} \cdots W_2 W_1 = W_{1:N}$, since $W_1, W_2, \ldots, W_N$ are all independent. \end{proof} For a dimension~$d\in\N$, radius~$r>0$, and $0 < h < r$, a {\it $(d,r)$-hyperspherical cap of height~$h$} is a subset $\MC \subset B^d(r)$ of the form $\{ x \in B^d(r) : \langle x, u \rangle \geq r-h\}$, where $u$ is any $d$-dimensional unit vector. We define the {\it area of a $(d,r)$-hyperspherical cap of height $h$}~---~$\MC$~---~to be $\sigma^{d,r}(\partial \MC \cap S^d(r))$. \begin{lemma} \label{lem:volcap} For $d \geq 20$, choose any $0 \leq h \leq 1$. Then, the area of a $(d,1)$-hyperspherical cap of height $h$ is at least $$ \frac{3 - 4 F((1-h) \sqrt{d-3})}{2}. $$ \end{lemma} \begin{proof} In \cite{chudnov_minimax_1986}, it is shown that the area of a $(d,1)$-hyperspherical cap of height $h$ is given by $\frac{1 - C_{d-2}(h)/C_{d-2}(0)}{2}$, where $$ C_d(h) := \int_0^{1-h} (1-t^2)^{(d-1)/2} dt. $$ Next, by the inequality $1-t^2 \geq \exp(-2t^2)$ for $0 \leq t \leq 1/2$, \begin{eqnarray} \int_0^1 (1-t^2)^{(d-3)/2} dt & \geq & \int_0^{1/2} \exp\left(2 \cdot \frac{-t^2(d-3)}{2} \right) dt\nonumber\\ & = & \sqrt{\pi/(d-3)} \cdot \frac{2F(\sqrt{(d-3)/2}) - 1}{2}\nonumber\\ \label{eq:cd0} & \geq & \sqrt{\pi/(d-3)} \cdot \frac{1 - 2 \exp(-(d-3)/4)}{2}, \end{eqnarray} where the last inequality follows from the standard estimate $F(x) \geq 1 - \exp(-x^2/2)$ for $x \geq 1$. Also, since $1-t^2 \leq \exp(-t^2)$ for all $t$, \begin{eqnarray} \int_0^{1-h} (1-t^2)^{(d-3)/2} dt & \leq & \int_0^{1-h} \exp\left( \frac{-t^2 (d-3)}{2} \right) dt \nonumber\\ \label{eq:cdh} & = & \sqrt{2\pi/(d-3)} \cdot \frac{2F((1-h) \sqrt{d-3}) - 1}{2}. \end{eqnarray} Therefore, for $d \geq 20$, by (\ref{eq:cd0}) and (\ref{eq:cdh}), \begin{eqnarray} \frac{1-C_{d-2}(h)/C_{d-2}(0)}{2} & \geq &\frac{1 - \frac{\sqrt{2} \cdot (2F((1-h)\sqrt{d-3}) - 1)}{1 - 2 \exp(-(d-3)/4)}}{2} \nonumber\\ & \geq & \frac{1 - \sqrt{2} \cdot (2F((1-h)\sqrt{d-3}) - 1) \cdot (1 + 4 \exp(-(d-3)/4))}{2}\nonumber\\ & \geq & \frac{3 - 4 F((1-h)\sqrt{d-3})}{2},\nonumber \end{eqnarray} where the second inequality has used $1/(1-y) \leq 1 + 2y$ for all $0 < y < 1/2$ (and where $y = 2\exp((-(d-3)/4)) < 2 \exp(-17/4) < 1/2$), and the final inequality uses $1 + 4\exp(-(d-3)/4) \leq \sqrt{2}$ for $d \geq 20$. The above chain of inequalities gives us the desired result. \end{proof} \begin{lemma} \label{lemma:haar_def} Let $d \in \BN, d \geq 20$; $a \geq 1$ be a real number (possibly depending on $d$); and $\Phi \in \BR^d$ be some vector. Set $r = \| \Phi\|_2 / \sqrt{ad}$, and suppose that $V \in S^d(r)$ is drawn according to the uniform measure. Then, with probability at least $\frac{3 - 4F(2/\sqrt{a})}{2}$, $V$ will have deficiency margin $\| \Phi\|_2 / (ad)$ with respect to $\Phi$. \end{lemma} \begin{proof} By rescaling, we may assume without loss of generality that $\| \Phi\|_2 = 1$, so that $r = 1/\sqrt{ad}$. Let $\MD$ denote the intersection of $B^{d}(r)$ with the open $d$-ball of radius $1-1/(ad)$ centered at $\Phi$. Let $\MC \subset B^{d}(r)$ denote the $(d,r)$-hyperspherical cap of height $r \cdot \big( 1 - 2/(\sqrt{ad})\big) = r - 2/(ad)$ whose base is orthogonal to the line between $\mathbf{0}$ and $\Phi$ (see Figure \ref{fig:initfig}). Note that ${\sigma}^{d,r}(\partial\MD \cap S^d(r))$, the Haar measure of the portion of $\partial\MD$ intersecting $S^d(r)$, gives the probability that $V$ belongs to the boundary of $\MD$. By Lemma \ref{lem:volcap} above (along with rescaling arguments), since $d \geq 20$, $\sigma^{d,r}(\partial \MC \cap S^d(r)) \geq \frac 12 \cdot (3 - 4F(2/\sqrt{a}))$, and therefore $V \in {\partial}\MC$ with at least this probability. We next claim that $\MC \subseteq \MD$. To see this, first let $\MT \subset \BR^d$ denote the $(d-1)$-sphere of radius $1-1/(ad)$ centered at $\Phi$ (see Figure \ref{fig:initfig}). Let $P$ be the intersection of $\MT$ with the line from $\mathbf{0}$ to $\Phi$, and $Q$ denote the intersection of this line with the unique hyperplane of codimension 1 containing $\MT \cap \partial B^{d}(r)$~---~we denote this hyperplane by $\MH$. If we can show that $\norm{P-Q}_2 \leq 1/(ad)$, then it follows that $\MC$ lies entirely on the other side of $\MH$ as $\mathbf{0}$, which will complete the proof that $\MC \subseteq \MD$. The calculation of $\norm{P-Q}_2$ is simply an application of the law of cosines: letting $\theta$ be the angle determining the intersection of $\partial B^{d}(r)$ and $\MT$ (see Figure~\ref{fig:initfig}), note that $$ (1-1/(ad))^2 = r^2 + 1^2 - 2r\cos \theta = 1/(ad) + 1 - 2/\sqrt{ad} \cdot \cos(\theta), $$ so $$ d(P,Q) = r \cos \theta - 1/(ad) = \frac 12 (1/(ad) - 1/(a^2d^2)) < 1/(ad), $$ as desired. Using that $\MC \subseteq \MD$, we continue with the proof. Notice the fact that $\MC\subseteq \MD$ is equivalent to $\partial \MC \cap S^d(r) \subseteq \partial \MD \cap S^d(r)$, by the structure of $\MC$ and~$\MD$. {Since the probability that $V$ lands in $\partial\MC$ is at least $\frac{3-4F(2/\sqrt{a})}{2}$, this lower bound applies to $V$ landing in $\partial\MD$ as well}. Since all $V \in \partial\MD$ have distance at most $1-1/(ad)$ from $\Phi$, and since $\sigma_{min}(\Phi) = \| \Phi\|_2 = 1$, it follows that for any $V \in \partial\MD$, $\| V - \Phi\|_2 \leq \sigma_{min}(\Phi) - 1/(ad)$. Therefore, with probability of at least $\frac{3-4F(2/\sqrt a)}{2}$, $V$ has deficiency margin $\| \Phi\|_2/(ad)$ with respect to~$\Phi$. \end{proof} \begin{figure} \begin{center} \begin{tikzpicture} \draw[very thin, pattern = north west lines, pattern color = gray, opacity=0.25] (2.73,-4.2) -- (2.73, 4.2) arc (56.65:-56.65:5cm) -- cycle; \draw[very thin, pattern = north west lines, pattern color = gray, opacity=0.25] (2.73,-4.2) -- (2.73, 4.2) arc (148:212:8cm) -- cycle; \draw (0,0) circle (5cm); \draw (0,0) -- (10,0); \draw (0,0) -- (2.7, 4.25); \draw [thick, blue] (2.7,-5) -- (2.7,5); \node [above] at (2.7,5) {\color{blue} $\MH$}; \node [above] at (10, 0) {$\Phi$}; \node [above] at (0,0) {$\mathbf{0}$}; \node [above left] at (1.5, 0) {$P$}; \node [above right] at (2.7, 0) {$Q$}; \node [above right] at (0.2, 0) {$\theta$}; \node [above] at (1.3, 2.1) {$r$}; \node [right] at (5.5, 6.92) {\color{red} $\MT$}; \draw [thick,red] (5.5,6.92) arc (120:240:8cm) \filldraw[fill opacity=0.2,fill=green] (3.8,-3.25) -- (3.8,3.25) arc (40.39:-40.39:5cm) -- cycle; \node [above] at (4.35,-0.9) {\color{darkspringgreen} $\MC$}; \draw[decoration={calligraphic brace,amplitude=5pt}, decorate, line width=1.25pt] (3.85,0.05) node {} -- (4.95,0.05); \node [above,align=center] at (4.4, 0.1) {\small Height \\ \small of $\MC$}; \end{tikzpicture} \end{center} \caption{Figure for proof of Lemma \ref{lemma:haar_def}. The dashed region denotes $\MD$. Not to scale.} \label{fig:initfig} \end{figure} \begin{lemma}[Lemma~\ref{lem:integrate_rad} restated] \label{lem:integrate_rad_restate} Let $d \in \BN, d \geq 20$; $b_2 > b_1 \geq 1$ be real numbers (possibly depending on $d$); and $\Phi \in \BR^d$ be a vector. Suppose that $\mu$ is a rotation-invariant distribution over $\BR^d$ with a well-defined density, such that, for some $0 < \ep < 1$, $$ \pr_{V \sim \mu} \left[ \frac{\| \Phi\|_2}{\sqrt{b_2d}} \leq \| V \|_2 \leq \frac{\| \Phi\|_2}{\sqrt{b_1d}}\right] \geq 1-\ep. $$ Then, with probability at least $(1-\ep) \cdot \frac{3 - 4F(2/\sqrt{b_1})}{2}$, $V$ will have deficiency margin $\| \Phi\|_2 / (b_2d)$ with respect to $\Phi$. \end{lemma} \begin{proof} By rescaling we may assume that $\| \Phi\|_2 = 1$ without loss of generality. Then the deficiency margin of $V$ is equal to $1 - \| V- \Phi\|_2$. $\mu$~has a well-defined density, so we can set $\hat{\mu}$ to be the probability density function of $\| V\|_2$. Since $\mu$ is rotation-invariant, we can integrate over spherical coordinates, giving \begin{eqnarray} && \pr[1 - \| V - \Phi\|_2 \geq {1} / (b_2d)] \nonumber\\ &=& \int_0^\infty \pr\big[1 - \| V - \Phi\|_2 \geq {1} / (b_2d) ~\big|~ \| V \|_2 = r \big]\hat \mu(r) dr\nonumber\\ & \geq & \int_{1/(\sqrt{b_2d})}^{1/(\sqrt{b_1d})} \frac{3 - 4F(2r\sqrt{d} )}{2} \hat \mu(r) dr \nonumber\\ & \geq & \frac{3 - 4F(2/\sqrt{b_1})}{2} \cdot \int_{1/(\sqrt{b_2d})}^{1/(\sqrt{b_1d})} \hat\mu(r) dr \nonumber\\ & \geq & \frac{3 - 4F(2/\sqrt{b_1})}{2} \cdot (1-\ep)\nonumber, \end{eqnarray} where the first inequlaity used Lemma \ref{lemma:haar_def} and the fact that the distribution of $V$ conditioned on $\| V\|_2 = r$ is uniform on $S^d(r)$. \end{proof} Now we are ready to prove Claim \ref{claim:margin_init}: \begin{proof}[Proof of Claim \ref{claim:margin_init}] We let $W \in \BR^{1 \times d_0} \simeq \BR^{d_0}$ denote the random vector $W_{1:N}$; also let $\mu$ denote the distribution of $W$, so that by Lemma \ref{lem:rotation_inv}, $\mu$ is rotation-invariant. Let $C_1$ be the constant from Lemma~\ref{lem:norm_concentrate} for $C_2 = 999/1000$. For some $a \geq 10^5$, the standard deviation of the entries of each $W_{j}$ is given by \begin{equation} \label{eq:define_s} s = \left( \frac{\| \Phi\|_2^2}{ad_0^3d_1 \cdots d_{N-1} C_1} \right)^{1/(2N)}. \end{equation} Then by Lemma \ref{lem:norm_concentrate}, $$ \pr \left[ \frac{ \| \Phi\|_2^2}{ad_0^3 C_1} \cdot \left( \frac{1}{C_1N} \right)^{2N} \leq \| W\|_2^2 \leq \frac{\| \Phi\|_2^2}{ad_0} \right] \geq {\frac{999}{1000}}. $$ Then Lemma~\ref{lem:integrate_rad_restate}, with $d = d_0$, $b_1 = a$ and $b_2 = ad_0^2 C_1 \cdot (C_1N)^{2N}$, implies that with probability at least $\frac{999}{1000} \cdot \frac{3 - 4F(2/\sqrt{a})}{2}$, $W$ has deficiency margin $\| \Phi\|_2 / (ad_0^3 C_1^{2N+1} N^{2N})$ with respect to $\Phi$. But $a \geq 10^5$ implies that this probability is at least $0.49$, and from (\ref{eq:define_s}), \be \label{eq:deficiency_margin_final} \frac{\| \Phi\|_2}{ad_0^3 C_1^{2N+1} N^{2N}} = \frac{s^{2N} d_1 \cdots d_{N-1}}{\| \Phi\|_2 (C_1N)^{2N}}. \ee Next recall the assumption in the hypothesis that $s\geq{C}_{1}N(c\cdot\norm{\Phi}_2/(d_1\cdots{d}_{N-1}))^{1/2N}$. Then the deficiency margin in (\ref{eq:deficiency_margin_final}) is at least $$ \frac{\left(C_1N(c \norm{\Phi}_2 / (d_1 \cdots d_{N-1}))^{1/(2N)}\right)^{2N} d_1 \cdots d_{N-1}}{\| \Phi\|_2 (C_1N)^{2N}} = c, $$ completing the proof. \end{proof} \subsection{Proof of Claim~\ref{claim:diverge_balance}} \label{app:proofs:diverge_balance} \begin{proof} The target matrices~$\Phi$ that will be used to prove the claim satisfy $\sigma_{min}(\Phi)=1$. We may assume without loss of generality that $c \geq 3/4$, the reason being that if a matrix has deficiency margin $c$ with respect to $\Phi$ and $c' < c$, it certainly has deficiency margin $c'$ with respect to $\Phi$. We first consider the case $d=1$, so that the target and all matrices are simply real numbers; we will make a slight abuse of notation in identifying $1\times1$ matrices with their unique entries. We set $\Phi = 1$. For all choices of $\eta$, we will set the initializations $W_1(0), \ldots, W_N(0)$ so that $W_{1:N}(0) = c$. Then $$ \| W_{1:N}(0) - \Phi\|_F = |W_{1:N}(0) - \Phi| = 1-c = \sigma_{min}(\Phi) - c, $$ so the initial end-to-end matrix $W_{1:N}(0) \in \BR^{1 \times 1}$ has deficiency margin $c$. Now fix $\eta$. Choose $A \in \BR$ with \begin{eqnarray} \label{eq:define_A} A &=& \max \left\{\sqrt{\eta N}, \frac{2}{\eta(1-c) c^{(N-1)/N}}, 2000,20/\eta,\left( \frac{20 \cdot 10^{2N-1}}{\eta^{2N}}\right)^{1/(2N-2)} \right\}. \end{eqnarray} We will set: \begin{equation} \label{eq:define_wj0} W_j(0) = \begin{cases} Ac^{1/N} \quad : \quad 1 \leq j \leq N/2 \\ c^{1/N}/A \quad : \quad N/2 < j \leq N, \end{cases} \end{equation} so that $W_{1:N}(0) = c$. Then since $L^N(W_1, \ldots, W_N) = \frac 12 (1 - W_N \cdots W_1)^2$, the gradient descent updates are given by $$ W_{j}(t+1) = W_j(t) - \eta (W_{1:N}(t) - 1) \cdot W_{1:j-1}(t) W_{j+1:N}(t), $$ where we view $W_1(t), \ldots, W_N(t)$ as real numbers. This gives \begin{equation} W_j(1) = \begin{cases} c^{1/N}A - \eta (c-1)c^{(N-1)/N}/A \quad : \quad 1 \leq j \leq N/2 \\ c^{1/N}/A - \eta (c-1) c^{(N-1)/N}A \quad : \quad N/2 < j \leq N. \end{cases}\nonumber \end{equation} Since $3/4 \leq c < 1$ and $-\eta(c-1)c^{(N-1)/N}A \geq 0$, we have that $A/2 \leq 3A/4 \leq W_j(1)$ for $1 \leq j \leq N/2$. Next, since $\frac{1-c}{1-c^{1/N}} \leq N$ for $0 \leq c < 1$, we have that $A^2 \geq \eta N \geq \frac{\eta (1-c)}{1-c^{1/N}}$, which implies that $A^2 \geq c^{1/N}A^2 + \eta (1-c)$, or $c^{1/N} A + \frac{\eta(1-c)}{A} \leq A$. Thus $W_j(1) \leq A$ for $N/2 < j \leq N$. Similarly, using the same bound $3/4 \leq c < 1$ and the fact that $\eta (1-c) c^{(N-1)/N} A \geq 2$ we get $\frac{3}{16} \eta A \leq W_j(1) \leq \eta A$ for $N/2 < j \leq N$. In particular, for all $1 \leq j \leq N$, we have that $\frac{\min\{\eta, 1\}}{10} A \leq W_j(1) \leq \max\{\eta, 1\} A$. We prove the following lemma by induction: \begin{lemma} \label{lemma:weight_explode} For each $t \geq 1$, the real numbers $W_1(t), \ldots, W_N(t)$ all have the same sign and this sign alternates for each integer $t$. Moreover, there are real numbers $2 \leq B(t) < C(t)$ for $t \geq 1$ such that for $1 \leq j \leq N$, $B(t) \leq |W_j(t)| \leq C(t)$ and $\eta B(t)^{2N-1} \geq 20 C(t)$. \end{lemma} \begin{proof} First we claim that we may take $B(1) = \frac{\min\{\eta, 1\}}{10} A$ and $C(1) = \max\{ \eta, 1\} A$. We have shown above that $B(1) \leq W_j(1) \leq C(1)$ for all $j$. Next we establish that $\eta B(1)^{2N-1} \geq 20 C(1)$. If $\eta \leq 1$, then $$ \eta B(1)^{2N-1} =\eta^{2N} \cdot (A/10)^{2N-1} \geq 20 A = 20 C(1), $$ where the inequality follows from $A \geq \left(\frac{20 \cdot 10^{2N-1}}{\eta^{2N}}\right)^{1/(2N-2)}$ by definition of $A$. If $\eta \geq 1$, then $$ \eta B(1)^{2N-1} = \eta (A/10)^{2N-1} \geq 20 \eta A = 20 C(1), $$ where the inequality follows from $A \geq 2000 \geq \left(20 \cdot 10^{2N-1}\right)^{1/(2N-2)}$ by definition of $A$. Now, suppose the statement of Lemma \ref{lemma:weight_explode} holds for some $t$. Suppose first that $W_j(t)$ are all positive for $1 \leq j \leq N$. Then for all $j$, as $B(t) \geq 2$, and $\eta B(t)^{2N-1} \geq 20 C(t)$, \begin{eqnarray} W_j(t+1) & \leq & C(t) - \eta \cdot (B(t)^N - 1) \cdot B(t)^{N-1} \nonumber\\ & \leq & C(t) - \frac{\eta}{2} B(t)^{2N-1}\nonumber\\ & \leq & -9 C(t)\nonumber, \end{eqnarray} which establishes that $W_j(t+1)$ is negative for all $j$. Moreover, \begin{eqnarray} W_j(t+1) & \geq & -\eta (C(t)^N - 1) \cdot C(t)^{N-1} \nonumber\\ & \geq & -\eta C(t)^{2N-1}. \nonumber \end{eqnarray} Now set $B(t+1) = 9C(t)$ and $C(t+1) = \eta C(t)^{2N-1}$. Since $N \geq 2$, we have that $$ \eta B(t+1)^{2N-1} = \eta (9C(t))^{2N-1} \geq \eta 9^3 C(t)^{2N-1} > 20 \eta C(t)^{2N-1} = 20 C(t+1). $$ The case that all $W_j(t)$ are negative for $1 \leq j \leq N$ is nearly identical, with the same values for $B(t+1), C(t+1)$ in terms of $B(t), C(t)$, except all $W_{j}(t+1)$ will be positive. This establishes the inductive step and completes the proof of Lemma \ref{lemma:weight_explode}. \end{proof} By Lemma \ref{lemma:weight_explode}, we have that for all $t \geq 1$, $L^N(W_1(t), \ldots, W_N(t)) = \frac 12 (W_{1:N}(t) - 1)^2 \geq \frac 12 (2^N - 1)^2 > 0$, thus completing the proof of Claim \ref{claim:diverge_balance} for the case where all dimensions are equal to 1. For the general case where $d_0 = d_1 = \cdots = d_N = d$ for some $d \geq 1$, we set $\Phi = I_d$, and given $c, \eta$, we set $W_j(0)$ to be the $d \times d$ diagonal matrix where all diagonal entries except the first one are equal to 1, and where the first diagonal entry is given by Equation~\eqref{eq:define_wj0}, where $A$ is given by Equation~\eqref{eq:define_A}. It is easily verified that all entries of $W_j(t)$, $1 \leq j \leq N$, except for the first diagonal element of each matrix, will remain constant for all $t \geq 0$, and that the first diagonal elements evolve exactly as in the 1-dimensional case presented above. Therefore the loss in the $d$-dimensional case is equal to the loss in the 1-dimensional case, which is always greater than some positive constant. \end{proof} We remark that the proof of Claim \ref{claim:diverge_balance} establishes that the loss $\ell(t):=L^N(W_1(t), \ldots, W_N(t))$ grows at least exponentially in $t$ for the chosen initialization. Such behavior, in which gradients and weights explode, indeed takes place in deep learning practice if initialization is not chosen with care. \subsection{Proof of Claim~\ref{claim:diverge_margin}} \label{app:proofs:diverge_margin} \begin{proof} We will show that a target matrix $\Phi\in\R^{d\times{d}}$ which is symmetric with at least one negative eigenvalue, along with identity initialization ($W_j(0) = I_d$, $\forall{j}\in\{1,\ldots,N\}$), satisfy the conditions of the claim. First, note that non-stationarity of initialization is met, as for any $1 \leq j \leq N$, $$ \frac{\partial L^N(W_1(0), \ldots, W_N(0))}{\partial W_j(0)} = W_{j+1:N}(0)^\top (W_{1:N}(0) - \Phi) W_{1:j-1}(0) = I_d - \Phi \neq \mathbf{0}, $$ where the last inequality follows since $\Phi$ has a negative eigenvalue. To analyze gradient descent we use the following result, which was established in \cite{bartlett2018gradient}: \begin{lemma}[\cite{bartlett2018gradient}, Lemma 6] \label{lemma:failure_idinit} If $W_1(0), \ldots, W_N(0)$ are all initialized to identity, $\Phi$ is symmetric, $\Phi = UDU^\top$ is a diagonalization of $\Phi$, and gradient descent is performed with any learning rate, then for each $t \geq 0$ there is a diagonal matrix $\hat D(t)$ such that $W_j(t) = U\hat D(t) U^\top$ for each $1 \leq j \leq N$. \end{lemma} By Lemma \ref{lemma:failure_idinit}, for any choice of learning rate $\eta$, the end-to-end matrix at time $t$ is given by $W_{1:N}(t) = U \hat D(t)^N U^\top$. As long as some diagonal element of $D$ is negative, say equal to $-\lambda < 0$, then $$ \ell(t)=L^N(W_1(t), \ldots, W_N(t)) = \frac 12 \| W_{1:N}(t) - \Phi\|_F^2 = \frac 12 \| \hat D(t)^L - D \|_F^2 \geq \frac 12 \lambda^2 > 0. $$ \end{proof} \section{$\ell_2$ Loss over Whitened Data} \label{app:whitened} Recall the $\ell_2$~loss of a linear predictor~$W\in\R^{d_y\times{d}_x}$ as defined in Section~\ref{sec:prelim}: $$L(W)=\frac{1}{2m}\|WX-Y\|_F^2 \text{\,,}$$ where~$X\in\R^{d_x\times{m}}$ and~$Y\in\R^{d_y\times{m}}$. Define $\Lambda_{xx}:=\tfrac{1}{m}XX^\top\in\R^{d_x\times{d}_x}$, $\Lambda_{yy}:=\tfrac{1}{m}YY^\top\in\R^{d_y\times{d}_y}$ and $\Lambda_{yx}:=\tfrac{1}{m}YX^\top\in\R^{d_y\times{d}_x}$. Using the relation $\norm{A}_F^2=\Tr(AA^\top)$, we have: \beas L(W)&=&\tfrac{1}{2m}\Tr\big((WX-Y)(WX-Y)^\top\big) \\[1mm] &=&\tfrac{1}{2m}\Tr(WXX^\top{W}^\top)-\tfrac{1}{m}\Tr(WXY^\top)+\tfrac{1}{2m}\Tr(YY^\top) \\[1mm] &=&\tfrac{1}{2}\Tr(W\Lambda_{xx}{W}^\top)-\Tr(W\Lambda_{yx}^\top)+\tfrac{1}{2}\Tr(\Lambda_{yy}) \text{\,.} \eeas By definition, when data is whitened, $\Lambda_{xx}$~is equal to identity, yielding: \beas L(W)&=&\tfrac{1}{2}\Tr(W{W}^\top)-\Tr(W\Lambda_{yx}^\top)+\tfrac{1}{2}\Tr(\Lambda_{yy}) \\[1mm] &=&\tfrac{1}{2}\Tr\big((W-\Lambda_{yx})(W-\Lambda_{yx})^\top\big)-\tfrac{1}{2}\Tr(\Lambda_{yx}\Lambda_{yx}^\top)+\tfrac{1}{2}\Tr(\Lambda_{yy}) \\[1mm] &=&\tfrac{1}{2}\norm{W-\Lambda_{yx}}_F^2+c \text{\,,} \eeas where~$c:=-\tfrac{1}{2}\Tr(\Lambda_{yx}\Lambda_{yx}^\top)+\tfrac{1}{2}\Tr(\Lambda_{yy})$ does not depend on~$W$. Hence we arrive at Equation~\eqref{eq:loss_whitened}. \subsubsection*{Acknowledgments} \else \section*{Acknowledgments} \fi \acknowledgments \fi \fi \section*{References} {\small \ifdefined\ICML \bibliographystyle{icml2018} \else \bibliographystyle{plainnat} \fi \section{Conclusion} \label{sec:conc} For deep linear neural networks, we have rigorously proven convergence of gradient descent to global minima, at a linear rate, provided that the initial weight matrices are approximately balanced and the initial end-to-end matrix has positive deficiency margin. The result applies to networks with arbitrary depth, and any configuration of input/output/hidden dimensions that supports full rank, \ie~in which no hidden layer has dimension smaller than both the input and output. Our assumptions on initialization~---~approximate balancedness and deficiency margin~---~are both necessary, in the sense that violating any one of them may lead to convergence failure, as we demonstrated explicitly. Moreover, for networks with output dimension~$1$ (scalar regression), we have shown that a balanced initialization, \ie~a random choice of the end-to-end matrix followed by a balanced partition across all layers, leads assumptions to be met, and thus convergence to take place, with constant probability. Rigorously proving efficient convergence with significant probability under customary layer-wise independent initialization remains an open problem. The recent work of~\citet{shamir2018exponential} suggests that this may not be possible, as at least in some settings, the number of iterations required for convergence is exponential in depth with overwhelming probability. This negative result, a theoretical manifestation of the ``vanishing gradient problem'', is circumvented by balanced initialization. Through simple experiments we have shown that the latter can lead to favorable convergence in deep learning practice, as it does in theory. Further investigation of balanced initialization, including development of variants for convolutional layers, is regarded as a promising direction for future research. The analysis in this paper uncovers special properties of the optimization landscape in the vicinity of gradient descent trajectories. We expect similar ideas to prove useful in further study of gradient descent on non-convex objectives, including training losses of deep non-linear neural networks. \section{Convergence Analysis} \label{sec:converge} In this section we establish convergence of gradient descent for deep linear neural networks (Equations~\eqref{eq:gd} and~\eqref{eq:lnn_obj}) by directly analyzing the trajectories taken by the algorithm. We begin in Subsection~\ref{sec:converge:balance_margin} with a presentation of two concepts central to our analysis: \emph{approximate balancedness} and \emph{deficiency margin}. These facilitate our main convergence theorem, delivered in Subsection~\ref{sec:converge:theorem}. We conclude in Subsection~\ref{sec:converge:balance_init} by deriving a convergence guarantee that holds with constant probability over a random initialization. \subsection{Approximate Balancedness and Deficiency Margin} \label{sec:converge:balance_margin} In our context, the notion of approximate balancedness is formally defined as follows: \vspace{1mm} \begin{definition} \label{def:balance} For~$\delta\geq0$, we say that the matrices $W_j\in\R^{d_j\times{d}_{j-1}}$, $j{=}1,\ldots,N$, are $\delta$-\emph{balanced} if: $$\norm{W_{j+1}^{\top}W_{j+1}-W_{j}W_j^\top}_F\leq\delta\quad,\,\forall{j}\in\{1,\ldots,N-1\} \text{\,.}$$ \end{definition} Note that in the case of $0$-balancedness, \ie~$W_{j+1}^{\top}W_{j+1}=W_{j}W_j^\top$, $\forall{j}\in\{1,\ldots,N-1\}$, all matrices~$W_j$ share the same set of non-zero singular values. Moreover, as shown in the proof of Theorem~1 in~\citet{arora2018optimization}, this set is obtained by taking the $N$-th root of each non-zero singular value in the end-to-end matrix~$W_{1:N}$. We will establish approximate versions of these facts for $\delta$-balancedness with $\delta>0$, and admit their usage by showing that if the weights of a linear neural network are initialized to be approximately balanced, they will remain that way throughout the iterations of gradient descent. The condition of approximate balancedness at initialization is trivially met in the special case of linear residual networks ($d_0=\cdots=d_N=d$ and $W_1(0)=\cdots=W_N(0)=I_d$). Moreover, as Claim~\ref{claim:balance_init} in Appendix~\ref{app:balance_margin_init} shows, for a given~$\delta>0$, the customary initialization via random Gaussian distribution with mean zero leads to approximate balancedness with high probability if the standard deviation is sufficiently small. \medskip The second concept we introduce~---~deficiency margin~---~refers to how far a ball around the target is from containing rank-deficient (\ie~low rank) matrices. \vspace{1mm} \begin{definition} \label{def:margin} Given a target matrix~$\Phi\in\R^{d_N\times{d}_0}$ and a constant $c>0$, we say that a matrix~$W\in\R^{d_N\times{d}_0}$ has \emph{deficiency margin~$c$ with respect to~$\Phi$} if:\note{ Note that deficiency margin~$c>0$ with respect to~$\Phi$ implies~$\sigma_{min}(\Phi)>0$, \ie~$\Phi$ has full rank. Our analysis can be extended to account for rank-deficient~$\Phi$ by replacing~$\sigma_{min}(\Phi)$ in Equation~\eqref{eq:margin} with the smallest positive singular value of~$\Phi$, and by requiring that the end-to-end matrix~$W_{1:N}$ be initialized such that its left and right null spaces coincide with those of~$\Phi$. Relaxation of this requirement is a direction for future work. } \be \norm{W-\Phi}_F\leq\sigma_{min}(\Phi)-c \label{eq:margin} \text{\,.} \ee \end{definition} The term ``deficiency margin'' alludes to the fact that if Equation~\eqref{eq:margin} holds, every matrix~$W'$ whose distance from~$\Phi$ is no greater than that of~$W$, has singular values $c$-bounded away from zero: \vspace{1mm} \begin{claim} \label{claim:margin_interp} Suppose~$W$ has deficiency margin~$c$ with respect to~$\Phi$. Then, any matrix~$W'$ (of same size as~$\Phi$ and~$W$) for which $\norm{W'-\Phi}_F\leq\norm{W-\Phi}_F$ satisfies~$\sigma_{min}(W')\geq{c}$. \end{claim} \vspace{-3mm} \begin{proof} Our proof relies on the inequality $\sigma_{min}(A+B)\geq\sigma_{min}(A)-\sigma_{max}(B)$~---~see Appendix~\ref{app:proofs:margin_interp}. \end{proof} \vspace{-2mm} We will show that if the weights $W_1,\ldots,W_N$ are initialized such that (they are approximately balanced and) the end-to-end matrix~$W_{1:N}$ has deficiency margin~$c>0$ with respect to the target~$\Phi$, convergence of gradient descent to global minimum is guaranteed.\note{ In fact, a deficiency margin implies that all critical points in the respective sublevel set (set of points with smaller loss value) are global minima. This however is far from sufficient for proving convergence, as sublevel sets are unbounded, and the loss landscape over them is non-convex and non-smooth. Indeed, we show in Appendix~\ref{app:fail} that deficiency margin alone is not enough to ensure convergence~---~without approximate balancedness, the lack of smoothness can cause divergence. } Moreover, the convergence will outpace a particular rate that gets faster when $c$ grows larger. This suggests that from a theoretical perspective, it is advantageous to initialize a linear neural network such that the end-to-end matrix has a large deficiency margin with respect to the target. Claim~\ref{claim:margin_init} in Appendix~\ref{app:balance_margin_init} provides information on how likely deficiency margins are in the case of a single output model (scalar regression) subject to customary zero-centered Gaussian initialization. It shows in particular that if the standard deviation of the initialization is sufficiently small, the probability of a deficiency margin being met is close to~$0.5$; on the other hand, for this deficiency margin to have considerable magnitude, a non-negligible standard deviation is required. \medskip Taking into account the need for both approximate balancedness and deficiency margin at initialization, we observe a delicate trade-off under the common setting of Gaussian perturbations around zero: if the standard deviation is small, it is likely that weights be highly balanced and a deficiency margin be met; however overly small standard deviation will render high magnitude for the deficiency margin improbable, and therefore fast convergence is less likely to happen; on the opposite end, large standard deviation jeopardizes both balancedness and deficiency margin, putting the entire convergence at risk. This trade-off is reminiscent of empirical phenomena in deep learning, by which small initialization can bring forth efficient convergence, while if exceedingly small, rate of convergence may plummet (``vanishing gradient problem''), and if made large, divergence becomes inevitable (``exploding gradient problem''). The common resolution of residual connections~\citep{he2016deep} is analogous in our context to linear residual networks, which ensure perfect balancedness, and allow large deficiency margin if the target is not too far from identity. \subsection{Main Theorem} \label{sec:converge:theorem} Using approximate balancedness (Definition~\ref{def:balance}) and deficiency margin (Definition~\ref{def:margin}), we present our main theorem~---~a guarantee for linear convergence to global minimum: \vspace{1mm} \begin{theorem} \label{theorem:converge} Assume that gradient descent is initialized such that the end-to-end matrix~$W_{1:N}(0)$ has deficiency margin~$c>0$ with respect to the target~$\Phi$, and the weights $W_1(0),\ldots,W_N(0)$ are $\delta$-balanced with $\delta=c^2\big/\big(256\cdot{N}^3\cdot\norm{\Phi}_{F}^{2(N-1)/N}\big)$. Suppose also that the learning rate~$\eta$ meets: \be \eta\leq\frac{c^{(4N-2)/N}}{6144\cdot{N}^3\cdot\norm{\Phi}_{F}^{(6N-4)/N}} \label{eq:descent_eta} \text{~.} \ee Then, for any~$\epsilon>0$ and: \be T\geq\frac{1}{\eta\cdot{c}^{2(N-1)/N}}\cdot\log\left(\frac{\ell(0)}{\epsilon}\right) \label{eq:converge_t} \text{\,,} \ee the loss at iteration~$T$ of gradient descent~---~$\ell(T)$~---~is no greater than~$\epsilon$. \end{theorem} \subsubsection{On the Assumptions Made} The assumptions made in Theorem~\ref{theorem:converge}~---~approximate balancedness and deficiency margin at initialization~---~are both necessary, in the sense that violating any one of them may lead to convergence failure. We demonstrate this in Appendix~\ref{app:fail}. In the special case of linear residual networks (uniform dimensions and identity initialization), a sufficient condition for the assumptions to be met is that the target matrix have (Frobenius) distance less than~$0.5$ from identity. This strengthens one of the central results in~\citet{bartlett2018gradient} (see Section~\ref{sec:related}). For a setting of random near-zero initialization, we present in Subsection~\ref{sec:converge:balance_init} a scheme that, when the output dimension is~$1$ (scalar regression), ensures assumptions are satisfied (and therefore gradient descent efficiently converges to global minimum) with constant probability. It is an open problem to fully analyze gradient descent under the common initialization scheme of zero-centered Gaussian perturbations applied to each layer independently. We treat this scenario in Appendix~\ref{app:balance_margin_init}, providing quantitative results concerning the likelihood of each assumption (approximate balancedness or deficiency margin) being met individually. However the question of how likely it is that both assumptions be met simultaneously, and how that depends on the standard deviation of the Gaussian, is left for future work. An additional point to make is that Theorem~\ref{theorem:converge} poses a structural limitation on the linear neural network. Namely, it requires the dimension of each hidden layer~($d_i,~i=1,\ldots,N-1$) to be greater than or equal to the minimum between those of the input~($d_0$) and output~($d_N$). Indeed, in order for the initial end-to-end matrix~$W_{1:N}(0)$ to have deficiency margin~$c>0$, it must (by Claim~\ref{claim:margin_interp}) have full rank, and this is only possible if there is no intermediate dimension~$d_i$ smaller than~$\min\{d_0,d_N\}$. We make no other assumptions on network architecture (depth, input/output/hidden dimensions). \subsubsection{Proof} The cornerstone upon which Theorem~\ref{theorem:converge} rests is the following lemma, showing non-trivial descent whenever $\sigma_{min}(W_{1:N})$ is bounded away from zero: \vspace{1mm} \begin{lemma} \label{lemma:descent} Under the conditions of Theorem~\ref{theorem:converge}, we have that for every $t=0,1,2,\ldots$~:\note{ Note that the term~$\frac{dL^1}{dW}(W_{1:N}(t))$ below stands for the gradient of~$L^1(\cdot)$~---~a convex loss over (directly parameterized) linear models (Equation~\eqref{eq:lin_obj})~---~at the point~$W_{1:N}(t)$~---~the end-to-end matrix of the network at iteration~$t$. It is therefore (see Equation~\eqref{eq:gd_loss}) non-zero anywhere but at a global minimum. } \be \ell(t+1)\leq\ell(t)-\frac{\eta}{2}\cdot\sigma_{min}\big(W_{1:N}(t)\big)^{\frac{2(N-1)}{N}}\cdot\norm{\frac{dL^1}{dW}\big(W_{1:N}(t)\big)}_F^2 \label{eq:descent} \text{~.} \ee \end{lemma} \begin{proof}[Proof of Lemma~\ref{lemma:descent} (in idealized setting; for complete proof see Appendix~\ref{app:proofs:descent})] \hspace{-0.5mm}We prove the lemma here for the idealized setting of perfect initial balancedness~($\delta=0$): $$W_{j+1}^{\top}(0)W_{j+1}(0)=W_{j}(0)W_j^\top(0)\quad,~\forall{j}\in\{1,\ldots,N-1\} \text{~,}$$ and infinitesimally small learning rate~($\eta\to0^+$)~---~\emph{gradient flow}: $$\dot{W}_j(\tau)=-\frac{\partial{L}^N}{\partial{W}_j}\big(W_1(\tau),\ldots,W_N(\tau)\big)\quad,~j=1,\ldots,N\quad,~\tau\in[0,\infty) \text{~,}$$ where~$\tau$ is a continuous time index, and dot symbol (in~$\dot{W}_j(\tau)$) signifies derivative with respect to time. The complete proof, for the realistic case of approximate balancedness and discrete updates~($\delta,\eta>0$), is similar but much more involved, and appears in Appendix~\ref{app:proofs:descent}. Recall that~$\ell(t)$~---~the objective value at iteration~$t$ of gradient descent~---~is equal to~$L^1(W_{1:N}(t))$ (see Equation~\eqref{eq:gd_loss}). Accordingly, for the idealized setting in consideration, we would like to show: \be \frac{d}{d\tau}L^1\left(W_{1:N}(\tau)\right)\leq-\frac{1}{2}\sigma_{min}\big(W_{1:N}(\tau)\big)^{\frac{2(N-1)}{N}}\cdot\norm{\frac{dL^1}{dW}\big(W_{1:N}(\tau)\big)}_F^2 \label{eq:descent_ideal} \text{\,.} \ee We will see that a stronger version of Equation~\eqref{eq:descent_ideal} holds, namely, one without the $1/2$ factor (which only appears due to discretization). By (Theorem~$1$ and Claim~$1$ in)~\citet{arora2018optimization}, the weights $W_1(\tau),\ldots,W_N(\tau)$ remain balanced throughout the entire optimization, and that implies the end-to-end matrix~$W_{1:N}(\tau)$ moves according to the following differential equation: \be vec\left(\dot{W}_{1:N}(\tau)\right)=-{P}_{W_{1:N}(\tau)}\cdot{vec}\left(\frac{dL^1}{dW}\left(W_{1:N}(\tau)\right)\right) \label{eq:e2e_gf} \text{\,,} \ee where~$vec(A)$, for an arbitrary matrix~$A$, stands for vectorization in column-first order, and $P_{W_{1:N}(\tau)}$ is a positive semidefinite matrix whose eigenvalues are all greater than or equal to $\sigma_{min}(W_{1:N}(\tau))^{2(N-1)/N}$. Taking the derivative of~$L^1(W_{1:N}(\tau))$ with respect to time, we obtain the sought-after Equation~\eqref{eq:descent_ideal} (with no $1/2$ factor): \beas \frac{d}{d\tau}L^1\left(W_{1:N}(\tau)\right)&=&\inprod{vec\left(\frac{dL^1}{dW}\big(W_{1:N}(\tau)\big)\right)}{vec\left(\dot{W}_{1:N}(\tau)\right)} \\ &=&\inprod{vec\left(\frac{dL^1}{dW}\big(W_{1:N}(\tau)\big)\right)}{-{P}_{W_{1:N}(\tau)}\cdot{vec}\left(\frac{dL^1}{dW}\left(W_{1:N}(\tau)\right)\right)} \\ &\leq&-\sigma_{min}\big(W_{1:N}(\tau)\big)^{\frac{2(N-1)}{N}}\cdot\norm{vec\left(\frac{dL^1}{dW}\big(W_{1:N}(\tau)\big)\right)}^2 \\ &=&-\sigma_{min}\big(W_{1:N}(\tau)\big)^{\frac{2(N-1)}{N}}\cdot\norm{\frac{dL^1}{dW}\big(W_{1:N}(\tau)\big)}_F^2 \text{\,.} \eeas The first transition here (equality) is an application of the chain rule; the second (equality) plugs in Equation~\eqref{eq:e2e_gf}; the third (inequality) results from the fact that the eigenvalues of the symmetric matrix $P_{W_{1:N}(\tau)}$ are no smaller than $\sigma_{min}(W_{1:N}(\tau))^{2(N-1)/N}$ (recall that~$\norm{\cdot}$ stands for Euclidean norm); and the last (equality) is trivial~---~$\norm{A}_F=\norm{vec(A)}$ for any matrix~$A$. \end{proof} With Lemma~\ref{lemma:descent} established, the proof of Theorem~\ref{theorem:converge} readily follows: \begin{proof}[Proof of Theorem~\ref{theorem:converge}] By the definition of~$L^1(\cdot)$ (Equation~\eqref{eq:lin_obj}), for any~$W\in\R^{d_N\times{d}_0}$: $$\frac{dL^1}{dW}(W)=W-\Phi\quad\implies\quad\norm{\frac{dL^1}{dW}(W)}_F^2=2\cdot{L}^1(W) \text{~.}$$ Plugging this into Equation~\eqref{eq:descent} while recalling that $\ell(t)=L^1(W_{1:N}(t))$ (Equation~\eqref{eq:gd_loss}), we have (by Lemma~\ref{lemma:descent}) that for every $t=0,1,2,\ldots$~: $$L^1\big(W_{1:N}(t+1)\big)\leq{L}^1\big(W_{1:N}(t)\big)\cdot\Big(1-\eta\cdot\sigma_{min}\big(W_{1:N}(t)\big)^{\frac{2(N-1)}{N}}\Big) \text{~.}$$ Since the coefficients $1-\eta\cdot\sigma_{min}(W_{1:N}(t))^{\frac{2(N-1)}{N}}$ are necessarily non-negative (otherwise would contradict non-negativity of $L^1(\cdot)$), we may unroll the inequalities, obtaining: \be L^1\big(W_{1:N}(t+1)\big)\leq{L}^1\big(W_{1:N}(0)\big)\cdot\prod\nolimits_{t'=0}^{t}\Big(1-\eta\cdot\sigma_{min}\big(W_{1:N}(t')\big)^{\frac{2(N-1)}{N}}\Big) \text{\,.} \label{eq:descent_concat} \ee Now, this in particular means that for every~$t'=0,1,2,\ldots$~: $$L^1\big(W_{1:N}(t')\big)\leq{L}^1\big(W_{1:N}(0)\big)\quad\implies\quad\norm{W_{1:N}(t')-\Phi}_F\leq\norm{W_{1:N}(0)-\Phi}_F \text{\,.}$$ Deficiency margin~$c$ of~$W_{1:N}(0)$ along with Claim~\ref{claim:margin_interp} thus imply $\sigma_{min}\big(W_{1:N}(t')\big)\geq{c}$, which when inserted back into Equation~\eqref{eq:descent_concat} yields, for every~$t=1,2,3,\ldots$~: \be L^1\big(W_{1:N}(t)\big)\leq{L}^1\big(W_{1:N}(0)\big)\cdot\Big(1-\eta\cdot{c}^{\frac{2(N-1)}{N}}\Big)^t \label{eq:descent_geo} \text{~.} \ee $\eta\cdot{c}^{\frac{2(N-1)}{N}}$ is obviously non-negative, and it is also no greater than~$1$ (otherwise would contradict non-negativity of $L^1(\cdot)$). We may therefore incorporate the inequality $1-\eta\cdot{c}^{2(N-1)/N}\leq\exp\big(-\eta\cdot{c}^{2(N-1)/N}\big)$ into Equation~\eqref{eq:descent_geo}: $$L^1\big(W_{1:N}(t)\big)\leq{L}^1\big(W_{1:N}(0)\big)\cdot\exp\big(-\eta\cdot{c}^{2(N-1)/N}\cdot{t}\big) \text{\,,}$$ from which it follows that $L^1(W_{1:N}(t))\leq\epsilon$ if: $$t\geq\frac{1}{\eta\cdot{c}^{2(N-1)/N}}\cdot\log\left(\frac{L^1(W_{1:N}(0))}{\epsilon}\right) \text{\,.}$$ Recalling again that $\ell(t)=L^1(W_{1:N}(t))$ (Equation~\eqref{eq:gd_loss}), we conclude the proof. \end{proof} \subsection{Balanced Initialization} \label{sec:converge:balance_init} We define the following procedure, \emph{balanced initialization}, which assigns weights randomly while ensuring perfect balancedness: \vspace{1mm} \begin{procedure}[Balanced initialization] \label{proc:balance_init} Given $d_0, d_1, \ldots, d_N \in \N$ such that $\min\{d_1,\ldots,d_{N-1}\}\geq\min\{d_0,d_N\}$ and a distribution~$\D$ over $d_N\times{d}_0$~matrices, a \emph{balanced initialization} of~$W_j\in\R^{d_j\times{d}_{j-1}}$, $j{=}1,\ldots,N$, assigns these weights as follows: \vspace{-1.5mm} \begin{enumerate}[label=(\roman*)] \item Sample $A\in\R^{d_N\times{d}_0}$ according to~$\D$. \vspace{-1mm} \item Take singular value decomposition~$A=U\Sigma{V}^\top$, where $U \in \R^{d_N \times \min\{d_0, d_N\}}$, $V \in \R^{d_0 \times \min\{d_0, d_N\}}$ have orthonormal columns, and $\Sigma \in \R^{\min\{d_0, d_N\} \times \min\{d_0, d_N\}}$ is diagonal and holds the singular values of $A$. \vspace{-1mm} \item Set $W_N\simeq{U}\Sigma^{1/N},W_{N-1}\simeq\Sigma^{1/N},\ldots,W_2\simeq\Sigma^{1/N},W_1\simeq\Sigma^{1/N}V^\top$, where the symbol~``$\simeq$'' stands for equality up to zero-valued padding.\note{ These assignments can be accomplished since $\min\{d_1,\ldots,d_{N-1}\}\geq\min\{d_0,d_N\}$. } \note{ By design $W_{1:N}=A$ and $W_{j+1}^{\top}W_{j+1}=W_{j}W_j^\top$, $\forall{j}\in\{1,\ldots,N{-}1\}$~---~these properties are actually all we need in Theorem~\ref{theorem:converge_balance_init}, and step~\emph{(iii)} in Procedure~\ref{proc:balance_init} can be replaced by any assignment that meets them. \label{note:balance_init_props} } \end{enumerate} \end{procedure} The concept of balanced initialization, together with Theorem~\ref{theorem:converge}, leads to a guarantee for linear convergence (applicable to output dimension~$1$~---~scalar regression) that holds with constant probability over the randomness in initialization: \vspace{1mm} \begin{theorem} \label{theorem:converge_balance_init} For any constant $0<p<1/2$, there are constants $d_0',a>0$ \note{ As shown in the proof of the theorem (Appendix \ref{app:proofs:converge_balance_init}), $d_0',a>0$ can take on any pair of values for which: \emph{(i)}~$d_0'\geq20$; and~\emph{(ii)}~$\big(1 - 2\exp(-d_0'/16)\big)\big(3-4F(2/\sqrt{a/2})\big) \geq 2p$, where $F(\cdot)$ stands for the cumulative distribution function of the standard normal distribution. For example, if~$p=0.25$, it suffices to take any~$d_0'\geq100,a\geq100$. We note that condition~\emph{(i)} here ($d_0'\geq20$) serves solely for simplification of expressions in the theorem. } such that the following holds. Assume $d_N=1,d_0\geq d_0'$, and that the weights $W_1(0),\ldots,W_N(0)$ are subject to balanced initialization (Procedure~\ref{proc:balance_init}) such that the entries in $W_{1:N}(0)$ are independent zero-centered Gaussian perturbations with standard deviation $s\leq \|\Phi\|_2/\sqrt{ad_0^2}$. Suppose also that we run gradient descent with learning rate $ \eta \leq (s^2d_0)^{4-2/N}\big/\big(10^5 N^3 \| \Phi\|_2^{10-6/N}\big)$. Then, with probability at least~$p$ over the random initialization, we have that for every~$\epsilon>0$ and: $$ T \geq \frac4\eta \left( \ln(4) \left(\frac{\|\Phi\|_2}{s^2d_0}\right)^{2-2/N} + \| \Phi\|_2^{2/N-2}\ln(\| \Phi\|_2^2/(8\ep)) \right) \text{\,,} $$ the loss at iteration~$T$ of gradient descent~---~$\ell(T)$~---~is no greater than~$\epsilon$. \end{theorem} \vspace{-2mm} \begin{proof} See Appendix \ref{app:proofs:converge_balance_init}. \end{proof} \section{Experiments} \label{sec:exper} Balanced initialization (Procedure~\ref{proc:balance_init}) possesses theoretical advantages compared with the customary layer-wise independent scheme~---~it allowed us to derive a convergence guarantee that holds with constant probability over the randomness of initialization (Theorem~\ref{theorem:converge_balance_init}). In this section we present empirical evidence suggesting that initializing with balancedness may be beneficial in practice as well. For conciseness, some of the details behind our implementation are deferred to Appendix~\ref{app:impl}. We began by experimenting in the setting covered by our analysis~---~linear neural networks trained via gradient descent minimization of $\ell_2$~loss over whitened data. The dataset chosen for the experiment was UCI Machine Learning Repository's ``Gas Sensor Array Drift at Different Concentrations'' \citep{vergara2012chemical,rodriguez2014calibration}. Specifically, we used the dataset's ``Ethanol'' problem~---~a scalar regression task with~$2565$ examples, each comprising~$128$ features (one of the largest numeric regression tasks in the repository). Starting with the customary initialization of layer-wise independent random Gaussian perturbations centered at zero, we trained a three layer network ($N=3$) with hidden widths ($d_1,d_2$) set to~$32$, and measured the time (number of iterations) it takes to converge (reach training loss within $\epsilon=10^{-5}$ from optimum) under different choices of standard deviation for the initialization. To account for the possibility of different standard deviations requiring different learning rates (values for~$\eta$), we applied, for each standard deviation independently, a grid search over learning rates, and recorded the one that led to fastest convergence. The result of this test is presented in Figure~\ref{fig:exper}(a). As can be seen, there is a range of standard deviations that leads to fast convergence (a few hundred iterations or less), below and above which optimization decelerates by orders of magnitude. This accords with our discussion at the end of Subsection~\ref{sec:converge:balance_init}, by which overly small initialization ensures approximate balancedness (small~$\delta$; see Definition~\ref{def:balance}) but diminishes deficiency margin (small~$c$; see Definition~\ref{def:margin})~---~``vanishing gradient problem''~---~whereas large initialization hinders both approximate balancedness and deficiency margin~---~``exploding gradient problem''. In that regard, as a sanity test for the validity of our analysis, in a case where approximate balancedness is met at initialization (small standard deviation), we measured its persistence throughout optimization. As Figure~\ref{fig:exper}(c) shows, our theoretical findings manifest themselves here~---~trajectories of gradient descent indeed preserve weight balancedness. In addition to a three layer network, we also evaluated a deeper, eight layer model (with hidden widths identical to the former~---~$N=8$, $d_1=\cdots=d_7=32$). In particular, using the same experimental protocol as above, we measured convergence time under different choices of standard deviation for the initialization. Figure~\ref{fig:exper}(a) displays the result of this test alongside that of the three layer model. As the figure shows, transitioning from three layers to eight aggravated the instability with respect to initialization~---~there is now a narrow band of standard deviations that lead to convergence in reasonable time, and outside of this band convergence is extremely slow, to the point where it does not take place within the duration we allowed ($10^6$~iterations). From the perspective of our analysis, a possible explanation for the aggravation is as follows: under layer-wise independent initialization, the magnitude of the end-to-end matrix~$W_{1:N}$ depends on the standard deviation in a manner that is exponential in depth, thus for large depths the range of standard deviations that lead to moderately sized~$W_{1:N}$ (as required for a deficiency margin) is limited, and within this range, there may not be many standard deviations small enough to ensure approximate balancedness. The procedure of balanced initialization (Procedure~\ref{proc:balance_init}) circumvents these difficulties~---~it assigns~$W_{1:N}$ directly (no exponential dependence on depth), and distributes its content between the individual weights $W_1,\ldots,W_N$ in a perfectly balanced fashion. Rerunning the experiment of Figure~\ref{fig:exper}(a) with this initialization replacing the customary layer-wise scheme (using same experimental protocol), we obtained the results shown in Figure~\ref{fig:exper}(b)~---~both the original three layer network, and the deeper eight layer model, converged quickly under virtually all standard deviations tried. As a final experiment, we evaluated the effect of balanced initialization in a setting that involves non-linear activation, softmax-cross-entropy loss and stochastic optimization (factors not accounted for by our analysis). For this purpose, we turned to the MNIST tutorial built into TensorFlow \citep{abadi2016tensorflow},\note{ \url{https://github.com/tensorflow/tensorflow/tree/master/tensorflow/examples/tutorials/mnist} } which comprises a fully-connected neural network with two hidden layers (width~$128$ followed by~$32$) and ReLU activation \citep{nair2010rectified}, trained through stochastic gradient descent (over softmax-cross-entropy loss) with batch size~$100$, initialized via customary layer-wise independent Gaussian perturbations centered at zero. While keeping the learning rate at its default value $0.01$, we varied the standard deviation of initialization, and for each value measured the training loss after $10$~epochs.\note{ As opposed to the dataset used in our experiments with linear networks, measuring the training loss with MNIST is non-trivial computationally (involves passing through $60K$~examples). Therefore, rather than continuously polling training loss until it reaches a certain threshold, in this experiment we chose to evaluate speed of convergence by measuring the training loss once after a predetermined number of iterations. } We then replaced the original (layer-wise independent) initialization with a balanced initialization based on Gaussian perturbations centered at zero (latter was implemented per Procedure~\ref{proc:balance_init}, disregarding non-linear activation), and repeated the process. The results of this experiment are shown in Figure~\ref{fig:exper}(d). Although our theoretical analysis does not cover non-linear activation, softmax-cross-entropy loss or stochasticity in optimization, its conclusion of balanced initialization leading to improved (faster and more stable) convergence carried over to such setting. \begin{figure} \vspace{-5mm} \subfloat[]{\includegraphics[scale=0.21]{std_vs_convtime_unbalanced.png}} \ \subfloat[]{\includegraphics[scale=0.21]{std_vs_convtime_balanced.png}} \ \subfloat[]{\includegraphics[scale=0.21]{{unbalanced_shallow_balances}.png}} \ \subfloat[]{\includegraphics[scale=0.21]{mnist_std_train_loss.png}} \ \vspace{-1mm} \caption{ Experimental results. \textbf{(a)}~Convergence of gradient descent training deep linear neural networks (depths~$3$ and~$8$) under customary initialization of layer-wise independent Gaussian perturbations with mean~$0$ and standard deviation~$s$. For each network, number of iterations required to reach $\epsilon=10^{-5}$ from optimal training loss is plotted as a function of~$s$ (missing values indicate no convergence within $10^6$~iterations). Dataset in this experiment is a numeric regression task from UCI Machine Learning Repository (details in text). Notice that fast convergence is attained only in a narrow band of values for~$s$, and that this phenomenon is more extreme with the deeper network. \textbf{(b)}~Same setup as in~(a), but with layer-wise independent initialization replaced by balanced initialization (Procedure~\ref{proc:balance_init}) based on Gaussian perturbations with mean~$0$ and standard deviation~$s$. Notice that this change leads to fast convergence, for both networks, under wide range of values for~$s$. Notice also that the shallower network converges slightly faster, in line with the results of~\citet{saxe2014exact} and~\citet{arora2018optimization} for $\ell_2$~loss. \textbf{(c)}~For the run in~(a) of a depth-$3$ network and standard deviation~$s={10^{-3}}$, this plot shows degree of balancedness (minimal~$\delta$ satisfying $\|W_{j+1}^{\top}W_{j+1}-W_{j}W_j^\top\|_F\leq\delta~,\,\forall{j}\in\{1,\ldots,N-1\}$) against magnitude of weights ($\min_{j=1,\ldots,N}\|W_{j}W_j^\top\|_F$) throughout optimization. Notice that approximate balancedness persists under gradient descent, in line with our theoretical analysis. \textbf{(d)}~Convergence of stochastic gradient descent training the fully-connected non-linear (ReLU) neural network of the MNIST tutorial built into TensorFlow (details in text). Customary layer-wise independent and balanced initializations~---~both based on Gaussian perturbations centered at zero~---~are evaluated, with varying standard deviations. For each configuration~$10$ epochs of optimization are run, followed by measurement of the training loss. Notice that although our theoretical analysis does not cover non-linear activation, softmax-cross-entropy loss and stochastic optimization, the conclusion of balanced initialization leading to improved convergence carries over to this setting. } \label{fig:exper} \vspace{-3mm} \end{figure} \section{Introduction} \label{sec:intro} Deep learning builds upon the mysterious ability of gradient-based optimization methods to solve related non-convex problems. Immense efforts are underway to mathematically analyze this phenomenon. The prominent \emph{landscape} approach focuses on special properties of critical points (\ie~points where the gradient of the objective function vanishes) that will imply convergence to global optimum. Several papers (\eg~\citet{ge2015escaping,lee2016gradient}) have shown that (given certain smoothness properties) it suffices for critical points to meet the following two conditions: \emph{(i)}~\emph{no poor local minima}~---~every local minimum is close in its objective value to a global minimum; and~\emph{(ii)}~\emph{strict saddle property}~---~every critical point that is not a local minimum has at least one negative eigenvalue to its Hessian. While condition~\emph{(i)} does not always hold (\cf~\citet{safran2018spurious}), it has been established for various simple settings (\eg~\citet{soudry2016no,kawaguchi2016deep}). Condition~\emph{(ii)} on the other hand seems less plausible, and is in fact provably false for models with three or more layers (\cf~\citet{kawaguchi2016deep}), \ie~for \emph{deep} networks. It has only been established for problems involving \emph{shallow} (two layer) models, \eg~matrix factorization (\citet{ge2016matrix,du2018algorithmic}). The landscape approach as currently construed thus suffers from inherent limitations in proving convergence to global minimum for deep networks. A potential path to circumvent this obstacle lies in realizing that landscape properties matter only in the vicinity of trajectories that can be taken by the optimizer, which may be a negligible portion of the overall parameter space. Several papers (\eg~\citet{saxe2014exact,arora2018optimization}) have taken this \emph{trajectory-based} approach, primarily in the context of \emph{linear neural networks}~---~fully-connected neural networks with linear activation. Linear networks are trivial from a representational perspective, but not so in terms of optimization~---~they lead to non-convex training problems with multiple minima and saddle points. Through a mix of theory and experiments, \citet{arora2018optimization}~argued that such non-convexities may in fact be beneficial for gradient descent, in the sense that sometimes, adding (redundant) linear layers to a classic linear prediction model can accelerate the optimization. This phenomenon challenges the holistic landscape view, by which convex problems are always preferable to non-convex ones. Even in the linear network setting, a rigorous proof of efficient convergence to global minimum has proved elusive. One recent progress is the analysis of~\citet{bartlett2018gradient} for \emph{linear residual networks}~---~a particular subclass of linear neural networks in which the input, output and all hidden dimensions are equal, and all layers are initialized to be the identity matrix (\cf~\citet{hardt2016identity}). Through a trajectory-based analysis of gradient descent minimizing $\ell_2$~loss over a whitened dataset (see Section~\ref{sec:prelim}), \citet{bartlett2018gradient}~show that convergence to global minimum at a \emph{linear rate}~---~loss is less than~$\epsilon>0$ after $\OO(\log\frac{1}{\epsilon})$~iterations~---~takes place if one of the following holds: \emph{(i)}~the objective value at initialization is sufficiently close to a global minimum; or \emph{(ii)}~a global minimum is attained when the product of all layers is positive definite. The current paper carries out a trajectory-based analysis of gradient descent for general deep linear neural networks, covering the residual setting of~\citet{bartlett2018gradient}, as well as many more settings that better match practical deep learning. Our analysis draws upon the trajectory characterization of~\citet{arora2018optimization} for gradient flow (infinitesimally small learning rate), together with significant new ideas necessitated due to discrete updates. Ultimately, we show that when minimizing $\ell_2$~loss of a deep linear network over a whitened dataset, gradient descent converges to the global minimum, at a linear rate, provided that the following conditions hold: \emph{(i)}~the dimensions of hidden layers are greater than or equal to the minimum between those of the input and output; \emph{(ii)}~layers are initialized to be \emph{approximately balanced} (see Definition~\ref{def:balance})~---~this is met under commonplace near-zero, as well as residual (identity) initializations; and \emph{(iii)}~the initial loss is smaller than any loss obtainable with rank deficiencies~---~this condition will hold with probability close to~$0.5$ if the output dimension is~$1$ (scalar regression) and standard (random) near-zero initialization is employed. Our result applies to networks with arbitrary depth and input/output dimensions, as well as any configuration of hidden layer widths that does not force rank deficiency (\ie~that meets condition~\emph{(i)}). The assumptions on initialization (conditions~\emph{(ii)} and~\emph{(iii)}) are necessary, in the sense that violating any one of them may lead to convergence failure. Moreover, in the case of scalar regression, they are met with constant probability under a random initialization scheme. We are not aware of any similarly general analysis for efficient convergence of gradient descent to global minimum in deep learning. \medskip The remainder of the paper is organized as follows. In Section~\ref{sec:prelim} we present the problem of gradient descent training a deep linear neural network by minimizing the $\ell_2$~loss over a whitened dataset. Section~\ref{sec:converge} formally states our assumptions, and presents our convergence analysis. Key ideas brought forth by our analysis are demonstrated empirically in Section~\ref{sec:exper}. Section~\ref{sec:related} gives a review of relevant literature, including a detailed comparison of our results against those of~\citet{bartlett2018gradient}. Finally, Section~\ref{sec:conc} concludes. \section{Gradient Descent for Deep Linear Neural Networks} \label{sec:prelim} We denote by~$\norm{\vv}$ the Euclidean norm of a vector~$\vv$, and by~$\norm{A}_F$ the Frobenius norm of a matrix~$A$. We are given a training set $\{(\x^{(i)},\y^{(i)})\}_{i=1}^{m}\subset\R^{d_x}\times\R^{d_y}$, and would like to learn a hypothesis (predictor) from a parametric family~$\HH:=\{h_\theta:\R^{d_x}\to\R^{d_y}\,|\,\theta\in\Theta\}$ by minimizing the $\ell_2$~loss:\note{ Much of the analysis in this paper can be extended to loss types other than~$\ell_2$. In particular, the notion of deficiency margin (Definition~\ref{def:margin}) can be generalized to account for any convex loss, and, so long as the loss is differentiable, a convergence result analogous to Theorem~\ref{theorem:converge} will hold in the idealized setting of perfect initial balancedness and infinitesimally small learning rate (see proof of Lemma~\ref{lemma:descent}). We leave to future work treatment of approximate balancedness and discrete updates in this general setting. } $$\min_{\theta\in\Theta}~L(\theta):=\frac{1}{2m}\sum\nolimits_{i=1}^m\|h_\theta(\x^{(i)})-\y^{(i)}\|^2 \text{\,.}$$ When the parametric family in question is the class of linear predictors, \ie~$\HH=\{\x\mapsto{W}\x\,|\,W\in\R^{d_y\times{d}_x}\}$, the training loss may be written as $L(W)=\frac{1}{2m}\|WX-Y\|_F^2$, where~$X\in\R^{d_x\times{m}}$ and~$Y\in\R^{d_y\times{m}}$ are matrices whose columns hold instances and labels respectively. Suppose now that the dataset is \emph{whitened}, \ie~has been transformed such that the empirical (uncentered) covariance matrix for instances~---~$\Lambda_{xx}:=\frac{1}{m}XX^\top \in\R^{d_x\times{d}_x}$~---~is equal to identity. Standard calculations (see Appendix~\ref{app:whitened}) show that in this case: \be L(W)=\frac{1}{2}\norm{W-\Lambda_{yx}}_F^2+c \label{eq:loss_whitened} \text{\,,} \ee where~$\Lambda_{yx}:=\tfrac{1}{m}YX^\top\in\R^{d_y\times{d}_x}$ is the empirical (uncentered) cross-covariance matrix between instances and labels, and $c$ is a constant (that does not depend on $W$). Denoting~$\Phi:=\Lambda_{yx}$ for brevity, we have that for linear models, minimizing $\ell_2$~loss over whitened data is equivalent to minimizing the squared Frobenius distance from a \emph{target matrix}~$\Phi$: \be \min\nolimits_{W\in\R^{d_y\times{d}_x}}~L^1(W):=\frac{1}{2}\norm{W-\Phi}_F^2 \text{\,.} \label{eq:lin_obj} \ee Our interest in this work lies on \emph{linear neural networks}~---~fully-connected neural networks with linear activation. A depth-$N$ ($N\in\N$) linear neural network with hidden widths $d_1,\ldots,d_{N-1}\in\N$ corresponds to the parametric family of hypotheses $\HH:=\{\x\mapsto{W}_{N}W_{N-1}\cdots{W}_1\x\,|\,W_j\in\R^{d_j\times{d}_{j-1}},\,j=1,\ldots,N\}$, where $d_0:=d_x,\,d_N:=d_y$. Similarly to the case of a (directly parameterized) linear predictor (Equation~\eqref{eq:lin_obj}), with a linear neural network, minimizing $\ell_2$~loss over whitened data can be cast as squared Frobenius approximation of a target matrix $\Phi$: \be \min\nolimits_{W_j\in\R^{d_j\times{d}_{j-1}},\,j=1,\ldots,N}~L^N(W_1,\ldots,W_N):=\frac{1}{2}\norm{W_{N}W_{N-1}\cdots{W}_1-\Phi}_F^2 \label{eq:lnn_obj} \text{\,.} \ee Note that the notation~$L^N(\cdot)$ is consistent with that of Equation~\eqref{eq:lin_obj}, as a network with depth~$N=1$ precisely reduces to a (directly parameterized) linear model. We focus on studying the process of training a deep linear neural network by \emph{gradient descent}, \ie~of tackling the optimization problem in Equation~\eqref{eq:lnn_obj} by iteratively applying the following updates: \be W_j(t+1)\leftarrow{W}_j(t)-\eta\frac{\partial{L}^N}{\partial{W}_j}\big(W_1(t),\ldots,W_N(t)\big)\quad,\,j=1,\ldots,{N}\quad,\,t=0,1,2,\ldots \label{eq:gd} \text{~~,} \ee where $\eta>0$ is a configurable learning rate. In the case of depth~$N=1$, the training problem in Equation~\eqref{eq:lnn_obj} is smooth and strongly convex, thus it is known (\cf~\citet{boyd2004convex}) that with proper choice of~$\eta$, gradient descent converges to global minimum at a linear rate. In contrast, for any depth greater than~$1$, Equation~\eqref{eq:lnn_obj} comprises a fundamentally non-convex program, and the convergence properties of gradient descent are highly non-trivial. Apart from the case~$N=2$ (shallow network), one cannot hope to prove convergence via landscape arguments, as the strict saddle property is provably violated (see Section~\ref{sec:intro}). We will see in Section~\ref{sec:converge} that a direct analysis of the trajectories taken by gradient descent can succeed in this arena, providing a guarantee for linear rate convergence to global minimum. \medskip We close this section by introducing additional notation that will be used in our analysis. For an arbitrary matrix~$A$, we denote by~$\sigma_{max}(A)$ and~$\sigma_{min}(A)$ its largest and smallest (respectively) singular values.\note{ If~$A\in\R^{d\times{d'}}$, $\sigma_{min}(A)$~stands for the $\min\{d,d'\}$-th largest singular value. Recall that singular values are always non-negative. } For~$d\in\N$, we use~$I_d$ to signify the identity matrix in~$\R^{d\times{d}}$. Given weights $W_1,\ldots,W_N$ of a linear neural network, we let~$W_{1:N}$ be the direct parameterization of the \emph{end-to-end} linear mapping realized by the network, \ie~$W_{1:N}:=W_{N}W_{N-1}\cdots{W}_1$. Note that $L^N(W_1,\ldots,W_N)=L^1(W_{1:N})$, meaning the loss associated with a depth-$N$ network is equal to the loss of the corresponding end-to-end linear model. In the context of gradient descent, we will oftentimes use~$\ell(t)$ as shorthand for the loss at iteration~$t$: \be \ell(t):=L^N(W_1(t),\ldots,W_N(t))=L^1(W_{1:N}(t)) \text{\,.} \label{eq:gd_loss} \ee \section{Related Work} \label{sec:related} \vspace{-2mm} Theoretical study of gradient-based optimization in deep learning is a highly active area of research. As discussed in Section~\ref{sec:intro}, a popular approach is to show that the objective landscape admits the properties of no poor local minima and strict saddle, which, by~\citet{ge2015escaping,lee2016gradient,panageas2017gradient}, ensure convergence to global minimum. Many works, both classic (\eg~\citet{baldi1989neural}) and recent (\eg~\citet{choromanska2015loss,kawaguchi2016deep,hardt2016identity,soudry2016no,haeffele2017global,nguyen2017loss,safran2018spurious,nguyen2018loss,laurent2018deep}), have focused on the validity of these properties in different deep learning settings. Nonetheless, to our knowledge, the success of landscape-driven analyses in formally proving convergence to global minimum for a gradient-based algorithm, has thus far been limited to shallow (two layer) models only (\eg~\citet{ge2016matrix,du2018power,du2018algorithmic}). An alternative to the landscape approach is a direct analysis of the trajectories taken by the optimizer. Various papers (\eg~\citet{brutzkus2017globally,li2017convergence,zhong2017recovery,tian2017analytical,brutzkus2018sgd,li2018algorithmic,du2018gradient,du2018convolutional,liao2018almost}) have recently adopted this strategy, but their analyses only apply to shallow models. In the context of linear neural networks, deep (three or more layer) models have also been treated~---~\cf~\citet{saxe2014exact} and~\citet{arora2018optimization}, from which we draw certain technical ideas for proving Lemma~\ref{lemma:descent}. However these treatments all apply to gradient flow (gradient descent with infinitesimally small learning rate), and thus do not formally address the question of computational efficiency. To our knowledge, \citet{bartlett2018gradient}~is the only existing work rigorously proving convergence to global minimum for a conventional gradient-based algorithm training a deep model. This work is similar to ours in the sense that it also treats linear neural networks trained via minimization of $\ell_2$~loss over whitened data, and proves linear convergence (to global minimum) for gradient descent. It is more limited in that it only covers the subclass of linear residual networks, \ie~the specific setting of uniform width across all layers ($d_0=\cdots=d_N$) along with identity initialization. We on the other hand allow the input, output and hidden dimensions to take on any configuration that avoids ``bottlenecks'' (\ie~admits $\min\{d_1,\ldots,d_{N-1}\}\geq\min\{d_0,d_N\}$), and from initialization require only approximate balancedness (Definition~\ref{def:balance}), supporting many options beyond identity. In terms of the target matrix~$\Phi$, \citet{bartlett2018gradient}~treats two separate scenarios:\note{ There is actually an additional third scenario being treated~---~$\Phi$~is asymmetric and positive definite~---~but since that requires a dedicated optimization algorithm, it is outside our scope. } \emph{(i)}~$\Phi$ is symmetric and positive definite; and~\emph{(ii)}~$\Phi$ is within distance~$1/10e$ from identity.\note{ $1/10e$~is the optimal (largest) distance that may be obtained (via careful choice of constants) from the proof of Theorem~$1$ in~\citet{bartlett2018gradient}. } Our analysis does not fully account for scenario~\emph{(i)}, which seems to be somewhat of a singularity, where all layers are equal to each other throughout optimization (see proof of Theorem~$2$ in~\citet{bartlett2018gradient}). We do however provide a strict generalization of scenario~\emph{(ii)}~---~our assumption of deficiency margin (Definition~\ref{def:margin}), in the setting of linear residual networks, is met if the distance between target and identity is less than~$0.5$.
{ "timestamp": "2019-10-29T01:06:05", "yymm": "1810", "arxiv_id": "1810.02281", "language": "en", "url": "https://arxiv.org/abs/1810.02281", "abstract": "We analyze speed of convergence to global optimum for gradient descent training a deep linear neural network (parameterized as $x \\mapsto W_N W_{N-1} \\cdots W_1 x$) by minimizing the $\\ell_2$ loss over whitened data. Convergence at a linear rate is guaranteed when the following hold: (i) dimensions of hidden layers are at least the minimum of the input and output dimensions; (ii) weight matrices at initialization are approximately balanced; and (iii) the initial loss is smaller than the loss of any rank-deficient solution. The assumptions on initialization (conditions (ii) and (iii)) are necessary, in the sense that violating any one of them may lead to convergence failure. Moreover, in the important case of output dimension 1, i.e. scalar regression, they are met, and thus convergence to global optimum holds, with constant probability under a random initialization scheme. Our results significantly extend previous analyses, e.g., of deep linear residual networks (Bartlett et al., 2018).", "subjects": "Machine Learning (cs.LG); Neural and Evolutionary Computing (cs.NE); Machine Learning (stat.ML)", "title": "A Convergence Analysis of Gradient Descent for Deep Linear Neural Networks", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.987375049232425, "lm_q2_score": 0.8104789040926008, "lm_q1q2_score": 0.8002466478302737 }
https://arxiv.org/abs/2211.15338
Learning Integrable Dynamics with Action-Angle Networks
Machine learning has become increasingly popular for efficiently modelling the dynamics of complex physical systems, demonstrating a capability to learn effective models for dynamics which ignore redundant degrees of freedom. Learned simulators typically predict the evolution of the system in a step-by-step manner with numerical integration techniques. However, such models often suffer from instability over long roll-outs due to the accumulation of both estimation and integration error at each prediction step. Here, we propose an alternative construction for learned physical simulators that are inspired by the concept of action-angle coordinates from classical mechanics for describing integrable systems. We propose Action-Angle Networks, which learn a nonlinear transformation from input coordinates to the action-angle space, where evolution of the system is linear. Unlike traditional learned simulators, Action-Angle Networks do not employ any higher-order numerical integration methods, making them extremely efficient at modelling the dynamics of integrable physical systems.
\section{Appendix} \subsection{Baseline Models} \label{sec:baselines} \textbf{Euler Update Network}: This model updates the latent state via the forward Euler method: \begin{align*} \hat{u}(t) &\gets \text{Encode}(u(t)) \\ \hat{u}(t + \Delta t) &\gets \hat{u}(t) + \Delta t \cdot {\mathcal{F}}(\hat{u}(t)) \\ u(t + \Delta t) &\gets \text{Decode}(\hat{u}(t + \Delta t)) \end{align*} \textbf{Neural Ordinary Differential Equations}: A generalization of Euler Update Networks, Neural ODEs \citep{chen:node} have demonstrated state-of-the-art performance on several time-series forecasting problems. They usually use a higher-order numerical integration scheme to update the latent state: \vspace{-0.4em} \begin{align*} \hat{u}(t) &\gets \text{Encode}(u(t)) \\ \hat{u}(t + \Delta t) &\gets \hat{u}(t) + {\int_{t}^{t + \Delta t}}{\mathcal{F}}(\hat{u}(s)) ds \\ u(t + \Delta t) &\gets \text{Decode}(\hat{u}(t + \Delta t)) \end{align*} \textbf{Hamiltonian Neural Network}: Hamiltonian Neural Networks (HNNs) \citep{hnn} models the Hamiltonian in a latent space explicitly and updates the latent coordinates via Hamilton's equations. \begin{align*} \hat{q}(t), \hat{p}(t) &\gets \text{Encode}(u(t)) \\ \hat{q}(t + \Delta t) &\gets \hat{q}(t) + {\int_{t}^{t + \Delta t}} \pdd{}{\hat{p}}H(\hat{q}(s), \hat{p}(s)) ds \\ \hat{p}(t + \Delta t) &\gets \hat{p}(t) - {\int_{t}^{t + \Delta t}} \pdd{}{\hat{q}}H(\hat{q}(s), \hat{p}(s)) ds \\ u(t + \Delta t) &\gets \text{Decode}(\hat{q}(t + \Delta t), \hat{p}(t + \Delta t)) \end{align*} Wherever applicable, we used the Dormand-Prince 5(4) solver, a 5th order Runge-Kutta method for numerical integration. All of these baselines can technically simulate the Action-Angle Network by encoding the canonical coordinates into the action-angle coordinates to linearize the dynamics. \subsection{Harmonic Oscillators} \label{sec:harmonic} We model a system of $n$ point masses, which are connected to a wall via springs of constant $k_w$ and to each other via springs of constant $k_p$. This can be described by the set of $n$ differential equations: \begin{align} \label{eqn:harmonic} m \frac{d^2q_i}{dt^2} = -k_w q_i + {\sum_{j \neq i}} k_p (q_j - q_i). \end{align} where $q_i(t)$ is the position of the $i$th point mass. This forms a system of coupled harmonic oscillators, where $k_p$ controls the strength of the coupling. Any solution to \autoref{eqn:harmonic} is a linear combination of the `normal modes' of the system \citep{morin:normal-modes}: \begin{align*} q(t) = {\textstyle \sum_{r = 1}^n} A_r c_r \cos{(\omega_r t + \phi_r)}. \end{align*} The $n$ angular frequencies $\omega$ and coefficients $c$ are found by solving the following eigenvalue-eigenvector equation $(M + \omega^2 \mathbb{I}_n) c = 0$, where $M$ is the matrix defined as: \begin{align} M_{ij} = \begin{cases} -\dfrac{k_w + (n - 1)k_p}{m_i} \ &\text{if} \ i = j \\ \dfrac{k_p}{m_i} \ &\text{if} \ i \neq j \end{cases} \end{align} \section*{Checklist} \begin{enumerate} \item For all authors... \begin{enumerate} \item Do the main claims made in the abstract and introduction accurately reflect the paper's contributions and scope? \answerYes{We have designed and evaluated Action-Angle Networks, showing how they can efficiently learn the dynamics of integrable systems.} \item Did you describe the limitations of your work? \answerYes{Action-Angle Networks can be applied to only integrable systems. Their performance on non-integrable Hamiltonian systems still needs investigation.} \item Did you discuss any potential negative societal impacts of your work? \answerYes{We have mentioned an overview in the Broader Impact statement. We do not expect any negative societal impacts of our work. } \item Have you read the ethics review guidelines and ensured that your paper conforms to them? \answerYes{Our paper does not utilise or expose any human-derived data. All data for the physical systems discussed here are synthetically generated via numerical integration.} \end{enumerate} \item If you are including theoretical results... \begin{enumerate} \item Did you state the full set of assumptions of all theoretical results? \answerNA{} \item Did you include complete proofs of all theoretical results? \answerNA{} \end{enumerate} \item If you ran experiments... \begin{enumerate} \item Did you include the code, data, and instructions needed to reproduce the main experimental results (either in the supplemental material or as a URL)? \answerYes{Yes, we have added a link to our code in \autoref{sec:experiments}.} \item Did you specify all the training details (e.g., data splits, hyperparameters, how they were chosen)? \answerYes{Yes, the entire training pipeline and configuration is available at the link for the code in \autoref{sec:experiments}.} \item Did you report error bars (e.g., with respect to the random seed after running experiments multiple times)? \answerNo{We plan to address this in a future version.} \item Did you include the total amount of compute and the type of resources used (e.g., type of GPUs, internal cluster, or cloud provider)? \answerYes{Experiments were run on the CPU nodes of MIT SuperCloud, and can be easily reproduced with minimal resources.} \end{enumerate} \item If you are using existing assets (e.g., code, data, models) or curating/releasing new assets... \begin{enumerate} \item If your work uses existing assets, did you cite the creators? \answerNA{} \item Did you mention the license of the assets? \answerYes{Our code linked in \autoref{sec:experiments} is licensed under the Apache License 2.0.} \item Did you include any new assets either in the supplemental material or as a URL? \answerYes{We have released code and simulation tools at the link in \autoref{sec:experiments}.} \item Did you discuss whether and how consent was obtained from people whose data you're using/curating? \answerNA{} \item Did you discuss whether the data you are using/curating contains personally identifiable information or offensive content? \answerNA{} \end{enumerate} \item If you used crowdsourcing or conducted research with human subjects... \begin{enumerate} \item Did you include the full text of instructions given to participants and screenshots, if applicable? \answerNA{} \item Did you describe any potential participant risks, with links to Institutional Review Board (IRB) approvals, if applicable? \answerNA{} \item Did you include the estimated hourly wage paid to participants and the total amount spent on participant compensation? \answerNA{} \end{enumerate} \end{enumerate} \section{Conclusion} Our preliminary experiments indicate that Action-Angle Networks can be a promising alternative for learning efficient physical simulators. Action-Angle Networks overcome many of the obstacles -- instability and inefficiency -- faced by state-of-the-art learned simulators today. That being said, our experiments here only model very simple integrable systems. Next, we plan to analyse the performance of Action-Angle Networks on modelling large-scale non-integrable systems (such as cosmological simulations and molecular dynamics trajectories) to better understand their tradeoffs. \section{Experiments} \label{sec:experiments} We compare the Action-Angle Network to three strong baseline models: the Euler Update Network (EUN), the Neural Ordinary Differential Equations (Neural ODE) \citep{chen:node}, and the physics-inspired Hamiltonian Neural Networks (HNN) \citep{hnn}. A comparison of these models can be found in \autoref{tab:model_comparison}. The baseline models are further described in \autoref{sec:baselines}. \begin{table*}[htbp] \centering \caption{Comparing different models.} \label{tab:model_comparison} \rowcolors{2}{}{gray!10} \begin{tabular}{*5c} \toprule & Action-Angle Network & EUN & Neural ODE & HNN \\ \midrule Parameter count & $\approx 8.5$K & $\approx 9$K & $\approx 100$K & $\approx 200$K \\ Inference time & $O(1)$ & $O(1)$ & $O(\Delta t)$ & $O(\Delta t)$ \\ Learns conservation laws & \checkmark & & & \checkmark \\ Learns linear dynamics & \checkmark & \checkmark & & \\ \bottomrule \end{tabular} \end{table*} As detailed in \autoref{sec:harmonic}, we simulate a system of coupled harmonic oscillators.\footnote{We have created animations of the system trajectory and model predictions at \href{https://ameya98.github.io/ActionAngleNetworks/webpage/index.html}{this webpage}. Our codebase to run all experiments and analyses is available at \url{https://github.com/ameya98/ActionAngleNetworks}.} \autoref{fig:performance_vs_samples} depicts the prediction error for each of the models as a function of training samples, showing that the Action-Angle Network is much more data-efficient than the other baselines. \autoref{fig:inference_times} shows that the Action-Angle Network can be queried much faster than the Neural ODE and the HNN, with an inference time that is independent of $\Delta t$. \autoref{fig:performance_vs_time} depicts the prediction error as a function of $\Delta t$, showing that the Action-Angle Network also scales much better with the jump size $\Delta t$, even for jump sizes larger than those seen during training $\Delta t_{\max} = 10$. Finally, \autoref{fig:angular_frequencies} shows that the predicted angular frequencies from the Action-Angle Network closely match the true angular frequencies. \begin{figure}[htbp] \centering \begin{tabular}[c]{cc} \begin{subfigure}[c]{0.38\textwidth} \centering \includegraphics[width=\textwidth]{figures/action_angle_prediction_error.pdf} \caption{Action-Angle Network} \label{fig:action_angle_prediction_error} \end{subfigure} \hfill \begin{subfigure}[c]{0.38\textwidth} \centering \includegraphics[width=\textwidth]{figures/euler_update_prediction_error.pdf} \caption{Euler Update Network} \label{fig:euler_update_prediction_error} \end{subfigure} \end{tabular} \\ \begin{tabular}[c]{cc} \begin{subfigure}[b]{0.38\textwidth} \centering \includegraphics[width=\textwidth]{figures/neural_ode_prediction_error.pdf} \caption{Neural ODE} \label{fig:neural_ode_prediction_error} \end{subfigure} \hfill \begin{subfigure}[b]{0.38\textwidth} \centering \includegraphics[width=\textwidth]{figures/hamiltonian_network_prediction_error.pdf} \caption{Hamiltonian Neural Network} \label{fig:hamiltonian_network_prediction_error} \end{subfigure} \end{tabular} \caption{Prediction errors on test data as a function of training samples.} \label{fig:performance_vs_samples} \end{figure} \begin{figure}[htbp] \centering \begin{tabular}[c]{cc} \begin{subfigure}[b]{0.45\textwidth} \centering \includegraphics[width=\textwidth]{figures/inference_times.pdf} \caption{Inference times as a function of jump $\Delta t$.} \label{fig:inference_times} \end{subfigure} \hfill \begin{subfigure}[b]{0.42\textwidth} \centering \includegraphics[width=\textwidth]{figures/performance_vs_time.pdf} \caption{Prediction error as a function of jump $\Delta t$.} \label{fig:performance_vs_time} \end{subfigure} \end{tabular} \caption{Evaluating each model on (a) inference time and (b) prediction error.} \label{fig:evaluation} \end{figure} \begin{figure}[!htbp] \centering \begin{tabular}[c]{cc} \begin{subfigure}[c]{0.4\textwidth} \centering \includegraphics[width=\textwidth]{figures/angular_frequency_0.pdf} \end{subfigure} \hfill \begin{subfigure}[c]{0.4\textwidth} \centering \includegraphics[width=\textwidth]{figures/angular_frequency_1.pdf} \end{subfigure} \end{tabular} \caption{Kernel density plot of angular frequencies predicted by the Action-Angle Network. True angular frequencies are indicated by dashed vertical lines.} \label{fig:angular_frequencies} \end{figure} \section*{Broader Impact Statement} In this paper, we have proposed a novel method for modelling the dynamics of Hamiltonian systems, with improvements in computational efficiency. There has been much interest in accelerating simulations of various kinds across scientific domains that could have impact on human lives (for example, simulating biological cell cycles and immune system responses). However, given the techniques here apply to highly-constrained physical systems, we do not anticipate any negative social implications of our work. \section{Introduction} \label{sec:intro} \subsection{Modelling Hamiltonian Systems} Hamiltonian systems are an important class of physical systems whose dynamics are governed by a scalar function $H$, called the Hamiltonian. The state of a Hamiltonian system is a 2-tuple $u(t) = (q(t), p(t))$, where $q \in \mathbb{R}^n$ are the positions and $p \in \mathbb{R}^n$ are the canonical momenta. We are interested in predicting the time evolution of a Hamiltonian system as a function of time $t$. In particular, we seek to learn a model ${\mathcal{M}}_\phi$ (parameterized by $\phi$) that can accurately predict the future state $u(t + \Delta t)$ given the current state $u(t)$: \vspace{-0.2em} \begin{align} {\mathcal{M}}_\phi(u(t), \Delta t) \approx u(t + \Delta t). \end{align} Previous efforts towards modelling Hamiltonian systems \citep{hnn,lnn,sympnets,chen:node,kidger:nde} have seen success with neural networks trained via backpropagation on the trajectory prediction objective: \begin{align*} \phi^* = \argmin_\phi \ \sum_{t} \norm{{\mathcal{M}}_\phi(u(t), \Delta t) - u(t + \Delta t))}^2. \end{align*} Such models are often constrained in some way to match the underlying physical evolution, which improves their accuracy. However, they suffer from several drawbacks that have restricted their effectiveness: \textbf{(1)} Their predictions tend to be unstable over long roll-outs (when $\Delta t$ is large). \textbf{(2)} They tend to require many parameters and need long training times. Further, the inference times of these models often scale with $\Delta t$. Based on the principle of \emph{action-angle coordinates} \citep{arnol'd} from classical mechanics, we introduce a new paradigm for learning physical simulators which incorporate an inductive bias for learning \emph{integrable} dynamics. Our Action-Angle Network learns an invertible transformation to action-angle coordinates to linearize the dynamics in this space. In this sense, the Action-Angle Network can be seen as a physics-informed adaptation of DeepKoopman \citep{lusch:koopman}. The Action-Angle Network is efficient both in parameter count and inference cost; it requires much fewer parameters than previous methods to reach similar performance and enjoys an inference time independent of $\Delta t$. \subsection{Action-Angle Coordinates} The time evolution of any Hamiltonian system is given by Hamilton's equations: \vspace{-0.2em} \begin{align} \dd{q}{t} = \pdd{H}{p}, \quad \dd{p}{t} = -\pdd{H}{q} \vspace{-0.5em} \end{align} where $H$ is the Hamiltonian of the system. The complexity for learning these dynamics arises solely because $H$ is not known and must be inferred; only samples from a trajectory over which $H$ is conserved is available. The core issue for all physical simulators is the non-linearity of the dynamics when expressed in the canonical coordinates $(q, p)$. However, for \emph{integrable} systems \citep{tong:dynamics} which possess significant symmetry, the dynamics are actually linear in a set of coordinates termed the \emph{action-angle} coordinates \citep{arnol'd}. The actions $I$ and angles $\theta$ are related to the canonical coordinates $(q, p)$ via a symplectic transformation ${\mathcal{T}}$ \citep{symplectic:meyer2013introduction} as $(I, \theta) = {\mathcal{T}}(q, p)$. In the action-angle coordinates, the Hamiltonian only depends on the actions $I$, not the angles $\theta$. Thus, Hamilton's equations in this basis tell us that: \vspace{-0.2em} \begin{align} \dd{I}{t} = -\pdd{H}{\theta} = 0, \quad \dd{\theta}{t} = \pdd{H}{I}. \label{eqn:theta-dot} \end{align} Thus, the actions $I$ are always constant across a trajectory, while the angles evolve linearly with constant rate $\dot{\theta} = \frac{\partial H}{\partial I}$. The action-angle space can be thought of as a torus $\mathbb{T}^n$, where the actions $I$ are a function of the radii, and the angles describe the individual phases living in $[0, 2\pi)$. Learning the mapping ${\mathcal{T}}$ from canonical coordinates $(q, p)$ to action-angle coordinates $(I, \theta)$ for several physical systems was first explored in \citep{learning-symmetries}, which did not focus on the complete dynamics of integrable systems. We leverage their framework to additionally incorporate a dynamics model ${\mathcal{F}}$ to learn a complete physical simulator. \cite{ishikawa:integrable} attempt to model the time evolution of actions for integrable systems by learning the Hamiltonian in action-angle coordinates. However, their overall objective and training setup are different from ours: they have samples of the true action-angle coordinates $(I, \theta)$ as input, while we only observe the canonical coordinates $(q, p)$. \section{Action-Angle Networks} \label{sec:method} \begin{figure}[htbp] \centering \includegraphics[width=0.78\textwidth]{figures/model_diagram.png} \caption{An overview of the Action-Angle Network.} \label{fig:model_diagram} \end{figure} Suppose we observe the system at time $t$, and we query our model ${\mathcal{M}}_\phi$ to predict the state at a future time $t + \Delta t$. The Action-Angle Network performs the following operations, illustrated in \autoref{fig:model_diagram}: \begin{itemize}[leftmargin=*] \item \textbf{Encode}: Convert the current state in canonical coordinates $q(t), p(t)$ to action-angle coordinates $I(t), \theta(t)$ via a learned map ${\mathcal{T}}$. \item \textbf{Evolve}: Compute the angular velocities $\dot{\theta}(t)$ at this instant with the dynamics model ${\mathcal{F}}$: \\$\dot{\theta}(t) \gets {\mathcal{F}}(I(t))$. Evolve the angles to time $t + \Delta t$ with these angular velocities: \begin{align} \theta(t + \Delta t) \gets \theta(t) + \dot{\theta}(t) \cdot \Delta t \pmod{2\pi} \end{align} \vspace{-1.5em} \item \textbf{Decode}: Convert the new state back to canonical coordinates $q(t + \Delta t), p(t + \Delta t)$ via ${\mathcal{T}}^{-1}$. \end{itemize} \textbf{Symplectic Normalizing Flows}: The transformation ${\mathcal{T}}$ from canonical coordinates to action-angle coordinates is guaranteed to be symplectic \citep{arnol'd}. To constrain our encoder to learn only symplectic transformations, we compose layers of symplectic normalizing flows \citep{sympflow,learning-symmetries,sympnets}. In particular, we found that the G-SympNet layers \citep{sympnets} with our modification described below performed well empirically. G-SympNet was shown to be universal; they can represent any symplectic transformation given sufficient width and depth. Similar to affine coupling layers \citep{nice}, each G-SympNet layer $\psi_i$ operates on only one of the coordinates keeping the other fixed, depending on the parity of $i$: \begin{align} \psi_{2k}(q, p) = (q, p + f(q)), \quad \psi_{2k + 1}(q, p) = (q + f(p), p) \end{align} where $f$ is of the following form $f(x) = Cx + W^T \text{diag}(A)\sigma(Wx + B)$ with learnable parameters $A \in \mathbb{R}^{d_o}, B \in \mathbb{R}^{d_o}, C \in \mathbb{R}$ and $W \in \mathbb{R}^{d_o \times n}$, where $d_o$ is a hyperparameter. Our minor modification above allows each layer to model the identity transformation, enabling the training of deeper models. \textbf{Action-Angles via Polar Coordinates}: We found that learning a mapping from canonical coordinates which live in $\mathbb{R}^{2n}$ directly to action-angle coordinates which live in the torus $\mathbb{T}^n$ was challenging for the network. We hypothesize that the differing topology of these spaces is a major obstacle because of the resulting singularities. To bypass this, we borrow a trick from \cite{learning-symmetries}; we have the G-SympNet instead output the components of the actions $(I^{(x)}, I^{(y)})$ in the Cartesian coordinate basis. These are then converted to action-angles $(I, \theta)$ in the polar coordinate basis, by the standard transformation ${\mathcal{T}}_{\text{polar}}$ applied to each pair of $(I^{(x)}_i, I^{(y)}_i)$ coordinates: \begin{align} I_i = \sqrt{(I^{(x)}_i)^2 + (I^{(y)}_i)^2}, \quad \theta_i = \arctan(I^{(y)}_i / I^{(x)}_i) \end{align} Our encoder can thus be described as ${\mathcal{T}} := {\mathcal{T}}_\text{Polar} \circ \text{G-SympNet}$. \textbf{Evolve}: From \autoref{eqn:theta-dot}, we know that the angular velocities $\dot{\theta}$ are only a function of the actions $I$ at the instant $t_0$, which we model as a simple multi-layer perceptron (MLP) ${\mathcal{F}}$. Since the angles $\theta$ are guaranteed to evolve linearly, we do not need to use higher-order numerical integration schemes. Instead, a single call to the forward Euler method is exact: \begin{align} \dot{\theta}(t_0) &\gets {\mathcal{F}}(I(t_0)) \quad \theta(t_0 + \Delta t) \gets \theta(t_0) + \dot{\theta}(t_0) \cdot \Delta t \pmod{2\pi} \end{align} \textbf{Decode}: As ${\mathcal{T}}$ is symplectic, ${\mathcal{T}}$ is invertible. Thus, our decoder is ${\mathcal{T}}^{-1} := \text{G-SympNet}^{-1} \circ {\mathcal{T}}_{\text{polar}}^{-1}$. \subsection{Training} We generate a trajectory $\text{Tr} = \{(q(t), p(t))\}_{t = 1}^T$ of $T = 1000$ time steps, by either evaluating the closed-form dynamics equations or via numerical integration. In practise, this trajectory can also be obtained from raw observations, and does not need to be regularly sampled in time. We train models upto the first $500$ time steps and evaluate performance on the last $500$ time steps, for each trajectory. \textbf{State Prediction Error}: We wish to learn parameters $\phi := (\phi_{{\mathcal{T}}}, \phi_{{\mathcal{F}}})$ that minimize the prediction loss $L_{\text{predict}}$: \vspace{-0.8em} \begin{align} L_{\text{predict}} = \textstyle\frac{1}{1 + \Delta t} \textstyle\sum_{t_0 \in \text{Tr}} \norm{{\mathcal{M}}_\phi(u(t_0), \Delta t) - u(t_0 + \Delta t)}^2. \end{align} \textbf{Regularization of Predicted Actions}: Additionally, we wish to enforce the fact that the predicted actions $I$ are constant across the trajectory. We do this by adding a regularizer $L_{\text{action}}$ on the variance of the predicted actions: \vspace{-0.3em} \begin{align} L_{\text{action}} &= \textstyle\widehat{\mathrm{Var}}(I) = \frac{1}{T}\sum_{t_0 \in \text{Tr}} \left(I(t_0) - \widehat{\mathbb{E}}[I]\right)^2 \text{where} \ \widehat{\mathbb{E}}[I] = \frac{1}{T}\sum_{t_0 \in \text{Tr}}I(t_0). \end{align} We find that this global regularizer works better than the loss proposed in \cite{learning-symmetries} that minimizes only the local pairwise differences between the predicted actions across the trajectory. As the Action-Angle Network is completely differentiable, the parameters $\phi$ can be obtained via gradient-descent-based minimization of the total loss: $ \phi^* = \text{argmin}_\phi \left(L_{\text{predict}} + \lambda L_{\text{action}}\right), $ where $\lambda$ is a hyperparameter that controls the strength of the regularization. \textbf{Schedule for $\Delta t$}: Increasing $\Delta t$ over the course of training helped, as it corresponded to gradually increasing the complexity of the prediction task. We set $\Delta t_{\text{max}} = 10$ and sampled $\Delta t$ according to: \begin{align} \Delta t \sim \text{Uniform}\left(0, (\text{Current training step})/(\text{Maximum training steps})\right) \cdot \Delta t_{\text{max}} \end{align} \vspace{-2em}
{ "timestamp": "2022-11-29T02:24:33", "yymm": "2211", "arxiv_id": "2211.15338", "language": "en", "url": "https://arxiv.org/abs/2211.15338", "abstract": "Machine learning has become increasingly popular for efficiently modelling the dynamics of complex physical systems, demonstrating a capability to learn effective models for dynamics which ignore redundant degrees of freedom. Learned simulators typically predict the evolution of the system in a step-by-step manner with numerical integration techniques. However, such models often suffer from instability over long roll-outs due to the accumulation of both estimation and integration error at each prediction step. Here, we propose an alternative construction for learned physical simulators that are inspired by the concept of action-angle coordinates from classical mechanics for describing integrable systems. We propose Action-Angle Networks, which learn a nonlinear transformation from input coordinates to the action-angle space, where evolution of the system is linear. Unlike traditional learned simulators, Action-Angle Networks do not employ any higher-order numerical integration methods, making them extremely efficient at modelling the dynamics of integrable physical systems.", "subjects": "Machine Learning (cs.LG); Computational Physics (physics.comp-ph)", "title": "Learning Integrable Dynamics with Action-Angle Networks", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9653811651448431, "lm_q2_score": 0.8289388125473629, "lm_q1q2_score": 0.8002419166907558 }
https://arxiv.org/abs/1610.07539
Origami Constructions of Rings of Integers of Imaginary Quadratic Fields
In the making of origami, one starts with a piece of paper, and through a series of folds along seed points one constructs complicated three-dimensional shapes. Mathematically, one can think of the complex numbers as representing the piece of paper, and the seed points and folds as a way to generate a subset of the complex numbers. Under certain constraints, this construction can give rise to a ring, which we call an origami ring. We will talk about the basic construction of an origami ring and further extensions and implications of these ideas in algebra and number theory, extending results of Buhler,this http URL. In particular, in this paper we show that it is possible to obtain the ring of integers of an imaginary quadratic field through an origami construction.
\section{Introduction} In origami, the artist uses intersections of folds as reference points to make new folds. This kind of construction can be extended to points on the complex plane. In \cite{Buhler2010}, the authors define one such mathematical construction. In this construction, one can think of the complex plane as representing the ``paper", and lines representing the ``folds". The question they explored is which points in the plane can be constructed through iterated intersections of lines, starting with a prescribed set of allowable angles and only the points 0 and 1. First, we say that the set $S=\{0,1\}$ is the set of \emph{seed points}. We fix a set $U$ of angles, or ``directions", determining which lines we can draw through the points in our set. Thus, we can define the ``fold" through the point $p$ with angle $u$ as the line given by $$L_u(p):=\{p+ru:r\in\mathbb R\}.$$ Notice that $U$ can also be comprised of points (thinking of $u\in U$ as defining a direction) on the unit circle, i.e. the circle group $\mathbb{T}$. Moreover, $u$ and $-u$ define the same line, so we can think of the directions as being in the quotient group $\mathbb{T}/\{\pm1\}$. Finally, if $u$ and $v$ in $U$ determine distinct folds, we say that $$I_{u,v}(p,q)=L_u(p)\cap L_v(q)$$ is the unique point of intersection of the lines $L_u(p)$ and $L_v(q)$. Let $R(U)$ be the set of points obtained by iterated intersections $I_{u,v}(p,q)$, starting with the points in $S$. Alternatively, we define $R(U)$ to be the smallest subset of $\mathbb{C}$ that contains $0$ and $1$ and $I_{u,v}(p,q)$ whenever it contains $p$ and $q$, and $u, v$ determine distinct folds. The main theorem of \cite{Buhler2010} is the following. \begin{theorem}\label{Buhler} If $U$ is a subgroup of $\mathbb{T}/\{\pm1\}$, and $|U|\geq3$, then $R(U)$ is a subring of $\mathbb{C}$. \end{theorem} Let $U_n$ denote the cyclic group of order $n$ generated by $e^{i\pi/n}\mod\{\pm1\}$. Then Buhler, et. al., obtain the following corollary. \begin{theorem} Let $n\geq3$. If $n$ is prime, then $R(U_n)=\mathbb{Z}[\zeta_n]$ is the cyclotomic integer ring. If $n$ is not prime, then $R(U_n)=\mathbb{Z}[\zeta_n,\frac{1}{n}]$. \end{theorem} In \cite{nedrenco}, Nedrenco explores whether $R(U)$ is a subring of $\mathbb C$ even if $U$ is not a group, and obtains a negative answer, and some necessary conditions for this to be true. The main result is that given the set of directions $U=\{1,e^{i\alpha},e^{i\beta}\}$, if $\alpha\not\equiv\beta\bmod\pi$ then $R(U)=\mathbb Z+z\mathbb Z$ for some $z\in\mathbb C$. Clearly, this will not always be a ring. In this paper, we explore the inverse problem, that is, given an ``interesting" subring of $\mathbb C$, can we obtain it via an origami construction? The answer is affirmative in the case of the ring of integers of an imaginary quadratic field. The next section of this paper delves deeper into the origami construction, in particular the intersection operator. Some properties in this section are crucial for understanding the proofs of our main results. This section also explores an example of an origami construction in more depth, that of the Gaussian integers, since it illustrates the geometric and algebraic approach through a very well known ring. Finally, in Section 3, we state and prove our main result. \begin{rem} Even though Nedrenco essentially proved one of our two results in \cite{nedrenco}, this was accomplished completely independently by us. In fact, our proof is different, since we are interested in the reverse of Nedrenco's question. That is, we explore whether a given ring can be an origami ring. Nedrenco explores the conditions for which his origami construction ends up being a ring of algebraic integers. This distinction is subtle, but important, and we want to clarify that all of what follows, unless otherwise indicated as coming from \cite{Buhler2010}, is original work. \end{rem} \subsection*{Acknowledgements.} This work was part of the first author's senior Honors Thesis at Bates College, advised by the second author, and we thank the Bates Mathematics Department, in particular Peter Wong, for useful feedback and discussions throughout the thesis process. We would also like to thank the rest of the Honors committee, Pamela Harris and Matthew Cot\'{e}, for the careful reading of the thesis and their insightful questions. Finally, we would like to thank the 2015 Summer Undergraduate Applied Mathematics Institute (SUAMI) at Carnegie Mellon University for inspiring the research problem. \section{Properties of Origami Rings} \subsection{The intersection operator} Let $U\subset \mathbb T$, as before. There are important properties of the $I_{u,v}(p,q)$ operator that are integral for us to prove our theorem. \par Let $u,v\in U$ be two distinct angles. Let $p,q$ be points in $R(U)$. Consider the pair of intersecting lines $L_u(p)$ and $L_v(q)$. In \cite{Buhler2010}, it is shown that we can express $I_{u,v}(p,q)$ as \begin{equation}\tag{$\ast$} I_{u,v}(p,q)=\frac{u\bar pv-\bar upv}{u\bar v-\bar uv}+\frac{q\bar vu-\bar qvu}{\bar uv-u\bar v}=\frac{[u,p]}{[u,v]}v+\frac{[v,q]}{[v,u]}u \label{closedform} \end{equation} where $[x,y]=\overline{x}y-x\overline{y}$. From the algebraic closed form (\ref{closedform}) of the intersection operator, we can see by straightforward computation that the following properties hold for for $p,q,u,v\in \mathbb C$. \begin{description} \item[Symmetry] $I_{u,v}(p,q)=I_{v,u}(q,p)$ \item[Reduction] $I_{u,v}(p,q)=I_{u,v}(p,0)+I_{v,u}(q,0)$ \item[Linearity] $I_{u,v}(p+q,0)=I_{u,v}(p,0)+I_{u,v}(q,0)$ and $rI_{u,v}(p,0)=I_{u,v}(rp,0)$ where $r\in\mathbb R$. \item[Projection] $I_{u,v}(p,0)$ is a projection of $p$ on the line $\{rv:r\in\mathbb R\}$ in the $u$ direction. \item[Rotation] For $w\in \mathbb{T}$, $wI_{u,v}(p,q)=I_{wu,wv}(wp,wq)$. \end{description} \subsection{An illustrative example} Let $S=\{0,1\}$ be our set of seed points. Now, let $U=\{1,e^{i\pi/4},i\}$. This is clearly not a group, since $e^{\frac{i\pi}{4}}i=e^{\frac{3i\pi}{4}}\not\in U$. In figure \ref{fig: gauss}, we show the different stages of the construction, obtained by iterated intersections. \begin{figure}[!h] \includegraphics[height=.3\textheight]{Gauss1} \includegraphics[height=.3\textheight]{Gauss2} \includegraphics[height=.3\textheight]{Gauss3} \includegraphics[height=.3\textheight]{Gauss4} \includegraphics[height=.3\textheight]{Gauss5} \includegraphics[height=.3\textheight]{Gauss6} \caption[Generational Expansion]{Each successive graph shows all possible intersections from the previous graph using $U=\{1,e^{i\pi/4},i\}$ as our set of allowable angles.}\label{fig: gauss} \end{figure} \begin{rem} All the figures in this document were created by coding the algorithm for iterated intersections into Maple \cite{maple}. \end{rem} Notice that a pattern seems to emerge: the points constructed all have the form $a+bi$ where $a,b \in \mathbb Z$. This seems to indicate that this origami construction generates the Gaussian integers, a subring of $\mathbb{C}$. In fact, this is a special case of the main result of \cite{nedrenco}, where $z=i$. We prove in the next section that the ring of algebraic integers of an imaginary quadratic field can always be obtained through an origami construction. \section{Constructing $\mathcal{O}(\mathbb{Q}(\sqrt{m}))$} A natural question, related to the previous section is this: Which subrings of $\mathbb C$ can be generated through an origami construction, that is, which subrings are origami rings? We have seen that the cyclotomic integers $\mathbb Z[\zeta_n]$, where $n$ is prime, are origami rings by Theorem \ref{Buhler}. Let $m<0$ be a square-free integer, so $\mathbb Q(\sqrt{m})$ is an imaginary quadratic field. Denote by $\mathcal{O}(\mathbb Q(\sqrt{m}))$ the ring of algebraic integers in $\mathbb Q(\sqrt{m})$. Recall that a complex number is an algebraic integer if and only if it is the root of some monic polynomial with coefficients in $\mathbb Z$. Then we have the following well-known theorem (for details see, for example, \cite[pg. 15]{marcus}). \begin{theorem} The set of algebraic integers in the quadratic field $\mathbb Q (\sqrt{m})$ is \begin{align*} \{a+b\sqrt{m}: a,b\in\mathbb Z\} & \text{ if $m\equiv 2$ or $3 \mod 4$}\\ \left\{\frac{a+b\sqrt{m}}{2}: a,b\in\mathbb Z , a\equiv b \mod 2\right\} & \text{ if $m\equiv 1 \mod 4$} \end{align*} \end{theorem} And so, we can state our main theorem. \begin{theorem}\label{ourtheorem} Let $m<0$ be a squarefree integer, and let $\theta=\arg(1+\sqrt{m})$. Then $\mathcal{O}(\mathbb Q(\sqrt{m}))=R (U)$ where \begin{enumerate} \item $U=\{1,i,e^{i\theta}\}$, if $m\equiv 2$ or $3 \mod 4$. \item $U=\{1, e^{i\theta},e^{i(\pi-\theta)}\}$, if $m\equiv 1 \mod 4$. \end{enumerate} \end{theorem} Notice that the Gaussian integers are a special case of Theorem \ref{ourtheorem}.1. \subsection{Proof of Theorem \ref{ourtheorem}.1} Let $m\equiv 2$ or $3 \mod 4$ and $m<0$. Let $U=\{1,i,e^{i\theta}\}$ where $\theta$ is the principal argument of $1+\sqrt{m}$. \begin{lemma}\label{cases1} $I_{u,v} (p,q)\in \mathbb Z[\sqrt{m}]$ whenever $u,v\in U$ and $p,q\in \mathbb Z[\sqrt{m}] $. \end{lemma} \begin{proof} Since there are three possible directions, there are ${3\choose 2}=6$ cases to consider. Let $p=a+b\sqrt{m}$ and $q=c+d\sqrt{m}$. Then \begin{enumerate} \item $I_{1,i} (p,q)=c+b\sqrt{m}$\\ \item $I_{1,e^{i\theta}} (p,q)=b+c-d+b\sqrt{m}$\\ \item $I_{i,1} (p,q)=a+d\sqrt{m}$\\ \item $I_{i,e^{i\theta}} (p,q)=a+ (a-c+d)\sqrt{m}$\\ \item $I_{e^{i\theta},1} (p,q)=a-b+d+d\sqrt{m}$\\ \item $I_{e^{i\theta},i} (p,q)=c+ (-a+b+c)\sqrt{m}$\\ \end{enumerate} In other words, if $p,q\in\mathbb Z[\sqrt{m}]$, then so is $I_{u,v} (p,q)$. All of these can be obtained from straightforward computations using equation (\ref{closedform}). For example, $$I_{1,i} (p,q)=\frac{[1,a+b\sqrt{m}]}{[1,i]}i+\frac{[i,c+d\sqrt{m}]}{[i,1]}=c+b\sqrt{m}.$$ \end{proof} \par This concludes the proof for the closure of the intersection operator. In other words, as long as our seed set starts with elements in $\mathbb Z[\sqrt{m}]$, then the intersections will also be in $\mathbb Z[\sqrt{m}]$. We can also express this claim as $R (U)\subseteq \mathbb Z[\sqrt{m}]$. It remains to be shown that any element in $\mathbb Z[\sqrt{m}]$ is also an element in $R (U)$. \begin{lemma} $\mathbb Z[\sqrt{m}]\subseteq R (U)$. \end{lemma} \begin{proof} \par Let $a+b\sqrt{m}$ be an element in in $\mathbb Z[\sqrt{m}]$. We want to show that it can be constructed from starting with $\{0,1\}$ and the given set $U$. We can reduce the problem by showing that given points $\{n+k\sqrt{m},n+1+k\sqrt{m}\}$ we can construct $$n-1+k\sqrt{m}, n+2+k\sqrt{m}, n+ (k-1)\sqrt{m} \hspace{3pt} \text{and} \hspace{3pt} n+1+ (k+1)\sqrt{m}.$$ In Figure \ref{redux} we give an illustration of this step of the proof. In essence, the following is the induction step to a double induction on the real and imaginary components of an arbitrary integer we are constructing. That is, we prove that for any two adjacent points in the construction, we can construct points that are adjacent in every direction. This, and the fact that our seed is $\{0,1\}$, is enough to show that we can construct all of the integers. \begin{figure}[!h] \includegraphics[ height=0.3\textheight]{redux12} \includegraphics[ height=0.3\textheight]{redux22} \caption[Reduction]{Given two points next to each other, we want to show that we can generate all the points immediately around them. This results in more points next to each other, upon which we can repeat the process. Notice that there are six adjacent points to an adjacent pair and we only prove this for four, but it is easy to see that this is enough. Think of this as two overlapping ``crosses". }\label{redux} \end{figure} We will now construct the desired points using the appropriate reference points. \begin{description} \item[Constructing $n+2+k\sqrt{m}$] Consider $$I_{i,1} (I_{1,e^{i\theta}} (I_{e^{i\theta},i} (n+k\sqrt{m},n+1+k\sqrt{m}),n+1+k\sqrt{m}),n+1+k\sqrt{m}).$$ Notice that we can evaluate this expression using the six cases enumerated in Lemma \ref{cases1}. In particular, we apply case (6) first to get $$I_{i,1} (I_{1,e^{i\theta}} (n+1+ (k+1)\sqrt{m},n+1+k\sqrt{m}),n+1+k\sqrt{m})$$ Next, we apply case (2) to get $$I_{i,1} (n+1+ (k+1)\sqrt{m},n+1+k\sqrt{m})$$ Finally, we use case (3) to get $$n+2+k\sqrt{m}$$ \item[Constructing $n-1+k\sqrt{m}$] Consider $$I_{i,1} (I_{1,e^{i\theta}} (I_{i,e^{i\theta}} (n+k\sqrt{m},n+1+k\sqrt{m}),n+k\sqrt{m}),n+k\sqrt{m})$$ First, we apply case (4) to get $$I_{i,1} (I_{1,e^{i\theta}} (n+ (k-1)\sqrt{m},n+k\sqrt{m}),n+k\sqrt{m})$$ Next, we apply case (2) to get $$I_{i,1} (n-1+ (k-1)\sqrt{m},n)$$ Finally, we apply case (3) to get $$n-1+k\sqrt{m}$$ \item[Constructing $n+ (k+1)\sqrt{m}$ and $n+1+ (k+1)\sqrt{m}$] Consider $$I_{e^{i\theta},i} (n+k\sqrt{m},n+1+k\sqrt{m})$$ Using case (6) we get $$n+1+ (k+1)\sqrt{m}$$ Now consider $$I_{i,1} (n+k\sqrt{m},n+1+ (k+1)\sqrt{m})$$ Using case (3) we get $$n+ (k+1)\sqrt{m}$$ \item[Constructing $n+ (k-1)\sqrt{m}$ and $n+1+ (k-1)\sqrt{m}$] Consider $$I_{i,e^{i\theta}} (n+k\sqrt{m},n+1+k\sqrt{m})$$ Using case (4) we get $$n+ (k-1)\sqrt{m}$$ Consider $$I_{i,1} (n+1+k\sqrt{m},n+ (k-1)\sqrt{m})$$ Using case (3) we get $$n+1+ (k-1)\sqrt{m}$$ \end{description} \end{proof} With this we have shown that $\mathbb Z[\sqrt{m}]\subseteq R (U)$, completing the proof. \subsection{Proof of \ref{ourtheorem}.2} The proof for \ref{ourtheorem}.2 employs the same strategy as the proof for \ref{ourtheorem}.1, with a subtle difference given by the slightly different structure of the ring. Let $m\equiv 1 \mod 4$ and $m<0$. Let $U=\{1, e^{i\theta},e^{i (\pi-\theta)}\}$ where $\theta$ is the principal argument of $1+\sqrt{m}$. \begin{lemma} $I_{u,v} (p,q)\in \mathcal{O}(\mathbb Q(\sqrt{m}))$ where $u,v\in U$ and $p,q\in \mathcal{O}(\mathbb Q(\sqrt{m}))$. \end{lemma} \begin{proof} Again, there are six cases to consider. Let $p=\frac{a+b\sqrt{m}}{2}$ and $q=\frac{c+d\sqrt{m}}{2}$, where $a\equiv b\bmod 2$ and $c\equiv d\bmod2$. \begin{enumerate} \item $I_{1,e^{i\theta}} (p,q)=b+c-d+b\sqrt{m}$\\ \item $I_{1,e^{i (\pi-\theta)}} (p,q)=c+d-b+b\sqrt{m}$\\ \item $I_{e^{i\theta},1} (p,q)=a-b+d+d\sqrt{m}$\\ \item $I_{e^{i\theta},e^{i (\pi-\theta)}} (p,q)=\frac{ (a-b+c+d)+ (b-a+c+d)\sqrt{m}}{2}$\\ \item $I_{e^{i (\pi-\theta)},1} (p,q)=a+b-d+d\sqrt{m}$\\ \item $I_{e^{i (\pi-\theta)},e^{i\theta}} (p,q)=\frac{ (a+b+c-d)+ (a+b-c+d)\sqrt{m}}{2}$\\ \end{enumerate} All of these cases can be obtained, again, by straightforward computation using (\ref{closedform}), and are left as a exercise. Notice that (1), (2), (3), and (5) all are clearly in the ring of algebraic integers. The only additional fact to show is that (4) and (6) are as well. But it's easy to see that $$a-b+c+d\equiv b-a+c+d \bmod 2,$$ since $a\equiv b\bmod 2$. And similarly $$a+b+c-d\equiv a+b-c+d \bmod 2,$$ because $c\equiv d\bmod 2$. \end{proof} \par This concludes the proof for the closure of the intersection operator. In other words, as long as our seed set starts with elements in $\mathcal{O}(\mathbb Q(\sqrt{m}))$, then the intersections will also be in $\mathcal{O}(\mathbb Q(\sqrt{m}))$. That is, $R (U)\subseteq \mathcal{O}(\mathbb Q(\sqrt{m}))$. It remains to be shown that any element in $\mathcal{O}(\mathbb Q(\sqrt{m}))$ is also an element in $R (U)$. \begin{lemma} $\mathcal{O}(\mathbb Q(\sqrt{m}))\subseteq R(U)$. \end{lemma} \begin{proof} \par Let $\dfrac{a+b\sqrt{m}}{2}$ be an element in in $\mathcal{O}(\mathbb Q(\sqrt{m}))$. As before, we can reduce the problem to one of double induction. This is done by showing that given points $$\left\{\frac{n+k\sqrt{m}}{2},\frac{n+2+k\sqrt{m}}{2}\right\}$$ we can construct $$\dfrac{n-2+k\sqrt{m}}{2}, \dfrac{n+4+k\sqrt{m}}{2}, \dfrac{n+1+ (k-1)\sqrt{m}}{2} \hspace{3pt} \text{and} \hspace{3pt}\dfrac{n+1+ (k+1)\sqrt{m}}{2}.$$ Figure \ref{redux2} is an illustration of the points we want to construct. \begin{figure}[!h] \includegraphics[ height=0.2\textheight]{redux12} \includegraphics[ height=0.2\textheight]{redux32} \includegraphics[ height=0.2\textheight]{redux312} \caption[Reduction]{Given two points next to each other, we want to show that we can generate all the points immediately around them. Notice that this results in more points next to each other, upon which we can repeat the process. As before, we do not need to actually make all eight adjacent points, since it suffices to make the two points on either side of where we start, and the two points above and below where we start. The graph on the left illustrates the starting points. The graph in the center are all points adjacent to the two starting points. The graph on the right shows the points whose construction is sufficient to prove the theorem.}\label{redux2} \end{figure} We will now show how to construct the desired points. \begin{description} \item[Constructing $\frac{n+1+ (k-1)\sqrt{m}}{2}$] Consider $$I_{e^{i \theta},e^{i (\pi-\theta)}} \left(\frac{n+k\sqrt{m}}{2},\frac{n+2+k\sqrt{m}}{2}\right)$$ By applying case (4) from above, we see that $$I_{e^{i \theta},e^{i (\pi-\theta)}} \left(\frac{n+k\sqrt{m}}{2},\frac{n+2+k\sqrt{m}}{2}\right)=\frac{n+1+ (k-1)\sqrt{m}}{2}$$ \item[Constructing $\frac{n+1+ (k+1)\sqrt{m}}{2}$] Consider $$I_{e^{i (\pi-\theta)},e^{i \theta}} \left(\frac{n+k\sqrt{m}}{2},\frac{n+2+k\sqrt{m}}{2}\right)$$ By applying case (6) from above, we see that $$I_{e^{i (\pi-\theta)},e^{i \theta}} \left(\frac{n+k\sqrt{m}}{2},\frac{n+2+k\sqrt{m}}{2}\right)=\frac{n+1+ (k+1)\sqrt{m}}{2}$$ \item[Constructing $\frac{n-2+k\sqrt{m}}{2}$] Consider $$I_{e^{i\theta},1} \left(I_{1,e^{i (\pi-\theta)}} \left(\frac{n+1+ (k+1)\sqrt{m}}{2},\frac{n+k\sqrt{m}}{2}\right),\frac{n+k\sqrt{m}}{2}\right)$$ By applying case (1) from above, we can reduce the previous expression to $$I_{e^{i\theta},1} \left(\frac{n-1+ (k+1)\sqrt{m}}{2},\frac{n+k\sqrt{m}}{2}\right)$$ We further reduce the expression using case (5) from above. The result is $$I_{e^{i\theta},1} \left(\frac{n-1+ (k+1)\sqrt{m}}{2},\frac{n+k\sqrt{m}}{2}\right)=\frac{n-2+k\sqrt{m}}{2}$$ \item[Constructing $\frac{n+4+k\sqrt{m}}{2}$] Consider $$I_{e^{i (\pi-\theta)},1} \left(I_{1,e^{\pi\theta}} \left(\frac{n+1+ (k+1)\sqrt{m}}{2},\frac{n+2+k\sqrt{m}}{2}\right),\frac{n+2+k\sqrt{m}}{2}\right)$$ By applying case (5) from above, we can reduce the previous expression to $$I_{e^{i (\pi-\theta)},1} \left(\frac{n+3+ (k+1)\sqrt{m}}{2},\frac{n+2+k\sqrt{m}}{2}\right)$$ We further reduce the expression using case (1) from above. The result is $$I_{e^{i (\pi-\theta)},1} \left(\frac{n+3+ (k+1)\sqrt{m}}{2},\frac{n+2+k\sqrt{m}}{2}\right)=\frac{n+4+k\sqrt{m}}{2}$$ \end{description} \end{proof} With this we have shown that $$\left\{\frac{a+b\sqrt{m}}{2}: a,b\in\mathbb Z , a\equiv b \mod 2\right\}\subseteq R (U)$$ completing the proof. \subsection{Some final illustrations and remarks} In Figure \ref{fig: gausspartial} we see that when we use $U=\{1,i,e^{i\theta}\}$ as our angle set, then the origami ring grows into the first and third quadrants, and bleeds into the others. As discussed, $R(U)=\mathbb Z[i]$. In Figure \ref{fig: partialhalfs} we see that when we use $U=\{1,e^{i\arg(1+\sqrt{-3})},e^{i(\pi-\arg(1+\sqrt{-3}))}\}$ as our angle set, then the origami ring grows along the real line, and slowly bleeds into the rest of the complex plane. This is an illustration of the second case of Theorem \ref{ourtheorem}. In this case, $R(U)=\mathcal{O}(\mathbb Q(\sqrt{-3}))$. Of course, $R(U)$ is assumed to be closed under the intersection operator, so the growth pattern doesn't matter in an abstract sense. However, computationally, it means that the number of steps it takes to construct a point is not related to that point's modulus. In fact, we get an entirely different measure of distance if we only consider the number of steps it takes to generate a point. One possible additional exploration is, given more general starting angles and points, to describe the dynamics of the iterative process. These examples also serve to illustrate the progression of the iterative process, which was coded into Maple to produce the graphs. \begin{figure}[!h] \includegraphics[ height=.3\textheight]{Gauss5} \caption[Five Generations of the Easy Case]{This graph depicts the first 5 generations of origami points using $U=\{1, i,e^{i\frac{\pi}{4}}\}$}\label{fig: gausspartial} \end{figure} \begin{figure}[!h] \includegraphics[ height=.3\textheight]{batman21} \caption[Five Generations of the Hard Case]{This graph depicts the first 5 generations of origami points using $$U=\{1,e^{i\arg(1+\sqrt{-3})},e^{i(\pi-\arg(1+\sqrt{-3}))}\}$$}\label{fig: partialhalfs} \end{figure} \pagebreak
{ "timestamp": "2016-10-25T02:12:43", "yymm": "1610", "arxiv_id": "1610.07539", "language": "en", "url": "https://arxiv.org/abs/1610.07539", "abstract": "In the making of origami, one starts with a piece of paper, and through a series of folds along seed points one constructs complicated three-dimensional shapes. Mathematically, one can think of the complex numbers as representing the piece of paper, and the seed points and folds as a way to generate a subset of the complex numbers. Under certain constraints, this construction can give rise to a ring, which we call an origami ring. We will talk about the basic construction of an origami ring and further extensions and implications of these ideas in algebra and number theory, extending results of Buhler,this http URL. In particular, in this paper we show that it is possible to obtain the ring of integers of an imaginary quadratic field through an origami construction.", "subjects": "Number Theory (math.NT)", "title": "Origami Constructions of Rings of Integers of Imaginary Quadratic Fields", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9902915220858951, "lm_q2_score": 0.8080672204860316, "lm_q1q2_score": 0.8002221177228309 }
https://arxiv.org/abs/1106.5444
An Extension of Young's Inequality
Young's inequality is extended to the context of absolutely continuous measures. Several applications are included.
\section{Introduction} Young's inequality \cite{Y1912} asserts that every strictly increasing continuous function $f:\left[ 0,\infty\right) \longrightarrow\left[ 0,\infty\right) $ with $f\left( 0\right) =0$ and $\underset{x\rightarrow \infty}{\lim}f\left( x\right) =\infty$ verifies an inequality of the following form \begin{equation} ab\leq\int_{0}^{a}f\left( x\right) dx+\int_{0}^{b}f^{-1}\left( y\right) dy, \label{youngineq \end{equation} whenever $a$ and $b\ $are nonnegative real numbers. The equality occurs if and only if $f\left( a\right) =b$. See \cite{HLP}, \cite{Mit}, \cite{NP2006} and \cite{RV} for details and significant applications. Several questions arise naturally in connection with this classical result. \begin{enumerate} \item[(Q1):] Is the restriction on strict monotonicity (or on continuity) really necessary? \item[(Q2):] Is there any weighted analogue of Young's inequality? \item[(Q3):] Can Young's inequality be improved? \end{enumerate} F. Cunningham Jr. and N. Grossman \cite{CG1971} noticed that the question (Q1) has a positive answer (correcting the prevalent belief that Young's inequality is the business of strictly increasing continuous functions). The aim of the present paper is to extend the entire discussion to the framework of locally absolutely continuous measures and to prove several improvements. As well known, Young's inequality is an illustration of the Legendre duality. Precisely, the function \[ F(a)=\int_{0}^{a}f\left( x\right) dx\text{ and }G(b)=\int_{0}^{b f^{-1}\left( x\right) dx, \] are both continuous and convex on $\left[ 0,\infty\right) $ and (\ref{youngineq}) can be restated a \begin{equation} ab\leq F(a)+G(b)\text{\quad for all }a,b\in\left[ 0,\infty\right) , \label{youngineq2 \end{equation} with equality if and only if $f\left( a\right) =b.$ Because of the equality case, the formula (\ref{youngineq2}) leads to the following connection between the functions $F$ and $G: \begin{equation} F(a)=\sup\left\{ ab-G(b):b\geq0\right\} \label{defF \end{equation} and \[ G(b)=\sup\left\{ ab-F(a):a\geq0\right\} . \] It turns out that each of these formulas produces a convex function (possibly on a different interval). Some details are in order. By definition, the \emph{conjugate} of a convex function $F$ defined on a nondegenerate interval $I$ is the functio \[ F^{\ast}:I^{\ast}\rightarrow\mathbb{R},\text{\quad}F^{\ast}(y)=\sup\left\{ xy-F(x):x\in I\right\} , \] with domain $I^{\ast}=\left\{ y\in\mathbb{R}:F^{\ast}(y)<\infty\right\} $. Necessarily $I^{\ast}$ is an non-empty interval and $F^{\ast}$ is a convex function whose level sets $\left\{ y:F^{\ast}(y)\leq\lambda\right\} $ are closed subsets of $\mathbb{R}$ for each $\lambda\in\mathbb{R}$ (usually such functions are called \emph{closed} convex functions). A convex function may not be differentiable, but it admits a good substitute for differentiability. The \emph{subdifferential\ }of a real function\emph{\ }$F$ defined on an interval $I$ is a multivalued function $\partial F:I\rightarrow\mathcal{P (\mathbb{R})$ defined b \[ \partial F(x)=\left\{ \lambda\in\mathbb{R}:F(y)\geq F(x)+\lambda(y-x)\text{, for every}\,\,y\in I\right\} . \] Geometrically, the subdifferential gives us the slopes of the supporting lines for the graph of $F$. The sub\allowbreak differential at a point\emph{\ }is always a convex set, possibly empty, but the convex functions $F:I\rightarrow \mathbb{R}$ have the remarkable property that $\partial F(x)\neq\emptyset$ at all interior points. It is worth noticing that $\partial F(x)=\left\{ F^{\prime}(x)\right\} $ at each point where $F$ is differentiable (so this formula works for all points of $I$ except for a countable subset). See \cite{NP2006}, page 30. \begin{lemma} \label{Lem1}\noindent\emph{(}Legendre duality, \emph{\cite{NP2006}}, page \emph{41)}. Let $F:I\rightarrow\mathbb{R}$\ be a closed convex function. Then its conjugate $F^{\ast}:I^{\ast}\rightarrow\mathbb{R}$ is also convex and closed and: $i)$ $xy\leq F(x)+F^{\ast}(y)$ for all $x\in I,$ $y\in I^{\ast};$ $ii)$ $xy=F(x)+F^{\ast}(y)$ if, and only if, $y\in\partial F(x);$ $iii)$ $\partial F^{\ast}=\,\left( \partial F\right) ^{-1}$ \emph{(}as graphs\emph{)}$;$ $iv)$ $F^{\ast\ast}=F.$ \end{lemma} Recall that the inverse of a graph $\Gamma$ is the set $\Gamma^{-1}=\left\{ \left( y,x\right) :(x,y)\in\Gamma\right\} .$ How far is Young's inequality from the Legendre duality? Surprisingly, they are pretty closed in the sense that in most cases the Legendre duality can be converted into a Young like inequality. Indeed, every continuous convex function admits an integral representation. \begin{lemma} \label{Lem2}\noindent\emph{(}See \emph{\cite{NP2006}}, page \emph{37)}. Let $F$\ be a continuous convex function defined on an interval $I$ and let $\varphi:I\rightarrow\mathbb{R}$ be a function such that $\varphi (x)\in\partial F(x)$ for every $x\in\,I.$\ Then for every $a<b$ in $I$ we have \[ F(b)-F(a)=\int_{a}^{b}\,\varphi(t)\,dt. \] \end{lemma} As a consequence, the heuristic meaning of the formula $i)$ in Lemma \ref{Lem1} is the following Young like inequality, \[ ab\leq\int_{a_{0}}^{a}\varphi\left( x\right) dx+\int_{b_{0}}^{b}\psi\left( y\right) dy\text{\quad for all }a\in I,\ b\in I^{\ast}, \] where $\varphi$ and $\psi$ are selection functions for $\partial F$ and respectively $\left( \partial F\right) ^{-1}$. Now it becomes clear that Young's inequality should work outside strict monotonicity (as well as outside continuity). The details are presented in Section 2. Our approach (based on the geometric meaning of integrals as areas) allows us to extend the framework of integrability to all positive measures $\rho$ which are locally absolutely continuous with respect to the planar Lebesgue measure $dxdy$. See Theorem \ref{ThmYoungNondecr}\ below. A special case of Young's inequality i \[ xy\leq\frac{x^{p}}{p}+\frac{y^{q}}{q}, \] which works for all $x,y\geq0$, and $p,q>1$ with $1/p+1/q=1$. Theorem \ref{ThmYoungNondecr} yields the following companion to this inequality in the case of Gaussian measure $\frac{4}{2\pi}e^{-x^{2}-y^{2}}dxdy$ on $[0,\infty)\times\lbrack0,\infty):$ \[ \operatorname{erf}(x)\operatorname{erf}(y)\leq\frac{2}{\sqrt{\pi}}\int_{0 ^{x}\operatorname{erf}\left( s^{p-1}\right) e^{-s^{2}}ds+\frac{2}{\sqrt{\pi }}\int_{0}^{y}\operatorname{erf}\left( t^{q-1}\right) e^{-t^{2}}dt, \] wher \begin{equation} \operatorname{erf}(x)=\frac{2}{\sqrt{\pi}}\int_{0}^{x}e^{-s^{2}}ds \label{erf \end{equation} is the \emph{Gauss error function} (or the erf function). The precision of our generalization of Young's inequality makes the objective of Section 3. In Section 4 we discuss yet another extension of Young's inequality, based on recent work done by J. Jak\v{s}eti\'{c} and J. E. Pe\v{c}ari\'{c} \cite{P}. The paper ends by noticing the connection of our result to the theory of $c$-convexity (that is, of convexity associated to a cost density function). Last but not the least, all results in this paper can be extended verbatim to the framework of nondecreasing functions $f:[a_{0},a_{1})\rightarrow\lbrack A_{0},A_{1})$ such that $a_{0}<a_{1}\leq\infty$ and $A_{0}<A_{1}\leq\infty,$ $f(a_{0})=A_{0}$ and $\lim_{x\rightarrow a_{1}}f(x)=A_{1}.$ In other words, the interval $[0,\infty)$ plays no special role in Young's inequality. Besides, there is a straightforward companion of Young's inequality for nonincreasing functions, but this is outside the scope of the present paper. \section{Young's inequality for weighted measures} In what follows $f:\left[ 0,\infty\right) \longrightarrow\left[ 0,\infty\right) $ will denote a nondecreasing function such that $f\left( 0\right) =0$ and $\underset{x\rightarrow\infty}{\lim}f\left( x\right) =\infty.$ Since $f$ is not necessarily injective we will attach to $f$ a \emph{pseudo-inverse} by the following formula \[ f_{\sup}^{-1}:\left[ 0,\infty\right) \longrightarrow\left[ 0,\infty\right) ,\quad f_{\sup}^{-1}\left( y\right) =\inf\{x\geq0:f(x)>y\}. \] Clearly, $f_{\sup}^{-1}$ is nondecreasing and $f_{\sup}^{-1}\left( f\left( x\right) \right) \geq x$ for all $x.$ Moreover, with the convention $f(0-)=0,$ \[ f_{\sup}^{-1}\left( y\right) =\sup\left\{ x:y\in\left[ f\left( x-\right) ,f\left( x+\right) \right] \right\} ; \] here $f\left( x-\right) $ and $f\left( x+\right) $ represent the lateral limits at $x$. When $f$ is also continuous \[ f_{\sup}^{-1}(y)=\max\left\{ x\geq0:y=f(x)\right\} . \] \begin{remark} $($F. Cunningham Jr. and N. Grossman \cite{CG1971}$)$. \emph{Since pseudo-inverses will be used as integrands, it is convenient to enlarge the concept of pseudo-inverse by referring to any function} $g$ \emph{such that \[ f_{\inf}^{-1}\leq g\leq f_{\sup}^{-1}, \] \emph{where} $f_{\inf}^{-1}(y)=\sup\{x\geq0:f(x)<y\}$. \emph{Necessarily,} $g$ \emph{is nondecreasing and any two} \emph{pseudo-inverses agree except on a countable set (so their integrals will be the same)}. \end{remark} Given $0\leq a<b,$ we define the \emph{epigraph} and the \emph{hypograph} of $f|_{[a,b]}$ respectively b \[ \operatorname{epi}f|_{[a,b]}=\left\{ \left( x,y\right) \in\left[ a,b\right] \times\left[ f\left( a\right) ,f\left( b\right) \right] :y\geq f\left( x\right) \right\} , \] an \[ \operatorname{hyp}f|_{[a,b]}=\left\{ \left( x,y\right) \in\left[ a,b\right] \times\left[ f\left( a\right) ,f\left( b\right) \right] :y\leq f\left( x\right) \right\} . \] Their intersection is the \emph{graph} of $f|_{[a,b]},$ \[ \operatorname*{graph}f|_{[a,b]}=\left\{ \left( x,y\right) \in\left[ a,b\right] \times\left[ f\left( a\right) ,f\left( b\right) \right] :y=f\left( x\right) \right\} . \] Notice that our definitions of epigraph and hypograph are not the standard ones, but agree with them in the context of monotone functions. We will next consider a measure $\rho$ on $\left[ 0,\infty\right) \times\left[ 0,\infty\right) ,$ which is locally absolutely continuous with respect to the Lebesgue measure $dxdy,$ that is, $\rho$ is of the form \[ \rho\left( A\right) =\int_{A}K\left( x,y\right) dxdy, \] where $K:\left[ 0,\infty\right) \times\left[ 0,\infty\right) \longrightarrow\lbrack0,\infty)\ $is a Lebesgue locally integrable function, and $A$ is any compact subset of $\left[ 0,\infty\right) \times\left[ 0,\infty\right) $. Clearly \begin{align*} \rho\left( \operatorname{hyp}f|_{[a,b]}\right) +\rho\left( \operatorname{epi}f|_{[a,b]}\right) & =\rho\left( \left[ a,b\right] \times\left[ f\left( a\right) ,f\left( b\right) \right] \right) \\ & =\int_{a}^{b}\int_{f\left( a\right) }^{f\left( b\right) }K\left( x,y\right) dydx. \end{align*} Moreover \[ \rho\left( \operatorname{hyp}f|_{[a,b]}\right) =\int_{a}^{b}\left( \int_{f\left( a\right) }^{f\left( x\right) }K\left( x,y\right) dy\right) dx. \] an \begin{equation} \rho\left( \operatorname{epi}f|_{[a,b]}\right) =\int_{f\left( a\right) }^{f\left( b\right) }\left( \int_{a}^{f_{\sup}^{-1}\left( y\right) }K\left( x,y\right) dx\right) dy.\nonumber \end{equation} The discussion above can be summarized as follows: \begin{lemma} \label{Lem3}Let $f:\left[ 0,\infty\right) \longrightarrow\left[ 0,\infty\right) $ be a nondecreasing function such that $f\left( 0\right) =0$ and $\underset{x\rightarrow\infty}{\lim}f\left( x\right) =\infty$. Then for every Lebesgue locally integrable function $K:\left[ 0,\infty\right) \times\left[ 0,\infty\right) \longrightarrow\lbrack0,\infty)$ and every pair of nonnegative numbers $a<b, \begin{multline*} \int_{a}^{b}\left( \int_{f\left( a\right) }^{f\left( x\right) }K\left( x,y\right) dy\right) dx+\int_{f\left( a\right) }^{f\left( b\right) }\left( \int_{a}^{f_{\sup}^{-1}\left( y\right) }K\left( x,y\right) dx\right) dy\\ =\int_{a}^{b}\int_{f\left( a\right) }^{f\left( b\right) }K\left( x,y\right) dydx. \end{multline*} \end{lemma} We can now state the main result of this section: \begin{theorem} \label{ThmYoungNondecr}\emph{(}Young's inequality for nondecreasing functions\textbf{\emph{)}. }Under the assumptions of Lemma $3$, for every pair of nonnegative numbers $a<b,$ and every number $c\geq f(a)$ we have \begin{multline*} \int_{a}^{b}\int_{f\left( a\right) }^{c}K\left( x,y\right) dydx\\ \leq\int_{a}^{b}\left( \int_{f\left( a\right) }^{f\left( x\right) }K\left( x,y\right) dy\right) dx+\int_{f\left( a\right) }^{c}\left( \int_{a}^{f_{\sup}^{-1}\left( y\right) }K\left( x,y\right) dx\right) dy. \end{multline*} If in addition $K$ is strictly positive almost everywhere, then the equality occurs if and only if $c\in\left[ f\left( b-\right) ,f\left( b+\right) \right] .$ \begin{proof} We start with the case where $f\left( a\right) \leq c\leq f\left( b-\right) $. See Figure \ref{fig1}. \ \begin{figure} [h] \begin{center} \includegraphics[ height=5.7464cm, width=6.491cm {fig1.jpg \caption{The geometry of Young's inequality when $f\left( a\right) \leq c\leq f\left( b-\right) .$ \label{fig1 \end{center} \end{figure} In this case \begin{multline*} \int_{a}^{b}\left( \int_{f\left( a\right) }^{f\left( x\right) }K\left( x,y\right) dy\right) dx+\int_{f\left( a\right) }^{c}\left( \int _{a}^{f_{\sup}^{-1}\left( y\right) }K\left( x,y\right) dx\right) dy\\ =\int_{a}^{f_{\sup}^{-1}\left( c\right) }\left( \int_{f\left( a\right) }^{f\left( x\right) }K\left( x,y\right) dy\right) dx+\int_{f\left( a\right) }^{c}\left( \int_{a}^{f_{\sup}^{-1}\left( y\right) }K\left( x,y\right) dx\right) dy\\ +\int_{f_{\sup}^{-1}\left( c\right) }^{b}\left( \int_{f\left( a\right) }^{f\left( x\right) }K\left( x,y\right) dy\right) dx\\ =\int_{a}^{f_{\sup}^{-1}\left( c\right) }\int_{f\left( a\right) ^{c}K\left( x,y\right) dydx+\int_{f_{\sup}^{-1}\left( c\right) ^{b}\left( \int_{c}^{f\left( x\right) }K\left( x,y\right) dy\right) dx\\ +\int_{f_{\sup}^{-1}\left( c\right) }^{b}\int_{f\left( a\right) ^{c}K\left( x,y\right) dydx\\ \geq\int_{a}^{b}\int_{f\left( a\right) }^{c}K\left( x,y\right) dydx, \end{multline*} with equality if and only if $\int_{f_{\sup}^{-1}\left( c\right) ^{b}\left( \int_{c}^{f\left( x\right) }K\left( x,y\right) dy\right) dx=0.$ When $K$ is strictly positive almost everywhere, this means that $c=f\left( b-\right) $. If $c\geq f\left( b+\right) ,$ the \begin{multline*} \int_{a}^{b}\left( \int_{f\left( a\right) }^{f\left( x\right) }K\left( x,y\right) dy\right) dx+\int_{f\left( a\right) }^{c}\left( \int _{a}^{f_{\sup}^{-1}\left( y\right) }K\left( x,y\right) dx\right) dy\\ =\int_{a}^{f_{\sup}^{-1}\left( c\right) }\left( \int_{f\left( a\right) }^{f\left( x\right) }K\left( x,y\right) dy\right) dx+\int_{f\left( a\right) }^{c}\left( \int_{a}^{f_{\sup}^{-1}\left( y\right) }K\left( x,y\right) dx\right) dy\\ -\int_{b}^{f_{\sup}^{-1}\left( c\right) }\left( \int_{f\left( a\right) }^{f\left( x\right) }K\left( x,y\right) dy\right) dx\\ =\int_{a}^{f_{\sup}^{-1}\left( c\right) }\int_{f\left( a\right) ^{c}K\left( x,y\right) dydx\\ -\left( \int_{b}^{f_{\sup}^{-1}\left( c\right) }\left( \int_{f\left( a\right) }^{f\left( c\right) }K\left( x,y\right) dy\right) dx-\int_{f\left( b+\right) }^{c}\left( \int_{a}^{f_{\sup}^{-1}\left( y\right) }K\left( x,y\right) dx\right) dy\right) \\ \geq\int_{a}^{b}\int_{f\left( a\right) }^{c}K\left( x,y\right) dydx. \end{multline*} Equality holds if and only if$\ \int_{f\left( b+\right) }^{c}\left( \int_{a}^{f_{\sup}^{-1}\left( y\right) }K\left( x,y\right) dx\right) dy$, that is, when $c=f\left( b+\right) $ (provided that $K$ is strictly positive almost everywhere). See Figure 2 \begin{figure} [ptb] \begin{center} \includegraphics[ height=5.3927cm, width=6.6558cm {fig2.jpg \caption{The case $c\geq f\left( b+\right) .$ \label{fig2 \end{center} \end{figure} If $c\in\left( f\left( b-\right) ,f\left( b+\right) \right) ,$ then $f_{\sup}^{-1}\left( c\right) =b$ and the inequality in the statement of Theorem \ref{ThmYoungNondecr} is actually an equality. See Figure \ref{fig3} \begin{figure} [h] \begin{center} \includegraphics[ height=6.1593cm, width=6.9216cm {fig3.jpg \caption{The equality case. \label{fig3 \end{center} \end{figure} \end{proof} \end{theorem} \begin{corollary} \label{CorContIncr}\emph{(}Young's inequality for continuous increasing functions\emph{)}\textbf{. }If $f:\left[ 0,\infty\right) \longrightarrow \left[ 0,\infty\right) $ is also continuous and increasing, the \begin{multline*} \int_{a}^{b}\int_{f\left( a\right) }^{c}K\left( x,y\right) dydx\\ \leq\int_{a}^{b}\left( \int_{f\left( a\right) }^{f\left( x\right) }K\left( x,y\right) dy\right) dx+\int_{f\left( a\right) }^{c}\left( \int_{a}^{f^{-1}\left( y\right) }K\left( x,y\right) dx\right) dy \end{multline*} for every real number $c\geq f(a)$. Assuming $K$ strictly positive almost everywhere, the equality occurs if and only if $\ c=f\left( b\right) .$ \end{corollary} If $K\left( x,y\right) =1$ for every $x,y\in\left[ 0,\infty\right) $, then Corollary \ref{CorContIncr} asserts tha \[ bc-af\left( a\right) <\int_{a}^{b}f\left( x\right) dx+\int_{f\left( a\right) }^{c}f^{-1}\left( y\right) dy\text{\quad for all }0<a<b\text{ and }c>f(a); \] equality occurs if and only if $c=f\left( b\right) $. In the special case where $a=f\left( a\right) =0$, this reduces to the classical inequality of Young. \begin{remark} $($The probabilistic companion of Theorem \ref{ThmYoungNondecr}$)$. \emph{Suppose there is given a nonnegative random variable} $X:[0,\infty )\rightarrow\lbrack0,\infty)$ \emph{whose cumulative distribution function} $F_{X}(x)=P\left( X\leq x\right) $ \emph{admits a density, that is, a nonnegative Lebesgue-integrable function} $\rho_{X}$ \emph{such} \emph{that} \[ P\left( x\leq X\leq y\right) =\int_{x}^{y}\rho_{X}(u)du\text{\quad for all }x\leq y. \] \emph{The} quantile function \emph{of the distribution function} $F_{X}$ $($\emph{also known as the} increasing rearrangement \emph{of the random variable} $X)$ \emph{is defined by \[ Q_{X}(x)=\inf\left\{ y:F_{X}(y)\geq x\right\} . \] \emph{Thus, a quantile function is nothing but a pseudo-inverse of} $F_{X}$. \emph{Motivated by Statistics, a number of fast algorithms were developed for computing the quantile functions with high accuracy. See} \cite{A}. \emph{Without entering the details, we recall here the remarkable formula} \emph{(due to} \emph{G.} \emph{Steinbrecher)} \emph{for the quantile function of the normal distribution: \[ \operatorname{erf}^{-1}(z)=\sum_{k=0}^{\infty}\frac{c_{k}\left( \frac {\sqrt{\pi}}{2}z\right) ^{2k+1}}{2k+1}, \] \emph{where} $c_{0}=1$ \emph{and} \[ c_{k}=\sum_{m=0}^{k-1}\frac{c_{m}c_{k-m-1}}{\left( m+1\right) \left( 2m+1\right) }\text{\quad\emph{for all} }k\geq1. \] \emph{According to Theorem} \ref{ThmYoungNondecr}, \emph{for every pair of continuous random variables} $Y,Z:[0,\infty)\rightarrow\lbrack0,\infty)$ \emph{with density} $\rho_{Y,Z},$ \emph{and every positive numbers} $b$ \emph{and} $c,$ \emph{the following inequality holds: \[ P\left( Y\leq b;Z\leq c\right) \leq\int_{0}^{b}\left( \int_{0}^{F_{X (x)}\rho_{Y,Z}\left( x,y\right) dy\right) dx+\int_{0}^{c}\left( \int _{0}^{Q_{X}(y)}\rho_{Y,Z}\left( x,y\right) dx\right) dy. \] \emph{This can be seen as a principle of uncertainty, since it shows that the functions} \[ x\rightarrow\int_{0}^{F_{X}(x)}\rho_{Y,Z}\left( x,y\right) dy\text{ and }y\rightarrow\int_{0}^{Q_{X}(y)}\rho_{Y,Z}\left( x,y\right) dx \] \emph{cannot be made simultaneously small.} \end{remark} \begin{remark} $($The higher dimensional analogue of Theorem $1).$ \emph{Consider a locally absolutely continuous} \emph{kernel} $K:\left[ 0,\infty\right) \times...\times\left[ 0,\infty\right) \longrightarrow\lbrack0,\infty ),\ K=K\left( s_{1},s_{2},...,s_{n}\right) ,$ \emph{and a family} $\phi _{1},...,\phi_{n}:[a_{i},b_{i}]\rightarrow\mathbb{R}\ $\emph{of nondecreasing functions defined on subintervals of} $\left[ 0,\infty\right) .$ \emph{Then} \begin{multline*} \int_{\phi_{1}\left( a_{1}\right) }^{\phi_{1}\left( b_{1}\right) \int_{\phi_{2}\left( a_{2}\right) }^{\phi_{2}\left( b_{2}\right) \cdots\int_{\phi_{n}\left( a_{n}\right) }^{\phi_{n}\left( b_{n}\right) }K\left( s_{1},s_{2},...,s_{n}\right) ds_{n}...ds_{2}ds_{1}\\ \le {\displaystyle\sum\limits_{i=1}^{n}} \int_{\phi_{i}\left( a_{i}\right) }^{\phi_{i}\left( b_{i}\right) }\left( \int_{\phi_{1}\left( a_{1}\right) }^{\phi_{1}\left( s\right) }\cdots \int_{\phi_{n}\left( a_{n}\right) }^{\phi_{n}\left( s\right) }K\left( s_{1},...,s_{n}\right) ds_{n}...ds_{i+1}ds_{i-1}...ds_{1}\right) ds.\ \end{multline*} \emph{The proof is based on mathematical induction (which is left to the reader). The above inequality cover the n-variable generalization of Young's inequality as obtained by Oppenheim \cite{O1927} (as well as the main result in \cite{Pa1992}).} \end{remark} The following stronger version of Corollary \ref{CorContIncr} incorporates the Legendre duality. \begin{theorem} \label{extYoung}Let $f:\left[ 0,\infty\right) \longrightarrow\left[ 0,\infty\right) $ be a continuous nondecreasing function and $\Phi :[0,\infty)\rightarrow\mathbb{R}$ a convex function whose conjugate is also defined on $[0,\infty)$. Then for all $b>a\geq0,$ $c\geq f(a),$ and $\varepsilon>0$ we have \begin{multline*} \int_{a}^{b}\Phi\left( \varepsilon\int_{f\left( a\right) }^{f\left( x\right) }K\left( x,y\right) dy\right) dx+\int_{f(a)}^{c}\Phi^{\ast }\left( \frac{1}{\varepsilon}\int_{a}^{f_{\sup}^{-1}\left( y\right) }K\left( x,y\right) dx\right) dx\\ \geq\int_{a}^{b}\int_{f\left( a\right) }^{c}K\left( x,y\right) dydx-(c-f(a))\Phi\left( \varepsilon\right) -(b-a)\Phi^{\ast}\left( 1/\varepsilon\right) . \end{multline*} \end{theorem} \begin{proof} According to the Legendre duality \begin{equation} \Phi(\varepsilon u)+\Phi^{\ast}(v/\varepsilon)\geq uv\text{\quad for all }u,v,\varepsilon\geq0. \label{fi \end{equation} For $u=\int_{f\left( a\right) }^{f\left( x\right) }K\left( x,y\right) dy$ and $v=1$ we ge \[ \Phi\left( \varepsilon\int_{f\left( a\right) }^{f\left( x\right) }K\left( x,y\right) dy\right) +\Phi^{\ast}\left( 1/\varepsilon\right) \geq\int_{f\left( a\right) }^{f\left( x\right) }K\left( x,y\right) dy, \] and by integrating both sides from $a$ to $b$ we obtain the inequalit \[ \int_{a}^{b}\Phi\left( \varepsilon\int_{f\left( a\right) }^{f\left( x\right) }K\left( x,y\right) dy\right) dx+(b-a)\Phi^{\ast}\left( 1/\varepsilon\right) \geq\int_{a}^{b}\left( \int_{f\left( a\right) }^{f\left( x\right) }K\left( x,y\right) dy\right) dx. \] In a similar manner, starting with $u=1$ and $v=\int_{a}^{f_{\sup}^{-1}\left( y\right) }K\left( x,y\right) dx,$ we arrive first at the inequality \[ \Phi\left( \varepsilon\right) +\Phi^{\ast}\left( \frac{1}{\varepsilon \int_{a}^{f_{\sup}^{-1}\left( y\right) }K\left( x,y\right) dx\right) \geq\int_{a}^{f_{\sup}^{-1}\left( y\right) }K\left( x,y\right) dx, \] and then t \begin{multline*} (c-f(a))\Phi\left( \varepsilon\right) +\int_{f(a)}^{c}\Phi^{\ast}\left( \frac{1}{\varepsilon}\int_{a}^{f_{\sup}^{-1}\left( y\right) }K\left( x,y\right) dx\right) dx\\ \geq\int_{f\left( a\right) }^{c}\left( \int_{a}^{f_{\sup}^{-1}\left( y\right) }K\left( x,y\right) dx\right) dy. \end{multline*} Therefore \begin{multline*} \int_{a}^{b}\Phi\left( \varepsilon\int_{f\left( a\right) }^{f\left( x\right) }K\left( x,y\right) dy\right) dx+\int_{f(a)}^{c}\Phi^{\ast }\left( \frac{1}{\varepsilon}\int_{a}^{f_{\sup}^{-1}\left( y\right) }K\left( x,y\right) dx\right) dx\\ \geq\int_{a}^{b}\left( \int_{f\left( a\right) }^{f\left( x\right) }K\left( x,y\right) dy\right) dx+\int_{f\left( a\right) }^{c}\left( \int_{a}^{f_{\sup}^{-1}\left( y\right) }K\left( x,y\right) dx\right) dy\\ -(b-a)\Phi^{\ast}\left( 1/\varepsilon\right) -(c-f(a))\Phi\left( \varepsilon\right) . \end{multline*} According to Theorem \ref{ThmYoungNondecr} \begin{multline*} \int_{a}^{b}\left( \int_{f\left( a\right) }^{f\left( x\right) }K\left( x,y\right) dy\right) dx+\int_{f\left( a\right) }^{c}\left( \int _{a}^{f_{\sup}^{-1}\left( y\right) }K\left( x,y\right) dx\right) dy\\ \geq\int_{a}^{b}\int_{f\left( a\right) }^{c}K\left( x,y\right) dydx, \end{multline*} and the inequality in the statement of Theorem \ref{extYoung} is now clear. \end{proof} In the special case where $K\left( x,y\right) =1,$ $a=f\left( a\right) =0$ and $\Phi(x)=x^{p}/p$ (for some\emph{\ }$p>1$), Theorem \ref{extYoung} yields the following inequality \[ \int_{0}^{b}f^{p}\left( x\right) dx+\int_{0}^{c}\left( f_{\sup}^{-1}\left( y\right) \right) ^{p}dy\geq pbc-\left( p-1\right) \left( b+c\right) ,\ \text{for every }b,c\geq0. \] This remark extends a result due to W. T. Sulaiman \cite{S}. We end this section by noticing the following result that complements Theorem \ref{ThmYoungNondecr}. \begin{proposition} \label{PropMerkle}Under the assumptions of Lemma \ref{Lem3} \begin{multline*} \int_{a}^{b}\left( \int_{f\left( a\right) }^{f\left( x\right) }K\left( x,y\right) dy\right) dx+\int_{f\left( a\right) }^{c}\left( \int _{a}^{f_{\sup}^{-1}\left( y\right) }K\left( x,y\right) dx\right) dy\\ \leq\max\left\{ \int_{a}^{b}\int_{f\left( a\right) }^{f\left( b\right) }K\left( x,y\right) dydx,\int_{a}^{f_{\sup}^{-1}\left( c\right) \int_{f\left( a\right) }^{c}K\left( x,y\right) dydx\right\} . \end{multline*} Assuming $K$ strictly positive almost everywhere, the equality occurs if and only if $c=f\left( b\right) .$ \end{proposition} \begin{proof} If $c<f\left( b\right) $, then from Lemma \ref{Lem3} we infer tha \begin{multline*} \int_{a}^{b}\left( \int_{f\left( a\right) }^{f\left( x\right) }K\left( x,y\right) dy\right) dx+\int_{f\left( a\right) }^{c}\left( \int _{a}^{f_{\sup}^{-1}\left( y\right) }K\left( x,y\right) dx\right) dy\\ =\int_{a}^{b}\left( \int_{f\left( a\right) }^{f\left( x\right) }K\left( x,y\right) dy\right) dx+\int_{f\left( a\right) }^{f\left( b\right) }\left( \int_{a}^{f_{\sup}^{-1}\left( y\right) }K\left( x,y\right) dx\right) dy\\ -\int_{c}^{f\left( b\right) }\left( \int_{a}^{f_{\sup}^{-1}\left( y\right) }K\left( x,y\right) dx\right) dy\\ \leq\int_{a}^{b}\int_{f\left( a\right) }^{f\left( b\right) }K\left( x,y\right) dydx \end{multline*} The other case, $c\geq f\left( b\right) $, has a similar approach. \end{proof} Proposition \ref{PropMerkle} extends a result due to M. J. Merkle \cite{Me}. \section{The precision in Young's inequality} The main result of this section is as follows: \begin{theorem} \label{ThmPrec}Under the assumptions of Lemma \ref{Lem3}, for all $b\geq a\geq0$ and $c\geq f(a), \begin{multline*} \int_{a}^{b}\left( \int_{f\left( a\right) }^{f\left( x\right) }K\left( x,y\right) dy\right) dx+\int_{f\left( a\right) }^{c}\left( \int _{a}^{f_{\sup}^{-1}\left( y\right) }K\left( x,y\right) dx\right) dy\\ -\int_{a}^{b}\int_{f\left( a\right) }^{c}K\left( x,y\right) dydx\leq \left\vert \int_{f_{\sup}^{-1}\left( c\right) }^{b}\int_{c}^{f\left( b\right) }K\left( x,y\right) dydx\right\vert \text{. \end{multline*} Assuming $K$ strictly positive almost everywhere, the equality occurs if and only if $c=f\left( b\right) $. \end{theorem} \begin{proof} The case where $f\left( a\right) \leq c\leq f\left( b-\right) $ is illustrated in Figure \ref{fig4}. The left-hand side of the inequality in the statement of Theorem \ref{ThmPrec} represents the measure of the cross-hatched curvilinear trapezium, while right-hand side is the measure of the $ABCD$ rectangle \begin{figure} [h] \begin{center} \includegraphics[ height=5.4696cm, width=6.9787cm {fig4.jpg \caption{The geometry of the case $f\left( a\right) \leq c\leq f\left( b-\right) .$ \label{fig4 \end{center} \end{figure} Therefore \begin{multline*} \int_{a}^{b}\left( \int_{f\left( a\right) }^{f\left( x\right) }K\left( x,y\right) dy\right) dx+\int_{f\left( a\right) }^{c}\left( \int _{a}^{f_{\sup}^{-1}\left( y\right) }K\left( x,y\right) dx\right) dy\\ -\int_{a}^{b}\int_{f\left( a\right) }^{c}K\left( x,y\right) dydx=\int _{f_{\sup}^{-1}\left( c\right) }^{b}\left( \int_{c}^{f\left( x\right) }K\left( x,y\right) dy\right) dx\\ \leq\int_{f_{\sup}^{-1}\left( c\right) }^{b}\int_{c}^{f\left( b\right) }K\left( x,y\right) dydx. \end{multline*} The equality holds if and only if $\int_{f_{\sup}^{-1}\left( c\right) ^{b}\left( \int_{c}^{f\left( x\right) }K\left( x,y\right) dy\right) dx=0,$ that is, when $f\left( b-\right) =c.$ The case where $c\geq f\left( b+\right) $ is similar to the precedent one. The first term will be \begin{multline*} \int_{a}^{b}\left( \int_{f\left( a\right) }^{f\left( x\right) }K\left( x,y\right) dy\right) dx+\int_{f\left( a\right) }^{c}\left( \int _{a}^{f_{\sup}^{-1}\left( y\right) }K\left( x,y\right) dx\right) dy\\ -\int_{a}^{b}\int_{f\left( a\right) }^{c}K\left( x,y\right) dydx=\int _{b}^{f_{\sup}^{-1}\left( c\right) }\left( \int_{f\left( b\right) }^{f\left( x\right) }K\left( x,y\right) dy\right) dx\\ \leq\int_{b}^{f_{\sup}^{-1}\left( c\right) }\int_{f\left( b\right) ^{c}K\left( x,y\right) dydx. \end{multline*} Equality holds if and only if $\int_{b}^{f_{\sup}^{-1}\left( c\right) \int_{f\left( b\right) }^{c}K\left( x,y\right) dydx=0,\ $so we must have $f\left( b+\right) =c$. The case where $c\in\left[ f\left( b-\right) ,f\left( b+\right) \right] $ is trivial, both sides of our inequality being equal to zero. \end{proof} \begin{corollary} \label{CorMing}\emph{(}E. Minguzzi\emph{\ }\cite{M}\emph{)}. If moreover $K\left( x,y\right) =1$ on $\left[ 0,\infty\right) \times\left[ 0,\infty\right) $, and $f$ is continuous and increasing, then \[ \int_{a}^{b}f\left( x\right) dx+\int_{f\left( a\right) }^{c}f^{-1}\left( y\right) dy\ -bc+af\left( a\right) \leq\left( f^{-1}\left( c\right) -b\right) \cdot\left( c-f\left( b\right) \right) . \] The equality occurs if and only if $c=f\left( b\right) $. \end{corollary} More accurate bounds can be indicated under the presence of convexity. \begin{corollary} \label{CorJP}Let $f$ be a nondecreasing continuous function, which is convex on the interval $\left[ \min\left\{ f_{\sup}^{-1}\left( c\right) ,b\right\} ,\max\left\{ f_{\sup}^{-1}\left( c\right) ,b\right\} \right] $. Then \begin{multline*} i)~\int_{a}^{b}\left( \int_{f\left( a\right) }^{f\left( x\right) }K\left( x,y\right) dy\right) dx+\int_{f\left( a\right) }^{c}\left( \int_{a}^{f_{\sup}^{-1}\left( y\right) }K\left( x,y\right) dx\right) dy\\ -\int_{a}^{b}\int_{f\left( a\right) }^{c}K\left( x,y\right) dydx\\ \leq\int_{f_{\sup}^{-1}\left( c\right) }^{b}\int_{c}^{c+\frac{f(b)-c {b-f_{\sup}^{-1}\left( c\right) }(x-f_{\sup}^{-1}\left( c\right) )}K\left( x,y\right) dydx\text{,\quad for every }c\leq f\left( b\right) ; \end{multline* \begin{multline*} ii)~\int_{a}^{b}\left( \int_{f\left( a\right) }^{f\left( x\right) }K\left( x,y\right) dy\right) dx+\int_{f\left( a\right) }^{c}\left( \int_{a}^{f_{\sup}^{-1}\left( y\right) }K\left( x,y\right) dx\right) dy\\ -\int_{a}^{b}\int_{f\left( a\right) }^{c}K\left( x,y\right) dydx\\ \geq\int_{b}^{f_{\sup}^{-1}\left( c\right) }\int_{f\left( b\right) }^{f(b)+\frac{c-f(b)}{f_{\sup}^{-1}\left( c\right) -b}(x-b)}K\left( x,y\right) dydx\text{,\quad for every }c\geq f\left( b\right) . \end{multline*} If $\ f$ is concave on the aforementioned interval, then the inequalities above work in the reverse way. Assuming $K$ strictly positive almost everywhere, the equality occurs if and only if$\ f$ is an affine function or $f\left( b\right) =c$. \begin{proof} We will restrict here to the case of convex functions, the argument for the concave functions being similar. The left-hand side term of each of the inequalities in our statement represents the measure of the cross-hatched surface. See Figure 5 and Figure 6. \medski \raisebox{-0cm}{\parbox[b]{5.9638cm}{\begin{center} \includegraphics[ height=4.9446cm, width=5.9638cm {fig5.jpg \\ Figure 5. The geometry of the case $c\leq f\left( b\right) . \end{center}} \raisebox{-0cm}{\parbox[b]{5.9594cm}{\begin{center} \includegraphics[ height=4.9776cm, width=5.9594cm {fig6.jpg \\ Figure 6. The geometry of the case $c\geq f\left( b\right) . \end{center}} \qquad\medskip As the points of the graph of the convex function $f$ (restricted to the interval of endpoints $b$ and $f_{\sup}^{-1}\left( c\right) )$ are under the chord joining $\left( b,f\left( b\right) \right) $ and $\left( f_{\sup }^{-1}\left( c\right) ,c\right) ,$ it follows that this measure is less than the measure of the enveloping triangle $MNQ$ when $c\leq f(b).$ This yields $i)$. The assertion $ii)$ follows in a similar way. \end{proof} \end{corollary} Corollary \ref{CorJP} extends a result due to J. Jak\v{s}eti\'{c} and J. E. Pe\v{c}ari\'{c}\textit{\ \cite{P}.} They considered the special case were $K\left( x,y\right) =1$ on $\left[ 0,\infty\right) \times\left[ 0,\infty\right) $ and $f:\left[ 0,\infty\right) \rightarrow\left[ 0,\infty\right) $ is increasing and differentiable, with an increasing derivative on the interval $\left[ \min\left\{ f^{-1}\left( c\right) ,b\right\} ,\max\left\{ f^{-1}\left( c\right) ,b\right\} \right] $ and $f(0)=0.$ In this case the conclusion of Corollary \ref{CorJP} reads as follows \begin{align*} i)\text{ }\int_{0}^{b}f\left( x\right) dx+\int_{0}^{c}f^{-1}\left( y\right) dy\ -bc & \leq\frac{1}{2}\left( f^{-1}\left( c\right) -b\right) \left( c-f\left( b\right) \right) \ \text{for }c<f\left( b\right) ;\\ ii)\text{ }\int_{0}^{b}f\left( x\right) dx+\int_{0}^{c}f^{-1}\left( y\right) dy\ -bc & \geq\frac{1}{2}\left( f^{-1}\left( c\right) -b\right) \left( c-f\left( b\right) \right) \ \text{for }c>f\left( b\right) . \end{align*} The equality holds if $f\left( b\right) =c$ or $f$ is an affine function. The inequality sign should be reversed if $f$ has a decreasing derivative on the interval \[ \left[ \min\left\{ f^{-1}\left( c\right) ,b\right\} ,\max\left\{ f^{-1}\left( c\right) ,b\right\} \right] . \] \section{The connection with $c$-convexity} Motivated by the mass transportation theory, several people \cite{D1988}, \cite{EN1974} drew a parallel to the classical theory of convex functions by extending the Legendre duality. Technically, given two compact metric spaces $X$ and $Y$ and a \emph{cost density} function $c:X\times Y\rightarrow \mathbb{R}$ (which is supposed to be continuous), we may consider the following generalization of the notion of convex function: \begin{definition} \label{cConv}A function $F:X\rightarrow\mathbb{R}$ is $c$-convex if there exists a function $G:Y\rightarrow\mathbb{R}$ such that \begin{equation} F(x)=\sup_{y\in Y}\left\{ c(x,y)-G(y)\right\} ,\;\text{for all }x\in X. \label{c-conv \end{equation} \end{definition} We abbreviate (\ref{c-conv}) by writing $F=G^{c}$. A useful remark is the equalit \[ F^{cc}=F, \] that is \begin{equation} F(x)=\sup_{y\in Y}\left\{ c(x,y)-F^{c}(y)\right\} ,\;\text{for all }x\in X. \label{cdual \end{equation} The classical notion of convex function corresponds to the case where $X$ is a compact interval and $c(x,y)=xy$. The details can be found in \cite{NP2006}, pp. 40-42. Theorem \ref{ThmYoungNondecr} illustrates the theory of $c$-convex functions for the spaces $X=[a,\infty]$, $Y=[f(a),\infty]$ (the Alexandrov one point compactification of $[a,\infty)$ and respectively $[f(a),\infty)$), and the cost function \begin{equation} c(x,y)=\int_{a}^{x}\int_{f(a)}^{y}K\left( s,t\right) dtds\text{.} \label{cKrelation \end{equation} In fact, under the hypotheses of this theorem, the function \[ F(x)=\int_{a}^{x}\left( \int_{f\left( a\right) }^{f\left( s\right) }K\left( s,t\right) dt\right) ds,\quad x\geq a, \] an \[ G(y)=\int_{f\left( a\right) }^{y}\left( \int_{a}^{f_{\sup}^{-1}\left( t\right) }K\left( s,t\right) ds\right) dt,\quad y\geq f(a), \] verify the relations $F^{c}=G$ and $G^{c}=F$ (due to the equality case as specified in the statement of Theorem \ref{ThmYoungNondecr}, so they are both $c$-convex. On the other hand, a simple argument shows that $F$ and $G$ are also convex in the usual sense. Let us call the functions $c$ that admits a representation of the form (\ref{cKrelation}) with $K\in L^{1}(\mathbb{R\times R}),$ \emph{absolutely} \emph{continuous in the hyperbolic sense}. With this terminology, Theorem \ref{ThmYoungNondecr} can be rephrased as follows: \begin{theorem} \label{ThmcConv}Suppose that $c:[a,b]\times\lbrack A,B]\rightarrow\mathbb{R}$ is an absolutely continuous function in the hyperbolic sense with mixed derivative $\frac{\partial^{2}c}{\partial x\partial y}\geq0,$ and $f:[a,b]\rightarrow\lbrack A,B]$ is a nondecreasing function such that $f(a)=A.$ The \begin{equation} c(x,y)-c(a,f(a))\leq\int_{a}^{x}\frac{\partial c}{\partial t}(t,f(t))dt+\int _{f(a)}^{y}\frac{\partial c}{\partial s}(f_{\sup}^{-1}(s),s)ds \label{cyineq \end{equation} for all $(x,y)\in\lbrack a,A]\times\lbrack b,B]$. If $\frac{\partial^{2}c}{\partial x\partial y}>0$ almost everywhere,\ then (\ref{cyineq}) becomes an equality if and only if $y\in\left[ f(x-),f(x+)\right] ;$ here we made the convention $f(a-)=f(a)$ and $f(b+)=f(b).$ \end{theorem} Necessarily, an absolutely continuous function $c$ in the hyperbolic sense, is continuous. It admits partial derivatives of the first order and a mixed derivative $\frac{\partial^{2}c}{\partial x\partial y}$ almost everywhere. Besides, the functions $y\rightarrow\frac{\partial c}{\partial x}(x,y)$ and $x\rightarrow\frac{\partial c}{\partial y}(x,y)$ are defined everywhere in their interval of definition and represent absolutely continuous functions; they are also nondecreasing provided that $\frac{\partial^{2}c}{\partial x\partial y}\geq0$ almost everywhere. A special case of Theorem \ref{ThmcConv} was proved by Zs. P\'{a}les \cite{Pa1990}, \cite{Pa1992} (assuming $c:[a,A]\times\lbrack b,B]\rightarrow \mathbb{R}$ a continuously differentiable function with nondecreasing derivatives $y\rightarrow\frac{\partial c}{\partial x}(x,y)$ and $x\rightarrow\frac{\partial c}{\partial y}(x,y),$ and $f:[a,b]\rightarrow \lbrack A,B]$ an increasing homeomorphism). An example which escapes his result but is covered by Theorem \ref{ThmcConv} is offered by the functio \[ c(x,y)=\int_{0}^{x}\left\{ \frac{1}{s}\right\} ds\int_{0}^{y}\left\{ \frac{1}{t}\right\} dt,\,\quad x,y\geq0, \] where $\left\{ \frac{1}{s}\right\} $ denotes the fractional part of $\frac{1}{s}$ if $s>0,$ and $\left\{ \frac{1}{s}\right\} =0$ if $s=0$. According to Theorem \ref{ThmcConv} \begin{multline*} \int_{0}^{x}\left\{ \frac{1}{s}\right\} ds\int_{0}^{y}\left\{ \frac{1 {t}\right\} dt\\ \leq\int_{0}^{x}\left( \left\{ \frac{1}{s}\right\} \int_{0}^{f(s)}\left\{ \frac{1}{t}\right\} dt\right) ds+\int_{0}^{y}\left( \left\{ \frac{1 {t}\right\} \int_{0}^{f_{\sup}^{-1}(t)}\left\{ \frac{1}{s}\right\} ds\right) dt, \end{multline*} for every nondecreasing function $f:[0,\infty)\rightarrow\lbrack0,\infty)$ such that $f(0)=0.$ \medskip \noindent\textbf{Acknowledgement.} The authors were supported by CNCSIS Grant PN2 ID\_$420.$
{ "timestamp": "2011-06-28T02:06:47", "yymm": "1106", "arxiv_id": "1106.5444", "language": "en", "url": "https://arxiv.org/abs/1106.5444", "abstract": "Young's inequality is extended to the context of absolutely continuous measures. Several applications are included.", "subjects": "Classical Analysis and ODEs (math.CA)", "title": "An Extension of Young's Inequality", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9787126457229185, "lm_q2_score": 0.8175744695262775, "lm_q1q2_score": 0.8001704721455747 }
https://arxiv.org/abs/2010.00315
Exact hyperplane covers for subsets of the hypercube
Alon and Füredi (1993) showed that the number of hyperplanes required to cover $\{0,1\}^n\setminus \{0\}$ without covering $0$ is $n$. We initiate the study of such exact hyperplane covers of the hypercube for other subsets of the hypercube. In particular, we provide exact solutions for covering $\{0,1\}^n$ while missing up to four points and give asymptotic bounds in the general case. Several interesting questions are left open.
\section{Introduction} A vector $v\in \mathbb{R}^n$ and a scalar $\alpha\in \mathbb{R}$ determine the hyperplane \[ \{x\in \mathbb{R}^n:\langle v,x\rangle \coloneqq v_1x_1+\dots+v_nx_n=\alpha\} \] in $\mathbb{R}^n$. How many hyperplanes are needed to cover $\{0,1\}^n$? Only two are required; for instance, $\{x:x_1=0\}$ and $\{x:x_1=1\}$ will do. What happens however if $0\in \mathbb{R}^n$ is not allowed on any of the hyperplanes? We can `exactly' cover $\{0,1\}^n\setminus \{0\}$ with $n$ hyperplanes: for example, the collections $\{\{x:x_i=1\}:i\in [n]\}$ or $\{\{x:\sum_{i=1}^nx_i=j\}:j \in [n]\}$ can be used, where $[n]:=\{1,2,\ldots,n\}$. Alon and F\"{u}redi \cite{AlonFuredi} showed that in fact $n$ hyperplanes are always necessary. Recently, a variation was studied by Clifton and Huang \cite{CliftonHuang}, in which they require that each point from $\{0,1\}^n\setminus \{0\}$ is covered at least $k$ times for some $k\in \mathbb{N}$ (while $0$ is never covered). Another natural generalisation is to put more than just $0$ to the set of points we wish to avoid in the cover. For $B\subseteq \{0,1\}^n$, the \emph{exact cover} of $B$ is a set of hyperplanes whose union intersects $\{0,1\}^n$ exactly in~$B$ (points from $\{0,1\}^n\setminus B$ are not covered). Let $\ec(B)$ denote the \emph{exact cover number} of $B$, i.e., the minimum size of an exact cover of $B$. We will usually write $B$ in the form $\{0,1\}^n \setminus S$ for some subset $S \subseteq \{0,1\}^n$. In particular, the result of Alon and F\"{u}redi \cite{AlonFuredi} states that $\ec(\{0,1\}^n\setminus \{0\})=n$. We first determine what happens if we remove up to four points. \begin{theorem} \label{thm:uptofour} Let $S\subseteq\{0,1\}^n$. \begin{itemize} \item If $|S|\in \{2,3\}$, then $\ec(\{0,1\}^n\setminus S)=n-1$. \item If $|S|=4$, then $\ec(\{0,1\}^n\setminus S)=n-1$ if there is a hyperplane $Q$ with $|Q\cap S|=3$ and $\ec(\{0,1\}^n\setminus S)=n-2$ otherwise. \end{itemize} \end{theorem} The upper bounds are shown by iteratively reducing the dimension of the problem by one using a single `merge coordinates' hyperplane; this allows us to reduce the question to the case $n\leq 7$, which we can handle exhaustively. Since the number of required hyperplanes seems to decrease, a natural question is whether this pattern continues. For $n\in \mathbb{N}$ and $k\in [2^n]$, we also introduce the exact cover numbers \begin{align*} \ec(n,k)&=\max\{\ec(\{0,1\}^n\setminus S):S\subseteq\{0,1\}^n,~ |S|=k\},\\ \ec(n)&=\max\{\ec(B):B\subseteq\{0,1\}^n\}. \end{align*} Our main result concerns the asymptotics of $\ec(n)$ and implies that $\ec(n,k)$ can be much larger than $n$. \begin{theorem} \label{thm:ec:arbitrary} For any positive integer $n$, $2^{n-2}/n^2\leq \ec(n)\leq 2^{n+1}/n$. \end{theorem} The lower bound uses a random construction and the upper bound uses the fact that we can efficiently cover the hypercube with Hamming spheres. We leave open whether $\ec(n,k)\leq n$ when $k$ is sufficiently large with respect to $n$, but can show that $\ec(n,k)$ is always at most a constant (depending on $k$) away from $n$. \begin{theorem} \label{thm:ec:fixed_size} For any positive integer $k$, \[n-\log_2(k) \leq \ec(n,k) \leq n-2^k+\ec(2^k,k).\] \end{theorem} The proof of this theorem uses the same techniques as the proof of Theorem \ref{thm:uptofour}. The problem of determining the asymptotics of $\ec(n)$ was also suggested by F\"{u}redi at Alon's birthday conference in 2016. \section{Covering all but up to four points} \label{sec:up_to_four} In this section, we determine $\ec(\{0,1\}^n\setminus S)$ for subsets $S$ of size 2, 3 and 4. For the lower bounds, we use the following result of Alon and F\"{u}redi \cite{AlonFuredi}. \begin{theorem}[Corollary 1 in \cite{AlonFuredi}] \label{thm:ATcor} If $n\geq m\geq 1$, then $m$ hyperplanes that do not cover all vertices of $\{0,1\}^n$ miss at least $2^{n-m}$ vertices. \end{theorem} For the upper bounds, it suffices to give an explicit construction of a collection of hyperplanes that exactly covers $\{0,1\}^n\setminus S$, for every subset $S$ of size 2, 3 or 4. We split the proof of Theorem \ref{thm:uptofour} into two cases, the case where $|S| \in \{2, 3\}$ and the case where $|S| = 4$. \begin{lemma} \label{lem:cov:23} Let $n \geq 2$ and $S\subseteq\{0,1\}^n$ with $|S|\in \{2,3\}$. Then $\ec(\{0,1\}^n\setminus S)=n-1$. \end{lemma} \begin{proof} For $n=2$ the statement is true, therefore let $n \geq 3$ and $S\subseteq\{0,1\}^n$ with $|S|\in \{2,3\}$. We first prove the lower bound $\ec(\{0,1\}^n\setminus S) \geq n-1$; this follows from applying the case of $m=n-2$ in Theorem \ref{thm:ATcor}. Indeed, this shows that any $n-2$ hyperplanes that do not cover all of $\{0,1\}^n$ miss at least $4$ vertices, and hence a minimum of $n-1$ hyperplanes are required to miss $2$ or $3$ vertices. For the upper bound, note that we may assume by vertex transitivity that $(0,\dots,0)\in S$. Consider first the case $|S|=2$. By relabelling the indices, we may assume the second vector $u$ in $S$ satisfies $\{i\in[n]:u_i=1\}=\{1,\dots,\ell\}$ for some $\ell\in \mathbb{N}$. We cover $\{0,1\}^n\setminus S$ by the collection of $n-1$ hyperplanes \[ \{\{x:x_i=1\}:i\in \{\ell+1,\dots,n\}\}\cup \left\{\left\{x:x_1+\dots+x_\ell=j\right\}:j\in [\ell-1]\right\}, \] noting none of these hyperplanes contain an element from $S$. Now consider the case $|S|=3$. We may assume the second and third vectors in $S$ correspond to the subsets $\{1,\dots,a+b\}$ and $\{1,\dots,a\}\cup\{a+b+1,\dots,a+b+c\}$ for some $a,b,c\in \mathbb{Z}_{\geq 0}$ with $a+b\geq 1$ and $c\geq 1$. We first add the $n-(a+b+c)$ hyperplanes of the form $\{x:x_i=1\}$ for $i\in \{a+b+c+1,\dots,n\}$. For $x\in S$, we have \begin{align*} &x_1+\dots+x_a\in \{0,a\},\\ &x_{a+1}+\dots+x_{a+b}\in \{0,b\},\\ &x_{a+b+1}+\dots+x_{a+b+c}\in \{0,c\}. \end{align*} If $a\geq 1$, we add the $a-1$ hyperplanes $\{x:x_1+\dots+x_a=i\}$ for $i\in [a-1]$. Analogously, we add the $b-1$ hyperplanes $\{x: x_{a+1}+\ldots+x_{a+b}=i\}$ for $i\in[b-1]$ if $b\geq 1$, and the $c-1$ hyperplanes $\{x: x_{a+b+1}+\ldots+x_{a+b+c}=i\}$ for $i\in [c-1]$. The only points of $\{0,1\}\setminus S$ that are yet to be covered satisfy the equations above and also satisfy $x_i=0$ for $i>a+b+c$. Suppose first that $a,b\geq 1$. In this case we have added $n-3$ hyperplanes so far. The problem has effectively been reduced to covering $\{0,1\}^3$ with three missing points $(0,0,0), (1,1,0)\text{ and } (1,0,1)$ using $2$ hyperplanes. Indeed, we may add the following two hyperplanes to our collection in order to exactly cover $\{0,1\}^n\setminus S$: \begin{align*} &\left\{x:\frac{x_1+\dots+x_a}a+\frac{x_{a+1}+\dots+x_{a+b}}b+\frac{x_{a+b+1}+\dots+x_{a+b+c}}c=1\right\},\\ &\left\{x:\frac{x_{a+1}+\dots+x_{a+b}}b+\frac{x_{a+b+1}+\dots+x_{a+b+c}}c=2\right\}. \end{align*} Suppose now that $a=0$ or $b=0$. Since $a+b\geq 1$ and $c\geq 1$, we have used $n-2$ hyperplanes so far. If $a=0$, we may add the hyperplane \[ \left\{x:\frac{x_{1}+\dots+x_{b}}b+\frac{x_{b+1}+\dots+x_{b+c}}c=2\right\} \] and, if $b=0$, we add \[ \left\{x:-\frac{x_1+\dots+x_a}a+\frac{x_{a+1}+\dots+x_{a+c}}c=1\right\}. \] In either case, the resulting collection covers $\{0,1\}\setminus S$ without covering any point in $S$. \end{proof} For the case of four missing points, we always need at least $n-2$ hyperplanes by Theorem \ref{thm:ATcor}. For $n=3$, we may need either $1$ or $2$ hyperplanes. For example, we may exactly cover $\{0,1\}^3\setminus (\{0\}\times \{0,1\}^2)$ by the single hyperplane $\{x:x_1=1\}$, but if $S$ does not lie on a hyperplane then we need two hyperplanes. The set $\{0\}\times \{0,1\}^2$ has the special property that there is no hyperplane that covers three of its points without covering the fourth. It turns out this condition is exactly what decides how many hyperplanes are required when removing four points. \begin{lemma} \label{lem:cov:4} Let $S\subseteq\{0,1\}^n$ with $|S|=4$. Then $\ec(\{0,1\}^n\setminus S)=n-1$ if there is a hyperplane $Q$ with $|Q\cap S|=3$ and $\ec(\{0,1\}^n\setminus S)=n-2$ otherwise. \end{lemma} \begin{proof} We know that $\ec(\{0,1\}^n\setminus S)\geq n-2$ from Theorem \ref{thm:ATcor}. If there is a hyperplane $Q$ intersecting $S$ in exactly three points, then $\ec(\{0,1\}^n\setminus S)\geq n-1$. Indeed, by vertex transitivity, we may assume that $0$ is the point of $S$ uncovered by $Q$. Any exact cover of $\{0,1\}^n\setminus S$ can be extended to an exact cover of $\{0,1\}^n\setminus \{0\}$ by adding the hyperplane $Q$ to the collection. We prove the claimed upper bounds by induction on $n$, handling the case $n\leq 7$ by computer search. Again, we may assume that $0\in S$. Let $u,v,w$ denote the other three vectors in $S$. For any $i$ with $u_i=v_i=w_i=0$, we can use a hyperplane of the form $\{x:x_i=1\}$ to reduce the covering problem to one of a lower dimension. (Note that dropping the coordinate $i$ in this case does not change whether three points in $S$ can be covered without covering the fourth.) Hence we may assume by induction that no such $i$ exists. After possibly permuting coordinates, we assume that $u_i=v_i=w_i=1$ on the first $a$ coordinates, $u_i=v_i=1$ and $w_i=0$ on the $b$ coordinates after that, and so on, i.e., sorted by decreasing Hamming weight and lexicographically within the same weight. In other words, our four vectors take the form \begin{equation} \label{eq:VennForm} \begin{pmatrix} 0\\ u\\ v\\ w\\ \end{pmatrix}= \begin{pmatrix} 0& 0 & 0& 0& 0 & 0& 0\\ 1 & 1 & 1& 0 &1 & 0 & 0 \\ 1 & 1 & 0 & 1 & 0 &1 & 0\\ 1 & 0 & 1 & 1 & 0 & 0 & 1\\ \end{pmatrix} ,\end{equation} where each column may be replaced with $0$ or more columns of its type. Since $n>7$, by the pigeonhole principle one of the columns must be repeated at least twice. We will show how to handle the case for which this is the first column (i.e. $a\geq 2$); the other cases are analogous. Our collection of hyperplanes will contain the hyperplanes \begin{equation} \label{eq:ec:a} \{\{x:x_1+\dots+x_a=i\}:i\in [a-1]\}. \end{equation} The only points $x$ which have yet to be covered have the property that $x_i$ takes the same value in $\{0,1\}$ for all $i\in [a]$. We now proceed similarly to the proof of Lemma \ref{lem:cov:23}. Informally, we wish to `merge' the first $a$ coordinates and then apply the induction hypothesis. For each $s\in S$, we define $\pi(s)=(s_{a},\dots,s_n)$. Let $\pi(S)=\{\pi(s):s\in S\}$. Then $|S|=|\pi(S)|=4$. Any hyperplane \[ P=\{y:v_1y_1+\dots+v_{n-a+1}y_{n-a+1}=\alpha\} \] in $\{0,1\}^{n-a+1}$ can be used to define a hyperplane \[ L(P)= \left\{x:v_1\frac{x_1+\dots+x_a}a+v_2x_{a+1}+\dots +v_{n-a+1} x_n=\alpha\right\} \] in $\{0,1\}^n$. For all $x\in \{0,1\}^n$ with $\sum_{i=1}^a x_i\in \{0,a\}$, we find that $\pi(x)\in P$ if and only if $x\in L(P)$. This shows that if $P_1,\dots,P_M$ form an exact cover for $\{0,1\}^{n-a+1}\setminus \pi(S)$, then $L(P_1),\dots,L(P_M)$, together with the hyperplanes from $(\ref{eq:ec:a})$, form an exact cover for $\{0,1\}^n\setminus S$. This proves \[ \ec(\{0,1\}^n\setminus S)\leq \ec(\{0,1\}^{n-a+1}\setminus \pi(S))+a-1. \] Since there is a hyperplane covering three points in $S$ without covering the fourth if and only if this is the case for $\pi(S)$, we find the claimed upper bounds by induction. Observe that the proof reduction works also in the case $n\leq 7$ if there are at least two coordinates of the same type in (\ref{eq:VennForm}). Thus, the computer verification is needed only in the case when each column in (\ref{eq:VennForm}) appears at most once. The code used to check the small cases is attached to the arXiv submission at \url{https://arxiv.org/abs/2010.00315}. \end{proof} Another natural variant on the original Alon-F\"{u}redi problem is to ask for the exact cover number of a single layer of a hypercube without one point. It turns out this can be easily solved by translating it to the original problem. \begin{proposition} Let $n\in \mathbb{N}$ and $i\in \{0,\dots,n\}$. Let $B$ be obtained by removing a single point from the $i$-th layer $\{x\in \{0,1\}^n:x_1+ \ldots + x_n=i\}$. Then $\ec(B)=\min\{i,n-i\}$. \end{proposition} \begin{proof} We may assume that $i\leq n/2$ and that $b=(1,\dots,1,0,\dots,0)$ is the missing point. The upper bound follows by taking the hyperplanes \[ \{\{x:x_1+\dots+x_i=j\}:j\in \{0,\dots,i-1\}\}. \] For the lower bound, we claim that we may find a cube of dimension $i$ within the $i$-th layer for which $b$ plays the role of the origin. Indeed, consider the affine map \[ \iota:\{0,1\}^i\to \{0,1\}^n: x \mapsto (1-x_1,1-x_2,\dots,1-x_i,0,\dots,0,x_i,x_{i-1},\dots,x_1). \] That is, we view the point $b$ as the origin and take the directions of the form $(-1,0,\dots,0,1)$, $(0,-1,0,\dots,0,1,0)$, etcetera, as the axes of the cube. Now $(B\setminus \{b\}) \cap \iota(\{0,1\}^i) = \iota(\{0,1\}^i\setminus\{0\})$, and hence we may convert any exact cover for $B\setminus\{b\}$ to an exact cover for $\{0,1\}^i\setminus\{0\}$. The lower bound follows from the result of Alon and F\"{u}redi \cite{AlonFuredi}. \end{proof} \section{Asymptotics} \label{sec:ec:asympt} We first consider the asymptotics of $\ec(n,k)$ when $k$ is held fixed. For the upper bound, we prove the following lemma. \begin{lemma} \label{lem:ec:k} For all $k\in \mathbb{N}$ and $n\geq 2^{k-1}$, $\ec(n,k)\leq 1+\ec(n-1,k)$. \end{lemma} \begin{proof} Fix $k\in \mathbb{N}$, $n\geq 2^{k-1}$ and a subset $S\subseteq \{0,1\}^n$ of size $|S|=k$. For any $i\in [n]$, let $S_{-i}\subseteq\{0,1\}^{n-1}$ be obtained from $S$ by deleting coordinate $i$ from each element of $S$. We claim that there exists an $i\in [n]$ such that $|S_{-i}|=k$ and \begin{equation} \label{eq:ec:i} \ec(\{0,1\}^n\setminus S)\leq 1+\ec(\{0,1\}^{n-1}\setminus S_{-i}). \end{equation} The lemma follows immediately from this claim. By vertex transitivity, we may assume that $0\in S$. Suppose first that there exists $i\in[n]$ for which $s_i=0$ for all $s\in S$. Then $|S_{-i}|=k$. From an exact cover for $\{0,1\}^{n-1}\setminus S_{-i}$, we may obtain an exact cover for $\{x \in \{0,1\}^n\setminus S:x_i=0\}$. Combining with the hyperplane $\{x:x_i=1\}$, this gives an exact cover for $\{0,1\}^n\setminus S$. This proves (\ref{eq:ec:i}). We henceforth assume that $0\in S$ and that the remaining $k-1$ elements of~$S$ cannot all be $0$ on the same coordinate. Hence there are at most $2^{k-1}-1$ possible values that $(s_i:s\in S)$ can take for $i\in [n]$. Since $n\geq 2^{k-1}$, by the pigeonhole principle, there must exist coordinates $1\leq i<j\leq n$ with $s_i=s_j$ for all $s\in S$. This implies that $|S_{-i}|=|S|=k$. We now show (\ref{eq:ec:i}) is satisfied. After permuting coordinates, we may assume that $(i,j)=(1,2)$. An exact cover for $\{0,1\}^{n-1}\setminus S_{-1}$ is converted to an exact cover for $\{0,1\}^n\setminus S$ as in the proof of Lemma \ref{lem:cov:4}: any hyperplane of the form \[ P=\{y:v_1y_1+\dots +v_{n-1}y_{n-1}=\alpha\} \] is converted to \[ L(P)=\left\{x:v_1\frac{x_1+x_2}2+v_2x_3+\dots +v_{n-1}x_n=\alpha\right\}, \] and we add the hyperplane $\{x:x_1+x_2=1\}$ to the adjusted collection. \end{proof} It is now easy to prove to that $\ec(n,k)=n+\Theta_k(1)$. \begin{proof}[Proof of Theorem \ref{thm:ec:fixed_size}] Let $k\in \mathbb{N}$. We need to prove that for all $n\geq 2^k$, \[ n-\log_2(k)\leq \ec(n,k)\leq n-2^k+\ec(2^k,k). \] The upper bound is vacuous for $n=2^k$ and follows from $n-2^k$ applications of Lemma \ref{lem:ec:k} for $n>2^k$. The lower bound follows from Theorem \ref{thm:ATcor}: if $n-\ell$ hyperplanes cover all but $k$ vertices, then $k\geq 2^\ell$, and hence $n-\ell\geq n-\log_2(k)$. (In fact, this shows $\ec(\{0,1\}^n\setminus S)\geq n-\log_2(k)$ for each subset $S\subseteq\{0,1\}^n$ of size $k$.) \end{proof} We now turn to the problem of comparing exact cover numbers for sets $S$ of different sizes. We use two auxiliary lemmas. For the lower bound, we use a random argument for which we need to know the approximate number of intersection patterns of the hypercube. An \textit{intersection pattern} of $\{0,1\}^n$ is a non-empty subset $P\subseteq \{0,1\}^n$ for which there exists a hyperplane $H$ with $H\cap \{0,1\}^n=P$. \begin{lemma} \label{lem:q_n_num_int_patterns} $\{0,1\}^n$ has at most $2^{n^2}$ possible intersection patterns. \end{lemma} \begin{proof} We will associate each intersection pattern with a unique element from $(\{0,1\}^n)^n$. Let $P\subseteq \{0,1\}^n$ be an intersection pattern with $P= H\cap \{0,1\}^n$ for $H$ a hyperplane. Then $|P|<2^n$. Let $x \in P$ be such that $\sum_{i=1}^nx_i2^i$ is minimal. Let $\oplus$ denote coordinate-wise addition modulo $2$ and write $x\oplus P=\{x\oplus p:p\in P\}\subseteq \{0,1\}^n$. Note that $0\in x\oplus P$ since $x\in P$, and that $x\oplus P$ is the intersection of a linear subspace of dimension $n-1$ with $\{0,1\}^n$. (The linear subspace can be obtained from $H$ by a series of reflections.) We greedily find $0\leq k\leq n-1$ linearly independent vectors $v_1,\dots,v_k\in x\oplus P$ whose linear span intersects $\{0,1\}^n$ in $x \oplus P$. We label $P$ with the $n$-tuple $(x,v_1,\dots,v_k,0,\dots,0)$, where we added $n-1-k$ copies of the vector $0$ at the end of the tuple. This associates each intersection pattern to a unique element from $(\{0,1\}^n)^n$. \end{proof} The above proof is rather crude, but in fact not far from the truth: the number of possible intersection patterns is $2^{(1+o(1))n^2}$ (see e.g. \cite[Lemma 4.3]{Baldi}). We also use an auxiliary result for the upper bound. The \textit{total domination number} of a graph $G$ is the minimum cardinality of a subset $D\subseteq V(G)$ such that each $v\in V(G)$ has a neighbour in $D$. \begin{lemma}[Theorem 5.2 in \cite{totaldomhypercube}] \label{lem:total_dom_num} The total domination number of the hypercube is at most $2^{n+1}/n$ for any $n \geq 1$. \end{lemma} Note that this bound must be close to tight since the hypercube is a regular graph of degree $n$, so any total dominating set has cardinality at least $2^n/n$. We are now ready to prove $2^{n-2}/n^2\leq \ec(n) \leq 2^{n+1}/n$. \begin{proof}[Proof of Theorem \ref{thm:ec:arbitrary}] For the lower bound, we need to give a subset $B\subseteq \{0,1\}^n$ that is difficult to cover exactly. We will find a subset $S$ for which all ``large" intersection patterns have a non-empty intersection with $S$. This means that to cover $\{0,1\}^n\setminus S$, we can only use hyperplanes with ``small" intersection patterns. We take a subset $S\subseteq \{0,1\}^n$ at random by including each point independently with probability $1/2$. Note that the lower bound is trivial for $n \leq 8$, so we may assume that $n > 8$. For any fixed intersection pattern $P$, the probability that it is disjoint from our random set $S$ is $\left(\frac12\right)^{|P|}$. By Lemma \ref{lem:q_n_num_int_patterns}, there are at most $2^{n^2}$ possible intersection patterns. Hence, by the union bound, the probability that there is an intersection pattern which has at least $2n^2$ elements and does not intersect with $S$, is at most $2^{n^2}\left(\frac12\right)^{2n^2}$, which is smaller than $1/2$ for $n \geq 2$. With probability at least $1/2$, our random set $S$ has at most $2^{n-1}$ points. Hence, there exists a subset $S$ of size $2^{n-1}$ that `hits' all intersection patterns of size at least $2n^2$. Any exact cover for $\{0,1\}^n \setminus S$ consists entirely of hyperplanes whose intersection pattern has size at most $2n^2$, and hence needs at least $|\{0,1\}^n\setminus S|/2n^2=2^{n-2}/n^2$ hyperplanes. We now prove the upper bound. The Hamming distance on $\{0,1\}^n$ is given by $d(x,y)=\sum_{i=1}^n |x_i-y_i|$. A Hamming sphere of radius 1 around a point $x\in \{0,1\}^n$ is given by $S(x)=\{y\in \{0,1\}^n:d(x,y)=1\}$. We claim that any subset of a Hamming sphere is an intersection pattern. Since the cube is vertex-transitive, it suffices to prove our claim for $S(0)$. The hyperplane $\{x:\sum_{i=1}^n x_i=1\}$ intersects $\{0,1\}^n$ in $S(0)$. Intersecting that hyperplane with hyperplanes of the form $\{x:x_j=0\}$ gives a lower-dimensional affine subspace, and we can construct such a subspace which intersects $S(0)$ in any subset we desire. In order to turn the affine subspace into a hyperplane with the same intersection pattern, we may add generic directions that do not yield new points in the hypercube (e.g. consider adding $(1,\pi,0,\dots,0)$). This proves each subset of a Hamming sphere is an intersection pattern. The hypercube has total domination number at most $2^{n+1}/n$ by Lemma~\ref{lem:total_dom_num}. Hence, we can find a subset $D$ of the cube such that each vertex has a neighbour in $D$. In particular, there are $M\leq 2^{n+1}/n$ Hamming spheres centered on the vertices in $D$ that cover the cube. For any $B\subseteq \{0,1\}^n$, we write $B=B_1\cup \dots \cup B_M$ such that each $B_i$ is covered by at least one of the Hamming spheres. This means that each $B_i$ is a intersection pattern, and we may cover $B$ exactly using $M$ hyperplanes. This gives the desired exact cover of $B$ with at most $2^{n+1}/n$ hyperplanes. \end{proof} Noga Alon pointed out the following improvement on the constant of the lower bound in Theorem~\ref{thm:ec:arbitrary}. There are at most $2^{n^2}$ possible intersection patterns by Lemma~\ref{lem:q_n_num_int_patterns}, so if all possible nonempty $B\subseteq \{0,1\}^n$ can be achieved by taking a union of $x$ of them, then $2^{n^2x}\geq 2^{2^n} -1$. The left-hand side of this inequality is even and the right-hand side is odd, hence $2^{n^2x}\geq 2^{2^n}$ and so $x\geq \frac{2^{n}}{n^2}$. \section{Conclusion} \label{sec:ec:concl} Based on the fact that $\ec(n,k)\leq n$ for $k=1,2,3,4$, one might hope to prove that in fact $\ec(n,k)\leq n+C$ for some constant $C>0$ (independent of~$k$). However, this is not true in general by Theorem \ref{thm:ec:arbitrary}. A~natural question is then whether this will be true for $n$ sufficiently large when $k$ is fixed. \begin{problem} Is there a constant $C>0$, such that for any $k\in \mathbb{N}$ there exists a $n_0(k)\in \mathbb{N}$ such that $\ec(n,k)\leq n+C$ for all $n\geq n_0(k)$? \end{problem} In an earlier version of this paper \cite{AaronsonGGJK20}, we conjectured that for any $S\subseteq \{0,1\}^r$ and $n\in \mathbb{N}$ with $n\geq r$, \[ \ec(\{0,1\}^{n}\setminus (S\times\{0\}^{n-r}))=\ec(\{0,1\}^r\setminus S) +n-r. \] This would have given a negative answer to the problem above, but the counterexample $S = \{1000, 1111,1001,1011,0110,0001,0010,0111\}$ when $n = 6$ was given by Adam Zsolt Wagner \cite{Adam}. \smallskip One approach to improving the lower bound in Theorem \ref{thm:ec:arbitrary} is to try to prove that, for some $\varepsilon\in (0,1)$, the number of hyperplanes containing $n^{1+\varepsilon}$ points is $O(2^{n^{1+\varepsilon}})$. Unfortunately, this is false: there are $2^{(1+o(1))n^2}$ possible intersection patterns of size at least $n^2$. This can be seen by considering intersection patterns of the form $\{0,1\}^{\log_2(n^2)}\times B$ for $B\subseteq\{0,1\}^{n-\log_2(n^2)}$. (If $B$ is a non-empty intersection pattern, then $\{0,1\}^{\log_2(n^2)}\times B$ is an intersection pattern containing at least $n^2$ points.) On the other hand, by taking every other layer we may intersect each `axis-aligned subcube' of the form $\{0,1\}^a\times \{x\}$, ensuring that no such intersection pattern can be used in a cover. However, there is a more general type of subcube to consider. We say a subset $A\subseteq \{0,1\}^n$ of size $|A|=2^d$ forms a $d$-dimensional \emph{subcube} if there are vectors $u,w_1,\dots,w_d\in \mathbb{R}^n$ such that \[ A=\{u+\alpha_1w_1+\dots+\alpha_d w_d: \alpha_1,\dots,\alpha_d \in \{0,1\}\}. \] A solution to the following problem might help improve either the upper or lower bound of Theorem \ref{thm:ec:arbitrary}. \begin{problem} Fix $n,d \in \mathbb{N}$. What is the smallest cardinality of a subset $S\subseteq\{0,1\}^n$ for which $A\cap S\neq \emptyset$ for all $d$-dimensional subcubes $A \subseteq \{0,1\}^n$? \end{problem} This is of a similar flavour to a problem proposed by Alon, Krech and Szab\'{o}~\cite{AlonKrechSzabo}, who asked instead for the asymptotics of the above problem when the cubes have to be axis-aligned. A $d$-dimensional axis-aligned subcube is of the form $\{0,1\}^d\times\{x\}$ after permuting coordinates. Let $g(n,d)$ denote the minimal cardinality of a subset that hits all $d$-dimensional axis-aligned subcubes in $\{0,1\}^n$. The best-known asymptotic bounds for $g(n,d)$ are from~\cite{AlonKrechSzabo}: \[ \frac{\log_2(d)}{2^{d+2}}\leq \lim_{n\to \infty}\frac{g(n,d)}{2^n} \leq \frac1{d+1}. \] Finally, we remark that we have already seen these subcubes come up in Lemma \ref{lem:cov:4} as well: the sets $S\subseteq \{0,1\}^n$ of size 4 with $\ec(\{0,1\}^n\setminus S)=n-2$ are exactly the 2-dimensional subcubes. \subsection*{Acknowledgements} We thank Noga Alon for pointing out to us that the problem had also been suggested by F\"{u}redi and for providing the reference~\cite{Baldi}. We would also like to thank Alex Scott for useful discussions and the Jagiellonian University for their hospitality in hosting us during the time this research was conducted. \bibliographystyle{abbrv}
{ "timestamp": "2021-07-02T02:25:57", "yymm": "2010", "arxiv_id": "2010.00315", "language": "en", "url": "https://arxiv.org/abs/2010.00315", "abstract": "Alon and Füredi (1993) showed that the number of hyperplanes required to cover $\\{0,1\\}^n\\setminus \\{0\\}$ without covering $0$ is $n$. We initiate the study of such exact hyperplane covers of the hypercube for other subsets of the hypercube. In particular, we provide exact solutions for covering $\\{0,1\\}^n$ while missing up to four points and give asymptotic bounds in the general case. Several interesting questions are left open.", "subjects": "Combinatorics (math.CO)", "title": "Exact hyperplane covers for subsets of the hypercube", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9932024678460111, "lm_q2_score": 0.8056321889812553, "lm_q1q2_score": 0.8001558782723668 }
https://arxiv.org/abs/2301.11638
One-dimensional integral Rellich type inequalities
The motive of this note is twofold. Inspired by the recent development of a new kind of Hardy inequality, here we discuss the corresponding Hardy-Rellich and Rellich inequality versions in the integral form. The obtained sharp Hardy-Rellich type inequality improves the previously known result. Meanwhile, the established sharp Rellich type integral inequality seems new.
\section{Introduction} In the celebrated paper \cite{hardy}, Godfrey H. Hardy first stated the famous inequality which reads as: let $1<p<\infty$ and $f$ be a $p$-integrable function on $(0, \infty)$, which vanishes at zero, then the function $r\longmapsto \frac{1}{r}\int_{0}^{r} f(t) \:{\rm d}t$ is $p$-integrable over $(0, \infty)$ and there holds \begin{align}\label{hardy} \int_{0}^{\infty}\bigg|\frac{1}{r}\int_{0}^{r} f(t)\:{\rm d}t\bigg|^p\:{\rm d}r\leq \bigg(\frac{p}{p-1}\bigg)^p \int_{0}^{\infty}|f(r)|^p\:{\rm d}r. \end{align} The constant on the right-hand side of \eqref{hardy} is sharp. The development of the famous Hardy inequality \eqref{hardy} during the period 1906-1928 has its own history and we refer to \cite{kuf} (also, see the preface of \cite{hardy-book-rs}). Recent progress by Frank-Laptev-Weidl \cite{nh} presents a novel one-dimensional inequality with the same sharp constant, which improves the classical Hardy inequality \eqref{hardy}. This new version looks as follows: let $1<p<\infty$, then for any $f\in L^p(0,\infty)$, which vanishes at zero, one has \begin{align}\label{new-hardy} \int_{0}^{\infty}\sup_{0<s<\infty}\bigg|\min\biggr\{\frac{1}{r},\frac{1}{s}\biggr\}\int_{0}^{s}f(t)\:{\rm d}t \bigg|^p\:{\rm d}r\leq \bigg(\frac{p}{p-1}\bigg)^p \int_{0}^{\infty}|f(r)|^p\:{\rm d}r. \end{align} Certainly, \eqref{new-hardy} gives an improvement of \eqref{hardy}. Recently, the multidimensional version in the supercritical case and the discrete version of \eqref{new-hardy} have been established in \cite{nhm} and \cite{nhd}, respectively. In the same spirit, one may ask about the possible structure of Hardy-Rellich and Rellich type inequalities. In this short note, we answer affirmatively about the possible forms of these two types of inequalities. Let us recall the one-dimensional Hardy-Rellich inequality. For $f\in C^2(0,\infty)$ with $f^\prime(0)=0$, there holds \begin{align}\label{hardy-rel} \int_{0}^{\infty}\frac{|f^\prime(r)|^2}{r^{2}}\:{\rm d}r \leq 4 \int_{0}^{\infty}|f^{\prime\prime}(r)|^2\:{\rm d}r. \end{align} Starting from it, there have been several articles in which the authors studied many improvements of inequality \eqref{hardy-rel}. Here we mention only a few of them \cite{bgr-22, coss,cazacu, ya,tz} and references therein. Now let us write \eqref{hardy-rel} in the integral form. Note that it can be derived from the weighted 1D classical Hardy inequality. This reads as follows: let $f\in C^1(0,\infty)$, then there holds \begin{align}\label{hardy-rel-int} \int_{0}^{\infty}\frac{|\int_{0}^{r}f^\prime(t)\:{\rm d}t|^2}{r^{2}}\:{\rm d}r\leq 4\int_{0}^{\infty}|f^{\prime}(r)|^2\:{\rm d}r. \end{align} Here the constant $4$ is sharp. We give an improved version of this inequality in Theorem \ref{hardy-rel-th}. Let us briefly mention another important function inequality the so-called Rellich inequality which was first introduced in \cite{rel}. It is worth recalling the one-dimensional Rellich inequality. The classical (1D-)Rellich inequality states that for $f\in C^2(0,\infty)$ with $f(0)=0$ and $f^\prime(0)=0$, there holds \begin{align}\label{rel} \int_{0}^{\infty}\frac{|f(r)|^2}{r^{4}}\:{\rm d}r\leq \frac{16}{9} \int_{0}^{\infty}|f^{\prime\prime}(r)|^2\:{\rm d}r. \end{align} Over the past few decades, there has been a constant effort to improve \eqref{rel}. Here are some references \cite{hinz,ozawa, BT, MNSS,bmo,rsadv, cm}. In this short contribution, we also obtain another type of the Rellich inequality (see Theorem \ref{rel-th} with $p=2$). To the best of our knowledge, the most recent progress in this direction was made in \cite{bmo}. However, a one-dimensional study is still missing. Thus, trying to fill this gap is another motivation for the present paper. Taking inspiration from there we obtain the following version of Rellich inequality, which reads as: for $f\in L^2(0,\infty)$ there holds \begin{align}\label{n-rel-2} &\int_{0}^{\infty}\frac{1}{r^{4}}\bigg(\int_{0}^{r}\int_{0}^{\tau}|f(t)|\:{\rm d}t\:{\rm d}\tau\bigg)^2\:{\rm d}r\nonumber\\&\leq\int_{0}^{\infty}\frac{1}{r^{4}}\bigg(\int_{0}^{r}\sup_{0<s<\infty}\min\biggl\{1,\frac{\tau}{s}\biggr\}\int_{0}^{s}|f(t)|\:{\rm d}t\:{\rm d}\tau\bigg)^2\:{\rm d}r\nonumber \\&\leq \frac{16}{9} \int_{0}^{\infty}|f(r)|^2\:{\rm d}r. \end{align} Moreover, we will establish the constant $16/9$ is a sharp constant. Therefore, \eqref{n-rel-2} can be compared with \eqref{rel}. \color{black} Note that we have mentioned only the $L^2(0,\infty)$ case but we will discuss the result for the general $L^p(0,\infty)$ case. The plan of this paper is simple. In Section \ref{prelm} we discuss a few preliminaries and then we state the main results. In Section \ref{proof} we complete the proofs of the main theorems. In the final section, we will briefly mention some more related inequalities. \section{Preliminaries and main results}\label{prelm} Let us begin this section with basic facts about \emph{symmetric decreasing rearrangement}. For more details, we refer to \cite[Chapter 3]{lieb}. We denote $f^*$ the symmetric decreasing rearrangement of $f$. It is well known that $f^*$ is a nonnegative, radially symmetric, and nonincreasing function. Irrespective of several properties of $f^*$, the useful properties in our context is the equimeasurability property, i.e. \begin{align*} \text{vol}(\{|f|>\tau\})=\text{vol}(\{f^*>\tau\}) \text{ for all }\tau\geq 0. \end{align*} By using the \emph{layer cake representation} and the above property, we have the following helpful identity: \begin{align}\label{norm-presv} \int_{0}^\infty|f(t)|^p\:{\rm d}t=\int_{0}^\infty|f^*(t)|^p\:{\rm d}t\:\: \text{ for all }p\geq 1. \end{align} Also it is clear that for any $s>0$ there holds \begin{align}\label{norm-comp} \int_0^s|f(t)|\:{\rm d}t\leq \int_0^sf^*(t)\:{\rm d}t. \end{align} These relations will be very much valuable in the proofs. Now, we are ready to state the following important observation. \begin{lemma}\label{rel-lem-1} Let $f$ be a locally absolutely continuous function on $(0,\infty)$. Then for a fixed $r>0$ the following identity holds: \begin{align}\label{rel-eq-1} \sup_{0<s<\infty}\min\biggl\{1,\frac{r}{s}\biggr\}\int_{0}^{s}f^*(t)\:{\rm d}t=\int_{0}^{r}f^*(t)\:{\rm d}t, \end{align} where $f^*$ is the non-increasing rearrangement of $f$. \end{lemma} \begin{proof} We wish to calculate the supremum by using the advantage of the non-increasing property of $f^*$. For any fixed $r>0$, we consider the following two cases: {\bf Case 1:} Let $0<s\leq r$. Then we obtain \begin{align*} \min\biggl\{1,\frac{r}{s}\biggr\}\int_{0}^{s}f^*(t)\:{\rm d}t=\int_{0}^{s}f^*(t)\:{\rm d}t\leq \int_{0}^{r}f^*(t)\:{\rm d}t. \end{align*} {\bf Case 2:} Let $r\leq s<\infty$. Then we have by change of variable \begin{align*} \min\biggl\{1,\frac{r}{s}\biggr\}\int_{0}^{s}f^*(t)\:{\rm d}t=\frac{r}{s}\int_0^{s}f^*(t)\:{\rm d}t\leq\frac{r}{s}\int_{0}^{s}f^*(tr/s)\:{\rm d}t=\int_{0}^{r}f^*(v)\:{\rm d}v. \end{align*} In both cases, we get \begin{align*} \min\biggl\{1,\frac{r}{s}\biggr\}\int_{0}^{s}f^*(t)\:{\rm d}t \leq \int_{0}^{r}f^*(t)\:{\rm d}t. \end{align*} Hence the supremum is attained at $s=r$ and we arrive at \begin{align*} \sup_{0<s<\infty}\min\biggl\{1,\frac{r}{s}\biggr\}\int_{0}^{s}f^*(t)\:{\rm d}t=\int_{0}^{r}f^*(t)\:{\rm d}t. \end{align*} \end{proof} Now we are ready to present an improvement of \eqref{hardy-rel-int}. That is, this gives a natural improvement of the Hardy-Rellich inequality in the integral form. Below we will describe the corresponding differential form which improves the original Hardy-Rellich inequality \eqref{hardy-rel} in a simple form. \begin{theorem}\label{hardy-rel-th} Let $g\in C^1(0,\infty)$, then there holds \begin{align}\label{eqn-hardy-rel-th} \int_{0}^{\infty}\sup_{0<s<\infty}\bigg|\min\biggl\{\frac{1}{r},\frac{1}{s}\biggr\}\int_{0}^{r}g^\prime(t)\:{\rm d}t\bigg|^2\:{\rm d}r\leq 4 \int_{0}^{\infty}|g^\prime(r)|^2\:{\rm d}r. \end{align} Moreover, the constant is sharp. \end{theorem} \begin{remark}\label{rem-hardy-rel} Let $f\in C^2(0,\infty)$, with $f^\prime(0)=0$. Then choosing $g(r)=f^\prime(r)$ in the improved Hardy-Rellich inequality \eqref{eqn-hardy-rel-th}, we obtain \begin{align}\label{dif-hr-1} \int_{0}^{\infty}\sup_{0<s<\infty}\bigg|\min\biggl\{\frac{1}{r},\frac{1}{s}\biggr\}f^\prime(s)\bigg|^2\:{\rm d}r\leq 4\int_{0}^{\infty}|f^{\prime\prime}(r)|^2\:{\rm d}r. \end{align} Moreover, it is straightforward to check the following identity \begin{align}\label{dif-hr-2} \sup_{0<s<\infty}\bigg|\min\biggl\{\frac{1}{r},\frac{1}{s}\biggr\}\;f^\prime(s)\bigg|^2=\max\biggl\{ \sup_{0<s\leq r}\frac{|f^\prime(s)|^2}{r^2}\:,\: \sup_{r\leq s<\infty}\frac{|f^\prime(s)|^2}{s^2}\biggr\}. \end{align} Then combining \eqref{dif-hr-1} and \eqref{dif-hr-2} we deduce \begin{align}\label{hardy-rel-impv} \int_{0}^{\infty}\frac{|f^\prime(r)|^2}{r^{2}}\:{\rm d}r\leq \int_{0}^{\infty}\max\biggl\{ \sup_{0<s\leq r}\frac{|f^\prime(s)|^2}{r^2}\:,\: \sup_{r\leq s<\infty}\frac{|f^\prime(s)|^2}{s^2}\biggr\}\:{\rm d}r \leq 4 \int_{0}^{\infty}|f^{\prime\prime}(r)|^2\:{\rm d}r. \end{align} Therefore, \eqref{hardy-rel-impv} appears as an immediate improvement of \eqref{hardy-rel} in the one-dimension differential form. The similar discussion gives the improvement of \eqref{hardy-rel-int} as well (cf. \cite[Theorem 1.3]{coss}). \end{remark} Now we are going to discuss the second main result of this note. Before presenting the statement first let us recall the classical one-dimensional $L^{p}$-Rellich inequality. This reads as follows: let $p>1$, $f\in C^2(0,\infty)$ with $f(0)=0$ and $f^\prime(0)=0$ there holds \begin{align}\label{p-rel} \int_{0}^{\infty}\frac{|f(r)|^p}{r^{2p}}\:{\rm d}r\leq \frac{p^{2p}}{(p-1)^p(2p-1)^p} \int_{0}^{\infty}|f^{\prime\prime}(r)|^p\:{\rm d}r. \end{align} Now we are ready to demonstrate the one-dimensional Rellich-type inequality in the following integral form. \begin{theorem}\label{rel-th} Let $f\in L^p(0,\infty)$, $p>1$. Then we have \begin{align}\label{eqn-rel-th} &\int_{0}^{\infty}\frac{1}{r^{2p}}\bigg(\int_{0}^{r}\int_{0}^{\tau}|f(t)|\:{\rm d}t\:{\rm d}\tau\bigg)^p\:{\rm d}r\nonumber\\&\leq\int_{0}^{\infty}\frac{1}{r^{2p}}\bigg(\int_{0}^{r}\sup_{0<s<\infty}\min\biggl\{1,\frac{\tau}{s}\biggr\}\int_{0}^{s}|f(t)|\:{\rm d}t\:{\rm d}\tau\bigg)^p\:{\rm d}r\nonumber \\&\leq \frac{p^{2p}}{(p-1)^p(2p-1)^p} \int_{0}^{\infty}|f(r)|^p\:{\rm d}r. \end{align} Moreover, the constant is sharp. \end{theorem} \section{Proofs of Theorems \ref{hardy-rel-th} and \ref{rel-th}}\label{proof} This section is concerned with the proofs of Theorems \ref{hardy-rel-th} and \ref{rel-th}. Before going further let us recall the following lemma. \begin{lemma}\cite[Lemma 3.1]{nhm}\label{hr-lemma} Let $1<p<\infty$. Let $g$ be any nonnegative function on $(0,\infty)$. Assume $h$ is a strictly positive non-decreasing function on $(0,\infty)$ such that $s h(r)\leq r h(s)$ for any $r,s\in(0,\infty)$ with $r\leq s$. Let $f$ be a locally absolutely continuous function on $(0,\infty)$. Then we have \begin{equation*} \int_{0}^{\infty}g(r)\sup_{0<s<\infty}\bigg|\min\biggl\{\frac{1}{h(r)},\frac{1}{h(s)}\biggr\}\int_{0}^{s}f(t)\:{\rm d}t\bigg|^p\:{\rm d}r \leq \int_{0}^{\infty}g(r)\bigg|\frac{1}{h(r)}\int_{0}^{r}f^*(t)\:{\rm d}t\bigg|^p\:{\rm d}r, \end{equation*} where $f^*$ is the non-increasing rearrangement of $f$. \end{lemma} Now, as a direct consequence of Lemma \eqref{hr-lemma}, we derive the proof of Theorem \ref{hardy-rel-th}. {\bf Proof of Theorem \ref{hardy-rel-th}:} Let us consider $g(r)=1$ and $h(r)=r$ be functions on $(0,\infty)$. Substituting these in Lemma \ref{hr-lemma} with $p=2$, then we have \begin{equation*} \int_{0}^{\infty}\sup_{0<s<\infty}\bigg|\min\biggl\{\frac{1}{r},\frac{1}{s}\biggr\}\int_{0}^{s}g^{\prime}(t)\:{\rm d}t\bigg|^2\:{\rm d}r \leq\int_{0}^{\infty}\frac{1}{r^{2}}\bigg|\int_{0}^{r}(g^{\prime})^*(t)\:{\rm d}t\bigg|^2\:{\rm d}r. \end{equation*} By using the Hardy-Rellich inequality in the form \eqref{hardy-rel-int} for the function $(g^\prime)^*$, we obtain \begin{align*} \int_{0}^{\infty}\sup_{0<s<\infty}\bigg|\min\biggl\{\frac{1}{r},\frac{1}{s}\biggr\}\int_{0}^{s}g^\prime(t)\:{\rm d}t\bigg|^2\:{\rm d}r&\leq 4\int_{0}^{\infty}|{(g^\prime)}^{*}(r)|^2\:{\rm d}r\\&=4\int_{0}^{\infty}|g^\prime(r)|^2\:{\rm d}r. \end{align*} In the last step, we have used norm preserving property \eqref{norm-presv}. The sharpness follows from the optimality of the constant in \eqref{hardy-rel-int}. It ends the proof. {\bf Proof of Theorem \ref{rel-th}:} The first inequality follows from the property of the supremum. Now taking the integral of \eqref{rel-eq-1} from $0$ to $r$ we have \begin{align}\label{rel-eq-2} \int_{0}^{r}\sup_{0<s<\infty}\min\biggl\{1,\frac{\tau}{s}\biggr\}\int_{0}^{s}f^*(t)\:{\rm d}t\:{\rm d}\tau=\int_{0}^{r}\int_{0}^{\tau}f^*(t)\:{\rm d}t\:{\rm d}\tau. \end{align} \begin{align*} &\int_{0}^{\infty}\frac{1}{r^{2p}}\bigg(\int_{0}^{r}\sup_{0<s<\infty}\min\biggl\{1,\frac{\tau}{s}\biggr\}\int_{0}^{s}|f(t)|\:{\rm d}t\:{\rm d}\tau\bigg)^p\:{\rm d}r\\&\overset{\eqref{norm-comp}}{\leq} \int_{0}^{\infty}\frac{1}{r^{2p}}\bigg(\int_{0}^{r}\sup_{0<s<\infty}\min\biggl\{1,\frac{\tau}{s}\biggr\}\int_{0}^{s}f^*(t)\:{\rm d}t\:{\rm d}\tau\bigg)^p\:{\rm d}r\\&\overset{\eqref{rel-eq-2}}{=}\int_{0}^{\infty}\frac{1}{r^{2p}}\bigg(\int_{0}^{r}\int_{0}^{\tau}f^*(t)\:{\rm d}t\:{\rm d}\tau\bigg)^p\:{\rm d}r\\&\overset{\eqref{p-rel}}{\leq}\frac{p^{2p}}{(p-1)^p(2p-1)^p}\int_{0}^{\infty}|f^*(r)|^p\:{\rm d}r\\&\overset{\eqref{norm-presv}}{=}\frac{p^{2p}}{(p-1)^p(2p-1)^p}\int_{0}^{\infty}|f(r)|^p\:{\rm d}r. \end{align*} This completes the proof of the new type Rellich inequality in the integral form. {\bf Optimality:} We set \begin{align}\label{rel-const} C_{p}:=\inf_{f\in L^p(0,\infty)\setminus\{0\}}\frac{\int_{0}^{\infty}\frac{1}{r^{2p}}\big(\int_{0}^{r}\int_{0}^{\tau}|f(t)|\:{\rm d}t\:{\rm d}\tau\big)^p\:{\rm d}r}{\int_{0}^{\infty}|f(r)|^p\:{\rm d}r}. \end{align} The validity of \eqref{eqn-rel-th} immediately implies \begin{align*} C_{p}\leq \frac{p^{2p}}{(p-1)^p(2p-1)^p}. \end{align*} So it remains to show other side and this will be done by giving a proper minimizing sequence. We divide the proof in some steps. {\bf Step 1.} Let us start with a cut-off function $\chi:[0,\infty)\rightarrow \mathbb{R}$ with the following properties: \begin{itemize} \item[1.] $\chi(r)\in [0,1]$ for all $r\in [0,\infty)$ and $\chi$ is smooth; \item[2.] $\chi$ satisfies the following \begin{equation*} \chi(r)= \begin{dcases} 1 & 0\leq r\leq 1, \\ 0 & 2\leq r< \infty; \\ \end{dcases} \end{equation*} \item[3.] $\chi$ is decreasing function, i.e. $\chi^\prime(r)\leq 0$ for all $r\in [0,\infty)$. \end{itemize} Now for a small $\epsilon>0$, let us define the minimizing functions $\{f_\epsilon\}$ as follows: \begin{align*} f_\epsilon(r):=r^{\frac{\epsilon-1}{p}}\chi(r). \end{align*} {\bf Step 2.} In this step we will estimate r.h.s. of \eqref{eqn-rel-th}. The denominator of \eqref{rel-const} gives \begin{align}\label{denom} \int_{0}^{\infty}|f_\epsilon(r)|^p\:{\rm d}r&=\int_0^\infty r^{\epsilon-1}\chi^p(r)\:{\rm d}r\nonumber \\&=\int_0^1 r^{\epsilon-1}\:{\rm d}r+\int_1^2r^{\epsilon-1}\chi^p(r)\:{\rm d}r\nonumber \\&=\frac{1}{\epsilon}+O(1). \end{align} Therefore, for a fixed positive $\epsilon$, we have $f_\epsilon\in L^p(0,\infty)$. {\bf Step 3.} In this part we will evaluate the numerator of \eqref{rel-const}. Consider the integration by parts and we compute \begin{align}\label{numeo} &\int_{0}^{\infty}\frac{1}{r^{2p}}\bigg(\int_{0}^{r}\int_{0}^{\tau}|f_\epsilon(t)|\:{\rm d}t\:{\rm d}\tau\bigg)^p\:{\rm d}r\nonumber \\&=\int_{0}^{\infty}\frac{1}{r^{2p}}\bigg(\int_{0}^{r}\int_{0}^{\tau}t^{\frac{\epsilon-1}{p}}\chi(t)\:{\rm d}t\:{\rm d}\tau\bigg)^p\:{\rm d}r\nonumber \\&=\bigg(\frac{p}{\epsilon-1+p}\bigg)^p\int_{0}^{\infty}\frac{1}{r^{2p}}\bigg[\int_0^r\chi(\tau)\tau^{\frac{\epsilon-1+p}{p}}\:{\rm d}\tau-\int_{0}^{r}\int_{0}^{\tau}t^{\frac{\epsilon-1+p}{p}}\chi^\prime(t)\:{\rm d}t\:{\rm d}\tau\bigg]^p\:{\rm d}r\nonumber \\&\geq \bigg(\frac{p}{\epsilon-1+p}\bigg)^p\int_{0}^{\infty}\frac{1}{r^{2p}}\bigg[\int_0^r\chi(\tau)\tau^{\frac{\epsilon-1+p}{p}}\:{\rm d}\tau\bigg]^p\:{\rm d}r\nonumber \\&=\bigg(\frac{p}{\epsilon-1+p}\bigg)^p\bigg(\frac{p}{\epsilon-1+2p}\bigg)^p\int_{0}^{\infty}\frac{1}{r^{2p}}\bigg[\chi(r)r^{\frac{\epsilon-1+2p}{p}}-\int_{0}^{r}\tau^{\frac{\epsilon-1+2p}{p}}\chi^\prime(\tau)\:{\rm d}\tau\bigg]^p\:{\rm d}r\nonumber \\&\geq \bigg(\frac{p}{\epsilon-1+p}\bigg)^p\bigg(\frac{p}{\epsilon-1+2p}\bigg)^p\int_{0}^{\infty}r^{\epsilon-1}\chi^p(r)\:{\rm d}r\nonumber \\&=\bigg(\frac{p}{\epsilon-1+p}\bigg)^p\bigg(\frac{p}{\epsilon-1+2p}\bigg)^p\bigg[\int_0^1 r^{\epsilon-1}\:{\rm d}r+\int_1^2r^{\epsilon-1}\chi^p(r)\:{\rm d}r\bigg]\nonumber \\&=\frac{1}{\epsilon}\bigg(\frac{p}{\epsilon-1+p}\bigg)^p\bigg(\frac{p}{\epsilon-1+2p}\bigg)^p+O(1). \end{align} In between, exploiting $\chi^\prime\leq 0$, we used a simple inequality $(a+b)^p\geq a^p$ twice, for nonnegative real numbers $a$ and $b$ with $p>1$. {\bf Step 4.} Finally, by using \eqref{denom} and \eqref{numeo} we deduce the ratio \begin{align*} &\frac{\int_{0}^{\infty}\frac{1}{r^{2p}}\big(\int_{0}^{r}\int_{0}^{\tau}|f(t)|\:{\rm d}t\:{\rm d}\tau\big)^p\:{\rm d}r}{\int_{0}^{\infty}|f(r)|^p\:{\rm d}r}\\ &\geq \frac{\frac{1}{\epsilon}\big(\frac{p}{\epsilon-1+p}\big)^p\big(\frac{p}{\epsilon-1+2p}\big)^p+O(1)}{\frac{1}{\epsilon}+O(1)}\rightarrow \frac{p^{2p}}{(p-1)^p(2p-1)^p}\; \text{ for }\epsilon\rightarrow 0. \end{align*} Hence $\{f_\epsilon\}$ is the required minimizing sequence and, in turn, we have $$C_p=\frac{p^{2p}}{(p-1)^p(2p-1)^p}.$$ \section{Few more remarks}\label{inq} Here we shortly revisit some more inequalities. It hints that rescaling above inequalities may generate new inequalities. Let us recall the differential form of the inequality \eqref{new-hardy}. \begin{lemma}\cite[Theorem 1]{nh} Let $1 < p < \infty$. Then, for any locally absolutely continuous function $f$ on $(0, \infty)$ with $\liminf_{r\rightarrow 0} |f(r)| = 0$, there holds \begin{align*} \int_{0}^{\infty}\max\biggl\{\sup_{0<s\leq r} \frac{|f(s)|^p}{r^p}, \sup_{r\leq s<\infty} \frac{|f(s)|^p}{s^p} \biggr\}\:{\rm d}r\leq \bigg(\frac{p}{p-1}\bigg)^p \int_{0}^{\infty}|f^\prime(r)|^p\:{\rm d}r. \end{align*} \end{lemma} Now by substituting $\frac{1}{r}\int_{0}^{r}f(t)\:{\rm d}t$ and $\int_{0}^{r}f(t)\:{\rm d}t$ instead of $f(r)$ one can have two immediate corollaries, respectively. \begin{corollary} Let $1 < p < \infty$. Then, for any locally absolutely continuous function $f$ on $(0, \infty)$, there holds \begin{align*} &\int_{0}^{\infty}\max\biggl\{\sup_{0<s\leq r} \frac{|\frac{1}{s}\int_{0}^{s}f(t)\:{\rm d}t|^p}{r^p}, \sup_{r\leq s<\infty} \frac{|\frac{1}{s}\int_{0}^{s}f(t)\:{\rm d}t|^p}{s^p} \biggr\}\:{\rm d}r\nonumber\\&\leq \bigg(\frac{p}{p-1}\bigg)^p 2^{p-1} \biggl\{\int_{0}^{\infty} \frac{|f(r)|^p}{r^p}\:{\rm d}r + \int_{0}^{\infty}\frac{1}{r^{2p}}\bigg(\int_{0}^{r}|f(t)|^p\:{\rm d}t\bigg)\:{\rm d}r\biggr\}. \end{align*} \end{corollary} \begin{corollary} Let $1 < p < \infty$. Then, for any locally absolutely continuous function $f$ on $(0, \infty)$, there holds \begin{align*} \int_{0}^{\infty}\max\biggl\{\sup_{0<s\leq r} \frac{|\int_{0}^{s}f(t)\:{\rm d}t|^p}{r^p}, \sup_{r\leq s<\infty} \frac{|\int_{0}^{s}f(t)\:{\rm d}t|^p}{s^p} \biggr\}\:{\rm d}r\leq \bigg(\frac{p}{p-1}\bigg)^p \int_{0}^{\infty}|f(r)|^p\:{\rm d}r \end{align*} \end{corollary} \medskip \section*{Acknowledgments} This work was partially supported by the NU program 20122022CRP1601. The first author is supported in part by JSPS Kakenhi 18KK0073, 19H00644. The second author is supported in part by National Theoretical Science Research Center Operational Plan V-Mathematics Field (2/5) (Project number 111-2124-M-002-014-G58-01). \medskip
{ "timestamp": "2023-01-30T02:10:02", "yymm": "2301", "arxiv_id": "2301.11638", "language": "en", "url": "https://arxiv.org/abs/2301.11638", "abstract": "The motive of this note is twofold. Inspired by the recent development of a new kind of Hardy inequality, here we discuss the corresponding Hardy-Rellich and Rellich inequality versions in the integral form. The obtained sharp Hardy-Rellich type inequality improves the previously known result. Meanwhile, the established sharp Rellich type integral inequality seems new.", "subjects": "Functional Analysis (math.FA); Classical Analysis and ODEs (math.CA)", "title": "One-dimensional integral Rellich type inequalities", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9843363549643401, "lm_q2_score": 0.8128673269042767, "lm_q1q2_score": 0.8001348616345624 }
https://arxiv.org/abs/1602.03445
Davenport constant for commutative rings
The Davenport constant is one measure for how "large" a finite abelian group is. In particular, the Davenport constant of an abelian group is the smallest $k$ such that any sequence of length $k$ is reducible. This definition extends naturally to commutative semigroups, and has been studied in certain finite commutative rings. In this paper, we give an exact formula for the Davenport constant of a general commutative ring in terms of its unit group.
\section{Introduction} The Davenport constant is an important concept in additive number theory. In particular, it measures the largest zero-free sequence of an abelian group. The Davenport constant was introduced by Davenport in 1966 \cite{Davenport}, but was actually studied prior to that in 1963 by Rogers \cite{Rogers}. The definition was first extended by Geroldinger and Schneider to abelian semigroups by \cite{GHK2006} as follows: \begin{definition} For an additive abelian semigroup $S$, let $d(S)$ denote the smallest $d\in \mathbb{N}_0\cup \{\infty\}$ with the following property: For any $m\in \mathbb{N}$ and $s_1, \dots, s_m\in S$ there exists a subset $J\subseteq [1, m]$ such that $\abs{J}\le d$ and \[ \sum_{j = 1}^m s_j = \sum_{j\in J} s_j. \] \end{definition} In addition, \cite{GHK2006} showed the following: \begin{proposition} \label{infprop} If $\abs{S} < \infty$, then $d(S) < \infty$. \end{proposition} Wang and Guo \cite{WangGuo2008} then gave the definition of the large Davenport constant in terms of reducible and irreducible sequences, as follows: \begin{definition} Let $S$ be a commutative semigroup (not necessarily finite). Let $A$ be a sequence of elements in $S$. We say that $A$ is \textit{reducible} if there exists a proper subsequence $B\subsetneq A$ such that the sum of the elements in $A$ is equal to the sum of the elements in $B$. Otherwise, we say that $A$ is \textit{irreducible}. \end{definition} \begin{definition} Let $S$ be a finite commutative semigroup. Define the \textit{Davenport constant} $D(S)$ of $S$ as the smallest $d\in \mathbb{N} \cup \{\infty\}$ such that every sequence $S$ of $d$ elements in $S$ is reducible. \end{definition} \begin{remark} $D$ and $d$ are related by the equation $D(S) = d(S) + 1$. \end{remark} Note that if $S$ is an abelian group, being irreducible is equivalent to being zero-sum free, so the definition of the Davenport constant here is equivalent to the classical definition of the Davenport constant for abelian groups. In all following sections, unless otherwise noted: \begin{itemize} \item All semigroups are unital and commutative. Furthermore, we will use multiplication notation for semigroups, as opposed to the additive convention used in \cite{GHK2006,WangGuo2008,Wang2015124,QuWangZhang}. \item Similarly, rings are unital and commutative. \item Sets represented by the capital letter $S$ are semigroups. \item Sets represented by the capital letter $T$ are ideals in a semigroup. \item Sets represented by the capital letters $A, B$ are sequences of elements in a semigroup. In addition, $\pi(A)$ denotes the product of the elements in $A$. \item Sets represented by the capital letter $R$ are commutative rings. \item By abuse of notation, when we write $D(R)$, we actually mean $D(S_R)$, where $S_R$ is the semigroup of $R$ {\it under multiplication}. \item $C_n$ denotes the cyclic group of order $n$; $\mathbb{Z} / n\mathbb{Z}$ denotes the ring with additive group $C_n$. \end{itemize} \section{Previous Results} So far, this general setting has been studied extensively in the cases where the semigroup is the semigroup of a (commutative) ring under multiplication. The general idea in this class of problems is to show that the Davenport constant of the semigroup of a ring $R$ under multiplication is ``close'' to the Davenport constant of its group of units $U(R)$. The reason this has been done is that very little is known about the precise value of $D(G)$ when $G$ is an abelian group of high rank, and often the unit group of these commutative rings will be a group of high rank. However, as we will show, if there is a general theorem for the Davenport constant of an abelian group, then by \Cref{thm1}, we will also have a general theorem for the Davenport constant of an arbitrary finite commutative ring. \begin{remark} Since $U(R)$ is a sub-semigroup of $R$, we clearly have $D(R)\ge D(U(R))$, so we would like to say something about the difference $D(R) - D(U(R))$. \end{remark} In their paper, Wang and Guo showed the following result: \footnote{\cite{WangGuo2008} actually erroneously claimed that for general $n_i$, $D(R) = D(U(R)) + \#\{1\le i\le r: 2\| n_i\}$. The author has corresponded with the authors of \cite{WangGuo2008} on this matter, and they offer the result given above.} \begin{theorem}[Wang, Guo, 2008, \cite{WangGuo2008}] \label{wangguo} If $R = \mathbb{Z}/n_1\mathbb{Z} \times \dots \times \mathbb{Z}/n_r\mathbb{Z}$, where each of the $n_1, \dots, n_r$ are odd. Then $D(R) = D(U(R))$. \end{theorem} Later, Wang showed the following theorem: \begin{theorem}[Wang, 2015, \cite{Wang2015124}] \label{wang} Suppose $q > 2$ is a prime power. If $R \neq \mathbb{F}_q[x]$ is a quotient ring of $\mathbb{F}_q[x]$, then $D(R) = D(U(R))$. \end{theorem} However, Wang left the case $q = 2$ open and gave an instance for which $D(R)\neq D(U(R))$. Later, Zhang, Wang, and Qu gave the following bound when $q = 2$: \begin{theorem}[Zhang, Wang, Qu, 2015, \cite{QuWangZhang}] \label{qwz} If $f\in \mathbb{F}_2[x]$ is nonconstant and $R = \mathbb{F}_2[x] / (f)$, then $D(U(R))\le D(R) \le D(U(R)) + \delta_f$, where $\delta_f = \deg[\gcd(f, x(x + 1))]$. \end{theorem} \section{Summary of New Results} First, we will show that the converse of \Cref{infprop} holds for rings: \begin{theorem} \label{thm4} If $R$ is a commutative ring and $D(R) < \infty$, then $\abs{R} < \infty$. \end{theorem} However, the main result of this paper will be to relate $D(R)$ and $D(U(R))$ for arbitrary finite rings $R$. As we will see, the reason why $D(R) = D(U(R))$ does not hold in general is closely related to the presence of certain index 2 ideals in $R$, which we can see immediately in the following exact formula for $D(R)$ in terms of the Davenport constants of certain subgroups of $U(R)$: \begin{theorem} \label{thm1} Suppose $R$ is of the form \[ (\mathbb{Z} / 2\mathbb{Z})^{k_1} \times (\mathbb{Z} / 4\mathbb{Z})^{k_2} \times (\mathbb{Z} / 8\mathbb{Z})^{k_3} \times (\mathbb{F}_2[x] / (x^2))^{k_4 } \times R', \] where $R'$ is a product of local rings not isomorphic to $\mathbb{Z} / 2\mathbb{Z}, \mathbb{Z} / 4\mathbb{Z}, \mathbb{Z} / 8\mathbb{Z}, \mathbb{F}_2[x]/(x^2)$. Then \[ D(R) = \max_{0\le a\le k_2 + k_4, 0\le b\le k_3}\bigg[D(U(R') \times {C_2}^{k_2 + k_4 + 2k_3 - a - 2b}) + k_1 + 2a + 3b \bigg]. \] \end{theorem} \begin{remark} Note that since $R$ is finite, $R$ is Artinian, so there is a unique decomposition of $R$ as a product as a product of local rings (\cite{atiyahmacdonald}, \S 8). Thus the quantities $k_1, k_2, k_3, k_4$ are well-defined as functions of $R$. \end{remark} As a corollary, we get the following bound on $D(R) - D(U(R))$: \begin{corollary} \label{thm2} Suppose $R$ is of the form \[ (\mathbb{Z} / 2\mathbb{Z})^{k_1} \times (\mathbb{Z} / 4\mathbb{Z})^{k_2} \times (\mathbb{Z} / 8\mathbb{Z})^{k_3} \times (\mathbb{F}_2[x] / (x^2))^{k_4 } \times R', \] where $R'$ is a product of local rings not isomorphic to $\mathbb{Z} / 2\mathbb{Z}, \mathbb{Z} / 4\mathbb{Z}, \mathbb{Z} / 8\mathbb{Z}, \mathbb{F}_2[x]/(x^2)$. Then \[ D(U(R)) + k_1 \le D(R) \le D(U(R)) + k_1 + k_2 + k_3 + k_4. \] In addition, equality holds on the right if $U(R)$ is a power of 2 or if $k_i = 0$ for $i = 2, 3, 4$. \begin{proof} For the left hand side, note that we can pick $a = b = 0$ to get \[ D(R)\ge D(U(R') \times {C_2}^{k_2 + k_4 + 2k_3}) + k_1 = D(U(R)) + k_1 \] For the right hand side, we use the well-known facts $D(G\times H) \ge D(G) + D(H) - 1$, and $D(C_2) = 2$ to get \begin{align*} D(R) &= \max_{0\le a\le k_2, 0\le b\le k_3}\bigg[D(U(R') \times {C_2}^{k_2 + k_4 + 2k_3 - a - 2b}) + k_1 + 2a + 3b \bigg] \\ &= \max_{0\le a\le k_2, 0\le b\le k_3}\bigg[D(U(R') \times {C_2}^{k_2 + k_3 + 2k_3 - a - 2b}) + (a + 2b) (D(C_2) - 1) + k_1 + a + b \bigg] \\ &\le \max_{0\le a\le k_2, 0\le b\le k_3}\bigg[D(U(R') \times {C_2}^{k_2 + k_4 + 2k_3}) + k_1 + a + b \bigg] \\ &= \max_{0\le a\le k_2, 0\le b\le k_3}\bigg[D(U(R)) + k_1 + a + b \bigg] \\ &= D(U(R)) + k_1 + k_2 + k_3 + k_4. \end{align*} When $k_2 + k_3 + k_4 = 0$, equality clearly holds as the left hand side and the right hand side are the same. On the other hand, when $\abs{U(R)}$ is a power of 2, then $U(R')$ is also a 2-group. However, by \cite{Olson1969}, if $G$ and $H$ are 2-groups, then $D(G\times H) = D(G) + D(H) - 1$. Thus \begin{align*} D(R) &\ge D(U(R')) + k_1 + 2k_2 + 3k_3 + 2k_4 \\ &= D(U(R')) + (k_2 + k_4 + 2k_3) (D(C_2) - 1) + k_1 + k_2 + k_3 + k_4 \\ &= D(U(R')\times {C_2}^{k_2 + k_4 + 2k_3}) + k_1 + k_2 + k_3 + k_4 \\ &= D(R) + k_1 + k_2 + k_3 + k_4 \ge D(R), \end{align*} so equality holds in this case as well. \end{proof} \end{corollary} We also have the following more concise bound on $D(R) - D(U(R))$ that does not depend on writing down a local ring product decomposition. \begin{corollary} \label{thm3} Suppose $R$ is a finite commutative ring, and let $n_2(R)$ be the number of index two (prime, maximal) ideals of $R$. Then $D(R)\le D(U(R)) + n_2(R)$. \begin{proof} Since $R$ is finite (and thus Artinian), $R$ can be expressed as a finite product of local rings. Thus in the setting of \Cref{thm2}, it suffices to show that $k_1 + k_2 + k_3 + k_4\le n_2(R)$. Since $R$ is Artinian, we have that $R\simeq \prod_{i = 1}^n R / \mathfrak{q}_i$, where $\{\sqrt{\mathfrak{q}_i}\}_{i = 1}^n = \Spec(R)$. In particular, the number of $R / \mathfrak{q}_i$ that can be isomorphic to $\mathbb{Z} / 2\mathbb{Z}, \mathbb{Z}/4\mathbb{Z}, \mathbb{Z}/8\mathbb{Z}$ or $\mathbb{F}_2[x] / (x^2)$ is at most the number of index 2 ideals in $\Spec(R)$, which is precisely $n_2(R)$. \end{proof} \end{corollary} Using \Cref{thm1} and \Cref{thm2,thm3}, we will give generalizations of each of the results in the previous section. The main line of attack to prove \Cref{thm1} will be to reduce the problem to the case of finite local rings and then talk about what happens when we ``glue'' local rings together via product. However, we will first discuss the gluing mechanism for a more general class of semigroups (``almost unit-stabilized''); this will require the notion of \textit{unit-stabilized pairs} and a \textit{relative Davenport constant}. Afterwards, we will show that finite local rings are almost unit-stabilized, and that more structure holds for all finite local rings other than $\mathbb{Z} / 2\mathbb{Z}, \mathbb{Z} / 4 \mathbb{Z}, \mathbb{Z} / 8\mathbb{Z}$, and $\mathbb{F}_2[x] / (x^2)$. \section{Davenport constant for semigroups} Let $S$ be a semigroup and $U(S)$ denote its group of units. \begin{definition} We say that a sequence $A$ with elements in $S$ is \textit{reducible} if there exists a proper subsequence $A'\subsetneq A$ such that $\pi(A') = \pi(A)$. Otherwise, $A$ is \textit{irreducible}. \end{definition} \begin{definition} The Davenport constant $D(S)$ is the minimum positive integer $k$ such that any sequence $A$ of size $k$ is reducible. \end{definition} \begin{definition} For $S$ a semigroup, let $\Delta(S) = D(S) - D(U(S))$. \end{definition} \begin{remark} Clearly $\Delta(S) \ge 0$. \end{remark} \begin{lemma} \label{prodlemma} Suppose $S$ and $S'$ are semigroups. Then \[ D(S\times S')\ge D(S) + D(S') - 1. \] \begin{proof} Let $\ell = D(S) - 1$ and $\ell' = D(S') - 1$. Then there exists an irreducible sequence $s_1, \dots, s_\ell$ in $S$ of length $\ell$ and an irreducible sequence $s_1',\dots s_{\ell'}'$ in $S'$ of length $\ell'$. Then the sequence \[ (s_1, 1_{S'}), \dots, (s_\ell, 1_{S'}), (1_S, s_1'), \dots, (1_S, s_{\ell'}') \] is an irreducible sequence in $S\times S'$ of length $\ell + \ell' = D(S) + D(S') - 2$, which means that $D(S\times S') \ge D(S) + D(S') - 1$. \end{proof} \end{lemma} \section{Proof of \texorpdfstring{\Cref{thm4}}{Theorem \ref{thm4}}} We will first handle the case where $R$ is infinite. In particular, we will show that if $\abs{R} = \infty$, then $D(R) = \infty$. \begin{remark} The corresponding theorem for semigroups clearly does not hold: let $S$ be the semigroup with underlying set $\mathbb{Z}^{\ge 0}$ and operation $a * b = \max(a, b)$. Then $S$ is infinite but $D(S) = 2$. \end{remark} \begin{lemma} \label{inflemma} Suppose $G$ is an infinite abelian group. Then $D(G) = \infty$. \begin{proof} Let $g_1, \dots, g_n$ be an irreducible sequence. Then $g_1, \dots, g_n, g$ is not irreducible if and only if $g^{-1}$ is not of the form $g_{i_1} \cdots g_{i_k}$. However, there are at most $2^n$ such $g$, which means that we can always extend an irreducible sequence to get a longer irreducible sequence, which means that $D(G) = \infty$. \end{proof} \end{lemma} \begin{lemma} Suppose $R$ is a ring and $I\subseteq R$ is an ideal. Then $D(R)\ge D(R / I)$. \begin{proof} This follows from the fact that any irreducible sequence in $R / I$ can be lifted to an irreducible sequence in $R$. \end{proof} \end{lemma} We are now ready to give the proof of \Cref{thm4}: \begin{theorem*} If $R$ is a commutative ring and $D(R) < \infty$, then $\abs{R} < \infty$. \begin{proof} Suppose $D(R) = k < \infty$. \begin{itemize} \item $\abs{R / \mathfrak{m}}$ is finite for every maximal ideal $\mathfrak{m} \subseteq R$. Otherwise, $U(R / \mathfrak{m})$ is an infinite abelian group, so $D(R / \mathfrak{m}) = \infty$, which means that $D(R) = \infty$ as well. \item $R$ has at most $k - 1$ maximal ideals. Suppose otherwise, that $R$ has $k$ maximal ideals $\mathfrak{m}_1, \dots, \mathfrak{m}_k$. Then by the Chinese Remainder theorem, there exists $a_1, \dots, a_k\in R$ such that $a_i \in \mathfrak{m}_i$ and $a_i - 1 \in \mathfrak{m}_j$ for $i\neq j$, and $a_1, \dots, a_k$ would be an irreducible sequence of length $k$, contradiction. \item Let $J(R)$ denote the Jacobson radical of $R$, the intersection of all of the maximal ideals of $R$. Since $R$ has a finite number of maximal ideals, by the Chinese Remainder Theorem, $R / J(R)\simeq \prod_{i} R / \mathfrak{m}_i$, which means that $R / J(R)$ is finite. \item On the other hand, $\abs{U(R)} < \infty$, for if $\abs{U(R)} = \infty$, then $D(U(R)) = \infty$, so $D(R) = \infty$ as well. \item Finally, $1 + J(R)\subseteq U(R)$, which means that $\abs{J(R)} < \infty$ as well. Thus $\abs{R} = \abs{R / J(R)} \cdot \abs{J(R)} < \infty$. \end{itemize} \end{proof} \end{theorem*} \section{Relative Davenport constant} In order to prove \Cref{thm1} (and its subsequent corollaries), we need to introduce the notion of the \textit{relative Davenport constant}, which measures the longest irreducible sequence in $S$ with product lying in an ideal $T\subseteq S$. \begin{definition} A set $T\subseteq S$ is an ideal if $S\cdot T\subseteq T$. ($T = \emptyset$ is allowed here.) \end{definition} Suppose $S_1, S_2$ are semigroups and $T_1\subseteq S_1, T_2\subseteq S_2$ are ideals. Then $T_1\times T_2\subseteq S_1\times S_2$ is also an ideal. \begin{definition} Given a semigroup $S$ and an ideal $T\subseteq S$, $A$ is a \textit{$T$-sequence} if $\pi(A)\in T$. \end{definition} \begin{definition} Given a semigroup $S$ and a (possibly empty) ideal $T\subseteq S$, define the \textit{relative Davenport constant} $D(S, T)$ to be the minimum positive integer $k$ such that any $T$-sequence $A$ of size $k$ is reducible. \end{definition} We can again define $d(S, T)$ as the maximum length of an irreducible $T$-sequence, and we once again have the relation $D(S, T) = d(S, T) + 1$. \begin{remark} If $T = \emptyset$ is the empty ideal, then $d(S, T)$ is tautologically 0 and so $D(S, T)$ is 1. \end{remark} Before we proceed, we first need the following helpful lemma. \begin{lemma} \label{rellemma} Let $S$ be a semigroup and $T$ be an ideal. Suppose $s\in S$ and $A$ is a sequence in $S$ of length $D(S, T)$ such that $s\cdot \pi(A) \in T$. Then there exists a proper subsequence $A'\subsetneq A$ such that $s\pi(A') = s\cdot \pi(A)$. \begin{proof} If $\pi(A)\in T$, then we are done by the definition of $D(S, T)$. Otherwise, consider the sequence $B = A\cup \{s\}$. Since $\abs{B} = D(S, T) + 1 > D(S, T)$, there exists a proper subsequence $B'\subsetneq B$ such that $\pi(B') = \pi(B) = s\cdot \pi(A)$. Note that if $s\notin B'$, then $B'\subseteq A$, which means that $\pi(B')\notin T$ (as $\pi(A)\notin T$ by definition), whereas $\pi(B) \in T$, contradiction. Thus $a\in B'$, so if we let $A' = B' \backslash \{s\}$, then $s\cdot \pi(A') = \pi(B') = \pi(B) = s\cdot \pi(A)$, as desired. \end{proof} \end{lemma} \begin{lemma} \label{grplem} Suppose $G$ is an abelian group and $H$ is a subgroup of $G$, and suppose $S$ is a semigroup. Then \[ D(G \times S, G \times T) \ge D((G / H) \times S, (G / H) \times T) + D(H) - 1. \] \begin{proof} Let $\ell_1 = D(G / H\times S, (G / H)\times T) - 1, \ell_2 = D(H) - 1$. Then there exists an irreducible ($(G/ H)\times T$)-sequence $(g_1, s_1), \dots (g_{\ell_1}, s_{\ell_1})$ in $(G / H) \times S$ of length $\ell_1$, and an irreducible sequence $h_1, \ldots, h_{\ell_2}$ in $H$ of length $\ell_2$. In this case, if we let $g_i'$ be any lift of $g_i$ from $G / H$ to $G$, then the sequence \[ (g_1', s_1), \ldots, (g_{\ell_1}', s_{\ell_1}), (h_1, 1_S), \ldots, (h_{\ell_2}, 1_S) \] is an irreducible $G\times T$-sequence in $G\times S$ of length \[ \ell_1 + \ell_2 = D((G / H)\times S, (G / H)\times T) + D(H) - 2. \] \end{proof} \end{lemma} \section{Unit-Stabilized Pairs} \begin{definition} An ordered pair $(S,T)$ of a semigroup $S$ and an ideal $T\subseteq S$ is a \textit{unit-stabilized pair} if for all $a, b\in S$ such that $ab\notin T$ and $\Stab_{U(S)}(a) = \Stab_{U(S)}(ab)$, then $ab = au$ for some $u \in U(S)$. \end{definition} \begin{definition} We say that $S$ is a \textit{unit-stabilized semigroup} if $(S, \emptyset)$ is a unit-stabilized pair. \end{definition} \begin{example} If $R$ is a finite field other than $\mathbb{F}_2$, then [the semigroup of] $R$ [under multiplication] is unit-stabilized. $(\mathbb{F}_2, 0)$ is a unit-stabilized pair. \end{example} It turns out most local rings $R$ are unit-stabilized. Unit-stabilized rings are nice because they satisfy $\Delta(R) = 0$ and are closed under the operation of taking products. However, if $R$ has residue field $\mathbb{F}_2$, then it is in fact not unit-stabilized: $U(R) = 1 + \mathfrak{m}$, where $\mathfrak{m}$ is the maximal ideal of $R$. Thus if $x\in R$ satisfies $\Ann_R(x) = \mathfrak{m}$, then $\Stab_{U(R)}(x) = \Stab_{U(R)}(0) = U(R)$, but clearly $0\neq ux$ for any unit $u$. As we can see, the unit-stabilized behavior breaks down around 0. However, even when $R$ is not unit-stabilized, it turns out that $(R, 0)$ will be a unit-stabilized pair. We have the following formula that helps for working with unit-stabilized pairs: \begin{theorem} \label{relthm} Suppose $(S, T)$ is a unit-stabilized pair and $S'$ is a semigroup and $T'$ is an ideal in $S'$. Then \begin{equation*} D(S\times S', S\times T') = \max[D(U(S)\times S', U(S)\times T'), D(S\times S', T \times T')]. \end{equation*} \begin{proof} First off, it is clear that $D(S\times S', S \times T')\ge D(U(S)\times S', U(S)\times T')$ and $D(S\times S', S\times T')\ge D(S\times S', T\times T')$, so it suffices to show the inequality in the other direction. Let \[ \ell = \max[D(U(S)\times S', U(S)\times T'), D(S\times S', T \times T')]. \] Suppose we have a sequence $A = \{a_1,\ldots,a_\ell\}$ with $\pi(A)\in S\times T'$, and suppose for the sake of contradiction that $A$ is irreducible. Let $P_{S}$ and $P_{S'}$ denote the projection maps from $S\times S'$ to $S$ and $S'$, respectively. If $P_S(\pi(A))\in T$, then $\pi(A)\in T\times T'$, and by the definition of $D(S\times S', T \times T')$ we automatically have that $A$ is not irreducible. Thus $P_S(\pi(A))\notin T$. Let $B$ be a minimal subsequence of $A$ such that \[ \Stab_{U(S)}(P_S(\pi(B))) = \Stab_{U(S)}(P_S(\pi(A))), \] and without loss of generality suppose that $B = \{a_1, \ldots, a_k\}$. Let $p_0 = (1_{S}, 1_S)$, and for $1\le i\le \ell$, $p_i = a_ip_{i - 1}$ and let $K_i = \Stab_{U(S)}(P_S(p_i))$. Then we have \[ 0 = K_0\subseteq K_1\subseteq \dots \subseteq K_\ell. \] In addition, for $1\le i\le k$, $K_{i-1}\neq K_i$, so by \Cref{grplem}, \begin{align*} D(K_i)\ge D(K_{i - 1}) + D(K_i / K_{i - 1}) - 1 \ge D(K_{i - 1}) + 1, \end{align*} which means that $D(K_k)\ge k + 1$. Applying \Cref{grplem} again, we have, \begin{align*} D((U(S) / K_k) \times S', (U(S) / K_k)\times T') \le D(U(S) \times S', U(S) \times T') - D(K_k) + 1 \le \ell - k. \end{align*} In addition, for $k < i \le \ell$, \[ \Stab(P_S(p_k)) \subseteq \Stab(P_S(a_ip_k)) \subseteq \Stab(P_S(p_\ell)) = \Stab(P_S(p_k)), \] so the inclusions are equalities. Since $p_\ell = P_S(\pi(A))\notin T$ and $(S, T)$ is unit-stabilized, for each $k < i \le \ell$ there exists a $u_i\in U(S)$ such that $u_iP_S(p_k) = a_iP_S(p_k)$ for $k < i \le \ell$. Let $t_i = (u_i, P_{S'}(a_i))$. We have $p_kt_i = p_ka_i$ for $k\le i\le \ell$. Since $t_{k+1}, \ldots, t_{\ell}$ is a sequence of length \[ \ell - k\ge D((U(S) / K_k) \times S', (U(S) / K_k)\times T') \] and $(1, \pi_{S'}(p_k))\cdot t_{k+1}\cdots t_{\ell} \in U(S)\times T'$, by \Cref{rellemma}, there exists a (possibly empty) proper subsequence $i_1, \ldots, i_m$ such that $ (1, \pi_{S'}(p_k))\cdot t_{i_1}\cdots t_{i_m}\cdot (u, 1_{S'}) = (1, \pi_{S'}(p_k))\cdot t_{k+1}\cdots t_{\ell}$ for some $u\in K_k$. Then since $u\in \Stab(P_S(p_k))$, $(u, 1) \cdot p_k = p_k$, so \begin{align*} a_1a_2\cdots a_k \cdot a_{i_1}\cdots a_{i_m} &= p_k \cdot a_{i_1}\cdots a_{i_m} \\ &= p_k \cdot t_{i_1}\cdots t_{i_m}\\ &= (1, P_{S'}(p_k))\cdot (P_S(p_k), 1) \cdot t_{i_1}\cdots t_{i_m} \\ &= (1, P_{S'}(p_k))\cdot (P_S(p_k), 1) \cdot (u, 1) \cdot t_{i_1}\cdots t_{i_m} \\ &= (1, P_{S'}(p_k))\cdot (P_S(p_k), 1) \cdot t_{k + 1}\cdots t_{\ell} \\ &= (1, P_{S'}(p_k))\cdot (P_S(p_k), 1) \cdot a_{k + 1}\cdots a_{\ell} \\ &= p_k \cdot a_{k+1}\cdot a_\ell \\ &= a_1\cdots a_\ell = \pi(A). \end{align*} This contradicts the assumption that $A$ was irreducible. Thus $D(S\times S', T \times T')\le \max[D(U(S)\times S', U(S)\times T'), D(S\times S', T \times T')]$, as desired. \end{proof} \end{theorem} \begin{corollary} Suppose $(S, T)$ is a unit-stabilized pair. Then $D(S) = \max[D(U(S)), D(S, T)]$. \begin{proof} Apply \Cref{relthm} to the case where $S'$ is the trivial group and $T' = S'$. \end{proof} \end{corollary} \begin{corollary} If $S$ is a unit-stabilized semigroup, then $D(S) = D(U(S))$. \end{corollary} \section{Almost Unit-Stabilized semigroups} \begin{lemma} \label{zerolemma} If $S$ has a 0 element, then for all semigroups $S'$, \[ D(S \times S', \{0\}\times T') = D(S', T') + D(S, \{0\}) - 1. \] \begin{proof} First we will show that $D(S\times S', \{0\}\times T)\le D(S', T') + D(S, \{0\}) - 1$. Let $\ell = D(S', T') + D(S, \{0\}) - 1$, and suppose we have an irreducible $\{0\}\times T$-sequence $(s_1, s_1'), \ldots, (s_\ell, s_\ell')$. Consider the smallest subset such that the second coordinate of the product is equal to 0. By the definition of $D(S, \{0\})$, this subset has size at most $k = D(S, \{0\}) - 1$. Without loss of generality suppose $s_1\cdots s_k = 0$, and let $p = s_1'\cdots s_k'$. Then the sequence $s_{k+1}',\dots, s_\ell'$ has length $D(S', T')$, so by \Cref{rellemma} there is some proper subsequence such that $ps_{i_1}'\cdots s_{i_r}' = ps_{k + 1}'\cdots s_{\ell}'$. Since the first coordinate of both sides must be zero, we have found a proper subsequence with the same product, contradiction. Now to show the reverse inequality, it suffices to construct an irreducible ($\{0\}\times T'$)-sequence of length $D(S', T') + D(S, \{0\}) - 2$. Let $\ell_1 = D(S', T') - 1$ and $\ell_2 = D(S, \{0\}) - 1$. By definition, there exists an irreducible $T'$-sequence $s_1', s_2', \ldots, s_{\ell_1}'\in S'$ and an irreducible $\{0\}$-sequence $s_1, s_2, \ldots, s_{\ell_2}$. Then $(1_{S}, s_1'), \dots, (1_{S}, s_{\ell_1}'), (s_{1}, 1_{S'}), \dots, (s_{\ell_2}, 1_{S'})$ is an irreducible $(\{0\}\times T')$-sequence of length $\ell_1 + \ell_2 = D(S', T') + D(S, \{0\}) - 2$, as desired. \end{proof} \end{lemma} \begin{definition} A semigroup $S$ is \textit{almost unit-stabilized} if $S$ has a 0 element and $(S, \{0\})$ is a unit-stabilized pair. \end{definition} \begin{remark} If $S$ is unit-stabilized and contains a 0 element, the $S$ is also almost unit-stabilized. \end{remark} \begin{lemma} \label{auslem} Suppose $S$ is almost unit-stabilized, and suppose $S'$ is any other semigroup. Then \begin{equation*} D(S\times S') = \max(D(U(S)\times S'), D(S, \{0\}) + D(S') - 1). \end{equation*} \begin{proof} This follows directly from \Cref{relthm} and \Cref{zerolemma}. \end{proof} \end{lemma} \begin{corollary} \label{auscor} Suppose $S$ is almost unit-stabilized and $\Delta(S) = 0$. Then for any semigroup $S'$, $D(S\times S') = D(U(S)\times S')$. \begin{proof} If $\Delta(S) = 0$, then $D(S, \{0\})\le D(U(S))$. Thus by \Cref{prodlemma}, \begin{align*} D(S, \{0\}) + D(S') - 1 \le D(U(S)) + D(S') - 1\le D(U(S)\times S'), \end{align*} which means that by \Cref{auslem}, \[ D(S\times S') = \max(D(U(S) \times S'), D(S, \{0\}) + D(S') - 1) = D(U(S) \times S'). \] \end{proof} \end{corollary} \begin{theorem} \label{austhm} Suppose $S_1, \dots, S_k$ are almost unit-stabilized semigroups. Suppose further that for $i > \ell$, $\Delta(S_i) = 0$. Then \begin{equation*} D(S_1\times\dots\times S_k) = \max_{I\subseteq [1,\ell]}\left[D\left(\prod_{i\in [1,k]\backslash I} U(S_i)\right) + \sum_{i\in I} (D(S_i, \{0\}) - 1)\right]. \end{equation*} \begin{proof} First, by $\ell$ applications of \Cref{auslem}, we have that for any semigroup $S'$, \begin{equation*} D(S_1\times\dots\times S_\ell\times S') = \max_{I\subseteq [1,\ell]}\left[D\left(\left(\prod_{i\in [1,\ell]\backslash I} U(S_i)\right)\times S'\right) + \sum_{i\in I} (D(S_i, \{0\}) - 1)\right]. \end{equation*} Now if we let $S' = S_{\ell + 1} \times \dots \times S_k$ and apply \Cref{auscor} $k - \ell$ times, we have \begin{equation*} D(S_1\times\dots\times S_k) = \max_{I\subseteq [1,\ell]}\left[D\left(\prod_{i\in [1,k]\backslash I} U(S_i)\right) + \sum_{i\in I} (D(S_i, \{0\}) - 1)\right]. \end{equation*} \end{proof} \end{theorem} \begin{corollary} If $T = S_1 \times\dots\times S_k$ is a product of almost unit-stabilized semigroups, then \begin{equation*} \Delta(T) \le \Delta(S_1)+\dots+\Delta(S_k). \end{equation*} \begin{proof} Applying \Cref{austhm} to the (trivial) case where $\ell = k$, we have \begin{equation*} D(T) = \max_{I\subseteq [1,k]}\left[D\left(\prod_{i\in [1,k]\backslash I} U(S_i)\right) + \sum_{i\in I} (D(S_i, \{0\}) - 1)\right]. \end{equation*} However, using the inequality $D(G)\ge D(G / H) + D(H) - 1$ (a special case of \Cref{grplem} where $S$ is the trivial group), we have that for any $I\subseteq [1,k]$, \begin{align*} D\left(\prod_{i\in [1,k]\backslash I} U(S_i)\right) + \sum_{i\in I} (D(U(S_i)) - 1) \le D\left(\prod_{i\in [1,k]} U(S_i)\right) = D(U(T)). \end{align*} Thus \begin{align*} D(T) &= \max_{I\subseteq [1,k]}\left[D\left(\prod_{i\in [1,k]\backslash I} U(S_i)\right) + \sum_{i\in I} (D(S_i, \{0\}) - 1)\right] \\ &\le \max_{I\subseteq [1,k]}\left[D(U(T)) + \sum_{i\in I} (D(S_i, \{0\}) - D(U(S_i))\right] \\ &= D(U(T)) + \sum_{i = 1}^k \max(0, D(S_i, \{0\}) - D(U(S_i)))\\ &= D(U(T)) + \sum_{i = 1}^k \Delta(S_i), \end{align*} from which the desired result immediately follows. \end{proof} \end{corollary} \begin{remark} It would be nice if it were true in general that $\Delta(S_1\times S_2) \le \Delta(S_1) + \Delta(S_2)$. However, this is not true, even when one of $S_1$ or $S_2$ is almost unit-stabilized! For an example, let $S_1 = \mathbb{Z} / 3\mathbb{Z}$ and let $S_2 = S\times S'$, where $S' = \mathbb{Z}/4\mathbb{Z}$ and $S = \{0, 1, 2, 4\}\subseteq \mathbb{Z}/7\mathbb{Z}$. We have that the following: \begin{itemize} \item $U(S_1)\simeq C_2, U(S)\simeq C_3, U(S')\simeq C_2$. \item $S_1, S, S'$ are unit-stabilized. \item $D(S_1) = D(U(S_1)) = 2$. \item $D(S) = D(U(S)) = 3$. \item $D(S') = D(S', \{0\}) = 3$. $D(U(S')) = 2$. \item $D(S_2) = D(S\times S') = D(U(S)\times S') = \max(D(U(S)\times U(S')), D(U(S)) + D(S', \{0\}) - 1) = \max(6, 3 + 3 - 1) = 6$, $D(U(S_2)) = D(C_6) = 6$. \end{itemize} However, \begin{align*} D(S_1\times S_2) = D(S_1\times S \times S') &= D(U(S_1)\times U(S)\times S') \\ &= \max(D(C_6\times U(S')), D(C_6) + D(S', \{0\} - 1) \\ &= \max(7, 6 + 3 - 1) = 8. \end{align*} On the other hand, \[ D(U(S_1\times S_2)) = D(U(S_1)\times U(S)\times U(S')) = D(C_6\times C_2) = 7. \] Thus $\Delta(S_1) = \Delta(S_2) = 0$ but $\Delta(S_1\times S_2) = 1$. In fact, it is possible to construct $S_1, S_2$ such that $\Delta(S_1) = \Delta(S_2) = 0$ and one of $S_1, S_2$ is almost unit-stabilized, but $\Delta(S_1\times S_2)$ is arbitrarily large. \end{remark} \section{Local Rings} In this section, we look at the case where $R$ is a (finite) local ring with maximal ideal $\mathfrak{m}$ and residue field $k$. It turns out local rings are all either unit-stabilized or almost unit-stabilized. \begin{lemma} For all $a\in R$ there exists a positive integer $n$ such that $a\mathfrak{m}^n = 0$. \begin{proof} Since $R$ is finite, the chain of ideals \[ aR\supseteq a\mathfrak{m}\supseteq a\mathfrak{m}^2\supseteq \cdots \] must stabilize, so $a\mathfrak{m}^n = a\mathfrak{m}^{n+1}$ for some positive integer $n$. Then by Nakayama's lemma we must have $a\mathfrak{m}^n = 0$. \end{proof} \end{lemma} \begin{lemma} If $a\neq 0$ and $\Ann_R(a) = \Ann_R(ab)$, then $b\in U(R)$. \begin{proof} Suppose instead that $b\notin U(R)$, so $b\in \mathfrak{m}$. Consider the minimum $n$ such that $a\mathfrak{m}^n = 0$. (Clearly $n > 0$.) Then $ab\mathfrak{m}^{n-1} \subseteq a\mathfrak{m}^n = 0$ but $a\mathfrak{m}^{n - 1}\neq 0$, which means that $\mathfrak{m}^{n-1}\not\subseteq \Ann_R(a)$ but $\mathfrak{m}^{n-1}\subseteq \Ann_R(ab)$, contradiction. \end{proof} \end{lemma} \begin{theorem} \label{goodrings} We have: \begin{enumerate} \item If $k \not \simeq \mathbb{F}_2$, then $R$ is unit stabilized. \item If $k \simeq \mathbb{F}_2$, then $R$ is almost unit-stabilized. \end{enumerate} \begin{proof} Note that for all $a\in R$, $\Stab_{U(R)}(a) = (1 + \Ann_R(a)) \cap U(R)$. In addition, if $a\neq 0$, then $\Ann_R{U(R)}$ is a proper ideal of $R$, which means that $\Ann_R{U(R)}\subseteq \mathfrak{m}$, so $1 + \Ann_R{U(R)}\subseteq U(R)$, which means that $\Stab_{U(R)}(a) = 1 + \Ann_R(a)$. Thus if $ab\neq 0$ and $\Stab_{U(R)}(a) = \Stab_{U(R)}(ab)$, then $\Ann_R(a) = \Ann_R(ab)$, which by the previous lemma implies $b\in U(R)$. In the case where $R / \mathfrak{m}\not\simeq \mathbb{F}_2$, then we have $1 + \mathfrak{m}\subsetneq U(R)$, so for all $a \neq 0$, $\Stab_{U(R)}(a)\subseteq 1 + \mathfrak{m} \subsetneq U(R) = \Stab_{U(R)}(0)$, which means that $R$ is unit-stabilized. \end{proof} \end{theorem} \begin{corollary} If $k\not \simeq \mathbb{F}_2$, then $\Delta(R) = 0$. \end{corollary} \begin{theorem} \label{badrings} Suppose $R$ is a finite local ring. Then $\Delta(R) \le 1$, and equality holds if and only if $R$ is isomorphic to one of the following rings: \begin{itemize} \item $\mathbb{Z} / 2\mathbb{Z}$ \item $\mathbb{Z} / 4\mathbb{Z}$ \item $\mathbb{Z} / 8\mathbb{Z}$ \item $\mathbb{F}_2[x] / (x^2)$ \end{itemize} \begin{proof} We have already taken care of the case where $k\not\simeq\mathbb{F}_2$. Suppose $R$ is a local ring with residue field $\mathbb{F}_2$. Then $\abs{R} = 2^n$ for a positive integer $n$. We will show the following: \begin{enumerate} \item $D(R, 0)\le n + 1$. \item If $D(R, 0) = n + 1$, then $\mathfrak{m}^{n - 1}\neq 0$. \item $D(U(R))\ge n$. \item If $D(U(R)) = n$, then $U(R)\simeq C_2^{n-1}$. \item If $n \ge 3$ and $\mathfrak{m}^{n - 1}\neq 0$ and $U(R)\simeq C_2^{n - 1}$, then $\mathfrak{m} = (2)$ and $R\simeq \mathbb{Z} / 2^n\mathbb{Z}$. \item If $n\ge 4$, then $D(U(\mathbb{Z} / 2^n\mathbb{Z})) \not \simeq C_2^{n - 1}$. \item If $n\le 2$, then $R\simeq \mathbb{Z} / 2\mathbb{Z}, \mathbb{Z} / 4\mathbb{Z}$, or $\mathbb{F}_2[x] / (x^2)$. \end{enumerate} To finish the proof, note that since $R$ is almost unit-stabilized, $D(R) = \max(D(U(R)), D(R, 0)$, so from (1) and (2) we immediately get $\Delta(R)\le 1$. Now suppose $\Delta(R) = 1$. Since $D(R, 0) \le n + 1$, $D(U(R))\ge n$, and $\Delta(R) = \max(0, D(R, 0) - D(U(R))$, equality must hold in both cases. Thus by (2), $\mathfrak{m}^{n - 1}\neq 0$; by (4), $U(R)\simeq C_2^{n - 1}$; by (5), this means that either $\abs{R}\le 4$ or $R\simeq \mathbb{Z} / 2^n\mathbb{Z}$. In the first case, by (7), $R$ must be either $\mathbb{Z} / 2\mathbb{Z}, \mathbb{Z} / 4\mathbb{Z}$, or $\mathbb{F}_2[x] / (x^2)$. Otherwise, by (6), we must have $n = 3$, in which case $R\simeq \mathbb{Z} / 8\mathbb{Z}$. Finally, we have the following values of $D(R)$ and $D(U(R))$ for $R = \mathbb{Z} / 2\mathbb{Z}, \mathbb{Z} / 4\mathbb{Z}, \mathbb{Z} / 8\mathbb{Z}, \mathbb{F}_2[x] / (x^2)$ \begin{center} \begin{tabular}{c|c|c} $R$ & $D(R)$ & $D(U(R))$ \\ \hline $\mathbb{Z} / 2\mathbb{Z}$ & 2 & 1 \\ $\mathbb{Z} / 4\mathbb{Z}$ & 3 & 2 \\ $\mathbb{Z} / 8\mathbb{Z}$ & 4 & 3 \\ $\mathbb{F}_2[x] / (x^2) $ & 3 & 2 \\ \end{tabular} \end{center} In each of these cases, we have $\Delta(R) = 1$. Here are proofs of the 7 claims: \begin{enumerate} \item Suppose $a_1, \dots, a_{n+1}$ is an irreducible sequence in $R$ with $a_1 \cdots a_{n + 1} = 0$. Then none of $a_1, \dots, a_{n+1}$ are in $U(R)$, or else they could be omitted. In addition, $a_1\cdots a_n\neq 0$. Thus if $p_i = a_1\cdots a_i$, we have $\Stab_{U(R)}(p_i)\subsetneq \Stab_{U(R)}(p_{i+1})$ for $0\le i< n$. In particular, this means that $\abs{\Stab_{U(R)}(p_i)}\ge 2^i$. However, when $i = n$ this gives $2^{i - 1} = \abs{U(R)}\ge \abs{\Stab_{U(R)}(p_i)} \ge 2^i$, contradiction. \item Suppose $D(R, 0) = n + 1$, and let $a_1, \dots, a_n$ be an irreducible sequence in $R$ with $a_1\cdots a_n = 0$. Then none of the $a_i$ are units, so $a_i\in \mathfrak{m}$. In addition, $a_1\cdots a_{n-1} \neq 0$, which means that $\mathfrak{m}^{n - 1}\neq 0$. \item We have that $\abs{U(R)} = 2^{n - 1}$. Thus we can write \[ U(R) = C_{2^{e_1}}\times \cdots \times C_{2^{e_r}}, \] where $e_1 + \dots + e_r = n - 1$. Since $R$ is a 2-group, by \cite{Olson1969}, \[ D(U(R)) = 1 + \sum_{i = 1}^r (D(C_{2^{e_i}}) - 1) = 1 + \sum_{i = 1}^r (2^{e_i} - 1) \ge 1 + \sum_{i = 1}^r e_i = n. \] Here we also used the inequality $2^n - 1\ge n$ for all $n\in \mathbb{Z}$. \item In the above equation, equality holds if and only if each of the $e_i$ is equal to 1, in which case $U(R)\simeq C_2^{n - 1}$. \item Since $\mathfrak{m}^{n- 1} \neq 0$, by Nakayama's lemma we have that $\mathfrak{m}^i\subsetneq \mathfrak{m}^{i - 1}$ for $1\le i\le n$, in which case $2\abs{\mathfrak{m}^i}\le \abs{\mathfrak{m}^{i - 1}}$. Thus we have \[ 2^{n} = \abs{\mathfrak{m}^0} \ge 2\abs{\mathfrak{m}^1} \ge \dots \ge 2^{n - 1} \abs{\mathfrak{m}^{n - 1}} \ge 2^{n -1 } \cdot 2 = 2^n, \] so equality holds in each of the above inequalities. In particular, we have $\abs{\mathfrak{m} / \mathfrak{m^2}} = 2$, so \[ \dim_k(\mathfrak{m} / \mathfrak{m^2}) = 1, \] which means that $\mathfrak{m}$ is principal. Suppose $\mathfrak{m} = (t)$. Then $t - 1$ is a unit, and since $U(R)\simeq C_2^{n - 1}$, we must have $(t - 1)^2 = 1$, or $t^2 = 2t$. I claim that $2\in \mathfrak{m} \backslash \mathfrak{m}^2$. Suppose otherwise, that $2\in \mathfrak{m}^2$. Then we have $t^2 = 2t\in \mathfrak{m}^3$, which means that $\mathfrak{m}^2 = (t^2) = \mathfrak{m}^3$. By Nakayama's lemma, this means that $\mathfrak{m}^2 = 0$, which also means $\mathfrak{m}^{n - 1} = 0$ as $n\ge 3$, contradiction. Thus $2\in \mathfrak{m} \backslash \mathfrak{m}^2$, which means that $2 = ut$ for some unit $u$, so $\mathfrak{m} = (t) = (2)$. Finally, note that \[ (2^{n - 1}) = \mathfrak{m}^{n - 1} \neq 0, \] so $2^{n - 1}\neq 0$. Thus in the additive group of $R$, 1 has order $2^n$, which means that we must have $R\simeq \mathbb{Z} / 2^n\mathbb{Z}$. \item If $n\ge 4$, then 5 is a unit in $\mathbb{Z} / 2^n\mathbb{Z}$, but $3^2\not \equiv 1\pmod{2^n}$, so $U(\mathbb{Z} / 2^n\mathbb{Z})\not\simeq C_2^{n - 1}$. \item When $n = 1$, there is a unique ring with two elements, $\mathbb{Z} / 2\mathbb{Z}$. Now if $n = 2$, $\abs{\mathfrak{m}} = 2$, so $\mathfrak{m} = \{0, x\}$ for some $x\neq 0, 1$. Then $R = \{0, 1, x, x - 1\}$. We have two cases: \begin{itemize} \item $2\neq 0$. Then we must have $x = 2$ and $x - 1 = 3$, so $R\simeq \mathbb{Z} / 4\mathbb{Z}$. \item 2 = 0. Then $1 = (x + 1)^2 = x^2 + 1$, so $x^2 = 0, (x+1)^2 = 1, x(x+1) = x$. In this case, $R\simeq \mathbb{F}_2[x] / x^2$. \end{itemize} \end{enumerate} \end{proof} \end{theorem} \section{Proof of \texorpdfstring{\Cref{thm1}}{Theorem \ref{thm1}}} Let $R_1 = \mathbb{Z} / 2\mathbb{Z}, R_2 = \mathbb{Z} / 4\mathbb{Z}, R_3 = \mathbb{Z} / 8\mathbb{Z}, R_4 = \mathbb{F}_2[x]/(x^2)$ denote the ``bad'' local rings. We have that $R$ is of the form \[ R_1^{k_1} \times R_2^{k_2} \times R_3^{k_3} \times R_4^{k_4 } \times R', \] where $R'$ is a product of local rings not isomorphic to $R_1, R_2, R_3$, or $R_4$. Then by \Cref{badrings}, $R'$ is a product of almost unit-stabilized local rings with $\Delta = 0$. In addition, we have \begin{itemize} \item $U(R_1) = 0$, $D(R_1) = 1$. \item $U(R_2)\simeq C_2$, $D(R_2) = 2$. \item $U(R_3)\simeq C_2 \times C_2$, $D(R_3) = 3$. \item $U(R_4)\simeq C_2$, $D(R_4) = 2$. \end{itemize} Plugging this all into \Cref{austhm}, we have \begin{align*} D(R) &= \max_{n_i\in [0, k_i]} \bigg[ D\left( {C_2}^{(k_2 - n_2) + 2(k_3 - n_3) + (k_4 - n_4)} \times U(R')\right) + n_1 + 2n_2 + 3n_3 + 2n_4 \bigg] \\ &= \max_{n_i\in [0, k_i]} \bigg[ D\left( {C_2}^{k_2 + k_4 + 2k_3 - (n_2 + n_4) - 2n_3} \times U(R')\right) + k_1 + 2(n_2 + n_4) + 3n_3 \bigg] \\ &= \max_{0\le a\le k_2 + k_4, 0\le b\le k_3} \bigg[ D\left( U(R') \times {C_2}^{k_2 + k_4 + 2k_3 - a - 2b} \right) + k_1 + 2a + 3b \bigg]. \end{align*} \section{A few more corollaries of \texorpdfstring{\Cref{thm1}}{Theorem \ref{thm1}}} \begin{theorem} \label{randomthms} $D(R) = D(U(R))$ in any of the following scenarios: \begin{enumerate} \item $\abs{R}$ is odd. \item $R$ is an $\mathbb{F}_q$ algebra, where $q$ is a power of 2 other than 2. \item $R = \mathbb{Z} / n\mathbb{Z}$, where $16\mid n$. \item $R = \mathbb{F}_2[x] / (f)$, where $v_x(f), v_{x + 1}(f)\notin \{1, 2\}$. \item $R$ is of the form $\mathbb{Z} / 4\mathbb{Z} \times R'$ or $\mathbb{F}_2[x] / (x^2) \times R'$, where $\abs{U(R')} > 1$ is odd and $\Delta(R') = 0$. \item $R = \mathbb{F}_2[x] / (x^2g)$, where $g$ is a product of (at least one) distinct irreducibles that are not $x$ or $x + 1$. \end{enumerate} \begin{proof} \ \begin{enumerate} \item This follows immediately from \Cref{thm3}, as a ring with odd order cannot have any ideals of index 2. \item Note that any quotient ring of $R$ is also an $\mathbb{F}_q$ algebra, which means that any proper ideal has index at least $q$ in $R$. Thus $n_2(R) = 0$, so by \Cref{thm3}, $\Delta(R) = 0$. \item In the setting of \Cref{thm2}, it is easy to see that for the ring $R = \mathbb{Z} / n\mathbb{Z}$ with $16\mid n$, so we have $k_1 = k_2 = k_3 = k_4 = 0$, which means that $\Delta(R) = 0$ as well. \item This is a direct consequence of \Cref{polythm} below, which is in turn a consequence of \Cref{thm2}. \item We first need the following lemma: \begin{lemma} Suppose $G$ is a nontrivial group of odd order. Then $D(G\times C_2) > D(G) + 1$. \begin{proof} (We will use additive notation for this proof). Let $(g_1, \dots, g_\ell)$ be an irreducible sequence in $G$, where $\ell = D(G) - 1$. Note that the map $g\mapsto g + g$ in $G$ is injective, which means that the map $(\times \frac12)$ is well-defined. Then the sequence \[ (g_1, 0), \dots, (g_{\ell - 1}, 0), (\frac12 g_\ell, 1), (\frac12 g_\ell, 1), (\frac12 g_\ell, 1) \] is an irreducible sequence in $G\times C_2$ of length $\ell + 2 = D(G) + 1$, so $D(G\times C_2) > D(G) + 1$. \end{proof} \end{lemma} Since $\abs{U(R')}$ is odd, the decomposition of $R'$ into a product of local rings cannot have any rings isomorphic to $\mathbb{Z} / 4\mathbb{Z}, \mathbb{Z} / 8\mathbb{Z}$, or $\mathbb{F}_2[x] / (x^2)$. In addition, since $\Delta(R') = 0$, by \Cref{thm2}, this decomposition cannot have any rings isomorphic to $\mathbb{Z} / 2\mathbb{Z}$. Thus the setting of \Cref{thm1} applies; plugging in both $R = \mathbb{Z} / 4\mathbb{Z} \times R'$ and $\mathbb{F}_2[x] / (x^2) \times R'$ gives \[ D(R) = \max(D(U(R')) + 2, D(U(R') \times C_2)) \] However, by the lemma, $D(U(R') \times C_2) > D(U(R')) + 1$, which means that $D(R) = D(U(R') \times C_2) = D(U(R))$. \item If $f\in \mathbb{F}_2$ is irreducible, then the unit group of $\mathbb{F}_2[x] / (f)$ has odd order (in particular, it has order $2^{\deg f} - 1$.) Thus if $g$ is a product of irreducibles other than $x, x + 1$, then $\abs{U(\mathbb{F}_2[x] / (g))} > 1$ is odd, and by \Cref{thm2}, $\Delta(\mathbb{F}_2[x] / (g)) = 0$. Thus this is a special case of (5), with $R' = \mathbb{F}_2[x] / (g)$ and $R\simeq \mathbb{F}_2[x] / (x^2) \times R'$, so $\Delta(R) = 0$. \end{enumerate} \end{proof} \end{theorem} Note that (1) in \Cref{randomthms} is a generalization of \Cref{wangguo} and (2) is a generalization of \Cref{wang}. In addition, we have the following refinement of \Cref{qwz}. \begin{theorem} \label{polythm} Let $f\in \mathbb{F}_2[x]$ be nonconstant and $R = \mathbb{F}_2[x] / (f)$. If $f = x^a (x + 1)^b g$, where $\gcd(g, x(x+1)) = 1$, then \[ \delta_{a1} + \delta_{b1}\le D(R) - D(U(R)) \le \delta_{a1} + \delta_{b1} + \delta_{a2} + \delta_{b2}, \] where $\delta_{ij}$ is the Kronecker $\delta$ function. \begin{proof} Note that if $f$ factors into irreducibles as $f = f_1^{e_1}\cdots f_n^{e_n}$, then the decomposition of $\mathbb{F}_2[x] / (f)$ into products of local rings is \[ \mathbb{F}_2[x]/(f) \simeq \mathbb{F}_2[x] / (f_1^{e_1}) \times \dots \times \mathbb{F}_2[x] / (f_n^{e_n}). \] In the setting of \Cref{thm2}, we have $k_1 = \delta_{a1} + \delta_{b1}$, $k_2 = k_3 = 0$, and $k_4 = \delta_{a2} + \delta_{b2}$. The desired result immediately follows. \end{proof} \end{theorem} Finally, we give a partial answer to the general problem stated by Wang and Guo in \cite{WangGuo2008}: \begin{theorem} If $R = \mathbb{Z}/n_1\mathbb{Z} \times \dots \times \mathbb{Z}/n_r\mathbb{Z}$, then \[ \#\{1\le i\le r: 2\| n_i\} \le D(R) - D(U(R)) \le \#\{1\le i\le r: 2\mid n_i, 16\nmid n_i\}. \] In addition, equality holds on the right when all of the $n_i$ are powers of 2. \begin{proof} Again in the setting of \Cref{thm2}, we have $k_4 = 0$, and for $1\le i\le 3$, \[ k_i = \#\{1\le i\le r: 2^i\| n_i\} \] Thus by \Cref{thm2}, \[ D(R) - D(U(R)) \ge k_1 = \#\{1\le i\le r: 2\| n_i\} \] and \begin{align*} D(R) - D(U(R))&\le k_1 + k_2 + k_3 + k_4 \\ &= \sum_{i = 1}^3 \#\{1\le i\le r: 2^i\| n_i\} \\ &= \#\{1\le i\le r: 2\mid n_i, 16\nmid n_i\}. \end{align*} In addition, if each of the $n_i$ is a power of 2, $U(R)$ is a 2-group, so by \Cref{thm2}, equality holds on the right. \end{proof} \end{theorem} \section{Further Direction} Unfortunately, we still do not have a complete classification of the finite rings $R$ for which $D(R) = D(U(R))$, even if we restrict to the simple cases where $R$ is of the form $\mathbb{Z} / n\mathbb{Z}$ or $\mathbb{F}_2[x] / (f)$. As we can see, these questions depend on the relation between $D(G)$ and $D(G\times C_2^e)$ for an abelian group $G$. This is what we know: If $G = C_{n_1} \times \dots \times C_{n_r}$, where $n_1 \mid \cdots \mid n_r$, then we can define $M(G) = 1 + \sum_{i = 1}^r (n_r - 1)$. Clearly $D(G)\ge M(G)$. However, we have the following: \begin{proposition} \label{gooddelta} $D(G) = M(G)$ in each of the following cases: \begin{enumerate} \item $G$ is a $p$-group. \cite{Olson1969} \item $G = C_n \times C_{kn}$. \cite{Olson1969} \item $G = C_{p^ak} \times H$, where $H$ is $p$-group and $p^a \ge M(H)$. \cite{vebii} \item $G = C_2 \times C_{2n} \times C_{2kn}$, where $n, k$ are odd and the largest prime dividing $n$ is less than 11. \cite{vebii} \item $G = C_2^3 \times C_{2n}$, where $n$ is odd. \cite{baayen} \end{enumerate} \end{proposition} (For a more complete list, see \cite{mazur}.) On the other hand, it is not true that $D(G) = M(G)$ for all $G$. \begin{proposition} \label{baddelta} $D(G) > M(G)$ if $G = C_2^e \times C_{2n} \times C_{2nk}$ for $n, k$ odd and $e\ge 3$. \cite{Geroldinger} \end{proposition} In what follows, we will try and compute exact values of $\Delta(R)$ when $R$ is of the form $\mathbb{Z} / n\mathbb{Z}$ or $\mathbb{F}_2[x] / (f)$. \subsection{\texorpdfstring{$R = \mathbb{Z} / n\mathbb{Z}$}{R = Z / nZ}} By \Cref{thm3}, we have $\Delta(\mathbb{Z} / n\mathbb{Z}) \le 1$. \begin{question} For $n\in\mathbb{Z}^+$, when is $\Delta(\mathbb{Z} / n\mathbb{Z})$ equal to 0 and when is it 1? \end{question} Note that we have already resolved this problem for most $n$. In particular, we have the following: \begin{itemize} \item If $n$ is odd or $16 \mid n$, $\Delta(R) = 0$. \item If $n = 2b$ where $b$ is odd, then $\Delta(R) = 1$. \item If $n = 4b$ where $b$ is odd, then by \Cref{thm1} we have \[ \Delta(R) = \max(0, D(U(\mathbb{Z} / b\mathbb{Z})) + 2 - D(U(\mathbb{Z} / n\mathbb{Z}))) \] However, $U(\mathbb{Z} / n\mathbb{Z}) = U(\mathbb{Z} / b\mathbb{Z}) \times C_2$, which means that \[ D(U(\mathbb{Z} / n\mathbb{Z})) \ge D(U(\mathbb{Z} / b\mathbb{Z})) + 1. \] Thus $\Delta( \mathbb{Z} / n\mathbb{Z}) = 1$ if and only if \[ D(U(\mathbb{Z} / n\mathbb{Z})) = D(U(\mathbb{Z} / b\mathbb{Z})) + 1. \] Using this, we can calculate $\Delta(R)$ in the following special cases using \Cref{gooddelta,baddelta} in conjunction with \Cref{thm1}: \begin{itemize} \item If $b$ is a prime power then $\Delta(R) = 1$. \item If $b$ is a product of distinct primes of the form $2^{2^n} + 1$, then $\Delta(R) = 1$. \item If $b = p_1^{e_1} p_2^{e_2}$ and $\gcd(\phi(p_1^{e_1}), \phi(p_2^{e_2}))$ has no prime factors greater than 7, then $\Delta(R) = 1$. \item If $b = p_1^{e_1}p_2^{e_2}p_3^{e_3}$ and $ \phi(p_1^{e_1})/2, \phi(p_2^{e_2})/2, \phi(p_3^{e_3})/2 $ are pairwise relatively prime, then $\Delta(R) = 1$. \item If $b = p_1^{e_1}p_2^{e_2}p_3^{e_3}p_4^{e_4}$, $p_1, p_2, p_3, p_4 \equiv 3\pmod{4}$, and $ \phi(p_1^{e_1})/2, \phi(p_2^{e_2})/2, \phi(p_3^{e_3})/2, \phi(p_4^{e_4})/2 $ are pairwise relatively prime, then $\Delta(R) = 0$. \end{itemize} \item If $n = 8b$ where $b$ is odd, then using similar logic as the previous case we have $\Delta(\mathbb{Z} / n\mathbb{Z}) = 1$ if and only if \[ D(U(\mathbb{Z} / n\mathbb{Z})) = D(U(\mathbb{Z} / b\mathbb{Z})) + 2. \] Thus we can calculate $\Delta(R)$ in the following special cases: \begin{itemize} \item If $b$ is a prime power then $\Delta(R) = 1$. \item If $b$ is a product of distinct primes of the form $2^{2^n} + 1$, then $\Delta(R) = 1$. \item If $b = p_1^{e_1} p_2^{e_2}$ and $\gcd(\phi(p_1^{e_1}), \phi(p_2^{e_2})) = 1$, then $\Delta(R) = 1$. \item If $b = p_1^{e_1}p_2^{e_2}p_3^{e_3}$, $p_1, p_2, p_3\equiv 3\pmod{4}$, and \[ \gcd[ \phi(p_1^{e_1})/2, \phi(p_2^{e_2})/2, \phi(p_3^{e_3})/2] = 2 \] and for $i\neq j$, $ \gcd[ \phi(p_i^{e_i})/2, \phi(p_j^{e_j})/2] $ has no prime factors greater than 7, then $\Delta(R) = 0$. \item If $b = p_1^{e_1}p_2^{e_2}p_3^{e_3}p_4^{e_4}$, $p_1, p_2, p_3, p_4 \equiv 3\pmod{4}$, and $ \phi(p_1^{e_1})/2, \phi(p_2^{e_2})/2, \phi(p_3^{e_3})/2, \phi(p_4^{e_4})/2 $ are pairwise relatively prime, then $\Delta(R) = 0$. \end{itemize} \end{itemize} However, the general cases $n = 4b, 8b$ remain open. \subsection{\texorpdfstring{$R = \mathbb{F}_2[x]/(f)$}{R = F2[x]/(f)}} In this case, \Cref{thm3} gives the bound $\Delta(\mathbb{F}_2[x] / (f)) \le 2$. \begin{question} For $f\in \mathbb{F}_2[x]$, when is $\Delta(\mathbb{F}_2[x] / (f))$ equal to 0, when is it 1, and when is it 2? \end{question} Again, we have already answered this question for most $f\in \mathbb{F}_2[x]$. Let $f = x^a(x+1)^b g$, where $\gcd(g, x(x + 1)) = 1$. Then \begin{itemize} \item If $g = 1$, then $\Delta(R) = \delta_{a1} + \delta_{a2} + \delta_{b1} + \delta_{b2}$. \item If $a, b\neq 2$, then $\Delta(R) = \delta_{a1} + \delta_{b1}$. \item If $a = 2$ and $b \le 1$ and $\abs{U(\mathbb{F}_2[x] / (g))}$ is odd, then $\Delta(R) = b$. (Note that this condition corresponds to $g$ being a product of distinct irreducibles.) \item If $a = b = 2$ and $U(\mathbb{F}_2[x] / (g))$ is cyclic (i.e. product of distinct irreducibles with pairwise relatively prime degree), then $\Delta(R) = 1$. \item If $a = 2$, $b\in [3, 14]\cup [17, 24] \cup[33, 40]$ and $U(\mathbb{F}_2[x]/(g))$ is cyclic, then $\Delta(R) = 0$. \end{itemize} \subsection{Recap} For general $n\in \mathbb{Z}$, the rank of $U(\mathbb{Z} / n\mathbb{Z})$ is equal to the number of prime factors of $n$, which gets arbitrarily large. Unfortunately, very little is known about $D(G) - M(G)$ for general $G$ of large rank. In a similar vein, the rank of $\mathbb{F}_2[x] / (f)$ for general $f\in \mathbb{F}_2[x] / (f)$ can get annoyingly large, and even the special case $\Delta(\mathbb{F}_2[x]/(x^2g))$ with $g\in \mathbb{F}_2[x]$ irreducible is tricky to calculate. For example, \cite{Geroldinger} gives for any $k > 0$, $D(C_{2k + 1} \times {C_2}^r) = 4k + 1 + r$ for $1\le r\le 4$ but $D(C_{2k + 1} \times {C_2}^5) > 4k + 7$. If $f$ is an irreducible polynomial in $\mathbb{F}_2[x]$ of degree $d$, one can check that $U(\mathbb{F}_2[x] / (f^2))\simeq C_{2^d - 1}\times {C_2}^d$. Thus if $R_d = \mathbb{F}_2[x] / (x^2f_d^2)$, where $f_d$ is an irreducible polynomial of degree $d$, applying \Cref{thm1} gives \[ D(R_d) = \max(D(C_{2^d - 1} \times {C_2}^d) + 2, D(C_{2^d - 1} \times {C_2}^{d+1})) \] which gives $\Delta(R_2) = \Delta(R_3) = 1$ but $\Delta(R_4) = 0$. The general problem of determining $\Delta(\mathbb{F}_2[x] / (f))$ for non-squarefree $f\in \mathbb{F}_2[x]$ is still open for reasons such as this. For more information about the growth of $D(C_{2k + 1} \times {C_2}^d)$, we refer the reader to \cite{mazur}. \subsection{Infinite Semigroups} Finally, we look at the case where $S$ is an infinite commutative semigroup. By \Cref{thm4} and \Cref{inflemma}, when $S$ is either \begin{itemize} \item a group, or \item the semigroup of a commutative ring $R$ under multiplication, \end{itemize} then $D(S)$ being finite implies $S$ is finite. However, there are semigroups $S$ such that $S$ is infinite but $S$ is finite. Thus we have the following question: \begin{question} For what other families $\mathcal{F}$ of (commutative, unital) semigroups is it true $S\in \mathcal{F}$ and $\abs{S} = \infty$ implies $D(S) = \infty$? \end{question} \section*{Acknowledgments} This research was conducted at the University of Minnesota Duluth REU and was supported by NSF grant 1358695 and NSA grant H98230-13-1-0273. The author thanks Joe Gallian for suggesting the problem and supervising the research, and Benjamin Gunby for helpful comments on the manuscript. \bibliographystyle{plain}
{ "timestamp": "2016-02-11T02:10:05", "yymm": "1602", "arxiv_id": "1602.03445", "language": "en", "url": "https://arxiv.org/abs/1602.03445", "abstract": "The Davenport constant is one measure for how \"large\" a finite abelian group is. In particular, the Davenport constant of an abelian group is the smallest $k$ such that any sequence of length $k$ is reducible. This definition extends naturally to commutative semigroups, and has been studied in certain finite commutative rings. In this paper, we give an exact formula for the Davenport constant of a general commutative ring in terms of its unit group.", "subjects": "Group Theory (math.GR); Commutative Algebra (math.AC); Combinatorics (math.CO)", "title": "Davenport constant for commutative rings", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9843363512883317, "lm_q2_score": 0.8128673223709251, "lm_q1q2_score": 0.8001348541841125 }
https://arxiv.org/abs/2010.05840
Cohomology fractals, Cannon-Thurston maps, and the geodesic flow
Cohomology fractals are images naturally associated to cohomology classes in hyperbolic three-manifolds. We generate these images for cusped, incomplete, and closed hyperbolic three-manifolds in real-time by ray-tracing to a fixed visual radius. We discovered cohomology fractals while attempting to illustrate Cannon-Thurston maps without using vector graphics; we prove a correspondence between these two, when the cohomology class is dual to a fibration. This allows us to verify our implementations by comparing our images of cohomology fractals to existing pictures of Cannon-Thurston maps.In a sequence of experiments, we explore the limiting behaviour of cohomology fractals as the visual radius increases. Motivated by these experiments, we prove that the values of the cohomology fractals are normally distributed, but with diverging standard deviations. In fact, the cohomology fractals do not converge to a function in the limit. Instead, we show that the limit is a distribution on the sphere at infinity, only depending on the manifold and cohomology class.
\section{Introduction} \begin{figure}[htb] \centering \subfloat[Cannon--Thurston map.]{ \label{Fig:CTVector} \includegraphics[width=0.47\textwidth]{Figures/match_up/two_colour2_rot/two_colour_Cannon-Thurston_match_up_vector_1000} } \thinspace \subfloat[Cohomology fractal.]{ \label{Fig:CTPixelColour} \includegraphics[width=0.47\textwidth]{Figures/match_up/two_colour2_rot/two_colour_Cannon-Thurston_match_up_colour_1000} } \subfloat[\reffig{CTPixelColour} in black and white.]{ \label{Fig:CTPixelBW} \includegraphics[width=0.47\textwidth]{Figures/match_up/two_colour2_rot/two_colour_Cannon-Thurston_match_up_bw_1000} } \thinspace \subfloat[\reffig{CTVector} overlaid on \reffig{CTPixelBW}.]{ \label{Fig:CTPixelBWandVector} \includegraphics[width=0.47\textwidth]{Figures/match_up/two_colour2_rot/two_colour_Cannon-Thurston_match_up_bw+vector_1000} } \caption{Matching up Cannon--Thurston map images with the cohomology fractal for \texttt{m004}, the figure-eight knot complement. Compare our \reffig{CTPixelBW} with Figure~10.11 of \emph{Indra's Pearls}~\cite[page 335]{IndrasPearls}, which was produced by paint-filling a vector graphics image~\cite{Wright19}.} \label{Fig:MatchUp} \end{figure} Cannon and Thurston discovered that Peano curves arise naturally in hyperbolic geometry~\cite{CannonThurston07}. They proved that for every closed hyperbolic three-manifold, equipped with a fibration over the circle, there is a map from the circle to the sphere that is continuous, finite to one, and surjective. Furthermore this \emph{Cannon--Thurston map} is equivariant with respect to the action of the fundamental group. We review their construction in \refsec{CannonThurston}; \reffig{CTVector} shows an approximation. In a previous expository paper~\cite{BachmanSchleimerSegerman20}, we introduced \emph{cohomology fractals}; these are images arising from a hyperbolic three-manifold $M$ equipped with a cohomology class $[\omega] \in H^1(M; \mathbb{R})$. See~\reffig{CTPixelColour}. In that paper we gave an overview of the construction; we also discussed some of the features of the three-manifold and cohomology class that can be seen in its cohomology fractal. We have also written an open-source~\cite{github_cohomology_fractals} real-time web application for exploring these fractals. This is available at \url{https://henryseg.github.io/cohomology_fractals/}. In the present work we give rigorous definitions of cohomology fractals, we relate them to Cannon--Thurston maps (see \reffig{CTPixelBWandVector}), we give technical details of our implementation, and we discuss their limiting behaviour. We now outline the contents of each section of the paper. Note that we include a glossary of notation in \refapp{Notation}. We begin by reviewing the definitions of ideal and material triangulations, and their hyperbolic geometry in \refsec{Triangulations}. In \refsec{CannonThurston} we define Cannon--Thurston maps. In \refsec{Graphics} we discuss the differences between vector and raster graphics. We also recall a vector graphics algorithm (\refalg{CTApprox}) used in previous work to illustrate Cannon--Thurston maps. In \refsec{CohomologyFractals} we give several equivalent definitions of the cohomology fractal. It depends on choices beyond the manifold $M$ and the cohomology class $[\omega]$: there is a choice of viewpoint $p \in M$ and a choice of a visual radius $R$. The fractal is a function $\Phi^{\omega,p}_{R} \from \UT{p}{M} \to \mathbb{R}$. Roughly, for each vector $v \in \UT{p}{M}$ we build the geodesic arc $\gamma$ of length $R$ from $p$ in the direction of $v$ and compute $\Phi^{\omega,p}_{R}(v) = \omega(\gamma)$. (Note that we repeatedly generalise the definition of the cohomology fractal throughout the paper; the decorations alter to remind the reader of the desired context.) In \reffig{MatchUp}, we see a cohomology fractal closely matching an approximation of a Cannon--Thurston map, as produced by \refalg{CTApprox}. In \refsec{Matching} we prove the following. \begin{restate}{Proposition}{Prop:LightDark} Cohomology fractals are dual to approximations of the Cannon--Thurston map. \end{restate} Thus we have a new representation of Cannon--Thurston maps. We also compare cohomology fractals with the \emph{lightning curves} of Dicks and various coauthors. (The name is due to Wright~\cite[page~324]{IndrasPearls}.) We experimentally observe that the lightning curve corresponds to some of the brightest points of the cohomology fractal. In \refsec{Implement} we describe the algorithms we use to produce images of cohomology fractals. Adding the ability to move through the manifold leads us to separate the viewpoint $p$ from a \emph{basepoint}, denoted $b$, of the cohomology fractal. We still trace rays starting at $p$, but then evaluate $\omega$ on any path in $\cover{M}$ from $b$ to the endpoint of $\gamma$. We also generalise the above \emph{material} view (with vectors $v$ in $\UT{p}{\cover{M}}$) to the \emph{ideal} and \emph{hyperideal} views (with vectors $v$ being perpendicular to a horosphere or geodesic plane, respectively). Each view is a subset $D \subset \UT{}{\cover{M}}$; our notation for the cohomology fractal becomes $\Phi^{\omega,b,D}_{R} \from D \to \mathbb{R}$. In \refsec{Cone} we discuss cohomology fractals for incomplete and closed manifolds. We draw cohomology fractals in the closed case in two ways. First, we deform the cohomology fractal for a surgery parent through Thurston's \emph{Dehn surgery space}. Second, we reimplement our algorithms using material triangulations. We also discuss possible sources of numerical error in our implementations. In \refsec{Experiments} we give a sequence of experiments exploring the dependence of cohomology fractals on the visual radius $R$. For any fixed $R$, the cohomology fractal is constant on regions with sizes roughly proportional to $\exp(-R)$. As $R$ increases, these regions subdivide, and intricate patterns come into focus. This suggests that there is a limiting object. The following shows that such a limit cannot be a function. \begin{restate}{Theorem}{Thm:NoPicture} Suppose that $M$ is a finite volume, oriented hyperbolic three-manifold. Suppose that $F$ is a transversely oriented surface. Then the limit \[ \lim_{R \to \infty} \Phi_R(v) \] does not exist for almost all $v \in \UT{}{\cover{M}}$. \end{restate} Indeed, experimentally, increasing $R$ leads to noisy pictures. However, this is due to undersampling. A heuristic argument (see \refrem{Embiggen}) shows that we can avoid noise if we increase the screen resolution as we increase $R$. We simulate this by computing \emph{supersampled} images. These, and further experiments, indicate that in contrast with \refthm{NoPicture}, the \emph{mean} of the cohomology fractal, taken over a pixel, converges. Its values appear to be normally distributed with standard deviation growing like $\sqrt{R}$. Motivated by this, in \refsec{CLT}, we show that the cohomology fractal obeys a central limit theorem. \begin{restate}{Theorem}{Thm:CLT} Fix a connected, orientable, finite volume, complete hyperbolic three-manifold $M$ and a closed, non-exact, compactly supported one-form $\omega \in \Omega_c^1(M)$. There is $\sigma > 0$ such that for all basepoints $b$, all views $D$ with area measure $\mu_D$, for all probability measures $\nu_D\ll \mu_D$, and for all $\alpha\in\mathbb{R}$, we have $$\lim_{T\to\infty} \nu_D\left[ v\in D : \frac{\Phi_T(v)}{\sqrt{T}} \leq \alpha \right] = \int_{-\infty}^\alpha \frac{1}{\sigma\sqrt{2\pi}} e^{-(s/\sigma)^2/2} ds$$ where $\Phi_T=\Phi^{\omega,b,D}_{T}$ is the associated cohomology fractal. \end{restate} \noindent That is, if we regard the cohomology fractal across a pixel as a random variable, divide it by $\sqrt{T}$, and take the limit, the result is a normal distribution of mean zero. The standard deviation of the normal distribution only depends on the manifold and cohomology class. The proof uses Sinai's central limit theorem for geodesic flows. In \refsec{Pixel}, we prove that treating the cohomology fractals as \emph{distributions} gives a well-defined limit. In this introduction, for simplicity, we focus on the case where $D$ is a material view. The \emph{pixel theorem} (\refthm{Pixel}) states that the limit \[ \Phi^{\omega,b,D}(\eta) = \lim_{T\to\infty} \int_{D} \Phi^{\omega,b,D}_{T} \eta \] is well-defined for any two-form $\eta \in \Omega^2(D)$. \refthm{Pixel} also states various transformation laws relating, for example, the distributions corresponding to different views. Thus there is a view-independent distribution related to the view-dependent distributions via the conformal isomorphism $i_D$ from $D$ to $\bdy \cover{M}$. \begin{restate}{Corollary}{Cor:BoundaryDistribution} Suppose that $M$ is a connected, orientable, finite volume, complete hyperbolic three-manifold. Fix a closed, compactly supported one-form $\omega \in \Omega_c^1(M)$ and a basepoint $b \in \cover{M}$. Then there is a distribution $\Phi^{\omega,b}$ on $\bdy_\infty \cover{M}$ so that, for any material view $D$ and for any $\eta \in \Omega^2(D)$, we have \[ \Phi^{\omega,b,D}(\eta) = \Phi^{\omega, b}((i_D^{-1})^* \eta) \] \end{restate} The above discussion addresses smooth test functions. We can also prove convergence for a wider class of test functions; these include the indicator functions of regions with piecewise smooth boundary. However, we do not know whether or not the cohomology fractal converges to a measure. We conclude with a few questions and directions for future work in \refsec{Questions}. \subsection*{Acknowledgements} This material is based in part upon work supported by the National Science Foundation under Grant No. DMS-1439786 and the Alfred P. Sloan Foundation award G-2019-11406 while the authors were in residence at the Institute for Computational and Experimental Research in Mathematics in Providence, RI, during the Illustrating Mathematics program. The fourth author was supported in part by National Science Foundation grant DMS-1708239. We thank Fran\c{c}ois Gu\'eritaud for suggesting we use ray-tracing to generate cohomology fractals. We thank Curt McMullen for suggesting that \refthm{Pixel} should be true and also for giving us permission to reproduce \reffig{McMullen}. We thank Mark Pollicott and Alex Kontorovich for guiding us through the literature on exponential mixing of the geodesic flow. We thank Ian Melbourne for enlightening conversations on central limit theorems. \section{Triangulations} \label{Sec:Triangulations} We briefly review the notions of material and ideal triangulations of three-manifolds. \subsection{Combinatorics} Suppose that $M$ is a compact, connected, oriented three-manifold. We will consider two cases. Either \begin{itemize} \item the boundary $\bdy M$ is empty; here we call $M$ \emph{closed}, or \item the boundary is non-empty, consisting entirely of tori; here we call $M$ \emph{cusped}. \end{itemize} Suppose that $\mathcal{T}$ is a \emph{triangulation}: that is, a collection of oriented model tetrahedra together with a collection of orientation-reversing face pairings. We allow distinct faces of a tetrahedron to be glued, but we do not allow a face to be glued to itself. The quotient space, denoted $|\mathcal{T}|$, is thus a CW--complex which is an oriented three-manifold away from its zero-skeleton. We say that $\mathcal{T}$ is a \emph{material} triangulation of $M$ if there is an orientation-preserving homeomorphism from $|\mathcal{T}|$ to $M$. We say that $\mathcal{T}$ is an \emph{ideal} triangulation of $M$ if there is an orientation-preserving homeomorphism from $|\mathcal{T}|$, minus a small open neighbourhood of its vertices, to $M$. Equivalently, $|\mathcal{T}|$ minus its vertices is homeomorphic to $M^\circ$, the interior of $M$. \begin{example} Suppose that $M$ is obtained from $S^3$ by removing a small open neighbourhood of the figure-eight knot. See \reffig{FigureEightKnot}. As discussed in~\cite[Chapter 1]{ThurstonNotes}, the knot exterior $M$ has an ideal triangulation with two tetrahedra. See \reffig{FigureEightTriangulation}. Here we have not truncated the model tetrahedra. Instead we draw the vertex link; in this case it is a torus in $M$. \end{example} \begin{figure}[htb] \centering \subfloat[{Tube around the figure-eight knot.}]{ \label{Fig:FigureEightKnot} \includegraphics[width = 0.33\textwidth]{Figures/fig_8_knot_triang_2d_pic_3k} }% \hfill% \subfloat[Triangulation of the figure-eight knot complement. The colours and arrows indicate the gluings.]{ \label{Fig:FigureEightTriangulation} \includegraphics[width = 0.61\textwidth]{Figures/Fig8_knot_complement} } \caption{The figure-eight knot complement. This manifold is known as \texttt{m004} in the SnapPy census. The black lines in \reffig{FigureEightKnot} cut the tube into triangles, corresponding to the eight green triangles in \reffig{FigureEightTriangulation}.} \end{figure} \subsection{Geometry} \label{Sec:Geometry} We deal with the geometry of the two types of triangulations separately. \subsubsection{Ideal triangulations} We give each model ideal tetrahedron $t$ a hyperbolic structure. That is, we realise $t$ as an ideal tetrahedron in $\mathbb{H}^3$ with geodesic faces. This can be constructed as the convex hull of four points on $\bdy_\infty \mathbb{H}^3$. We require that the face pairings be orientation reversing isometries. For these hyperbolic tetrahedra to combine to give a complete hyperbolic structure on the manifold $M^\circ$ requires certain conditions to be satisfied. Very briefly: consider a loop in the dual one-skeleton of the triangulation. This visits the tetrahedra in some order. The product of the corresponding sequence of isometries must give the identity if the loop is trivial in the fundamental group. If the loop is peripheral then the product must be a parabolic element. These conditions reduce to a finite set of algebraic constraints. These are Thurston's \emph{gluing equations}, see~\cite[Section~4.2]{ThurstonNotes} and~\cite[Section~4.2]{PurcellKnotTheory}. Using the upper half-space model of $\mathbb{H}^3$ we define the \emph{shape} of each ideal hyperbolic tetrahedron to be the cross-ratio of its four ideal points. The gluing equations impose a finite number of polynomial conditions on these shapes. For our implementation, we also require that the shapes have positive imaginary part. This ensures that the ideal hyperbolic tetrahedra glue together to give a complete, finite volume hyperbolic structure on $M^\circ$. Furthermore, the model orientations of all of the tetrahedra agree with the orientation on $M$. In particular, when a geodesic ray crosses a face, it has a sensible continuation. \subsubsection{Material triangulations} \label{Sec:MaterialGeometry} To find a hyperbolic structure for material triangulations, we replace Thurston's gluing equations with a construction due to Andrew Casson \cite{casson:geo} and Damian Heard \cite{Orb,heardThesis}. To specify a hyperbolic structure on a material triangulation, it suffices to assign lengths to its edges. This is because the isometry class of an oriented material hyperbolic tetrahedron is determined by its six edge lengths. There are two conditions that must be satisfied. First, for each model tetrahedron, there is a collection of inequalities that must be satisfied for its edges. Second, for each edge of the triangulation, the dihedral angles about it must sum to $2\pi$. If these inequalities and equalities hold, then we obtain a hyperbolic structure on the three-manifold. See \cite[Section~2]{matthiasVerifyingFinite} for further details. \section{Cannon--Thurston maps} \label{Sec:CannonThurston} \begin{wrapfigure}[15]{r}{0.36\textwidth} \vspace{-21pt} \begin{minipage}{0.355\textwidth} \[ \begin{tikzcd} \bdy_\infty \cover{F} \arrow[r, "\Psi"] \arrow[d, hook'] & \bdy_\infty \cover{M} \arrow[d, hook'] \\ \closure{F} \arrow[r, "\closure{\alpha}"] & \closure{M} \\ \cover{F} \arrow[r, "\cover{\alpha}"] \arrow[u, hook] \arrow[d] & \cover{M} \arrow[u, hook] \arrow[d] \\ F \arrow[r, "\alpha"] & M \end{tikzcd} \] \end{minipage} \caption{The various spaces and maps involved in constructing the Cannon--Thurston map $\Psi$.} \label{Fig:CannonThurston} \end{wrapfigure} Here we sketch Cannon and Thurston's construction; see \reffig{CannonThurston} for an overview. We refer to~\cite{CannonThurston07} for the details. See also~\cite{Mj18}. Suppose that $F=F^2$ and $M=M^3$ are connected, compact, oriented two- and three-manifolds. Suppose that $F^\circ$ and $M^\circ$ admit complete hyperbolic metrics of finite area and volume respectively. In an abuse of notation, we will conflate $F$ with $F^\circ$, and similarly $M$ with $M^\circ$. We call a proper embedding $\alpha \from F \to M$ a \emph{fibre} if there is a map $\rho \from M \to S^1$ so that for all $t \in S^1$ the preimage $\rho^{-1}(t)$ is a surface properly isotopic to $\alpha(F)$. Let $\cover{F}$ and $\cover{M}$ be the universal covers of $F$ and $M$ respectively. Since $F$ and $M$ are hyperbolic, their covers are identified with hyperbolic two- and three-space respectively. Let $\bdy_\infty \cover{F} \homeo \bdy_\infty\mathbb{H}^2 \homeo S^1$ and $\bdy_\infty \cover{M} \homeo \bdy_\infty\mathbb{H}^3 \homeo S^2$ be their ideal boundaries. We set \[ \closure{F} = \cover{F} \cup \bdy_\infty \cover{F} \qquad \mbox{and} \qquad \closure{M} = \cover{M} \cup \bdy_\infty \cover{M} \] Each union is equipped with the unique topology that makes the group action continuous. Note that $\closure{F}$ and $\closure{M}$ are homeomorphic to a closed two- and three-ball, respectively. We compose the covering map from $\cover{F} \to F$ with the embedding $\alpha$ and then lift to obtain an equivariant map $\cover{\alpha} \from \cover{F} \to \cover{M}$. We call $\cover{\alpha}$ an \emph{elevation} of $F$. \reffig{Elevation} shows an elevation of the fibre of the figure-eight knot complement. Cannon and Thurston gave the first proof of the following theorem in the closed case~\cite[page 1319]{CannonThurston07}. The cusped case follows from work of Bowditch~\cite[Theorem 0.1]{Bowditch07}. \begin{theorem} \label{Thm:CannonThurston} Suppose $M$ is a connected, oriented, finite volume hyperbolic three-manifold. Suppose that $\alpha \from F \to M$ is a fibre of a surface bundle structure on $M$. Then there is an extension of $\cover{\alpha}$ to a continuous and equivariant (with respect to the fundamental group of $M$) map $\closure{\alpha} \from \closure{F} \to \closure{M}$. The restriction of $\closure{\alpha}$ to $\bdy_\infty \cover{F}$ gives a sphere-filling curve. \qed \end{theorem} We will use the notation $\Psi \from \bdy_\infty \cover{F} \to \bdy_\infty \cover{M}$ for the restriction of $\closure{\alpha}$ to $S^1_\infty$. We call this a \emph{Cannon--Thurston map}. We now turn to the task of visualising $\Psi$. \section{Illustrating Cannon--Thurston maps} \label{Sec:Graphics} The standard joke (see~\cite[page~373]{Thurston82} and~\cite[page~335]{IndrasPearls}) is that it is straightforward to draw an accurate picture of a Cannon--Thurston map; it is solid black. \begin{figure}[htb] \centering \subfloat[An elevation of the fibre. The fibre is a pleated surface made from two ideal triangles. The three pleating angles are $\pi/3$, $\pi$, and $5\pi/3$.]{ \includegraphics[width = 0.47\textwidth]{Figures/m004_elevation} \label{Fig:Elevation} } \subfloat[An approximation of the Cannon--Thurston map, reproduced from Figure~8 of~\cite{Thurston82}.] { \includegraphics[width = 0.47\textwidth]{Figures/bending_the_circle_fig_8_part_6} \label{Fig:Thurston} } \caption{Views in the universal cover of the figure-eight knot complement. } \end{figure} The first (more instructive) illustration of a Cannon--Thurston map is due to Thurston. He gives a sequence of approximations to the sphere-filling curve in~\cite[Figure~8]{Thurston82}. We reproduce the last of these in \reffig{Thurston}. A striking version of this image by Wright also appears in \emph{Indra's Pearls}~\cite[Figure~10.11]{IndrasPearls}. In this example both $M$ and $F$ are non-compact; $M$ is the complement of the figure-eight knot and $F$ is a Seifert surface. \subsection{Vector and raster graphics} In this section, we outline the technique used by Thurston and Wright to generate images of Cannon--Thurston maps, in order to contrast it with our algorithm. Our algorithm generates an image by producing a colour for each pixel on a screen. In other words, its output is a map from a grid of pixels in the \emph{image plane} into a space of possible colours. We call such a map a \emph{raster graphics image}. In contrast, \refalg{CTApprox} (below) produces \emph{vector graphics} -- that is a description of an image as a collection of various primitive objects in the (euclidean) image plane. An example of a primitive is a line segment, specified by the coordinates of its end points. Other primitives include arcs, circles, and so on. Note that we generally need to convert vector graphics to raster graphics to make a physical representation of an image. To \emph{rasterise} a vector graphics image, we need to decide which pixels are coloured by which primitives. For example, a disk colours all of the pixels whose coordinates are close enough to the centre of the disk. Rasterisation is necessary for most output devices, such as screens or printers. The exceptions include plotters, laser cutters, and cathode-ray oscilloscopes. There are advantages to deferring rasterisation and saving the vector graphics to a file (in the PDF, PostScript, or SVG format, for example). For example, deferred rasterisation can take the resolution of the output device into account. Design programs such as Inkscape and Adobe Illustrator allow editing the geometric primitives in a vector graphics image. Rasterisation is usually carried out by a black-box general purpose algorithm, the details of which are hidden from the user. \begin{algorithm} \label{Alg:CTApprox} (Approximate a Cannon--Thurston map) We are given a fibre $F$ of the three-manifold $M$. We choose an elevation $\cover{F} \subset \cover{M}$ of $F$. As described in \refsec{CannonThurston}, the map \[ \Psi \from S^1 \homeo \bdy_\infty \cover{F} \to \bdy_\infty \cover{M} \homeo S^2 \] is sphere-filling. To approximate $\Psi$, we first choose a large disk $D \subset \cover{F}$. Typically, $M$ is described with an ideal triangulation $\mathcal{T}$, with $F$ realised as a surface carried by the two-skeleton $\mathcal{T}^{(2)}$. Therefore $\cover{F}$ is given as a surface carried by $\cover{\mathcal{T}}^{(2)}$. The disk $D$ then consists of some finite collection of ideal hyperbolic triangles in $\cover{\mathcal{T}}^{(2)}$. The boundary of $D$ consists of a loop of geodesics in $\mathbb{H}^3 \cup \bdy_\infty \mathbb{H}^3$. We now define $\Psi_D$ to be the loop in $\bdy_\infty \mathbb{H}^3$ obtained by projecting each arc of $\bdy D$ to an arc in $\bdy_\infty \mathbb{H}^3$. \end{algorithm} Note that the algorithm produces a circularly ordered collection of points in $\bdy_\infty \mathbb{H}^3$ spanning geodesics in $\mathbb{H}^3$. However, conventional vector graphics require primitives to be in the euclidean plane. Thus, we must make two choices of projections. The first projection from $\mathbb{H}^3$ to $\bdy_\infty\mathbb{H}^3$ takes the geodesics to arcs in $\bdy_\infty\mathbb{H}^3$ and the second projection takes these arcs in $\bdy_\infty\mathbb{H}^3$ to arcs in the euclidean image plane. We draw our pictures in the ``ideal view''. That is, we use the upper half space model of $\mathbb{H}^3$ and project $\bdy D$ down to $\mathbb{C}$ (viewed as the boundary of $\mathbb{H}^3$). The arcs between vertices now simply become straight lines in $\mathbb{C}$. This is also the choice made by Thurston in~\reffig{Thurston}, as well as Wada in his program \emph{OPTi}~\cite{opti}. Other depictions by McMullen~\cite{McMullenWeb} and Calegari~\cite[Figure~1.14]{Calegari07} project outward from the origin to the boundary of the Poincar\'e ball model of $\mathbb{H}^3$ (and then to the image plane using a perspective or orthogonal projection). \begin{remark} The images in \emph{Indra's Pearls}~\cite{IndrasPearls} have been rasterised using a customised rasteriser that illustrates further features of the Cannon--Thurston map. For example, one side of the polygonal path of line segments is filled, or the line segments are coloured using some combinatorial condition. See Figures~10.11 and~10.13 of~\cite{IndrasPearls}. \end{remark} \subsection{Motivating raster graphics} Our work here began when we asked if we could avoid vector graphics when illustrating Cannon--Thurston maps. This is less natural, but would allow us to take advantage of extremely fast graphics processing unit (GPU) calculation. We were inspired in part by work of Vladimir Bulatov~\cite{Bulatov18} and also of Roice Nelson and the fourth author~\cite{visualizing_hyperbolic_honeycombs}, using reflection orbihedra. (See also the work of Peter Stampfli~\cite{Stampfli19}.) They all use raster graphics strategies to draw tilings of $\mathbb{H}^2$ and $\mathbb{H}^3$. A number of others have also used raster graphics to explore kleinian groups, outside of the setting of reflection orbihedra. They include Peter Liepa~\cite{Liepa11}, Jos Leys~\cite{Leys17}, and Abdelaziz Nait Merzouk~\cite{Knighty} (also see~\cite{Christensen}). \section{Cohomology fractals} \label{Sec:CohomologyFractals} We give a sequence of more-or-less equivalent definitions of cohomology fractals, beginning with the conceptually simplest (for us), and moving towards versions that are most convenient for our implementation or our proofs. Fixing notation, we take $M$ to be a riemannian manifold, $\cover{M}$ its universal cover, and $\pi \from \UT{}{M} \to M$ its unit tangent bundle. \begin{definition} \label{Def:Basic} Suppose that we are given the following data. \begin{itemize} \item A connected, complete, oriented riemannian manifold $M^n$, \item a cocycle $\omega \in Z^1(M; \mathbb{Z})$, \item a point $p \in M$, and \item a radius $R \in \mathbb{R}_{>0}$. \end{itemize} From these, we define the \emph{cohomology fractal} on the unit sphere $\UT{p}{M} \homeo S^{n-1}$. This is a function $\Phi_R = \Phi^{\omega,p}_{R} \from \UT{p}{M} \to \mathbb{Z}$ defined as follows. Suppose that $v \in \UT{p}{M}$ is a unit tangent vector. Let $\gamma$ be the unique geodesic segment starting at $p$ with initial direction $v$ and of length $R$. Let $q$ be the endpoint of $\gamma$. Choose any shortest path $\gamma'$ from $q$ to $p$ (on a set of full measure $\gamma'$ is unique). Thus $\gamma \cup \gamma'$ is a one-cycle. We define $\Phi_R(v) = \omega(\gamma \cup \gamma')$. \end{definition} Our next definition moves in the direction of concrete examples: \begin{definition} \label{Def:Dual} Here we further assume that $M$ is a three-manifold. Let $F \subset M$ be a properly embedded, transversely oriented surface. We choose $F$ so that $p \notin F$. We define $\gamma$ and $q$ as above. Now take $\gamma'$ to be the shortest path from $q$ to $p$ in the complement of $F$. We now define $\Phi_R(v) = \Phi^{F,p}_{R}(v)$ to be the algebraic intersection number between $F$ and $\gamma \cup \gamma'$. \end{definition} We modify once again to obtain a definition very close to our implementation. \begin{definition} \label{Def:UsingTriangulation} We equip $M$ with a material (or ideal) triangulation $\mathcal{T}$. We properly homotope the surface $F$ to lie in the two-skeleton $\mathcal{T}^{(2)}$. For each face $f$ this gives us a weight $\omega(f)$. This is the signed number of sheets of $F$ running across $f$. We dispense with $\gamma'$; we take $\Phi_R(v)$ to be the sum of the weighted intersections between $\gamma$ and the faces of the triangulation. \end{definition} To aid in comparing cohomology fractals to Cannon--Thurston maps (in \refsec{Matching}), we lift to the universal cover, $\cover{M}$. \begin{definition} \label{Def:UniversalCover} Since cochains pull back, let $\cover{\omega}$ be the lift of $\omega$. Let $\cover{p}$ be a fixed lift of the point $p$. Since $\cover{\omega}$ is a coboundary, it has a primitive, say $W$; we choose $W$ so that $W(\cover{p}) = 0$. We form $\cover{\gamma}$ as before and let $\cover{q}$ be its endpoint. We define $\Phi_R(v) = W(\cover{q})$. \end{definition} To analyse the behaviour of the cohomology fractal as $R$ tends to infinity, we rephrase our definition in a dynamical setting. Here, the radius $R$ is replaced by a time $T$. \begin{definition} \label{Def:OneForm} Suppose that $\omega \in \Omega^1(M,\mathbb{R})$ is a closed one-form. Let $\varphi_t \from \UT{}{M} \to \UT{}{M}$ be the geodesic flow for time $t$. We define \[ \Phi_T(v) = \Phi_{T}^{\omega,p}(v) = \int_0^T \omega(\varphi_t(v))\, dt \qedhere \] \end{definition} \begin{figure}[htb!] \centering \subfloat[$R=e^{0.5}$]{ \label{Fig:exp0.5} \includegraphics[width=0.47\textwidth]{Figures/m004_exp0p5} } \subfloat[$R=e^1$]{ \includegraphics[width=0.47\textwidth]{Figures/m004_exp1} } \\ \subfloat[$R=e^{1.5}$]{ \includegraphics[width=0.47\textwidth]{Figures/m004_exp1p5} } \subfloat[$R=e^{2}$]{ \includegraphics[width=0.47\textwidth]{Figures/m004_exp2} } \caption{Cohomology fractals for \texttt{m004}, with various values of $R$.} \label{Fig:VisSphereRadii} \end{figure} In \refsec{Implement} we will discuss how to calculate $\Phi_R$ in practice. Before giving those details, we show the reader what $\Phi_R$ looks like for a few values of $R$. See \reffig{VisSphereRadii}. The map $\Phi_R$ maps into $\mathbb{R}$; we indicate the value of $\Phi_R(v)$ by brightness. For each value of $R$, we draw the value of $\Phi_R(v)$ for a small square subset of the unit tangent vectors, $\UT{p}{M}$. Here we are using \refdef{UsingTriangulation}, our manifold $M$ is the complement of the figure-eight knot, and the surface $F$ is a fibre of $M$. Note that when $R$ is small, as in \reffig{exp0.5}, $\Phi_R$ is constant on large regions of the sphere. As $R$ increases, the value of $\Phi_R$ on nearby rays becomes less correlated, and we see a fractal structure come into focus. \begin{remark} This complicated behaviour is a consequence of the hyperbolic geometry of our manifold. Consider instead the example where $M$ is the three-torus $S^1 \times S^1 \times S^1$ and the surface $F$ is an essential torus embedded in $M$. Again, $\Phi_R$ counts the number of elevations of $F$ the ray $\gamma$ passes through. Since the elevations are parallel planes in $\cover{M} \homeo \mathbb{R}^3$, the value of $\Phi_R$ is constant on circles in $\UT{p}{M}$ parallel to these planes. Here $\Phi_R$ is much simpler. \end{remark} \begin{remark} Some of the geometry and topology of the manifold $M$ can be seen from the cohomology fractal $\Phi_R$. Our recent expository paper on cohomology fractals~\cite{BachmanSchleimerSegerman20} gives many such examples, including the appearances of cusps, totally geodesic subsurfaces, and loxodromic elements of the fundamental group. \end{remark} \section{Matching figures} \label{Sec:Matching} In this section we compare our cohomology fractals to Cannon--Thurston maps. \begin{example} Suppose that $M$ is the figure-eight knot complement. Suppose that $F$ is a fibre of the fibration of $M$. Suppose that $\Psi$ is the associated Cannon--Thurston map. \reffig{CTVector} shows an approximation $\Psi_D \from S^1 \to S^2 \homeo \bdy_\infty \mathbb{H}^3$ of $\Psi$; we produced this image using \refalg{CTApprox}. (Note that the vector graphics image in \reffig{CTVector} has been converted to a raster graphics image to save on file size and rendering time.) \reffig{CTPixelColour} shows a cohomology fractal $\Phi_R$ corresponding to $F$ and looking towards the same part of $\bdy_\infty \mathbb{H}^3$. \reffig{CTPixelBW} shows $\Phi_R$ again, but with the contrast increased and colour scheme simplified. Here the colour associated to a vector $v$ is either white or grey, according to whether $\Phi(v)$ is negative or not. \reffig{CTPixelBWandVector} shows \ref{Fig:CTVector} overlaid on \ref{Fig:CTPixelBW}. We see that the red curve of $\Psi_D$ is almost the common boundary of the white and grey regions of $\Phi_R$. There are several small areas where $\Psi_D$ does not track the boundary. These only appear close to fairly large cusps; they exist because implementations of \refalg{CTApprox} generally have trouble approaching cusps from the side. In the cohomology fractal $\Phi_R$ we see that there are chains of ``octopus heads'' that reach almost all the way towards each cusp. \end{example} This behaviour is generally true for fibrations, as follows. Here we follow \refdef{UniversalCover}. \begin{proposition} \label{Prop:LightDark} Suppose that $M$ is a connected, oriented, finite volume hyperbolic three-manifold. Suppose that $\alpha \from F \to M$ is a fibre of a surface bundle structure on $M$. Fix $p$ in $F$, and a lift $\cover{p} \in \cover{M}$. Fix any $R > 0$ and let $\Phi_R = \Phi^F_R$ be the resulting cohomology fractal for $F$. Let $Z = \Phi_R^{-1}(\mathbb{R}_{\geq 0})$. Then there is a disk $D \subset \cover{F}$, containing $\cover{p}$, so that the Cannon--Thurston map approximation $\Psi_D$ is a component of $\bdy Z$ (with error at most $\exp(-R)$). \end{proposition} \begin{proof}[Proof sketch] We assume we are in the setting of \refdef{UsingTriangulation}. Let $\omega$ be the one-cocycle dual to $F$. Let $W$ be a primitive for $\cover{\omega}$. Consider the two regions of $\cover{M}$ where $W$ is negative or, respectively, non-negative. The common boundary of these is exactly an elevation $\cover{F}$ of the fibre $F$. Let $B^3_R$ be the ball in $\mathbb{H}^3$ of radius $R$ with centre $\cover{p}$. Let $S^2_R$ be the boundary of $B^3_R$. Thus the intersection $\cover{F} \cap S^2_R$ is a collection of curves; these separate the points $\cover{q} \in S^2_R$ where $W$ is negative from those where it is non-negative. Finally, note that $\pi \circ \varphi_R \from \UT{\cover{p}}{\mathbb{H}^3} \to S^2_R$ is the exponential map. We deduce that $\Phi_R = W \circ \pi \circ \varphi_R$ is, up to change of domain, the same as $W | S^2_R$. Recall that $\cover{F}$ is a union of triangles. Assume that one of these contains $\cover{p}$. Let $E$ be the collection of triangles of $\cover{F}$ having at least one edge meeting the ball $B^3_R$. Let $D$ be the connected component of (the union over) $E$ that contains $\cover{p}$. Let $\rho \from \mathbb{H}^3 - \{\cover{p}\} \to \UT{\cover{p}}{\mathbb{H}^3}$ be the inward central projection. For any set $K \subset \mathbb{H}^3 - \{\cover{p}\}$ we call the diameter of $\rho(K)$ the \emph{visual diameter} of $K$. This is measured with respect to the fixed metric on the unit two-sphere $\UT{\cover{p}}{\mathbb{H}^3}$. Suppose that $e$ is a bi-infinite geodesic in $\mathbb{H}^3$. If $e$ lies outside of $B^3_R$ then the visual diameter of $e$ is small; in fact, for large $R$ the visual diameter of $e$ is less than $2\exp(-R)$. Likewise, if $e$ meets $S^2_R$, then for either component $e'$ of $e - B^3_R$ the visual diameter of $e'$ is less than $\exp(-R)$. Now suppose that $T$ is a triangle of $E$. Using the above, we deduce that the visual diameter of each component of $T - B^3_R$ is small. Thus the inward central projections of $\bdy D$ and $D \cap S^2_R$ have Hausdorff distance bounded by a small multiple of $\exp(-R)$. So let $\Psi_D$ be the curve in $\bdy_\infty \mathbb{H}^3$ obtained by projecting $\bdy D$ outwards. By the above, each arc of the resulting polygonal curve has small visual diameter. Also by the above, the Hausdorff distance between the curves $\rho \circ \Psi_D$ and $\rho \circ (\cover{F} \cap S^2_R)$ is small. \end{proof} \begin{remark} Suppose that $F$ is totally geodesic or, more generally, quasi-fuchsian. In this case, the Cannon--Thurston map $\Psi$ is a circle or quasi-circle, respectively. Note however that different elevations now give distinct Cannon--Thurston maps. It is natural to take their union and obtain a circle (or quasi-circle) packing. Now, if $F$ is also Thurston-norm minimising then we still obtain matches. For example in~\cite[Figure~6]{BachmanSchleimerSegerman20}, we see how, for a totally geodesic surface in the Whitehead link complement, the cohomology fractal matches the associated circle packing. On the other hand, if $[F]$ is trivial in $H_2(M, \bdy M)$ then the cohomology fractal $\Phi_R$ is bounded and oscillates as $R$ tends to infinity. \end{remark} \subsection{Lightning curves} \label{Sec:Lightning} \begin{figure}[htb] \centering \subfloat[The lightning curve superimposed on the cohomology fractal.]{ \includegraphics[width = 0.47\textwidth]{Figures/match_up_lightning_1000} \label{Fig:LightningMatchUp} } \subfloat[A black and white version of \reffig{CTPixelColour} highlighting only the brightest pixels.]{ \includegraphics[width = 0.47\textwidth]{Figures/match_up_lightning_threshhold_1000} \label{Fig:LightningThreshhold} } \caption{Matching up the lightning curve, reproduced from \cite[Figure~7]{DicksEtAl06}, with the cohomology fractal for \texttt{m004}. } \end{figure} Suppose that $M$ is a cusped, fibred three-manifold, with fibre $F$. Dicks with various co-authors defines and studies the \emph{lightning curves}~\cite{AlperinEtAl99, DicksPorti02, DicksEtAl02, DicksEtAl06, DicksEtAl10, DicksEtAl12}; these are certain fractal arcs in the plane. In more detail; suppose that $c$ and $d$ are distinct cusps of an elevation $\cover{F}$ of $F$. Let $[c, d]^\rcurvearrowup$ be the arc of $\bdy_\infty \cover{F}$ that is between $c$ and $d$ and anti-clockwise of $c$. The Cannon--Thurston map $\Psi$ sends the arc $[c, d]^\rcurvearrowup$ to a union of disks in $\bdy_\infty \cover{M}$ meeting only along points. The boundary of any one of these disks is a \emph{lightning curve}. Since the lightning curve is defined in terms of the Cannon--Thurston map, it is not too surprising that we can also see something of the lightning curve in the cohomology fractal. In \reffig{LightningMatchUp} we show a segment of the lightning curve for the figure-eight knot complement generated by Cannon and Dicks~\cite[Figure~7]{DicksEtAl06} overlaid on \reffig{CTPixelColour}. The lightning curve seems to follow some of the brightest pixels in the cohomology fractal. \reffig{LightningThreshhold} is a black and white version of \reffig{CTPixelColour}, with a relatively high threshold set for a pixel to be white -- the lightning curve seems to be there, but this is nowhere near as clear as it was for the approximations to the Cannon--Thurston map. We do not fully understand the correspondence here. We note that for clarity, \reffig{LightningMatchUp} shows only one segment of the lightning curve. There is another segment, symmetrical with the shown segment under a 180 degree rotation about the centre of the image. This second segment seems to follow the darkest pixels of the cohomology fractal. \section{Implementation} \label{Sec:Implement} In this section we give an overview of an implementation of cohomology fractals. Our implementation is written in Javascript and GLSL; the code is available at \cite{github_cohomology_fractals}. We are in the process of making cohomology fractals available in SnapPy. The second author has already implemented ray-tracing in hyperbolic three-manifolds~\cite{snappy}. We now follow \refdef{UsingTriangulation}. Suppose that $M$ is a connected, oriented, finite volume hyperbolic three-manifold. Let $\mathcal{T}$ be a material or ideal triangulation of $M$. We are given a weighting $\omega \from \mathcal{T}^{(2)} \to \mathbb{R}$ for the faces of the two-skeleton. We represent the triangulation $\mathcal{T}$ as a collection $\{t_i\}$ of model hyperbolic tetrahedra. Each tetrahedron $t_i$ has four faces $f^i_m$ lying in four geodesic planes $P^i_m$ in the hyperboloid model of $\mathbb{H}^3$. Suppose that $t_j \in \mathcal{T}$ is another model tetrahedron, with faces $f^j_n$. If the face $f^i_m$ is glued to $f^j_n$, then we have isometries $g^i_m$ and $g^j_n$ realising the gluings. Note that $g^i_m$ and $g^j_n$ are inverses. We are given a camera location $p$ in $M$; this is realised as a point (again called $p$) in some tetrahedron $t_i$. \begin{remark} The reader familiar with computer graphics will note that we also require a frame at the camera location. To simplify the exposition, we will mostly suppress this detail. \end{remark} We are also given a radius $R$ as well as a maximum allowed step count $S$. \subsection{Ray-tracing} For each pixel of the screen, we generate a corresponding unit tangent vector $u$ in the tangent space to the current tetrahedron $t_i$. We then \emph{ray-trace} through $\mathcal{T}$. That is, we travel along the geodesic starting at $p$, in the direction $u$, for distance $R$, taking at most $S$ steps. \reffig{DesiredRayPath} shows a toy example, where we replace the three-dimensional hyperbolic triangulation $\mathcal{T}$ of $M$ with a two-dimensional euclidean triangulation of the two-torus. \begin{figure}[htb] \centering \subfloat[\emph{The desired ray path, starting from the pair $(p, u)$ of length $R$.}]{ \label{Fig:DesiredRayPath} \labellist \small\hair 2pt \pinlabel $p$ [r] at 25 14 \pinlabel $u$ [b] at 32.5 18 \endlabellist \includegraphics[height=1.67in]{Figures/ray_trace_schematic_developed} } \subfloat[\emph{The implementation of the ray path. The iterations of the loop are labelled with integers. }]{ \label{Fig:ImplementationRayPath} \labellist \small\hair 2pt \pinlabel 1 at 32 20 \pinlabel 3 at 20 27.5 \pinlabel 5 at 23.5 34 \pinlabel 9 at 13.5 10.5 \pinlabel 7 at 43 5.5 \pinlabel \rotatebox{300}{6} at 90 30 \pinlabel \rotatebox{300}{4} at 94 20 \pinlabel \rotatebox{300}{2} at 84 14 \pinlabel \rotatebox{300}{8} at 72 5 \pinlabel \rotatebox{300}{10} at 82 5 \endlabellist \includegraphics[height=1.67in]{Figures/ray_trace_schematic_implementation} } \caption{A toy example of developing a ray through a tiling of a euclidean torus. Note that the geodesic segments passing through a tile are parallel; this is only because the geometry is euclidean. In a hyperbolic tiling the segments are much less ordered. } \end{figure} It is perhaps most natural to think of ray-tracing as occurring in $\cover{M}$, the universal cover of the manifold, as shown in \reffig{DesiredRayPath}. However, the na\"ive floating-point implementation in the hyperboloid model quickly loses precision. We instead ray-trace in the manifold, as illustrated in \reffig{ImplementationRayPath}. Thus, all points we calculate lie within our fixed collection of ideal hyperbolic tetrahedra $\{t_i\}$. For each pixel, we do the following. \begin{enumerate} \item The following initial data are given: an index $i$ of a tetrahedron, a point $p$ in $t_i$, and a tangent vector $u$ at $p$. Initialise the following variables. \begin{itemize} \item The total distance travelled: $r \gets 0$. \item The number of steps taken: $s \gets 0$. \item The current tetrahedron index: $j \gets i$. \item The current position: $q \gets p$. \item The current tangent vector: $v \gets u$. \end{itemize} \item \label{Itm:LoopStart} Let $\gamma$ be the geodesic ray starting at $p$ in the direction of $u$. Find the index $n$ so that $\gamma$ exits $t_j$ through the face $f^j_n$. Let $t_k$ be the other tetrahedron glued to face $f^j_n$. \item \label{Itm:LoopHitFace} Calculate the position $q'$ and tangent vector $v'$ where $\gamma$ intersects $f^j_n$. Let $r'$ be the distance from $q$ to $q'$. Set $r \gets r + r'$ and set $s \gets s + 1$. \item If $r > R$ or $s > S$ then stop. \item Set $j \gets k$, set $q \gets g^j_n(q')$, and set $v \gets Dg^j_n(v')$. \item Go to step \refitm{LoopStart}. \end{enumerate} This implements the ray-tracing part of the algorithm. In our toy example, this is shown in \reffig{ImplementationRayPath}. \subsection{Integrating} To determine the colour of the pixel, we also track the total signed weight we accumulate along the ray. For this, we add the following steps to the loop above. \begin{itemize} \item[(1b)] An initial weight $w_0$ is given. Initialise the following. \begin{itemize} \item[$\bullet$] The current weight: $w_c \gets w_0$. \end{itemize} \item[(5b)] Let $f$ be the face between $t$ and $t'$, co-oriented towards $t'$. Set $w_c \gets w_c + \omega(f)$. \end{itemize} At the end of the loop, the value of $w_c$ gives the brightness of the current pixel. (In fact, we apply a function very similar to the arctan function to remap the possible values of $w_c$ to a bounded interval. We then apply a gradient that passes through a number of different colours. This helps the eye see finer differences between values than a direct map to brightness.) \subsection{Moving the camera} In our applications, we enable the user to fly through the manifold $M$. Depending on the keys pressed by the user at each time step, we apply an isometry $g$ to $p$. We also track an orthonormal frame for the user; this determines how tangent vectors correspond to pixels of the screen. We also apply the isometry $g$ to this frame. When the user flies out of a face $f$ of the tetrahedron they are in, we apply the corresponding isometry $g^i_k$ to the position $p$ and the user's frame. We also add $w(f)$ to the initial weight $w_0$. Without this last step, the overall brightness of the image would change abruptly as the user flies through a face with non-zero weight. \begin{remark} \label{Rem:Basepoint} With this last modification, the cohomology fractal depends on a choice of basepoint $b \in \cover{M}$. The point $p \in M$ must now also be replaced by $p \in \cover{M}$ (abusing notation, we use the same symbol for both points). We add $b$ to the notation, and now write the cohomology fractal as \[ \Phi_R^{\omega, b, p} \from \UT{p}{\cover{M}} \to \mathbb{R} \qedhere \] \end{remark} \begin{remark} \label{Rem:DependenceOnb} The dependence of the cohomology fractal on $b$ is minor: If we change $b$ to $b'$, then the value of $\Phi_R^{\omega, b, p}(v)$ changes by the weight we pick up along any path from $b'$ to $b$. \end{remark} \subsection{Material, ideal, and hyperideal views} \label{Sec:Views} The above discussion describes the \emph{material view}; the geodesic rays emanate radially from $p$. To render an image, we place a rectangle in the tangent space at $p$. For each pixel of the screen, we take the tangent vector $u$ to be the corresponding point of the rectangle. See \reffig{ViewsMaterial}. \begin{definition} The \emph{field of view} of a material image is the angle between the tangent vectors pointing at the midpoints of the left and right sides of the image. \end{definition} \begin{remark} The material view suffers from perspective distortion. This is most noticeable towards the edges of the image, and is worse when the field of view is large. \end{remark} To generalise the material view to the ideal and hyperideal, we introduce the following terminology. We say that a subset $D \subset \UT{}{\cover{M}}$ is a \emph{view} if it is one of the following. \begin{enumerate} \item In the \emph{material view}, $D$ is a fibre $\UT{p}{\cover{M}}\homeo S^2$. \item In the \emph{ideal view}, we take $D$ to be the collection of outward normals to a horosphere $H$. That is, the vectors point away from $\bdy_\infty H$. To render an image we place a rectangle in $D$. For each pixel of the screen we set the initial vector $v$ to be the corresponding point of the rectangle. The starting point is then $\pi(v) \in \cover{M}$, the basepoint of $v$. Finally, we set $w_0$ to be the total weight accumulated, along the arc from $b$ to $\pi(v)$, as we pass through faces of the triangulation. See \reffig{ViewsIdeal}. \item In the \emph{hyperideal view}, we take $D$ to be the collection of normals to a transversely oriented geodesic plane $P$. We draw $P$ on the euclidean rectangle of the screen using the Klein model. The algorithm is otherwise identical to the ideal view case. See \reffig{ViewsHyperideal}. \end{enumerate} \begin{remark} The ideal view in hyperbolic geometry is the analogue of an orthogonal view in euclidean geometry. In both cases this is the limit of backing the camera away from the subject while simultaneously zooming in. \end{remark} \begin{remark} The hyperideal view suffers from an ``inverse'' form of perspective distortion. Towards the edges of the image, round circles look like ellipses, with the minor axis along the radial direction. \end{remark} \begin{definition} \label{Def:View} Let $D \subset \UT{}{\cover{M}}$ be a view, as discussed above. In the notation for the cohomology fractal, we replace $p$ by $D$: \[ \Phi_R^{\omega, b, D} \from D \to \mathbb{R} \qedhere \] \end{definition} \begin{figure}[htb] \centering \subfloat[Material.]{% \includegraphics[width = 0.32\textwidth]{Figures/comparison_material}% \label{Fig:ViewsMaterial} } \subfloat[Ideal.]{% \includegraphics[width = 0.32\textwidth]{Figures/comparison_ideal}% \label{Fig:ViewsIdeal} } \subfloat[Hyperideal.]{% \includegraphics[width = 0.32\textwidth]{Figures/comparison_hyperideal}% \label{Fig:ViewsHyperideal}% } \caption{Comparison between different views of the cohomology fractal for \texttt{m004}. } \end{figure} \subsection{Edges} We give the user the option to see the edges of the triangulation. The user selects an edge thickness $\varepsilon > 0$. The web application implements this in a lightweight fashion: In step \refitm{LoopHitFace}, if the distance from the point $q'$ to one of the three edges of the face we have intersected is less than $\varepsilon$, then we exit the loop early. Depending on user choice, the pixel is either coloured by the weight $w_c$ or by the distance $d$. See \reffig{Edges}. In SnapPy, we compute the intersection of the ray with a cylinder about the edge in addition to the intersection with the faces. \begin{figure}[htb] \centering \subfloat[Coloured by cohomology fractal.]{ \includegraphics[width = 0.47\textwidth]{Figures/edges_cohom_fractal_ideal} } \subfloat[Coloured by distance.]{ \includegraphics[width = 0.47\textwidth]{Figures/edges_distance_ideal} } \caption{Edges of the ideal triangulation of \texttt{m004}, as seen in the material view.} \label{Fig:Edges} \end{figure} \subsection{Elevations} We also give the user the option to see several elevations of the surface $F$. The user selects a weight $w_{\max} > 0$. In step (5b), if $w_0 < 0$ but $w_c>0$, then we have crossed the elevation at weight zero. In this case we exit the loop, and colour the pixel by the distance $d$. Similarly, if $w_0 > w_{\max}$ but $w_c < w_{\max}$, then we have crossed the elevation at weight $w_{\max}$, and again we stop and colour by distance. Finally, if $0<w_0<w_{\max}$, then we stop if $w_c$ has changed from $w_0$. \reffig{Elevation} shows a single elevation. \subsection{Triangulations, geometry, and cocycles} We obtain our triangulations and their hyperbolic shapes from the SnapPy census. We put some effort into choosing good representative cocycles; the choice here makes very little difference to the appearance of the cohomology fractal, but it makes a large difference to the appearance of the elevations. That is, a poor choice of cocycle gives a ``noisy'' elevation. For example, adding the boundary of a tetrahedron to the Poincar\'e dual surface may perform a one-three move to its triangulation. This adds unnecessary ``spikes'' to the elevations. When our manifold has Betti number one, there is only one cohomology class of interest. Here we searched for taut ideal structures dual to this class~\cite{Lackenby00}. When the SnapPy triangulation did not admit such a taut structure, we randomly searched for one that did. A taut structure gives a Poincar\'e dual surface with the minimum possible Euler characteristic. When the Betti number is larger than one, we used tnorm~\cite{tnorm20} to find initial simplicial representatives of vertices of the Thurston norm ball~\cite{Thurston86} in $H_2(M,\bdy M)$. We then greedily performed Pachner moves to reduce the complexity of the cocycles. We often, but not always, realised the minimum possible Euler characteristic. \subsection{Discussion} Any visualisation of a hyperbolic tiling suffers from the mismatch between the hyperbolic metric of the tiling and the euclidean metric of the image. The tools for generating more of the tiling involve applying hyperbolic isometries. The tiles thus shrink exponentially in size while growing exponentially in number. This makes it difficult for the tiles to cleanly approach $\bdy_\infty \mathbb{H}^2$ or $\bdy_\infty \mathbb{H}^3$. Approaching a ``parabolic'' point at infinity is even more difficult. In the vector graphics approach, one must be careful to avoid wasting time generating huge numbers of invisible objects: tiles may be too small or their aspect ratios too large. The ray-tracing approach (and any similar raster graphics approach) deals with this mismatch directly. Here we start with the pixel that is to be coloured and then generate only the hyperbolic geometry needed to determine its colour. A disadvantage of the ray-tracing approach is that we generate the hyperbolic geometry necessary for each pixel independently, meaning that much work is duplicated. However, the massive parallelism in modern graphics processing units mitigates, and is in fact designed to deal with, this kind of issue. It often turns out to be faster to duplicate work in many parallel processes rather than compute once then transmit the result to all processes requiring it. \section{Incomplete structures and closed manifolds} \label{Sec:Cone} Suppose that $M$ is a cusped hyperbolic manifold. Recall that we generate cohomology fractals for $M$ by using an ideal triangulation $\mathcal{T}$. Associated to $\mathcal{T}$ there is the \emph{shape variety}; that is we impose the gluing equations outlined in \refsec{Geometry}, omitting the peripheral ones. This gives us a space of deformations of the complete hyperbolic structure to incomplete hyperbolic structures; see~\cite[Section~4.4]{ThurstonNotes} and~\cite[Section~6.2]{PurcellKnotTheory}. If we deform correctly, we reach an incomplete structure whose completion has the structure of a hyperbolic manifold. The result is a \emph{hyperbolic Dehn filling} of the original cusped manifold. \subsection{Incomplete structures} Suppose that $(M, \mathcal{T})$ is an ideally triangulated manifold. Let $Z^s$ be a path in the shape variety, where $Z^\infty$ is the complete structure and the completion of $Z^1$ is a closed hyperbolic three-manifold obtained by Dehn filling $M$. Between the two endpoints, we have \emph{incomplete structures} $M_s$ on the manifold $M$. In an incomplete geometry, there are geodesic segments that cannot be extended indefinitely. Suppose that, as in our algorithm, we only consider geodesic segments emanating from $p$ of length at most $R$. The endpoints of the rays that do not extend to distance $R$ form the \emph{incompleteness locus} $\Sigma_s$ in the ball $B^3_R \subset \mathbb{H}^3$. It follows from work of Thurston that $\Sigma_s$ is a discrete collection of geodesic segments, for generic values of $s$~\cite{Thurston82}. Suppose that $\omega \from \mathcal{T}^{(2)} \to \mathbb{R}$ is the given weight function dual to a properly embedded surface $F$ in $M$. We assume that the boundary of $F$ (if any) gives loops in the filled manifold that, there, bound disks. Thus $F$ also gives a cohomology fractal in the filled manifold. \begin{remark} Note that there is no canonical way of transferring a base point $b$ or view $D$ between two different geometric structures $M_s$ and $M_{s'}$. However, we can choose $b$ and $D$ for each $M_s$ in a way that gives us continuously varying pictures. We do not dwell on the details here. \end{remark} \begin{figure}[htb!] \centering \subfloat[$s=10$]{ \includegraphics[width=0.31\textwidth]{Figures/FiguresMatthiasVer2/m122_4t_-t/cone_10} } \subfloat[$s=5$]{ \includegraphics[width=0.31\textwidth]{Figures/FiguresMatthiasVer2/m122_4t_-t/cone_5} } \subfloat[$s=4$]{ \includegraphics[width=0.31\textwidth]{Figures/FiguresMatthiasVer2/m122_4t_-t/cone_4} } \\ \subfloat[$s=3$]{ \includegraphics[width=0.31\textwidth]{Figures/FiguresMatthiasVer2/m122_4t_-t/cone_3} } \subfloat[$s=2$]{ \includegraphics[width=0.31\textwidth]{Figures/FiguresMatthiasVer2/m122_4t_-t/cone_2} } \subfloat[$s=1.8$]{ \includegraphics[width=0.31\textwidth]{Figures/FiguresMatthiasVer2/m122_4t_-t/cone_1p8} } \\ \subfloat[$s=1.6$]{ \includegraphics[width=0.31\textwidth]{Figures/FiguresMatthiasVer2/m122_4t_-t/cone_1p6} } \subfloat[$s=1.4$]{ \includegraphics[width=0.31\textwidth]{Figures/FiguresMatthiasVer2/m122_4t_-t/cone_1p4} } \subfloat[$s=1.2$]{ \includegraphics[width=0.31\textwidth]{Figures/FiguresMatthiasVer2/m122_4t_-t/cone_1p2} } \caption{Cohomology fractals for \texttt{m122($4s, -s$)} as $s$ varies.} \label{Fig:Bending} \end{figure} \reffig{Bending} shows cohomology fractals for various $M_s$. We see a kind of branch cut in the background to either side of the incompleteness locus $\Sigma_s$. As we vary $s$, the background appears to bend along the geodesic. Other paths in the shape variety will give shearing as well as (or instead of) bending. When we reach a Dehn filling, the two sides again match, and we see the structure of the closed filled manifold. See \reffig{m122_4_-1}. (The two sides can also match before we reach the Dehn filling due to symmetries of the cusped manifold lining up with the cone structure.) \begin{figure}[htb] \centering \includegraphics[width = 0.65\textwidth]{Figures/FiguresMatthiasVer2/m122_4_-1/without_evil} \caption{Cohomology fractal for the Dehn filling \texttt{m122(4,-1)}. This gives a final image for \reffig{Bending}, with $s = 1$.} \label{Fig:m122_4_-1} \end{figure} \subsection{Numerical instability near the incompleteness locus} \label{Sec:NumericalInstability} Our algorithm, given in \refsec{Implement}, does not require completeness. However, a ray from $p$ to $\Sigma_s$ necessarily meets infinitely many tetrahedra. This is because near $\Sigma_s$ we are far from the thick part of any tetrahedron, and the thin parts of the tetrahedra are almost ``parallel'' to $\Sigma_s$. Thus the innermost loop of the algorithm will always halt by reaching the maximum step count; it follows that we cannot ``see through'' a neighbourhood of $\Sigma_s$. \reffig{m122_4_-1_with_evil} shows the cohomology fractal drawn with a small maximum step count, making such a neighbourhood visible. \begin{figure}[htb] \centering \includegraphics[width = 0.65\textwidth]{Figures/FiguresMatthiasVer2/m122_4_-1/with_evil} \caption{Cohomology fractal for the Dehn filling \texttt{m122(4,-1)} drawn with an incomplete structure on an ideal triangulation. Here the maximum number of steps $S$ is 55. Compare with \reffig{m122_4_-1}.} \label{Fig:m122_4_-1_with_evil} \end{figure} Increasing the maximum step count shrinks the opaque neighbourhood of $\Sigma_s$. However, as a ray approaches $\Sigma_s$, its segments within the model tetrahedra tend to their ideal vertices. Thus the coordinates blow up; this appears to lead to numerically unstable behaviour. See \reffig{EvilCloseUp}. In the next section we describe a method to eliminate these numerical defects; we use this to produce \reffig{WithoutEvilCloseUp}. \begin{figure}[htb] \centering \subfloat[Numerical instability near the incompleteness locus.]{ \includegraphics[width = 0.47\textwidth]{Figures/FiguresMatthiasVer2/m122_4_-1/with_evil_detail} \label{Fig:EvilCloseUp} } \subfloat[The same view with a material triangulation implementation.]{ \includegraphics[width = 0.47\textwidth]{Figures/FiguresMatthiasVer2/m122_4_-1/without_evil_detail} \label{Fig:WithoutEvilCloseUp} } \caption{A view of the cohomology fractal for the manifold \texttt{m122(4,-1)} near the incompleteness locus. On the left we have taken the maximum number of steps $S$ sufficiently large to ensure that all rays reach distance $R$.} \end{figure} Note that numerical instability caused by rays approaching the ideal vertices also occurs for the complete structure on a cusped manifold. It is less noticeable in this case however, because these errors occur in a small part of the visual sphere for typical positions. \subsection{Material triangulations} \label{Sec:MaterialTriangulations} In order to remove instability around the incompleteness locus, we remove it. That is, we abandon (spun) ideal triangulations in favour of material triangulations. There is no change to the algorithm in \refsec{Implement}; we only alter the input data (the planes $P^i_k$ and face-pairing matrices $g^i_k$): Given the edge lengths (see \refsec{MaterialGeometry}) for a material triangulation, \cite[Lemma~3.4]{matthiasVerifyingFinite} assigns hyperbolic isometries to the edges of a doubly truncated simplex (also known as permutahedron). These can be used to switch a tetrahedron between different standard positions (as defined in \cite[Definition~3.2]{matthiasVerifyingFinite}) where one of its faces is in the $\mathbb{H}^2\subset\mathbb{H}^3$ plane. We assume that every tetrahedron is in $(0,1,2,3)$--standard position. Given a face-pairing, we apply the respective isometries to each of the two tetrahedra such that the faces in question line up in the $\mathbb{H}^2\subset\mathbb{H}^3$ plane. The face-pairing matrix $g^i_k$ is now given by composing the inverse of the first isometry with the second isometry. For example, let face 3 of one tetrahedron be paired with face 2 of another tetrahedron via the permutation $(0,1,2,3)\mapsto(0,1,3,2)$. To line up the faces, we need to bring the second tetrahedron from the default $(0,1,2,3)$--standard position into $(0,1,3,2)$--standard position by applying $\gamma_{012}$ from \cite[Lemma~3.4]{matthiasVerifyingFinite} which will thus be the face-pairing matrix, see \cite[Figure~4]{matthiasVerifyingFinite}. It is left to compute the planes $P^i_k$. Note that $P^i_3$ (for each $i$) is the canonical copy of $\mathbb{H}^2 \subset \mathbb{H}^3$. All other $P^i_k$ can be obtained by applying the isometries from \cite[Lemma~3.4]{matthiasVerifyingFinite} again. \subsection{Cannon--Thurston maps in the closed case} \begin{figure}[htb] \centering \subfloat[McMullen's illustration~\cite{McMullen19}. See also~\cite{McMullenWeb}.]{ \includegraphics[width = 0.47\textwidth]{Figures/McMullen_fiber_refl_1000} \label{Fig:McMullen} } \subfloat[Cohomology fractal - hyperideal view.]{ \includegraphics[width = 0.47\textwidth]{Figures/mcmullenFractal} \label{Fig:OrbifoldCohomFractal} } \caption{Views of \texttt{m004(0,2)}.} \label{Fig:Orbifold} \end{figure} Cannon and Thurston's original proof was in the closed case. Thurston's original images and all subsequent renderings, with one notable exception, are in the cusped case. With some minor modifications, \refprop{LightDark} applies in the closed case; thus the cohomology fractals again approximate Cannon--Thurston maps. We are aware of only one previous example in the closed case, due to McMullen~\cite{McMullenWeb}. In \reffig{Orbifold} we give a rasterisation of his original vector graphics image~\cite{McMullen19}, and our version of the same view. The filling \texttt{m004(0,2)} of the figure-eight knot complement has an incomplete hyperbolic metric. The completion is a hyperbolic orbifold $\mathcal{O}$ with angle $\pi$ about the orbifold locus; the universal cover is $\mathbb{H}^3$. Since the filling is a multiple of the longitude, the orbifold $\mathcal{O}$ is again fibred. An elevation of this fibre to $\mathbb{H}^3$ gives a Cannon--Thurston map. Our image, \reffig{OrbifoldCohomFractal} is the cohomology fractal for the fibre in $\mathcal{O}$, in the hyperideal view. This is implemented using a material triangulation of an eight-fold cover $M$. Since $M$ with its fibre, is commensurable with $\mathcal{O}$ with its fibre, we obtain the same image. McMullen's image, reproduced in \reffig{McMullen} was generated using his program \texttt{lim}~\cite{lim}. Briefly, let $\mathcal{O}^\infty$ be the infinite cyclic cover of $\mathcal{O}$. McMullen produces a sequence $\mathcal{O}^n$ of quasi-fuchsian orbifolds that converge in the geometric topology to $\mathcal{O}^\infty$. In each of these the convex core boundary is a pleated surface. The supporting planes of this pleated surfaces give round circles in $\bdy \mathbb{H}^3$. His image then is obtained by taking $n$ fairly large, passing to the universal cover of $\mathcal{O}^n$, and drawing the boundaries of many supporting planes~\cite{McMullen19}. \subsection{Accumulation of floating point errors} \label{Sec:Accumulate} Our implementation uses single-precision floating point numbers. As we saw in \refsec{NumericalInstability}, this can cause problems when rays approach the vertices of ideal tetrahedra. However, floating point errors can accumulate for large values of $R$ whether or not rays approach the vertices. This can therefore also affect material triangulations. With these problems in mind, we cannot claim that our images are rigorously correct. However, for small values of $R$ we can be confident that our images are accurate. For very small values the endpoints of our rays all sit within the same tetrahedron, and so all pixels are the same colour. As we increase $R$ (as in \reffig{VisSphereRadii}), we see regions of constant colour, separated by arcs of circles. This is provably correct: (horo-)spheres meet the totally geodesic faces of tetrahedra in circles. If we zoom in whilst increasing $R$, eventually floating point errors become visible. \reffig{NumNoise} shows the results of an experiment to determine when this happens, for a material triangulation. At around $R = 11$, the circular arcs separating regions of the same colour become stippled. At around $R = 13$, the regions are no longer distinct. \begin{remark} \label{Rem:Quasigeodesic} Perhaps surprisingly, this accumulation of error does not mean that our pictures are inaccurate. Suppose that the side lengths of our pixels are on a somewhat larger scale than the precision of our floating point numbers. For each pixel, our implementation produces a piecewise geodesic, starting in the direction through the centre of the pixel, but with small angle defect at each vertex. Due to the nature of hyperbolic geometry, this piecewise geodesic cannot curve away from the true geodesic fast enough to leave the visual cone on the desired pixel. Thus, as long as the pixel size is not too small, each pixel is coloured according to some sample within that pixel. \end{remark} \begin{figure}[htb] \centering \subfloat[$R=10.5,$ field of view $\sim 0.015^\circ$]{ \includegraphics[width=0.45\textwidth]{Figures/FiguresMatthiasVer2/m122_4_-1/dist_10_5.png} } \thinspace \subfloat[$R=11.5,$ field of view $\sim 0.005^\circ$]{ \includegraphics[width=0.45\textwidth]{Figures/FiguresMatthiasVer2/m122_4_-1/dist_11_5.png} } \subfloat[$R=12.5,$ field of view $\sim 0.002^\circ$]{ \includegraphics[width=0.45\textwidth]{Figures/FiguresMatthiasVer2/m122_4_-1/dist_12_5.png} } \thinspace \subfloat[$R=13.5,$ field of view $\sim 0.0007^\circ$]{ \includegraphics[width=0.45\textwidth]{Figures/FiguresMatthiasVer2/m122_4_-1/dist_13_5.png} } \caption{We zoom into the cohomology fractal for \texttt{m122(4,-1)} while increasing $R$. The field of view of the image is proportional to $e^{-R}$. Noise due to rounding errors becomes visible at $R\sim 10.5$ and completely dominates the picture when $R\sim13.5$. } \label{Fig:NumNoise} \end{figure} \section{Experiments} \label{Sec:Experiments} The sequence of images in \reffig{VisSphereRadii} suggests that some form of fractal object is coming into focus. When $R$ is small, the function $\Phi_R = \Phi^{F,b,D}_{R}$ is constant on large regions of $D$. As $R$ increases, these regions subdivide, producing intricate structures. As we have defined it so far, the cohomology fractal depends on $R$. A natural question is whether or not there is a limiting object that does not depend on $R$. In this section we describe a sequence of experiments we undertook to explore this question. Inspired by these, in Sections~\ref{Sec:CLT} and~\ref{Sec:Pixel} we provide mathematical explanations of our observations. \subsection{These pictures do not exist} A na\"ive guess might be that the cohomology fractal converges to a function as $R$ tends to infinity. However, consider a ray following a closed geodesic $\gamma$ in $M$ that has positive algebraic intersection with the surface $F$. Choosing $D$ so that it contains a tangent direction $v$ along $\gamma$, we see that $\Phi^{F}_{R}(v)$ diverges to infinity as $R$ tends to infinity. The issue is not restricted to the measure zero set of rays along closed geodesics. Suppose that $v$ is a generic vector in a material view $D$. Recall that the geodesic flow is ergodic~\cite[Hauptsatz~7.1]{Hopf39}. Thus the ray starting from $v$ hits $F$ infinitely many times. So $\Phi^{F}_{R}(v)$ again diverges. Thus we have the following theorem. \begin{theorem} \label{Thm:NoPicture} Suppose that $M$ is a finite volume, oriented hyperbolic three-manifold. Suppose that $p$ is any point of $M$. Suppose that $F$ is a compact, transversely oriented surface. Then the limit \[ \lim_{R \to \infty} \Phi_R(v) \] does not exist for almost all $v \in \UT{p}{M}$. \qed \end{theorem} \begin{remark} To generalise \refthm{NoPicture} from finite volume to infinite volume manifolds, we must replace Hopf's ergodicity theorem by some other dynamical property. For example, Rees~\cite[Theorem~4.7]{Rees81} proves the ergodicity of the geodesic flow on the infinite cyclic cover of a hyperbolic surface bundle. This is generalised to the bounded geometry case by Bishop and Jones~\cite[Corollary~1.4]{BishopJones97}. Both of these works rely in a crucial fashion on Sullivan's equivalent criteria for ergodicity~\cite[page~172]{Sullivan79}. \end{remark} One might hope that as $R$ tends to infinity, nearby points diverge in similar ways. If so, we might be able to rescale and have, say, $\Phi_{R}/R$ or $\Phi_{R}/\sqrt{R}$ converge. However, increasing $R$ in our implementation produces the sequence of images shown in \reffig{VisSphereRadii2}. We see that, as we increase $R$, the images become noisy as neighbouring pixels appear to decorrelate. Eventually the fractal structure is washed away. Dividing the cohomology fractal by, say, some power of $R$ only changes the contrast. Depending on this power, the limit is either almost always zero or does not exist. \begin{figure}[htb!] \centering \subfloat[$R=e^{2.5}$]{ \includegraphics[width=0.225\textwidth]{Figures/m004_exp2p5} } \subfloat[$R=e^{3}$]{ \includegraphics[width=0.225\textwidth]{Figures/m004_exp3} } \subfloat[$R=e^{4}$]{ \includegraphics[width=0.225\textwidth]{Figures/m004_exp4} } \subfloat[$R=e^{5}$]{ \includegraphics[width=0.225\textwidth]{Figures/m004_exp5} } \caption{Cohomology fractals for \texttt{m004}, with larger values of $R$. Each image here and in \reffig{VisSphereRadii} has $1000\times1000$ pixels. } \label{Fig:VisSphereRadii2} \end{figure} \reffig{VisSphereRadii2} also demonstrates that \refrem{Quasigeodesic}, while valid, is misleading; it is true that for large $R$, every ray ends up somewhere within its pixel, but the colour one obtains is random noise. This noise is due to undersampling. In our images each pixel $U$ is coloured using a single ray passing (almost, as we saw in \refsec{Accumulate}) through its centre. When $R$ is small relative to the side length of $U$ the function $\Phi_{R}|U$ is generally constant; thus any sample is a good representative. As $R$ becomes larger the function $\Phi_{R}|U$ varies more and more wildly; thus a single sample does not suffice. \subsection{Take a step back and look from afar} Let $D$ be an ideal view in the sense of \refsec{Views}. We identify $\pi(D)$, isometrically, with the euclidean plane $\mathbb{C}$. Using this identification, we may refer to the vectors of $D$ as $z_D$ for $z \in \mathbb{C}$. Let $E$ be the ideal view obtained from $D$ by flowing outwards by a distance $d$. Thus, $\varphi_d(D) = E$. We similarly identify $\pi(E)$ with the euclidean plane, in such a way that for each $z \in \mathbb{C}$, we have $\varphi_d(z_D) = (e^d z)_E$. We may now state the following. \begin{figure}[htbp] \labellist \small\hair 2pt \pinlabel {$\bdy \mathbb{H}^3$} [r] at 0 0 \pinlabel {$D$} [r] at 34 218 \pinlabel {$E$} [r] at 173 74 \pinlabel {$R$} [l] at 340 50 \pinlabel {$R+d$} [l] at 493 122.5 \endlabellist \includegraphics[width = 0.6\textwidth]{Figures/two_ideal_views} \caption{Side view of ``screens'' (in red) for two ideal views, drawn in the upper half space model. The outward pointing normals to each horosphere point down in the figure. } \label{Fig:TwoIdealViews} \end{figure} \begin{lemma} \label{Lem:MovingIsScaling} Suppose that $D$ is an ideal view and $E = \varphi_d(D)$. Then the cohomology fractal based at $b$ satisfies \[ \Phi_{R+d}^{\omega, b,D}\left(z_D\right) = \Phi_{R}^{\omega, b,E}\left((e^d z)_E\right) \] \end{lemma} \begin{proof} Consider \reffig{TwoIdealViews}. \end{proof} Said another way, if we fly backwards a distance $d$ and replace $R$ with $R+d$, we see the exact same image, scaled down by a factor of $e^d$. As a consequence, in the ideal view we have the following. \begin{remark} Each small part of a cohomology fractal with large $R$ is the same as the cohomology fractal for a smaller $R$ with a different view. \end{remark} \begin{remark} \label{Rem:Embiggen} Since we know that we can make non-noisy images for small enough values of $R$, we can therefore make a non-noisy image of a cohomology fractal for any value of $R$, as long as we are willing to use a screen with high enough resolution. \end{remark} The natural question then is how the \emph{perceived} image changes as we simultaneously increase the resolution and increase $R$. This convergence question is different from the convergence of the cohomology fractal to a function as in \refthm{NoPicture}: when we look at a very large screen from far away, our eyes average the colours of nearby pixels. Thus, we move away from thinking of the limit as a function evaluated at points, towards thinking of it as a measure evaluated by integrating over a region. As we will see later, in fact the correct limiting object is a distribution. \subsection{Supersampling} \begin{figure}[htbp] \begin{tabular}{c m{0.275\textwidth} m{0.275\textwidth} m{0.275\textwidth}} & \multicolumn{1}{c}{$1\times 1$} & \multicolumn{1}{c}{$2\times2$} & \multicolumn{1}{c}{$128\times 128$} \\ 4 & \includegraphics[width=0.29\textwidth]{Figures/FiguresMatthiasVer2/m122_4_-1_sub4/b_subsampling_0400_0001.png} & \includegraphics[width=0.29\textwidth]{Figures/FiguresMatthiasVer2/m122_4_-1_sub4/b_subsampling_0400_0002.png} & \includegraphics[width=0.29\textwidth]{Figures/FiguresMatthiasVer2/m122_4_-1_sub4/b_subsampling_0400_0128.png} \\ 6 & \includegraphics[width=0.29\textwidth]{Figures/FiguresMatthiasVer2/m122_4_-1_sub4/b_subsampling_0600_0001.png} & \includegraphics[width=0.29\textwidth]{Figures/FiguresMatthiasVer2/m122_4_-1_sub4/b_subsampling_0600_0002.png} & \includegraphics[width=0.29\textwidth]{Figures/FiguresMatthiasVer2/m122_4_-1_sub4/b_subsampling_0600_0128.png} \\ 8 & \includegraphics[width=0.29\textwidth]{Figures/FiguresMatthiasVer2/m122_4_-1_sub4/b_subsampling_0800_0001.png} & \includegraphics[width=0.29\textwidth]{Figures/FiguresMatthiasVer2/m122_4_-1_sub4/b_subsampling_0800_0002.png} & \includegraphics[width=0.29\textwidth]{Figures/FiguresMatthiasVer2/m122_4_-1_sub4/b_subsampling_0800_0128.png} \\ 10 & \includegraphics[width=0.29\textwidth]{Figures/FiguresMatthiasVer2/m122_4_-1_sub4/b_subsampling_1000_0001.png} & \includegraphics[width=0.29\textwidth]{Figures/FiguresMatthiasVer2/m122_4_-1_sub4/b_subsampling_1000_0002.png} & \includegraphics[width=0.29\textwidth]{Figures/FiguresMatthiasVer2/m122_4_-1_sub4/b_subsampling_1000_0128.png} \\ 12 & \includegraphics[width=0.29\textwidth]{Figures/FiguresMatthiasVer2/m122_4_-1_sub4/b_subsampling_1200_0001.png} & \includegraphics[width=0.29\textwidth]{Figures/FiguresMatthiasVer2/m122_4_-1_sub4/b_subsampling_1200_0002.png} & \includegraphics[width=0.29\textwidth]{Figures/FiguresMatthiasVer2/m122_4_-1_sub4/b_subsampling_1200_0128.png} \\ \end{tabular} \caption{\texttt{m122(4,-1)}. Field of view: $12.8^\circ$. $128\times 128$ pixels. For each image, the visual radius $R$ is given at the start of its row, while the number of samples per pixel is given at the top of its column.} \label{Fig:RagainstNumSamples} \end{figure} To investigate this without requiring ever larger screens to view the results, we sample the cohomology fractal at many vectors in a grid within each pixel and average the results to give a colour for the pixel. That is, we employ supersampling. See \reffig{RagainstNumSamples}. Here we draw cohomology fractals with $R$ ranging from $4$ to $12$, and with either $1$, $2^2$, or $128^2$ subsamples for each pixel. Each image has resolution $128 \times 128$. \begin{remark*} Note that some pdf readers do not show individual pixels with sharp boundaries: they automatically blur the image when zooming in. To combat this blurring and see the pixels clearly, we have scaled each image by a factor of three, so each pixel of our result is represented by nine pixels in these images. \end{remark*} With one sample per pixel, as we increase $R$ the fractal structure comes into focus but then is lost to noise. This matches our observations in Figures~\ref{Fig:VisSphereRadii} and~\ref{Fig:VisSphereRadii2}. Taking subsamples and averaging makes little difference for small $R$: the only advantage is an anti-aliasing effect on the boundaries between regions of constant value. However, subsamples help greatly with reducing noise for larger $R$. With $2\times 2$ subsamples, we see much less noise at $R=10$, becoming more noticeable at $R=12$. Taking $128\times128$ samples seems to be very stable: there is almost no difference between the images with $R=10$ and $R=12$. This suggests that the perceived images converge. \subsection{Mean and variance within a pixel} To better understand how subsampling interacts with increasing $R$, in \reffig{SubPixEvolution} we graph the average value within a selection of pixel-sized regions as $R$ increases. \begin{figure}[htb] \includegraphics[width=0.3\textwidth]{Figures/FiguresMatthiasVer2/m122_4_-1/pixel_evolution_series_1_0_0} \includegraphics[width=0.3\textwidth]{Figures/FiguresMatthiasVer2/m122_4_-1/pixel_evolution_series_1_0_1} \includegraphics[width=0.3\textwidth]{Figures/FiguresMatthiasVer2/m122_4_-1/pixel_evolution_series_1_0_2}\\ \includegraphics[width=0.3\textwidth]{Figures/FiguresMatthiasVer2/m122_4_-1/pixel_evolution_series_1_1_0} \includegraphics[width=0.3\textwidth]{Figures/FiguresMatthiasVer2/m122_4_-1/pixel_evolution_series_1_1_1} \includegraphics[width=0.3\textwidth]{Figures/FiguresMatthiasVer2/m122_4_-1/pixel_evolution_series_1_1_2}\\ \includegraphics[width=0.3\textwidth]{Figures/FiguresMatthiasVer2/m122_4_-1/pixel_evolution_series_1_2_0} \includegraphics[width=0.3\textwidth]{Figures/FiguresMatthiasVer2/m122_4_-1/pixel_evolution_series_1_2_1} \includegraphics[width=0.3\textwidth]{Figures/FiguresMatthiasVer2/m122_4_-1/pixel_evolution_series_1_2_2} \caption{The graph of the average value of the cohomology fractal for \texttt{m122(4,-1)} for various square regions $U$ with field of view $0.1^\circ$. Thus, these are the same size as the pixels of \reffig{RagainstNumSamples}. These are each computed by taking $1000\times 1000$ samples. We also show the envelopes of 0.5, 1.0 and 1.5 standard deviations. } \label{Fig:SubPixEvolution} \end{figure} When $R$ is small, the graphs are more-or-less step functions, as much of the time the pixel $U$ is inside of a constant value region of the cohomology fractal. The graphs are also very similar for small $R$. This is because the pixels are close to each other, so all of their rays initially cross the same sequence of faces of the triangulation. Around $R=6$, we reach the ``last step'' of the step function, then the regions of constant value become smaller than $U$. For $R \geq 10$, the mean seems to settle down, while the standard deviation appears to grow like $\sqrt{R}$. Again this suggests that the perceived images converge. However, if the standard deviation continues to increase with $R$, then eventually any number of subsamples within each pixel will succumb to noise. \subsection{Histograms} We have looked at the standard deviation of a sample of values within a pixel. Next, we analyse the distribution of these values in more detail. See \reffig{m122Histogram}. \begin{figure}[htb] \centering \subfloat[Histogram of weights and the normal distribution with the same mean and standard deviation.]{ \includegraphics[width = 6.5cm]{Figures/FiguresMatthiasVer2/m122_4_-1/distribution} \label{Fig:m122(4,1)Dist} } \subfloat[Cohomology fractal.]{ \includegraphics[height = 5cm]{Figures/FiguresMatthiasVer2/m122_4_-1/without_evil} \label{Fig:m122(4,1)CohomFrac} } \caption{Statistics for a cohomology fractal of \texttt{m122(4,-1)} for a square region with field of view $20^\circ$ and $R=e^2$.} \label{Fig:m122Histogram} \end{figure} We fix $R = e^2$. We sample $\Phi_R$ at each point of a $1000 \times 1000$ grid within a square of a material view with field of view $20^\circ$. We chose a relatively large field of view here so that we get an ``in focus'' image of the cohomology fractal with a relatively small value of $R$. Here we are being cautious to get good data, avoiding potential problems that our implementation has with large values of $R$ as discussed in \refsec{Accumulate}. We histogram the resulting data with appropriate choices of bucket widths. In \reffig{m122(4,1)Dist} we show the histogram and the normal distribution with the same mean and standard deviation for our closed example, \texttt{m122(4,-1)}. In \reffig{m122(4,1)CohomFrac} we show the sample data as a 1000 by 1000 pixel image. We also draw the normal distribution with the same mean and standard deviation; the data seems to fit this well. \begin{figure}[htb] \centering \subfloat[Histogram of weights and the normal distribution with the same mean and standard deviation.]{ \includegraphics[width = 6.5cm]{Figures/FiguresMatthiasVer2/s789/distribution} \label{Fig:s789Dist} } \subfloat[Cohomology fractal.]{ \includegraphics[height = 5cm]{Figures/FiguresMatthiasVer2/s789/view} } \caption{Statistics for a cohomology fractal of \texttt{s789} for a class vanishing on cusp.} \label{Fig:s789Histogram} \end{figure} \begin{figure}[htb] \centering \subfloat[Histogram of weights and the normal distribution with the same mean and standard deviation.]{ \includegraphics[width = 6.5cm]{Figures/FiguresMatthiasVer2/s789/distribution2} } \subfloat[Cohomology fractal.]{ \includegraphics[height = 5cm]{Figures/FiguresMatthiasVer2/s789/view2} } \caption{Statistics for a cohomology fractal of \texttt{s789} for a class not vanishing on cusp.} \label{Fig:s789Histogram2} \end{figure} We repeat this experiment back in the cusped case with \texttt{s789}. See Figures~\ref{Fig:s789Histogram} and~\ref{Fig:s789Histogram2}. Here we show the cohomology fractal for two different cohomology classes $[\omega]\in H^1(M)$. The cohomology class shown in \reffig{s789Histogram} vanishes when restricted to $\bdy M$, while in \reffig{s789Histogram2} it does not. The distribution appears to be normal when the cohomology class vanishes on $\partial M$. When $[\omega]$ does not vanish on $\bdy M$, something more complicated appears to be happening. One feature here is that the tails are much too long for a normal distribution. A heuristic explanation for this is that in a neighbourhood of the cusp, a geodesic ray crosses the surface repeatedly in the same direction. This allows it to gain a linear weight in logarithmic distance. \section{The central limit theorem} \label{Sec:CLT} In this section, we prove a central limit theorem for the values of the cohomology fractal $\Phi_T$ across a pixel. That is, the distribution of the values of the scaled cohomology fractal $R_T=\Phi_T/\sqrt{T}$ converges to a normal distribution with mean zero. \subsection{Setup} \label{Sec:formalDefView} We recall the framework introduced in \refsec{Views} that unifies the material, ideal, and hyperideal views. Let $M$ be a connected, orientable, finite volume, complete hyperbolic three-manifold. We will use the abbreviations $X=\UT{}{M}$ and $\cover{X}=\UT{}{\cover{M}}$. We call a two-dimensional subset $D \subset \cover{X}$ a \emph{view} if it is of one of the following.% \begin{itemize} \item For a material view, fix a basepoint $p\in\cover{M}$ and let $D=\UT{p}{\cover{M}}$. Note that $D$ can be identified isometrically with $S^2$. \item For the ideal view, fix a horosphere $H\subset\cover{M}$ and let $D$ be the set of outward normals to $H$. \item For the hyperideal view, fix a hyperbolic plane $H\subset\cover{M}$ and let $D$ be the set of normals to $H$ facing one of the two possible directions. \end{itemize} Note that $D\subset \cover{X}$ has a riemannian metric induced from the riemannian metric on $\cover{X}$. This metric also endows $D$ with an area two-form and associated measure denoted by $\zeta=\zeta_D$ and $\mu=\mu_D$. Recall that $\pi\from\cover{X}\to \cover{M}$ is the projection to the base space. \begin{remark} \label{Rem:extrinsicCurvCorr} Note that there is another riemannian metric on $D$ for the ideal and hyperideal view coming from isometrically identifying the horosphere or hyperbolic plane $H=\pi(D)$ with $\mathbb{E}^2$, respectively, $\mathbb{H}^2$. Up to a constant factor, this metric is the same as the above metric. The factor is trivial for the hyperideal view and $\sqrt{2}$ for the ideal view and arises as $\sqrt{1+K_H^2}$ from the extrinsic curvature $K_H$ of $H$. We have $K_H=1$ so that adding $K_H$ to the ambient curvature $-1$ of $\mathbb{H}^3$ gives the horosphere's intrinsic curvature 0. \end{remark} In this notation, the definition of the cohomology fractal, for a given closed one-form $\omega\in\Omega^1(M)$ and basepoint $b\in\cover{M}$, becomes the following. For $v\in D$, we have \begin{equation} \label{eq:cohomFractView} \Phi^{\omega, b, D}_{T}(v) = \int_0^T \omega(\varphi_t(v)) dt + \int_{b}^{\pi(v)} \omega \end{equation} For the second integral, any path from $b$ to $\pi(v)$ in $\cover{M}$ can be chosen as $\omega$ is closed. This integral is constant in $v$ for the material view since $\pi(D)=p$. Choosing $W\in\Omega^0(\cover{M})$ so that $dW=\cover{\omega}$ and $W(b) = 0$, we can simply write \[ \Phi_T(v)=\Phi^{\omega,b,D}_{T} (v) = W(\varphi_T(v)) \] The central limit theorem will apply to probability measures $\nu_D$ that are absolutely continuous with respect to the area measure $\mu_D$ on $D$. We use the usual notation $\nu_D\ll\mu_D$ for absolute continuity. By the Radon-Nikodym theorem, this is equivalent to saying that the measure $\nu_D$ is given by \begin{equation} \label{Eqn:RN} \nu_D(U) = \int_U h\cdot d\mu_D \end{equation} where $h\geq 0$ is measurable with $\int_D h\cdot d\mu_D = 1$. \begin{remark} \label{Rem:MeasuresForms} In \refsec{Pixel}, we will switch from measures $\nu_D$ to forms $\eta_D$, for the following reason. Here, in \refsec{CLT} we follow the well-established notation of \cite{Sinai60,zweimuellerInfMeasurePreserving2007}. However, the transformation laws in \refsec{Pixel} are better stated in the language of forms. In both cases we consider probability measures or two-forms that are ``products'': namely of a suitable function $h\from D\to\mathbb{R}$ with the area measure $\mu_D$ or two-form $\zeta_D$ respectively. The function $h$ should be thought of as an indicator (or a kernel) function for a pixel. \end{remark} \subsection{The statement of the central limit theorem} \label{Sec:StatementCLT} The goal of this section is to prove the following. \begin{theorem} \label{Thm:CLT} Fix a connected, orientable, finite volume, complete hyperbolic three-manifold $M$ and a closed, non-exact, compactly supported one-form $\omega\in\Omega_c^1(M)$. There is $\sigma > 0$ such that for all basepoints $b$, for all views $D$ with area measure $\mu_D$, for all probability measures $\nu_D\ll \mu_D$, and for all $\alpha \in \mathbb{R}$, we have \[ \lim_{T\to\infty} \nu_D\left[ v\in D : \frac{\Phi_T(v)}{\sqrt{T}} \leq \alpha \right] = \int_{-\infty}^\alpha \frac{1}{\sigma\sqrt{2\pi}} e^{-(s/\sigma)^2/2} ds \] where $\Phi_T=\Phi^{\omega,b,D}_{T}$ is the associated cohomology fractal. \end{theorem} Let us recall some notions from probability to clarify what this means. Let $(P, \nu)$ be a probability space. For each $T \in \mathbb{R}_{>0}$, let $R_T \from P \to \mathbb{R}$ be a measurable function. For each $T$, the probability measure $\nu$ on $P$ induces a probability measure $\nu \circ R_T^{-1}$ on $\mathbb{R}$ telling us how the values of the random variable $R_T$ are distributed when sampling $P$ with respect to $\nu$. Let $\psi$ be a probability measure on $\mathbb{R}$. \begin{definition} \label{Def:ConvergeInDistribution} We say that the random variables $R_T$ \emph{converges in distribution} to $\psi$ if the measures $\nu \circ R_T^{-1}$ converge in measure to $\psi$. This is denoted by $\nu \circ R_T^{-1} \Rightarrow \psi$. \end{definition} Here, by the Portmanteau theorem, we can use any of several equivalent definitions of weak convergence of measures. We are only interested in the case where $\psi$ is absolutely continuous with respect to Lebesgue measure on $\mathbb{R}$; that is, we can write $\psi(V) = \int_V p(x) dx$ for any measurable $V \subset \mathbb{R}$. Note that here $p \from \mathbb{R} \to \mathbb{R}_{\geq 0}$ is the \emph{probability density function} for $\psi$. Convergence in distribution $\nu \circ R_T^{-1} \Rightarrow \psi$ is then equivalent to saying that for all $\alpha$ we have \[ \lim_{T\to \infty} \nu\left[ x\in P: R_T(x) \leq \alpha \right] = \int_{-\infty}^\alpha p(s) \cdot ds \] We define \[ n_\sigma(s)=\frac{1}{\sigma\sqrt{2\pi}} e^{-(s/\sigma)^2/2} \quad\mbox{and}\quad \psi_\sigma(V) = \int_V n_\sigma(x) dx \] The latter is the \emph{normal distribution} with mean zero and standard deviation $\sigma$. \begin{example} \newcommand{\mathrm{head}}{\mathrm{head}} \newcommand{\mathrm{tail}}{\mathrm{tail}} The process of flipping coins can be modelled as follows. Set $P=\{\mathrm{head}, \mathrm{tail}\}^\mathbb{N}$. We define a measure $\nu_P$ on $P$ as follows. Given any prefix $v$ of length $n$, the set of all infinite words in $P$ starting with $v$ has measure $2^{-n}$. Let $S_i\from P\to \mathbb{R}$ be the random variable $S_i(w) = \pm 1$ as $w_i$ is heads or tails respectively. Define $\Sigma_N = S_0 + S_1 + \cdots + S_{N-1}$. The classical central limit theorem states that \[ \nu_P \circ R_N^{-1} \Rightarrow \psi_1\quad\mbox{where}\quad R_N=\frac{\Sigma_N}{\sqrt{N}}\from P\to\mathbb{R} \qedhere \] \end{example} We can now restate \refthm{CLT} as \[ \nu_D \circ R_T^{-1}\Rightarrow\psi_\sigma \quad\mbox{where}\quad R_T=\frac{\Phi_T}{\sqrt{T}}\from D\to\mathbb{R} \] \subsection{Sinai's theorem} \label{Sec:Sinai} Our proof of \refthm{CLT} starts with Sinai's central limit theorem for geodesic flows~\cite{Sinai60}. We use the following version of Sinai's theorem which is adopted from \cite[Theorem~VIII.7.1 and subsequent Nota Bene]{FranchiLeJan}. This applies to functions that are not derivatives in the following sense. Recall that $X = \UT{}{M}$. \begin{definition} \label{Def:lieDerivative} Let $f\from X\to\mathbb{R}$ be a smooth function. We say that $f$ is a \emph{derivative} if there is a smooth function $F\from X\to\mathbb{R}$ such that \[ f(v)= \left.\frac{dF(\varphi_t(v))}{d t}\right|_{t=0} \qedhere \] \end{definition} \noindent Let $\mu_X = \mu_{\operatorname{Haar}}/\mu_{\operatorname{Haar}}(X)$ be the normalised Haar measure. \begin{theorem}[Sinai-Le Jan's Central Limit Theorem] \label{Thm:sinai} Fix a connected, orientable, finite volume, complete hyperbolic three-manifold $M$. Let $f \from X = \UT{}{M} \to \mathbb{R}$ be a compactly supported, smooth function with $\int_X f \cdot d\mu_X = 0$. Assume $f$ is not a derivative. Let \[ R_T(v) =\frac{\int_0^T f(\varphi_t(v))dt}{\sqrt{T}} \] Then there is a $\sigma>0$ such that $\mu_X \circ R_T^{-1} \Rightarrow \psi_\sigma$. \qedhere \end{theorem} In fact, the constant $\sigma$ appearing in \refthm{sinai} is the square root of the \emph{variance} of $f$ which Franchi--Le Jan denote by $\mathcal{V}(f)$. They give a formula for $\mathcal{V}(f)$ in \cite[Theorem~VIII.7.1]{FranchiLeJan} and state that $\mathcal{V}(f)$ vanishes if and only if $f$ is a derivative. \begin{remark} To relate \refthm{sinai} to~\cite[Theorem~VIII.7.1]{FranchiLeJan}, note that Franchi--Le Jan think of $f$ as a function on the frame bundle of $\cover{M}$ that is both $\Gamma$ and $\SO(2)$--invariant. Since the smooth function $f$ is compactly supported, $f$ has bounded and H\"olderian derivatives as required by Franchi--Le Jan. Note that they also require $f$ to not be a derivative (denoted by $\mathcal{L}_0h$, see \cite[(VIII.1)]{FranchiLeJan}) of a function $h$ but allow $h$ to be a function on the frame bundle. However, if an $\SO(2)$--invariant $f$ is the derivative of a function $h$ on the frame bundle, it is also the derivative of an $\SO(2)$--invariant function on the frame bundle. \end{remark} We deduce \refthm{CLT} from Sinai's theorem in three steps. \begin{enumerate} \item \refthm{sinaiOne} generalises Sinai's theorem to arbitrary probability measures $\nu_X \ll \mu_X$ on the five-dimensional $X = \UT{}{M}$. \item \refthm{sinaiTwo} restricts from $X$ to the two-dimensional view $D$ using a measure $\nu_D \ll \mu_D$. \item Finally, we show that the term $\int_{b}^{\pi(v)}\omega$ from \eqref{eq:cohomFractView} can be added to obtain $\Phi_T$. \end{enumerate} \subsection{Generalising Sinai's Theorem} We begin with a definition. \begin{definition} \label{Def:ConvInProb} Let $(P, \mu)$ be a finite measure space. For $n \in \mathbb{N}$, let $Q_n\from P\to \mathbb{R}$ be a measurable function. We say that \emph{$Q_n$ converges to zero in probability} and write $\mu \circ Q_n^{-1}\to 0$ if for all $\varepsilon >0$ we have \[ \lim_{n\to\infty} (\mu\circ Q_n^{-1}) ((-\infty, -\varepsilon) \cup (\varepsilon,\infty)) \to 0 \qedhere \] \end{definition} The following result is called \emph{strong distributional convergence}, see \cite[Proposition~3.4]{zweimuellerInfMeasurePreserving2007}. \begin{theorem} \label{Thm:strongErgodicConvergence} Let $(P, \mu)$ be a finite measure space and $T \from P \to P$ be an ergodic, measure-preserving transformation. For all $n\in\mathbb{N}$, let $R_n \from P \to \mathbb{R}$ be a measurable function. Let $Q_n = R_n\circ T-R_n$ and suppose that $\mu\circ Q_n^{-1}\to 0$. Let $\psi$ be a probability measure on $\mathbb{R}$. If we have $\nu\circ R_n^{-1}\Rightarrow \psi$ for some probability measure $\nu\ll \mu$, then we have $\nu\circ R_n^{-1}\Rightarrow \psi$ for all probability measures $\nu\ll \mu$. \qed \end{theorem} \begin{remark} We have specialised Zweim\"uller's Proposition~3.4 in \cite{zweimuellerInfMeasurePreserving2007} to finite measure spaces. To obtain Zweim\"uller's result for $\sigma$--finite measure spaces $(P,\mu)$, we need to replace the requirement $\mu\circ Q_n^{-1}\to 0$ by the following weaker requirement denoted by $Q_n\xrightarrow{\mu} 0$ in \cite[Footnote~3]{zweimuellerInfMeasurePreserving2007}: for all probability measures $\nu\ll \mu$ we have $\nu\circ Q_n^{-1}\to 0$. To see that $Q_n\xrightarrow{\mu} 0$ is weaker, we can use the following standard result: for any $\nu \ll \mu$ we have \[ \sup \{ \nu(A) : \mbox{$A$ measurable with $\mu(A)\leq \varepsilon$} \} \to 0\quad\mbox{as}\quad\varepsilon\to 0 \] The requirements $\mu\circ Q_n^{-1}\to 0$ and $Q_n\xrightarrow{\mu} 0$ are equivalent if $\mu$ is finite. \end{remark} Using this, we now give our first variant of Sinai's theorem. \begin{theorem} \label{Thm:sinaiOne} With the same hypotheses as in \refthm{sinai}, we have the following. There is a $\sigma>0$ such that for any probability measure $\nu_X \ll \mu_X$ we have $\nu_X \circ R_T^{-1} \Rightarrow \psi_\sigma$. \end{theorem} \begin{proof} By \refthm{sinai}, there is a $\sigma$ such that $\mu_X \circ R_T^{-1} \Rightarrow \psi_\sigma$ as $T\to\infty$. Note that the random variables in \refthm{strongErgodicConvergence} are indexed by $n\in\mathbb{N}$ instead of $T\in\mathbb{R}$ but it is easy to see that sequential convergence and convergence in distribution are equivalent. In other words, it is suffices to show that for any sequence $(T_n)$ with $T_n\to\infty$ the random variables $S_n=R_{T_n}$ satisfy $\nu_X \circ S_n^{-1}\Rightarrow \psi_\sigma$. In order to apply \refthm{strongErgodicConvergence}, we need the following two claims. \begin{claim} The time-one geodesic flow $\varphi_1$ is ergodic. \end{claim} \begin{proof} Set $G=\mathbb{Z}$. By \cite{PughShub}, the transformation $\varphi_1$ is an \emph{Anosov element}. By \cite[Theorem~1.1]{PughShub}, it is ergodic. \end{proof} \begin{claim} Let $Q_n=S_n\circ \varphi_1-S_n$. Then, $\nu_X \circ Q_n^{-1}\to 0$. \end{claim} \begin{proof} We will prove a stronger statement: $\| Q_n \|_\infty \to 0$. \begin{align*} \left| Q_n(v)\right| & = \left| \frac{\int_1^{T_n+1} f(\varphi_t(v)) dt }{\sqrt{T_n}} - \frac{\int_0^{T_n} f(\varphi_t(v)) dt}{ \sqrt{T_n}}\right| \\ & = \left| \frac{\int_{T_n}^{T_n+1} f(\varphi_t(v)) dt}{\sqrt{T_n}} - \frac{\int_0^1 f(\varphi_t(v)) dt }{ \sqrt{T_n}}\right| \\ & \leq \frac{2 \| f\|_\infty}{\sqrt{T_n}} \qedhere \end{align*} \end{proof} We can now finish the proof of \refthm{sinaiOne}. We apply \refthm{strongErgodicConvergence} with $R_n$ replaced by $S_n$ and $T$ replaced by $\varphi_1$. \end{proof} \subsection{Coordinates} \label{Sec:coordinatesView} Given a view $D$, we introduce coordinates for a neighbourhood of $D$ in $\cover{X}=\UT{}{\cover{M}}\homeo\UT{}{\mathbb{H}^3}$ as follows; it may be helpful to consult \reffig{localCoordinates}. Fix $r \in \cover{X}$. If $r$ is close enough to $D$ in $\cover{X}$, then there is an $x_{\sfu} \in D$ such that the rays emanating from $x_{\sfu}$ and $r$ converge to the same ideal point in $\bdy_\infty\cover{M}$. Consider the set $H = H^{\sf{s}}(x_{\sfu}) \subset \cover{X}$ such that \begin{itemize} \item $x_{\sfu} \in H$, \item $\pi(H)$ is a horosphere, and \item $H$ are the ``inward pointing'' normals to $\pi(H)$. \end{itemize} Let $x_{\sfs}$ be the intersection of $H$ with the line through $r$. Let $x_{\sff}$ be the signed distance from $x_{\sfs}$ to $r$ along this line. The triple \[ (x_{\sfu}, x_{\sff}, x_{\sfs}) \qquad \mbox{with $x_{\sfu} \in D$, $x_{\sff} \in \mathbb{R}$, and $x_{\sfs} \in H^{\sf{s}}(x_{\sfu})$} \] determines the vector $r \in \cover{X}$ uniquely. \begin{figure}[htbp] \centering \subfloat[Coordinates.]{ \labellist \small\hair 2pt \pinlabel {$H$} at 173 210 \pinlabel {$x_{\sfu}$} at 157 174 \pinlabel {$x_{\sfs}$} at 114 162 \pinlabel {$x_{\sff}$} at 76 175 \pinlabel {$r$} at 90 204 \pinlabel {$r_s$} at 113 128 \pinlabel {$D$} at 223 135 \endlabellist \includegraphics[width = 0.47\textwidth]{Figures/localCoordinates} \label{Fig:localCoordinates}} \hspace{-0.035\textwidth} \subfloat[Flow.]{ \labellist \small\hair 2pt \pinlabel {$H$} at 207 190 \pinlabel {$x_{\sfu}$} at 136 142 \endlabellist \includegraphics[width = 0.47\textwidth]{Figures/stableDirection} \label{Fig:StableDirection}} \caption{Coordinates for $\cover{X}=\UT{}{\cover{M}}=\UT{}{\mathbb{H}^3}$ and flow of a box $(x_{\sfu}, (-\varepsilon,\varepsilon), B^{\sf{s}}_\varepsilon(\pi(x_{\sfu})))$.} \end{figure} Suppose $N$ is a submanifold. Let $d_N(p,q)$ denote the length of the shortest curve in $N$ connecting $p$ and $q$. Given $x_{\sfu}\in D$, let \[ B^{\sf{s}}_\varepsilon(x_{\sfu}) = \{ x_{\sfs}\in H : d_H(x_{\sfu}, x_{\sfs}) \leq \varepsilon\} \quad\mbox{where}\quad H = H^{\sf{s}}(x_{\sfu}) \] Let \[ D_\varepsilon = \{ (x_{\sfu}, x_{\sff}, x_{\sfs}) : x_{\sfu}\in D, x_{\sff} \in (-\varepsilon, \varepsilon), x_{\sfs} \in B^{\sf{s}}_\varepsilon(x_{\sfu}) \} \] \begin{remark} \label{Rem:product} Note that the subscripts appearing in the coordinates $(x_{\sfu}, x_{\sff}, x_{\sfs})$ refer to the unstable, flow, and stable foliations. \begin{itemize} \item The points $(x_{\sfu}, 0, x_{\sfu})$ give a copy of $D$, which is unstable. \item If we fix $x_{\sfu}$ and $x_{\sfs}$, and vary $x_{\sff}$, then we obtain a geodesic flow line. \item Also, if we fix $x_{\sfu}$ and $x_{\sff}$, and vary $x_{\sfs}$, then we obtain a stable horosphere. \end{itemize} Note that each $H = H^{\sf{s}}(x_{\sfu})$ is isometric to a (pointed) copy of $\mathbb{C}$. However, the coordinates above do not live in a geometric product $D \times \mathbb{R} \times \mathbb{C}$. They instead form a smooth fibre bundle over $D$. (In the material case, the view $D$ is a copy of $S^2$. If we factor away the flow direction from our coordinates, what remains is isomorphic to the non-trivial bundle $\mathrm{T} S^2$.) Thus we will only locally appeal to a ``product structure'' on these coordinates. \end{remark} The following lemma is deduced from the exponential convergence inside of stable leaves. See \reffig{StableDirection}. \begin{lemma} \label{Lem:distFlowStable} For all $t\geq 0$, $1>\varepsilon>0$ and all $(x_{\sfu}, x_{\sff}, x_{\sfs}) \in D_\varepsilon$, we have \[ d_\cover{X}(\varphi_t(x_{\sfu}),\varphi_t(x_{\sfu},x_{\sff},x_{\sfs})) \leq 2 \varepsilon \] \end{lemma} \begin{proof} This follows from \begin{align*} d_\cover{X}((x_{\sfu},t,x_{\sfu}), (x_{\sfu},t,x_{\sfs})) \leq d_{\varphi_t(H)}((x_{\sfu},t,x_{\sfu}), (x_{\sfu},t,x_{\sfs})) & \\ = e^{-t} d_H((x_{\sfu},0,x_{\sfu}), (x_{\sfu},0,x_{\sfs})) &\leq e^{-t} \varepsilon \leq \varepsilon \\ \mbox{and} \quad d_\cover{X}((x_{\sfu},t,x_{\sfs}), (x_{\sfu},t+x_{\sff},x_{\sfs})) &= |x_{\sff}| \leq \varepsilon \end{align*} where $H=H^{\sf{s}}(x_{\sfu})$. \end{proof} \subsection{Proof of the central limit theorem} We now use these coordinates to continue with the proof of \refthm{CLT}. \begin{lemma} \label{Lem:almostEqualRandomVariables} Let $(P,\nu)$ be a probability space. Let $R_T, S_T\from P\to \mathbb{R}$ be two one-parameter families of measurable functions. Assume that $\nu\circ R_T^{-1}\Rightarrow \psi$ where $\psi$ is a probability measure on $\mathbb{R}$ with bounded probability distribution function $p\from\mathbb{R}\to\mathbb{R}_{\geq 0}$. Assume that there is a monotonically growing family $U_T\subset P$ of measurable sets such that $\left\| (R_T - S_T)|U_T\right\|_\infty\to 0$ and $\nu(P-U_T)\to 0$. Then $\nu\circ S_T^{-1}\Rightarrow \psi$. \end{lemma} \begin{proof} Fix $\alpha$ and let \[ P_T = \nu \left[x\in P : S_T(x) \leq \alpha \right]\quad\mbox{and}\quad Q_T = \nu \left[x\in P : S_T(x) > \alpha \right] \] We need to show that for every $\varepsilon > 0$, there is a $T_0$ such that for all $T\geq T_0$ we have \[ \int_{-\infty}^{\alpha} p(s) \cdot ds - \varepsilon \leq P_T \leq \int_{-\infty}^{\alpha} p(s) \cdot ds + \varepsilon \] We only deal with the second inequality since the first inequality can be derived in an analogous way using $P_T + Q_T = 1$. We have the following estimate. \[ P_T \leq \nu\left[ x\in P:R_T(x) \leq \alpha + \left\| (R_T - S_T) |U_T\right\|_\infty \right] + \nu(P-U_T) \] Fix $\varepsilon > 0$. Let $\delta=\varepsilon / (3 \|p\|_\infty)$. By hypothesis, we have for all large enough $T$ \[ P_T \leq \nu\left[ x\in P:R_T(x) \leq \alpha + \delta \right] + \frac{\varepsilon}{3} \] Because $R_T \xRightarrow{\nu} R$, we furthermore have for all large enough $T$ \begin{align*} P_T &\leq \frac{\varepsilon}{3} + \int_{-\infty}^{\alpha + \delta} p(s) \cdot ds + \frac{\varepsilon}{3}\\ & \leq \frac{\varepsilon}{3} + \int_{-\infty}^{\alpha} p(s) \cdot ds + \|p \|_\infty \delta + \frac{\varepsilon}{3} = \int_{-\infty}^{\alpha} p(s) \cdot ds + \varepsilon \qedhere \end{align*} \end{proof} We return to the case of interest where $f$ is given by a one-form $\omega$. \begin{lemma} \label{Lem:lieDerivativeNonVanishing} Let $\omega \in \Omega^1(M)$ be closed but not exact. Then $\omega\from X\to\mathbb{R}$ is not a derivative in the sense of \refdef{lieDerivative}. \end{lemma} \begin{proof} We prove the contrapositive: that is, if $\omega$ is a derivative in the sense of \refdef{lieDerivative} then $\omega = d W$ for a function $W \from M \to \mathbb{R}$. Fix a basepoint $p \in M$. We define $W(q) = \int_\gamma \omega$. Here $\gamma$ is a path from $p$ to $q$. All that is left is to show that $W$ is well-defined. So, suppose that $\gamma'$ is another path from $p$ to $q$. Thus $z = \gamma - \gamma'$ is a cycle. Let $z^*$ be the geodesic representative of $z$. Since $\omega$ is closed we have $\omega(z) = \omega(z^*)$. Since $\omega$ is a derivative we have $\omega(z^*) = 0$ and we are done. \end{proof} \begin{theorem} \label{Thm:sinaiTwo} With the same hypotheses as in \refthm{CLT}, we have the following. There is $\sigma > 0$ such that for all views $D$ with area measure $\mu_D$, for all probability measures $\nu_D\ll \mu_D$, and for all $\alpha\in\mathbb{R}$, we have \[ \lim_{T\to\infty} \nu_D\left[ v\in D : R_T(v) \leq \alpha \right] = \int_{-\infty}^\alpha \frac{1}{\sigma\sqrt{2\pi}} e^{-(s/\sigma)^2/2} ds \] where $R_T= \int_0^T \omega(\varphi_t(v)) dt / \sqrt{T}$. \end{theorem} \begin{proof} The one-form $\omega$ is not a derivative by \reflem{lieDerivativeNonVanishing}. Taking $f=\omega$, let $\sigma$ be as in \refthm{sinaiOne}. Fix a probability measure $\nu_D\ll\mu_D$. We define a measure $\nu_\cover{X}$ on $\cover{X}$ using the coordinates $(x_{\sfu}, x_{\sff}, x_{\sfs})$ by taking the product of, in order, \begin{itemize} \item $\nu_D$ \item the Lebesgue measure on $\mathbb{R}$ restricted to $[-1,1]$ \item the Lebesgue measure on $\mathbb{C}\cong H^{\sf{s}}(x_{\sfu})$ restricted to the unit disk $B^{{\sf{s}}}_1(x_{\sfu})$ \end{itemize} We scale $\nu_\cover{X}$ to be a probability measure. Note that the Lebesgue measure on $H^{\sf{s}}(x_{\sfu})$ does not depend on the isometric identification of $\mathbb{C}$ with $H^{\sf{s}}(x_{\sfu})$. Thus $\nu_\cover{X}$ is well-defined. By summing over fundamental domains, the probability measure $\nu_\cover{X}$ descends to a probability measure $\nu_X\ll \mu_X$ on $X$. Given that $R_T\from\cover{X}\to\mathbb{R}$ is $\pi_1(M)$--invariant, \refthm{sinaiOne} yields $\nu_\cover{X}\circ R_T^{-1}\Rightarrow \psi_\sigma$. Note that $\nu_\cover{X}$ is supported in the closure of $D_1$ (as defined before \refrem{product}). We have a projection $p\from D_1 \to D$ where $p(x_{\sfu},x_{\sff},x_{\sfs})=(x_{\sfu},0,x_{\sfu})$. By construction, we have $\nu_\cover{X}(p^{-1}(U)) = \nu_D(U)$ for any measurable set $U\subset D$. \begin{claim} \label{Clm:projConv} We have \[ \nu_\cover{X}\circ S_T^{-1}\Rightarrow \psi_\sigma\quad\mbox{where}\quad S_T = R_T\circ p \] \end{claim} \begin{proof} We take $P=D$ and we take $U_T=D$ for all $T$. Applying \reflem{almostEqualRandomVariables}, it is left to show that $\| R_T - S_T \|_\infty \to 0$. Let $W\from\cover{M}\to\mathbb{R}$ be a primitive of $\cover{\omega}$. That is $dW=\cover{\omega}$. In an abuse of notation, we abbreviate $W\circ \pi$ as $W$. We can now write \begin{align*} R_T(x_{\sfu},x_{\sff},x_{\sfs}) & = \frac{W(\varphi_T(x_{\sfu},x_{\sff},x_{\sfs})) - W(x_{\sfu},x_{\sff},x_{\sfs})}{\sqrt{T}} \\ S_T(x_{\sfu},x_{\sff},x_{\sfs}) & = \frac{W(\varphi_T(x_{\sfu},0,x_{\sfu})) - W(x_{\sfu},0,x_{\sfu})}{\sqrt{T}} \end{align*} Since $1/\sqrt{T}\to 0$, it is sufficient to show that both of \begin{align*} W(\varphi_T(x_{\sfu},x_{\sff},x_{\sfs})) & -W(\varphi_T(x_{\sfu},0,x_{\sfu})) \textrm{ and }\\ W(x_{\sfu},x_{\sff},x_{\sfs}) & -W(x_{\sfu},0,x_{\sfu}) \end{align*} are bounded by twice the Lipschitz constant of $W$. This follows from \reflem{distFlowStable} when replacing $\varepsilon$ by $1$ and setting $t$ to either $T$ or $0$. \end{proof} Fix $\alpha$. \refthm{sinaiTwo} follows from \begin{align*} \nu_D[x\in D: R_T(x) \leq \alpha] & = \nu_D[x\in D: S_T(x) \leq \alpha] \\ &= \nu_\cover{X}(p^{-1}(\{x\in D: S_T(x) \leq \alpha\}))\\ &= \nu_\cover{X}[x\in D_1:S_T(x)\leq \alpha] \end{align*} converging to $\int^\alpha_{-\infty} n_\sigma(s) ds$ by \refclm{projConv}. \end{proof} \begin{proof}[Proof of \refthm{CLT}] Note that \refthm{sinaiTwo} shows convergence for $$R_T(v)=\frac{\int_0^T f(\varphi_t(v))dt}{\sqrt{T}}$$ but we need to show convergence for $Q_T(v)= \Phi_T(v)/\sqrt{T}$, the difference being \[ \Delta(v) = R_T(v)-Q_T(v) = \frac{\int_b^{\pi(v)} \omega}{\sqrt{T}}=\frac{\int_b^{\pi(u)} \omega + \int_{\pi(u)}^{\pi(v)} \omega}{\sqrt{T}} \] where $u\in D$ is a fixed basepoint. Thus, we need to show that \reflem{almostEqualRandomVariables} applies when taking $P=D$. Denote the constant $\int_b^{\pi(u)} \omega$ by $C$. Let $C'$ be a bound on the absolute value of $\omega\from X\to\mathbb{R}$. It is convenient to let $$U_T = \{ v\in D : d_D(u, v) \leq \sqrt[4]{T} \}$$ Then, $\| \Delta | U_T \|_\infty \leq (C + C' \sqrt[4]{T}) /\sqrt{T} \to 0$ for $v\in U_T$. Since $U_T$ exhausts $D$ and $\nu$ is a finite measure, $\nu(D-U_T)\to 0$. \end{proof} \section{The pixel theorem} \label{Sec:Pixel} In this section, we prove that the cohomology fractal gives rise to a distribution at infinity. That is, integrating against the cohomology fractal then taking the limit as $R$ tends to infinity, gives a continuous linear functional on smooth, compactly supported functions. \subsection{Motivation} Throughout the paper, we have drawn many images of cohomology fractals, always depending on a visual radius $R$. The obvious question is whether there is a limiting image as $R$ tends to infinity. It turns out that the answer critically hinges on the question of what a pixel is. As we showed in \refthm{NoPicture}, thinking of a pixel as a sampled point does not work. After realising this, our next thought was that the cohomology fractal might converge to a signed measure $\mu$. We managed to prove this for squares (as well as for regions with piecewise smooth boundary). However our proof does not generalise to arbitrary measurable sets. See \refsec{RemarkMeasureHard} for a discussion. We finally arrived at the notion of thinking of a pixel as a smooth test function; see~\cite{Smith95}. The cohomology fractal now assigns to a pixel its weighted ``average value''; in other words, we obtain a well-defined \emph{distribution}. This distribution satisfies various transformation laws; these describe how it changes as we alter the chosen cocycle, basepoint, or view. To prove these we rely heavily on the \emph{exponential mixing} of the geodesic flow. \subsection{Background and statement} \label{Sec:StatementPixel} Before stating the theorem we establish our notation. We define $\omega$, $b$, $D$, and $T$ as in \refsec{formalDefView}. However, as mentioned in \refrem{MeasuresForms}, we switch from using the area measure $\mu_D$ to the area two-form $\zeta_D$ and from a probability measure $\nu_D$ to a compactly supported two-form $\eta_D$. To obtain $\eta = \eta_D\in\Omega^2_c(D)$, we set $\eta_D = h\zeta_D$; here $h\in\Omega^0_c(D)$ is compactly supported and smooth. That is, $h$ is Hodge dual to $\eta$. The function $h$ should be thought of as the kernel function for a pixel. The discussion below could be phrased completely in terms of $h$. However, using $\eta$ allows us to neatly express the transformation laws between different views. \begin{definition} \label{Def:Distribution} For a compactly supported two-form $\eta \in \Omega^2_c(D)$, we define \[ \Phi^{\omega,b,D}_{T}(\eta) = \int_{D} \Phi^{\omega,b,D}_{T} \eta \quad \mbox{and} \quad \Phi^{\omega,b,D}(\eta)= \lim_{T\to\infty} \Phi^{\omega,b,D}_{T}(\eta) \qedhere \] \end{definition} As we shall see, $\Phi^{\omega,b,D}$ is a \emph{distribution}: a continuous linear functional on $\Omega^2_c(D)$. We recall the topology on $\Omega^2_c(D)$ in the proof of \refthm{Pixel}. We will use $\int_{D}$ to denote the \emph{canonical distribution} $\eta \mapsto \int_{D} \eta$. To give a transformation law between views $D$ and $E$, we will need a way to relate one to the other. Recall that $\cover{M}$ is isometric to $\mathbb{H}^3$; thus we have $\bdy_\infty \cover{M} \homeo \CP^1$. As $t$ tends to infinity, the flow $\varphi_t$ takes a unit tangent vector $v \in D$ to some point $\varphi_\infty(v) \in \bdy_\infty\cover{M}$. This induces a conformal embedding $i_D$ of $D$ into $\bdy_\infty \cover{M}$. We define $i_E$ similarly. We take $i_{E,D} = i_D^{-1} \circ i_E$ where it is defined. This is a \emph{conformal isomorphism} from (a subset of) $E$ to (a subset of) $D$. We can now state the main result of this section. \begin{theorem}[Pixel theorem] \label{Thm:Pixel} Suppose that $M$ is a connected, orientable, finite volume, complete hyperbolic three-manifold. Fix a closed, compactly supported one-form $\omega \in \Omega_c^1(M)$, a basepoint $b \in \cover{M}$, and a view $D$. \begin{enumerate} \item \label{Itm:wellDefDist} Then $\Phi^{\omega, b, D}$ is well-defined and is a distribution. \item \label{Itm:cocycleIndep} Given $\omega' \in \Omega_c^1(M)$ with $[\omega]=[\omega']$, there is a constant $C$ so that we have \[ \Phi^{\omega', b, D} - \Phi^{\omega, b, D}= C \cdot \int_D \] \item \label{Itm:basePointPixelThm} Given another basepoint $b' \in \cover{M}$, we have \[ \Phi^{\omega, b', D} - \Phi^{\omega, b, D}= \left[ \int_b^{b'} \omega \right] \cdot \int_D \] \item \label{Itm:viewTransform} Given another view $E$ and a two-form $\eta \in \Omega^2_c(\operatorname{image}(i_{E,D}))$, we have \[ \Phi^{\omega,b,D}(\eta) = \Phi^{\omega, b, E}(i_{E,D}^* \, \eta) \] \end{enumerate} \end{theorem} The last property gives us a distribution at infinity as follows. \begin{corollary} \label{Cor:BoundaryDistribution} Suppose that $M$ is a connected, orientable, finite volume, complete hyperbolic three-manifold. Fix a closed, compactly supported one-form $\omega \in \Omega_c^1(M)$ and a basepoint $b \in \cover{M}$. Then there is a distribution $\Phi^{\omega,b}$ on $\bdy_\infty \cover{M}$ so that, for any view $D$ and any $\eta \in \Omega^2_c(D)$, we have \[ \pushQED{\qed} \Phi^{\omega,b,D}(\eta) = \Phi^{\omega, b}((i_D^{-1})^* \eta) \qedhere \popQED \] \end{corollary} \subsection{Proof of the pixel theorem} \label{Sec:PixelProof} We now describe some background necessary to prove the theorem. Throughout this section, we fix a connected, orientable, finite volume, complete hyperbolic three-manifold $M$. We use the abbreviations $X=\UT{}{M}$ and $\cover{X}=\UT{}{\cover{M}}$. Let \begin{equation} \label{Eqn:yt} Y_t = \int_D (\omega \circ \varphi_t) \eta \end{equation} Fubini's theorem implies that \begin{equation} \Phi^{\omega,b,D}_{T}(\eta) = \Phi^{\omega,b,D}_{0}(\eta) + \int_0^T Y_t dt \end{equation} so much of the proof boils down to obtaining exponential decay of $Y_t$. In this section we will use the Haar measure $\mu_{\operatorname{Haar}}$ for integrals over $X = \UT{}{M}$. We will use the shorthand $dv = d\mu_{\operatorname{Haar}}(v)$ throughout. We also need to introduce the Sobolev norm $S_m = S_{m,\infty}$ for smooth functions $f$ on homogenous spaces. First consider functions $f \from G \to \mathbb{R}$ where $G$ is a Lie group. Fix a basis for the Lie algebra of $G$; we think of the elements in this basis as left-invariant vector fields on $G$. The Sobolev norm $S_{m}(f)$ is the maximum of all $L_\infty$--norms of functions obtained by differentiating $f$ up to $m$ times using these vector fields in any order. Suppose that $\Gamma$ and $H$ are (respectively) discrete and compact subgroups of $G$. The Sobolev norm of $f \from \Gamma \backslash G \slash H \to \mathbb{R}$ is the Sobolev norm of the lift of $f$ to $G$. As usual, we have $M = \Gamma \backslash \mathbb{H}^3 \homeo \Gamma \backslash \PSL(2,\mathbb{C}) \slash \operatorname{PSU}(2)$. Likewise, we have $X = \UT{}{M} \homeo \Gamma \backslash \PSL(2,\mathbb{C}) \slash \operatorname{PSO}(2)$. For any of the three views, material, ideal, or hyperideal, we can also express $D$ in this fashion. For example, in the material view we have $D \homeo S^2 \homeo \operatorname{PSU}(2) \slash \operatorname{PSO}(2)$. For $\eta \in \Omega^2_c(D)$, define $S_m(\eta) = S_m(h)$ where $h$ is the Hodge dual of $\eta$. Note that the Sobolev norm depends on our choice of basis; however, changing the basis changes the resulting norm by a bounded factor and thus only changes the constant $C$ in the following lemma. \begin{lemma} \label{Lem:expoDecay} Let $M$ be a connected, orientable, finite volume, complete hyperbolic three-manifold. There is a constant $m \in \mathbb{N}$ such that the following is true. Fix a view $D$ in $M$ and a smooth, compactly supported function $f \from X =\UT{}{M} \to \mathbb{R}$ with $\int_X f(v) dv = 0$. Fix a compact set $K \subset D$. There are constants $C > 0$ and $c > 0$ such that for all two-forms $\eta \in \Omega_K^2(D)$ supported in $K$ and for all $t\geq 0$, we have \[ \left| \int_D (f \circ \varphi_t) \eta \right| \leq Ce^{-ct} S_m(\eta) \] \end{lemma} To prove this, we use the exponential decay of correlation coefficients for geodesic flows. This is a much studied area. We will rely on \cite{KelmerOh} because they explicitly give the dependence of the decay on the Sobolev norms of the functions involved. For hyperbolic, finite volume three-manifolds, their theorem can be simplified to the following. \begin{theorem} \label{Thm:mixing} Let $M$ be a connected, orientable, finite volume, complete hyperbolic three-manifold. Then there exists $m\in\mathbb{N}$, $C>0$ and $c>0$ with the following property. For any smooth functions $f,g\from X=\UT{}{M}\to\mathbb{R}$ with $\int_{X} f(v)dv = 0$ and for all $t\geq 0$, we have \begin{equation} \left|\int_{X} f(\varphi_t(v)) g(v) dv\right| \leq C e^{-ct} S_{m}(f) S_{m}(g) \end{equation} \end{theorem} \begin{proof} The more general \cite[Theorem~3.1]{KelmerOh} relates to \refthm{mixing} as follows. They integrate over the frame bundle $\Gamma\backslash\PSL(2,\mathbb{C})$ using the Bowen-Margulis-Sullivan-measure. However, we can think of a function $X\to\mathbb{R}$ as an $\operatorname{PSO}(2)$--invariant function on the frame bundle and the BMS-measure is simply the Haar measure in the case of a hyperbolic, finite volume three-manifold $\Gamma\backslash \mathbb{H}^3$. Note that \cite[Theorem~3.1]{KelmerOh} requires the functions $f$ and $g$ to be supported on a unit neighbourhood of the preimage of the convex core of $M$. However, for finite volume $M$, the convex core is just $M$. Furthermore, conventions for the Sobolev norm $S_m$ differ in whether to take the sum or maximum of the $L_\infty$--norms of derivatives; however the resulting norms are equivalent because they differ by a constant factor. \end{proof} To prove \reflem{expoDecay} using \refthm{mixing}, we construct test functions $h_\varepsilon \from \cover{X}\to\mathbb{R}$ that tend to the given two-form $\eta\in\Omega^2_c(D)$ in the sense that $Y_t$ can be approximated by \begin{equation} \label{Eqn:blurredIndicator} Y_{t,\varepsilon}=\int\limits_\cover{X} \cover{f}(\varphi_t(v)) h_\varepsilon(v) dv \end{equation} Note that there are several incompatibilities between $\eta$ and $h_\varepsilon$: \begin{enumerate} \item $\eta\in\Omega^2_c(D)$ is a two-form but $h_\varepsilon$ has to be a function. \item $D\subset\cover{X}$ but the integral \refthm{mixing} is over $X$. \item $D$ is two-, not five-dimensional. \end{enumerate} The first issue is solved by using the Hodge dual $h\in\Omega^0_c(D)$. That is, $\eta = h\zeta$ where $\zeta=\zeta_D$ is the area form on $D$. For the second issue, we reformulate \refthm{mixing} as follows: \begin{theorem} \label{Thm:mixing2} Let $M$ be a connected, orientable, finite volume, complete hyperbolic three-manifold. There is a constant $m\in\mathbb{N}$ such that the following is true. Fix a smooth, compactly supported function $f\from X=\UT{}{M}\to\mathbb{R}$ with $\int_X f(v)dv = 0$ and a compact set $K\subset \cover{X}=\UT{}{\cover{M}}$. There exists $C>0$ and $c>0$ such that for all smooth functions $g\from\cover{X}\to\mathbb{R}$ supported in $K$ and all $t\geq 0$, we have \begin{equation} \left|\int_\cover{X} \cover{f}(\varphi_t(v)) g(v) dv\right| \leq C e^{-ct} S_{m}(g) \end{equation} \end{theorem} \begin{proof} Note that $$\int_\cover{X} \cover{f}(\varphi_t(v)) g(v) dv = \int_X f(\varphi_t(v)) g_{\sum}(v) dv$$ where $g_{\sum}(v)$ is the sum of all $g(\cover{v})$ where $\cover{v}\in\cover{X}$ is a preimage of $v\in X$. Since $K$ can only cover a finite number of copies of a fundamental domain of $M$, the sum $g_{\sum}(v)$ has a finite and bounded number of terms. Since $f$ has compact support, $S_m(f)$ is finite and the result follows from \refthm{mixing}. \end{proof} To address the third issue, we make the following definition. \begin{definition} \label{Def:Bump} Define an \emph{$\varepsilon$--bump function} by $$b_\varepsilon(x)= \begin{cases} \exp\left(\frac{1}{(x/\varepsilon)^2 - 1}\right) & \mbox{if}\quad |x|<\varepsilon, \\ 0 & \mbox{otherwise} \end{cases}$$ and set $B=\left[\int_{-\infty}^{\infty} b_1(x)dx\right] \left[\int_0^{\infty} b_1(r) 2\pi r dr\right]$. \end{definition} We again use the coordinates on $\cover{X}$ already introduced in \refsec{coordinatesView}. Recall that $H=H^{\sf{s}}(x_{\sfu})$. We define $r_s=d_H(x_{\sfu},x_{\sfs})$. Again, see \reffig{localCoordinates}. We now define \begin{equation} \label{Eqn:defHEps} h_\varepsilon(r)=h(x_{\sfu}) \cdot \frac{b_\varepsilon(x_{\sff})b_\varepsilon(r_s)}{B \varepsilon^3} \end{equation} Using Fubini's theorem, we can now write $$Y_{t,\varepsilon} = \! \int_D \int_\mathbb{R} \int_{H^{\sf{s}}(x_{\sfu})} \cover{f}(\varphi_t(x_{\sfu},x_{\sff},x_{\sfs})) h_\varepsilon(x_{\sfu},x_{\sff},x_{\sfs}) J_D(x_{\sfu}, x_{\sff}, x_{\sfs}) dx_{\sfs} dx_{\sff} dx_{\sfu}$$ where $dx_{\sfu}$ and $dx_{\sfs}$ are using the area measures on $D$ and $H^{\sf{s}}(x_{\sfu})$, respectively, and where $J_D(x_{\sfu},x_{\sff},x_{\sfs})$ is the smooth function such that \begin{equation} \label{Eqn:ScaleFactor} d\mu_X = J_D(x_{\sfu},x_{\sff},x_{\sfs}) dx_{\sfs} dx_{\sff} dx_{\sfu} \end{equation} Note that, by construction, $J_D$ is invariant under isometries fixing $D$. In particular, $J_D(x_{\sfu},0,x_{\sfu})$ is a positive constant. We set $J_0 = J_D(x_{\sfu},0,x_{\sfu})$. Defining \begin{equation} \label{Eqn:SliceBlurredIndicator} Y^{\sf{u}}_{t,\varepsilon}(x_{\sfu}) = \int_\mathbb{R} \int_{H^{\sf{s}}(x_{\sfu})} \cover{f}(\varphi_t(x_{\sfu},x_{\sff},x_{\sfs})) J_D(x_{\sfu}, x_{\sff}, x_{\sfs}) \frac{b_\varepsilon(x_{\sff})b_\varepsilon(r_s)}{B \varepsilon^3} dx_{\sfs} dx_{\sff}, \end{equation} we can write $$Y_{t,\varepsilon} = \int_D Y^{\sf{u}}_{t,\varepsilon}(x_{\sfu}) h(x_{\sfu}) dx_{\sfu}$$ \begin{lemma} \label{Lem:approxStableAndFlow} For any smooth, compactly supported function $f\from X\to\mathbb{R}$ and any view $D$, there is a $C>0$ such that for all $t\geq 0$, for all $1>\varepsilon>0$, and for all $x_{\sfu}\in D$, we have $$\left| \cover{f}(\varphi_t(x_{\sfu})) J_0 - Y^{\sf{u}}_{t,\varepsilon}(x_{\sfu}) \right| \leq C\varepsilon$$ \end{lemma} \begin{proof} Let $$b_\varepsilon(x_{\sfu}, x_{\sff}, x_{\sfs}) = \frac{b_\varepsilon(x_{\sff}) b_\varepsilon(r_s)}{B\varepsilon^3}$$ Note that the support of $b_\varepsilon(x_{\sfu}, x_{\sff}, x_{\sfs})$ is the closure of $D_\varepsilon$ and that $\iint b_\varepsilon(x_{\sfu},x_{\sff},x_{\sfs}) dx_{\sfs} dx_{\sff} = 1$. Thus, it suffices to show that there is a $C>0$ such that for all $t\geq 0$, for all $1>\varepsilon>0$, and for all $(x_{\sfu},x_{\sff},x_{\sfs})\in D_\varepsilon$, we have \begin{equation} \label{Eq:lemApprox} \left| \cover{f}(\varphi_t(x_{\sfu})) J_0 - \cover{f}(\varphi_t(x_{\sfu},x_{\sff},x_{\sfs})) J_D(x_{\sfu}, x_{\sff}, x_{\sfs})\right| \leq C \varepsilon \end{equation} \begin{claim} Let $f \from X \to \mathbb{R}$ be a smooth, compactly supported function. Let $L=L(f)$ be a Lipschitz constant for $f$ and let $\cover{f}\from\cover{X}\to\mathbb{R}$ be its lift. For all $t \geq 0$, for all $1 > \varepsilon > 0$, and for all $(x_{\sfu}, x_{\sff}, x_{\sfs}) \in D_\varepsilon$, we have \[ \left| \cover{f}(\varphi_t(x_{\sfu})) - \cover{f}(\varphi_t(x_{\sfu},x_{\sff},x_{\sfs}))\right| \leq 2L\varepsilon \enspace \mbox{and} \enspace \left|\cover{f}(\varphi_t(x_{\sfu}))\right| \leq \| f \|_\infty \] \end{claim} \begin{proof} The first inequality follows from \reflem{distFlowStable}. The second is by definition. \end{proof} \begin{claim} The function $J_D$ has a finite Lipschitz constant $L'$ when restricted to $D_1$. Thus, for all $1>\varepsilon>0$ and for all $(x_{\sfu},x_{\sff},x_{\sfs})\in D_\varepsilon$, we have \[ \left| J_0 - J_D(x_{\sfu},x_{\sff},x_{\sfs})\right| \leq 2L'\varepsilon \] \end{claim} \begin{proof} Since $J_D$ is invariant under isometries preserving $D$, we can assume that $x_{\sfu}$ is fixed. Then, the domain $(x_{\sfu},(-1,1),B^{{\sf{s}}}_1(x_{\sfu}))$ has compact closure and thus $J_D$ has a finite Lipschitz constant $L'$ when restricted to it. The claim now follows from \reflem{distFlowStable}. \end{proof} Using the above two claims, the left hand side of \eqref{Eq:lemApprox} is bounded by $2L\varepsilon \cdot J_0 +\|f\|_\infty \cdot 2L'\varepsilon + 2L\varepsilon \cdot 2L' \varepsilon$. Thus setting $C=2L J_0 + 2L' \|f\|_\infty + 4L L'$ suffices to prove \reflem{approxStableAndFlow}. \end{proof} \begin{lemma} \label{Lem:approxStableFlow} Let $M$ be a connected, orientable, finite volume, complete hyperbolic three-manifold and $f\from X\to\mathbb{R}$ a smooth, compactly supported function. Fix a view $D$ and a compact set $K\subset D$. There is a $C>0$ such that for all smooth $h\from D\to\mathbb{R}$ supported in $K$ and for all $t\geq 0$, $1>\varepsilon>0$, we have $$\left| Y_t J_0 - Y_{t,\varepsilon} \right| \leq C\varepsilon \| h\|_\infty$$ \end{lemma} \begin{proof} We have $$Y_t J_0 - Y_{t,\varepsilon} = \int_D \left( \cover{f}(\varphi_t(x_{\sfu})) J_0 - Y^{\sf{u}}_{t,\varepsilon}(x_{\sfu})\right) h(x_{\sfu}) dx_{\sfu} $$ which is bounded by $$C\varepsilon \int_D |h(x_{\sfu})| dx_{\sfu}$$ by \reflem{approxStableAndFlow}. The result now follows since $K$ has finite area. \end{proof} \begin{lemma} \label{Lem:sobNormBound} Let $M$ be a connected, orientable, finite volume, complete hyperbolic three-manifold. Fix $m\in\mathbb{N}$. Fix a view $D$ and a compact set $K\subset D$. There is a $C>0$ such that for all smooth $h\from D\to\mathbb{R}$ supported in $K$ and all $1>\varepsilon>0$, we have $$S_m(h_\varepsilon) \leq C \varepsilon^{-(m+3)} S_m(h).$$ \end{lemma} \begin{proof} We estimate $S_m(h_\varepsilon)$ by using that $h_\varepsilon$ is separable as defined in \eqref{Eqn:defHEps}. In suitable coordinates, the second factor can be written as $$g_\varepsilon\from\mathbb{R}^3\to\mathbb{R},\quad (x, y, z)\mapsto \frac{b_\varepsilon(x) b_\varepsilon(\sqrt{y^2+z^2})}{B\varepsilon^3}$$ We have $g_\varepsilon(u)=g_1(u/\varepsilon) / \varepsilon^{3}$ so $S_m(g_\varepsilon) = \varepsilon^{-(m+3)} S_m(g_1)$. Using that all $h_\varepsilon$ are supported in a common compact set, the lemma follows from the following fact about Sobolev norms. Recall that a Sobolev norm requires a choice of vector fields that pointwise span the tangent space of the manifold (or a bundle over the manifold when using the Sobolev norm defined earlier by lifting a function $f\from \Gamma \backslash G \slash H\to \mathbb{R}$ to $G$). However, any two such choices yield Sobolev norms that differ by a bounded factor when restricting to a small enough neighbourhood or compact set. In particular, up to a bounded factor, we can estimate the Sobolev norm $S_m(h_\varepsilon)$ by Sobolev norms using local coordinates in which $h_\varepsilon$ is separable. \end{proof} \begin{proof}[Proof of \reflem{expoDecay}] Let $m$ be as in \refthm{mixing2}. Fix a smooth, compactly supported function $f$, a view $D$, and a compact $K\subset D$. \refthm{mixing2} states that there is a $C_0$ and $c_0$ such that for all smooth $h\from D\to\mathbb{R}$ supported in $K$ and all $1 > \varepsilon > 0$, we may set $g=h_\varepsilon$ and have $$|Y_{t,\varepsilon} | = \left| \int_\cover{X} \cover{f}(\varphi_t(v)) h_\varepsilon(v) dv \right| \leq C_0 e^{-c_0t} S_m(h_\varepsilon)$$ Applying \reflem{approxStableFlow} to the left-hand side and \reflem{sobNormBound} to the right hand-side, there are $C_1$ and $C_2$ such that \[ |Y_t J_0 | \leq C_1 \varepsilon \|h\|_\infty + C_0 e^{-c_0 t} C_2 \varepsilon^{-(m+3)} S_m(h) \] Since this holds for all $1 > \varepsilon > 0$, we can set $\varepsilon = e^{-c_0t/2(m+3)}$ and obtain \[ |Y_t | \leq \frac{C_1 + C_0 C_2}{J_0} e^{-c_0t/2} S_m(h) \qedhere \] \end{proof} \begin{proof}[Proof of \refthm{Pixel}] Recall that $$\Phi^{\omega,b,D}(\eta)=\int_0^\infty Y_t + \int_D \left[\int_b^{\pi(v)} \omega\right] \eta$$ We now discuss the topology on $\Omega^2_c(D)$. Many textbooks on distributions define the topology on $C^\infty_c(\mathbb{R}^n)$. This suffices in the ideal and hyperideal cases, where the view $D$ is diffeomorphic to $\mathbb{R}^2$. When $D$ is a two-sphere, we instead rely on the theory of distributions on generic smooth manifolds. We point the reader to \cite[Section~2.3]{BanCrainic} where such distributions are called ``generalised sections''. In particular, the definition of $\mathcal{D}(M;E)$ in \cite{BanCrainic} gives the correct topology on $\Omega^0_c(D)\cong C^\infty_c(D)$ and thus on $\Omega^2_c(D)\cong\Omega^0_c(D)$. Using \cite[Corollary~2.2.1]{BanCrainic} and the definition of the topology on $\Omega^2_c(D)$, it is not hard to see that statement~\refitm{wellDefDist} is equivalent to the following claim. \begin{claim} For any compact $K\subset D$, there is a $C>0$ such that for all $\eta\in\Omega^2_K(D)$, the integrals in $\Phi^{\omega,b,D}(\eta)$ exists and we have $|\Phi^{\omega,b,D}(\eta)| \leq C S_m(\eta)$. \end{claim} \begin{proof} Taking $f=\omega\from X\to\mathbb{R}$, \reflem{expoDecay} implies that there is a $C_0>0$ such that for all $\eta\in\Omega^2_K(D)$, the integral $\int_0^\infty Y_t$ exists and is bounded by $C_0 S_m(\eta)$. Let $C_1=\int_K \zeta_D$ be the area of $K$ and $C_2$ be the maximum of $\int_b^{\pi(v)} \omega$ over $v\in K$. Recall that $h\in\Omega^0_K(D)$ denotes the Hodge dual to $\eta=h\zeta_D$. Then, \[ \left| \Phi^{\omega,b,D}(\eta) \right| \leq C_0 S_m(\eta) + C_1 C_2 \| h\|_\infty \leq (C_0 + C_1 C_2) S_m(\eta) \qedhere \] \end{proof} Let $W\from \cover{M}\to\mathbb{R}$ be a primitive such that $dW=\omega$. In an abuse of notation, we again abbreviate $W\circ \pi$ as $W$. We have \begin{align} \Phi^{\omega,b,D}(\eta)&=\lim_{T\to\infty} \int_{v\in D} \left(W(\varphi_T(v)) - W(b)\right) \eta(v) \nonumber \\ & = \left[\lim_{T\to\infty} \int_D (W\circ \varphi_T) \eta\right] - W(b) \int_D \eta \label{eq:splitOffCanonical} \end{align} Statement~\refitm{cocycleIndep} is equivalent to the following claim. \begin{claim} If $[\omega] = 0$, there is a $C$ such that $\Phi^{\omega,b,D} = C \cdot \int_D $ \end{claim} \begin{proof} Since $\omega$ is a coboundary, the primitive descends to a map $W\from M\to\mathbb{R}$. Add a constant to $W$ to arrange $\int_M W dv = 0$; here we integrate with respect to the volume measure. Taking $f=W$, \reflem{expoDecay} now applies to \eqref{eq:splitOffCanonical} and we have $\Phi^{\omega,b,D} = -W(b) \cdot \int_D$. \end{proof} Statement~\refitm{basePointPixelThm} follows from from \refrem{DependenceOnb}. It is left to show Statement~\refitm{viewTransform}. Recall that we defined maps $i_D\from D\to\partial_\infty\cover{M}$ and $i_E\from E\to\partial_\infty\cover{M}$. Let $U=\operatorname{image}(i_D)\cap \operatorname{image}(i_E) \subset \partial_\infty\cover{M}$. Furthermore, let us define two distributions on $U$ by \[ \Phi^{\omega,b}_{\leftarrow D}(\delta) = \Phi^{\omega,b,D}(i_D^*(\delta)) \qquad\mbox{and}\qquad \Phi^{\omega,b}_{\leftarrow E} (\delta)= \Phi^{\omega,b,E}(i_E^*(\delta)) \] For every $\eta \in \Omega^2_c(\operatorname{image}(i_{E,D}))$, there is a compactly supported two-form $\delta$ on $U$ such that $\eta = i_D^* \delta$ and $i^*_{E,D} \eta = i_E^* \delta$. Thus, it is enough to show that \begin{equation} \label{Eqn:limViewSame} \Phi^{\omega,b}_{\leftarrow D} = \Phi^{\omega,b}_{\leftarrow E} \end{equation} \begin{figure}\centering \labellist \small\hair 2pt \pinlabel $s$ [B] at 89 179 \pinlabel $\varphi_T(i_D^{-1}(N))$ [Br] at 48 175 \pinlabel $\varphi_{T+\Delta T}(i_E^{-1}(N))$ [Bl] at 131 175 \pinlabel $i_D^{-1}(s)$ [tr] at 81 80 \pinlabel $i_E^{-1}(s)$ [tl] at 126 72 \pinlabel $\Delta T_s$ [bl] at 116 88 \pinlabel $r_s$ [b] at 104 79 \pinlabel $y$ [b] at 117 116 \pinlabel $H$ [Bl] at 128 115 \endlabellist \includegraphics[height=8cm]{Figures/viewsConverging} \caption{The regions $\varphi_T(i_D^{-1}(N))$ and $\varphi_{T+\Delta T}(i_E^{-1}(N))$ for two material views. The distance between corresponding points in the two regions is bounded by a constant for all large enough $T$. The constant can be made arbitrarily small by making the neighbourhood $N$ about $p$ small enough. \label{Fig:convergenceViews}} \end{figure} \begin{claim} \label{Clm:convergenceDistance} For every $p\in U$ and $\varepsilon>0$, there is a neighbourhood $N\subset U$ of $p$, a number $\Delta T\in\mathbb{R}$ and $T_0\geq 0$ such that for all $s\in N$ and $T\geq T_0$, we have \[|W(\varphi_T(i_D^{-1}(s))) - W(\varphi_{T+\Delta T}(i_E^{-1}(s)))| \leq \varepsilon\] \end{claim} \begin{proof} It might be helpful to consult \reffig{convergenceViews}. Let $L$ be a Lipschitz constant of $W\from X\to\mathbb{R}$. It is enough to show that \[d_\cover{X}(\varphi_T(i_D^{-1}(s)), \varphi_{T+\Delta T}(i_E^{-1}(s)))\leq \varepsilon / L\] Given $s\in U\subset S^2$, the rays corresponding to $i_D^{-1}(s)$ and $i_E^{-1}(s)$ both converge to $s$. There is a horosphere $H$ about $s$ containing the initial point $\pi(i_D^{-1}(s))$ of the first ray. The other ray intersects $H$ orthogonally as well. Let us denote this intersection by $y$. Let $\Delta T_s$ be the signed distance from $\pi(i_E^{-1}(s))$ to $y$ and let $r_s=d_H(y, \pi(i_D^{-1}(s)))$. Use the given $p\in U\subset S^2$ to set $\Delta T=\Delta T_p$. When flowing the first ray by $T$ and the second by $T+\Delta T$, we obtain the following distance estimate (see \refrem{extrinsicCurvCorr} to explain the factor $\sqrt{2}$) $$d_{\cover{X}}(\varphi_T(i_D^{-1}(s)), \varphi_{T+\Delta T}(i_E^{-1}(s))) \leq \sqrt{2} r_s e^{-T} + |\Delta T - \Delta T_s|$$ There is a neighbourhood $N$ of $p$ such that for all $s\in N$, we have $|\Delta T - \Delta T_s| \leq \varepsilon / 2L$. Note that $r_s$ is bounded on $N$ so there is a $T_0$ such that $\sqrt{2} r_s e^{-T_0}\leq \varepsilon / 2L.$ \end{proof} We define a semi-norm on $\Omega^2_c(U)$ as follows. Identify $\partial_\infty \cover{M}$ with $S^2$. Such an identification induces an area form $\zeta_\infty$ on $\partial_\infty \cover{M}$. Given $\delta \in\Omega^2_c(U)$, let $d$ be the Hodge dual, that is $\delta=d\cdot \zeta_\infty$. Let \[ \| \delta \|_1 = \int_U |d| \cdot \zeta_\infty \] Note that $\|\delta\|_1$ does not depend on the identification of $\partial_\infty \cover{M}$ with $S^2$. \begin{claim} \label{Clm:convergenceRelatedInts} For every $p\in U$ and $\varepsilon > 0$, there is a neighbourhood $N \subset U$ of $p$ such that for all $\delta\in \Omega^2_c(N)$, we have \[ \left| \Phi^{\omega,b}_{\leftarrow D}(\delta) - \Phi^{\omega,b}_{\leftarrow E}(\delta) \right| \leq \varepsilon \|\delta\|_1 \] \end{claim} \begin{proof} Let $N$, $\Delta T$ and $T_0$ as in \refclm{convergenceDistance}. Add a constant to the primitive $W$ of $\omega$ such that $W(b)=0$. Then, the left hand-side can be expressed using $W$ as follows. \begin{align*} & \left| \lim_{T\to\infty}\int_U W(\varphi_T(i_D^{-1}(s))) \cdot \delta - \lim_{T\to\infty}\int_U W(\varphi_T(i_E^{-1}(s))) \cdot \delta \right| = \\ & \left| \lim_{T\to\infty}\int_U W(\varphi_T(i_D^{-1}(s))) \cdot \delta - \lim_{T\to\infty}\int_U W(\varphi_{T+\Delta T}(i_E^{-1}(s))) \cdot \delta\right| = \\ & \left| \lim_{T\to\infty}\int_U \left[ W(\varphi_T(i_D^{-1}(s))) - W(\varphi_{T+\Delta T}(i_E^{-1}(s))\right] \cdot \delta \right| \end{align*} \refclm{convergenceDistance} implies that the integral is bounded by $\varepsilon \|\delta\|_1$ for all $T_0\geq T$, thus the limit is bounded by $\varepsilon \|\delta\|_1$. \end{proof} We now verify \eqref{Eqn:limViewSame}. Fix a smooth, compactly supported $\delta\in\Omega^2_c(U)$. Also fix $\varepsilon>0$. For each $p\in U$, pick a neighbourhood $N$ as in \refclm{convergenceRelatedInts}. The support $\supp(\delta)$ can be covered by finitely many of these neighbourhoods. Consider a partition of unity with respect to this finite cover. By multiplying $\delta$ by the partition functions, we obtain smooth two-forms $\delta_1,\dots, \delta_n$ with $\sum \delta_i = \delta$ and $\sum \|\delta_i\|_1 = \|\delta\|_1$ Since $\sum \varepsilon \|\delta_i\|_1 = \varepsilon \|\delta\|_1$, \refclm{convergenceRelatedInts} implies \[ \left| \Phi^{\omega,b}_{\leftarrow D}(\delta) - \Phi^{\omega,b}_{\leftarrow E}(\delta) \right| \leq \varepsilon \|\delta\|_1 \] Since $\varepsilon$ was arbitrary, we are done. \end{proof} \begin{remark} McMullen~\cite{McMullen19} suggests another approach to \refthm{Pixel}. Arrange matters so that $W$, the primitive of $\omega$, is harmonic with respect to the hyperbolic metric. Prove that $W$ has suitably bounded growth as it approaches $\bdy_\infty \mathbb{H}^3$, in terms of the euclidean metric in the Poincar\'e ball model. Now prove and then apply an appropriate version (of part (iii) implies part (i)) of~\cite[Theorem~1.1]{Straube84}. \end{remark} \subsection{The cohomology fractal as a measure} \label{Sec:RemarkMeasureHard} We do not know whether the cohomology fractal converges to a signed measure; that is whether or not $$\lim_{T\to\infty} \int_U \Phi^{\omega,b,D}_T \cdot d\mu_D$$ converges for every measurable $U\subset D$. Instead, we have the following partial result. \begin{theorem}[Square pixel theorem] \label{Thm:SquarePixel} Suppose that $M$ is a connected, orientable, finite volume, complete hyperbolic three-manifold. Fix a closed, compactly supported one-form $\omega \in \Omega_c^1(M)$, a basepoint $b \in \cover{M}$, and a view $D$. Suppose that $U \subset D$ is bounded and open and $\partial U$ consists of finitely many smooth curves. Then the following limit exists \[ \lim_{T\to\infty} \int_U \Phi^{\omega,b,D}_T \cdot d\mu_D \] \end{theorem} \begin{remark} Recall that the pictures in \reffig{RagainstNumSamples} were generated by uniformly sampling across a pixel square. \refthm{SquarePixel} finally proves that this technique (with enough samples) will give accurate images in some non-empty range of visual radii. Note that the earlier \refthm{Pixel} is not sufficient; it requires a smooth filter function. \end{remark} \begin{remark} To generalise \refthm{SquarePixel} to measurable subsets seems difficult. Our proof does not apply, for example, to an open set $U \subset D$ bounded by several Osgood arcs~\cite{Osgood03}. \end{remark} Suppose that $h\from D \to \mathbb{R}$ is a bounded measurable function with compact support. We define $h^{\sf{u}}_\varepsilon\from D \to \mathbb{R}$ by setting \begin{equation} \label{Eqn:hEpsUnstable} h^{\sf{u}}_\varepsilon(x_{\sfu}) = \frac{1}{B_\varepsilon} \int_D h \cdot b_\varepsilon(d_D(x_{\sfu}, v)) \cdot d\mu_D(v) \end{equation} Here $B_\varepsilon$ is a normalisation factor such that $\int_D h \cdot d\mu_D=\int_D h^{\sf{u}}_\varepsilon \cdot d\mu_D$. We call $h^{\sf{u}}_\varepsilon$ the \emph{$\varepsilon$--mollification (in the unstable direction)} of $h$. \begin{lemma} \label{Lem:technicalSquarePixel} Let $h\from D\to\mathbb{R}$ be a bounded measurable function with compact support. Assume that there is a $c>0$ and $C>0$ such that for all $1 > \varepsilon >0$, we have \[ \| h - h^{\sf{u}}_\varepsilon \|_1 \leq C\varepsilon^c \] Then the following limit exists \[ \lim_{T\to\infty} \int_D \Phi^{\omega,b,D}_T \cdot h \cdot d\mu_D \] \end{lemma} \begin{proof}[Proof sketch] We show that $Y_t = \int_D (\omega \circ \varphi_t) \cdot h\cdot d\mu_D$ decays exponentially. Let $\eta_\varepsilon = h^{\sf{u}}_\varepsilon \zeta_D$ and $Y_{t,\varepsilon} = \int_D (\omega \circ \varphi_t) \cdot \eta_\varepsilon$. Regarding $\omega$ as function $X\to\mathbb{R}$, we have \[ |Y_t - Y_{t,\varepsilon}| \leq \| h - h^{\sf{u}}_\varepsilon \|_1 \cdot \| \omega\|_\infty \leq C\varepsilon^c \|\omega\|_\infty \] Since $h$ is bounded, the Sobolev norm of the convolution $h^{\sf{u}}_\varepsilon$ of $h$ can be estimated from the Sobolev norm of the convolution kernel. This can be done similarly to \reflem{sobNormBound}. Thus, there is a $C_0>0$ such that for all $1>\varepsilon > 0$, we have $S_m(\eta_\varepsilon) \leq C_0 \varepsilon^{-(m+2)}$. Taking $f = \omega$ and $\eta = \eta_\varepsilon$, \reflem{expoDecay} states that there is a $C_1>0$ and $c_1>0$ such that for all $t\geq 0$ and $1>\varepsilon>0$, we have \[ |Y_{t,\varepsilon}| \leq C_1 e^{-c_1t} \varepsilon^{-(m+2)} \] Thus, we have \[ |Y_t| \leq C \varepsilon^c \| \omega\|_\infty + C_1e^{-c_1 t} \varepsilon^{-(m+2)} \] We obtain exponential decay when setting $\varepsilon = e^{-c_1 t / 2 (m+2)}.$ \end{proof} \begin{proof}[Proof sketch of \refthm{SquarePixel}] The theorem follows from \reflem{technicalSquarePixel} when taking $h$ to the indicator function $\chi_U$ of $U$. It is left to show that there is a $C$ such that $\| h - h^{\sf{u}}_\varepsilon\|_1 \leq C \varepsilon$. Note that $h-h^{\sf{u}}_\varepsilon$ is bounded by one, thus it is enough to show that the area where $h-h^{\sf{u}}_\varepsilon$ is non-trivial is bounded by $C\varepsilon$. But the area of this neighbourhood can be approximated by the length of $\partial U$ multiplied by $\varepsilon$. \end{proof} \section{Questions and projects} \label{Sec:Questions} \begin{question} Suppose that $F$ is a surface in a hyperbolic three-manifold $M$. \refthm{CLT} tells us that the standard deviation $\sigma$ is a topological invariant of the pair $(M,F)$. What are the number theoretic (or other) properties of $\sigma$? Fixing $M$, does $\sigma$ ``see'' the shape of the Thurston norm ball? \end{question} \begin{question} Suppose that $F$ is a fibre of a closed, connected hyperbolic surface bundle $M$. In \refprop{LightDark} we showed that approximations of the Cannon--Thurston map are (components of) level sets of the cohomology fractal. Is there some more precise sense in which the Cannon--Thurston map $\Psi$ is a ``level set'' of the distributional cohomology fractal $\Phi^{F, b}$? \end{question} \begin{question} Can the cohomology class $[\omega]$ be recovered from the distributional cohomology fractal $\Phi^{\omega, b}$? \end{question} \begin{question} \reffig{SubPixEvolution} suggests that in the example of \texttt{m122(4,-1)} the mean has settled down at around $R=10$ for a pixel size of $0.1^\circ$. We also see this in \reffig{RagainstNumSamples} in that there is hardly any difference between the images at $R = 10$ and $R=12$ with $128\times 128$ samples. \refthm{SquarePixel} tells us that given enough samples we can produce an accurate picture of the distributional cohomology fractal. Can one calculate effective bounds that would allow us to produce a provably correct image? \end{question} \begin{question} In \reffig{RagainstNumSamples}, the image with $1 \times 1$ samples and $R = 8$ is very similar to the image with $128 \times 128$ samples and $R = 12$. However, we do not understand how an image generated with only one sample per pixel can so closely approximate the limiting object. The manifold \texttt{m122(4,-1)} is small; as a result, perhaps the geodesic flow mixes rapidly enough? Does this fail in larger manifolds? \end{question} \begin{question}[Mark Pollicott] We consider lowering the dimension of $F$ and $M$ by one. Let $F$ be a non-separating curve in a closed, connected hyperbolic surface $M$. Fix a point $p \in M$. Let $P$ be a ``pixel'' -- that is, a closed arc in $\UT{p}{M} \homeo S^1$ with centre $c_P$ and radius $r_P$. The distributional cohomology fractal $\Phi^F$ exists and in fact gives a ``signed measure'' to each such pixel $P$ (see \refsec{RemarkMeasureHard}). By taking the pair $(c_P, r_P)$ to the measure of $P$, we obtain a function from $S^1 \times (0, \pi]$ to $\mathbb{R}$. What does its graph look like? For example, what happens if we fix $c_P$ and allow the radius to vary? How does the graph behave as $r_P$ approaches zero? \end{question} \begin{question} The histogram in \reffig{m122(4,1)Dist} is low near the mean. Increasing $R$ (within the range that we trust our experiments, see \refsec{Accumulate}) reduces, but does not remove, this gap. Why is it there? (This does not seem to happen in \reffig{s789Dist}, where the surface is closed but the manifold is not.) \end{question} \begin{question} \label{Que:Cusped} Consider the experiment shown in \reffig{s789Histogram2}. Here the support of the cocycle $\omega$ is not compact. We see that the distribution of the cohomology fractal, over a pixel, appears to not be normal. Further experiments show that it depends sensitively on the choice of pixel. Can one verify rigorously that it is not normal? We suspect that some version of ``subtracting the largest excursion'' (see the remarks immediately before \cite[Theorem~1]{DiamondVaaler86}) will yield a more reasonable distribution. \end{question} \begin{question} \refthm{Pixel} applies to cocycles $\omega$ with compact support. Consider a cusped manifold and a cocycle $\omega$ such that the pullback of $[\omega]$ to the cusp torus is non-trivial. In this case, the Sobolev norm of $\omega$ is infinite; thus \refthm{mixing} does not apply. Are there modifications, perhaps as indicated in \refque{Cusped}, so that we again obtain a distribution at infinity for the cohomology fractal? \end{question} \begin{question} In the fibred case, what is the relationship between the cohomology fractal and the lightning curve? See \refsec{Lightning}. \end{question} We end with some ideas for future software projects. \begin{project} One could use material triangulations in the closed case to draw an approximation $\Psi_D$ to the Cannon--Thurston map, following \refalg{CTApprox}. By \refprop{LightDark}, these match the cohomology fractal. Motivated by Figures~\ref{Fig:m122_4_-1} and~\ref{Fig:Orbifold}, we anticipate that $\Psi_D$ will look significantly different from Cannon--Thurston maps in the cusped case; the ``mating dendrites'' that approximate $\Psi$ have bounded branching at all points. \end{project} \begin{project} In \refsec{Cone}, we discussed cohomology fractals for incomplete structures along a line in Dehn surgery space. As discussed in \refsec{NumericalInstability}, all of these suffer from numerical defects along the incompleteness locus $\Sigma_s$. These defects are visible (although small) in \reffig{Bending}. When the slope $s$ is integral, we use material triangulations (\refsec{MaterialTriangulations}) to remove these defects. For general $s$, material triangulations are not available. Instead, one could \emph{accelerate} through tubes about $\Sigma_s$. That is, we modify the cellulation of the manifold by truncating each tetrahedron, replacing the lost volume with a solid torus ``cell'' around $\Sigma_s$. \end{project}
{ "timestamp": "2020-10-13T02:40:56", "yymm": "2010", "arxiv_id": "2010.05840", "language": "en", "url": "https://arxiv.org/abs/2010.05840", "abstract": "Cohomology fractals are images naturally associated to cohomology classes in hyperbolic three-manifolds. We generate these images for cusped, incomplete, and closed hyperbolic three-manifolds in real-time by ray-tracing to a fixed visual radius. We discovered cohomology fractals while attempting to illustrate Cannon-Thurston maps without using vector graphics; we prove a correspondence between these two, when the cohomology class is dual to a fibration. This allows us to verify our implementations by comparing our images of cohomology fractals to existing pictures of Cannon-Thurston maps.In a sequence of experiments, we explore the limiting behaviour of cohomology fractals as the visual radius increases. Motivated by these experiments, we prove that the values of the cohomology fractals are normally distributed, but with diverging standard deviations. In fact, the cohomology fractals do not converge to a function in the limit. Instead, we show that the limit is a distribution on the sphere at infinity, only depending on the manifold and cohomology class.", "subjects": "Geometric Topology (math.GT); Dynamical Systems (math.DS)", "title": "Cohomology fractals, Cannon-Thurston maps, and the geodesic flow", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9843363503693294, "lm_q2_score": 0.8128673201042492, "lm_q1q2_score": 0.8001348512059141 }
https://arxiv.org/abs/2007.01985
Limits of almost homogeneous spaces and their fundamental groups
We say that a sequence of proper geodesic spaces $X_i$ consists of \textit{almost homogeneous spaces} if there is a sequence of discrete groups of isometries $G_i \leq Iso(X_i)$ such that diam$(X_i/G_i)\to 0$ as $i \to \infty$.We show that if a sequence $X_i$ of almost homogeneous spaces converges in the pointed Gromov--Hausdorff sense to a space $X$, then $X$ is a nilpotent locally compact group equipped with an invariant geodesic metric.Under the above hypotheses, we show that if $X$ is semi-locally-simply-connected, then it is a nilpotent Lie group equipped with an invariant Finsler or sub-Finsler metric, and for large enough $i$, there are subgroups $\Lambda_i \leq \pi_1(X_i) $ with surjective morphisms $\Lambda_i\to \pi_1(X)$.
\section{Introduction} We say that a sequence of proper geodesic spaces $X_i$ consists of \textit{almost homogeneous spaces} if there is a sequence of discrete groups of isometries $G_i \leq Iso(X_i)$ such that diam$(X_i/G_i)\to 0$ as $i \to \infty$. \begin{remark} \rm A sequence of homogeneous spaces $X_i$ does not necessarily consist of almost homogeneous spaces, since the groups $Iso (X_i ) $ are not necessarily discrete. \end{remark} \begin{example} \rm Let $Z_i$ be a sequence of compact geodesic spaces with diam$(Z_i)\to 0$ as $i \to \infty$. If $\tilde{Z}_i \to Z_i$ is a sequence of Galois covers, then the sequence $\tilde{Z}_i$ consists of almost homogeneous spaces. \end{example} The goal of this paper is to understand the Gromov--Hausdorff limits of sequences of almost homogeneous spaces. In the case when the sequence consists of blow-downs of a single space, the problem was solved by Mikhail Gromov and Pierre Pansu (\cite{GromovPG}, \cite{Pansu}). \begin{theorem} \rm (Gromov--Pansu) Let $(X,x_0)$ be a pointed proper geodesic space, and $G \leq Iso (X)$ a discrete group of isometries with diam$(X/G) < \infty$. If for some sequence of positive numbers $\lambda_i \to \infty$, the sequence \[ \lim_{i \to \infty} \dfrac{1}{\lambda_i}(X, x_0 ) \to (Y,y_0) \] converges in the pointed Gromov--Hausdorff sense, then $Y$ is a simply connected nilpotent Lie group equipped with a Carnot--Caratheodory metric (a Carnot--Caratheodory metric is a special kind of invariant sub-Finsler metric satisfying that for any $\lambda > 0$, the space $\lambda Y$ is isometric to $Y$). \end{theorem} When the limit is compact, Alan Turing solved the finite dimensional case \cite{Tur}, and using Turing's result, Tsachik Gelander solved the infinite dimensional case \cite{Gel}. \begin{theorem} \rm (Turing) Let $X_i$ be a sequence of almost homogeneous spaces. If the sequence $X_i$ converges in the Gromov--Hausdorff sense to a compact finite dimensional space $X$, then $X$ is a finite dimensional torus equipped with an invariant Finsler metric. \end{theorem} \begin{theorem} \rm (Gelander) Let $X_i$ be a sequence of almost homogeneous spaces. If the sequence $X_i$ converges in the Gromov--Hausdorff sense to a compact space $X$, then $X$ is a (possibly infinite dimensional) torus equipped with an invariant metric. \end{theorem} The main result of this paper deals with the case in which the limit is non-compact. \begin{theorem}\label{MAIN} \rm Let $X_i$ be a sequence of almost homogeneous spaces. If the sequence $X_i$ converges in the pointed Gromov--Hausdorff sense to a space $X$, then $X$ is a nilpotent group equipped with an invariant metric. Furthermore, if $X$ is semi-locally-simply-connected, then it is a Lie group equipped with a Finsler or sub-Finsler metric, and for large enough $i$ there are subgroups $\Lambda_i \leq \pi_1(X_i)$ with surjective morphisms \begin{equation}\label{Li} \Lambda_i \twoheadrightarrow \pi_1(X). \end{equation} \end{theorem} \begin{remark}\label{Montzip} \rm The hypothesis of $X$ being semi-locally-simply-connected can be replaced by $X$ having finite topological dimension, because of the following result (solution to Hilbert's fifth problem) by Deane Montgomery and Leo Zippin \cite{MZ}. \end{remark} \begin{theorem}\label{MZHil} \rm (Montgomery--Zippin) Let $X$ be a homogeneous proper geodesic space. If $X$ has finite topological dimension, then it is homeomorphic to a topological manifold, and its isometry group is a Lie group. \end{theorem} \subsection{Lower semi-continuity of $\pi_1$} Under the conditions of Theorem \ref{MAIN}, Equation \ref{Li} implies that if the spaces $X_i$ are simply connected, and $X$ is semi-locally-simply-connected, then $X$ is simply connected as well. This is an instance of a more general phenomenon (see \cite{GromovMS}, Section 3E). \begin{theorem}\label{LSOFFGC} \rm (Folklore) Let $X_i$ be a sequence of compact geodesic spaces converging to a compact semi-locally-simply-connected space $X$. Then for large enough $i$, there are surjective morphisms \begin{equation} \pi_1 (X_i) \twoheadrightarrow \pi_1 (X). \end{equation} \end{theorem} This property is further studied by Christina Sormani and Guofang Wei in \cite{SWHau}, \cite{SWUni}, \cite{SWCov}. This result fails if the limit is not compact (see \cite{SWUni}, Example 1.2), or not semi-locally-simply-connected (see \cite{SWHau}, Example 2.6). The following example shows that the conclusion of Theorem \ref{MAIN} fails if one works with homogeneous spaces instead of almost homogeneous spaces. \begin{example}\label{BergZ} \rm Let $Y$ be $\mathbb{S}^1$ with its standard metric of length $2 \pi$ and $Z_i $ be $\mathbb{S}^3$ with the round (bi-invariant) metric of constant curvature $1/i$. Let $X_i $ be the quotient $(Y \times Z_i ) / \mathbb{S}^1$ where $\mathbb{S}^1$ acts on $Y \times Z_i$ as follows: $$ z(w,q) = (wz ^{-1}, z q): z, w \in \mathbb{S}^1, q \in \mathbb{S}^3. $$ Then $X_i$ is isometric to $\mathbb{S}^3$ equipped with a re-scaled Berger metric. The sequence $X_i$ consists of simply connected homogeneous spaces, but its pointed Gromov--Hausdorff limit is $\mathbb{S}^1 \times \mathbb{R}^2$, which is not simply connected. \end{example} Under the hypotheses of Theorem \ref{MAIN}, assuming $X$ is semi-locally-simply-connected, one may wonder whether one could obtain the conclusion of Theorem \ref{LSOFFGC}, which is slightly stronger than Equation \ref{Li}. The following example shows that it is not the case. \begin{example}\label{BergZ} \rm Define a ``dot product'' $\mathbb{R}^4 \times \mathbb{R}^4 \to \mathbb{R}^6$ in $\mathbb{R}^4$ as \[ (a \cdot x) : = (a_1x_2, a_1x_3, a_1x_4, a_2x_3, a_2x_4, a_3x_4), \text{ } a,x \in \mathbb{R}^4 . \] With it, define a group structure on $\mathbb{R}^4 \times \mathbb{R}^6$ as \[ (a,b) \cdot (x,y) : = (a+x , b+y+ (a\cdot x)), \text{ }a,x \in \mathbb{R}^4 , \text{ } b,y \in \mathbb{R}^6 . \] Let $G$ be the above group equipped with a left invariant Riemannian metric. For each $i$, define the subgroups $K \leq K_i \triangleleft G_i \leq G \cong \mathbb{R}^4 \times \mathbb{R}^6 $ as: \begin{eqnarray*} K & : = & \{ 0 \} \times \mathbb{Z}^6 \\ K_i & : = & ( i \mathbb{Z}^4 ) \times \mathbb{Z}^6 \\ G_i & : = & \left( \frac{1}{i} \mathbb{Z}^4 \right) \times \left( \frac{1}{i^2} \mathbb{Z}^6 \right) . \end{eqnarray*} The sequence $X_i : = G / K_i$ converges in the pointed Gromov--Hausdorff sense to $X:= G/K$. Since the sequence of finite groups $G_i / K_i$ acts on $X_i$ with diam$(X_i/(G_i/K_i))\to 0$, the sequence $X_i$ consists of almost homogeneous spaces. A direct computation shows that the abelianization of $\pi_1(X_i ) = K_i$ is isomorphic to \[ \mathbb{Z}^4 \oplus \left( \mathbb{Z} / i^2 \mathbb{Z} \right)^6 . \] Since $\pi_1(X) = K \cong \mathbb{Z}^6$, for no $i$ there are surjective morphisms \[ \pi_1(X_i ) \twoheadrightarrow \pi_1(X).\] \end{example} \subsection{Existence of the limit} Given a sequence $X_i$ of almost homogeneous spaces, one may wonder which conditions guarantee the existence of a convergent subsequence. If the spaces $X_i$ are $n$-dimensional Riemannian manifolds with a uniform lower Ricci curvature bound, Gromov compactness criterion implies the existence of a convergent subsequence \cite{GromovMS}. Itai Benjamini, Hilary Finucane and Romain Tessera found another sufficient condition for this partial limit to exist when the spaces $X_i$ are graphs \cite{BFT}. \begin{theorem}\label{BFTExist} \rm (Benjamini--Finucane--Tessera) Let $D_i \leq \Delta _i$ be two sequences going to infnity, and let $X_i$ be a sequence of graphs. Assume there is a sequence of discrete groups $G_i \leq Iso (X_i)$ acting transitively on the sets of vertices. If the balls of radius $D_i$ in $X_i$ satisfy $$\vert B^{X_i}(x,D_i)\vert =O(D_i^q) $$ for some $q>0$, then the sequence of almost homogeneous spaces \[ \dfrac{1}{\Delta_i} X_i \] has a subsequence converging the pointed Gromov--Hausdorff sense to a nilpotent Lie group. \end{theorem} \begin{remark} \rm Recently, Romain Tessera and Matthew Tointon showed that the hypothesis in Theorem \ref{BFTExist} of the groups $G_i$ being discrete can be removed \cite{TT}. Moreover, one could define a sequence of \textit{weakly almost homogeneous spaces} to be a sequence of proper geodesic spaces $X_i$ with groups of isometries $G_i \leq Iso (X_i)$ acting with discrete orbits and such that diam$(X_i / G_i ) \to 0$. Their results imply that Theorem \ref{MAIN} holds under the weaker assumption that the spaces $X_i$ are weakly almost homogeneous. \end{remark} \subsection{Further problems} There are two natural strengthenings of Theorem \ref{MAIN}. One could ask if a weaker conclusion holds if one removes the hypotheses of the groups $G_i$ acting almost transitively, or the limit $X$ being semi-locally-simply-connected. \begin{con} \rm Let $X_i$ be a sequence of simply connected proper geodesic spaces. Assume there is a sequence of discrete groups of isometries $G_i \leq Iso (X_i)$ with diam$(X_i/G_i) \leq C$ for some $C>0$. If $X_i$ converges in the pointed Gromov--Hausdorff sense to a semi-locally-simply-connected space $X$, then $X$ is simply connected. \end{con} \begin{con} \rm Let $X_i$ be a sequence of simply connected almost homogeneous spaces. If $X_i$ converges in the pointed Gromov--Hausdorff sense to a space $X$, then $X$ is simply connected. \end{con} \subsection{Summary} The proof of Theorem \ref{MAIN} consists of three parts. In the first part, we show that $X$ is a nilpotent group equipped with an invariant metric (Theorem \ref{PI}). In the second part, we show that if $X$ is semi-locally-simply-connected, then it is a Lie group (Theorem \ref{PII}). In the third part, we show that if $X$ is a Lie group, then for large enough $i$, there are subgroups $\Lambda_i \leq \pi_1(X_i)$ with surjective morphisms $\Lambda_i \twoheadrightarrow \pi_1 (X)$ (Theorem \ref{PIII}). In Section \ref{Prelim}, we present the relevant definitions and preliminary results required for the proof of Theorem \ref{MAIN}. In Section \ref{PartI} we prove Theorem \ref{PI} by repeated applications of a Margulis Lemma by Breuillard--Green--Tao \cite{BGT}, Gleason--Yamabe structure theory of locally compact groups (\cite{Gleason}, \cite{Yamabe}), and a result of Berestovskii about groups of isometries of homogeneous spaces \cite{BerBus}. In Section \ref{PartII} we prove Theorem \ref{PII} using elementary Lie theory and algebraic topology techniques. Sections \ref{Nilsection} to \ref{Rootsection} contain the proof of Theorem \ref{PIII}, finishing the proof of Theorem \ref{MAIN}. In Section \ref{Nilsection}, we reduce Theorem \ref{PIII} to the case when the groups $G_i$ are uniformly almost nilpotent (see Lemma \ref{anilLie}). We use again the Margulis Lemma of Breuillard--Green--Tao \cite{BGT}, paired with a local uniform doubling of $X$ by Nagel--Stein--Wainger \cite{NSW}. After this reduction, in Section \ref{FreeSection}, we use commutator estimates similar to the ones in (\cite{BK}, \cite{GrAF}) to prove that the groups $G_i$ act ``almost by translations'' on the spaces $X_i$. In Section \ref{Torsection}, we use the escape norm from \cite{BGT} to find small normal subgroups $W_i \triangleleft G_i$ with the property that the spaces $X_i$ and $X_i/ W_i$ are Gromov--Hausdorff close, and the groups $Z_i := G_i/W_i$ contain large subsets $Y_i$ without non-trivial subgroups. In Section \ref{Nilprogsection}, we use the space $X$ as a model (as defined by Hrushovski \cite{Hr}) for the algebraic ultralimit \[Y := \lim\limits_{i \to \alpha} Y_i.\] This enables us to find large nice subsets $P_i$ of $Z_i$ (nilprogressions in $C_0$-regular form, as defined by Breuillard--Green--Tao \cite{BGT}). In Section \ref{Globalsection}, we use the Malcev Embedding Theorem \cite{Malcev} to find groups $\Gamma_{P_i}$ isomorphic to lattices in simply connected Lie groups, with isometric actions $$ \Phi_i: \Gamma_{P_i} \to Iso ( X_i /W_i).$$ Using elementary algebraic topology, we show that the kernels $Ker(\Phi_i)$ of those actions are isomorphic to quotients of $\pi_1(X_i).$ Finally, in section \ref{Rootsection}, we find subgroups of $Ker (\Phi_i)$ isomorphic to $\pi_1 (X)$, finishing the proof of Theorem \ref{PIII}, and consequently Theorem \ref{MAIN}. \subsection{Acknowledgments} The author would like to thank Vladimir Finkelshtein, Enrico LeDonne, Adriana Ortiz, Anton Petrunin, and Burkhard Wilking for helpful comments and discussions. This research was supported in part by the International Centre for Theoretical Sciences (ICTS) during a visit for participating in the program - Probabilistic Methods in Negative Curvature (Code: ICTS/pmnc 2019/03). \section{Preliminaries}\label{Prelim} \subsection{Notation} For $H,K$ subgroups of a group $G$, we define their \textit{commutator subgroup} $[H,K]$ to be the group generated by the elements $[h,k]:=hkh^{-1}k^{-1}$ with $h \in H, k \in K$. Define $G^{(0)} $ as $G$, and $G^{(j+1)}$ inductively as $G^{(j+1)} := [G^{(j)}, G] $. If $G^{(s)} = \{ e\}$ for some $s \in \mathbb{N}$, we say that $G$ is \textit{nilpotent of step} $\leq s$. For any element $g$ in a group $G$, we denote by $L_g$ the left shift $G \to G$ given by $L_g(h) = gh$. If $G$ is abelian, we may denote $L_g$ by $+g$. We say that a set $A \subset G$ is \textit{symmetric} if $A=A^{-1} $ and $e \in A$. For subsets $A_1, \ldots, A_k \subset G$, we denote by $A_1 \cdots A_k $ the set of all products $$ \{ a_1 \cdots a_k \vert a_i \in A_i \} \subset G, $$ and by $A_1 \times \ldots \times A_k $ the set of all sequences $$ \{ (a_1, \ldots, a_k) \vert a_i \in A_i \} \subset G^k. $$ If $A_i = A $ for $i = 1, \ldots , k$, we will also denote $ A_1 \cdots A_k $ by $A^k$, and $ A_1 \times \ldots \times A_k $ by $A^{ \times k }$.\\ Let $Y$ be a topological space and $\beta, \gamma : [0,1] \to Y$ two curves. We denote by $\overline{\beta} : [0,1] \to Y$ the curve given by $\overline{\beta} (t) = \beta (1-t) $. And if $\beta(1)= \gamma (0)$, we denote by $\beta \ast \gamma : [0,1] \to Y$ the concatenation \begin{center} $ \beta \ast \gamma (t) = \begin{cases} \beta (2t) & \text{ if } t \leq 1/2 \\ \gamma (2t-1) & \text{ if }t \geq 1/2. \end{cases} $ \end{center} If $\beta(1)\neq \gamma (0)$, we say that $\beta \ast \gamma$ is undefined. We call $\beta \ast \gamma$ the \textit{concatenation} of $\beta$ and $\gamma$. We will write $\beta \simeq \gamma$ whenever $\beta $ and $\gamma$ are homotopic relative to their endpoints. If $Y$ is a geodesic space, we say that a curve is an $\varepsilon$\textit{-lasso nailed at }$\beta (0)$ if it is of the form $ \beta \ast \gamma \ast \overline{\beta } $, with $\beta (1) = \gamma (0)$, and $\gamma$ a loop contained in a closed ball of radius $\varepsilon$.\\ In a metric space $(Y,d)$, we will denote the closed ball of center $q\in Y$ and radius $r > 0$ as $$B_{d}^Y(q,r) : = \{ y \in Y \vert d(y, q) \leq r \}. $$ We will sometimes omit $d$ or $Y$ and write $B(q,r)$ if the metric space we are considering is clear from the context. \subsection{Uniform distance} \begin{definition} \rm Let $A, B$ be metric spaces and $f,h : A \to B$ a function. For a subset $C \subset A $, we define the \textit{uniform distance} between $f$ and $h$ in $C$ as $$ d_U(f,h; C) : = \sup\limits_{c \in C} d( f(c), h(c) ).$$ \end{definition} \begin{definition} \rm Let $A, B$ be two metric spaces and $f: A \to B$ a function. We define the \textit{distortion} of $f$ as $$ Dis(f) : = \sup\limits_{a_1, a_2 \in A} \vert d(f(a_1), f(a_2)) - d (a_1, a_2) \vert . $$ \end{definition} The following proposition follows immediately from the triangle inequality in the spaces $B_1$ and $B_2$. \begin{proposition}\label{UnifProp} \rm Let $A, B_1, B_2$ be metric spaces, $f, g, h : A \to B_1$, $f_1, g_1 : B_1 \to B_2$, and $C \subset A$. Then \begin{center} $ d_U(f,g; C) \leq d_U(f,h;C) + d_U(h,g; C) , $ $ d_U(f_1f ,g_1g; C) \leq d_U(f,g; C) + d_U(f_1,g_1;g( C)) + Dis( f_1 \vert_{f(C) \cup g(C)} ) . $ \end{center} \end{proposition} \subsection{Gromov--Hausdorff convergence} In this subsection we present the basic results about Gromov--Hausdorff convergence. We refer the reader to (\cite{BBI}, Chapter 7) for proofs and further discussion. \begin{definition} \rm Let $A,B $ be subsets of a metric space $(Z,d)$. The \textit{Hausdorff distance} between $A$ and $B$ is defined as the infimum of the numbers $r>0$ such that for all $a_0 \in A$, $b_0 \in B$ there are $a_1 \in A$, $ b_1 \in B $ with $d(a_0,b_0)<r$ and $ d(a_1, b_1)<r$. \end{definition} \begin{definition} \rm Let $A,B $ be metric spaces, and $f: A \to B$. We say that $f$ is an $\varepsilon$\textit{-approximation} if $Dis (f) \leq \varepsilon$, and the Hausdorff distance between $f(A)$ and $B$ is at most $\varepsilon$. \end{definition} \begin{definition} \rm Let $X_i$ be a sequence of proper metric spaces. We say that $X_i$ \textit{converges in the Gromov--Hausdorff sense} to a complete metric space $X$ if there is a sequence of $\varepsilon_i$-approximations $f_i : X_i \to X$ with $\varepsilon_i \to 0$ as $i \to \infty$. \end{definition} \begin{theorem}\label{gromov--hausdorff--distance} \rm (Gromov) In the class of proper metric spaces modulo isometry, there is a metric $d_{GH}$ such that for any pair of spaces $X$, $Y$, the following holds: \begin{itemize} \item If $d_{GH}(X,Y) \leq \varepsilon $, then there is a $10\varepsilon$-approximation $X \to Y$. \item If there is an $\varepsilon$-approximation $X \to Y$, then $d_{GH}(X,Y) \leq 10 \varepsilon$. \end{itemize} \end{theorem} \begin{definition} \rm Let $(X_i, p_i)$ be a sequence of pointed proper metric spaces. We say that $(X_i, p_i)$ converges in the pointed Gromov--Hausdorff sense to a pointed complete metric space $(X,p)$ if there are two sequences of functions $f_i : X_i \to X$ and $h_i : X \to X_i$ with $f_i(p_i) =p$, $h_i (p) = p_i$ and such that for all $R > 0$, we have \begin{equation}\label{pGH1} \lim_{i \to \infty} Dis \left( f_i \vert_{B(p_i, R)} \right) = \lim_{i \to \infty} Dis \left( h_i \vert_{B(p, R)} \right) = 0 \end{equation} and \begin{equation}\label{pGH2} \lim_{i \to \infty} d_U \left( h_if_i, Id_{X_i} ; B^{X_i}(p_i,R) \right) = \lim_{i \to \infty} d_U \left( f_ih_i, Id_{X} ; B^{X}(p,R) \right) = 0. \end{equation} \end{definition} \begin{remark}\label{fihi} \rm Throughout this paper, whenever we have a sequence of proper metric spaces $(X_i, p_i)$ converging in the pointed Gromov--Hausdorff sense to a pointed metric space $(X,p) $, we assume they come equipped with functions $f_i : X_i \to X$ and $h_i : X \to X_i$ with $f_i (p_i)=p$, $h_i(p)=p_i$, and satisfying equations \ref{pGH1} and \ref{pGH2} for all $R > 0$. \end{remark} \begin{lemma} \rm If $X_i $ is a sequence of proper metric spaces converging in the Gromov--Hausdorff sense to a complete space $X$, then for all $p \in X$ there is a sequence $p_i \in X_i $ such that the sequence $(X_i, p_i)$ converges in the pointed Gromov--Hausdorff sense to $(X,p)$. \end{lemma} \begin{lemma}\label{proper-geodesic} \rm If $(X_i, p_i) $ is a sequence of proper geodesic spaces converging in the pointed Gromov--Hausdorff sense to a complete space $(X,p)$, then $X$ is a proper geodesic space. \end{lemma} \subsection{Groups of isometries} Here we summarize the main results about groups acting on metric spaces we will need. \begin{remark}\label{topology-basis} \rm Throughout this paper, for a proper metric space $X$, we use the compact-open topology on the group of isometries $Iso (X)$. With this topology, $Iso(X)$ is a locally compact, first countable, Hausdorff group. For any $p\in X$, a basis of compact neighborhoods of the identity with respect to this topology is given by the family of sets \[ U_{R,\varepsilon}^X : = \{ g \in Iso (X) \vert d_U(g,Id_X; B(p,R))\leq \varepsilon \} . \] With respect to this topology, a group $\Gamma \leq Iso(X)$ is \textit{discrete} if and only if its action has discrete orbits and finite stabilizers. \end{remark} \begin{lemma}\label{short-generators} \rm Let $(Y,q)$ be a pointed proper geodesic space, and $G \leq Iso (Y)$ a closed subgroup. If diam$(Y/G) < \infty$, then the set \[ S : = \{ g \in G \vert d(gq,q ) \leq 3\cdot \text{diam}(Y/G) \} \] generates $G$. \end{lemma} \begin{proof} Let $g \in G$. Since $Y$ is geodesic, there is a sequence of points $q=q_0, q_1, \ldots , q_k = gq \in Y$ satisfying that $d(q_j,q_{j-1}) \leq $diam$(Y/G)$ for each $j$. Choose a sequence $g_1, \ldots, g_{k-1} \in G$ such that $d(g_jq,q_j)\leq $diam$(Y/G)$ for each $j$, and let $g_k : = g$. From the triangle inequality, for each $j$ we have $g_{j-1}^{-1}g_j \in S$. Then \[ g = g_1 ( g_1^{-1}g_2) \cdots (g_{k-1}^{-1}g_k) \in S^k . \] \end{proof} \begin{lemma}\label{diameter-subgroup} \rm Let $X$ be a proper geodesic space, and $H \leq G \leq Iso (X)$ be closed subgroups. If diam$(X/G) $ and $[G:H]$ are finite, then \[ \text{diam}(X/H) \leq 3 [G: H ]\cdot \text{diam}(X/G) . \] \end{lemma} \begin{proof} Let $x \in X$, and define \[ S : = \{ g \in G \vert d(gx,x) \leq 3 \cdot \text{diam} (X/G) \}, \] which generates $G$ by Lemma \ref{short-generators}. If we have $S^{k+1} H = S^k H$ for some $k \in \mathbb{N}$, then an inductive argument implies that \[S^kH = \bigcup_{n \in \mathbb{N}} S^n H = G .\] Since there are only $[G:H]$ cosets, then $S^{k+1}H = S^k H$ for $k \geq [G:H]-1$. This implies that $S^{[G:H]-1}$ intersects each $H$-coset. For any $y \in X$, there is $g \in G$ with $d(x,gy)\leq $ diam$(X/G)$, and by the above analysis, there is $u \in S^{[G:H]-1}$ with $u^{-1}g \in H$. This implies that \begin{eqnarray*} d(x,u^{-1}gy) & = & d(ux,gy)\\ & \leq & d(ux,x) + d(x,gy)\\ & \leq & 3 ([G:H]-1) \cdot \text{diam}(X/G) + \text{diam}(X/G). \end{eqnarray*} Since $x,y \in X$ were arbitrary, the lemma follows. \end{proof} \begin{lemma}\label{normal-stabilizers} \rm Let $(Y,q)$ be a pointed proper geodesic space, $G \leq Iso (Y)$ a closed group acting transitively, and $H \triangleleft G$ a normal subgroup. If for some $\varepsilon > 0 $, \[ H \subset \{ g \in G \vert d(gq,q) \leq \varepsilon \}, \] then \[ H \subset \{ g \in G \vert d(gy,y) \leq \varepsilon \text{ for all }y \in Y \}. \] In particular, if $H$ is contained in the group $\{ g \in G \vert gq=q \}$, then $H$ is trivial. \end{lemma} \begin{proof} If $ y \in Y$, then there is $g \in G$ with $gq=y$. Since $H$ is normal in $G$, then for any $h \in H$ we have $g^{-1}hg \in H$, and \[ d(hy,y) = d(hgq, gq) = d (g^{-1}h g q,q) \leq \varepsilon. \] \end{proof} The following result by Valerii Berestovskii concerns groups that act transitively on geodesic spaces (\cite{BerBus}, Theorem 1). \begin{theorem}\label{berestovskii-open} \rm (Berestovskii) Let $X$ be a proper geodesic space, and $G$ a closed subgroup of $Iso (X)$ acting transitively. If $\mathcal{O} \leq G$ is an open subgroup, then $\mathcal{O}$ also acts transitively on $X$. \end{theorem} \subsection{Hilbert's fifth problem and locally compact groups} Hilbert's fifth problem consists of understanding which locally compact Hausdorff groups are Lie groups. One satisfactory answer is Theorem \ref{MZHil}. On a different direction, Andrew Gleason and Hidehiko Yamabe found that any locally compact Hausdorff group is not far from being a Lie group (\cite{Gleason}, \cite{Yamabe}). \begin{theorem}\label{gleason-yamabe} \rm (Gleason--Yamabe) Let $G$ be a locally compact Hausdorff group. Then there is an open subgroup $\mathcal{O}\leq G$ with the following property: for any open neighborhood of the identity $U \subset \mathcal{O}$, there is a compact normal subgroup $K \triangleleft \mathcal{O}$ with $K \subset U$ such that $\mathcal{O}/K$ is a connected Lie group. \end{theorem} The following result by Victor Glushkov implies that the set of compact normal subgroups with the property that the corresponding quotient is a connected Lie group is closed under finite intersections \cite{Glu}. \begin{theorem}\label{glushkov} \rm (Glushkov) Let $\mathcal{O}$ be a locally compact Hausdorff group, and $K_1, K_2$ compact normal subgroups such that both $\mathcal{O}/K_1$ and $\mathcal{O}/K_2$ are connected Lie groups. Then $\mathcal{O}/(K_1 \cap K_2)$ is a connected Lie group. \end{theorem} \begin{corollary}\label{glushkov-corollary} \rm Let $\mathcal{O}$ be a connected, locally compact, first countable, Hausdorff group that is not a Lie group. Then it contains a sequence of compact, normal, non-trivial subgroups $K_1 \geq K_2 \geq \ldots$ such that \[ \bigcap_{j=1}^{\infty} K_j = 0, \] and $\mathcal{O}/K_j$ is a connected Lie group for all $j$. \end{corollary} \subsection{Lie groups} In this section we introduce the tools from Lie theory we will need. \begin{definition} \rm Let $\mathfrak{g}$ be a nilpotent Lie algebra. We say that an ordered basis $\{ v_1, \ldots , v_n \} \hookrightarrow \mathfrak{g}$ is a \textit{strong Malcev basis} if for all $k \in \{1, \ldots ,n \} $, the vector subspace $J_j \leq \mathfrak{g}$ generated by $\{ v_{k+1}, \ldots , v_n \}$ is an ideal, and $v_k + J_k$ is in the center of $\mathfrak{g}/J_k$. \end{definition} The following theorem encompasses the results and discussion in (\cite{Cor}, Section 1.2). \begin{theorem}\label{simply-connected-nilpotent} \rm Let $G$ be a simply connected nilpotent Lie group and $\mathfrak{g} $ its Lie algebra. Let $Z$ be the center of $G$ and $\mathfrak{z}$ the center of $\mathfrak{g}$. Then \begin{itemize} \item $\exp : \mathfrak{g} \to G$ is a diffeomorphism. \item $Z= \exp ( \mathfrak{z} )$. \item Any compact subgroup of $G$ is trivial. \item If $\{ v_1, \ldots , v_n \} \hookrightarrow \mathfrak{g}$ is a strong Malcev basis, then the map $ \psi : \mathbb{R}^n \to G$ given by \[ \psi (x_1, \ldots , x_n) := \exp(x_1v_1) \cdots \exp (x_n v_n) \] is a diffeomorphism. \item After identifying $\mathfrak{g} $ with $\mathbb{R}^n$ through a strong Malcev basis, the maps $ ( \log \circ $ $ \psi ) : \mathfrak{ g} \to \mathfrak{g}$ and $ ( \psi ^{-1} \circ \exp ) : \mathfrak{ g} \to \mathfrak{g}$ are polynomial of degree bounded by a number depending only on $n$. \end{itemize} \end{theorem} \begin{corollary}\label{aspherical} \rm Let $G$ be a connected nilpotent Lie group. Then $G$ is aspherical. That is, $\pi_k(X) =0$ for all $k \geq 2$. \end{corollary} \begin{proof} For smooth manifolds, being aspherical is equivalent to having a contractible universal cover. By Theorem \ref{simply-connected-nilpotent}, the universal cover $\tilde{G} \to G$, being a simply connected nilpotent Lie group, is contractible. \end{proof} \begin{lemma}\label{discrete-normal-central} \rm Let $G$ be a connected Lie group, and $\Lambda \leq G$ a discrete normal subgroup. Then $\Lambda$ is central in $G$. \end{lemma} \begin{proof} For fixed $\lambda \in \Lambda$, the map $ \phi_{\lambda}:G \to \Lambda$ given by $ \phi_{\lambda}(g) := g \lambda g^{-1}$ is a continuous map from a connected space to a discrete space, hence it is constant. Since $\phi_{\lambda}(e) = \lambda$, we have $g \lambda = \lambda g$ for all $g \in G$. Since $\lambda \in \Lambda$ was arbitrary, the lemma follows. \end{proof} \begin{lemma}\label{compact-nilpotent-1} \rm Let $G$ be a connected nilpotent Lie group, and $K \leq G$ a compact subgroup. Then $K$ is central in $G$. \end{lemma} \begin{proof} Let $\rho : \tilde{G} \to G$ be the universal cover of $G$ and $Z\leq \tilde{G}$ be the center of $\tilde{G}$. By Theorem \ref{simply-connected-nilpotent}, $G/Z$ is a simply connected nilpotent group. Since $\rho (Z)$ is contained in the center of $G$, the lemma is a consequence of the following claim. \begin{center} \textbf{Claim:} The preimage $\rho^{-1}(K)$ is contained in $Z$ \end{center} By Lemma \ref{discrete-normal-central}, $\Lambda $ is contained in $Z$, and we have a projection from $G= \tilde{G} / \Lambda$ to $\tilde{G}/Z$ sending $K$ to the compact group $\rho^{-1}(K)/(Z\cap \rho ^{-1}(K)) \leq \tilde{G}/Z$. By Theorem \ref{simply-connected-nilpotent}, $\rho^{-1}(K)/(Z\cap \rho ^{-1}(K))$ is trivial and the claim follows. \end{proof} \begin{corollary}\label{compact-nilpotent-2} \rm A connected compact nilpotent Lie group is abelian. \end{corollary} \begin{definition} \rm We say that a continuous map $Y \to Z$ between path connected topological spaces \textit{has no content} or \textit{has trivial content} if the induced map $\pi_1(Y) \to \pi_1 (Z) $ is trivial. Otherwise, we say that the map \textit{has non-trivial content}. \end{definition} \begin{lemma}\label{LongHomo} \rm Let $G$ be a connected nilpotent Lie group and $K$ a compact subgroup such that its connected component $K_0$ is nontrivial. Then the inclusion $K_0 \to G$ has non-trivial content. \end{lemma} \begin{proof} Since $K_0$ is a connected compact nilpotent Lie group, it is homeomorphic to a torus, and $\pi_1(K_0)$ is non-trivial. From the long exact homotopy sequence of the fibration $ G \to G/K_0$, we extract the portion $$ \pi_2 (G/K_0) \to \pi_1 (K_0) \to \pi_1 (G) $$ By Lemma \ref{compact-nilpotent-1}, $K_0$ is central in $G$. Hence, $G/K_0$ is a connected nilpotent Lie group, and aspherical by Corollary \ref{aspherical}. This implies that $\pi_2(G/K_0)=0$, and the map $\pi_1 (K_0) \to \pi_1 (G) $ is non-trivial. \end{proof} \begin{lemma}\label{FinalCover} \rm Let $G$, $\tilde{G}$ be connected Lie groups such that $\tilde{G}$ is a discrete extension of $G$ (i.e. there is a surjective continuous morphism $f: \tilde{G} \to G$ with discrete kernel). Assume $G$ and $\tilde{G}$ are equipped with invariant geodesic metrics for which $f$ is a local isometry. Let $\delta > 0$ be such that $B^{G}(e,\delta)$ contains no non-trivial subgroups, and the inclusion $B^{G}(e, \delta) \to G$ has no content. Then $B^{\tilde{G}}(e,\delta)$ contains no non-trivial subgroups and the inclusion $B^{\tilde{G}}(e, \delta) \to \tilde{G}$ has no content. \end{lemma} \begin{proof} If there is a group $H \subset B^{\tilde{G}}(p, \delta)$, then its image $f(H)$ is a subgroup of $G$ contained in $B^{G}(e, \delta)$, so $f(H) = \{e \} \leq G$. If there is a non-trivial element $h \in H \backslash \{e \}$, we can take a shortest path $\gamma : [0,1] \to \tilde{G}$ from $e$ to $h$. The projection $f \circ \gamma$ would be a non-contractible loop in $G$ contained in $B^{G}(e, \delta )$, contradicting the hypothesis that $B^{G}(e, \delta) \to G$ has no content. Also, we can consider the commutative diagram \[ \begin{tikzcd}[column sep=3em] B^{\tilde{G}}(e, \delta) \arrow{r}{ } \arrow{d}{} & \tilde{G} \arrow{d}{f} \\ B^{G}(e,\delta) \arrow{r}{ } & G \end{tikzcd} \] Since $ f$ is a covering map, the right vertical arrow induces an injective map at the level of fundamental groups, and the bottom horizontal arrow has no content by hypothesis. Therefore the top horizontal arrow has trivial content as well. \end{proof} \subsection{Lifting curves} In this section we present some techniques we use to lift curves from one space to another. \begin{lemma}\label{LipApprox} \rm Let $Y$ be a proper geodesic space, and $\gamma : [0,1] \to Y$ be a continuous curve. Then for every $\varepsilon > 0$ there is a Lipschitz curve $\beta : [0,1] \to Y$ with $\beta (0)= \gamma (0)$, $\beta (1) = \gamma (1)$, and $$d_U(\gamma , \beta , [0,1]) \leq \varepsilon .$$ \end{lemma} \begin{proof} Let $k > 0$ such that $\vert x_1 - x_2 \vert \leq 3/k$ implies $ d (\gamma (x_1), \gamma (x_2)) \leq \varepsilon /10$ for $x_1, x_2 \in [0,1]$. Then choose $\beta : [0,1]\to Y$ to be such that $\beta \vert_{[j/k, (j+1)/k]}$ is a constant speed minimizing curve between $ \gamma (j/k) $ and $\gamma ( (j+1)/k)$ for $j = 0, 1, \ldots , k-1$. It is easy to see that $\beta$ satisfies the desired properties. \end{proof} \begin{lemma}\label{ShortLoopNTI} \rm Let $f:Y \to Z$ be a continuous map between proper geodesic spaces. Assume that $Y$ is semi-locally-simply-connected, and the composition $B(y,r) \to Y \to Z$ has non-trivial content for some $y \in Y$, $r > 0$. Then there is a loop $\beta : [0,1] \to Y$ based at $y$ of length $\leq 3r $ such that $f \circ \beta$ is non-contractible in $Z$. \end{lemma} \begin{proof} Since $Y$ is semi-locally-simply-connected, there is $\varepsilon >0$ such that any two closed curves $\beta_1, \beta_2 : [0,1] \to B(y,2r)$ with $d_U(\beta_1 , \beta_2, [0,1] ) \leq \varepsilon$, are homotopic to each other in $Y$. By hypothesis, there is a loop $\gamma_1 :[0,1] \to Y $ based at $y$ whose image is contained in $B(y,r)$ and such that $f \circ \gamma_1$ is non-contractible in $Z$. By Lemma \ref{LipApprox}, there is a Lipschitz curve $\gamma : [0,1] \to B(y,(1.1)r) $ such that $f \circ \gamma $ is non-contractible in $Z$. Let $L$ be the Lipschitz constant of $\gamma$, and pick $\ell \in \mathbb{N}$ with $\ell \geq 2L /r$. For each $j = 0, 1 \ldots , \ell$, set $\sigma_{j}$ to be a minimizing path from $y$ to $\gamma (j/\ell)$, and define $\beta_j $ as the concatenation $\overline{\sigma}_{j+1} \ast \gamma \vert_{[j/\ell , (j+1 )/ \ell ]} \ast \sigma_j$. Since $\gamma$ is homotopic to the concatenation $\beta_{\ell-1} \ast \ldots \ast \beta_0$, at least one of the curves $\beta_j$ is satisfies that $f \circ \beta_j$ is non-contractible in $Z$, and all of them have length $\leq 3r$. \end{proof} \begin{corollary}\label{ShortLoopNT} \rm Let $Y$ be a proper semi-locally-simply-connected geodesic space. Assume the inclusion $B(y,r) \to Y$ has nontrivial content for some $y \in Y$, $r > 0$. Then there is a non-contractible loop based at $y$ of length $\leq 3r $. \end{corollary} \begin{proof} Apply Lemma \ref{ShortLoopNTI} with $f = Id_Y$. \end{proof} \begin{definition} \rm Let $f : Y_1 \to Y_2$ be a map between proper geodesic spaces. We say that $f$ is a \textit{metric submersion} if for every $y \in Y_1, $ and $r > 0$, we have $f(B^{Y_1}(y,r)) = B^{Y_2}(f(y) ,r)$ \end{definition} \begin{lemma}\label{Submer} \rm Let $Y$ be a proper geodesic space, and $\Gamma \leq Iso (Y)$ a closed subgroup. Then the quotient map $f: Y \to Y/ \Gamma$ is a metric submersion. \end{lemma} \begin{proof} Let $y \in Y $, $r > 0 $, and $z \in Y/\Gamma$ with $d_{Y/\Gamma}( f(y), z) \leq r$. Since $\Gamma $ is closed, the orbits are closed, and since $Y$ is proper, there is $z_1 \in f^{-1}(z)$ with $d(y,z_1) = d(f(y), z ) \leq r$. This proves $ B^{Y/ \Gamma}(f(y), r) \subset f(B^Y(y,r)) $. The other contention is immediate from the definition of the metric in $Y/ \Gamma$. \end{proof} \begin{lemma}\label{HoriLift} \rm Let $f: Y_1 \to Y_2$ be a metric submersion between proper geodesic spaces, $\gamma : [0,1] \to Y_2$ a Lipschitz curve, and $q_0 \in f^{-1}(\gamma (0))$. Then there is a curve $\tilde{\gamma} : [0,1] \to Y_1$ with $f \circ \tilde{ \gamma} = \gamma ,$ $\tilde{\gamma}(0) = q_0$, and length$(\tilde{\gamma}) = $length$ (\gamma )$. \end{lemma} \begin{proof} For each $j \in \mathbb{N}$, let $D_j : = \{ 0, \frac{1}{2^j}, \ldots , \frac{2^j-1}{2^j} , 1 \} $, and define $h_j : D_j \to Y_1$ as follows: Let $h_j(0)=q_0$, and inductively, let $h_j(x + 1/2^j)$ be a point in $f^{-1}( \gamma (x + 1/2^j ) )$, such that $$ d_{Y_1} ( h_j ( x + 1/2^j), h_j ( x ) ) = d_{Y_2} ( \gamma ( x + 1/2^j ), \gamma (x) ), \text{ for }x \in D_j \backslash \{ 1 \} . $$ Using Cantor's diagonal argument, we can find a subsequence of $h_j$ that converges for every dyadic rational. Since the maps $h_j$ are uniformly Lipschitz, we can extend this map to a Lipschitz map $\tilde{ \gamma} : [0,1] \to Y_1$. It is easy to check that $\tilde{\gamma}$ satisfies the desired properties. \end{proof} \subsection{Ultralimits} In this section we discuss the ultrafilter tools we will use during the proof of Theorem \ref{PIII}, including metric ultralimits and algebraic ultraproducts. We refer the reader to (\cite{AKP}, Chapter 4) , (\cite{BFT}, Section 2.1), (\cite{BGT}, Appendix A) for proofs and further discussions. \begin{definition} \rm Let $\wp (\mathbb{N})$ denote the power set of the natural numbers and $\alpha : \wp (\mathbb{N}) \to \{ 0,1 \}$ be a function. We say that $\alpha$ is a \textit{non-principal ultrafilter} if it satisfies: \begin{itemize} \item $\alpha (\mathbb{N}) = 1$. \item $\alpha (A \cup B )= \alpha (A) + \alpha (B)$ for all disjoint $A , B \subset \mathbb{N}$. \item $\alpha (F )=0$ for all finite $F \subset \mathbb{N}$. \end{itemize} \end{definition} Using Zorn's Lemma it is not hard to show that nonprincipal ultrafilters exist. We will choose one $(\alpha )$ and fix it for the rest of this paper. Sets $A \subset \mathbb{N}$ with $\alpha (A) =1$ are called $\alpha$-large. For a property $P : \mathbb{N}\to \{0,1\}$, if $\alpha ( P^{-1} (1) ) =1$ we say that ``$P$ holds for $i$ sufficiently $\alpha$-large'', or ``$P$ holds for $i$ sufficiently close to $\alpha$''. \begin{definition} \rm Let $A_i$ be a sequence of sets. In the product $$A^{\prime} : = \prod _{i=1}^{\infty} A_i,$$ we say that two sequences $\{ a_i \} , \{ a_i^{\prime} \}$ are $\alpha $\textit{-equivalent} if $$ \alpha \left( \{ i \vert a_i = a_i^{\prime} \} \right) =1. $$ The set $A^{\prime}$ modulo this equivalence relation is called the \textit{algebraic ultraproduct} of the sets $A_i$ and is denoted by $$A_{\alpha} : = \lim\limits_{i \to \alpha} A_i. $$ \end{definition} If $A_i = \mathbb{R}$ for each $i$, an element in $A_{\alpha}$ is called a \textit{non-standard real number}. For $x= \{ x_i\}$, $y=\{ y_i \}$ non-standard real numbers, we say that $x \leq y$ if $\alpha \left( \{ i \in \mathbb{N} \vert x_i \leq y_i \} \right) =1$. Let $ x_i $ be a sequence in a metric space $X$. We say that the sequence \textit{ultraconverges} to a point $x_{\infty} \in X$ if for every $\varepsilon > 0$, $$ \alpha \left( \{ i \vert d(x_i, x_{\infty}) < \varepsilon \} \right) = 1 .$$ If this is the case, the point $x_{\infty} $ is called the \textit{ultralimit} of the sequence, and we write $x_i \xrightarrow{\alpha} x_{\infty}$ or $$\lim\limits_{i \to \alpha} x_i = x_{\infty}. $$ It is easy to show that if a sequence has an ultralimit, then it is unique. Furthermore, if $X$ is compact, then any sequence in $X$ ultraconverges. \begin{definition} \rm Let $ (X_i, p_i) $ be a sequence of pointed metric spaces. Let $X_{\alpha}^{\prime}$ be the set of sequences $ x_i \in X_i$ such that $$ \sup\limits_{i \in \mathbb{N}} d(x_i, p_i) < \infty.$$ Equip $X_{\alpha}^{\prime}$ with the pseudometric $$d( \{ x_i \}, \{ y_i \} ) := \lim\limits_{i \to \alpha} d(x_i, y_i).$$ Let $X_{\alpha}$ be the metric space corresponding to the pseudometric space $X_{\alpha}^{\prime}$. The pointed metric space $(X_{\alpha}, p_{\alpha} )$, where $p_{\alpha} $ is the class of the sequence $\{ p_i \} $, is called the \textit{metric ultralimit} of the sequence $ (X_i, p_i)$. \end{definition} It is straightforward to show that $X_{\alpha} $ is always a complete metric space. \begin{theorem}\label{Subseq} \rm If a sequence of proper metric spaces $(X_i, p_i)$ converges in the pointed Gromov--Hausdorff sense to a space $(X,p)$, then $(X_{\alpha}, p_{\alpha}) $ and $(X,p)$ are isometric. Conversely, if the sequence $(X_i, p_i)$ is precompact in the pointed Gromov--Hausdorff topology, then there is a subsequence that converges to $(X_{\alpha}, p_{\alpha}) $ in the pointed Gromov--Hausdorff sense. \end{theorem} \begin{remark} \rm To define an algebraic ultraproduct $A_{\alpha} = \lim_{i \to \alpha} A_i $ or a metric ultraproduct $X_{\alpha} = \lim_{i \to \alpha}X_i$, we don't require the sets $A_i $ or $X_i$ to be defined for all $i$, but only for all $i$ in an $\alpha$-large set. \end{remark} \subsection{Approximate isometries, equivariant convergence} If we have a sequence of pointed proper metric spaces $(X_i, p_i)$ converging in the pointed Gromov--Hausdorff sense to a space $(X,p)$, then any sequence of groups of isometries $G_i \leq Iso (X_i)$ converges in a suitable sense to a group $G_{\alpha} \leq Iso (X)$. In this section we present a precise statement of this result. The proofs follow from the discussion in (\cite{GromovPG}, Section 6). \begin{definition} \rm Let $ (X_i, p_i)$ be a sequence of pointed proper metric spaces and $Y_i$ a sequence of proper metric spaces. We say that two sequences of functions $\phi_i, \psi_i : X_i \to Y_i$ are \textit{ultraequivalent} if for every $R>0$, \begin{center} $\lim\limits_{i\to \alpha} d_U ( \phi_i , \psi_i , B(p_i, R) ) =0 $. \end{center} \end{definition} \begin{lemma}\label{ALISON} \rm Let $(X,p)$ be a pointed proper metric space and $\phi_i : X \to X$ a sequence of functions for which the sequence $ d ( \phi_i p, p ) $ is bounded. Assume that for every $R > 0$, $Dis \left( \phi_i \vert_{B(p, R)} \right) \to 0 $ as $i \to \infty$, and consider the map $\phi_{\alpha}:X \to X $ given by $$ \phi_{\alpha} (x) : = \lim\limits_{i \to \alpha} \phi_i (x). $$ Then the sequence $\phi_i$ is ultraequivalent to the constant sequence $\phi_{\alpha}$, and $\phi_{\alpha}$ is an isometric embedding. We call $\phi_{\alpha}$ the \textit{ultralimit} of the sequence $\phi_i$. Furthermore, if there is a sequence $\psi_i: X \to X$ with $Dis \left( \psi_i \vert_{B(p,R)} \right) \to 0$ as $i \to 0$ for all $R>0$, and such that the sequence $\phi_i\psi_i$ is ultraequivalent to the identity $Id_X$, then $\phi_{\alpha} $ is a surjective isometry. \end{lemma} Let $(X_i, p_i)$ be a sequence of proper geodesic spaces, converging in the pointed Gromov--Hausdorff sense to a space $(X, p)$, and consider a sequence of groups of isometries $G_i \leq Iso (X_i)$. We say that a sequence $g_i \in G_i$ is \textit{stable} if the sequence $ d(g_i(p_i), p_i) $ is bounded. The set $G_{\alpha}$ of stable sequences modulo ultraequivalence is called the \textit{equivariant ultralimit } of the sequence $G_i$. By definition, we have sequences of maps $f_i : X_i \to X$ and $h_i : X \to X_i$ with $f_i(p_i ) = p$, $h_i (p) = p_i$, and satisfying equations \ref{pGH1} and \ref{pGH2} for all $R>0$. We say that a sequence $g_i \in G_i$ \textit{ultraconverges} to a map $g: X \to X$ if the sequence of maps $(f_i \circ g_i \circ h_i ) $ is ultraequivalent to the constant sequence $g$. By Lemma \ref{ALISON}, any stable sequence has a unique ultralimit. \begin{theorem}\label{equivariant-ultralimit} \rm (Gromov) Let $(X_i,p_i)$ be a sequence of pointed proper metric spaces converging in the pointed Gromov--Hausdorff sense to a complete space $(X,p)$. Consider a sequence of groups of isometries $G_i \leq Iso (X_i)$ and let $G_{\alpha}$ be their equivariant ultralimit. The map $ \Phi : G_{\alpha} \to Iso (X)$ that sends a sequence to its ultralimit is well defined (that is, the ultralimit doesn't depend on the representative in the equivalence class), is injective (that is, if two stable sequences have the same equivariant ultralimit, then the sequences are ultraequivalent), and has closed image. Furthermore, if diam$(X_i/G_i) \to 0$, then $G_{\alpha} $ acts transitively on $X$. \end{theorem} \begin{remark} \rm The image $\Phi (G_{\alpha})$ may depend on the choice of the functions $f_i, h_i$ used to obtain the ultralimit of a stable sequence $g_i \in G_i$. However, as pointed out in Remark \ref{fihi}, we always assume that the functions $f_i$ and $h_i$ are already given in the hypothesis that the spaces $X_i$ converge to $X$. \end{remark} \begin{remark} \rm The set $G_{\alpha }$ has two equivalent group structures. The one obtained by pulling back the group structure in $Iso(X)$ through $\Phi$: \begin{center} $a \ast b := \Phi^{-1} (\Phi (a) \circ \Phi (b)). $ \end{center} And the one given by the groups $G_i$: \begin{center} $\{ g_i \} \ast \{ g_i^{\prime} \} : = \{ g_i \circ g_i ^{\prime} \} $. \end{center} \end{remark} \begin{remark} \rm The group $G_{\alpha}$ has two equivalent topologies. The one obtained by pulling back the topology of $Iso (X)$ through $\Phi$, yielding the following family of sets as a compact neighborhood basis of the identity: \begin{center} $U_{R, \varepsilon}^{\alpha} : = \Phi ^{-1} \left( U_{R, \varepsilon}^X\right) $. \end{center} And the one given by the groups $G_i$, yielding the following family of sets as a compact neighborhood basis of the identity: \begin{center} $\tilde{U}_{R,\varepsilon}^{\alpha}:= \left\{ \{ g_i \} \in G_{\alpha} \big\vert \text{ } \alpha \left( \left\{ i \vert g_i \in U_{R, \varepsilon} ^{X_i} \right\} \right) =1 \right\} .$ \end{center} \end{remark} \subsection{Local groups} In this section we present the elementary theory of local groups an approximate groups we will use. We refer the reader to (\cite{BGT}, Appendix B) for proofs and further discussion. \begin{definition} \rm Let $(G,e)$ be a pointed topological space. We say that $G$ is a \textit{local group}, if there are continuous maps $(\cdot )^{-1}: G \to G$ and $\cdot : \Omega \to G$ for some $ \Omega \subset G \times G$ such that \begin{itemize} \item $\Omega$ is an open set containing $(G \times \{ e\} ) \cup ( \{ e \} \times G ) $. \item For all $g \in G$, we have $(g,(g)^{-1}), ((g)^{-1},g)\in \Omega$ and $g \cdot e = e \cdot g = g$. \item For all $g \in G$, we have $g \cdot (g)^{-1} = (g)^{-1} \cdot g = e$. \item For all $g,h,k \in G$ such that $(g,h) , (gh,k), (h,k), (g, hk) \in \Omega$, we have $g (hk)= (gh)k$. \end{itemize} \end{definition} \begin{definition} \rm We say that a local group $G$ is a \textit{local Lie group} if it is a smooth manifold, and the maps $(\cdot )^{-1}: G \to G$ and $\cdot : \Omega \to G$ are smooth. \end{definition} \begin{definition} \rm Let $G$ be a local group. We say that a subset $A \subset G$ is \textit{symmetric} if $e \in A$ and $g^{-1} \in A$ for all $g \in A$. \end{definition} \begin{definition} \rm Let $G$ and $H$ be two local groups. We say a function $\phi : G \to H$ is a morphism if the following holds: \begin{itemize} \item $\phi (e) = e$. \item For all $g \in G$, we have $\phi (g^{-1}) = \left[ \phi (g) \right] ^{-1}$. \item If $g,h \in G$ are such that $g\cdot h$ is defined in $G$, then $\phi (g) \cdot \phi (h) $ is defined in $H$, and $\phi (g\cdot h ) = \phi (g) \phi (h)$. \end{itemize} \end{definition} \begin{definition} \rm Let $(G,e$) be a local group, and $U\subset G$ a symmetric subset. We say that $U$ is a \textit{restriction} of $G$ if when restricting $(\cdot )^{-1} $ to $U$, and $\cdot : \Omega \to G$ to $\left[ ( \cdot )^{-1}(U)\right] \cap (U \times U) $, we obtain a local group structure on $U$. \end{definition} \begin{definition} \rm Let $G$ be a local group, and $g_1, \ldots , g_m \in G$. We say that the product $g_1 \cdots g_m$ is \textit{well defined in} $G$ if for each $1 \leq j\leq k \leq m$, we can find an element $g_{[j,k]} \in G$ such that \begin{itemize} \item For all $j \in \{ 1, \ldots , m \} $, we have $g_{[j,j]}= g_j$ \item For all $1 \leq j \leq k \leq \ell \leq m$, the pair $\left( g_{[j,k]} , g_{[k+1,\ell ]} \right) $ lies in $\Omega$, and $g_{[j,k]} \cdot g_{[k+1,\ell ]} = g_{[j, \ell ]}$. \end{itemize} For sets $A_1, \ldots , A_m \in G$, we say that the product $A_1\cdots A_m$ is \textit{well defined} if for all choices of $g_j \in A_j$, the product $g_1 \cdots g_m$ is well defined. \end{definition} \begin{definition} \rm Let $G$ be a local group. We say that a subset $A \subset G$ is a \textit{multiplicative set} if it is symmetric, and $A^{200}$ is well defined in $G$. \end{definition} \begin{definition} \rm We say that a local group $G$ is \textit{cancellative} if the follwing holds: \begin{itemize} \item For all $g,h,k \in G$ such that $gh$ and $gk$ are well defined and equal to each other, we have $h=k$. \item For all $g,h,k \in G$ such that $hg$ and $kg$ are well defined and equal to each other, we have $h=k$. \item For all $g,h \in G$ such that $gh $ and $h^{-1}g^{-1}$ are well defined, then $(gh)^{-1} = h^{-1}g^{-1}$. \end{itemize} \end{definition} \begin{definition} \rm Let $G, G^{\prime}$ be local groups such that $G^{\prime}$ is a restriction of $G$. We say that $G^{\prime}$ is a \textit{sub-local group} of $G$ if there is an open set $V\subset G$ containing $G^{\prime}$ with the property that for all $a,b \in G^{\prime}$ such that $ab$ is well defined in $V$, then $ab \in G^{\prime}$. If $V$ also satisfies that for all $a \in G^{\prime} $, $b \in V$ such that $bab^{-1}$ is well defined in $V$, then $bab^{-1}\in G^{\prime}$, we say that $G^{\prime}$ is a \textit{normal} sub-local group of $G$, and $V$ is called a \textit{normalizing neighborhood} of $G^{\prime}$. \end{definition} \begin{lemma} \rm Let $G$ be a cancellative group and $H$ be a normal sub-local group with normalizing neighborhood $V$. Let $W \subset G$ be an open symmetric subset such that $W^6 \subset V$. Then there is a cancellative local group $W/H $ equipped with a surjective morphism $\phi : W \to W / H$ such that, for all $g,h \in W$, one has $\phi (g) = \phi (h)$ if and only if $gh^{-1}\in H$, and for any $E \subset W / H$, one has that $E$ is open if and only if $\phi^{-1}(E)$ is open. \end{lemma} \subsection{Ultraconvergence of polynomials} \begin{definition} \rm Let $ Q_i : \mathbb{R}^{m} \to \mathbb{R}^n $ be a sequence of polynomials of bounded degree. We say that the sequence converges \textit{well} to a polynomial $Q: \mathbb{R}^{m} \to \mathbb{R}^n$ if the sequence $Q_i$ is ultraequivalent to the constant sequence $Q$. Equivalently, the sequence $Q_i$ converges well to $Q$ if the sequences of coefficients of $Q_i$ ultraconverge to the corresponding coefficients of $Q$. \end{definition} \begin{lemma}\label{PreWell} \rm For each $d\in \mathbb{N}$, there is $N_0 \in \mathbb{N}$ such that the following holds. Let $I_{N_0}:= \{ -1, \ldots,\frac{-1}{N_0}, 0, \frac{1}{N_0}, \ldots , 1 \} $, and assume we have polynomials $Q_i, Q: \mathbb{R}^m \to \mathbb{R}^m$ of degree $\leq d$ such that $Q_i(x) \xrightarrow{\alpha} Q(x) $ for all $x\in \left( I_{N_0} \right)^{\times m}$. Then $Q_i$ converges well to $Q$. \end{lemma} \begin{proof} Working on each coordinate, we may assume that $n=1$. We proceed by induction on $m$, the case $m =1$ being elementary Lagrange interpolation. Name the variables $x_1, \ldots , x_m$. Since $$ \mathbb{R}[x_1, \ldots , x_m] = (\mathbb{R}[x_1])[x_2, \ldots , x_{m }], $$ we can consider the polynomials $Q_i, Q$ as polynomials $\tilde{Q}_i, \tilde{Q}$ in the variables $x_2, \ldots, x_{m }$ with coefficients in $\mathbb{R}[x_1]$. If $Q_i(x) \xrightarrow{\alpha} Q(x) $ for all $x\in \left( I_{N_0}\right)^{\times m}$, we would have $\tilde{Q}_i(q, x^{\prime}) \xrightarrow{\alpha} \tilde{Q}(q,x^{\prime})$ for all $q \in I_{N_0}$ and $x^{\prime} \in ( I_{N_0})^{\times (m -1)}$. By the induction hypothesis, if $N_0$ was large enough, depending on $d$, the coefficients of $\tilde{Q}_i $, which are polynomials in $\mathbb{R}[x_{1}]$, ultraconverge to the coefficients of $\tilde{Q}$ whenever $x_{1}\in I_{N_0}$. By the case $m =1$, if $N_0$ was large enough, the coefficients of $Q_i$ ultraconverge to the coefficients of $Q$. \end{proof} \begin{lemma}\label{Well} \rm Let $Q_i: \mathbb{R}^n \times \mathbb{R}^n \to \mathbb{R}^n$ be a sequence of polynomial group structures in $\mathbb{R}^n$ of bounded degree. Assume $Q_i $ converges well to a polynomial group structure $Q: \mathbb{R}^n \times \mathbb{R}^n \to \mathbb{R}^n$. Then the corresponding sequence of Lie algebra structures on $\mathbb{R}^n$ converges well to the Lie algebra structure of $Q$. \end{lemma} \begin{proof} This follows from the fact that the structure coefficients of the Lie algebras depend continuously on the second derivatives of $Q_i$, which by hypothesis, ultraconverge to the corresponding second derivatives of $Q$. \end{proof} \begin{definition}\label{quasil} \rm For $x \in \mathbb{R}^r$, we define its \textit{support} as $$ \text{supp} ( x_1, \ldots , x_r) := \{ i \in \{ 1, \ldots , r \} \vert x_i \neq 0 \} . $$ For $x,y \in \mathbb{R}^r$, we say that $x \preceq y$ if $i \leq j$ for every $i \in \text{supp} (x) $, $j \in \text{supp} (y) $. We say that a polynomial group structure $Q : \mathbb{R}^r \times \mathbb{R}^r\to \mathbb{R}^r$ is \textit{quasilinear} if $$Q(x,y) = Q(x, 0) + Q(0,y) = x+y$$ whenever $x \preceq y$. \end{definition} Note that for any quasilinear group structure $Q:\mathbb{R}^r \times \mathbb{R}^r \to \mathbb{R}^r$, the coordinate axes are one-parameter subgroups, and the exponential map $$\exp : T_0 \mathbb{R}^r = \mathbb{R}^r \to \mathbb{R}^r$$ is the identity when restricted to such axes. Moreover, by the Baker--Campbell--Hausdorff formula, for $x= (x_1, \ldots , x_r) \in \mathbb{R}^r$, the expression \begin{eqnarray*} \log ( x ) & = & \log \left( x_{ 1} e_1 + \ldots + x_{i}e_r \right) \\ & = & \log \left( ( x_{ 1} e_1 ) \ldots ( x_{r}e_r) \right) \\ & = & \log ( \exp ( \log ( x_{1}e_1 ) ) \ldots \exp( \log ( x_{r}e_r ) ) ) \\ & = & \log ( \exp (x_{1 }e_1) \ldots \exp( x_{r}e_r) ) \end{eqnarray*} is a polynomial on the variables $x_1, \ldots , x_r$, and its coefficients depend continuously on the structure coefficients of the Lie algebra associated to $(\mathbb{R}^r, Q)$. This, together with Lemma \ref{Well}, implies the following result. \begin{lemma}\label{Quasilinear Continuity} \rm Consider quasilinear polynomial nilpotent group structures $Q_i , Q : \mathbb{R}^r\times \mathbb{R}^r \to \mathbb{R}^r$ of bounded degree. Let $\log_i ,\log : \mathbb{R}^r \to \mathbb{R}^r = T_0 \mathbb{R}^r$ denote the logarithm maps for the group structures $Q_i$ and $Q$, respectively. Assume the sequence $Q_i$ converges well to $Q$, and a sequence $x_i \in \mathbb{R}^r$ ultraconverges to a point $x \in \mathbb{R}^r$. Then \[ \lim\limits_{i \to \alpha} \log_i (x_i) = \log (x). \] \end{lemma} \subsection{Constructing covering spaces}\label{CoverConstruction} In this section we present the constructions of covering spaces we will need. \begin{definition} \rm Let $\varepsilon > 0$. We say that a covering map $\tilde{Y} \to Y$ of geodesic spaces is $\varepsilon$\textit{-wide} if for every $y \in Y$, the ball $B^Y(y, \varepsilon ) $ is an evenly covered neighborhood of $y$. \end{definition} The following result is obtained via the standard construction of covering spaces (see \cite{Munk}, Theorem 77.1). \begin{lemma}\label{cover-characterization} \rm Let $(Y,y_0)$ be a pointed proper geodesic space, $\Delta$ a group, and $\varepsilon > 0$. Then the following are equivalent: \begin{itemize} \item There is a Galois $\varepsilon$-wide covering map $\tilde{Y} \to Y$ with Galois group $\Delta$. \item There is a surjective morphism $\pi _ 1 (Y,y_0) \to \Delta$ whose kernel contains all the classes containing $\varepsilon$-lassos nailed at $y_0$. \end{itemize} \end{lemma} The following result by Sormani and Wei implies that if two geodesic spaces are sufficiently close in the Gromov--Hausdorff sense, then one can transfer Galois covers from one of them to the other (\cite{SWHau}, Theorem 3.4). \begin{theorem}\label{sormani-cover-copy} \rm (Sormani--Wei) Let $A, B$ be proper geodesic spaces with $$d_{GH} (A,B ) \leq \varepsilon/200, $$ and $\tilde{B} \to B$ a Galois $\varepsilon$-wide covering map with Galois group $\Delta$. Then there is a Galois cover $\tilde{A} \to A$ with Galois group $\Delta$. \end{theorem} Consider a pointed proper geodesic space $(Z,z_0)$, and a discrete group of isometries $\Gamma \leq Iso (Z)$, with diam$(Z/\Gamma) \leq \rho /10$ for some $\rho > 0$. We define \begin{center} $B:= B(z_0, \rho ),$ $ S := \{ g \in \Gamma \vert d(gz_0, z_o) \leq 2 \rho \} = \{ g \in \Gamma \vert gB \cap B \neq \emptyset \} . $ \end{center} Let $ \tilde{\Gamma } $ be the abstract group generated by $S$, with relations \begin{center} $s= s_1 s_2$ in $ \tilde{\Gamma } $, whenever $s, s_1, s_2 \in S$ and $s= s_1s_2$ in $\Gamma$. \end{center} If we denote the canonical embedding $S \hookrightarrow \tilde{\Gamma} $ as $(s \to s^{\sharp } )$, then there is a unique group morphism $\Phi : \tilde{\Gamma } \to \Gamma$ that satisfies $\Phi (s^{\sharp }) = s$ for all $s \in S$. This map is surjective by Lemma \ref{short-generators}. Equip $ \tilde{\Gamma } $ with the discrete topology, and consider the topological space \begin{center} $\tilde{Z} := \left( \tilde{\Gamma } \times B \right) / \sim$, \end{center} where $\sim$ is the minimal equivalence relation such that \begin{center} $ (gs^{\sharp}, x) \sim (g,s x) $ whenever $s \in S$, $x , sx \in B$. \end{center} We obtain a continuous map $\Psi : \tilde{Z} \to Z$ given by \begin{center} $\Psi (g, x) : = \Phi (g)(x) .$ \end{center} \begin{theorem}\label{monodromy} \rm The map $\Psi$ is a Galois $(\rho /3)$-wide covering map with Galois group $Ker (\Phi )$. \end{theorem} The proof of Theorem \ref{monodromy} is obtained from a sequence of lemmas. \begin{lemma}\label{relation-is-good} \rm Let $(a,x), (b,y ) \in \tilde{\Gamma} \times B$. The following are equivalent: \begin{itemize} \item there is $s \in S$ with $b= as^{\sharp}$, $x=sy$. \item there is $t \in S$ with $a=bt^{\sharp}$, $y=tx$. \item $(a,x) \sim (b,y)$. \end{itemize} \end{lemma} \begin{proof} The first two conditions are equivalent by taking $t $ to be $s^{-1}$, and they imply the third one by definition. Using the fact that the first two conditions are equivalent, the third condition implies that there is a sequence $s_1, \ldots , s_k \in S$ such that \[(a,x) \sim (as_1^{\sharp} , s_1^{-1}x) \sim \ldots \sim (as_1^{\sharp} \cdots s_k^{\sharp} , s_k^{-1} \cdots s_1^{-1} x ) = (b,y). \] This implies that $ \left( s_1^{-1} \right) , \left( s_2^{-1}s_1^{-1} \right) , \ldots , \left( s_k^{-1}\ldots s_1^{-1} \right) \in S$, allowing us to prove by induction on $j$ that $(s_1\cdots s_j)^{\sharp} s_{j+1}^{\sharp} = (s_1 \cdots s_{j+1} )^{\sharp}$ in $\tilde{\Gamma }$. This implies the first condition by taking $s$ to be $(s_1 \cdots s_k) \in S$. \end{proof} We say that a subset $A \subset \tilde{\Gamma } \times B$ is \textit{saturated} if it is a union of equivalence classes of the relation $\sim$. By definition, if $A$ is open and saturated, then its image in $\tilde{Z}$ is open. Let $U \subset Z$ be an open ball of radius $\rho /2$. Since $\Phi$ is surjective and diam$(Z/\Gamma ) \leq \rho / 10$, there is $g_0 \in \tilde{\Gamma}$ such that \[V : = \Phi(g_0^{-1})(U) \subset B^Z(z_0, 2\rho /3).\] \begin{lemma}\label{monodromy1} \rm The preimage of $U$ is given by \begin{equation}\label{preimage-u} \Psi ^{-1}(U) = \bigsqcup_{g \in Ker(\Phi)} \left( \bigcup_{s \in S} \left( \{ g_0gs^{\sharp} \} \times \left(( s^{-1}V) \cap B \right) \right) \right) / \sim . \end{equation} \end{lemma} \begin{proof} By direct evaluation, if $x \in V \cap sB$, then \[ \Psi (g_0gs^{\sharp}, s^{-1}x) = \Phi (g_0g)x = \Phi(g_0) x \in \Phi (g_0) (V) = U. \] On the other hand, if $(h,x) \in \tilde{\Gamma} \times B$ is such that $\Phi(h) (x) \in U$, then $\Phi(g_0^{-1}h)(x) \in V$, and $\Phi(h) \in \Phi( g_0) S $, implying that $h = g_0gs^{\sharp}$ for some $g \in Ker (\Phi)$, $s \in S$. Also, $s(x) = \Phi ( g_0^{-1} g_0 s^{\sharp} )(x) = \Phi(g_0^{-1}h)(x) \in V$, proving that the class of $(h,x)$ belongs to the right hand side of Equation \ref{preimage-u}. \end{proof} \begin{lemma}\label{monodromy2} \rm As $g$ ranges through $ Ker (\Phi )$, the sets \[ W_g : = \bigcup_{s \in S} \left( \{ g_0gs^{\sharp} \} \times \left(( s^{-1}V) \cap B \right) \right) \subset \tilde{\Gamma } \times B \] are open, disjoint, and saturated. \end{lemma} \begin{proof} The fact that they are open is straightforward, since $(s^{-1}V) \cap B$ is open in $B$ for each $s \in S$. To prove that they are disjoint, assume that \[ (g_0g_1s_1^{\sharp},x ) \sim (g_0 g_2 s_2 ^{\sharp} , y) \] for some $g_1, g_2 \in Ker(\Phi),$ $s_1s_2 \in S , $ $ x\in (s_1^{-1}V) \cap B , y \in (s_2^{-1}V ) \cap B . $ Lemma \ref{relation-is-good} implies that there is $t \in S$ with \begin{equation}\label{mono-equation} g_0g_1s_1^{\sharp} t^{\sharp} = g_0g_2s_2^{\sharp} , \text{ } x=ty . \end{equation} By Taking $\Phi$ on both sides of the first equation, we get $s_2 = s_1t$, and hence $s_2^{\sharp} = s_1^{\sharp}t^{\sharp}$. Canceling this in Equation \ref{mono-equation}, we get $g_1=g_2$, proving that the sets $W_g$ are disjoint. To prove that they are saturated, assume that for some $(h,x ) \in \tilde{\Gamma} \times B$, $g \in Ker (\Phi) $, $s \in S$, $y \in (s^{-1} V )\cap B$, we have $(h,x) \sim (g_0gs^{\sharp} , y)$. By Lemma \ref{relation-is-good}, there is $t \in S$ with \[ h = g_0gs^{\sharp}t^{\sharp}, \text{ } y = tx. \] This implies that $st (x)= s(y) \in V \subset B$, hence $st \in S$ and $(st)^{\sharp}=s^{\sharp} t ^{\sharp} $. Then \[ (h,x) \in \{ g_0 g (st)^{\sharp} \} \times \left( ((st)^{-1}V) \cap B \right) , \] proving that $W_g$ is saturated. \end{proof} \begin{lemma}\label{monodromy3} \rm For each $g \in Ker (\Phi )$, the image of $W_g$ in $\tilde{Z}$ is sent homeomorphically via $\Psi$ onto $U$. \end{lemma} \begin{proof} Surjectivity is traightforward, since $\Phi ( g_0 g ) (V) = U$. To check injectivity, assume that for some $s,t \in S$, $x\in (s^{-1}V) \cap B$, $y \in (t^{-1}V) \cap B$, we have \[ \Phi(g_0gs^{\sharp})(x) = \Phi (g_0gt^{\sharp})(y). \] Then $x=s^{-1}ty \in B$, implying that $s^{-1}t \in S$, and consequently, $(s^{-1}t)^{\sharp} = (s^{\sharp})^{-1}t^{\sharp}$. Hence \[ g_0gs^{\sharp} (s^{-1}t)^{\sharp} = g_0gt^{\sharp} , \text{ }x= (s^{-1}t)y, \] obtaining injectivity. To check that $\Psi \vert_{W_g}$ is open, take $\mathcal{O} \subset W_g$ open and saturated containing the class of \[(g_0g, x) \in \{g_0g\} \times V . \] Then $\Psi $ sends $ ( \{g_0g\} \times V ) \cap \mathcal{O} $ to an open neighborhood of $\Phi(g_0)(x)$. Since $(g_0g,x) $ was arbitrary, $\Psi \vert _{W_g} $ is open. \end{proof} By lemmas \ref{monodromy1}, \ref{monodromy2}, and \ref{monodromy3}, $U$ is an evenly covered neighborhood. Since $U$ was arbitrary, Theorem \ref{monodromy} follows. \section{Nilpotent groups of isometries}\label{PartI} In this section, we begin the proof of Theorem \ref{MAIN} with the following result. \begin{theorem}\label{PI} \rm Let $(X_i,p_i)$ be a sequence of almost homogeneous spaces that converges in the pointed Gromov--Hausdorff sense to a space $(X,p)$. Then $X$ is a nilpotent locally compact group equipped with an invariant metric. \end{theorem} By hypothesis, there are discrete groups of isometries $G_i \leq Iso (X_i)$ with diam$(X_i/G_i) \to 0$. To prove Theorem \ref{PI} we first reduce it to the case when the groups $G_i$ are almost nilpotent. \begin{lemma}\label{anil} \rm Under the hypotheses of Theorem \ref{PI}, there are discrete groups of isometries $\Gamma_i \leq Iso (X_i)$ with diam$(X_i/\Gamma_i)\to 0$, satisfying that for each $\varepsilon > 0$, there is $N = N(\varepsilon ) \in \mathbb{N} $ such that for all $i \geq N$, and $g \in \Gamma_i ^{(N)}$, we have $d(gp_i,p_i ) < \varepsilon $. \end{lemma} This Lemma is an elementary consequence of the following result, which is one of the strongest versions of the Margulis Lemma \cite{BGT}. It states that if a sufficiently large ball in a Cayley graph can be covered by a controlled number of balls of half its radius, then the corresponding group is virtually nilpotent. \begin{theorem}\label{Margulis} \rm (Breuillard--Green--Tao) For $C>0$, there is $N(C) \in \mathbb{N}$ such that the following holds: Let $A$ be a finite symmetric subset of a group $G$, which is in turn generated by a finite symmetric set $S$. If $S^N\subset A$, and $A^2$ can be covered by $C$ left translates of $A$, then there is a subgroup $ G^{\prime} \leq G$ with \begin{itemize} \item $[G:G^{\prime}]\leq N$ \item $\left( G^{\prime}\right)^{(N)} \subset A^4 $ \end{itemize} \end{theorem} \begin{proof}[Proof of Lemma \ref{anil}:] Fix $k \in \mathbb{N}$, and define \[ A_i : = \{ g \in G_i \vert d(gp_i,p_i) \leq 2/k \} . \] It is straightforward to verify that $A_i$ is finite and symmetric. Since $X$ is proper, there is $C > 0 $ such that $B(p,4/k)$ can be covered by $C\in \mathbb{N}$ balls of radius $1/k$. That is, there are $\{ q_1, \ldots , q_C \} \in X$ such that \[ B(p , 4 /k) \subset \bigcup_{j=1}^C B(q_j, 1/k). \] From the definition of pointed Gromov--Hausdorff convergence, there are sequences of functions $f_i : X_i \to X$, $h_i : X \to X_i$ with $f_i(p_i)=p$, $h_i(p)=p_i$, and satisfying equations \ref{pGH1} and \ref{pGH2} for all $R>0$. Since diam$(X_i/G_i ) \to 0$, there are group elements $\{ g_{1,i}, \ldots , g_{C,i} \} \in G_i $ such that $d(g_{j,i}p_i, h_iq_j)\to 0$ as $i \to \infty$ for each $j\in \{1, \ldots, C \}$. This implies that for large enough $i$, \[ A_i^ 2 \subset \{ g \in G_i \vert d(gp_i,p_i)\leq 4 /k \} \subset \bigcup_{j=1}^C g_{j,i} A_i . \] One can also define \[ S_i : = \{ g \in G_i \vert d(gp_i,p_i) \leq 3 \cdot \text{diam}(X_i/G_i) \} , \] which generates $G_i$ by Lemma \ref{short-generators}. Let $N_k:=N(C)\in \mathbb{N}$ be given by Theorem \ref{Margulis}. Since diam$(X_i/G_i) \to 0$, there is $M_k\in \mathbb{N}$ such that for all $i \geq M_k$, we have \[ \text{diam}(X_i/G_i) \leq \frac{1}{4kN_k}.\] This immediately implies $S_i^{N_k} \subset A_i$. Hence there are subgroups $ G_i^{\prime} (k) \leq G_i$ which satisfy, by Lemma \ref{diameter-subgroup}, \begin{itemize} \item diam$(X_i/(G_i^{\prime}(k)))\leq 3 N_k \cdot $ diam$ (X_i/G_i) $ \item If $i \geq M_k$, then all $g \in \left( G_i^{\prime} (k) \right)^{(N_k)}$ satisfy $d(gp_i,p_i) \leq 8 /k$ \end{itemize} By inductively repeating this construction with $k \in \mathbb{N}$, we can assume that $N_{k+1} \geq N_k$ and $M_{k+1} \geq M_k$ for all $k$. With this, we define $\Gamma_i : = G_i^{\prime}(k)$, where $k$ is the largest integer such that $i \geq M_k$. The verify the conclusion of the lemma, for any positive $\varepsilon $ we define $N(\varepsilon) : = \max \{N_k,M_k \} $ with $k$ satisfying $\varepsilon k \geq 8$. \end{proof} \begin{proof}[Proof of Theorem \ref{PI}:] Let $\Gamma_i$ be given by Lemma \ref{anil}. By Lemma \ref{proper-geodesic} and Theorem \ref{equivariant-ultralimit}, $X$ is a proper geodesic space and the equivariant ultralimit $\Gamma_{\alpha} \leq Iso (X)$ is a closed subgroup acting transitively. By Lemma \ref{anil}, for any $\varepsilon > 0$, there is $N \in \mathbb{N}$, such that \[ \Gamma_{\alpha}^{(N)} \subset \{ g \in \Gamma _{\alpha} \vert d(gp,p) \leq \varepsilon \} . \] Since $\Gamma_{\alpha}$ acts transitively on $X$, Lemma \ref{normal-stabilizers} implies \begin{equation}\label{nilalpha} \Gamma_{\alpha}^{(N)} \subset \{ g \in \Gamma_{\alpha} \vert d(g q,q) \leq \varepsilon \text{ for all }q \in X \} . \end{equation} By Remark \ref{topology-basis} and Theorem \ref{gleason-yamabe}, there is an open subgroup $\mathcal{O} \leq \Gamma_{\alpha}$ with the property that for any $R , \varepsilon > 0$, there is a compact normal subgroup $K \triangleleft \mathcal{O}$ with $K \subset U_{R, \varepsilon}^{\alpha}$ and such that $\mathcal{O}/K$ is a connected Lie group. Fix $K_0 \triangleleft \mathcal{O}$ a compact normal subgroup satisfying that $ \mathcal{O}/K_0$ is a connected Lie group, and denote by $\pi_{K_0}: \mathcal{O} \to \mathcal{O}/K_0$ the projection. Let $U \subset \mathcal{O}/K_0$ be a small open neighborhood of the identity, such that any subgroup of $\mathcal{O}/K_0$ contained in $U$ is trivial. Since $\pi_{K_0}^{-1}(U) \subset \mathcal{O}$ is an open neighborhood of the identity, there is $\varepsilon > 0 $ such that \begin{equation}\label{NSSNil} \{ g \in \mathcal{O} \vert d(gq,q) \leq \varepsilon \text{ for all }q \in X \} \subset \pi_{K_0}^{-1}(U) . \end{equation} Let $N(\varepsilon) \in \mathbb{N} $ be given by Lemma \ref{anil}. Then Equations \ref{nilalpha} and \ref{NSSNil} imply that \[ \left( \mathcal{O}/K_0 \right)^{(N)} = \mathcal{O}^{(N)}/K_0 = \pi_{K_0}\left( \mathcal{O}^{(N)}\right) \subset U , \] and hence $\mathcal{O}/K_0$ is nilpotent of step $\leq N$. Notice that we also proved that if a compact normal subgroup $K_1 \triangleleft \mathcal{O}$ satisfies that $\mathcal{O} /K_1 $ is a connected Lie group, then $\mathcal{O}/K_1$ is nilpotent. \begin{center} \textbf{Claim:} $\mathcal{O}$ is nilpotent of step $\leq N+1$. \end{center} By contradiction, assume there is $g \in \mathcal{O}^{(N+1)} $ with $gq \neq q$ for some $q \in X$. We choose $R := 2 d (p,q) +1$, $\varepsilon_0 : = d(gq,q) /2 $, and take a compact normal subgroup $K \triangleleft \mathcal{O}$ with $K \leq U_{R,\varepsilon_0}^{\alpha}$ and such that $\mathcal{O}/K$ is a connected Lie group. By Theorem \ref{glushkov}, $\mathcal{O}/(K \cap K_0)$ is a connected Lie group, and it fits in an exact sequence of connected nilpotent Lie groups \[ 0 \to K_0/(K \cap K_0) \to \mathcal{O}/ (K \cap K_0 ) \to \mathcal{O} / K_0 \to 1 . \] The group $K_0/(K\cap K_0)$ is a compact subgroup of the connected nilpotent Lie group $\mathcal{O}/(K \cap K_0)$, so by Lemma \ref{compact-nilpotent-1}, it is central. This implies that \begin{eqnarray*} (\mathcal{O}/(K \cap K_0))^{(N+1)} & = & [ (\mathcal{O}/(K \cap K_0) )^{(N)}: \mathcal{O}/(K \cap K_0) ] \\ & \leq & [ K_0/(K \cap K_0 ) : \mathcal{O}/(K \cap K_0) ]\\ & = & 0. \end{eqnarray*} This implies that $\mathcal{O}^{(N+1)} \leq K \cap K_0 \leq K \subset U_{R, \varepsilon_0}^{\alpha} $, but by construction, there is $g \in \mathcal{O}^{N+1} \backslash U_{R, \varepsilon_0}^{\alpha} $, which is a conradiction, proving the claim. All that is left is proving that $\mathcal{O}$ acts freely and transitively on $X$, showing that $X \cong \mathcal{O}$. The fact that it acts transitively, follows from Theorem \ref{berestovskii-open}. To prove that the action is free, for $q \in X$ consider $\mathcal{O}_q : = \{ g \in \mathcal{O} \vert gq=q \}$. \begin{center} \textbf{Claim:} $\mathcal{O}_p$ is a central subgroup of $\mathcal{O}$. \end{center} Assume there are $g \in \mathcal{O}_p$, $h \in \mathcal{O}$, with $[g,h] \neq e$. Then there are $R, \varepsilon$ such that $[g,h]$ is not in $ U_{R, \varepsilon}^{\alpha}$, and a compact normal subgroup $K \triangleleft \mathcal{O}$ with $K \subset U_{R, \varepsilon }^{\alpha}$ and such that $\mathcal{O}/K$ is a connected Lie group. $\mathcal{O}_p / K$ is a compact subgroup of the connected nilpotent Lie group $\mathcal{O}/K$. Hence by Lemma \ref{compact-nilpotent-1}, $[g,h] \in K \subset U_{R, \varepsilon}^{\alpha}$, but by construction, $[g,h] $ is not in $U_{R, \varepsilon}^{\alpha}$, which is a contradiction, proving the claim. By Lemma \ref{normal-stabilizers}, $\mathcal{O}_p$ is trivial. Also, since $\mathcal{O}$ acts transitively on $X$, then for any $q \in X$, there is $h \in \mathcal{O}$ with $hp=q$. This implies that \[ \mathcal{O}_q = h \mathcal{O}_p h^{-1} = \mathcal{O}_p = \{ Id_X \} . \] \end{proof} \section{Semi-locally-simply-connected groups}\label{PartII} In this section, we prove the second part of Theorem \ref{MAIN}, consisting of the following result. \begin{theorem}\label{PII} \rm Let $X_i$ be a sequence of almost homogeneous spaces, converging in the pointed Gromov--Hausdorff sense to a space $X$. If $X$ is semi-locally-simply-connected, then $X$ is a Lie group equipped with an invariant Finsler or sub-Finsler metric. \end{theorem} Due to the following result of Valerii Berestovskii (\cite{BerII}, Theorem 3), all we need to show is that $X$ is a Lie group. \begin{theorem}\label{BerestovskiiFinsler} \rm (Berestovskii) Let $Y$ be a proper geodesic space whose isometry group acts transitively. If $Y$ is homeomorphic to a topological manifold, then its metric is given by a Finsler or a sub-Finsler structure. \end{theorem} \begin{proof}[Proof of Theorem \ref{PII}:] By Theorem \ref{PI}, $X$ is a connected nilpotent group. By Corollary \ref{glushkov-corollary}, if $X$ is not a Lie group, then it contains a sequence of compact subgroups $K_1 \geq K_2 \geq \ldots $ with \begin{equation}\label{trivial-intersection} \bigcap_{j=1}^{\infty} K_j = 0 \end{equation} and such that $H_j : = X/K_j$ is a connected nilpotent Lie group. \begin{center} \textbf{Case 1:} Only for finitely many $j$, the connected component of $K_j/K_{j+1}$ is non-trivial. \end{center} Let $j_0$ be such that $K_j/K_{j+1} $ is discrete for all $j \geq j_0$. Let $\delta > 0 $ be small enough so that $B^{H_{j_0}} (e, \delta) $ contains no non-trivial subgroups, and $B^{H_{j_0}} (e, \delta) \to H_{j_0}$ has no content. By Lemma \ref{FinalCover} and induction, the balls $B^{H_j } (e,\delta)$ contain no non-trivial subgroups for $j \geq j_0$. From Equation \ref{trivial-intersection}, there is $M \in \mathbb{N}$ such that for all $\ell \geq M$, we have $K_{\ell} \subset B^{X}(p, \delta)$, and \[ K_{\ell} / K_{\ell + 1} \subset B^{H_{\ell + 1} }(e, \delta ) . \] This implies $K_{\ell} = K _{\ell + 1}$ for $\ell \geq M$, and $X= H_M$. \begin{center} \textbf{Case 2:} For infinitely many $j$, the connected component. of $K_j/K_{j+1}$ is non-trivial \end{center} Since $X$ is semi-locally-simply-connected, there is $\delta > 0$ such that the inclusion $B^X(p , \delta ) \to X $ has no content. By our assumption, there is $j \in \mathbb{N}$ with $ K_j \subset B^X(p, \delta / 12) $, and the connected component of $K_j/K_{j+1} $ is non-trivial. By Lemma \ref{LongHomo} and Corollary \ref{ShortLoopNT}, there is a non-contractible loop $\gamma_0: [0,1] \to H_{j+1}$ based at $e$, with length$(\gamma_0)\leq \delta / 4$. A contradiction will arise by generating from this loop a small non-contractible loop in $X$. Since $H_{j+1}$ is a Lie group, there is $ 0 < \varepsilon < \delta /4 $ such that if two closed curves in $H_{j+1}$ are at uniform distance less than $\varepsilon$, then they are homotopic to each other. Choose $\ell \in \mathbb{N}$ large enough so that $K_{\ell}\subset B^X(p,\varepsilon / 2)$ and let $f: H_{\ell} \to H_{j+1} $ be the natural map, which by Lemma \ref{Submer} is a metric submersion. Using Lemma \ref{HoriLift}, we get a curve $\tilde{\gamma}_1: [0,1] \to H_{\ell}$ with $\tilde{\gamma}_1(0) = e$, $f \circ \tilde{\gamma}_1 = \gamma_0$, and length$(\tilde{\gamma}_1) \leq \delta / 4$. Since $f(\tilde{\gamma}_1(1) ) = e$, we have $\tilde{\gamma}_1 (1) \in K_{j+1}/K_{\ell} $. For each $m \in \mathbb{N}$, define the curve $\tilde{\gamma}_m : [0,1]\to H_{\ell}$ as \[\tilde{\gamma}_m(t) : = \left[ \tilde{\gamma}_1(1) \right]^{m-1} \tilde{\gamma}_1(t). \] Observe that $\tilde{\gamma}_{m}(1) = \tilde{\gamma}_{m+1}(0)$ for each $m$, so we can define the curves $\beta_m : = \tilde{\gamma}_1 \ast \ldots \ast \tilde{ \gamma }_m$. Let $m_0 $ be the smallest integer such that $\beta_{m_0}(1) $ lies in the connected component of the identity of $ K_{j+1}/K_{\ell}$, and let $\beta : [0,1]\to K_{j+1}/K_{\ell}$ be a (not necessarily rectifiable) curve from $\beta_{m_0} (1) $ to $e$. Then the curve $\beta_{m_0}\ast \beta$ is a curve whose image lies in $B^{H_{\ell}}(e, \delta / 3)$, and the composition $f \circ (\beta_{m_0 } \ast \beta)$ is homotopic to $\gamma_0 \ast \ldots \ast \gamma_0$ ($m_0$ times). Since $\pi_1 (H_{j+1} ) $ has no torsion, $f \circ (\beta_{m_0} \ast \beta)$ is non-contractible in $H_{j+1}$. By Lemma \ref{ShortLoopNTI}, there is a loop $\tilde{\gamma} $ in $H_{\ell}$ based at $e$ with length$(\tilde{\gamma}) \leq \delta $ such that $f \circ \tilde{ \gamma} $ is non-contractible in $H_{j+1}$. By Lemma \ref{Submer}, the quotient map $\rho : X \to H_{\ell}$ is a metric submersion, and hence by Lemma \ref{HoriLift}, there is a curve $c_1 : [0,1] \to B^X(p, \delta ) $ with $c_1(0)=p$, and $\rho \circ c_1 = \tilde{\gamma}$. Let $c_2: [0,1] \to X$ be a minimizing path from $c_1(1)$ to $p$. Then the curve $c: =c_1 \ast c_2$ is a loop in $ B^X(p, \delta )$ such that \[ d_U ( (f \circ \rho \circ c ),( f \circ \rho \circ \tilde{c}_1); [0,1] ) \leq \varepsilon , \] for some suitable reparametrization $\tilde{c}_1 $ of $c_1$. This implies that $f \circ \rho \circ c$ is non-contractible in $H_{j+1} $, and the composition $B^X(p, \delta) \to X \to H_j$ has non-trivial content, which is a contradiction. \end{proof} \section{Discrete groups with uniform doubling}\label{Nilsection} As stated in the Summary, in Sections \ref{Nilsection} to \ref{Rootsection} we prove the following theorem, which is the last part of Theorem \ref{MAIN}. \begin{theorem}\label{PIII} \rm Let $(X_i, p_i)$ be a sequence of almost homogeneous spaces, converging in the pointed Gromov--Hausdorff sense to a Lie group $X$. Then for large enough $i$ there are subgroups $\Lambda_i \leq \pi_1(X_i)$ with surjective morphisms \begin{equation*} \Lambda _i \to \pi_1 (X). \end{equation*} \end{theorem} The first step is to obtain the following refinement of Lemma \ref{anil}. \begin{lemma}\label{anilLie} \rm Under the hypotheses of Theorem \ref{PIII}, there are discrete groups of isometries $\Gamma_i \leq Iso (X_i)$ with diam$(X_i/\Gamma_i)\to 0$, and $N \in \mathbb{N}$, satisfying that for each $\varepsilon > 0$, there is $M = M(\varepsilon ) \in \mathbb{N} $ such that for all $i \geq M$, and $g \in \Gamma_i ^{(N)}$, we have $d(gp_i,p_i ) < \varepsilon $. \end{lemma} To prove Lemma \ref{anilLie}, we need a result by Alexander Nagel, Elias Stein, and Stephen Waigner about how small balls behave in manifolds with invariant metrics (\cite{NSW}, Section 3). \begin{theorem}\label{small-balls} \rm (Nagel--Stein--Waigner) Let $Y$ be a manifold equipped with a Finsler or sub-Finsler metric, and assume $ Iso (Y)$ acts transitively. Then the Hausdorff dimension of $Y$ is an integer, and the corresponding Hausdorff measure $\mu$ on $Y$ is positive on open sets and finite on compact sets. Furthermore, there is a constant $C>0$ such that for all $q \in Y$, $s \in (0, 1]$, one has \[ \dfrac{\mu ( B(q, 10s) )}{ \mu ( B(q, s))} \leq C .\] \end{theorem} \begin{lemma}\label{CCDoubl} \rm Let $Y$ be a manifold equipped with a Finsler or sub-Finsler metric, and assume $ Iso (Y)$ acts transitively. Then there is a constant $C > 0$ such that for every $q\in Y$, $r \in (0, 1]$, there are elements $q_1, \ldots , q_m \in Y$, with $m \leq C$ such that $$ B(q, 4r) \subset \bigcup_{j=1}^m B(q_j , r ). $$ \end{lemma} \begin{proof} For $q\in Y$, $r \in (0, 1]$, pick a maximal set $q_1, \ldots , q_m \in Y$ such that \begin{center} $q_1, \ldots , q_m \in B(q, 4r)$ \end{center} and the family of balls \begin{center} $ \left\{ B\left( q_j, \frac{r}{2} \right) \right\}_{j=1}^m $ \end{center} is disjoint. By Theorem \ref{small-balls}, $m \leq C$ for some $C$ independent of $r$. If the balls $\{ B(q_j , r ) \}_{j=1}^m$ do not cover $B(q, 4r) $, then there is $q^{\prime} \in B(q, 4r ) $ with $$ \min_{j=1, \ldots, m} d(q^{\prime }, q_j) > r, $$ contradicting the maximality of the family $\{q_1, \ldots , q_m \}.$ \end{proof} \begin{proof}[Proof of Lemma \ref{anilLie}:] Just like in the proof of Lemma \ref{anil}, we construct discrete subgroups $G_i^{\prime}(k)\leq Iso (X_i) $, and integers $N_k, M_k$ such that \begin{itemize} \item diam$(X_i/(G_i^{\prime}(k)))\leq 3 N_k \cdot $ diam$ (X_i/G_i) $ \item If $i \geq M_k$, then all $g \in \left( G_i^{\prime} (k) \right)^{(N_k)}$ satisfy $d(gp_i,p_i) \leq 8 /k$ \end{itemize} However, by Theorem \ref{BerestovskiiFinsler} and Lemma \ref{CCDoubl}, $C$ does not depend on $k$, and hence $N : = N_k$ does not depend on $k$ either. We define $\Gamma_i : = G_i^{\prime}(k)$, where $k$ is the largest integer such that $i \geq M_k$. To verify the conclusion of the lemma, for any positive $\varepsilon $ we take $M(\varepsilon) : = M_k $ with $k$ satisfying $\varepsilon k \geq 8$. \end{proof} \section{Almost translational behavior}\label{FreeSection} Let $\Gamma_i$ be the sequence of groups given by Lemma \ref{anilLie}, and let $\Gamma_{\alpha}$ be its equivariant ultralimit, which by Theorem \ref{MZHil}, is a Lie group. With $N \in \mathbb{N}$ given by Lemma \ref{anilLie}, the normal subgroup $\Gamma_{\alpha}^{(N)}$ is contained in the subgroup $\{ g \in \Gamma_{\alpha } \vert gp=p\}$, so by Lemma \ref{normal-stabilizers}, $\Gamma_{\alpha}^{(N)}$ is trivial. By Theorem \ref{berestovskii-open}, the connected component of the identity $\Gamma_0$ also acts transitively. The group $ \{ g \in \Gamma_0 \vert gp=p \} $ is compact, so by Lemma \ref{compact-nilpotent-1}, it is central in $\Gamma_0$. Thus Lemma \ref{normal-stabilizers} implies that $ \{ g \in \Gamma_0 \vert gp=p \} $ is trivial, and we can identify $\Gamma_0 \cong X $. In this section, we use commutator estimates similar to the ones in (\cite{BK}, \cite{GrAF}) to prove that $\Gamma_{\alpha}$ is connected, and from that we deduce that the groups $\Gamma_i$ act, in some sense, almost by translations. \begin{proposition}\label{alpha=x} \rm $\Gamma_{\alpha}$ is connected. \end{proposition} \begin{proof} Since $\Gamma_0$ acts transitively on $X$, each connected component of $\Gamma_{\alpha}$ does too. This implies that $\Gamma_{\alpha}$ is connected if and only if the group \[ \Gamma_p : = \{ g \in \Gamma_{\alpha} \vert gp=p \} \] is trivial. To study this group, we require a result by Enrico LeDonne and Alessandro Ottazzi about isometries of nilpotent Lie groups with invariant metrics (\cite{LeDO}, Theorem 1.2). \begin{theorem}\label{subFInf} \rm (LeDonne--Ottazzi) Let $G$ be a nilpotent Lie group equipped with an invariant Finsler or sub-Finsler metric, and $f : G \to G$ an isometry. Then $f$ is smooth, and uniquely defined by its first order data at the identity ($f(e)$, $d_ef$). \end{theorem} By Theorem \ref{subFInf}, the compact group $\Gamma_p$ consists of diffeomorphisms. This implies that there is a $\Gamma_p$-invariant inner product $\langle \cdot , \cdot \rangle_0$ in $T_pX$. Then for each $g \in \Gamma_p$ there is an $\langle \cdot , \cdot \rangle_0$-orthonormal basis $$ a_1, b_1, \ldots , a_{k_1} , b_{k_1}, c_1, \ldots , c_{k_2}, d_1 , \ldots , d_{k_3} \in T_pX $$ and angles $$\theta_1, \ldots , \theta_{k_1} \in \mathbb{S}^1 \backslash \{ 1 \}$$ such that the derivative $d_pg : T_pX \to T_pX$ satisfies: \begin{eqnarray*} d_pg (a_j) &=& \cos \theta_j a_j + \sin \theta_j b_j, \\ d_pg (b_j) &= &-\sin \theta_j a_j + \cos \theta_j b_j, \\ d_pg (c_j)& =& -c_j ,\\ d_pg(d_j ) & = & d_j. \end{eqnarray*} Assume by contradiction that there exists $g \in \Gamma_p \backslash \{ Id _X\}$, then again by Theorem \ref{subFInf}, $d_pg \neq Id_{T_pX}$, so $k_1 + k_2 > 0$. Let $d_0$ denote the left invariant Riemannian metric in $X$ given by $\langle \cdot , \cdot \rangle_0$ and $d_U$ denote the uniform distance with respect to $d_0$. We deal first with the case $k_1 > 0 $. By the Baker--Cambell--Hausdorff formula, for every $\varepsilon > 0 $ there is $\delta > 0$ such that \begin{equation}\label{Taylor1} d_U\left( L_{\exp (\delta a_1)} , \exp \circ ( + \delta a_1) \circ \left( \exp^{-1} \right); B_{d_0}(p, 100 N 2^N \delta) \right) < \varepsilon \delta \end{equation} and \begin{equation}\label{Taylor2} d_U \left( g, \exp \circ \left( d_pg \right) \circ \left( \exp^{-1} \right) ; B_{d_0}( p , 100 N 2^N \delta) \right) < \varepsilon \delta . \end{equation} By equations \ref{Taylor1} and \ref{Taylor2}, and repeated applications of Proposition \ref{UnifProp}, one can find $C^{\prime}>0$ such that the step $N$ commutators satisfy that the uniform distance between \[ [\ldots [ L_{\exp(\delta a_1)}, g ], \ldots ], g ] : B_{d_0}^{X}(p, \delta) \to X \] and \[ \exp \circ [\ldots [ +\delta a_1, d_pg ], \ldots ], d_p g] \circ \left( \exp^{-1} \right) : B_{d_0}^{X}(p, \delta) \to X \] is bounded above by $C^{\prime} \varepsilon \delta$. However, by direct computation, $$d_0 \left( \exp \circ [\ldots [ +\delta a_1, d_pg ], \ldots ], d_pg] \circ \left( \exp^{-1}\right) (p), p \right) = \delta \vert \theta_1 - 1 \vert ^N + o(\delta).$$ So, if $\varepsilon > 0 $ was chosen small enough depending on $C^{\prime}$, as $\delta \to 0$ one obtains $$[\ldots [ L_{\exp(\delta a_1)}, g ], \ldots ], g](p) \neq p,$$ contradicting the fact that every step $N$ commutator is trivial. The case $k_2 > 0$ is similar, but using $c_1$ instead of $a_1$. \end{proof} By Proposition \ref{alpha=x}, $\Gamma_{\alpha} = \Gamma_0 \cong X$. This means that we can assume the action of $\Gamma_{\alpha}$ on $X$ is given by left multiplications. In order to translate this information back into the sequence $\Gamma_i$, we need to consider almost morphisms from the groups $\Gamma_i$ to $X$ and to $Iso(X)$. We define the \textit{holonomy maps} $t : \Gamma_i \to X$ and $\hat{t}: \Gamma _i \to Iso (X)$ as \[ t(g):= f_i (g( h_i (p))) ,\text{ } \hat{t}(g) : = L_{t(g)}. \] By the above discussion, for any stable sequence $g_i \in \Gamma_i$, its ultratimit $g_{\alpha} \in \Gamma_{\alpha}$ coincides with \[ \lim\limits_{i\to \alpha} \left( f_i \circ g_i \circ h_i \right) =L_{\lim\limits_{i\to \alpha} t(g_i)} . \] Besides that, for any bounded sequence $x_i \in X$, one has \[ \lim\limits_{i\to \alpha} L_{x_i} =L_{\lim\limits_{i\to \alpha} x_i} . \] Thus we obtain, using Proposition \ref{UnifProp} repeatedly, the following results. \begin{proposition}\label{holonomy1} \rm For any stable sequence $g_i \in \Gamma_i$, the sequence $ \left( f_i \circ g_i \circ h_i \right) $ is ultraequivalent to the sequence $\hat{t}(g_i)$. \end{proposition} \begin{proposition}\label{Holohomo} \rm For any pair of stable sequences $g_i , g_i ^{\prime} \in \Gamma_i$, the sequence $ \hat{t}(g_ig_i^{\prime})$ is ultraequivalent to the sequence $\hat{t}(g_i)\hat{t}(g_i^{\prime}) . $ \end{proposition} \begin{corollary}\label{holonomy2} \rm For any pair of stable sequences $g_i, g_i^{\prime} \in \Gamma_i , $ we have, $$ \lim\limits_{i\to \alpha } \hat{t} (g_ig_i^{\prime}) = \lim\limits_{i\to \alpha } \hat{t} (g_i) \hat{t} (g_i^{\prime}) = \lim\limits_{i\to \alpha } \hat{t} (g_i) \lim\limits_{i\to \alpha } \hat{t} (g_i^{\prime}) . $$ \end{corollary} \section{Getting rid of the torsion}\label{Torsection} Now we want to identify and get rid of the torsion elements of $\Gamma_i$. Unfortunately, for $g \in \Gamma_i $, the point $t(g) $ being close or even equal to $p$, may not imply that high powers of $g$ do not ``escape''. This happens because the definition of $t: \Gamma_i \to X$ contains an ``error'' coming from the fact that $f_i$ and $h_i$ are not an actual isometries. To deal with the torsion elements, we use the \textit{escape norm} and techniques from \cite{BGT}. Since $X$ is locally contractible, there is $\varepsilon_0 \in (0,1)$ such that every loop of length $\leq 10^6 \varepsilon_0$ is nullhomotopic. Let $B $ be a small open convex symmetric set in the Lie algebra $\mathfrak{g} $ of $\Gamma_{\alpha}$ such that \[ \exp(B )\subset B(p, \varepsilon_0) . \] Fix $R > 10^6\varepsilon_0 ,$ and define the sets \begin{eqnarray*} \Theta_i & : = & \{ g \in \Gamma_i \vert d(gp_i , p_i ) \leq R \} \\ \hat{ A} _i & : = & \{ g \in \Theta_i \vert t(g) \in \exp (B) \} ,\\ \hat{S}_i & := & \{ g \in \Theta_i \vert t(g) \in \exp \left( B / 10^5 C^3 \right) \} , \\ A_i & := & \hat{A}_i \cup \hat{A}_i^{-1} ,\\ S_i & : = & \hat{S}_i \cup \hat{S}_i^{-1}, \end{eqnarray*} where $C \in \mathbb{N}$ was obtained from Lemma \ref{CCDoubl}. \begin{definition}\label{approxgr} \rm Let $A$ be a finite symmetric subset of a multiplicative set and $C > 0$. We say that $A$ is a $C$\textit{-approximate group} if $A^2$ can be covered by $\leq C$ left translates of $A$. \end{definition} \begin{definition} \rm Let $A $ be a $C$-approximate group. We say that $A$ is a \textit{strong} $C$-approximate group if there is a symmetric set $S \subset A$ satisfying the following: \begin{itemize} \item $\left( \{ asa^{-1} \vert a \in A^4, s \in S\} \right) ^{10^3C^3} \subset A$. \item If $g, g^2, \ldots , g^{1000} \in A^{100}$, then $g \in A$. \item If $g, g^2 , \ldots , g^{10^6 C^3} \in A$, then $g \in S$. \end{itemize} \end{definition} By the Baker--Campbell--Hausdorff formula, Proposition \ref{holonomy1}, and Corollary \ref{holonomy2}, we see that if $B$ was chosen small enough, for $i$ sufficiently close to $\alpha$, we have \begin{center} $\{ asa^{-1} \vert a \in A_i^4, s \in S_i \} \subset S_i^2,$ \end{center} and all three conditions of a strong global approximate group hold. \begin{lemma}\label{Strong} \rm For $i$ sufficiently close to $\alpha$, the set $A_i$, thanks to $S_i$, is a strong $C$-approximate group. \end{lemma} \begin{definition} \rm Let $A $ be a subset of a multiplicative set $G$. For $g\in G$, we define the escape norm as \begin{center} $\Vert g \Vert_{A} : = \inf \left\{ \dfrac{1}{m+1} \bigg| \text{ } e, g, g^2, \ldots , g^m \in A \right\}$. \end{center} \end{definition} In strong approximate groups, the escape norm satisfies really nice properties, which Breuillard, Green, and Tao call the ``Gleason lemmas''. \begin{theorem} \label{Gleason} \rm (Gleason--Breuillard--Green--Tao) Let $A$ be a strong $C$-approximate group. Then for $g_1, g_2, \ldots , g_n \in A^{10}$, we have \begin{itemize} \item $\Vert g_1 \Vert_A = \Vert g_1 ^{-1} \Vert _A$. \item $\Vert g_1 ^k \Vert _A \leq \vert k \vert \Vert g_1 \Vert_A $ \item $\Vert g_2g_1g_2^{-1} \Vert_A \leq 10^3 \Vert g_1 \Vert_A$. \item $\Vert g_1g_2 \cdots g_n \Vert_A \leq O_C\left( \sum_{j=1}^n \Vert g_j \Vert_A \right) $. \item $\Vert [g_1,g_2]\Vert _A \leq O_C\left( \Vert g_1 \Vert_A \Vert g_2 \Vert_A \right)$. \end{itemize} \end{theorem} \begin{proof} The first three properties are immediate. We refer the reader to (\cite{BGT}, Section 8) for a proof of the other two properties. \end{proof} Lemma \ref{Strong} and Theorem \ref{Gleason} imply that for $i$ sufficiently close to $\alpha$, $$ W_i := \{ g \in A_i \vert \Vert g \Vert _{A_i}=0 \} $$ is a subgroup of $\Gamma_i$ normalized by $A_i$. By Lemma \ref{short-generators}, for $i$ sufficiently $\alpha$-large, $ W_i $ is a normal subgroup of $\Gamma_i$ such that for any sequence $w_i \in W_i$, one has $$ \lim\limits_{i \to \alpha} t(w_i ) =p. $$ Therefore, by Lemma \ref{normal-stabilizers}, the quotient map $X_i \to X_i/W_i$ is an $\varepsilon_i$-approximation with $\varepsilon_i \to 0$ as $i \to \alpha$, and \begin{equation}\label{quotient-gh-close} \lim\limits_{i \to \alpha} d_{GH} \left( X_i, X_i/W_i \right) =0. \end{equation} The quotients $Z_i := \Gamma_i/W_i $ are particularly useful to us since they satisfy the no-small-subgroup property. Let $Y_i := \hat{\pi} (A_i)$, where $\hat{\pi} : \Gamma_i \to Z_i$ is the standard quotient map. For $[g] \in Y_i \backslash \{ e \},$ we have $\Vert g \Vert_{A_i}\neq 0$, and therefore $g^m $ is not in $A_i^2 \supset A_i W_i $ for some $m>0$. This implies that $ \hat{\pi} (g^m)= [g]^m $ does not belong to $Y_i$. In other words, every non-trivial element in $Y_i$ eventually ``escapes'' from $Y_i$. We still have the map \begin{center} $\overline{t}:\hat{\pi}(\Theta_i) \to X = \Gamma_{\alpha}$ \end{center} given by \begin{center} $\overline{t}\left( [g]\right) := t(g)$. \end{center} Of course, to make this map well defined, we have to choose one representative from each class in $ \hat{\pi}(\Theta_i) $. However, different choices of representatives only change the value of $\overline{t}$ by an error which goes to $0$ as $i \to \alpha$. More precisely, if one considers two sequences $g_i, g_i^{\prime} \in \Theta_i $ with $\hat{\pi}(g_i) = \hat{\pi}(g_i^{\prime}) $ for $\alpha$-large enough $i$, then there is a sequence $w_i \in W_i$ satisfying $ g_i = g_i^{\prime} w_i $ for $\alpha$-large enough $i$, and by Corollary \ref{holonomy2}, \[ \lim\limits_{i \to \alpha} \hat{t} ( g_i ) = \lim\limits_{i \to \alpha} \hat{t} ( g_i ^{\prime}) \hat{t} ( w_i) = \lim\limits_{i \to \alpha} \hat{t} ( g_i ^{\prime}). \] Let $\tilde{X} $ be the universal cover of $X$. By our choice of $B$, for $\alpha$-large enough $i$, and $g \in Y_i^8$, $$\overline{t}( g ) \in (\exp (B))^{10} \subset B^X(p, 10 \varepsilon_0) \cong B^{\tilde{X}}(e, 10 \varepsilon_0) ,$$ and we can think of $\overline{t} $ as a map from $ Y_i^8 $ to $\tilde{X}$. We will denote this map by $\overline{t}_{\tilde{X}} : Y^8_i \to \tilde{X}$. \section{The nilprogressions}\label{Nilprogsection} In this section, we apply a short basis procedure by Breuillard, Green, and Tao to find nice subsets in the local groups $Y_i$. \begin{definition} \rm Let $A_i$ be a sequence of finite subsets of multiplicative sets. If there is a $C>0$ such that $A_i$ are local $C$-approximate groups for $i $ sufficiently close to $\alpha$, we say that the algebraic ultraproduct $A= \lim_{i \to \alpha} A_i $ is an \textit{ultra approximate group}. If for $\alpha$-large enough $i$, the approximate groups $A_i$ do not contain non-trivial subgroups, we say that $A$ is an \textit{NSS} (no small subgroups) ultra aproximate group. For subsets $A^{\prime}_i \subset A_i^4$ with the property $(A^{\prime}_i)^4 \subset A_i^4$, we say that the algebraic ultraproduct $A^{\prime}=\lim_{i \to \alpha} A^{\prime}_i $ is a \textit{sub-ultra approximate group} of $A$ if it is an ultra approximate group, and there is a constant $C^{\prime}\in \mathbb{N}$ such that $A_i$ can be covered by $C_0$ many translates of $A_i^{\prime}$ for $i$ sufficiently close to $\alpha$. \end{definition} Let $Y$ be the algebraic ultraproduct $\lim_{i \to \alpha} Y_i$. By the discussion in the previous section, it is an NSS ultra approximate group. Consider the map \begin{center} $\tilde{t} : Y^8 \to X$ \end{center} given by the metric ultralimit \begin{center} $\tilde{t}( \{ g_i \} ) := \lim\limits_{i \to \alpha} \overline{t}(g_i)$. \end{center} Corollary \ref{holonomy2} implies that $\tilde{t} $ is a homomorphism, but moreover, it is a \textit{good model}, as defined by Ehud Hrushovski \cite{Hr}. \begin{definition} \rm Let $A=\lim_{i \to \alpha} A_i$ be an ultra approximate group. A \textit{good Lie model} for $A$ is a connected local Lie group $L$, together with a morphism $\sigma:A^8 \to L $ satisfying: \begin{itemize} \item There is an open neighborhood $U_0 \subset L$ of the identity with $U_0 \subset \sigma (A) $ and $\sigma^{-1}(U_0) \subset A$. \item $\sigma (A)$ is precompact. \item For $F \subset U \subset U_0$ with $F $ compact and $U$ open, there is an algebraic ultraproduct $A^{\prime}= \lim_{i \to \alpha} A_i^{\prime}$ of finite sets $A_i^{\prime} \subset A_i$ with $\sigma^{-1}(F ) \subset A^{\prime} \subset \sigma^{-1}(U)$. \end{itemize} \end{definition} \begin{definition} \rm Let $B$ be a local group, $u_1, u_2, \ldots , u_r \in B$, and $N_1, N_2, \ldots , N_r $ $\in \mathbb{R}^+$. The set $P(u_1, \ldots , u_r ; N_1, \ldots , N_r)$ is defined as the set of words in the $u_i$'s and their inverses such that the number of appearances of $u_i$ and $u_i^{-1}$ is not more than $N_i$. We say that $P(u_1, \ldots , u_r ; N_1, \ldots , N_r)$ is well defined if every word in it is well defined in $B$. When that is the case, we call it a \textit{progression of rank} $r$ (a progression of rank $0$ is defined to be the trivial subgroup). We say a progression $P(u_1, \ldots , u_r ; N_1, \ldots , N_r)$ is a \textit{nilprogression in $C_0$-regular form} for some $C_0>0$ if it also satisfies the following properties: \begin{itemize} \item For all $1 \leq i \leq j \leq r$, and all choices of signs, we have \begin{center} $ [ u_i^{\pm 1} , u_j^{\pm 1} ] \in P \left( u_{j+1} , \ldots , u_r ; \dfrac{C_0N_{j+1}}{N_iN_j} , \ldots , \dfrac{C_0N_r}{N_iN_j} \right). $ \end{center} \item The expressions $ u_1 ^{n_1} \ldots u_r^{n_r} $ represent distinct elements as $n_1, \ldots , n_r$ range over the integers with $\vert n_1 \vert \leq N_1/C_0 , \ldots , \vert n_r \vert \leq N_r/C_0$. \end{itemize} For a nilprogression $P$ in $C_0$-regular form, and $\varepsilon \in (0,1)$, it is easy to see that $P( u_1, \ldots , u_r ; \varepsilon N_1, \ldots , \varepsilon N_r )$ is also a nilprogression in $C_0$-regular form. We denote it by $\varepsilon P$. We define the \textit{thickness} of $P$ as the minimum of $N_1, \ldots , N_r$ and we denote it by thick$(P)$. The set $\{ u_1 ^{n_1 }\ldots u_r^{n_r} \vert \vert n_i \vert \leq N_i/C _0 \}$ is called the \textit{grid part of }$P$, and is denoted by $G(P)$. Let $P_i$ be a sequence of sets. If for $\alpha$-large enough $i$, $P_i$ is a nilprogression of rank $r$ in $C_0$-regular form for some $r \in \mathbb{N}$, $C_0>0$, independent of $i$, we say that the algebraic ultraproduct $ P = \lim_{i \to \alpha} P_i$ is an \textit{ultra nilprogression of rank $r$ in $C_0$-regular form}. We denote $\lim_{i \to \alpha} \varepsilon P_i$ as $\varepsilon P$. If $(\text{thick}(P_i))_i$ is unbounded, we say that $P$ is a \textit{non-degenerate ultra nilprogression}. The algebraic ultraproduct $G(P) :=\lim_{i \to \alpha} G(P_i)$ is called the \textit{grid part of }$P$. \end{definition} The main result of this section is the following theorem, which is slightly stronger than (\cite{BGT}, Theorem 9.3). \begin{theorem}\label{Nilprog} \rm Let $A=\lim_{i \to \alpha} A_i$ be an NSS ultra approximate group. Assume there is a good Lie model $\sigma: A^8 \to L$. Then $A^4$ contains a nondegenerate ultra nilprogression $P$ of rank $r := dim(L)$ in $C_0$-regular form, with the property that for all standard $\varepsilon \in (0,1)$, there is an open set $U_{\varepsilon} \subset L$ with $\sigma^{-1}(U_{\varepsilon}) \subset G(\varepsilon P)$. \end{theorem} \begin{proof} The proof is done by induction on $r$. In the case $r=0$, $L$ is a trivial group, and necessarily $U_0 = L$. The first property of being a good model implies that $ A^8 = \sigma^{-1} (L)=\sigma^{-1}(U_0) \subset A $. Hence for $\alpha$-large enough $i$, $A_i$ is a group, which is trivial by the NSS property, and there is nothing to show. In order to prove the induction step with $r \geq 1$, we are going to follow the construction performed in (\cite{BGT}, Section 9), and use the results they provide Let $\hat{B}$ be a small open convex symmetric set in $\mathfrak{ l}$, the Lie algebra of $L$. Let $A^{\prime \prime \prime} \subset A^{\prime \prime} \subset A^{\prime } \subset A$ be sub ultra approximate groups of $A$ such that \begin{center} $ \sigma ^{-1} (\exp (\hat{B} )) \subset A^{\prime} \subset \sigma^{-1} ( \exp ((1.001) \hat{B}) ), $ $ \sigma ^{-1} (\exp (\delta \hat{B})) \subset A^{\prime \prime } \subset \sigma^{-1} ( \exp ((1.001) \delta \hat{B}) ) , $ $ \sigma ^{-1} (\exp ( \delta \hat{B} /10)) \subset A^{\prime \prime \prime } \subset \sigma^{-1} ( \exp ((1.001) \delta \hat{B}/10) ), $ \end{center} where $\delta \in (0,1)$ will be chosen later. Notice that if $\hat{B}$ was chosen small enough, then $A^{\prime} $, $A^{\prime \prime}$, $A^{\prime \prime\prime} $ are strong ultra approximate groups. Let $u \in A^{\prime } \backslash \{ e \}$ be such that minimizes $ \Vert u \Vert _{A^{\prime}}$ (in this setting, $\Vert \cdot \Vert_{A^{\prime}}$ is a nonstandard real number). Then, by Theorem \ref{Gleason}, if $\delta$ was chosen small enough, for all $x \in (A^{\prime \prime})^{10}$ we have $\Vert x \Vert_{A^{ \prime}} \leq 100 \delta $, and \begin{center} $ \Vert [u, x] \Vert _{A^{\prime}} = O\left( \Vert u \Vert _{A^{\prime}} \Vert x \Vert _{A^{\prime}} \right) < \Vert u \Vert _{A^{\prime}} $. \end{center} Since $\Vert u \Vert_{A^{\prime}}$ was minimal, $u $ commutes with every element in $(A^{\prime \prime })^{10}$. Consequently, if we define \[Z := \{ u^n \vert \vert n \vert \leq 1/ \Vert u \Vert _{A^{\prime}} \} ,\] then every element of $Z$ will commute with every element of $(A^{\prime \prime })^{10}$. Since $\left( A^{\prime \prime} \right)^6$ is well defined, by Theorem ``local group quotient'' we can form the quotients $A^{\prime \prime} /Z : = \lim_{i \to \alpha}(A^{\prime \prime }_i/Z_i) $ and $A^{\prime \prime \prime } /Z : = \lim_{i \to \alpha}(A^{\prime \prime \prime }_i/Z_i) $. \begin{proposition} \rm (Breuillard--Green--Tao) The image $\sigma (Z)$ is of the form $\phi ( [-1,1] )$, with $\phi (t) := \exp (tv)$, for some non-zero $v$ in the center of $\mathfrak{ l}$. By choosing a small open neighborhood of the identity $U\subset L$, we can form the quotient $U/\sigma (Z) $, and by choosing $\delta $ small enough, one can guarantee that \begin{itemize} \item $U/\sigma (Z)$ is a connected local Lie group of dimension $r-1$. \item $A^{\prime\prime} / Z$ and $A^{\prime\prime\prime}/Z$ are NSS ultra-approximate groups. \item $\overline{\sigma} \left(A^{\prime\prime }/Z \right)^8 \to U / \sigma (Z)$ is a good model. \end{itemize} \end{proposition} By this proposition, we can apply the induction hypothesis to $A^{\prime \prime \prime}/Z$ and conclude that there is a non-degenerate nilprogression $\overline{P}(\overline{u}_1, \ldots , \overline{ u}_{r-1} ; \overline{N_1}, \ldots , $ $\overline{N}_{r-1} )$ in $\overline{C}$-regular form such that $\overline{P} \subset (A^{\prime \prime \prime}/ Z)^4 \subset A^{\prime \prime}/Z$ with the property that for all standard $\varepsilon > 0$, there is an open set $V_{\varepsilon} \subset U / \sigma (Z)$ with $( \overline{\sigma})^{-1} (V_{\varepsilon}) \subset G(\varepsilon \overline{P})$. Let $\pi: A^{\prime \prime} \to A^{\prime \prime }/Z$ be the natural projection. To properly ``lift'' $\overline{P}$ to $A^{\prime \prime}$ through $\pi$, we need the following result. \begin{lemma}\label{lifting} \rm (Breuillard--Green--Tao) For every $w \in A^{\prime \prime }/ Z$, there is $w^{\prime} \in A^{\prime \prime}$ with $\pi (w^{\prime}) =w$, and $\Vert w^{\prime } \Vert _{A^{\prime \prime} } = O( \Vert w \Vert_{A^{\prime \prime}/Z} )$. \end{lemma} Construct $P : =P( u_1, \ldots , u_{r};$ $ N_1, \ldots , N_{r})$, where $u_j \in A^{\prime \prime}$ is a lift of $\overline{u}_j$ that minimizes $\Vert u_j \Vert _{A^{\prime \prime }}$ for $j=1, \ldots , r-1$, \begin{center} $N_j : = \delta_0 \overline{N}_j$ for $j=1, \ldots , r-1$, $u_{r} : = u$, $N_{r }: = \delta_0 / \Vert u \Vert _{A^{\prime \prime}}$. \end{center} \begin{proposition} \rm (Breuillard--Green--Tao) If $\delta_0 >0$ is small enough, then $P$ is a non-degenerate ultra-nilprogression in $C_0$-regular form for some $C_0 > 0$. \end{proposition} The only thing left to prove is that for all $\varepsilon > 0$, there is an open $U_{\varepsilon} \subset L$ such that $\sigma ^{-1} (U_{\varepsilon}) \subset G( \varepsilon P)$. By contradiction, assume that for some $\varepsilon > 0$, the element $x$ of $A^{\prime \prime }\backslash G( \varepsilon P)$ with minimal norm $\Vert x \Vert _{A^{\prime \prime}}$ satisfies $\sigma (x) = e_L$. If that is the case, $ \overline{\sigma } (x) = e_{L/ \sigma (Z)}$, and by our induction hypothesis, for all standard $\eta >0$, we have $\pi (x) \in G( \eta \overline{P})$. Therefore $x = u_1 ^{n_1} \ldots u_{r} ^{n_r}$, with \begin{center} $\vert n_j \vert \leq \eta \overline{N_j} /\overline{C}$ for $j=1, \ldots , r-1$, $\vert n_{r} \vert \leq 1/ \Vert u_{r } \Vert_{A^{\prime }} $. \end{center} Also, from Theorem \ref{Gleason}, Lemma \ref{lifting}, and the fact that $\overline{N}_j = O(1/\Vert \overline{u}_j \Vert _{A^{\prime \prime }/Z})$, we get \begin{eqnarray*} \Vert u_1^{n_1} \ldots u_{r-1}^{n_{r-1}} \Vert _{A^{\prime \prime}} &= & O \left( \sum_{j=1}^{r-1} \Vert u_j ^{n_j} \Vert_{A^{\prime \prime}} \right)\\ & = & O\left( \sum_{j=1}^{r-1} \vert n_j \vert \Vert u_j \Vert_{A^{\prime \prime}} \right) \\ & = & O \left( \eta \sum_{j=1}^{r-1} \overline{N_j} \Vert \overline{u}_j \Vert_{A^{\prime \prime}/Z} \right) \\ & = & O (\eta). \end{eqnarray*} Since $\eta$ was arbitrary, we obtain that $\Vert u_1^{n_1} \ldots u_{r-1}^{n_{r-1}} \Vert _{A^{\prime \prime}}$ is infinitesimal. Using once more Theorem \ref{Gleason}, \begin{center} $\Vert u_{r} ^{n_{r}} \Vert _{ A^{\prime \prime} } = O( \Vert x \Vert_{A^{\prime \prime}} + \Vert u_1^{n_1} \ldots u_{r-1}^{n_{r-1}} \Vert _{A^{\prime \prime}} ) $. \end{center} This implies that $ \Vert u_{r } ^{n_{r}} \Vert _{ A^{\prime \prime} } $ is infinitesimal and $\vert n_{r} \vert = o (N_{r} ) \leq \varepsilon N_{r} / C_0 $. Also, since $\eta$ was arbitrary, $\vert n_j \vert \leq \varepsilon N_j /C_0$ for $j=1, \ldots , r-1$. Therefore $x \in G(\varepsilon P)$, which is a contradiction. \end{proof} \begin{remark}\label{strong-basis} \rm Note that from the proof of Theorem \ref{Nilprog}, the group $L$ is nilpotent and we can assume the basis $\{ l_1, \ldots , l_r \} $ of $ \mathfrak{ l}$ given by $$ \exp ( l_j ) = \sigma \left( u_j ^{\left\lfloor N_j /C_0 \right\rfloor } \right) \text{ for }j= 1, \ldots , r, $$ is a strong Malcev basis. \end{remark} Let $r: = dim (X)$. We conclude this section with the following proposition, obtained by applying Theorem \ref{Nilprog} to the good model $\tilde{t}: Y ^8\to X $. \begin{proposition}\label{nilprog2} \rm There is $C_0>0,$ and for each $i \in \mathbb{N}$, elements $u_{1,i}, \ldots , u_{r,i} \in Y_i$, $N_1 (i), \ldots , N_r (i) \in \mathbb{N}$ with $ N_j (i) \xrightarrow{\alpha} \infty $ for each $j = 1, \ldots , r ,$ such that for $i$ close enough to $\alpha$, \begin{center} $P_i:=P(u_{1,i} , \ldots , u_{r,i} ; N_1 (i) , \ldots , N_r (i))$ \end{center} is a nilprogression in $C_0$-regular form contained in $Y_i^4$, and such that for every $\varepsilon , \delta \in (0, 1)$, there are $\varepsilon_0, \delta_0 >0$ with \[ \{ g \in \hat{\pi}(\Theta_i) \vert d( \overline{t}(g), p ) < \delta _0 \} \subset G(\varepsilon P_i) \] \[ G(\varepsilon_0 P_i) \subset \{ g \in \hat{\pi}(\Theta_i) \vert d( \overline{t}(g), p ) < \delta \}. \] \end{proposition} \section{Malcev theory}\label{Globalsection} In this section, we use the nilprogressions constructed in Proposition \ref{nilprog2} to build a sequence of simply connected nilpotent Lie groups $H_i$ converging in a suitable sense to $\tilde{X}$. The proof of the following lemma can be found in (\cite{BGT}, Appendix C). \begin{lemma}\label{DoublingNilprogression} \rm Let $r\in \mathbb{N}, C_0>1$, then there exist $M_0, \delta_0 > 0$ such that, for any nilprogression $P$ of rank $r$ in $C_0$-regular form with thick$(P) > M_0$, we have $$ G(\delta_0 P )^{2} \subset G(P) . $$ \end{lemma} Let $\mathfrak{g} $ be the lie algebra of $\tilde{X}$, and let $\delta_0 >0$ be given by Lemma \ref{DoublingNilprogression} with $r,C_0,$ from Proposition \ref{nilprog2}. By Remark \ref{strong-basis}, we can assume the basis $\{ v_1, \ldots , v_r \} \subset \mathfrak{g}$ given by \begin{equation}\label{malcev-basis-construction} \exp ( v_j) := \lim\limits_{i\to \alpha} \overline{t}_{\tilde{X}} \left( u_{j,i}^{\left\lfloor \frac{\delta_0 N_j(i)}{C_0} \right\rfloor} \right) \text{ for each }j = 1, \ldots , r, \end{equation} is a strong Malcev basis, and hence the map $\psi : \mathbb{R}^r \to \tilde{X}$ given by $$ \psi (x_1, \ldots, x_r ) : = \exp (x_1 v_1) \ldots \exp (x_r v_r) $$ is a diffeomorphism. \begin{lemma}\label{Yiren} \rm The group structure $Q: \mathbb{R}^r \times \mathbb{R}^r \to \mathbb{R}^r$ given by $$ Q(x,y) := \psi^{-1} ( \psi (x) \psi (y) ) $$ is a quasilinear polynomial of degree $\leq d(r)$. We will denote the group $(\mathbb{R}^r, Q)$ as $H$. \end{lemma} \begin{proof} By the Baker--Campbell--Hausdorff formula, after identifying $\mathfrak{g}$ with $\mathbb{R}^r$ via the basis $\{ v_1, \ldots , v_r \}$, the map $\mathbb{R}^r \times \mathbb{R}^r \to \mathbb{R}^r$ given by $$ (x,y) \to \exp ^{-1} ( \psi (x) \psi (y) ) $$ is polynomial of degree $\leq r$. Also, by Theorem \ref{simply-connected-nilpotent}, the map $\mathbb{R}^r \to \mathbb{R}^r$ given by $$x \to \psi ^{-1} ( \exp (x) )$$ is polynomial of degree bounded by a number depending only on $r$. Therefore the composition is also polynomial of degree $\leq d(r)$. Quasilinearity is immediate from the definition. \end{proof} \begin{definition} \rm Let $P(v_1 , \ldots , v_r ; N_1 , \ldots , N_r)$ be a nilprogression in $C_0$-regular form. Assume $N_j\geq C_0$ for each $j$, and set $\Gamma_P$ to be the abstract group generated by $\gamma_1, \ldots , \gamma_r$ with relations $[\gamma _j , \gamma_k] = \gamma_{k+1}^{\beta_{j, k}^{ k+1}} \ldots \gamma_r^{\beta_{j, k}^r} $ whenever $j < k $, where $[v _j , v_k] = v_{k+1}^{\beta_{j, k}^{ k+1}} \ldots v_r^{\beta_{j, k}^{ r}} $ and $\vert \beta_{j, k}^{l } \vert \leq \frac{C_0N_l}{N_j N_k}$. We say that $P$ is \textit{good} if $\Gamma_P$ is isomorphic to a lattice in a simply connected nilpotent Lie group, and each element of $\Gamma _P$ has a unique expression of the form \begin{center} $ \gamma_1^{n_1}\ldots \gamma_r^{n_r}, $ with $n_1, \ldots, n_r \in \mathbb{Z}$. \end{center} \end{definition} The proof of the following result can be found in (\cite{Cle}, Section 4.2). \begin{theorem}\label{Embed} \rm (Malcev) Let $r \in \mathbb{N} , C_0 >0 $, and $P(v_1 , \ldots , v_r ;$ $ N_1 , \ldots , N_r)$ a nilprogression in $C_0$-regular form in a group $\Gamma$. If thick$(P)$ is large enough depending on $r$ and $ C$, then $P$ is good and the map $v_j \to \gamma_j$ extends to an embedding $ \sharp : G(P) \to \Gamma_P.$ For $A \subset G(P)$, we will denote its image under this embedding by $A^{\sharp}$. Furthermore, there is a quasilinear polynomial group structure (see Definition \ref{quasil}) $$Q: \mathbb{R}^{r}\times \mathbb{R}^r \to \mathbb{R}^r $$ of degree $\leq d(r)$ such that the multiplication in $\Gamma_P$ is given by \begin{center} $ \gamma_1^{n_1}\ldots \gamma_r^{n_r} \gamma_1^{m_1}\ldots \gamma_r^{m_r} = \gamma_1^{(Q(n,m))_1}\ldots \gamma_r^{(Q(n,m))_r} $ for $n,m \in \mathbb{Z}^r.$ \end{center} $Q$ is called the \textit{Malcev polynomial} of $P$, and $(\mathbb{R}^r, Q)$ the \textit{Malcev Lie group} of $P$. $\Gamma_P$ is isomorphic, via $\gamma_j \to e_j$, to the lattice $(\mathbb{Z}^r, Q\vert_{\mathbb{Z}^r \times \mathbb{Z}^r} )$. \end{theorem} By Proposition \ref{nilprog2} and Theorem \ref{Embed}, for $\alpha$-large enough $i$, the nilprogressions $P_i$ are good with Malcev polynomials $\hat{Q}_i$. Let $N_0\in \mathbb{N}$ be given by Lemma \ref{PreWell} with $d (r)$ given by lemmas \ref{Yiren} and \ref{Embed}, and define $\xi : \mathbb{N}\to \mathbb{N}$ as $$ \xi (n) : = N_0 \left\lfloor \frac{\delta_0 n}{C_0N_0} \right\rfloor. $$ For $i\in \mathbb{N}$, consider $ \kappa_i : \mathbb{R}^r \to \mathbb{R}^r $ given by $$ \kappa_i(x_1, \ldots , x_r) : = (x_1 \xi (N_1 (i)), \ldots , x_r \xi (N_r (i))) .$$ Let $H_i$ be the group $(\mathbb{R}^r , Q_i)$, where $Q_i : \mathbb{R}^r \times \mathbb{R}^r \to \mathbb{R}^r $ is the group structure given by $$ Q_i(x,y) : = \kappa_i^{-1} (\hat{Q}_i ( \kappa_i(x) ,\kappa_i(y)) ) . $$ The following is the main result of this section. \begin{proposition}\label{Goooooood} \rm The sequence of quasilinear polynomial group structures $Q_i$ converges well to $Q$. \end{proposition} We define $\Omega \subset \mathbb{R}^r$ as $$ \Omega : = \left\{ -1, \ldots,\frac{-1}{N_0}, 0, \frac{1}{N_0}, \ldots , 1 \right\} ^{\times r} . $$ Consider the maps $\omega_{\alpha} : \Omega \to \tilde{X}$ and $\omega_i : \Omega \to \Gamma_{P_i}$ defined as $\omega_{\alpha} : = \psi \vert _{ \Omega }$ and \begin{center} $ \omega_i ( x_1, \ldots, x_r) := \gamma_1^{ \xi (x_1 N_1(i)) }\ldots \gamma_r^{ \xi (x_r N_r(i)) } . $ \end{center} We also define maps $\overline{t}^{\flat } : G(P_i )^{\sharp } \to X$ and $\overline{t}^{\flat }_{\tilde{X}} : G(P_i)^{\sharp} \to \tilde{X}$ as \begin{center} $ \overline{t}^{\flat} ( x^{\sharp } ) := \overline{t} (x)$ and $ \overline{t}_{\tilde{X}}^{\flat} ( x^{\sharp } ) := \overline{t}_{\tilde{X}} (x).$ \end{center} Consider the following diagram. \begin{displaymath} \begin{tikzcd}[column sep=3em] \Omega \times \Omega \arrow{r}{ \omega_i } \arrow{d}{ Id} & (G (\delta_0 P_i )^{\sharp})^{\times 2} \arrow{r}{\ast} \arrow{d}{ \overline{t} _{\tilde{X}} ^{\flat}}& G (P_i )^{\sharp} \arrow{d}{ \overline{t} _{\tilde{X}} ^{\flat} } \arrow{r}{\kappa_i^{-1}} & \mathbb{R}^r \arrow{d}{Id} \\ \Omega \times \Omega \arrow{r}{\omega_{\alpha} } & \tilde{X}\times \tilde{X} \arrow{r}{\ast }& \tilde{X} \arrow{r}{ \psi^{-1} } & \mathbb{R}^r \end{tikzcd} \end{displaymath} The first row of the diagram is the polynomial $Q_i$, while the second row is the polynomial $Q$. Commutativity of the diagram does not hold in general, but it holds in the limit, as the following proposition shows. \begin{proposition}\label{CD} \rm For every $x,y \in \Omega$, \[ \lim\limits_{i \to \alpha} \kappa_i^{-1} \left( \omega_i(x) \omega_i(y) \right) = \psi^{-1} ( \omega_{\alpha}(x)\omega_{\alpha}(y)) .\] \end{proposition} \begin{proof} We will first show that for any $x \in \Omega$, we have \begin{equation}\label{omegas} \lim_{ i \to \alpha } \overline{t}_{\tilde{X}}^{\flat }\left( \omega_i (x) \right) = \omega_{\alpha} (x). \end{equation} Using Equation \ref{malcev-basis-construction} and Corollary \ref{holonomy2}, we obtain \begin{eqnarray*} \omega_{\alpha }(x) & = & \exp (x_1 v_1) \ldots \exp (x_r v_r)\\ & = & \lim_{i \to \alpha} \overline{t}_{\tilde{X}}^{\flat } \left( \gamma_1^{ \xi ( x_1 N_1 (i)) } \right) \ldots \lim_{i \to \alpha} \overline{t}_{\tilde{X}}^{\flat } \left( \gamma_{r}^{ \xi ( x_r N_r (i)) } \right)\\ & = & \lim_{i \to \alpha} \overline{t}_{\tilde{X}}^{\flat } (\omega_i(x)). \end{eqnarray*} We then show that for any sequence $x_i^{\sharp} \in G(P_i)^{\sharp}$, we have \begin{equation}\label{PreCD} \psi \left( \lim\limits_{i \to \alpha} \kappa_i^{-1}(x_i^{\sharp} ) \right) = \lim\limits_{i\to \alpha} \overline{t}_{\tilde{X}}^{\flat} ( x_i^{\sharp} ) . \end{equation} We can decompose the sequence $x_i = u_{1,i}^{p_{1,i}}\ldots u_{r,i}^{p_{r,i}} \in G(P_i)$ as \begin{center} $ x_i = x_{1,i}\ldots x_{r,i} , $ with $x_{j,i} = u_{j,i}^{p_{j,i}} \in G(P_i)$. \end{center} Then, by Equation \ref{malcev-basis-construction} and Corollary \ref{holonomy2}, \begin{eqnarray*} \lim\limits_{i \to \alpha}\overline{t}_{\tilde{X}} (x_i ) & = & \lim\limits_{i \to \alpha} \overline{t}_{\tilde{X}} (x_{1,i}) \ldots \lim\limits_{i \to \alpha} \overline{t}_{\tilde{X}} (x_{r,i})\\ & = & \exp \left( \lim\limits_{i \to \alpha} \dfrac{C_0p_{1,i}}{\delta_0N_1(i)} v_1 \right)\ldots \exp \left( \lim\limits_{i \to \alpha} \dfrac{C_0p_{r,i}}{\delta_0N_r(i)} v_r \right)\\ & = & \psi \left( \dfrac{C_0}{\delta_0} \lim\limits_{i \to \alpha} \left( \dfrac{p_{1,i}}{N_1(i)}, \ldots , \dfrac{p_{r,i}}{N_r(i)} \right) \right) \\ & = & \psi \left( \lim\limits_{i \to \alpha} \kappa_i^{-1} ( x_i ) \right) . \end{eqnarray*} From equations \ref{omegas} and \ref{PreCD}, we conclude \begin{eqnarray*} \omega_{\alpha}(x) \omega_{\alpha}(y) & = & \lim_{i \to \alpha}\overline{t}_{\tilde{X}}^{\flat} ( \omega_i (x) ) \lim_{i \to \alpha}\overline{t}_{\tilde{X}}^{\flat} ( \omega_i (y) ) \\ & = & \lim_{i \to \alpha} \overline{t}_{\tilde{X}}^{\flat} (\omega_i (x) \omega_i (y))\\ & = & \psi \left( \lim\limits_{i \to \alpha} \kappa_i^{-1} \left( \omega_i(x) \omega_i(y) \right) \right). \end{eqnarray*} \end{proof} Proposition \ref{Goooooood} follows from Lemma \ref{PreWell} and Proposition \ref{CD}. By Lemma \ref{Well}, the sequence of Lie algebras of the groups $H_i$ converges well to $\mathfrak{g}$. Consider the morphisms $ \natural : \Gamma_{P_i} \to H_i $ and $\natural : \tilde{X} \to H$ given by $g^{\natural } : = \kappa_i^{-1}(g)$ for $g \in \Gamma_{P_i}$, and $g^{\natural}: = \psi^{-1}(g) $ for $g \in \tilde{X}$. These identifications allow us to have all these group structures in the same underlying set $\mathbb{R}^r$. \section{Almost torsion elements}\label{Rootsection} In this section we finish the proof of Theorem \ref{PIII} by establishing the following result. \begin{proposition}\label{final} \rm For $\alpha$-large enough $i$, there are groups $\Lambda_i \leq \pi_1(X_i)$ with surjective morphisms $\Lambda_i \to \pi_1(X)$. \end{proposition} Recall that $Z_i := \Gamma_i / W_i$, and define $q_i : = W_i (p_i) \in X_i / W_i $. Let $\eta > 0$ be small enough so that for $\alpha$-large enough $i$, $$ S_i : = \{ g \in Z_i \vert d(g(q_i) , q_i ) < \eta \} \subset G(\delta_0 P_i). $$ Let $\tilde{Z}_i$ be the abstract group generated by $S_i$, with relations \begin{center} $ s= s_1s_2 \in \tilde{Z}_i $ whenever $ s,s_1, s_2 \in S_i$ and $s=s_1s_2$ in $Z_i$. \end{center} Notice that for $\alpha$-large enough $i$, we have $\tilde{Z}_i = \Gamma_{P_i}$, and by Theorem \ref{monodromy}, there is a Galois $(\eta /3)$-wide covering map $\tilde{X}_i^{\prime} \to X_i/W_i$ whose Galois group is the kernel of the canonical map $\Phi_i : \Gamma_{P_i} \to Z_i$. Proposition \ref{final} will follow from the following result. \begin{proposition}\label{final2} \rm For $\alpha$-large enough $i$, there are subgroups $\Lambda_{i}^{\prime} \leq Ker (\Phi_i)$ isomorphic to $\pi_1(X)$. \end{proposition} \begin{proof}[Proof of Proposition \ref{final}:] By Theorem \ref{sormani-cover-copy} and Equation \ref{quotient-gh-close}, for $\alpha$-large enough $i$, there are Galois covering spaces $\tilde{X}_i \to X_i$ with Galois groups $Ker (\Phi_i)$. By Lemma \ref{cover-characterization}, for $\alpha$-large enough $i$, there are surjective morphisms $\rho _i : \pi_1 (X_i) \to Ker (\Phi_i)$. Proposition \ref{final} is obtained by defining $\Lambda_i : = \rho^{-1}_i( \Lambda_i^{\prime} ) \leq \pi_1(X_i)$, where $\Lambda_i^{\prime}$ are given by Proposition \ref{final2}. \end{proof} \begin{remark}\label{phi-locally-injective} \rm Notice that if $i$ is $\alpha$-large enough, then for every $g \in \Gamma_{P_i} $ with $$d ( \Phi_i (g)( q_i) , q_i ) < \eta ,$$ there is a unique $w \in G( \delta_0 P_i )^{\sharp}$ such that $gw \in Ker (\Phi_i)$. Also, $Ker(\Phi_i ) \cap G(P_i)^{\sharp } = \{ e_{\Gamma_{P_i}} \}$. \end{remark} \begin{proof}[Proof of Proposition \ref{final2}:] Let $\Phi : \tilde{X} \to X$ denote the canonical projection. Then by Lemma \ref{discrete-normal-central}, $Ker (\Phi) \cong \pi_1(X) $ is central, and since $\tilde{X}$ has no torsion, $Ker (\Phi )$ is free abelian. Let $\{ \lambda_1 , \ldots , \lambda_{\ell} \} $ be a basis of $Ker (\Phi )$ as a free abelian group. Pick $M \in \mathbb{N}$ large enough so that the $M$-th roots of the $\lambda_j$'s lie in the ball $B^{\tilde{X}}(e, \eta /2)$. For each $j \in \{ 1, \ldots , \ell \}$ pick a sequence $$ \lambda _j (i) \in G(P_i)^{\sharp} \subset \Gamma_{P_i} $$ with $$ \lim\limits_{i \to \alpha} \overline{t}_{\tilde{X}}^{\flat} (\lambda_j (i)) = \dfrac{\lambda_j}{M}. $$ Since $Q_i$ converges well to $Q$, Equation \ref{PreCD} implies that \begin{equation}\label{well-lambda} \lim\limits_{i \to \alpha} (\lambda_j (i)^{\natural})^M = \lambda_j^{\natural} . \end{equation} Also, for each $j \in \{ 1, \ldots , \ell \}$ we have \begin{eqnarray*} \lim\limits_{i \to \alpha} \overline{t} \left( \Phi_i \left( \lambda_j(i) \right) ^M \right) & = & \left( \lim\limits_{i \to \alpha} \overline{t} \left( \Phi_i \left( \lambda_j(i) \right) \right)\right)^M\\ & = & \left( \lim\limits_{i\to \alpha} \overline{t}^{\flat} ( \lambda_j (i) ) \right)^M \\ & = & \Phi \left( \dfrac{\lambda_j}{M} \right)^M\\ & = & e_X. \end{eqnarray*} By Remark \ref{phi-locally-injective}, for $\alpha$-large enough $i$, and all $j \in \{ 1 , \ldots , \ell \}$, there are $w_{j,i}\in G( P_i)^{\sharp}$ such that $$ \lambda_j (i)^M w_{j,i} \in Ker (\Phi_i).$$ Since $Q_i$ converges well to $Q$, we deduce from Equation \ref{well-lambda} that $$ \lim\limits_{i \to \alpha} \left[ \lambda_j(i) ^M w_{j,i}\right]^{\natural} = \lambda_j^{\natural} , $$ and hence by Lemma \ref{Quasilinear Continuity}, $$ \lim\limits_{i \to \alpha} \log_i \left( \left[ \lambda_j(i) ^M w_{j,i}\right]^{\natural} \right) = \log (\lambda_j^{\natural}) \text{ for each }j \in \{ 1, \ldots , \ell \} , $$ where $\log_i, \log,$ denote the logarithm maps with respect to $H_i$ and $H$, respectively. Therefore for $\alpha$-large enough $i$, the set $$\left\{ \log_i \left( \left[ \lambda_1(i) ^M w_{1,i}\right]^{\natural} \right) , \ldots , \log_i \left( \left[ \lambda_{\ell}(i) ^M w_{\ell ,i}\right]^{\natural} \right) \right\} $$ is linearly independent. Also, for $j,k \in \{ 1, \ldots , \ell \}$, \begin{eqnarray*} \lim\limits_{i \to \alpha} [ \lambda_j (i)^M w_{j,i},\lambda_k(i) ^M w_{k,i} ]^{\natural} &=& \left[ \lim\limits_{i \to \alpha}\lambda_j(i)^Mw_{j,i}, \lim\limits_{i \to \alpha}\lambda_k(i)^Mw_{k,i} \right]^{\natural} \\ & = & \left( \lambda_j\lambda_k\lambda_j^{-1} \lambda_k^{-1} \right)^{\natural} \\ & = & e_{\tilde{X}}^{\natural}. \end{eqnarray*} By Remark \ref{phi-locally-injective}, this implies that for $\alpha$-large enough $i$, $$ [ \lambda_j (i)^M w_{j,i},\lambda_k(i) ^M w_{k,i} ] =e_{\Gamma_{P_n}},$$ and the group $$ \langle \lambda_1 (i)^M w_{1,i}, \ldots , \lambda_{\ell} (i)^M w_{\ell , i} \rangle \leq Ker \left( \Phi_i \right) $$ is a free abelian group of rank $\ell$, hence isomorphic to $ Ker (\Phi) $ and $\pi_1(X)$. \end{proof} \begin{proof}[Proof of Theorem \ref{PIII}:] Assume by contradiction that the theorem fails for some sequence $X_i$. After taking a subsequence, we can assume that for no $i$, there is a subgroup $\Lambda_i \leq \pi_1(X_i)$ with a surjective morphism $\Lambda_i \to \pi_1(X)$. This contradicts Proposition \ref{final} for $\alpha$-large enough $i$. \end{proof}
{ "timestamp": "2021-12-01T02:13:58", "yymm": "2007", "arxiv_id": "2007.01985", "language": "en", "url": "https://arxiv.org/abs/2007.01985", "abstract": "We say that a sequence of proper geodesic spaces $X_i$ consists of \\textit{almost homogeneous spaces} if there is a sequence of discrete groups of isometries $G_i \\leq Iso(X_i)$ such that diam$(X_i/G_i)\\to 0$ as $i \\to \\infty$.We show that if a sequence $X_i$ of almost homogeneous spaces converges in the pointed Gromov--Hausdorff sense to a space $X$, then $X$ is a nilpotent locally compact group equipped with an invariant geodesic metric.Under the above hypotheses, we show that if $X$ is semi-locally-simply-connected, then it is a nilpotent Lie group equipped with an invariant Finsler or sub-Finsler metric, and for large enough $i$, there are subgroups $\\Lambda_i \\leq \\pi_1(X_i) $ with surjective morphisms $\\Lambda_i\\to \\pi_1(X)$.", "subjects": "Metric Geometry (math.MG); Group Theory (math.GR)", "title": "Limits of almost homogeneous spaces and their fundamental groups", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9843363522073338, "lm_q2_score": 0.8128673178375734, "lm_q1q2_score": 0.8001348504687964 }
https://arxiv.org/abs/1311.6176
Inverse questions for the large sieve
Suppose that an infinite set $A$ occupies at most $\frac{1}{2}(p+1)$ residue classes modulo $p$, for every sufficiently large prime $p$. The squares, or more generally the integer values of any quadratic, are an example of such a set. By the large sieve inequality the number of elements of $A$ that are at most $X$ is $O(X^{1/2})$, and the quadratic examples show that this is sharp. The simplest form of the inverse large sieve problem asks whether they are the only examples. We prove a variety of results and formulate various conjectures in connection with this problem, including several improvements of the large sieve bound when the residue classes occupied by $A$ have some additive structure. Unfortunately we cannot solve the problem itself.
\subsection[#1]{\sc #1}} \newcommand\E{\mathbb{E}} \newcommand\Z{\mathbb{Z}} \newcommand\R{\mathbb{R}} \newcommand\T{\mathbb{T}} \newcommand\C{\mathbb{C}} \newcommand\N{\mathbb{N}} \newcommand\A{\mathscr{A}} \newcommand\B{\mathscr{B}} \newcommand\G{\mathbf{G}} \newcommand\SL{\operatorname{SL}} \newcommand\Upp{\operatorname{Upp}} \newcommand\im{\operatorname{im}} \newcommand\bad{\operatorname{bad}} \newcommand\bes{\operatorname{bes}} \newcommand\GL{\operatorname{GL}} \newcommand\Mat{\operatorname{Mat}} \newcommand\Alg{\operatorname{Alg}} \newcommand\Ad{\operatorname{Ad}} \newcommand\tr{\operatorname{tr}} \newcommand\dist{\operatorname{dist}} \newcommand\vol{\operatorname{vol}} \newcommand\res{\operatorname{res}} \newcommand\nonres{\operatorname{nonres}} \newcommand\unif{\operatorname{unif}} \renewcommand\P{\mathbb{P}} \newcommand\F{\mathbb{F}} \newcommand\Q{\mathbb{Q}} \renewcommand\b{{\bf b}} \newcommand\g{\mathfrak{g}} \newcommand\h{\mathfrak{h}} \newcommand\n{\mathfrak{n}} \renewcommand\a{\mathfrak{a}} \newcommand\p{\mathfrak{p}} \newcommand\q{\mathfrak{q}} \renewcommand\b{\mathfrak{b}} \renewcommand\r{\mathfrak{r}} \newcommand\eps{\varepsilon} \newcommand\id{{\operatorname{id}}} \newcommand\st{{\operatorname{st}}} \newcommand\diam{{\operatorname{diam}}} \newcommand\proofsymb{{\hfill{\usebox{\proofbox}}\vspace{11pt}}} \renewcommand{\labelenumi}{(\roman{enumi})} \newcommand{{{}^*}}{{{}^*}} \onehalfspace \parindent 5mm \parskip 0mm \begin{document} \title{Inverse questions for the large sieve} \author[Green]{Ben Green} \address{Mathematical Institute\\ Radcliffe Observatory Quarter\\ Woodstock Road\\ Oxford OX2 6GG\\ England } \email{[email protected]} \author[Harper]{Adam J Harper} \address{Jesus College \\ Cambridge CB5 8BL\\ England } \email{[email protected]} \begin{abstract} Suppose that an infinite set $\A$ occupies at most $\frac{1}{2}(p+1)$ residue classes modulo $p$, for every sufficiently large prime $p$. The squares, or more generally the integer values of any quadratic, are an example of such a set. By the large sieve inequality the number of elements of $\A$ that are at most $X$ is $O(X^{1/2})$, and the quadratic examples show that this is sharp. The simplest form of the \emph{inverse large sieve problem} asks whether they are the only examples. We prove a variety of results and formulate various conjectures in connection with this problem, including several improvements of the large sieve bound when the residue classes occupied by $\A$ have some additive structure. Unfortunately we cannot solve the problem itself. \end{abstract} \maketitle \begin{center}\emph{To Roger Heath-Brown on his 60th birthday}\end{center} \setcounter{tocdepth}{1} \tableofcontents \section{Introduction} \textsc{Notation.} Most of our notation is quite standard. When dealing with infinite sets $\A$, we write $\A[X]$ for the intersection of $\A$ with the initial segment $[X] := \{1,\dots, X\}$. Our primary aim in this paper is to study sets $\A$ of integers with the property that the reduction $\A \md{p}$ occupies at most $\frac{1}{2}(p+1)$ residue classes modulo $p$ for all sufficiently large primes $p$. It follows from the large sieve that $|\A[X]| \ll X^{1/2}$ for all $X$ (we will recall the details of this argument below). This is clearly sharp up to the value of the implied constant, as shown by taking $\A$ to be the set of squares or more generally the set of integer values taken by any rational quadratic, that is to say quadratic with rational coefficients. It has been speculated, most particularly by Helfgott and Venkatesh \cite[pp 232--233]{helfgott-venkatesh} and by Walsh \cite{walsh}, that quadratics provide the only examples of sets for which the large sieve bound is essentially sharp. See also \cite[Problem 7.4]{croot-lev}. One might call problems of this type the ``inverse large sieve problem''. Unfortunately, we have not been able to prove any statement of this kind, and our aims here are more modest. Suppose that $\A \md{p} \subset S_p$ for all sufficiently large primes $p$. Our first set of results consists of improvements to the large sieve bound when $S_p$ looks very much unlike a quadratic set modulo $p$, for example by having some additive structure. \begin{theorem}\label{thm1.4} Suppose that for each prime $p \leq X^{1/2}$ one has a set $S_p \subset \Z/p\Z$ of size $(p+1)/2$. Suppose there is some $\delta > 0$ such that, for each $p$, $S_p$ has at least $(\frac{1}{16} + \delta) p^3$ additive quadruples, that is to say quadruples $(s_1,s_2,s_3,s_4)$ with $s_1 + s_2 = s_3 + s_4$. Suppose that $\A \mdlem{p} \subset S_p$ for all $p$. Then $|\A[X]| \ll X^{1/2 - c\delta^2}$, where $c > 0$ is an absolute constant.\end{theorem} The condition of having at least $(\frac{1}{16} + \delta) p^3$ additive quadruples will {\em not} be satisfied by a generic (e.g. randomly selected) set $S_p \subset \Z/p\Z$ of size $(p+1)/2$, which will have $(\frac{1}{16} + o(1)) p^3$ such quadruples for large $p$. But it is a rather general condition corresponding to $S_p$ being additively structured, and certainly we are not aware of any previous improvements to the large sieve bound under comparably general conditions. An extreme case of the preceding theorem is that in which $S_p$ is in fact an interval. Here a simple calculation, reproduced later, shows that Theorem \ref{thm1.4} is applicable with the choice $\delta=1/48$, but we can do rather better. \begin{theorem}\label{interval-sieve} Suppose that $\A$ is a set of integers and that, for each prime $p$, the set $\A \mdlem{p}$ lies in some interval $I_p$. Let $\eps > 0$ be arbitrary. Then \begin{enumerate} \item If $|I_p| \leq (1 - \eps) p$ for at least a proportion $\eps$ of the primes in each dyadic interval $[Z,2Z]$ then $|\A[X]| \ll_{\eps} (\log \log X)^{C\log(1/\eps)}$, where $C > 0$ is some absolute constant; \item If $|I_p| \leq \frac{p}{2}$ for all primes then $|\A[X]| \ll (\log \log X)^{\gamma + o(1)}$, where $\gamma = \frac{\log 18}{\log(3/2)} \approx 7.129$; \item If $|I_p| = [\alpha p, \beta p]$ for all primes $p$ and for fixed $0 \leq \alpha < \beta < 1$ \textup{(}not depending on $p$\textup{)} then $|\A| = O_{\beta - \alpha}(1)$; \item There are examples of infinite sets $\A$ with $|I_p| \leq (\frac{1}{2} + \eps) p$ for all $p$. \end{enumerate} \end{theorem} This improves on the results of an unpublished preprint~\cite{green} by the first author, in which it was shown that one has $|\A[X]| \ll X^{1/3+o(1)}$ under the condition (ii). Theorem \ref{thm1.4} also covers the case in which $S_p$ is an arithmetic progression of length $\frac{1}{2}(p+1)$, where the common difference of this arithmetic progression may depend on $p$. Here again one could apply Theorem \ref{thm1.4} with the choice $\delta=1/48$, but we can also handle this situation with a less restrictive condition on the size of $S_p$. \begin{theorem}\label{prog-thm} Let $\eps > 0$. Suppose that $\A$ is a set of integers and that, for each prime $p \leq X^{1/2}$, the set $\A \mdlem{p}$ lies in some arithmetic progression $S_p$ of length $(1-\eps)p$. Then $|\A[X]| \ll_{\eps} X^{1/2 - \eps'}$, where $\eps' > 0$ depends on $\eps$ only. \end{theorem} We are not aware of any previous results improving the large sieve bound $X^{1/2}$ when the $S_p$ are arbitrary arithmetic progressions, even for $\eps = 1/2$. \vspace{11pt} After proving the foregoing results, we turn to the ``robustness'' of the inverse large sieve problem. The aim of these results is to show that if $|\A \md{p}| \leq \frac{1}{2}(p+1)$ (or if similar conditions hold), if $|\A[X]| \approx X^{1/2}$, and if $\A$ is even vaguely close to quadratic in structure, then it must in fact approximate a quadratic very closely. Our proof methods here lead to some complicated dependencies between parameters, so we do not state and prove the most general result possible, settling instead for a couple of statements that have relatively clean formulations. The first and main one concerns finite sets. Here, and henceforth in the paper, we say that a rational quadratic $\psi$ has \emph{height at most $H$} if it can be written as $\psi(x) = \frac{1}{d}(ax^2 + bx + c)$ with $a,b,c,d \in \Z$ and $\max(|a|, |b|, |c|, |d|) \leq H$. \begin{theorem}\label{stability-thm} Let $X_0 \in \N$, and let $\eps > 0$. Let $X \in \N$ be sufficiently large in terms of $X_0$ and $\eps$, and suppose that $H \leq X^{1/8}$. Suppose that $A,B \subset [X]$ and that $|A \md{p}| + |B \md{p}| \leq p+1$ for all $p \in [X_0, X^{1/4}]$. Then, for some absolute constant $c > 0$, one of the following holds: \begin{enumerate} \item \textup{(Better than large sieve)} Either $|A \cap [X^{1/2}]|$ or $|B \cap [X^{1/2}]|$ is $\leq X^{1/4 - c\eps^3}$; \item \textup{(Behaviour with quadratics)} Given any two rational quadratics $\psi_A,\psi_B$ of height at most $H$, either $|A \setminus \psi_{A}(\Q)|$ and $|B\setminus \psi_{B}(\Q)| \leq HX^{1/2 - c}$, or else at least one of $|A \cap \psi_{A}(\Q)|$ and $|B \cap \psi_{B}(\Q)|$ is bounded above by $HX^{1/4 + \eps}$. \end{enumerate} \end{theorem} We expect that if the large sieve bound is close to sharp for $A$ and $B$, then there must exist rational quadratics of ``small'' height containing ``many'' points of $A$ and $B$. Together with Theorem \ref{stability-thm}, this provides some motivation for making the following conjecture of the form ``almost equality in the large sieve implies quadratic structure''. \begin{conjecture}\label{stability-conj} Let $X_0 \in \N$, and let $\rho > 0$. Let $X \in \N$ be sufficiently large in terms of $X_0$ and $\rho$. Suppose that $A,B \subset [X]$ and that $|A \md{p}| + |B \md{p}| \leq p+1$ for all $p \in [X_0, X^{1/4}]$. Then there exists a constant $c = c(\rho) > 0$ such that one of the following holds: \begin{enumerate} \item \textup{(Better than large sieve)} Either $|A \cap [X^{1/2}]|$ or $|B \cap [X^{1/2}]|$ is $\leq X^{1/4 - c}$; \item \textup{(Quadratic structure)} There are two rational quadratics $\psi_A,\psi_B$ of height at most $X^{\rho}$ such that $|A \setminus \psi_{A}(\Q)|$ and $|B\setminus \psi_{B}(\Q)| \leq X^{1/2 -c}$. \end{enumerate} \end{conjecture} The contents of Theorem \ref{stability-thm} and of Conjecture \ref{stability-conj} are perhaps a little hard to understand on account of the parameters $H, X, \rho$ and $\eps$. As a corollary we establish the following more elegant statement involving infinite sets. \begin{theorem}\label{stability-thm-symmetric} Suppose that $\A$ is a set of positive integers and that $|\A \md{p}| \leq \frac{1}{2}(p+1)$ for all sufficiently large primes $p$. Then one of the following options holds: \begin{enumerate} \item \textup{(Quadratic structure)} There is a rational quadratic $\psi$ such that all except finitely many elements of $\A$ are contained in $\psi(\Q)$; \item \textup{(Better than large sieve)} For each integer $k$ there are arbitrarily large values of $X$ such that $|\A[X]| < \frac{X^{1/2}}{\log^k X}$; \item \textup{(Far from quadratic structure)} Given any rational quadratic $\psi$, for all $X$ we have $|\A[X] \cap \psi(\Q)| \leq X^{1/4 + o_{\psi}(1)}$. \end{enumerate} \end{theorem} We conjecture that option (iii) is redundant. This is another conjecture of inverse large sieve type, rather cleaner than Conjecture \ref{stability-conj}. \begin{conjecture}\label{stability-conj-symmetric} Suppose that $\A$ is a set of positive integers and that $|\A \md{p}| \leq \frac{1}{2}(p+1)$ for all sufficiently large primes $p$. Then one of the following options holds: \begin{enumerate} \item \textup{(Quadratic structure)} There is a rational quadratic $\psi$ such that all except finitely many elements of $\A$ are contained in $\psi(\Q)$; \item \textup{(Better than large sieve)} For each integer $k$ there are arbitrarily large values of $X$ such that $|\A[X]| < \frac{X^{1/2}}{\log^k X}$. In particular, $\lim\inf_{X \rightarrow \infty} X^{-1/2} |\A[X]| = 0$. \end{enumerate} \end{conjecture} We remark that some very simple properties of rational quadratics are laid down in Appendix \ref{rat-quad-app}. In particular we draw attention to the fact that given a rational quadratic $\psi$ there are further rational quadratics $\psi_1,\psi_2$ such that $\psi_1(\Z) \subset \psi(\Q) \cap \Z \subset \psi_2(\Z)$. In particular, $|\psi(\Q) \cap [X]| \ll_{\psi} X^{1/2}$. \vspace{11pt} Our final task in the paper is to show, elaborating on ideas of Elsholtz \cite{elsholtz}, that Conjecture \ref{stability-conj} would resolve the currently unsolved ``inverse Goldbach problem'' of Ostmann \cite[p. 13]{ostmann} (and see also \cite[p. 62]{erdos}). This asks whether the set of primes can be written as a sumset $\A + \B$ with $|\A|, |\B| \geq 2$, except for finitely many mistakes. Evidently the answer should be that it cannot be so written. \begin{theorem}\label{ils-ostmann} Assume Conjecture \ref{stability-conj}. Let $\A, \B$ be two sets of positive integers, with $|\A|, |\B| \geq 2$, such that $\A + \B$ contains all sufficiently large primes. Then $\A + \B$ also contains infinitely many composite numbers. \end{theorem} We remark that much stronger statements that would imply this should be true, but we do not know how to prove them. For example, it is reasonable to make the following conjecture. \begin{conjecture} Let $\delta > 0$. Then if $X$ is sufficiently large in terms of $\delta$, the following is true. Let $A, B \subset [X]$ be any sets with $|A|, |B| \geq X^{\delta}$. Then $A + B$ contains a composite number. \end{conjecture} We do not know how to prove this for any $\delta \leq \frac{1}{2}$. If one had it for any $\delta < \frac{1}{2}$, the inverse Goldbach problem would follow. \vspace{11pt} The proofs of the above theorems are rather diverse and use the large sieve, Gallagher's ``larger sieve'', and several other tools from harmonic analysis and analytic number theory. The proofs of Theorems \ref{thm1.4} and \ref{prog-thm}, which involve lifting the additive structure of the $S_p$ to additive structure on $\A[X]$, also involve some ideas of additive combinatorial flavour, although we do not need to import many results from additive combinatorics to prove them. With very few exceptions (for example, Lemma \ref{progr-energy} depends on standard Fourier arguments given in detail in Lemma \ref{large-fourier}), Sections 3,4,5,6 and 7 may be read independently of one another. The situation considered in the majority of this paper, in which $\A \md{p}$ is small for \emph{every} prime $p$ (or at least for every prime $p \leq X^{1/2}$), may seem rather restrictive. It would be possible to adapt our arguments and prove many of our theorems under weaker conditions, and we leave this to the reader who has need of such results. However, it seems possible that any set $\A$ for which $|\A \md{p}| \leq (1 - c)p$ for a decent proportion of primes $p$ and for which $|\A[X]| \geq X^c$ for infinitely many $X$ has at least some ``algebraic structure''. Moreover such statements may well be true in finitary settings, in which $\A$ is restricted to some finite interval $[X]$ and $p$ is only required to range over some (potentially quite small) subinterval of $[X]$. Unfortunately none of our methods come close to establishing such strong results. \vspace{11pt} \textsc{Acknowledgements.} The authors are very grateful to Jean Bourgain for allowing us to use his ideas in Section \ref{interval-sieve-sec}. The first-named author is supported by ERC Starting Grant 279438 \emph{Approximate Algebraic Structure and Applications}. He is very grateful to the University of Auckland for providing excellent working conditions on a sabbatical during which this work was completed. The second-named author was supported by a Doctoral Prize from the EPSRC when this work was started; by a postdoctoral fellowship from the Centre de Recherches Math\'ematiques, Montr\'eal; and by a research fellowship at Jesus College, Cambridge when the work was completed. \section{The large sieve and the larger sieve} \textsc{The large sieve}. Let us begin by briefly recalling a statement of the large sieve bound. The following may be found in Montgomery \cite{montgomery-early}. \begin{proposition}\label{classicls} Let $\mathscr{A}$ be a set of positive integers with the property that $\mathscr{A} \mdlem{p} \subset S_p$ for each prime $p$. Then for any $Q,X$ we have the bound $$ |\mathscr{A}[X]| \leq \frac{X + Q^{2}}{\sum_{q \leq Q} \mu^{2}(q) \prod_{p \mid q} \frac{|S_p^c|}{|S_p|}} \leq \frac{X + Q^2}{\sum_{p \leq Q} \frac{|S_p^c|}{p}}. $$ where $\mu(q)$ denotes the M\"{o}bius function. \end{proposition} The second bound is a little crude but has the virtue of being simple: we will use it later on. In the particular case that $|S_p| \leq \frac{1}{2}(p+1)$ for all $p$, discussed in the introduction, the first bound implies upon setting $Q := X^{1/2}$ that $$ |\mathscr{A}[X]| \leq \frac{2X}{\sum_{q \leq X^{1/2}} \mu^{2}(q) \prod_{p \mid q} \frac{p-1}{p+1}} \ll X^{1/2}, $$ as we claimed. The large sieve may also be profitably applied to ``small sieve'' situations in which $|S_p| = p - O(1)$ (as opposed to ``large sieve'' situations in which $p - |S_p|$ is large). We will need one such result later on, in \S \ref{stab-sec}. \begin{lemma}\label{small-from-large} Suppose that $\mathscr{B} \subset \Z$ is a set with the property that $\mathscr{B} \mdlem{p}$ misses $w(p)$ residue classes, for every prime $p$. Suppose that the function $w$ has average value $k$ in the \textup{(}fairly weak\textup{)} sense that $\frac{1}{Z}\sum_{Z \leq p \leq 2Z} w(p) \log p = k + O(\frac{1}{\log^2 Z})$ for all $Z$. Then $|\mathscr{B} \cap [-N, N]| \ll N (\log N)^{-k}$ for all $N$. \end{lemma} \begin{proof} In view of Proposition \ref{classicls}, it would suffice to know that $$ \sum_{q \leq N^{1/2}} \mu^2(q) \prod_{p | q} \frac{w(p)}{p - w(p)} \gg \log^k N$$ for all large $N$. If we define a multiplicative function $g(n)$, supported on squarefree integers, by $g(p) := w(p)/p$, then it would obviously suffice to know that $\sum_{q \leq N^{1/2}} g(q) \gg \log^k N$ for all large $N$. However there is an extensive theory, dating back to Hal\'{a}sz, Wirsing and others, that gives asymptotics and bounds for sums of multiplicative functions. For example, partial summation and the assumption that $\sum_{Z \leq p \leq 2Z} w(p) \log p = k Z + O(\frac{Z}{\log^2 Z})$ show that $g$ satisfies the conditions of Theorem A.5 of Friedlander and Iwaniec~\cite{opera-cribo}, and therefore $$ \sum_{q \leq N} g(q) \sim c_{g} \log^k N \;\;\; \text{as} \; N \rightarrow \infty , $$ for a certain constant $c_{g} > 0$. \end{proof} \textsc{The larger sieve}. The ``larger sieve'' was introduced by Gallagher \cite{gallagher}. A pleasant discussion of it may be found in chapter 9.7 of Friedlander and Iwaniec \cite{opera-cribo}. We will apply it several times in the paper, and we formulate a version suitable for those applications. \begin{theorem}\label{larger-sieve} Let $0 < \delta \leq 1$, and let $Q > 1$. Let $\mathscr{P}$ be a set of primes. For each prime $p \in \mathscr{P}$, suppose that one is given a set $S_p \subset \Z/p\Z$, and write $\sigma_p := |S_p|/p$. Suppose that there is some set $\A'_p \subset \A$, $|\A'_p[X]| \geq \delta |\A[X]|$, such that $\A'_p \md{p} \subset S_p$. Then \[ |\A[X]| \ll \frac{Q}{\delta^2 \sum_{p \in \mathscr{P}, p \leq Q} \frac{\log p}{\sigma_p p} - \log X},\] provided that the denominator is positive. \end{theorem} \emph{Remark.} In this paper we will always have $\delta$ at least some absolute constant, not depending on $X$, and very often we will have $\delta \approx 1$. \begin{proof} We examine the expression \[ I := \sum_{\substack{p \in \mathscr{P} \\ p \leq Q}} \sum_{\substack{x,y \in \A[X] \\ x \neq y}} 1_{p | x - y} \log p.\] On the one hand we have \[ \sum_{p \in \mathscr{P}} 1_{p | n} \log p \leq \sum_{p} 1_{p | n} \log p \leq \log n,\] and therefore \[ I \leq |\A[X]|^2 \log X.\] On the other hand, writing $\A(a,p;X)$ for the number of $x \in \A[X]$ with $x \equiv a \md{p}$, we have \[ I = \sum_{\substack{p \in \mathscr{P}\\ p \leq Q}} \sum_{a \mdsub{p}} |\A(a,p;X)|^2 \log p - |\A[X]| \sum_{\substack{p \in \mathscr{P}\\ p \leq Q}} \log p.\] Comparing these facts yields \begin{equation}\label{to-use} \sum_{\substack{p \in \mathscr{P} \\ p \leq Q}} \sum_{a \mdsub{p}} |\A(a,p;X)|^2 \log p \leq |\A[X]|^2 \log X + O(|\A[X]|Q),\end{equation} and so of course, since $\A'_p \subset \A$, \[ \sum_{\substack{p \in \mathscr{P} \\ p \leq Q}} \sum_{a \mdsub{p}} |\A'_p(a,p;X)|^2 \log p \leq |\A[X]|^2 \log X + O(|\A[X]|Q).\] However by the Cauchy--Schwarz inequality and the fact that $\A'_p \md{p} \subset S_p$ we have \[ \sum_{a \mdsub{p}} |\A'_p(a,p;X)|^2 \geq \frac{|\A'_p[X]|^2}{\sigma_p p} \geq \delta^2 \frac{|\A[X]|^2}{\sigma_p p}.\] Summing over $p$ and rearranging, we obtain the result. \end{proof} The larger sieve bound can be a little hard to get a feel for, so we give an example. Suppose that $\mathscr{P}$ consists of all primes and that $\sigma_p = \alpha$ for all $p$. Take $\delta = 1$. Then, since $\sum_{p \leq Q} \frac{\log p}{p} = \log Q + O(1)$, the larger sieve bound is essentially \[ |\A[X]| \ll \frac{Q}{\frac{1}{\alpha} \log Q - \log X}.\] Taking $Q$ a little larger than $X^{\alpha}$, we obtain the bound $|\A[X]| \ll X^{\alpha + o(1)}$. This beats an application of the large sieve when $\alpha < \frac{1}{2}$, that is to say when we are sieving out a majority of residue classes (hence the terminology ``larger sieve''). However in the type of problems we are generally considering in this paper, where $\alpha = \frac{1}{2}$, we only recover the bound $|\A[X]| \ll X^{1/2 + o(1)}$, as of course we must since nothing has been done to exclude the example where $\A$ is a set of values of a quadratic. In actual fact one of our three applications of the larger sieve (in the proof of Theorem \ref{thm1.4}) requires an inspection of the above proof, rather than an application of the result itself. This is the observation that when $\sigma_p \approx \frac{1}{2}$ the larger sieve \emph{does} beat the bound $|\A[X]| \ll X^{1/2 + o(1)}$ unless $\A$ satisfies a certain ``uniform fibres'' condition. Recall that if $\A$ is a set of integers then $\A(a,p;X)$ is the number of $x \in \A[X]$ with $x \equiv a \md{p}$. \begin{lemma}\label{lem1.1} Let $\kappa > 0$ and $\eta > 0$ be small parameters. Suppose that $\A$ is a set of integers occupying at most $(p+1)/2$ residue classes modulo $p$ for all $p$. Say that $\A[X]$ has {\em $\eta$-uniform fibres above $p$} if \[ \sum_{a \mdsublem{p}} |\A(a, p;X)|^2 \leq (2 + \eta) |\A[X]|^2/p.\] Let $\mathscr{P}_{\unif}$ be the set of primes above which $\A[X]$ has $\eta$-uniform fibres. Then either $|\A[X]| \leq X^{1/2 - \kappa}$ or else ``most'' fibres are $\eta$-uniform in the sense that \[ \sum_{\substack{p \leq X^{1/2}, \\ p \notin \mathscr{P}_{\unif}}} \frac{\log p}{p} \leq \frac{3\kappa \log X + O(1)}{\eta}.\] \end{lemma} \begin{proof} Let $\mathscr{P}$ be the set of all primes, and let $Q := X^{1/2 - \kappa}$. We proceed as in the proof of the larger sieve until \eqref{to-use}, which was the inequality \[ \sum_{p \leq Q} \sum_{a \mdsub{p}} |\A(a,p;X)|^2 \log p \leq |\A[X]|^2 \log X + O(|\A[X]|Q).\] Now by the Cauchy--Schwarz inequality we have \[ \sum_{a \mdsub{p}} |\A(a,p;X)|^2 \geq 2|\A[X]|^2/(p+1)\] for all $p$. Using this and the estimate $\sum_{p \leq Q} \log p/p = \log Q + O(1)$, we see that the left-hand side of \eqref{to-use} is at least \[ 2|\A[X]|^2 (\log Q + O(1)) + \eta |\A[X]|^2 \sum_{\substack{p \leq Q \\ p \notin \mathscr{P}_{\unif}}} \frac{\log p}{p}.\] Therefore \[ \eta \sum_{\substack{p \leq Q \\ p \notin \mathscr{P}_{\unif}}} \frac{\log p}{p} \leq \log X - 2 \log Q + O(1) + O(\frac{Q}{|\A[X]|}).\] If $|\A[X]| \leq X^{1/2 - \kappa}$ then we are done; otherwise, the term $O(Q/|\A[X]|)$ may be absorbed into the $O(1)$ term and, after a little rearrangement, we obtain \[ \sum_{\substack{p \leq Q \\ p \notin \mathscr{P}_{\unif}}} \frac{\log p}{p} \leq \frac{2 \kappa \log X + O(1)}{\eta}.\] Since \[ \sum_{Q \leq p \leq X^{1/2}} \frac{\log p}{p} = \kappa \log X + O(1),\] the claimed bound follows. \end{proof} \section{Sieving by additively structured sets} Our aim in this section is to establish Theorem \ref{thm1.4}. Let $A$ be a finite set of integers. As is standard, we write $E(A,A)$ for the {\em additive energy} of $A$, that is to say the number of quadruples $(a_1,a_2,a_3,a_4) \in A^4$ with $a_1 + a_2 = a_3 + a_4$. If $p$ is a prime, write $E_p(A, A)$ for the number of quadruples with $a_1 + a_2 \equiv a_3 + a_4 \md{p}$. It is easy to see that $E_p(A,A) \geq |A|^4/p$. In situations where this inequality is not tight, we can get a lower bound for the additive energy $E(A,A)$. To do this we will use the {\em analytic large sieve inequality}, which is something like an approximate version of Bessel's inequality (and which leads, in a non-obvious way, to the large sieve bound that we stated as Proposition \ref{classicls}). We cite the following version, which is best possible in various aspects, from Chapter 9.1 of Friedlander and Iwaniec \cite{opera-cribo}. \begin{proposition} Let $0 < \delta \leq 1/2$, and suppose that $\theta_{1}, \dots , \theta_{R} \in \R/\Z$ form a $\delta$-spaced set of points, in the sense that $\Vert \theta_{r} - \theta_{s} \Vert \geq \delta$ for all $ r \neq s$ where $\Vert \cdot \Vert$ denotes distance to the nearest integer. Suppose that $(a(x))_{M < x \leq M+X}$ are any complex numbers, where $X$ is a positive integer. Then \[ \sum_{r=1}^{R} \bigg|\sum_{M < x \leq M+X} a(x) e(\theta_{r}x) \bigg|^{2} \leq (X - 1 + \delta^{-1}) \sum_{M < x \leq M+X} |a(x)|^{2}, \] where as usual $e(\theta) := \exp\{2\pi i \theta\}$. \end{proposition} Using the analytic large sieve inequality, we shall prove the following lemma. \begin{lemma}[Lifting additive energy]\label{lift-lem} Suppose that $A \subset [X]$. Then we have \[ \sum_{p \leq X^{1/2}} p \big( E_p(A,A) - \frac{|A|^4}{p} \big) \leq 3 X E(A,A).\] \end{lemma} \begin{proof} Write $r(x)$ for the number of representations of $x$ as $a_1 + a_2$ with $a_1,a_2 \in A$. Then \[ E(A,A) = \sum_{x \leq 2X} r(x)^2,\] whilst \[ E_p(A, A) = \sum_{\substack{x,x' \leq 2X \\ x \equiv x' \mdsub{p}}} r(x) r(x') = \frac{1}{p}\sum_{a \mdsub{p}} |\sum_{x \leq 2X} r(x) e(ax/p)|^2.\] It follows that \begin{align*} \sum_{p \leq X^{1/2}} p E_p(A,A) & = \sum_{p \leq X^{1/2}} \sum_{a \mdsub{p}} |\sum_{x \leq 2X} r(x) e(ax/p)|^2 \\ & = \sum_{p \leq X^{1/2}} \sum_{\substack{a \mdsub{p} \\ a \neq 0}} |\sum_{x \leq 2X} r(x) e(ax/p)|^2 + \sum_{p \leq X^{1/2}} p \frac{|A|^4}{p}.\end{align*} Now the fractions $a/p$ are $1/X$-spaced, as $a, p$ range over all pairs with $p \leq X^{1/2}$ prime and $1 \leq a \leq p-1$. By the analytic form of the large sieve it follows that \[ \sum_{p \leq X^{1/2}} \sum_{a \mdsub{p} a \neq 0} |\sum_{x \leq 2X} r(x) e(ax/p)|^2 \leq 3 X \sum_{x \leq 2X} r(x)^2.\] Putting all these facts together gives the result.\end{proof} \begin{corollary}\label{cor1.3} Let $\eta,\delta > 0$ be small real numbers with $\eta \leq \delta^2$. Suppose that $A \subset [X]$ is a set. Let $\mathscr{P}$ be a set of primes satisfying $36\delta^{-2} \leq p \leq X^{1/2}$, and suppose that the following are true whenever $p \in \mathscr{P}$: \begin{enumerate} \item $A \mdlem{p}$ lies in a set $S_p$ of cardinality at most $\frac{1}{2}(p+1)$; \item $S_p$ has at least $(\frac{1}{16} + \delta) p^3$ additive quadruples; \item $A$ has $\eta$-uniform fibres mod $p$, in the sense that $\sum_{a \mdsublem{p}} |A(a; p)|^2 \leq (2 + \eta) |A|^2/p$, where $A(a;p)$ is the number of $x \in A$ with $x \equiv a \mdlem{p}$. \end{enumerate} Then $E(A,A) \geq \frac{\delta |A|^4}{3X}|\mathscr{P}|$. \end{corollary} \begin{proof} Suppose that $p \in \mathscr{P}$. We will obtain a lower bound for $E_p(A,A)$ which beats the trivial bound of $E_p(A,A) \geq |A|^4/p$. The corollary will then follow quickly from Lemma \ref{lift-lem}. First of all we apply the variance identity \[ \sum_{m = 1}^M (t_m - \frac{1}{M}\sum_{i = 1}^M t_i)^2 = \sum_{i=1}^M t_i^2 - \frac{1}{M} (\sum_{i=1}^M t_i)^2\] with $M := |S_p|$ and the $t_i$ being the $A(a;p)$ with $a \in S_p$. This and the uniform fibres assumption yields \[ \sum_{a \in S_p} \big(|A(a;p)| - \frac{|A|}{|S_p|}\big)^2 \leq \frac{(2 + \eta) |A|^2}{p} - \frac{2|A|^2}{p+1} \leq \frac{|A|^2}{p}\big(\eta + \frac{2}{p}\big).\] Write $f : \Z/p\Z \rightarrow \R$ for the function $f(a) := |A(a;p)|$, and $g : \Z/p\Z \rightarrow \R$ for the function which is $|A|/|S_p|$ on $S_p$ and zero elsewhere. We have shown\footnote{The normalisations here are the ones standard in additive combinatorics. Write $\Vert F \Vert_2 = (\frac{1}{p}\sum_{x \in \Z/p\Z} |F(x)|^2)^{1/2}$, but $\Vert \hat{F} \Vert_2 = (\sum_r |\hat{F}(r)|^2)^{1/2}$ on the Fourier side. Then we have the Parseval identity $\Vert F \Vert_2 = \Vert \hat{F} \Vert_2$ and Young's inequality $\Vert \hat{F} \Vert_{\infty} \leq \Vert F \Vert_1$, both of which are used here.} that \[ \Vert f - g \Vert_2 \leq \frac{|A|}{p} \sqrt{\eta + \frac{2}{p}}.\] Note also that, by the Cauchy--Schwarz inequality, \[ \Vert f - g \Vert_1 \leq \Vert f - g \Vert_2 \leq \frac{|A|}{p} \sqrt{\eta + \frac{2}{p}} ,\] and hence \[ \Vert \hat{f} - \hat{g} \Vert_4^4 \leq \Vert \hat{f} - \hat{g} \Vert_{\infty}^{2} \Vert \hat{f} - \hat{g} \Vert_2^{2} \leq \Vert f - g \Vert_1^2 \Vert f - g \Vert_2^2 \leq \frac{|A|^4}{p^4}(\eta + \frac{2}{p})^{2},\] which of course implies that \[ \Vert \hat{f} - \hat{g} \Vert_4 \leq \frac{|A|}{p} (\eta^{1/2} + \sqrt{\frac{2}{p}}),\] where $\hat{f}(r) := \frac{1}{p}\sum_{a \in \Z/p\Z} f(a)e(-ar/p)$, similarly for $\hat{g}$. It follows that \[ \Vert \hat{f} \Vert_4 \geq \Vert \hat{g} \Vert_4 - \frac{\eta^{1/2}|A|}{p} - \frac{\sqrt{2}|A|}{p^{3/2}}.\] Note, however, that \[ E_p(A,A) = \sum_{a_1 + a_2 = a_3 + a_4} f(a_1)f(a_2) f(a_3) f(a_4) = p^3 \Vert \hat{f} \Vert_4^4,\] whilst \[ \Vert \hat{g} \Vert_4^4 = \frac{|A|^4}{|S_p|^4} \frac{E_p(S_p, S_p)}{p^{3}} \geq \frac{|A|^4}{16|S_p|^4} (1 + 16\delta)\] and so \[ \Vert \hat{g} \Vert_4 \geq \frac{|A|}{2|S_p|}(1 + 2\delta) \geq \frac{|A|}{p+1}(1 + 2\delta).\] Putting these facts together, and remembering that $\eta \leq \delta^{2}$ and $p \geq 36\delta^{-2}$, yields \[ \Vert \hat{f} \Vert_4 \geq \frac{|A|}{p}(1 + 2\delta - \eta^{1/2} - 3/p^{1/2}) \geq \frac{|A|}{p}(1 + \delta/2)\] and so $E_p(A,A) \geq \frac{|A|^4}{p}(1 + \delta)$ whenever $p \in \mathscr{P}$. The result now follows immediately from Lemma \ref{lift-lem}. \end{proof} \begin{corollary}\label{cor3.3} Let $\kappa > 0$ be a small parameter. Suppose that $A \subset [X]$ and that, for every prime $p \leq X^{1/2}$, the set $A \mdlem{p}$ lies in a set $S_p$ of cardinality at most $\frac{1}{2}(p+1)$ and with at least $(\frac{1}{16} + \delta) p^3$ additive quadruples. Then either $|A| \ll X^{1/2 - \kappa}$ or else $E(A,A) \geq \delta X^{-9 \kappa/\delta^2} |A|^3$. \end{corollary} \begin{proof} Suppose that $|A| \geq X^{1/2 - \kappa}$ and that $\kappa \log X$ is large enough. (If $\kappa \log X$ is small then $|A| \ll X^{1/2} \ll X^{1/2 - \kappa}$ by the usual large sieve bound.) Set $\eta := \delta^2$. By Lemma \ref{lem1.1} we either have $|A| \leq X^{1/2 - \kappa}$ or else $A$ has $\eta$-uniform fibres on a set $\mathscr{P} \subset [X^{1/2}]$ of primes satisfying \[ \sum_{\substack{p \leq X^{1/2} \\ p \notin \mathscr{P}}} \frac{\log p}{p} \leq \frac{4\kappa \log X}{\eta}.\] This implies that $|\mathscr{P}| \geq X^{1/2 - 8\kappa/\eta}$, and so by Corollary \ref{cor1.3} and the fact that $|A| \geq X^{1/2 - \kappa}$ we have $E(A,A) \geq \delta X^{-9\kappa/\eta} |A|^3$, which gives the claimed bound. \end{proof} The main task for the rest of this section will be to prove the following. \begin{proposition}[Differenced larger sieve]\label{diff-largersieve} Let $X$ be large, and let $A \subset [X]$ be a set with the property that $A \mdlem{p}$ lies in a set $S_p$ of size at most $\frac{1}{2}(p+1)$ for all primes $p \leq X^{1/2}$. Suppose that $E(A,A) \geq |A|^3/K$. Then $|A| \leq K X^{1/2 - c_0}$, where $c_0 > 0$ is an absolute constant. \end{proposition} Let us pause to see how this and Corollary \ref{cor3.3} combine to establish Theorem \ref{thm1.4}. \begin{proof}[Proof of Theorem \ref{thm1.4} given Proposition \ref{diff-largersieve}] Let $\kappa > 0$ be a parameter to be specified shortly. Suppose that $A \subset [X]$, and that $A \mdlem{p} \subset S_p$ for all primes $p$. Suppose furthermore that $|S_p| = \frac{1}{2}(p+1)$ and that $S_p$ has at least $(\frac{1}{16} + \delta) p^3$ additive quadruples for all $p$. By Corollary \ref{cor3.3} we see that either $|A| \ll X^{1/2 - \kappa}$, or else $E(A,A) \geq \delta X^{-9\kappa/\delta^2} |A|^3$. In this second case it follows from Proposition \ref{diff-largersieve} that $|A| \ll_{\delta} X^{\frac{1}{2} + 9 \kappa/\delta^2 - c_0}$. Choosing $\kappa := c_0\delta^2/10$ gives the result. \end{proof} It remains to prove Proposition \ref{diff-largersieve}. As the reader will soon see, the proof might be thought of as a ``differenced larger sieve'' argument, in which the larger sieve is not applied to $A$ directly, but rather to intersections of shifted copies of $A$ (as in Lemma \ref{lem2.7}) and to a set $H$ of pairwise differences of elements of $A$ (as in Lemma \ref{lem2.8}). The assumption that $A$ has large additive energy allows one to recover bounds on $A$ from that information (as in Lemma \ref{lem3.5}). \emph{Remark.} It is possible to prove Proposition \ref{diff-largersieve} with a quite respectable value of the constant $c_{0}$. Unfortunately the quality of the final bound in Theorem \ref{thm1.4} is not really determined by the value of $c_{0}$, but by the much poorer bounds that we achieved when trying to force the set $A$ to have uniform fibres mod $p$. We believe that by reworking Corollary \ref{cor1.3} a little one could prove Theorem \ref{thm1.4} with an improved bound $|\A[X]| \ll X^{1/2 - c\sqrt{\delta}}$, but this is presumably very far from optimal. \begin{proof}[Proof of Proposition \ref{diff-largersieve}] The argument is a little involved, so we begin with a sketch. Suppose that $E(A,A) \approx |A|^3$. Then it is not hard to show that $|A \cap (A + h)| \approx |A|$ for $h \in H$, where $|H| \approx |A|$. Modulo $p$, the set $A \cap (A + h)$ is contained in $S_p \cap (S_p + h)$. If, for some $h \in H$, we have $|S_p \cap (S_p + h)| < (\frac{1}{2} - c)p$ then an application of the larger sieve implies that $|A \cap (A + h)| < X^{1/2 - c'}$, and hence $|A| \lessapprox X^{1/2 - c'}$. The alternative is that $|S_p \cap (S_p + h)| \approx \frac{1}{2}p$ for many $p$, for all $h \in H$. Using this we can show that there is \emph{some} $p$ for which $|S_p \cap (S_p + h)| \approx \frac{1}{2}p$ for many $h$. By a result of Pollard, there is no such set $S_p$. Let us turn now to the details, formulating a number of lemmas which correspond to the above heuristic discussion. From now on, the assumptions are as in Proposition \ref{diff-largersieve}. \begin{lemma}\label{lem3.5} Then there is a set $H \subset [-X, X]$, $|H| \geq |A|/2K$ such that $|A \cap (A + h)| \geq |A|/2K$ for all $h \in H$. \end{lemma} \begin{proof} This is completely standard additive combinatorics and is a consequence, for example, of the inequalities in \cite[\S 2.6]{tao-vu}. It is no trouble to give a self-contained proof: note that $E(A,A) = \sum_x |A \cap (A + x)|^2$ and that we have the trivial bound $|A \cap (A + x)| \leq |A|$ for all $x$. If $H$ is the maximal set with the stated property then \[ E(A, A) = \sum_{x \in H} |A \cap (A + x)|^2 + \sum_{x \notin H} |A \cap (A + x)|^2 \leq |H| |A|^2 + \frac{|A|}{2K}\sum_{x} |A \cap (A + x)| = |H| |A|^2 + \frac{|A|^3}{2K},\] from which the statement follows immediately. \end{proof} \begin{lemma}\label{lem2.7} Let $c > 0$ be a small constant. Set $Q := X^{1/2 - c/2}$, and suppose that there is some $h \in H$ such that \[ \sum_{p \leq Q} \frac{\log p}{p} \frac{|S_p \cap (S_p + h)|}{p} < (\frac{1}{2} - c) \log Q.\] Then $|A| \ll K X^{1/2 - c/4}$. \end{lemma} \begin{proof} Note that $A \cap (A + h) \subseteq S_p \cap (S_p + h)$ for all $p$. Write $\sigma_p := |S_p \cap (S_p + h)|/p$ and apply the larger sieve, Theorem \ref{larger-sieve}, with $\delta = 1$ and $A$ replaced by $A \cap (A + h)$. We obtain the bound \begin{equation}\label{eq77denom} |A \cap (A + h)| \ll \frac{Q}{\sum_{p \leq Q} \frac{\log p}{p\sigma_p} - \log X},\end{equation} provided that the denominator is positive. Our assumption is that \[ \sum_{p \leq Q} \frac{\sigma_p \log p}{p} \leq (\frac{1}{2} - c) \log Q.\] Since $4t + 1/t \geq 4$ for all $t > 0$, it follows that \[ \sum_{p \leq Q} \frac{\log p}{\sigma_p p} \geq 4\log Q + O(1) - 4(\frac{1}{2} - c)\log Q = (2 + 4c)\log Q + O(1).\] It is easy to check that the denominator of \eqref{eq77denom} is indeed positive, since $Q = X^{1/2 - c/2}$. We obtain the bound \[ |A \cap (A + h)| \ll X^{1/2 - c/2 + o(1)} \ll X^{1/2 - c/4}.\] Since $|A \cap (A + h)| \geq |A|/2K$, the lemma follows. \end{proof} Before stating the next lemma, let us isolate a fact which will be needed in the proof. This is basically due to Pollard. \begin{lemma}[Pollard]\label{add-comb} Let $\eps > 0$ be small, and let $S \subset \Z/p\Z$ be a non-empty set such that $|S| < (1-2\eps)p$. Then there are at most $4\eps |S| + 1$ values of $h \in \Z/p\Z$ such that $|S \cap (S + h)| \geq (1 - \eps)|S|$. \end{lemma} \begin{proof} This follows quickly from a well-known result of Pollard \cite{pollard}. Writing $N_i$ for the number of $h$ such that $|S \cap (S + h)| \geq i$, Pollard's result in our setting implies that $N_1 + \dots + N_r \geq r(2|S|-r)$ for all $2|S| - p \leq r \leq |S|$. Temporarily write $H$ for the set of all $h \in \Z/p\Z$ such that $|S \cap (S + h)| \geq (1 - \eps)|S|$, and also let $R := |S| - 2\lfloor \eps|S| \rfloor$ and $U := |S| - \lfloor \eps|S| \rfloor$, where $\lfloor \cdot \rfloor$ denotes the floor function. Then \[ N_{R+1} + \dots + N_{|S|} \geq N_{R+1} + \dots + N_{U} \geq |H|(U - R) = |H| \lfloor \eps|S| \rfloor .\] Pollard's result tells us that \[ N_1 + \dots + N_{R} \geq R(2|S|-R) = |S|^{2} - 4\lfloor \eps|S| \rfloor^{2} .\] On the other hand we trivially have \[ N_1 + \dots + N_{|S|} = |S|^2.\] Combining all these facts leads to the result provided that $|S| \geq 1/\epsilon$. Alternatively, if $|S| < 1/\epsilon$ then $|S \cap (S + h)| \geq (1 - \eps)|S|$ only if $S \cap (S + h) = S$, in which case $S \cap (S + nh) = S$ for every $n$. Since $S$ is a proper subset of $\Z/p\Z$, this can only happen when $h=0$. \end{proof} We will also require a simple and standard averaging principle, the proof of which we include here for completeness. \begin{lemma}\label{simple-avg} Let $\eps, \eps'$ be real numbers with $0 < \eps \leq \eps'$. Let $X$ be a finite set, let $(\lambda(x))_{x \in X}$ be nonnegative weights, and suppose that $f : X \rightarrow [0,1]$ is a function such that $\sum_{x \in X} \lambda(x) f(x) \geq (1 - \eps) \sum_{x \in X} \lambda(x)$. Let $X' \subset X$ be the set of all $x \in X$ such that $f(x) \geq 1 - \eps'$. Then $\sum_{x \in X'} \lambda(x) \geq (1 - \frac{\eps}{\eps'}) \sum_{x \in X} \lambda(x)$. In particular if $\sum_{x \in X} f(x) \geq (1 - \eps) |X|$ then there are at least $(1 - \frac{\eps}{\eps'}) |X|$ values of $x$ such that $f(x) \geq 1 - \eps'$. \end{lemma} \begin{proof} We have \[ (1 - \eps) \sum_{x \in X} \lambda(x) \leq \sum_{x \in X} \lambda(x) f(x) = \sum_{x \in X'} \lambda(x) f(x) + \sum_{x \in X \setminus X'} \lambda(x) f(x) \leq \sum_{x \in X'} \lambda(x) + (1 - \eps') \sum_{x \in X \setminus X'} \lambda(x).\] Rearranging this inequality gives the first result. The second one follows by taking all the weights $\lambda(x)$ to be 1. \end{proof} \begin{lemma}\label{lem2.8} Let $c > 0$ be a sufficiently small absolute constant. Let $H \subset [-X,X]$ be a set of size $X^{1/2 - c/4}$, and let $Q = X^{1/2 - c/2}$. Then there is some $h \in H$ such that \[ \sum_{p \leq Q} \frac{\log p}{p} \frac{|S_p \cap (S_p + h)|}{p} < (\frac{1}{2} - c) \log Q.\] \end{lemma} \begin{proof} Suppose not. Then certainly \[ \sum_{h \in H}\sum_{p \leq Q} \frac{\log p}{p} \frac{|S_p \cap (S_p + h)|}{p} \geq (\frac{1}{2} - c) |H| \log Q \geq (\frac{1}{2} - c) |H| \sum_{p \leq Q} \frac{\log p}{p}.\] Write $\mathscr{P}$ for the set of primes $p \leq Q$ such that \begin{equation}\label{averagehyp} \sum_{h \in H} \frac{|S_p \cap (S_p + h)|}{p} \geq (\frac{1}{2} - c^{1/2}) |H| . \end{equation} By Lemma \ref{simple-avg} applied with $X$ the set of primes $p \leq Q$, $\lambda(p) = \frac{\log p}{2p}|H|$, $f(p) = \frac{2}{|H|} \frac{1}{p}\sum_{h \in H} |S_p \cap (S_p + h)|$, $\epsilon = 2c$ and $\epsilon' = 2c^{1/2}$ we have\footnote{The alert reader will observe that our applications of Lemma \ref{simple-avg} are slightly bogus, since we have $f(p) \leq (p+1)/p$ rather than $f(p) \leq 1$, as required in the lemma. This can be corrected by instead setting $\lambda(p) = \frac{\log p}{2p} \frac{p+1}{p} |H|$ and $f(p) = \frac{2}{|H|} \frac{1}{p+1}\sum_{h \in H} |S_p \cap (S_p + h)|$, which makes no essential difference to the conclusions about $\mathscr{P}$ and $H_p$.} \[ \sum_{p \in \mathscr{P}} \frac{\log p}{p} \geq (1 - c^{1/2} ) \sum_{p \leq Q} \frac{\log p}{p}.\] Note that $|S_p \cap (S_p + h)| \leq \frac{1}{2}(p+1)$ always. It also follows from Lemma \ref{simple-avg} applied to the inequality (\ref{averagehyp}) that, for all $p \in \mathscr{P}$, there is a set $H'_p \subset H$ with $|H'_p| \geq (1 - c^{1/4}) |H|$ such that $|S_p \cap (S_p + h)| \geq (\frac{1}{2} - c^{1/4})p$ for all $h \in H'_p$. On the other hand, by Lemma \ref{add-comb} it follows that $|H'_p \md{p}| \leq 4 c^{1/4} p + 1$, and so all but $c^{1/4} |H|$ elements of $H$ reduce to lie in a set of size $4 c^{1/4} p + 1 < \frac{1}{3}p$ modulo $p$, for all $p \in \mathscr{P}$, a set which satisfies $\sum_{p \in \mathscr{P}} \frac{\log p}{p} \geq (1 - c^{1/2}) (\log Q + O(1))$. We may apply the larger sieve, Theorem \ref{larger-sieve}, to this situation, taking $\delta = 1 - c^{1/4}$ and $\sigma_p = 1/3$ for all $p \in \mathscr{P}$. This gives the bound \[ |H| \ll \frac{Q}{3(1 - c^{1/4})^{2} (1 - c^{1/2})(\log Q + O(1)) -\log X }\] provided that the denominator is positive. If $c$ is sufficiently small then the denominator will be positive with our choice of $Q$, namely $X^{1/2 - c/2}$, and we get the bound $|H| \ll X^{1/2 - c/2 + o(1)}$. This is contrary to assumption. \end{proof} We may now conclude the proof of Proposition \ref{diff-largersieve}. As in the hypothesis of the proposition, let $A \subset [X]$ be a set such that $A \md{p} \subset S_p$ for all $p$, where $|S_p| \leq \frac{1}{2}(p+1)$. Suppose additionally that $E(A,A) \geq |A|^3/K$. By Lemma \ref{lem3.5} there is a set $H \subset [-X,X]$, $|H| \geq |A|/2K$, such that $|A \cap (A + h)| \geq |A|/2K$ for all $h \in H$. If $|H| < X^{1/2 - c/4}$ then the proposition follows, so suppose this is not the case. Then Lemma \ref{lem2.8} applies and we may conclude that there is an $h \in H$ such that \[ \sum_{p \leq Q} \frac{\log p}{p} \frac{|S_p \cap (S_p + h)|}{p} < (\frac{1}{2} - c) \log Q,\] where $Q = X^{1/2 - c/2}$. Finally, by Lemma \ref{lem2.7}, it follows that $|A| \ll K X^{1/2 - c/4}$, thereby concluding the proof of the proposition. \end{proof} \section{Sieving by intervals}\label{interval-sieve-sec} Our aim in this section is to establish Theorem \ref{interval-sieve}. We begin by recalling the statement of it. \begin{interval-sieve-repeat} Suppose that $\A$ is a set of integers and that, for each prime $p$, the set $\A \mdlem{p}$ lies in some interval $I_p$. Let $\eps > 0$ be arbitrary. Then \begin{enumerate} \item If $|I_p| \leq (1 - \eps) p$ for at least a proportion $\eps$ of the primes in each dyadic interval $[Z,2Z]$ then $|\A[X]| \ll_{\eps} (\log \log X)^{C\log(1/\eps)}$, where $C > 0$ is some absolute constant; \item If $|I_p| \leq \frac{p}{2}$ for all primes then $|\A[X]| \ll (\log \log X)^{\gamma + o(1)}$, where $\gamma = \frac{\log 18}{\log(3/2)} \approx 7.129$; \item If $|I_p| = [\alpha p, \beta p]$ for all primes $p$ and for fixed $0 \leq \alpha < \beta < 1$ \textup{(}not depending on $p$\textup{)} then $|\A| = O_{\beta - \alpha}(1)$; \item There are examples of infinite sets $\A$ with $|I_p| \leq (\frac{1}{2} + \eps) p$ for all $p$. \end{enumerate} \end{interval-sieve-repeat} The proof of parts (i) and (ii) relies on the following basic lemma. \begin{lemma}\label{large-fourier} Suppose that $p$ is a prime, that $I_p \subset \Z/p\Z$ is an interval of length at most $(1 - \eps) p$, and that $A \subset [X]$ is a set with $A \mdlem{p} \subset I_p$. Then there is some integer $k$, $1 \leq k \leq \lceil 2/\eps^2 \rceil$, such that \[ |\sum_{x \leq X} 1_A(x) e(k x/p) | \geq \eps |A|/32.\] If $|I_p| \leq p/2$ then we have the following more precise conclusion: there is an integer $k \in \{1,2\}$ such that \[ |\sum_{x \leq X} 1_A(x) e(k x/p)| \geq |A|/3.\] \end{lemma} \begin{proof} We claim that there is a $1$-periodic real-valued function \[ f(\theta) = 1 + \sum_{0 < |k| \leq \lceil 2/\eps^2 \rceil} c_k e(k\theta)\] such that $f(\theta) \leq 0$ when $|\theta| \geq \eps/2$. To construct $f(\theta)$, consider first the convolution $\psi(\theta) := 1_{|\theta| \leq \eps/4} \ast 1_{|\theta| \leq \eps/4} = \int_{\R/\Z} 1_{|\theta - \phi| \leq \eps/4} 1_{|\phi| \leq \eps/4} d\phi$. We have \[ |\hat{\psi}(k)| = |\widehat{1_{|\phi| \leq \eps/4}}(k)|^{2} = \big|\int_{\R/\Z} 1_{|\phi| \leq \eps/4} e(-k\phi) d\phi\big|^{2} \leq \min(\eps, \frac{1}{\pi |k|})^2.\] From the Fourier inversion formula it follows that \[ \frac{8}{\eps^2} \psi(\theta) = 2 + \sum_{k \neq 0} c_k e(k \theta),\] where $|c_k| \leq \min(8, \frac{1}{\eps^2 |k|^2})$. Furthermore, by construction, $\psi(\theta) = 0$ for $|\theta| \geq \eps/2$. Define \[ f(\theta) := 1 + \sum_{0 < |k| \leq K} c_k e(k \theta),\] where $K := \lceil 2/\eps^2 \rceil$. Since \[ \sum_{|k| > K} |c_k| \leq \sum_{|k| \geq K+1} \frac{1}{\eps^2 |k|^2} \leq \frac{2}{\eps^2 K} \leq 1,\] it follows that $f$ has the required properties. Now there is some $\beta \in [0,1]$ (depending on $I_p$) such that $\Vert \frac{x}{p} + \beta \Vert \geq \eps/2$ whenever $x \in A$, where $\Vert \cdot \Vert$ denotes distance to the nearest integer. This means that $f(\frac{x}{p} + \beta) \leq 0$, and so \[ 1 + \sum_{0 < |k| \leq \lceil 2/\eps^2 \rceil} c_k e( k (\frac{x}{p} + \beta)) \leq 0.\] It follows that \[ |A| \leq -\sum_{0 < |k| \leq \lceil 2/\eps^2 \rceil} c_k \sum_{x \leq X} 1_A(x) e( k (\frac{x}{p} + \beta)).\] Using the triangle inequality, one obtains \[ |A| \leq \sum_{0 < |k| \leq \lceil 2/\eps^2 \rceil} |c_k| | \sum_{x \leq X} 1_A(x) e(k x/p) |.\] To conclude the proof of the lemma, we observe that \[ \sum_{0 < |k| \leq K} |c_k| \leq \frac{32}{\eps},\] an estimate that follows upon splitting into the ranges $0 < |k| \leq 1/\eps$ and $|k| > 1/\eps$. For the second statement, simply note that the function $f(\theta) = 1 - 2\cos \theta + \cos 2\theta$ satisfies $f(\theta) \leq 0$ when $|\theta| \leq \pi/2$; rewriting the left-hand side as $2\cos\theta (\cos \theta - 1)$, this becomes clear. The rest of the argument proceeds as before.\end{proof} We turn now to the proof of Theorem \ref{interval-sieve} (i). The general scheme of the argument, and in particular the use of Vinogradov's estimate (Proposition \ref{vino-est} below) was suggested to us by Jean Bourgain. We are very grateful to him for allowing us to include it here. The heart of the matter is the proof of the following lemma, from which Theorem \ref{interval-sieve} (i) follows rather easily by an iteration argument (or equivalently induction on $X$). \begin{lemma}\label{it-step} Suppose that $A \subset [X]$ and that $A \md{p}$ lies in an interval $I_p$ of length at most $(1-\eps) p$, for at least $\eps$ of all primes in each dyadic interval. Suppose that $X > X_0(\eps)$. Then there is a subinterval of $[X]$ of length $\exp (\log^{7/10} X)$ containing at least $c \eps^5 |A|$ points of $A$, where $c > 0$ is a small absolute constant. \end{lemma} Indeed, before proving this lemma let us explain how it implies Theorem \ref{interval-sieve} (i). We set $X_{0} = X$ and $A_{0} = \A[X]$, and by repeated application of the lemma we construct numbers $X_{i}$ and sets $A_{i}$ such that $A_{i} \subset [X_{i}]$, $A_{i} \md{p}$ lies in an interval $I_p$ of length at most $(1-\eps) p$, for at least $\eps$ of all primes in each dyadic interval, $\log X_{i+1} = \log^{7/10} X_{i}$ and $|A_{i+1}| \geq c \eps^5 |A_{i}|$. This procedure terminates when we first have $X_{i+1} \leq X_{0}(\eps)$, which will happen after $\ll \log\log\log X$ iterations. Consequently we have $|\A[X]| \leq (c^{-1}\eps^{-5})^{O(\log\log\log X)} X_{0}(\eps) \ll_{\eps} (\log \log X)^{C\log(1/\eps)}$, as claimed\footnote{As we have written things, we need to have $X_{0}(\eps) = \exp(C\log^{100}(1/\epsilon))$ (say) in order for the parameter $Y$ in the proof of Lemma \ref{it-step} to be large enough. But we remark that by taking more care of the final iterations in the proof of Theorem \ref{interval-sieve} (i), one could obtain a bound $|\A[X]| \ll (\log \log X)^{C\log(1/\eps)}$ for all $X$, with an absolute implied constant (not depending on $\eps$).} in Theorem \ref{interval-sieve}. \begin{proof}[Proof of Lemma \ref{it-step}] Suppose that $p$ is a prime such that $A \md{p} \subset I_p$. By Lemma \ref{large-fourier}, there is some $k$, $1 \leq k \leq \lceil 2/\eps^2 \rceil$, such that \begin{equation}\label{eq997} |\sum_{a \in A} e(k a/p)| \geq \eps |A|/32.\end{equation} Let $Y$, $1 \ll Y \ll X$, be a parameter to be selected later (we will in fact take $Y = \exp(c\log^{7/10} X)$). We may choose a single $k$ so that the preceding estimate holds for $\gg \eps^2$ of the primes in $[Y, 2Y]$ for which we know that $A \md{p} \subset I_p$, that is for $\gg \eps^3$ of all the primes in $[Y, 2Y]$. Now we use the following fact: there is a weight function $w : [Y, 2Y] \rightarrow \R_{\geq 0}$ such that \begin{enumerate} \item $w(p) \geq 1$ for all primes $p \in [Y, 2Y]$; \item $\sum_{Y \leq n \leq 2Y} w(n) \leq 10 \pi(Y)$; \item $w(n) = \sum_{d| n: d \leq Y^{1/2}} \lambda_d$, where $\sum_{d \leq Y} \frac{|\lambda_d|}{d} \ll \log^{3} Y$. \end{enumerate} Such a function can be constructed in the form $w(n) = (\sum_{d | n} \mu(d) \psi(\frac{\log d}{\log Y}))^2$, where $\psi \in C^{\infty}(\R)$ is supported on $|x| \leq \frac{1}{4}$, is bounded in absolute value by 1, and $\psi(0) = 1$. Property (i) is then clear, whilst bound (ii) can be verified by expanding out and interchanging the order of summation. To check (iii), we note that it is clear that $|\lambda_d| \leq \sum_{[d_1, d_2] = d} 1 \leq \tau_3(d)$, the number of ways of writing $d$ as a product of three nonnegative integers. The claimed bound is then an easy exercise. It follows from \eqref{eq997} and the above properties that \begin{equation}\label{eq527} \sum_{Y \leq n \leq 2Y} w(n) |\sum_{a \in A} e(ka/n)|^2 \geq c \eps^5 \pi(Y) |A|^2.\end{equation} Expanding out and applying the triangle inequality yields \begin{equation}\label{eq37} \sum_{a, a' \in A} |\sum_{Y \leq n \leq 2Y} w(n) e(\frac{k(a - a')}{n}) | \geq c\eps^5 \pi(Y) |A|^2.\end{equation} We now claim that, if $Y$ is chosen judiciously, the contribution to this from those pairs $a,a'$ with $|a - a'| \geq Y^{10}$ (say) can be ignored. Indeed suppose, on the contrary, that \begin{equation}\label{eq843} |\sum_{Y \leq n \leq 2Y} w(n) e(\frac{x}{n})| \geq \frac{c}{100}\eps^5 \pi(Y),\end{equation} for some $x := k(a - a')$ satisfying $Y^{10} \leq x \ll \eps^{-2}X$. By property (iii) of $w(n)$ and the triangle inequality, this implies that \[ \sum_{d \leq Y^{1/2}} |\lambda_d| |\sum_{Y/d \leq n' \leq 2Y/d} e(\frac{x}{dn'})| \geq \frac{c}{100} \eps^5 \pi(Y).\] By the upper bound (iii) for $\sum |\lambda_d| / d$ it follows that there is some $d \leq Y^{1/2}$ such that \begin{equation}\label{eq478} |\sum_{Y/d \leq n' \leq 2Y/d} e(\frac{x}{d n'})| \gg \eps^5 \log^{-4} Y \frac{Y}{d}.\end{equation} At this point we invoke the following powerful estimate of Vinogradov. \begin{proposition}\label{vino-est} Let $\delta > 0$ be small and $Y$ be large, and suppose that $x \geq Y^5$. Suppose that $|\sum_{Y \leq n \leq 2Y} e(x/n)| \geq \delta Y$. Then $x \geq \exp( c\log^{3/2} Y/\log^{1/2}(1/\delta))$. \end{proposition} \begin{proof} Using e.g. Theorem 8.25 of Iwaniec and Kowalski \cite{iwaniec-kowalski}, one obtains that \[ |\sum_{Y \leq n \leq 2Y} e(x/n)| \ll Y e^{-c\frac{\log^{3}Y}{\log^{2}x}}. \] Thus we must have $1/\delta \gg \exp(c\log^{3}Y/\log^{2}x)$, from which the conclusion of the proposition quickly follows. \end{proof} Applying this Proposition to \eqref{eq478} leads to a contradiction unless \[ \frac{x}{d} > \exp \bigg(c \frac{\log^{3/2} Y}{(\log \log Y)^{1/2} + (\log(1/\epsilon))^{1/2}}\bigg), \] which would imply that $X \gg \epsilon^{2} x > \exp (c (\log^{3/2} Y)/((\log \log Y)^{1/2} + (\log(1/\epsilon))^{1/2}))$. With $Y = \exp(c\log^{7/10} X)$ and $X \geq \exp(C\log^{100}(1/\epsilon))$, say, this will not be so. It follows that we were wrong to assume \eqref{eq843}, and so indeed the contribution to \eqref{eq37} of those pairs $a,a'$ with $|a - a'| \geq Y^{10}$ may be ignored. Thus we have \begin{equation}\label{eq372} \sum_{\substack{a, a' \in A \\ |a - a'| \leq Y^{10}}} \bigg|\sum_{Y \leq n \leq 2Y} w(n) e(\frac{k(a - a')}{n}) \bigg| \geq \frac{c}{2}\eps^5 \pi(Y) |A|^2.\end{equation} Finally, we may apply the trivial bound to the inner sum, recalling from (ii) above that $\sum_{Y \leq n \leq 2Y} w(n) \leq 10 \pi(Y)$. We obtain \[ \sum_{\substack{a, a' \in A \\ |a - a'| \leq Y^{10}}} 1 \gg \eps^5 |A|^2,\] which implies that there is a subinterval of $[X]$ of length $Y^{10}$ containing $\gg \eps^5 |A|$ elements of $A$. This concludes the proof of Lemma \ref{it-step}, and hence of Theorem \ref{interval-sieve} (i). \end{proof} \emph{Proof of Theorem \ref{interval-sieve} (ii)} (sketch). We proceed as above, with the following changes. \begin{itemize} \item Use the second conclusion of Lemma \ref{large-fourier} to conclude that there is some $k \in \{1,2\}$ such that \[ \sum_{Y \leq p \leq 2Y} | \sum_{a \in A} e(ka/p)|^{2} \geq \frac{1}{18} (\pi(2Y) - \pi(Y)) |A|^{2}.\] This takes the place of \eqref{eq527}. \item Expand out as in \eqref{eq37} to get \[ \sum_{a, a' \in A} |\sum_{Y \leq p \leq 2Y} e(\frac{k(a - a')}{p})| \geq \frac{1}{18} (\pi(2Y) - \pi(Y)) |A|^2.\] \item Choose $Y = \exp(\log^{2/3+o(1)} X)$, and use Jutila \cite[Theorem 2]{jutila} (which is a Vinogradov-type estimate for $\sum_{p \leq P} e(x/p)$) to show that the contribution from those pairs with $|a - a'| \geq Y^{10}$ can be ignored, so we have \[ \sum_{\substack{a, a' \in A \\ |a - a'| \leq Y^{10}}} |\sum_{Y \leq p \leq 2Y} e(\frac{k(a - a')}{p})| \geq (\frac{1}{18} - o(1)) \pi(Y) |A|^2.\] \item Conclude that there is some interval of length $\sim Y^{10}$ containing at least $(\frac{1}{18} - o(1))|A|$ points of $A$, and proceed iteratively as before. \end{itemize} \emph{Remark.} We could have used Jutila's bound in the proof of Theorem \ref{interval-sieve} (i) as well, instead of using the weight function $w$ . We chose not to do this in the interests of self-containment and of variety. Note that Jutila's paper predates Vaughan's identity \cite{vaughan} for prime number sums, and his argument would be a little more accessible if this device were used. A model for such an argument may be found in the paper of Granville and Ramar\'e \cite{granville-ramare}. \vspace{11pt} \emph{Proof of Theorem \ref{interval-sieve}} (iii). This is essentially a consequence of Jutila \cite[Corollary, p126]{jutila}. A slight variant of that Corollary shows that the number of $p \in [x^{1/2}, 2x^{1/2}]$ with $\alpha \leq \{ x/p \} \leq \beta$ is $\sim (\beta - \alpha) (\pi(2x^{1/2}) - \pi(x^{1/2}))$, and so all elements of $\A$ are bounded by $O_{\beta - \alpha}(1)$.\vspace{11pt} \emph{Proof of Theorem \ref{interval-sieve}} (iv). We take $\A$ to consist of the numbers $a_i = \prod_{p \leq X_i} p$, for some extremely rapidly-growing sequence $X_1 < X_2 < \dots$. Given a prime $p$, suppose that $X_i \leq p \leq X_{i+1}$. Then $a_{i+1},a_{i+2},\dots$ all reduce to zero $\md{p}$, and so $\A \md{p} = \{0,a_1,\dots, a_i\}$. By choosing the $X_i$ sufficiently rapidly growing we may ensure that $0 < a_1 < \dots < a_{i-1} < \eps p$. Regardless of the value of $a_i \md{p}$ (which we cannot usefully control) the set $\A \md{p}$ will be contained in some interval of length at most $(\frac{1}{2} + \eps) p$.\vspace{11pt} \emph{Remark.} With $\A$ as constructed above, the shape of $|\A[X]|$ is $\log_* X$. Thus there is still a considerable gap between the bound of (i) and the construction given here. We expect, however, that the correct bound in (i) is of $\log_* $ type, which would follow assuming vaguely sensible conjectures on exponential sums $\sum_{n \leq Y} e(x/n)$. If, for example, the conclusion of Proposition \ref{vino-est} were instead that $x \geq \exp(Y^{1/10})$ then we would get a $\log_*$-type bound on $\A[X]$ in this case. \section{Sieving by arithmetic progressions} In this section we shall prove Theorem \ref{prog-thm}, whose statement was as follows. \begin{prog-thm-rpt} Let $\eps > 0$. Suppose that $\A$ is a set of integers and that, for each prime $p \leq X^{1/2}$, the set $\A \mdlem{p}$ lies in some arithmetic progression $S_p$ of length $(1-\eps)p$. Then $|\A[X]| \ll_{\eps} X^{1/2 - \eps'}$, where $\eps' > 0$ depends on $\eps$ only. \end{prog-thm-rpt} Suppose to begin with that $|S_{p}|=\frac{1}{2}(p+1)$ for each odd prime $p$. Then (dilating $S_p$ to the interval $\{1,\dots, \frac{1}{2}(p+1)\}$, which preserves the additive energy) $$ E_{p}(S_{p},S_{p}) = (\frac{p+1}{2})^{2} + 2\sum_{h=1}^{\frac{1}{2}(p-1)} (\frac{p+1}{2} - h)^{2} \geq \frac{p^{3}}{12} . $$ Thus Theorem \ref{thm1.4} is applicable with the choice $\delta = 1/48$, and the result follows in this case. Now we turn to the proof of Theorem \ref{prog-thm} for arbitrary $\eps > 0$. We begin with a result which should be compared to Corollary \ref{cor1.3}, but which is simpler to state and prove than that result. \begin{lemma}\label{progr-energy} Let $\eps > 0$ be small, and suppose that $A \subset [X]$ is a set and $\mathscr{P} \subset [X^{1/2}]$ is a set of primes such that $|\mathscr{P}| \geq 2/\eps^{2}$. If $S_{p} \subset \Z/p\Z$ is an arithmetic progression of length at most $(1-\eps)p$, and if $A \mdlem{p} \subset S_p$ for each prime $p \in \mathscr{P}$, then $E(A,A) \gg \frac{\eps^{4} |A|^{4}}{X} |\mathscr{P}|$, where the constant implicit in the $\gg$ notation is absolute. \end{lemma} \begin{proof} In view of Lemma \ref{lift-lem}, and our assumption that $|\mathscr{P}| \geq 2/\eps^{2}$, it will suffice to show that $$ E_{p}(A,A) - \frac{|A|^{4}}{p} \gg \frac{\eps^{4} |A|^{4}}{p} $$ for each prime $p \in \mathscr{P}$ that is greater than $2/\eps^{2}$ (that being a positive proportion of all the primes in $\mathscr{P}$). As in the proof of Corollary \ref{cor1.3} we have $$ E_{p}(A,A) = p^3 \sum_{r \in \Z/p\Z} \big|\frac{1}{p} \sum_{x \leq X} 1_A(x) e(-xr/p) \big|^{4} . $$ The contribution from the $r=0$ term is evidently equal to $|A|^{4}/p$. By Lemma \ref{large-fourier} (which was stated in the case $S_p$ is an interval, but may easily be adapted to the case $S_p$ a progression by dilation), if $p > \lceil 2/\eps^2 \rceil$ then there is some nonzero $r$ satisfying $|\sum_{x \leq X} 1_A(x) e(rx/p)| \geq \eps |A|/32$. The result follows immediately. \end{proof} The other major ingredient that we shall need is an analogue of Lemma \ref{lem2.8} that applies when the sets $S_{p}$ have size at most $(1-\eps)p$, rather than size at most $\frac{1}{2}(p+1)$ as in that lemma. The following result provides this. \begin{lemma}\label{progr-goodh} Let $c > 0$ be a sufficiently small absolute constant. Let $H \subset [-X,X]$ be a set of size $X^{1/2 - c/4}$, and let $Q = X^{1/2 - c/2}$. Suppose that for each prime $p \in \mathscr{P}'$ we have a subset $S_p \subset \Z/p\Z$ such that $\frac{1}{10}p \leq |S_p| \leq (1-\eps)p$, where $\mathscr{P}'$ is a subset of the primes $p \leq Q$ satisfying $\sum_{p \in \mathscr{P}'} \frac{\log p}{p} \geq \frac{1}{3}\log Q$. Then there is some $h \in H$ such that \[ \sum_{p \in \mathscr{P}'} \frac{\log p}{p} \frac{|S_p \cap (S_p + h)|}{p} < (1 - c\eps) \sum_{p \in \mathscr{P}'} \frac{\log p}{p} \frac{|S_p|}{p} .\] \end{lemma} \begin{proof} The proof is quite close to the proof of Lemma \ref{lem2.8}, so we shall give a fairly brief account. If the conclusion were false then we would have $$ \sum_{h \in H} \sum_{p \in \mathscr{P}'} \frac{\log p}{p} \frac{|S_p \cap (S_p + h)|}{p} \geq (1 - c\eps) |H| \sum_{p \in \mathscr{P}'} \frac{\log p}{p} \frac{|S_p|}{p}. $$ In this case, if we let $\mathscr{P}$ denote the subset of primes $p \in \mathscr{P}'$ for which $$ \sum_{h \in H} \frac{|S_p \cap (S_p + h)|}{p} \geq (1 - c^{1/2}\eps) |H| \frac{|S_p|}{p} , $$ then applying Lemma \ref{simple-avg} with $X = \mathscr{P}'$, $\lambda_{p} = \frac{\log p}{p^2} |H| |S_p|$ and $f(p) = |H|^{-1}\sum_{h \in H} \frac{|S_p \cap (S_p + h)|}{|S_p|}$ yields $$ \sum_{p \in \mathscr{P}} \frac{\log p}{p} \frac{|S_p|}{p} \geq (1 - c^{1/2}) \sum_{p \in \mathscr{P}'} \frac{\log p}{p} \frac{|S_p|}{p} \geq \frac{1}{40}\log Q. $$ As in the proof of Lemma \ref{lem2.8}, it also follows from Lemma \ref{simple-avg} that, for all $p \in \mathscr{P}$, there is a set $H'_p \subset H$ with $|H'_p| \geq (1 - c^{1/4}) |H|$ such that $|S_p \cap (S_p + h)| \geq (1 - c^{1/4}\eps)|S_p|$ for all $h \in H'_p$. Finally, using Lemma \ref{add-comb} applied to $S_{p}$ (and with $\epsilon$ replaced by $c^{1/4}\epsilon$) it follows that $|H'_p \md{p}| \leq 4 c^{1/4} \eps p + 1$, and so all but $c^{1/4} |H|$ elements of $H$ reduce to lie in a set of size $4 c^{1/4} \eps p + 1 < \frac{1}{100}p$ modulo $p$, for all $p \in \mathscr{P}$, a set which satisfies $\sum_{p \in \mathscr{P}} \log p/p \geq \frac{1}{40}\log Q$. We may apply the larger sieve, Theorem \ref{larger-sieve}, to this situation, taking $\delta = 1 - c^{1/4}$ and $\sigma_p = 1/100$ for all $p \in \mathscr{P}$. This gives the bound \[ |H| \ll \frac{Q}{100(1 - c^{1/4})^{2} (1/40)\log Q -\log X }\] provided that the denominator is positive. If $c$ is sufficiently small then this will be so with our choice of $Q$, namely $X^{1/2 - c/2}$, and we get the bound $|H| \ll X^{1/2 - c/2 + o(1)}$. This is contrary to assumption. \end{proof} Now we can prove Theorem \ref{prog-thm}, by applying the foregoing lemmas repeatedly to the intersection of $A$ (and of the sets $S_{p}$) with shifted copies of itself. The key point here is that, since any subset of $A$ will lie in the arithmetic progression $S_{p}$ when reduced modulo $p$, we can use Lemma \ref{progr-energy} throughout this ``intersecting process'' to obtain a lower bound on additive energy. In particular, we don't need to keep track of any ``uniformity of fibres'' throughout the process. Eventually we will obtain a subset of $A$ that has cardinality quite close to $|A|$, but lies modulo $p$ in a multiply intersected copy of $S_p$ having size $< (1/2-c)p$ (for most primes $p$). The theorem will then follow immediately from the larger sieve. \begin{proof}[Proof of Theorem \ref{prog-thm}] We assume that $\eps$ is small, and that $X$ is sufficiently large in terms of $\eps$. This is certainly permissible for proving the theorem. We will prove that $|A| < X^{1/2-c(\eps)}$, where $c(\eps) = K^{-1/\eps}$ for a large absolute constant $K > 0$. Suppose, for a contradiction, that $|A| \geq X^{1/2-c(\eps)}$. We proceed iteratively, setting $A_0 = A$ and constructing a sequence of sets $A_i \subset A$, $i = 1,2,3,\dots$ such that \begin{equation}\label{progr-size} |A_i| \geq X^{1/2-3^{i}K^{-1/\eps}}. \end{equation} The sets $A_i$ will satisfy $A_i \md{p} \subset S_p^i$, where $S_p^i \subset S_p$, and where \begin{equation}\label{progr-decay} \sum_{p \leq Q} \frac{\log p}{p} \frac{|S_p^i|}{p} < (1-c\eps/2)^{i}(\log Q + O(1)). \end{equation} Here $c$ is the absolute constant from Lemma \ref{progr-goodh}, and $Q=X^{1/2-c/2}$. After $O(1/\eps)$ steps we will, in particular, have $$ \sum_{p \leq Q} \frac{\log p}{p} \frac{|S_p^i|}{p} < (1/2 - c)\log Q , $$ and therefore, writing $\sigma^i_p := |S_p^i|/p$, we will have $$ \sum_{p \leq Q} \frac{\log p}{\sigma_p^i p} \geq 4\log Q + O(1) - 4(\frac{1}{2} - c)\log Q = (2 + 4c)\log Q + O(1) , $$ as argued in the proof of Lemma \ref{lem2.7}. Using the larger sieve, this implies that $|A_i| \ll X^{1/2-c/2}$, which contradicts the lower bound (\ref{progr-size}) provided that $K$ was chosen large enough. We will show how the set $A_{i+1}$ is obtained from $A_i$, and verify that it satisfies the size bound (\ref{progr-size}) and that its reductions modulo primes $p$ satisfy the bound (\ref{progr-decay}). Firstly, if $A_i \subset A$ satisfies (\ref{progr-size}) then Lemma \ref{progr-energy} implies that $$ E(A_i,A_i) \gg \frac{\eps^{4} |A_i|}{X^{1/2} \log X} |A_i|^3 \gg \frac{\eps^{4} X^{-3^{i}K^{-1/\eps}}}{\log X} |A_i|^3 \geq 2X^{-2 \cdot 3^{i}K^{-1/\eps}} |A_i|^3 , $$ provided that $X$ is large enough in terms of $\eps$. Using Lemma \ref{lem3.5}, it follows that there is a set $H \subset [-X,X]$ such that $|H| \geq |A_i|X^{-2 \cdot 3^{i}K^{-1/\eps}} \geq X^{1/2-3^{i+1}K^{-1/\eps}}$, and such that $$ |A_i \cap (A_i + h)| \geq X^{1/2-3^{i+1}K^{-1/\eps}} $$ for all $h \in H$. The set $A_{i+1}$ will be of the form $A_i \cap (A_i + h)$, for suitably chosen $h \in H$. Note that any such choice will indeed satisfy the size bound (\ref{progr-size}). In view of (\ref{progr-decay}), we may assume that the sets $S_p^i$ satisfy $$ (1/2 - c)\log Q \leq (1-c\eps/2)^{i+1}(\log Q + O(1)) \leq \sum_{p \leq Q} \frac{\log p}{p} \frac{|S_p^i|}{p} < (1-c\eps/2)^{i}(\log Q + O(1)), $$ the lower bound holding because if it failed we could simply set $A_{i+1}=A_i$. Now let $\mathscr{P}'$ denote the set of primes $p \leq Q$ for which $|S_p^i| \geq \frac{1}{10}p$. We must have \[ \sum_{p \in \mathscr{P}'} \frac{\log p}{p} \geq \frac{1}{3}\log Q \qquad \textrm{and} \qquad \sum_{p \in \mathscr{P}'} \frac{\log p}{p} \frac{|S_p^i|}{p} \geq \frac{1}{2} \sum_{p \leq Q} \frac{\log p}{p} \frac{|S_p^i|}{p}, \] say, because otherwise the lower bound we just assumed would be violated. Thus we can apply Lemma \ref{progr-goodh}, deducing that for some $h \in H$ we have $$ \sum_{p \in \mathscr{P}'} \frac{\log p}{p} \frac{|S_p^i \cap (S_p^i + h)|}{p} < (1 - c\eps) \sum_{p \in \mathscr{P}'} \frac{\log p}{p} \frac{|S_p^i|}{p} \leq \sum_{p \in \mathscr{P}'} \frac{\log p}{p} \frac{|S_p^i|}{p} - \frac{c\eps}{2}\sum_{p \leq Q} \frac{\log p}{p} \frac{|S_p^i|}{p} . $$ Finally, if we set $A_{i+1} = A_i \cap (A_i + h)$ for this choice of $h$, and correspondingly set $S_p^{i+1} = S_p^i \cap (S_p^i + h)$, then the upper bound (\ref{progr-decay}) will indeed be satisfied. \end{proof} \section{Robustness of the inverse large sieve problem}\label{stab-sec} In this section we prove Theorem \ref{stability-thm}. Let us begin by recalling the statement. \begin{stability-thm-repeat} Let $X_0 \in \N$, and let $\eps > 0$. Let $X \in \N$ be sufficiently large in terms of $X_0$ and $\eps$, and suppose that $H \leq X^{1/8}$. Suppose that $A,B \subset [X]$ and that $|A \md{p}| + |B \md{p}| \leq p+1$ for all $p \in [X_0, X^{1/4}]$. Then, for some absolute constant $c > 0$, one of the following holds: \begin{enumerate} \item \textup{(Better than large sieve)} Either $|A \cap [X^{1/2}]|$ or $|B \cap [X^{1/2}]|$ is $\leq X^{1/4 - c\eps^3}$; \item \textup{(Behaviour with quadratics)} Given any two rational quadratics $\psi_A,\psi_B$ of height at most $H$, either $|A \setminus \psi_{A}(\Q)|$ and $|B\setminus \psi_{B}(\Q)| \leq HX^{1/2 - c}$, or else at least one of $|A \cap \psi_{A}(\Q)|$ and $|B \cap \psi_{B}(\Q)|$ is bounded above by $HX^{1/4 + \eps}$. \end{enumerate} \end{stability-thm-repeat} The proof of Theorem \ref{stability-thm} requires a number of preliminary ingredients, which we assemble now. We start with some results concerning \emph{quasisquares}, that is to say squarefree integers that are quadratic residues modulo ``many'' primes (see, for example, \cite[Section 12.14]{dekoninck-luca}). The first result is not the one actually required later on (which is Lemma \ref{pseudosquares}), but it may be of independent interest and its proof motivates the proof of the result needed later. \begin{lemma}\label{quasi-square} Let $\eta > 0$, and suppose that $Y \geq Z > 2$. Suppose that $\mathscr{P} \subset [Z,2Z]$ is a set consisting of at least $\eta Z/\log Z$ of the primes in $[Z,2Z]$. Then the number of squarefree $q \in [1,Y]$ such that $\left( \frac{q}{p} \right) = 1$ for all $p \in \mathscr{P}$ is at most $8 (6 \log Y/\eta)^{3\log Y/\log Z}$. \end{lemma} \begin{proof} First of all, let $\mathscr{Q}$ denote the set of all $q$ that are squarefree, lie in $[1,Y]$ and are a quadratic residue modulo all $p \in \mathscr{P}$. Each $q \in \mathscr{Q}$ is either congruent to $1 \md{4}$, or it is congruent to $2$ or $3 \md{4}$. Let us assume that at least half of the elements of $\mathscr{Q}$ are congruent to $2$ or $3 \md{4}$, and henceforth redefine $\mathscr{Q}$ to consist of those elements only, and aim to show that $|\mathscr{Q}| \leq 4 (6 \log Y/\eta)^{3\log Y/\log Z}$. (The proof when at least half of the elements are congruent to $1 \md{4}$ is essentially the same.) Let $k \geq 3$ be the smallest integer for which $Z^k > Y^2$. Then if $n \in [Z^k, 2^k Z^k]$ is any product of $k$ distinct primes from $\mathscr{P}$, and if $q \in \mathscr{Q}$, we have that the Jacobi symbol $\left( \frac{q}{n} \right) = \prod_{p \mid n} \left( \frac{q}{p} \right) = 1$. Let $\mathscr{S}$ be the set of all such $n$; then clearly \begin{equation}\label{s-lower} |\mathscr{S}| = \binom{|\mathscr{P}|}{k} \geq \frac{|\mathscr{P}|^{k}}{k^{k}} \geq \left(\frac{\eta}{k\log Z}\right)^k Z^k . \end{equation} (Of course this is only true if $|\mathscr{P}| \geq k$, but otherwise the bound we shall derive is trivial anyway.) Finally, note that if $q$ is squarefree and congruent to $2$ or $3 \md{4}$, and if $n \in \mathscr{S}$ (so, in particular, $n$ is odd), then $\left( \frac{q}{n} \right) = \left( \frac{4q}{n} \right) = \chi_{4q}(n)$, where $\chi_{4q}(n)$ is a primitive character modulo $4q$. (It is the primitive quadratic character corresponding to the fundamental discriminant $4q$.) The multiplicative form of the large sieve \cite[Theorem 7.13]{iwaniec-kowalski} implies that \[ \sum_{\substack{q \leq Q, \\ q \equiv 2 \text{ or } 3 \md 4, \\ q \; \text{squarefree}}} \left|\sum_{n \in \mathscr{S}} a_{n} \chi_{4q}(n) \right|^2 \leq (16Q^2 + N) \sum_{n \in \mathscr{S}} |a_{n}|^{2} , \] for any set $\mathscr{S} \subset [N]$ and for any $Q$ and any coefficients $a_{n}$. Applying this with our particular set $\mathscr{S}$, and with $a_{n} = 1$, yields \[ |\mathscr{Q}||\mathscr{S}|^2 \leq (16Y^2 + 2^k Z^k)|\mathscr{S}| < 2^{k+2} Z^k |\mathscr{S}|,\] and therefore by \eqref{s-lower} \[ |\mathscr{Q}| \leq \frac{2^{k+2} Z^k}{|\mathscr{S}|} \leq 4 (\frac{2k\log Z}{\eta})^k.\] Noting that $k \leq \frac{2\log Y}{\log Z} + 1 \leq \frac{3\log Y}{\log Z}$, the result follows. \end{proof} \emph{Remarks.} The conclusion of Lemma \ref{quasi-square} is nontrivial when $Y$ is any fixed power of $Z$, and even for somewhat larger values of $Y$. It seems to us that the bound obtained here is stronger than could (straightforwardly) be obtained using the real character sum estimate of Heath-Brown \cite{hb}, which comes with an unspecified factor of $(QN)^{\eps}$. Now we present the result we need later on. Of course more general statements are possible, but we leave their formulation as an exercise to the interested reader. \begin{lemma}\label{pseudosquares} Suppose that $Z \geq Y^{1/5}$ are large. The number of squarefree $q \in [1,Y]$ which are squares modulo 95\% of the primes in $[Z, 2Z]$ is $O(\log^{10} Z)$. Furthermore, all of these $q$ apart from $q=1$ are at least $Y^\varrho$, for a certain absolute constant $\varrho > 0$. \end{lemma} \begin{proof} The argument for the first part closely follows the preceding. Write $\mathscr{Q}$ for the set of all squarefree $q \in [1,Y]$ which are squares modulo at least 95\% of the primes $p \in [Z,2Z]$. Write $\mathscr{P}_q$ for the set of these primes; note carefully that $\mathscr{P}_q$ may depend on $q$. Write $\mathscr{S}$ for the set of products of $10$ distinct primes from $[Z,2Z]$, and write $\mathscr{S}_q$ for the set of products of $10$ distinct primes from $\mathscr{P}_q$. Note that $\left(\frac{q}{n} \right) = 1$ whenever $n \in \mathscr{S}_q$. Furthermore, \[ |\mathscr{S}_q| = \binom{|\mathscr{P}_q|}{10} \geq (1 - o(1)) (1 - \frac{1}{20})^{10} \binom{|\mathscr{P}|}{10} > 0.59\binom{|\mathscr{P}|}{10} = 0.59|\mathscr{S}|\] for every $q \in \mathscr{Q}$ (the key point here is that $0.59 > 0.5$). It follows that for every $q \in \mathscr{Q}$ we have $\sum_{n \in \mathscr{S}} \left( \frac{q}{n} \right) \geq 0.18|\mathscr{S}|$. The proof now concludes as before. To prove the second part, we use a form of the prime number theorem for the real character $\chi(m) = \left( \frac{q}{m} \right)$ (which, provided $q > 1$ is squarefree, is always a non-principal character of conductor at most $4q$). This tells us (see e.g. Theorem 7 in Gallagher's paper~\cite{gallagherdensity}) that \[ \sum_{Z \leq m \leq 2Z} \Lambda(m) \chi(m) = O(Z \max ( e^{-c\sqrt{\log Z}}, e^{-c\log Z/\log q} )) \] if $\chi$ has no exceptional zero, and \[ \sum_{Z \leq m \leq 2Z} \Lambda(m) \chi(m) = \frac{Z^{\beta} - (2Z)^{\beta}}{\beta} + O(Z \max ( e^{-c\sqrt{\log Z}}, e^{-c\log Z/\log q} )),\] if $\chi$ does have an exceptional zero $\beta \in (\frac{1}{2}, 1)$. Either way, we have the 1-sided inequality \[ \sum_{Z \leq m \leq 2Z} \Lambda(m) \chi(m) \leq O(Z \max ( e^{-c\sqrt{\log Z}}, e^{-c\log Z/\log q} ) ).\] Since \[ \sum_{Z \leq m \leq 2Z} \Lambda(m) = Z(1 + o(1))\] by the prime number theorem, it follows that if $q \leq Y^{\varrho}$ with $\varrho$ small enough then at most $95$\% of the primes in $[Z,2Z]$ are such that $(q | p) = \chi(p) = 1$ .\end{proof} \begin{proof}[Proof of Theorem \ref{stability-thm}] A key role will be played by primes $p$ that are close to $X^{1/4}$. It is convenient to introduce some terminology concerning them. Let $C$ be a large absolute constant to be specified later. We say that $p \sim X^{1/4}$ if $X^{1/4 - C\eps} < p < X^{1/4}$. Furthermore, we will say that a certain property holds ``for at least 1\% of primes $p \sim X^{1/4}$'' if the set $\mathscr{P}$ of primes for which this property holds satisfies \[ \sum_{p \sim X^{1/4} : p \in \mathscr{P}} \frac{\log p}{p} \geq 0.01 \sum_{p \sim X^{1/4}} \frac{\log p}{p}.\] The weighting of $\log p / p$ is included with some later applications of the larger sieve in mind. We begin with some preliminary analysis using the larger sieve, strongly based on the work of Elsholtz \cite{elsholtz}. \begin{lemma}\label{large-sieve-app} Let $A, B$ be sets such that $|A \md{p}| + |B \md{p}| \leq p+1$ for all primes $p \in [X_0, X^{1/4}]$. Let $\eps > 0$, and suppose that $X$ is sufficiently large in terms of $X_0$ and $\eps$. Then either $A \cap [X^{1/2}]$ or $B \cap [X^{1/2}]$ has size at most $X^{1/4 - c\eps^3}$, or else for at least 99\% of primes $p \sim X^{1/4}$ we have both $|A \md{p}| \leq (\frac{1}{2} + \eps)p$ and $|B\md{p}| \leq (\frac{1}{2} + \eps) p$. We may take $c = 2^{-10}$. \end{lemma} \begin{proof} Write $\alpha_p := |A \md{p}|/p$, $\beta_p := |B \md{p}|/p$. Thus we are assuming that $\alpha_p + \beta_p \leq 1 + \frac{1}{p}$. Suppose the final statement of the lemma is false. Then \[ \sum_{p \sim X^{1/4}} \frac{\log p}{p} \big( (1 - 2\alpha_p)^2 + (1 - 2\beta_p)^2\big) \geq 2^{-5}\eps^2 \sum_{p \sim X^{1/4}} \frac{\log p}{p} \geq 2^{-5} \eps^3 \log X.\] If $c < 2^{-8}$, we can remove the contribution from $X^{1/4 - 2c\eps^3} \leq p \leq X^{1/4}$ trivially to get \[ \sum_{p < X^{1/4 - 2c\eps^3}} \frac{\log p}{p}\big( (1 - 2\alpha_p)^2 + (1 - 2\beta_p)^2\big) \geq 2^{-6} \eps^3 \log X.\] We claim that if $a,b$ are positive real numbers with $a + b \leq 1$ then \[ \frac{1}{a} + \frac{1}{b} \geq 4 + 2(1 - 2a)^2 + 2(1 - 2b)^2.\] To see this, apply the inequality \[ \frac{1}{x} + \frac{1}{1 - x} = 4 + \frac{(1 - 2x)^2}{x (1 - x)} \geq 4 + 4(1 - 2x)^2\] with $x = a$ and $x = b$ in turn, and add the results. Applying this together with the above, we obtain \[ \sum_{X_0 \leq p \leq X^{1/4 - 2c\eps^3}} \frac{\log p}{p} (\frac{1}{\alpha_p} + \frac{1}{\beta_p}) \geq (1 - 8c\eps^3)\log X + 2^{-5}\eps^3 \log X - O(1) \geq (1- 8c\eps^3 + 2^{-6} \eps^3)\log X .\] Here we used the fact (an estimate of Mertens) that $\sum_{X_0 \leq p \leq Z} \frac{\log p}{p} = \log Z + O_{X_0}(1)$. Note also that we only have $\alpha_p + \beta_p \leq 1 +\frac{1}{p}$, and not $\alpha_p + \beta_p \leq 1$; the introduction of the $O(1)$ term takes care of this as well, the full justification of which we leave to the reader\footnote{We are working with the condition $|A \md{p}| + |B \md{p}| \leq p+1$, rather than the cleaner condition $|A \md{p}| + |B \md{p}| \leq p$, so that we can formulate Theorem \ref{stability-thm-symmetric} to include the case in which $\A$ is the set of values of a quadratic. Note, however, that in this case Lemma \ref{large-sieve-app} is vacuous anyway. Therefore this small point really can be ignored.}. Without loss of generality the contribution from the $\alpha_p$ is at least that from the $\beta_p$, so \[ \sum_{X_0 \leq p \leq X^{1/4 - 2c\eps^3}} \frac{\log p}{p} \frac{1}{\alpha_p} \geq (\frac{1}{2} - 4c\eps^3+ 2^{-7}\eps^3)\log X > (\frac{1}{2} + 2^{-8}\eps^3) \log X\] if $c \leq 2^{-10}$. Then, however, the larger sieve implies that \[ |A \cap [X^{1/2}]| \ll \frac{X^{1/4 - 2c\eps^3}}{\eps^3 \log X} < X^{1/4 - c\eps^3},\] and the result follows. \end{proof} Suppose now that the hypotheses are as in Theorem \ref{stability-thm}. Replace $\eps$ by $\eps/2C$ (the statement of the theorem does not change). Let $\psi_A,\psi_B$ be rational quadratics of height at most $H$. If option (ii) of the theorem does not hold then at least $HX^{1/4 + 2C\eps}$ elements of $A$ lie in $\psi_{A}(\Q)$ and at least $HX^{1/4 + 2C\eps}$ elements of $B$ lie in $\psi_{B}(\Q)$. Suppose also that option (i) of Theorem \ref{stability-thm} does not hold. Then we may apply Lemma \ref{large-sieve-app} to conclude that both $|A \md{p}|$ and $|B \md{p}|$ are $\leq (\frac{1}{2} + \eps)p$ for at least 99\% of all primes $p \sim X^{1/4}$. (We urge the reader to recall the special meaning of this notation.) Using this information together with the fact that $|A \md{p}| + |B \md{p}| \leq p+1$ for all $p \sim X^{1/4}$, we will deduce that in fact \begin{equation}\label{weak} |A \setminus \psi_{A}(\Q)|, |B \setminus \psi_{B}(\Q)| \ll HX^{1/2 - c}, \end{equation} this being the other conclusion of Theorem \ref{stability-thm} (ii). It suffices to prove this for $A$, the proof for $B$ being identical. Write $\psi = \psi_{A}$, and suppose that $p \sim X^{1/4}$. Set \[ T_p := A \md{p} \cap (\psi(\Q) \cap \Z) \md p,\] \[ U_p := A \md{p} \setminus (\psi(\Q) \cap \Z) \md{p}.\] We know that $|A \md{p}| \leq (\frac{1}{2} + \eps) p$ for at least 99\% of all primes $p \sim X^{1/4}$. For these primes, then, \[ |T_p| + |U_p| \leq (\frac{1}{2} + \eps)p.\] We claim that $|U_p| \leq 2\eps p$ for at least 98\% of all primes $p \sim X^{1/4}$. If this failed, we would have \begin{equation}\label{tp-bound}|T_p| \leq (\frac{1}{2} - \eps) p\end{equation} for at least $1\%$ of the primes $p \sim X^{1/4}$. Write $\psi(x) = \frac{1}{d}(a x^2 + bx + c)$ with $a,b,c,d$ having no common factor and $|a|, |b|, |c|, |d| \leq H$, and note that $\psi^{-1}(\Z) \subset \frac{1}{a}\Z$. Note furthermore that \begin{equation}\label{35} \{ x \in \Z : \psi(\frac{x}{a}) \in A \cap \psi(\Q)\} \subset \bigcap_{p \sim X^{1/4}} \{ x \in [-C_1HX^{1/2}, C_1HX^{1/2}] : \psi(\frac{x}{a}) \md{p} \in T_p\} ,\end{equation} where $C_1$ is some absolute constant. If $p \sim X^{1/4}$, the condition that $\psi(\frac{x}{a}) \md{p} \in T_p$ forces $x \md{p}$ to lie in some set $S_p \subset \Z/p\Z$ of residue classes with $|S_p| \leq 2|T_p|$. We now have a large sieve problem to which Proposition \ref{classicls} may be applied. If $p$ is such that \eqref{tp-bound} holds then we have $|S_p^c|/|S_p| \geq \eps$, and so \[ \sum_{q \leq X^{1/4}} \mu^2(q) \prod_{p | q} \frac{|S_p^c|}{|S_p|} \geq \sum_{p \sim X^{1/4}} \frac{|S_p^c|}{|S_p|} \gg \eps \frac{X^{1/4 -C\eps}}{\log X}.\] Here, $X^{1/4 - C\eps}/\log X$ is a crude lower bound for the size of a set of primes constituting 1\% of all $p \sim X^{1/4}$, this lower bound being attained when all the primes congregate at the bottom of the interval $X^{1/4 - C\eps} \leq p \leq X^{1/4}$. It follows from \eqref{35} and Proposition \ref{classicls} that $|A\cap \psi(\Q)| < HX^{1/4 + 2C\eps}$ if $X$ is large enough, contrary to assumption. Now let us write $A= A_{\psi} \cup E$, where $A_{\psi}$ consists of those $x \in A$ for which $x \md{p} \in T_p$ for at least $97\%$ of $p \sim X^{1/4}$, and $E$ consists of those $x \in A$ such that $x \md{p} \in U_p$ for at least $3\%$ of $p \sim X^{1/4}$. The idea here is that $A_{\psi}$ satisfies a large number of local conditions suggesting that its elements lie in $\psi(\Q)$. We would like to relate $A_{\psi}$ to $A \cap \psi(\Q)$, and show that $E$ is small. With this idea in hand, we can divide the task of proving \eqref{weak} into two subclaims, namely \begin{equation}\label{claim} |A_{\psi} \setminus \psi(\Q)| \ll H X^{1/2 - c} \qquad \mbox{and} \qquad |E| \ll X^{1/4} .\end{equation} Of course, we could tolerate a weaker bound for $|E|$, but as it turns out we need not settle for one. We start with the first claim, which is quite straightforward given results we established earlier. Let $\Delta$ be the discriminant of $\psi$. If $x$ is an integer then so is $4ad x + \Delta d^2$, and furthermore if $x =\psi(n)$ then $4a d x + \Delta d^2 = (2a n + b)^2$. Therefore if $x \in A_{\psi}$ then $4a d x + \Delta d^2$ is an integer which is a square modulo $p$ for at least 97\% of all $p \sim X^{1/4}$. By a simple averaging argument, $4ad x + \Delta d^2$ is a square modulo at least 95\% of all $p\in [Z,2Z]$ for some $Z \sim X^{1/4}$. It follows from Lemma \ref{pseudosquares} that \begin{equation}\label{false-eq}4a d x + \Delta d^2 = n_i s^2,\end{equation} where $s \in \Z$ and $n_i$ is one of at most $O(\log^{10} X)$ squarefree integers, with $n_1 = 1$ and $n_i \geq X^{\varrho}$ if $i \geq 2$, where $\varrho > 0$ is an absolute constant. The number of $x \in [X]$ for which \eqref{false-eq} holds for a given $i$ is $\ll H\sqrt{X/n_i}$, and so the number of $x$ for which this holds for \emph{some} $i \geq 2$ is $\ll H \log^{10} X \cdot X^{(1- \varrho)/2} \ll H X^{1/2-\varrho/4}$. If $i = 1$, so that $n_1 = 1$, then $4a d x + \Delta d^2$ is a square and so $x \in \psi(\Q)$. This concludes the proof of the first bound in \eqref{claim}. We turn now to the proof of the second bound in \eqref{claim}. Recall first of all that $|U_p| \leq 2\eps p$ for at least 98\% of all $p \sim X^{1/4}$, and also that for every $x \in E$ we have $x \md{p} \in U_p$ for at least 3\% of all $p \sim X^{1/4}$. For every $x \in E$, both of these events occur for at least 1\% of all $p \sim X^{1/4}$. Write $E_p$ for the subset of $E$ whose elements belong to $U_p$ modulo $p$. By the preceding facts we have \[ \sum_{\substack{p \sim X^{1/4} \\ |U_p| \leq 2\eps p}} \frac{\log p}{p}|E_p| = \sum_{x \in E} \sum_{\substack{p \sim X^{1/4} \\ |U_p| \leq 2\eps p}} \frac{\log p}{p}1_{x \md{p} \in U_p} \geq \frac{1}{100}|E|\sum_{p \sim X^{1/4}} \frac{\log p}{p}. \] Writing $\mathscr{P} := \{ p \sim X^{1/4} : |U_p| \leq 2\eps p \; \text{and} \; |E_p| \geq \textstyle\frac{1}{200}|E| \}$, it follows that \[ \sum_{p \in \mathscr{P}} \frac{\log p}{p} \geq \frac{1}{200} \sum_{p \sim X^{1/4}} \frac{\log p}{p}.\] Applying the larger sieve (that is Theorem \ref{larger-sieve}) with the choices $\delta = \frac{1}{200}$, $Q = X^{1/4}$ and $\sigma_{p} = 2\eps$, we obtain $|E| \ll X^{1/4}(\frac{1}{2^{20}\eps}\sum_{p \sim X^{1/4}} \frac{\log p}{p} - \log X)^{-1}$, provided that the term in parentheses is positive. That term is $> (2^{-21} C - 1)\log X$, which is positive if $C > 2^{22}$ (say). Thus we get the bound $|E| \ll X^{1/4}$. This completes the proof of \eqref{claim}, and hence \eqref{weak} and Theorem \ref{stability-thm}.\end{proof} We turn now to the proof of Theorem \ref{stability-thm-symmetric}, the stability theorem for a single infinite set $\A$. Again, we begin by recalling the statement. \begin{stability-thm-symmetric-repeat} Suppose that $\A$ is a set of positive integers and that $|\A \md{p}| \leq \frac{1}{2}(p+1)$ for all sufficiently large primes $p$. Then one of the following options holds: \begin{enumerate} \item \textup{(Quadratic structure)} There is a rational quadratic $\psi$ such that all except finitely many elements of $\A$ are contained in $\psi(\Q)$; \item \textup{(Better than large sieve)} For each integer $k$ there are arbitrarily large values of $X$ such that $|\A[X]| < \frac{X^{1/2}}{\log^k X}$; \item \textup{(Far from quadratic structure)} Given any rational quadratic $\psi$, for all $X$ we have $|\A[X] \cap \psi(\Q)| \leq X^{1/4 + o_{\psi}(1)}$. \end{enumerate} \end{stability-thm-symmetric-repeat} \begin{proof} Suppose that neither item (ii) nor item (iii) holds. Then there is an $\eps > 0$ and a rational quadratic $\psi$ such that $|\A[X] \cap \psi(\Q)| > X^{1/4 + \eps}$ for arbitrarily large values of $X$. For any such $X$ we may apply Theorem \ref{stability-thm} with $A = \A[X]$ to conclude that either option (ii) of our present theorem holds, or else for infinitely many $X$ we have $|\A[X] \setminus \psi(\Q)| \ll X^{1/2 - c}$. (We note in passing that Lemma \ref{large-sieve-app} is redundant inside the proof of Theorem \ref{stability-thm} in this setting, being trivially true.) We will deduce from this and further applications of the fact that $|\A \md{p}| \leq \frac{1}{2}(p+1)$ for $p$ sufficiently large that either (i) or (ii) of Theorem \ref{stability-thm-symmetric} holds. Let $\tilde \psi$ be a rational quadratic satisfying the conclusions of Lemma \ref{tedious-local-lemma}, thus $\psi(\Q) \cap \Z \subset \tilde\psi(\Z)$ and for all sufficiently large primes $p$ the reductions $\md{p}$ of $\psi(\Q) \cap \Z$ and of $\tilde\psi(\Z)$ are the same. Suppose that option (ii) of Theorem \ref{stability-thm-symmetric} does not hold for the infinitely many values of $X$ for which we have $|\A[X] \setminus \psi(\Q)| \ll X^{1/2 - c}$. Then there is some integer $k$ such that (letting $X$ range through these values) \begin{equation}\label{weak2} \lim\sup_{X \rightarrow \infty} X^{-1/2}\log^k X|\A[X] \cap \psi(\Q)| = \infty.\end{equation} We claim that this implies statement (i) of Theorem \ref{stability-thm}, and in fact the stronger conclusion $|\A \setminus \psi(\Q)\big| \leq k+1$. Suppose this statement is false. Then there are elements $x_1,\dots,x_{k+1}$ in $\A$ but not in $\psi(\Q)$. Since $x$ lies in $\psi(\Q)$ if and only if $4ad x + \Delta d^2$ is the square of a rational number, it follows that none of $4ad x_i + \Delta d^2$ is a square. Set $m_i := 4ad x_i + \Delta d^2$, and suppose that $p$ is a prime such that $(m_i | p ) = -1$. If $p$ is sufficiently large then $x_i \notin (\psi(\Q) \cap \Z) \md{p}$ and hence $x_i \notin \tilde\psi(\Z) \md{p}$. For each prime $p$, let $k(p)$ be the number of indices $i \in \{1,\dots, k+1\}$ such that $(m_i | p) = -1$. From the above reasoning and the assumption that $x_i \in \mathscr{A}$ it follows that $\A \cap \tilde\psi(\Z) \md{p}$ must occupy a set of size at most $\frac{1}{2}(p+1) - k(p)$ for all sufficiently large primes $p$. Define a set $\mathscr{B} \subset \Z$ by $\A \cap \tilde\psi(\Z) = \tilde\psi(\mathscr{B})$. Thus $|\tilde\psi(\mathscr{B}) \md{p}| \leq \frac{1}{2}(p+1) - k(p)$ for all sufficiently large primes $p$, which implies that $|\mathscr{B} \md{p}| \leq p - 2k(p) + 1$. Note also that $\A[X] \cap \psi(\Q) \subset \tilde\psi{(\mathscr{B} \cap [c_1 \sqrt{X}, c_2 \sqrt{X}])}$ for some constants $c_1,c_2$ depending only on $\tilde\psi$. We may now apply Lemma \ref{small-from-large} to the set $\mathscr{B}$. In that lemma we may take $w(p) = 2k(p) - 1$, where $k(p)$ is the number of $i$ for which $(m_i | p) = -1$, or equivalently (if $p > 2$) for which $\chi_i(p) = -1$ where $\chi_i(n) = (4m_i | n)$ is a real Dirichlet character and $( \, | \, )$ denotes the Kronecker symbol. Thus $k(p) = \frac{1}{2}(k+1) - \frac{1}{2}\sum_{i=1}^{k+1} \chi_i(p)$, and so $w(p) = k -\sum_{i = 1}^{k+1} \chi_i(p)$. The conditions of Lemma \ref{small-from-large} are easily satisfied by the prime number theorem for characters with a fairly crude error term. It follows from Lemma \ref{small-from-large} and the above discussion that \[ |\mathscr{A}[X] \cap \psi(\Q)| \leq |\mathscr{B} \cap [c_1 \sqrt{X}, c_2 \sqrt{X}]| \ll X^{1/2}(\log X)^{-k},\] contrary to \eqref{weak2}. \end{proof} \section{Composite numbers in $\A + \B$} Recall from the introduction the following conjecture of ``inverse large sieve'' type. \begin{stability-conj-repeat} Let $X_0 \in \N$, and let $\rho > 0$. Let $X \in \N$ be sufficiently large in terms of $X_0$ and $\rho$. Suppose that $A,B \subset [X]$ and that $|A \md{p}| + |B \md{p}| \leq p+1$ for all $p \in [X_0, X^{1/4}]$. Then there exists a constant $c = c(\rho) > 0$ such that one of the following holds: \begin{enumerate} \item \textup{(Better than large sieve)} Either $|A \cap [X^{1/2}]|$ or $|B \cap [X^{1/2}]|$ is $\leq X^{1/4 - c}$; \item \textup{(Quadratic structure)} There are two rational quadratics $\psi_A,\psi_B$ of height at most $X^{\rho}$ such that $|A \setminus \psi_{A}(\Q)|$ and $|B\setminus \psi_{B}(\Q)| \leq X^{1/2 -c}$. \end{enumerate} \end{stability-conj-repeat} Our aim in this section is to prove Theorem \ref{ils-ostmann}, which is the following statement. \begin{ils-ostmann-repeat} Assume Conjecture \ref{stability-conj}. Let $\A, \B$ be two sets of positive integers, with $|\A|, |\B| \geq 2$, such that $\A + \B$ contains all sufficiently large primes. Then $\A + \B$ also contains infinitely many composite numbers. \end{ils-ostmann-repeat} This follows quite straightforwardly from the following fact. \begin{lemma}\label{lem6.2} There is $\rho > 0$ with the following property. Suppose that $\psi_{A},\psi_{B}$ are two rational quadratics of height at most $X^{\rho}$ and that $A \subset \psi_{A}(\Q) \cap [X]$ and $B \subset \psi_{B}(\Q) \cap [X]$ are sets of positive integers. Suppose that $A + B$ contains no composite number. Then at least one of $A$ and $B$ has cardinality $\ll X^{1/3}$. \end{lemma} \begin{proof}[Proof of Theorem \ref{ils-ostmann} given Lemma \ref{lem6.2}] Let $\A, \B$ be two sets of positive integers with $|\A|, |\B| \geq 2$ such that $\A + \B$ coincides with the set of primes on $[X_0,\infty)$. We claim that there are infinitely many $X$ such that either $|\A[X]|$ or $|\B[X]|$ has cardinality at most $X^{1/2-c}$. This, however, is contrary to a theorem\footnote{The state-of-the-art here is $|\A[X]| \gg X^{1/2}/\log X \log\log X$: see \cite{elsholtz-harper}.} of Elsholtz \cite{elsholtz}, which implies that $|\A[X]|, |\B[X]| \gg X^{1/2} \log^{-5} X$ for all sufficiently large $X$. It remains to prove the claim. Let $\rho$ be as in Lemma \ref{lem6.2}. For each $X$, write $A := \A \cap (X^{1/4},X]$ and $B := \B \cap (X^{1/4}, X]$. If $p \in [X_0, X^{1/4}]$ then $A + B$ contains no multiple of $p$, since any such number would be a nontrivial multiple of $p$ (and hence composite) and lies in $\A + \B$. For these primes $p$, then, we have $|A \md{p}| + |B \md{p}| \leq p$, since $B \md{p}$ cannot intersect $(-A) \md{p}$. Assuming Conjecture \ref{stability-conj}, for each $X$ one of the two options (i) or (ii) of that conjecture holds. If (i) holds for infinitely many $X$ then without loss of generality we have $|A \cap [X^{1/2}]| \ll X^{1/4 - c}$ for infinitely many $X$. By Elsholtz \cite{elsholtz} we have $|\A[X^{1/4}]| \ll X^{1/8 + o(1)}$, and therefore $|\A[X^{1/2}]| \ll X^{1/4 - c}$ for infinitely many $X$, thereby establishing the claim. Suppose, then, that (ii) holds for all sufficiently large $X$. That is, there are rational quadratics $\psi_{A}, \psi_{B}$ of height $X^{\rho}$ such that $|A\setminus \psi_{A}(\Q)|, |B \setminus \psi_{B}(\Q)| \leq X^{1/2-c}$. Write $A' := A \cap \psi_{A}(\Q)$ and $B' := B \cap \psi_{B}(\Q)$. Certainly $A' + B'$ contains no composite numbers. By Lemma \ref{lem6.2}, for all $X$ at least one of $A', B'$ has cardinality $\ll X^{1/3}$, and this means that indeed either $\A[X]$ or $\B[X]$ has size at most $X^{1/2-c}$ for infinitely many $X$. \end{proof} \begin{proof}[Proof of Lemma \ref{lem6.2}] Write $\psi_{A}(x) = \frac{1}{d_A}(a_A x^2 + b_A x + c_A)$, $\psi_{B}(x) = \frac{1}{d_B}(a_B x^2 + b_B x + c_B)$. Here $a_A, a_B, b_A, b_B, c_A, c_B, d_A, d_B $ are integers, all of magnitude at most $H = X^{\rho}$. Set $Y := (H^{2}X)^{1/4}$. If $A + B$ contains no composite number then the set $(A \cap (Y, X]) + (B \cap (Y, X])$ contains no multiple of any prime $p \leq Y$. Note also that $\psi_{A}^{-1}(\Z) \subset \frac{1}{a_A}\Z$ and $\psi_{B}^{-1}(\Z) \subset \frac{1}{a_B} \Z$. Set \[ S_A := \{ x \in \Z : \psi_{A}(\frac{x}{a_A}) \in A \cap (Y, X] \}, \quad S_B := \{ x \in \Z : \psi_{B}(\frac{x}{a_B}) \in B \cap (Y, X] \},\] and note that $\psi_{A}(\frac{x_A}{a_A}) + \psi_{B}(\frac{x_B}{a_B}) \neq 0 \md{p}$ whenever $x_A \in S_A$, $x_B \in S_B$, and $p \leq Y$ is a prime. To prove Lemma \ref{lem6.2}, it suffices to show that either $|S_A|$ or $|S_B|$ has size $\ll X^{1/3}$. Note furthermore (by completing the square) that $S_A,S_B \subset [-4(H^{2}X)^{1/2}, 4(H^{2}X)^{1/2}]$. We will focus attention only on those primes $p \leq Y$ for which $(-a_A a_B d_A d_B | p) = 1$, that is to say for which $-a_A a_B d_A d_B$ is a square modulo $p$. We look for such primes amongst the $p \leq Y$ with $p \equiv 1 \md{8}$. Since both $-1$ and $2$ are squares modulo such a prime, it certainly suffices to additionally ensure that $(q_1 |p) = 1, \dots, (q_k | p) = 1$, where $q_1,\dots, q_k$ are the distinct odd primes appearing in $a_A a_B d_A d_B$. This is equivalent to the union of $(q_1 - 1) \dots (q_k - 1)/2^k$ congruence conditions modulo $q_1 \dots q_k$. Together with the condition $p \equiv 1\md{8}$, we get the union of at least $2^{-k-2} \phi(q)$ congruence conditions modulo $q := 8 q_1 \dots q_k$. Since $|a_A|, |a_B|, |d_A|, |d_B| \leq H$ we have $q \leq 8 H^4$. Now we invoke \cite[Corollary 18.8]{iwaniec-kowalski}, a quantitative version of Linnik's theorem on the least prime in an arithmetic progression, which implies that there are at least $\frac{Y}{\phi(q)\sqrt{q}\log Y}$ primes $p \leq Y$ satisfying each of these congruence conditions, and hence $\gg \frac{2^{-k}Y}{\sqrt{q}\log Y}$ such primes in total. (Here we used the fact that $H = X^{\rho}$ with $\rho$ sufficiently small.) Write $\mathscr{P}$ for the set of such primes, thus $|\mathscr{P}| \gg X^{1/4 - o(1)}H^{-2}$. Now suppose that $p$ is such a prime and that $-a_A d_A/a_B d_B \equiv m^2 \md{p}$. Then we have \begin{align*} \psi_A(\frac{x_A}{a_A}) + \psi_B (\frac{x_B}{a_B}) = \frac{1}{a_A d_A} (x_{A}^{2} + b_{A}x_{A} + a_{A}c_{A}) & + \frac{1}{a_B d_B} (x_{B}^{2} + b_{B}x_{B} + a_{B}c_{B}) \\ & \equiv \frac{1}{a_A d_A} \big(( x_A + c_1)^2 - (mx_B + c_2)^2 + c_3\big), \end{align*} where $c_1,c_2,c_3$ do not depend on $x_A, x_B$. Therefore for each prime $p \in \mathscr{P}$ we have one of the following alternatives. \begin{enumerate} \item $c_3 \equiv 0$ modulo $p$. Then whenever $x_A \in S_A \md{p}$ we must have $\frac{1}{m}(x_A + c_1 - c_2) \notin S_B \md{p}$, whence $|S_A \md{p}| + |S_B \md{p}| \leq p$. \item $c_3 \not\equiv 0$ modulo $p$. Then we have $\psi_{A}(\frac{x_A}{a_A}) + \psi_B(\frac{x_B}{a_B}) \equiv 0 \md{p}$ whenever \[ (x_A + mx_B + c_1 + c_2)(x_A - mx_B + c_1 - c_2) \equiv -c_3,\] an equation which has $p-1$ solutions $(x_A, x_B)$. This solution set must be disjoint from $S_A \times S_B$. Since to each $x_A$ there are at most two $x_B$, and to each $x_B$ there are at most two $x_A$, this forces at least one of $S_A \md{p}, S_B \md{p}$ to have size $\leq 7p/8$, say. \end{enumerate} In both cases at least one of $S_A \md{p}, S_B \md{p}$ has size $\leq 7p/8$. Without loss of generality the first holds for at least half the elements of $\mathscr{P}$. Finally the large sieve, as in Proposition \ref{classicls}, tells us that $|S_A| \ll \frac{(H^{2}X)^{1/2}}{|\mathscr{P}|} \ll H^3X^{1/4 + o(1)}$. If $\rho$ is small enough then this is $\ll X^{1/3}$, as required. \end{proof}
{ "timestamp": "2013-11-26T02:10:27", "yymm": "1311", "arxiv_id": "1311.6176", "language": "en", "url": "https://arxiv.org/abs/1311.6176", "abstract": "Suppose that an infinite set $A$ occupies at most $\\frac{1}{2}(p+1)$ residue classes modulo $p$, for every sufficiently large prime $p$. The squares, or more generally the integer values of any quadratic, are an example of such a set. By the large sieve inequality the number of elements of $A$ that are at most $X$ is $O(X^{1/2})$, and the quadratic examples show that this is sharp. The simplest form of the inverse large sieve problem asks whether they are the only examples. We prove a variety of results and formulate various conjectures in connection with this problem, including several improvements of the large sieve bound when the residue classes occupied by $A$ have some additive structure. Unfortunately we cannot solve the problem itself.", "subjects": "Number Theory (math.NT)", "title": "Inverse questions for the large sieve", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9843363485313248, "lm_q2_score": 0.8128673178375735, "lm_q1q2_score": 0.8001348474806889 }
https://arxiv.org/abs/2106.10388
Upper bounds for critical probabilities in Bernoulli Percolation models
We consider bond and site Bernoulli Percolation in both the oriented and the non-oriented cases on $\mathbb{Z}^d$ and obtain rigorous upper bounds for the critical points in those models for every dimension $d \geq 3$.
\section{Introduction} The study of Bernoulli percolation on $\mathbb{Z}^d$ is more than 60 years old and the existence of a non-trivial phase transition for $d\geq 2$ is well established for the model and several of its variants, but the exact value of the critical parameter $p_c$ is seldom known. A celebrated result of Kesten (see \cite{K3}) proved that the critical probability in Bernoulli bond percolation on $\mathbb{Z}^2$ is $1/2$. Beyond that, a handful of planar lattices had their critical probability established and planarity was always the key factor. On the other hand, for dimensions $d\geq 3$, there is not much hope of finding exact values for the critical probability and the best we can expect are numerical results via Monte Carlo Methods (see the very efficient algorithm in \cite{NZ}), statistical estimates based on a comparison with dependent percolation (see Section 6.2 of \cite{BR} for an overview) or rigorous bounds (see for instance \cite {W,WP}). In this paper, we focus on finding rigorous upper bounds for site and bond Bernoulli Percolation on $\mathbb{Z}^d$ for every $d\geq 3$, in both the oriented and the non-oriented cases. The primary tools we use are couplings between the models we seek to understand and models where bounds or precise values for the critical probabilities are known. Although those bounds are still far from the values obtained from Monte Carlo simulations, we believe that most of them are the best rigorous upper bounds in the literature. The remainder of the text is organized as follows: in Section \ref{sec:model} we define more precisely the models and state the main results, in Section \ref{sec:couplings} we establish some dynamical couplings, in Section \ref{sec:proofs} we prove the theorems and in Section \ref{sec:comments} we give a numerical table with upper bounds for the critical probabilities in homogeneous Bernoulli percolation models for dimensions up to $d=9$. \section{The models and main results}\label{sec:model} For $d \geq 1$, the underlying graph for all the models will be the $d$-dimensional hypercubic lattice in which the set of vertices is $\mathbb{Z}^d$ and the set of edges is the set of non ordered pairs $E(\mathbb{Z}^d) : = \{ \langle v, u \rangle : v, u \in \mathbb{Z}^d ~ \text{and} ~ |v-u| =1\}$. We abuse notation and denote this graph simply by $\mathbb{Z}^d$. \subsection{Bond percolation} Given $p \in [0,1]$, consider a family of independent random variables $\{X_e\}_{e \in E(\mathbb{Z}^d)}$, where, for each $e \in E(\mathbb{Z}^d)$, $X_e$ has Ber($p$) distribution. Let $\mu_e$ be the law of $X_e$, and let $\mathbb{P}_p := \prod_{e \in E(\mathbb{Z}^d)} \mu_e$ be the resulting product measure. We declare an edge $e$ to be {\it open} if $X_e=1$ and {\it closed} otherwise. We first consider the non-oriented case and denote by $\{x \leftrightarrow y\}$ the event where $x, y \in \mathbb{Z}^d$ are connected by an open path, i.e., there exist $x_0, \dots, x_n$ such that $x_0 = x$, $x_n =y$ and each $\langle x_{j-1}, x_j \rangle$ belongs to $ E(\mathbb{Z}^d)$ and is open for $j=1, \dots, n$. Let $\mathcal{C}^b_0 := \{ x \in \mathbb{Z}^d: 0 \leftrightarrow x\}$ be the open cluster of the origin, and $|\mathcal{C}^b_0|$ its size. We define the percolation probability by $\theta^{b}_d(p) := \mathbb{P}_p(|\mathcal{C}^b_0|= \infty)$. The critical point for the {\it non-oriented bond} Bernoulli percolation model will be denoted by \[p_c^{b}(d)=\sup\{p \geq 0 : \theta^b_d(p)=0\}.\] We now consider the oriented case. Let $\{e_1, \dots, e_d\}$ be the set of positive unit vectors of $\mathbb{Z}^d$. We denote by $\{x \rightarrow y\}$ the event where $x, y \in \mathbb{Z}^d$ are connected by an oriented open path, i.e., there exist $x_0, \dots, x_n$ such that $x_0 = x$, $x_n =y$ and for each $j=1, \dots, n$, we have $x_j = x_{j-1} + e$, for some $e \in \{e_1, \dots, e_d\}$ and $\langle x_{j-1}, x_j\rangle$ is open. Let $\vv{\mathcal{C}}^b_0 := \{ x \in \mathbb{Z}^d : 0 \rightarrow x\}$ be the oriented open cluster of the origin, and $|\vv{\mathcal{C}}^b_0|$ its size. Analogously, we define $\vv{\theta}^{\, b}_d(p) := \mathbb{P}_p(|\vv{\mathcal{C}}^b_0|= \infty)$ the corresponding oriented percolation probability, and we denote the critical point for the {\it oriented bond} Bernoulli percolation model by \[\vv{p}_c^{\, b}(d)=\sup\{p \geq 0 : \vv{\theta}^{\,b}_d(p)=0\}.\] Our results are the following: \begin{theorem}\label{theo:bounds} Consider non-oriented bond Bernoulli percolation and let $p^{\ast}(d)$ be the unique solution in $(0,1)$ of \[\prod_{i=0}^2 \left(1-(1-p)^{\lfloor\frac{d+i}{3}\rfloor}\right)=2-\sum_{i=0}^2(1-p)^{\lfloor\frac{d+i}{3}\rfloor}.\] Then, for every $d\geq 3$, we have $p_c(d) \leq p^{\ast}(d)$. \end{theorem} \begin{theorem}\label{theo:bondO} Consider oriented bond Bernoulli percolation on $\mathbb{Z}^d$. 1) If $d$ is even, then $ \vv{p}^{\,b}_c(d)\leq 1-(1/3)^{2/d}$; 2) For $d \geq 4$, we have that \[\vv{p}^{\,b}_c(d)\leq \frac{1}{d}+\frac{C_d}{d^2},\] where \[C_d=1+\frac{8}{d}+\frac{d^{5/2}}{(\sqrt{2\pi})^{d-1}}\left[\frac{d-1}{d-3}\right]e^{\frac{1}{12d}}.\] 3) For any dimension $d\geq 2$, we have that $\vv{p}^{\,b}_c(d+1)\leq f(d)$, where $f(d)$ is the unique solution in $(0,1)$ of \[ p = \vv{p}^{\,b}_c(d) \left[p + (1-p)^{(d+1)/d}\right]. \] \end{theorem} {\bf Remark:} It is known (see \cite{CD}) that $\vv{p}^{\,b}_c(d)\sim 1/d$, hence the third upper bound above is asymptotically sharp. \subsection{Site percolation} Given a parameter $p \in [0,1]$, we consider a family $\{X_v\}_{v \in \mathbb{Z}^d}$ of independent Bernoulli random variables with parameter $p$. As before, $\mathbb{P}_p$ will denote the resulting product measure. A vertex $v \in \mathbb{Z}^d$ is declared to be {\it open} if $X_v = 1$ and {\it closed} otherwise. Now, the sequence of vertices $(x_0, \dots, x_n)$ is said to be an open path if all the vertices are open and for each $j = 1, \dots, n$, $\langle x_{j-1}, x_j\rangle \in E(\mathbb{Z}^d)$. If in addition to these conditions, for each $j = 1, \dots, n$, we have $x_j = x_{j-1}+ e$ for some $e \in \{e_1, \dots, e_d\}$, the sequence is said to be an oriented open path. Having these definitions we define $\mathcal{C}^s_0$, $\vv{\mathcal{C}}^s_0$, $\theta_d^{\, s}(p)$ and $\vv{\theta}_d^{\, s}(p)$ accordingly. Finally, the critical points for {\it non-oriented} and {\it oriented} site Bernoulli percolation are respectively given by \[ p_c^{s}(d)=\sup\{p \geq 0 : \theta^s_d(p)=0\} \quad \quad \text{and} \quad \quad \vv{p}_c^{\, s}(d)=\sup\{p \geq 0 : \vv{\theta}^{\,s}_d(p)=0\} .\] For site percolation models, our results are the following \begin{theorem}\label{theo:siteNO} Consider non-oriented site Bernoulli percolation on $\mathbb{Z}^d$. 1) If $d$ is even, then $ p_c^{s}(d)\leq 1-(0,32)^{2/d}$; 2) If $d$ is divisible by 3, then $p_c^{s}(d)\leq 1-(0,5)^{3/d}$; 3) For any dimension $d\geq 2$, we have that $p_c^{s}(d+1)\leq g(d)$, where $g(d)$ is the unique solution in $(0,1)$ of \[ p = p^{\,s}_c(d) \left[p + (1-p)^{2d/(2d-1)}\right]. \] \end{theorem} \begin{theorem}\label{theo:siteO} Consider oriented site Bernoulli percolation on $\mathbb{Z}^d$. 1) If $d$ is even, then $ \vv{p}_c^{\,s}(d) \leq 1-(0,25)^{2/d}$; 2) For any dimension $d \geq 2$, we have that $\vv{p}_c^{\,s}(d+1) \leq h(d)$, where $h(d)$ is the unique solution in $(0,1)$ of \[p = \vv{p}^{\,s}_c(d) \left[p + (1-p)^{(d+1)/d}\right]. \] \end{theorem} \section{The Dynamical Couplings}\label{sec:couplings} Although we are only interested in homogeneous percolation, some of the tools we will use are related to couplings between {\it anisotropic} bond percolation models. We now define the anisotropic (or inhomogeneous) non-oriented and oriented bond percolation models. For each $i=1, \dots, d$, let $E_i = \{ \langle x, x + e_i \rangle: x \in \mathbb{Z}^d\}$ be the set of edges parallel to $e_i$. Given $p_1, \dots, p_d \in [0,1]$, we consider a family of independent random variables $\{X_e\}_{e \in E(\mathbb{Z}^d)}$, but now, for each $e \in E_i$, $X_e$ has Ber($p_i$) distribution, $i=1, \dots, n$. The open cluster of the origin $\mathcal{C}^b_0$ and the oriented open cluster of the origin $\vv{\mathcal{C}}^b_0$ are defined analogously. The probabilities that $|\mathcal{C}^b_0|$, $|\vv{\mathcal{C}}^b_0|$ are infinite, will be denoted by $\theta_d(p_1, \dots, p_d)$ and $\vv{\theta}_d (p_1, \dots, p_d)$ respectively. The first coupling is the content of Proposition 1 in \cite{GPS2}: \begin{proposition} \label{prop:coupling} Consider inhomogeneous non-oriented Bernoulli bond percolation in $\mathbb{Z}^d$. Let $p_1, \dots, p_{d+1} \in [0,1]$ and let $\Tilde{p}_d \in [0,1]$ be such that \[(1-\Tilde{p}_d) = (1-p_d)(1-p_{d+1}).\] Then, $\theta_{d+1}(p_1, \dots,p_d, p_{d+1}) \geq \theta_{d}(p_1, \dots, p_{d-1}, \Tilde{p}_{d}) .$ \end{proposition} Now we define bond and site percolation models in the triangular lattice $\mathbf{T}$. This lattice is simply $\mathbb{Z}^2$ with extra edges of the form $\langle v,v+(1,1)\rangle$. That is, $\mathbf{T} = (V_{\mathbf{T}}, E_{\mathbf{T}})$ where $V_{\mathbf{T}} = \mathbb{Z}^2$ and $E_{\mathbf{T}}$ is the set of non ordered pairs $\{\langle v, u\rangle : v-u = (1,0), (0,1) \text{ or } (1,1) \}$. We will denote by $\theta^b_{\mathbf{T}}$ and $\theta^s_{\mathbf{T}}$ the corresponding percolation probability for bond and site models, respectively. We will consider inhomogeneous bond percolation on $\mathbf{T}$, where the corresponding parameters $(p_1,p_2,p_3)$ will refer to edges of the form $\langle v,v+(1,0)\rangle$, $\langle v,v+(0,1)\rangle$ and $\langle v,v+(1,1)\rangle$, respectively. In the next proposition, we construct a monotonic coupling between inhomogeneous bond percolation on the triangular lattice $\mathbf{T}$ with parameters $(p_1,p_2,p_3)$ and on $\mathbb{Z}^3$ with the same parameters. \begin{proposition}\label{prop:triangular} Let $(p_1,p_2,p_3)\in [0,1]$ and consider two inhomogeneous bond Bernoulli percolation processes on the triangular lattice and on $\mathbb{Z}^3$, both with parameters $(p_1,p_2,p_3)$. Then \[\theta^b_{\mathbf{T}}(p_1,p_2,p_3)\leq\theta^b_3(p_1,p_2,p_3).\] \end{proposition} \begin{proof} We will construct a dynamic coupling between the percolation process on $\mathbb{Z}^{3}$ with parameters $(p_1,p_2, p_3)$ and an infection process over $\mathbf{T}$. We will do it in such a way that the law of infected sites in $\mathbf{T}$ is the same as the law of the open cluster of the origin for anisotropic percolation on $\mathbb{Z}^3$ and also that, if the infection process survives, the open cluster of the origin of the process in $\mathbb{Z}^3$ must be infinite. The coupling will be built based on a susceptible-infected strategy described as follows. First, we declare the origin of $\mathbf{T}$ as the \textit{initial} infected component. Next, at each time-step, we possibly grow the infected component and associate each new vertex $v$ of the infected component in $\mathbf{T}$ to a vertex $x(v)$ in the open cluster of the origin in $\mathbb{Z}^3$. More precisely, consider a vertex $v$ in the infected component of $\mathbf{T}$ and a neighbor $v+u$ out of the infected component. The vertex $v$, in the infected component in $\mathbf{T}$, will be associated with some vertex $x(v)$ in the open cluster of the origin in $\mathbb{Z}^3$. According to $u = \pm(1,0) , \pm(0,1)$ or $\pm(1,1)$ we denote $\tau(u) = \pm(1,0,0)$, $\tau(u) = \pm(0,1,0)$ or $\tau(u) = \pm(0,0,1)$ respectively. If $\langle x(v), x(v)+\tau(u) \rangle$, is open, we infect $v+u$. (and write $x(v+u) = x(v) + \tau(u)$). Let's define the sequence of sets $(I_n, x(I_n), R_n, S_n)_{n \geq 0}$. Here, $I_n$ represents the \textit{infected vertices} in $\mathbf{T}$ and $x(I_n)$ represents the vertices in $\mathbb{Z}^{3}$ associated with the infected vertices. $R_n$ represents the \textit{removed edges} of $\mathbf{T}$. Finally, given $I_n$, $x(I_n)$ and $R_n$, the \textit{susceptible edges set} is given by \begin{equation*} S_{n} := \{ \langle v, u \rangle : v \in I_{n} \: \mbox{and} \: u \notin I_{n} \} \cap R_{n}^C. \end{equation*} At time $n=0$, we set \begin{itemize} \item $I_0 = \{0\} \subset V_{\mathbf{T}}$; \item $R_0 = \emptyset \subset E_{\mathbf{T}}$; \item $x(0) = 0 \in \mathbb{Z}^{3}$; \item $S_0 := \{ \langle v, u \rangle : v \in I_{0} \: \mbox{and} \: u \notin I_{0} \} \cap R_{0}^C \:\: = \{ \langle 0, \pm (1,0)\rangle, \langle 0,\pm (1,0)\rangle ,\langle 0,\pm (1,1)\rangle \}$. \end{itemize} This means that, at time $n=0$, only the vertex $0$ is infected, and it can potentially infect any of its neighbours, so all edges containing the origin are susceptible. After that, in each step, an infected vertex tries to infect a non-infected vertex through a susceptible edge (if the latter exists). From now on, we choose an arbitrary, but fixed, ordering of the edges in $\mathbf{T}$. Suppose that $I_n, x(I_n), R_n$ and $S_n$ are already defined. If there is no susceptible edge then the process stops. More specifically, if $S_n = \emptyset$, then for all $k \geq 1$, \begin{itemize} \item $I_{n+k} = I_n$; \item $R_{n+k} = R_n$; \item $S_{n+k} = S_n$. \end{itemize} Otherwise, if there exists some susceptible edge, then the infected vertex in the smallest (in the previously fixed ordering) such edge tries to infect its non-infected neighbour. Let us write it in symbols. Suppose $S_n \neq \emptyset$ and let $g_n$ be the smallest edge in $S_n$. Since $g_n \in S_n$, it has to be equal to some $\langle v, v+u_n \rangle$, where $v \in I_n$, $v+u_n \notin I_n$, and $u_n \in \{ \pm (0,1), \pm (1,0), \pm (1,1)\}$. We set $\tau(\pm(1,0))=\pm (1,0,0)$, $\tau(\pm(0,1))=\pm (0,1,0)$ and $\tau(\pm(1,1))=\pm (0,0,1)$ . Then, $v$ \textit{infects} $v+u_n$ if $\langle x(v), x(v)+\tau (u_n) \rangle$ is open in $\mathbb{Z}^{3}$. More precisely, if $\langle x(v), x(v)+\tau (u_n) \rangle$ is open in $\mathbb{Z}^{3}$ then we write \[ I_{n+1} := I_n \cup \{v+u_n\},\] and define \[ x(v+u_n) := x(v) + \tau(u_n). \] Otherwise, if $\langle x(v), x(v)+ \tau(u_n) \rangle$ is closed in $\mathbb{Z}^{3}$, we set $I_{n+1}:= I_n$. Now that we have explored $g_n$, we \textit{remove} it and write \[ R_{n+1} := R_n \cup \{ g_n \}.\] Next, to conclude our induction step, we set \begin{equation*} S_{n+1} := \{ \langle v, u \rangle : v \in I_{n+1} \: \mbox{and} \: u \notin I_{n+1} \} \cap R_{n+1}^C. \end{equation*} Observe that the function $x: \cup_n I_n \longrightarrow \mathbb{Z}^{3}$ is injective. In fact, if $v= (v_1, v_2) \in I_n$, then by construction, we have that $x(v) = (x_1,x_2,x_3)$ satisfies \[ v_1=x_1+x_3 \mbox{ and } v_2=x_2+x_3.\] Now, observe that the image of $x$ is contained in the open cluster of the origin $\mathcal{C}^b_0$ of $\mathbb{Z}^{3}$. Since $x$ is injective, $|\cup_n I_n| \leq |\mathcal{C}^b_0|$. Also, note that $\cup_{n} I_n$ has the same law as $\mathcal{C}^{\mathbf{T}}_0$, where $\mathcal{C}^{\mathbf{T}}_0$ is the open cluster of the origin in $\mathbf{T}$ with parameters $p_1, p_2,p_3$. Therefore, \begin{equation*} \theta^b_{\mathbf{T}}(p_1,p_2,p_3) \leq \theta_{3}^b (p_1,p_2,p_3), \end{equation*} and the proof of Proposition \ref{prop:triangular} follows. \end{proof} To conclude this section, we construct two couplings that will be used to prove Item 3 of Theorem \ref{theo:bondO}, along with its site version, which will be used in the proofs of Item 3 of Theorem \ref{theo:siteNO} and Item 2 of Theorem \ref{theo:siteO}. These couplings are reminiscent of the coupling in \cite{GSS} and give an upper bound for the critical probability for any percolation model in $\mathbb{Z}^{d+1}$ as a function of the corresponding critical probability in $\mathbb{Z}^d$. Since our goal is to obtain better upper bounds, we need to improve the coupling. \begin{proposition}\label{prop:crossover} Consider oriented bond Bernoulli percolation on $\mathbb{Z}^{d+1}$ and suppose that \begin{equation} \label{eq:crossbond} p > \vv{p}^{\,b}_c(d) \left[p + (1-p)^{(d+1)/d}\right]. \end{equation} Then, $\vv{\theta}^{\,b}_{d+1}(p)>0$. \end{proposition} \begin{proof} Let ${\mathcal{E}} = \{e_1, \dots, e_d\}$ denote the set of positive unit vectors of $\mathbb{Z}^d$. To avoid ambiguities, we denote the set of positive unit vectors of $\mathbb{Z}^{d+1}$ by $\{ u_1, \dots, u_{d+1}\}$. Consider the multigraph $\mathbb{Z}^{d+1}_{\mathcal{E}}$ defined as follows: the set of vertices is $\mathbb{Z}^{d+1}$ and the set of edges is given by $E(\mathbb{Z}^{d+1}_{\mathcal{E}}):= \left( \cup_{i=1}^d E_i \right) \cup E_{\mathcal{E}}$ where $E_i = \{ \langle x, x + u_i \rangle: x \in \mathbb{Z}^{d+1}\}$, $i=1, \dots, d$, and we define $E_{\mathcal{E}} := \{\langle v, v+u_{d+1}\rangle_e : v \in \mathbb{Z}^{d+1} , e \in {\mathcal{E}}\}$. In words, each edge of $\mathbb{Z}^{d+1}$ parallel to $u_{d+1}$ is partitioned into another $|{\mathcal{E}}|$ edges indexed by ${\mathcal{E}}$ in $\mathbb{Z}^{d+1}_{\mathcal{E}}$, while edges parallel to all other directions remain unmodified. We prove this proposition in two steps. First, we will define a multigraph $\mathbb{Z}^{d+1}_{\mathcal{E}}$ and show that, with certain parameters, the model on $\mathbb{Z}^{d+1}_{\mathcal{E}}$ is equivalent to the homogeneous model on $\mathbb{Z}^{d+1}$ with parameter $p$. To conclude, we will show that if $p$ satisfies Inequality \eqref{eq:crossbond}, then this new model on $\mathbb{Z}^{d+1}_{\mathcal{E}}$ dominates a supercritical model on $\mathbb{Z}^{d}$. Consider now inhomogeneous oriented bond Bernoulli percolation on $\mathbb{Z}^{d+1}_{\mathcal{E}}$, where edges in $\cup_{i=1}^d E_i$ are open with probability $p$ and edges in $E_{\mathcal{E}}$ are open with probability $q$, where $q$ is such that $(1-p) = (1-q)^{|{\mathcal{E}}|}$. Clearly, the distribution of the open cluster in this model is the same as in the homogeneous model on $\mathbb{Z}^{d+1}$ with parameter $p$. We will construct a coupling showing that the model on $\mathbb{Z}_{\mathcal{E}}^{d+1}$ with parameters $p$ and $q$ as above, dominates the homogeneous model on $\mathbb{Z}^d$ with parameter $p/[1 - (1-p)q]$. First, for each $e_i \in \mathcal{E}$, let $ \sigma(e_i):= u_i \in \mathbb{Z}^{d+1}$. Then, for each $v \in \mathbb{Z}^{d+1}$ and each $e \in {\mathcal{E}}$, let $A(v,e)$ be the event where either the edge $\langle v, v+\sigma(e) \rangle$ is open or, for some $k \geq 1$, \begin{itemize} \item the edges $\langle v+ iu_{d+1}, v + (i+1)u_{d+1} \rangle_e$, $i = 0, \dots, k-1$, are open and \item the edges $\langle v + iu_{d+1}, v + iu_{d+1} + \sigma(e) \rangle$, $i = 0, \dots, k-1$ are closed, and \item $\langle v + ku_{d+1}, v + ku_{d+1} + \sigma(e) \rangle$ is open. \end{itemize} In the event $A(v,e)$, we define $u(v,e) = v+\sigma(e)$ if $\langle v, v+\sigma(e) \rangle$ is open, or $u(v,e) = v + k u_{d+1} + \sigma(e)$ if $k \geq 1$ is such that the above three conditions are met. Observe that \begin{eqnarray*} \mathbb{P}(A(v,e)) &=& p\sum_{i = 0}^{\infty} \left[q(1-p)\right]^i = \frac{p}{1 - (1-p)q}\\ &=& \frac{p}{1-(1-p)(1-(1-p)^{1/d})} \\ &=& \frac{p}{ p + (1-p)^{(d+1)/d} }. \end{eqnarray*} Similarly to what was done in Proposition \ref{prop:triangular}, we now build the sequence of sets $(I_n, x(I_n), R_n, S_n)_{n \geq 0}$. For $n =0$, we set \begin{itemize} \item $I_0 = \{0\} \subset \mathbb{Z}^d$; \item $R_0 = \emptyset \subset E(\mathbb{Z}^d)$; \item $x(0) = 0 \in \mathbb{Z}_{\mathcal{E}}^{d+1}.$ \end{itemize} Suppose that, for some $n \geq 0$, $I_n, x(I_n)$ and $R_n$ are already defined. Then $S_n \subset E(\mathbb{Z}^d)$ is given by \begin{equation* S_{n} := \{ \langle v, u \rangle : v \in I_{n} \: \mbox{and} \: u \notin I_{n} \} \cap R_{n}^C. \end{equation*} If $S_n = \emptyset$, our sequence becomes constant, that is, \begin{equation}\label{eq:stop} (I_{n+k}, x(I_{n+k}), R_{n+k}, S_{n+k}) = (I_n, x(I_n), R_n, S_n), \quad \forall k \geq 1. \end{equation} Otherwise, let $g_n = \langle v , v + e \rangle$ be the smallest (according to a prefixed ordering, as in Proposition~\ref{prop:triangular}) edge of $S_n$, where $v \in I_n$, $e \in {\mathcal{E}}$, and $v+e \notin I_n$. We set $R_{n+1} = R_n \cup \{g_n\}$. We also set \begin{equation*} I_{n+1} = \begin{cases} I_n \cup \{v+e\}, &\text{if } A(x(v),e) \text{ occurs};\\ I_n, & \text{otherwise}. \end{cases} \end{equation*} In case $A(x(v),e)$ occurs, we set $x(v+e) = u(x(v),e)$. Once our sequence $(I_n, x(I_n), R_n, S_n)_ {n \geq 0}$ is built, note that by construction, the function $x: \cup_n I_n \longrightarrow \mathbb{Z}^{d+1}_{\mathcal{E}}$ is injective. In fact, for each $v \in \cup_n I_n$, the projection of $x(v)$ into $\mathbb{Z}^d$ is equal to $v$. The conclusion of the proof follows in a similar way to that of Proposition \ref{prop:triangular}. \end{proof} \begin{proposition}\label{prop:crossoversite} Consider non-oriented site Bernoulli percolation on $\mathbb{Z}^{d+1}$ and suppose that \begin{equation}\label{eq:cross1} p > p^s_c(d) \left[p + (1-p)^{2d/(2d-1)}\right]. \end{equation} Then $\theta^s_{d+1}(p)>0$. \end{proposition} \begin{proof} The strategy of the proof is similar to the previous proposition. As before, let $\{e_1, \dots, e_d\}$ denote the set of positive unit vectors of $\mathbb{Z}^d$ and let $\{ u_1, \dots, u_{d+1}\}$ denote the set of positive unit vectors of $\mathbb{Z}^{d+1}$. We will consider a graph where each vertex of $\mathbb{Z}^{d+1}$ will be partitioned into another $2d-1$ vertices. For each $v \in \mathbb{Z}^{d+1}$, we define the set of split vertices $V_v := \{v^{(1)}, \dots, v^{(2d-1)}\}$. Let $\mathbb{Z}^{d+1}_V$ be the graph with vertex set $\cup_{v \in \mathbb{Z}^{d+1}} V_v$ and edge set \begin{equation*} E(\mathbb{Z}^{d+1}_V) := \left\{ \langle x,y \rangle ~:~ x \in V_v \text{ and } y \in V_u, \text{ for some } \langle v,u \rangle \in E(\mathbb{Z}^{d+1}) \right\}. \end{equation*} Given $p$ satisfying \eqref{eq:cross1}, let $q$ be such that $(1-p) = (1-q)^{2d-1}$. We will consider the non-oriented site Bernoulli percolation model on $\mathbb{Z}^{d+1}_V$ with parameter $q$. Note that the distribution of the open cluster of the origin in this new model is the same as in the model on $\mathbb{Z}^{d+1}$ with parameter $p$. To build the coupling between $\mathbb{Z}^{d+1}_V$ with parameter $q$ and $\mathbb{Z}^d$ with parameter $p/[1 - (1-p)q]$, we again construct a sequence of sets $(I_n, x(I_n), R_n, S_n)_{n \geq 0}$. Since we are considering a site model, for each $n \geq 0$, the sets $R_n$ and $S_n$ will be, respectively, the sets of {\it removed} and {\it susceptible} vertices (instead of edges) at step $n$. We start with two infected vertices (otherwise the origin should have been split into $2d$ vertices, instead of $2d-1$). For $n = 0$, we set \begin{itemize} \item $I_0 = \{0, e_1\} \subset \mathbb{Z}^d$; \item $R_0 = \emptyset \subset \mathbb{Z}^d$; \item $x(0) = 0^{(1)} \in \mathbb{Z}_V^{d+1} \text{ and } x(e_1) = u_1^{(1)} \in \mathbb{Z}_V^{d+1}. $ \end{itemize} Inductively, if for some $n \geq 0$, the sets $I_n, x(I_n)$ and $R_n$ are already defined, we set \begin{equation}\label{eq:SnSITE} S_n = \{u \in \mathbb{Z}^d : \langle v,u \rangle \in E(\mathbb{Z}^d) \text{ for some } v \in I_n\} \cap (I_n \cup R_n)^C. \end{equation} If $S_n = \emptyset$, the sequence becomes constant, as in \eqref{eq:stop}. Otherwise, let $a_n$ be the smallest (in a preset ordering) vertex of $S_n$. We can write $a_n = v+ e$, such that $v \in I_n$ and $e \in \{\pm e_1, \dots, \pm e_d\}$. In this case, let $j_n$ denote the number of susceptible neighbors of $v$ (including $a_n$), that is \begin{equation*} j_n := \left\vert \{ u \in S_n : \langle v, u \rangle \in E(\mathbb{Z}^d) \} \right\vert . \end{equation*} Since we start with two infected vertices, necessarily $|S_n| \neq 2d$, and then $1 \leq j_n \leq 2d-1$. For each $i \in \{1, \dots, d\}$, we set the notation $\sigma(\pm e_i) = \pm u_i$. Consider the event $A_n$ where, either, for some $\ell \in \{1, \dots, 2d-1 \}$, the vertex $(x(v)+\sigma(e))^{(\ell)}$ is open or, for some $k \geq 1$ and $\ell \in \{1, \dots, 2d-1\}$, \begin{itemize} \item the vertices $\left( x(v) + iu_{d+1} \right)^{(j_n)}$, $i = 1, \dots, k$, are open, and \item the vertices $\left( x(v) + i u_{d+1} + \sigma(e) \right)^{(j)}$, are closed $\forall j \in \{1, \dots, 2d-1 \}$, $i = 0, \dots. k-1$, and \item the vertex $\left( x(v) + ku_{d+1} + \sigma(e) \right)^{(\ell)}$ is open. \end{itemize} Note that $A_n$ has probability \begin{eqnarray*} \mathbb{P}(A_n) &=& \left(1 - (1-q)^{2d-1}\right) \sum_{i=0}^{\infty} \left[q(1-q)^{2d-1}\right]^i \\ &=& \frac{p}{1 - (1-p)q}\\ &=& \frac{p}{1-(1-p)\left(1-(1-p)^{1/(2d-1)}\right)} \\ &=& \frac{p}{ p + (1-p)^{2d/2d-1} }. \end{eqnarray*} To conclude our induction step, we set \begin{itemize} \item $R_{n+1} = \begin{cases} R_n \cup \{v+e\}, &\text{if } A_n \text{ does not occur};\\ R_n, & \text{otherwise}; \end{cases}$ \item $I_{n+1} = \begin{cases} I_n \cup \{v+e\}, &\text{if } A_n \text{ occurs};\\ I_n, & \text{otherwise}. \end{cases}$ \end{itemize} In the event $A_n$, if for some $\ell \in \{1, \dots, 2d-1 \}$ the vertex $(x(v)+\sigma(e))^{(\ell)}$ is open, we set $x(v+e) =( x(v)+\sigma(e) )^{(\ell)}$. Otherwise, let $k \geq 1$ and $\ell \in \{1, \dots, 2d-1\}$ be such that the above three conditions are satisfied, in this case, we set $x(v+e) =( x(v)+ k u_{d+1} + \sigma(e) )^{(\ell)}$. By construction, for each $v \in \cup_n I_n$, the projection of $x(v)$ onto $\mathbb{Z}^d$ is equal to $v$. Therefore $x: \mathbb{Z}^d \longrightarrow \mathbb{Z}^{d+1}_V$ is also injective. It follows that the site percolation model in $\mathbb{Z}^{d+1}_V$ (with $e_1^{(1)}$ declared open), dominates the infection process $(I_n)_{n \geq 0}$. Since $|\cup_n I_n|$ has the same distribution as the size of the open cluster of $\{0, e_1\}$ in a supercritical percolation model on $\mathbb{Z}^d$, the proof follows. \end{proof} \section{Proof of Theorems}\label{sec:proofs} In this section, we prove the theorems using the couplings of the last section \subsection{Proof of Theorem \ref{theo:bounds}} \begin{proof} Given Propositions \ref{prop:coupling} and \ref{prop:triangular}, the proof of Theorem \ref{theo:bounds} is quite straightforward. Recall that for bond percolation on the triangular lattice with parameters $(p_1,p_2,p_3)$, assuming $p_1,p_2,p_3<1$, we have (see Theorem 11.116 in \cite{Grim}) \begin{equation*} \theta_{\mathbf{T}}(p_1,p_2,p_3)>0 \Leftrightarrow p_1+p_2+p_3-p_1p_2p_3>1. \end{equation*} Consider now the partition \begin{equation*} d=\left\lfloor\frac{d}{3}\right\rfloor+\left\lfloor\frac{d+1}{3}\right\rfloor+\left\lfloor\frac{d+2}{3}\right\rfloor. \end{equation*} Let $p>p^{\ast}(d)$ and set, for $i=1,2,3$, \begin{equation*} p_i=1-(1-p)^{\lfloor \frac{d+i-1}{3}\rfloor}. \end{equation*} Then, iteratively applying Proposition \ref{prop:coupling} and finally Proposition \ref{prop:triangular}, we have \begin{equation*} \theta_d(p)\geq\theta_3(p_1,p_2,p_3)\geq \theta_{\mathbf{T}}(p_1,p_2,p_3)>0, \end{equation*} which concludes the proof. \end{proof} \subsection{Proof of Theorem \ref{theo:bondO}} \begin{proof} First, we remark that Proposition \ref{prop:coupling} holds {\it mutatis mutandis} for inhomogeneous oriented bond Bernoulli percolation. As a consequence, we have the following corollary which we state without proof. \begin{lemma}\label{lem:Cor} Let $k \in \mathbb{N}$ be such that $d$ is divisible by $k$. Given $p \in [0,1]$, let $\tilde{p}$ be such that \[(1-\tilde{p}) = (1-p)^{d/k}.\] Then, $\vv{\theta}^{\, b}_d(p) \geq \vv{\theta}_k^{\, b}(\tilde{p})$. \end{lemma} Taking $k=2$ in Lemma \ref{lem:Cor} and using Liggett's upper bound $\vv{p}_c^{\, b}(2) \leq 2/3$ (see \cite{L}), the first item of Theorem \ref{theo:bondO} follows. The third item is equivalent to Proposition \ref{prop:crossover}. Finally, the second item is implicitly proved in \cite{GPS}. There, the authors defined a quantity $\lambda(1/d, \dots, 1/d)$ and showed that (see Equation (3.2)), if $p \in [0,1]$ satisfies \[\phi(d) := \lambda(1/d, \dots, 1/d) \leq dp -1,\] then $\vv{\theta}^{\, b}_d(p) > 0$. In particular, \[ \vv{p}_c^{\, b}(d) \leq \frac{1+\phi(d)}{d}.\] The conclusion follows from the estimate given in Subsection 3.1 of \cite{GPS} (see the equation above (3.5)), that is \[\phi(d) \leq \frac{1}{d}+\frac{8}{d^2}+\frac{d^{3/2}}{(\sqrt{2\pi})^{d-1}}\left[\frac{d-1}{d-3}\right]e^{\frac{1}{12d}}. \] \begin{comment} Let $\{S_n^1\}_{n \geq 0}$ and $\{S_n^2\}_{n \geq 0}$ be two independent, oriented, simple and symmetrical random walks on $\mathbb{Z}^d$, starting on the origin. In the paper \cite{GPS} was shown that, if $p \in [0,1]$ is such that \[\phi(d) := \sum_{n=1}^{\infty} \mathbb{P}\left( S_n^1 = S_n^2 \right) \leq dp -1,\] then $\vv{\theta}^{\, b}_d(p) > 0$. In particular, \[ \vv{p}_c^{\, b}(d) \leq \frac{1+\phi(d)}{d}.\] The conclusion follows from the estimate given in Subsection 3.1 of \cite{GPS} (see the equation above 3.5), \[\phi(d) \leq \frac{1}{d}+\frac{8}{d^2}+\frac{d^{3/2}}{(\sqrt{2\pi})^{d-1}}\left[\frac{d-1}{d-3}\right]e^{\frac{1}{12d}}. \] \end{comment} \end{proof} \subsection{Proof of Theorem \ref{theo:siteNO}} Inhomogeneous percolation is not well defined for site models, and therefore, we are not able to formulate a version of Proposition \ref{prop:coupling} for site percolation. But note that Lemma \ref{lem:Cor} only involves homogeneous models. In the following, we give its site version. \begin{lemma}\label{lem:ksite} Consider non-oriented site Bernoulli percolation on $\mathbb{Z}^{d}$. Let $k \in \mathbb{N}$ be such that $d$ is divisible by $k$. Given $p \in [0,1]$, let $\tilde{p}$ be such that \[(1-\tilde{p}) = (1-p)^{d/k}.\] Then, we have ${\theta}^s_d(p) \geq {\theta}^s_k(\tilde{p})$. \end{lemma} \begin{proof} The lemma follows by a coupling between non-oriented site Bernoulli percolation on $\mathbb{Z}^d$, with parameter $p$, and on $\mathbb{Z}^{k}$, with parameter $\tilde{p}$. To establish this coupling, we again construct a sequence of vertex sets $(I_n, x(I_n), R_n, S_n)_{n \geq 0}$. First, we recall that $\mathcal{E} = \{e_1, \dots, e_d\}$ is the set of positive unit vectors of $\mathbb{Z}^d$ and let $(D_{u_1}, \dots, D_{u_k})$ be a uniform partition of $\mathcal{E}$ into $k$ subsets indexed by the set $\{u_1, \dots, u_k \}\subset\mathbb{Z}^k$ of positive unit vectors of $\mathbb{Z}^k$. For each $u \in \{u_1, \dots, u_k \} $, let $D_{-u} = -D_u$. Consider the non-oriented site Bernoulli percolation model on $\mathbb{Z}^d$, with parameter $p$. For each $v \in \mathbb{Z}^d$ and each $u \in \{\pm u_1, \dots, \pm u_k\}$, we define the following event \begin{equation*} B(v,u) := \{ v+e \text{ is open, for some } e \in D_u \}. \end{equation*} If the event $B(v,u)$ occurs, we set $e(v,u) = v+e$, where $e$ is an open vertex that guarantee the occurrence of $B(v,u)$. Note that $\mathbb{P}(B(v,u)) = \tilde{p}.$ For $n = 0$, we define \begin{itemize} \item $I_0 = \{0\} \subset \mathbb{Z}^k$; \item $R_0 = \emptyset \subset \mathbb{Z}^k$; \item $x(0) = 0 \in \mathbb{Z}^d$. \end{itemize} If for some $n \geq 0$, the sets $I_n, x(I_n)$ and $R_n$ are already defined, let $S_n$ be as given in \eqref{eq:SnSITE}. If $S_n = \emptyset$, the sequence becomes constant as in \eqref{eq:stop}. Otherwise, let $v_n$ be the smallest vertex of $S_n$ (according to a preset ordering of $\mathbb{Z}^k$). In this case, we write $v_n = v+u \notin I_n$, where $v \in I_n$ and $u \in \{\pm u_1, \dots, \pm u_k\}$. To conclude the induction step, we define \begin{itemize} \item $R_{n+1} = \begin{cases} R_n \cup \{v+u\}, &\text{if } B(x(v),u) \text{ does not occurs};\\ R_n, & \text{otherwise}; \end{cases}$ \item $I_{n+1} = \begin{cases} I_n \cup \{v+u\}, &\text{if } B(x(v),u) \text{ occurs};\\ I_n, & \text{otherwise}. \end{cases}$ \end{itemize} If $B(x(v),u)$ occurs, we define $x(v+u) = e(x(v),u)$. Note that, by construction, the function $x: \cup_n I_n \longrightarrow \mathbb{Z}^d$ is such that, for each $v = v_{u_1}u_1+ \cdots + v_{u_k}u_k \in \cup_n I_n$, writing $x(v) = x_{e_1} e_1 + \cdots + x_{e_d} e_d$, we have \begin{equation*} \sum_{e \in D_u} x_e = v_u, \quad \forall u \in \{u_1, \dots, u_k\}. \end{equation*} Therefore, $x: \cup_n I_n \longrightarrow \mathbb{Z}^d$ is injective. The conclusion of the lemma follows as in previous couplings. \end{proof} With Lemma \ref{lem:ksite} we are now able to conclude the goal of this section. \begin{proof}[Proof of Theorem \ref{theo:siteNO}] The first item of Theorem \ref{theo:siteNO} follows by taking $k=2$ in Lemma \ref{lem:ksite} toghether with Wierman's upper bound $p_c^s(2)\leq 0.68$ (see \cite{W2}). Taking $k=3$ in Lemma \ref{lem:ksite} and using the upper bound $p_c^s(3) < 1/2$ of Campanino-Russo (see \cite{CR}), the second item follows. The third item follows directly by Proposition \ref{prop:crossoversite}. \end{proof} \subsection{Proof of Theorem \ref{theo:siteO}} \begin{proof} First, we remark that Lemma \ref{lem:ksite} and Proposition \ref{prop:crossoversite} hold {\it mutatis mutandis} for oriented site Bernoulli percolation. Therefore to conclude the proof of Theorem \ref{theo:siteO}, the last ingredient is Liggett's upper bound $\vv{p}^{\,s}_c(2) \leq 3/4$ (see \cite{L}). \end{proof} \section{Explicit bounds}\label{sec:comments} In this section, we present a table with upper bounds for bond and site Bernoulli Percolation in $\mathbb{Z}^ d$, up to $d=9$, in both the oriented and the non-oriented cases. In each column we give a numerical upper bound rounded to four decimals for the critical probability in the head of the column. \begin{center} \begin{tabular}{|c|c|c|c|c|} \hline \textbf{Dimension} & $p_c^b(d)$ & $\vv{p_c}^b(d)$ & $p_c^s(d)$ & $\vv{p_c}^s(d)$\\ \hline 3 & 0,3473 &0,5680 &0,5000 &0,6422 \\ \hline 4 & 0,2788 &0,4227 &0,4344 &0,5000 \\ \hline 5 & 0,2284 & 0,3926 &0,4156& 0,4615 \\ \hline 6 & 0,1922 & 0,2734 & 0,2929 & 0,3701 \\ \hline 7 & 0,1682 & 0,2028 & 0,2866 & 0,3533\\ \hline 8 & 0,1486 &0,1627 & 0,2479 &0,2929\\ \hline 9 & 0,1326 &0,1371 &0,2063 &0,2844\\ \hline \end{tabular} \end{center} \vspace{0.5cm} To obtain each bound, we used the following: \vspace{0.5cm} \textbf{Non-oriented bond percolation.} All bounds were obtained from Theorem \ref{theo:bounds}. \textbf{Oriented bond percolation} All bounds came from Theorem \ref{theo:bondO}. For $d=4$ we used Item 1), for $d=5$ we used the upper bound for $d= 4$, and Item 3). For $d\geq 6$ we used Item 2). \textbf{Non-oriented site percolation.} The bound for $d=3$ follows from Campanino-Russo (see \cite{CR}). The others follow from Theorem \ref{theo:siteNO}. For $d=4$ and $d=8$ we used Item 1). For $d=6$ and $d=9$ we used Item 2). For $d=5$ and $d=7$ we used Item 3) along with the upper bounds obtained for $d=4$ and $d=6$. \textbf{Oriented site percolation.} All the bounds came from Theorem \ref{theo:siteO}. For even dimensions the bounds follow from Item 1), and for odd dimensions the bounds follow from Item 2) using the upper bounds for even dimensions. \begin{comment}\begin{tabular}{|c|c|c|c|c|} \hline \textbf{Dimension} & $p_c^b(d)$ & $\vv{p_c}^b(d)$ & $p_c^s(d)$ & $\vv{p_c}^s(d)$\\ \hline 3 & 0,3473 &0,5680 &0,5000 &0,6422 \\ \hline 4 & 0,2787 &0,4227 &0,4344 &0,5528\\ \hline 5 & 0,2283 & 0,3926 &0,4156& 0,5086 \\ \hline 6 & 0,1921 & 0,2733 & 0,2929 & 0,3701 \\ \hline 7 & 0,1681 & 0,2028 & 0,2796 & 0,3533\\ \hline 8 & 0,1485 &0,1627 & 0,2479 &0,2929\\ \hline 9 & 0,1326 &0,1370 &0,2063 &0,2843\\ \hline \end{tabular} \end{comment} \vspace{1.2cm} {\bf \Large Acknowledgements} The authors thank Roger Silva for valuable comments on the manuscript. P.A. Gomes has been supported by São Paulo Research Foundation (FAPESP), grant 2020/02636-3 and grant 2017/10555-0. R. Sanchis has been partially supported by Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq), CAPES and by FAPEMIG (PPM 00600/16). \thebibliography{} \bibitem{BR} Bollob\'as, B., and Riordan, O. {\it Percolation}, Cambridge University Press (2006). \bibitem{CR} Campanino, M., and Russo, L. {\it An upper bound on the critical percolation probability for the three-dimensional cubic lattice}, Ann. Prob. 13 (2), 478-491, (1985). \bibitem{CD} Cox, J.T., and Durrett, R. {\it Oriented percolation in dimensions $d \geq 4$: bounds and asymptotic formulas.} Math Proc. Camb. Phil. Soc. Vol. 93(1), pp.151-162, (1983). \bibitem{GPS} Gomes, P.A., Pereira, A., and Sanchis, R. {\it Anisotropic oriented percolation in high dimensions}. ALEA Lat. Am. J. Probab. Math. Stat. 17 (1), 531-543, (2020). \bibitem{GPS2} Gomes, P.A., Pereira, A., and Sanchis, R. {\it Anisotropic percolation in high dimensions: the non-oriented case}. ArXiv:2106.09083. \bibitem{GSS} Gomes P.A., Sanchis, R. and Silva, R.W.C. {\it A note on the dimensional crossover critical exponent.} Lett. Math. Phys. 110 (12), 3427–3434, (2020). \bibitem{Grim} Grimmett, G.R. {\it Percolation.} Second edition. Springer-Verlag, Berlin, (1999). \bibitem{K3} Kesten, H. {\it The Critical Probability of Bond Percolation on the Square Lattice Equals 1/2}, Commun. Math. Phys. 74 (1), 41-59, (1980). \bibitem{L} Liggett, T.M. {\it Survival of discrete time growth models with applications to oriented percolation.} Ann. Appl. Probab. 5 (3), 613-636, (1995). \bibitem{NZ} Newman, R.E.J., and Ziff, R.M. {\it Fast Monte Carlo algorithm for site or bond percolation}, Phys. Rev. E, 64 (1), 016706, (2001). \bibitem{W} Wierman, J.C. {\it Bond percolation critical probability bounds for the Kagomé lattice by a substitution method}. In: Disorder in Physical Systems, Grimmett, G. and Welsh, D. J. A. (eds.), pp.349-360, (1990). \bibitem{W2} Wierman, J.C. {\it Substitution method critical probability bounds for the square lattice site percolation model}. Combin. Probab. Comput. 4 (2), 181-188 (1995). \bibitem{WP} Wierman, J.C., and Parviainen, R. {\it Ordering Bond Percolation Critical Probabilities}, UUDM Report, (2002) \url{http://www.ams.jhu.edu/~wierman/Papers/Bond-Pc-ordering.pdf} \vspace{2cm} \begin{minipage}{0.45\textwidth} Pablo Almeida Gomes\\ Universidade de S\~ao Paulo, Brasil\\ E-mail: \url{[email protected]} \end{minipage} \vspace{0.5cm} \begin{minipage}{0.45\textwidth} Alan Pereira\\ Universidade Federal de Alagoas, Brasil\\ E-mail: \url{[email protected]} \end{minipage} \vspace{0.5cm} \begin{minipage}{0.5\textwidth} R\'emy Sanchis\\ Universidade Federal de Minas Gerais, Brasil\\ E-mail: \url{[email protected]} \end{minipage} \end{document}
{ "timestamp": "2021-06-22T02:04:44", "yymm": "2106", "arxiv_id": "2106.10388", "language": "en", "url": "https://arxiv.org/abs/2106.10388", "abstract": "We consider bond and site Bernoulli Percolation in both the oriented and the non-oriented cases on $\\mathbb{Z}^d$ and obtain rigorous upper bounds for the critical points in those models for every dimension $d \\geq 3$.", "subjects": "Probability (math.PR)", "title": "Upper bounds for critical probabilities in Bernoulli Percolation models", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9843363476123224, "lm_q2_score": 0.8128673155708975, "lm_q1q2_score": 0.8001348445024903 }
https://arxiv.org/abs/2103.11443
Bipartite biregular Moore graphs
A bipartite graph $G=(V,E)$ with $V=V_1\cup V_2$ is biregular if all the vertices of a stable set $V_i$ have the same degree $r_i$ for $i=1,2$. In this paper, we give an improved new Moore bound for an infinite family of such graphs with odd diameter. This problem was introduced in 1983 by Yebra, Fiol, and Fàbrega.\\ Besides, we propose some constructions of bipartite biregular graphs with diameter $d$ and large number of vertices $N(r_1,r_2;d)$, together with their spectra. In some cases of diameters $d=3$, $4$, and $5$, the new graphs attaining the Moore bound are unique up to isomorphism.
\section{Introduction} The {\emph{degree/diameter problem}} for graphs consists in finding the largest order of a graph with prescribed degree and diameter. We call this number the \emph{Moore bound}, and a graph whose order coincides with this bound is called a {\emph{Moore graph}}. There is a lot of work related to this topic (see a survey by Miller and \v{S}ir\'a\v{n} \cite{MS16}), and also some restrictions of the original problem. One of them is related to the bipartite Moore graphs. In this case, the goal is finding regular bipartite graphs with maximum order and fixed diameter. In this paper, we study the problem, proposed by Yebra, Fiol and F\`abrega \cite{YFF83} in 1983, that consists in finding biregular bipartite Moore graphs. A bipartite graph $G=(V,E)$ with $V=V_1\cup V_2$ is biregular if, for $i=1,2$, all the vertices of a stable set $V_i$ have the same degree. We denote $[r,s;d]$-bigraph a bipartite biregular graph of degrees $r$ and $s$ and diameter $d$; and by $[r,s;d]$-bimoore graph the bipartite biregular graph of diameter $d$ that attains the Moore bound, which is denoted $M(r,s;d)$. Notice that constructing these graphs is equivalent to construct block designs, where one partite set corresponds to the points of the block design, and the other set corresponds to the blocks of the design. Moreover, each point is in a fixed number $s$ of blocks, and the size of each block is equal to $r$. The incidence graph of this block design is an $[r,s;d]$-biregular bipartite graph of diameter $d$. This type of graph is often used as an alternative to a hypergraph in modeling some interconnection networks. Actually, several applications deal with the study of bipartite graphs such that all vertices of every partite set have the same degree. For instance, in an interconnection network for a multiprocessor system, where the processing elements communicate through buses, it is useful that each processing element is connected to the same number of buses and also that each bus is connected to the same number of processing elements to have a uniform traffic through the network. These networks can be modeled by hypergraphs (see Bermond, Bond, Paoli, and Peyrat \cite{BeBoPaPe83}), where the vertices indicate the processing elements and the edges indicate the buses of the system. They can also be modeled by bipartite graphs with a set of vertices for the processing elements, another one for the buses, and edges that represent the connections between processing elements and buses since all vertices of each set have the same degree. The degree/diameter problem is strongly related to the {\em degree\/}/{\em girth problem\/} (also known as {\em the cage problem\/}) that consists in finding the smallest order of a graph with prescribed degree and girth (see the survey by Exoo and Jajcay \cite{ExooJaj08}). Note that when for an even girth of the graph, $g=2d$, the lower bound of this value coincides with the Moore bound for bipartite graphs (the largest order of a bipartite regular graph with given diameter $d$). In the bipartite biregular problem, we have the same situation. In 2019, Filipovski, Ramos-Rivera and Jajcay \cite{FilRamRivJaj19} introduced the concept of bipartite biregular Moore cages and presented lower bounds on the orders of bipartite biregular $(m,n;g)$-graphs. The bounds when $g=2d$ and $d$ even also coincide with the bounds given by Yebra, Fiol, and F\`abrega in \cite{YFF83}. Note that these bounds only coincide when the diameter is even. The cases for odd diameter and girth $g=2d$ are totally different, even for the extreme values. The contents of the paper are as follows. In the rest of this introductory section, we recall the Moore-like bound $M(r,s;d)$ derived in Yebra, Fiol, and F\`abrega \cite{YFF83} on the order of a bipartite biregular graph with degrees $r$ and $s$, and diameter $d$. Following the same problem of obtaining good bounds, in Section \ref{sec:improvedMoore} we prove that, for some cases of odd diameter, the Moore bound of \cite{YFF83} can be improved (see the new bounds in Tables \ref{tab:d=3} and \ref{tab:d=5}). In the two following sections, we basically deal with the case of even diameter because known constructions provide optimal (or very good) results. Thus, Section \ref{gen-pols} is devoted to the Moore bipartite biregular graphs associated with generalized polygons. In Section \ref{gen-cons}, we propose two general graph constructions: the subdivision graphs giving Moore bipartite biregular graphs with even diameter, and the semi-double graphs that, from a bipartite graph of any given diameter, allows us to obtain another bipartite graph with the same diameter but with a greater number of vertices. For these two constructions, we also give the spectrum of the obtained graphs. Finally, a numeric construction of bipartite biregular Moore graphs for diameter $d=3$ and degrees $r$ and $3$ is proposed in Section \ref{sec:d=3}. \subsection{Moore-like bounds} Let $G=(V,E)$, with $V=V_1\cup V_2$, be a $[r,s;d]$-bigraph, where each vertex of $V_1$ has degree $r$, and each vertex of $V_2$ has degree $s$. Note that, counting in two ways the number of edges of $G$, we have \begin{equation} \label{basic=} r N_1=s N_2, \end{equation} where $N_i=|V_i|$, for $i=1,2$. Moreover, since $G$ is bipartite with diameter $d$, from one vertex $u$ in one stable set, we must reach {\bf all} the vertices of the other set in at most $k-1$ steps. Suppose first that the diameter is even, say, $k=2m$ (for $m\ge 1$). Then, by simple counting, if $u\in V_1$, we get \begin{equation} \label{N2-even} N_2\le r+r(r-1)(s-1)+\cdots+r[(r-1)(s-1)]^{m-1}= r\frac{[(r-1)(s-1)]^{m}-1}{(r-1)(s-1)-1}, \end{equation} and, if $u\in V_2$, \begin{equation} \label{N1-even} N_1\le s\frac{[(r-1)(s-1)]^{m}-1}{(r-1)(s-1)-1}. \end{equation} In the case of equalities in \eqref{N2-even} and \eqref{N1-even}, condition \eqref{basic=} holds, and the Moore bound is \begin{equation} \label{Moore-even} M(r,s;2m)= (r+s)\frac{[(r-1)(s-1)]^{m}-1}{(r-1)(s-1)-1}. \end{equation} The Moore bounds for $2\le s\le r\le 10$ and $d=4,6$ are shown in Tables \ref{tab:d=4} and \ref{tab:d=6}, where values in boldface are known to be attainable. \begin{table}[t] \begin{center} \begin{tabular}{|c||c|c|c|c|c|c|c|c|c|} \hline $r\setminus s$ & 2 & 3 & 4 & 5 & 6 & 7 & 8 & 9 & 10 \\ \hline \hlin 2 & $\textbf{8}^{\bullet}$ \\ \cline{1-3} 3 & $\textbf{15}^{* \diamond}$ & \textbf{30} \\ \cline{1-4} 4 & $\textbf{24}^{* \diamond}$ & 49 & \textbf{80} \\ \cline{1-5} 5 & $\textbf{35}^*$ & $\textbf{72}^{\bullet}$ & 117 & \textbf{170} \\ \cline{1-6} 6 & $\textbf{48}^*$ & 99 & 160 & 231 & \textbf{312} \\ \cline{1-7} 7 & $\textbf{63}^*$ & 130 & 209 & 300 & 403 & $\stackrel{(518)}{516}$ \\ \cline{1-8} 8 & $\textbf{80}^*$ & 165 & 264 & 377 & 504 & 645 & \textbf{800} \\ \cline{1-9} 9 & $\textbf{99}^*$ & 204 & 325 & 462 & 615 & 784 & 969 & \textbf{1170} \\ \cline{1-10} 10 & $\textbf{120}^*$ & 247 & $\textbf{292}^{\bullet}$ & 555 & 736 & 935 & 1152 & 1387 & \textbf{1640} \\ \hline \end{tabular} \end{center} \vskip-.25cm \caption{Moore bounds for diameter $d=4$. The attainable known values are in boldface. The asterisks correspond to the subdivision graphs $S(K_{r,r})$, see Section \ref{sec:subdiv-graphs}. The symbol `$\bullet$' indicates the orders of the graphs according to Theorem \ref{bb-Moore} and the diamonds correspond to unique graphs.} \vskip1cm \label{tab:d=4} \end{table} \begin{table}[t] \begin{center} \begin{tabular}{|c||c|c|c|c|c|c|c|c|c|} \hline $r\setminus s$ & 2 & 3 & 4 & 5 & 6 & 7 & 8 & 9 & 10 \\ \hline \hlin 2 & $\textbf{12}^{\bullet}$ \\ \cline{1-3} 3 & \textbf{35*} & \textbf{126} \\ \cline{1-4} 4 & \textbf{78*} & 301 & \textbf{728} \\ \cline{1-5} 5 & \textbf{147*} & 584 & 1431 & \textbf{2730} \\ \cline{1-6} 6 & \textbf{248*} & 999 & 2410 & 4631 & \textbf{7812} \\ \cline{1-7} 7 & 387 & 1570 & 3773 & 7212 & 12103 & $\stackrel{(18662)}{18660}$ \\ \cline{1-8} 8 & \textbf{570*} & 2321 & 5556 & 10569 & 17654 & 27105 & \textbf{39216} \\ \cline{1-9} 9 & \textbf{803*} & 3276 & $\textbf{7813}^{\bullet}$ &14798 & 24615 & 37648 & 54281 & \textbf{74898} \\ \cline{1-10} 10 & \textbf{1092*} & 4459 & 10598 & 19995 & 33136 & 50507 & 72594 & 99883 & \textbf{132860} \\ \hline \end{tabular} \end{center} \vskip-.25cm \caption{Moore bounds for diameter $d=6$. The attainable known values are in boldface. The asterisks correspond to the bimoore graphs of Proposition \ref{bb(s=2,d=6)-Moore}. The `$\bullet$' indicates the order of the graph according to Theorem \ref{bb-Moore}.} \label{tab:d=6} \end{table} Similarly, if the diameter is odd, say, $k=2m+1$ (for $m\ge 1$), and $u\in V_1$, we have \begin{equation} \label{N1-odd} N_1\le 1+r(s-1)+\cdots + r(s-1)[(r-1)(s-1)]^{m-1}=1+r(s-1)\frac{[(r-1)(s-1)]^{m}-1}{(r-1)(s-1)-1}=N_1', \end{equation} whereas, if $u\in V_2$, \begin{equation} \label{N2-odd} N_2\le 1+s(r-1)\frac{[(r-1)(s-1)]^{m}-1}{(r-1)(s-1)-1}=N_2'. \end{equation} But, in this case, $N_1'r\neq N_2's$ and, hence, the Moore bound must be smaller than $N_1'+N_2'$. In fact, it was proved in Yebra, Fiol, and F\`abrega \cite{YFF83} that, assuming $r>s$, \begin{equation} \label{N1-N2-odd} N_1\le \left\lfloor \frac{N_2'}{\rho}\right\rfloor \sigma\qquad\mbox{and} \qquad N_2\le \left\lfloor \frac{N_2'}{\rho}\right\rfloor \rho, \end{equation} where $\rho=\frac{r}{\gcd\{r,s\}}$ and $\sigma=\frac{s}{\gcd\{r,s\}}$. Then, in this case, we take the Moore bound \begin{equation} \label{Moore-odd} M(r,s;2m+1)= \left\lfloor\frac{1+s(r-1)\frac{[(r-1)(s-1)]^{m}-1}{(r-1)(s-1)-1}}{\rho}\right\rfloor (\rho+\sigma). \end{equation} Two bipartite biregular graphs with diameter three attaining the Moore bound \eqref{Moore-odd} were given in \cite{YFF83}. Namely, in Figure \ref{3unics}$(a)$, with $r=4$ and $s=3$, we would have the unattainable values $(N_1',N_2')=(9,10)$, whereas we get $(N_1,N_2)=(6,8)$, giving $M(4,3;3)=14$. In Figure \ref{3unics}$(b)$, with $r=5$ and $s=3$, $(N_1',N_2')=(11,13)$, and $(N_1,N_2)=(6,10)$, now corresponding to $M(5,3;3)=16$. \begin{figure}[t] \centering \includegraphics[width=\linewidth]{3unics.pdf} \vskip-6.75cm \caption{$(a)$ The only [4,3;3]-bimoore graph on $14$ vertices; $(b)$ One of the two [5,3;3]-bimoore graph on $16$ vertices; $(c)$ The only [6,3;3]-bimoore graph on $21$ vertices.} \label{3unics} \end{figure} As shown in Section \ref{sec:improvedMoore}, the bound \eqref{Moore-odd} can be improved for some values of the degree $r$. So, we display there the tables of the new Moore bounds for small values of $r,s$ and diameters $d=3$ and $d=5$. Recall that if $G=(V,E)$ is an $r$-regular graph of diameter $d$, then its \emph{defect} is $\delta=\delta(G)=M(r;d)-|V|$, where $M(r;d)$ stands for the corresponding Moore bound. Thus, in this paper, the defect of a $[r,s;d]$-bigraph $G=(V_1\cup V_2,E)$ is defined as $\delta=M(r,s;d)-|V_1\cup V_2|$. \section{An improved Moore bounds for odd diameter} \label{sec:improvedMoore} Let us begin with a simple result concerning the girth of the possible biregular bipartite graphs attaining the bound \eqref{Moore-odd} for odd diameter. \begin{lemma} \label{g<=4m} Every biregular bipartite graph $G$ of odd diameter $d=2m+1$ with order attaining the Moore bound \eqref{Moore-odd} has girth $g\le 4m$. \end{lemma} \begin{proof} Since $G$ is bipartite, we must have $g\le 2d=4m+2$. Consider the trees $T_1$ rooted at $u\in V_1$ , and $T_2$ rooted at $v\in V_2$, of vertices at distance at most $2m$ of their roots. If $G$ has girth $g=4m+2$, all vertices of $T_1$ must be different (otherwise, $T_1$ could not have the maximum number of vertices), and the same holds for all the vertices of $T_2$. This occurs if and only if $T_1$ and $T_2$ have numbers of vertices $N_1'$ and $N_2'$ given by \eqref{N1-odd} and \eqref{N2-odd}, respectively. But this is not possible because the Moore bound \eqref{Moore-odd} is obtained from \eqref{N1-N2-odd} as $\lfloor N_2'/\rho\rfloor(\rho+\sigma) <N_1'+N_2'$ (since $r>s\Rightarrow N_2'>N_1'$). Hence, the girth of $G$ is at most $4m$. \end{proof} \subsection{The case of diameter three} As a consequence of the following result, we prove in Corollary \ref{coro:optimal} that the $[6,3;3]$-graph of Figure \ref{3unics}$(c)$ has the maximum possible order. \\ \begin{proposition} \label{nonupperbound} If $\rho=\frac{r}{\gcd\{r,s\}}$ divides $s-1$, then there is no $[r,s;3]$-graph with order attaining the Moore-like bound in \eqref{Moore-odd}. Instead, the new improved Moore bound is \begin{equation} \label{Moore-odd-improved(d=3)} M^*(r,s;3)= (1+s(r-1)-\rho)\left(1+\frac{\sigma}{\rho}\right), \end{equation} with \begin{equation} \label{N1-N2-(d=3)} N_1\le [1+s(r-1)-\rho]\frac{\sigma}{\rho} \qquad \mbox{and}\qquad N_2\le 1+s(r-1)-\rho, \end{equation} where $\rho=\frac{r}{\gcd\{r,s\}}$ and $\sigma=\frac{s}{\gcd\{r,s\}}$. \end{proposition} \begin{proof} Suppose that, under the hypothesis, there exists a $[r,s;3]$-graph $G=(V_1\cup V_2,E)$ that attains the upper bound in \eqref{Moore-odd}. Then, with $r>s$, $|V_1|$ is the number of vertices of degree $r$, $ |V_2|$ is the number of vertices of degree $s$, and $N_1=|V_1|< |V_2|=N_2$. Thus, for diameter $d=3$, we have $$ N_2=\left\lfloor \frac{1+s(r-1)}{\rho}\right\rfloor\rho=1+s(r-1)=N_2'. $$ This means that there is only one shortest path of length at most 2 from any vertex $v\in V_2$ to all the vertices of $V_2$. Hence, the girth of $G$ is larger than 4. If not, we would have a cycle $u_1\sim v_1\sim u_2\sim v_2 (\sim u_1)$, with $u_i\in V_1$ and $v_i\in V_2$ for $i=1,2$. So, there would be 2 shortest 2-paths between $v_1$ and $v_2$. This is a contradiction with Lemma \ref{g<=4m} and, hence, the upper bound \eqref{Moore-odd} cannot be attained. \end{proof} Otherwise, if $\rho$ does not divide $s-1$, we have the bound \eqref{Moore-odd} with $m=1$, where \eqref{N1-N2-odd} yields \begin{equation} \label{N1-N2(d=3)not-divide} N_1\le \left\lfloor s\cdot \gcd\{r,s\}-\frac{s-1}{\rho}\right\rfloor \sigma\qquad\mbox{and}\qquad N_2\le \left\lfloor s\cdot \gcd\{r,s\}-\frac{s-1}{\rho}\right\rfloor\rho. \end{equation} In Table \ref{tab:d=3}, we show the values of the Moore bounds in \eqref{Moore-odd} and \eqref{Moore-odd-improved(d=3)} for $s\le 2\le r\le 11$ and diameter $d=3$. The values between parenthesis correspond to the old bound \eqref{Moore-odd} that was given in \cite{YFF83}. The attainable known values are in boldface. The asterisks indicate the values obtained in this paper (see Proposition \ref{propo:(r,3)}). The $[6,3;3]$-bigraph with $21$ vertices of Figure \ref{3unics}$(c)$ can be shown to be unique, up to isomorphisms. \begin{table}[t] \begin{center} \begin{tabular}{|c||c|c|c|c|c|c|c|c|c|c|} \hline $r\setminus s$ & 2 & 3 & 4 & 5 & 6 & 7 & 8 & 9 & 10 & 11 \\ \hline \hlin 2 & \textbf{6} \\ \cline{1-3} 3 & \textbf{5} & \textbf{14} \\ \cline{1-4} 4 & \textbf{9} & $\textbf{14}^{* \diamond}$ & \textbf{26} \\ \cline{1-5} 5 & \textbf{7} & \textbf{16*} & 27 & \textbf{42} \\ \cline{1-6} 6 & \textbf{12} & $\stackrel{(24)}{\textbf{21}^{\diamond}}$ & $\stackrel{(35)}{30}$ & 44 & \textbf{62} \\ \cline{1-7} 7 & \textbf{9} & \textbf{20*} & 33 & 48 & 65 & 86 \\ \cline{1-8} 8 & \textbf{15} & \textbf{22*} & 42 & 52 & 70 & 90 & \textbf{114} \\ \cline{1-9} 9 & \textbf{11} & 32 & 39 & 56 & 80 & 96 & 119 & \textbf{146} \\ \cline{1-10} 10 & \textbf{18} & \textbf{26*} & 49 & $\stackrel{(69)}{66}$ & $\stackrel{(88)}{80}$ & 102 & 126 & 152 & \textbf{182} \\ \cline{1-11} 11 & \textbf{13} & \textbf{28*} & 45 & 64 & 85 & 108 & 133 & 160 & 189 & 222 \\ \hline \end{tabular} \end{center} \vskip-.25cm \caption{Best Moore bounds for diameter $d=3$. The attainable known values are in boldface. The asterisks correspond to the graphs obtained according to Proposition \ref{propo:(r,3)}, and the diamonds correspond to unique graphs. The values between parenthesis correspond to the old (unattainable) bound \eqref{Moore-odd}.} \label{tab:d=3} \vskip1cm \end{table} As a consequence of the above results, we get the following Moore bounds when the degree $r$ is a multiple of the degree $s$. \begin{corollary} For $s\ge 2$, the best Moore bounds for the orders $N_1$ and $N_2$ of a $[\rho s,s;3]$-graph are as follows: \begin{itemize} \item[$(i)$] If $\rho|(s-1)$, then \begin{equation*} N_1 \le (s^2-1)-\frac{s-1}{\rho},\qquad\mbox{and}\qquad N_2 \le \rho(s^2-1)-(s-1). \end{equation*} \item[$(i)$] If $\rho\nmid (s-1)$, then \begin{equation*} N_1 \le s^2-\left\lceil s/\rho\right\rceil,\qquad\mbox{and}\qquad N_2 \le \rho(s^2-\left\lceil s/\rho\right\rceil). \end{equation*} \end{itemize} \end{corollary} \begin{proof} Note that, under the hypothesis, $\gcd\{r,s\}=s$, $\rho=\frac{r}{s}$, and $s=1$. Then, $(i)$ follows from \eqref{N1-N2-(d=3)}. Concerning $(ii)$, the values in \eqref{N1-N2(d=3)not-divide} become $$ N_1\le \left\lfloor s^2-\frac{s-1}{\rho}\right\rfloor\qquad\mbox{\rm and}\qquad N_2\le \left\lfloor s^2-\frac{s-1}{\rho}\right\rfloor\rho, $$ which are expressions equivalent to those given above. \end{proof} Assuming that $\rho=2$ and $s$ is odd, we get the following consequence. \begin{corollary} \label{coro:2s-s} There is no $[2s,s;3]$-bimoore graph for $s$ odd. \label{coro:optimal} \end{corollary} \begin{proof} From the first statement, the Moore bound $M(6,3;3)=24$ given in \eqref{Moore-odd} is not attained. With this bound, we have $N_2'=16$ vertices of degree $3$, and $N_1'=8$ vertices of degree $6$. So, the first possible values are $N_2=14$ vertices of degree $3$ and $N_1=7$ vertices of degree $6$, which corresponds to the graph depicted in Figure \ref{3unics}$(c)$. \end{proof} \subsection{The general case} Proposition \ref{nonupperbound} can be extended for any odd diameter, as shown in the following result. \begin{theorem}\label{nonupperbound2m+1} If $\rho=\frac{r}{\gcd\{r,s\}}$ divides $s\frac{[(r-1)(s-1)]^{m}-1}{(r-1)(s-1)-1}-1$, then there is no $[r,s;2m+1]$-graph with order attaining the Moore-like bound in \eqref{Moore-odd}. Instead, the new improved Moore bound is \begin{equation} \label{Moore-odd-improved} M^*(r,s;2m+1)= \left(\frac{1+s(r-1)\frac{[(r-1)(s-1)]^{m}-1}{(r-1)(s-1)-1}}{\rho}-1\right) (\rho+\sigma), \end{equation} where, as before, $\rho=\frac{r}{\gcd\{r,s\}}$ and $\sigma=\frac{s}{\gcd\{r,s\}}$. \end{theorem} \begin{proof} The proof that the bound in \eqref{Moore-odd} is not attainable follows the same reasoning as in Proposition \ref{nonupperbound}. Indeed, from the hypothesis, if there exists a $[r,s;2m+1]$-graph $G=(V_1\cup V_2,E)$ with order attaining the upper bound in \eqref{Moore-odd}, we would have $N_2=N_2'$, and the girth would be larger than $4m$, a contradiction. Then, as before, the new possible bounds in \eqref{N1-N2-odd} are $$ N_1\le \left( \frac{N_2'}{\rho}-1\right) \sigma\qquad\mbox{and} \qquad N_2\le \left(\frac{N_2'}{\rho}-1\right) \rho, $$ as claimed. \end{proof} \begin{table}[t] \begin{center} \begin{tabular}{|c||c|c|c|c|c|c|c|c|c|} \hline $r\setminus s$ & 2 & 3 & 4 & 5 & 6 & 7 & 8 & 9 & 10 \\ \hline \hlin 2 & 10 \\ \cline{1-3} 3 & $\textbf{15}^\diamond$ & 62 \\ \cline{1-4} 4 & 36 & $\stackrel{(112)}{105}$ & 242 \\ \cline{1-5} 5 & 56 & 168 & 369 & 682 \\ \cline{1-6} 6 & 80 & $\stackrel{(249)}{246}$ & $\stackrel{(535)}{530}$ & 957 & 1562 \\ \cline{1-7} 7 & 108 & 330 & 715 & $\stackrel{(1284)}{1272}$ & 2067 & 3110 \\ \cline{1-8} 8 & 140 & 429 & 924 & $\stackrel{(1651)}{1638}$ & 2646 & 3945 & 5602 \\ \cline{1-9} 9 & 176 & 544 & $\stackrel{(1157)}{1144}$ & 2044 & 3280 & 4880 & 6885 & 9362 \\ \cline{1-10} 10 & 216 & 663 & 1407 & $\stackrel{(2499)}{2496}$ & $\stackrel{(3976)}{3968}$ & 5882 & 8289 & 11229 & 14762 \\ \hline \end{tabular} \end{center} \vskip-.25cm \caption{Best Moore bounds for diameter $d=5$. The diamond corresponds to unique graphs. The values between parenthesis correspond to old (unattainable) bounds \eqref{Moore-odd}. } \label{tab:d=5} \end{table} Table \ref{tab:d=5} shows the values of the Moore bounds in \eqref{Moore-odd} and \eqref{Moore-odd-improved} for $s\le 2\le r\le 10$ and diameter $d=5$. As before, the values between parenthesis correspond to old (unattainable) bounds, as those in \eqref{Moore-odd} \subsection{Computational results} The enumeration of bigraphs with maximum order can be done with a computer whenever the Moore bound is small enough. To this end, given a diameter $d$, first we generate with {\em Nauty} \cite{MP13} all bigraphs with maximum orders $N_1=|V_1|$ and $N_2=|V_2|$ allowed by the Moore bound $M(r,s;d)$. Second, we filter the generated graphs keeping those with diameter $d$ using the library {\em NetworkX} from Python. Computational resources forces to study the cases where $\max\{N_1,N_2\} \leq 24$ and $n=N_1+N_2 \leq 32$. Computational results are shown in Table \ref{tab:comput}. \begin{table}[t] \begin{center} \begin{tabular}{|c||c|c|c|c|c|} \hline $[r,s;d]$ & $n$ & $N_1$ & $N_2$ & Generated graphs & Graphs with diameter $d$ \\ \hline\hline $[4,3;3]$ & ${\bf 14}^\diamond$ & 6 & 8 & 18 & 1 \\ \hline $[5,3;3]$ & {\bf 16} & 6 & 10 & 45 & 2 \\ \hline $[6,3;3]$ & 24 & 8 & 16 & 977278 & 0 \\ & ${\bf 21}^\diamond$ & 7 & 14 & 7063 & 1 \\ \hline $[7,3;3]$ & {\bf 20} & 6 & 14 & 344 & 4 \\ \hline $[8,3;3]$ & {\bf 22} & 6 & 16 & 950 & 10 \\ \hline $[9,3;3]$ & 32 & 8 & 24 & $>$122996904 & ? \\ & 28 & 7 & 21 & 2262100 & 1 \\ \hline $[10,3;3]$ & {\bf 26} & 6 & 20 & 6197 & 19 \\ \hline $[11,3;3]$ & {\bf 28} & 6 & 22 & 14815 & 16 \\ \hline $[5,4;3]$ & 27& 12 & 15 & 822558 & 0 \\ & {\bf 18} & 8 & 10 & 3143 & 583 \\ \hline $[3,2;4]$ & ${\bf 15}^\diamond$ & 6 & 9 & 6 & 1 \\ \hline $[4,2;4]$ & ${\bf 24}^\diamond$ & 8 & 16 & 204 & 1 \\ \hline $[3,2;5]$ & 20 & 8 & 12 & 20 & 0 \\ & ${\bf 15}^\diamond$ & 6 & 9 & 6 & 1 \\ \hline \end{tabular} \end{center} \vskip-.25cm \caption{Complete enumeration of bipartite biregular Moore graphs for some (small) cases of the degrees $r,s$ and diameter $d$.}\label{tab:comput} \end{table} Even with the computational limitations on the order of the sets $V_1$ and $V_2$, some of the optimal values given in Tables \ref{tab:d=4}, \ref{tab:d=3}, and \ref{tab:d=5} have been found with this method. When there is no graph for the largest values $N_1$ and $N_2$, the following lower feasible pair values may decrease dramatically, producing a large number of optimal graphs, as it happens in the case $[5,4;3]$. The particular case $[9,3;3]$ is computationally very hard, and it is out of our computational resources (which are very limited), but our guess is that there is no Moore graph in this case. Finally, we point out that these results encourage us to study the case $s=d=3$ and $3 \nmid r$ more in detail (see Section \ref{sec:d=3}), where computational evidence shows that there is always a Moore graph. \section{Bipartite biregular Moore graphs from generalized polygons} \label{gen-pols} In the first part of this section, we recall the connection between Moore graphs and generalized polygons that was extensively studied (see, for instance, Bamberg, Bishnoi, and Royle \cite{BamBisRoy}) because we will use it for the rest of the paper. In fact, our first result is an immediate consequence of the result proved by Araujo-Pardo, Jajcay, and Ramos in \cite{AraJajRam}, and the analysis given in the introduction about the coincidence of the bounds for bipartite biregular cages and bipartite biregular Moore graphs when $d$ is even. \begin{theorem}\cite{AraJajRam} Whenever a generalized quadrangle, hexagon, or octagon ${\mathcal G}$ of order $(s,t)$ exists, its point-line incidence graph is an $(s+1,t+1;8)$-, $(s+1,t+1;12)$- or $(s+1,t+1;16)$-cage, respectively. Hence, there exist infinite families of bipartite biregular $(n+1,n^2+1;8)$-, $(n^2+1,n^3+1;8)$-, $(n,n+2;8)$-, $(n+1,n^3+1;12)$- and $(n+1,n^2+1;16)$-cages. \end{theorem} Then, immediately we conclude the following result. \begin{theorem} \label{bb-Moore} There exists infinite families of bipartite biregular $[r^2+1,r+1;4]$-, $[r^3+1,r^2+1;4]$-, $[r+2,r;4]$-, $[r^3+1,r+1;6]$- and $[r^2+1,r+1;8]$-bimoore graphs. \end{theorem} In the following, we give some results related to generalized $n$-gons that we use in the rest of the paper. \begin{lemma}[\cite{VM98}, Lemma 1.3.6]\label{VMLemma} A geometry $\Gamma = (\cal{P},\cal{L},{\bf I})$ is a (weak) generalized $n$-gon if and only if the incidence graph of $\Gamma$ is a connected bipartite graph of diameter $d=n$ and girth $g=2d$, such that each vertex is incident with at least three (at least two) edges. \end{lemma} As every generalized polygon $\Gamma$ can be associated with a pair $(r,s)$, called the {\em order} of $\Gamma$, such that every line is incident with $r+1$ points, and every point is incident with $s+1$ lines (see van Maldeghem \cite{VM98}). This means, in particular, that the incidence graph of $\Gamma$ is a bipartite biregular graph with degrees $r+1$ and $s+1$. Besides, the following result determines the orders of both partite sets. \begin{theorem}[\cite{VM98}, Corollary 1.5.5]\label{PL} If there exists a $\Gamma = (\cal{P},\cal{L},{\bf I})$ (weak) generalized $n$-gon of order $(r,s)$ for $n \in \{3,4,6,8\}$, then \begin{itemize} \item For $n=3$, $|{\cal{P}}|= r^2+r+1$\ and\ $|{\cal{L}}|=s^2+s+1$. \item For $n=4$, $|{\cal{P}}|=(1+r)(1+rs)$\ and\ $|{\cal{L}}|=(1+s)(1+rs)$. \item For $n=6$, $|{\cal{P}}|=(1+r)(1+rs+r^2s^2)$\ and\ $|{\cal{L}}=(1+s)(1+rs+r^2s^2)$. \item For $n=8$, $|{\cal{P}}|=(1+r)(1+rs)(1+r^2s^2)$\ and\ $|{\cal{L}}|=(1+s)(1+rs)(1+r^2s^2)$. \end{itemize} \end{theorem} In 1964, Feit and Higman proved that finite generalized $n$-gons exist only for $n=\{3,4,6,8\}$. When $n=3$, we have the projective planes; when $n=4$, we have the generalized quadrangles, which are known to exist for parameter pairs $(q,q), (q,q^2), (q^2,q),(q^2,q^3), $ $(q^3,q^2), (q-1,q+1),(q+1,q-1) $; when $n=6$, we have the generalized hexagons with parameters $ (q,q),(q,q^3),(q^3,q)$, in both cases for $q$ prime power; and, finally, for $n=8$, we have the generalized octagons, which are only known to exist for the pairs $(q,q^2),(q^2,q)$, where $q$ is an odd power of $2$. \section{Two general constructions} \label{gen-cons} In this section, we construct some infinite families of Moore or large semiregular bipartite graphs derived from two general constructions: the subdivision graphs and the semi-double graphs \subsection{The subdivision graphs} \label{sec:subdiv-graphs} Given a graph $G=(V,E)$, its {\em subdivision graph} $S(G)$ is obtained by inserting a new vertex in every edge of $G$. So, every edge $e=uv\in E$ becomes two new edges, $ux$ and $xv$, with new vertex $x$ of degree two ($\deg(x)=2$). In our context, we have the following result. \begin{proposition} \label{propo:subdiv} Let $G=(V_1\cup V_2,E)$ be an $r$-regular bipartite graph with $n$ vertices, $m$ edges, diameter $D(\ge 2)$, and spectrum $\spec G=\{\lambda_0^{m_0},\lambda_1^{m_1},\ldots,\lambda_{d-1}^{m_{d-1}},\lambda_d^{m_d}\}$, where $\lambda_i=-\lambda_{d-i}$ and $m_i=m_{d-i}$ for $i=0,\ldots,\lfloor d/2\rfloor$. Then, its {\em subdivision graph} $S(G)$, is a bipartite biregular graph with degrees $(r,2)$, $n+m$ vertices, $2m$ edges, has diameter $2D$, and spectrum \begin{equation} \label{spec-double} \spec S(G)= \pm\sqrt{\spec G+r} \cup \{0^{m-n}\}. \end{equation} \end{proposition} \begin{proof} Let $S(G)$ have partite sets $U_1$ and $U_2$, with $U_1=V_1\cup V_2$ (the old vertices of $G$), and $U_2$ being the new vertices of degree 2. Let $x_{uv}\in U_2$ denote the vertex inserted in the previous edge $uv$. The first statement is obvious. To prove that the diameter of $S(G)$ is $2m$, we consider three cases: \begin{itemize} \item[$(i)$] Since $G$ has diameter $d$, there is a path of length $\ell \le d$ from any vertex $u\in U_1$ to any other vertex $v\in U_1$, say $u_0(=u),u_1,u_2,\ldots,u_{\ell}$. This path clearly induces a path of length $2\ell$ in $S(G)$. Namely, $u_0(=u),x_{u_0u_1},u_1,x_{u_1u_2}u_2,\ldots,x_{u_{\ell-1}u_{\ell}}u_{\ell}$. \item[$(ii)$] Since $G$ is bipartite, there is a path of length $\ell\le d-1$ between any two vertices in the same partite set (if $d$ is odd) or in different partite sets (if $d$ is even). Thus, from a vertex $x_{u_1u_2}\in U_2$, there is a path of length $2\ell+1\le 2d-1$ to any vertex $v\in U_1$. (This is because either $u_1$ or $u_2$ are at distance $\ell\le d-1$.) \item[$(iii)$] Finally, a path of length at most $2d$ between vertices $x_{u_1u_2},x_{v_1v2}\in U_2$ is obtained by considering first the path from $x_{u_1u_2}$ to $v_i$, for some $i\in\{1,2\}$ (which, according to $(ii)$, has length at most $2d-1$), together with the edge $v_ix_{v_1v2}$. \end{itemize} The result about the spectrum of $S(G)$ follows from a result by Cvetkovi\'{c}~\cite{Cv75}, who proved that, if $G$ is an $r$-regular graph with $n$ vertices and $m\big(=\frac{1}{2}nr\big)$ edges, then the characteristic polynomials of $S(G)$ and $G$ satisfy $\phi_{S(G)}(x) = x^{m-n} \phi_{G}(x^2-r)$. \end{proof} For instance, the $[2,r;4]$-Moore graphs proposed in Yebra, Fiol, and F\`abrega \cite{YFF83}, with $N_1=2r$ and $N_2=r^2$ can be obtained as the subdividing graphs $S(K_{r,r})$ (see the values in column of $s=2$ in Table \ref{tab:d=4}). For larger diameters, we can use the same construction with the known Moore bipartite graphs, which correspond to the incidence graphs of generalized polygons with $r=s$. \begin{proposition} \label{bb(s=2,d=6)-Moore} For any value of $r\ge 3$, with $r-1$ a prime power, there exist three infinite families of bimoore graphs with corresponding parameters $[r,2;2m]$ for $m\in\{3,4,6\}$. \end{proposition} \begin{proof} According to \eqref{Moore-even}, a bipartite biregular Moore graph with degrees $r$ and $2$ and diameter $d=2m$ has order $M(r,2;d)=\frac{r+2}{r-2}[(r-1)^m -1]$. Then, from Proposition \ref{propo:subdiv}, these are the parameters obtained when considering the subdividing graph $S(G)$ of a bipartite Moore graph $G$ of degree $r$ and diameter $m$. (For diameter $d=6$, see the values in the column $s=2$ of Table \ref{tab:d=6}.) \end{proof} \subsection{The semi-double graphs} Let $G$ be a bipartite graph with stable sets $V_1$ and $V_2$. Given $i\in \{1,2\}$, the {\em semi-double(-$V_i$) graph} $G^{2V_i}$ is obtained from $G$ by doubling each vertex of $V_i$, so that each vertex $u\in V_i$ gives rise to another vertex $u'$ with the same neighborhood as $u$, $G(u')=G(u)$. Thus, assuming, without loss of generality, that $i=1$, the graph $G^{2V_1}$ is bipartite with stable sets $V_1\cup V_1'$ and $V_2$, and satisfies the following result. \begin{theorem} \label{th:semi-double} Let $G=(V_1\cup V_2,E)$ be a bipartite graph on $n=n_1+n_2=|V_1|+|V_2|$ vertices, diameter $D(\ge 2)$, and spectrum $\spec G$. Then, its semi-double graph $G^{2V_1}$, on $N=2n_1+n_2$ vertices, has the same diameter $D$, and spectrum \begin{equation} \label{spec-semi-double} \spec G^{2V_1}=\sqrt{2}\cdot \spec G \cup \{0^{n_1}\}. \end{equation} \end{theorem} \begin{proof} Let $p : u_1(=u),u_2,\ldots,u_{\delta-1},u_{\delta}(=v)$ be a shortest path in $G$ between vertices $u$ and $v(\neq u')$, for $2\le \delta\le D$. If $u,v\in V_1\cup V_2$, $p$ also is a shortest path in $G^{2V_1}$. Otherwise, the following also are shortest paths in $G^{2V_1}$: \begin{align*} & u_1(=u),u_2,\ldots,u_{\delta-1},u_{\delta}'(=v');\\ & u_1'(=u'),u_2,\ldots,u_{\delta-1},u_{\delta}(=v);\\ & u_1'(=u'),u_2,\ldots,u_{\delta-1},u_{\delta}'(=v'). \end{align*} Finally, the distance from $u$ to $u'$ is clearly two. \\ To prove \eqref{spec-semi-double}, notice that the adjacency matrices $\mbox{\boldmath $A$}$ and $\mbox{\boldmath $A$}^{[1]}$ of $G$ and $G^{2V_1}$, are $$ \mbox{\boldmath $A$}=\left( \begin{array}{c|c} \mbox{\normalfont\Large\bfseries 0} & \mbox{\boldmath $N$} \\ \hline \\ [-.4 cm] \mbox{\boldmath $N$}^{\top} & \mbox{\normalfont\Large\bfseries 0} \end{array} \right)\qquad \mbox{and}\qquad \mbox{\boldmath $A$}^{[1]} =\left( \begin{array}{c|c|c} \mbox{\normalfont\Large\bfseries 0} & \mbox{\normalfont\Large\bfseries 0} & \mbox{\boldmath $N$} \\ \hline \\ [-.4 cm] \mbox{\normalfont\Large\bfseries 0} & \mbox{\normalfont\Large\bfseries 0} & \mbox{\boldmath $N$}\\ \hline \\ [-.4 cm] \mbox{\boldmath $N$}^{\top} & \mbox{\boldmath $N$}^{\top} & \mbox{\normalfont\Large\bfseries 0} \end{array} \right), $$ respectively, where $\mbox{\boldmath $N$}$ is an $n_1\times n_2$ matrix. Now, we claim that, if $\mbox{\boldmath $v$}=(\mbox{\boldmath $v$}_1 | \mbox{\boldmath $v$}_2)^{\top}$ is a $\lambda$-eigenvector of $\mbox{\boldmath $A$}$, then $\mbox{\boldmath $v$}^{[1]}=(\mbox{\boldmath $v$}_1 | \mbox{\boldmath $v$}_1| \sqrt{2}\mbox{\boldmath $v$} _2)^{\top}$ is a $\sqrt{2}\lambda$-eigenvector of $\mbox{\boldmath $A$}^{[1]}$. Indeed, from \begin{equation*} \mbox{\boldmath $A$}\mbox{\boldmath $v$}=\left( \begin{array}{c|c} \mbox{\normalfont\Large\bfseries 0} & \mbox{\boldmath $N$} \\ \hline \\ [-.4 cm] \mbox{\boldmath $N$}^{\top} & \mbox{\normalfont\Large\bfseries 0} \end{array} \right) \left( \begin{array}{c} \mbox{\boldmath $v$}_1\\ \hline \mbox{\boldmath $v$}_2 \end{array} \right) =\lambda \left( \begin{array}{c} \mbox{\boldmath $v$}_1\\ \hline \mbox{\boldmath $v$}_2 \end{array} \right) \qquad \Rightarrow \qquad \mbox{\boldmath $N$}\mbox{\boldmath $v$} _2=\lambda \mbox{\boldmath $v$} _1\quad{\rm and}\quad \mbox{\boldmath $N$}^{\top}\mbox{\boldmath $v$} _1=\lambda \mbox{\boldmath $v$} _2, \end{equation*} we have $$ \mbox{\boldmath $A$}^{[1]}\mbox{\boldmath $v$}^{[1]}= \left( \begin{array}{c|c|c} \mbox{\normalfont\Large\bfseries 0} & \mbox{\normalfont\Large\bfseries 0} & \mbox{\boldmath $N$} \\ \hline \\ [-.4 cm] \mbox{\normalfont\Large\bfseries 0} & \mbox{\normalfont\Large\bfseries 0} & \mbox{\boldmath $N$} \\ \hline \\ [-.4 cm] \mbox{\boldmath $N$}^{\top} & \mbox{\boldmath $N$}^{\top} & \mbox{\normalfont\Large\bfseries 0} \end{array} \right) \left( \begin{array}{c} \mbox{\boldmath $v$}_1\\ \hline \mbox{\boldmath $v$}_1\\ \hline \\ [-.45 cm] \sqrt{2}\mbox{\boldmath $v$} _2 \end{array} \right)= \left( \begin{array}{c} \sqrt{2}\mbox{\boldmath $N$}\mbox{\boldmath $v$} _2\\ \hline \sqrt{2}\mbox{\boldmath $N$}\mbox{\boldmath $v$} _2\\ \hline \\ [-.45 cm] 2\mbox{\boldmath $N$}^{\top}\mbox{\boldmath $v$} _1 \end{array} \right \left( \begin{array}{c} \sqrt{2}\lambda \mbox{\boldmath $v$} _1\\ \hline \sqrt{2}\lambda \mbox{\boldmath $v$} _1\\ \hline 2\lambda \mbox{\boldmath $v$} _2 \end{array} \right) =\sqrt{2}\lambda \mbox{\boldmath $v$} ^{[1]}. $$ Moreover, if $\mbox{\boldmath $U$}=\left( \begin{array}{c} \mbox{\boldmath $U$}_1\\ \hline \mbox{\boldmath $U$}_2 \end{array} \right)$ is a matrix whose columns are the $n$ independent eigenvectors of $\mbox{\boldmath $A$}$, then the $n$ columns of the matrix $\mbox{\boldmath $U$}'=\left( \begin{array}{c} \mbox{\boldmath $U$}_1\\ \hline \mbox{\boldmath $U$}_1\\ \hline \\ [-.45 cm] \sqrt{2}\mbox{\boldmath $U$}_2 \end{array} \right)$ also are independent since, clearly, $\rank \mbox{\boldmath $U$}'=\rank \mbox{\boldmath $U$}=n_1+n_2$. Consequently, $\spec \mbox{\boldmath $A$}\subset \spec \mbox{\boldmath $A$}^{[1]}$. Finally, each of the remaining $n_1$ eigenvalues $0$ corresponds to an eigenvector with $u$-th component $+1$ and $u'$-th component $-1$ for a given $u\in V_1$, and $0$ elsewhere. This is because the matrix $\mbox{\boldmath $U$}^{[1]}$ obtained by extending $\mbox{\boldmath $U$}'$ with such eigenvalues, that is, $$ \mbox{\boldmath $U$}^{[1]}=\left( \begin{array}{c|c} \mbox{\boldmath $U$}_1 & \mbox{\boldmath $I$} \\ \hline \mbox{\boldmath $U$}_1 & -\mbox{\boldmath $I$} \\ \hline \\ [-.45 cm] \sqrt{2}\mbox{\boldmath $U$}_2 & \mbox{\normalfont\Large\bfseries 0} \end{array} \right) $$ has rank $n=2n_1+n_2$, as required. \end{proof} In general, we can consider the $k$-tuple graph $G^{kV_i}$, which is defined as expected by replacing each vertex $u\in V_i$ of $G$ by $k$ vertices $u_1,\ldots,u_k$ with the same adjacencies as $u$. Then, similar reasoning as in the proof of Theorem \ref{th:semi-double} leads to the following result. \\ \begin{theorem} \label{th:k-tuple} Let $G=(V_1\cup V_2,E)$ be a bipartite graph on $n=n_1+n_2=|V_1|+|V_2|$ vertices, diameter $D(\ge 2)$, and spectrum $\spec G$. Then, its $k$-tuple graph $G^{kV_1}$, on $N=kn_1+n_2$ vertices, has the same diameter $D$, and spectrum \begin{equation} \label{spe-k-yuble} \spec G^{kV_1}=\sqrt{k}\cdot \spec G \cup \{0^{(k-1)n_1}\}. \end{equation} \end{theorem} As a consequence of Theorem \ref{th:semi-double}, we introduce a family of $[r,2r;d]$-graphs for $d=\{3,4,6\}$ using the existence of bipartite Moore graphs of order $M(r;d)$ for $r-1$ a prime power and these values of $d$. That is, the incidence graphs of the mentioned generalized polygons. \begin{theorem} \label{r2r} The following are $[r,2r;d]$-biregular bipartite graphs for $r\geq 3$, $r-1$ a prime power, and diameter $d\in \{3,4,6\}$. \begin{itemize} \item [(i)] A $[r,2r;3]$-biregular bipartite graph has order $n=3r^2-3r+3$ with defect $\delta=\frac{3}{2}(r-1)$ for odd $r$, and $\delta=\frac{3r}{2}-3$ for even $r$. \item [$(ii)$] A $[r,2r;4]$-biregular bipartite graph has order $n=3r^3-6r^2+10r$ with defect $\delta=3r^3-3r^2-2r$. \item [$(iii)$] A $[r,2r;6]$-biregular bipartite graph has order $n=2r^5-8r^4+14r^3-12r^2+6r$ with defect $\delta=10r^5-28r^4+31r^3-15r^2-3r$. \end{itemize} \end{theorem} \begin{proof} $(i)$ Let $G$ be a $[r,3]$-Moore graph, that is, the incidence graph of a projective plane of order $r-1$. As already mentioned, $G$ is bipartite, has $2r^2-2r+1$ vertices, and diameter $3$. Then, by Theorem \ref{th:semi-double}, the semi-double graph $G^{[1]}$ (or $G^{[2]}$) is a bipartite graph on $3r^2-3r+3$ vertices, biregular with degrees $r$ and $2r$, and diameter $3$. Thinking on the projective plane, this corresponds to duplicate, for instance, each line, so that each point is on $2r$ lines, and each line has $r$ points, as before. Then, $G^{[1]}$ is the incidence graph of this new incidence geometry ${\cal{I}}_G$. Moreover, the Moore bound \eqref{Moore-even} is $M(2r,r;3)=3r^2-\frac{3}{2}(r-1)$ if $r$ is odd, and $M(2r,r;3)=3r^2-\frac{3}{2}r$ if $r$ is even. Thus, a simple calculation gives that the defect of $G$ is equal to $\delta=\frac{3}{2}(r-1)$ for $r$ odd and $\delta=\frac{3}{2}r-3$ for $r$ even. Figure \ref{3unics}$(c)$ depicts the only $[6,3;3]$-biregular bipartite graph of order $21$ obtained from the Heawood graph (that is, the incidence graph of the Fano plane). Notice that, by Corollary \ref{coro:2s-s}, this is a Moore graph since it has the maximum possible number of vertices (see Table \ref{tab:d=3}). $(ii)$ In this case, we apply Theorem \ref{th:semi-double} to $G$ being the $[r;4]$-Moore graph (that is, the incidence graph of a generalized quadrangle of order $r-1$). Now the semi-double graph $G^{[1]}$ is $[r,2r;4]$-bipartite biregular with $3[(r-1)^3+(r-1)^2+(r-1)+1]=3r^3-6r^2+10r$ vertices, whereas the Moore bound in \eqref{Moore-even} is $M(2r,r;3)=6r^3-9r^2+6r$. Hence, the defect is $\delta=3r^3-3r^2-2r$. For example, the $[3,6;4]$-bigraph of order $45$ obtained from the Tutte graph (that is, the incidence graph of the generalized quadrangle of order $2$) has defect $\delta=54$. $(iii)$ Finally, to obtain a $[r,2r;6]$-bigraph of order $2r^5-8r^4+14r^3-12r^2+6r$, we apply Theorem \ref{th:semi-double} to the $[r;6]$-Moore graph (the incidence graph of a generalized hexagon of order $r-1$). Then, we obtain a $[r,2r;6]$-bigraph with $2r^5-8r^4+14r^3-12r^2+6r$ vertices, whereas the Moore bound in \eqref{Moore-even} is $M(2r,r;6)=12r^5-36r^4+45r^3-27r^2+9r$, yielding a defect $\delta=10r^5-28r^4+31r^3-15r^2-3r$. \end{proof} Notice that, in Theorem \ref{r2r}, we only state results for Moore graphs constructed with regular generalized quadrangles; clearly, it is possible to do the same starting with biregular bipartite Moore graphs of diameters $6$ and $8$ given as in Theorem \ref{bb-Moore}, but since the bounds are far to be tight, we restricted the details to the best cases. \section{Bipartite biregular Moore graphs of diameter $3$} \label{sec:d=3} We have already seen some examples of large biregular graphs with diameter three. Namely, the Moore graphs with degrees $(3,4)$, $(3,5)$, and $(3,6)$ of Figure \ref{3unics} (the last one in Theorem \ref{r2r}$(i)$). Now, we begin this section with the simple case of bimoore graphs with degrees $r$ (even) and $2$, and diameter $3$. For these values, the Moore bounds in \eqref{Moore-odd} and \eqref{Moore-even} turn out to be $M(r,2;3)=2+r$ when $r(>1)$ is odd, and $M(r,2;3)=3\left(1+\frac{r}{2}\right)$ ($N_1=3$ and $N_2=3r/2$) when $r(>2)$ is even, respectively. In the first case, the bound is attained by the complete bipartite graph $K_{2,r}$ and, hence, the diameter is, in fact, $2$. In the second case, the Moore graphs are obtained via Theorem \ref{th:k-tuple}. Let $G$ be a hexagon (or $6$-cycle) with vertex set $V_1\cup V_2=\{u_1,u_3,u_5\}\cup \{2,4,6\} $. Then, the $k$-tuple $G^{kV_1}$ with $k=r/2$ is a $[2,2r;3]$-bimoore graph on $3(r+1)$ vertices (see Table \ref{tab:d=3}). In general, and in terms of designs, to guarantee that the diameter is equal to $3$, it is necessary that any pair of points shares a block, and any pair of blocks has a non-empty intersection. It is well known that this kind of structure exists when $r=m$ and $r-1$ is a prime power. Namely, the so-called projective plane of order $r-1$. In this case, any pair of blocks (or lines) intersects in one and only one point and, for any pair of points, they share one and only one block (or line). For more details about projective planes, you can consult, for instance, Coxeter \cite{Cox93}. In our constructions of block designs giving graphs of diameter $3$, this condition is not necessary, and two blocks can be intersected in more than one point, and two points can share one or more blocks. Moreover, we may have the same block appearing more than once. \subsection{Existence of bipartite biregular Moore graphs of diameter 3 and degrees $r\ge 3$ and $s=3$} \label{sec:d=3} Let us consider the diameter $3$ case together with $r>s=3$. According to Eqs. \eqref{N1-odd} and \eqref{N2-odd}, the maximum numbers of vertices for the partite sets, in this case, are $N_1'=2r+1$ and $N_2'=3r-2$. We recall that the equality $N_1'r=3N_2'$ does not hold when the diameter is odd. Since $$ \rho=\frac{r}{\gcd\{r,3\}}=\left\{\begin{array}{cl} r & \textrm{if} \ \ 3 \nmid r \\ \frac{r}{3} & \textrm{otherwise} \end{array}\right. \quad \textrm{and} \quad \sigma=\frac{3}{\gcd\{r,3\}}=\left\{\begin{array}{cl} 3 & \textrm{if} \ \ 3 \nmid r \\ 1 & \textrm{otherwise} \end{array}\right. $$ we obtain the Moore-like bounds (see \eqref{N1-N2-odd}): \[ N_1 \leq \left\lfloor \frac{3r-2}{\rho}\right\rfloor \sigma = \left\{\begin{array}{cl} 3\omega & \textrm{if} \ \ 3 \nmid r \\ \omega' & \textrm{otherwise} \end{array}\right. \quad \textrm{and} \quad N_2\le \left\lfloor \frac{3r-2}{\rho}\right\rfloor \rho = \left\{\begin{array}{cl} r\omega & \textrm{if} \ \ 3 \nmid r \\ \frac{r\omega'}{3} & \textrm{otherwise} \end{array}\right. \] where $\omega=\left\lfloor \frac{3r-2}{r}\right\rfloor=2$ and $\omega'=\left\lfloor \frac{3r-2}{r/3}\right\rfloor=8$. As a consequence, whenever $3 \nmid r$, we obtain Moore-like bounds $N_1\leq 6$ and $N_2 \leq 2r$. Otherwise, when $3|r$, we have the bounds $N_1\leq 8$ and $N_2 \leq \frac{8}{3}r$. The following graphs have orders that either attain or are close to such Moore bounds. Given an integer $n\ge 6$, let $G_{6+n}=(V_1 \cup V_2, E)$ be the bipartite graph with independent sets $V_1=\{(0,j) \ | \ j \in \mathbb{Z}_6\}$, $V_2=\{(1,i) \ | \ i \in \mathbb{Z}_{n}\}$, and where $(1,i) \sim (0,j)$ for all $i \in \mathbb{Z}_{n}$ if and only if $j \equiv i \pmod{6}$ or $j \equiv i+1 \pmod{6}$ or $j \equiv i+3 \pmod{6}$. Thus, $G_{6+n}$ is a bipartite graph with $n+6$ vertices, where every vertex $v\in V_2$ has degree $3$ and, assuming that $n=6k+\rho$ ($n\equiv \rho \pmod{6}${\color{blue})}, vertex $u\in V_1$ has the degree indicated in Table \ref{tab:degreesV1}. Indeed, let us consider the case $\rho=2$ (the other cases are analogous), where $n=6k+2$ and, for simplicity, let $V_2= \{0,1,\ldots,6k-1\}$. According to the adjacency rules, the vertex $(0,i)\in V_1$, for $i\in \ns{Z}_6$, is adjacent to all the vertices $ j\in V_2$ with $j\equiv i,i-1,i-3 \pmod{6}$. Then, we have the following cases: \begin{itemize} \item If $i=0$, $V_2$ contains $k+1$ numbers $j\equiv 0 \pmod6$, $k$ numbers $j\equiv-1\equiv 5 \pmod6$, and $k$ numbers $j\equiv-3\equiv 3 \pmod6$. Thus, the degree of $(0,0)$ is $3k+1$. \item If $i=1$, $V_2$ contains $k+1$ numbers $j\equiv 1 \pmod6$, $k+1$ numbers $j\equiv0 \pmod6$, and $k$ numbers $j\equiv-2\equiv 4 \pmod6$. Thus, the degree of $(0,2)$ is $3k+2$. \item If $i=2$, $V_2$ contains $k$ numbers $j\equiv 2 \pmod6$, $k+1$ numbers $j\equiv1 \pmod6$, and $k$ numbers $j\equiv-1\equiv 5 \pmod6$. Thus, the degree of $(0,3)$ is $3k+1$. \item[ \mbox{\boldmath $v$} dots] \item If $i=5$, $V_2$ contains $k$ numbers $j\equiv 5 \pmod6$, $k$ numbers $j\equiv4 \pmod6$, and $k$ numbers $j\equiv2 \pmod6$. Thus, the degree of $(0,5)$ is $3k$. \end{itemize} \begin{table}[h] \begin{center} \begin{tabular}{|c||c|c|c|c|c|c|} \hline $\rho\setminus u$ & (0,0) & (0,1) & (0,2) & (0,3) & (0,4) & (0,5) \\ \hline 0 & $3k$ & $3k$ & $3k$ & $3k$ & $3k$ & $3k$ \\ 1 & $3k+1$ & $3k+1$ & $3k$ & $3k+1$ & $3k$ & $3k$ \\ 2 & {\boldmath $3k+1$} & {\boldmath $3k+2$} & {\boldmath $3k+1$} & {\boldmath $3k+1$} & {\boldmath $3k+1$} & {\boldmath $3k$} \\ 3 & $3k+1$ & $3k+2$ & $3k+2$ & $3k+2$ & $3k+1$ & $3k+1$ \\ 4 & {\boldmath $3k+2$} & {\boldmath $3k+2$} & {\boldmath $3k+2$} & {\boldmath $3k+3$} & {\boldmath $3k+2$} & {\boldmath $3k+1$} \\ 5 & $3k+2$ & $3k+3$ & $3k+2$ & $3k+3$ & $3k+3$ & $3k+2$ \\ \hline \end{tabular} \end{center} \caption{Degrees of the vertices $u=(0,j)\in V_1$ when $|V_2|=n=6k+\rho$. } \label{tab:degreesV1} \end{table} \begin{proposition} \label{G(6+n)} The diameter of the bipartite graph $G_{6+n}=(V_1\cup V_2, E)$, on $n+6$ vertices, is $d=3$. \end{proposition} \begin{proof} It suffices to prove that every pair of different vertices $u,u'\in V_1$, and every pair of different vertices $v,v'\in V_2$, are at distance two. In the first case, notice that $u=(0,j)\in V_1$ is adjacent to every vertex $v=(1,i)\in V_2$ such that $i\equiv j,j-1, j-3 \pmod6$. Moreover, vertex $(1,j)\in V_2$ is adjacent to vertices $(0,j+1),(0,j+3)\in V_1$, vertex $(1,j-1)\in V_2$ is adjacent to vertices $(0,j-1),(0,j+2)\in V_1$, and vertex $(1,j-3)\in V_2$ is adjacent to vertices $(0,j-3),(0,j-2)\in V_1$. Schematically (with all arithmetic modulo 6), \begin{align} (0,j)\quad & \sim \quad (1, j), (1,j-1), (1, j-3) \label{u-u'}\\ & \sim \quad (0,j+1),(0,j+3),(0,j-1),(0,j+2),(0,j-3),(0,j-2) \nonumber, \end{align} which are all vertices of $V_1$ different from $(0,j)$. Similarly, starting from a vertex of $V_2$, we have \begin{align*} (1,i)\quad & \sim \quad (0, i), (0,i+1), (0, i+3)\\ & \sim \quad (1,i), (1,i-1),(1,i-3),(1,i+1),(1,i-2),(1,i+3),(1,i+1), \end{align*} with the last two representing all vertices of $V_2$ because the second entries cover all values modulo 6. This completes the proof. \end{proof} The following result proves the existence of an infinite family of Moore bipartite biregular graphs with diameter three. \begin{proposition} \label{propo:(r,3)} For any integer $r\ge 6$ such that $3 \nmid r$, there exists a bipartite graph $G$ on $2r+6$ vertices (the Moore bound for the case of degrees $(r,3)$ and diameter $3$) with degrees $3,r,r\pm 1$ and diameter $d=3$. Moreover, when $r\equiv 2 \mod3$, there exists a Moore bipartite biregular graph with degrees $(r,3)$ and diameter $3$. \end{proposition} \begin{proof} From the comments at the beginning of this subsection, if $3 \nmid r$, the graph $G_{6+n}$ of Proposition \ref{G(6+n)} with $n=2r$ has maximum order for diameter $3$. However, as shown before, not every vertex of $u\in V_1$ has degree $r$ since it is required to become a bipartite biregular Moore graph. More precisely, if $2r=6k+\rho$ (with $\rho=2$ or $\rho=4$, since $3 \nmid r$), from the two rows in boldface of Table \ref{tab:degreesV1}, we have $$ \deg(u)=\left\{\begin{array}{cl} r & \textrm{if} \ \ u \in \{(0,0),(0,2),(0,3),(0,4)\}, \\ r+1 & \textrm{if} \ \ u=(0,\rho-1), \\ r-1 & \textrm{if} \ \ u=(0,5). \\ \end{array}\right. $$ This proves the first statement. When $r\equiv 2 \pmod3$, that is $\rho=4$, we obtain biregularity by modifying only one adjacency of $G_{6+n}$, as follows. Let $G'_r$ be the graph $G_{6+r}$ defined above, but where the edge $(0,3)\sim(1,i)$, for some $i\equiv 0 \pmod 3$, is switched to $(0,5)\sim(1,i)$. Then, $G'_r$ is a bipartite semiregular Moore graph of diameter $3$ for all $r\ge 5$. To prove that $G'_r$ has diameter $d=3$, let us check again that every pair of different vertices $u,u'\in V_1$, and every pair of different vertices $v,v'\in V_2$, are at distance two. \begin{itemize} \item If $u,u'\in V_1$, we only need to consider the case when $u=(0,3)$. Then, the paths are as follows (where $(1,i)$ represents any vertex $(1,i')$ with $i'\equiv i \pmod6$): \begin{align*} (0,3)\quad & \sim \quad (1, 3), (1,2), (1, 0)\quad \sim \quad (0,4),(0,0),(0,2),(0,5),(0,1), \end{align*} because $(0,3)$ was initially adjacent to more than one vertex of type $(0,i)$ with $i\equiv 0 \pmod 3$. Thus, all vertices of $V_0$ are reached from $(0,3)$. \item If $v,v'\in V_2$, assume that $v=(1,i)$ with $i\equiv 0 \pmod6$ (the case where $i\equiv 3 \pmod6$ is similar). Then, \begin{itemize} \item If $v=(1,0)$, we have the paths \begin{align*} (1,0)\quad & \sim \quad (0, 0), (0,1), (0,5) \quad \sim \quad (1,0),(1,3),(1,1),(1,4),(1,2). \end{align*} Notice that, in this case, the first step is not $(1,0)\sim (0,3)$ (deleted edge), but $(1,0)\sim (0,5)$. Despite this, we still reach all vertices of $V_2$, as required. \item If $v=(1,i)$ with $i\neq 0$, the first step is \begin{align*} (1,i)\quad & \sim \quad (0, i), (0,i+1), (0,i+3) \end{align*} So, the only problem would be when some $i,i+1,i+3$ is $3$, since we have not the adjacency $(0,3)\sim (1,0)$. But, if so, we have the following alternative adjacencies: if $i=0,3$, we have $(0,0)\sim (1,0)$; and if $i=2$, we have $(0,5)\sim (1,0)$. Again, all vertices of $V_2$ are reached, completing the proof. \end{itemize} \end{itemize} \vskip-.5cm \end{proof}
{ "timestamp": "2021-03-23T01:21:11", "yymm": "2103", "arxiv_id": "2103.11443", "language": "en", "url": "https://arxiv.org/abs/2103.11443", "abstract": "A bipartite graph $G=(V,E)$ with $V=V_1\\cup V_2$ is biregular if all the vertices of a stable set $V_i$ have the same degree $r_i$ for $i=1,2$. In this paper, we give an improved new Moore bound for an infinite family of such graphs with odd diameter. This problem was introduced in 1983 by Yebra, Fiol, and Fàbrega.\\\\ Besides, we propose some constructions of bipartite biregular graphs with diameter $d$ and large number of vertices $N(r_1,r_2;d)$, together with their spectra. In some cases of diameters $d=3$, $4$, and $5$, the new graphs attaining the Moore bound are unique up to isomorphism.", "subjects": "Combinatorics (math.CO)", "title": "Bipartite biregular Moore graphs", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9843363531263359, "lm_q2_score": 0.8128673087708699, "lm_q1q2_score": 0.8001348422911373 }
https://arxiv.org/abs/2002.01597
Berge cycles in non-uniform hypergraphs
We consider two extremal problems for set systems without long Berge cycles. First we give Dirac-type minimum degree conditions that force long Berge cycles. Next we give an upper bound for the number of hyperedges in a hypergraph with bounded circumference. Both results are best possible in infinitely many cases.
\section{Introduction} \subsection{Classical results on longest cycles in graphs} The {\em circumference} $c(G)$ of a graph $G$ is the length of its longest cycle. In particular, if a graph has a cycle $C$ which covers all of its vertices, $V(C)=V(G)$, we say it is {\em hamiltonian}. A classical result of Dirac states that high minimum degree in a graph forces hamiltonicity. \begin{thm}[Dirac~\cite{D}]\label{th:D} Let $n \geq 3$, and let $G$ be an $n$-vertex graph with minimum degree $\delta(G)$. If $\delta(G) \geq n/2$, then $G$ contains a hamiltonian cycle. If $G$ is 2-connected, then $c(G)\geq \min\{n, 2\delta(G)\}$. \end{thm} Inspired by this theorem, it is common in extremal combinatorics to refer to results in which a minimum degree condition forces some structure as a {\em Dirac-type condition}. The second part of Theorem~\ref{th:D} cannot be extended to non 2-connected graphs: let ${\mathcal F}_{n,k}$ be the family of graphs in which each block (inclusion maximal 2-connected subgraph) of the graph is a copy of $K_{k-1}$. Every $F \in {\mathcal F}_{n,k}$ has minimum degree $k-2$, but its longest cycle has length $k-1$. \begin{thm}[Erd\H{o}s, Gallai~\cite{EG}]\label{th:EG} Let $G$ be an $n$-vertex graph with no cycle of length $k$ or longer. Then $e(G) \leq \frac{n-1}{k-2}{k-1 \choose 2}$. \end{thm} So the graphs in ${\mathcal F}_{n,k}$ have the maximum number of edges among the $n$-vertex graphs with circumference $k-1$. They also maximize the number of cliques of any size: \begin{thm}[Luo~\cite{luo}]\label{cliques} Let $G$ be an $n$-vertex graph with no cycle of length $k$ or longer. Then the number of copies of $K_r$ in $G$ is at most $\frac{n-1}{k-2}{k-1 \choose r}$. \end{thm} \subsection{Known results on cycles in hypergraphs} A hypergraph ${\mathcal H}$ is a set system. We often refer to the ground set as the set of vertices $V({\mathcal H})$ of ${\mathcal H}$ and to the sets as the hyperedges $E({\mathcal H})$ of ${\mathcal H}$. When there is no ambiguity, we may also refer to the hyperedges as edges. In this paper, we prove versions of Theorems~\ref{th:D} and~\ref{th:EG} for hypergraphs with no restriction on edge sizes. Namely, we seek long {\em Berge cycles}. A {\bf Berge cycle} of length $\ell$ in a hypergraph is a set of $\ell$ distinct vertices $\{v_1, \ldots, v_\ell\}$ and $\ell$ distinct edges $\{e_1, \ldots, e_\ell\}$ such that $\{ v_{i}, v_{i+1} \}\subseteq e_i$ with indices taken modulo $\ell$. The vertices $\{v_1, \ldots, v_\ell\}$ are called {\bf representative vertices} of the Berge cycle. A {\bf Berge path} of length $\ell$ in a hypergraph is a set of $\ell+1$ distinct vertices $\{v_1, \ldots, v_{\ell+1}\}$ and $\ell$ distinct hyperedges $\{e_1, \ldots, e_{\ell}\}$ such that $\{ v_{i}, v_{i+1} \}\subseteq e_i$ for all $1\leq i\leq \ell$. The vertices $\{v_1, \ldots, v_{\ell+1}\}$ are called {\bf representative vertices} of the Berge path. For a hypergraph ${\mathcal H}$, the {\bf 2-shadow} of ${\mathcal H}$, denoted $\partial_2 {\mathcal H}$, is the graph on the same vertex set such that $xy \in E(\partial_2{\mathcal H})$ if and only if $\{x,y\}$ is contained in an edge of ${\mathcal H}$. Note that if we require no conditions on multiplicities of hyperedges, then we can arbitrarily add hyperedges of size $1$ without creating new Berge cyles or Berge paths. From now on, we only consider {\em simple} hypergraphs, i.e., those without multiple edges (except if it is stated otherwise). Bermond, Germa, Heydemann, and Sotteau~\cite{bermond} were among the first to prove Dirac-type results for uniform hypergraphs without long Berge cycles: Let $k > r$ and ${\mathcal H}$ be an $r$-uniform hypergraph with minimum degree $\delta({\mathcal H}) \geq {k-2 \choose r-1} + (r-1)$, then ${\mathcal H}$ contains a Berge cycle of length at least $k$. For large $n$, generalizations and results for linear hypergraphs are proved by Jiang and Ma~\cite{JM}. Ma, Hou, and Gao~\cite{MHG_r3} studied $3$-uniform hypergraphs. Coulson and Perarnau~\cite{rainbow} proved that if ${\mathcal H}$ is an $r$-uniform hypergraph on $n$ vertices, $r = o(\sqrt{n})$, and ${\mathcal H}$ has minimum degree $\delta({\mathcal H}) > {\lfloor (n-1)/2 \rfloor \choose r-1}$, then ${\mathcal H}$ contains a Berge hamiltonian cycle. Our new results differ from these in several aspects. We consider non-uniform hypergraphs, prove exact formulas, prove results for every $n$ (or every $n> 14$), and use only classical tools mentioned above and in Section~\ref{ss0}. \section{New results} Our first result is a Dirac-type condition that forces hamiltonian Berge cycles. \begin{thm}\label{diracH} Let $n \geq 15$ and let ${\mathcal H}$ be an $n$-vertex hypergraph such that $\delta({\mathcal H}) \geq 2^{(n-1)/2} + 1$ if $n$ is odd, or $\delta({\mathcal H}) \geq 2^{n/2 -1} + 2$ if $n$ is even. Then ${\mathcal H}$ contains a Berge hamiltonian cycle. \end{thm} The following four constructions show that Theorem~\ref{diracH} is the best possible. \newline ${}$\enskip --- \enskip Let $n$ be odd. Let ${\mathcal H}$ be the $n$-vertex hypergraph on the ground set $[n]$ with edges $\{A: A \subseteq [(n+1)/2]\} \cup \{B: B \subseteq \{(n+1)/2, \ldots n\}\}$. Then $\delta({\mathcal H}) = 2^{(n-1)/2}$ and ${\mathcal H}$ has no hamiltonian Berge cycle (because it has a cut vertex). \newline ${}$\enskip --- \enskip Let $n$ be even. Let ${\mathcal H}$ be the $n$-vertex hypergraph on the ground set $[n]$ with edges $\{A: A \subseteq [n/2]\} \cup \{B: B \subseteq \{(n/2 +1, \ldots n\}\}$ and the set $[n]$. Then $\delta({\mathcal H}) = 2^{n/2 -1}+1$ and ${\mathcal H}$ has no hamiltonian Berge cycle (because it has a cut edge, $[n]$). \newline ${}$\enskip --- \enskip Let $n$ be odd. Let ${\mathcal H}$ be the $n$-vertex hypergraph on the ground set $[n]$ obtained by taking all hyperedges with at most one vertex in $[(n+1)/2]$. Then $\delta({\mathcal H}) = 2^{(n-1)/2}$, and ${\mathcal H}$ cannot contain a Berge cycle with two consecutive representative vertices in $[(n+1)/2]$. \newline ${}$\enskip --- \enskip Let $n$ be even. Let ${\mathcal H}$ be the $n$-vertex hypergraph on the ground set $[n]$ obtained by taking all hyperedges with at most one vertex in $[n/2 + 1]$ and the edge $[n]$. Then $\delta({\mathcal H}) = 2^{n/2 - 1} + 1$, and ${\mathcal H}$ cannot contain a Berge cycle with two instances of two consecutive representative vertices in $[n/2 + 1]$ (because only one edge of ${\mathcal H}$ contains multiple vertices in $[n/2+1]$). Next, we consider hypergraphs without long Berge paths or cycles. \begin{thm}\label{th:mindeg_path} Let $k \geq 2$ and let ${\mathcal H}$ be a hypergraph such that $\delta({\mathcal H}) \geq 2^{k-2}+1$. Then ${\mathcal H}$ contains a Berge path with $k$ base vertices. \end{thm} A vertex disjoint union of complete hypergraphs of $k-1$ vertices shows that this bound is best possible for $n:=|V({\mathcal H})|$ divisible by $(k-1)$. It would be interesting to find $\max \delta({\mathcal H})$ for other values of $n$, and also for the cases when ${\mathcal H}$ is connected or 2-connected. \begin{thm}\label{th:mindeg_cycle} Let $k \geq 3$ and let ${\mathcal H}$ be a hypergraph such that $\delta({\mathcal H}) \geq 2^{k-2}+2$. Then ${\mathcal H}$ contains a Berge cycle of length at least $k$. \end{thm} The following constructions show that the bound in Theorem~\ref{th:mindeg_cycle} is best possible when $n$ is divisible by $(k-1)$ and also when $n\equiv 1 \mod (k-1)$ for $n>(k-1)(2^{k-2}+1)$. In the first case, take a vertex disjoint union of complete hypergraphs with $k-1$ vertices and add one more set, namely $[n]$. In the other case, take $m:=(n-1)/(k-1) \geq 2^{k-2}+1$ disjoint $(k-1)$-sets $A_1, \dots, A_m$ and an element $x$ such that $[n]=(\cup _{1\leq i\leq m}A_i)\cup \{ x\} $. Then define ${\mathcal H}$ as the union of complete hypergraphs on the sets $A_i$'s together with the hyperedges of the form $A_i \cup \{ x\} $. If we do not insist on connectedness, then $(2^{k-2}+1)$-regular examples can be constructed for \emph{all} $n\geq k^2 2^{k-2}$. Finally, we prove a hypergraph version of Theorem~\ref{th:EG}. \begin{thm}\label{th:EGh}Let $n \geq k \geq 3$ and let ${\mathcal H}$ be an $n$-vertex hypergraph with no Berge cycle of length $k$ or longer. Then \[e({\mathcal H}) \leq 2+ \frac{n-1}{k-2}\left(2^{k-1}-2\right) .\] \end{thm} The bound in Theorem~\ref{th:EGh} is best possible when $n \equiv 1 \mod (k-2)$. Take $m:= (n-1)/(k-2)$ and disjoint sets $A_1 , \ldots, A_m$ of size $k-2$. Let $x$ be a new element, and set $[n] = (\cup _{1\leq i\leq m}A_i)\cup \{ x\}$. Define ${\mathcal H}$ to be the union of all sets $A$ such that there exists an $i$ with $A \setminus \{x\} \subseteq A_i$. Note that the 2-shadow $\partial_2({\mathcal H})$ is in the family ${\mathcal F}_{n,k}$ defined before Theorem~\ref{th:EG}. There are many exact results concerning the maximum size of uniform hypergraphs avoiding Berge paths and cycles, see the recent results of Ergemlidze et al.~\cite{EGyMSTZ} or one by the present authors~\cite{FKL}. \section{Dirac type conditions for hamiltonian hypergraphs} In this section, we present a proof for Theorem~\ref{diracH}. The proof method relies on reducing the hypergraph to a dense nonhamiltonian graph. In the next three subsections we collect some results about such graphs. Subsections~\ref{ss33} and~\ref{ss34} contain the proof for hypergraphs. \subsection{Classical tools}\label{ss0} Let $G$ be an $n$-vertex graph. The {\em hamilton-closure} of $G$ is the unique graph $C(G)$ of order $n$ that can be obtained from $G$ by recursively joining nonadjacent vertices with degree-sum at least $n$. \begin{thm}[Bondy, Chv\'atal~\cite{BC}]\label{BCthm} If $C(G)$ is hamiltonian, then so is $G$. \end{thm} A graph $G$ is called {\em hamiltonian-connected} if for any pair of vertices $x,y \in V(G)$ there is a hamiltonian $(x,y)$-path. The following corollary can be obtained from Theorem~\ref{BCthm} or from the classical result of P\'osa~\cite{Posa}: If for every pair of nonadjacent vertices $x,y \in V(G)$ we have $d(x) + d(y) \geq |V(G)| + 1$, then $G$ is hamiltonian-connected. \begin{cor}\label{cor1} If $e(G)\geq \binom{n}{2}-2$ and $n\geq 5$ then $G$ is hamiltonian-connected. \hfill\ifhmode\unskip\nobreak\fi\quad\ifmmode\Box\else\hfill$\Box$\fi \end{cor} We will need the following result about the structure of matchings in bipartite graphs. It is a well known fact in the theory of transversal matroids (but one can also give a short, direct proof finding an $M_3\subseteq M_1\cup M_2$). \begin{thm}\label{matroid} Let $G[X,Y]$ be a bipartite graph. Suppose that there is a matching $M_1$ in $G$ joining the vertices of $X_1\subseteq X$ and $Y_1\subseteq Y$. Suppose also that we have another matching $M_2$ with end vertices $X_2\subseteq X$ and $Y_2 \subseteq Y$ such that $Y_2\subseteq Y_1$. Then there exists a third matching $M_3$ from $X_3 \subseteq X$ to $Y_3 \subseteq Y$ such that \begin{equation*} Y_3=Y_1 \quad \text{and} \quad X_3\supseteq X_2. \end{equation*} \end{thm} \begin{thm}[Erd\H{o}s~\cite{Erdos}]\label{Erdos} Let $n, d$ be integers with $1 \leq d \leq \left \lfloor \frac{n-1}{2} \right \rfloor$, and set $h(n,d):={n-d \choose 2} + d^2$. If $G$ is a nonhamiltonian graph on $n$ vertices with minimum degree $\delta(G) \geq d$, then \[e(G) \leq \max\left\{ h(n,d),h(n, \left \lfloor \frac{n-1}{2} \right \rfloor)\right\}=:e(n,d).\] \end{thm} \subsection{A lemma for nonhamiltonian graphs}\label{ss11} The lemma below follows from a result of Voss~\cite{Voss} (and from the even more detailed descriptions by Jung~\cite{Jung} and Jung, Nara~\cite{JungNara}). We only state and use a weaker version and for completeness include a short proof. Define five classes of nonhamiltonian graphs. ${}$\enskip --- \enskip Let $n=2k+2$, $V=V_1\cup V_2$, $|V_1|=|V_2|=k+1$, ($V_1\cap V_2=\emptyset$). We say that $G\in {\mathcal G}_1$ if its edge set is the union of two complete graphs with vertex sets $V_1$ and $V_2$ and it contains at most one further edge $e_0$ (joining $V_1$ and $V_2$); \newline ${}$\enskip --- \enskip Let $n=2k+1$, $V=V_1\cup V_2$, $|V_1|=|V_2|=k+1$, $V_1\cap V_2=\{ x_0\}$. We say that $G\in {\mathcal G}_2$ if its edge set is the union of two complete graphs with vertex sets $V_1$ and $V_2$; \newline ${}$\enskip --- \enskip Let $n=2k+2$, $V=V_1\cup V_2$, $|V_1|=k+1$, $|V_2|=k+2$, $V_1\cap V_2=\{ x_0\}$. We say that $G\in {\mathcal G}_3$ if its edge set is the union of a complete graph with vertex set $V_1$ and a $2$-connected graph $G_2$ with vertex set $V_2$ such that $\deg_G(v)\geq k$ for every vertex $v\in V$; \newline ${}$\enskip --- \enskip Let $n=2k+1$, $V=V_1\cup V_2$, $|V_1|=k$, $|V_2|=k+1$, ($V_1\cap V_2=\emptyset$). We say that $G\in {\mathcal G}_4$ if $V_2$ is an independent set, and its edge set contains all edges joining $V_1$ and $V_2$; \newline ${}$\enskip --- \enskip Let $n=2k+2$, $V=V_1\cup V_2$, $|V_1|=k$, $|V_2|=k+2$, ($V_1\cap V_2=\emptyset$). We say that $G\in {\mathcal G}_5$ if $V_2$ contains at most one edge $e_0$ and $\deg_G(v)\geq k$ for every vertex $v\in V$ (so its edge set contains all but at most two edges joining $V_1$ and $V_2$). \begin{lem}\label{5G}Let $k\geq 3$ be an integer, $n \in \{ 2k+1, 2k+2\} $. Suppose that $G$ is an $n$-vertex nonhamiltonian graph with $\delta(G) \geq k = \lfloor(n-1)/2\rfloor$, $V:= V(G)$. Then $G\in {\mathcal G}_1\cup \dots \cup {\mathcal G}_5$. \end{lem} \begin{proof} Suppose first that $G$ is not 2-connected. Then there exist two blocks $B_1, B_2$ of $G$ (i.e., $B_i$ is a maximal 2-connected subgraph or a $K_2$) which are {\em endblocks}, i.e., for $i=1,2$ there is a vertex $v_i\in B_i$ such that $V(B_i)\setminus \{ v_i\}$ does not meet any other block. Then $\{v\} \cup N(v) \subset V(B_i)$ for all $v\in V(B_i)\setminus \{ v_i\}$, so an endblock has at least $k+1$ vertices and if $|V(B_i)|=k+1$ then it is a clique. If $B_1$ and $B_2$ are disjoint then we get $n=2k+2$, and $G\in {\mathcal G}_1$. If $B_1$ and $B_2$ meet, then $G$ has no other blocks, and $G\in {\mathcal G}_2\cup {\mathcal G}_3$. Suppose now that $G$ is 2-connected. By the second part of Dirac's theorem (Theorem~\ref{th:D}), the length of a longest cycle $C$ of $G$ is at least $2k$. If $|V(C)| = n-1$, assume $C=v_1 \ldots v_{n-1} v_1$ and $v_n \notin V(C)$. Then $v_n$ has at least $k$ neighbors in $C$, with no two of them appearing consecutively (otherwise we could extend $C$ to a hamiltonian cycle). Without loss of generality, let $N(v_n) = \{v_1, v_3, \ldots, v_{2k-1}\}$. If for some $i<j$ such that $v_i, v_j \in N(v_n)$, $v_{i+1}v_{j+1} \in E(G)$, then we obtain the hamiltonian cycle $v_1 v_2 \ldots v_i v_n v_j v_{j-1} \ldots v_{i+1}v_{j+1} v_{j+2} \ldots v_{n-1}v_1$. Therefore the vertices in $C$ of even parity, together with $v_n$, form an independent set. In case of $n=2k+1$ we got $G\in {\mathcal G}_4$. If $n=2k+2$ then in the same way we get that $\{v_{2k+1}\}\cup \{ v_2, v_4, \dots, v_{2k-2} \}$ together with $v_n$ is also independent, so the set $\{ v_2, ..., v_{2k-2}\} \cup \{ v_{2k}, v_{2k+1}, v_n \}$ contains only the edge $v_{2k}v_{2k+1}$, $G\in {\mathcal G}_5$. Finally, consider the case that $|V(C)| = n-2$, (i.e., $n=2k+2$) and let $x, y \notin V(C)$. We claim that $xy\notin E(G)$. Indeed, suppose to the contrary, that $xy \in E(G)$. Without loss of generality, $A:=\{v_1, v_3, \ldots, v_{2k-3}\} \subseteq N(x)$ or $(A\setminus \{ v_{2k-3}\}) \cup \{ v_{2k-2}\})\subseteq N(x)$. Note that for any $v_i \in N(x)$, $\{v_{i-2}, v_{i-1}, v_{i+1}, v_{i+2}\} \cap N(y) = \emptyset$ (indices are taken modulo $2k$), because we can remove a segment of $C$ with at most 3 vertices and replace it with a segment with at least 4 containing the edge $xy$. This leads to a contradiction because there is not enough room on the $2k$-cycle $C$ to distribute the at least $k-1$ vertices of $N(y)-x$. If $xy \notin E(G)$ then without loss of generality $N(x) = \{v_1, v_3,\ldots v_{2k-1}\}$. Then the set $\{x\} \cup \{v_2, \ldots, v_{2k}\}$ is an independent set. If $y v_i \in E(G)$ for some $i \in \{2, 4, \ldots, {2k}\}$, then because $y$ has $k$ neighbors in $C$ and no two of them appear consecutively, $N(y) = \{v_2, v_4, \ldots, v_{2k}\}$, and we obtain a hamiltonian cycle by replacing the segment $v_1v_2v_3v_4$ of $C$ with the path $v_1 x v_3 v_2 y v_4$. Therefore $V_2:=\{v_2, v_4, \ldots, v_{2k}\} \cup \{x, y\}$ is an independent set of size $k+2$, and so $G\in {\mathcal G}_5$. \end{proof} \subsection{A maximality property of the graphs in ${\mathcal G}_1\cup\ldots\cup{\mathcal G}_5$} Let $G\in {\mathcal G}_1\cup \dots \cup {\mathcal G}_5$ be a graph. Delete a set of edges ${\mathcal A}$ from $E(G)$ where $|{\mathcal A}|\leq 1$ for $G\in {\mathcal G}_2\cup {\mathcal G}_3\cup {\mathcal G}_4$ and $|{\mathcal A}|\leq 2$ for $G\in {\mathcal G}_1\cup {\mathcal G}_5$. Then add a set of new edges ${\mathcal B}$ as defined below: \newline ${}$\enskip --- \enskip For $G\in {\mathcal G}_1$, $|{\mathcal B}|=2$ and it consists of any two disjoint pairs joining $V_1$ and $V_2$; \newline ${}$\enskip --- \enskip for $G\in {\mathcal G}_2\cup{\mathcal G}_3$, $|{\mathcal B}|=1$ and it consists of any pair $x_1x_2$ joining $V_1\setminus \{ x_0\}$ and $V_2\setminus \{ x_0\}$ (here $x_i\in V_i$); \newline ${}$\enskip --- \enskip for $G\in {\mathcal G}_4$, $|{\mathcal B}|=1$ and it consists of any pair contained in $V_2$; \newline ${}$\enskip --- \enskip and for $G\in {\mathcal G}_5$, $|{\mathcal B}|=2$ and it consists of any two distinct pairs contained in $V_2$. \begin{lem}\label{5Gmodified} If $k\geq 6$, then the graph $\left(E(G)\setminus {\mathcal A}\right) \cup {\mathcal B}$ defined by the above process is hamiltonian, except if $G\in {\mathcal G}_3$, $x_0$ has exactly two neighbors $x_2$ and $y_2$ in $V_2$, ${\mathcal A}= \{ x_0y_2\}$, ${\mathcal B}= \{ x_1x_2\}$, and $G[V_2\setminus \{ x_0\}]$ is either a $K_{k+1}$ or misses only the edge $x_2y_2$. \end{lem} \begin{proof} If $G\in {\mathcal G}_1$ and we add two disjoint edges $x_1x_2$ and $y_1y_2$ joining $V_1$ and $V_2$ ($x_1, y_1\in V_1$) then to form a hamiltonian cycle we need an $x_1x_2$ path $P_1$, and a $y_1y_2$ path $P_2$ of length $k$, $V(P_i)=V_i$ and $E(P_i)\subset E(G)\setminus {\mathcal A}$. Such paths exist because the graph $G[V_i]\setminus {\mathcal A}$ has at least $\binom{k+1}{2}-2$ edges, so it satisfies the condition of Corollary~\ref{cor1}. If $G\in {\mathcal G}_2\cup {\mathcal G}_3$ and we add an edge $x_1x_2$ joining $V_1\setminus\{x_0\}$ and $V_2\setminus\{x_0\}$ then we need paths $P_1$, $P_2$ of length $|V_i|-1$ joining $x_i$ to $x_0$, $V(P_i)=V_i$ and $E(P_i)\subset E(G)\setminus {\mathcal A}$. If $G[V_i]\setminus {\mathcal A}$ satisfies the condition of Corollary~\ref{cor1} then we can find $P_i$. The only missing case is when $|V_2|=k+2$ (so $G\in {\mathcal G}_3$). Let $G_2$ be the graph on $|V_2|+1$ vertices obtained from $G[V_2]\setminus {\mathcal A}$ by adding a new vertex $x_2'$ and two edges $x_0x_2'$ and $x_2x_2'$. If $G_2$ has a hamiltonian cycle $C$ then it should contain $x_0x_2'$ and $x_2x_2'$ so the rest of the edges of $C$ can serve as $P_2$ we are looking for. Consider the hamilton-closure $C(G_2)$ and apply Theorem~\ref{BCthm} to $G_2$. Since the degrees of $V_2\setminus \{ x_0\}$ in $G_2$ are at least $k-1$ and $2(k-1)\geq k+3= |V(G_2)|$, $C(G_2)$ is a complete graph on $V_2\setminus \{ x_0\}$. So $C(G_2)$ is hamiltonian unless the only neighbors of $x_0$ in $G_2$ are $x_2$ and $x_2'$. Hence $N_G(x_0)\cap V_2= \{ x_2, y_2\}$ and ${\mathcal A}=\{ x_0y_2\}$. The last case is when $G\in {\mathcal G}_5$, $|{\mathcal A}|=2$, ${\mathcal B}= \{ e_1,e_2\}$ (two distinct edges inside $V_2$). (The proofs of the other cases, especially when $G\in {\mathcal G}_4$ are easier). We create a graph $H_0$ from $G$ as follows: Delete the edge $e_0$ (if it exists), delete the edges of ${\mathcal A}$ joining $V_1$ and $V_2$, add two new vertices $z_1, z_2$ to $V_1$ and join $z_i$ to the endpoints of $e_i$. We obtain the graph $H$ by adding all possible $\binom{k+2}{2}$ pairs from $V_1\cup \{ z_1,z_2\}$ to $H_0$. If $H$ is hamiltonian then its hamiltonian cycle must use only edges of $H_0$ (because $V_2$ is an independent set of size $k+2$ in $H$). If the graph $H_0$ is hamiltonian then its hamiltonian cycle must use the two edges of the degree $2$ vertex $z_i$, so $\left(G\setminus (\{ e_0\}\cup {\mathcal A})\right) \cup {\mathcal B}$ is hamiltonian as well. So it is sufficient to show that $H$ has a hamiltonian cycle. Let $A$ be the graph on $V(H)$ consisting of the edges of ${\mathcal A}$ joining $V_1$ and $V_2$ together with the (at most) two missing pairs $E(K(V_1, V_2))\setminus E(G)$. We will again apply Theorem~\ref{BCthm} to $H$, so consider the hamilton-closure $C(H)$. The degree $\deg_H(x)$ of an $x\in V_1$ is $(2k+3)-\deg_A(x)$. The degree $\deg_H(y)$ of a $y\in V_2$ is at least $|V_1|-\deg_A(y)= k-\deg_A(y)$. Since $\deg_A(x)+\deg_A(y)\leq |E(A)|+1\leq 5$ we get for $k\geq 6$ that \[ \deg_H(x)+ \deg_H(y)\geq (3k+3)-\left(\deg_A(x)+\deg_A(y)\right)\geq 3k-2 \geq 2k+4= |V(H)|. \] So $C(H)$ contains the complete bipartite graph $K(V_1,V_2)=K_{k,k+2}$. Then it is really a simple task to find a hamiltonian cycle in $C(H)$ and therefore $\left(E(G)\setminus {\mathcal A}\right) \cup {\mathcal B}$ is hamiltonian. \end{proof} \subsection{Proof of Theorem~\ref{diracH}, reducing the hypergraph to a dense graph}\label{ss33} Fix ${\mathcal H}$ to be an $n$ vertex hypergraph satisfying the minimum degree condition. We will find a hamiltonian Berge cycle in ${\mathcal H}$. Recall that $H=\partial_2({\mathcal H})$ denote the 2-shadow of ${\mathcal H}$, a graph on $V=V({\mathcal H})$. Define a bipartite graph $B:= B[E({\mathcal H}), E(H)]$ with parts $E({\mathcal H})$ and $E(H)$ and with edges $\{ h, xy\}$ where a hyperedge $h\in E({\mathcal H})$ is joined to the graph edge $xy\in E(H)$ if $\{ x, y\} \subseteq h$. In the case of $\{ x,y\} \in {\mathcal H}$ we consider the edge $xy\in E(H)$ and $\{ x,y\} \in E({\mathcal H})$ as two distinct objects of $B$, so $B$ is indeed a bipartite graph (with $|E({\mathcal H})|+| E(H)|$ vertices and no loops). Let $M$ be a maximum matching of $B$. So $M$ can be considered as a partial injection of maximum size, i.e., a bijection $\phi$ between two subsets ${\mathcal M}\subseteq E({\mathcal H})$ and ${\mathcal E}\subseteq E(H)$ such that $|{\mathcal M}|=|{\mathcal E}|$, $\phi(m)\subseteq m$ for $m\in {\mathcal M}$ (and $\phi(m_1)\neq \phi(m_2)$ for $m_1\neq m_2$). Consider the subgraph $G=(V,{\mathcal E})$ of $H$. Then $G$ does not have a hamiltonian cycle, otherwise by replacing the edges of a hamiltonian cycle with their corresponding matched hyperedges in $M$, we obtain a hamiltonian Berge cycle in ${\mathcal H}$ (with representative vertices in the same order). In this subsection we are going to prove that \begin{equation}\label{eq331} \delta(G) \geq \lfloor (n-1)/2 \rfloor:=k. \end{equation} Since $G$ has no hamiltonian cycle and $k\geq 7$, if~\eqref{eq331} holds, then by Lemma~\ref{5G}, $G\in {\mathcal G}_1\cup \dots\cup {\mathcal G}_5$. We will consider this case and prove the remainder of Theorem~\ref{diracH} in the next subsection. Let ${\mathcal H}_2:= E({\mathcal H})\cap \partial_2({\mathcal H})$, the set of 2-element edges of ${\mathcal H}$. We may assume that among all maximum sized matchings of $B$ the matching $M$ maximizes $|{\mathcal M}\cap {\mathcal H}_2|$. \begin{claim}\label{cl33} ${\mathcal H}_2\subseteq {\mathcal M}$, $\partial_2({\mathcal M})= E(H)$, and every $m\in E({\mathcal H})\setminus {\mathcal M}$ induces a complete graph in $G$. \end{claim} \begin{proof} If $m\in E({\mathcal H})$ contains an edge $e\in E(H)\setminus E(G)$ then one can enlarge the matching $M$ by adding $\{m,e\}$ to it, if it is possible. Since $M$ is maximal, it cannot be enlarged, so $m\in {\mathcal M}$. This implies the second and the third statements. We also obtained that if $\{x,y\}\in E({\mathcal H})$ then $xy\in E(G)$, so $\phi(m)=xy$ for some $m\in {\mathcal M}$. In case of $|m|>2$ we can replace the pair $\{ m, xy\}$ by the pair $\{ \{ x,y\}, xy\}$ in $M$ and the new matching covers more edges from ${\mathcal H}_2$ than $M$ does (in the graph $B$). So $|m|=2$, all members of ${\mathcal H}_2$ must belong to ${\mathcal M}$. \end{proof} To continue the proof of Theorem~\ref{diracH}, let $d:= \delta (G)$, $v\in V$ such that $D:=N_G(v)$, $|D|=d$. Since $G$ is not hamiltonian, Theorem~\ref{th:D} gives $d\leq k$. Let ${\mathcal H}_v$ denote the set of hyperedges of ${\mathcal H}$ containing the vertex $v$, ($\deg_{\mathcal H}(v)=|{\mathcal H}_v|$), and split it into two parts, ${\mathcal H}_v= {\mathcal D}\cup {\mathcal L}$ where ${\mathcal D}:= \{ e\in E({\mathcal H}): v\in e \subseteq \{ v\} \cup D\}$ and ${\mathcal L}:= {\mathcal H}_v\setminus {\mathcal D}$. Split ${\mathcal D}$ further into three parts according to the sizes of its edges, ${\mathcal D}={\mathcal D}^- \cup {\mathcal D}_2\cup {\mathcal D}_3$ where ${\mathcal D}_i:= \{ e\in {\mathcal D}: |e|=i\}$ (for $i=2,3$) and ${\mathcal D}^-:= {\mathcal D}\setminus \left({\mathcal D}_2\cup {\mathcal D}_3\right)$. Since ${\mathcal D}$ can have at most $2^d$ members and we handle ${\mathcal D}_2$ and ${\mathcal D}_3$ separately we get \begin{equation}\label{eq333} |{\mathcal D}| \leq 2^d-d-\binom{d}{2} +|{\mathcal D}_2|+ |{\mathcal D}_3|. \end{equation} Recall that the matching $M$ in the bipartite graph $B$ can be considered as a bijection $\phi: {\mathcal M} \to {\mathcal E}$, where ${\mathcal M}\subseteq E({\mathcal H})$ and ${\mathcal E}\subseteq E(H)$. Define another matching $M_2$ in $B$ by an injection $\phi_2: {\mathcal D}_2\cup {\mathcal D}_3\to E(G)$ as follows. If $m\in {\mathcal M} \cap ({\mathcal D}_2\cup {\mathcal D}_3)$ then $\phi_2(m):= \phi(m)$. In particular, since ${\mathcal D}_2\subseteq {\mathcal M}$, if $\{ v,x\} \in {\mathcal D}_2$ then $\phi_2(\{ v,x\})= vx$. If $m= \{ v,x,y\} \in {\mathcal D}_3 \setminus {\mathcal M}$ then let $\phi_2(m):= xy$. Since $\phi_2({\mathcal D}_2\cup {\mathcal D}_3)\subseteq E(G)$ we can apply Theorem~\ref{matroid} to the matchings $M$ and $M_2$ in $B$ with $X_1:= {\mathcal M}$, $Y_1:= E(G)$, and $X_2:= {\mathcal D}_2\cup {\mathcal D}_3$. So there exists a subfamily ${\mathcal L}_3\subseteq {\mathcal H} \setminus ({\mathcal D}_2\cup {\mathcal D}_3)$ and a bijection $\phi_3: ({\mathcal D}_2\cup {\mathcal D}_3 \cup {\mathcal L}_3)\to E(G)$. The matching $M'$ defined by $\phi_3$ is also a largest matching of $B$. Every $m\in {\mathcal L}$ has an element $x\notin D$, so $vx\notin E(G)$. If $m$ is not matched in $M'$, then we add $\{m, vx\}$ to $M'$ to get a larger matching. Hence $m \in {\mathcal L}_3$. These yield \begin{equation}\label{eq334} |{\mathcal L}|\leq |{\mathcal L}_3| = e(G)-|{\mathcal D}_2|-|{\mathcal D}_3|. \end{equation} Summing up~\eqref{eq333} and~\eqref{eq334}, then using the lower bound for $|{\mathcal H}_v|$ and the upper bound of Theorem~\ref{Erdos} for $e(G)$ we obtain \begin{equation* 2^k +1 \leq \deg_{\mathcal H}(v) \leq 2^d-\binom{d+1}{2} +e(n,d) \end{equation*} The inequality $2^k +1 \leq 2^d-\binom{d+1}{2} +e(n,d)$ does not hold for $n\geq 15$ and $d< k$, e.g., for $(n,k,d) = (16, 7, 6)$, the right hand side is only $64-21+85 = 128$. This completes the proof of $d=k$. \subsection{Proof of Theorem~\ref{diracH}, the end}\label{ss34} We may assume that $G\in {\mathcal G}_1\cup \dots\cup {\mathcal G}_5$ by Lemma~\ref{5G}, $\phi$ is a bijection $\phi: {\mathcal M} \to E(G)$ with $\phi(m)\subseteq m$ where ${\mathcal M}\subseteq E({\mathcal H})$, and Claim~\ref{cl33} holds. Let ${\mathcal L}_v$ denote the set of edges $m\in {\mathcal H}$ containing an edge $vy$ of $E(H)\setminus E(G)$. Note that ${\mathcal L}_v\subseteq {\mathcal M}$. If $\deg_G(v)=k$, then the family ${\mathcal L}_v$ is non-empty, otherwise $\deg_{{\mathcal H}}(v) \leq 2^k$. Call a graph $F$ with vertex set $V$ a {\em Berge graph of} ${\mathcal H}$ if $E(F)\subseteq E(H)$, and there exists a subhypergraph ${\mathcal F}\subseteq {\mathcal H}$, and a bijection $\psi:{\mathcal F} \to E(F)$ such that $\psi(m)\subseteq m$ for each $m\in {\mathcal F}$. We are looking for a Berge graph of ${\mathcal H}$ having a hamiltonian cycle. In particular, the graph $G$ is a Berge graph of ${\mathcal H}$ and it is almost hamiltonian. We will show that a slight change to $G$ yields a hamiltonian Berge graph of ${\mathcal H}$. If $G\in {\mathcal G}_2\cup {\mathcal G}_3$ then choose any $v\in V_1\setminus \{ x_0\}$ and let $m\in {\mathcal L}_v$. There exists an edge $vy\in \left(E(H)\setminus E(G)\right)$ contained in $m$. Then $y\in V_2\setminus \{ x_0\}$. The graph $\left(E(G)\setminus \{ \phi(m) \}\right)\cup \{ vy\} $ is a Berge graph of ${\mathcal H}$ (we map $m$ to the edge $vy$ instead of $\phi(m)$). According to Lemma~\ref{5Gmodified} (with ${\mathcal A}:= \{ \phi(m)\}$ and ${\mathcal B}:= \{ vy\}$) it is hamiltonian except if we run into the only exceptional case: $x_0$ has exactly two $G$-neighbors $x_2$ and $y_2$ in $V_2$, $vy=vx_2$, and $\phi(m)=x_0y_2$. In this case $m$ contains $\{ x_0,v, x_2, y_2\}$ so it can be avoided by choosing $y:= y_2$ instead of $y=x_2$. If $G\in {\mathcal G}_4$ then we argue in a very similar way. Choose any $v\in V_2$ and let $m\in {\mathcal L}_v$ containing an edge $vy\in \left(E(H)\setminus E(G)\right)$. Then $y\in V_2$ and the graph $\left(E(G)\setminus \{ \phi(m) \}\right)\cup \{ vy\} $ is a Berge graph of ${\mathcal H}$ that is hamiltonian by Lemma~\ref{5Gmodified} with ${\mathcal A}:= \{ \phi(m)\}$ and ${\mathcal B}:= \{ vy\}$. From now on we may suppose that $n=2k+2$ so $|{\mathcal L}_v|\geq 2$ for $\deg_G(v)=k$. If $G\in {\mathcal G}_1$ then define ${\mathcal M}_{1,2}$ as the members of ${\mathcal M}$ meeting both $V_1$ and $V_2$. The minimum degree condition on ${\mathcal H}$ implies that $|{\mathcal M}_{1,2}|\geq 2$. Since ${\mathcal M}_{1,2}$ can have at most one member of size $2$, we can choose an $m_1$, $|m_1|\geq 3$. By symmetry we may suppose that $|m_1\cap V_1|\geq 2$ and let $x_2\in V_2\cap m_1$. Choose an element $y\in V_2$, $y\notin e_0$, $y\neq x_2$. Since $|{\mathcal L}_y|\geq 2$ we can choose an $m_2\in {\mathcal M}_{1,2}$ such that $m_1\neq m_2$ and $y\in m_2$. Take any pair $\{ y_1,y\}\subseteq m_2$ with $y_1\in V_1$. Then one can choose an $x_1\in m_1\cap V_1$ so that $x_1\neq x_2$. So the pairs $\{ x_1, x_2\}\subseteq m_1$ and $\{ y_1, y\}\subseteq m_2$ are disjoint. Lemma~\ref{5Gmodified} with ${\mathcal A}:= \{ \phi(m_1), \phi(m_2)\}$ and ${\mathcal B}:= \{ x_1x_2, y_1y\}$ implies that the graph $\left(E(G)\setminus {\mathcal A} \right)\cup {\mathcal B} $ is a hamiltonian Berge graph of ${\mathcal H}$. If $G\in {\mathcal G}_5$ then $|{\mathcal L}_v|\geq 2$ for any $v\in V_2\setminus e_0$ and for all members $m$ of ${\mathcal L}_v$ we have $|m\cap V_2|\geq 2$. Fix $v\in V_2\setminus e_0$ and let $m_1$ be an arbitrary member of ${\mathcal L}_v$. Choose a pair $\{ v,v'\} \subseteq m_1\cap V_2$. Fix another vertex $u\in V_2\setminus (e_0\cup \{ v,v'\})$ and let $m_2$ be an arbitrary member of ${\mathcal L}_u$. Choose a pair $\{ u,u'\} \subseteq m_2\cap V_2$. Then $u\notin \{ v,v'\}$ so the pairs $\{ u,u'\}$ and $\{ v,v'\} $ are distinct. Again, apply Lemma~\ref{5Gmodified} with ${\mathcal A}:= \{ \phi(m_1), \phi(m_2)\}$ and ${\mathcal B}:= \{ uu', vv'\}$. This completes the proof of Theorem~\ref{diracH}. \hfill\ifhmode\unskip\nobreak\fi\quad\ifmmode\Box\else\hfill$\Box$\fi \begin{rem} We can also show that all extremal examples are slight modifications of the four types of the sharpness examples described after Theorem~\ref{diracH}. \end{rem} \section{Dirac-type conditions for long Berge cycles}\label{secCk} In this section we prove Theorem~\ref{th:mindeg_path} for Berge paths and Theorem~\ref{th:mindeg_cycle} for Berge cycles. In fact we prove the two statements simultaneously. \begin{proof}[Proof of Theorems~\ref{th:mindeg_path} and~\ref{th:mindeg_cycle}] Suppose that $\delta({\mathcal H})\geq 2^{k-2}+1$, $k\geq 3$ and that ${\mathcal H}$ has no Berge cycle of length $k$ or longer. We will show that it contains a Berge path of length $k-1$ (thus establishing Theorem~\ref{th:mindeg_path}) and then that $\delta({\mathcal H})= 2^{k-2}+1$ (which completes the proof of Theorem~\ref{th:mindeg_cycle}). Choose a longest Berge path in ${\mathcal H}$ according the following rules. We say that a Berge path with edges $\{e_1,\dots,e_s\}$ is \emph{better} than a Berge path with edges $\{f_1,\dots,f_t\}$ if \newline\indent${}$\quad a) $s>t$ or \newline\indent${}$\quad b) $s=t$ and $\sum |e_i|<\sum |f_j|$. Consider a best Berge path ${\mathcal P}$ in ${\mathcal H}$. Let the base vertices of the path be $v_1, v_2, \dots, v_p$. Let $e_1, \dots ,e_{p-1}$ be the edges of the path ($v_{i}, v_{i+1}\in e_i$). First, we show that $p\geq k-1$. (In fact, $p\geq k$ follows but that will be proved later). Indeed, let ${\mathcal H}^{(p)}$ be the hypergraph consisting of the edges of ${\mathcal H}$ containing $v_p$, contained in $\{ v_1, \dots, v_p \}$ and also the edges of the path, i.e., \[ E({\mathcal H}^{(p)}):=\{ e\in E({\mathcal H}): v_p\in e\subseteq \{ v_1, \dots, v_p\}\} \cup \{e_1, \dots ,e_{p-1} \}. \] Then for $p\leq k-2$ (and $k\geq 3$) we have \[ |E({\mathcal H}^{(p)})|\leq 2^{p-1}+ (p-1)\leq 2^{k-2}< \delta({\mathcal H})\leq \deg_{\mathcal H}(v_p). \] So there exits an edge $f$ in $E({\mathcal H})\setminus E({\mathcal H}^{(p)})$ containing $v_p$. Then $e_1, \dots ,e_{p-1},f$ form a Berge path longer than ${\mathcal P}$, a contradiction. Now we have $p\geq k-1$, so we can define $W:=\{ v_1, \dots, v_{k-1}\}$. Let ${\mathcal P}_1$ be the subhypergraph consisting of the first $k-1$ edges of ${\mathcal P}$, $E({\mathcal P}_1):= \{ e_1, ..., e_{k-1}\}$ (if $p=k-1$ we take ${\mathcal P}_1:={\mathcal P}$). Let ${\mathcal H}_1$ be the subhypergraph of ${\mathcal H}$ consisting of the edges incident to $v_1$. \begin{claim}\label{cl:52} Every edge $f\in E({\mathcal H}_1)\setminus E({\mathcal P}_1)$ is contained in $W:=\{ v_1, \dots, v_{k-1}\}$. \end{claim} \begin{proof} First, we show that every edge $f\in E({\mathcal H}_1)\setminus E({\mathcal P}_1)$ avoids $\{v_k, \dots, v_p\}$. Otherwise, if there exists an edge $f\in E({\mathcal H}_1)\setminus E({\mathcal P}_1)$ such that $f\cap \{v_k, \dots, v_p\} \neq \emptyset$, then suppose that $v_i$ has the minimum index ($k\leq i\leq p $) such that $v_i$ is a vertex of such an $f$. Then $e_1, \dots, e_{i-1}$ and $f$ are forming a Berge cycle of length $i$, since these hyperedges are all distinct and $v_1,v_i\in f$. Finally, suppose that there is an edge $f\in E({\mathcal H}_1)\setminus E({\mathcal P}_1)$ such that $v\in f$, $v\notin W$. Then $v\notin \{ v_1, \dots, v_p\}$ so the path $f, e_1, \dots, v_p$ is longer than ${\mathcal P}$, a contradiction. \end{proof} Let ${\mathcal K}$ be the family of all $2^{k-2}$ subsets of $W$ that contain $v_1$. We claim there is a one-to-one mapping $\varphi$ from ${\mathcal H}_1 \setminus e_{k-1}$ to ${\mathcal K}$. The existence of such a $\varphi$ implies \begin{equation}\label{eq41} \delta({\mathcal H})\leq \deg_{\mathcal H}(v_1) \leq 2^{k-2} + 1. \end{equation} If an edge $e$ of ${\mathcal H}_1$ satisfies $e \subseteq W$, then let $\varphi(e) = e$. Otherwise, let ${\mathcal A} \subseteq {\mathcal H}_1$ be the set of the edges of ${\mathcal H} \setminus \{e_{k-1}\}$ that contain both $v_1$ and some vertex outside of $W$. By Claim~\ref{cl:52}, each $e \in {\mathcal A}$ must be some edge $e_i$ in ${\mathcal P}_1$. Hence it remains to show that all elements of ${\mathcal A}$ can be mapped to distinct elements of ${\mathcal K}$ that are not edges of ${\mathcal H}$. Observe that if $e_i\in {\mathcal A}$ then $\{v_i,v_{i+1}\}\notin {\mathcal H}$. Otherwise, we get a better path by replacing $e_i$ by $\{v_i,v_{i+1}\}$. Also, for $1\leq i\leq k-2$, $e_i\in {\mathcal A}$ implies $v_1\in e_i$ and $\{v_i,v_{i+1}\}\subset e_i$. Since $e_i\not \subset W$ we get $|e_i|\geq 4$ for $i\geq 2$. We also obtain that in case of $i\geq 3$, $e_i\in {\mathcal A}$ we have $\{v_1, v_i,v_{i+1}\}\notin {\mathcal P} $, and moreover $\{v_1,v_i,v_{i+1}\} \notin {\mathcal H}$ since otherwise we get a better path by replacing $e_i$ by $\{v_1, v_i,v_{i+1}\}$. For $3\leq i\leq k-2$ (and $e_i\in {\mathcal A}$) define $\varphi(e_i)$ as $\{v_1, v_i,v_{i+1}\}$. If $e_2\in {\mathcal A}$ and $\{ v_1, v_2, v_3 \}\not \in {\mathcal H}$ then we proceed as above, $\varphi(e_2):=\{ v_1, v_2, v_3 \}$. Otherwise, if $e_2\in {\mathcal A}$ (so $|e_2|\geq 4$) and $\{ v_1, v_2, v_3 \}\in {\mathcal H}$ then $\{ v_1, v_2, v_3 \}\in {\mathcal P}$ too (otherwise, we get a better path by replacing $e_2$ by $\{v_1, v_2,v_3\}$). We get $e_1=\{ v_1, v_2, v_3 \}$ (and $e_1 \subset e_2$). We claim that $\{ v_1, v_3\}\notin {\mathcal H}$. Otherwise we rearrange the base vertices of the path ${\mathcal P}$ by exchanging $v_1$ and $v_2$ (and get the order $v_2, v_1, v_3, \dots, v_{p}$) and observe that the Berge path $\{ v_2, v_1, v_3\}, \{ v_1, v_3 \}, e_3, \dots, e_{p-1}$ is better than ${\mathcal P}$, a contradiction. So in this case $\varphi(e_2):= \{ v_1, v_3 \}$. Finally, if $e_1\in {\mathcal A}$ then $\varphi(e_1):= \{ v_1, v_2 \}$, and the definition of $\varphi$ is complete. We have shown that $\deg_{\mathcal H}(v_1) \leq |{\mathcal H}_1 \setminus \{e_{k-1}\}| +1 \leq 2^{k-2} +1$. Equality holds, so $v_1 \in e_{k-1}$. In particular $e_{k-1}$ must exist, so ${\mathcal P}$ was a Berge path of length at least $k-1$. \end{proof} \medskip Our method works for multihypergraphs as well. If the maximum multiplicity of an edge is $\mu$, then the corresponding necessary bounds on the minimum degrees are $\mu 2^{k-2}+1$ or $\mu 2^{k-2}+2$, respectively. Indeed, suppose that $\delta({\mathcal F})\geq \mu 2^{k-2}+1$, $k\geq 3$ and that ${\mathcal F}$ has no Berge cycle of length $k$ or longer. Let ${\mathcal H}$ be the simple hypergraph obtained from ${\mathcal F}$ by keeping one copy from the multiple edges. We have $\delta({\mathcal H})\geq 2^{k-2}+1$. Then Theorems~\ref{th:mindeg_path} implies that ${\mathcal H}$ (and ${\mathcal F}$ as well) contain a Berge path with $k$ base vertices. As in the proof of Theorem~\ref{th:mindeg_cycle}, consider a best Berge path ${\mathcal P}$ in ${\mathcal H}$ with base vertices $v_1, v_2, \dots, v_p$ and edges $e_1, \dots ,e_{p-1}$. We have $p\geq k$. Then~\eqref{eq41} gives $\deg_{\mathcal H}(v_1)= 2^{k-2} + 1$ and we get $\deg_{\mathcal H}(v_1)= |{\mathcal H}_1 \setminus \{e_{k-1}\}| +1$. Since we also obtained $\{v_1, v_{k-1}, v_k\} \subset e_{k-1}$, the multiplicity of $e_{k-1}$ could not exceed $1$. So $\delta({\mathcal F})$ could not exceed $\mu 2^{k-2}+1$. \section{Maximum number of edges} \noindent{\em Proof of Theorem~\ref{th:EGh}.}\enskip Suppose that among all $n$-vertex hypergraphs with $c({\mathcal H})<k$ and $e({\mathcal H})$ edges our ${\mathcal H}$ is chosen so that $\sum_{e \in E({\mathcal H})} |e|$ is minimized. We claim that ${\mathcal H}$ is a downset, that is, for any $e \in E({\mathcal H})$ and $e' \subset e$, $e' \in E({\mathcal H})$. Indeed, if there exists a set $e'$ and a hypergedge $e$ such that $e' \subset e$ such that $e' \notin E({\mathcal H})$ and $e \in E({\mathcal H})$, then the hypergraph obtained by replacing $e$ with $e'$ also does not contain a Berge cyle of length $k$ or longer. This contradicts the choice of ${\mathcal H}$. Let $H = \partial_2{\mathcal H}$ be the 2-shadow of ${\mathcal H}$. Suppose that $H$ contains a cycle $C = v_1 v_2 \ldots v_\ell v_1$. Every edge $v_iv_{i+1}$ of $C$ is contained in a hyperedge of ${\mathcal H}$. But since ${\mathcal H}$ is a downset, the hyperedge $\{v_i, v_{i+1}\}$ is also contained in $E({\mathcal H})$. Therefore ${\mathcal H}$ also contains a (Berge) cycle of length $\ell$. Hence the graph $H$ contains no cycles of length at least $k$. Let $e_r({\mathcal H})$ be the number of hyperedges of ${\mathcal H}$ of size $r$. In $H$, every hyperedge $e$ of ${\mathcal H}$ is represented by a clique of order $|e|$, and so $e_r({\mathcal H})$ is at most the number of cliques of size $r$ in $H$. Since $c(H)<k$, each hyperedge contains at most $k-1$ vertices. By Theorem~\ref{cliques}, \[e({\mathcal H}) = e_0({\mathcal H})+ e_1({\mathcal H})+ \sum_{r=2}^{k-1} e_r({\mathcal H}) \leq 1+ n+ \sum_{r=2}^{k-1}\frac{n-1}{k-2}{k-1 \choose r} = 2+ \frac{n-1}{k-2}\left(2^{k-1}-2\right). \hfill\ifhmode\unskip\nobreak\fi\quad\ifmmode\Box\else\hfill$\Box$\fi \]
{ "timestamp": "2020-02-06T02:04:47", "yymm": "2002", "arxiv_id": "2002.01597", "language": "en", "url": "https://arxiv.org/abs/2002.01597", "abstract": "We consider two extremal problems for set systems without long Berge cycles. First we give Dirac-type minimum degree conditions that force long Berge cycles. Next we give an upper bound for the number of hyperedges in a hypergraph with bounded circumference. Both results are best possible in infinitely many cases.", "subjects": "Combinatorics (math.CO)", "title": "Berge cycles in non-uniform hypergraphs", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9814534376578004, "lm_q2_score": 0.8152324938410783, "lm_q1q2_score": 0.8001127335706679 }
https://arxiv.org/abs/1206.4740
Leinartas's partial fraction decomposition
These notes describe Leinartas's algorithm for multivariate partial fraction decompositions and employ an implementation thereof in Sage.
\section{Introduction} In \cite{Lein1978}, Le{\u\i}nartas\ gave an algorithm for decomposing multivariate rational expressions into partial fractions. In these notes I re-present Le{\u\i}nartas's\ algorithm, because it is not well-known, because its English translation \cite{Lein1978} is difficult to find, and because it is useful e.g. for computing residues of multivariate rational functions; see \cite[Chapter 3]{AiYu1983} and \cite{RaWi2012}. Along the way I include examples that employ an open-source implementation of Le{\u\i}nartas's\ algorithm that I wrote in Sage \cite{Sage}. The code can be downloaded from \href{http://www.alexraichev.org/research.html}{my website} and is currently under peer review for incorporation into the Sage codebase. For a different type of multivariate partial fraction decomposition, one that uses iterated univariate partial fraction decompositions, see \cite{Stou2008}. \section{Algorithm} Henceforth let $K$ be a field and $\overline{K}$ its algebraic closure. We will work in the factorial polynomial rings $K[X]$ and $\overline{K}[X]$, where $X = X_1,\ldots, X_d$ with $d \ge 1$. Le{\u\i}nartas's\ algorithm is contained in the constructive proof of the following theorem, which is \cite[Theorem 1]{Lein1978}\footnote{ Le{\u\i}nartas\ used $K = \mathbb{C}$, but that is an unnecessary restriction. By the way, Le{\u\i}nartas's\ article contains typos in equation (c) on the second page, equation (b) on the third page, and the equation immediately after equation (d) on the third page: the right sides of those equations should be multiplied by $P$. }. \begin{theorem}[Le{\u\i}nartas\ decompositon]\label{leinartas-decomp} Let $f = p/q$, where $p, q \in K[X]$. Let $q = q_1^{e_1} \cdots q_m^{e_m}$ be the unique factorization of $q$ in $K[X]$, and let $V_i = \{x \in \overline{K}^d : q_i(x) = 0 \}$, the algebraic variety of $q_i$ over $\overline{K}$. The rational expression $f$ can be written in the form \[ f = \sum_A \frac{p_A}{\prod_{i \in A} q_i^{b_i}}, \] where the $b_i$ are positive integers (possibly greater than the $e_i$), the $p_A$ are polynomials in $K[X]$ (possibly zero), and the sum is taken over all subsets $A \subseteq \{1,\ldots, m\}$ such that $\cap_{i \in A} V_i \neq \emptyset$ and $\{q_i : i \in A\}$ is algebraically independent (and necessarily $|A| \le d$). \end{theorem} Let us call a decomposition of the form above a \hl{Le{\u\i}nartas\ decomposition}. An immediate consequence of the theorem is the following. \begin{corollary} Every rational expression in $d$ variables can be represented as a sum of rational expressions each of whose denominators contains at most $d$ unique irreducible factors. \qed \end{corollary} Now for a constructive proof of the theorem. It involves two steps: decomposing $f$ via the Nullstellensatz and then decomposing each resulting summand via algebraic dependence. We need a few lemmas. The following lemma is a strengthening of the weak Nullstellensatz and is proved in \cite[Lemma 3.2]{DLMM2008}. \begin{lemma}[Nullstellensatz certificate]\label{null-cert} A finite set of polynomials $\{q_1, \ldots, q_m \} \subset K[X]$ has no common zero in $\overline{K}^d$ iff there exist polynomials $h_1, \ldots, h_m \in K[X]$ such that \[ 1 = \sum_{i=1}^m h_i q_i. \] Moreover, if $K$ is a computable field, then there is a computable procedure to check whether or not the $q_i$ have a common zero in $\overline{K}^d$ and, if not, return the $h_i$. \qed \end{lemma} Let us call a sequence of polynomials $h_i$ satisfying the equation above a \hl{Nullstellensatz certificate} for the $q_i$. Note that in contrast to the usual weak Nullstellensatz, here the polynomials $h_i$ are in $K[X]$ and not just in $\overline{K}[X]$. Some examples of computable fields are finite fields, $\mathbb{Q}$, finite degree extensions of $\mathbb{Q}$, and $\overline{\mathbb{Q}}$. Applying Lemma~\ref{null-cert} we get the following lemma \cite[Lemma 3]{Lein1978}. \begin{lemma}[Nullstellensatz decomposition]\label{null-decomp} Under the hypotheses of Theorem~\ref{leinartas-decomp}, the rational expression $f$ can be written in the form \[ f = \sum_A \frac{p_A}{\prod_{i \in A} q_i^{e_i}}, \] where the $p_A$ are polynomials in $K[X]$ (possibly zero) and the sum is taken over all subsets $A \subseteq \{1,\ldots, m\}$ such that $\cap_{i \in A} V_i \neq \emptyset$. \end{lemma} \begin{proof} If $\cap_{i=1}^m V_i \neq \emptyset$, then the result holds. Suppose now that $\cap_{i=1}^m V_i = \emptyset$. Then the polynomials $q_i^{e_i}$ have no common zero in $\overline{K}^d$. So by Lemma~\ref{null-cert} \[ 1 = h_1 q_1^{e_1} + \cdots + h_m q_m^{e_m} \] for some polynomials $h_i$ in $K[X]$. Multiplying both sides of the equation by $p/q$ yields \begin{align*} f &= \frac{p (h_1 q_1^{e_1} + \cdots + h_m q_m^{e_m})}{q_1^{e_1} \cdots q_m^{e_m}} \\ &= \sum_{i=1}^m \frac{p h_i}{q_1^{e_1} \cdots \widehat{q_i^{e_i}} \cdots q_m^{e_m}} \end{align*} Note that $p h_i \in K[X]$. Next we check each summand $p h_i/(q_1^{e_1} \cdots \widehat{q_i^{e_i}} \cdots q_m^{e_m})$ to see whether $\cap_{j \neq i } V_j \neq \emptyset$. If so, then stop. If not, then apply Lemma~\ref{null-cert} to ${q_1^{e_1}, \ldots \widehat{q_i^{e_i}}, \ldots q_m^{e_m}}$. Repeating this procedure until it stops yields the desired result. The procedure must stop, because each $V_i \neq \emptyset$ since each $q_i$ is irreducible in $K[X]$ and hence not a unit in $K[X]$. \end{proof} Let us call a decomposition of the form above a \hl{Nullstellensatz decomposition}. \begin{example} Consider the rational expression \[ f := \frac{X^2 Y + X Y^2 + X Y + X + Y}{X Y (X Y + 1)} \] in $\mathbb{Q}(X,Y)$. Let $p$ denote the numerator of $f$. The irreducible polynomials $X, Y, XY + 1 \in \mathbb{Q}[X, Y]$ in the denominator have no common zero in $\overline{\mathbb{Q}}^2$. So they have a Nullstellensatz certificate, e.g. $(-Y, 0, 1)$: \[ 1 = (-Y)X + (0)X + (1)(XY + 1). \] Applying the algorithm in the proof of Lemma~\ref{null-decomp} gives us a Nullstellensatz decomposition for $f$ in one iteration: \begin{align*} f =& \frac{p(-Y)}{Y(XY + 1)} + \frac{p(1)}{XY} \\ =& \frac{-p}{XY + 1} + \frac{p}{XY} \\ =& -X -Y -1 + \frac{1}{X Y + 1} + X + Y + 1 + \frac{X + Y}{XY} \\ & \text{(after applying the division algorithm)} \\ =& \frac{1}{X Y + 1} + \frac{X + Y}{XY}. \end{align*} Notice that \[ f = \frac{1}{X} + \frac{1}{Y} + \frac{1}{XY + 1} \] is also a Nullstellensatz decomposition for $f$. So Nullstellensatz decompositions are not unique. \end{example} The next lemma is a classic in computational commutative algebra; see e.g. \cite{Kaya2009}. \begin{lemma}[Algebraic dependence certificate]\label{algdep-cert} Any set $S$ of polynomials in $K[X]$ of size $> d$ is algebraically dependent. Moreover, if $K$ is a computable field and $S$ is finite, then there is a computable procedure that checks whether or not $S$ is algebraically dependent and, if so, returns an annihilating polynomial over $K$ for $S$. \qed \end{lemma} The next lemma is \cite[Lemma 1]{Lein1978}. \begin{lemma}\label{algdep-powers} A finite set of polynomials $\{q_1, \ldots, q_m\} \subset K[X]$ is algebraically dependent iff for all positive integers $e_1, \ldots, e_m$ the set of polynomials $\{q_1^{e_1}, \ldots, q_m^{e_m}\}$ is algebraically dependent. \end{lemma} \begin{proof} A set of polynomials $\{q_1, \ldots, q_m\} \subset K[X]$ is algebraically independent iff the $m \times d$ Jacobian matrix $J(q_1, \ldots, q_m) := \left( \frac{\partial q_i}{\partial X_j}\right)$ over the vector space $K(X)^d$ has rank $m$ (by the Jacobian criterion; see e.g. \cite{EhRo1993}) iff for all positive integers $e_i$ the matrix $\left(e_i q_i^{e_i -1} \frac{\partial q_i}{\partial X_j}\right) = J(q_1^{e_1}, \ldots, q_m^{e_m})$ over the vector space $K(X)^d$ has rank $m$ (since we are just taking scalar multiples of rows) iff the set of polynomials $q_1^{e_1}, \ldots, q_m^{e_m}$ is algebraically independent (by the Jacobian criterion). Moreover, if $\{q_1, \ldots, q_m\}$ is algebraically dependent, then any member of the (necessarily nonempty) elimination ideal \[ \langle Y_1 - q_1, \ldots, Y_m - q_m, Y_1^{e_1} - Z_1, \ldots, Y_m^{e_m} - Z_m \rangle_{K[X,Y,Z]} \cap K[Z_1, \ldots, Z_m], \] is an annihilating polynomial for $q_1^{e_1}, \ldots, q_m^{e_m}$. Moreover a finite basis for the elimination ideal can be computed using Groebner bases; see e.g. \cite[Chapter 3]{CLO2007}. \end{proof} Applying the previous two lemmas we get our final lemma \cite[Lemma 2]{Lein1978}. \begin{lemma}[Algebraic dependence decomposition]\label{algdep-decomp} Under the hypotheses of Theorem~\ref{leinartas-decomp}, the rational expression $f$ can be written in the form \[ f = \sum_A \frac{p_A}{\prod_{i \in A} q_i^{b_i}}, \] where the $b_i$ are positive integers (possibly greater than the $e_i$), the $p_A$ are polynomials in $K[X]$ (possibly zero), and the sum is taken over all subsets $A \subseteq \{1,\ldots, m\}$ such that $\{q_i : i \in A\}$ is algebraically independent (and necessarily $|A| \le d$). \end{lemma} \begin{proof} If $\{q_1, \ldots, q_m\}$ is algebraically independent, then the result holds. Notice that in this case $m \le d$ by Lemma~\ref{algdep-cert}. Suppose now that $\{q_1, \ldots, q_m\}$ is algebraically dependent. Then so is $\{q_1^{e_1}, \ldots, q_m^{e_m}\}$ by Lemma~\ref{algdep-powers}. Let $g = \sum_{\nu \in S} c_\nu Y^\nu \in K[Y_1, \ldots, Y_m]$ be an annihilating polynomial for $\{q_1^{e_1}, \ldots, q_m^{e_m}\}$, where $S \subset \mathbb{N}^m$ is the set of multi-indices such that $c_\nu \neq 0$. Choose a multi-index $\alpha \in S$ of smallest norm $||\alpha|| = \alpha_1 + \cdots + \alpha_m$. Then at $Q:= (q_1^{e_1}, \ldots, q_m^{e_m})$ we have \begin{align*} g(Q) &= 0 \\ c_\alpha Q^\alpha &= -\sum_{\nu \in S \setminus{\{\alpha\}}} c_\nu Q^\nu \\ 1 &= \frac{-\sum_{\nu \in S \setminus{\{\alpha\}}} c_\nu Q^\nu}{c_\alpha Q^\alpha}. \end{align*} Multiplying both sides of the last equation by $p / q$ yields \begin{align*} \frac{p}{q} &= \sum_{\nu \in S \setminus{\{\alpha\}}} \frac{-p c_\nu Q^\nu}{c_\alpha Q^{\alpha + 1}} \\ &= \sum_{\nu \in S \setminus{\{\alpha\}}} \frac{-p c_\nu}{c_\alpha} \prod_{i=1}^m \frac{q_i^{e_i \nu_i}}{q_i^{e_i(\alpha_i + 1)}} \\ \end{align*} Since $\alpha$ has the smallest norm in $S$ it follows that for any $\nu \in S \setminus{\{\alpha\}}$ there exists $i$ such that $\alpha_i + 1 \le \nu_i$, so that $e_i(\alpha_i + 1) \le e_i \nu_i$. So for each $\nu \in S \setminus{\{\alpha\}}$, some polynomial $q_i^{e_i (\alpha_i + 1)}$ in the denominator of the right side of the last equation cancels. Repeating this procedure yields the desired result. \end{proof} Let us call a decomposition of the form above an \hl{algebraic dependence decomposition}. \begin{example} Consider the rational expression \[ f := \frac{(X^2 Y^2 + X^2 Y Z + X Y^2 Z + 2 X Y Z + X Z^2 + Y Z^2)}{X Y Z (X Y + Z)} \] in $\mathbb{Q}(X,Y,Z)$. Let $p$ denote the numerator of $f$. The irreducible polynomials $X, Y, Z, XY + Z \in \mathbb{Q}[X,Y,Z]$ in the denominator are four in number, which is greater than the number of ring indeterminates, and so they are algebraically dependent. An annihilating polynomial for them is $g(A,B,C,D) = AB + C - D$. Applying the algorithm in the proof of Lemma~\ref{algdep-decomp} gives us an algebraic dependence decomposition for $f$ in one iteration: \begin{align*} f =& \sum_{\nu \in S \setminus{\{\alpha\}}} \frac{-p c_\nu Q^\nu}{c_\alpha Q^{\alpha + 1}} \\ & \text{where $Q = (X,Y,Z,XY + Z)$ and $\alpha = (0,0,0,1)$} \\ =& \frac{pQ^{(1,1,0,0)}}{Q^{(1,1,1,2)}} + \frac{pQ^{(0,0,1,0)}}{Q^{(1,1,1,2)}} \\ =& \frac{p}{Q^{(0,0,1,2)}} + \frac{p}{Q^{(1,1,0,2)}} \\ =& \frac{p}{Z (XY + Z)^2} + \frac{p}{XY(XY + Z)^2}. \end{align*} Notice that in this example the exponent 2 of the irreducible factor $XY + Z$ in the denominators of the decomposition is larger than the exponent 1 of $XY + Z$ in the denominator of $f$. Notice also that \[ f = \frac{1}{X} + \frac{1}{Y} + \frac{1}{Z} + \frac{1}{XY + Z} \] is also an algebraic dependence decomposition for $f$. So algebraic dependence decompositions are not unique. \end{example} Finally, here is Le{\u\i}nartas's\ algorithm. \begin{proof}[Proof of Theorem~\ref{leinartas-decomp}] First find the irreducible factorization of $q$ in $K[X]$. This is a computable procedure if $K$ is computable. Then decompose $f$ via Lemma~\ref{null-decomp}. Finally decompose each summand of the result via Lemma~\ref{algdep-decomp}. As highlighted above, the last two steps are computable if $K$ is. \end{proof} \begin{example} Consider the rational expression \[ f := \frac{2X^2 Y + 4X Y^2 + Y^3 - X^2 - 3 X Y - Y^2}{X Y (X + Y) (Y - 1)} \] in $\mathbb{Q}(X,Y)$. Computing a Nullstellensatz decomposition according to the proof of Lemma~\ref{null-decomp} with Nullstellensatz combination $1 = 0(X) + 1(Y) + 0(X + Y) -1(Y - 1)$ yields \begin{align*} f =& X - Y + \frac{Y^3 + X^2 - Y^2 + X}{X(Y-1)} + \frac{X^2 Y - 2X^2 - XY}{(X + Y)(Y - 1)} +\\ & \frac{-2X^3 - Y^3 - 2X^2 + Y^2}{X(X + Y)} + \frac{2X^2 Y - Y^3 + X^2 + 3X Y + Y^2}{XY(X + Y)}. \end{align*} Computing an algebraic dependence decomposition for the last term according to the proof of Lemma~\ref{algdep-decomp} with annihilating polynomial $g(A,B,C) = A + B - C$ for $(X, Y, X + Y)$ yields \begin{align*} & \frac{2X^2 Y - Y^3 + X^2 + 3X Y + Y^2}{XY(X + Y)} \\ &= 1 + \frac{2X^2 Y - Y^3 + X^2 + 3X Y + Y^2}{XY^2} + \frac{-2X^2 Y - XY^2 - X^2 - 3XY - Y^2}{Y^2 (X + Y)}. \end{align*} The two equalities taken together give us a Le{\u\i}nartas\ decomposition for $f$. Notice that \[ f = \frac{1}{X} + \frac{1}{Y} + \frac{1}{X + Y} + \frac{1}{Y - 1} \] is also a Le{\u\i}nartas\ decomposition of $f$. So Le{\u\i}nartas\ decompositions are not unique. \end{example} \begin{remark} In case $d=1$, Le{\u\i}nartas\ decompositions are unique once the fractions are written in lowest terms (and one disregards summand order). To see this, note that a Le{\u\i}nartas\ decomposition of a univariate rational expression $f = p/q$ must have fractions all of the form $p_i/q_i^{e_i}$, where $q = q_1^{e_1} \cdots q_m^{e_m}$ is the unique factorization of $q$ in $K[X]$. This is because two or more univariate polynomials are algebraically dependent (by Lemma~\ref{algdep-cert}). Assume without loss of generality here that $\deg(p) < \deg(q)$. It follows that if we have two Le{\u\i}nartas's\ decompositions of $p/q$, then we can write them in the form $a_1/q' + a_2/q'' = b_1/q' + b_2/q''$, where $q = q'q''$ with $q'$ and $q''$ coprime, $\deg(a_1), \deg(b_1) < \deg(q')$, and $\deg(a_2), \deg(b_2) < \deg(q'')$. Multiplying the equality by $q$ we get $a_1q'' + a_2q' = b_1q'' + b_2q'$. So $a_1 \equiv b_1 \pmod{q'}$ and $a_2 \equiv b_2 \pmod{q''}$. Thus $a_1 = b_1$ and $a_2 = b_2$. This observation used inductively demonstrates uniqueness. This argument fails in case $d \ge 2$, because then a Le{\u\i}nartas\ decomposition might not have fractions all of the form $p_i/q_i^{e_i}$. \end{remark} \begin{remark} A rational expression already with $\cap_{i=1}^m V_i \neq \emptyset$ and $\{q_1, \ldots, q_m\}$ algebraically independent, can not necessarily be decomposed further into partial fractions. For example, \[ f = \frac{1}{X_1 X_2 \cdots X_m} \in K(X_1, X_2, \ldots, X_d), \] with $m \le d$ can not equal a sum of rational expressions whose denominators each contain fewer than $m$ of the $X_i$. Otherwise, multiplying the equation by $X_1 X_2 \cdots X_m$ would yield \[ 1 = \sum_{i\in B} h_i X_i \] for some $h_i \in K[X]$ and some nonempty subset $B\subseteq \{1, 2, \ldots, m\}$, a contradiction to Lemma~\ref{null-cert} since $\{X_i : i\in B\}$ have a common zero in $\overline{K}^d$, namely the zero tuple. \end{remark} \bibliographystyle{amsalpha}
{ "timestamp": "2012-06-26T02:07:37", "yymm": "1206", "arxiv_id": "1206.4740", "language": "en", "url": "https://arxiv.org/abs/1206.4740", "abstract": "These notes describe Leinartas's algorithm for multivariate partial fraction decompositions and employ an implementation thereof in Sage.", "subjects": "Commutative Algebra (math.AC); Combinatorics (math.CO)", "title": "Leinartas's partial fraction decomposition", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9814534365728416, "lm_q2_score": 0.8152324938410783, "lm_q1q2_score": 0.8001127326861742 }
https://arxiv.org/abs/1001.2949
A Property of the Frobenius Map of a Polynomial Ring
Let R be a ring of polynomials in a finite number of variables over a perfect field k of characteristic p>0 and let F:R\to R be the Frobenius map of R, i.e. F(r)=r^p. We explicitly describe an R-module isomorphism Hom_R(F_*(M),N)\cong Hom_R(M,F^*(N)) for all R-modules M and N. Some recent and potential applications are discussed.
\section{Introduction} The main result of this paper is Theorem \ref{main} which is a type of adjointness property for the Frobenius map of a polynomial ring over a perfect field. The interest in this fairly elementary result comes from its striking recent applications (see \cite[Sections 5 and 6]{WZ} and \cite{YZ}) and also from the fact that despite extensive inquiries we have not been able to find it in the published literature. Especially interesting is the application in \cite{YZ} where a striking new result on local cohomology modules in characteristic $p>0$ is deduced from Theorem \ref{main}. There is no doubt that the same result is true in characteristic 0, but the only currently known (to us) proof is in characteristic $p>0$, based on Theorem \ref{main}. This has motivated what promises to be a very interesting search for some new technique to extend the result in \cite{YZ} to characteristic 0. We believe the results of this paper hold a potential for further applications and we discuss some of them in the last section. \section{Preliminaries} Let $R$ be a regular (but not necessarily local) UFD, let $R_s$ and $R_t$ be two copies of $R$ (the subscripts stand for {\it source} and {\it target}) and let $F:R_s\to R_t$ be a finite ring homomorphism. Let $$F_*:R_t{\rm -mod}\to R_s{\rm -mod}$$ be the restriction of scalars functor (i.e. $F_*(M)$ for every $R_t$-module $M$ is the additive group of $M$ regarded as an $R_s$-module via $F$) and let $$F^!, F^*:R_s{\rm -mod}\to R_t{\rm -mod}$$ be the functors defined by $F^!(N)={\rm Hom}_{R_s}(R_t,N)$ and $F^*(N)=R_t\otimes_{R_s}N$ for every $R_s$-module $N$. $F^*(N)$ and $F^!(N)$, for an $R_s$-module $N$, have a structure of $R_s$-module via the natural $R_s$-action on $R_t$, while $F_*(M)$, for an $R_t$-module $M$, retains its old $R_t$-module structure (from before the restriction of scalars). Thus $F^*(N), F^!(N)$, and $F_*(M)$ are both $R_s$- and $R_t$-modules. It is well-known that $F^*$ is left-adjoint to $F_*$, i.e. there is an isomorphism of $R_t$-modules \begin{alignat}{2}\label{adj4} {\rm Hom}_{R_t}(F^*(N), M)&\cong &&{\rm Hom}_{R_s}(N,F_*(M))\\ f&\mapsto &&(n\mapsto f(1\otimes n))\notag \\ (r\otimes n\mapsto rf(n))&\leftarrow &&f\notag \end{alignat} which is functorial in $M$ and $N$ (see, for example, \cite[II.5, p.110]{Ha}). This holds without any restrictions on $R$. The main result of this paper is based on a different type of adjointness (see \ref{adj3} below) which is certainly not well-known (we could not find a reference). Unlike (\ref{adj4}) it is not quite canonical but depends on a choice of a certain isomorphism $\phi$ (described in (\ref{phi}) below). And it requires the conditions we imposed on $R$, namely, a regular UFD. Since $R_t$, being regular, is locally Cohen-Macaulay, it follows from the Auslander-Buchsbaum theorem that the projective dimension of $R_t$ as $R_s$-module is zero, i.e. $R_t$ is projective, hence locally free, as an $R_s$-module. It is a standard fact that in this case $F^!$ is right adjoint to $F_*$, \cite[Ch. 3, Exercise 6.10]{Ha} i.e. for every $R_t$-module $M$ and every $R_s$-module $N$ there is an $R_t$-module isomorphism \begin{alignat}{1}\label{adj1} {\rm Hom}_{R_s}(F_*(M),N)\cong&{\rm Hom}_{R_t}(M,F^!(N))\\ (m\mapsto \mathfrak f(1))\leftarrow& f\notag \end{alignat} where $\mathfrak f=f(m):R_t\to N$. This isomorphism is functorial both in $M$ and in $N$. Consider the $R_t$-module $H\stackrel{\rm def}{=}{\rm Hom}_{R_s}(R_t,R_s)$. Since $R_t$ is a locally free $R_s$-module of finite rank, it is a standard fact \cite[Ch 2, Exercise 5.1(b)]{Ha} that there is an isomorphism of functors \begin{align}\label{f^!} F^!\cong H\otimes_{R_s}-, \end{align} i.e. for every $R_s$-module $N$ there is an $R_t$-module isomorphism \begin{align}\label{f^!'} H\otimes_{R_s}N&\cong {\rm Hom}_{R_s}(R_t,N)\\ \notag h\otimes n&\mapsto (r\mapsto h(r)n) \end{align} and this isomorphism is functorial in $N$. Replacing $F^!$ by $H\otimes_{R_s}-$ in (\ref{adj1}) produces an $R_t$-module isomorphism \begin{align}\label{adj2} {\rm Hom}_{R_s}(F_*(M),N)\cong {\rm Hom}_{R_t}(M, H\otimes_{R_s}N). \end{align} which is functorial in $M$ and $N$. It follows from \cite[Kor. 5.14]{KM} that locally $H$ is the canonical module of $R_t$. In particular, the rank of $H$ as $R_t$-module is 1 and therefore $H$ is $R_t$-module isomorphic to some ideal $I$ of $R_t$. According to \cite[Kor. 6.13]{KM} the quotient $R_t/I$ is locally Gorenstein of dimension dim$R_t-1$, hence $I$ has pure height 1. Since $R_t$ is a UFD, $I$ is principal, i.e. there is an $R_t$-module isomorphism \begin{align}\label{phi} \phi:R_t\to H. \end{align} Hence $\phi$ induces an isomorphism of functors \begin{align}\label{f^*} R_t\otimes_{R_s}-\stackrel{\phi\otimes{\rm id}}{\cong} H\otimes_{R_s}-. \end{align} Replacing $H\otimes_{R_s}-$ by $F^*=R_t\otimes_{R_s}-$ in (\ref{adj2}) via (\ref{f^*}) produces an $R_t$-module isomorphism \begin{align}\label{adj3} {\rm Hom}_{R_s}(F_*(M),N)\cong {\rm Hom}_{R_t}(M, F^*(N)) \end{align} which is functorial in $M$ and $N$. While the isomorphisms (\ref{adj1}), (\ref{f^!}), (\ref{f^!'}) and (\ref{adj2}) are canonical, the isomorphisms (\ref{phi}), (\ref{f^*}) and (\ref{adj3}) depend on a choice of $\phi$. Every $R_t$-module isomorphism $\phi':R_t\to H$ is obtained from a fixed $\phi$ by multiplication by an invertible element of $R_t$, i.e. $\phi'=c\cdot \phi$ where $c\in R_t$ is invertible. Therefore the isomorphisms (\ref{f^*}) and (\ref{adj3}) are defined up to multiplication by an invertible element of $R_t$. Since the element $1\in R_t$ generates the $R_s$-submodule $F(R_s)$ of $R_t$ and does not belong to $\mathfrak mR_t$ for any maximal ideal $\mathfrak m$ of $R_s$, the $R_s$-module $R_t/F(R_s)$ is projective. Hence applying the functor Hom$_{R_s}(-,R_s)$ to the injective map $F:R_s\to R_t$ produces a surjection $H\to {\rm Hom}_{R_s}(R_s, R_s).$ Composing it with the standard $R_s$-module isomorphism ${\rm Hom}_{R_s}(R_s, R_s)\stackrel{\psi\mapsto \psi(1)}{\longrightarrow}R_s$ produces an $R_s$-module surjection $H\to R_s$. Composing this latter map with the isomorphism $\phi$ from (\ref{phi}) produces an $R_s$-module surjection \begin{align}\label{psi} \psi:R_t\to R_s. \end{align} If $N$ is an $R_s$-module, applying $-\otimes_{R_s}N$ to $\psi$ produces an $R_s$-module surjection \begin{align}\label{psiN} \psi_N:R_t\otimes_{R_s}N\to N. \end{align} It is not hard to check that the isomorphism (\ref{adj3}) sends $g\in {\rm Hom}_{R_t}(M,F^*(N))$ to $\psi_N\circ g\in {\rm Hom}_{R_s}(F_*(M),N).$ \section{The Main Result} For the rest of this paper $R$ is a ring of polynomials in a finite number of variables over a perfect field $k$ of characteristic $p>0$ and $F:R_s\to R_t$ is the standard Frobenius map, i.e. $F(r)=r^p$. The main result of this paper (Theorem \ref{main}) is an explicit description of the isomorphism (\ref{adj3}) in terms of polynomial generators of $R$. The recent applications \cite{WZ, YZ} crucially depend on this explicit description. We keep the notation of the preceding section. Let $x_1,\dots, x_n$ be some polynomial generators of $R$ over the field $k$, i.e. $R=k[x_1,\dots,x_n]$. We denote the multi-index $i_1,\dots, i_n$ by $\bar i$. Since $k$ is perfect, $R_t$ is a free $R_s$-module on the $p^n$ monomials $e_{\bar i}\stackrel{\rm def}{=}x_1^{i_1}\cdots x_n^{i_n}$ where $0\leq i_j<p$ for every $j$. If $i_j<p-1$, then $x_je_{\bar i}=e_{\bar i'}$ where $\bar i'$ is the multi-index $i_1,\dots, i_{j-1},i_j+1, i_{j+1},\dots, i_n$. If $i_j=p-1$, then $x_je_{\bar i}=x_j^pe_{\bar i'}$ where $\bar i'$ is the multi-index $i_1,\dots, i_{j-1},0,i_{j+1},\dots,i_n$. Let $\{f_{\bar i}\in H|0\leq i_j<p\ {\rm for\ every}\ j\}$ be the dual basis of $H$, i.e. $f_{\bar i}(e_{\bar {i'}})=1$ if $\bar i=\bar {i'}$ and $f_{\bar i}(e_{\bar {i'}})=0$ otherwise. If $i_j>0$, then $x_jf_{\bar i}=f_{\bar i'}$ where $\bar i'$ is the multi-index $i_1,\dots, i_{j-1},i_j-1,i_{j+1},\dots, i_n$. If $i_j=0$, then $x_jf_{\bar i}=x_j^pf_{\bar i'}$ where $\bar i'$ is the multi-index $i_1,\dots,i_{j-1},p-1,i_{j+1},\dots,i_n$. Denote the multi-index $p-1,\dots, p-1$ by $\overline{p-1}$ and let $\overline{p-1}-\bar i$ be the multi-index $p-1-i_1,\dots, p-1-i_n$. \begin{proposition}\label{Fphi}$(\rm cf.$ \cite[Remark 3.11]{BSTZ}$)$ The $R_s$-linear isomorphism $\phi:R_t\to H$ that sends $e_{\bar i}$ to $f_{\overline{p-1}-\bar i}$ is $R_t$-linear. \end{proposition} \emph{Proof.} All we have to show is that $\phi(x_je_{\bar i})=x_j\phi(e_{\bar i})$ for all indices $j$ and multi-indices $\bar i$. This is straightforward from the definition of $\phi$ and the above description of the action of $x_j$ on $e_{\bar i}$ and $f_{\bar i}$.\qed \smallskip Clearly, $F^*(N)=R_t\otimes_{R_s}N=\oplus_{\bar i}(e_{\bar i}\otimes_{R_s}N)$, as $R_s$-modules. Thus every $R_s$-linear map $g:M\to F^*(N)$ has the form $g=\oplus_{\bar i}(e_{\bar i}\otimes_{R_s}g_{\bar i})$ where $g_{\bar i}:M\to N$ are $R_s$-linear maps (i.e. $g_{\bar i}:F_*(M)\to N$ because $M$ with its $R_s$-module structure is $F_*(M)$). \begin{lemma} \label{F^*} An $R_s$-linear map $g:M\to F^*(N)$ as above is $R_t$-linear if and only if $g_{\bar i}(-)=g_{\overline{p-1}}(e_{\overline{p-1}-\bar i}(-))$ for every $\bar i$ (here $e_{\overline{p-1}-\bar i}\in R_t$ acts on $(-)\in F_*(M)$ via the $R_t$-module structure on $F_*(M)$). \end{lemma} \emph{Proof.} Assume $g$ is $R_t$-linear. Then $g$ commutes with multiplication by every element of $R_t$ and in particular with multiplication by $e_{\overline{p-1}-\bar i}$. That is $g(e_{\overline{p-1}-\bar i}(-))=e_{\overline{p-1}-\bar i}g(-)$. Since $e_{\overline{p-1}-\bar i}e_{\bar i}=e_{\overline{p-1}}$, the $\overline{p-1}$-component of $e_{\overline{p-1}-\bar i}g(-)$ is $e_{\overline{p-1}}\otimes_{R_s}g_{\bar i}(-)$ while the $\overline{p-1}$-component of $g(e_{\overline{p-1}-\bar i}(-))$ is $e_{\overline{p-1}}\otimes_{R_s}g_{\overline{p-1}}(e_{\overline{p-1}-\bar i}(-)).$ Since the two $\overline{p-1}$-components are equal, $g_{\bar i}(-)=g_{\overline{p-1}}(e_{\overline{p-1}-\bar i}(-))$. Conversely, assume $g_{\bar i}(-)=g_{\overline{p-1}}(e_{\overline{p-1}-\bar i}(-))$ for every $\bar i$. To show that $g$ is $R_t$-linear all one has to show is that $g$ commutes with the action of every $x_j\in R_t$, i.e. the $\bar i$-components of $g(x_j(-))$ and $x_jg(-)$ are the same for all $\bar i$. If $i_j>0$, then the $\bar i$-component of $x_jg(-)$ is $e_{\bar i}\otimes_{R_s}g_{\bar i'}(-)$ where $\bar i'$ is the index $i_1,\dots,i_{j-1},i_j-1,i_{j+1},\dots,i_n$ while the $\bar i$-component of $g(x_j(-))$ is $e_{\bar i}\otimes_{R_s}g_{\bar i}(x_j(-))$. But the fact that $g_{\bar i}(-)=g_{\overline{p-1}}(e_{\overline{p-1}-\bar i}(-))$ for every $\bar i$ implies $g_{\bar i'}(-)=g_{\bar i}(x_j(-))$. If $i_j=0$, then the $\bar i$-component of $x_jg(-)$ is $e_{\bar i''}{\otimes_{R_s}}(_sx_jg_{\bar i''}(-))$ where $_sx_j$ denotes the element of $R_s$ corresponding to $x_j\in R_t$ (i.e. $F(_sx_j)=x_j^p$) and $\bar i''$ is the index $i_1,\dots, i_{j-1},p-1,i_{j+1},\dots, i_n$ while the $\bar i$-component of $g(x_j(-))$ is $g_{\bar i}(x_j(-))$. But the fact that $g_{\bar i}(-)=g_{\overline{p-1}}(e_{\overline{p-1}-\bar i}(-))$ for every $\bar i$ implies $_sx_jg_{\bar i''}(-)=g_{\bar i}(x_j(-))$.\qed \smallskip Finally we are ready for the main result of the paper which is the following explicit description of the isomorphism (\ref{adj3}) for the Frobenius map. \begin{theorem}\label{main} For every $R_t$-module $M$ and every $R_s$-module $N$ there is an $R_t$-linear isomorphism \begin{alignat}{2} {\rm Hom}_{R_s}(F_*(M),N)&\cong& &{\rm Hom}_{R_t}(M,F^*(N))\notag\\ g_{\overline{p-1}}(-)&\leftarrow &&(g=\oplus_{\bar i}(e_{\bar i}\otimes_{R_s}g_{\bar i}(-)))\\ g&\mapsto &&\oplus_{\bar i}(e_{\bar i}\otimes_{R_s}g(e_{\overline{p-1}-\bar i}(-))). \end{alignat} \end{theorem} \emph{Proof.} As is pointed out at the end of the preceding section, the isomorphism (\ref{adj3}) sends $g\in {\rm Hom}_{R_t}(M,F^*(N))$ to $\psi_N\circ g\in {\rm Hom}_R(F_*(M),N).$ It is straightforward to check that with $\phi$ as in Proposition \ref{Fphi} the map $\psi$ of (\ref{psi}) sends $e_{\overline{p-1}}$ to 1. This implies that $\psi_N\circ g=g_{\overline{p-1}}$ and finishes the proof that formula (11) produces the isomorphism of (\ref{adj3}) in one direction. The fact that the other direction of this isomorphism is according to formula (12) has essentially been proven in Lemma \ref{F^*}. \qed \section{Potential Applications} The notion of $F$-finite modules was introduced in \cite{L}. An $F$-finite module is determined by a generating morphism, i.e. an $R$-module homomorphism $\beta:M\to F^*(M)$ where $M$ is a finite $R$-module. For simplicity assume the $R$-module $M$ has finite length, i.e. the dimension of $M$ as a vector space over $k$, which we denote by $d$, is finite. Then the dimension of $F^*(M)$ as a vector space over $k$ equals $p^n\cdot d$. The number $p^n$ can be huge even for quite modest values of $p$ and $n$. Thus the target of $\beta$ may be a huge-dimensional vector space even if $d, p$ and $n$ are fairly small. But the $R$-module $F_*(M)$ has dimension $d$ as a $k$-vector space, hence the map $\tilde\beta:F_*(M)\to M$ that corresponds to $\beta$ under the isomorphism of Theorem \ref{main}, is a map between two $d$-dimensional vector spaces. Huge-dimensional vector spaces do not appear! This should make the map $\tilde\beta$ easier to manage computationally than the map $\beta$. Of course the isomorphism of Theorem \ref{main} means that many properties of $\beta$ could be detected in $\tilde\beta$; for example, $\beta$ is the zero map if and only if $\tilde\beta$ is. Therein lies the potential for using Theorem \ref{main} to make computations more manageable. The functors ${\rm Ext}^i_R(-, R)$ commute with both $F^*$ and $F_*$. More precisely, for every finitely generated $R_t$-module $M$ and every finitely generated $R_s$-module $N$ there exist functorial $R_t$-module isomorphisms $$\kappa_i:{\rm Ext}^i_{R_t}(F^*(N),R_t)\cong F^*( {\rm Ext}^i_{R_s}(N,R_s))$$$$\lambda_i:F_*({\rm Ext}^i_{R_t}(M,R_t))\cong {\rm Ext}^i_{R_s}(F_*(M),R_s).$$ Indeed, for $i=0, M=R_t$ and $N=R_s$ a straightforward composition of the $R_t$-module isomorphism $R_t\otimes_{R_s}R_s\stackrel{r'\otimes r\mapsto r'r^p}{\longrightarrow}R_t$ and the standard $R$-module isomorphisms ${\rm Hom}_{R}(R,R)\cong R$ for $R=R_t, R_s$ produce $\kappa_0$ while an additional $R_t$-module isomorphism ${\rm Hom}_{R_s}(R_t,R_s)=H\cong R_t$ produces $\lambda_0$. This implies that if $-$ stands for a complex of finite free $R$-modules, then ${\rm Hom}_R(-,R)$ commutes with $F^*$ and $F_*$. Taking now finite free resolutions of $M$ and $N$ in the categories of $R_t$- and $R_s$-modules respectively and considering that $F^*$ and $F_*$, being exact, commute with the operation of taking the (co)homology of complexes, we get $\kappa_i$ and $\lambda_i$ for every $i$. It is straightforward to check that there is a commutative diagram $$ \begin{CD} {\rm Hom}_{R_t}(F^*(N), M)@>>> {\rm Hom}_{R_s}(N,F_*(M))\\ @VVV @VVV\\ {\rm Hom}_{R_t}({\rm Ext}^i_{R_t}(M,R_t),{\rm Ext}^i_{R_t}(F^*(N),R_t))@.{\rm Hom}_{R_s}({\rm Ext}^i_{R_s}(F_*(M), R_s),{\rm Ext}^i_{R_s}(N,R_s))\\ @VVV @VVV\\ {\rm Hom}_{R_t}({\rm Ext}^i_{R_t}(M,R_t),F^*({\rm Ext}^i_{R_s}(N,R_s)))@>>>{\rm Hom}_{R_s}(F_*({\rm Ext}^i_{R_t}(M, R_t)),{\rm Ext}^i_{R_s}(N,R_s)) \end{CD} $$ where the top horizontal map is the isomorphism (\ref{adj4}), the bottom horizontal map is the isomorphism of Theorem \ref{main}, the bottom vertical maps are induced by $\kappa_i$ and $\lambda_i$ and, finally, the top vertical maps are defined by sending every $f\in {\rm Hom}(L,L')$ to the map ${\rm Ext}^i(L', R)\to {\rm Ext}^i(L,R)$ functorially induced by $f$. In other words, if a pair of maps $F^*(N)\to M$ and $N\to F_*(M)$ correspond to each other under (\ref{adj4}), then the induced maps ${\rm Ext}^i_{R_t}(M,R_t)\to F^*({\rm Ext}^i_{R_s}(N,R_s))$ and $F_*({\rm Ext}^i_{R_t}(M, R_t)) \to {\rm Ext}^i_{R_s}(N,R_s)$ correspond to each other under isomorphism of Theorem \ref{main}. Of course there is a similar diagram with the isomorphism of Theorem \ref{main} in the top row; we are not going to use it in the rest of the paper. \smallskip An important example of $F$-finite modules are local cohomology modules $H^i_I(R)$ of $R$ with support in an ideal $I\subset R$. A generating morphism of $H^i_I(R)$ is the composition $$f:{\rm Ext}^i_R(R/I, R)\to {\rm Ext}^i_R(F^*(R/I), R)\stackrel{\kappa_i}{\cong}F^*({\rm Ext}^i_R(R/I, R))$$ where the first map is induced by the isomorphism $F^*(R/I)\stackrel{r'\otimes\bar r\mapsto r'\bar r^p}{\cong}R/I^{[p]}$ followed by the natural surjecton $R/I^{[p]}\to R/I$. The following proposition holds the potential for simplifying computations involving local cohomology modules. \begin{proposition} Let $I\subset R$ be an ideal and let the composition $$f:{\rm Ext}^i_R(R/I, R)\to {\rm Ext}^i_R(F^*(R/I), R)\stackrel{\kappa_i}{\cong}F^*({\rm Ext}^i_R(R/I, R))$$ be as above.The map that corresponds to $f$ under the isomorphism of Theorem \ref{main} is the composition $$g:F_*({\rm Ext}^i_R(R/I, R))\stackrel{\lambda_i}{\cong}{\rm Ext}^i_R(F_*(R/I), R)\to {\rm Ext}^i_R(R/I, R)$$ where the second map in the composition is nothing but the map induced on ${\rm Ext}^i_R(-,R)$ by the natural Frobenius map $R/I\stackrel{r\mapsto r^p}{\cong} F_*(R/I)$. \end{proposition} \emph{Proof.} This is immediate from the above commutative diagram considering that the maps $F^*(R/I)\stackrel{r'\otimes r\mapsto r'r^p}{\longrightarrow} R/I$ and $R/I\stackrel{r\to r^p}{\rightarrow}F_*(R/I)$ correspond to each other under the isomorphism (\ref{adj4}).\qed \smallskip In addition to potential use in computation, the material of this section has already been used in a proof of a theoretical result \cite[Section 5]{WZ}.
{ "timestamp": "2010-01-18T04:36:20", "yymm": "1001", "arxiv_id": "1001.2949", "language": "en", "url": "https://arxiv.org/abs/1001.2949", "abstract": "Let R be a ring of polynomials in a finite number of variables over a perfect field k of characteristic p>0 and let F:R\\to R be the Frobenius map of R, i.e. F(r)=r^p. We explicitly describe an R-module isomorphism Hom_R(F_*(M),N)\\cong Hom_R(M,F^*(N)) for all R-modules M and N. Some recent and potential applications are discussed.", "subjects": "Commutative Algebra (math.AC)", "title": "A Property of the Frobenius Map of a Polynomial Ring", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9814534360303622, "lm_q2_score": 0.8152324915965392, "lm_q1q2_score": 0.8001127300410168 }
https://arxiv.org/abs/1706.06630
Improved upper bounds in the moving sofa problem
The moving sofa problem, posed by L. Moser in 1966, asks for the planar shape of maximal area that can move around a right-angled corner in a hallway of unit width. It is known that a maximal area shape exists, and that its area is at least 2.2195... - the area of an explicit construction found by Gerver in 1992 - and at most $2\sqrt{2}=2.82...$, with the lower bound being conjectured as the true value. We prove a new and improved upper bound of 2.37. The method involves a computer-assisted proof scheme that can be used to rigorously derive further improved upper bounds that converge to the correct value.
\section{Introduction} The \textbf{moving sofa problem} is a well-known unsolved problem in geometry, first posed by Leo Moser in 1966 \cite{unsolved-problems, moser}. It asks: \begin{quote} \textit{What is the planar shape of maximal area that can be moved around a right-angled corner in a hallway of unit width?} \end{quote} We refer to a connected planar shape that can be moved around a corner in a hallway as described in the problem as a \textbf{moving sofa shape}, or simply a \textbf{moving sofa}. It is known \cite{gerver} that a moving sofa of maximal area exists. The shape of largest area currently known is an explicit construction found by Joseph Gerver in 1992 \cite{gerver} (see also \cite{romik} for a recent perspective on Gerver's results), known as \textbf{Gerver's sofa} and shown in Figure~\ref{fig:gerver}. Its area is \textbf{Gerver's constant} $$\mu_\textrm{G} = 2.21953166\ldots,$$ an exotic mathematical constant that is defined in terms of a certain system of transcendental equations but which does not seem to be expressible in closed form. \begin{figure} \begin{center} \scalebox{0.85}{\includegraphics{gerversofa.pdf}} \caption{Gerver's sofa, conjectured to be the solution to the moving sofa problem. Its boundary is made up of 18 curves, each given by a separate analytic formula; the tick marks show the points of transition between different analytic pieces of the boundary.} \label{fig:gerver} \end{center} \end{figure} Gerver conjectured that $\mu_\textrm{G}$ is the largest possible area for a moving sofa, a possibility supported heuristically by the local-optimality considerations from which his shape was derived. Gerver's construction provides a lower bound on the maximal area of a moving sofa. In the opposite direction, it was proved by Hammersley \cite{hammersley} in 1968 that a moving sofa cannot have an area larger than $2\sqrt{2}\approx 2.82$. It is helpful to reformulate these results by denoting $$ \mu_\textrm{MS} = \max \Big\{ \operatorname{area}(S)\,:\, S \textrm{ is a moving sofa shape} \Big\}, $$ the so-called \textbf{moving sofa constant}.\footnote{See Finch's book \cite[Sec.~8.12]{finch}; note that Finch refers to Gerver's constant $\mu_\textrm{G}$ as the ``moving sofa constant,'' but this terminology currently seems unwarranted in the absence of a proof that the two constants are equal.} The above-mentioned results then translate to the statement that $$ \mu_\textrm{G} \le \mu_\textrm{MS} \le 2\sqrt{2}. $$ The main goal of this paper is to derive improved upper bounds for $\mu_\textrm{MS}$. We prove the following explicit improvement to Hammersley's upper bound from 1968. \begin{thm}[New area upper bound in the moving sofa problem] \label{thm:new-upperbound} We have the bound \begin{equation} \label{eq:upperbound} \mu_\textrm{MS} \le 2.37. \end{equation} \end{thm} More importantly than the specific bound $2.37$, our approach to proving Theorem~\ref{thm:new-upperbound} involves the design of a computer-assisted proof scheme that can be used to rigorously derive even sharper upper bounds; in fact, our algorithm can produce a sequence of rigorously-certified bounds that converge to the true value $\mu_\textrm{MS}$ (see Theorems~\ref{thm:conv-moving-sofa} and~\ref{thm:asym-sharpness} below). An implementation of the scheme we coded in \texttt{C++} using exact rational arithmetic certifies $2.37$ as a valid upper bound after running for 480 hours on one core of a 2.3 GHz Intel Xeon E5-2630 processor. Weaker bounds that are still stronger than Hammersley's bound can be proved in much less time---for example, a bound of $2.7$ can be proved using less than one minute of processing time. Our proof scheme is based on the observation that the moving sofa problem, which is an optimization problem in an infinite-dimensional space of shapes, can be relaxed in many ways to arrive at a family of finite-dimensional optimization problems in certain spaces of polygonal shapes. These finite-dimensional optimization problems are amenable to attack using a computer search. Another of our results establishes new restrictions on a moving sofa shape of largest area. Gerver \cite{gerver} proved that such a largest area shape must undergo rotation by an angle of at least $\pi/3$ as it moves around the corner, and does not need to rotate by an angle greater than $\pi/2$. As explained in the next section, Gerver's argument actually proves a slightly stronger result with $\pi/3$ replaced by the angle $\beta_0 = \sec^{-1}(\mu_\textrm{G}) \approx 63.22^\circ$. We will prove the following improved bound on the angle of rotation of a moving sofa of maximal area. \begin{thm}[New rotation lower bound in the moving sofa problem] \label{thm:angle-bound} Any moving sofa shape of largest area must undergo rotation by an angle of at least $\sin^{-1}(84/85) \approx 81.203^\circ$ as it moves around the corner. \end{thm} There is no reason to expect this bound to be sharp; in fact, it is natural to conjecture that any largest area moving sofa shape must undergo rotation by an angle of $\pi/2$. As with the case of the bound \eqref{eq:upperbound}, our techniques make it possible in principle to produce further improvements to the rotation lower bound, albeit at a growing cost in computational resources. The paper is arranged as follows. Section~\ref{sec:theory} below defines the family of finite-dimensional optimization problems majorizing the moving sofa problem and develops the necessary theoretical ideas that set the ground for the computer-assisted proof scheme. In Section~\ref{sec:algorithm} we build on these results and introduce the main algorithm for deriving and certifying improved bounds, then prove its correctness. Section~\ref{sec:numerical} discusses specific numerical examples illustrating the use of the algorithm, leading to a proof of Theorems~\ref{thm:new-upperbound} and~\ref{thm:angle-bound}. The Appendix describes \texttt{SofaBounds}, a software implementation we developed as a companion software application to this paper \cite{sofabounds}. \paragraph{Acknowledgements.} Yoav Kallus was supported by an Omidyar Fellowship at the Santa Fe Institute. Dan Romik thanks Greg Kuperberg for a key suggestion that was the seed from which Proposition~\ref{prop:sofa-fg-bounds} eventually grew, and John Sullivan, Joel Hass, Jes\'us De Loera, Maria Trnkova and Jamie Haddock for helpful discussions. \section{A family of geometric optimization problems} \label{sec:theory} In this section we define a family of discrete-geometric optimization problems that we will show are in a sense approximate versions of the moving sofa problem for polygonal regions. Specifically, for each member of the family, the goal of the optimization problem will be to maximize the area of the intersection of translates of a certain finite set of polygonal regions in $\mathbb{R}^2$. It is worth noting that such optimization problems have been considered more generally in the computational geometry literature; see, e.g., \cite{harpeled-roy, mount-silverman}. We start with a few definitions. Set \begin{align*} H &= \mathbb{R} \times [0,1], \\ V &= [0,1] \times \mathbb{R}, \\ L_{\textrm{horiz}} &= (-\infty,1]\times[0,1], \\ L_{\textrm{vert}} &= [0,1]\times (-\infty,1], \\ L_0 &= L_{\textrm{horiz}} \cup L_{\textrm{vert}}. \end{align*} For an angle $\alpha \in [0,\pi/2]$ and a vector $\mathbf{u}=(u_1,u_2)\in\mathbb{R}^2$, denote \begin{align*} L_\alpha(\mathbf{u}) &= \Big\{ (x,y)\in\mathbb{R}^2\,:\, u_1 \le x\cos \alpha + y\sin \alpha \le u_1+1 \\& \qquad\qquad\qquad\qquad\qquad\qquad \textrm{ and \ } -x\sin\alpha + y \cos\alpha \le u_2+1 \Big\} \\ & \quad \cup \Big\{ (x,y)\in\mathbb{R}^2\,:\, x\cos \alpha + y\sin \alpha \le u_1+1 \\ & \qquad\qquad\qquad\qquad\qquad\qquad \textrm{ and } u_2\le -x\sin\alpha + y \cos\alpha \le u_2+1 \Big\}, \end{align*} For angles $\beta_1, \beta_2$, denote \begin{align*} B(\beta_1,\beta_2) &= \Big\{ (x,y)\in\mathbb{R}^2\,:\, 0 \le x\cos \beta_1 + y\sin \beta_1 \\& \qquad\qquad\qquad\qquad\qquad\qquad \textrm{ and \ } x\cos \beta_2 + y\sin \beta_2 \le 1 \Big\} \\ & \quad \cup \Big\{ (x,y)\in\mathbb{R}^2\,:\, x\cos \beta_1 + y\sin \beta_1 \le 1 \\& \qquad\qquad\qquad\qquad\qquad\qquad \textrm{ and \ } 0 \le x\cos \beta_2 + y\sin \beta_2 \Big\}, \end{align*} Geometrically, $L_\alpha(\mathbf{u})$ is the $L$-shaped hallway $L_0$ translated by the vector~$\mathbf{u}$ and then rotated around the origin by an angle of $\alpha$; and $B(\beta_1,\beta_2)$, which we nickname a ``butterfly set,'' is a set that contains a rotation of the vertical strip $V$ around the origin by an angle $\beta$ for all $\beta\in[\beta_1,\beta_2]$. See Fig.~\ref{fig:geometric-sets}. \begin{figure} \begin{center} \begin{tabular}{ccc} \scalebox{0.5}{\includegraphics{rot-trans-L.pdf}} & \scalebox{0.5}{\includegraphics{butterfly.pdf}} \\ (a) & (b) \end{tabular} \caption{ a) The rotated and translated $L$-shaped corridor $L_\alpha(u_1,u_2)$. (b) The ``butterfly set'' $B(\beta_1,\beta_2)$. } \label{fig:geometric-sets} \end{center} \end{figure} Next, let $\lambda$ denote the area measure on $\mathbb{R}^2$ and let $\lambda^*(X)$ denote the maximal area of any connected component of $X\subset\mathbb{R}^2$. Given a vector $\boldsymbol{\alpha} = (\alpha_1,\ldots,\alpha_k)$ of angles $0< \alpha_1<\ldots < \alpha_k < \pi/2$ and two additional angles $\beta_1,\beta_2 \in (\alpha_k,\pi/2]$ with $\beta_1\le\beta_2$, define \begin{align} g_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}(\mathbf{u}_1,\ldots,\mathbf{u}_k) &= \lambda^*\left( H \cap \bigcap_{j=1}^k L_{\alpha_j}(\mathbf{u}_j) \cap B(\beta_1,\beta_2) \right) \quad (\mathbf{u}_1,\ldots,\mathbf{u}_k \in \mathbb{R}^2), \label{eq:def-little-g} \\[5pt] \label{eq:def-G} G_{\boldsymbol{\alpha}}^{\beta_1,\beta_2} &= \sup \left\{ g_{\boldsymbol{\alpha}}(\mathbf{u}_1,\ldots,\mathbf{u}_k) \,:\, \mathbf{u}_1,\ldots,\mathbf{u}_k \in \mathbb{R}^{2} \right\}. \end{align} An important special case is $G_{\boldsymbol{\alpha}}^{\pi/2,\pi/2}$, which we denote simply as $G_{\boldsymbol{\alpha}}$. Note that $B(\beta_1,\pi/2)\cap H = H$, so in that case the inclusion of $B(\beta_1,\beta_2)$ in \eqref{eq:def-little-g} is superfluous. The problem of computing $G_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}$ is an optimization problem in $\mathbb{R}^{2k}$. The following lemma shows that the optimization can be performed on a compact subset of $\mathbb{R}^{2k}$ instead. \begin{lem}\label{lem:supmax} There exists a box $\Omega_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}=[a_1,b_1]\times\ldots \times [a_{2k},b_{2k}] \subset \mathbb{R}^{2k}$, with the values of $a_i,b_i$ being explicitly computable functions of $\boldsymbol{\alpha}, \beta_1, \beta_2$, such that \begin{equation} \label{eq:supmax} G_{\boldsymbol{\alpha}}^{\beta_1,\beta_2} = \max \left\{ g_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}(\mathbf{u}) \,:\, \mathbf{u} \in \Omega_{\boldsymbol{\alpha}}^{\beta_1,\beta_2} \right\}. \end{equation} \end{lem} \begin{proof} We will show that any value of $g_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}(\mathbf{u})$ attained for some $\mathbf{u}\in\mathbb{R}^2$ is matched by a value attained inside a sufficiently large box. This will establish that $g_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}(\mathbf{u})$ is bounded from above; the fact that it attains its maximum follows immediately, since $g_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}(\mathbf{u})$ is easily seen to be an upper semicontinuous function. Start by observing that for every interval $[x_1,x_2]$ and $0<\alpha<\pi/2$, there are intervals $I$ and $J$ such that if $(u,v)\in \mathbb{R}^2\setminus I\times J$, the set $\big([x_1,x_2]\times[0,1]\big)\cap L_\alpha(u,v)$ is either empty or is identical to $\big([x_1,x_2]\times[0,1]\big)\cap L_\alpha(u',v')$ for some $(u',v')\in I\times J$. Indeed, it can be checked that this is correct with the choices \begin{align*} I&=[x_1 \cos\alpha-1,x_2 \cos\alpha+\sin\alpha], \\ J&=[-x_2 \sin\alpha-1,-x_1 \sin\alpha+\cos\alpha]. \end{align*} We now divide the analysis into two cases. First, if $\beta_2<\pi/2$, then $H\cap B(\beta_1,\beta_2)\subseteq[-\tan \beta_2,\sec \beta_2]\times[0,1]$. Therefore, if we define $\Omega_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}=I_1\times J_1 \times I_2 \times J_2 \times \cdots\times I_k \times J_k$, where for each $1\le i\le k$, $I_i$ and $J_i$ are intervals $I,J$ as described in the above observation as applied to the angle $\alpha=\alpha_i$, then $g_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}(u_1,\ldots,u_{2k})$ is guaranteed to attain its maximum on $\Omega_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}$, since any value attained outside $\Omega_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}$ is either zero or matched by a value attained inside~$\Omega_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}$. Second, if $\beta_2=\pi/2$, then $H\cap B(\beta_1,\beta_2) = H$. The optimization objective function $g_\mathbf{\alpha}^{\beta_1,\beta_2}(\mathbf{u}_1,\ldots,\mathbf{u}_k)$ is invariant to translating all the rotated $L$-shaped hallways horizontally by the same amount (which corresponds to translating each variable $\mathbf{u}_j$ in the direction of the vector $(\cos\alpha_j, -\sin\alpha_j)$). Therefore, fixing an arbitrary $1\le j\le k$, any value of $g_\mathbf{\alpha}^{\beta_1,\beta_2}(\mathbf{u}_1,\ldots,\mathbf{u}_k)$ attained on $\mathbb{R}^{2k}$ is also attained at some point satisfying $\mathbf{u}_j=(0,u_{j,2})$. Furthermore, we can constrain $u_{j,2}$ as follows: first, when $u_{j,2}<-\tan\alpha_j-1$, then $L_{\alpha_j}(0,u_{j,2})\cap H$ is empty. Second, when $u_{j,2}>\sec\alpha_j$, then $L_{\alpha_j}(0,u_{j,2})\cap H$ is the union of two disconnected components, one of which is a translation of $\rotmat{\alpha_j}(H)\cap H$ and the other is a translation of $\rotmat{\alpha_j}(V)\cap H$. Since the largest connected component of $H \cap \bigcap_{j=1}^k L_{\alpha_j}(\mathbf{u}_j)$ is contained in one of these two rhombuses, and since the translation of the rhombus does not affect the maximum attained area, we see that any objective value attained with $\mathbf{u}_j=(0,u_{j,2})$, where $u_{j,2}>\sec\alpha_j$, can also be attained with $\mathbf{u}_j=(0,\sec\alpha_j)$. So, we may restrict $\mathbf{u}_j\in I_j\times J_j$, where $I_j = \{0\}$ and $J_j=[-\tan\alpha_j-1,\sec\alpha_j]$. Finally, when $\mathbf{u}_j\in I_j\times J_j$, we have $H\cap L_{\alpha_j}(\mathbf{u})\subseteq[\csc\alpha_j,\sec\alpha_j]\times[0,1]$, so we can repeat a procedure similar to the one used in the case $\beta=\pi/2$ above to construct intervals $I_i$ and $J_i$ for all $i\neq j$ to ensure that \eqref{eq:supmax} is satisfied. \end{proof} We now wish to show that the function $G_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}$ relates to the problem of finding upper bounds in the moving sofa problem. The idea is as follows. Consider a sofa shape $S$ that moves around the corner while rotating continuously and monotonically (in a clockwise direction, in our coordinate system) between the angles $0$ and $\beta \in [0,\pi/2]$. A key fact proved by Gerver \cite[Th.~1]{gerver} is that in the moving sofa problem it is enough to consider shapes being moved in this fashion. By changing our frame of reference to one in which the shape stays fixed and the $L$-shaped hallway $L_0$ is dragged around the shape while being rotated, we see (as discussed in \cite{gerver, romik}) that $S$ must be contained in the intersection \begin{equation} \label{eq:sx-intersections} S_{\mathbf{x}} = L_{\textrm{horiz}} \cap \bigcap_{0\le t\le \beta} L_t(\mathbf{x}(t)) \cap \Big(\mathbf{x}(\beta)+\rotmat{\beta}(L_{\textrm{vert}}) \Big), \end{equation} where $\mathbf{x}:[0,\beta]\to\mathbb{R}^2$ is a continuous path satisfying $\mathbf{x}(0)=(0,0)$ that encodes the path by which the hallway is dragged as it is being rotated, and where $\rotmat{\beta}(L_{\textrm{vert}})$ denotes $L_{\textrm{vert}}$ rotated by an angle of $\beta$ around $(0,0)$ (more generally, here and below we use the notation $\rotmat{\beta}(\cdot)$ for a rotation operator by an angle of $\beta$ around the origin). We refer to such a path as a \textbf{rotation path}, or a \textbf{$\beta$-rotation path} when we wish to emphasize the dependence on $\beta$. Thus, the area of $S$ is at most $\lambda^*(S_\mathbf{x})$, the maximal area of a connected component of $S_\mathbf{x}$, and conversely, a maximal area connected component of $S_\mathbf{x}$ is a valid moving sofa shape of area $\lambda^*(S_\mathbf{x})$. Gerver's result therefore implies that \begin{equation} \label{eq:sofaconst-characterization} \mu_\textrm{MS} = \sup \Big\{ \lambda^*(S_\mathbf{x}) \,:\, \mathbf{x}\textrm{ is a $\beta$-rotation path for some }\beta\in[0,\pi/2] \Big\}. \end{equation} It is also convenient to define $$ \mu_*(\beta) = \sup \Big\{ \lambda^*(S_\mathbf{x}) \,:\, \mathbf{x}\textrm{ is a $\beta$-rotation path} \Big\} \qquad (0 \le \beta \le \pi/2). $$ so that we have the relation \begin{equation} \label{eq:sofaconst-beta} \mu_\textrm{MS} = \sup_{0 < \beta \le \pi/2} \mu_*(\beta). \end{equation} Moreover, as Gerver pointed out in his paper, $\mu_*(\beta)$ is bounded from above by the area of the intersection of the horizontal strip $H$ and the rotation of the vertical strip $V$ by an angle $\beta$, which is equal to $\sec(\beta)$. Since $\mu_\textrm{MS} \ge \mu_\textrm{G}$, and $\sec(\beta)\ge \mu_\textrm{G}$ if and only if $\beta\in[\beta_0,\pi/2]$, where we define $\beta_0 = \sec^{-1}(\mu_\textrm{G}) \approx 63.22^\circ$, we see that in fact \begin{equation} \label{eq:sofaconst-beta0} \mu_\textrm{MS} = \sup_{\beta_0 \le \beta \le \pi/2} \mu_*(\beta), \end{equation} and furthermore, $\mu_\textrm{MS} > \mu_*(\beta)$ for any $0<\beta<\beta_0$, i.e., any moving sofa of maximal area has to rotate by an angle of at least $\beta_0$. (Gerver applied this argument to claim a slightly weaker version of this result in which the value of $\beta_0$ is taken as $\pi/3 = \sec^{-1}(2)$; see \cite[p.~271]{gerver}). Note that it has not been proved, but seems natural to conjecture, that $\mu_\textrm{MS} = \mu_*(\pi/2)$---an assertion that would follow from Gerver's conjecture that the shape he discovered is the moving sofa shape of largest area, but may well be true even if Gerver's conjecture is false. The relationship between our family of finite-dimensional optimization problems and the moving sofa problem is made apparent by the following result. \begin{prop} \label{prop:sofa-fg-bounds} (i) For any $\boldsymbol{\alpha}=(\alpha_1,\ldots,\alpha_k)$ and $\beta$ with $0<\alpha_1<\ldots<\alpha_k \le \beta\le\pi/2$, we have \begin{equation} \mu_*(\beta) \le G_{\boldsymbol{\alpha}}. \label{eq:sofa-fg-bound1} \end{equation} \noindent (ii) For any $\boldsymbol{\alpha}=(\alpha_1,\ldots,\alpha_k)$ with $0<\alpha_1<\ldots<\alpha_k \le \beta_0$, we have \begin{equation} \mu_\textrm{MS} \le G_{\boldsymbol{\alpha}}. \label{eq:sofa-fg-bound2} \end{equation} \noindent (iii) For any $\boldsymbol{\alpha}=(\alpha_1,\ldots,\alpha_k)$ and $\beta_1,\beta_2$ with $0<\alpha_1<\ldots<\alpha_k \le \beta_1 < \beta_2 \le \pi/2$, we have \begin{equation} \sup_{\beta\in[\beta_1,\beta_2]} \mu_*(\beta) \le G_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}. \label{eq:sofa-fg-bound3} \end{equation} \end{prop} \begin{proof} Start by noting that, under the assumption that $0 < \alpha_1 <\ldots < \alpha_k \le \beta \le \pi/2$, if $\mathbf{x}$ is a $\beta$-rotation path then the values $\mathbf{x}(\alpha_1),\ldots,\mathbf{x}(\alpha_k)$ may potentially range over an arbitrary $k$-tuple of vectors in $\mathbb{R}^2$. It then follows that \begin{align*} \mu_*(\beta) &= \sup \Big\{ \lambda^*(S_\mathbf{x}) \,:\, \mathbf{x}\textrm{ is a $\beta$-rotation path} \Big\} \nonumber \\ &= \sup \left\{ \lambda^*\left(L_{\textrm{horiz}} \cap \bigcap_{0\le t\le \beta} L_t(\mathbf{x}(t)) \cap \Big(\mathbf{x}(\beta)+\rotmat{\beta}(L_{\textrm{vert}}) \Big)\right) \,:\, \right. \nonumber \\ & \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \quad \ \ \left. \mathbf{x}\textrm{ is a $\beta$-rotation path} \vphantom{\lambda^*\left(L_{\textrm{horiz}} \cap \bigcap_{0\le t\le \beta} L_t(\mathbf{x}(t)) \cap \Big(\mathbf{x}(\beta)+\rotmat{\beta}(L_{\textrm{vert}}) \Big)\right)} \right\} \nonumber \\ &\le \sup \left\{ \lambda^*\left(H \cap \bigcap_{j=1}^k L_{\alpha_j}(\mathbf{x}(\alpha_j)) \right) \,:\, \mathbf{x}\textrm{ is a $\beta$-rotation path} \right\} \nonumber \\ &= \sup \left\{ \lambda^*\left(H \cap \bigcap_{j=1}^k L_{\alpha_j}(\mathbf{x}_j) \right) \,:\, \mathbf{x}_1,\ldots,\mathbf{x}_k\in\mathbb{R}^2 \right\} = G_{\boldsymbol{\alpha}}. \end{align*} This proves claim (i) of the Proposition. If one further assumes that $\alpha_k\le \beta_0$, \eqref{eq:sofa-fg-bound2} also follows immediately using \eqref{eq:sofaconst-beta0}, proving claim (ii). The proof of claim (iii) follows a variant of the same argument used above; first, note that we may assume that $\beta_2<\pi/2$, since the case $\beta_2=\pi/2$ already follows from part (i) of the Proposition. Next, observe that \begin{align*} \mu_*(\beta) &= \sup \left\{ \lambda^*\left(L_{\textrm{horiz}} \cap \bigcap_{0\le t\le \beta} L_t(\mathbf{x}(t)) \cap \Big(\mathbf{x}(\beta)+\rotmat{\beta}(L_{\textrm{vert}}) \Big)\right) \,:\, \right. \nonumber \\ & \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \quad \ \ \left. \mathbf{x}\textrm{ is a $\beta$-rotation path} \vphantom{\lambda^*\left(L_{\textrm{horiz}} \cap \bigcap_{0\le t\le \beta} L_t(\mathbf{x}(t)) \cap \Big(\mathbf{x}(\beta)+\rotmat{\beta}(L_{\textrm{vert}}) \Big)\right)} \right\} \\ & \le \sup_{\mathbf{x}_1,\ldots,\mathbf{x}_{k+1}\in\mathbb{R}^2} \left[ \lambda^*\left(H \cap \bigcap_{j=1}^k L_{\alpha_j}(\mathbf{x}_j) \cap \Big(\mathbf{x}_{k+1}+\rotmat{\beta}(V) \Big)\right) \right]. \\ & = \sup_{\mathbf{y}_1,\ldots,\mathbf{y}_{k}\in\mathbb{R}^2} \left[ \lambda^*\left(H \cap \bigcap_{j=1}^k L_{\alpha_j}(\mathbf{y}_j) \cap \rotmat{\beta}(V) \right) \right], \\\end{align*} where the last equality follows by expressing $\mathbf{x}_{k+1}$ in the form $\mathbf{x}_{k+1}=a (1,0)+b (-\sin \beta,\cos \beta)$, making the substitution $\mathbf{x}_j = \mathbf{y}_j+a(1,0)$ ($1\le j\le k$), and using the facts that $H+a(1,0)=H$ and $\rotmat{\beta}(V)+b(-\sin \beta,\cos\beta) = \rotmat{\beta}(V)$. Finally, as noted after the definition of $B(\beta_1,\beta_2)$, this set has the property that if $\beta \in [\beta_1,\beta_2]$ then $\rotmat{\beta}(V) \subset B(\beta_1,\beta_2)$. We therefore get for such $\beta$ that \begin{align*} \mu_*(\beta) & \le \sup_{\mathbf{y}_1,\ldots,\mathbf{y}_{k}\in\mathbb{R}^2} \left[ \lambda^*\left(H \cap \bigcap_{j=1}^k L_{\alpha_j}(\mathbf{y}_j) \cap B(\beta_1,\beta_2) \right) \right] = G_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}, \end{align*} which finishes the proof. \end{proof} \paragraph{Example.} In the case of a vector $\boldsymbol{\alpha}=(\alpha)$ with a single angle $0<\alpha<\pi/2$, a simple calculation, which we omit, shows that \begin{equation} \label{eq:g-alpha-explicit} G_{(\alpha)} = \sec\alpha+\csc\alpha. \end{equation} Taking $\alpha=\pi/4$ and using Proposition~\ref{prop:sofa-fg-bounds}(ii), we get the result that $\mu_\textrm{MS}\le 2\sqrt{2}$, which is precisely Hammersley's upper bound for $\mu_\textrm{MS}$ mentioned in the introduction (indeed, this application of the Proposition is essentially Hammersley's proof rewritten in our notation). \bigskip We conclude this section with a result that makes precise the notion that the optimization problems involved in the definition of $G_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}$ are finite-dimensional approximations to the (infinite-dimensional) optimization problem that is the moving sofa problem. \begin{thm}[Convergence to the moving sofa problem] \label{thm:conv-moving-sofa} Let \begin{align*} \boldsymbol{\alpha}(n,k) &= (\tfrac1n\tfrac\pi2,\tfrac2n\tfrac\pi2,\tfrac3n\tfrac\pi2,\ldots,\tfrac{n-k-1}{n}\tfrac\pi2)\text, \\ \gamma_1(n,k)&=\tfrac{n-k}{n}\tfrac\pi2, \\ \gamma_2(n,k)&=\tfrac{n-k+1}{n}\tfrac\pi2, \end{align*} and let \begin{equation} \label{eq:def-wn} W_n = \max_{k=1,\ldots,\lceil n/3\rceil} G_{\boldsymbol{\alpha}(n,k)}^{\gamma_1(n,k),\gamma_2(n,k)}\text. \end{equation} Then $\lim_{n\to\infty} W_n = \mu_\textrm{MS}$. \end{thm} \begin{proof} For each $n\ge 6$, denote by $k^*_n$ the smallest value of $k$ for which the maximum in the definition $W_n$ is attained, let $\beta_n' = \gamma_1(n,k_n^*)$, and let $\beta_n''=\gamma_2(n,k_n^*)$. There is subsequence of $W_n$ converging to its limit superior. Furthermore, the values of $\beta_n'$ in this subsequence have an accumulation point. Therefore, let $n_m$ denote the indices of a subsequence of $W_n$ converging to its limit superior such that $\beta_n'$ (and therefore also $\beta_n''$) converges to some limiting angle $\beta \in [\pi/3,\pi/2]$. Let $\mathbf{u}^{(n)}=\left(u_1^{(n)},\ldots,u_{2(n-{k^*_n}-1)}^{(n)}\right)$ denote a point in $\Omega_{\boldsymbol{\alpha}(n,k^*_n)}^{\beta'_n,\beta''_n}$ where $g_{\boldsymbol{\alpha}(n,{k^*_n})}^{\beta_n',\beta_n''}$ attains its maximum value, which (it is immediate from the definitions) is equal to $G_{\boldsymbol{\alpha}(n,{k^*_n})}^{\beta_n',\beta_n''}=W_n$. Moreover, as more than one point may attain this value, let $\mathbf{u}^{(n)}$ be chosen to be minimal under the coordinatewise partial order with respect to this property. Let $P_n$ be a largest area connected component of $$H\cap \bigcap_{j=1}^{n-{k^*_n}-1} L_{j\pi/2n}\left(u^{(n)}_{2j-1},u^{(n)}_{2j}\right)\cap B\left(\beta_n',\beta_n''\right).$$ Again by the definitions, $\lambda(P_n)=W_n$. Note that the diameter of $P_n$ is bounded from above by a universal constant; this is easy to check and is left as an exercise. Now, of course $P_n$ is not a moving sofa shape, but we will show that it approximates one along a suitable subsequence. To this end, define functions $U^{(n)}:[0,\beta]\to \mathbb{R}$, $V^{(n)}:[0,\beta]\to \mathbb{R}$ by \begin{align*} U^{(n)}(t) &= \max_{(x,y)\in P_n} \left( x\cos t+y\sin t - 1 \right), \\ V^{(n)}(t) &= \max_{(x,y)\in P_n} \left( -x\sin t+y\cos t - 1 \right). \end{align*} We note that $U^{(n)}(t)$ and $V^{(n)}(t)$ have the following properties: first, they are Lipschitz-continuous with a uniform (independent of $n$) Lipschitz constant; see pp.\ 269--270 of Gerver's paper \cite{gerver} for the proof, which uses the fact that the diameters of $P_n$ are bounded. Second, the fact that $P_n \subset H$ implies that $\displaystyle V^{(n)}(0) = \max_{(x,y)\in P_n} (y-1) \le 0$. This in turn implies that \begin{align} P_n &\subseteq (-\infty,U^{(n)}(0)+1] \times [0,V^{(n)}(0)+1] \nonumber \\ & \subseteq (-\infty,U^{(n)}(0)+1] \times [V^{(n)}(0),V^{(n)}(0)+1] \nonumber \\&= (U^{(n)}(0),V^{(n)}(0)) + L_{\textrm{horiz}}. \label{eq:third-observation} \end{align} Third, we have $$ \left(U^{(n)}(j \pi/2n), V^{(n)}(j \pi/2n) \right) = \left(u^{(n)}_{2j-1}, u^{(n)}_{2j} \right) $$ for all $j=1,\ldots,n-{k^*_n}$; that is, $U^{(n)}(t)$ and $V^{(n)}(t)$ continuously interpolate the odd and even (respectively) coordinates of the vector $\mathbf{u}^{(n)}$. Indeed, the relation $P_n \subset L_{j\pi/2n}\left(u^{(n)}_{2j-1},u^{(n)}_{2j}\right)$ implies trivially that $$U^{(n)}(j \pi/2n) \le u^{(n)}_{2j-1} \ \ \textrm{ and }\ \ V^{(n)}(j \pi/2n) \le u^{(n)}_{2j}.$$ However, if we had a strict inequality $U^{(n)}(j \pi/2n) < u^{(n)}_{2j-1}$ (respectively, $V^{(n)}(j \pi/2n) < u^{(n)}_{2j}$), that would imply that replacing $L_{j\pi/2n}\left(u^{(n)}_{2j-1},u^{(n)}_{2j}\right)$ by $L_{j\pi/2n}\left(u^{(n)}_{2j-1}-\epsilon,u^{(n)}_{2j}\right)$ (respectively, $L_{j\pi/2n}\left(u^{(n)}_{2j-1},u^{(n)}_{2j}-\epsilon\right)$) for some small positive $\epsilon$ in the definition of $P_n$ would increase its area slightly, in contradiction to the maximality property defining $\mathbf{u}^{(n)}$. We now define a smoothed version $T_n$ of the polygon $P_n$ by letting \begin{align*} T_n =& P_n\cap\left(L_{\textrm{horiz}}+(U^{(n)}(0),V^{(n)}(0))\right) \cap\bigcap_{0\le t\le\beta} L_t(U^{(n)}(t),V^{(n)}(t))\\ &\cap \left( \rotmat{\beta}\Big(L_{\textrm{vert}}\Big)+(U^{(n)}(\beta),V^{(n)}(\beta)) \right) \\ =& P_n\cap\bigcap_{0\le t\le\beta} L_t(U^{(n)}(t),V^{(n)}(t))\cap \left( \rotmat{\beta}\Big(L_{\textrm{vert}}\Big)+(U^{(n)}(\beta),V^{(n)}(\beta)) \right), \end{align*} where the first equality sign is a definition, and the second equality follows from \eqref{eq:third-observation}. We claim that the Hausdorff distance between the sets $T_{n_m}$ and $P_{n_m}$ goes to $0$ as $n\to\infty$. To see this, let $(x,y)\in P_{n_m}\setminus T_{n_m}$. From the fact that $(x,y)\in P_{n_m}$ we have that \begin{equation} \label{eq:ineqPnm1} \begin{aligned} x\cos t + y\sin t &\ge U^{(n_m)}(t) \ \textrm{ or } \ -x\sin t + y\cos t \ge V^{(n_m)}(t) \end{aligned} \end{equation} for all $t = j\pi/2n_m$, where $j=1,\ldots,n_m-k_{n_m}^*-1$. Moreover, we have $y>V^{(n_m)}(0)$ and \begin{equation} \label{eq:ineqPnm2} x\cos t + y\sin t \ge U^{(n_m)}(t) \end{equation} for $t = \beta'_{n_m} = (n_m - k_{n_m}*)\pi/2n_m$. We want to show that there exists $\delta_m\to 0$ such that \begin{equation} \begin{aligned} \label{eq:ineqTnm1} x\cos t' + (y+\delta_m)\sin t' &\ge U^{(n_m)}(t') \ \textrm{ or } \\ -x\sin t' + (y+\delta_m)\cos t' &\ge V^{(n_m)}(t'). \end{aligned} \end{equation} for all $0<t'<\beta$ and \begin{equation} \label{eq:ineqTnm2} x\cos \beta + (y+\delta_m)\sin \beta \ge U^{(n_m)}(\beta)\text. \end{equation} We claim that $\delta_m = C((1/n_m) + |\beta-\beta'_{n_m}|)^{1/2}$ suffices, where $C$ is some constant. First, if $\beta<\pi/2$, then for $t'\in[(1/n_m)^{1/2},\beta]$, we have \eqref{eq:ineqTnm1} from the uniform Lipschitz continuity of $U^{(n_m)}(t')$, $V^{(n_m)}(t')$, and the other terms as functions of $t'$ (recall $x$ and $y$ are uniformly bounded), from the fact that we have \eqref{eq:ineqPnm1} for some $t$ with $|t-t'|<((1/n_m) + |\beta-\beta'_{n_m}|)$, and from $|\sin t'|, |\cos t'| > \tfrac12((1/n_m) + |\beta-\beta'_{n_m}|)^{1/2}$. For $t'<(1/n_m)^{1/2}$, the fact that $y>V^{(n_m)}(0)$ and Lipschitz continuity suffice to give the second clause of \eqref{eq:ineqTnm1}. Finally, \eqref{eq:ineqTnm2} is satisfied due to Lipschitz continuity and the inequality \eqref{eq:ineqPnm2}. The case of $\beta=\pi/2$ can be worked out similarly. Therefore, for every $(x,y)\in P_{n_m}\setminus T_{n_m}$, we can construct a point $(x,y')\in T_{n_m}$, with $|y'-y|\le \delta_m$ with $\delta_m\to0$. We now use the fact that the vector-valued function $(U^{(n_m)}(t),V^{(n_m)}(t))$ is uniformly Lipschitz to conclude using the Arzel\`a-Ascoli theorem that it has a subsequence (which we still denote by $n_m$, to avoid clutter) such that the ``anchored'' version of the function $(U^{(n_m)}(t),V^{(n_m)}(t))$, namely $$(U^{(n_m)}(t)-U^{(n_m)}(0), V^{(n_m)}(t)-V^{(n_m)}(0))$$ converges in the supremum norm to some limiting function $\mathbf{x}:[0,\beta]\to\mathbb{R}^2$, with the same Lipschitz constant, which satisfies $\mathbf{x}(0)=(0,0)$; that is, the limiting function $\mathbf{x}$ is a $\beta$-rotation path. Now let \begin{equation*} \begin{aligned} T_\infty =& L_{\textrm{horiz}} \cap\bigcap_{0\le t\le\beta} L_t(\mathbf{x}(t)) \cap \left(\rotmat{\beta}\left(L_{\textrm{vert}}\right)+\mathbf{x}(\beta)\right). \end{aligned} \end{equation*} Since the Hausdorff distances of $P_{n_m}$ to $T_{n_m}$ and of $T_{n_m}-(U^{(n)}(0),V^{(n)}(0))$ to $T_\infty\cap \left(T_{n_m}-(U^{(n)}(0),U^{(n)}(0))\right)$ both approach zero as $m\to\infty$, we have that the largest connected component of $T_\infty$ has an area at least as large as $\lim_{m\to\infty} \lambda(P_{n_m}) = \lim\sup_{n\to\infty} W_n$. On the other hand, $T_\infty$ is of the form \eqref{eq:sx-intersections} for a $\beta$-rotation path $\mathbf{x}(t)$, so, by \eqref{eq:sofaconst-characterization}, its area is bounded from above by $\mu_\textrm{MS}$. Therefore, $\lim\sup_{n\to\infty} W_n\le\mu_\textrm{MS}$. We also have that $W_n\ge \mu_*(\beta)$ for all $\pi/3\le\beta\le\pi/2$ from Proposition~\ref{prop:sofa-fg-bounds}(iii), so $W_n \ge \mu_\textrm{MS}$ for all $n$. We conclude that $\lim_{n\to\infty} W_n=\mu_\textrm{MS}$, as claimed. \end{proof} We remark that in view of Theorem~\ref{thm:angle-bound}, it is easy to see that Theorem~\ref{thm:conv-moving-sofa} remains true if we replace the range $1\le k\le \lceil n/3 \rceil$ of values of $k$ in \eqref{eq:def-wn} with the smaller (and therefore computationally more efficient) range $1\le k\le \lceil n/9 \rceil$. \section{An algorithmic proof scheme for moving sofa area bounds} \label{sec:algorithm} The theoretical framework we developed in the previous section reduces the problem of deriving upper bounds for $\mu_\textrm{MS}$ to that of proving upper bounds for the function $G_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}$. Since this function is defined in terms of solutions to a family of optimization problems in finite-dimensional spaces, this is already an important conceptual advance. However, from a practical standpoint it remains to develop and implement a practical, efficient algorithm for solving optimization problems in this class. Our goal in this section is to present such an algorithm and establish its correctness. Our computational strategy is a variant of the \textbf{geometric branch and bound} optimization technique \cite{ratscheck}. Recall from Lemma~\ref{lem:supmax} that the maximum of the function $g_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}$ is attained in a box (a Cartesian product of intervals) $\Omega_{\boldsymbol{\alpha}}^{\beta_1,\beta_2} \subset \mathbb{R}^{2k}$. Our strategy is to break up $\Omega_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}$ into sub-boxes. On each box $E$ being considered, we will compute a quantity $\Gamma_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}(E)$, which is an upper bound for $g_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}(\mathbf{u})$ that holds uniformly for all $\mathbf{u} \in E$. In many cases this bound will not be an effective one; in such a case the box will be subdivided into two further boxes $E_1$ and $E_2$, which will be inserted into a queue to be considered later. Other boxes lead to effective bounds and need not be considered further. By organizing the computation efficiently, practical bounds can be established in a reasonable time, at least for small values of~$k$. To make the idea precise, we introduce a few more definitions. Given two intervals $I=[a,b], J=[c,d] \subseteq \mathbb{R}$ and $\alpha \in[0,\pi/2]$, define $$ \widehat{L}_\alpha(I,J) = \bigcup_{u \in I, v \in J} L_\alpha(u,v). $$ Note that $\widehat{L}_\alpha(I,J)$ can also be expressed as a Minkowski sum of $L_\alpha(0,0)$ with the rotated rectangle $\rotmat{\alpha}(I\times J)$; in particular, it belongs to the class of planar sets known as \textbf{Nef polygons} (see the Appendix for further discussion of Nef polygons and their relevance to our software implementation of the algorithm). Now, for a box $E=I_1\times \ldots \times I_{2k} \subset \mathbb{R}^2$, define \begin{equation}\label{eq:defcalF} \Gamma_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}(E) = \lambda^*\left( H \cap \bigcap_{j=1}^k \widehat{L}_{\alpha_j}(I_{2j-1},I_{2j}) \cap B(\beta_1,\beta_2) \right)\text. \end{equation} Thus, by the definitions we have trivially that \begin{equation} \label{eq:upperbound-trivially} \sup_{\mathbf{u}\in E} g_{\boldsymbol{\alpha}}(\mathbf{u}) \le \Gamma_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}(E). \end{equation} Next, given a box $E=I_1\times\ldots\times I_{2k}$ where $I_j=[a_j,b_j]$, let $$ P_{\textrm{mid}}(E) = \left( \frac{a_1+b_1}{2}, \ldots ,\ldots,\frac{a_{2k}+b_{2k}}{2}\right) $$ denote its midpoint. We also assume that some rule is given to associate with each box $E$ a coordinate $i=\operatorname{ind}(E) \in \{1,\ldots,2k\}$, called the \textbf{splitting index} of $E$. This index will be used by the algorithm to split $E$ into two sub-boxes, which we denote by $\operatorname{split}_{i,1}(E)$ and $\operatorname{split}_{i,2}(E)$, and which are defined as \begin{align*} \operatorname{split}_{i,1}(E) &= I_1\times \ldots \times I_{i-1} \times \left[a_i,\tfrac12(a_i+b_i)\right] \times I_{i+1}\times \ldots I_{2k}, \\ \operatorname{split}_{i,2}(E) &= I_1\times \ldots \times I_{i-1} \times \left[\tfrac12(a_i+b_i),b_i\right] \times I_{i+1}\times \ldots I_{2k}. \end{align*} We assume that the mapping $E\mapsto \operatorname{ind}(E)$ has the property that, if the mapping $E \mapsto \operatorname{split}_{\operatorname{ind}(E),j}(E)$ is applied iteratively, with arbitrary choices of $j\in\{1,2\}$ at each step and starting from some initial value of $E$, the resulting sequence of splitting indices $i_1,i_2,\ldots$ contains each possible coordinate infinitely many times. A mapping satisfying this assumption is referred to as a \textbf{splitting rule}. The algorithm is based on the standard data structure of a \textbf{priority queue} \cite{CLRS} used to hold boxes that are still under consideration. Recall that in a priority queue, each element of the queue is associated with a numerical value called its priority, and that the queue realizes operations of pushing a new element into the queue with a given priority, and popping the highest priority element from the queue. In our application, the priority of each box $E$ will be set to a value denoted $\Pi(E)$, where the mapping $E\mapsto \Pi(E)$ is given and is assumed to satisfy \begin{equation} \label{eq:priority-map-condition} \Pi(E) \ge \Gamma_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}(E). \end{equation} Aside from this requirement, the precise choice of mapping is an implementation decision. (A key point here is that setting $\Pi(E)$ \emph{equal} to $\Gamma_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}(E)$ is conceptually the simplest choice, but from the practical point of view of minimizing programming complexity and running time it may not be optimal; see the Appendix for further discussion of this point.) Note that, since boxes are popped from the queue to be inspected by the algorithm in decreasing order of their priority, this ensures that the algorithm pursues successive improvements to the upper bound it obtains in a greedy fashion. The algorithm also computes a lower bound on $G_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}$ by evaluating $g_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}(\mathbf{u})$ at the midpoint of every box it processes and keeping track of the largest value observed. This lower bound is used to discard boxes in which it is impossible for the maximum to lie. The variable keeping track of the lower bound is initialized to some number $\ell_0$ known to be a lower bound for $G_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}$. In our software implementation we used the value $$ \ell_0 = \begin{cases} 0 & \textrm{if }\beta_2 < \pi/2, \\ 11/5 & \textrm{if }\beta_2 = \pi/2, \end{cases} $$ this being a valid choice thanks to the fact that (by Proposition~\ref{prop:sofa-fg-bounds}(iii)) $G_{\boldsymbol{\alpha}}^{\beta_1,\pi/2} \ge \mu_*(\pi/2) \ge \mu_\textrm{G} = 2.2195\ldots > 2.2=11/5$. Note that simply setting $\ell_0=0$ in all cases would also result in a valid algorithm, but would result in a slight waste of computation time compared to the definition above. With this setup, we can now describe the algorithm, given in pseudocode in Listing~\ref{alg:branchandbound}. The next proposition is key to proving the algorithm's correctness. \begin{listing} \begin{mdframed}[backgroundcolor=codebgcolor] \begin{algorithmic} \State $\varname{box\_queue} \gets $ an empty priority queue of boxes \State $\varname{initial\_box} \gets $ box representing $\Omega_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}$, computed according to the \State \phantom{$\varname{initial\_box} \gets $} function of $\boldsymbol{\alpha}, \beta_1, \beta_2$ described in Lemma~\ref{lem:supmax} \smallskip \State \keyword{push} $\varname{initial\_box}$ into $\varname{box\_queue}$ with priority $\Pi(\varname{initial\_box})$ \smallskip \State $\varname{best\_lower\_bound\_so\_far} \gets $ the initial lower bound $\ell_0$ \medskip \While{true} \medskip \State \keyword{pop} highest priority element of $\varname{box\_queue}$ into $\varname{current\_box}$ \State $\varname{current\_box\_lower\_bound} \gets g_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}(P_{\textrm{mid}}(\varname{current\_box}))$ \State $\varname{best\_upper\_bound\_so\_far} \gets \Pi(\varname{current\_box})$ \medskip \If{$\varname{current\_box\_lower\_bound} > \varname{best\_lower\_bound\_so\_far}$} \State $\varname{best\_lower\_bound\_so\_far} \gets \varname{current\_box\_lower\_bound}$ \EndIf \medskip \State $\varname{i} \gets \operatorname{ind}(\varname{current\_box})$ \smallskip \smallskip \For{$\varname{j}=1,2$} \State $\varname{new\_box} \gets \operatorname{split}_{\varname{i},\varname{j}}(\varname{current\_box})$ \If{$\Pi(\varname{new\_box}) \ge \varname{best\_lower\_bound\_so\_far}$ } \State \keyword{push} $\varname{new\_box}$ into $\varname{box\_queue}$ with priority $\Pi(\varname{new\_box})$ \EndIf \EndFor \medskip \State \textbf{Reporting point:} print the values of $\varname{best\_upper\_bound\_so\_far}$ \State \phantom{\textbf{Reporting point:}} and $\varname{best\_lower\_bound\_so\_far}$ \medskip \EndWhile \end{algorithmic} \end{mdframed} \caption{The algorithm for computing bounds for $G_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}$.} \label{alg:branchandbound} \end{listing} \begin{prop}\label{prop:lead-box} Any box $\tilde E$ which is the highest priority box in the queue $\varname{box\_queue}$ at some step satisfies \begin{equation} \label{eq:lead-box} \sup\{g_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}(\mathbf{u}):\mathbf{u}\in\tilde E\} \le G_{\boldsymbol{\alpha}}^{\beta_1,\beta_2} \le \Pi(\tilde E)\text. \end{equation} \end{prop} \begin{proof} First, the lower inequality holds for \emph{all} boxes in the queue, simply because the value $g_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}(\mathbf{u})$ for any $\mathbf{u}\in\Omega_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}$ is a lower bound on its maximum over all $\mathbf{u}\in\Omega_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}$. Next, let $Q_n$ denote the collection of boxes in the priority queue after $n$ iterations of the \texttt{while} loop (with $Q_0$ being the initialized queue containing the single box $\Omega_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}$), and let $D_n$ denote the collection of boxes that were discarded (not pushed into the priority queue during the execution of the \texttt{if} clause inside the \texttt{for} loop) during the first $n$ iterations of the \texttt{while} loop. Then we first note that for all $n$, the relation \begin{equation} \label{eq:confspace-decom} \Omega_{\boldsymbol{\alpha}}^{\beta_1,\beta_2} = \bigcup_{E \in Q_n\cup D_n} E \end{equation} holds. Indeed, this is easily proved by induction on $n$: if we denote by $X$ the highest priority element in $Q_n$, then during the $(n+1)$th iteration of the \texttt{while} loop, $X$ is subdivided into two boxes $X=X_1\cup X_2$, and each of $X_1, X_2$ is either pushed into the priority queue (i.e., becomes an element of $Q_{n+1}$) or discarded (i.e., becomes an element of $D_{n+1}$), so we have that $\Omega_{\boldsymbol{\alpha}}^{\beta_1,\beta_2} = \bigcup_{E \in Q_{n+1}\cup D_{n+1}} E$, completing the inductive step. Second, note that for any box $X\in D_n$, since $X$ was discarded during the $k$th iteration of the \texttt{while} loop for some $1\le k\le n$, we have that $\Pi(X)$ is smaller than the value of $\varname{best\_lower\_bound\_so\_far}$ during that iteration. But $\varname{best\_lower\_bound\_so\_far}$ is always assigned a value of the form $g_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}(\mathbf{u})$ for some $\mathbf{u}\in R$ and is therefore bounded from above by $G_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}$, so we have established that \begin{equation} \label{eq:xupperbound-le} \Pi(X) < G_{\boldsymbol{\alpha}}^{\beta_1,\beta_2} \qquad (X \in D_n). \end{equation} The relations \eqref{eq:supmax}, \eqref{eq:upperbound-trivially}, \eqref{eq:priority-map-condition}, \eqref{eq:confspace-decom}, and \eqref{eq:xupperbound-le} now imply that \begin{align*} G_{\boldsymbol{\alpha}}^{\beta_1,\beta_2} &= \max \left\{ g_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}(\mathbf{u}) \,:\, \mathbf{u}\in \Omega_{\boldsymbol{\alpha}}^{\beta_1,\beta_2} \right\} \\ &= \max_{E \in Q_n\cup D_n} \left( \sup \left\{ g_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}(\mathbf{u}) \,:\, \mathbf{u}\in E \right\} \right) \le \max_{E \in Q_n\cup D_n} \Gamma_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}(E) \\ &\le \max_{E \in Q_n\cup D_n} \Pi(E) = \max_{E \in Q_n} \Pi(E). \end{align*} Finally, $\max_{E \in Q_n} \Pi(E) = \Pi(\tilde E)$, since $\tilde E$ was assumed to be the box with highest priority among the elements of $Q_n$, so we get the upper inequality in \eqref{eq:lead-box}, which finishes the proof. \end{proof} We immediately have the correctness of the algorithm as a corollary: \begin{thm}[Correctness of the algorithm] Any \,\!\!\! value of the variable \ $\varname{best\_upper\_bound\_so\_far}$ reported by the algorithm is an upper bound for $G_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}$. \end{thm} Note that the correctness of the algorithm is not dependent on the assumption we made on the splitting index mapping $E\mapsto \operatorname{ind}(E)$ being a splitting rule. The importance of that assumption is explained by the following result, which also explains one sense in which assuming an equality in \eqref{eq:priority-map-condition} rather than an inequality provides a benefit (of a theoretical nature at least). \begin{thm}[Asymptotic sharpness of the algorithm] \label{thm:asym-sharpness} Assume that the priority mapping $E\mapsto \Pi(E)$ is taken to be \begin{equation} \label{eq:priority-map-sharpness} \Pi(E) = \Gamma_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}(E). \end{equation} Then the upper and lower bounds output by the algorithm both converge to $G_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}$. \end{thm} \begin{proof} As one may easily check, the upper bound used in the calculation under the assumption \eqref{eq:priority-map-sharpness}, $\Gamma_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}(E)$, approaches the actual supremum of $g_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}(\mathbf{u})$ over $E$ as the diameter of $E$ approaches zero. That is $|\Gamma_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}(E) - \sup\{g_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}(\mathbf{u}):\mathbf{u}\in E\}|$ is bounded by a function of the diameter of $E$ that approaches zero when the diameter approaches zero. The same is true of the variation in each box, $|\sup\{g_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}(\mathbf{u}):\mathbf{u}\in E\} - \inf\{g_{\boldsymbol{\alpha}}^{\beta_1,\beta_2}(\mathbf{u}):\mathbf{u}\in E\}|$. When using a valid splitting rule, the diameter of the leading box approaches zero as $n$ approaches infinity, and Proposition~\ref{prop:lead-box} completes the proof. \end{proof} As with the case of the choice of priority mapping and the value of the initial lower bound $\ell_0$, the specific choice of splitting rule to use is an implementation decision, and different choices can lead to algorithms with different performance. A simple choice we tried was to use the index of the coordinate with the largest variation within $E$ (i.e., the ``longest dimension'' of $E$). Another choice, which we found gives superior performance and is the rule currently used in our software implementation \texttt{SofaBounds}, is to let the splitting index be the value of $i$ maximizing $\lambda(D_i\cap S(E))$, where $S(E)$ is the argument of $\lambda^*$ in \eqref{eq:defcalF}, and \begin{align*} D_i = \begin{cases} \displaystyle \bigcup_{u\in {I_{2j-1}}} \widehat{L}_{\alpha_j}(u,I_{2j}) \setminus \bigcap_{u\in {I_{2j-1}}} \widehat{L}_{\alpha_j}(u,I_{2j}) & \textrm{if }i=2j-1, \\[14pt] \displaystyle \bigcup_{u\in {I_{2j}}} \widehat{L}_{\alpha_j}(I_{2j-1},u) \setminus \bigcap_{u\in {I_{2j}}} \widehat{L}_{\alpha_j}(I_{2j-1},u) & \textrm{if }i=2j. \end{cases} \end{align*} \section{Explicit numerical bounds} \label{sec:numerical} We report the following explicit numerical bounds obtained by our algorithm, which we will then use to prove Theorems~\ref{thm:new-upperbound} and~\ref{thm:angle-bound}. \begin{thm} \label{thm:explicit-bounds} Define angles \begin{align*} \alpha_1 &= \sin^{-1}\tfrac{7}{25} \approx 16.26^\circ, \\ \alpha_2 &= \sin^{-1}\tfrac{33}{65} \approx 30.51^\circ, \\ \alpha_3 &= \sin^{-1}\tfrac{119}{169} \approx 44.76^\circ, \\ \alpha_4 &= \sin^{-1} \tfrac{56}{65} = \pi/2-\alpha_2 \approx 59.59^\circ, \\ \alpha_5 &= \sin^{-1}\tfrac{24}{25} = \pi/2-\alpha_1 \approx 73.74^\circ, \\ \alpha_6 &= \sin^{-1} \tfrac{60}{61} \approx 79.61^\circ, \\ \alpha_7 &= \sin^{-1} \tfrac{84}{85} \approx 81.2^\circ. \end{align*} Then we have the inequalities \begin{align} G_{(\alpha_1,\alpha_2,\alpha_3,\alpha_4,\alpha_5)} &\le 2.37, \label{eq:numerical-bound1} \\[5pt] G_{(\alpha_1,\alpha_2,\alpha_3)}^{\alpha_4,\alpha_5} &\le 2.21, \label{eq:numerical-bound2} \\ G_{(\alpha_1,\alpha_2,\alpha_3)}^{\alpha_5,\alpha_6} &\le 2.21, \label{eq:numerical-bound3} \\ G_{(\alpha_1,\alpha_2,\alpha_3,\alpha_4)}^{\alpha_6,\alpha_7} &\le 2.21. \label{eq:numerical-bound4} \end{align} \end{thm} \begin{proof} Each of the inequalities \eqref{eq:numerical-bound1}--\eqref{eq:numerical-bound4} is certified as correct using the \texttt{SofaBounds} software package by invoking the \texttt{run} command from the command line interface after loading the appropriate parameters. For \eqref{eq:numerical-bound1}, the parameters can be loaded from the saved profile file \texttt{thm9-bound1.txt} included with the package (see the Appendix below for an illustration of the syntax for loading the file and running the computation). Similarly, the inequalities \eqref{eq:numerical-bound2}, \eqref{eq:numerical-bound3}, \eqref{eq:numerical-bound4} are obtained by running the software with the profile files \texttt{thm9-bound2.txt}, \texttt{thm9-bound3.txt}, and \texttt{thm9-bound4.txt}, respectively. Table~\ref{table:benchmarking} in the Appendix shows benchmarking results with running times for each of the computations. \end{proof} \begin{proof}[Proof of Theorem~\ref{thm:new-upperbound}] For angles $0\le \beta_1<\beta_2 \le \pi/2$ denote $$M(\beta_1,\beta_2) = \sup_{\beta_1\le \beta \le \beta_2} \mu_*(\beta).$$ By \eqref{eq:sofaconst-beta0}, we have \begin{equation} \label{eq:sofaconst-tworanges} \mu_\textrm{MS} = M(\beta_0,\pi/2) = \max\Big( M(\beta_0,\alpha_5), M(\alpha_5,\pi/2) \Big). \end{equation} By Proposition~\ref{prop:sofa-fg-bounds}(iii), $M(\beta_0,\alpha_5)$ is bounded from above by $G_{(\alpha_1,\alpha_2,\alpha_3)}^{\alpha_4,\alpha_5}$, and by Proposition~\ref{prop:sofa-fg-bounds}(i), $M(\alpha_5,\pi/2)$ is bounded from above by $G_{(\alpha_1,\alpha_2,\alpha_3,\alpha_4,\alpha_5)}$. Thus, combining \eqref{eq:sofaconst-tworanges} with the numerical bounds \eqref{eq:numerical-bound1}--\eqref{eq:numerical-bound2} proves \eqref{eq:upperbound}. \end{proof} \begin{proof}[Proof of Theorem~\ref{thm:angle-bound}] Using the same notation as in the proof of Theorem~\ref{thm:new-upperbound} above, we note that $$ M(0,\alpha_7) = \max\Big( M(0,\alpha_4), M(\alpha_4,\alpha_5), M(\alpha_5,\alpha_6), M(\alpha_6,\alpha_7) \Big). $$ Now, by Gerver's observation mentioned after the relation \eqref{eq:sofaconst-beta}, we have that $M(0,\alpha_4) \le \sec(\alpha_4) < \sec(\pi/3) = 2$. By Proposition~\ref{prop:sofa-fg-bounds}(iii) coupled with the numerical bounds \eqref{eq:numerical-bound2}--\eqref{eq:numerical-bound4}, the remaining three arguments $M(\alpha_4,\alpha_5)$, $M(\alpha_5,\alpha_6)$, and $M(\alpha_6,\alpha_7)$ in the maximum are all bounded from above by $2.21$, so in particular we get that $M(0,\alpha_7)\le 2.21 < \mu_\textrm{G} \approx 2.2195$. On the other hand, we have that $$ \mu_\textrm{MS} = M(0,\pi/2) = \max\Big( M(0,\alpha_7), M(\alpha_7,\pi/2) \Big) \ge \mu_\textrm{G}. $$ We conclude that $\mu_\textrm{MS} = M(\alpha_7,\pi/2)$ and that $\mu_*(\beta) \le 2.21 < \mu_\textrm{MS}$ for all $\beta<\alpha_7$. This proves that a moving sofa of maximal area has to undergo rotation by an angle of at least $\alpha_7$, as claimed. \end{proof} \section{Concluding remarks} The results of this paper represent the first progress since Hammersley's 1968 paper \cite{hammersley} on deriving upper bounds for the area of a moving sofa shape. Our techniques also enable us to prove an improved lower bound on the angle of rotation a maximal area moving sofa shape must rotate through. Our improved upper bound of $2.37$ on the moving sofa constant comes much closer than Hammersley's bound to the best known lower bound $\mu_\textrm{G} \approx 2.2195$ arising from Gerver's construction, but clearly there is still considerable room for improvement in narrowing the gap between the lower and upper bounds. In particular, some experimentation with the initial parameters used as input for the \texttt{SofaBounds} software should make it relatively easy to produce further (small) improvements to the value of the upper bound. More ambitiously, our hope is that a refinement of our methods---in the form of theoretical improvements and/or speedups in the software implementation, for example using parallel computing techniques---may eventually be used to obtain an upper bound that comes very close to Gerver's bound, thereby providing supporting evidence to his conjecture that the shape he found is the solution to the moving sofa problem. Some supporting evidence of this type, albeit derived using a heuristic algorithm, was reported in a recent paper by Gibbs \cite{gibbs}. Alternatively, a failure of our algorithm (or improved versions thereof) to approach Gerver's lower bound may provide clues that his conjecture may in fact be false. Our methods should also generalize in a fairly straightforward manner to other variants of the moving sofa problem. In particular, in a recent paper \cite{romik}, one of us discovered a shape with a piecewise algebraic boundary that is a plausible candidate to being the solution to the so-called \textbf{ambidextrous moving sofa problem}, which asks for the largest shape that can be moved around a right-angled turn \textit{either to the left or to the right} in a hallway of width~1 (Fig.~\ref{fig:romik-sofa}(a)). The shape, shown in Fig.~\ref{fig:romik-sofa}(b), has an area given by the intriguing explicit constant \begin{align*} \mu_\textrm{R} &=\sqrt[3]{3+2 \sqrt{2}}+\sqrt[3]{3-2 \sqrt{2}}-1 +\arctan\left[ \frac{1}{2} \left( \sqrt[3]{\sqrt{2}+1}- \sqrt[3]{\sqrt{2}-1}\, \right) \right] \nonumber \\ & \qquad\qquad\qquad = 1.64495521\ldots \end{align*} As with the case of the original (non-ambidextrous) moving sofa problem, the constant $\mu_\textrm{R}$ provides a lower bound on the maximal area of an ambidextrous moving sofa shape; in the opposite direction, any upper bound for the original problem is also an upper bound for the ambidextrous variant of the problem, which establishes $2.37$ as a valid upper bound for that problem. Once again, the gap between the lower and upper bounds seems like an appealing opportunity for further work, so it would be interesting to extend the techniques of this paper to the setting of the ambidextrous moving sofa problem so as to obtain better upper bounds on the ``ambidextrous moving sofa constant.'' \begin{figure} \begin{center} \begin{tabular}{cc} \scalebox{0.3}{\includegraphics{ambisofa-in-hallway.pdf}} & \raisebox{10pt}{\scalebox{0.5}{\includegraphics{ambisofa.pdf}}} \\[3pt] (a) & (b) \end{tabular} \caption{(a) The ambidextrous moving sofa problem involves maximization of the area of moving sofa shapes that can navigate a hallway with right-angled turns going both ways, as shown in the figure; (b) a shape discovered by Romik \cite{romik} that was derived as a possible solution to the ambidextrous moving sofa problem. The boundary of the shape is a piecewise algebraic curve; the tick marks in the figure delineate the transition points between distinct parts of the boundary.} \label{fig:romik-sofa} \end{center} \end{figure} \section*{Appendix: The \texttt{SofaBounds} software} We implemented the algorithm described in Section~\ref{sec:algorithm} in the software package \texttt{SofaBounds} we developed, which serves as a companion package to this paper and whose source code is available to download online \cite{sofabounds}. The package is a Unix command line tool written in \texttt{C++} and makes use of the open source computational geometry library \texttt{CGAL} \cite{cgal}. All computations are done in the exact rational arithmetic mode supported by \texttt{CGAL} to ensure that the bounds output by the algorithm are mathematically rigorous. For this reason, the software only works with angles $\gamma$ for which the vector $(\cos\gamma, \sin\gamma)$ has rational coordinates, i.e., is a rational point $(a/c,b/c)$ on the unit circle; clearly such angles are parametrized by Pythagorean triples $(a,b,c)$ such that $a^2+b^2=c^2$, and it is using such triples that the angles are entered into the program as input from the user. For example, to approximate an angle of $45$ degrees, we used the Pythagorean triple $(119,120,169)$, which corresponds to an angle of $\sin^{-1}(119/169) \approx 44.76^\circ$. The software uses the \textbf{Nef polygon} geometric primitive implemented in \texttt{CGAL}; recall that Nef polygons are planar sets that can be obtained from a finite set of half-planes by applying set intersection and complementation operations. It is easy to see that all the planar sets manipulated by the algorithm belong to this family and can be readily calculated using elementary geometry and the features of \texttt{CGAL}'s Nef polygon sub-library \cite{seel}. Our implementation uses the priority rule $$ \Pi(E) = \lambda\left( H \cap \bigcap_{j=1}^k \widehat{L}_{\alpha_j}(I_{2j-1},J_{2j}) \cap B(\beta_1,\beta_2) \right)\text, $$ i.e., we use the total area of the intersection as the priority instead of the area of the largest connected component as in \eqref{eq:defcalF}; this is slightly less ideal from a theoretical point of view, since Theorem~\ref{thm:asym-sharpness} does not apply, but simplified the programming and in practice probably results in better computational performance. The software runs our algorithm on a single Unix thread, since the parts of the CGAL library we used are not thread-safe; note however that the nature of our algorithm lends itself fairly well to parallelization, so a multithreading or other parallelized implementation could yield a considerable speedup in performance, making it more practical to continue to improve the bounds in Theorems~\ref{thm:new-upperbound} and \ref{thm:angle-bound}. To illustrate the use of the software, Listing~\ref{code-listing} shows a sample working session in which the upper bound $2.5$ is derived for $G_{\boldsymbol{\alpha}}$ with \begin{equation} \label{eq:angles-30-45-60} \boldsymbol{\alpha}=\left(\sin^{-1}\frac{33}{65},\sin^{-1}\frac{119}{169}, \sin^{-1}\frac{56}{65}\right) \approx (30.51^\circ, 44.76^\circ, 59.49^\circ). \end{equation} The numerical bounds \eqref{eq:numerical-bound1}--\eqref{eq:numerical-bound4} used in the proofs of Theorems~\ref{thm:new-upperbound} and~\ref{thm:angle-bound} were proved using \texttt{SofaBounds}, and required several weeks of computing time on a desktop computer. Table~\ref{table:benchmarking} shows some benchmarking information, which may be useful to anyone wishing to reproduce the computations or to improve upon our results. \newcommand{\ignore}[1]{{}} \begin{listing} \begin{mdframed}[backgroundcolor=codebgcolor] \begin{alltt} Users/user/SofaBounds$ \inputline{SofaBounds} SofaBounds version 1.0 Type "help" for instructions. > \inputline{load example-30-45-60.txt} File 'example-30-45-60.txt' loaded successfully. > \inputline{settings} Number of corridors: 3 Slope 1: 33 56 65 (angle: 30.5102 deg) Slope 2: 119 120 169 (angle: 44.7603 deg) Slope 3: 56 33 65 (angle: 59.4898 deg) Minimum final slope: 1 0 1 (angle: 90 deg) Maximum final slope: 1 0 1 (angle: 90 deg) Reporting progress every: 0.01 decrease in upper bound > \inputline{run} <iterations=0> <iterations=1 | upper bound=3.754 | time=0:00:00> <iterations=7 | upper bound=3.488 | time=0:00:01> <iterations=9 | upper bound=3.438 | time=0:00:01> \vspace{7pt} \ignore{<iterations=13 | upper bound=3.428 | time=0:00:02> <iterations=14 | upper bound=3.416 | time=0:00:02> <iterations=16 | upper bound=3.405 | time=0:00:02> <iterations=18 | upper bound=3.397 | time=0:00:03> <iterations=19 | upper bound=3.380 | time=0:00:03> <iterations=24 | upper bound=3.360 | time=0:00:04> <iterations=25 | upper bound=3.301 | time=0:00:04> <iterations=30 | upper bound=3.273 | time=0:00:05> <iterations=35 | upper bound=3.248 | time=0:00:06> <iterations=36 | upper bound=3.222 | time=0:00:07> <iterations=37 | upper bound=3.202 | time=0:00:07> <iterations=41 | upper bound=3.189 | time=0:00:08> <iterations=45 | upper bound=3.162 | time=0:00:09> <iterations=46 | upper bound=3.140 | time=0:00:09> <iterations=47 | upper bound=3.043 | time=0:00:09> <iterations=51 | upper bound=3.010 | time=0:00:10> <iterations=53 | upper bound=2.996 | time=0:00:11> <iterations=56 | upper bound=2.983 | time=0:00:11> <iterations=64 | upper bound=2.970 | time=0:00:14> <iterations=67 | upper bound=2.957 | time=0:00:14> <iterations=73 | upper bound=2.943 | time=0:00:16> <iterations=79 | upper bound=2.928 | time=0:00:18> <iterations=87 | upper bound=2.918 | time=0:00:20> <iterations=92 | upper bound=2.903 | time=0:00:21> <iterations=95 | upper bound=2.870 | time=0:00:22> <iterations=102 | upper bound=2.857 | time=0:00:24> <iterations=111 | upper bound=2.848 | time=0:00:27> <iterations=116 | upper bound=2.838 | time=0:00:28> <iterations=118 | upper bound=2.827 | time=0:00:29> <iterations=127 | upper bound=2.817 | time=0:00:31> <iterations=134 | upper bound=2.809 | time=0:00:34> <iterations=139 | upper bound=2.796 | time=0:00:35> <iterations=143 | upper bound=2.790 | time=0:00:36> <iterations=151 | upper bound=2.778 | time=0:00:38> <iterations=165 | upper bound=2.763 | time=0:00:42> <iterations=173 | upper bound=2.748 | time=0:00:44> <iterations=181 | upper bound=2.738 | time=0:00:47> <iterations=204 | upper bound=2.730 | time=0:00:53> <iterations=220 | upper bound=2.719 | time=0:00:58> <iterations=240 | upper bound=2.709 | time=0:01:03> <iterations=278 | upper bound=2.699 | time=0:01:14> <iterations=300 | upper bound=2.689 | time=0:01:21> <iterations=326 | upper bound=2.680 | time=0:01:28> <iterations=361 | upper bound=2.670 | time=0:01:38> <iterations=413 | upper bound=2.660 | time=0:01:53> <iterations=462 | upper bound=2.650 | time=0:02:07> <iterations=548 | upper bound=2.640 | time=0:02:31> <iterations=619 | upper bound=2.630 | time=0:02:52> <iterations=724 | upper bound=2.620 | time=0:03:22> <iterations=812 | upper bound=2.610 | time=0:03:48> <iterations=945 | upper bound=2.600 | time=0:04:27> <iterations=1100 | upper bound=2.590 | time=0:05:12> <iterations=1290 | upper bound=2.580 | time=0:06:07> <iterations=1513 | upper bound=2.570 | time=0:07:18> } \textnormal{[\textit{... 54 output lines deleted ...}]}\vspace{7pt} <iterations=1776 | upper bound=2.560 | time=0:08:43> <iterations=2188 | upper bound=2.550 | time=0:10:48> <iterations=2711 | upper bound=2.540 | time=0:13:23> <iterations=3510 | upper bound=2.530 | time=0:18:18> <iterations=4620 | upper bound=2.520 | time=0:24:54> <iterations=6250 | upper bound=2.510 | time=0:34:52> <iterations=8901 | upper bound=2.500 | time=0:50:45> \end{alltt} \end{mdframed} \caption{A sample working session of the \texttt{SofaBounds} software package proving an upper bound for $G_{\boldsymbol{\alpha}}$ with $\boldsymbol{\alpha}$ given by \eqref{eq:angles-30-45-60} . User commands are colored in \inputline{\textrm{blue}}. The session loads parameters from a saved profile file \texttt{example-30-45-60.txt} (included with the source code download package) and rigorously certifies the number $2.5$ as an upper bound for $G_{\boldsymbol{\alpha}}$ (and therefore also for $\mu_\textrm{MS}$, by Proposition~\ref{prop:sofa-fg-bounds}(ii)) in about $50$ minutes of computation time on a laptop with a 1.3 GHz Intel Core M processor. } \label{code-listing} \end{listing} \begin{table} \begin{center} \begin{tabular}{|c|c|c|c|} \hline Bound & Saved profile file & Num.\ of iterations & Computation time \\ \hline \eqref{eq:numerical-bound1} & \texttt{thm9-bound1.txt} & 7,724,162 & 480 hours \\ \eqref{eq:numerical-bound2} & \texttt{thm9-bound2.txt} & \phantom{0,000,}917 & 2 minutes \\ \eqref{eq:numerical-bound3} & \texttt{thm9-bound3.txt} & \phantom{,00}26,576 & 1:05 hours \\ \eqref{eq:numerical-bound4} & \texttt{thm9-bound4.txt} & \phantom{0,}140,467 & 6:23 hours \\ \hline \end{tabular} \caption{Benchmarking results for the computations used in the proof of the bounds \eqref{eq:numerical-bound1}--\eqref{eq:numerical-bound4}. The computations for \eqref{eq:numerical-bound1} were performed on a 2.3 GHz Intel Xeon E5-2630 processor, and the computations for \eqref{eq:numerical-bound2}--\eqref{eq:numerical-bound4} were performed were performed on a 3.4 GHz Intel Core i7 processor.} \label{table:benchmarking} \end{center} \end{table} Additional details on \texttt{SofaBounds} can be found in the documentation included with the package. \clearpage
{ "timestamp": "2018-01-08T02:15:09", "yymm": "1706", "arxiv_id": "1706.06630", "language": "en", "url": "https://arxiv.org/abs/1706.06630", "abstract": "The moving sofa problem, posed by L. Moser in 1966, asks for the planar shape of maximal area that can move around a right-angled corner in a hallway of unit width. It is known that a maximal area shape exists, and that its area is at least 2.2195... - the area of an explicit construction found by Gerver in 1992 - and at most $2\\sqrt{2}=2.82...$, with the lower bound being conjectured as the true value. We prove a new and improved upper bound of 2.37. The method involves a computer-assisted proof scheme that can be used to rigorously derive further improved upper bounds that converge to the correct value.", "subjects": "Metric Geometry (math.MG); Optimization and Control (math.OC)", "title": "Improved upper bounds in the moving sofa problem", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9814534349454033, "lm_q2_score": 0.815232489352, "lm_q1q2_score": 0.8001127269536122 }
https://arxiv.org/abs/math/0606320
Remarks on the Cayley Representation of Orthogonal Matrices and on Perturbing the Diagonal of a Matrix to Make it Invertible
This note contains two remarks. The first remark concerns the extension of the well-known Cayley representation of rotation matrices by skew symmetric matrices to rotation matrices admitting -1 as an eigenvalue and then to all orthogonal matrices. We review a method due to Hermann Weyl and another method involving multiplication by a diagonal matrix whose entries are +1 or -1. The second remark has to do with ways of flipping the signs of the entries of a diagonal matrix, C, with nonzero diagonal entries, obtaining a new matrix, E, so that E + A is invertible, where A is any given matrix (invertible or not).
\section{The Cayley Representation of Orthogonal Matrices} \label{sec1} Given any rotation matrix, $R\in \mathbf{SO}(n)$, if $R$ does not admit $-1$ as an eigenvalue, then there is a unique skew symmetric matrix, $S$, ($\transpos{S} = -S$) so that \[ R = (I - S)(I + S)^{-1}. \] This is a classical result of Cayley \cite{Cayley} (1846) and $R$ is called the {\it Cayley transform of $S$\/}. Among other sources, a proof can be found in Hermann Weyl's beautiful book {\sl The Classical Groups\/} \cite{Weyl46}, Chapter II, Section 10, Theorem 2.10.B (page 57). \medskip As we can see, this representation misses rotation matrices admitting the eigenvalue $-1$, and of course, as $\det((I - S)(I + S)^{-1}) = +1$, it misses improper orthogonal matrices, i.e., those matrices $R\in \mathbf{O}(n)$ with $\det(R) = -1$. \medskip\noindent {\bf Question 1}. Is there a way to extend the Cayley representation to all rotation matrices (matrices in $\mathbf{SO}(n)$)? \medskip\noindent {\bf Question 2}. Is there a way to extend the Cayley representation to all orthogonal matrices (matrices in $\mathbf{O}(n)$)? \medskip\noindent {\bf Answer}: Yes in both cases! \medskip An answer to Question 1 is given in Weyl's book \cite{Weyl46}, Chapter II, Section 10, Lemma 2.10.D (page 60): \begin{prop} (Weyl) \label{Weyl1} Every rotation matrix, $R\in \mathbf{SO}(n)$, can be expressed as a product \[ R = (I - S_1)(I + S_1)^{-1}(I - S_2)(I + S_2)^{-1}, \] where $S_1$ and $S_2$ are skew symmetric matrices. \end{prop} \medskip Thus, if we allow two Cayley representation matrices, we can capture orthogonal matrices having an even number of $-1$ as eigenvalues. Actually, proposition \ref{Weyl1} can be sharpened slightly as follows: \begin{prop} \label{prop2} Every rotation matrix, $R\in \mathbf{SO}(n)$, can be expressed as \[ R = \Bigl((I - S)(I + S)^{-1}\Bigr)^2 \] where $S$ is a skew symmetric matrix. \end{prop} \medskip Proposition \ref{prop2} can be easily proved using the following well-known normal form for orthogonal matrices: \begin{prop} \label{prop3} For every orthogonal matrix, $R\in \mathbf{O}(n)$, there is an orthogonal matrix $P$ and a block diagonal matrix $D$ such that $R = PD\,\transpos{P}$, where $D$ is of the form \[ D = \amsdiagmat{D}{p} \] such that each block $D_i$ is either $1$, $-1$, or a two-dimensional matrix of the form $$D_i = \amsmata{\cos\theta_i}{-\sin\theta_i}{\sin\theta_i}{\cos\theta_i}$$ where $0 < \theta_i < \pi$. \end{prop} \medskip In particular, if $R$ is a rotation matrix ($R\in \mathbf{SO}(n)$), then it has an even number of eigenvalues $-1$. So, they can be grouped into two-dimensional rotation matrices of the form \[ \amsmata{-1}{0}{0}{-1}, \] i.e., we allow $\theta_i = \pi$ and we may assume that $D$ does not contain one-dimensional blocks of the form $-1$. \medskip A proof of Proposition \ref{prop3} can be found in Gantmacher \cite{Gantmacher1}, Chapter IX, Section 13 (page 285), or Berger \cite{Berger90}, or Gallier \cite{Gallbook2}, Chapter 11, Section 11.4 (Theorem 11.4.5). \medskip Now, for every two-dimensional rotation matrix \[ T = \amsmata{\cos\theta}{-\sin\theta}{\sin\theta}{\cos\theta} \] with $0 < \theta \leq \pi$, observe that \[ T^{\frac{1}{2}} = \amsmata{\cos(\theta/2)}{-\sin(\theta/2)}{\sin(\theta/2)}{\cos(\theta/2)} \] does not admit $-1$ as an eigenvalue (since $0 < \theta/2 \leq\pi/2$) and $T = \left(T^{\frac{1}{2}}\right)^2$. Thus, if we form the matrix $R^{\frac{1}{2}}$ by replacing each two-dimensional block $D_i$ in the above normal form by $D_i^{\frac{1}{2}}$, we obtain a rotation matrix that does not admit $-1$ as an eigenvalue, $R = \left(R^{\frac{1}{2}}\right)^2$ and the Cayley transform of $R^{\frac{1}{2}}$ is well defined. Therefore, we have proved Proposition \ref{prop2}. $\mathchoice\sqr76\sqr76\sqr{2.1}3\sqr{1.5}3$ \bigskip Next, why is the answer to Question 2 also yes? \medskip This is because \begin{prop} \label{prop4} For any orthogonal matrix, $R\in \mathbf{O}(n)$, there is some diagonal matrix, $E$, whose entries are $+1$ or $-1$, and some skew-symmetric matrix, $S$, so that \[ R = E(I - S)(I + S)^{-1}. \] \end{prop} \medskip As such matrices $E$ are orthogonal, all matrices $E(I - S)(I + S)^{-1}$ are orthogonal, so we have a Cayley-style representation of all orthogonal matrices. \medskip I am not sure when Proposition \ref{prop4} was discovered and originally published. Since I could not locate this result in Weyl's book \cite{Weyl46}, I assume that it was not known before 1946, but I did stumble on it as an exercise in Richard Bellman's classic \cite{Bellman}, first published in 1960, Chapter 6, Section 4, Exercise 11, page 91-92 (see also, Exercises, 7, 8, 9, and 10). \medskip Why does this work? \medskip\noindent {\bf Fact E}: Because, for every $n\times n$ matrix, $A$ (invertible or not), there some diagonal matrix, $E$, whose entries are $+1$ or $-1$, so that $I + EA$ is invertible! \medskip This is Exercise 10 in Bellman \cite{Bellman} (Chapter 6, Section 4, page 91). Using Fact E, it is easy to prove Proposition \ref{prop4}. \medskip\noindent {\it Proof of Proposition \ref{prop4}\/}. Let $R\in \mathbf{O}(n)$ be any orthogonal matrix. By Fact E, we can find a diagonal matrix, $E$ (with diagonal entries $\pm1$), so that $I + ER$ is invertible. But then, as $E$ is orthogonal, $ER$ is an orthogonal matrix that does not admit the eigenvalue $-1$ and so, by the Cayley representation theorem, there is a skew symmetric matrix, $S$, so that \[ ER = (I - S)(I + S)^{-1}. \] However, notice that $E^2 = I$, so we get \[ R = E(I - S)(I + S)^{-1}, \] as claimed. $\mathchoice\sqr76\sqr76\sqr{2.1}3\sqr{1.5}3$ \medskip But Why does Fact E hold? \medskip As we just observed, $E^2 = I$, so by multiplying by $E$, \[ \hbox{$I + EA$ is invertible iff $E + A$ is.} \] \medskip Thus, we are naturally led to the following problem: If $A$ is any $n\times n$ matrix, is there a way to perturb the diagonal entries of $A$, i.e., to add some diagonal matrix, $C = \mathrm{diag}(c_1, \ldots, c_n)$, to $A$ so that $C + A$ becomes invertible? \medskip Indeed this can be done, and we will show in the next section that what matters is not the magnitude of the perturbation but the signs of the entries being added. \section{Perturbing the Diagonal of a Matrix to Make it Invertible} \label{sec2} In this section we prove the following result: \begin{prop} \label{prop5} For every $n\times n$ matrix (invertible or not), $A$, and every any diagonal matrix, $C = \mathrm{diag}(c_1, \ldots, c_n)$, with $c_i \not= 0$ for $i = 1, \ldots, n$, there an assignment of signs, $\epsilon_i = \pm 1$, so that if $E = \mathrm{diag}(\epsilon_1 c_1, \ldots, \epsilon_n c_n)$, then $E + A$ is invertible. \end{prop} \noindent{\it Proof\/}.\enspace Let us evaluate the determinant of $C + A$. We see that $\Delta = \det(C + A)$ is a polynomial of degree $n$ in the variables $c_1, \ldots, c_n$ and that all the monomials of $\Delta$ consist of products of distinct variables (i.e., every variable occurring in a monomial has degree $1$). In particular, $\Delta$ contains the monomial $c_1 \cdots c_n$. In order to prove Proposition \ref{prop5}, it will suffice to prove \begin{prop} \label{prop6} Given any polyomial, $P(x_1, \ldots, x_n)$, of degree $n$ (in the indeterminates $x_1, \ldots, x_n$ and over any integral domain of characteristic unequal to $2$), if every monomial in $P$ is a product of distinct variables, for every $n$-tuple $(c_1, \ldots, c_n)$ such that $c_i \not= 0$ for $i = 1, \ldots, n$, then there is an assignment of signs, $\epsilon_i = \pm 1$, so that \[ P(\epsilon_1 c_1, \ldots, \epsilon_n c_n) \not= 0. \] \end{prop} \medskip Clearly, any assignment of signs given by Proposition \ref{prop6} will make $\det(E + A) \not= 0$, proving Proposition \ref{prop5}. $\mathchoice\sqr76\sqr76\sqr{2.1}3\sqr{1.5}3$ \medskip It remains to prove Proposition \ref{prop6}. \bigskip\noindent {\it Proof of Proposition \ref{prop6}\/}. We proceed by induction on $n$ (starting with $n = 1$). For $n = 1$, the polynomial $P(x_1)$ is of the form $P(x_1) = a + bx_1$, with $b\not= 0$ since $\mathrm{deg}(P) = 1$. Obviously, for any $c\not= 0$, either $a + bc\not= 0$ or $a - bc \not= 0$ (otherwise, $2bc = 0$, contradicting $b\not= 0$, $c\not= 0$ and the ring being an integral domain of characteristic $\not= 2$). \medskip Assume the induction hypothesis holds for any $n\geq 1$ and let $P(x_1, \ldots, x_{n+1})$ be a polynomial of degree $n+1$ satisfying the conditions of Proposition \ref{prop6}. Then, $P$ must be of the form \[ P(x_1, \ldots, x_n, x_{n+1}) = Q(x_1, \ldots, x_n) + S(x_1, \ldots, x_n)x_{n+1}, \] where both $Q(x_1, \ldots, x_n)$ and $S(x_1, \ldots, x_n)$ are polynomials in $x_1, \ldots, x_n$ and $S(x_1, \ldots, x_n)$ is of degree $n$ and all monomials in $S$ are products of distinct variables. By the induction hypothesis, we can find $(\epsilon_1, \ldots, \epsilon_n)$, with $\epsilon_i = \pm 1$, so that \[ S(\epsilon_1 c_1, \ldots, \epsilon_n c_n) \not= 0. \] But now, we are back to the case $n = 1$ with the polynomial \[ Q(\epsilon_1 c_1, \ldots, \epsilon_n c_n) + S(\epsilon_1 c_1, \ldots, \epsilon_n c_n)x_{n+1}, \] and we can find $\epsilon_{n+1} = \pm 1$ so that \[ P(\epsilon_1 c_1, \ldots, \epsilon_n c_n, \epsilon_{n+1}c_{n+1}) = Q(\epsilon_1 c_1, \ldots, \epsilon_n c_n) + S(\epsilon_1 c_1, \ldots, \epsilon_n c_n)\epsilon_{n+1}c_{n+1} \not= 0, \] establishing the induction hypothesis. $\mathchoice\sqr76\sqr76\sqr{2.1}3\sqr{1.5}3$ \medskip Note that in Proposition \ref{prop5}, the $c_i$ can be made arbitrarily small or large, as long as they are not zero. Thus, we see as a corollary that any matrix can be made invertible by a very small perturbation of its diagonal elements. What matters is the signs that are assigned to the perturbation. \medskip Another nice proof of Fact E is given in a short note by William Kahan \cite{Kahan2004}. Due to its elegance, we feel compelled to sketch Kahan's proof. This proof uses two facts: \begin{enumerate} \item[(1)] If $A = (A_1, \ldots, A_{n-1}, U)$ and $B = (A_1, \ldots, A_{n-1}, V)$ are two $n\times n$ matrices that differ in their last column, then \[ \det(A + B) = 2^{n-1}(\det(A) + \det(B)). \] This is because determinants are multilinear (alternating) maps of their columns. Therefore, if $\det(A) = \det(B) = 0$, then $\det(A + B) = 0$. Obviously, this fact also holds whenever $A$ and $B$ differ by just one column (not just the last one). \item[(2)] For every $k = 0, \ldots 2^{n} -1$, if we write $k$ in binary as $k = k_n \cdots k_1$, then let $E_k$ be the diagonal matrix whose $i$th diagonal entry is $-1$ iff $k_i = 1$, else $+1$ iff $k_i = 0$. For example, $E_0 = I$ and $E_{2^n -1} = -I$. Observe that $E_k$ and $E_{k+1}$ differ by exactly one column. Then, it is easy to see that \[ E_0 + E_1 + \cdots + E_{2^n -1} = 0. \] \end{enumerate} The proof proceeds by contradiction. Assume that $\det(I + E_kA) = 0$, \\ for $k = 0, \ldots, 2^n -1$. The crux of the proof is that \[ \det(I + E_0A + I + E_1A + I + E_2A + \cdots + I + E_{2^n -1}A) = 0. \] However, as $E_0 + E_1 + \cdots + E_{2^n -1} = 0$, we see that \[ I + E_0A + I + E_1A + I + E_2A + \cdots + I + E_{2^n -1}A = 2^n I, \] and so, \[ 0 = \det(I + E_0A + I + E_1A + I + E_2A + \cdots + I + E_{2^n -1}A) = \det(2^n I) = 2^n \not= 0, \] a contradiction! \medskip To prove that $\det(I + E_0A + I + E_1A + I + E_2A + \cdots + I + E_{2^n -1}A) = 0$, we observe using fact (2) that, \[ \det(I + E_{2i}A + I + E_{2i+1}A) = \det(I + E_{2i}A) + \det(I + E_{2i+1}A) = 0, \] for $i = 0, \ldots, 2^{n - 1} -1$; similarly, \[ \det(I + E_{4i}A + I + E_{4i+1}A + I + E_{4i+2}A + I + E_{4i+3}A) = 0, \] for $i = 0, \ldots, 2^{n - 2} -1$; by induction, we get \[ \det(I + E_0A + I + E_1A + I + E_2A + \cdots + I + E_{2^n -1}A) = 0, \] which concludes the proof. \bigskip\noindent {\bf Final Questions\/}: \begin{enumerate} \item[(1)] When was Fact E first stated and by whom (similarly for Proposition \ref{prop4})? \item[(2)] Can Proposition \ref{prop5} be generalized to non-diagonal matrices (in an interesting way)? \end{enumerate}
{ "timestamp": "2006-06-13T23:47:20", "yymm": "0606", "arxiv_id": "math/0606320", "language": "en", "url": "https://arxiv.org/abs/math/0606320", "abstract": "This note contains two remarks. The first remark concerns the extension of the well-known Cayley representation of rotation matrices by skew symmetric matrices to rotation matrices admitting -1 as an eigenvalue and then to all orthogonal matrices. We review a method due to Hermann Weyl and another method involving multiplication by a diagonal matrix whose entries are +1 or -1. The second remark has to do with ways of flipping the signs of the entries of a diagonal matrix, C, with nonzero diagonal entries, obtaining a new matrix, E, so that E + A is invertible, where A is any given matrix (invertible or not).", "subjects": "Numerical Analysis (math.NA); General Mathematics (math.GM)", "title": "Remarks on the Cayley Representation of Orthogonal Matrices and on Perturbing the Diagonal of a Matrix to Make it Invertible", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.990140146733106, "lm_q2_score": 0.8080672181749422, "lm_q1q2_score": 0.80009979397395 }
https://arxiv.org/abs/0911.4204
Maximal independent sets and separating covers
In 1973, Katona raised the problem of determining the maximum number of subsets in a separating cover on n elements. The answer to Katona's question turns out to be the inverse to the answer to a much simpler question: what is the largest integer which is the product of positive integers with sum n? We give a combinatorial explanation for this relationship, via Moon and Moser's answer to a question of Erdos: how many maximal independent sets can a graph on n vertices have? We conclude by showing how Moon and Moser's solution also sheds light on a problem of Mahler and Popken's about the complexity of integers.
\section{Introduction} We begin with a simply stated problem, which has made numerous appearances in mathematics competitions:% \footnote{In particular, the 1976 IMO asked for the $n=1976$ case, the 1979 Putnam asked for the $n=1979$ case, and on April 23rd 2002, the 3rd Community College of Philadelphia Colonial Mathematics Challenge asked for the $n=2002$ case.} % what is the largest number which can be written as the product of positive integers that sum to $n$? We denote this number by $\ell(n)$. A moment's thought shows that one should use as many $3$s as possible; if $m\ge 5$ appears in the product then it can be replaced by $3(m-3)>m$, and while $2$s and $4$s can occur in the product, the latter can occur at most once since $4\cdot 4<2 \cdot 3 \cdot 3$ and the former at most twice since $2 \cdot 2 \cdot 2<3 \cdot 3$. This shows that for $n\ge 2$, $$ \ell(n) = \left\{ \begin{array}{ll} 3^i & \mbox{if $n=3i$,}\\ 4 \cdot 3^{i-1} & \mbox{if $n=3i+1$,}\\ 2 \cdot 3^i & \mbox{if $n=3i+2$,}\\ \end{array} \right. $$ while $\ell(1)=1$. Note that it follows from the combinatorial definition of $\ell$ that this function is strictly increasing and {\it super-multiplicative\/}, meaning that it satisfies $\ell(n_1)\ell(n_2)\le\ell(n_1+n_2)$. In 1973, G. O. H. Katona~\cite[Problem 8, p. 306]{katona:combinatorial-s:} posed a problem which looks completely unlike the determination of $\ell(n)$. A {\it separating cover\/}% \footnote{We make this slight deviation from Katona's original formulation so that $s(1)=1$.} over the ground set $X$ is a collection $\S$ of subsets of $X$ which satisfies two properties: \begin{itemize} \item the union of the sets in $\S$ is all of $X$, and \item for every pair of distinct elements $x,y\in X$ there are disjoint sets $S,T\in\S$ with $x\in S$ and $y\in T$. \end{itemize} Katona asked about the function $$ s(m)=\min\{n : \mbox{there is a separating cover on $m$ elements with $n$ sets\}}. $$ M.-C. Cai and A. C. C. Yao gave independent solutions several years later. \begin{theorem}[Cai~\cite{cai:solutions-to-ed:} and Yao~\cite{yao:on-a-problem-of:}, independently]\label{sss} For all $m\ge 2$, $$ s(m) = \left\{ \begin{array}{ll} 3i & \mbox{if $2 \cdot 3^{i-1}<m\le 3^i$,}\\ 3i+1 & \mbox{if $3^i<m\le 4 \cdot 3^{i-1}$,}\\ 3i+2 & \mbox{if $4 \cdot 3^{i-1}<m\le 2 \cdot 3^i$,}\\ \end{array} \right. $$ while $s(1)=1$. \end{theorem} Thus $s(\ell(n))=n$ for all positive integers $n$ --- in other words, $s$ is a left inverse of $\ell$. Ironically, the question we began with appears at the beginning of R. Honsberger's {\it Mathematical Gems III\/}~\cite{honsberger:mathematical-ge:}, while Katona's problem occurs at the end, where Honsberger describes the proof as ``long and much more complicated than the arguments in the earlier chapters.'' We present a short combinatorial explanation for the equivalence of these two problems. \section{A Combinatorial Interpretation of ${\ell}$} In order to give a combinatorial explanation for why $s(\ell(n))=n$, we first need a combinatorial interpretation of $\ell$. We use a graph-theoretic interpretation, although several others are available.% \footnote{Another --- in terms of integer complexity --- is given later in this note. Additionally, $\ell(n)$ is the order of the largest abelian subgroup of the symmetric group of order $n$; see Bercov and Moser~\cite{bercov:on-abelian-perm:}.} % Let $G$ be a graph over the vertex set $V(G)$. A subset $I\subseteq V(G)$ is {\it independent\/} if there is no edge between any two vertices of $I$, and it is a {\it maximal independent set (MIS)\/} if it is not properly contained in any other independent set. In the 1960s, P. Erd\H{o}s asked how many MISes a graph on $n$ vertices could have, which we define as $$ g(n)=\max\{m : \mbox{there is graph on $n$ vertices with $m$ MISes\}}. $$ Let us denote by $m(G)$ the number of MISes in the graph $G$. This quantity is particularly easy to compute when $G$ is a disjoint union: \begin{proposition}\label{mis-product} The disjoint union of the graphs $G$ and $H$ has $m(G)m(H)$ MISes. \end{proposition} \begin{proof} For any MIS $M$ of this union, $M\cap V(G)$ must be an MIS of $G$ and $M\cap V(H)$ must be an MIS of $H$. Conversely, if $M_G$ and $M_H$ are MISes of $G$ and $H$, respectively, then $M_G\cup M_H$ is an MIS of the disjoint union of $G$ and $H$. \end{proof} Because the complete graph on $n$ vertices has $n$ MISes, Proposition~\ref{mis-product} implies that $g(n)\ge\ell(n)$ for all positive integers $n$; we need only take a disjoint union of edges, triangles, and complete graphs on $4$ vertices to achieve this lower bound. In 1965, J. W. Moon and L. Moser proved that this is best possible. \begin{theorem}[Moon and Moser~\cite{moon:on-cliques-in-g:}]\label{mis} For all positive integers $n$, $g(n)=\ell(n)$. \end{theorem} Indeed, Moon and Moser showed that the only extremal graphs (the graphs with $g(n)$ MISes) are those built by taking disjoint copies of edges, triangles, and complete graphs on $4$ vertices in the quantities suggested by the formula for $\ell$. (In the case $n=3i+1\ge 4$ there are two extremal graphs, one with $i-1$ triangles and two disjoint edges, the other with $i-1$ triangles and a complete graph on $4$ vertices.) \section{A Short Proof of Theorem~\ref{mis}} Before demonstrating the relationship between MISes and separating covers, we pause to present a short proof of Moon and Moser's theorem. First we need a definition: for a set $X\subseteq V(G)$, we denote by $G-X$ the graph obtained by removing the vertices $X$ from $G$ and all edges incident to vertices in $X$. When $X=\{v\}$, we abbreviate this notation to $G-v$. Our proof makes extensive use of the following upper bound. \begin{proposition}\label{mis-recurse} For any graph $G$ and vertex $v\in V(G)$, we have $$ m(G)\le m(G-v)+m(G-N[v]), $$ where $N[v]$ denotes the {\it closed neighborhood\/} of $v$, i.e., $v$ together with its neighbors. \end{proposition} \begin{proof} The map $M\mapsto M-v$ gives a bijection between MISes of $G$ containing $v$ and MISes of $G-N[v]$. The proof is completed by noting that every MIS of $G$ that does not contain $v$ is also an MIS of $G-v$. \end{proof} \newenvironment{mis-proof}{\medskip\noindent {\it Proof of Theorem~\ref{mis}.\/}}{\qed\bigskip} \begin{mis-proof} Our proof is by induction on $n$, and we prove the stronger statement which characterizes the extremal graphs. It is easy to check the theorem for graphs with five or fewer vertices, so take $G$ to be a graph on $n\ge 6$ vertices, and assume the theorem holds for graphs with fewer than $n$ vertices. If $G$ contains a vertex of degree $0$, that is, an isolated vertex, then clearly $m(G)\le g(n-1)=\ell(n-1)<\ell(n)$. If $G$ contains a vertex $v$ of degree $1$ then, letting $w$ denote the sole vertex adjacent to $v$, we have by Proposition~\ref{mis-recurse} that $$ m(G) \le m(G-w)+m(G-N[w]) \le 2\ell(n-2) = \left\{\begin{array}{ll} 8\cdot 3^{i-2}&\mbox{if $n=3i$,}\\ 4\cdot 3^{i-1}&\mbox{if $n=3i+1$,}\\ 2\cdot 3^{i}&\mbox{if $n=3i+2$.} \end{array}\right. $$ In all three cases we have an upper bound of at most $\ell(n)$, with equality if and only if $n=3i+1$ and $G$ is a disjoint union of $i-1$ triangles and two edges, or $n=3i+2$ and $G$ is a disjoint union of $i$ triangles and an edge. If $G$ contains a vertex $v$ of degree $3$ or greater, then we have $$ m(G) \le m(G-v)+m(G-N[v]) \le \ell(n-1)+\ell(n-4) = \left\{\begin{array}{ll} 8\cdot 3^{i-2}&\mbox{if $n=3i$,}\\ 4\cdot 3^{i-1}&\mbox{if $n=3i+1$,}\\ 16\cdot 3^{i-2}&\mbox{if $n=3i+2$.} \end{array}\right. $$ Again, all three cases give an upper bound of at most $\ell(n)$, with equality if and only if $n=3i+1$ and $G$ is a disjoint union of $i-1$ triangles together with a complete graph on $4$ vertices. This leaves us to consider the case where every vertex of $G$ has degree $2$, which implies that $G$ consists of a disjoint union of cycles. If each of these cycles is a triangle, then $n=3i$ and $G$ is a disjoint union of $i$ triangles, as desired. Thus we may assume that at least one connected component of $G$ is a cycle of length $j\ge 4$, which we denote by $C_j$. Our goal in this case is to show that $G$ is not extremal (i.e., $m(G)<\ell(n)$), and by the super-multiplicativity of $\ell$, it suffices to show that this single cycle of length $j$ is not extremal. It is easy to check that $m(C_4)=2<4=\ell(4)$ and $m(C_5)=5<6=\ell(5)$, it therefore suffices to show that $m(C_j)<\ell(j)$ for $j\ge 6$. (In fact, F\"uredi~\cite{furedi:the-number-of-m:} found $m(C_j)$ exactly --- it is the $j$th Perrin number.) Label the vertices of our cycle on $j\ge 6$ vertices as $u,v,w,\dots$ so that $u$ is adjacent to $v$ which is in turn adjacent to $w$. By applying Proposition~\ref{mis-recurse} twice, we see that for $j\ge 6$, \begin{eqnarray*} m(C_j) &\le& m(C_j-w)+m(C_j-N[w])\\ &\le& m(C_j-w-u)+m(C_j-w-N[u])+m(C_j-N[w])\\ &\le& 2\ell(j-3)+\ell(j-4), \end{eqnarray*} which is strictly less that $3\ell(j-3)=\ell(j)$, completing the proof. \end{mis-proof} \section{A Combinatorial Explanation for ${s(\ell(n))=n}$} With Moon and Moser's Theorem~\ref{mis} proved, we are now ready to explain the connection to separating covers. Propositions~\ref{mis2sss} and \ref{sss2mis} illuminate the connection between separating covers and MISes, and then Proposition~\ref{main} gives a combinatorial explanation for why $s$ is a left inverse of $\ell=g$. \begin{proposition}\label{mis2sss} From a graph on $n$ vertices with $m$ MISes one can construct a separating cover on $m$ elements with at most $n$ sets. \end{proposition} \begin{proof} Take $G$ to be a graph with $n$ vertices and $m$ MISes and let $\mathcal{M}$ denote the collection of MISes in $G$. The separating cover promised consists of the family of sets $\{S_v : v\in V(G)\}$ where $$ S_v=\{M\in\mathcal{M} : v\in M\}. $$ Clearly this is a family with $m$ elements (the MISes $\mathcal{M}$) and $n$ (not necessarily distinct) sets (one for each vertex of $G$), and this family covers the set $\mathcal{M}$ because each MIS lies in at least one $S_v$, so it remains to check only that it is separating. Take distinct sets $M, N\in\mathcal{M}$. Because $M$ and $N$ are both maximal there is some vertex $u\in M\setminus N$ . By the maximality of $N$, it must contain a vertex $v$ adjacent to $u$. Therefore $M\in S_u$, $N\in S_v$, and because $u$ and $v$ are adjacent, $S_u\cap S_v=\emptyset$, completing the proof. \end{proof} \begin{proposition}\label{sss2mis} From a separating cover on $m$ elements with $n$ sets one can construct a graph on $n$ vertices with at least $m$ MISes. \end{proposition} \begin{proof} Let $\S$ be such a cover over the ground set $X$. We define a graph $G$ on the vertices $\S$ where $S\in\S$ is adjacent to $T\in\S$ if and only if they are disjoint. For each $x\in X$, the set $$ I_x=\{S\in\S : x\in S\} $$ is an independent set in $G$. For each $x\in X$, choose an MIS $M_x\supseteq I_x$. We have only to show that these MISes are distinct. Take distinct elements $x,y\in X$. Because $\S$ is separating, there are disjoint sets $S,T\in\S$ with $x\in S$ and $y\in T$. Therefore $S\in M_x$, $T\in M_y$, and since $S$ and $T$ are disjoint they are adjacent in $G$, so $T\notin M_x$, and thus $M_x\neq M_y$. \end{proof} \begin{proposition}\label{main} For all positive integers $m$ and $n$, \begin{eqnarray*} s(m)&=&\min\{n : g(n)\ge m\},\\ g(n)&=&\max\{m : s(m)\le n\}. \end{eqnarray*} \end{proposition} \begin{proof} First observe that $s$ and $g$ are both nondecreasing. The proof then follows from the two claims \begin{enumerate} \item[(1)] If $s(m)\le n$ then $g(n)\ge m$, and \item[(2)] If $g(n)\ge m$ then $s(m)\le n$ \end{enumerate} To prove (1), suppose that $s(m)\le n$. Then there is a separating cover with $m$ elements and at most $n$ sets, so by Proposition~\ref{sss2mis}, there is a graph with at most $n$ vertices and at least $m$ MISes. This and the fact that $g$ is nondecreasing establish that $g(n)\ge m$. Now suppose that $g(n)\ge m$. Then there is a graph with $n$ vertices and at least $m$ MISes, so by Proposition~\ref{mis2sss}, there is a separating cover with at least $m$ elements and at most $n$ sets. Because $s$ is nondecreasing, we conclude that $s(m)\le n$, proving (2). \end{proof} \section{Integer Complexity} We conclude with another appearance of $g$. The {\it complexity\/}, $c(m)$, of the integer $m$ is the least number of $1$s needed to represent it using only $+$s, $\cdot$s, and parentheses. For example, the complexity of $10$ is $7$, and there are essentially three different minimal expressions: $$ 10=(1+1+1)(1+1+1)+1=(1+1)(1+1+1+1+1)=(1+1)((1+1)(1+1)+1), $$ Figure~\ref{fig-complexity} shows a plot of the complexities of the first $1000$ integers. \begin{figure} \begin{center} \savedata{\plotdata}[{{1,1},{2,2},{3,3},{4,4},{5,5},{6,5},{7,6},{8,6},{9,6},{10,7},{11,8},{12,7},{13,8},{14,8},{15,8},{16,8},{17,9},{18,8},{19,9},{20,9},{21,9},{22,10},{23,11},{24,9},{25,10},{26,10},{27,9},{28,10},{29,11},{30,10},{31,11},{32,10},{33,11},{34,11},{35,11},{36,10},{37,11},{38,11},{39,11},{40,11},{41,12},{42,11},{43,12},{44,12},{45,11},{46,12},{47,13},{48,11},{49,12},{50,12},{51,12},{52,12},{53,13},{54,11},{55,12},{56,12},{57,12},{58,13},{59,14},{60,12},{61,13},{62,13},{63,12},{64,12},{65,13},{66,13},{67,14},{68,13},{69,14},{70,13},{71,14},{72,12},{73,13},{74,13},{75,13},{76,13},{77,14},{78,13},{79,14},{80,13},{81,12},{82,13},{83,14},{84,13},{85,14},{86,14},{87,14},{88,14},{89,15},{90,13},{91,14},{92,14},{93,14},{94,15},{95,14},{96,13},{97,14},{98,14},{99,14},{100,14},{101,15},{102,14},{103,15},{104,14},{105,14},{106,15},{107,16},{108,13},{109,14},{110,14},{111,14},{112,14},{113,15},{114,14},{115,15},{116,15},{117,14},{118,15},{119,15},{120,14},{121,15},{122,15},{123,15},{124,15},{125,15},{126,14},{127,15},{128,14},{129,15},{130,15},{131,16},{132,15},{133,15},{134,16},{135,14},{136,15},{137,16},{138,15},{139,16},{140,15},{141,16},{142,16},{143,16},{144,14},{145,15},{146,15},{147,15},{148,15},{149,16},{150,15},{151,16},{152,15},{153,15},{154,16},{155,16},{156,15},{157,16},{158,16},{159,16},{160,15},{161,16},{162,14},{163,15},{164,15},{165,15},{166,16},{167,17},{168,15},{169,16},{170,16},{171,15},{172,16},{173,17},{174,16},{175,16},{176,16},{177,17},{178,17},{179,18},{180,15},{181,16},{182,16},{183,16},{184,16},{185,16},{186,16},{187,17},{188,17},{189,15},{190,16},{191,17},{192,15},{193,16},{194,16},{195,16},{196,16},{197,17},{198,16},{199,17},{200,16},{201,17},{202,17},{203,17},{204,16},{205,17},{206,17},{207,17},{208,16},{209,17},{210,16},{211,17},{212,17},{213,17},{214,18},{215,17},{216,15},{217,16},{218,16},{219,16},{220,16},{221,17},{222,16},{223,17},{224,16},{225,16},{226,17},{227,18},{228,16},{229,17},{230,17},{231,17},{232,17},{233,18},{234,16},{235,17},{236,17},{237,17},{238,17},{239,18},{240,16},{241,17},{242,17},{243,15},{244,16},{245,17},{246,16},{247,17},{248,17},{249,17},{250,17},{251,18},{252,16},{253,17},{254,17},{255,17},{256,16},{257,17},{258,17},{259,17},{260,17},{261,17},{262,18},{263,19},{264,17},{265,18},{266,17},{267,18},{268,18},{269,19},{270,16},{271,17},{272,17},{273,17},{274,18},{275,17},{276,17},{277,18},{278,18},{279,17},{280,17},{281,18},{282,18},{283,19},{284,18},{285,17},{286,18},{287,18},{288,16},{289,17},{290,17},{291,17},{292,17},{293,18},{294,17},{295,18},{296,17},{297,17},{298,18},{299,19},{300,17},{301,18},{302,18},{303,18},{304,17},{305,18},{306,17},{307,18},{308,18},{309,18},{310,18},{311,19},{312,17},{313,18},{314,18},{315,17},{316,18},{317,19},{318,18},{319,19},{320,17},{321,18},{322,18},{323,18},{324,16},{325,17},{326,17},{327,17},{328,17},{329,18},{330,17},{331,18},{332,18},{333,17},{334,18},{335,19},{336,17},{337,18},{338,18},{339,18},{340,18},{341,19},{342,17},{343,18},{344,18},{345,18},{346,19},{347,20},{348,18},{349,19},{350,18},{351,17},{352,18},{353,19},{354,18},{355,19},{356,19},{357,18},{358,19},{359,20},{360,17},{361,18},{362,18},{363,18},{364,18},{365,18},{366,18},{367,19},{368,18},{369,18},{370,18},{371,19},{372,18},{373,19},{374,19},{375,18},{376,19},{377,19},{378,17},{379,18},{380,18},{381,18},{382,19},{383,20},{384,17},{385,18},{386,18},{387,18},{388,18},{389,19},{390,18},{391,19},{392,18},{393,19},{394,19},{395,19},{396,18},{397,19},{398,19},{399,18},{400,18},{401,19},{402,19},{403,19},{404,19},{405,17},{406,18},{407,19},{408,18},{409,19},{410,18},{411,19},{412,19},{413,20},{414,18},{415,19},{416,18},{417,19},{418,19},{419,20},{420,18},{421,19},{422,19},{423,19},{424,19},{425,19},{426,19},{427,19},{428,20},{429,19},{430,19},{431,20},{432,17},{433,18},{434,18},{435,18},{436,18},{437,19},{438,18},{439,19},{440,18},{441,18},{442,19},{443,20},{444,18},{445,19},{446,19},{447,19},{448,18},{449,19},{450,18},{451,19},{452,19},{453,19},{454,20},{455,19},{456,18},{457,19},{458,19},{459,18},{460,19},{461,20},{462,19},{463,20},{464,19},{465,19},{466,20},{467,21},{468,18},{469,19},{470,19},{471,19},{472,19},{473,20},{474,19},{475,19},{476,19},{477,19},{478,20},{479,21},{480,18},{481,19},{482,19},{483,19},{484,19},{485,19},{486,17},{487,18},{488,18},{489,18},{490,19},{491,20},{492,18},{493,19},{494,19},{495,18},{496,19},{497,20},{498,19},{499,20},{500,19},{501,20},{502,20},{503,21},{504,18},{505,19},{506,19},{507,19},{508,19},{509,20},{510,19},{511,19},{512,18},{513,18},{514,19},{515,20},{516,19},{517,20},{518,19},{519,20},{520,19},{521,20},{522,19},{523,20},{524,20},{525,19},{526,20},{527,20},{528,19},{529,20},{530,20},{531,20},{532,19},{533,20},{534,20},{535,21},{536,20},{537,21},{538,21},{539,20},{540,18},{541,19},{542,19},{543,19},{544,19},{545,19},{546,19},{547,20},{548,20},{549,19},{550,19},{551,20},{552,19},{553,20},{554,20},{555,19},{556,20},{557,21},{558,19},{559,20},{560,19},{561,20},{562,20},{563,21},{564,20},{565,20},{566,21},{567,18},{568,19},{569,20},{570,19},{571,20},{572,20},{573,20},{574,19},{575,20},{576,18},{577,19},{578,19},{579,19},{580,19},{581,20},{582,19},{583,20},{584,19},{585,19},{586,20},{587,21},{588,19},{589,20},{590,20},{591,20},{592,19},{593,20},{594,19},{595,20},{596,20},{597,20},{598,20},{599,21},{600,19},{601,20},{602,20},{603,20},{604,20},{605,20},{606,20},{607,21},{608,19},{609,20},{610,20},{611,21},{612,19},{613,20},{614,20},{615,20},{616,20},{617,21},{618,20},{619,21},{620,20},{621,20},{622,21},{623,21},{624,19},{625,20},{626,20},{627,20},{628,20},{629,20},{630,19},{631,20},{632,20},{633,20},{634,21},{635,20},{636,20},{637,20},{638,21},{639,20},{640,19},{641,20},{642,20},{643,21},{644,20},{645,20},{646,20},{647,21},{648,18},{649,19},{650,19},{651,19},{652,19},{653,20},{654,19},{655,20},{656,19},{657,19},{658,20},{659,21},{660,19},{661,20},{662,20},{663,20},{664,20},{665,20},{666,19},{667,20},{668,20},{669,20},{670,21},{671,21},{672,19},{673,20},{674,20},{675,19},{676,20},{677,21},{678,20},{679,20},{680,20},{681,21},{682,21},{683,22},{684,19},{685,20},{686,20},{687,20},{688,20},{689,21},{690,20},{691,21},{692,21},{693,20},{694,21},{695,21},{696,20},{697,21},{698,21},{699,21},{700,20},{701,21},{702,19},{703,20},{704,20},{705,20},{706,21},{707,21},{708,20},{709,21},{710,21},{711,20},{712,21},{713,22},{714,20},{715,20},{716,21},{717,21},{718,22},{719,23},{720,19},{721,20},{722,20},{723,20},{724,20},{725,20},{726,20},{727,21},{728,20},{729,18},{730,19},{731,20},{732,19},{733,20},{734,21},{735,20},{736,20},{737,21},{738,19},{739,20},{740,20},{741,20},{742,21},{743,22},{744,20},{745,21},{746,21},{747,20},{748,21},{749,22},{750,20},{751,21},{752,21},{753,21},{754,21},{755,21},{756,19},{757,20},{758,20},{759,20},{760,20},{761,21},{762,20},{763,20},{764,21},{765,20},{766,21},{767,22},{768,19},{769,20},{770,20},{771,20},{772,20},{773,21},{774,20},{775,21},{776,20},{777,20},{778,21},{779,21},{780,20},{781,21},{782,21},{783,20},{784,20},{785,21},{786,21},{787,22},{788,21},{789,22},{790,21},{791,21},{792,20},{793,21},{794,21},{795,21},{796,21},{797,22},{798,20},{799,21},{800,20},{801,21},{802,21},{803,21},{804,21},{805,21},{806,21},{807,22},{808,21},{809,22},{810,19},{811,20},{812,20},{813,20},{814,21},{815,20},{816,20},{817,21},{818,21},{819,20},{820,20},{821,21},{822,21},{823,22},{824,21},{825,20},{826,21},{827,22},{828,20},{829,21},{830,21},{831,21},{832,20},{833,21},{834,21},{835,22},{836,21},{837,20},{838,21},{839,22},{840,20},{841,21},{842,21},{843,21},{844,21},{845,21},{846,21},{847,21},{848,21},{849,22},{850,21},{851,22},{852,21},{853,22},{854,21},{855,20},{856,21},{857,22},{858,21},{859,22},{860,21},{861,21},{862,22},{863,23},{864,19},{865,20},{866,20},{867,20},{868,20},{869,21},{870,20},{871,21},{872,20},{873,20},{874,21},{875,21},{876,20},{877,21},{878,21},{879,21},{880,20},{881,21},{882,20},{883,21},{884,21},{885,21},{886,22},{887,23},{888,20},{889,21},{890,21},{891,20},{892,21},{893,22},{894,21},{895,22},{896,20},{897,21},{898,21},{899,22},{900,20},{901,21},{902,21},{903,21},{904,21},{905,21},{906,21},{907,22},{908,22},{909,21},{910,21},{911,22},{912,20},{913,21},{914,21},{915,21},{916,21},{917,22},{918,20},{919,21},{920,21},{921,21},{922,22},{923,22},{924,21},{925,21},{926,22},{927,21},{928,21},{929,22},{930,21},{931,21},{932,22},{933,22},{934,23},{935,21},{936,20},{937,21},{938,21},{939,21},{940,21},{941,22},{942,21},{943,22},{944,21},{945,20},{946,21},{947,22},{948,21},{949,21},{950,21},{951,22},{952,21},{953,22},{954,21},{955,22},{956,22},{957,22},{958,23},{959,22},{960,20},{961,21},{962,21},{963,21},{964,21},{965,21},{966,21},{967,22},{968,21},{969,21},{970,21},{971,22},{972,19},{973,20},{974,20},{975,20},{976,20},{977,21},{978,20},{979,21},{980,21},{981,20},{982,21},{983,22},{984,20},{985,21},{986,21},{987,21},{988,21},{989,22},{990,20},{991,21},{992,21},{993,21},{994,22},{995,22},{996,21},{997,22},{998,22},{999,20},{1000,21}}] \psset{xunit=0.00474525in, yunit=0.0520833in} \begin{pspicture}(0,-3)(1053.684211,28.80000000) \psaxes[dy=5,Dy=5,dx=100,Dx=100,showorigin=false](0,0)(999,23) \rput[c](0,25.4){$c(m)$} \rput[l](1026.34,0){$m$} \dataplot[plotstyle=dots,dotstyle=*,dotsize=3\psxunit]{\plotdata} \end{pspicture} \end{center} \caption{The complexities of the first 1000 integers.}\label{fig-complexity} \end{figure} This definition was first considered by Mahler and Popken~\cite{mahler:on-a-maximum-pr:}, and while a straightforward recurrence, $$ c(m)=\min \{c(d)+c(m/d) : d\divides m\}\cup\{c(i)+c(m-i) : 1\le i\le m-1\}, $$ is easy to verify, several outstanding conjectures and questions remain, for which we refer to R. K. Guy~\cite{guy:unsolved-proble:}. In that article, Guy mentions that J. Selfridge gave an inductive proof of the following result. \begin{proposition}[Selfridge~{[unpublished]}]\label{prop-integer-complexity} The greatest integer of complexity $n$ is $g(n)$. \end{proposition} One direction of Selfridge's proposition is clear: the problem we began with shows that $\ell(n)=g(n)$ has complexity at most $n$. In a final demonstration of the surprising versatility of Moon and Moser's Theorem~\ref{mis}, we show how it implies the other direction, via the following construction. \begin{proposition}\label{comp2mis} From an expression of the integer $m$ with $n$ $1$s one can construct a graph on $n$ vertices with $m$ MISes. \end{proposition} \begin{figure} \begin{center} \psset{xunit=0.007in, yunit=0.007in} \psset{linewidth=0.02in} \begin{pspicture}(0,0)(120,80) \pscircle*(0,20){0.04in} \pscircle*(0,60){0.04in} \pscircle*(40,20){0.04in} \pscircle*(40,60){0.04in} \pscircle*(80,40){0.04in} \pscircle*(120,20){0.04in} \pscircle*(120,60){0.04in} \psline(0,20)(0,60) \psline(40,20)(40,60) \psline(40,20)(120,60) \psline(40,60)(120,20) \psline(120,20)(120,60) \end{pspicture} \end{center} \caption{The construction described in the proof of Proposition~\ref{prop-integer-complexity}, applied to the expression $$ 10=(1+1)((1+1)(1+1)+1). $$ There are graphs on $7$ vertices with more MISes than the graph shown because $10$ is not the greatest integer of complexity $7$ ($12$ is).} \label{prop-integer-complexity-figure} \end{figure} \begin{proof} Before describing our inductive construction we need a definition. Given graphs $G$ and $H$, their {\it join\/} is the graph $G+H$ obtained from their disjoint union $G\cup H$ by adding all edges connecting vertices of $G$ with vertices of $H$. We know already from Proposition~\ref{mis-product} that $m(G\cup H)=m(G)m(H)$, and a similar formula for joins is easy to verify: $m(G+H)=m(G)+m(H)$ because every MIS in $G+H$ is either an MIS of $G$ or an MIS of $H$. Now suppose we have an expression of the integer $m$ with $n$ $1$s. If $n=1$, then there is only one such expression, $1$, and we associate to this expression the one vertex graph. If $n\ge 2$, then any such expression must decompose as either $e_1+e_2$ or $e_1e_2$, where $e_1$ and $e_2$ are expressions with fewer $1$s. If our expression is $e_1+e_2$ then we associate it to the join of the graphs associated to $e_1$ and $e_2$, and if our expression is $e_1e_2$ then we associate it to the disjoint union of the graphs associated to $e_1$ and $e_2$. Figure~\ref{prop-integer-complexity-figure} shows an example. It follows that the resulting graph has precisely as many vertices as the expression has $1$s, and precisely $m$ MISes. \end{proof} \minisec{Acknowledgements} I am grateful to the referees for their detailed and insightful comments. In particular, the change from ``separating families'' to ``separating covers,'' which simplified many of these results, was suggested by one of the referees. \bibliographystyle{acm}
{ "timestamp": "2010-08-30T02:00:22", "yymm": "0911", "arxiv_id": "0911.4204", "language": "en", "url": "https://arxiv.org/abs/0911.4204", "abstract": "In 1973, Katona raised the problem of determining the maximum number of subsets in a separating cover on n elements. The answer to Katona's question turns out to be the inverse to the answer to a much simpler question: what is the largest integer which is the product of positive integers with sum n? We give a combinatorial explanation for this relationship, via Moon and Moser's answer to a question of Erdos: how many maximal independent sets can a graph on n vertices have? We conclude by showing how Moon and Moser's solution also sheds light on a problem of Mahler and Popken's about the complexity of integers.", "subjects": "Combinatorics (math.CO)", "title": "Maximal independent sets and separating covers", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9901401455693092, "lm_q2_score": 0.8080672135527632, "lm_q1q2_score": 0.800099788456919 }
https://arxiv.org/abs/2104.05092
Remark on the Chain rule of fractional derivative in the Sobolev framework
A chain rule for power product is studied with fractional differential operators in the framework of Sobolev spaces. The fractional differential operators are defined by the Fourier multipliers. The chain rule is considered newly in the case where the order of differential operators is between one and two. The study is based on the analogy of the classical chain rule or Leibniz rule.
\section{Introduction} \section{Introduction} The chain rule or Leibniz rule is an essential tool to study nonlinear differential equations. In the study of nonlinear partial differential equations(PDEs), fractional differential operators are also known as powerful tools. So, in order to analyze nonlinear PDEs, chain rules for fractional differential operators are naturally required. Even though fractional differential operators may be non-local unlike classical operators, estimates for fractional derivative has been studied on the analogy of classical chain rules. The history of this study can go back at least to the work of Kato and Ponce \cite{KP88}. In this paper, we consider a chain rule corresponding to the identity: \[ \frac{d}{dx} F(u) = F'(u) u', \] in the framework of Riesz potential space $\dot H_p^s = D^{-s} L^p(\mathbb R^d)$, where $s \in \mathbb R$, $d \geq 1$, and $L^p$ is the usual Lebesgue space. $\dot H_p^s$ is also called as homogeneous Sobolev space. The fractional differential operator $D^s = (-\Delta)^{s/2}$ is recognized as a Fourier multiplier by $D^{s} = \mathfrak F^{-1} |\cdot|^s \mathfrak F$, where $\mathfrak F$ is the standard Fourier transform. Especially, we study the case where $F$ behaves like power product, that is, $F(z) \sim |z|^{p-1} z$. In \cite{CW91}, Christ and Weinstein showed the following estimate: \begin{Proposition}[{\cite[Proposition:3.1]{CW91}}] \label{Proposition:1.1} Let $d \geq 1$ and $s \in (0,1)$. Let $F \in C(\mathbb C)$ and $G \in C(\mathbb C:[0,\infty))$ satisfy \[ |F(u) - F(v)| \leq (G(u)+G(v))|u-v|. \] Let $1 \leq p < \infty$, $1 < r, q < \infty$ satisfy \begin{align} \frac 1 p = \frac 1 q + \frac 1 r. \label{eq:1.1} \end{align} The estimate \[ \| F(u) \|_{\dot H_p^s} \leq \| G(u) \|_{L^q} \| u \|_{\dot H_r^s}. \] holds for any $u \in \dot H_r^s$ with $G(u) \in L^q$. \end{Proposition} Roughly speaking, Proposition \ref{Proposition:1.1} asserts that $D^s F(u)$ behaves like $F'(u) D^s u$. Since the Riesz operator $R = D^{-1} \nabla$ is bounded on $L^p$ when $p \in (1,\infty)$, $\| F(u) \|_{\dot H_p^s}$ may be estimated even when $s \geq 1$ by combining the classical chain rule, Proposition \ref{Proposition:1.1}, and the H\"older inequality. On the other hand, when $F(z) = |z|^{\rho-1} z$, Proposition \ref{Proposition:1.1} may be insufficient to handle general cases. For example, when $\rho \in (1,2)$ and $s \in (1,\rho)$, one may regard \[ D^s F(u) \sim D^{s-1} (|u|^{\rho-1} \nabla u). \] By using the fractional Leibniz rule (for example, see \cite[Theorem 1]{GO14}, \cite{ES20}, \cite{FGO18}, and references therein), one can distribute $D^{s-1}$ like $\nabla$ to $\nabla u$ and $|u|^{\rho-1}$. Since $f(z) = |z|^{\rho-1}$ is not differentiable, one cannot control $D^{s-1} |u|^{\rho-1}$ by applying Proposition \ref{Proposition:1.1} directly. However, on the analogy of the fact that $f^\rho \in C^1$ when $f \in C^{\lceil \rho \rceil}$, $D^s (|u|^{\rho-1} u)$ is expected to be controlled similarly when $s < \rho$. Here $\lceil \cdot \rceil$ is the ceiling function defined by $\lceil s \rceil = \min\{m \in \mathbb Z, m \geq s\}$. In order to see this natural expectation, one may need to extend Proposition \ref{Proposition:1.1} with $s \in (1,2)$. We generalize this expectation slightly. With $\rho > 1$, we put $F_\rho \in C^{1}(\mathbb C)$ satisfying $F(0) = F'(0)=0$ and \begin{align} |F_\rho(u) - F_\rho(v)| &\lesssim \max(|u|,|v)^{\rho-1} | u - v|, \nonumber\\ |F_\rho'(u) - F_\rho'(v)| &\lesssim \begin{cases} \max(|u|,|v)^{\rho-2} | u - v| &\mathrm{if} \quad \rho \geq 2,\\ | u - v|^{\rho -1} &\mathrm{if} \quad \rho < 2, \end{cases} \label{eq:1.2} \end{align} where $a \lesssim b$ stands for $a \leq C b$ with some positive constant $C$. We also denote $a \sim b$ when $a \lesssim b$ and $b \lesssim a$. We use these notation though this paper. In \cite{GOV94}, Ginibre, Ozawa, and Velo showed the expectation above in the framework of homogeneous Besov spaces $\dot B_{p,q}^s$. \begin{Proposition}[{\cite[Lemma:3.4]{GOV94}}] \label{Proposition:1.2} Let $d \geq 1$, $\rho \geq 1$, and $s \in (0,\min(2,\rho))$. Let $1 \leq p, r \leq \infty$ and $(\rho-1)^{-1} \leq q \leq \infty$ satisfy \eqref{eq:1.1}. If $u \in \dot B_{r,2}^s \cap L^{(\rho-1)q}$, then \begin{align} \| F_\rho(u) \|_{\dot B_{p,2}^s} \lesssim \| u \|_{L^{(\rho-1)q}}^{\rho-1} \| u \|_{\dot B_{r,2}^s}. \label{eq:1.3} \end{align} \end{Proposition} Since Besov spaces may play a role in useful auxiliary spaces, Proposition \ref{Proposition:1.2} has been used to study even $\dot H_p^s$-valued solutions to some nonlinear PDEs. The advantage to consider the fractional chain rule in the framework of Besov space is the following representation of homogeneous Besov norms: \[ \| u \|_{\dot B_{r,q}^s} \sim \bigg( \int_0^\infty \lambda^{-s q-1} \sup_{|y| < \lambda} \|(\tau_y - 2 + \tau_{-y}) u \|_{L^r}^q d \lambda, \bigg)^{1/q}, \] where $\tau_y u = u(\cdot + y)$, $s \in (0,2)$, and $1 \leq p, q \leq \infty$ (See \cite[6.2.5. Theorem]{BL}, for example.). With this representation, the term $(\tau_y - 2 + \tau_{-y}) u$ gives a clear explanation of the connection between the classical and fractional chain rules. We remark that even though the fractional differential operators are defined by Fourier multipliers, it is nontrivial why $F'(u) \sim |u|^{\rho-1}$ appears as an upper bound of \eqref{eq:1.3} from the view point Fourier multipliers. We further remark that when $\rho =2$, the chain rule may be shown by the argument of Fourier multipliers. See \cite{FGO18}. Proposition \ref{Proposition:1.2} is an effective estimate but it seems more handy to close the argument only with Sobolev spaces. So one of the main purpose of this paper is to show a similar estimate to Proposition \ref{Proposition:1.2} in the framework of $\dot H_p^s$. This is one of the main statement of this paper. \begin{Proposition} \label{Proposition:1.3} Let $\rho > 1$ and $s \in (1,\rho)$. Let $1 < p, r < \infty$ and $(\rho-1)^{-1} < q < \infty$ satisfy \eqref{eq:1.1}. The estimate \[ \| F_\rho(u) \|_{\dot H_p^s} \lesssim \| u \|_{L^{(\rho-1) q}}^{\rho-1} \| u \|_{\dot H_r^s} \] holds for any $u \in \dot H_r^s \cap L^{(\rho-1)q}$. \end{Proposition} We remark that the case when $p$ or $r=1, \infty$ is included in the assumption of Proposition \ref{Proposition:1.2} but not in that of Proposition \ref{Proposition:1.3}. It is because our proof is based on the boundedness of Hardy-Littlewood maximal operator. See next section. The main idea for the proof of Proposition \ref{Proposition:1.3} is to express $\| F(u) \|_{\dot H_p^s}$ with $(\tau_y - 2 + \tau_{-y}) F(u)$ like the proof of Proposition \ref{Proposition:1.2}. Since $\dot H_p^s$ seems not to be expressed like \eqref{eq:1.3}, we rewrite each dyadic components of $F(u)$ by $(\tau_y - 2 + \tau_{-y}) F(u)$. Namely, the identity \begin{align} Q_j F(u)(x) = \frac 1 2 \int (\tau_y - 2 + \tau_{-y}) F(u)(x) \psi_j(y) dy \label{eq:1.4} \end{align} plays a critical role in this paper, where $Q_j$ is the standard $j$-th Littlewood-Paley dyadic operator and $\psi_j$ is the corresponding kernel. The details of $Q_j$ and $\psi_j$ are stated in next section. We note that the identity \[ Q_j F(u)(x) = \int (\tau_y - 1) F(u)(x) \psi_j(y) dy \] plays an essential role in \cite{CW91} as well. In this paper, we deploy a similar but more careful approach with \eqref{eq:1.4}. Our second purpose is to extend Corollary \ref{Corollary:1.4} below following from Proposition \ref{Proposition:1.1}. Here we put $a_+ = \max(a,0)$ and $a_- = \min(a,0)$ for $a \in \mathbb R$. \begin{Corollary} \label{Corollary:1.4} Let $s \in (0,1)$, $\rho \geq 1$. Let $1 \leq p < \infty$ and $1 < q, r < \infty$ satisfy \eqref{eq:1.1}. Then the estimate \begin{align*} &\| F_\rho (u) - F_\rho (v) \|_{\dot H^s_p}\\ &\lesssim (\| u \|_{L^q(\rho-1)} + \| v\|_{L^q(\rho-1)} )^{\rho-1} \| u - v\|_{\dot H_r^s}\\ &+ (\| u \|_{\dot H_r^s} + \| v\|_{\dot H_r^s} ) (\| u \|_{L^q(\rho-1)} + \| u \|_{L^q(\rho-1)})^{(\rho-2)_+} \| u - v\|_{L^q(\rho-1)}^{\min(\rho-1,1)} \end{align*} holds for any $u \in \dot H_r^s \cap L^{q(\rho-1)}$. \end{Corollary} Corollary \ref{Corollary:1.4} corresponds to the identity \[ \nabla (u^\rho - v^\rho) = \rho u^{\rho-1} \nabla(u-v) + \rho ( u^{\rho-1} - v^{\rho-1} ) \nabla v . \] Corollary \ref{Corollary:1.4} is as worthy as Proposition \ref{Proposition:1.1} to show the locally-in-time well-posedness of semilinear PDEs. We remark that Corollary \ref{Corollary:1.4} is not necessarily to construct solutions with a contraction argument because one may deploy a contraction argument in the ball of $\dot H_p^s$ with the distance of $L^p$, for example. However, the proof of the continuous dependence of solution maps for initial data may require Corollary \ref{Corollary:1.4}. Our second purpose is to extend Corollary \ref{Corollary:1.4} to the case where $s \in (1,2)$. Namely, the second main statement of this paper is \begin{Proposition} \label{Proposition:1.5} Let $\rho \in (1,2)$ and $s \in (1,\rho)$. Let $1 \leq p < \infty$, $1 < r < \infty$, $(\rho-1)^{-1} < q < \infty$ satisfy \eqref{eq:1.1}. Let $d_{\rho,s} \in (0, \rho - s)$. Then \begin{align*} &\| F_\rho (u) - F_\rho (v) \|_{\dot H^s_p}\\ &\lesssim (\| u \|_{L^{q(\rho-1)}} + \| v\|_{L^{q(\rho-1)}} )^{\rho-1} \| u - v\|_{\dot H_r^s}\\ &+ (\| u \|_{\dot H_r^s} + \| v\|_{\dot H_r^s} ) (\| u \|_{L^{q(\rho-1)}} + \| v\|_{L^{q(\rho-1)}} )^{\rho - 1 - d_{\rho,s}} \| u - v\|_{L^{q(\rho-1)}}^{d_{\rho,s}}. \end{align*} \end{Proposition} We note that an extension of Corollary \ref{Corollary:1.4} in the framework of Besov spaces was given in \cite[Proposition 2.1]{ON99} and \cite[Lemma 6.2]{FIW}. We remark that on the analogy of classical chain rules, it is expected that Proposition \ref{Proposition:1.5} may hold with $d_{\rho,s} = \rho -s$ but a similar technical difficulty arises in both Sobolev and Besov frameworks. We also remark that the case where $\rho = s$ is hopeless to extend Proposition \ref{Proposition:1.5} because of the failure of classical analogy, say $|x| \not \in C^1$, for example. In the next section, we collect some notation and basic estimates. In section \ref{section:3}, we revisit the proof of Proposition \ref{Proposition:1.1} for the completeness. In section \ref{section:4}, we give the proofs of Propositions \ref{Proposition:1.3} and \ref{Proposition:1.5} \section{Preliminary} \subsection{Notation} Here we collect some notation. Let $\mathcal S$ be the Schwartz class. Let $\psi \in \mathcal S$ be radial and satisfy \[ \mathrm{supp} \ \mathfrak F \psi \subset \{ \xi \ \mid \ 1/2 \leq |\xi| \leq 2 \} \] and \[ \sum_{j} \mathfrak F \psi(2^{-j} \xi) = 1 \] for $\xi \neq 0$. For $j \in \mathbb Z$, let $\psi_j = 2^{jn} \psi(2^{j} \cdot)$ so that $\| \psi_j \|_{L^1} = \| \psi \|_{L^1}$ and let $Q_j = \psi_j \ast$. We also put $\widetilde Q_j = Q_{j-1} + Q_j + Q_{j+1}$ and $\widetilde \psi_j = \psi_{j-1} + \psi_j + \psi_{j+1}$. We remark that $\widetilde Q_j Q_j = Q_j$ holds for any $j$. It is known that for $s \in \mathbb R$ and $1 < p,q < \infty$, the homogeneous Sobolev and Triebel-Lizorkin norms are equivalent, that is, \[ \| f \|_{\dot H_p^s} \sim \| 2^{js} Q_j f \|_{L^p(\ell^2)}. \] For the details of this equivalence, we refer the reader to \cite[Theorem 5.1.2]{G}, for example. For $f \in L_{\mathrm{loc}}^1$, we define the Hardy-Littlewood maximal operator by \[ Mf(x) = \sup_{r > 0} \frac{1}{|B(r)|} \int_{B(r)} |f(x+y)| dy \] and \[ M^{(\eta)} f(x) = M(|f|^\eta)^{1/\eta} \] for $\eta > 0$, where $B(r) \subset \mathbb R^d$ is the ball with radius $r$ centered at the origin. It is known that $M$ is bounded operator on $L^p$ and $L^p(\ell^q)$ for $1 < p,q < \infty$ (See \cite[Theorems 2.1.6 and 4.6.6]{G} and reference therein). Moreover, for $1 \leq p < \infty$, we define weighted $L^p$ norm with weight function $w$ by \[ \| f \|_{L_w^p} = \bigg( \int |f|^p w dx \bigg)^{1/p}. \] \subsection{Basic Estimates} Here we collect some estimates. \begin{Lemma}[{\cite[Theorem :2.1.10]{G}}] \label{Lemma:2.1} Let $w \in L^1$ be a positive radially decreasing function. Then the estimate \[ |w \ast g(x)| \leq \| w \|_{L^1} Mg(x) \] holds for any $x \in \mathbb R^n$. \end{Lemma} \begin{Lemma} \label{Lemma:2.2} \begin{align*} | \widetilde Q_j g(x+y) - \widetilde Q_j g(x) | &\lesssim M g(x) + M g(x+y),\\ | \widetilde Q_j g(x+y) - 2 \widetilde Q_j g(x) + \widetilde Q_j g(x-y) | &\lesssim M g(x+y) + Mg(x) + M g(x-y). \end{align*} Moreover, if $|y| < 2^{-j}$, then we have \begin{align*} | \widetilde Q_j g(x+y) - \widetilde Q_j g(x) | &\lesssim 2^j |y| Mg(x),\\ | \widetilde Q_j g(x+y) - 2 \widetilde Q_j g(x) + \widetilde Q_j g(x-y) | &\lesssim 2^{2j} |y|^2 Mg(x). \end{align*} \end{Lemma} \begin{proof} The estimates for $| \widetilde Q_j g(x+y) - \widetilde Q_j g(x) |$ is given in \cite{CW91} but for completeness, we give the proof. The first two estimate follows directly from Lemma \ref{Lemma:2.1}. By the fundamental theorem calculus implies that the following identities hold: \[ \psi_j(x+y) - \psi_j(x) = \int_0^1 (\nabla \psi)_j(x + \theta y) d \theta \cdot 2^j y \] and \begin{align*} &\psi_j(x+y) - 2 \psi_j(x) + \psi_j(x-y)\\ &= \int_0^1 ((\nabla \psi)_j(x+\theta y) - (\nabla \psi)_j(x-\theta y)) d \theta \cdot 2^j y\\ &= \sum_{|\alpha|=2} \int_0^1 \int_0^1 (\partial^\alpha \psi)_j(x+ (2\theta'-1) \theta y)) d \theta' d \theta \cdot 2^{2j} y^2. \end{align*} Since for $|\alpha| \leq 2$, if $|y| < 2^{-j}$, \[ (\partial^\alpha \psi)_j(x+y) \lesssim 2^j (2+2^j|x+y|)^{-n-1} \leq 2^j (1+2^j|x|)^{-n-1}, \] Then the third and fourth estimates are obtained by combining this and Lemma \ref{Lemma:2.1}. \end{proof} \begin{Lemma} \label{Lemma:2.3} For $0 < S < S'$, $P \geq 1$, and $a \in \ell^P$, the following estimate holds: \[ \| 2^{jS} \sum_{k} 2^{(k-j)_-S'} a_k \|_{\ell_j^P} \leq \bigg( \frac{1}{2^S-1} + \frac{1}{2^{S'-S}-1} \bigg) \| 2^{ks} a_k \|_{\ell_k^P}. \] \end{Lemma} \begin{proof} By the Minkowski inequality, we have \begin{align*} \| 2^{jS} \sum_{k < j} 2^{(k-j)_-S'} a_k \|_{\ell_j^P} &= \| 2^{jS} \sum_{\ell < 0} 2^{\ell s'} a_{\ell + j} \|_{\ell_j^P} + \| 2^{jS} \sum_{\ell \geq 0} a_{\ell + j} \|_{\ell_j^P}\\ &\leq \sum_{\ell < 0} 2^{\ell (S'-S)} \| 2^{s(\ell+j)} a_{\ell + j} \|_{\ell_j^P} + \sum_{\ell < 0} 2^{- \ell S} \| 2^{S(\ell+j)} a_{\ell + j} \|_{\ell_j^P}\\ &\leq \frac{2}{2^{S'-S}-1} \| 2^{Sk} a_{k} \|_{\ell_k^P} + \frac{1}{2^S-1} \| 2^{S k} a_{k} \|_{\ell_k^P}. \end{align*} \end{proof} \begin{Lemma} \label{Lemma:2.4} Let $0 < S < 1$ and $1 \leq P,Q < \infty$. The following estimate holds: \begin{align*} &\| 2^{jS} \| v(y+x) (u(y + x) - u(x)) \|_{L_{\psi_j,y}^P} \|_{\ell_j^Q}\\ & \lesssim M^{(P)} v(x) \| 2^{ks} M Q_k u \|_{\ell_k^Q} + \| 2^{ks} M^{(P)} (v M Q_k u)(x) \|_{\ell_k^Q} \end{align*} \end{Lemma} \begin{proof} Let $w \in L^1$ be smooth radially decreasing function satisfying \[ w(x) \geq (1+|x|)^P |\psi(x)| \] and put $w_j = 2^{jn} w(2^{j} \cdot)$. Lemmas \ref{Lemma:2.1} and \ref{Lemma:2.2} imply that we have \begin{align*} &\| v(y+x) (u(y + x) - u(x)) \|_{L_{\psi_j,y}^P}\\ &= \sum_k \| v(y+x)(\widetilde Q_k Q_k u(y + x) - \widetilde Q_k Q_k u(x)) \|_{L_{\psi_j,y}^P}\\ &\leq \sum_{k < j} 2^{k} \| y v(y+x) \|_{L_{\psi_j,y}^P} M Q_k u(x)\\ &+ \sum_{k \geq j} \bigg( \| v(y+x) MQ_ku(x+y) \|_{L_{\psi_j,y}^P} + M Q_k u(x) \| v(y+x) \|_{L_{\psi_j,y}^P}\bigg)\\ &\leq \| w \|_{L^1} \sum_{k < j} 2^{k-j} M^{(P)} v(x) M Q_k u(x)\\ &+ \| w \|_{L^1} \sum_{k \geq j} \bigg( M^{(P)}( v MQ_ku) (x) + M^{(p)} v(x) M Q_k u(x) \bigg). \end{align*} Therefore, Lemma \ref{Lemma:2.4} follows from these estimates above and Lemma \ref{Lemma:2.3}. \end{proof} \begin{Corollary} \label{Corollary:2.5} Let $0 < S < 1$, $1 \leq P_0 < 2$. Let $P_0 < P <\infty$ and $1 < Q,R < \infty$ satisfy \[ \frac 1 P = \frac 1 Q + \frac 1 R. \] The estimates \[ \| 2^{jS} \| (u(y + \cdot) - u(\cdot)) \|_{L_{\psi_j,y}^{P_0}} \|_{L^P(\ell_j^2)} \lesssim \| u \|_{\dot H_P^s} \] and \[ \| 2^{jS} \| v(y+\cdot) (u(y + \cdot) - u(\cdot)) \|_{L_{\psi_j,y}^{P_0}} \|_{L^P(\ell_j^2)} \lesssim \| v \|_{L^Q} \| u \|_{\dot H_R^S} \] hold. \end{Corollary} \begin{proof} For $P_1 > P_0$, the boundedness of $M$ implies that we have \[ \| M^{(P_0)} f \|_{L^{P_1}} = \| M|f|^{P_0} \|_{L^{P_1/P_0}}^{1/P_0} \lesssim \| |f|^{P_0} \|_{L^{P_1/P_0}}^{1/P_0} = \| f \|_{L^{P_1}}. \] Since $P_0 < P < Q$, Corollary \ref{Corollary:2.5} follows from Lemma \ref{Lemma:2.4} and the H\"older inequality. \end{proof} \section{Revisit of the Chain rules when $s \in (0,1)$} \label{section:3} In this section, we revisit the proof of Proposition \ref{Proposition:1.1} and Corollary \ref{Corollary:1.4}. The proofs are essentially given in \cite{CW91} but for completeness, we give the proofs. \begin{proof}[Proof of Proposition \ref{Proposition:1.1}] Since $\int \psi(y) dy = 0$, the identity \[ Q_j F(u)(x) = \int F(u)(x+y) - F(u)(y) \psi_j(y) dy \] holds for any $x \in \mathbb R^n$ and $j \in \mathbb Z$. Therefore, Proposition \ref{Proposition:1.1} follows from Corollary \ref{Corollary:2.5} with $(v,P_0,P,Q,R,S)$ replaced by $(G(u),1,p,q,r,s)$. \end{proof} \begin{proof}[Proof of Corollary \ref{Corollary:1.4}] The identity \begin{align*} &Q_j F_\rho(u)(x) - Q_j F_\rho(v)(x)\\ &= \int (F_\rho(u)(x+y) - F_\rho(u)(y) + F_\rho(v)(x+y) - F_\rho(v)(x)) \psi_j(y) dy \end{align*} holds. The fundamental theorem of calculus implies that the identity \begin{align*} &F_\rho(u)(x+y) - F_\rho(u)(y) + F_\rho(v)(x+y) - F_\rho(v)(x)\\ &= \{ (u - v)(x+y) - (u-v)(x) \} \int_0^1 F_\rho'(\theta u(x+y) + (1-\theta) u(x)) d \theta\\ &+ \{ v(x+y) - v(x) \} \\ &\cdot \int_0^1 \{ F_\rho'(\theta u(x+y) + (1-\theta) u(x)) - F_\rho'(\theta v(x+y) + (1-\theta) v(x)) \} d \theta. \end{align*} holds. The identity above and \eqref{eq:1.2} imply that the estimate \begin{align*} &| F_\rho(u)(x+y) - F_\rho(u)(y) + F_\rho(v)(x+y) - F_\rho(v)(x)|\\ &\lesssim ( |u(x+y)|^{\rho-1} + |u(x)|^{\rho-1}) | (u - v)(x+y) - (u-v)(x) |\\ &+ ( |u(x+y)| + |u(x)| + |v(x+y)| + |v(x)|)^{(\rho-2)_+}\\ &\cdot (|(u-v)(x+y)| + |(u-v)(x)|)^{\rho-1} | (u - v)(x+y) - (u-v)(x) | \end{align*} holds. Therefore, Corollary \ref{Corollary:1.4} follows from Corollary \ref{Corollary:2.5} and the estimates above. \end{proof} \section{Proofs of the Chain rules when $s \in (1,2)$} \label{section:4} \begin{proof}[Proof of Proposition \ref{Proposition:1.3}] We note that it is shown in \cite[(3.26)]{GOV94} that the estimate \begin{align} |(\tau_y - 2 + \tau_{-y}) F_\rho(u)(x)| &\lesssim |u(x)|^{\rho-1} |u(x+y) - 2 u(x) + u(x-y)| \nonumber\\ &+ |u(x+y) - u(x)|^\rho + |u(x-y) - u(x)|^\rho \label{eq:4.1} \end{align} holds. Moreover, the symmetry of $\psi$ and $\int \psi = 0$ imply that we have \begin{align} Q_j F_\rho(u)(x) &= \int F_\rho(u)(x+y) \psi_j(y) dy \nonumber\\ &= \int F_\rho(u)(x-y) \psi_j(y) dy \nonumber\\ &= \frac 1 2 \int (F_\rho(u)(x+y) + F_\rho(u)(x-y)) \psi_j(y) dy \nonumber\\ &= \frac 1 2 \int (\tau_y - 2 + \tau_{-y}) F_\rho(u)(x) \psi_j(y) dy \label{eq:4.2} \end{align} From \eqref{eq:4.1} and \eqref{eq:4.2}, the estimate \begin{align} |Q_j F(u)(x)| &\lesssim |u(x)|^{\rho - 1} \int |u(x+y) - 2 u(x) + u(x-y)| | \psi_j(y)| dy \nonumber\\ &+ \int |u(x+y) - u(x)|^p |\psi_j(y)| dy \label{eq:4.3} \end{align} follows. Put $w (x)$ be radially decreasing function with $w(x) \geq (1+|x|^2) |\psi(x)|$. By Lemma \ref{Lemma:2.2}, when $k<j$, the first term on the RHS of \eqref{eq:4.3} is estimated by \begin{align*} &\int_{|y| < C 2^{-k}} | \widetilde Q_k Q_k u(x+y) - 2 \widetilde Q_k Q_k u(x) + \widetilde Q_k Q_k u(x-y) | |\psi_j(y)| dy\\ &\leq M Q_k u(x) \int_{|y| < C 2^{-k}} 2^{2(k-j)} \omega_j(y) dy\\ &\lesssim 2^{2(k-j)} M Q_k u(x). \end{align*} Similarly, when $k \geq j$, it is estimated by \begin{align*} &\int_{|y| > 2^{-k}} | \widetilde Q_k Q_k u(x+y) - 2 \widetilde Q_k Q_k u(x) + \widetilde Q_k Q_k u(x-y) | |\psi_j(y) |dy\\ &\lesssim 2^{2(k-j)} \int_{|y| > 2^{-k}} (2 M Q_k u(x+y) + 2 M Q_k u(x) | \omega_j(y) dy\\ &\lesssim 2^{2(k-j)} (M Q_k u(x) + M^2 Q_k u(x)). \end{align*} Therefore, the estimates above and Lemma \ref{Lemma:2.3} imply that we have \begin{align} &\| 2^{js} |u|^{\rho-1} \int |u(\cdot+y) - 2 u(\cdot) + u(\cdot-y)| | \psi_j(y)| dy \|_{L^p(\ell^2)} \nonumber\\ &\lesssim \| 2^{ks} |u|^{\rho-1} (M Q_k u + M^2 Q_k u) \|_{L^p(\ell^2)} \nonumber\\ &\lesssim \| u \|_{L^{q(\rho-1)}}^{\rho-1} \| u \|_{\dot H_r^s}. \label{eq:4.4} \end{align} Moreover, by the Lemma \ref{Lemma:2.4}, the second term of the RHS of \eqref{eq:4.3} satisfy the estimate \begin{align} \| 2^{js} \|u(\cdot + y) - u(\cdot) \|_{L_{\psi_j}^\rho}^\rho \|_{L^p(\ell^2)} &= \| 2^{js/\rho} \|u(\cdot + y) - u(\cdot) \|_{L_{\psi_j}^\rho}\|_{L^{\rho p}(\ell^2)}^{\rho} \nonumber\\ &\lesssim \| 2^{ks/\rho} M^{(\rho)} Q_k u \|_{L^{\rho p}(\ell^2)}^{\rho} \nonumber\\ &\lesssim \| u \|_{\dot H_{\rho p}^{s/\rho}}^{\rho} \nonumber\\ &\lesssim \| u \|_{L^{q(\rho-1)}}^{\rho-1} \| u \|_{\dot H_r^s}, \label{eq:4.5} \end{align} where we have used the Gagliardo-Nirenberg inequality to obtain the last estimate. For the details of the Gagliardo-Nirenberg inequality, we refer the reader to \cite[Corollary 2.4]{HMOW11} and reference therein. Combining \eqref{eq:4.4} and \eqref{eq:4.5}, we obtain Proposition \ref{Proposition:1.3}. \end{proof} \begin{proof}[Proof of Proposition \ref{Proposition:1.5}] For simplicity, we denote $d_{\rho,s}$ by $d$. We note that the estimate \begin{align} &| (\tau_y -2 + \tau_{-y})(F_\rho (u) - F_\rho (v))(x)| \nonumber\\ &\lesssim \max_{z \in \{x, x+y, x-y\} } |u(z)|^{\rho-1} | (\tau_y -2 + \tau_{-y}) (z-w)(x)| \nonumber\\ & + \max_{z \in \{x, x+y, x-y\}} |u(z)|^{\rho-1} | (\tau_y -2 + \tau_{-y}) w(x)| \nonumber\\ & + (|(\tau_y-1)u(x)| + |(\tau_{-y}-1)u(x)|)^{\rho-1} |(\tau_y-1)(u-v)(x)| \nonumber\\ & + |(\tau_{-y}-1)u(x)| \min( \sigma, \max_{z \in \{x, x+y, x-y\}} |(u-v)(z)| )^{\rho-1} \label{eq:4.6} \end{align} holds, where \[ \sigma = |(\tau_y-1)u(x)| + |(\tau_{-y}-1)u(x)| +|(\tau_y-1)v(x)| + |(\tau_{-y}-1)v(x)|. \] For the details of the estimate above, see \cite[Lemma:A.1]{FIW}. Therefore, the proof of proposition \ref{Proposition:1.5}, it is sufficient to show \begin{align} &\| 2^{js} \| B(u,v)(\cdot,y) \|_{L_{\psi_j,y}^1} \|_{L^p(\ell^2)} \nonumber\\ &\lesssim (\| u \|_{\dot H_r^s} + \| v\|_{\dot H_r^s} ) (\| u \|_{L^{q(\rho-1)}} + \| v\|_{L^{q(\rho-1)}} )^{\rho - 1 - d_{\rho,s}} \| u - v\|_{L^{q(\rho-1)}}^{d_{\rho,s}}. \label{eq:4.7} \end{align} where \[ B(u,v)(x,y) = | u(x+y) - u(x)| |v(x+y)-v(x)|^{\rho- 1 - d} |u(x+y)-v(x+y)|^d. \] The other terms on the RHS of \eqref{eq:4.6} may be treated like the argument above so we omit the detail. Put \[ q_0 = q \frac{\rho-1}{d} \qquad \mathrm{and} \qquad r_0 = \frac{p q_0}{q_0-p}. \] Then the H\"older inequality and Lemma \ref{Lemma:2.1} imply that we have \begin{align*} &\| B(u,v)(x,y) \|_{L^1_{\psi_j,y}}\\ &\leq \| |(u-v)(x+y)|^d \|_{L^{q_0/p}_{\psi_j,y}} \| |u(x+y)-u(x)| | v(x+y) - v(x)|^{\rho-1 - d} \|_{L^{r_0/p}_{\psi_j,y}}\\ &\lesssim (M^{(q (\rho-1)/p)}|u-v|(x))^d\\ &\cdot \| u(x+y)-u(x) \|_{L_{\psi_j,y}^{(\rho-d)r_0/p}} \| v(x+y)-v(x) \|_{L_{\psi_j,y}^{(\rho-d)r_0/p}}^{\rho- d-1}. \end{align*} Then Lemma \ref{Lemma:2.1} and Corollary \ref{Corollary:2.5} imply that the estimates \begin{align*} &\| 2^{js} \| B(u,v)(\cdot,y) \|_{L_{\psi_j,y}^1} \|_{L^p(\ell^2)}\\ &\lesssim \| M^{(q (\rho-1)/p)}|u-v| \|_{L^{dq_0}}^d\\ &\cdot \| 2^{j s/(\rho-d)} \| u(\cdot+y)-u(\cdot) \|_{L_{\psi_j,y}^{(\rho-d)r_0/p}} \|_{L^{r_0(\rho-d)}(\ell^2)}\\ &\cdot \| 2^{j s/(\rho-d)} \| v(\cdot+y)-v(\cdot) \|_{L_{\psi_j,y}^{(\rho-d)r_0/p}} \|_{L^{r_0(\rho-d)}(\ell^2)}^{\rho - d - 1}\\ &\lesssim \| u-v \|_{L^{q(\rho-1)}}^d \| u \|_{\dot H_{r_0(\rho-d)}^{s/(\rho-d)}} \| v \|_{\dot H_{r_0(\rho-d)}^{s/(\rho-d)}}^{\rho-d-1} \end{align*} hold. Then \eqref{eq:4.7} follows from the estimates above and the Gagliardo-Nirenberg inequality. \end{proof} \begin{Remark} In the proof above, one cannot take $d = \rho - s$ because \[ \| 2^{j} \| u(\cdot+y)-u(\cdot) \|_{L_{\psi_j,y}^{(\rho-d)r_0/p}} \|_{L^{r_0(\rho-d)}(\ell^2)} \] is not controlled by Lemma \ref{Lemma:2.4}. \end{Remark} \section*{Acknowledgment} The author is supported by JSPS Early-Career Scientists 20K14337.
{ "timestamp": "2021-04-13T02:21:37", "yymm": "2104", "arxiv_id": "2104.05092", "language": "en", "url": "https://arxiv.org/abs/2104.05092", "abstract": "A chain rule for power product is studied with fractional differential operators in the framework of Sobolev spaces. The fractional differential operators are defined by the Fourier multipliers. The chain rule is considered newly in the case where the order of differential operators is between one and two. The study is based on the analogy of the classical chain rule or Leibniz rule.", "subjects": "Functional Analysis (math.FA)", "title": "Remark on the Chain rule of fractional derivative in the Sobolev framework", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9871787876194204, "lm_q2_score": 0.810478913248044, "lm_q1q2_score": 0.8000875909713094 }
https://arxiv.org/abs/2211.06484
A regular $n$-gon spiral
We construct a polygonal spiral by arranging a sequence of regular $n$-gons such that each $n$-gon shares a specified side and vertex with the $(n+1)$-gon in the construction. By offering flexibility for determining the size of each $n$-gon in the spiral, we show that a number of different analytical and asymptotic behaviors can be achieved.
\section{Introduction} Spirals are pervasive in fluid motion, biological structures, and engineering, affording numerous applications of mathematical spirals in modeling real-world phenomena \cite[ch.~11-14]{Thompson} \cite{Davis, Hammer}. Polygonal spirals are loosely defined as spirals that can be constructed using a sequence of polygons obeying a geometric relation. For example, in Fig. \ref{theo_spiral} we show how to join together right triangles in an infinite sequence to obtain the spiral of Theodorus---one of the oldest and most well-studied polygonal spirals \cite{Brink, Davis}. \begin{figure}[!htb] \begin{centering} \includegraphics[width=0.45\textwidth]{Theodorus_Spiral2.png} \caption{Plot of the first 7 triangles in the spiral of Theodorus. The sequence of shared vertex points $\vn{}{n}$ and the interpolation curve proposed by Davis \cite[p.~38]{Davis} are also shown.} \label{theo_spiral} \end{centering} \end{figure} In Fig. \ref{theo_spiral}, we let $\vn{}{n}$ denote the unique non-origin shared vertex of the triangle with hypotenuse $\sqrt{n}$ and the triangle with hypotenuse $\sqrt{n+1}$. Considerable attention has been given to studying the sequence $\vn{}{n}$ \cite{Brink, Davis, Gautschi, Hlawka}. The \textit{key idea} in the spiral of Theodorus construction is to define a sequence of polygons (i.e., right triangles) such that consecutive polygons share a unique side and at least one unique vertex point. By changing right triangles to similar triangles, or other types of similar polygons, we find that this idea is ubiquitous in the literature on polygonal spirals \cite{Anatriello, Strizhevsky, Yap} \cite[ch.~34]{Hammer}. In this paper, we make use of this key idea to construct a spiral made out of a sequence of regular $n$-gons. In particular, we start by considering the case where all the $n$-gons are normalized to have perimeter $1$ (see Fig. \ref{lam1}). We begin the construction with an equilateral triangle with side length $\frac{1}{3}$ in the first quadrant of the complex plane. Along the upper right edge of the triangle, we construct a square with side length $\frac{1}{4}$. We continue in this way, following Definition \ref{dcon}, to construct the pentagon, hexagon, etc., leading to the construction shown below. \begin{figure}[!htb] \begin{centering} \includegraphics[width= 0.6\textwidth]{spiral_s=1.png} \caption{Plot of the first few polygons of the Definition \ref{dcon} construction (choosing $l(n) = \frac{1}{n}$). The shared vertex points $\vn{}{n}$ (see Lemma \ref{lamv}) and a smooth interpolation curve defined in equation \ref{interp} are also shown.} \label{lam1} \end{centering} \end{figure} \noindent We desire to find the $n \to \infty$ limit behavior of the sequence $V(n)$ shown in the Fig. \ref{lam1} construction. In particular, we pose a slightly more general question: if the side length of each $n$-gon in the Fig. \ref{lam1} construction is instead given by $\ell_s(n) = n^{-s}$, for which values of $s \in \mathbb{R}$ does the sequence $\vn{\ell_s}{n}$ converge as $n \to \infty$? The answer to this question is one of the main results of this paper: \begin{theorem}[Convergence of $\vn{\ell_s}{n}$] \label{thm1} \begin{flalign*} \text{ As } n \to \infty, \hspace{1mm} \vn{\ell_s}{n} & \begin{cases} \text{converges to a point when } s > 0.\\ \text{converges to a circular orbit when } s=0.\\ \text{diverges when } s < 0. \end{cases} \end{flalign*} \end{theorem} We conclude the introduction by outlining the rest of the paper. In Section \ref{construct}, we formally define the geometric construction of the $n$-gon spiral (Definition \ref{dcon}) with an arbitrary length function, leading to a formula for the shared polygon vertex points (Lemma \ref{lamv}). In Section \ref{converge}, we use Lemma \ref{lamv} to prove Theorem \ref{thm1}, and then give a corollary to the theorem. In Section \ref{telescoping_spiral}, we introduce a special case of the $n$-gon spiral which admits particularly nice algebraic expressions, including a closed form analytic continuation of the discrete spiral to real values of $n$ (Theorem \ref{thm_analytic}) and an appearance of the golden ratio (Remark \ref{rem1}). \section{The formal construction} \label{construct} \begin{definition}[Construction of the $n$-gon spiral $\pp{l}$]\ \label{dcon} \begin{enumerate} \item Starting with $n=2$, the $n$th polygon in the spiral construction is a regular $n$-gon with side length $l(n)$, where $l: \mathbb{N} \to \mathbb{R}$. \item Consecutive polygons share a side and at least one vertex. Only consecutive polygons may share a side. \item Shared vertices of each polygon are consecutive vertices, and the side connecting two shared vertices is not a shared side. \item Consecutive polygons do not overlap (i.e. the shared area between consecutive polygons is 0). \item \textit{Notation:} let $\pp{l}$ represent the infinite sequence of polygons defined via $1$--$4$ above. \begin{itemize} \item In this sequence, $\pn{l}{n}$ represents the $n$-gon in $\pp{l}$. \item Let $\vv{l}$ denote the infinite sequence of shared vertex points of $\pp{l}$. In this sequence, $\vn{l}{n}$ represents the vertex of $\pn{l}{n}$ and $\pn{l}{n+1}$ that is shared for all choices of $l(n)$. \item Let $\cc{l}$ denote the infinite sequence of polygon centers of $\pp{l}$. In this sequence, $\cn{l}{n}$ represents the center of $\pn{l}{n}$. \end{itemize} \end{enumerate} \end{definition} Once the \textit{length function} $l(n)$ is chosen and the coordinates of the $2$-gon and $3$-gon are given, Definition \ref{dcon} fully determines $\pp{l}$. We will see in the next section that there is an algebraically simplest orientation of the construction in the complex plane, which we will adopt. With the rules for construction laid out, we desire to obtain a formula for the shared vertex point $\vn{l}{n}$. We begin by following Definition \ref{dcon} to draw an arbitrary $n$-gon and $(n+1)$-gon in the $\pp{l}$ construction. Using that the interior angles of a regular $n$-gon are $\alpha_n := \frac{\pi(n-2)}{n}$ radians, we want to find the angle with respect to the horizontal ($\theta_n$) that must be followed to get from the shared vertex point $\vn{l}{n-1}$ to $\vn{l}{n}$, pictured in Fig. \ref{angles}. In this diagram, we assume $\theta_n, \theta_{n+1} < \pi$. Equivalent results (mod $2\pi$) are obtained by considering the other relevant cases. \begin{figure}[!htb] \begin{centering} \includegraphics[width=0.48\textwidth]{geom_deriv.png} \caption{$\pp{l}$ angular relationships.} \label{angles} \end{centering} \end{figure} In the Fig. \ref{angles} diagram, we observe that the horizontal (dotted) lines for defining $\theta_n$ and $\theta_{n+1}$ are cut by a transversal (a side of the $n$-gon). Therefore, the angles $\theta_n$ and $\beta = \alpha_n + \alpha_{n+1} - \theta_{n+1}$ in Fig. \ref{angles} are supplementary, which affords an angular recurrence: \begin{equation} \theta_n + \beta = \pi \implies \theta_{n+1} = \theta_{n} + \frac{(n-2)\pi}{n} + \frac{(n-1)\pi}{n+1} - \pi, \hspace{2mm} n > 1. \label{t1_recurrence} \end{equation} To solve the recurrence for $\theta_n$, we need a starting value for $\theta_2$---a choice which determines the orientation of $\pp{l}$ in the complex plane. Appealing to simplicity, we find that choosing $\theta_2 = -3 \pi$ ensures that there is no nontrivial constant rotation applied to all elements of $\pp{l}$ (i.e. $\theta_n$ mod $2\pi$ has no constant additive term). This choice affords the orientation shown in Fig. \ref{lam1}. With $\theta_2$ in hand, we note that (\ref{t1_recurrence}) is a first order linear recurrence, and thus $\theta_{n+1}$ can be rewritten as a sum. Simplification of the resulting sum yields the angular relation for $\pp{l}$: \begin{equation} \label{t1} \theta_n = 2\pi\Big(\frac{n}{2}+\frac{1}{n}-2 H_n\Big), \hspace{2mm} n > 1, \end{equation} \noindent where $H_n$ is the $n$th harmonic number. Using (\ref{t1}), we are now ready to write down analytical expressions for the shared vertex points $\vn{l}{n}$. Examining Fig. \ref{angles}, we find that each shared vertex point $\vn{l}{n}$ can be written as a the sum over the integers $2 \leq k \leq n$ of the $k$-gon side length $l(k)$ multiplied by the corresponding rotation, $e^{i \theta_k}$. We define $\vn{l}{2} := 0$ to start the spiral from the origin. \begin{lem}[Expression for shared vertex points of the $n$-gon spiral]\ \label{lamv} \begin{equation*} \vn{l}{n} = \sum_{k=3}^n l(k) \hspace{1mm} e^{i \theta_k} = \sum_{k=3}^n (-1)^k l(k) \hspace{1mm} e^{2 \pi i (\frac{1}{k} - 2H_k)}, \hspace{2mm} n \in \mathbb{N}_{> 1}. \end{equation*} \end{lem} Since $\pp{l}$ consists of regular polygons, there is a natural relationship between the sequences $\pp{l},\vv{l},$ and $\cc{l}$. We can explicitly solve for $\cn{l}{n}$ by constructing the isosceles triangle with vertices $\vn{l}{n-1}$ and $\vn{l}{n}$ and an apex angle of $\frac{2 \pi}{n}$ radians. This recovers the expression for $\vn{l}{n}$ plus a much simpler function, $Q_l(n)$: \begin{lem}[Expression for polygon centers of the $n$-gon spiral]\ \label{lamc} \label{centers} \begin{equation*} \cn{l}{n} = \vn{l}{n} + Q_l(n), \hspace{ 2mm} \text{where } Q_l(n) = \frac{(-1)^n l(n) e^{2 \pi i (\frac{1}{n} -2H_n)}}{e^{\frac{2 \pi i}{n}} - 1}. \end{equation*} \end{lem} Lemma \ref{lamc} makes it clear that understanding of the behavior of $\vv{l}$ and $Q_l$ is sufficient for understanding the behavior of $\cc{l}$, and hence $\pp{l}$. \textit{As such, we will frequently refer to the sequence of shared vertex points $\vv{l}$ as ``the $n$-gon spiral."} \noindent We use the standard technique to obtain a smooth continuation of $\vn{l}{n}$ to real $n$: \begin{equation} \label{interp} \tilde{V}_l(n) := \sum_{k=3}^{\infty} l(k) \hspace{1mm} e^{i \theta_k} - l(k-2+n) \hspace{1mm} e^{i \theta_{k-2+n}}, \hspace{2mm} n \in \mathbb{R}_{> 1}. \end{equation} \section{Spiral convergence} \label{converge} As a step toward understanding the convergence properties of the $n$-gon spiral, we will analyze convergence in the case where the length function is $\ell_s(n) = n^{-s}$, $s \in \mathbb{R}$. For conciseness, we assign $W(s) := \lim_{n \to \infty} \vn{\ell_s}{n}$ to denote the desired limit. \begin{figure}[!htb] \begin{subfigure}{0.59\textwidth} \centering\captionsetup{width=.9\linewidth} \includegraphics[width=\textwidth]{spiral_s=0.png} \caption{The first few polygons of the spiral $\pp{\ell_0}$ $(s=0)$ plotted with the sequence of shared vertices $\vv{\ell_0}$ and the interpolation curve given by (\ref{interp}). We observe the convergence of $\vv{\ell_0}$ to a circular orbit (Theorem \ref{thm1}). Using Euler's transform for alternating series, we find numerically that the center of this circular orbit is at $\lim_{s \to 0^+} W(s) = 1.21711960256553... + i \hspace{1mm} 2.68541404871695...$} \label{n=0_spiral} \end{subfigure} \begin{subfigure}{0.37\textwidth} \centering\captionsetup{width=.94\linewidth} \includegraphics[width=\textwidth]{parametric_spiral_med2.png} \caption{The convergence points $W(s)$ for $s \in \mathbb{R}^+$ trace out a curve in the complex plane (black). We also plot ten spirals (from $s=0.0000726$ to $s=1.77$, interpolated by (\ref{interp})) with convergence points that are equally spaced along the curve.} \label{conv_curve} \end{subfigure} \caption{Illustration of Theorem \ref{thm1}.} \label{spiral_comparison} \end{figure} \begin{proof}[Proof of Theorem \ref{thm1}] We aim to characterize the limit behavior of the sequence of shared vertex points, \begin{equation} W(s) = \sum_{k=3}^{\infty} \frac{(-1)^k e^{2 \pi i (\frac{1}{k} - 2H_k)}}{k^s}. \end{equation} \noindent \underline{Case 1: $s < 0$.} The terms of the series do not approach 0, hence $W(s)$ diverges. \bigskip \noindent \underline{Case 2: $s > 1$.} $W(s)$ is an absolutely convergent series. \bigskip \noindent \underline{Case 3: $0 < s \leq 1$.} Letting $f(k):=\exp\big(2\pi i\big(\frac{1}{k}-2H_k\big)\big)$, we first accelerate convergence of $\vn{\ell_s}{n}$ by adding consecutive pairs of terms: \begin{equation} \label{Saccel} \vn{\ell_s}{2n} = \sum_{k=3}^{2n} \frac{(-1)^k f(k)}{k^s} = \sum_{j=2}^{n} F(j), \text{ where } F(j) := \frac{(2j-1)^sf(2j)-(2j)^sf(2j-1)}{((2j-1)(2j))^s} \end{equation} \noindent We now manipulate $F(j)$ in order to show $\vn{\ell_s}{2n}$ is absolutely convergent. To this end, we invoke the identity $H_{2j-1} = H_{2j} - \frac{1}{2j}$ to write $f(2j-1)$ in terms of $f(2j)$: \begin{equation} f(2j-1) = \exp\Big(2\pi i\Big(\frac{1}{2j-1}+\frac{1}{2j}\Big)\Big) f(2j). \label{f2j-1} \end{equation} \noindent Plugging (\ref{f2j-1}) and $(2j-1)^s = (2j)^s(1-\frac{1}{2j})^s$ into $F(j)$, we multiply the numerator and denominator of $F(j)$ by $2j-1$, which affords \begin{equation} \label{Fj} F(j) = \frac{f(2j)\Big((2j-1)\big(1-\frac{1}{2j}\big)^s - (2j-1) \exp\big(2\pi i\big(\frac{1}{2j-1}+\frac{1}{2j}\big)\big)\Big)}{(2j-1)^{1+s}} =: \frac{T(j)}{(2j-1)^{1+s}}. \end{equation} \noindent Since $\sum_{j=2}^{\infty} \frac{1}{(2j-1)^{1+s}}$ is convergent for $s \in (0,1)$, showing $\abs*{T(j)}$ is bounded gives the desired convergence result. By the triangle inequality, \begin{flalign} \abs*{T(j)} & < \max \hspace{1mm}\abs*{ (2j-1)\Big(1-\Big(1-\frac{1}{2j}\Big)^s\Big)} + \max \hspace{1mm}\abs*{-(2j-1)\Big(1-\exp\Big(2\pi i \Big(\frac{1}{2j-1} + \frac{1}{2j}\Big)\Big)\Big)} \nonumber\\ & =: \max(A(j,s)) + \max(B(j)) \label{Tj} \end{flalign} \begin{lem} \label{maxA} \begin{equation*} \max(A(j,s)) = \lim_{j\to\infty} A(j,s) = s. \end{equation*} \end{lem} \begin{proof} We obtain the $j \to \infty$ limit by direct calculation. To show this is a maximum, we make use of the following Bernoulli-type inequality (which can be readily proved using the binomial theorem): \begin{equation} \label{bern} 1-sy < (1+y)^{-s} < 1 \text{ for } y \in (0,1), \hspace{1mm} s \in (0,1]. \end{equation} Plugging $y := \frac{1}{2j-1}$ into (\ref{bern}), we simplify to obtain: \begin{equation} \label{bernsimp} \frac{s}{2j-1} > 1-\Big(1+\frac{1}{2j-1}\Big)^{-s} > 0. \end{equation} \noindent Observing that $(1+\frac{1}{2j-1})^{-s} = (1-\frac{1}{2j})^s$, we multiply (\ref{bernsimp}) by $2j-1$ to afford the desired bound: \begin{equation} s > A(j,s) = (2j-1)\Big(1-\Big(1-\frac{1}{2j}\Big)^s\Big) > 0. \end{equation} \end{proof} \begin{lem} \label{lemBj} \begin{equation*} \max(B(j)) = \lim_{j \to \infty} B(j) = 4\pi. \end{equation*} \end{lem} \begin{proof} The $j \to \infty$ limit is obtained by direct calculation. Defining $x(j):=\pi \big(\frac{1}{2j-1}+\frac{1}{2j}\big)$, we have that $x(j) \in (0, \pi)$ for all $j \geq 2$ since $x(2) < \pi$ and $x(j)$ is a monotonically decreasing function. From this, we can simplify $B(j)$ to obtain \begin{equation} B(j):= \abs*{(2j-1)(1-\exp(2 i x))} = 2(2j-1) \sin(x) \text{ for } j \in [2,\infty). \end{equation} Calculation of $B'(j)$ reveals \begin{equation} B'(j) = \frac{\pi}{j^2} + 4(\sin x - x \cos x). \end{equation} Since $\sin x - x \cos x$ is positive on $(0, \pi)$, $B(j)$ is increasing for all $j \geq 2$. \end{proof} \noindent Plugging Lemmas \ref{maxA} and \ref{lemBj} into (\ref{Tj}) and (\ref{Fj}) implies \begin{equation} \sum_{j=2}^{\infty} \abs*{F(j)} < (4 \pi + s)\hspace{1mm} 2^{-1-s}\zeta\Big(1+s,\frac{3}{2}\Big), \end{equation} hence $\sum_{j=2}^{\infty} F(j)$ is absolutely convergent for $s \in (0,1]$. Since the summand of $\vn{\ell_s}{n}$ vanishes as $n \to \infty$, $\lim_{n \to \infty} \vn{\ell_s}{2n+1} = \lim_{n \to \infty} \vn{\ell_s}{2n} = \sum_{j=2}^{\infty} F(j)$, so $W(s)$ is convergent. \bigskip \noindent \underline{Case 4: $s = 0$}. As in Case 3, we accelerate convergence of $\vn{\ell_0}{n}$ by pairing terms: \begin{equation} \vn{\ell_0}{2n} = \sum_{j=2}^{n} \big( f(2j) - f(2j-1) \big) =: U(n), \hspace{1mm}\text{where } f(k) := e^{2\pi i (\frac{1}{k}-2H_k)}. \label{Saccel2} \end{equation} Let $r \geq 1$ be a rational constant such that $nr \in \mathbb{N}$. We desire to show that the distance between two arbitrary, evenly indexed spiral points $U(nr)$ and $U(n)$ is a sinusoidal function of $r$ in the $n \to \infty$ limit. Hence, we desire to calculate $B(r) := \lim_{n \to \infty} \abs*{U(nr)-U(n)}$. Plugging the asymptotic harmonic series expansion $H_n = \gamma + \log(n) + \frac{1}{2n} + \mathcal{O}(n^{-2})$ into $B(r)$, we simplify and apply the Euler-Maclaurin formula to obtain \begin{equation} B(r) = \lim_{n \to \infty} \Bigg\lvert \int_{n}^{nr} \Big(e^{-4\pi i \log(2x)} - e^{-4\pi i \log(2x-1)}\Big)dx + \text{error terms} \hspace{1mm}\Bigg\rvert. \label{EMbound} \end{equation} Using standard bounds \cite{Lehmer}, we show that the error terms vanish as $n \to \infty$. After evaluating the integral in (\ref{EMbound}), we make use of Euler's formula to separate real and imaginary parts. Evaluating the desired limit in Mathematica affords \begin{equation} B(r) = \abs*{\sin(2\pi \log r)}. \label{Blim} \end{equation} As claimed, we find that the distance between arbitrary points $U(nr)$ and $U(n)$ in the $n \to \infty$ limit is sinusoidally dependent on the distance parameter $r$ (\ref{Blim}). Since $U(n) = \vn{\ell_0}{2n}$, equation \ref{Blim} shows that all evenly indexed points on the spiral converge to a circle with diameter 1 (Fig. \ref{n=0_spiral}). By explicitly constructing the circle through $\vn{\ell_0}{2n-2}, \vn{\ell_0}{2n},$ and $\vn{\ell_0}{2n+2},$ we can readily show that $\vn{\ell_0}{2n+1}$ is on this circle in the $n \to \infty$ limit, which confirms that the odd spiral points converge to the same circle. \bigskip \noindent This concludes the proof of Theorem \ref{thm1}. \end{proof} Theorem \ref{thm1} provides the limiting behavior of $\vn{\ell_s}{n}$ for $\ell_s(n) = n^{-s}$, which may be used to find the limiting behavior of $V(n)$ for geometrically significant length functions which are asymptotic to $\ell_s$. For example, we can consider a Definition \ref{dcon} spiral construction where the length function is determined by each $n$-gon being inscribed inside a circle of radius $n^{-s}$, $s \in \mathbb{R}$. \begin{corollary}[Corollary of Theorem \ref{thm1}] Let $\pp{insc}$ and $\pp{circ}$ denote the $n$-gon spiral constructions where each $n$-gon is inscribed and circumscribed with respect to a circle of radius $n^{-s}$, $s \in \mathbb{R}$, respectively. Then $\lim_{n \to \infty} \vn{insc}{n}$ and $\lim_{n \to \infty} \vn{circ}{n}$ converge for $s > -1$. \end{corollary} \begin{proof} The side lengths of the inscribed and circumscribed $n$-gons are given by $\ell_{insc}(n) = 2n^{-s} \sin(\frac{\pi}{n}) = \frac{2\pi}{n^{1+s}} + \mathcal{O}(\frac{1}{n^{3+s}})$ and $\ell_{circ}(n) = 2n^{-s} \tan(\frac{\pi}{n}) = \frac{2\pi}{n^{1+s}} + \mathcal{O}(\frac{1}{n^{3+s}})$, respectively. The result follows by Theorem \ref{thm1}. \end{proof} \noindent A similar result can be shown for the $n$-gon spiral where each $n$-gon has area $n^{-s}$. \section{The telescoping spiral} \label{telescoping_spiral} Here, we present a special choice of the length function that leads to closed form formulae for the discrete spiral points as well as their analytic continuation. \begin{theorem}[The telescoping spiral] \label{thm_analytic}\ \noindent The $n$-gon spiral with length function $L(k) = 2 \cos (\frac{2\pi}{k})$ admits the following closed form analytic continuation for $n \in \mathbb{R}_{>1}$: \begin{flalign} \vn{L}{n} & = -1 + (-1)^n e^{-4 \pi i(\gamma + \psi(n+1))}, \nonumber\\ Q_L(n) & = (-1)^n e^{-4 \pi i(\gamma + \psi(n+1))} \bigg(e^{\frac{2 \pi i}{n}} + \frac{e^{\frac{2 \pi i}{n}}+1}{e^{\frac{2 \pi i}{n}}-1}\bigg). \nonumber \end{flalign} \end{theorem} \begin{figure}[!h] \begin{subfigure}{.585\textwidth} \centering\captionsetup{width=\linewidth} \includegraphics[width=\textwidth]{degenerate_spiral_interpolated2.png} \caption{Plot of the polygon spiral $\pp{L}$ up to the $12$-gon. $\vn{L}{n}$ (see Theorem \ref{thm_analytic}) traces out the unit circle centered at $-1$. The polygon centers are also marked, with their analytic continuation $\cn{L}{n} = \vn{L}{n} + Q_L(n)$ plotted from $n = 1.05$.} \label{degen} \end{subfigure} \begin{subfigure}{.40\textwidth} \centering\captionsetup{width=0.8\linewidth} \includegraphics[width=0.9\textwidth]{Qspiral_1-35.png} \caption{$Q_L(n)$ (see Theorem \ref{thm_analytic}) plotted from $n=1.02$ to $n=35$. In the limit as $n \to 1$, Re$(Q_L(n)) \to 4(1-\frac{\pi^2}{6})$.} \label{Qspiral} \end{subfigure} \caption{The telescoping spiral construction and analytic continuation.} \label{tele_spiral} \end{figure} \begin{proof} Writing $L(k)$ in exponential form, we plug it into Lemma \ref{lamv} and simplify to obtain a telescoping series! \begin{equation} \label{tele1} \vn{L}{n} = \sum_{k=3}^n (-1)^k \big(e^{-4 \pi i H_{k-1}} + e^{-4 \pi i H_{k}} \big) = -1 + (-1)^n e^{-4 \pi i H_n} \end{equation} The harmonic numbers can be analytically continued to complex values of $n$ via $ H_n = \gamma + \psi(n+1)$, where $\psi(x) = \frac{\Gamma'(x)}{\Gamma(x)}$ is the digamma function, and $\gamma = 0.5772...$ is Euler's constant. This provides a direct analytic continuation of $\vn{L}{n}$ to $n \in \mathbb{R}_{>1}$, where $n \leq 1$ is excluded because we cannot construct a $1$-gon to obey Definition \ref{dcon} (and $\cn{L}{1}$ is not defined). By plugging $L(n)$ into Lemma \ref{centers}, we also obtain a closed-form analytic continuation of $Q_L(n)$, and hence $\cn{L}{n}$. \end{proof} From Theorem \ref{thm_analytic}, we see that all the values of $\vn{L}{n}$ lie on the unit circle centered at $-1$ (Fig. \ref{degen}), hence $\vn{L}{n}$ forms a degenerate spiral for $n \in (1, \infty)$. On the interval $n \in (1,\infty)$, $L(n)$ is negative for $n \in (\frac{4}{3}, 4)$, zero at $n= \frac{4}{3}, 4$, and positive everywhere else, resulting in 3 regions where the spirals $Q_L(n)$ and $\cn{L}{n}$ exhibit different winding behavior. The $n$-gon centers and vertices coincide at the zeros of $L(n)$ (Fig. \ref{degen}). \begin{rem}[Golden ratio intersection] \label{rem1} In Fig. \ref{degen}, we observe that $\cn{L}{n}$ intersects itself inside the unit disk centered at $-1$. This intersection occurs at $n = \varphi := \frac{\sqrt{5} + 1}{2}$ and $n = \varphi + 1$. At this point, $\cn{L}{\varphi} = -i e^{-\pi i(4(\gamma + \psi(\varphi))+\varphi)} \cot(\pi \varphi) - 1$. \end{rem} \section{Further Directions} There is much additional work to be done toward understanding the $n$-gon spiral introduced here. We showed that the sequence of shared polygon vertex points in Fig. \ref{lam1} converges, but is it possible to find a closed-form expression for this convergence value? Additionally, Theorem \ref{thm1} only describes the $n \to \infty$ behavior of the $n$-gon spiral for a particular family of length functions. Can one develop a convergence condition that applies to arbitrary length functions? The closed-form analytic continuation of the telescoping spiral (Theorem \ref{thm_analytic}) makes this spiral a particularly attractive choice for further study. How does the asymptotic behavior of the spirals $\cc{L}$ and $Q_L$ compare with classical spirals in the literature? We identified self-intersections of $Q_L(n)$ and $\cn{L}{n}$ at $n = \frac{4}{3}, 4$ and $n = \varphi, \varphi + 1$, respectively---do these spirals have other self-intersections at algebraic values of $n$? Can we use the analytic continuation of $\vv{L}$ and $\cc{L}$ to define a natural, continuous geometric transformation from a regular $n$-gon to a regular $(n+1)$-gon? If so, what would a ``regular polygon'' with a non-integer number of sides look like? The notion of arranging polygons with increasing numbers of sides is not limited to the polygonal spiral discussed here. In particular, making changes to each of the rules of Definition \ref{dcon} offers flexibility for discovering a range of intriguing polygonal constructions.
{ "timestamp": "2022-11-15T02:01:39", "yymm": "2211", "arxiv_id": "2211.06484", "language": "en", "url": "https://arxiv.org/abs/2211.06484", "abstract": "We construct a polygonal spiral by arranging a sequence of regular $n$-gons such that each $n$-gon shares a specified side and vertex with the $(n+1)$-gon in the construction. By offering flexibility for determining the size of each $n$-gon in the spiral, we show that a number of different analytical and asymptotic behaviors can be achieved.", "subjects": "Metric Geometry (math.MG); Classical Analysis and ODEs (math.CA)", "title": "A regular $n$-gon spiral", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9871787849789998, "lm_q2_score": 0.8104789063814617, "lm_q1q2_score": 0.8000875820527599 }
https://arxiv.org/abs/1305.5394
A note on the Waring ranks of reducible cubic forms
Let $W_3(n)$ be the set of Waring ranks of reducible cubic forms in $n+1$ variables. We prove that $W_3(n)\subseteq \lbrace 1,..., 2n+1\rbrace$.
\section{Introduction} \indent Let $K$ be an algebraically closed field of characteristic zero, let $V$ be a $(n+1)$-dimensional $K$-vector space and $F\in S^d V$, namely a homogeneous polynomial of degree $d$ in $n+1$ indeterminates. The {\bf Waring problem for polynomials} asks for the least value $s$ such that there exist linear forms $L_1, \ldots, L_s,$ for which $F$ can be written as a sum \begin{equation} F=L_1^d+\ldots+L_s^d. \end{equation} \noindent This value $s$ is called the \textit{Waring rank}, or simply the \textit{rank}, of the form $F$, and here it will be denoted by $rk(F)$. The Waring problem for a \textit{general form} $F$ of degree $d$ was solved by Alexander and Hirschowitz, in their celebrated paper \cite {AH}. \begin{theorem}[{\bf Alexander-Hirschowitz \cite{AH}}]\label{alexander-hirschowitz} A general form $F$ of degree $d$ in $n+1$ variables is the sum of $\lceil \frac{1}{n+1}\binom{n+d}{d}\rceil$ powers of linear forms, unless \begin{description} \item $d=2$, $s=n+1$ instead of $\lceil \frac{n+2}{2}\rceil$; \item $d=3$, $n=4$ and $s=8$ instead of $7$; \item $d=4$, $n=2,3,4$ and $s=6,10,15$ instead of $5,9,14$ respectively. \end{description} \end{theorem} \begin{remark} The assumption on the characteristic is not necessary, see \cite{IK} for more details. \end{remark} \indent The Waring problem in the case of a given homogeneous polynomial is far from being solved. A major development in this direction is made in \cite{CCG} where the rank of any monomial and the rank of any sum of pairwise coprime monomials are computed. \\ The present paper concerns with the Waring rank of reducible cubic forms. The main result of this work is the following theorem. \begin{theorem}\label{theo1} Let $W_3(n)$ be the set of ranks of reducible cubic forms in $n+1$ variables, then $$ W_3(n)\subseteq \lbrace 1,\ldots,2n+1\rbrace. $$ \end{theorem} \section{The Apolarity} In this section, we recall basic definitions and facts; see \cite{IK} and \cite{RS} for details. \indent Let $K$ be an algebraically closed field of characteristic zero, $S=\bigoplus_{i\geq 0} S_i=K[x_0, \ldots, x_n]$ and $T=\bigoplus_{i\geq 0} T_i=K[\partial_0, \ldots, \partial_n]$ be the \textit{dual ring} of $S$ (i.e. the ring of differential operators over $K$). $T$ is an $S$-module acting on $S$ by differentiation $$ \partial^\alpha (x^\beta)=\left\{\begin{array}{rl} \alpha ! \binom{\beta}{\alpha} x^{\beta-\alpha} &\mbox{ if } \beta \geq \alpha\\ 0 &\mbox{ otherwise. } \\ \end{array}\right. $$ \noindent where $\alpha$ and $\beta$ are multi-indices. The action of $T$ on $S$ is classically called {\bf apolarity}. Note that $S$ can also act on $T$ with a (dual) differentiation, defined by $$ x^\beta (\partial^\alpha)=\beta ! \binom{\alpha}{\beta} \partial^{\alpha-\beta}, $$ if $\alpha \geq \beta$ and $0$ otherwise.\\ \noindent In this way, we have a non-degenerate pairing between the forms of degree $d$ and the homogeneous differential operators of order $d$. Let us recall some basic definitions. \begin{definition} Let $F \in S$ be a form and $D\in T$ be a homogeneous differential operator. Then $D$ is \textit{apolar} to $F$ if $D(F)=0$. \end{definition} \begin{definition} For any $F\in S^dV$, we define the ideal $F^\perp=\lbrace D\in T | D(F)=0 \rbrace \subset T$, called the \textit{principal system} of $F$. If $F\in S^dV$, for every homogeneous operator $D\in T$ of degree $\geq d+1$, we have $D(F)=0$, or equivalently $D\in F^\perp$. The principal system of $F$ is a \textit{Gorenstein ideal}. \end{definition} \begin{definition} Given a homogeneous ideal $I\subset T$, the Hilbert function $\mbox{HF}$ of $T/I$ is defined as $$ \mbox{HF}(T/I,i)=\dim_K T_i -\dim_K I_i. $$ \noindent The first difference function $\Delta \mbox{HF}$ of the Hilbert function of $T/I$ is defined as $$ \Delta \mbox{HF}(T/I,i)=\mbox{HF}(T/I,i)-\mbox{HF}(T/I,i-1), $$ \noindent where $\mbox{HF}(T/I,-1)$ is set to be zero. \end{definition} \noindent Now, we recall the key result of this section. \begin{lemma}[Apolarity lemma]\label{ApolarityLemma} A form $F\in S^dV$ can be written as \begin{equation} F=\sum_{i=1}^s L_i^d, \end{equation} \noindent where $L_i$ are linear forms pairwise linearly indipendent, if and only if there exists an ideal $I \subset F^\perp$ such that $I$ is the ideal of a set of $s$ distinct points in $\mathbb P^n$, where these $s$ points are the corresponding points of the linear forms $L_i$ in the dual space $\mathbb P^{n*}$. \end{lemma} \indent For a proof of apolarity lemma \ref{ApolarityLemma} see for instance \cite{IK}. We will refer to the $s$ points of this lemma as \textit{decomposition points}. \section{Classification of Ranks of Reducible Cubic Forms in $\mathbb P^n$} \indent In this section we give the classification of the ranks of reducible cubic forms. Since the rank is invariant under projective transformations, we only need to check the projective equivalence classes of cubic forms. Let $W_3(n)$ be the set of values of ranks of reducible cubic forms in $n+1$ variables, namely forms of type $F=LQ$, where $L,Q\in S$ are linear and quadratic forms respectively. In order to give a classification, note that $W_3(n-1)\subset W_3(n)$. Indeed, every form in $n$ indeterminates is also a form in the ring of polynomials in $n+1$ indeterminates and the ranks as polynomial in $n$ variables and as polynomial in $n+1$ variables are equal. The subset $W_3(n-1)\subset W_3(n)$ is the set of the ranks of reducible cones in $n+1$ variables. The forms $F=LQ$ which are not cones (up to projective equivalence) are the following. \begin{itemize} \item (Type $A$) $Q$ is not a cone and $L$ is not tangent to $Q$. \begin{center} \includegraphics[scale=0.1]{TypeA.png} \end{center} \item (Type $B$) $Q$ is a cone and $L$ does not pass through any vertex of $Q$. \begin{center} \includegraphics[scale=0.08]{TypeB.png} \end{center} \item (Type $C$) $Q$ is not a cone and $L$ is tangent to $Q$. \begin{center} \includegraphics[scale=0.1]{TypeC.png} \end{center} \end{itemize} \newpage We have the following result. \begin{theorem}\label{Theo2} The ranks of reducible cubic forms $A$, $B$ and $C$ in $n+1$ variables are the following. \begin{center} \begin{tabular}{ | l | l |} \hline Type & Rank \\ \hline $A$ & $= 2n$ \\ \hline $B$ & $= 2n$ \\ \hline $C$ & $\geq 2n,\leq 2n+1$ \\ \hline \end{tabular} \end{center} \end{theorem} The ranks of cubic forms of type $A$ and $B$ are given by [\cite{LT}, Proposition $7.2$]. B. Segre proved that the cubic surface in $\mathbb P^3$ of type $C$ has rank $7$ \cite{Segre}. \subsection{Type $C$} Cubic forms of type $C$ are projectively equivalent to the cubic form \begin{equation}\label{cubicform} F=x_0(x_0x_1+x_2x_3+x_4^2+\ldots +x_n^2). \end{equation} \begin{Notation} We denote by $\int G dx_i$ a suitable choice of a primitive of $G$ (that will be specified any time it is needed), namely a form $H$ such that $\partial_i H=G$, where $\partial_i$ denotes the usual partial derivative with respect to the variable $x_i$. \end{Notation} First, note that if $n=2$, we have this proposition. \begin{proposition}\label{prop6} The cubic form $F=x_0(x_0x_1+x_2^2)$ has rank $\leq 5$. \begin{proof} Consider the coordinate system given by the following linear transformation. \begin{equation}\label{coordinate1} \left\{\begin{array}{rl} x_0=y_1\\ x_1=\frac{1}{3}y_1+y_3\\ x_2=y_2\\ \end{array}\right. \end{equation} \noindent By this, we have $F=\frac{1}{3}y_1^3+y_1^2y_3+y_1y_2^2$. Let $K_1=\int \partial_2 F dy_2$ be the primitive of $\partial_2F$ given by $K_1=\frac{1}{6}[(y_1+y_2)^3+(y_1-y_2)^3]=\frac{1}{3}y_1^3+y_1y_2^2$. Thus $F=K_1+y_1^2y_3$, where $y_1^2y_3=\frac{1}{6}[(y_1+y_3)^3-(y_1-y_3)^3-2y_3^3]$. Then $rk(F)\leq 5$, which proves the statement. \end{proof} \end{proposition} \noindent It is straightforward to generalize this fact as follows. \begin{proposition}\label{prop7} The cubic form $F=x_0(x_0x_1+x_2x_3+x_4^2+\ldots +x_n^2)$ has rank $\leq 2n+1$. \begin{proof} We prove it by induction on $n$. The proposition holds for $n=2$ by Proposition \ref{prop6}. Let us suppose the proposition true for all $i\leq n-1$ and prove the case $i=n$. Introduce the coordinate system given by the following linear transformation. \begin{equation}\label{coordinate3} \left\{\begin{array}{rll} x_0=y_1\\ x_1=y_3\\ x_2=y_0+y_2\\ x_3=y_0-y_2\\ x_4=y_4\\ \vdots \ \ \ \ \ \\ x_n=y_n\\ \end{array}\right. \end{equation} \noindent Then, the cubic becomes $F=y_0^2y_1-y_1y_2^2+y_1^2y_3+y_1y_4^2+\ldots +y_1y_n^2$. Setting $G=\int \partial_0 F dy_0=y_0^2y_1+\frac{1}{3} y_1^3$, we take $F=G - \frac{1}{3} y_1^3-y_1y_2^2+y_1^2y_3+y_1y_4^2+\ldots+y_1y_n^2$. We have that $rk(G)=2$. Let $H=-\frac{1}{3}y_1^3-y_1y_2^2+y_1^2y_3+y_1y_4^2+\ldots +y_1y_n^2$. Since $H$ is a cubic form in $\mathbb P^{n-1}$ decomposed into a smooth quadric $Q$ and a tangent space $L$ to a point of $Q$ (and hence it is of type $C$), by inductive assumption $rk (H)\leq 2(n-1)+1$. Thus $rk (F)\leq rk (G) + rk (H)\leq 2+2(n-1)+1=2n+1$. Repeating the argument, one obtains a decomposition for $F$. \end{proof} \end{proposition} \begin{remark} By [\cite{LT}, Theorem $1.3$], the rank of the cubic forms of type $C$ is $\geq 2n$. \end{remark} \begin{remark} The ranks of the reducible cubic forms are quite different from the generic rank of cubic forms given by the Alexander-Hirschowitz Theorem \ref{alexander-hirschowitz}: for sufficiently large values of $n$, the ranks of reducible cubics are smaller than the rank of the generic cubic. \end{remark} \section{Proof of Theorem \ref{theo1}} \begin{proof} We prove it by induction on $n$. If $n=1$, it is well known that cubic forms (actually, forms of any degree) in $2$ variables have rank at most their degree; in this case the set of ranks is exactly $W_3(1)$. Suppose that the statement holds for $i\leq n-1$ and we want to show it for $i=n$. Consider $ W_3(n)\setminus W_3(n-1)$; applying Theorem $\ref{Theo2}$, there exist forms of ranks $2n$ and of rank at most $2n+1$. By induction, $W_3(n-1)\subseteq \lbrace 1,\ldots,2n-1\rbrace$, and so $W_3(n)\subseteq \lbrace 1,\ldots,2n+1\rbrace$. \end{proof} Motivated by the result of Segre \cite{Segre}, we state the following \begin{conjecture}\label{conj on cubic form of type C} The Waring rank of the reducible cubic forms of type $C$ in $n+1$ variables is $2n+1$. \end{conjecture} \begin{remark}\label{remark on HF} The conjecture above states that $F=y_0^2y_1-y_1y_2^2+y_1^2y_3+y_1y_4^2+\ldots +y_1y_n^2$ has rank $\geq 2n+1$. The ideal $F^\perp$ is minimally generated by $\partial_i\partial_3$ (for $i\neq 1$), $\partial_1\partial_3-\partial_i^2$ (for $i\neq 1,2,3$), $\partial_1\partial_3+\partial_2^2$, $\partial_i\partial_j$ (for $i,j\neq 1,3$), $\partial_i^3$ (for $i \neq 3$), $\partial_1^2\partial_i$ (for $i\neq 3$). \\ \indent The degree of a zero-dimensional scheme can be computed using Hilbert functions. Let $\mathbb X$ be a set of decomposition points of $F$ and set $I=I(\mathbb X)\subset F^\perp$. Let us suppose that $\mathbb X$ has no points on $\lbrace \partial_3=0\rbrace$. In this case, $\partial_3$ is not a zero-divisor in $T/I$, which is crucial here. Then the degree of $\mathbb X$ is given by $$ \deg \mathbb X=\sum_{i\geq 0} \Delta \mbox{HF}(T/I,i)=\sum_{i\geq 0} \mbox{HF}(T/(I+\langle \partial_3\rangle),i)\geq \sum_{i\geq 0} \mbox{HF}(T/(F^\perp+\langle \partial_3\rangle),i)=2n+1, $$ \noindent where the Hilbert function $\mbox{HF}$ of $F^\perp+\langle\partial_3\rangle$ is the sequence $(1,n,n,0,-\cdots)$.\\ The case when $\mathbb X$ has points on $\lbrace \partial_3=0\rbrace$ requires a more careful analysis which we show for $n=2$. \end{remark} We propose a technique based on apolarity and Hilbert functions that might be generalized to higher dimensions. We will show it dealing with the known case $n=2$. \begin{example n=2} Let us denote $T=\mathbb C[\partial_1,\partial_2,\partial_3]$. In this case, we have $F=y_1(y_1y_3+y_2^2)$. The principal system of $F$ is the ideal $F^\perp=\langle \partial_1\partial_3-\partial_2^2,\partial_2\partial_3,\partial_3^2,\partial_1^3,\partial_1^2\partial_2,\partial_2^3\rangle$. Let $\mathbb X$ be a set of decomposition points of $F$ and let us set $I=I(\mathbb X)$.\\ \indent If $\mathbb X$ has no points on $\lbrace\partial_3=0\rbrace$ then $\partial_3$ is not a zero-divisor of $T/I$. Then $$\deg \mathbb X=\sum_{i\geq 0} \Delta \mbox{HF}(T/I,i)=\sum_{i\geq 0} \mbox{HF}(T/(I+\langle\partial_3\rangle),i)\geq \sum_{i\geq 0} \mbox{HF}(T/(F^\perp+\langle\partial_3\rangle),i)=5.$$ \noindent Indeed, $I+\langle\partial_3\rangle\subset F^\perp+\langle\partial_3\rangle=\langle \partial_3,\partial_1\partial_2,\partial_2^2,\partial_1^3,\partial_2^3 \rangle$ and the Hilbert function of $T/(F^\perp+\langle\partial_3\rangle)$ is the sequence $(1,2,2,0,-\cdots)$, as in Remark \ref{remark on HF} above.\\ \indent Let us assume that $\mathbb X$ has some point on $\lbrace\partial_3=0\rbrace$. If $\dim I_2\leq 1$ then the Hilbert function of $T/I$ is the sequence $(1,3,m\geq 5,\ldots)$ and hence again $\deg \mathbb X\geq 5$. So let us assume $\dim I_2\geq 2$. Note that $I_2\subset F^\perp_2=\langle \partial_1\partial_3-\partial_2^2,\partial_2\partial_3,\partial_3^2\rangle$. There exists a two-dimensional subspace of conics $L\subset I_2\subset F^\perp_2$. Either this space $L$ is the pencil $a\partial_3^2+b\partial_2\partial_3$, and the base locus of this pencil is $\lbrace\partial_3=0\rbrace$, or $I_2$ contains some irreducible conic of equation $\partial_1\partial_3-\partial_2^2+a\partial_3^2+b\partial_2\partial_3$, whose only common intersection with $\lbrace\partial_3=0\rbrace$ is the point $(1:0:0)$. The first case is not possible, since otherwise $\mathbb X\subset \lbrace\partial_3=0\rbrace$, namely $\partial_3 F=0$, which is false. Hence we have $\mathbb X\cap\lbrace\partial_3=0\rbrace=\{(1:0:0)\}.$ This implies that $\mathbb X\cap\lbrace\partial_3=0\rbrace\subset \mathbb X\cap\lbrace\partial_2=0\rbrace$. Then $\partial_3$ does not vanish at any point of $\mathbb X\cap\lbrace\partial_2\not=0\rbrace=\mathbb X'$. Note that $\deg \mathbb X'\leq \deg \mathbb X-1$ because the point $(1:0:0)$ does not belong to $\mathbb X'$. Setting $J=(I:\partial_2)$ the ideal of $\mathbb X'$, we have that $\partial_3$ is not a zero-divisor of $T/J$, so we can compute $$\deg \mathbb X'=\sum_{i\geq 0} \mbox{HF}(T/(J+\langle\partial_3\rangle),i)\geq \sum_{i\geq 0} \mbox{HF}(T/((F^\perp:\partial_2)+\langle\partial_3\rangle),i)\geq 4,$$ \noindent since $(F^\perp:\partial_2)+\langle\partial_3\rangle=\langle \partial_3,\partial_1^2,\partial_2^2\rangle$. Finally $\deg \mathbb X\geq\deg \mathbb X'+1\geq 5$, which says that the rank of $F$ is at least $5$ using the apolarity lemma \ref{ApolarityLemma}. \end{example n=2} \begin{Acknowledgement} This paper is part of author's Master thesis at University of Catania. I would like to thank Riccardo Re for his support and encouragement. I also would like to thank Enrico Carlini and Zach Teitler for insightful discussions. \end{Acknowledgement}
{ "timestamp": "2014-12-18T02:12:34", "yymm": "1305", "arxiv_id": "1305.5394", "language": "en", "url": "https://arxiv.org/abs/1305.5394", "abstract": "Let $W_3(n)$ be the set of Waring ranks of reducible cubic forms in $n+1$ variables. We prove that $W_3(n)\\subseteq \\lbrace 1,..., 2n+1\\rbrace$.", "subjects": "Algebraic Geometry (math.AG); Commutative Algebra (math.AC)", "title": "A note on the Waring ranks of reducible cubic forms", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9838471663618993, "lm_q2_score": 0.8333245932423308, "lm_q1q2_score": 0.8198640397211494 }
https://arxiv.org/abs/1809.05769
Differentiation Matrices for Univariate Polynomials
We collect here elementary properties of differentiation matrices for univariate polynomials expressed in various bases, including orthogonal polynomial bases and non-degree-graded bases such as Bernstein bases and Lagrange \& Hermite interpolational bases.
\section{Introduction} The transformation of the (possibly infinite) vector of coefficients $\mathbf{a}={\{a_k\}}_{k\geq0}$ in the expansion \begin{equation} f(x)=\sum_{k\geq0} a_k\phi_k(x) \end{equation} \noindent to the vector of coefficients $\mathbf{b} = \{b_k\}_{k\geq0}$ in the expansion \begin{equation} f'(x) = \sum_{k\geq0} a_k\phi_k(x) \end{equation} \noindent is, of course, linear because the differentiation operator is linear\footnote{We are assuming that $f(x)$ is a differentiable function and that the set $\{\phi_k\}_{k\geq0}$, which we will now sometimes collect in a row vector $\mathbf{\phi}$, is a complete basis. However, the bulk of this paper will be about finite $\mathbf{\phi}$ representing polynomials of degree at most~$n$. Note that we represent $f'(x)$ and $f(x)$ in the same basis.}. Here the $\phi_k(x)$ are univariate polynomials. The matrix representation of the linear transformation from $\mathbf{a}$ to $\mathbf{b}$, denoted $\mathbf{b}=\mathbf{D}_{\phi}\mathbf{a}$, is \begin{equation} \mathbf{D}_{\phi}=[d_{ji}] \end{equation} \noindent where $d_{ij} = \frac{\partial\phi_i'(x)}{\partial\phi_j(x)}$ (note the transposition); that is, the $d_{ij}$ are from \begin{equation} \phi_i'(x)=\sum_{j\geq0}d_{ij}\phi_j(x)\>. \end{equation} $\mathbf{D}_{\phi}$ is called a \textsl{differentiation matrix}. In vector form, $f(x)=\boldsymbol{\phi}(x)\mathbf{a}$ (where $\boldsymbol{\phi}=[\phi_0(x),\phi_1(x),\dots]$) so \begin{align} f'(x)=&\boldsymbol\phi'(x)\mathbf{a} \nonumber \\ =&\boldsymbol\phi(x)\mathbf{D}_{\phi}\mathbf{a} \nonumber\\ =&\boldsymbol{\phi}(x)\mathbf{b}\>. \end{align} Alternatively, we might work with $f'(x) = \mathbf{b}^T\boldsymbol{\phi}^T(x)$ and in that case use the transpose of $\ensuremath{\mat{D}}$, in $\mathbf{b}^T = \mathbf{a}^T\ensuremath{\mat{D}}^T$. The most familiar differentiation matrix is of course that of the monomial basis $\phi_k(x)=x^k$. The $4\times 4$ differentiation matrix, for polynomials of degree at most 3, is in this basis, \begin{equation} \ensuremath{\mathbf{D}_{\mathrm{monomial}}} = \left[ \begin{array}{cccc} 0 & 1 & 0& 0 \\ 0& 0 & 2 & 0 \\ 0 & 0 & 0 & 3\\ 0 & 0 & 0 & 0\\ \end{array} \right]. \end{equation} This generalizes easily to the degree $n$ case. This operation is so automatic that it's only rarely realized that it even has a matrix representation. If we are truncating to polynomials of degree at most~$n$ then the finite matrix $\mathbf{D}_{\textrm{monomial}}$ is defined by: \begin{equation}\label{DM} \mathbf{D}_{\textrm{monomial}} = \left[ \begin{array}{cccccc} 0 & 1 & 0&\cdots\\ 0 & 0 & 2 & 0 &\cdots \\ & & 0 & 3 & 0 &\\ & & & \ddots & \ddots & \\ & & & & & n\\ &&&&& 0\\ \end{array} \right], \end{equation} an $n+1$ by $n+1$ matrix. Differentiation matrices in other bases, such as the Chebyshev basis, Lagrange interpolational basis, or the Bernstein basis, are also useful in practice and we will see several explicit examples. \subsection{Reasons for studying Differentiation Matrices} Differentiation matrices are used in spectral methods for the numerical solution of ordinary and partial differential equations, going back to their implicit use by Lanczos with the Chebyshev basis\footnote{Actually, Lanczos used the generalized inverse, $\ensuremath{\mathbf{D}_{\mathrm{Chebyshev}}}^+$ which turns out to be bidiagonal for the Chebyshev basis; this is simpler for hand computation. We will see this later in this section.}. They can be used for quadrature, especially Filon or Levin quadrature, for highly oscillatory integrands. The first serious study seems to be~\cite{Don(1995)}. One of the present authors is working on this now~\cite{CorlessTrivedi:2018}. See also~\cite{Weideman},~\cite{olver2013fast}, and Chapter 11 of~\cite{corless2013graduate}. In this paper we study differentiation matrices that occur when using various polynomial bases. We confine ourselves to using one fixed basis $\{\phi_k\}_{k\geq0}$ for both $f(x)$ and $f'(x)$, but sometimes there are advantages to using different bases for $f'$ than for $f$: see~\cite{olver2013fast} for an example. The reasons for using the Chebyshev basis or the Lagrange basis include superior conditioning of expressions for functions in those bases, and sometimes superior convergence. The reasons for studying general properties and basis-independent properties, as this paper does, include the power of abstraction and the potential to apply to the results of other bases perhaps more suited to the problem of your current interest. Another purpose is to see the relationships among the various bases. It helps exposition to have some example bases in mind, in order to make the general theory intelligible and interesting, so we describe the differentiation matrices for a few polynomial bases in the next section. \subsection{Example Differentiation Matrices} Before we give examples, we repeat the following general observation: The columns of $\mathbf{D_{\phi}}$ are the coefficients of the derivatives $\phi_k'$ expressed in the $ \{\phi_k\}$. \\ \noindent\textsl{Proof}: If $b=\phi_k(x)$ then $\mathbf{b}=\mathbf{e}_k$ and $\mathbf{D}\mathbf{b}=\mathbf{d}_k$ the $k$-th column of $\mathbf{D}$; but $\mathbf{b}'(x)=\sum_{j=0}^n c_j\phi_j(x)$ for some $c_j$, and $d_k=[c_0, c_1, ..., c_n]^T$, by definition. \\ \noindent\textsl{Corollary}: if $\{\phi_k\}$ is degree-graded (\textsl{i.e.} $\mathrm{deg}\phi_k=k$), then $\mathbf{D}$ is strictly upper triangular. This is not true, of course, if $\{\phi_k\}$ is not degree-graded, \textsl{e.g.} $\phi_k$ is a Bernstein, Lagrange or Hermite interpolational basis. \subsubsection{Chebyshev Polynomials} One of the first kinds of differentiation matrices to be studied was for Chebyshev polynomials, \textsl{i.e}. $T_0(x)=1$, $T_1(x) = x$ and $T_{k+1} = 2xT_k(x)-T_{k-1}(x)$; alternatively, $T_k(x) = \cos(k\arccos(x))$ on $-1\leq x \leq1$. See for instance~\cite{olver2013fast} or (more briefly) Chapter 2 of~\cite{corless2013graduate}. For a thorough and modern introduction with application to the Chebfun software project see~\cite{berrut2004barycentric}. The derivative of $T_k(x)$ is explicitly given in terms of $T_0, T_1,...,T_{k-1}$ as a sum, in~\cite{olver2013fast} and as a Maple program in~\cite{corless2013graduate}. \begin{equation} \frac{dT_k(x)}{dx}= \begin{cases} 0 & k=0 \\ k(\frac{1+(-1)^{k-1}}{2})T_0+2k\sum_{j=0}^{\floor{\frac{k-1}{2}}} T_{k-1-2j}(x)& k\geq 1 \>.\\ \end{cases} \end{equation} Here the notation $\lfloor x\rfloor$ means the floor of $x$, the largest integer not greater than~$x$. From this formula we may construct the infinite differentiation matrix $\ensuremath{\mathbf{D}_{\mathrm{Chebyshev}}}$, defined by \begin{equation}\label{DC} \ensuremath{\mathbf{D}_{\mathrm{Chebyshev}}} = \left[ \begin{array}{ccccccccc} 0 & 1 & 0 & 3 & 0 & 5 & 0 & 7 & \cdots\\ & 0 & 4 & 0 & 8 & 0 & 12 & 0 & \cdots\\ & & 0 & 6 & 0 & 10 & 0 & 14 &\cdots\\ &&& 0 & 8 & 0& 12 & 0 &\cdots\\ &&&& 0 & 10 & 0 & 14&\cdots \\ &&&& & 0 & 12 & 0 &\cdots\\ &&&&&& 0 & 14 & \ddots \\ &&&&&&& 0 & \ddots \\ &&&&&&& & \ddots \end{array} \right]. \end{equation} As we see, the matrix is strictly upper triangular, just as the monomial basis matrix was; this is because the degree of $T_k'$ is $k-1$. Finite order differentiation matrices for Chebyshev polynomials are merely truncations of this. For a recent application of this matrix to the solution of pantograph equations, see~\cite{YANG2018132}. \medskip \goodbreak \textbf{Remark} Lanczos thought that this was cumbersome, and preferred the more compact \textsl{antiderivative} formulation (see~\cite{corless2013graduate} pp 125-126) \begin{equation} \int T_k(x)dx=\frac{1}{2(k+1)} T_{k+1}(x)-\frac{1}{2(k-1)}T_{k-1}(x)+\frac{k\sin{k\pi/2}}{k^2-1}\>, \end{equation} (giving a correct value $\frac{(T_2(x)+T_0(x))}{4}$ in the limit as $k\rightarrow1$; also $\int T_0(x)dx=T_1(x)).$ This allows a more simple transformation from the \textsl{derivative} \begin{equation} f'(x)=\sum_{k\geq0} b_kT_k(x) \end{equation} to its \textsl{antiderivative} \begin{equation} f(x)=\sum_{k\geq0} a_kT_k(x) \end{equation} \noindent by what we will see is a generalized inverse of $\ensuremath{\mathbf{D}_{\mathrm{Chebyshev}}}$: The infinite tridiagonal matrix $\ensuremath{\mathbf{D}_{\mathrm{Chebyshev}}}^+$, derived from equation (7), is except for the first row \begin{equation} \ensuremath{\mathbf{D}_{\mathrm{Chebyshev}}}^{+} = \left[ \begin{array}{cccccc} 0 & 0 & & & &\\ 1 & 0 & -1/2 & & &\\ & 1/4 & 0 & -1/4 & &\\ & & 1/6 & 0 & -1/6 & \\ &&& 1/8 & 0 & \ddots\\ &&&& 1/10 &\ddots\\ \end{array} \right]\>. \end{equation} This matrix is tridiagonal (with $0$ on main diagonal). Here the first row is $0$, meaning that an arbitrary constant can be added to the integral. We will see that the first row and the final column of truncations of this matrix will not matter for antiderivatives of degree $n-1$ polynomials. \subsubsection{Legendre Polynomials} The Legendre polynomials $\{P_n\}_n$ satisfy, $P_0(x)=1, P_1(x)=x$, \begin{equation} \int_{-1}^1P_n(x)P_m(x)dx=\frac{2}{n+1}[n=m] \end{equation} This is a combinatorial notation for what elsewhere is termed the Kronecker Delta function. Here $[n=m]$ is 1 when $n=m$ and 0 otherwise. This is called Iverson's convention in~\cite{knuth1992two}. For a discussion of the merit of this notation, see~\cite{knuth1992two}. The Legendre polynomials satisfy the three term recursion relation \begin{equation} (n+1)P_{n+1}-(2n+1)P_n+nP_{n-1}=0 \end{equation} By inspection, the differentiation matrix for polynomials $p(x)=\sum_{k\geq 0}c_kP_k(x)$ is, if $p'(x)=\sum_{k\geq 0}d_kP_k(x),$ \begin{gather}\label{DL} \begin{bmatrix} d_0 \\ d_1 \\ d_2 \\ \vdots \end{bmatrix} = \begin{bmatrix} 0 & 1 & 0 & 1 & 0 & 1 & 0 & 1 \cdots \\ 0 & 0 & 3 & 0 & 3 & 0 & 3 & 0 \cdots \\ 0 & 0 & 0 & 5 & 0 & 5 & 0 & 5 \cdots \\ 0 & 0 & 0 & 0 & 7 & 0 & 7 & 0 \cdots \\ & & & & \ddots & \ddots & \ddots & \ddots &\\ \end{bmatrix} \begin{bmatrix} c_ 0\\ c_1 \\ c_2 \\ \vdots \end{bmatrix} \end{gather} and \begin{gather} \begin{bmatrix} K \\ c_1 \\ c_2 \\ c_3\\ \vdots \end{bmatrix} = \begin{bmatrix} 0 & -\frac{1}{3} \\ 1 & 0 & -\frac{1}{5} \\ & \frac{1}{3} & 0 & -\frac{1}{7} \\ & & \frac{1}{5} & 0 & -\frac{1}{9} \\ & & & \frac{1}{7 }& 0 & \ddots \\ & & & & \frac{1}{9} & \ddots \\ & & & & & \ddots \\ \end{bmatrix} \begin{bmatrix} d_ 0\\ d_1 \\ d_2 \\ \vdots \end{bmatrix}. \end{gather} Like the matrix for the Chebyshev polynomials, the generalized inverse of $\ensuremath{\mathbf{D}_{\mathrm{Legendre}}}$ is tridiagonal. The simplicity of these matrices recommend them. \subsubsection{General Differentiation Matrix for Degree-Graded Polynomial Bases} Real polynomials $\{\phi_n(x)\}_{n=0}^{\infty}$ with $\phi_n(x)$ of degree $n$ which are orthonormal on an interval of the real line (with respect to some nonnegative weight function) necessarily satisfy a three-term recurrence relation (see Chapter 10 of \cite{Davis(1963)}, for example). These relations can be written in the form \begin{equation} \label{eq.rec} x \phi_j(x)=\alpha_j \phi_{j+1}(x) +\beta_j \phi_j(x) + \gamma_j\phi_{j-1}(x), \quad \quad j=0,1,\ldots, \end{equation} where the $\alpha_j,\;\beta_j,\;\gamma_j$ are real, $\alpha_j\ne0$, $\phi_{-1}(x)\equiv 0$ $\phi_0(x)\equiv 1$. Besides orthogonal polynomials, one can easily observe that the standard basis and Newton basis also satisfy (\ref{eq.rec}) with $\alpha_j= 1,\;\beta_j= 0,\;\gamma_j= 0$ and $\alpha_j= 1,\;\beta_j= z_j,\;\gamma_j= 0$, respectively where the $z_j$ are the nodes. \textbf{Lemma}: $\ensuremath{\mathbf{D}_{\mathrm{Degree-Graded}}}$ has the following structure: \begin{equation}\label{DG1} \ensuremath{\mathbf{D}_{\mathrm{Degree-Graded}}} = \left[ \begin{array}{cccc} 0&0&\cdots&0\\ & & & \vdots\\ & \Huge{\mathbf{Q}}&& \\ & & &0 \end{array} \right]^{T}, \end{equation} where \begin{equation}\label{DG} \mathbf{Q}_{i, j}= \left\{\begin{array}{cl} \frac{i}{\alpha_{i-1}}, & i=j\\ \frac{1}{\alpha_{i-1}}((\beta_{j-1}-\beta_{i-1})\mathbf{Q}_{i-1, j}+ \alpha_{j-2}\mathbf{Q}_{i-1, j-1}+ \gamma_j \mathbf{Q}_{i-1, j+1}- \gamma_{i-1}\mathbf{Q}_{i-2, j}). & i>j \end{array}\right.\end{equation} Any entry of $\mathbf{Q}$, with a negative or zero index is not considered in the above formula.\\ \noindent\textsl{Proof}: We provide the sketch of proof here. The proof itself is straightforward, but time-consuming. Taking the derivative of (\ref{eq.rec}) with respect to $x$, we have \begin{equation} \label{deq.rec} x \phi'_j(x)+ \phi_j(x)=\alpha_j \phi'_{j+1}(x) +\beta_j \phi'_j(x) +\gamma_j\phi'_{j-1}(x), \quad \quad j=0,1,\ldots, \end{equation} We let $j= 0$ in (\ref{deq.rec}) and simplify to get $$\phi'_1(x)= \frac{1}{\alpha_0}\phi_0(x).$$ We then let $j= 1$ in (\ref{deq.rec}) and simplify using (\ref{eq.rec}) with $j= 0$ and the result from the previous step to get $$\phi'_2(x)= \frac{\beta_0-\beta_1}{\alpha_0\alpha_1}\phi_0(x)+ \frac{2}{\alpha_1}\phi_1(x).$$ If we continue like this, and write the results in a matrix-vector form, the pattern stated in (\ref{DG}) will emerge. \\ We can now find the matrices that we have for the monomial basis in (\ref{DM}), Chebyshev basis (\ref{DC}) and Legendre basis (\ref{DL}) directly from (\ref{DG1}) simply by plugging in the specific values for the $\alpha_j$, $\beta_j$, and $\gamma_j$ for each of them. Another important degree-graded basis of this kind is the Newton basis. In the simplest case, let a polynomial $P(x)$ be specified by the data $\{ \left(z_j,{P}_j\right) \}_{j=0}^{n}$ where the $z_j$'s are distinct. The \textsl{Newton polynomials} are then defined by setting $N_0(x)=1$ and, for $k=1,\cdots,n,$ $$ N_k(x)=\prod_{j=0}^{k-1}(x-z_j)\>. $$ Then we may express \begin{equation} P(x)= \left[ \begin{array}{ccccc} a_0&a_1& \cdots &a_{n-1}&a_n \end{array} \right]\left[ \begin{array}{ccccc} N_0(x)\\N_1(x)\\\vdots\\N_{n-1}(x)\\N_n(x) \end{array} \right]\>. \end{equation} For $j=0,\cdots,n$, the $a_j$ can be found by divided differences as follows. \begin{equation} a_j=[P_0,P_1,\cdots,P_{j-1}], \end{equation} where we have $[P_j]=P_j$, and \begin{equation}\label{divdif} [P_{i},\cdots,P_{i+j}]=\frac{[P_{i+1},\cdots,P_{i+j}]-[P_i,\cdots,P_{i+j-1}]}{z_{i+j}-z_i}. \end{equation} A similar expression is possible even if the $z_j$ are not distinct, if we use \textsl{confluent} divided differences. We return to this later, but note that the Newton polynomials are well-defined for $z_j$ that are not distinct. Indeed, if they are all equal, say $z_j = a$, we recover Taylor polynomials $(z-a)^{j-1}$. If in (\ref{eq.rec}), we let $\alpha_j=1$, $\beta_j=z_j$ and $\gamma_j=0$, it will become the Newton basis. For $n= 4$, $\ensuremath{\mathbf{D}_{\mathrm{Newton}}}$, as given by (\ref{DG1}), has the following form. \begin{small}\begin{equation}\label{Nex}\hspace*{-0.5cm} \ensuremath{\mathbf{D}_{\mathrm{Newton}}}=\left[ \begin{array}{ccccc} 0&0&0&0&0\\1&0&0&0&0 \\z_{0}-z_{1}&2&0&0&0\\ (z_{0}-z_{2})(z_{0}-z_{1})& -2z_{2 }+z_{1}+z_{0}&3&0&0\\ (z_{0}-z_{3})(z_{0}-z_{2})(z_{0}-z_{1})& (z_{1}-z_{3})( z_{1}-2z_{2}+z_{0}) +(z_{0}-z_{2})(z_{0}-z_{1}) &-3z_{3}+z_{2}+z_{1}+z_{{0}}&4&0 \end{array} \right]^{T}\end{equation}\end{small} \subsubsection{Lagrange Bases} \bigbreak Differentiation matrices for Lagrange bases are particularly useful. See \cite{corless2013graduate} Chapter 2 for a detailed derivation. We give a summary here to establish notation. We suppose that function values $\rho_k$ are given at distinct nodes $\tau_k$ (that is, $\tau_k=\tau_i \Leftrightarrow i=k,$ for $0\leq k\leq n$). Then the barycentric weights $\beta_k$ are found once and for all from the partial fraction expansion \begin{equation} \frac{1}{w(\mathbf{z})}=\frac{1}{\prod_{k=0}^n (z-\tau_k)}=\sum_{k=0}^n \frac{\beta_k}{z-\tau_k}, \end{equation} giving \begin{equation} \beta_k=\prod_{\substack{j=0\\ j\neq k\\}}^n(\tau_k-\tau_j)^{-1}\>. \end{equation} These can be computed in a numerically stable fashion \cite{olver2013fast}, and once this has been done, the polynomial interpolant can be stably evaluated either by the first barycentric form \begin{equation} \rho(z)=w(z)\sum_{k=0}^n \frac{\beta_k\rho_k}{z-\tau_k} \end{equation} or the second, \begin{equation} \rho(z)=\frac{\sum_{k=0}^n \frac{\beta_k\rho_k}{z-\tau_k}}{\sum_{k=0}^n \frac{\beta_k}{z-\tau_k}}\>. \end{equation} See \cite{berrut2004barycentric} for details. Here we are concerned with the differentiation matrix \begin{equation} \ensuremath{\mathbf{D}_{\mathrm{Lagrange}}}=[{d}_{ij}] \end{equation} (as derived in many places, but for instance see the aforementioned Chapter 11 of \cite{corless2013graduate}).\\ We have that \begin{equation} d_{ij}=\frac{\beta_j}{\beta_i(\tau_i-\tau_j)} \text{ for } i\neq j \end{equation} and \begin{equation} d_{ii}=-\sum_{j\neq i}d_{ij} \end{equation} Construction of this matrix is an $O(n^2)$ process, and evaluation of the vector of polynomial derivatives $b$ by \begin{equation} \mathbf{b}=\ensuremath{\mathbf{D}_{\mathrm{Lagrange}}}\rho\>. \end{equation} is also an $O(n^2)$ process. Once this has been done, then $\rho'(z)$ can be evaluated stably by re-using the previously computed barycentric weights: \begin{equation} \rho'(z)=w(z)\sum_{k=0}^n \frac{\beta_kb_k}{z-\tau_k}\>. \end{equation} If the derivative is to be evaluated \textsl{very} frequently, it may be cost-effective to modify the weights and throw away one node. This is usually not worth the bother. \par\medskip\noindent \textsl{Example}(taken from chapter 11 in \cite{corless2013graduate}) Note that if $\tau=[-1, -\frac{1}{3},\frac{1}{3}, 1]$ then the differentiation matrix is \begin{equation} \ensuremath{\mathbf{D}_{\mathrm{Lagrange}}}= \left[ \begin{array}{cccc} -11 & \phantom{-}18 & -9 & \phantom{-}2\\ -2 & -3 & \phantom{-}6 & -1\\ \phantom{-}1 & -6 & \phantom{-}3 & \phantom{-}2\\ -2 & \phantom{-}9 & -18 & \phantom{-}11\\ \end{array} \right], \end{equation} so, \begin{equation} \ensuremath{\mathbf{D}_{\mathrm{Lagrange}}}^+= \frac{1}{360}\left[ \begin{array}{cccc} -81 & -147 & -123 & -9\\ -41 & -53 & -77 & -31\\ \phantom{-}31 & \phantom{-}77 & \phantom{-}53 & -41\\ \phantom{-}9 & \phantom{-}123 & \phantom{-}147 & \phantom{-}81\\ \end{array} \right]\>. \end{equation} If instead $\tau=[-1, -1/2, 1/2, 1]$, then it follows that \begin{equation} \ensuremath{\mathbf{D}_{\mathrm{Lagrange}}}= \frac{1}{6}\left[ \begin{array}{cccc} -19 & 24 & -8 & 3\\ -6 & 2 & 6 & -2\\ 2 & -6 & -2 & 6\\ -3 & 8 & -24 & 19\\ \end{array} \right], \end{equation} and \begin{equation} \ensuremath{\mathbf{D}_{\mathrm{Lagrange}}}^+ = \frac{1}{720}\left[ \begin{array}{cccc} -94 & -347 & -293 & 14\\ 94 & -193 & -247 & -14\\ 14 & 247 & 193 & -94\\ -14 & 293 & 347 & 94\\ \end{array} \right]. \end{equation} These matrices were displayed explicitly to demonstrate that, unlike the degree-graded case, the differentiation matrices are full, and their properties not very obvious\footnote{The row sums are zero, by design: the constant function has a constant vector representation, and its derivative should be (must be) zero. This is why $D_{ii}$ is the negative sum of all other entires.}. If $\tau=[1, i, -1, -i]$, \begin{equation} \ensuremath{\mathbf{D}_{\mathrm{Lagrange}}} = \frac{1}{2}\left[ \begin{array}{cccc} 3 & -1+i & -1 & -1-i\\ -1+i & -3i & 1+i & i\\ 1 & 1+i & -3 & 1-i\\ -1-i & -i & 1-i & 3i\\ \end{array} \right], \end{equation} and \begin{equation} \ensuremath{\mathbf{D}_{\mathrm{Lagrange}}}^+ = \frac{1}{24}\left[ \begin{array}{cccc} 11 & 4-3i & 5 & 4+3i\\ -3+4i & 11i & 3+4i & 5i\\ -5 & -4-3i & -11 & -4+3i\\ -3-4i& -i5 & 3-4i & -11i\\ \end{array} \right], \end{equation} which again has no obvious pattern (but see Theorem 11.3 of~\cite{corless2013graduate}: at least the singular values are simple). \subsubsection{Hermite Interpolational Bases} A Hermite interpolational basis is likely to be a bit less familiar to the reader than the Lagrange basis. They can be derived from Lagrange bases by letting two or more distinct nodes ``flow together'' (from whence the word confluency comes). Many methods to compute Hermite interpolational basis representations of polynomials are known that fit consecutive function values and derivative values (i.e. $f(\tau_i)$, $f'(\tau_i)/1!$, $\ldots$, $f^{(s_i-1)}(\tau_i)/(s_i-1)!$ are consecutive scaled values of the derivatives of $f$ at a particular node $\tau_i$, which is said to have confluency $s_i$, a non-negative integer). Many people use divided differences to express polynomials that fit confluent data, but this does not result in a Hermite interpolational basis (as pointed out in an earlier section, we would instead call that a Newton basis). We can solve the Hermite interpolation problem using a Newton basis, which is a degree-graded basis, and its differentiation matrix can be found through equation~\eqref{DG1}. Let's assume that at each node, $z_j$, we have the value and the derivatives of $P(x)$ up to the $s_j$-th order. The nodes at which the derivatives are given are treated as extra nodes. In fact we pretend that we have $s_j+1$ nodes, $z_j$, at which the value is $P_j$ and remember that $\sum_{i=0}^{k-1}s_i=n+1-k$. As such, the first $s_0+1$ nodes are $z_0$, the next $s_1+1$ nodes are $z_1$ and so on. Using the divided differences technique, as given by equation~\eqref{divdif}, to find the $a_j$, whenever we get $[P_j,P_j,\cdots,P_j]$ where $P_j$ is repeated $m$ times, we have \begin{equation} [P_j,P_j,\cdots,P_j]=\frac{P^{(m-1)}_j}{(m-1)!}, \end{equation} and all the values $P'_j$ to $P^{(s_j)}_j$ for $j=0,\cdots,k-1$ are given. For more details see e.g.~\cite{LJR(1983)}. For this confluent Newton basis, like the simple Newton basis, $\alpha_j= 1$, $\beta_j= z_j$, and $\gamma_j= 0$, but some of the $\beta_j$ are repeated. Other than that, the differentiation matrix can be found for this basis in a manner identical to $\ensuremath{\mathbf{D}_{\mathrm{Newton}}}$. This approach was used in~\cite{AFS2017} to find the coefficients of the Birkhoff interpolation. There are other advantages to solving the Hermite interpolation problem by using divided differences, for low degrees; the derivative is then almost directly available, for instance, and one does not really \textsl{need} a differentiation matrix. But there are numerical stability disadvantages to the confluent Newton basis. The main one is related to the relatively poor conditioning of the basis itself, for high-degree interpolants. [This does not matter much if the degree is low.] The next most important disadvantage is that the condition number of the polynomial expressed in this basis can be different if a different ordering of the nodes is used (it is usually true that the Leja ordering is good, but even so the condition number can be bad). See~\cite{Corless(2004)} for numerical experiments that confirm this. Another well-known solution to the Hermite interpolation problem involves constructing a basis that generalizes the Lagrange property, where each basis element is $1$ at one and only one node, and zero at all the others, which allows a direct sum to give the desired interpolant. One possible such definition (there are many variations) for a Hermite interpolational basis is to define it as a set of polynomials $H_{i,j}(z)$ with the index $i$ corresponding to the node indices, so if the nodes are $\tau_i$ with $0 \le i \le n$ then again for $H_{i,j}(z)$ we would have $0 \le i \le n$. The second index $j$ looks after the confluency at each node: $0 \le j \le s_i - 1$. Importantly, one needs consecutive derivative data at each node (else one has a Birkhoff interpolation problem~\cite{AFS2017}\cite{butcher2011polynomial}). Then we have the property (again written with the Iverson convention) \begin{equation} {H_{i,j}^{(k)}(\tau_\ell)} = [i=\ell][j=k]; \end{equation} that is, unless \textsl{both} the node indices are the same and the derivative indices are the same, the given derivative of basis polynomial is zero at the given node; if both the node indices and the derivative indices \textsl{are} the same, then the (scaled) Hermite basis element takes the value $1$. Using this definition, one can write the interpolant as a linear combination of this Hermite interpolational basis: $p(x) = \sum_{i=0}^n\sum_{j=0}^{s_i-1} \rho_{i,j}H_{i,j}(x)$. But there is a better way, that uses a stable partial fraction decomposition to get a collection of generalized barycentric weights $\ensuremath{\beta}_{i,j}$ that can be used to write down an efficient barycentric formula for evaluation of the polynomial. To be explicit, form the generalized node polynomial \begin{equation} w(z) = \prod_{i=0}^n (z-\tau_i)^{s_i}\>, \end{equation} which is exactly what you would get from the Lagrange node polynomial on letting each group of $s_i \ge 1$ distinct nodes flow together. Then the barycentric weights from the partial fraction decomposition of $1/w(z)$ must now account for the confluency: \begin{equation} \frac{1}{w(z)} = \sum_{i=0}^n \sum_{j=0}^{s_i-1} \frac{\beta_{i,j}}{(z-\tau_i)^{j+1}}\>. \label{eq:genparfrac} \end{equation} We will speak of the numerical computation of these $\beta_{i,j}$ shortly. Once we have them, we may simply write down barycentric forms of the polynomial that solves the Hermite interpolational problem: the first form is \begin{equation} p(z) = w(z) \sum_{i=0}^n \sum_{j=0}^{s_i-1} \sum_{k=0}^j \frac{\beta_{i,j}\rho_{i,k}}{(z-\tau_i)^{j+1-k}}\>. \end{equation} This form is simple to evaluate, and, provided the confluencies are not too large, numerically stable. This form can be manipulated into a second barycentric form by replacing $w(z)$ with the reciprocal of its partial fraction expansion, equation~\ref{eq:genparfrac}. The second form allows scaling of the generalized barycentric weights, which can prevent overflow. Incidentally, this allows us to give explicit expressions for the $H_{i,j}$ above: \begin{equation} H_{i,j}(z) = \sum_{k=0}^{s_i-1} \beta_{i,j+k} w(z)(z-\tau_i)^{-k-1}\>. \end{equation} [Equivalent expressions are given in the occasional textbook, but not all works on interpolation do so; the formula seems to be rediscovered frequently.] Given this apparatus, it makes sense to try to directly find the appropriate values of the derivatives at the nodes directly from the given function values and derivative values at the nodes; that is, by finding the differentiation matrix. Rather than give the derivation (a complete one can be found in chapter 11 of~\cite{corless2013graduate}) we point to both Maple code and Matlab code that implements those formulas, at \url{http://www.nfillion.com/coderepository/Graduate_Introduction_to_Numerical_Methods/} in \texttt{BHIP.mpl} and \texttt{genbarywts.m}, respectively. Evaluation in {\sc Matlab}\ can be done with the code \texttt{hermiteeval.m}. We give an example below. If a polynomial is known at three points, say $[-1,0,1]$, and the values of $p$, $p'$, and $p''/2$ are known at $-1$, while the values of $p$, $p'$, $p''/2$, and $p'''/6$ are known at $0$, and the values of $p$ and $p'$ are known at $1$, then the differentiation matrix is found to be \begin{equation} \left[ \begin {array}{ccccccccc} 0&1&0&0&0&0&0&0&0 \\ \noalign{\medskip}0&0&2&0&0&0&0&0&0\\ \noalign{\medskip}-{\frac{201 }{2}}&-{\frac{177}{4}}&-15&96&-60&24&-12&9/2&-3/4\\ \noalign{\medskip}0 &0&0&0&1&0&0&0&0\\ \noalign{\medskip}0&0&0&0&0&2&0&0&0 \\ \noalign{\medskip}0&0&0&0&0&0&3&0&0\\ \noalign{\medskip}{\frac{83}{ 4}}&6&1&-24&12&-12&4&{\frac{13}{4}}&-1/2\\ \noalign{\medskip}0&0&0&0&0 &0&0&0&1\\ \noalign{\medskip}35&11&2&0&48&0&16&-35&11\end {array} \right] \>. \end{equation} Applying this to the vector of values known at the nodes gives us the values of $[p'(-1)$, $p''(-1)$, $p'''(-1)/2$, $p'(0)$, $p''(0)$, $p'''(0)/2$ , $p^{(iv)}(0)/6$, $p'(1)$, $p''(1)]^T$, which describe $p'(z)$ on these nodes in the same way that $p(z)$ was described. Notice that some rows are essentially trivial, and just move known values into their new places. Notice that the nontrivial rows will, when multiplied by vectors representing constants (that is, $[c, 0, 0, c, 0, 0, 0, c, 0]^T$) give the zero vector. The nontrivial rows are constructed by recurrence relations from the generalized barycentric weights $\ensuremath{\beta}_{i,j}$, which are themselves merely the coefficients in the partial fraction expansion of the node polynomial. There is more than one way to compute the generalized barycentric weights $\beta_{i,j}$. The fastest way that we know is the algorithm of~\cite{schneider1991hermite}, which internally uses a confluent Newton basis. Unfortunately, because it does so, it inherits the poor numerical stability of that approach. The codes referred to above use a direct local Laurent series expansion method instead, as outlined in~\cite{Henrici(1964)} for instance; this method is slower but much more stable. As discussed in~\cite{corless2013graduate}, however, it becomes less stable for higher confluency and cannot be perfectly backward stable even for $s_i \ge 3$. We will see an example in section~\ref{sec:HermiteExample}. \subsubsection{Bernstein Polynomials} The Bernstein differentiation matrix is a tridiagonal matrix. Its entries are as follows: \begin{equation} [\mathbf{D_B}]_{i,j} = \begin{cases} 2i-n & i=j \\ -i & j = i-1\\ n-i & j=i+1 \end{cases}\>. \end{equation} Here the row and column indices $i$ and $j$ run from $0$ to $n$. For polynomials of degree at most $n=4$ expressed in the Bernstein basis, the matrix is explicitly \begin{equation} \left[ \begin {array}{ccccc} -4&4&0&0&0\\ \noalign{\medskip}-1&-2&3&0 &0\\ \noalign{\medskip}0&-2&0&2&0\\ \noalign{\medskip}0&0&-3&2&1 \\ \noalign{\medskip}0&0&0&-4&4\end {array} \right] \>. \end{equation} This is slightly different to the differentiation formulation seen in the Computer-Aided Geometric Design literature (e.g.~\cite{farin2014curves}), in that we \textsl{preserve the basis} to express the derivative in, even though that derivative is (nominally only) one degree too high. Degrees of polynomials expressed in Bernstein bases can be \textsl{elevated}, however, and when they are too high, they can be \textsl{lowered} or \textsl{reduced}. Indeed finding the \textsl{actual} degree of a polynomial expressed in a Bernstein basis can be, if there is noise in the coefficients, nontrivial. Here we simply keep the basis that we use to express $p(x)$, namely \begin{equation} p(x) = \sum_{i=0}^n c_i B_i^n(x) \end{equation} where \begin{equation} B_i^n(x) = {n \choose i} x^i (1-x)^{n-i}\>. \end{equation} By explicit computation, we find that the first column of the differentiation matrix (containing $-n$ in the zeroth row and $-1$ in the first row) correctly expresses the derivative of $B^n_0(x)$: \begin{align} -nB^n_0(x) - B^n_1(x) &= -n(1-x)^{n} - n x(1-x)^{n-1} \nonumber\\ &= -n(1-x)^{n-1}(1-x + x) \nonumber \\ &= \frac{d}{dx}B^n_0(x)\>. \end{align} Similarly, for $1 \le i \le n-1$, \begin{align} (n-i+1)B^n_{i-1} + (2i-n)B^n_i - (i+1)B^n_{i+1} &= x^{i-1}(1-x)^{n-i-1}{n \choose i}\left( i(1-x)^2 + (2i-n)x(1-x) - (n-i)x^2 \right) \nonumber \\ &= x^{i-1}(1-x)^{n-i-1}{n \choose i}\left( i-nx \right) \nonumber \\ &= \frac{d}{dx}B^n_i(x)\>. \end{align} By the reflection symmetry of $B^n_n(x)$ with $B^n_0(x)$, the final column is also correct. \begin{remark} As with the Lagrange polynomial bases, the pseudo-inverse of the Bernstein basis differentiation matrix is full. Also as with the Lagrange case, because $1 = \sum B^n_k(x)$ (that is, the Bernstein basis forms a partition of unity), application of the Bernstein differentiation matrix to a constant vector must return the zero vector and hence the row sums must be zero. \end{remark} \section{Basic Properties} \textsl{Definition}: Let $\mathbf{X}_{\phi}^k$ be the vector of coefficients of $x^k$ in the basis $\phi$. That is, if \begin{equation} x^k =b_{k,0}\phi_0(x)+b_{k,1}\phi_1(x)+\dots+b_{k,n}\phi_n(x)\>, \end{equation} then \begin{equation} \mathbf{X}_{\phi}^k=[b_{k,0},b_{k,1},\dots,b_{k,n}]^T\>.\ \end{equation} Set $\mathbf{1}_{\phi}=\mathbf{X}_{\phi}^0$. Let $\mathbf{V}$ be the matrix whose $k$-th column (numbering from zero) is $\frac{1}{k!}\mathbf{X}_{\phi}^k$. \bigbreak \subsection{Eigendecomposition of Differentiation Matrices} Let $\mathbf{D}_{\phi}$ be the differentiation matrix for polynomials of degree at most $n$, expressed in the basis $\{\phi_k\}_{k=0}^n$. Note that if \begin{equation} b=\rho_0\phi_0+\rho_1\phi_1+...+\rho_n\phi_n \end{equation} then \begin{equation} \rho'=b_0\phi_0+b_1\phi_1+...+b_n\phi_n\>. \end{equation} same basis; for degree graded basis, $b_n=0$. Then $\mathbf{D}_{\phi}\mathbf{p}=\mathbf{b}$ where \begin{align} \mathbf{\rho}=\begin{bmatrix} \rho_{0} \\ \rho_{1} \\ \vdots \\ \rho_{n}\> \end{bmatrix} \end{align} and \begin{align} \mathbf{b}=\begin{bmatrix} b_{0} \\ b_{1} \\ \vdots \\ b_{n} \end{bmatrix}. \end{align} \textbf{Lemma}: $\mathbf{D}$ is nilpotent.\\ \noindent\textsl{Proof}: $\mathbf{D}^{n+1}\mathbf{\rho}(x)=0$ for every polynomial of degree at most $n$; hence $\mathbf{D}^{n+1}\mathbf{p}(x)=0$ as required.\\ \noindent$\mathbf{Remark}$: Therefore all eigenvalues are zero.\\ \noindent\textsl{Proposition} $\mathbf{D}_{\phi}\mathbf{V}=\mathbf{VJ}$, where \begin{equation} \mathbf{J}= \left[ \begin{array}{cccc} 0 & 1 & &\\ 0 & 0 & 1 & \\ \vdots&\vdots&\ddots&1\\ 0&&&0\\ \end{array} \right], \end{equation} is the Jordan Canonical Form of the differentiation matrix $\mathbf{D}_{\phi}$. \bigbreak \noindent\textsl{Proof}: $\mathbf{D}_{\phi}(\frac{1}{k!}\mathbf{X}_{\phi}^k)=\frac{1}{(k-1)!}\mathbf{X}_{\phi}^{k-1}$ for $k\geq 1$, by construction. Moreover the columns $\mathbf{X}_{\phi}^k$ are linearly independent because the monomials $1,x,x^2,\dots,x^n$ and $\phi$ are a polynomial basis. Thus $\mathbf{V}$ is invertible.\\ \noindent\textbf{Remark} The isomorphism of the polynomial representation by coefficient vectors (of the basis $\phi$) is complete for addition, subtraction, differentiation, and scalar multiplication; but the representation of $p\cdot q$ is possible only if $\mathrm{deg} p+\mathrm{deg} q\leq n$. The multiplication rules are interesting as well; we get the usual Cauchy convolution for the monomial basis. \bigbreak \subsection{Pseudoinverse} \textsl{Observation} As long as $\mathrm{deg} p<n$, anti-differentiation works by using the pseudo inverse; one then adds a constant times $\mathbf{1}_{\phi}$. Call this anti-differentiation matrix $\mathbf{S}$. Then we want $\mathbf{S1}_{\phi}=\mathbf{X}_{\phi},$ and $\mathbf{S}\frac{\mathbf{X}^{k-1}}{(k-1)!}=\frac{\mathbf{X}^k}{k!}$.\\ \bigbreak \noindent Therefore, \begin{equation}\mathbf{S}\mathbf{V}(1:n)=[\mathbf{X},\frac{\mathbf{X}^2}{2},\dots,\frac{\mathbf{X}^{n-1}}{(n-1)!},\frac{\mathbf{X}^n}{n!}]\>,\end{equation} \noindent and thus, \begin{equation} \mathbf{VJ^+V^{-1}[V_n]=VJ^+[0, 0, ..., 0, 1]^T=[0,0, ...., 0]^T\>.} \end{equation} \textbf{Lemma}: \textsl{The Moore-Penrose pseudo-inverse of} \begin{equation} \mathbf{J}= \left[ \begin{array}{cccc} 0 & 1 & &\\ 0 & 0 & 1 & \\ \vdots&\vdots&\ddots&1\\ 0&&&0\\ \end{array} \right], \end{equation} is \begin{equation} \mathbf{J}^+=\mathbf{J^T}= \left[ \begin{array}{cccccc} 0 & 0 & & & &\\ 1 & 0 & & & & \\ &1 & \ddots & & & \\ && \ddots && &\\ &&&&1&0\\ \end{array} \right]. \end{equation} \textsl{Proof}: We need to verify that $\mathbf{JJ^TJ=J},\mathbf{J^TJ=J}$ and that both $\mathbf{J^TJ}$ and $\mathbf{JJ^T}$ are symmetric. The last two are trivial. Computation shows \begin{equation} \mathbf{JJ^T=J^TJ}= \left[ \begin{array}{ccccc} 0 & 0 & 0 &&\\ 0 & 1 & 0 & & \\ &0 &1&&\\ &&&\ddots&\\ 0&&&&1\\ \end{array} \right], \end{equation} so \begin{equation} \mathbf{JJ^TJ}= \left[ \begin{array}{cccc} 0 & 1 & &\\ 0 & 0 & 1 & \\ \vdots&\vdots&\ddots&1\\ 0&&&0\\ \end{array} \right] = \mathbf{J}\>. \end{equation} Similarly $\mathbf{J^TJJ^T=J}$.\\ $\mathbf{Proposition}$: The matrix $\mathbf{D}^+=\mathbf{VJ^+V^{-1}}$ is a generalized inverse of $\mathbf{D}$. \textsl{Proof}: It suffices to verify only the first two of the Moore-Penrose conditions: $\mathbf{D^+DD^+}=\mathbf{D^+}$ and $\mathbf{DD^+D}=\mathbf{D}$. These follow immediately. Interestingly $\mathbf{D^+}$ is not (in general) a Moore-Penrose inverse: neither $\mathbf{D^+D}$ nor $\mathbf{DD^+}$ need be Hermitian. The matrix $\mathbf{V}$ in a Lagrange basis is \begin{equation} \mathbf{V} = \left[ \begin{array}{ccccc} 1 & \tau_0 & \cdots & \cdots & \frac{\tau_0^n}{n!} \\ 1 & \tau_1& & & \\ \vdots & \vdots &&&\\ 1& \tau_n &&& \frac{\tau_n^k}{n!} \end{array} \right]\>. \end{equation} This is the product of a Vandermonde matrix and \begin{equation} \left[ \begin{array}{ccccc} 1 & & &&\\ & 1 & & & \\ & & \frac{1}{2} &&\\ &&& \ddots &\\ &&&& \frac{1}{n!} \end{array} \right]. \end{equation} This is likely to be extraordinarily ill-conditioned. However, this gives an explicit JCF for differentiation matrices on Lagrange bases. \section{Accuracy and Numerical Stability} There are several questions regarding numerical stability (and, unfortunately, the answers vary with the basis used, and with the degree). For the orthogonal polynomial bases and the Bernstein bases, the differentiation matrices have integer or rational entries, and there are no numerical difficulties in constructing them, only (perhaps) with their use. For the Lagrange and the Hermite interpolational bases, the (generalized) barycentric weights need to be constructed from the nodes, and then the entries of the differentiation matrix constructed from the weights. In floating-point arithmetic, this can be problematic for some sets of nodes (especially equally-spaced nodes); higher-precision construction of the weights, or use of symmetries as with Chebyshev nodes, may be needed. High or variable confluency can also be a difficulty. Use of higher precision in construction of the barycentric weights and of the differentiation matrix may be worth it, if the matrix is to be used frequently. For all differentiation matrices, there is the question of accuracy of computation of the polynomial derivative by matrix multiplication. In general, differentiation is infinitely ill-conditioned: the derivative of $f(x) + \varepsilon v(x)$ can be arbitrarily different to the derivative of $f(x)$. However, if both $f$ and the perturbation are restricted to be \textsl{polynomial}, then the ill-conditioning is finite, and the absolute condition number is bounded by the norm of the differentiation matrix $\mathbf{D}$. This is Theorem 11.2 of~\cite{corless2013graduate}, which we state formally below. \begin{theorem} If $f(x)$ and $\Delta f(x)$ are both polynomials of degree at most $n$, and are both expressed in a polynomial basis $\phi$, then \begin{equation} \| \Delta f'(x) \| \le \| \mathbf{D}_{\phi} \| \| \Delta f (x) \| \end{equation} where the norms $\| \Delta f' \|$ and $\| \Delta f \|$ are vector norms of their coefficients in $\phi$ and the norm of the differentiation matrix is the corresponding subordinate matrix norm. \end{theorem} One should check the norm $\|\ensuremath{\mat{D}}\|$ whenever one uses a differentiation matrix. We remark that the norms of powers of $\ensuremath{\mat{D}}$ can grow very large. For instance, for the Bernstein basis of dimension $n+1$ we find\footnote{We have no proof, only experimental evidence; it should be possible to prove this but we have not done so.} that $\| \ensuremath{\mat{D}}^n \|_\infty = 2^n n!$. The next power gives the zero matrix, of course. To give a sense of scale, we have $\| \ensuremath{\mat{D}} \|_\infty = 2n$ and hence this norm to the $n$th power is much larger yet, being $(2n)^n$ so a factor $n^n/n! \approx \exp(n)/\sqrt{2\pi n}$ larger. As a corollary, from the results discussed in~\cite{Embree:HLA:2013} the $\varepsilon$-pseudospectral radius of the $n+1$-dimensional Bernstein $\ensuremath{\mat{D}}$ matrix must then at least be $(2^n n!)^{1/(n+1)} \varepsilon^{1/(n+1)} \sim 2n \varepsilon^{1/(n+1)}/e$ as $n \to \infty$, for any $\varepsilon > 0$. This implies that for large enough dimension, matrices very near to $\ensuremath{\mat{D}}$ will have eigenvalues larger than $1$ in magnitude. We believe that similar results hold for other bases, indicating that higher-order derivatives are hard to compute accurately by using repeated application of multiplication by differentiation matrices (as is to be expected). \subsection{A Hermite interpolational example \label{sec:HermiteExample}} Consider interpolating the simple polynomial that is identically $1$ on the interval $-1 \le z \le 1$, using nodes with confluency three. That is, at each node we supply the value of the function ($1$), the value of the first derivative ($0$), and the value of the second derivative divided by $2$, which is also in this case just $0$. We consider taking $n+1$ nodes $\tau_j$ for $0 \le j \le n$, which gives us $3(n+1)$ pieces of data and thus a polynomial of degree at most $3n+2$. We then plot the error $p(z)-1$ on this interval. We also compute the differentiation matrix $\ensuremath{\mat{D}}$ on these nodes with this confluency, and multiply $\ensuremath{\mat{D}}$ by the vector containing the data for the constant function $1$. This should give us an identically $0$ vector (call it $\vec{Z}$), but will not, because of rounding error. We compute the infinity norm of $\vec{Z}$ and the infinity norm of the matrix $\ensuremath{\mat{D}}$. We take two sets of nodes: first the Chebyshev nodes $\tau_j = \cos(\pi(n-j)/n)$, and second the equally-spaced nodes $\tau_j = -1 + 2j/n$. We take $n=3$, $5$, $8$, $\ldots$, $55$ (Fibonacci numbers). In figure~\ref{fig:Dnorm} we find a log-log plot of the norms of $\ensuremath{\mat{D}}$ for these $n$. Remember that the degree of the interpolant is at most $3n+2$. We see that the norm of $\ensuremath{\mat{D}}$ grows extremely rapidly for equally-spaced nodes (as we would expect). For Chebyshev nodes there is still substantial growth (for confluency $3$; for confluency $2$ there is less growth, and for confluency $4$ there is more), but for $n=55$ and confluency $3$ at all nodes we have $\|\ensuremath{\mat{D}}\| $ approximately $10^{10}$ which still gives some accuracy in $\vec{Z}$. In figure~\ref{fig:Znorm} we see the corresponding norms of $\vec{Z}$. The behaviour is as predicted. \par\medskip\noindent \textbf{Remark}. The confluency really matters. If we use just simple Lagrange interpolation, that is confluency $s_i=1$ at each node, then the interpolation on $n=55$ Chebyshev nodes is in error by no more than $3.5\cdot 10^{-12}$. Of course, the nominal degree is much lower than it was in the Hermite case with confluency $3$. When we up the degree to $165$, the Lagrange error is no more than $1.5\cdot 10^{-11}$. When the confluency is $3$, and $n=55$ which is comparable, the error is $1.4\cdot 10^{-5}$. \begin{figure}[h] \centering \includegraphics[width=0.5\textwidth]{Dnorm.png} \caption{A comparison of norms of the differentiation matrices for Hermite interpolational basis on $n+1$ nodes, of confluency $3$, between equally-spaced nodes (solid boxes) and Chebyshev nodes (circles). We see growth in $n$ for both sets of nodes, but much more rapid growth for equally-spaced nodes.} \label{fig:Dnorm} \end{figure} \begin{figure}[h] \centering \includegraphics[width=0.5\textwidth]{Znorm.png} \caption{A comparison of norms of the vector $\vec{Z} = \ensuremath{\mat{D}} \mathbf{1}$ in Hermite interpolational bases. Equally-spaced nodes (solid box) and Chebyshev nodes (circles). As expected, we have $\|\vec{Z}\| \approx \|\ensuremath{\mat{D}}\|\cdot 10^{-16}$ when working in double precision.} \label{fig:Znorm} \end{figure} \section{Concluding Remarks} Expressing a polynomial in a particular basis reflects a choice taken by a mathematical modeller. We believe that choice should be respected. Indeed, changing bases can be ill-conditioned, often at least exponentially in the degree. There are exceptions, of course: interpolation on roots of unity with a Lagrange basis can be changed to a monomial basis by using the DFT, and the conversion is perfectly well-conditioned; similarly changing from a Lagrange basis on Chebyshev-Lobatto points to the Chebyshev basis is also perfectly well-conditioned. But, usually, one wants to continue to work in the basis chosen by the modeller. This is particularly true of the Bernstein basis, which has an optimal conditioning property: out of all bases that are nonnegative on the interval $[0,1]$, the Bernstein basis expression has the optimal condition number~\cite{Farouki(1996)}. This property was extended to bases nonnegative on a set of discrete points by~\cite{Corless(2004)}, who proved that Lagrange bases can be better even than Bernstein bases. See also~\cite{carnicer2017optimal}, who independently proved the same. Differentiation is a fundamental operation, and it is helpful to be able to differentiate polynomials without changing bases. This paper has examined the properties of the matrices for accomplishing this. We found several of the results presented here to be surprising, notably that the Jordan Canonical Form for all the differentiation matrices considered here was the same. Likewise, that there is a uniform formula for a pseudo-inverse of all differentiation matrices of the type considered here was also a surprise. One can extend this work in several ways. One of the first might be to look at differentiation matrices for \textsl{compact finite differences}. These are no longer always exact, and the matrices arising are no longer nilpotent (though they have null spaces corresponding to the polynomials of low enough degree that they are exact for). There are also some further experiments to run on the differentiation matrices we have studied in this paper already. For instance, it would be interesting to know theoretically the growth of $\|\ensuremath{\mat{D}}^k\|$ for various dimensions $n$; we found that for the Bernstein basis of dimension $n+1$ we had $\| \ensuremath{\mat{D}}^n \|_\infty = 2^n n!$. Since for the monomial basis we have $\| \ensuremath{\mat{D}}^n \|_\infty = n!$, this suggests that the natural scale for such a comparison is to divide by $n!$ and indeed that seems logical, because then in essence we are comparing the size of Taylor coefficients instead of comparing the size of derivatives and it is the Taylor coefficients that have a geometric interpretation in terms of location of nearby singularities. We leave this study of the dependence of the norm of $\ensuremath{\mat{D}}^k$ in different bases for future work. \section*{Acknowledgements} This work was supported by a Summer Undergraduate NSERC Scholarship for the third author. The second author was supported by an NSERC Discovery Grant. We also thank ORCCA and the Rotman Institute of Philosophy. \bibliographystyle{plain}
{ "timestamp": "2018-09-18T02:07:36", "yymm": "1809", "arxiv_id": "1809.05769", "language": "en", "url": "https://arxiv.org/abs/1809.05769", "abstract": "We collect here elementary properties of differentiation matrices for univariate polynomials expressed in various bases, including orthogonal polynomial bases and non-degree-graded bases such as Bernstein bases and Lagrange \\& Hermite interpolational bases.", "subjects": "Numerical Analysis (math.NA)", "title": "Differentiation Matrices for Univariate Polynomials", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9838471647042428, "lm_q2_score": 0.8333245932423308, "lm_q1q2_score": 0.8198640383397836 }
https://arxiv.org/abs/1704.03913
Higher-order clustering in networks
A fundamental property of complex networks is the tendency for edges to cluster. The extent of the clustering is typically quantified by the clustering coefficient, which is the probability that a length-2 path is closed, i.e., induces a triangle in the network. However, higher-order cliques beyond triangles are crucial to understanding complex networks, and the clustering behavior with respect to such higher-order network structures is not well understood. Here we introduce higher-order clustering coefficients that measure the closure probability of higher-order network cliques and provide a more comprehensive view of how the edges of complex networks cluster. Our higher-order clustering coefficients are a natural generalization of the traditional clustering coefficient. We derive several properties about higher-order clustering coefficients and analyze them under common random graph models. Finally, we use higher-order clustering coefficients to gain new insights into the structure of real-world networks from several domains.
\section{Derivation of higher-order clustering coefficients} In this section, we derive our higher-order clustering coefficients and some of their basic properties. We first present an alternative interpretation of the classical clustering coefficient and then show how this novel interpretation seamlessly generalizes to arrive at our definition of higher-order clustering coefficients. We then provide some probabilistic interpretations of higher-order clustering coefficients that will be useful for our subsequent analysis. \subsection{Alternative interpretation of the classical clustering coefficient} Here we give an alternative interpretation of the clustering coefficient that will later allow us to generalize it and quantify clustering of higher-order network structures (this interpretation is summarized in Fig.~\ref{fig:ccf_def}). Our interpretation is based on a notion of clique expansion. First, we consider a $2$-clique $K$ in a graph $G$ (that is, a single edge $K$; see Fig.~\ref{fig:ccf_def}, row $C_2$, column 1). Next, we \emph{expand} the clique $K$ by considering any edge $e$ adjacent to $K$, i.e., $e$ and $K$ share exactly one node (Fig.~\ref{fig:ccf_def}, row $C_2$, column 2). This expanded subgraph forms a wedge, i.e., a length-$2$ path. The classical global clustering coefficient $\gccf{}$ of $G$ (sometimes called the transitivity of $G$~\cite{boccaletti2006complex}) is then defined as the fraction of wedges that are \emph{closed}, meaning that the $2$-clique and adjacent edge induce a $(2 + 1)$-clique, or a triangle (Fig.~\ref{fig:ccf_def}, row $C_2$, column 3)~\cite{barrat2000properties,luce1949method}. The novelty of our interpretation of the clustering coefficient is considering it as a form of clique expansion, rather than as the closure of a length-$2$ path, which is key to our generalizations in the next section. Formally, the classical global clustering coefficient is \begin{equation}\label{eq:global_ccf} \gccf{} = \frac{6 \lvert K_3 \rvert}{\lvert W \rvert}, \end{equation} where $K_3$ is the set of $3$-cliques (triangles), $W$ is the set of wedges, and the coefficient $6$ comes from the fact that each $3$-clique closes 6 wedges---the 6 ordered pairs of edges in the triangle. We can also reinterpret the local clustering coefficient~\cite{watts1998collective} in this way. In this case, each wedge again consists of a $2$-clique and adjacent edge (Fig.~\ref{fig:ccf_def}, row $C_2$, column 2), and we call the unique node in the intersection of the $2$-clique and adjacent edge the \emph{center} of the wedge. The \emph{local clustering clustering coefficient} of a node $u$ is the fraction of wedges centered at $u$ that are closed: \begin{equation}\label{eq:local_ccf} \lccf{}{u} = \dfrac{2\lvert K_3(u) \rvert}{\lvert W(u) \rvert}, \end{equation} where $K_3(u)$ is the set of $3$-cliques containing $u$ and $W(u)$ is the set of wedges with center $u$ (if $\lvert W(u) \rvert = 0$, we say that $\lccf{}{u}$ is undefined). The \emph{average clustering coefficient} $\accf{}$ is the mean of the local clustering coefficients, \begin{equation}\label{eq:avg_ccf} \accf{} = \frac{1}{\lvert \widetilde V \rvert}\sum_{u \in \widetilde V} \lccf{}{u}, \end{equation} where $\widetilde V$ is the set of nodes in the network where the local clustering coefficient is defined. \subsection{Generalizing to higher-order clustering coefficients}\label{sec:generalization} Our alternative interpretation of the clustering coefficient, described above as a form of clique expansion, leads to a natural generalization to higher-order cliques. Instead of expanding $2$-cliques to $3$-cliques, we expand $\ell$-cliques to $(\ell + 1)$-cliques (Fig.~\ref{fig:ccf_def}, rows $C_3$ and $C_4$). Formally, we define an $\ell$-wedge to consist of an $\ell$-clique and an adjacent edge for $\ell \ge 2$. Then we define the global $\ell$th-order clustering coefficient $\gccf{\ell}$ as the fraction of $\ell$-wedges that are closed, meaning that they induce an $(\ell + 1)$-clique in the network. We can write this as \begin{equation}\label{eq:global_ccf_l} \gccf{\ell} = \frac{(\ell^2 + \ell)\lvert K_{\ell + 1} \rvert}{\lvert W_{\ell} \rvert}, \end{equation} where $K_{\ell + 1}$ is the set of $(\ell + 1)$-cliques, and $W_{\ell}$ is the set of $\ell$-wedges. The coefficient $\ell^2 + \ell$ comes from the fact that each $(\ell + 1)$-clique closes that many wedges: each $(\ell+1)$-clique contains $\ell + 1$ $\ell$-cliques, and each $\ell$-clique contains $\ell$ nodes which may serve as the center of an $\ell$-wedge. Note that the classical definition of the global clustering coefficient given in Eq.~\ref{eq:global_ccf} is equivalent to the definition in Eq.~\ref{eq:global_ccf_l} when $\ell = 2$. We also define higher-order local clustering coefficients: \begin{equation} \label{eq:def_lccf} \lccf{\ell}{u} = \frac{\ell \lvert K_{\ell + 1}(u) \rvert}{\lvert W_{\ell}(u) \rvert}, \end{equation} where $K_{\ell + 1}(u)$ is the set of $(\ell + 1)$-cliques containing node $u$, $W_{\ell}(u)$ is the set of $\ell$-wedges with center $u$ (where the center is the unique node in the intersection of the $\ell$-clique and adjacent edge comprising the wedge; see Fig.~\ref{fig:ccf_def}), and the coefficient $\ell$ comes from the fact that each $(\ell + 1)$-clique containing $u$ closes that many $\ell$-wedges in $W_{\ell}(u)$. The $\ell$th-order clustering coefficient of a node is defined for any node that is the center of at least one $\ell$-wedge, and the average $\ell$th-order clustering coefficient is the mean of the local clustering coefficients: \begin{equation}\label{eq:def_accf} \accf{\ell} = \frac{1}{\lvert \widetilde V_\ell \rvert}\sum_{u \in \widetilde V_\ell} \lccf{\ell}{u}, \end{equation} where $\widetilde V_{\ell}$ is the set of nodes that are the centers of at least one $\ell$-wedge. To understand how to compute higher-order clustering coefficients, we substitute the following useful identity \begin{equation}\label{eq:wedge_identity} \lvert W_{\ell}(u) \rvert = \lvert K_{\ell}(u) \rvert \cdot (d_u - \ell + 1), \end{equation} where $d_u$ is the degree of node $u$, into Eq.~\ref{eq:def_lccf} to get \begin{equation} \label{eq:deriv_lccf} \lccf{\ell}{u} = \frac{\ell \cdot \lvert K_{\ell+1}(u) \rvert}{(d_u - \ell + 1) \cdot \lvert K_{\ell}(u) \rvert }. \end{equation} From Eq.~\ref{eq:deriv_lccf}, it is easy to see that we can compute all local $\ell$th-order clustering coefficients by enumerating all $(\ell + 1)$-cliques and $\ell$-cliques in the graph. The computational complexity of the algorithm is thus bounded by the time to enumerate $(\ell + 1)$-cliques and $\ell$-cliques. Using the Chiba and Nishizeki algorithm~\cite{chiba1985arboricity}, the complexity is $O(\ell a^{\ell-2}m)$, where $a$ is the arboricity of the graph, and $m$ is the number of edges. The arboricity $a$ may be as large as $\sqrt{m}$, so this algorithm is only guaranteed to take polynomial time if $\ell$ is a constant. In general, determining if there exists a single clique with at least $\ell$ nodes is NP-complete~\cite{karp1972reducibility}. For the global clustering coefficient, note that \begin{equation} \lvert W_{\ell} \rvert = \sum_{u \in V} \lvert W_\ell(u) \rvert. \end{equation} Thus, it suffices to enumerate $\ell$-cliques (to compute $\lvert W_{\ell} \rvert$ using Eq.~\ref{eq:wedge_identity}) and to count the total number of $\ell$-cliques. In practice, we use the Chiba and Nishizeki to enumerate cliques and simultaneously compute $\gccf{\ell}$ and $\lccf{\ell}{u}$ for all nodes $u$. This suffices for our clustering analysis with $\ell = 2, 3, 4$ on networks with over a hundred million edges in Section~\ref{sec:empirical}. \subsection{Probabilistic interpretations of higher-order clustering coefficients} To facilitate understanding of higher-order clustering coefficients and to aid our analysis in Section~\ref{sec:theoretical}, we present a few probabilistic interpretations of the quantities. First, we can interpret $\lccf{\ell}{u}$ as the probability that a wedge $w$ chosen uniformly at random from all wedges centered at $u$ is closed: \begin{equation} \label{eq:prob_interp1} \lccf{\ell}{u} = \probof{w \in K_{\ell + 1}(u)}. \end{equation} The variant of this interpretation for the classical clustering case of $\ell=2$ has been useful for graph algorithm development~\cite{seshadhri2013triadic}. For the next probabilistic interpretation, it is useful to analyze the structure of the 1-hop neighborhood graph $\nbr{1}{u}$ of a given node $u$ (not containing node $u$). The vertex set of $\nbr{1}{u}$ is the set of all nodes adjacent to $u$, and the edge set consists of all edges between neighbors of $u$, i.e., $ \{ (v, w) \;\vert\; (u, v), (u, w), (v, w) \in E \}$, where $E$ is the edge set of the graph. Any $\ell$-clique in $G$ containing node $u$ corresponds to a unique $(\ell - 1)$-clique in $\nbr{1}{u}$, and specifically for $\ell = 2$, any edge $(u, v)$ corresponds to a node $v$ in $\nbr{1}{u}$. Therefore, each $\ell$-wedge centered at $u$ corresponds to an $(\ell-1)$-clique $K$ and one of the $d_u - \ell + 1$ nodes outside $K$ (i.e., in $\nbr{1}{u} \backslash K$). Thus, Eq.~\ref{eq:deriv_lccf} can be re-written as \begin{equation}\label{eq:lccf_nbrs} \frac{\ell \cdot \lvert K_{\ell}(\nbr{1}{u}) \rvert}{(d_u - \ell + 1) \cdot \lvert K_{\ell-1}(\nbr{1}{u}) \rvert }, \end{equation} where $K_{k}(\nbr{1}{u})$ denotes the number of $k$-cliques in $\nbr{1}{u}$. If we uniformly at random select an $(\ell - 1)$-clique $K$ from $\nbr{1}{u}$ and then also uniformly at random select a node $v$ from $\nbr{1}{u}$ outside of this clique, then $\lccf{\ell}{u}$ is the probability that these $\ell$ nodes form an $\ell$-clique: \begin{equation} \label{eq:prob_interp2} \lccf{\ell}{u} = \probof{K \cup \{ v \} \in K_{\ell}(\nbr{1}{u})}. \end{equation} Moreover, if we condition on observing an $\ell$-clique from this sampling procedure, then the $\ell$-clique itself is selected uniformly at random from all $\ell$-cliques in $\nbr{1}{u}$. Therefore, $\lccf{\ell-1}{u} \cdot \lccf{\ell}{u}$ is the probability that an $(\ell - 1)$-clique and two nodes selected uniformly at random from $\nbr{1}{u}$ form an $(\ell + 1)$-clique. Applying this recursively gives \begin{equation} \prod_{j=2}^{\ell}\lccf{j}{u} = \frac{\lvert K_{\ell}(\nbr{1}{u}) \rvert}{{d_u \choose \ell}}. \end{equation} In other words, the product of the higher-order local clustering coefficients of node $u$ up to order $\ell$ is the $\ell$-clique density amongst $u$'s neighbors. \section{Discussion} We have proposed higher-order clustering coefficients to study higher-order closure patterns in networks, which generalizes the widely used clustering coefficient that measures triadic closure. Our work compliments other recent developments on the importance of higher-order information in network navigation~\cite{rosvall2014memory,scholtes2017network} and on temporal community structure~\cite{sekara2016fundamental}; in contrast, we examine higher-order clique closure and only implicitly consider time as a motivation for closure. Prior efforts in generalizing clustering coefficients have focused on shortest paths~\cite{fronczak2002higher}, cycle formation~\cite{caldarelli2004structure}, and triangle frequency in $k$-hop neighborhoods~\cite{andrade2006neighborhood,jiang2004topological}. Such approaches fail to capture closure patterns of cliques, suffer from challenging computational issues, and are difficult to theoretically analyze in random graph models more sophisticated than the Erd\H{o}s-R\'enyi model. On the other hand, our higher-order clustering coefficients are simple but effective measurements that are analyzable and easily computable (we only rely clique enumeration, a well-studied algorithmic task). Furthermore, our methodology provides new insights into the clustering behavior of several real-world networks and random graph models, and our theoretical analysis provides intuition for the way in which higher-order clustering coefficients describe local clustering in graphs. Finally, we focused on higher-order clustering coefficients as a global network measurement and as a node-level measurement, and in related work we also show that large higher-order clustering implies the existence of mesoscale clique-dense community structure~\cite{yin2017local}. \section{Experimental results on real-world networks}\label{sec:empirical} We now analyze the higher-order clustering of real-world networks. We first study how the higher-order global and average clustering coefficients vary as we increase the order $\ell$ of the clustering coefficient on a collection of 20 networks from several domains. After, we concentrate on a few representative networks and compare the higher-order clustering of real-world networks to null models. We find that only some networks exhibit higher-order clustering once the traditional clustering coefficient is controlled. Finally, we examine the local clustering of real-world networks. \subsection{Higher-order global and average clustering} \input{macroscopic-ccfs-table.tex} We compute and analyze the higher-order clustering for networks from a variety of domains (Table~\ref{tab:all_ccfs}). We briefly describe the collection of networks and their categorization below: \begin{enumerate} \item Two synthetic networks---a random instance of an $\textnormal{Erd\H{o}s-R\'enyi}$ graph with $n=1,000$ nodes and edge probability $p=0.2$ and a small-world network with $n=20,000$ nodes, $k = 10$, and rewiring probability $p=0.1$; \item Four neural networks---the complete neural systems of the nematode worms \emph{P.\ pacificus} and \emph{C.\ elegans} as well as the neural connections of the Drosophila medulla and mouse retina; \item Four online social networks---two Facebook friendship networks between students at universities from 2005 (fb-Stanford, fb-Cornell) and two complete online friendship networks (Pokec and Orkut); \item Four collaboration networks---two co-authorship networks constructed from arxiv submission categories (arxiv-AstroPh and arxiv-HepPh), a co-authorship network constructed from DBLP, and the co-committee membership network of United States congresspersons (congress-committees); \item Four human communication networks---two email networks (email-Enron-core, email-Eu-core), a Facebook-like messaging network from a college (CollegeMsg), and the edits of user talk pages by other users on Wikipedia (wiki-Talk); and \item Four technological systems networks---three autonomous systems (oregon2-010526, as-caida-20071105, as-skitter) and a peer-to-peer connection network (p2p-Genutella31). \end{enumerate} In all cases, we take the edges as undirected, even if the original network data is directed. Table~\ref{tab:all_ccfs} lists the $\ell$th-order global and average clustering coefficients for $\ell=2,3,4$ as well as the fraction of nodes that are the center of at least one $\ell$-wedge (recall that the average clustering coefficient is the mean only over higher-order local clustering coefficients of nodes participating in at least one $\ell$-wedge; see \citet{kaiser2008mean} for a discussion on how this can affect network analyses). We highlight some important trends in the raw clustering coefficients, and in the next section, we focus on higher-order clustering compared to what one gets in a null model. Propositions~\ref{prop:ccf_er} and \ref{prop:ccf_sw} say that we should expect the higher-order global and average clustering coefficients to decrease as we increase the order $\ell$ for both the $\textnormal{Erd\H{o}s-R\'enyi}$ and small-world models, and indeed $\accf{2} > \accf{3} > \accf{4}$ for these networks. This trend also holds for most of the real-world networks (mouse-retina, congress-committees, and oregon2-010526 are the exceptions). Thus, when averaging over nodes, higher-order cliques are overall less likely to close in both the synthetic and real-world networks. \input{ccfs-table} The relationship between the higher-ordrer global clustering coefficient $\gccf{\ell}$ and the order $\ell$ is less uniform over the datasets. For the three co-authorship networks (arxiv-HepPh, arxiv-AstroPh, and DBLP) and the three autonomous systems networks (oregon2-010526, as-caida-20071105, and as-skitter), $\gccf{\ell}$ increases with $\ell$, although the base clustering levels are much higher for co-authorship networks. This is not simply due to the presence of cliques---a clique has the same clustering for any order (Fig.~\ref{fig:ccf_diffs}, left). Instead, these datasets have nodes that serve as the center of a star and also participate in a clique (Fig.~\ref{fig:ccf_diffs}, right; see also Proposition~\ref{prop:ccf_bounds}). On the other hand, $\gccf{\ell}$ decreases with $\ell$ for the two email networks and the two nematode worm neural networks. Finally, the change in $\gccf{\ell}$ need not be monotonic in $\ell$. In three of the four online social networks, $\gccf{3} < \gccf{2}$ but $\gccf{4} > \gccf{3}$. Overall, the trends in the higher-order clustering coefficients can be different within one of our dataset categories, but tend to be uniform within sub-categories: the change of $\accf{\ell}$ and $\gccf{\ell}$ with $\ell$ is the same for the two nematode worms within the neural networks, the two email networks within the communication networks, and the three co-authorship networks within the collaboration networks. These trends hold even if the (classical) second-order clustering coefficients differ substantially in absolute value. While the raw clustering values are informative, it is also useful to compare the clustering to what one expects from null models. We find in the next section that this reveals additional insights into our data. \subsection{Comparison against null models} For one real-world network from each dataset category, we also measure the higher-order clustering coefficients with respect to two null models (Table~\ref{tab:data_summary}). First, we compare against the Configuration Model (CM) that samples uniformly from simple graphs with the same degree distribution~\cite{bollobas1980probabilistic,milo2003uniform}. In real-world networks, $\accf{2}$ is much larger than expected with respect to the CM null model. We find that the same holds for $\accf{3}$. Second, we use a null model that samples graphs preserving both degree distribution and $\accf{2}$. Specifically, these are samples from an ensemble of exponential graphs where the Hamiltonian measures the absolute value of the difference between the original network and the sampled network~\cite{park2004statistical}. Such samples are referred to as as Maximally Random Clustered Networks (MRCN) and are sampled with a simulated annealing procedure~\cite{colomer2013deciphering}. Comparing $\accf{3}$ between the real-world and the null network, we observe different behavior in higher-order clustering across our datasets. Compared to the MRCN null model, \emph{C. elegans} has significantly less than expected higher-order clustering (in terms of $\accf{3}$), the Facebook friendship and autonomous system networks have significantly more than expected higher-order clustering, and the co-authorship and email networks have slightly (but not significantly) more than expected higher-order clustering (Table~\ref{tab:data_summary}). Put another way, all real-world networks exhibit clustering in the classical sense of triadic closure. However, the higher-order clustering coefficients reveal that the friendship and autonomous systems networks exhibit significant clustering beyond what is given by triadic closure. These results suggest the need for models that directly account for closure in node neighborhoods~\cite{bhat2016densification,lambiotte2016structural}. Our finding about the lack of higher-order clustering in \emph{C. elegans} agrees with previous results that 4-cliques are under-expressed, while open 3-wedges related to cooperative information propagation are over-expressed~\cite{benson2016higher,milo2002network,varshney2011structural}. This also provides credence for the ``3-layer'' model of \emph{C. elegans}~\cite{varshney2011structural}. The observed clustering in the friendship network is consistent with prior work showing the relative infrequency of open $\ell$-wedges in many Facebook network subgraphs with respect to a null model accounting for triadic closure~\cite{ugander2013subgraph}. Co-authorship networks and email networks are both constructed from ``events'' that create multiple edges---a paper with $k$ authors induces a $k$-clique in the co-authorship graph and an email sent from one address to $k$ others induces $k$ edges. This event-driven graph construction creates enough closure structure so that the average third-order clustering coefficient is not much larger than random graphs where the classical second-order clustering coefficient and degree sequence is kept the same. \begin{figure*}[tb] \phantomsubfigure{fig:ccfsA} \phantomsubfigure{fig:ccfsB} \phantomsubfigure{fig:ccfsC} \phantomsubfigure{fig:ccfsD} \phantomsubfigure{fig:ccfsE} \includegraphics[width=2\columnwidth]{local.eps} \caption{Top row: Joint distributions of ($\lccf{2}{u}$, $\lccf{3}{u}$) for (A) \emph{C. elegans} (B) Facebook friendship, (C) arxiv co-authorship, (D) email, and (E) autonomous systems networks. Each blue dot represents a node, and the red curve tracks the average over logarithmic bins. The upper trend line is the bound in Eq.~\ref{eq:Bound_Kappa3}, and the lower trend line is expected Erd\H{o}s-R\'enyi behavior from Proposition~\ref{prop:ccf_er_cond}. Bottom row: Average higher-order clustering coefficients as a function of degree. }\label{fig:ccfs} \end{figure*} \begin{figure}[tb] \phantomsubfigure{fig:ccfs_nullA} \phantomsubfigure{fig:ccfs_nullB} \includegraphics[width=0.9\columnwidth]{local-null.eps} \caption{Analogous plots of Fig.~\ref{fig:ccfs} for synthetic (A) $\textnormal{Erd\H{o}s-R\'enyi}$ and (B) small-world networks. Top row: Joint distributions of ($\lccf{2}{u}$, $\lccf{3}{u}$). Bottom row: Average higher-order clustering coefficients as a function of degree. }\label{fig:ccfs_null} \end{figure} We emphasize that simple clique counts are not sufficient to obtain these results. For example, the discrepancy in the third-order average clustering of \emph{C. elegans} and the MRCN null model is not simply due to the presence of 4-cliques. The original neural network has nearly twice as many 4-cliques (2,010) than the samples from the MRCN model (mean 1006.2, standard deviation 73.6), but the third-order clustering coefficient is larger in MRCN. The reason is that clustering coefficients normalize clique counts with respect to opportunities for closure. Thus far, we have analyzed global and average higher-order clustering, which both summarize the clustering of the entire network. In the next section, we look at more localized properties, namely the distribution of higher-order local clustering coefficients and the higher-order average clustering coefficient as a function of node degree. \subsection{Higher-order local clustering coefficients and degree dependencies} We now examine more localized clustering properties of our networks. Figure~\ref{fig:ccfs} (top) plots the joint distribution of $\lccf{2}{u}$ and $\lccf{3}{u}$ for the five networks analyzed in Table~\ref{tab:data_summary}, and Fig.~\ref{fig:ccfs_null} (top) provides the analogous plots for the $\textnormal{Erd\H{o}s-R\'enyi}$ and small-world networks. In these plots, the lower dashed trend line represents the expected $\textnormal{Erd\H{o}s-R\'enyi}$ behavior, i.e., the expected clustering if the edges in the neighborhood of a node were configured randomly, as formalized in Proposition~\ref{prop:ccf_er_cond}. The upper dashed trend line is the maximum possible value of $\lccf{3}{u}$ given $\lccf{2}{u}$, as given by Proposition~\ref{prop:ccf_bounds}. For many nodes in \emph{C. elegans}, local clustering is nearly random (Fig.~\ref{fig:ccfsA}, top), i.e., resembles the $\textnormal{Erd\H{o}s-R\'enyi}$ joint distribution (Fig.~\ref{fig:ccfs_nullA}, top). In other words, there are many nodes that lie on the lower trend line. This provides further evidence that \emph{ C. elegans} lacks higher-order clustering. In the arxiv co-authorship network, there are many nodes $u$ with a large value of $\lccf{2}{u}$ that have an even larger value of $\lccf{3}{u}$ near the upper bound of Eq.~\ref{eq:Bound_Kappa3} (see the inset of Fig.~\ref{fig:ccfsC}, top). This implies that some nodes appear in both cliques and also as the center of star-like patterns, as in Fig.~\ref{fig:ccf_diffs}. On the other hand, only a handful of nodes in the Facebook friendships, Enron email, and Oregon autonomous systems networks are close to the upper bound (insets of Figs.~\ref{fig:ccfsB},\ref{fig:ccfsD}, and \ref{fig:ccfsE}, top). Figures~\ref{fig:ccfs} and \ref{fig:ccfs_null} (bottom) plot higher-order average clustering as a function of node degree in the real-world and synthetic networks. In the $\textnormal{Erd\H{o}s-R\'enyi}$, small-world, \emph{C. elegans}, and Enron email networks, there is a distinct gap between the average higher-order clustering coefficients for nodes of all degrees. Thus, our previous finding that the average clustering coefficient $\accf{\ell}$ decreases with $\ell$ in these networks is independent of degree. In the Facebook friendship network, $\lccf{2}{u}$ is larger than $\lccf{3}{u}$ and $\lccf{4}{u}$ on average for nodes of all degrees, but $\lccf{3}{u}$ and $\lccf{4}{u}$ are roughly the same for nodes of all degrees, which means that 4-cliques and 5-cliques close at roughly the same rate, independent of degree, albeit at a smaller rate than traditional triadic closure (Fig.~\ref{fig:ccfsB}, bottom). In the co-authorship network, nodes $u$ have roughly the same $\lccf{\ell}{u}$ for $\ell = 2$, $3$, $4$, which means that $\ell$-cliques close at about the same rate, independent of $\ell$ (Fig.~\ref{fig:ccfsC}, bottom). In the Oregon autonomous systems network, we see that, on average, $\lccf{4}{u} > \lccf{3}{u} > \lccf{2}{u}$ for nodes with large degree (Fig.~\ref{fig:ccfsE}, bottom). This explains how the global clustering coefficient increases with the order, but the average clustering does not, as observed in Table~\ref{tab:all_ccfs}. \section{Introduction} Networks are a fundamental tool for understanding and modeling complex physical, social, informational, and biological systems~\cite{newman2003structure}. Although such networks are typically sparse, a recurring trait of networks throughout all of these domains is the tendency of edges to appear in small clusters or cliques~\cite{rapoport1953spread,watts1998collective}. In many cases, such clustering can be explained by local evolutionary processes. For example, in social networks, clusters appear due to the formation of triangles where two individuals who share a common friend are more likely to become friends themselves, a process known as \emph{triadic closure}~\cite{rapoport1953spread,granovetter1973strength}. Similar triadic closures occur in other networks: in citation networks, two references appearing in the same publication are more likely to be on the same topic and hence more likely to cite each other~\cite{wu2009modeling} and in co-authorship networks, scientists with a mutual collaborator are more likely to collaborate in the future~\cite{jin2001structure}. In other cases, local clustering arises from highly connected functional units operating within a larger system, e.g., metabolic networks are organized by densely connected modules~\cite{ravasz2003hierarchical}. The \emph{clustering coefficient} quantifies the extent to which edges of a network cluster in terms of triangles. The clustering coefficient is defined as the fraction of length-2 paths, or \emph{wedges}, that are closed with a triangle~\cite{watts1998collective,barrat2000properties} (Fig.~\ref{fig:ccf_def}, row $\gccf{2}$). In other words, the clustering coefficient measures the probability of triadic closure in the network. However, the clustering coefficient is inherently restrictive as it measures the closure probability of just one simple structure---the triangle. Moreover, higher-order structures such as larger cliques are crucial to the structure and function of complex networks~\cite{benson2016higher,yaverouglu2014revealing,rosvall2014memory}. For example, 4-cliques reveal community structure in word association and protein-protein interaction networks~\cite{palla2005uncovering} and cliques of sizes 5--7 are more frequent than triangles in many real-world networks with respect to certain null models~\cite{slater2014mid}. However, the extent of clustering of such higher-order structures has not been well understood nor quantified. \begin{figure}[tb] \begin{tabular}{l c c c} & 1.~Start with~ & 2.~Find an adjacent edge & 3.~Check for an \\ & an $\ell$-clique & to form an $\ell$-wedge & $(\ell+1)$-clique \\ \\ $\gccf{2}$ & \tpone{intro2_part1} & \tptwo{intro2_part1}{intro2_part2} & \tpthree{intro2_part1}{intro2_part2}{intro2_part3} \\ \\ $\gccf{3}$ & \tpone{intro3_part1} & \tptwo{intro3_part1}{intro3_part2} & \tpthree{intro3_part1}{intro3_part2}{intro3_part3} \\ \\ $\gccf{4}$ & \tpone{intro4_part1} & \tptwo{intro4_part1}{intro4_part2} & \tpthree{intro4_part1}{intro4_part2}{intro4_part3} \\ \end{tabular} \caption{Overview of higher-order clustering coefficients as clique expansion probabilities. The $\ell$th-order clustering coefficient $\gccf{\ell}$ measures the probability that an $\ell$-clique and an adjacent edge, i.e., an $\ell$-wedge, is closed, meaning that the $\ell - 1$ possible edges between the $\ell$-clique and the outside node in the adjacent edge exist to form an $(\ell + 1)$-clique.} \label{fig:ccf_def} \end{figure} Here, we provide a framework to quantify higher-order clustering in networks by measuring the normalized frequency at which higher-order cliques are closed, which we call \emph{higher-order clustering coefficients}. We derive our higher-order clustering coefficients by extending a novel interpretation of the classical clustering coefficient as a form of clique expansion (Fig.~\ref{fig:ccf_def}). We then derive several properties about higher-order clustering coefficients and analyze them under the $G_{n,p}$ and small-world null models. Using our theoretical analysis as a guide, we analyze the higher-order clustering behavior of real-world networks from a variety of domains. We find that each domain of networks has its own higher-order clustering pattern, which the traditional clustering coefficient does not show on its own. Conventional wisdom in network science posits that practically all real-world networks exhibit clustering; however, we find that not all networks exhibit higher-order clustering. More specifically, once we control for the clustering as measured by the classical clustering coefficient, some networks do not show significant clustering in terms of higher-order cliques. In addition to the theoretical properties and empirical findings exhibited in this paper, our related work also demonstrates a connection between higher-order clustering and community detection~\cite{yin2017local}. \section{} \end{document} \section{Theoretical analysis and higher-order clustering in random graph models}\label{sec:theoretical} We now provide some theoretical analysis of our higher-order clustering coefficients. We first give some extremal bounds on the values that higher-order clustering coefficients can take given the value of the traditional (second-order) clustering coefficient. After, we analyze the values of higher-order clustering coefficients in two common random graph models---the $G_{n,p}$ and small-world models. The theory from this section will be a useful guide for interpreting the clustering behavior of real-world networks in Section~\ref{sec:empirical}. \subsection{Extremal bounds} \begin{figure}[tb] \begin{tabular}{l @{\hskip 12pt} c @{\hskip 12pt} c @{\hskip 12pt} c} & \tpone{ccf_equal} & \tpone{ccf_lower} & \tpone{ccf_upper} \\ \\ $\lccf{2}{u}$ & $1$ & $\dfrac{d}{2(d - 1)} \approx \dfrac{1}{2}$ & $\frac{d - 2}{4d - 4} \approx \dfrac{1}{4}$ \\ \\ $\lccf{3}{u}$ & $1$ & $0$ & $\dfrac{d - 4}{2d - 4} \approx \dfrac{1}{2}$ \\ \\ $\lccf{4}{u}$ & $1$ & $0$ & $\dfrac{d - 6}{2d - 6} \approx \dfrac{1}{2}$ \end{tabular} \caption{Example 1-hop neighborhoods of a node $u$ with degree $d$ with different higher-order clustering. Left: For cliques, $\lccf{\ell}{u} = 1$ for any $\ell$. Middle: If $u$'s neighbors form a complete bipartite graph, $\lccf{2}{u}$ is constant while $\lccf{\ell}{u} = 0$, $\ell \ge 3$. Right: If half of $u$'s neighbors form a star and half form a clique with $u$, then $\lccf{\ell}{u} \approx \sqrt{\lccf{2}{u}}$, which is the upper bound in Proposition~\ref{prop:ccf_bounds}.} \label{fig:ccf_diffs} \end{figure} We first analyze the relationships between local higher-order clustering coefficients of different orders. Our technical result is Proposition~\ref{prop:ccf_bounds}, which provides essentially tight lower and upper bounds for higher-order local clustering coefficients in terms of the traditional local clustering coefficient. The main ideas of the proof are illustrated in Fig.~\ref{fig:ccf_diffs}. \begin{proposition}\label{prop:ccf_bounds} For any fixed $\ell \geq 3$, \begin{equation} \label{eq:Bound_Kappa3} 0 \leq \lccf{\ell}{u} \leq \sqrt{\lccf{2}{u}}. \end{equation} Moreover, \begin{enumerate} \item There exists a finite graph $G$ with a node $u$ such that the lower bound is tight and $\lccf{2}{u}$ is within $\epsilon$ of any prescribed value in $[0, \frac{\ell-2}{\ell-1}]$. \item There exists a finite graph $G$ with a node $u$ such that $\lccf{\ell}{u}$ is within $\epsilon$ of the upper bound for any prescribed value of $\lccf{2}{u} \in [0, 1]$. \end{enumerate} \end{proposition} \begin{proof} Clearly, $0 \leq \lccf{\ell}{u}$ if the local clustering coefficient is well defined. This bound is tight when $\nbr{1}{u}$ is $(\ell - 1)$-partite, as in the middle column of Fig.~\ref{fig:ccf_diffs}. In the $(\ell - 1)$-partite case, $\lccf{2}{u} = \frac{\ell-2}{\ell-1}$. By removing edges from this extremal case in a sufficiently large graph, we can make $\lccf{2}{u}$ arbitrarily close to any value in $[0, \frac{\ell-2}{\ell-1}]$. To derive the upper bound, consider the 1-hop neighborhood $\nbr{1}{u}$, and let \begin{equation} \delta_\ell(\nbr{1}{u}) = \frac{\lvert K_{\ell}(\nbr{1}{u}) \rvert}{{d_u \choose \ell}} \end{equation} denote the $\ell$-clique density of $\nbr{1}{u}$. The Kruskal-Katona theorem~\cite{kruskal1963number,katona1966theorem} implies that \begin{align*} &\delta_{\ell}(\nbr{1}{u}) \leq [\delta_{\ell - 1}(\nbr{1}{u})]^{\ell / (\ell - 1)} \\ &\delta_{\ell - 1}(\nbr{1}{u}) \leq [\delta_{2}(\nbr{1}{u})]^{(\ell - 1) / 2}. \end{align*} Combining this with Eq.~\ref{eq:deriv_lccf} gives \begin{align*} \lccf{\ell}{u} &\leq [\delta_{\ell - 1}(\nbr{1}{u})]^{\frac{1}{\ell - 1}} \leq \sqrt{\delta_{2}(\nbr{1}{u})} = \sqrt{\lccf{2}{u}}, \end{align*} where the last equality uses the fact that $\lccf{2}{u}$ is the edge density of $\nbr{1}{u}$. The upper bound becomes tight when $\nbr{1}{u}$ consists of a clique and isolated nodes (Fig.~\ref{fig:ccf_diffs}, right) and the neighborhood is sufficiently large. Specifically, let $\nbr{1}{u}$ consist of a clique of size $c$ and $b$ isolated nodes. When $\ell = 2$, \begin{equation*} \lccf{\ell}{u} = \frac{{c \choose 2}}{{c + b \choose 2}} = \frac{(c - 1)c}{(c + b - 1)(c + b)} \to \left(\frac{c}{c + b}\right)^2 \end{equation*} and by Eq.~\ref{eq:lccf_nbrs}, when $3 \le \ell \le c$, \begin{align*} \lccf{\ell}{u} &= \frac{\ell \cdot {c \choose \ell}}{(c + b - \ell + 1) \cdot {c \choose \ell - 1}} = \frac{c - \ell + 1}{c + b - \ell + 1} \to \frac{c}{c + b}. \end{align*} By adjusting the ratio $c / (b + c)$ in $\nbr{1}{u}$, we can construct a family of graphs such that $\lccf{2}{u}$ takes any value in the interval $[0,1]$ as $d_u \to \infty$ and $\lccf{\ell}{u} \to \sqrt{\lccf{2}{u}}$ as $d_u \to \infty$. \end{proof} The second part of the result requires the neighborhoods to be sufficiently large in order to reach the upper bound. However, we will see later that in some real-world data, there are nodes $u$ for which $\lccf{3}{u}$ is close to the upper bound $\sqrt{\lccf{2}{u}}$ for several values of $\lccf{2}{u}$. Next, we analyze higher-order clustering coefficients in two common random graph models: the Erd\H{o}s-R\'enyi model with edge probability $p$ (i.e., the $G_{n,p}$ model~\cite{erdos1959random}) and the small-world model~\cite{watts1998collective}. \subsection{Analysis for the $G_{n,p}$ model} Now, we analyze higher-order clustering coefficients in classical Erd\H{o}s-R\'enyi random graph model, where each edge exists independently with probability $p$ (i.e., the $G_{n,p}$ model~\cite{erdos1959random}). We implicitly assume that $\ell$ is small in the following analysis so that there should be at least one $\ell$-wedge in the graph (with high probability and $n$ large, there is no clique of size greater than $(2 + \epsilon)\log n / \log (1 / p)$ for any $\epsilon > 0$~\cite{bollobas1976cliques}). Therefore, the global and local clustering coefficients are well-defined. In the $G_{n, p}$ model, we first observe that any $\ell$-wedge is closed if and only if the $\ell - 1$ possible edges between the $\ell$-clique and the outside node in the adjacent edge exist to form an $(\ell+1)$-clique. Each of the $\ell - 1$ edges exist independently with probability $p$ in the $G_{n, p}$ model, which means that the higher-order clustering coefficients should scale as $p^{\ell - 1}$. We formalize this in the following proposition. \begin{proposition}\label{prop:ccf_er} Let $G$ be a random graph drawn from the $G_{n, p}$ model. For constant $\ell$, \begin{enumerate} \item $\expectover{G}{\gccf{\ell}} = p^{\ell - 1}$ \item $\expectover{G}{\lccf{\ell}{u} \;\vert\; W_{\ell}(u) > 0} = p^{\ell - 1}$ for any node $u$ \item $\expectover{G}{\accf{\ell}} = p^{\ell - 1}$ \end{enumerate} \end{proposition} \begin{proof} We prove the first part by conditioning on the set of $\ell$-wedges, $W_{\ell}$: \begin{align*} \expect{\gccf{\ell}} &=\textstyle \expectover{G}{\expectover{W_{\ell}}{\gccf{\ell} \;\vert\; W_{\ell}}} \\ &=\textstyle \expectover{G}{\expectover{W_{\ell}}{\frac{1}{\lvert W_{\ell}\rvert}\sum_{w \in W_{\ell}}\probof{w \text{ is closed}}}} \\ &=\textstyle \expectover{G}{\expectover{W_{\ell}}{\frac{1}{\lvert W_{\ell}\rvert}\sum_{w \in W_{\ell}}p^{\ell - 1}}} \\ &=\textstyle \expectover{G}{p^{\ell - 1}} \\ &=\textstyle p^{\ell - 1}. \end{align*} As noted above, the second equality is well defined (with high probability) for small $\ell$. The third equality comes from the fact that any $\ell$-wedge is closed if and only if the $\ell - 1$ possible edges between the $\ell$-clique and the outside node in the adjacent edge exist to form an $(\ell+1)$-clique. The proof of the second part is essentially the same, except we condition over the set of possible cases where $W_{\ell}(u) > 0$. Recall that $\widetilde V$ is the set of nodes at the center of at least one $\ell$-wedge. To prove the third part, we take the conditional expectation over $\widetilde V$ and use our result from the second part. \end{proof} The above results say that the global, local, and average $\ell$th order clustering coefficients decrease exponentially in $\ell$. It turns out that if we also condition on the second-order clustering coefficient having some fixed value, then the higher-order clustering coefficients still decay exponentially in $\ell$ for the $G_{n,p}$ model. This will be useful for interpreting the distribution of local clustering coefficients on real-world networks. \begin{proposition}\label{prop:ccf_er_cond} Let $G$ be a random graph drawn from the $G_{n, p}$ model. Then for constant $\ell$, \begin{align*} & \expectover{G}{\lccf{\ell}{u} \;\vert\; \lccf{2}{u}, W_{\ell}(u) > 0} \\ &= \left[\lccf{2}{u} - (1 - \lccf{2}{u}) \cdot O(1 / d_u^2)\right]^{\ell - 1} \approx (\lccf{2}{u})^{\ell - 1}. \end{align*} \end{proposition} \begin{proof} Similar to the proof of Proposition~\ref{prop:ccf_er_cond}, we look at the conditional expectation over $W_{\ell}(u) > 0$: \begin{align*} &\textstyle \expectover{G}{\lccf{\ell}{u} \;\vert\; \lccf{2}{u}, W_{\ell}(u) > 0} \\ &=\textstyle \expectover{G}{\expectover{W_{\ell}(u) > 0}{\lccf{\ell}{u} \;\vert\; \lccf{2}{u},\; W_{\ell}(u)}} \\ &=\textstyle \expectover{G}{\expectover{W_{\ell}(u) > 0}{\frac{1}{\lvert W_{\ell}(u)\rvert}\sum_{w \in W_{\ell}(u)}\probof{w \text{ closed} \;\vert\; \lccf{2}{u}}}}. \end{align*} Now, note that $\nbr{1}{u}$ has $m = \lccf{2}{u} \cdot {d_u \choose 2}$ edges. Knowing that $w \in W_{\ell}(u)$ accounts for ${\ell - 1 \choose 2}$ of these edges. By symmetry, the other $q = m - {\ell - 1 \choose 2}$ edges appear in any of the remaining $r = {d_u \choose 2} - {\ell - 1 \choose 2}$ pairs of nodes uniformly at random. There are ${r \choose q}$ ways to place these edges, of which ${r - \ell + 1 \choose q - \ell + 1}$ would close the wedge $w$. Thus, \begin{align*} &\probof{w \text{ is closed} \;\vert\; \lccf{2}{u}} \\ &=\textstyle \frac{{r - \ell + 1 \choose q - \ell + 1}}{{r \choose q}} = \frac{(r - \ell + 1)!q!}{(q - \ell + 1)!r!} = \frac{(q - \ell + 2)(q - \ell + 3) \cdots q}{(r - \ell + 2)(r - \ell + 3) \cdots r}. \end{align*} Now, for any small nonnegative integer $k$, \begin{align*} \frac{q - k}{r - k} &=\textstyle \frac{ \lccf{2}{u} \cdot {d_u \choose 2} - {\ell -1 \choose 2} - k }{ {d_u \choose 2} - {\ell - 1 \choose 2} - k } \\ &=\textstyle \lccf{2}{u} - (1 - \lccf{2}{u})\left[\frac{{\ell -1 \choose 2} + k}{{d_u \choose 2} - {\ell - 1 \choose 2} - k}\right] \\ &=\textstyle \lccf{2}{u} - (1 - \lccf{2}{u}) \cdot O(1 / d_u^2). \end{align*} (Recall that $\ell$ is constant by assumption, so the big-O notation is appropriate). The above expression approaches $(\lccf{2}{u})^{\ell - 1}$ when $\lccf{2}{u} \to 1$ as well as when $d_u \to \infty$. \end{proof} Proposition~\ref{prop:ccf_er_cond} says that even if the second-order local clustering coefficient is large, the $\ell$th-order clustering coefficient will still decay exponentially in $\ell$, at least in the limit as $d_u$ grows large. By examining higher-order clique closures, this allows us to distinguish between nodes $u$ whose neighborhoods are ``dense but random" ($\lccf{2}{u}$ is large but $\lccf{\ell}{u} \approx (\lccf{2}{u})^{\ell - 1}$) or ``dense and structured" ($\lccf{2}{u}$ is large \emph{and} $\lccf{\ell}{u} > (\lccf{2}{u})^{\ell - 1}$). Only the latter case exhibits higher-order clustering. We use this in our analysis of real-world networks in Section~\ref{sec:empirical}. \subsection{Analysis for the small-world model} We also study higher-order clustering in the small-world random graph model~\cite{watts1998collective}. The model begins with a ring network where each node connects to its $2k$ nearest neighbors. Then, for each node $u$ and each of the $k$ edges $(u, v)$ with $v$ following $u$ clockwise in the ring, the edge is rewired to $(u, w)$ with probability $p$, where $w$ is chosen uniformly at random. \begin{figure}[tb] \begin{centering} \includegraphics[width=1.0\columnwidth]{smallworld-ccs-updated} \end{centering} \caption{Average higher-order clustering coefficient $\accf{\ell}$ as a function of rewiring probability $p$ in small-world networks for $\ell = 2, 3, 4$ ($n = 20,000$, $k = 5$). Proposition~\ref{prop:ccf_sw} shows that the $\ell$th-order clustering coefficient when $p = 0$ predicts that the clustering should decrease modestly as $\ell$ increases.} \label{fig:sw_ccfs} \end{figure} With no rewiring ($p = 0$) and $k \ll n$, it is known that $\accf{2} \approx 3/4$~\cite{watts1998collective}. As $p$ increases, the average clustering coefficient $\accf{2}$ slightly decreases until a phase transition near $p = 0.1$, where $\accf{2}$ decays to $0$~\cite{watts1998collective} (also see Fig.~\ref{fig:sw_ccfs}). Here, we generalize these results for higher-order clustering coefficients. \begin{proposition}\label{prop:ccf_sw} In the small-world model without rewiring ($p = 0$), \begin{equation* \accf{\ell} \rightarrow (\ell + 1) / (2\ell) \end{equation*} for any constant $\ell \geq 2$ as $k \to \infty$ and $n \to \infty$ while $2k < n$. \end{proposition} \begin{proof} Applying Eq.~\ref{eq:deriv_lccf}, it suffices to show that \begin{equation} \label{eq:K_SW} \lvert K_{\ell}(u) \rvert = \frac{\ell}{(\ell - 1)!}\cdot k^{\ell -1 } + O(k^{\ell - 2}) \end{equation} as \begin{align*} \lccf{\ell}{u} = \frac{\ell \cdot \frac{(\ell + 1) k^{\ell}}{\ell!}}{(2k - \ell + 1) \cdot \frac{\ell k^{\ell -1 }}{(\ell - 1)!} }, \end{align*} which approaches $\frac{\ell + 1}{2\ell}$ as $k \to \infty$. Now we give a derivation of Eq.~\ref{eq:K_SW}. We first label the $2k$ neighbors of $u$ as $1, 2, \ldots, 2k$ by their clockwise ordering in the ring. Since $2k < n$, these nodes are unique. Next, define the \emph{span} of any $\ell$-clique containing $u$ as the difference between the largest and smallest label of the $\ell-1$ nodes in the clique other than $u$. The span $s$ of any $\ell$-clique satisfies $s \leq k-1$ since any node is directly connected with a node of label difference no greater than $k-1$. Also, $s \geq \ell-2$ since there are $\ell-1$ nodes in an $\ell$-clique other than $u$. For each span $s$, we can find $2k-1-s$ pairs of $(i, j)$ such that $1\leq i$, $j \leq 2k$ and $j - i = s$. Finally, for every such pair $(i, j)$, there are ${s - 1 \choose \ell - 3}$ choices of $\ell-3$ nodes between $i$ and $j$ which will form an $\ell$-clique together with nodes $u$, $i$, and $j$. Therefore, \begin{align*} \lvert \cliqueL{\ell}{u} \rvert &=\textstyle \sum_{s = \ell-2}^{k-1} (2k-1-s) \cdot {s - 1 \choose \ell - 3} \\ &=\textstyle \sum_{s = \ell-2}^{k-1} (2k-1-s) \cdot \frac{(s-1)(s - 2) \cdots (s - \ell + 3)}{(\ell - 3)!} \\ &=\textstyle \sum_{t = 1}^{k-\ell+2} (2k + 2 -t - \ell) \cdot \frac{t (t + 1) \cdots (t + \ell - 4)}{(\ell - 3)!}. \end{align*} If we ignore lower-order terms $k$ and note that $t = O(k)$, we get \begin{align*} \textstyle \lvert \cliqueL{\ell}{u} \rvert &= \textstyle \sum_{t = 1}^{k} \left[ \frac{(2k - t)t^{\ell - 3}}{(\ell - 3)!} + O(k^{\ell - 3}) \right] \\&=\textstyle \frac{1}{(\ell - 3)!}\sum_{t = 1}^{k} (2kt^{\ell - 3} - t^{\ell - 2}) + O(k^{\ell - 2}). \\ & = \textstyle \frac{1}{(\ell - 3)!}\left[2k \cdot \frac{k^{\ell - 2}}{\ell - 2} - \frac{k^{\ell - 1}}{\ell - 1} \right] + O(k^{\ell - 2}), \\ & = \textstyle \frac{\ell }{(\ell - 1)!} \cdot k^{\ell -1 } + O(k^{\ell - 2}) . \end{align*} \end{proof} Proposition~\ref{prop:ccf_sw} shows that, when $p = 0$, $\accf{\ell}$ decreases as $\ell$ increases. Furthermore, via simulation, we observe the same behavior as for $\accf{2}$ when adjusting the rewiring probability $p$ (Fig.~\ref{fig:sw_ccfs}). Regardless of $\ell$, the phase transition happens near $p = 0.1$. Essentially, once there is enough rewiring, all local clique structure is lost, and clustering at all orders is lost. This is partly a consequence of Proposition~\ref{prop:ccf_bounds}, which says that $\lccf{\ell}{u} \to 0$ as $\lccf{2}{u} \to 0$ for any $\ell$.
{ "timestamp": "2018-01-08T02:13:21", "yymm": "1704", "arxiv_id": "1704.03913", "language": "en", "url": "https://arxiv.org/abs/1704.03913", "abstract": "A fundamental property of complex networks is the tendency for edges to cluster. The extent of the clustering is typically quantified by the clustering coefficient, which is the probability that a length-2 path is closed, i.e., induces a triangle in the network. However, higher-order cliques beyond triangles are crucial to understanding complex networks, and the clustering behavior with respect to such higher-order network structures is not well understood. Here we introduce higher-order clustering coefficients that measure the closure probability of higher-order network cliques and provide a more comprehensive view of how the edges of complex networks cluster. Our higher-order clustering coefficients are a natural generalization of the traditional clustering coefficient. We derive several properties about higher-order clustering coefficients and analyze them under common random graph models. Finally, we use higher-order clustering coefficients to gain new insights into the structure of real-world networks from several domains.", "subjects": "Social and Information Networks (cs.SI); Statistical Mechanics (cond-mat.stat-mech); Physics and Society (physics.soc-ph); Machine Learning (stat.ML)", "title": "Higher-order clustering in networks", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.97631052387569, "lm_q2_score": 0.8397339676722393, "lm_q1q2_score": 0.8198411098942956 }
https://arxiv.org/abs/1006.4176
Unknotting Unknots
A knot is an an embedding of a circle into three-dimensional space. We say that a knot is unknotted if there is an ambient isotopy of the embedding to a standard circle. By representing knots via planar diagrams, we discuss the problem of unknotting a knot diagram when we know that it is unknotted. This problem is surprisingly difficult, since it has been shown that knot diagrams may need to be made more complicated before they may be simplified. We do not yet know, however, how much more complicated they must get. We give an introduction to the work of Dynnikov who discovered the key use of arc--presentations to solve the problem of finding a way to detect the unknot directly from a diagram of the knot. Using Dynnikov's work, we show how to obtain a quadratic upper bound for the number of crossings that must be introduced into a sequence of unknotting moves. We also apply Dynnikov's results to find an upper bound for the number of moves required in an unknotting sequence.
\section{Introduction} When one first delves into the theory of knots, one learns that knots are typically studied using their diagrams. The first question that arises when considering these knot diagrams is: how can we tell if two knot diagrams represent the same knot? Fortunately, we have a partial answer to this question. Two knot diagrams represent the same knot in $\mathbb{R}^3$ if and only if they can be related by the Reidemeister moves, pictured below. Reidemeister proved this theorem in the 1920's~ \cite{Reid}, and it is the underpinning of much of knot theory. For example, J. W. Alexander based the original definition of his celebrated polynomial on the Reidemeister moves~\cite{Alex}. \begin{figure}[h] \begin{center} \includegraphics[trim = 0mm 0mm 0mm 0mm, clip, height=1.5in]{ReidemeisterMoves.eps} \end{center} \vspace{-.3in} \caption{The three Reidemeister moves} \end{figure} Now imagine that you are presented with a complicated diagram of an unknot and you would like to use Reidemeister moves to reduce it to the trivial diagram that has no crossings. In considering a problem of this sort, you stumble upon a curious fact. Given a diagram of an unknot to be unknotted, it might be necessary to make the diagram more complicated before it can be simplified. We call such a diagram a {\em hard unknot diagram} \cite{KL}. A nice example of this is the Culprit, shown in Figure~\ref{culprit_intro}. If you look closely, you'll find that no simplifying type I or type II Reidemeister moves and no type III moves are available. Yet this is indeed the unknot. In order to unknot it, we need to introduce new crossings with Reidemeister I and II moves. In Figure~\ref{unknotCulprit}, we see that we can unknot the Culprit by making the diagram larger by two crossings (via a Reidemeister move of type two) and that it takes a total of ten Reidemeister moves to accomplish the unknotting. \bigbreak \begin{figure}[h] \begin{center} \includegraphics[height=1.2in, angle=90]{Culprit.eps} \end{center}\label{culprit_intro} \caption{The Culprit} \end{figure} \begin{figure}[h] \begin{center} \includegraphics[height=2in]{UndoCulprit.eps} \end{center} \caption{The Culprit Undone}\label{unknotCulprit} \end{figure} \begin{figure}[h] \begin{center} \includegraphics[height=2in]{Smallest.eps} \end{center} \caption{The smallest hard unknots} \end{figure} \begin{figure}[h] \begin{center} \includegraphics[height=1in]{Goeritz.eps} \end{center} \caption{The Goeritz unknot} \end{figure} In Figures 4 and 5 we indicate more examples of hard unknot diagrams. In Figure 4 we show the smallest possible such examples. In Figure 5 we show the very first such example, discovered by Goeritz in 1934 ~\cite{Goeritz}. \bigbreak At this point, we ask ourselves: how much more complicated does a diagram need to become before it can be simplified? Moreover, how many Reidemeister moves do we need to trivialize our picture? In this paper, we give a technique for finding upper bounds for these answers. In particular, we will prove the following theorem. \bigbreak \noindent {\bf Theorem 4.} {\em Suppose $K$ is a diagram (in Morse form) of the unknot with crossing number $cr(K)$ and number of maxima $b(K)$. Let $M=2b(K)+cr(K)$. Then the diagram can be unknotted by a sequence of Reidemeister moves so that no intermediate diagram has more than $(M-2)^2$ crossings.} \bigbreak The definition of Morse form for a diagram will be given in the body of the paper. In the case of our Culprit, we have that $cr(K) = 10$ and $b(K) = 5. $ Thus $M = 20$ and $(M-2)^2 = 18^{2} = 324.$ In actuality we only needed a diagram with $12$ crossings in our unknotting sequence. The theory of these bounds needs improvement, but it is, in fact, remarkable that there is a theory at all for such questions. Along with this theorem we will also give bounds on the number of Reidemeister moves needed for unknotting. We point the reader towards more results related to this question. As a disclaimer, we warn the reader that the difference between the lower bounds and upper bounds that are known is still vast. The quest for a satisfying answer to these questions continues. \section{Preliminaries} The method we present to find upper bounds makes use of a powerful result proven by Dynnikov in~\cite{dynnikov} regarding arc--presentations of knots. Here, we provide an overview of the theory of arc--presentations. \begin{definition} An \emph{arc--presentation} of a knot is a knot diagram comprised of horizontal and vertical line segments such that at each crossing in the diagram, the horizontal arc passes under the vertical arc. Furthermore, we require that no two edges in an arc--diagram are colinear. Two arc--presentations are \emph{combinatorially equivalent} if they are isotopic in the plane via an ambient isotopy of the form $h(x,y)=(f(x),g(y))$. The \emph{complexity} $c(L)$ of an arc--presentation is the number of vertical arcs in the diagram. We say more generally that a link diagram is {\em rectangular} if it has only vertical and horizontal edges. In Figure~\ref{arc} we give an example of a rectangular diagram that is an arc--presentation and another example of a rectangular diagram that is not an arc--presentation. \end{definition} Note that a rectangular diagram can naturally be drawn on a rectangular grid. If we start with such a grid and represent rectangular diagrams on the grid we have called these knots {\em mosaic knots} and used them to define a notion of {\em quantum knot.} See ~\cite{Mosaic} for more about quantum knots. For now, we focus our attention on arc--presentations. \begin{figure}[h] \begin{center} \includegraphics[height=1.5in]{trefoil.eps}\includegraphics[height=1.5in]{notarc.eps} \end{center} \vspace{-.3in} \caption{ \small{The picture on the left is an example of an arc--presentation of a trefoil. The picture on the right is an example that is \emph{not} an arc--presentation (since not all horizontal arcs pass under vertical arcs).}} \label{arc} \end{figure} \begin{proposition}[Dynnikov] Every knot has an arc--presentation. Any two arc--presentations of the same knot can be related to each other by a finite sequence of \emph{elementary moves}, pictured in Figures~\ref{stab} and~\ref{exch}. \end{proposition} The proof of this proposition is elementary, based on the Reidemeister moves. A sketch is provided in~\cite{dynnikov}. We will show how to convert a usual knot diagram to an arc--presentation in the next few paragraphs, making use of the concept of Morse diagrams of knots. \begin{figure}[h] \begin{center} \includegraphics[trim = 0mm 30mm 0mm 30mm, clip, height=1in]{stab1.eps} \includegraphics[trim = 0mm 30mm 0mm 30mm, clip, height=1in]{stab2.eps} \end{center} \caption{ \small{Elementary (de)stabilization moves. Stabilization moves increase the complexity of the arc--presentation while destabilization moves decrease the complexity.}} \label{stab} \end{figure} \begin{figure}[h] \begin{center} \includegraphics[trim = 0mm 30mm 0mm 30mm, clip, height=.9in]{exch2.eps}\includegraphics[trim = 0mm 30mm 0mm 30mm, clip, height=.9in]{arrow.eps}\includegraphics[trim = 0mm 30mm 0mm 30mm, clip, height=.9in]{exch1.eps} \end{center} \begin{center} \includegraphics[trim = 0mm 30mm 0mm 30mm, clip, height=.9in]{exch3.eps}\includegraphics[trim = 0mm 30mm 0mm 30mm, clip, height=.9in]{arrow.eps}\includegraphics[trim = 0mm 30mm 0mm 30mm, clip, height=.9in]{exch4.eps} \end{center} \caption{ \small{Some examples of exchange moves. Other allowed exchange moves include switching the heights of two horizontal arcs that lie in distinct halves of the diagram.}} \label{exch} \end{figure} \begin{definition} A knot diagram is in \emph{Morse form} if it has \begin{enumerate} \item no horizontal lines, \item no inflection points, \item a single singularity at each height, and \item each crossing is oriented to create a 45 degree angle with the vertical axis. \end{enumerate} \end{definition} We note that converting an arbitrary knot diagram into a diagram in Morse form requires no Reidemeister moves, only ambient isotopies of the plane. More information about Morse diagrams can be found in~\cite{lou}. \begin{lemma}\label{arclemma} Suppose a knot (or link) diagram $K$ in Morse form has $cr(K)$ crossings and $b(K)$ maxima. Then there is an arc--presentation $L_K$ of $K$ with complexity $c(L_K)$ at most $2b(K)+cr(K)$ that can be obtained by ambient isotopies of the plane (without the use of Reidemeister moves). \end{lemma} \begin{proof} We begin with a diagram in Morse form and convert this diagram into a piecewise linear diagram composed of lines with slope $\pm1$ with a vertex corresponding to each maximum and minimum (possibly with additional vertices---at most one for each pair of successive extrema). If we rotate this diagram by 45 degrees, we have a diagram composed entirely of horizontal and vertical arcs with complexity at most $2b(K)$. This diagram may fail to be an arc--presentation of $K$ if any crossing has a horizontal overpass. If more than half of the crossings in $K$ have horizontal overpasses, we rotate the diagram by 90 degrees. Now, at least half of the crossings are in the proper form. Any remaining crossings containing a horizontal overpass may locally be rotated 90 degrees to form our arc--presentation $L_K$, as shown in Figure~\ref{rotation}. For each crossing that requires this move, the complexity of the rectangular diagram increases by at most 2. Thus, the overall complexity of our diagram increases by at most $2(\frac{1}{2}cr(K))=cr(K)$. It follows that $c(L_K)\leq 2b(K)+cr(K)$. Note that neither converting a Morse diagram into a piecewise linear diagram nor locally rotating a crossing uses Reidemeister moves. These are ambient isotopies of the plane. \end{proof} \begin{figure}[h] \begin{center} \includegraphics[trim = 0mm 20mm 0mm 20mm, clip,height=1in]{morse.eps}\includegraphics[trim = 0mm 20mm 0mm 20mm, clip,height=1in]{notarc.eps} \end{center} \caption{ \small{A Morse diagram of a knot and a corresponding rectangular diagram.}} \label{morse} \end{figure} \begin{figure}[h] \begin{center} \includegraphics[trim = 0mm 30mm 0mm 30mm, clip, height=.7in]{rotation.eps} \end{center} \caption{ \small{Rotating a crossing to convert a rectangular diagram into an arc--presentation.}} \label{rotation} \end{figure} \begin{figure}[h] \begin{center} \includegraphics[height=1.3in]{notarc.eps}\includegraphics[height=1.3in]{arrow.eps}\includegraphics[height=1.3in, angle=90]{notarc.eps}\includegraphics[height=1.3in]{arrow.eps}\includegraphics[height=1.3in]{ugly.eps} \end{center} \caption{ \small{Converting a rectangular diagram into an arc--presentation by rotating the diagram then rotating a crossing. Note that the resulting diagram can be reduced to the arc--presentation from Figure 1 with an exchange move that doesn't require any Reidemeister moves.}} \label{morse} \end{figure} \section{Bounds on Crossings Needed to Simplify the Unknot} Our motivation for using Dynnikov's work to find upper bounds for an unknotting Reidemeister sequence began with the following theorem from~\cite{dynnikov}. \begin{theorem}[Dynnikov]\label{triv} If $L$ is an arc--presentation of the unknot, then there exists a finite sequence of exchange and destabilization moves $$L\rightarrow L_1\rightarrow L_2\rightarrow \cdots\rightarrow L_m$$ such than $L_m$ is trivial.\end{theorem} What is particularly interesting about this result is that the unknot can be simplified \textbf{without increasing the complexity of the arc--presentation}, that is, without the use of stabilization moves. This gives a useful physical bound on how large a diagram can be. Furthermore, if we apply Dynnikov's method to a knotted knot, it will stop on a diagram that is not a planar circle. Thus Dynnikov can detect the unknot. \bigbreak The problem of detecting the unknot has been investigated by many people. For example, the papers by Birman and Hirsch ~\cite{Alg} and Birman and Moody ~\cite{Obstr} give such methods. More recently it has been shown that Heegard Floer Homology (a generalization of the Alexander polynomial) not only detects the unknot, but can be used to calculate the least genus of an orientable spanning surface for any knot. This is an outstanding result and we recommend that the reader examine the paper by Manolescu, Oszvath, Szabo and Thurston ~\cite{Heegard} for more information. In that work, the Heegard Floer homology is expressed via a chain complex that is associated to a rectangular diagram of just the type that Dynnikov uses. \bigbreak Returning to the task at hand, we immediately derive a quadratic upper bound on the crossing number of diagrams in an unknotting sequence. A similar result can be found in~\cite{dynnikov}. \begin{theorem} Suppose $K$ is a diagram (in Morse form) of the unknot with crossing number $cr(K)$ and number of maxima $b(K)$. Then, for every $i$, the crossing number $cr(K_i)$ is no more than $(M-2)^2$ where $M=2b(K)+cr(K)$ and $K=K_0, K_1, K_2, ..., K_N$ is a sequence of knot diagrams such that $K_{i+1}$ is obtained from $K_i$ by a single Reidemeister move and $K_N$ is a trivial diagram of the unknot. \end{theorem} \begin{proof} To begin, we notice that $K$ can be viewed as an arc--presentation of complexity $M$ by a simple ambient isotopy of the plane, as shown in Lemma~\ref{arclemma}. By Theorem~\ref{triv}, there is a sequence of arc--presentations beginning with $K$ and ending with the trivial arc--presentation each having complexity no more than $M$ such that a diagram and its successor are related by an exchange or a destabilization move. Each destabilization move either preserves or reduces the number of crossings in the diagram. In the case that a destabilization move reduces the number of crossings, it can be viewed as a simplifying Redemeister I move. Otherwise, it can be viewed as a simple ambient isotopy of the plane. When an exchange move is performed, on the other hand, its analogous Reidemeister sequence may require type II and type III Reidemeister moves. (See Figure~\ref{factor}.) At most one type II Reidemeister move is required for any given exchange move, so an exchange move factors through a Reidemeister sequence of moves that adds at most two crossings (since type III moves preserve the crossing number). However, it is important to note that a Reidemeister II move is needed if and only if the exchange move itself increases the number of crossings by two in the arc--presentation. Thus, no more crossings are added when factoring an exchange move through a Reidemeister sequence than are added in the exchange move itself. \begin{figure}[h] \begin{center} \includegraphics[height=1.3in]{exch2.eps}\includegraphics[height=1.2in]{arrow.eps}\includegraphics[height=1.3in]{exch2a.eps} \includegraphics[height=1.2in]{arrow.eps} \vspace{-.3in} \includegraphics[height=1.3in]{exch2b.eps}\includegraphics[height=1.2in]{arrow.eps}\includegraphics[height=1.2in]{dots.eps}\includegraphics[height=1.2in]{arrow.eps}\includegraphics[height=1.4in]{exch2c.eps} \vspace{-.2in} \includegraphics[trim = 0mm 20mm 0mm 0mm, clip,height=1.2in]{arrow.eps}\includegraphics[height=1.3in]{exch1.eps} \end{center} \caption{ \small{Factoring an exchange move through a type II and multiple type III Reidemeister moves.}} \label{factor} \end{figure} It is straightforward to show that the maximum number of crossings that may occur in an arc-presentation with complexity less than or equal to $M$ is bounded above by $(M-2)^2$. If we translate an arc--presentation sequence of moves in a canonical fashion into a sequence of Reidemeister moves to unknot our unknot, many knot diagrams in the Reidemeister sequence will be arc--presentations and, as such, will have fewer than $(M-2)^2$ crossings. Furthermore, diagrams in this sequence that are not arc--presentations have no more crossings than their arc--presentation relatives. Thus, there exists a sequence of Reidemeister moves that unknots our original diagram $K$ that does not increase the crossing number to more than $(M-2)^2.$ \end{proof} \section{Bounds on Reidemeister Moves Needed to Simplify the Unknot} To find our upper bound on the number of Reidemeister moves, we must first specify an upper bound on the number $m$ of exchange and destabilization moves required to trivialize an arc--presentation. This bound will depend on the complexity $c(L)=n$ of the arc--diagram $L$. We also must provide an upper bound on the number of Reidemeister moves required for a destabilization or exchange move. In~\cite{dynnikov}, Dynnikov provides the following bounds on the number of combinatorially distinct arc--presentations of complexity $n$. \begin{proposition} Let $N(n)$ denote the number of combinatorially distinct arc-- presentations of complexity $n$. Then the following inequality holds. $$N(n)\leq \frac{1}{2}n [(n-1)!]^2$$ \end{proposition} \begin{proof} Suppose we want to create an arc--presentation on the $n\times n$ integer lattice. Let us choose a starting point in the lattice. There are $\frac{n}{2}=\frac{n^2}{2n}$ ways to choose this point since there are $n^2$ lattice points, $2n$ of which lie on a given diagram. From this point, we create a vertical arc ending at another point in the integer lattice. There are $n-1$ choices for this endpoint. From our new point, we want to create a horizontal arc with endpoint in the lattice. There are $n-1$ choices for this endpoint as well. Next, we make another vertical arc, choosing one of the $n-2$ possible endpoints. (There are only $n-2$ choices since no two arcs in the diagram should be colinear.) Similarly, we have $n-2$ choices for the endpoint of our next horizontal arc. Continuing in this fashion, we see that the number of distinct choices we must make is $[(n-1)!]^2$. Multiplying this quantity by $\frac{n}{2}$ to account for the initial choice of starting point, we get $\frac{1}{2}n[(n-1)!]^2$ \end{proof} Using this count on the number of distinct arc--presentations of a given size, we can find a bound (albeit a large one) on the number of arc--presentation moves we need. This is simply by virtue of the fact that any reasonable sequence of moves will contain mutually distinct arc--presentations that don't exceed the complexity of the original, and there are a limited number of such diagrams. \begin{lemma}\label{size} The number of terms, $m$, in the monotonic simplification of arc--presentation $L$ with $c(L)=n$ is bounded above by $\sum_{i=2}^n\frac{1}{2}i[(i-1)!]^2$. \end{lemma} \begin{proof} Suppose that an arc--presentation $L$ has complexity $n$. Since each $L_k$ from Theorem~\ref{triv} is combinatorially distinct from any other $L_j$ with $k\neq j$, we know that the number $m$ of arc--presentations in the sequence must be at most $\sum_{i=2}^nN(i)$ which is no greater than $\sum_{i=2}^n\frac{1}{2}i[(i-1)!]^2$. \end{proof} We should note that, if we start with an arc--presentation of the unknot, every arc--presentation in our simplification sequence must be a diagram of the unknot. As $n$ gets larger, we recognize that far fewer arc--presentations of complexity $n$ are unknots. Thus, in practice, $m$ will be much lower than the upper bound provided here. The authors would be interested to know what the probability is that an arc--presentation of complexity $n$ is the unknot. Using this probability, we could tighten the upper bound we found above. We return now to our second question: how many Reidemeister moves does it take to make an arc--presentation move? \begin{lemma}\label{reid} No more than $n-2$ Reidemeister moves are required to perform an exchange or destabilization move on an arc--presentation $L$ with complexity $c(L)=n$. \end{lemma} \begin{proof} Clearly, a destabilization move requires at most one Reidemeister move, a type I move. Now consider the first exchange move pictured in Figure~\ref{exch}. Let $d$ be the number of vertical strands intersecting both of the horizontal strands to be switched. Then the move requires $d$ type III moves and one type II move. Thus, the exchange move requires $d+1$ Reidemeister moves. We note that $d<a$, if $a$ is the length of the shorter horizontal arc. But $a$ cannot be greater than $n-2$, so the number of Reidemeister moves required is less than or equal to $n-2$. Similarly, the second exchange move pictured above requires $d$ type III moves but no type II moves. Thus, both pictured exchange moves require no more than $n-2$ Reidemeister moves. We note that other versions of the exchange moves (where the horizontal arcs lie in distinct halves of the arc--presentation) require no Reidemeister moves. \end{proof} For the finale, we put our two results together. \begin{theorem} Suppose $K$ is a diagram (in Morse form) of the unknot with crossing number $cr(K)$ and number of maxima $b(K)$. Let $M=2b(K)+cr(K)$. Then the number of Reidemeister moves required to unknot $K$ is less than or equal to $$\sum_{i=2}^M\frac{1}{2}i[(i-1)!]^2(M-2).$$ \end{theorem} \begin{proof} Suppose the arc--presentation $L_K$ of our knot diagram $K$ has complexity $c(L_K)=n$. Then at most $m (n-2)$ Reidemeister moves are required to produce the trivial (complexity 2) arc--presentation, where $m$ is the number of moves in the monotonic simplification of $L_K$. By our lemma, this quantity is bounded above by $$\sum_{i=2}^n\frac{1}{2}i[(i-1)!]^2(n-2).$$ But we showed that $n\leq 2b(K)+cr(K)=M$, thus the number of Reidemeister moves required to unknot $K$ is less than or equal to $$\sum_{i=2}^M\frac{1}{2}i[(i-1)!]^2(M-2).$$ \end{proof} \section{A Detour: Bounds for Untangling Links}\label{links} To illustrate that similar questions may be extended to families of knots and links beyond the unknot, we take a short detour to the world of non-trivial knots and links. In keeping with our theme, we make use of the work of Dynnikov. He proved two other results regarding the simplification of certain link diagrams~\cite{dynnikov}. In a fashion analogous to the previous section, we may use Dynnikov's results to bound the number of Reidemeister moves and the number of crossings needed to simplify certain types link diagrams. Before we state these theorems, however, let us clearly define our terms. \begin{figure}[h] \begin{center} \includegraphics[height=1.3in]{SplitLink.eps} \end{center} \caption{ \small{A split link.}}\label{split} \end{figure} \begin{figure}[h] \begin{center} \includegraphics[height=1.5in]{ConnectSum.eps} \end{center} \caption{ \small{A composite knot.}}\label{composite} \end{figure} \begin{definition} A link diagram $L$ is said to be \emph{split} if there is a line not intersecting $L$ such that there are components of the diagram lying on both sides of the line. A link (or knot) diagram $L$ is \emph{composite} if it can be viewed as a connect sum of two nontrivial links, i.e. if there is a line intersecting the link at two points such that the tangles on either side of the line are non-trivial. In general, a link is said to be split or composite if there exists a diagram of the link that is split or composite. Figures ~\ref{split} and~\ref{composite} give examples illustrating these definitions. \end{definition} Let's review the pertinent results from~\cite{dynnikov}. \begin{theorem}[Dynnikov] If $L$ is an arc--presentation of a split link, then there exists a finite sequence of exchange and destabilization moves $$L\rightarrow L_1\rightarrow L_2\rightarrow \cdots\rightarrow L_m$$ such than $L_m$ is split.\end{theorem} \begin{theorem}[Dynnikov] If $L$ is an arc--presentation of a non-split composite link, then there exists a finite sequence of exchange and destabilization moves $$L\rightarrow L_1\rightarrow L_2\rightarrow \cdots\rightarrow L_m$$ such than $L_m$ is composite.\end{theorem} We note that the statements of Lemmas ~\ref{arclemma},~\ref{size} and~\ref{reid} hold for arbitrary links as well as diagrams of the unknot. Thus, the following result is an immediate consequence of the previous theorems. \begin{theorem} Suppose $L$ is a diagram (in Morse form) of a split (resp. non-split composite) link with crossing number $cr(L)$ and number of maxima $b(L)$. Let $M=2b(L)+cr(L)$. Then the number of Reidemeister moves required to transform $L$ into a split (resp. composite) diagram is less than or equal to $$\sum_{i=2}^M\frac{1}{2}i[(i-1)!]^2(M-2).$$ \end{theorem} Similarly, we have the following extension of our results regarding maximum crossing numbers in a simplifying Reidemeister sequence. \begin{theorem} Suppose $L$ is a diagram (in Morse form) of a split (resp. non-split composite) link with crossing number $cr(L)$ and number of maxima $b(L)$. Then for every $i$, the crossing number $cr(L_i)$ is no more than $(M-2)^2$ where $M=2b(L)+cr(L)$ and $L=L_0, L_1, L_2, ..., L_N$ is a sequence of link diagrams such that $L_{i+1}$ is obtained from $L_i$ by a single Reidemeister move, $L_N$ is split (resp. composite). \end{theorem} \section{Hard Unknots} We have provided several upper bounds regarding the complexity of the Reidemeister sequence required to simplify an unknot. The bound that Dynnikov's work helps us obtain for the number of Reidemeister moves required to unknot an unknot is superexponential. Using a different technique, Hass and Lagarias were able to find a bound that is exponential in the crossing number of the diagram~\cite{hl}. They use the same technique to find an exponential bound for the number of crossings required for unknotting. For bounds of this second sort, the one presented here is a comparatively sharper estimate. Regarding lower bounds, it was recently shown in ~\cite{hn} that there are unknot diagrams for which the number of Reidemeister moves required for unknotting is quadratic in the crossing number of the initial diagram. In~\cite{hayashi}, similar quadratic lower bounds are given for links. On the other hand, little is known about how many additional crossings an unknot diagram might require in order to become unknotted. While the upper bound on the number of crossings needed in a Reidemeister sequence is merely quadratic in the crossing number of the initial unknot diagram, it nonetheless seems likely that this bound is far from being tight. Let us return to our friend, the Culprit. This famous hard unknot diagram was originally discovered by Ken Millett and introduced in~\cite{Culprit}. Recall that hard unknots are difficult to unknot by virtue of the fact that no simplifying type I or type II Reidemeister moves and no type III moves are available. In Figure~\ref{culprit}, we picture a Morse diagram of the Culprit, its corresponding rectangular diagram, and its arc-presentation obtained by rotating crossings where the over-strand was horizontal. Note that we need not specify crossing information in the arc--presentation, for it is assumed that all vertical lines pass over horizontal lines. \begin{figure}[h] \begin{center} \includegraphics[height=1.5in]{Culprit.eps}\includegraphics[height=1.5in]{CulpritGrid.eps}\includegraphics[height=1.5in]{CulpritArc.eps} \end{center} \caption{ \small{The Culprit with its rectangular diagram and arc--presentation.}}\label{culprit} \end{figure} We saw that the Culprit may be unknotted with ten Reidemeister moves. (See also ~\cite{KL}.) The maximum crossing number of all diagrams in the given Reidemeister sequence is 12, two more than the number of crossings in the Culprit. On the other hand, we can compute our upper bound on the number of crossings required for unknotting as follows. Since the crossing number $cr(K)=10$ and the number of maxima in the diagram is $b(K)=5$, we see that $M = cr(K) + 2b(K)=20$. Thus, our bound is $(M-2)^2=18^2=324.$ We can also use $M$ to find our bound for the number of Reidemeister moves required to unknot the Culprit. $$\sum_{i=2}^M\frac{1}{2}i[(i-1)!]^2(M-2)=9\sum_{i=2}^{20}i[(i-1)!]^2.$$ The largest term in this expression is roughly $10^{35}$, unfortunately quite a bit larger than ten. We challenge the reader to find examples where the maximum crossing number is closer to our bound and where the number of needed Reidemeister moves is large in comparison to the number of crossings in the original diagram. \section{Conclusions} We've considered the phenomenon that it may be quite hard to unknot a trivial knot and have provided several upper and lower bounds on the number of Reidemeister moves and the number of crossings needed to do the job. Known hard unknots like the Culprit and examples from~\cite{hn} illustrate that unknotting can be tricky, but not as tricky as the upper bounds that are known would have us believe. To answer the questions we've posed, there is much more to be done. \section{Acknowledgements} The authors would like to thank Jeffrey Lagarias and John Sullivan for their valuable comments. \bibliographystyle{abbrv}
{ "timestamp": "2011-11-08T02:00:36", "yymm": "1006", "arxiv_id": "1006.4176", "language": "en", "url": "https://arxiv.org/abs/1006.4176", "abstract": "A knot is an an embedding of a circle into three-dimensional space. We say that a knot is unknotted if there is an ambient isotopy of the embedding to a standard circle. By representing knots via planar diagrams, we discuss the problem of unknotting a knot diagram when we know that it is unknotted. This problem is surprisingly difficult, since it has been shown that knot diagrams may need to be made more complicated before they may be simplified. We do not yet know, however, how much more complicated they must get. We give an introduction to the work of Dynnikov who discovered the key use of arc--presentations to solve the problem of finding a way to detect the unknot directly from a diagram of the knot. Using Dynnikov's work, we show how to obtain a quadratic upper bound for the number of crossings that must be introduced into a sequence of unknotting moves. We also apply Dynnikov's results to find an upper bound for the number of moves required in an unknotting sequence.", "subjects": "Geometric Topology (math.GT)", "title": "Unknotting Unknots", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9863631619124993, "lm_q2_score": 0.8311430499496095, "lm_q1q2_score": 0.8198088867498952 }
https://arxiv.org/abs/2007.09846
Pure metric geometry: introductory lectures
We discuss domestic affairs of metric spaces, keeping away from any extra structure. Topics include universal spaces, injective spaces, Hausdorff and Gromov--Hausdorff convergences, and ultralimits.
\chapter{Space of sets} \section{Hausdorff distance} Let $\spc{X}$ be a metric space. Given a subset $A\subset \spc{X}$, consider the distance function to $A$ $$\distfun_A: \spc{X} \to [0,\infty)$$ defined as $$\distfun_A(x) \df \inf_{a\in A}\{\,\dist ax{\spc{X}}\,\}.$$ \begin{thm}{Definition}\label{def:hausdorff-convergence} Let $A$ and $B$ be two compact subsets of a metric space $\spc{X}$. Then the \index{Hausdorff distance}\emph{Hausdorff distance} between $A$ and $B$ is defined as $$|A-B|_{\Haus\spc{X}} \df \sup_{x\in \spc{X}}\{\,|\distfun_A(x)-\distfun_B(x)|\,\}. $$ \end{thm} The following observation gives a useful reformulation of the definition: \begin{thm}{Observation}\label{obs:Haus-nbhds} Suppose $A$ and $B$ be two compact subsets of a metric space $\spc{X}$. Then $|A-B|_{\Haus\spc{X}}< R$ if and only if and only if $B$ lies in an $R$-neighborhood of $A$, and $A$ lies in an $R$-neighborhood of~$B$. \end{thm} Note that the set of all nonempty compact subsets of a metric space $\spc{X}$ equipped with the Hausdorff metric forms a metric space. This new metric space will be denoted as $\Haus\spc{X}$. \begin{thm}{Exercise}\label{ex:diam} Let $\spc{X}$ be a metric space. Given a subset $A\subset \spc{X}$ define its \index{diameter}\emph{diameter} as $$\diam A\df\sup_{a,b\in A} |a-b|.$$ Show that $$\diam\:\Haus\spc{X}\to \RR$$ is a \index{Lipschitz function}\emph{$2$-Lipschitz function}; that is, \[|\diam A-\diam B|\le 2\cdot\dist{A}{B}{\Haus\spc{X}}\] for any two compact nonempty sets $A,B\subset\spc{X}$. \end{thm} \begin{thm}{Exercise}\label{ex:Hausdorff-bry} Let $A$ and $B$ be two compact subsets in the Euclidean plane $\RR^2$. Assume $|A-B|_{\Haus\RR^2}<\eps$. \begin{subthm}{ex:Hausdorff-bry:conv} Show that $|\Conv A-\Conv B|_{\Haus\RR^2}<\eps$, where $\Conv A$ denoted the convex hull of $A$. \end{subthm} \begin{subthm}{ex:Hausdorff-bry:bry} Is it true that $|\partial A-\partial B|_{\Haus\RR^2}<\eps$, where $\partial A$ denotes the boundary of $A$. Does the converse hold? That is, assume $A$ and $B$ be two compact subsets in $\RR^2$ and $|\partial A-\partial B|_{\Haus\RR^2}<\eps$; is it true that $|A-B|_{\Haus\RR^2}\z<\eps$? \end{subthm} \end{thm} Note that part \ref{SHORT.ex:Hausdorff-bry:conv} implies that $A\mapsto \Conv A$ defines a short map $\Haus\RR^2\to \Haus\RR^2$. \begin{thm}{Exercise}\label{ex:Haus-func} Let $A$ and $B$ be two compact subsets in metric space $\spc{X}$. Show that \[\dist{A}{B}{\Haus\spc{X}}=\sup_f\, \{\,\max_{a\in A}\{f(a)\}-\max_{b\in B}\{f(b)\,\},\] where the least upper bound is taken for all $1$-Lipschitz functions $f$. \end{thm} \section{Hausdorff convergence} \begin{thm}{Blaschke selection theorem}\label{thm:compact+Hausdorff} A metric space $\spc{X}$ is compact if and only if so is $\Haus\spc{X}$. \end{thm} The Hausdorff metric can be used to define convergence. Namely, suppose $K_1,K_2,\dots$, and $K_\infty$ are compact sets in a metric space $\spc{X}$. If $|K_\infty-K_n|_{\Haus\spc{X}}\to0$ as $n\to\infty$, then we say that the sequence $(K_n)$ {}\emph{converges} to $K_\infty$ \index{convergence in the sense of Hausdorff}\emph{in the sense of Hausdorff}; or we can say that $K_\infty$ is \emph{Hausdorff limit} of the sequence $(K_n)$. Note that the theorem implies that from any sequence of compact sets in $\spc{X}$ one can select a subsequence that converges in the sense of Hausdorff; for that reason, it is called a \emph{selection} theorem. \parit{Proof; ``only if'' part.} Consider the map $\iota$ that sends point $x\in \spc{X}$ to the one-point subset $\{x\}$ of $\spc{X}$. Note that $\iota\:\spc{X}\to \Haus\spc{X}$ is distance-preserving. Suppose that $A\subset \spc{X}$. Note that $\diam A=0$ if and only if $A$ is a one-point set. Therefore, from Exercise~\ref{ex:diam}, it follows that $\iota(\spc{X})$ is a closed subset of the compact space $\Haus\spc{X}$. Whence $\iota(\spc{X})$, and therefore $\spc{X}$, are compact. \qeds To prove the ``if'' part we will need the following two lemmas. \begin{thm}{Monotone convergence}\label{lem:decreasing-converges} Let $K_1\supset K_2\supset\dots$ be a nested sequence of nonempty compact sets in a metric space $\spc{X}$ then $K_\infty\z=\bigcap_n K_n$ is the Hausdorff limit of $K_n$; that is, $|K_\infty-K_n|_{\Haus\spc{X}}\to0$ as $n\to\infty$. \end{thm} \parit{Proof.} By finite intersection property, $K_\infty$ is a nonempty compact set. If the assertion were false, then there is $\eps>0$ such that for each $n$ one can choose $x_n\in K_n$ such that $\distfun_{K_\infty}(x_n)\ge\eps$. Note that $x_n\in K_1$ for each $n$. Since $K_1$ is compact, there is a \index{partial limit}\emph{partial limit}% \footnote{Partial limit is a limit of a subsequence.} $x_\infty$ of $x_n$. Clearly, $\distfun_{K_\infty}(x_\infty)\ge \eps$. On the other hand, since $K_n$ is closed and $x_m\in K_n$ for $m\ge n$, we get $x_\infty\in K_n$ for each $n$. It follows that $x_\infty\in K_\infty$ and therefore $\distfun_{K_\infty}(x_\infty)=0$ --- a contradiction.\qeds \begin{thm}{Lemma}\label{lem:complete+Hausdorff} If $\spc{X}$ is a compact metric space, then $\Haus\spc{X}$ is complete. \end{thm} \parit{Proof.} Let $(Q_n)$ be a Cauchy sequence in $\Haus\spc{X}$. Passing to a subsequence of $Q_n$ we may assume that $$|Q_n-Q_{n+1}|_{\Haus\spc{X}}\le \tfrac1{10^n}\eqlbl{eq:eps=1/10}$$ for each $n$. Denote by $K_n$ the closed $\tfrac1{10^n}$-neighborhood of $Q_n$; that is, \begin{align*} K_n&= \set{x\in \spc{X}}{\distfun_{Q_n}(x)\le \tfrac1{10^n}} \end{align*} Since $\spc{X}$ is compact so is each $K_n$. By \ref{obs:Haus-nbhds}, $|Q_n-K_n|_{\Haus\spc{X}}\le \tfrac1{10^n}$. From \ref{eq:eps=1/10}, we get $K_n\supset K_{n+1}$ for each $n$. Set $$K_\infty=\bigcap_{n=1}^\infty K_n.$$ By the monotone convergence (\ref{lem:decreasing-converges}), $|K_n-K_\infty|_{\Haus\spc{X}}\to 0$ as $n\to\infty$. Since $|Q_n-K_n|_{\Haus\spc{X}}\le \tfrac1{10^n}$, we get $|Q_n-K_\infty|_{\Haus\spc{X}}\to 0$ as $n\to\infty$ --- hence the lemma. \qeds \begin{thm}{Exercise}\label{ex:closure-union} Let $\spc{X}$ be a complete metric space and $K_1,K_2,\dots$ be a sequence of compact sets that converges in the sense of Hausdorff. Show that the union $K_1\cup K_2\cup\dots$ is a compact closure. Use this statement to show that in Lemma~\ref{lem:complete+Hausdorff} compactness of $\spc{X}$ can be exchanged to completeness. \end{thm} \parit{Proof of ``if'' part in \ref{thm:compact+Hausdorff}.} According to Lemma~\ref{lem:complete+Hausdorff}, $\Haus\spc{X}$ is complete. It remains to show that $\Haus\spc{X}$ is totally bounded (\ref{totally-bounded}); that is, given $\eps>0$ there is a finite $\eps$-net in $\Haus\spc{X}$. Choose a finite $\eps$-net $A$ in $\spc{X}$. Denote by $B$ the set of all subsets of $A$. Note that $B$ is a finite set in $\Haus\spc{X}$. For each compact set $K\subset \spc{X}$, consider the subset $K'$ of all points $a\in A$ such that $\distfun_K(a)\le \eps$. Observe that $K' \in B$ and $|K-K'|_{\Haus\spc{X}}\le\eps$. In other words, $B$ is a finite $\eps$-net in $\Haus\spc{X}$. \qeds \begin{thm}{Exercise}\label{ex:Haus-length} Let $\spc{X}$ be a complete metric space. Show that $\spc{X}$ is a length space if and only if so is $\Haus\spc{X}$. \end{thm} \section{An application} The following statement is called \index{isoperimetric inequality}\emph{isoperimetric inequality in the plane}. \begin{thm}{Theorem}\label{thm:isoperimetric} Among the plane figures bounded by closed curves of length at most $\ell$ the round disk has the maximal area. \end{thm} In this section, we will sketch a proof of the isoperimetric inequality that uses the Hausdorff convergence. It is based on the following exercise. \begin{thm}{Exercise}\label{ex:Huas-perimeter-area} Let $\spc{C}$ be a subspace of $\Haus\RR^2$ formed by all compact convex subsets in $\RR^2$. Show that perimeter\footnote{If the set degenerates to a line segment of length $\ell$, then its perimeter is defined as $2\cdot \ell$.} and area are continuous on~$\spc{C}$. That is, if a sequence of convex compact plane sets $X_n$ converges to $X_\infty$ in the sense of Hausdorff, then \[\perim X_n\to \perim X_\infty\quad\text{and}\quad\area X_n\to\area X_\infty\] as $n\to\infty$. \end{thm} \parit{Semiproof of \ref{thm:isoperimetric}.} It is sufficient to consider only convex figures of the given perimeter; if a figure is not convex, pass to its convex hull and observe that it has a larger area and smaller perimeter. Note that the selection theorem (\ref{thm:compact+Hausdorff}) together with the exercise imply the existence of figure $D$ with perimeter $\ell$ and maximal area. It remains to show that $D$ is a round disk. This is a problem in elementary geometry. Let us cut $D$ along a chord $[ab]$ into two lenses, $L_1$ and $L_2$. Denote by $L_1'$ the reflection of $L_1$ across the perpendicular bisector of $[ab]$. Note that $D$ and $D'=L_1'\cup L_2$ have the same perimeter and area. That is, $D'$ has perimeter $\ell$ and maximal possible area; in particular, $D'$ is convex. The following exercise will finish the proof. \qeds { \begin{wrapfigure}{o}{57 mm} \vskip-5mm \centering \includegraphics{mppics/pic-405} \end{wrapfigure} \begin{thm}{Exercise}\label{ex:round-disc} Suppose $D$ is a convex figure such that for any chord $[ab]$ of $D$ the above construction produces a convex figure $D'$. Show that $D$ is a round disk. \end{thm} } Another popular way to prove that $D$ is a round disk is given by the so-called {}\emph{Steiner's 4-joint method} \cite{blaschke}. \section{Remarks}\label{sec:H-variation} It seems that Hausdorff convergence was first introduced by Felix Hausdorff~\cite{hausdorff}. A couple of years later an equivalent definition was given by Wilhelm Blaschke~\cite{blaschke}. The following refinement of the definition was introduced by Zdeněk Frolík \cite{frolik}, later it was rediscovered by Robert Wijsman~\cite{wijsman}. This refinement is also called \index{Hausdorff convergence}\emph{Hausdorff convergence}; in fact, it takes an intermediate place between the original Hausdorff convergence and {}\emph{closed convergence}, also introduced by Hausdorff in \cite{hausdorff}. \begin{thm}{Definition}\label{def:gen-Haus-conv} Let $(A_n)$ be a sequence of closed sets in a metric space $\spc{X}$. We say that $(A_n)$ converges to a closed set $A_\infty$ in the sense of Hausdorff if for any $x\in\spc{X}$, we have $\distfun_{A_n}(x)\to \distfun_{A_\infty}(x)$ as $n\to\infty$. \end{thm} For example, suppose $\spc{X}$ is the Euclidean plane and $A_n$ is the circle with radius $n$ and center at the point $(n,0)$. If we use the standard definition (\ref{def:hausdorff-convergence}), then the sequence $(A_n)$ diverges, but it converges to the $y$-axis in the sense of Definition~\ref{def:gen-Haus-conv}. The following exercise is analogous to the Blaschke selection theorem (\ref{thm:compact+Hausdorff}) for the modified Hausdorff convergence. \begin{thm}{Exercise}\label{ex:generalized-selection} Let $\spc{X}$ be a proper metric space and $(A_n)_{n=1}^\infty$ be a sequence of closed sets in~$\spc{X}$. Assume that for some (and therefore any) point $x\in\spc{X}$, the sequence $a_n=\distfun_{A_n}(x)$ is bounded. Show that the sequence $(A_n)_{n=1}^\infty$ has a convergent subsequence in the sense of Definition~\ref{def:gen-Haus-conv}. \end{thm} \chapter{Space of spaces} \section{Gromov--Hausdorff metric} The goal of this section is to cook up a metric space out of metric spaces. More precisely, we want to define the so-called Gromov--Hausdorff metric on the set of {}\emph{isometry classes} of compact metric spaces. (Being isometric is an equivalence relation, and an isometry class is an equivalence class with respect to this equivalence relation.) The obtained metric space will be denoted by $\GH$. Given two metric spaces $\spc{X}$ and $\spc{Y}$, denote by $[\spc{X}]$ and $[\spc{Y}]$ their isometry classes; that is, $\spc{X}'\in [\spc{X}]$ if and only if $\spc{X}'\iso \spc{X}$. Pedantically, the Gromov--Hausdorff distance from $[\spc{X}]$ to $[\spc{Y}]$ should be denoted as $|[\spc{X}]-[\spc{Y}]|_{\GH}$; but we will write it as $|\spc{X}\z-\spc{Y}|_{\GH}$ and say (not quite correctly) ``$|\spc{X}\z-\spc{Y}|_{\GH}$ is the Gromov--Hausdorff distance from $\spc{X}$ to $\spc{Y}$''. In other words, from now on the term {}\emph{metric space} might also stand for its {}\emph{isometry class}. The metric on $\GH$ is defined as the maximal metric such that {}\emph{the distance between subspaces in a metric space is not greater than the Hausdorff distance between them}. Here is a formal definition: \begin{thm}{Definition}\label{def:GH} Let $\spc{X}$ and $\spc{Y}$ be compact metric spaces. The Gromov--Hausdorff distance $|\spc{X}-\spc{Y}|_{\GH}$ is defined by the following relation. Given $r > 0$, we have that $|\spc{X}-\spc{Y}|_{\GH} < r$ if and only if there exist a metric space $\spc{Z}$ and subspaces $\spc{X}'$ and $\spc{Y}'$ in $\spc{Z}$ that are isometric to $\spc{X}$ and $\spc{Y}$ respectively and such that $|\spc{X}'-\spc{Y}'|_{\Haus\spc{Z}} < r$. (Here $|\spc{X}'-\spc{Y}'|_{\Haus\spc{Z}}$ denotes the Hausdorff distance between sets $\spc{X}'$ and $\spc{Y}'$ in $\spc{Z}$.) \end{thm} Note that passing to the subspace $\spc{X}'\cup\spc{Y}'$ of $\spc{Z}$ does not affect the definition. Therefore we can always assume that $\spc{Z}$ is compact. \begin{thm}{Theorem}\label{thm:GH-is-a-metric} The set of isometry classes of compact metric spaces equipped with Gromov--Hausdorff metric forms a metric space (which is denoted by $\GH$). In other words, for arbitrary compact metric spaces $\spc{X}$, $\spc{Y}$ and $\spc{Z}$ the following conditions hold: \begin{subthm}{GH-1} $|\spc{X}-\spc{Y}|_{\GH}\ge 0$; \end{subthm} \begin{subthm}{GH-2} $|\spc{X}-\spc{Y}|_{\GH}=0$ if and only if $\spc{X}$ is isometric to $\spc{Y}$; \end{subthm} \begin{subthm}{GH-3} $|\spc{X}-\spc{Y}|_{\GH}=|\spc{Y}-\spc{X}|_{\GH}$; \end{subthm} \begin{subthm}{GH-4} $|\spc{X}-\spc{Y}|_{\GH}+|\spc{Y}-\spc{Z}|_{\GH}\ge |\spc{X}-\spc{Z}|_{\GH}$. \end{subthm} \end{thm} Note that \ref{SHORT.GH-1}, \ref{SHORT.GH-3}, and the ``if''-part of \ref{SHORT.GH-2} follow directly from Definition \ref{def:GH}. Part \ref{SHORT.GH-4} will be proved in Section~\ref{sec:GH-approx}. The ``only-if''-part of \ref{SHORT.GH-2} will be proved in Section~\ref{sec:alm-isom}. Recall that $a\cdot\spc{X}$ denotes $\spc{X}$ \index{scaled space}\emph{scaled} by factor $a>0$; that is, $a\cdot\spc{X}$ is a metric space with the underlying set of $\spc{X}$ and the metric defined by \[\dist{x}{y}{a\cdot\spc{X}}\df a\cdot\dist{x}{y}{\spc{X}}.\] \begin{thm}{Exercise}\label{ex:d_GH-and-diam} Let $\spc{X}$ be a compact metric space, $\spc{P}$ be the one-point metric space. Prove that \begin{subthm}{ex:d_GH-and-diam:point} \[|\spc{X}-\spc{P}|_{\GH}=\tfrac12\cdot \diam \spc{X}.\] \end{subthm} \begin{subthm}{ex:d_GH-and-diam:scale} \[|a\cdot\spc{X}-b\cdot \spc{X}|_{\GH}=\tfrac12\cdot|a-b|\cdot\diam\spc{X}.\] \end{subthm} \end{thm} \begin{thm}{Exercise}\label{ex:rectangle} Let $\spc{A}_r$ be a rectangle $1$ by $r$ in the Euclidean plane and $\spc{B}_r$ be a closed line interval of length $r$. Show that \[|\spc{A}_r-\spc{B}_r|_{\GH}>\tfrac1{10}\] for all large $r$. \end{thm} \begin{thm}{Advanced exercise}\label{ex:GH-inj} Let $\spc{X}$ and $\spc{Y}$ be compact metric spaces; denote by $\hat{\spc{X}}$ and $\hat{\spc{Y}}$ their injective envelopes (see the definition on page \pageref{page:InjX}). Show that \[|\hat{\spc{X}}-\hat{\spc{Y}}|_{\GH}\le 2\cdot|\spc{X}- \spc{Y}|_{\GH}.\] \end{thm} \section{Approximations}\label{sec:GH-approx} \begin{thm}{Definition}\label{ex:defGHR} Let $\spc{X}$ and $\spc{Y}$ be two metric spaces. A relation $\approx$ between points in $\spc{X}$ and $\spc{Y}$ is called $\eps$-approximation if the following conditions hold: \begin{itemize} \item For any $x\in \spc{X}$ there is $y\in \spc{Y}$ such that $x\approx y$. \item For any $y\in \spc{Y}$ there is $x\in \spc{X}$ such that $x\approx y$. \item If for some $x, x'\in \spc{X}$ and $y,y'\in \spc{Y}$ we have $x\approx y$ and $x'\approx y'$, then \[\bigl|\dist{x}{x'}{\spc{X}}-\dist{y}{y'}{\spc{Y}}\bigr|<2\cdot\eps.\] \end{itemize} \end{thm} \begin{thm}{Exercise}\label{ex:H-R} Let $\spc{X}$ and $\spc{Y}$ be two compact metric spaces. Show that \[\dist{\spc{X}}{\spc{Y}}{\GH}<\eps\] if and only if there is an $\eps$-approximation between $\spc{X}$ and $\spc{Y}$. In other words $\dist{\spc{X}}{\spc{Y}}{\GH}$ is the greatest lower bound of values $\eps>0$ such that there is an $\eps$-approximation between $\spc{X}$ and $\spc{Y}$. \end{thm} \parit{Proof of \ref{GH-4}.} Suppose that \begin{itemize} \item $\approx_1$ is a relation between points in $\spc{X}$ and $\spc{Y}$, \item $\approx_2$ is a relation between points in $\spc{Y}$ and $\spc{Z}$. \end{itemize} Consider the relation $\approx_3$ between points in $\spc{X}$ and $\spc{Z}$ such that $x\approx_3 z$ if and only if there is $y\in \spc{Y}$ such that $x\approx_1 y$ and $y\approx_2 z$. It is straightforward to check that if $\approx_1$ is an $\eps_1$-approximation and $\approx_2$ is an $\eps_2$-approximation, then $\approx_3$ is an $(\eps_1+\eps_2)$-approximation. Applying \ref{ex:H-R}, we get that if \[|\spc{X}-\spc{Y}|_{\GH}<\eps_1 \quad\text{and}\quad |\spc{Y}-\spc{Z}|_{\GH}<\eps_2 \] then \[|\spc{X}-\spc{Z}|_{\GH}<\eps_1+\eps_2.\] Hence \ref{GH-4} follows. \qeds \section{Almost isometries}\label{sec:alm-isom} \begin{thm}{Definition} Let $\spc{X}$ and $\spc{Y}$ be metric spaces and $\eps>0$. A map\footnote{possibly noncontinuous} $f\: \spc{X} \z\to \spc{Y}$ is called an \index{almost isometry}\emph{$\eps$-isometry} if $f(\spc{X})$ is an $\eps$-net in $\spc{Y}$ and \[\bigl|\dist{x}{x'}{\spc{X}}-\dist{f(x)}{f(x')}{\spc{Y}}\bigr|<\eps.\] for any $x,x'\in \spc{X}$. \end{thm} \begin{thm}{Exercise}\label{ex:eps-isom} Let $\spc{X}$ and $\spc{Y}$ be compact metric spaces. \begin{subthm}{ex:eps-isom:GH>isom} If $\dist{\spc{X}}{\spc{Y}}{\GH}<\eps$, then there is a $2\cdot\eps$-isometry $f\:\spc{X}\to\spc{Y}$. \end{subthm} \begin{subthm}{ex:eps-isom:isom>GH} If there is an $\eps$-isometry $f\:\spc{X}\to\spc{Y}$, then $\dist{\spc{X}}{\spc{Y}}{\GH}<\eps$. \end{subthm} \end{thm} \parit{Proof of the ``only if''-part in \ref{GH-2}.} \label{page:GH-2-proof} Let $\spc{X}$ and $\spc{Y}$ be compact metric spaces. Suppose that $\dist{\spc{X}}{\spc{Y}}{\GH}<\eps$ for any $\eps>0$; we need to show that there is an isometry $\spc{X}\to\spc{Y}$. By \ref{ex:eps-isom:GH>isom}, for each positive integer $n$, we can choose a $\tfrac1n$-isometry $f_n\:\spc{X}\to\spc{Y}$. Since $\spc{X}$ is compact, we can choose a countable dense set $S$ in~$\spc{X}$. Applying the diagonal procedure if necessary, we can assume that for every $x \in S$ the sequence $(f_n(x))$ converges in $\spc{Y}$. Consider the pointwise limit map $f_\infty \: S \to \spc{Y}$, $$f_\infty(x) \df \lim_{n\to\infty} f_n (x)$$ for every $x \in S$. Since $$|f_n (x)- f_n (x')|_{\spc{Y}}\lg |x- x'|_\spc{X} \pm\tfrac1n,$$ we have $$|f_\infty(x)-f_\infty (x')|_{\spc{Y}} = \lim_{n\to\infty} |f_n(x)-f_n (x')|_{\spc{Y}} = |x -x'|_\spc{X}$$ for all $x, x' \in S$; that is, the map $f_\infty\:S\to \spc{Y}$ is distance-preserving. Therefore, $f_\infty$ can be extended to a distance-preserving map from the whole $\spc{X}$ to $\spc{Y}$. The latter can be done by setting $$f_\infty(x)=\lim_{n\to\infty} f_\infty(x_n)$$ for some sequence of points $(x_n)$ in $S$ that converges to $x$ in $\spc{X}$. Indeed, if $x_n\to x$, then $(x_n)$ is Cauchy. Since $f_\infty$ is distance-preserving, $y_n=f_\infty(x_n)$ is also a Cauchy sequence in $\spc{Y}$; therefore it converges. It remains to observe that this construction does not depend on the choice of the sequence $(x_n)$. This way we obtain a distance-preserving map $f_\infty\:\spc{X}\to \spc{Y}$. It remains to show that $f_\infty$ is surjective; that is, $f_\infty(\spc{X})=\spc{Y}$. The same argument produces a distance-preserving map $g_\infty\:\spc{Y}\z\to \spc{X}$. If $f_\infty$ is not surjective, then neither is the composition $f_\infty\z\circ g_\infty\:\spc{Y}\to \spc{Y}$. So $f_\infty \z\circ g_\infty$ is a distance-preserving map from a compact space to itself which is not an isometry. The latter contradicts \ref{ex:non-contracting-map}. \qeds \section{Convergence} The Gromov--Hausdorff metric is used to define Gromov--Hausdorff convergence. Namely, a sequence of compact metric spaces $\spc{X}_n$ converges to compact metric spaces $\spc{X}_\infty$ in the sense of Gromov--Hausdorff if \[\dist{\spc{X}_n}{\spc{X}_\infty}{\GH}\to 0\quad\text{as}\quad n\to\infty.\] This convergence is more important than the metric --- in all applications, we use only the topology on $\GH$ and we do not care about the particular value of Gromov--Hausdorff distance between spaces. The following observation follows from \ref{ex:eps-isom}: \begin{thm}{Observation}\label{obs:GH-e-isom} A sequence of compact metric spaces $(\spc{X}_n)$ converges to $\spc{X}_\infty$ in the sense of Gromov--Hausdorff if and only if there is a sequence $\eps_n\to0+$ and an $\eps_n$-isometry $f_n\:\spc{X}_n\to \spc{X}_\infty$ for each $n$. \end{thm} In the following exercises {}\emph{converge} means {}\emph{converge in the sense of Gromov--Hausdorff}. \begin{thm}{Exercise}\label{ex:GH-SC} \begin{subthm}{ex:GH-SC:circle} Show that a sequence of compact simply connected length spaces cannot converge to a circle. \end{subthm} \begin{subthm}{ex:GH-SC:nonsc-limit} Construct a sequence of compact simply connected length spaces that converges to a compact non-simply connected space. \end{subthm} \end{thm} \begin{thm}{Exercise}\label{ex:sphere-to-ball} \begin{subthm}{ex:sphere-to-ball:2} Show that a sequence of length metrics on the 2-sphere cannot converge to the unit disk. \end{subthm} \begin{subthm}{ex:sphere-to-ball:3} Construct a sequence of length metrics on the 3-sphere that converges to a unit 3-ball. \end{subthm} \end{thm} Given two metric spaces $\spc{X}$ and $\spc{Y}$, we will write $\spc{X}\le \spc{Y}$ if there is a noncontracting map $f\:\spc{X}\to \spc{Y}$; that is, if $$ |x-x'|_{\spc{X}}\le|f(x)-f(x')|_{\spc{Y}}$$ for any $x,x'\in \spc{X}$. Further, given $\eps>0$, we will write $\spc{X}\le \spc{Y}+\eps$ if there is a map $f\:\spc{X}\to \spc{Y}$ such that $$|x-x'|_{\spc{X}}\le|f(x)-f(x')|_{\spc{Y}}+\eps$$ for any $x,x'\in \spc{X}$. \section{Uniformly totally bonded families} \begin{thm}{Definition}\label{def:utb} A family $\spc{Q}$ of (isometry classes) of compact metric spaces is called \index{uniformly totally bonded family}\emph{uniformly totally bonded} if it meets the following two conditions: \begin{subthm}{} spaces in $\spc{Q}$ have uniformly bounded diameters; that is, there is $D\in\RR$ such that \[\diam\spc{X}\le D\] for any space $\spc{X}$ in $\spc{Q}$. \end{subthm} \begin{subthm}{} For any $\eps>0$ there is $n\in\NN$ such that any space $\spc{X}$ in $\spc{Q}$ admits an $\eps$-net with at most $n$ points. \end{subthm} \end{thm} \begin{thm}{Exercise}\label{ex:utb+pack} Let $\spc{Q}$ be a family of compact spaces with uniformly bounded diameters. Show that $\spc{Q}$ is uniformly totally bonded if for any $\eps>0$ there is $n\in\NN$ such that \[\pack_\eps\spc{X}\le n\] for any space $\spc{X}$ in $\spc{Q}$. \end{thm} Fix a real constant $C$. A Borel measure $\mu$ on a metric space $\spc{X}$ is called \index{doubling space}\emph{$C$-doubling} if \[\mu[\oBall(p,2\cdot r)]< C\cdot\mu[\oBall(p,r)]\] for any point $p\in \spc{X}$ and any $r>0$. A Borel measure is called \index{doubling measure}\emph{doubling} if it is {}\emph{$C$-doubling} for some real constant $C$. \begin{thm}{Exercise}\label{pr:doubling} Let $\spc{Q}(C,D)$ be the set of all the compact metric spaces with diameter at most $D$ that admit a $C$-doubling measure. Show that $\spc{Q}(C,D)$ is totally bounded. \end{thm} Recall that we write $\spc{X}\le\spc{Y}$ if there is a distance-nondecreasing map $\spc{X}\to\spc{Y}$. \begin{thm}{Exercise}\label{pr:under} \begin{subthm}{pr:under:if} Let $\spc{Y}$ be a compact metric space. Show that the set of all spaces $\spc{X}$ such that $\spc{X}\le\spc{Y}$ is uniformly totally bounded. \end{subthm} \begin{subthm}{pr:under:only-if} Show that for any uniformly totally bounded set $\spc{Q}\subset\GH$ there is a compact space $\spc{Y}$ such that $\spc{X}\le\spc{Y}$ for any $\spc{X}$ in $\spc{Q}$. \end{subthm} \end{thm} \section{Gromov's selection theorem} The following theorem is analogous to Blaschke selection theorems (\ref{thm:compact+Hausdorff}). \begin{thm}{Gromov selection theorem}\label{thm:gromov-compactness} Let $\spc{Q}$ be a closed subset of $\GH$. Then $\spc{Q}$ is compact if and only if it is totally bounded. \end{thm} \begin{thm}{Lemma}\label{lem:GH-complete} The space $\GH$ is complete. \end{thm} Let us define gluing of metric spaces that will be used in the proof of the lemma. Suppose $\spc{U}$ and $\spc{V}$ are metric spaces with isometric closed sets $A\subset\spc{U}$ and $A'\subset\spc{V}$; let $\iota\:A\to A'$ be an isometry. Consider the space $\spc{W}$ of all equivalence classes in $\spc{U}\sqcup\spc{V}$ with the equivalence relation given by $a\sim\iota(a)$ for any $a\in A$. It is straightforward to check that the following defines a metric on~$\spc{W}$: \begin{align*} \dist{u}{u'}{\spc{W}}&\df\dist{u}{u'}{\spc{U}} \\ \dist{v}{v'}{\spc{W}}&\df\dist{v}{v'}{\spc{V}} \\ \dist{u}{v}{\spc{W}}&\df\min\set{\dist{u}{a}{\spc{U}}+\dist{v}{\iota(a)}{\spc{V}}}{a\in A} \end{align*} where $u,u'\in \spc{U}$ and $v,v'\in \spc{V}$. The space $\spc{W}$ is called the \index{gluing}\emph{gluing} of $\spc{U}$ and $\spc{V}$ along~$\iota$; briefly, we can write $\spc{W}=\spc{U}\sqcup_\iota\spc{V}$. If one applies this construction to two copies of one space $\spc{U}$ with a set $A\subset \spc{U}$ and the identity map $\iota\:A\to A$, then the obtained space is called the \index{double}\emph{double} of $\spc{U}$ along~$A$; this space can be denoted by $\sqcup_A^2\spc{U}$. Note that the inclusions $\spc{U}\hookrightarrow \spc{W}$ and $\spc{V}\hookrightarrow \spc{W}$ are distance preserving. Therefore we can and will conside $\spc{U}$ and $\spc{V}$ as the subspaces of $\spc{W}$; this way the subsets $A$ and $A'$ will be identified and denoted further by~$A$. Note that $A=\spc{U}\cap \spc{V}\subset \spc{W}$. \parit{Proof.} Let $(\spc{X}_n)$ be a Cauchy sequence in $\GH$. Passing to a subsequence if necessary, we can assume that $|\spc{X}_n-\spc{X}_{n+1}|_{\GH}<\tfrac1{2^n}$ for each~$n$. In particular, for each $n$ there is a metric space $\spc{V}_n$ with distance preserving inclusions $\spc{X}_n\hookrightarrow \spc{V}_n$ and $\spc{X}_{n+1}\hookrightarrow \spc{V}_n$ such that \[|\spc{X}_n-\spc{X}_{n+1}|_{\Haus\spc{V}_n}<\tfrac1{2^n}\] for each $n$. Moreover, we may assume that $\spc{V}_n=\spc{X}_n\cup\spc{X}_{n+1}$. Let us glue $\spc{V}_1$ to $\spc{V}_2$ along $\spc{X}_2$; to the obtained space glue $\spc{V}_3$ along $\spc{X}_3$, and so on. The obtained metric space $\spc{W}$ has an underlying set formed by the disjoint union of all $\spc{X}_n$ such that each inclusion $\spc{X}_n\z\hookrightarrow\spc{W}$ is distance preserving and \[|\spc{X}_n-\spc{X}_{n+1}|_{\Haus\spc{W}}<\tfrac1{2^n}\] for each $n$. In particular, \[|\spc{X}_m-\spc{X}_n|_{\Haus\spc{W}}<\tfrac1{2^{n-1}}\eqlbl{eq:|x_m-X_n|}\] if $m>n$. Denote by $\bar{\spc{W}}$ the completion of $\spc{W}$. Observe that the union $\spc{X}_1\z\cup \spc{X}_2\cup\z\dots\cup \spc{X}_n$ is compact and \ref{eq:|x_m-X_n|} implies that it forms a $\tfrac1{2^{n-1}}$-net in $\bar{\spc{W}}$. Whence $\bar{\spc{W}}$ is compact; see \ref{totally-bounded} and \ref{ex:compact-net}. Applying Blaschke selection theorem (\ref{thm:compact+Hausdorff}), we can pass to a subsequence of $(\spc{X}_n)$ that converges in $\Haus\bar{\spc{W}}$; denote its limit by $\spc{X}_\infty$. It remains to observe that $\spc{X}_\infty$ is the Gromov--Hausdorff limit of $(\spc{X}_n)$. \qeds \parit{Proof of \ref{thm:gromov-compactness}; ``only if'' part.} Suppose that there is no sequence $\eps_n\to0$ as described in \ref{def:utb}. Observe that in this case there is a sequence of spaces $\spc{X}_n\in\spc{Q}$ such that $$\pack_\delta \spc{X}_n\to\infty \quad\text{as}\quad n\to\infty$$ for some fixed $\delta>0$. Since $\spc{Q}$ is compact, this sequence has a partial limit, say $\spc{X}_\infty\in\spc{Q}$. Observe that $\pack_{\delta} \spc{X}_\infty=\infty$. Therefore, $\spc{X}_\infty$ is not compact --- a contradiction. \parit{``If'' part.} Suppose sequence $(\eps_n)$ as in the definition of uniformly totally bonded families (\ref{def:utb}). Note that $\diam \spc{X}\le \eps_1$ for any $\spc{X}\in \spc{Q}$. Given a positive integer $n$ consider the set of all metric spaces $\spc{W}_n$ with the number of points at most $n$ and diameter $\le \eps_1$. Note that $\spc{W}_n$ is a compact set in $\GH$ for each $n$. Further, a subspace formed by a maximal $\eps_n$-net of any $\spc{X}\in\spc{Q}$ belongs to $\spc{W}_n$. Therefore, $\spc{W}_n\cap\spc{Q}$ is a compact $\eps_n$-net in $\spc{Q}$. That is, $\spc{Q}$ has a compact $\eps$-net for any $\eps>0$. Since $\spc{Q}$ is closed in a complete space $\GH$, it implies that $\spc{Q}$ is compact. \qeds \begin{thm}{Exercise}\label{ex-GH-length} Show that the space $\GH$ is \begin{subthm}{ex-GH-length:length} length, \end{subthm} \begin{subthm}{ex-GH-length:geodesic} geodesic. \end{subthm} \end{thm} \begin{thm}{Exercise}\label{ex:GH-po} \begin{subthm}{ex:GH-po:a} Show that $$\dist{\spc{X}}{\spc{Y}}{\GH'}=\inf\set{\eps>0}{\spc{X}\le \spc{Y}+\eps \quad\text{and}\quad \spc{Y}\le \spc{X}+\eps}$$ defines a metric on the space of (isometry classes) of compact metric spaces. \end{subthm} \begin{subthm}{ex:GH-po:b} Moreover $\dist{*}{*}{\GH'}$ is equivalent to the Gromov--Hausdorff metric; that is, $$|\spc{X}_n-\spc{X}_\infty|_{\GH}\to 0 \quad\iff\quad \dist{\spc{X}_n}{\spc{X}_\infty}{\GH'}\to 0$$ as $n\to\infty$. \end{subthm} \end{thm} \section{Universal ambient space} Recall that a metric space is called universal if it contains an isometric copy of any separable metric space (in particular, any compact metric space). Examples of universal spaces include Urysohn space and $\ell^\infty$ --- the space of bounded infinite sequences with the metric defined by $\sup$-norm; see \ref{prop:sep-in-urys} and \ref{ex:frechet}. The following proposition says that the space $\spc{W}$ in Definition~\ref{def:GH} can be exchanged to a fixed universal space. \begin{thm}{Proposition}\label{prop:GH-with-fixed-Z} Let $\spc{U}$ be a universal space. Then for any compact metric spaces $\spc{X}$ and $\spc{Y}$ we have $$|\spc{X}-\spc{Y}|_{\GH} = \inf \{|\spc{X}'-\spc{Y}'|_{\Haus\spc{U}}\}$$ where the greatest lower bound is taken over all pairs of sets $\spc{X}'$ and $\spc{Y}'$ in $\spc{U}$ which isometric to $\spc{X}$ and $\spc{Y}$ respectively. \end{thm} \parit{Proof of \ref{prop:GH-with-fixed-Z}.} By the definition (\ref{def:GH}), we have that \[|\spc{X}-\spc{Y}|_{\GH} \le \inf \{|\spc{X}'-\spc{Y}'|_{\Haus\spc{U}}\};\] it remains to prove the opposite inequality. Suppse $|\spc{X}-\spc{Y}|_{\GH}<\eps$; let $\spc{X}'$, $\spc{Y}'$ and $\spc{Z}$ be as in \ref{def:GH}. We can assume that $\spc{Z}=\spc{X}'\cup\spc{Y}'$; otherwise pass to the subspace $\spc{X}'\cup\spc{Y}'$ of~$\spc{Z}$. In this case, $\spc{Z}$ is compact; in particular, it is separable. Since $\spc{U}$ is universal, there is a distance-preserving embedding of $\spc{Z}$ in $\spc{U}$; let us keep the same notation for $\spc{X}'$, $\spc{Y}'$, and their images. It follows that \[|\spc{X}'-\spc{Y}'|_{\Haus\spc{U}}<\eps,\] --- hence the result. \qeds \begin{thm}{Exercise}\label{ex:GH-urysohn} Let $\spc{U}_\infty$ be the Urysohn space. Given two compact set $A$ and $B$ in $\spc{U}_\infty$ define \[\|A-B\|=\inf\{|A-\iota(B)|_{\Haus\spc{U}_\infty}\},\] where the greatest lower bound is taken for all isometrics $\iota$ of $\spc{U}_\infty$. Show that $\|{*}\z-{*}\|$ defines a pseudometric% \footnote{The value $\|A-B\|$ is called \emph{Hausdorff distance up to isometry} from $A$ to $B$ in $\spc{U}_\infty$.} on nonempty compact subsets of $\spc{U}_\infty$ and its corresponding metric space is isometric to $\GH$. \end{thm} \section{Remarks} Suppose $\spc{X}_n\GHto \spc{X}_\infty$, then there is a metric on the disjoint union \[\bm{X}=\bigsqcup_{n\in \NN\cup\{\infty\}} \spc{X}_n\] that satisfies the following property: \begin{thm}{Property}\label{propery:GH} The restriction of metric on each $\spc{X}_n$ and $\spc{X}_\infty$ coincides with its original metric and $\spc{X}_n\Hto \spc{X}_\infty$ as subsets in $\bm{X}$. \end{thm} Indeed, since $\spc{X}_n\GHto \spc{X}_\infty$, there is a metric on $\spc{V}_n=\spc{X}_n\sqcup \spc{X}_\infty$ such that the restriction of metric on each $\spc{X}_n$ and $\spc{X}_\infty$ coincides with its original metric and $\dist{\spc{X}_n}{\spc{X}_\infty}{\Haus\spc{V}_n}<\eps_n$ for some sequence $\eps_n\to 0$. Gluing all $\spc{V}_n$ along $\spc{X}_\infty$, we obtain the required space $\bm{X}$. In other words, the metric on $\bm{X}$ defines the convergence $\spc{X}_n\z\GHto \spc{X}_\infty$. This metric makes it possible to talk about limits of sequences $x_n\in \spc{X}_n$ as $n\to\infty$, as well as weak limits of a sequence of Borel measures $\mu_n$ on $\spc{X}_n$ and so on. For that reason, it is useful to define convergence by specifying the metric on $\bm{X}$ that satisfies the property for the variation of Hausdorff convergence described in Section~\ref{sec:H-variation}. This approach is very flexible; in particular, it can be used to define Gromov--Hausdorff convergence of arbitrary metric spaces (net necessarily compact). In this case, a limit space for this generalized convergence is not uniquely defined. \begin{figure}[h!] \vskip-0mm \centering \includegraphics{mppics/pic-500} \end{figure} For example, if each space $\spc{X}_n$ in the sequence is isometric to the half-line, then its limit might be isometric to the half-line or the whole line. The first convergence is evident and the second could be guessed from the diagram. Often the isometry class of the limit can be fixed by marking a point $p_n$ in each space $\spc{X}_n$, it is called \index{pointed convergence}\emph{pointed Gromov--Hausdorff convergence} --- we say that $(\spc{X}_n,p_n)$ converges to $(\spc{X}_\infty,p_\infty)$ if there is a metric on $\bm{X}$ such that $\spc{X}_n\Hto \spc{X}_\infty$ and $p_n\to p_\infty$. For example, the sequence $(\spc{X}_n,p_n)=(\RR_+,0)$ converges to $(\RR_+,0)$, while $(\spc{X}_n,p_n)=(\RR_+,n)$ converges to $(\RR,0)$. The pointed convergence works nicely only for proper metric spaces; the following theorem is an analog of Gromov's selection theorem for this convergence. \begin{thm}{Theorem}\label{thm:pointed-gromov-compactness}% Let $\spc{Q}$ be a set of isometry classes of pointed proper metric spaces $(\spc{X},p)$. Assume that for any $R>0$, the $R$-balls in the spaces centered at the marked points form a uniformly totally bounded family of spaces. Then $\spc{Q}$ is precompact with respect to pointed Gromov--Hausdorff convergence. \end{thm} \chapter{Injective spaces} {}\emph{Injective spaces} (also known as {}\emph{hyperconvex spaces}) are the metric analog of convex sets. This lecture is based on a paper by John Isbell \cite{isbell}. \section{Admissible and extremal functions} Let $\spc{X}$ be a metric space. A function $r\:\spc{X}\to\RR$ is called \label{page:admissible function}\index{admissible function}\emph{admissible} if the following inequality \[r(x)+r(y)\ge \dist{x}{y}{\spc{X}}\eqlbl{eq:admissible}\] holds for any $x,y\in \spc{X}$. \begin{thm}{Observation}\label{obs:admissible} \begin{subthm}{obs:admissible:nonnegative} Any admissible function is nonnegative. \end{subthm} \begin{subthm}{obs:admissible:balls} If $\spc{X}$ is a geodesic space, then a function $r\:\spc{X}\to\RR$ is admissible if and only if \[\cBall[x,r(x)]\cap\cBall[y,r(y)]\ne \emptyset\] for any $x,y\in \spc{X}$. \end{subthm} \end{thm} \parit{Proof.} For \ref{SHORT.obs:admissible:nonnegative}, take $x=y$ in \ref{eq:admissible}. Part \ref{SHORT.obs:admissible:balls} follows from the triangle inequality and the existence of a geodesic $[xy]$. \qeds A minimal admissible function will be called \label{page:extremal function}\index{extremal function}\emph{extremal}. More precisely, an admissible function $r\:\spc{X}\to\RR$ is extremal if for any admissible function $s\:\spc{X}\to\RR$ we have \[s\le r\quad\Longrightarrow\quad s=r.\] \begin{thm}{Key exercise}\label{ex:+-c} Let $r$ be an extremal function and $s$ an admissible function on a metric space $\spc{X}$. Suppose that $r\ge s-c$ for some constant $c$. Show that $c\ge 0$ and $r\le s+c$. \end{thm} \begin{thm}{Observations}\label{obs:extremal} Let $\spc{X}$ be a metric space. \begin{subthm}{obs:extremal:distfun} For any point $p\in\spc{X}$ the distance function $r\z=\distfun_p$ is extremal. \end{subthm} \begin{subthm}{lem:extremal-lipschitz} Any extremal function $r$ on $\spc{X}$ is \index{1-Lipschitz function}\emph{1-Lipschitz}; that is, \[|r(p)-r(q)|\le \dist{p}{q}{}\] for any $p,q\in\spc{X}$. In other words, any extremal function is an extension function; see the definition on page \pageref{page:extension function}. \end{subthm} \begin{subthm}{lem:opposite} Let $r$ be an extremal function on $\spc{X}$. Then for any point $p\in\spc{X}$ and any $\delta>0$, there is a point $q\in \spc{X}$ such that \[r(p)+r(q)<\dist{p}{q}{\spc{X}}+\delta.\] Moreover, if $\spc{X}$ is compact, then there is $q$ such that \[r(p)+r(q)=\dist{p}{q}{\spc{X}}.\] \end{subthm} \begin{subthm}{obs:extremal:below} For any admissible function $s$ there is an extremal function $r$ such that $r\le s$. \end{subthm} \end{thm} \parit{Proof; \ref{SHORT.obs:extremal:distfun}.} By the triangle inequality, \ref{eq:admissible} holds; that is, $r=\distfun_p$ is an admissible function. Further, if $s\le r$ is another admissible function, then $s(p)=0$ and \ref{eq:admissible} implies that $s(x)\z\ge\dist{p}{x}{}$. Whence $s=r$. \parit{\ref{SHORT.lem:extremal-lipschitz}.} By \ref{SHORT.obs:extremal:distfun}, $\distfun_p$ is admissible. Since $r$ is admissible, we have that \[r\ge \distfun_p-r(p).\] Since $r$ is extremal, \ref{ex:+-c} implies that \[r\le \distfun_p+r(p),\] or, equivalently, \[r(q)-r(p)\le \dist{p}{q}{}\] for any $p,q\in\spc{X}$. The same way we can show that $r(p)-r(q)\le \dist{p}{q}{}$. Whence the statement follows. \parit{\ref{SHORT.lem:opposite}.} Again, by \ref{SHORT.obs:extremal:distfun}, $\distfun_p$ is an extremal function. Arguing by contradiction, assume \[r(q)\ge \distfun_p(q)-r(p)+\delta\] for any $q$. By \ref{ex:+-c}, we get that \[r(q)\le \distfun_p(q)+r(p)-\delta\] for any $q$. Taking $q=p$, we get $r(p)\le r(p)-\delta$, a contradiction. Suppose $\spc{X}$ is compact. Denote by $q_n$ the point provided by the first part of \ref{SHORT.lem:opposite} for $\delta=\tfrac1n$. Let $q$ be a partial limit of $q_n$. Then \[r(p)+r(q)\le\dist{p}{q}{\spc{X}}.\] Since $r$ is admissible, the opposite inequality holds; whence the second part of \ref{SHORT.lem:opposite} follows. \parit{\ref{SHORT.obs:extremal:below}.} Follows by Zorn's lemma. \qeds \section{Injective spaces} \begin{thm}{Definition}\label{def:injective} A metric space $\spc{Y}$ is called \index{injective space}\emph{injective} if for any metric space $\spc{X}$ and any of its subspaces $\spc{A}$ any short map $f\:\spc{A}\to \spc{Y}$ can be extended to a short map $F\:\spc{X}\to \spc{Y}$; that is, $f=F|_{\spc{A}}$. \end{thm} \begin{thm}{Exercise}\label{ex:inj=complete-geodesic-contractible} Show that any injective space is \begin{subthm}{ex:inj=complete-geodesic-contractible:complete} complete, \end{subthm} \begin{subthm}{ex:inj=complete-geodesic-contractible:geodesic} geodesic, and \end{subthm} \begin{subthm}{ex:inj=complete-geodesic-contractible:contractible} contractible. \end{subthm} \end{thm} \begin{thm}{Exercise}\label{ex:injective-spaces} Show that the following spaces are injective: \begin{subthm}{ex:injective-spaces:R} the real line; \end{subthm} \begin{subthm}{ex:injective-spaces:tree} complete metric tree; \end{subthm} \begin{subthm}{ex:injective-spaces:ell-infty} coordinate plane with the metric induced by the $\ell^\infty$-norm. \end{subthm} \end{thm} The following exercise deals with metric spaces that in some sense dual to the injective spaces. \begin{thm}{Exercise}\label{ex:ultrametric} Suppose that a metric space $\spc{X}$ satisfies the following property: For any subspace $\spc{A}$ in $\spc{X}$ and any other metric space $\spc{Y}$, any short map $f\:\spc{A}\to \spc{Y}$ can be extended to a short map $F\:\spc{X}\to \spc{Y}$. Show that $\spc{X}$ is an \index{ultrametric space}\emph{ultrametric space}; that is, the following strong version of the triangle inequality \[\dist{x}{z}{\spc{X}} \le \max\{\,\dist{x}{y}{\spc{X}},\dist{y}{z}{\spc{X}}\,\}\] holds for any three points $x,y,z\in \spc{X}$. \end{thm} \begin{thm}{Theorem}\label{thm:injective=hyperconvex} For any metric space $\spc{Y}$ the following condition are equivalent: \begin{subthm}{thm:injective=hyperconvex:injective} $\spc{Y}$ is injective \end{subthm} \begin{subthm}{thm:injective=hyperconvex:extremal} If $r\:\spc{Y}\to\RR$ is an extremal function, then there is a point $p\in \spc{Y}$ such that \[\dist{p}{x}{}\le r(x)\] for any $x\in \spc{Y}$. \end{subthm} \begin{subthm}{thm:injective=hyperconvex:balls} $\spc{Y}$ is \index{hyperconvex space}\emph{hyperconvex}; that is, if $\{\cBall[x_\alpha,r_\alpha]\}_{\alpha\in\IndexSet}$ is a family of closed balls in $\spc{Y}$ such that \[r_\alpha+r_\beta\ge \dist{x_\alpha}{x_\beta}{}\] for any $\alpha,\beta\in \IndexSet$, then all the balls in the family $\{\cBall[x_\alpha,r_\alpha]\}_{\alpha\in\IndexSet}$ have a common point. \end{subthm} \end{thm} \parit{Proof.} We will prove implications \ref{SHORT.thm:injective=hyperconvex:injective}$\Rightarrow$\ref{SHORT.thm:injective=hyperconvex:extremal}$\Rightarrow$\ref{SHORT.thm:injective=hyperconvex:balls}$\Rightarrow$\ref{SHORT.thm:injective=hyperconvex:injective}. \parit{\ref{SHORT.thm:injective=hyperconvex:injective}$\Rightarrow$\ref{SHORT.thm:injective=hyperconvex:extremal}.} Let us apply the definition of injective space to a one-point extension of $\spc{Y}$. It follows that for any extension function $r\:\spc{Y}\to\RR$ there is a point $p\in \spc{Y}$ such that \[\dist{p}{x}{}\le r(x)\] for any $x\in \spc{Y}$. By \ref{lem:extremal-lipschitz}, any extremal function is an extension function, whence the implication follows. \parit{\ref{SHORT.thm:injective=hyperconvex:extremal}$\Rightarrow$\ref{SHORT.thm:injective=hyperconvex:balls}.} By \ref{obs:admissible:balls}, part \ref{SHORT.thm:injective=hyperconvex:balls} is equivalent to the following statement: \begin{itemize} \item If $r\:\spc{Y}\to\RR$ is an admissible function, then there is a point $p\in \spc{Y}$ such that \[\dist{p}{x}{}\le r(x)\eqlbl{eq:|p-x|=<r(x)}\] for any $x\in \spc{Y}$. \end{itemize} Indeed, set $r(x)\df\inf\set{r_\alpha}{x_\alpha=x}$. (If $x_\alpha\ne x$ for any $\alpha$, then $r(x)=\infty$.) The condition in \ref{SHORT.thm:injective=hyperconvex:balls} implies that $r$ is admissible. It remains to observe that $p\in \cBall[x_\alpha,r_\alpha]$ for every $\alpha$ if and only if \ref{eq:|p-x|=<r(x)} holds. By \ref{obs:extremal:below}, for any admissible function $r$ there is an extremal function $\bar r\le r$; whence \ref{SHORT.thm:injective=hyperconvex:extremal}$\Rightarrow$\ref{SHORT.thm:injective=hyperconvex:balls}. \parit{\ref{SHORT.thm:injective=hyperconvex:balls}$\Rightarrow$\ref{SHORT.thm:injective=hyperconvex:injective}.} Arguing by contradiction, suppose $\spc{Y}$ is not injective; that is, there is a metric space $\spc{X}$ with a subset $\spc{A}$ such that a short map $f\:\spc{A}\to \spc{Y}$ cannot be extended to a short map $F\:\spc{X}\to \spc{Y}$. By Zorn's lemma, we may assume that $\spc{A}$ is a maximal subset; that is, the domain of $f$ cannot be enlarged by a single point.% \footnote{In this case, $\spc{A}$ must be closed, but we will not use it.} Fix a point $p$ in the complement $\spc{X}\backslash \spc{A}$. To extend $f$ to $p$, we need to choose $f(p)$ in the intersection of the balls $\cBall[f(x),r(x)]$, where $r(x)=\dist{p}{x}{}$. Therefore, this intersection for all $x\in \spc{A}$ has to be empty. Since $f$ is short, we have that \begin{align*} r(x)+r(y)&\ge \dist{x}{y}{\spc{X}}\ge \\ &\ge \dist{f(x)}{f(y)}{\spc{Y}}. \end{align*} Therefore, by \ref{SHORT.thm:injective=hyperconvex:balls} the balls $\cBall[f(x),r(x)]$ have a common point --- a contradiction. \qeds \begin{thm}{Exercise}\label{ex:one-point-gluing} Suppose a length space $\spc{W}$ has two subspaces $\spc{X}$ and $\spc{Y}$ such that $\spc{X}\cup\spc{Y}=\spc{W}$ and $\spc{X}\cap\spc{Y}$ is a one-point set. Assume $\spc{X}$ and $\spc{Y}$ are injective. Show that $\spc{W}$ is injective \end{thm} \begin{thm}{Exercise}\label{ex:urysohn-hyperconvex} Show that the $d$-Urysohn space is {}\emph{finitely hyperconvex} but not {}\emph{countably hyperconvex}; that is, the condition in \ref{thm:injective=hyperconvex:balls} holds for any finite family of balls, but may not hold for a countable family. Conclude that the $d$-Urysohn space is not injective. Try to do the same for the Urysohn space. \end{thm} \section{Space of extremal functions} Let $\spc{X}$ be a metric space. Consider the space $\Inj \spc{X}$ of extremal functions on $\spc{X}$ equipped with sup-norm; \label{page:InjX} that is, \[\dist{f}{g}{\Inj \spc{X}}\df\sup\set{|f(x)-g(x)|}{x\in \spc{X}}.\] Recall that by \ref{obs:extremal:distfun}, any distance function is extremal. It follows that the map $x\mapsto \distfun_x$ produces a distance-preserving embedding $\spc{X}\hookrightarrow\Inj \spc{X}$. So we can (and will) treat $\spc{X}$ as a subspace of $\Inj \spc{X}$, or, equivalently, $\Inj \spc{X}$ as an extension of $\spc{X}$. Since any extremal function is 1-Lipschitz, for any $f\in \Inj \spc{X}$ and $p\in \spc{X}$, we have that $f(x)\le f(p)+\distfun_p(x)$. By \ref{ex:+-c}, we also get $f(x)\ge -f(p)+\distfun_p(x)$. Therefore \[ \begin{aligned} \dist{f}{p}{\Inj \spc{X}}&=\sup\set{|f(x)-\distfun_p(x)|}{x\in \spc{X}}= \\ &=f(p). \end{aligned} \eqlbl{eq:f(p)=|f-p|} \] In particular, the statement in \ref{lem:opposite} can be written as \[\dist{f}{p}{\Inj\spc{X}}+\dist{f}{q}{\Inj\spc{X}}<\dist{p}{q}{\Inj\spc{X}}+\delta.\] \begin{thm}{Exercise}\label{ex:Inj(compact)} Let $\spc{X}$ be a metric space. Show that $\Inj\spc{X}$ is compact if and only if so is $\spc{X}$. \end{thm} \begin{thm}{Exercise}\label{ex:tripod+square} Describe the set of all extremal functions on a metric space $\spc{X}$ and the metric space $\Inj \spc{X}$ in each of the following case: \begin{subthm}{ex:tripod+square:tripod} $\spc{X}$ is a metric space with exactly three points $a,b,c$ such that \[\dist{a}{b}{\spc{X}}=\dist{b}{c}{\spc{X}}=\dist{c}{a}{\spc{X}}=1.\] \end{subthm} \begin{subthm}{ex:tripod+square:square} $\spc{X}$ is a metric space with exactly four points $p,q,x,y$ such that \[\dist{p}{x}{\spc{X}}=\dist{p}{y}{\spc{X}}=\dist{q}{x}{\spc{X}}=\dist{q}{x}{\spc{X}}=1\] and \[\dist{p}{q}{\spc{X}}=\dist{x}{y}{\spc{X}}=2.\] \end{subthm} \end{thm} \begin{thm}{Proposition}\label{prop:InjX-is-injective} For any metric space $\spc{X}$, its extension $\Inj\spc{X}$ is injective. \end{thm} \begin{thm}{Lemma}\label{lem:r|X-extremal} Let $\spc{X}$ be a metric space. Suppose that $r$ is an extremal function on $\Inj \spc{X}$. Then $r|_\spc{X}\in \Inj \spc{X}$; that is, the restriction of $r$ to $\spc{X}$ is an extremal function. \end{thm} \parit{Proof.} Arguing by contradiction, suppose that there is an admissible function $s\:\spc{X}\to \RR$ such that $s(x)\le r(x)$ for any $x\in\spc{X}$ and $s(p)\z< r(p)$ for some point $p\in\spc{X}$. Consider another function $\bar r\:\Inj \spc{X}\to\RR$ such that $\bar r(f)\df r(f)$ if $f\ne p$ and $\bar r(p)\df s(p)$. Let us show that $\bar r$ is admissible; that is, \[\dist{f}{g}{\Inj \spc{X}}\le\bar r(f)+\bar r(g) \eqlbl{r-admissible}\] for any $f,g\in \Inj \spc{X}$. Since $r$ is admissible and $\bar r= r$ on $(\Inj \spc{X})\backslash \{p\}$, it is sufficient to prove \ref{r-admissible} if $f\ne g=p$. By \ref{eq:f(p)=|f-p|}, we have $\dist{f}{p}{\Inj \spc{X}}=f(p)$. Therefore, \ref{r-admissible} boils down to the following inequality \[r(f)+s(p)\ge f(p).\eqlbl{eq:r(f)+s(p)>=f(p)}\] for any $f\in\Inj \spc{X}$. Fix small $\delta>0$. Let $q\in\spc{X}$ be the point provided by \ref{lem:opposite}. Then \begin{align*} r(f)+s(p)&\ge [r(f)-r(q)]+[r(q)+s(p)]\ge \intertext{since $r$ is 1-Lipschitz, and $r(q)\ge s(q)$, we can continue} &\ge -\dist{q}{f}{\Inj \spc{X}}+[s(q)+s(p)]\ge \intertext{by \ref{eq:f(p)=|f-p|} and since $s$ is admissible} &\ge -f(q)+\dist{p}{q}{}> \intertext{by \ref{lem:opposite}} &> f(p)-\delta. \end{align*} Since $\delta>0$ is arbitrary, \ref{eq:r(f)+s(p)>=f(p)} and \ref{r-admissible} follow. Summarizing: the function $\bar r$ is admissible, $\bar r\le r$ and $\bar r(p)<r(p)$; that is, $r$ is not extremal --- a contradiction. \qeds \parit{Proof of \ref{prop:InjX-is-injective}.} Choose an extremal function $r\:\Inj\spc{X}\to\RR$. Set $s\z\df r|_{\spc{X}}$. By \ref{lem:r|X-extremal}, $s\in \Inj\spc{X}$; that is, $s$ is extremal. By \ref{thm:injective=hyperconvex:extremal}, it is sufficient to show that \[r(f)\ge\dist{s}{f}{\Inj\spc{X}} \eqlbl{eq:r(f)>=| r-f|}\] for any $f\in\Inj\spc{X}$. Since $r$ is $1$-Lipschitz (\ref{lem:extremal-lipschitz}) we have that \[ s(x)-f(x)=r(x)-\dist{f}{x}{\Inj \spc{X}}\le r(f). \] for any $x\in\spc{X}$. Since $r$ is admissible we have that \[ s(x)-f(x)=r(x)-\dist{f}{x}{\Inj \spc{X}}\ge -r(f). \] for any $x\in\spc{X}$. That is, $|s(x)-f(x)|\le r(f)$ for any $x\in \spc{X}$. Recall that \[\dist{s}{f}{\Inj \spc{X}} \df \sup\set{|s(x)-f(x)|}{x\in\spc{X}};\] hence \ref{eq:r(f)>=| r-f|} follows. \qeds \begin{thm}{Exercise}\label{ex:4-on-a-line} Let $\spc{X}$ be a compact metric space. Show that for any two points $f,g\in\Inj \spc{X}$ lie on a geodesic $[pq]$ with $p,q\in \spc{X}$. \end{thm} \section{Injective envelope} An extension $\spc{E}$ of a metric space $\spc{X}$ will be called its \index{injective envelope}\emph{injective envelope} if $\spc{E}$ is an injective space and there is no proper injective subspace of $\spc{E}$ that contains $\spc{X}$. Two injective envelopes $e\:\spc{X}\hookrightarrow \spc{E}$ and $f\:\spc{X}\hookrightarrow \spc{F}$ are called equivalent if there is an isometry $\iota\: \spc{E}\to\spc{F}$ such that $f=\iota\circ e$. \begin{thm}{Theorem}\label{thm:inj-envelope} For any metric space $\spc{X}$, its extension $\Inj\spc{X}$ is an injective envelope. Moreover, any other injective envelope of $\spc{X}$ is equivalent to $\Inj\spc{X}$. \end{thm} \parit{Proof.} Suppose $S\subset \Inj\spc{X}$ is an injective subspace containing $\spc{X}$. Since $S$ is injective, there is a short map $w\:\Inj\spc{X}\to S$ that fixes all points in $\spc{X}$. Suppose that $w\:f\mapsto f'$; observe that $f(x)\ge f'(x)$ for any $x\in \spc{X}$. Since $f$ is extremal, $f=f'$; that is, $w$ is the identity map and therefore $S=\Inj\spc{X}$. Assume we have another injective envelope $e\:\spc{X}\hookrightarrow \spc{E}$. Then there are short maps $v\:\spc{E}\to \Inj\spc{X}$ and $w\:\Inj\spc{X}\to \spc{E}$ such that $x=v\circ e(x)$ and $e(x)=w(x)$ for any $x\in\spc{X}$. From above, the composition $v\circ w$ is the identity on $\Inj\spc{X}$. In particular, $w$ is distance-preserving. The composition $w\circ v\:\spc{E}\to \spc{E}$ is a short map that fixes points in $e(\spc{X})$. Since $e\:\spc{X}\hookrightarrow \spc{E}$ is an injective envelope, the composition $w\circ v$ and therefore $w$ are onto. Whence $w$ is an isometry. \qeds \begin{thm}{Exercise}\label{ex:d-p-inclusion} Suppose $\spc{X}$ is a subspace of a metric space $\spc{U}$. Show that the inclusion $\spc{X}\hookrightarrow\spc{U}$ can be extended to a distance-preserving inclusion $\Inj\spc{X}\hookrightarrow\Inj\spc{U}$. \end{thm} \section{Remarks} Injective spaces were introduced by Nachman Aronszajn and Prom Panitchpakdi \cite{aronszajn-panitchpakdi}. The injective envelope was introduced by John Isbell \cite{isbell}. It was rediscovered a couple of times since then; as a result the injective envelope has many other names including \index{tight span}\emph{tight span} and \index{hyperconvex hull}\emph{hyperconvex hull}. \chapter{Definitions} In this lecture we give some conventions used further and remind some definitions related to metric spaces. We assume some prior knowledge of metric spaces. For a more detailed introduction, we recommend the first couple of chapters in the book by Dmitri Burago, Yuri Burago, and Sergei Ivanov \cite{burago-burago-ivanov}. \section{Metric spaces} \label{sec:metric spaces} The distance between two points $x$ and $y$ in a metric space $\spc{X}$ will be denoted by $\dist{x}{y}{}$ or $\dist{x}{y}{\spc{X}}$. The latter notation is used if we need to emphasize that the distance is taken in the space~${\spc{X}}$. Let us recall the definition of metric. \begin{thm}{Definition}\label{def:metric} A \index{metric}\emph{metric} on a set $\spc{X}$ is a real-valued function $(x,y)\z\mapsto\dist{x}{y}{\spc{X}}$ that satisfies the following conditions for any three points $x,y,z\in \spc{X}$: \begin{subthm}{metric>=0} $\dist{x}{y}{\spc{X}}\ge 0$, \end{subthm} \begin{subthm}{metric=0} $\dist{x}{y}{\spc{X}}= 0$ $\iff$ $x=y$, \end{subthm} \begin{subthm}{metric:sym} $\dist{x}{y}{\spc{X}}=\dist{y}{x}{\spc{X}}$, \end{subthm} \begin{subthm}{metric:triangle} $\dist{x}{y}{\spc{X}}+\dist{y}{z}{\spc{X}}\ge\dist{x}{z}{\spc{X}}$. \end{subthm} \end{thm} Recall that a \index{metric space}\emph{metric space} is a set with a metric on it. The elements of the set are called \emph{points}. Most of the time we keep the same notation for the metric space and its underlying set. The function \[\distfun_x\:y\mapsto \dist{x}{y}{}\] is called the \index{distance function}\emph{distance function} from~$x$. Given $R\in[0,\infty]$ and $x\in \spc{X}$, the sets \begin{align*} \oBall(x,R)&=\{y\in \spc{X}\mid \dist{x}{y}{}<R\}, \\ \cBall[x,R]&=\{y\in \spc{X}\mid \dist{x}{y}{}\le R\} \end{align*} are called, respectively, the \index{open ball}\emph{open} and the \index{closed ball}\emph{closed balls} of radius $R$ with center~$x$. If we need to emphasize that these balls are taken in the metric space $\spc{X}$, we write \[\oBall(x,R)_{\spc{X}}\quad\text{and}\quad\cBall[x,R]_{\spc{X}}.\] \begin{thm}{Exercise}\label{ex:quad-inq} Show that \[\dist{p}{q}{\spc{X}}+\dist{x}{y}{\spc{X}}\le\dist{p}{x}{\spc{X}}+\dist{p}{y}{\spc{X}}+\dist{q}{x}{\spc{X}}+\dist{q}{y}{\spc{X}}\] for any four points $p$, $q$, $x$, and $y$ in a metric space $\spc{X}$. \end{thm} \section{Variations of definition} \parbf{Pseudometris.} A metric for which the distance between two distinct points can be zero is called a \index{pseudometric}\emph{pseudometric}. In other words, to define pseudometric, we need to remove condition \ref{SHORT.metric=0} from \ref{def:metric}. The following observation shows that nearly any question about pseudometric spaces can be reduced to a question about genuine metric spaces. Assume $\spc{X}$ is a pseudometric space. Consider an equivalence relation $\sim$ on $\spc{X}$ defined by \[x\sim y\quad\iff\quad\dist{x}{y}{}=0.\] Note that if $x\sim x'$, then $\dist{y}{x}{}=\dist{y}{x'}{}$ for any $y\in\spc{X}$. Thus, $\dist{*}{*}{}$ defines a metric on the quotient set $\spc{X}/{\sim}$. This way we obtain a metric space $\spc{X}'$. The space $\spc{X}'$ is called the \emph{corresponding metric space} for the pseudometric space $\spc{X}$. Often we do not distinguish between $\spc{X}'$ and~$\spc{X}$. \parbf{$\bm{\infty}$-metrics.} One may also consider metrics with values in $\RR\cup\{\infty\}$; we might call them \index{metric!$\infty$-metric}\emph{$\infty$-metrics}, but most of the time we use the term {}\emph{metric}. Again nearly any question about $\infty$-metric spaces can be reduced to a question about genuine metric spaces. Set \[x\approx y\quad\iff\quad\dist{x}{y}{}<\infty;\] it defines another equivalence relation on $\spc{X}$. The equivalence class of a point $x\in\spc{X}$ will be called the \index{metric component}\emph{metric component} of $x$; it will be denoted by $\spc{X}_x$. One could think of $\spc{X}_x$ as $\oBall(x,\infty)_{\spc{X}}$ --- the open ball centered at $x$ and radius $\infty$ in $\spc{X}$. It follows that any $\infty$-metric space is a {}\emph{disjoint union} of genuine metric spaces --- the metric components of the original $\infty$-metric space. \begin{thm}{Exercise}\label{ex:pseudo-infty-metric} Given two sets $A$ and $B$ on the plane, set \[\dist{A}{B}{}=\mu(A\bigtriangleup B),\] where $\mu$ denotes the Lebesgue measure and $\bigtriangleup$ denotes symmetric difference \[A\bigtriangleup B\df(A\cup B)\setminus(B\cap A).\] \begin{subthm}{ex:pseudo-infty-metric:pseudo} Show that $\dist{*}{*}{}$ is a pseudometric on the set of bounded closed subsets. \end{subthm} \begin{subthm}{ex:pseudo-infty-metric:infty} Show that $\dist{*}{*}{}$ is an $\infty$-metric on the set of all open subsets. \end{subthm} \end{thm} \section{Completeness} A metric space $\spc{X}$ is called \index{complete space}\emph{complete} if every Cauchy sequence of points in $\spc{X}$ converges in $\spc{X}$. \begin{thm}{Exercise}\label{ex:almost-min} Suppose that $\rho$ is a positive continuous function on a complete metric space $\spc{X}$ and $\eps>0$. Show that there is a point $x\in \spc{X}$ such that \[\rho(x)<(1+\eps)\cdot\rho(y)\] for any point $y\in \oBall(x,\rho(x))$. \end{thm} Most of the time we will assume that a metric space is complete. The following construction produces a complete metric space $\bar{\spc{X}}$ for any given metric space $\spc{X}$. \parbf{Completion.} Given a metric space $\spc{X}$, consider the set $\spc{C}$ of all Cauchy sequences in $\spc{X}$. Note that for any two Cauchy sequences $(x_n)$ and $(y_n)$ the right-hand side in \ref{eq:cauchy-dist} is defined; moreover, it defines a pseudometric on~$\spc{C}$ \[\dist{(x_n)}{(y_n)}{\spc{C}}\df\lim_{n\to\infty}\dist{x_n}{y_n}{\spc{X}}.\eqlbl{eq:cauchy-dist}\] The corresponding metric space $\bar{\spc{X}}$ is called \index{completion}\emph{completion} of $\spc{X}$. Note that the original space $\spc{X}$ forms a dense subset in its completion $\bar{\spc{X}}$. More precisely, for each point $x\in\spc{X}$ one can consider a constant sequence $x_n=x$ which is Cauchy. It defines a natural map $\spc{X}\to \bar{\spc{X}}$. It is easy to check that this map is distance-preserving. In particular, we can (and will) consider $\spc{X}$ as a subset of $\bar{\spc{X}}$. \begin{thm}{Exercise}\label{ex:complete-completion} Show that the completion of a metric space is complete. \end{thm} \section{Compact spaces} Let us recall few equivalent definitions of compact metric spaces. \begin{thm}{Definition}\label{def:compact} A metric space $\spc{K}$ is compact if and only if one of the following equivalent conditions hold: \begin{subthm}{} Every open cover of $\spc{K}$ has a finite subcover. \end{subthm} \begin{subthm}{} For any open cover of $\spc{K}$ there is $\eps>0$ such that any $\eps$-ball in $\spc{K}$ lies in one element of the cover. (The value $\eps$ is called a \index{Lebesgue number}\emph{Lebesgue number} of the covering.) \end{subthm} \begin{subthm}{} Every sequence of points in $\spc{K}$ has a subsequence that converges in $\spc{K}$. \end{subthm} \begin{subthm}{totally-bounded} The space $\spc{K}$ is complete and \index{totally bounded space}\emph{totally bounded}; that is, for any $\eps>0$, the space $\spc{K}$ admits a finite cover by open $\eps$-balls. \end{subthm} \end{thm} A subset $N$ of a metric space $\spc{K}$ is called \index{net}\emph{$\eps$-net} if any point $x\in \spc{K}$ lies on the distance less than $\eps$ from a point in $N$. Note that totally bounded spaces can be defined as spaces that admit a finite $\eps$-net for any $\eps>0$. \begin{thm}{Exercise}\label{ex:compact-net} Show that a space $\spc{K}$ is totally bounded if and only if it contains a compact $\eps$-net for any $\eps>0$. \end{thm} Let $\pack_\eps\spc{X}$ be the exact upper bound on the number of points $x_1,\z\dots,x_n\in \spc{X}$ such that $\dist{x_i}{x_j}{}\ge\eps$ if $i\ne j$. If $n=\pack_\eps\spc{X}<\infty$, then the collection of points $x_1,\dots,x_n$ is called a \index{maximal packing}\emph{maximal $\eps$-packing}. Note that $n$ is the maximal number of open disjoint $\tfrac\eps2$-balls in $\spc{X}$. \begin{thm}{Exercise}\label{ex:pack-net} Show that any maximal $\eps$-packing is an $\eps$-net. Conclude that a complete space $\spc{X}$ is compact if and only if $\pack_\eps\spc{X}\z<\infty$ for any $\eps>0$. \end{thm} \begin{thm}{Exercise}\label{ex:non-contracting-map} Let $\spc{K}$ be a compact metric space and \[f\:\spc{K}\z\to \spc{K}\] be a distance-nondecreasing map. Prove that $f$ is an \index{isometry}\emph{isometry}; that is, $f$ is a distance-preserving bijection. \end{thm} A metric space $\spc{X}$ is called \index{locally compact space}\emph{locally compact} if any point in $\spc{X}$ admits a compact neighborhood; in other words, for any point $x\in\spc{X}$ a closed ball $\cBall[x,r]$ is compact for some $r>0$. \section{Proper spaces} A metric space $\spc{X}$ is called \index{proper space}\emph{proper} if all closed bounded sets in $\spc{X}$ are compact. This condition is equivalent to each of the following statements: \begin{itemize} \item For some (and therefore any) point $p\in \spc{X}$ and any $R<\infty$, the closed ball $\cBall[p,R]_{\spc{X}}$ is compact. \item The function $\distfun_p\:\spc{X}\to\RR$ is \index{proper function}\emph{proper} for some (and therefore any) point $p\in \spc{X}$; that is, for any compact set $K\subset \RR$, its inverse image \[\distfun_p^{-1}(K)=\set{x\in \spc{X}}{\dist{p}{x}{\spc{X}}\in K}\] is compact. \end{itemize} \begin{thm}{Exercise}\label{ex:loc-compact-not-proper} Give an example of space which is locally compact but not proper. \end{thm} \section{Geodesics} \label{sec:geods} Let $\spc{X}$ be a metric space and $\II$\index{$\II$} a real interval. A~globally isometric map $\gamma\:\II\to \spc{X}$ is called a \index{geodesic}\emph{geodesic}% \footnote{Different authors call it differently: {}\emph{shortest path}, {}\emph{minimizing geodesic}. Also, note that the meaning of the term \emph{geodesic} is different from what is used in Riemannian geometry, altho they are closely related.}; in other words, $\gamma\:\II\to \spc{X}$ is a geodesic if \[\dist{\gamma(s)}{\gamma(t)}{\spc{X}}=|s-t|\] for any pair $s,t\in \II$. If $\gamma\:[a,b]\to \spc{X}$ is a geodesic and $p=\gamma(a)$, $q=\gamma(b)$, then we say that $\gamma$ is a geodesic from $p$ to $q$. In this case, the image of $\gamma$ is denoted by $[p q]$\index{$[{*}{*}]$}, and, with abuse of notations, we also call it a \index{geodesic}\emph{geodesic}. We may write $[p q]_{\spc{X}}$ to emphasize that the geodesic $[p q]$ is in the space ${\spc{X}}$. In general, a geodesic from $p$ to $q$ need not exist and if it exists, it need not be unique. However, once we write $[p q]$ we assume that we have chosen such geodesic. A \index{geodesic path}\emph{geodesic path} is a geodesic with constant-speed parameterization by the unit interval $[0,1]$. \section{Geodesic spaces and metric trees} A metric space is called \index{geodesic}\emph{geodesic} if any pair of its points can be joined by a geodesic. \begin{thm}{Exercise}\label{ex:pogorelov} Let $f$ be a centrally symmetric positive continuous function on $\mathbb{S}^2$. Given two points $x,y\in \mathbb{S}^2$, set \[\|x-y\|=\int_{B(x,\frac \pi2)\backslash B(y,\frac\pi2)}f.\] Show that $(\mathbb{S}^2,\|{*}-{*}\|)$ is a geodesic space and the geodesics in $(\mathbb{S}^2,\|{*}-{*}\|)$ run along great circles of $\mathbb{S}^2$. \end{thm} A geodesic space $\spc{T}$ is called a \index{metric tree}\emph{metric tree} if any pair of points in $\spc{T}$ are connected by a unique geodesic, and the union of any two geodesics $[xy]_{\spc{T}}$, and $[yz]_{\spc{T}}$ contain the geodesic $[xz]_{\spc{T}}$. In other words any triangle in $\spc{T}$ is a tripod; that is, for any three geodesics $[xy]$, $[yz]$, and $[zx]$ have a common point. \begin{thm}{Exercise}\label{ex:4-point-trees} Let $p,x,y,z$ be points in a metric tree. Consider three numbers \begin{align*} a&=|p-x|+|y-z|, & b&=|p-y|+|z-x|, & c&=|p-z|+|x-y|. \end{align*} Suppose that $a\le b\le c$. Show that $b=c$. \end{thm} Recall that the set \[S_r(p)_{\spc{X}}=\set{x\in\spc{X}}{\dist{p}{x}{\spc{X}}=r}\] is called a \emph{sphere} with center $p$ and radius $r$ in a metric space $\spc{X}$. \begin{thm}{Exercise}\label{ex:spheres-in-trees} Show that spheres in metric trees are ultrametric spaces. That is, \[\dist{x}{z}{} \le \max\{\,\dist{x}{y}{},\dist{y}{z}{}\,\}\] for any $x,y,z\in S_r(p)_{\spc{T}}$. \end{thm} \section{Length} A \index{curve}\emph{curve} is defined as a continuous map from a real interval $\II$ to a metric space. If $\II=[0,1]$, that is, if the interval is unit, then the curve is called a \index{path}\emph{path}. \begin{thm}{Definition} Let $\spc{X}$ be a metric space and $\alpha\: \II\to \spc{X}$ be a curve. We define the \index{length}\emph{length} of $\alpha$ as \[ \length \alpha \df \sup_{t_0\le t_1\le\ldots\le t_n}\sum_i \dist{\alpha(t_i)}{\alpha(t_{i-1})}{}. \] A curve $\alpha$ is called \index{rectifiable curve}\emph{rectifiable} if $\length \alpha<\infty$. \end{thm} \begin{thm}{Theorem}\label{thm:length-semicont} Length is a lower semi-continuous with respect to the pointwise convergence of curves. More precisely, assume that a sequence of curves $\gamma_n\:\II\to \spc{X}$ in a metric space $\spc{X}$ converges pointwise to a curve $\gamma_\infty\:\II\to \spc{X}$; that is, for any fixed $t \in \II$, $\gamma_n(t)\z\to\gamma_\infty(t)$ as $n\to\infty$. Then $$\liminf_{n\to\infty} \length\gamma_n \ge \length\gamma_\infty.\eqlbl{eq:semicont-length}$$ \end{thm} \begin{wrapfigure}{o}{20 mm} \vskip-0mm \centering \includegraphics{mppics/pic-100} \end{wrapfigure} Note that the inequality \ref{eq:semicont-length} might be strict. For example, the diagonal $\gamma_\infty$ of the unit square can be approximated by stairs-like polygonal curves $\gamma_n$ with sides parallel to the sides of the square ($\gamma_6$ is on the picture). In this case \[\length\gamma_\infty=\sqrt{2}\quad \text{and}\quad \length\gamma_n=2\] for any $n$. \parit{Proof.} Fix a sequence $t_0\le t_1\le\dots\le t_k$ in $\II$. Set \begin{align*}\Sigma_n &\df |\gamma_n(t_0)-\gamma_n(t_1)|+\dots+|\gamma_n(t_{k-1})-\gamma_n(t_k)|. \\ \Sigma_\infty &\df |\gamma_\infty(t_0)-\gamma_\infty(t_1)|+\dots+|\gamma_\infty(t_{k-1})-\gamma_\infty(t_k)|. \end{align*} Note that for each $i$ we have \[|\gamma_n(t_{i-1})-\gamma_n(t_i)|\to|\gamma_\infty(t_{i-1})-\gamma_\infty(t_i)|\] and therefore \[\Sigma_n\to \Sigma_\infty\] as $n\to\infty$. Note that \[\Sigma_n\le\length\gamma_n\] for each $n$. Hence, $$\liminf_{n\to\infty} \length\gamma_n \ge \Sigma_\infty.$$ Since the partition was arbitrary, by the definition of length, the inequality \ref{eq:semicont-length} is obtained. \qeds \section{Length spaces}\label{sec:intrinsic} If for any $\eps>0$ and any pair of points $x$ and $y$ in a metric space $\spc{X}$, there is a path $\alpha$ connecting $x$ to $y$ such that \[\length\alpha< \dist{x}{y}{}+\eps,\] then $\spc{X}$ is called a \index{length space}\emph{length space} and the metric on $\spc{X}$ is called a \index{length metric}\emph{length metric}.\label{page:length metric} An $\infty$-metric space is a length space if each of its metric components is a length space. In other words, if $\spc{X}$ is an $\infty$-metric space, then in the above definition we assume in addition that $\dist{x}{y}{\spc{X}}<\infty$. Note that any geodesic space is a length space. The following example shows that the converse does not hold. \begin{thm}{Example} Suppose a space $\spc{X}$ is obtained by gluing a countable collection of disjoint intervals $\{\II_n\}$ of length $1+\tfrac1n$, where for each $\II_n$ the left end is glued to $p$ and the right end to~$q$. Observe that the space $\spc{X}$ carries a natural complete length metric with respect to which $\dist{p}{q}{\spc{X}}=1$ but there is no geodesic connecting $p$ to~$q$. \end{thm} \begin{thm}{Exercise}\label{ex:no-geod} Give an example of a complete length space $\spc{X}$ such that no pair of distinct points in $\spc{X}$ can be joined by a geodesic. \end{thm} Directly from the definition, it follows that if $\alpha\:[0,1]\to\spc{X}$ is a path from $x$ to $y$ (that is, $\alpha(0)=x$ and $\alpha(1)=y$), then \[\length\alpha\ge \dist{x}{y}{}.\] Set \[\yetdist{x}{y}{}=\inf\{\length\alpha\}\] where the greatest lower bound is taken for all paths from $x$ to~$y$. It is straightforward to check that $(x,y)\mapsto \yetdist{x}{y}{}$ is an $\infty$-metric; moreover, $(\spc{X},\yetdist{*}{*}{})$ is a length space. The metric $\yetdist{*}{*}{}$ is called the \index{induced length metric}\emph{induced length metric}. \begin{thm}{Exercise}\label{ex:compact+connceted} Let $\spc{X}$ be a complete length space. Show that for any compact subset $K$ in $\spc{X}$ there is a compact path-connected subset $K'$ that contains $K$. \end{thm} \begin{thm}{Exercise}\label{ex:compact=>complete} Suppose $(\spc{X},\dist{*}{*}{})$ is a complete metric space. Show that $(\spc{X},\yetdist{*}{*}{})$ is complete. \end{thm} Let $A$ be a subset of a metric space $\spc{X}$. Given two points $x,y\in A$, consider the value \[\dist{x}{y}{A}=\inf_{\alpha}\{\length\alpha\},\] where the greatest lower bound is taken for all paths $\alpha$ from $x$ to $y$ in~$A$. In other words $\dist{*}{*}{A}$ denotes the induced length metric on the subspace $A$.% \footnote{The notation $\dist{*}{*}{A}$ conflicts with the previously defined notation for distance $\dist{x}{y}{\spc{X}}$ in a metric space $\spc{X}$. However, most of the time we will work with ambient length spaces where the meaning will be unambiguous.} Let $x$ and $y$ be points in a metric space $\spc{X}$. \begin{enumerate}[(i)] \item A point $z\in \spc{X}$ is called a \index{midpoint}\emph{midpoint} between $x$ and $y$ if \[\dist{x}{z}{}=\dist{y}{z}{}=\tfrac12\cdot\dist[{{}}]{x}{y}{}.\] \item Assume $\eps\ge 0$. A point $z\in \spc{X}$ is called an \index{$\eps$-midpoint}\emph{$\eps$-midpoint} between $x$ and $y$ if \[\dist{x}{z}{},\quad\dist{y}{z}{}\le\tfrac12\cdot\dist[{{}}]{x}{y}{}+\eps.\] \end{enumerate} Note that a $0$-midpoint is the same as a midpoint. \begin{thm}{Lemma}\label{lem:mid>geod} Let $\spc{X}$ be a complete metric space. \begin{subthm}{lem:mid>length} Assume that for any pair of points $x,y\in \spc{X}$, and any $\eps>0$, there is an $\eps$-midpoint~$z$. Then $\spc{X}$ is a length space. \end{subthm} \begin{subthm}{lem:mid>geod:geod} Assume that for any pair of points $x,y\in \spc{X}$, there is a midpoint~$z$. Then $\spc{X}$ is a geodesic space. \end{subthm} \end{thm} \parit{Proof.} We first prove \ref{SHORT.lem:mid>length}. Let $x,y\in \spc{X}$ be a pair of points. Set $\eps_n=\frac\eps{4^n}$, $\alpha(0)=x$ and $\alpha(1)=y$. Let $\alpha(\tfrac12)$ be an $\eps_1$-midpoint between $\alpha(0)$ and $\alpha(1)$. Further, let $\alpha(\frac14)$ and $\alpha(\frac34)$ be $\eps_2$-midpoints between the pairs $(\alpha(0),\alpha(\tfrac12))$ and $(\alpha(\tfrac12),\alpha(1))$ respectively. Applying the above procedure recursively, on the $n$-th step we define $\alpha(\tfrac{k}{2^n})$, for every odd integer $k$ such that $0<\tfrac k{2^n}<1$, as an $\eps_{n}$-midpoint of the already defined $\alpha(\tfrac{k-1}{2^n})$ and $\alpha(\tfrac{k+1}{2^n})$. In this way we define $\alpha(t)$ for $t\in W$, where $W$ denotes the set of dyadic rationals in $[0,1]$. Since $\spc{X}$ is complete, the map $\alpha$ can be extended continuously to $[0,1]$. Moreover, \[\begin{aligned} \length\alpha&\le \dist{x}{y}{}+\sum_{n=1}^\infty 2^{n-1}\cdot\eps_n\le \\ &\le \dist{x}{y}{}+\tfrac\eps2. \end{aligned} \eqlbl{eq:eps-midpoint} \] Since $\eps>0$ is arbitrary, we get \ref{SHORT.lem:mid>length}. To prove \ref{SHORT.lem:mid>geod:geod}, one should repeat the same argument taking midpoints instead of $\eps_n$-midpoints. In this case, \ref{eq:eps-midpoint} holds for $\eps_n=\eps\z=0$. \qeds Since in a compact space a sequence of $\tfrac1n$-midpoints $z_n$ contains a convergent subsequence, \ref{lem:mid>geod} immediately implies the following. \begin{thm}{Proposition}\label{prop:length+proper=>geodesic} Any proper length space is geodesic. \end{thm} \begin{thm}{Hopf--Rinow theorem}\label{thm:Hopf-Rinow} Any complete, locally compact length space is proper. \end{thm} Before reading the proof, it is instructive to solve \ref{ex:loc-compact-not-proper}. \parit{Proof.} Let $\spc{X}$ be a locally compact length space. Given $x\in \spc{X}$, denote by $\rho(x)$ the least upper bound of all $R>0$ such that the closed ball $\cBall[x,R]$ is compact. Since $\spc{X}$ is locally compact, $$\rho(x)>0 \quad\text{for any}\quad x\in \spc{X}.\eqlbl{eq:rho>0}$$ It is sufficient to show that $\rho(x)=\infty$ for some (and therefore any) point $x\in \spc{X}$. \begin{clm}{} If $\rho(x)<\infty$, then $B=\cBall[x,\rho(x)]$ is compact. \end{clm} Indeed, $\spc{X}$ is a length space; therefore for any $\eps>0$, the set $\cBall[x,\rho(x)-\eps]$ is a compact $\eps$-net in~$B$. Since $B$ is closed and hence complete, it must be compact. \claimqeds \begin{clm}{} $|\rho(x)-\rho(y)|\le \dist{x}{y}{\spc{X}}$ for any $x,y\in \spc{X}$; in particular, $\rho\:\spc{X}\to\RR$ is a continuous function. \end{clm} Indeed, assume the contrary; that is, $\rho(x)+|x-y|<\rho(y)$ for some $x,y\in \spc{X}$. Then $\cBall[x,\rho(x)+\eps]$ is a closed subset of $\cBall[y,\rho(y)]$ for some $\eps>0$. Then compactness of $\cBall[y,\rho(y)]$ implies compactness of $\cBall[x,\rho(x)+\eps]$, a contradiction.\claimqeds Set $\eps=\min\set{\rho(y)}{y\in B}$; the minimum is defined since $B$ is compact and $\rho$ is continuous. From \ref{eq:rho>0}, we have $\eps>0$. Choose a finite $\tfrac\eps{10}$-net $\{a_1,a_2,\dots,a_n\}$ in $B=\cBall[x,\rho(x)]$. The union $W$ of the closed balls $\cBall[a_i,\eps]$ is compact. Clearly, $\cBall[x,\rho(x)+\frac\eps{10}]\subset W$. Therefore, $\cBall[x,\rho(x)+\frac\eps{10}]$ is compact, a contradiction. \qeds \begin{thm}{Exercise}\label{exercise from BH} Construct a geodesic space $\spc{X}$ that is locally compact, but whose completion $\bar{\spc{X}}$ is neither geodesic nor locally compact. \end{thm} \begin{thm}{Advanced exercise}\label{ex:gross} Show that for any compact length-metric space $\spc{X}$ there is a number $\ell$ such that for any finite collection of points there is a point $z$ that lies on average distance $\ell$ from the collection; that is, for any $x_1,\dots,x_n\in \spc{X}$ there is $z\in \spc{X}$ such that \[\tfrac1n\cdot\sum_i|x_i-z|_{\spc{X}}=\ell.\] \end{thm} \chapter{Semisolutions} \input{metric-sol} \input{uryson-sol} \input{injective-sol} \input{converge-sol} \input{ultralimit-sol} {\small\sloppy \input{pure-metric.ind} \printbibliography[heading=bibintoc] \fussy } \end{document} \chapter{Ultralimits} Ultralimits provide a very general way to pass to a limit. This procedure works for \emph{any} sequence of metric spaces, its result reminds limit in the sense of Gromov Gausdorff, but has some strange features; for example, the limit of a constant sequence does \emph{not} coincide with this constant (see \ref{ex:ultrapower:compact}). In geometry, ultralimits are used only as a canonical way to pass to a convergent subsequence. It is a useful thing in the proofs where one needs to repeat ``pass to convergent subsequence'' too many times. This lecture is based on the introduction to the paper by Bruce Kleiner and Bernhard Leeb \cite{kleiner-leeb}. \section{Faces of ultrafilters} Recall that $\NN$ denotes the set of natural numbers, $\NN=\{1,2,\dots\}$ \begin{thm}{Definition} A finitely additive measure $\omega$ on $\NN$ is called an \index{ultrafilter}\emph{ultrafilter} if it satisfies the following condition: \begin{subthm}{} $\omega(\NN)=1$ and $\omega(S)=0$ or $1$ for any subset $S\subset \NN$. \end{subthm} An ultrafilter $\omega$ is called \emph{nonprincipal}\index{ultrafilter!nonprincipal ultrafilter}\index{nonprincipal ultrafilter} if in addition \begin{subthm}{} $\omega(F)=0$ for any finite subset $F\subset \NN$. \end{subthm} \end{thm} If $\omega(S)=0$ for some subset $S\subset \NN$, we say that $S$ is \index{$\omega$-small}\emph{$\omega$-small}. If $\omega(S)=1$, we say that $S$ contains \index{$\omega$-almost all}\emph{$\omega$-almost all} elements of $\NN$. \parbf{Classical definition.} More commonly, a nonprincipal ultrafilter is defined as a collection, say $\mathfrak{F}$, of sets in $\NN$ such that \begin{enumerate} \item\label{filter:supset} if $P\in \mathfrak{F}$ and $Q\supset P$, then $Q\in \mathfrak{F}$, \item\label{filter:cap} if $P, Q\in \mathfrak{F}$, then $P\cap Q\in \mathfrak{F}$, \item\label{filter:ultra} for any subset $P\subset\NN$, either $P$ or its complement is an element of $\mathfrak{F}$. \item\label{filter:non-prin} if $F\subset \NN $ is finite, then $F\notin \mathfrak{F}$. \end{enumerate} Setting $P\in\mathfrak{F}\Leftrightarrow\omega(P)=1$ makes these two definitions equivalent. A nonempty collection of sets $\mathfrak{F}$ that does not include the empty set and satisfies only conditions \ref{filter:supset} and \ref{filter:cap} is called a \index{filter}\emph{filter}; if in addition $\mathfrak{F}$ satisfies condition \ref{filter:ultra} it is called an \index{ultrafilter}\emph{ultrafilter}. From Zorn's lemma, it follows that every filter contains an ultrafilter. Thus there is an ultrafilter $\mathfrak{F}$ contained in the filter of all complements of finite sets; clearly, this ultrafilter $\mathfrak{F}$ is nonprincipal. \parbf{Stone--\v{C}ech compactification.} Given a set $S\subset \NN$, consider subset $\Omega_S$ of all ultrafilters $\omega$ such that $\omega(S)=1$. It is straightforward to check that the sets $\Omega_S$ for all $S\subset \NN$ form a topology on the set of ultrafilters on $\NN$. The obtained space is called \index{Stone--\v{C}ech compactification}\emph{Stone--\v{C}ech compactification} of $\NN$; it is usually denoted as $\beta\NN$. Let $\omega_n$ denotes the principal ultrafilter such that $\omega_n(\{n\})=1$; that is, $\omega_n(S)=1$ if and only if $n\in S$. Note that $n\mapsto\omega_n$ defines a natural embedding $\NN\hookrightarrow\beta\NN$. Using the described embedding, we can (and will) consider $\NN$ as a subset of $\beta\NN$. The space $\beta\NN$ is the maximal compact Hausdorff space that contains $\NN$ as an everywhere dense subset. More precisely, for any compact Hausdorff space $\spc{X}$ and a map $f\:\NN\to \spc{X}$ there is a unique continuous map $\bar f\:\beta\NN\to X$ such that the restriction $\bar f|_\NN$ coincides with $f$. \section{Ultralimits of points} \label{ultralimits}\index{ultralimit} Further, we will need the existence of a nonprincipal ultrafilter $\omega$, which we fix once and for all. Assume $(x_n)$ is a sequence of points in a metric space $\spc{X}$. Let us define the \index{$\omega$-limit}\emph{$\omega$-limit} of $(x_n)$ as the point $x_\omega$ such that for any $\eps>0$, $\omega$-almost all elements of $(x_n)$ lie in $\oBall(x_\omega,\eps)$; that is, \[\omega\set{n\in\NN}{\dist{x_\omega}{x_n}{}<\eps}=1.\] In this case, we will write \[x_\omega=\lim_{n\to\omega} x_n \ \ \text{or}\ \ x_n\to x_\omega\ \text{as}\ n\to\omega.\] For example, if $\omega$ is the {}\emph{principal} ultrafilter such that $\omega\{n\}=1$ for some $n\in\NN$, then $x_\omega=x_n$. Alternatively, the sequence $(x_n)$ can be regarded as a map $\NN\to\spc{X}$. In this case, the map $\NN\to\spc{X}$ can be extended to a continuous map $\beta\NN\to\spc{X}$ from the Stone--\v{C}ech compactification $\beta\NN$ of $\NN$. Then the $\omega$-limit $x_\omega$ can be regarded as the image of $\omega$. Note that $\omega$-limits of a sequence and its subsequence may differ. For example, sequence $y_n=-(-1)^n$ is a subsequence of $x_n=(-1)^n$, but for any ultrafilter $\omega$, we have \[\lim_{n\to\omega}x_n \ne \lim_{n\to\omega}y_n.\] \begin{thm}{Proposition}\label{prop:ultra/partial} Let $\omega$ be a nonprincipal ultrafilter. Assume $(x_n)$ is a sequence of points in a metric space $\spc{X}$ and $x_n\to x_\omega$ as $n\to\omega$. Then $x_\omega$ is a partial limit of the sequence $(x_n)$; that is, there is a subsequence $(x_n)_{n\in S}$ that converges to $x_\omega$ in the usual sense. \end{thm} \parit{Proof.} Given $\eps>0$, set $S_\eps=\set{n\in\NN}{\dist{x_n}{x_\omega}{}<\eps}$. Note that $\omega(S_\eps)=1$ for any $\eps>0$. Since $\omega$ is nonprincipal, the set $S_\eps$ is infinite. Therefore, we can choose an increasing sequence $(n_k)$ such that $n_k\in S_{\frac1k}$ for each $k\in \NN$. Clearly, $x_{n_k}\to x_\omega$ as $k\to\infty$. \qeds The following proposition is analogous to the statement that any sequence in a compact metric space has a convergent subsequence; it can be proved the same way. \begin{thm}{Proposition}\label{prop:ultra/compact} Let $\spc{X}$ be a compact metric space. Then any sequence of points $(x_n)$ in $\spc{X}$ has a unique $\omega$-limit $x_\omega$. In particular, a bounded sequence of real numbers has a unique $\omega$-limit. \end{thm} The following lemma is an ultralimit analog of the Cauchy convergence test. \begin{thm}{Lemma}\label{lem:X-X^w} Let $(x_n)$ be a sequence of points in a complete space~$\spc{X}$. Assume for each subsequence $(y_n)$ of $(x_n)$, the $\omega$-limit \[y_\omega=\lim_{n\to\omega}y_{n}\in \spc{X}\] is defined and does not depend on the choice of subsequence, then the sequence $(x_n)$ converges in the usual sense. \end{thm} \parit{Proof.} If $(x_n)$ is not a Cauchy sequence, then for some $\eps>0$, there is a subsequence $(y_n)$ of $(x_n)$ such that $\dist{x_n}{y_n}{}\ge\eps$ for all $n$. It follows that $\dist{x_\omega}{y_\omega}{}\ge \eps$, a contradiction.\qeds Ultralimits could shorten some proofs in the previous lecture. The following exercise provides an example. \begin{thm}{Exercise}\label{ex:prop:eps-isometry=isometry} Use ultralimits to give a shorter proof of \ref{GH-2} (page~\pageref{page:GH-2-proof}). \end{thm} \begin{thm}{Exercise}\label{ex:linear} Denote by $S$ the space of bounded sequences of real numbers. Show that there is a linear functional $L\:S\to\RR$ such that for any sequence $\bm{s}=(s_1,s_2,\dots)\in S$ the image $L(\bm{s})$ is a partial limit of $s_1,s_2,\dots$. \end{thm} \section{Ultralimits of spaces}\label{sec:Ultralimit of spaces} Recall that $\omega$ denotes a nonprincipal ultrafilter on the set of natural numbers. Let $\spc{X}_n$ be a sequence of metric spaces. Consider all sequences of points $x_n\in \spc{X}_n$. On the set of all such sequences, define a pseudometric by \[\dist{(x_n)}{(y_n)}{} = \lim_{n\to\omega} \dist{x_n}{y_n}{\spc{X}_n}. \eqlbl{eq:olim-dist}\] Note that the $\omega$-limit on the right-hand side is always defined and takes a value in $[0,\infty]$. (The $\omega$-convergence to $\infty$ is defined analogously to the usual convergence to $\infty$). Set $\spc{X}_\omega$ to be the corresponding metric space; that is, the underlying set of $\spc{X}_\omega$ is formed by classes of equivalence of sequences of points $x_n\in\spc{X}_n$ defined by \[(x_n)\sim(y_n) \ \Leftrightarrow\ \lim_{n\to\omega} \dist{x_n}{y_n}{}=0\] and the distance is defined by \ref{eq:olim-dist}. The space $\spc{X}_\omega$ is called the \index{$\omega$-limit space}\emph{$\omega$-limit} of $\spc{X}_n$. Typically $\spc{X}_\omega$ will denote the $\omega$-limit of sequence $\spc{X}_n$; we may also write \[\spc{X}_n\to\spc{X}_\omega\ \ \text{as}\ \ n\to\omega\ \ \text{or}\ \ \spc{X}_\omega=\lim_{n\to\omega}\spc{X}_n.\] Given a sequence $x_n\in \spc{X}_n$, we will denote by $x_\omega$ its equivalence class which is a point in $\spc{X}_\omega$; it can be written as \[x_n\to x_\omega \ \ \text{as}\ \ n\to\omega,\ \ \text{or}\ \ x_\omega=\lim_{n\to\omega} x_n.\] \begin{thm}{Observation}\label{obs:ultralimit-is-complete} The $\omega$-limit of any sequence of metric spaces is complete. \end{thm} \parit{Proof.} Let $\spc{X}_n$ be a sequence of metric spaces and $\spc{X}_n\to\spc{X}_\omega$ as $n\to\omega$. Choose a Cauchy sequence $x_1,x_2,\dots\in\spc{X}_\omega$. Passing to a subsequence, we can assume that $\dist{x_k}{x_{m}}{\spc{X}_\omega}<\tfrac1{k}$ for any $k<m$. Let us choose a double sequence $x_{n,m}\in \spc{X}_n$ such that for any fixed $m$ we have $x_{n,m}\to x_m$ as $n\to\omega$. Note that for any $k<m$ the inequality $\dist{x_{n,k}}{x_{n,m}}{}<\tfrac1{k}$ holds for $\omega$-almost all $n$. It follows that we can choose a nested sequence of sets \[\NN= S_1\supset S_2\supset\dots\] such that \begin{itemize} \item $\omega(S_m)=1$ for each $m$, \item $\bigcap_m S_m=\emptyset$, and \item $\dist{x_{n,k}}{x_{n,l}}{}<\tfrac1{k}$ for $k<l\le m$ and $n\in S_m$. \end{itemize} Consider the sequence $y_n=x_{n,m(n)}$, where $m(n)$ is the largest value such that $n\in S_{m(n)}$. Denote by $y_\omega\in \spc{X}_\omega$ the $\omega$-limit of $y_n$. Observe that $|y_m-x_{n,m}|<\tfrac1{m}$ for $\omega$-almost all $n$. It follows that $|x_m-y_\omega|\le \tfrac1{m}$ for any $m$. Therefore, $x_n\to y_\omega$ as $n\to \infty$. That is, any Cauchy sequence in $\spc{X}_\omega$ converges. \qeds \begin{thm}{Observation}\label{obs:ultralimit-is-geodesic} The $\omega$-limit of any sequence of length spaces is geodesic. \end{thm} \parit{Proof.} If $\spc{X}_n$ is a sequence of length spaces, then for any sequence of pairs $x_n, y_n\in X_n$ there is a sequence of $\tfrac1n$-midpoints $z_n$. Let $x_n\to x_\omega$, $y_n\to y_\omega$ and $z_n\to z_\omega$ as $n\to \omega$. Note that $z_\omega$ is a midpoint of $x_\omega$ and $y_\omega$ in $\spc{X}_\omega$. By Observation~\ref{obs:ultralimit-is-complete}, $\spc{X}_\omega$ is complete. Applying Lemma~\ref{lem:mid>geod} we get the statement. \qeds \begin{thm}{Exercise}\label{ex:lim(tree)} Show that an ultralimit of metric trees is a metric tree. \end{thm} \section{Ultrapower} If all the metric spaces in the sequence are identical $\spc{X}_n=\spc{X}$, its $\omega$-limit $\lim_{n\to\omega}\spc{X}_n$ is denoted by $\spc{X}^\omega$ and called $\omega$-power of $\spc{X}$. \begin{thm}{Exercise}\label{ex:ultrapower} For any point $x\in \spc{X}$, consider the constant sequence $x_n=x$ and set $\iota(x)=\lim_{n\to\omega}x_n\in\spc{X}^\omega$. \begin{subthm}{ex:ultrapower:a} Show that $\iota\:\spc{X}\to\spc{X}^\omega$ is distance-preserving embedding. (So we can and will consider $\spc{X}$ as a subset of $\spc{X}^\omega$.) \end{subthm} \begin{subthm}{ex:ultrapower:compact} Show that $\iota$ is onto if and only if $\spc{X}$ compact. \end{subthm} \begin{subthm}{ex:ultrapower:proper} Show that if $\spc{X}$ is proper, then $\iota(\spc{X})$ forms a metric component of $\spc{X}^\omega$; that is, a subset of $\spc{X}^\omega$ that lie at a finite distance from a given point. \end{subthm} \end{thm} Note that \ref{SHORT.ex:ultrapower:compact} implies that the inclusion $\spc{X}\hookrightarrow\spc{X}^\omega$ is not onto if the space $\spc{X}$ is not compact. However, the spaces $\spc{X}$ and $\spc{X}^\omega$ might be isometric; here is an example: \begin{thm}{Exercise}\label{ex:isom-ultrapower} Let $\spc{X}$ be a countable set with discrete metric; that is $\dist{x}{y}{\spc{X}}=1$ if $x\ne y$. Show that \begin{subthm}{ex:isom-ultrapower:no} $\spc{X}^\omega$ is not isometric to $\spc{X}$. \end{subthm} \begin{subthm}{ex:isom-ultrapower:yes} $\spc{X}^\omega$ is isometric to $(\spc{X}^\omega)^\omega$. \end{subthm} \end{thm} \begin{thm}{Observation}\label{obs:ultrapower-is-geodesic} Let $\spc{X}$ be a complete metric space. Then $\spc{X}^\omega$ is geodesic space if and only if $\spc{X}$ is a length space. \end{thm} \parit{Proof.} The ``if''-part follows from \ref{obs:ultralimit-is-geodesic}; it remains to prove the ``only-if''-part Assume $\spc{X}^\omega$ is geodesic space. Then any pair of points $x,y\in \spc{X}$ has a midpoint $z_\omega\in\spc{X}^\omega$. Fix a sequence of points $z_n\in \spc{X}$ such that $z_n\to z_\omega$ as $n\to \omega$. Note that $\dist{x}{z_n}{\spc{X}}\to \tfrac12\cdot \dist{x}{y}{\spc{X}}$ and $\dist{y}{z_n}{\spc{X}}\to \tfrac12\cdot \dist{x}{y}{\spc{X}}$ as $n\to\omega$. In particular, for any $\eps>0$, the point $z_n$ is an $\eps$-midpoint of $x$ and $y$ for $\omega$-almost all $n$. It remains to apply \ref{lem:mid>geod}. \qeds \begin{thm}{Exercise}\label{ex:two-geodesics-in-ultrapower} Assume $\spc{X}$ is a complete length space and $p,q\in\spc{X}$ cannot be joined by a geodesic in $\spc{X}$. Then there are at least two distinct geodesics between $p$ and $q$ in the ultrapower $\spc{X}^\omega$. \end{thm} \begin{thm}{Exercise}\label{ex:notproper-limit} Construct a proper metric space $\spc{X}$ such that $\spc{X}^\omega$ is not proper; that is, there is a point $p\in \spc{X}^\omega$ and $R<\infty$ such that the closed ball $\cBall[p,R]_{\spc{X}^\omega}$ is not compact. \end{thm} \section{Tangent and asymptotic spaces} Choose a space $\spc{X}$ and a sequence of $\lambda_n>0$. Consider the sequence of \index{scaled space}\emph{scalings} $\spc{X}_n=\lambda_n\cdot\spc{X}=(\spc{X},\lambda_n\cdot\dist{*}{*}{\spc{X}})$. Choose a point $p\in \spc{X}$ and denote by $p_n$ the corresponding point in $\spc{X}_n$. Consider the $\omega$-limit $\spc{X}_\omega$ of $\spc{X}_n$ (one may denote it by $\lambda_\omega\cdot \spc{X}$); set $p_\omega$ to be the $\omega$-limit of $p_n$. If $\lambda_n\to 0$ as $n\to\omega$, then the metric component of $p_\omega$ in $\spc{X}_\omega$ is called \index{$\lambda_\omega$-tangent space}\emph{$\lambda_\omega$-tangent space} at $p$ and denoted by $\T_p^{\lambda_\omega}\spc{X}$ (or $\T_p^{\omega}\spc{X}$ if $\lambda_n=n$).\label{page:ultratangent space} If $\lambda_n\to \infty$ as $n\to\omega$, then the metric component of $p_\omega$ is called \index{$\lambda_\omega$-asymptotic space}\emph{$\lambda_\omega$-asymptotic space}\footnote{Often it is called \emph{asymptotic cone} despite that it is not a cone in general; this name is used since in good cases it has a cone structure.} and denoted by $\Asym\spc{X}$ or $\Asym^{\lambda_\omega}\spc{X}$. Note that the space $\Asym\spc{X}$ and its point $p_\omega$ does not depend on the choice of $p\in \spc{X}$. The following exercise states that the tangent and asymptotic spaces depend on the sequence $\lambda_n$ and a nonprincipal ultrafilter $\omega$. \begin{thm}{Exercise}\label{ex:ultraT} Construct a metric space $\spc{X}$ with a point $p$ such that the tangent space $\T_p^{\lambda_\omega}\spc{X}$ depends on the sequence $\lambda_n$ and/or ultrafilter~$\omega$. \end{thm} For nice spaces, different choices may give the same space. \begin{thm}{Exercise}\label{ex:Asym(Lob)} Let $\spc{L}$ be the Lobachevsky plane; $\spc{T}=\Asym\spc{L}$. \begin{subthm}{ex:Asym(Lob):metric-tree} Show that $\spc{T}$ is a complete metric tree. \end{subthm} \begin{subthm}{ex:Asym(Lob):homogeneous} Show that $\spc{T}$ is one-point homogeneous; that is, given two points $s,t\in \spc{T}$ there is an isometry of $\spc{T}$ that maps $s$ to $t$. \end{subthm} \begin{subthm}{ex:Asym(Lob):continuum} Show that $\spc{T}$ has {}\emph{continuum degree} at any point; that is, for any point $t\in \spc{T}$ the set of connected components of the complement $\spc{T}\backslash\{t\}$ has cardinality continuum. \end{subthm} \begin{subthm}{ex:Asym(Lob):others} Prove \ref{SHORT.ex:Asym(Lob):metric-tree}--\ref{SHORT.ex:Asym(Lob):continuum} if $\spc{L}$ is Lobachevsky space and/or for the infinite 3-regular% \footnote{that is, the degree of any vertex is 3.} metric tree with unit edge length. \end{subthm} \end{thm} \section{Remarks} A nonprincipal ultrafilter $\omega$ is called \emph{selective}\index{ultrafilter!selective ultrafilter}\index{selective ultrafilter} if for any partition of $\NN$ into sets $\{C_\alpha\}_{\alpha\in\IndexSet}$ such that $\omega(C_\alpha)\z=0$ for each $\alpha$, there is a set $S\subset \NN$ such that $\omega(S)=1$ and $S\cap C_\alpha$ is a one-point set for each $\alpha\in\IndexSet$. The existence of a selective ultrafilter follows from the continuum hypothesis \cite{rudin}. For a selective ultrafilter $\omega$, there is a stronger version of Proposition~\ref{prop:ultra/partial}; namely we can assume that the subsequence $(x_n)_{n\in S}$ can be chosen so that $\omega(S)=1$. So, if needed, one may assume that the ultrafilter $\omega$ is chosen to be selective and use this stronger version of the proposition. \chapter{Universal spaces} This lecture is based on the discussion of Urysohn space in the book of Mikhael Gromov \cite{gromov-2007}. \section{Embedding in a normed space} Recall that a function $v\mapsto |v|$ on a vector space $\spc{V}$ is called \index{norm}\emph{norm} if it satisfies the following condition for any two vectors $v,w\in \spc{V}$ and a scalar $\alpha$: \begin{itemize} \item $|v|\ge 0$; \item $|\alpha\cdot v|=|\alpha|\cdot |v|$; \item $|v|+|w|\ge|v+w|$. \end{itemize} As an example, consider the space of real sequences equipped with \index{sup norm}\emph{sup norm} denoted by the $\ell^\infty$; that is, $\ell^\infty$-norm of $\bm{a}=a_1,a_2,\dots$ is defined by \[|\bm{a}|_{\ell^\infty}=\sup_n\{\,|a_n|\,\}.\] It is straightforward to check that for any normed space the function $(v,w)\mapsto |v-w|$ defines a metric on it. Therefore, any normed space is an example of metric space (in fact, it is a geodesic space). Often we do not distinguish between normed space and the corresponding metric space.\footnote{By Mazur--Ulam theorem, the metric remembers the linear structure of the space; a slick proof of this statement was given by Jussi V\"{a}is\"{a}l\"{a} \cite{vaisala}.} The following lemma implies that any compact metric space is isometric to a subset of a fixed normed space. Recall that \index{diameter}\emph{diameter} of a metric space $\spc{X}$ (briefly $\diam \spc{X}$) is defined as least upper bound on the distances between pairs of its points; that is, \[\diam \spc{X}=\sup\set{\dist{x}{y}{\spc{X}}}{x,y\in \spc{X}}.\] \begin{thm}{Lemma}\label{lem:frechet} Suppose $\spc{X}$ is a \index{bounded space}\emph{bounded} \index{separable space}\emph{separable} metric space; that is, $\diam\spc{X}$ is finite and $\spc{X}$ contains a countable, dense set $\{w_n\}$. Given $x\in \spc{X}$, set $a_n(x)=\dist{w_n}{x}{\spc{X}}$. Then \[\iota\:x\mapsto (a_1(x), a_2(x),\dots)\] defines a distance-preserving embedding $\iota\:\spc{X}\hookrightarrow \ell^\infty$. \end{thm} \parit{Proof.} By the triangle inequality \[|a_n(x)-a_n(y)|\le \dist{x}{y}{\spc{X}}.\eqlbl{eq:a-a=<dist}\] Therefore, $\iota$ is \index{short map}\emph{short} (in other words, $\iota$ is distance non-increasing). Again by triangle inequality we have \[|a_n(x)-a_n(y)|\ge \dist{x}{y}{\spc{X}}-2\cdot\dist{w_n}{x}{\spc{X}}.\] Since the set $\{w_n\}$ is dense, we can choose $w_n$ arbitrarily close to $x$. Whence the value $|a_n(x)-a_n(y)|$ can be chosen arbitrarily close to $\dist{x}{y}{\spc{X}}$. In other words \[\sup_n\{\,\left|\dist{w_n}{x}{\spc{X}}-\dist{w_n}{y}{\spc{X}}\right|\,\}\ge \dist{x}{y}{\spc{X}}.\] Hence \[\sup_n\{\,|a_n(x)-a_n(y)|\,\}\ge \dist{x}{y}{\spc{X}};\eqlbl{eq:a-a>=dist}\] that is, $\iota$ is distance non-contracting. Finally, observe that \ref{eq:a-a=<dist} and \ref{eq:a-a>=dist} imply the lemma. \qeds \begin{thm}{Exercise}\label{ex:compact-length} Show that any compact metric space $\spc{K}$ is isometric to a subspace of a compact geodesic space. \end{thm} The following exercise generalizes the lemma to arbitrary separable spaces. \begin{thm}{Exercise}\label{ex:frechet} Suppose $\{w_n\}$ is a countable, dense set in a metric space $\spc{X}$. Choose $x_0\in \spc{X}$; given $x\in \spc{X}$, set \[a_n(x)=\dist{w_n}{x}{\spc{X}}-\dist{w_n}{x_0}{\spc{X}}.\] Show that $\iota\:x\mapsto (a_1(x), a_2(x),\dots)$ defines a distance-preserving embedding $\iota\:\spc{X}\hookrightarrow \ell^\infty$. \end{thm} The following lemma implies that {}\emph{any metric space is isometric to a subset of a normed vector space}; its proof is nearly identical to the proof of \ref{ex:frechet}. \begin{thm}{Lemma}\label{lem:kuratowski} Let $\spc{X}$ be arbitrary metric space. Denote by $\ell^\infty(\spc{X})$ the space of all bounded functions on $\spc{X}$ equipped with sup-norm. Then for any point $x_0\in \spc{X}$, the map $\iota\:\spc{X}\to \ell^\infty(\spc{X})$ defined by \[\iota\:x\mapsto (\distfun_x-\distfun_{x_0})\] is distance-preserving. \end{thm} \section{Extension property} If a metric space $\spc{X}$ is a subspace of a pseudometric space $\spc{X}'$, then we say that $\spc{X}'$ is an \index{extension}\emph{extension} of $\spc{X}$. If in addition $\diam\spc{X}'\le d$, then we say that $\spc{X}'$ is a {}\emph{$d$-extension}. If the complement $\spc{X}'\backslash \spc{X}$ contains a single point, say $p$, we say that $\spc{X}'$ is a \index{one-point extension}\emph{one-point extension} of $\spc{X}$. In this case, to define metric on $\spc{X}'$, it is sufficient to specify the distance function from $p$; that is, a function $f\:\spc{X}\to\RR$ defined by \[f(x)=\dist{p}{x}{\spc{X}'}.\] The function $f$ cannot be taken arbitrary --- the triangle inequality implies that \[f(x)+f(y)\ge \dist{x}{y}{\spc{X}}\ge |f(x)-f(y)|\] for any $x,y\in \spc{X}$. In particular, $f$ is a non-negative 1-Lipschitz function on $\spc{X}$. For a $d$-extension, we need to assume in addition that $\diam\spc{X}\z\le d$ and $f(x)\le d$ for any $x\in \spc{X}$. Any function $f$ of that type will be called \index{extension function}\emph{extension function}\label{page:extension function} or {}\emph{$d$-extension function} respectively. \begin{thm}{Definition}\label{def:finite+1} A metric space $\spc{U}$ meets the \index{extension property}\emph{extension property} if for any finite subspace $\spc{F}\subset\spc{U}$ and any extension function $f\:\spc{F}\to\RR$ there is a point $p\in \spc{U}$ such that $\dist{p}{x}{}=f(x)$ for any $x\in \spc{F}$. If we assume in addition that $\diam \spc{U}\le d$ and instead of extension functions we consider only $d$-extension functions, then we arrive at a definition of {}\emph{$d$-extension property}. If in addition $\spc{U}$ is separable and complete, then it is called \index{Urysohn space}\emph{Urysohn space} or {}\emph{$d$-Urysohn space} respectively. \end{thm} \begin{thm}{Proposition}\label{prop:univeral-separable} There is a separable metric space with the ($d$-) extension property (for any $d\ge 0$). \end{thm} \parit{Proof.} Choose $d\ge 0$. Let us construct a separable metric space with the $d$-extension property. Let $\spc{X}$ be a compact metric space such that $\diam\spc{X}\le d$. Denote by $\spc{X}^d$ the space of all $d$-extension functions on $\spc{X}$ equipped with the metric defined by the sup-norm. Note that the map $\spc{X} \to \spc{X}^d$ defined by $x\mapsto\distfun_x$ is a distance-preserving embedding, so we can (and will) treat $\spc{X}$ as a subspace of $\spc{X}^d$, or, equivalently, $\spc{X}^d$ is an extension of $\spc{X}$. Let us iterate this construction. Start with a one-point space $\spc{X}_0$ and consider a sequence of spaces $(\spc{X}_n)$ defined by $\spc{X}_{n+1}\z=\spc{X}_n^d$. Note that the sequence is nested; that is, $\spc{X}_0\subset \spc{X}_1\subset\dots$ and the union \[\spc{X}_\infty=\bigcup_n\spc{X}_n;\] comes with metric such that $\dist{x}{y}{\spc{X}_\infty} = \dist{x}{y}{\spc{X}_n}$ if $x,y\in\spc{X}_n$. Note that if $\spc{X}$ is compact, then so is $\spc{X}^d$. It follows that each space $\spc{X}_n$ is compact. In particular, $\spc{X}_\infty$ is a countable union of compact spaces; therefore $\spc{X}_\infty$ is separable. Any finite subspace $\spc{F}$ of $\spc{X}_\infty$ lies in some $\spc{X}_n$ for $n<\infty$. By construction, there is a point $p\in \spc{X}_{n+1}$ that meets the condition in \ref{def:finite+1} for any extension function $f\:\spc{F}\to\RR$. That is, $\spc{X}_\infty$ has the $d$-extension property. The construction of a separable metric space with the extension property requires only minor changes. First, the sequence should be defined by $\spc{X}_{n+1}\z=\spc{X}_n^{d_n}$, where $d_n$ is an increasing sequence such that $d_n\to\infty$. Second, the point $p$ should be taken in $\spc{X}_{n+k}$ for sufficiently large $k$, so that $d_{n+k}>\max\{f(x)\}$. \qeds \begin{thm}{Proposition}\label{prop:completion-univeral} If a metric space $\spc{V}$ meets the ($d$-) extension property, then so does its completion. \end{thm} \parit{Proof.} Let us assume $\spc{V}$ meets the extension property. We will show that its completion $\spc{U}$ meets the extension property as well. The $d$-extension case can be proved along the same lines. Note that $\spc{V}$ is a dense subset in a complete space $\spc{U}$. Observe that $\spc{U}$ has the {}\emph{approximate extension property}; that is, if $\spc{F}\z\subset\spc{U}$ is a finite set, $\eps>0$, and $f\:\spc{F}\to \RR$ is an extension function, then there exists $p\in \spc{U}$ such that \[\dist{p}{x}{}\lg f(x)\pm\eps.\] for any $x\in\spc{F}$. Therefore, there is a sequence of points $p_n\in \spc{U}$ such that for any $x\in \spc{F}$, \[\dist{p_n}{x}{}\lg f(x)\pm\tfrac1{2^n}.\] Moreover, we can assume that \[\dist{p_n}{p_{n+1}}{} < \tfrac1{2^n}\eqlbl{eq:|pn-pn|}\] for all large $n$. Indeed, consider the sets $\spc{F}_n=\spc{F}\cup\{p_n\}$ and the functions $f_n\:\spc{F}_n\to\RR$ defined by $f_n(x)=f(x)$ if $x\ne p_n$ and \[f_n(p_n)=\max\set{\bigl|\dist{p_n}{x}{}- f(x)\bigr|}{x\in \spc{F}}.\] Observe that $f_n$ is an extension function for large $n$ and $f_n(p_n)\z<\tfrac1{2^n}$. Therefore, applying the approximate extension property recursively we get~\ref{eq:|pn-pn|}. By \ref{eq:|pn-pn|}, $(p_n)$ is a Cauchy sequence and its limit meets the condition in the definition of extension property (\ref{def:finite+1}). \qeds Note that \ref{prop:univeral-separable} and \ref{prop:completion-univeral} imply the following: \begin{thm}{Theorem}\label{thm:urysohn-exists} Urysohn space and $d$-Urysohn space for any $d>0$ exist. \end{thm} \section{Universality} A metric space will be called \emph{universal} if it includes as a subspace an isometric copy of any separable metric space. In \ref{ex:frechet}, we proved that $\ell^\infty$ is a universal space. The following proposition shows that an Urysohn space is universal as well. Unlike $\ell^\infty$, Urysohn spaces are separable; so it might be considered as a \emph{better} universal space. Theorem \ref{thm:compact-homogeneous} will give another reason why Urysohn spaces are better. \begin{thm}{Proposition}\label{prop:sep-in-urys} An Urysohn space is universal. That is, if $\spc{U}$ is an Urysohn space, then any separable metric space $\spc{S}$ admits a distance-preserving embedding $\spc{S}\hookrightarrow\spc{U}$. Moreover, for any finite subspace $\spc{F}\subset \spc{S}$, any distance-preserving embedding $\spc{F}\hookrightarrow \spc{U}$ can be extended to a distance-preserving embedding $\spc{S}\hookrightarrow\spc{U}$. A $d$-Urysohn space is $d$-universal; that is, the above statements hold provided that $\diam\spc{S}\le d$. \end{thm} \parit{Proof.} We will prove the second statement; the first statement is its partial case for $\spc{F}=\emptyset$. The required isometry will be denoted by $x\mapsto x'$. Choose a dense sequence of points $s_1,s_2,\dotsc\in\spc{S}$. We may assume that $\spc{F}=\{s_1,\dots,s_n\}$, so $s_i'\in \spc{U}$ are defined for $i\le n$. The sequence $s_i'$ for $i>n$ can be defined recursively using the extension property in $\spc{U}$. Namely, suppose that $s_1',\dots,s_{i-1}'$ are already defined. Since $\spc{U}$ meets the extension property, there is a point $s_i'\in \spc{U}$ such that \[\dist{s_i'}{s_j'}{\spc{U}}=\dist{s_i}{s_j}{\spc{S}}\] for any $j<i$. The constructed map $s_i\mapsto s_i'$ is distance-preserving. Therefore it can be continuously extended to whole $\spc{S}$. It remains to observe that the constructed map $\spc{S}\hookrightarrow\spc{U}$ is distance-preserving. \qeds \begin{thm}{Exercise}\label{ex:geodesics-urysohn} Show that any two distinct points in an Urysohn space can be joined by an infinite number of geodesics. \end{thm} \begin{thm}{Exercise}\label{ex:compact-extension} Modify the proofs of \ref{prop:completion-univeral} and \ref{prop:sep-in-urys} to prove the following theorem. \end{thm} \begin{thm}{Theorem}\label{thm:compact-extension} Let $K$ be a compact set in a separable space $\spc{S}$. Then any distance-preserving map from $K$ to an Urysohn space can be extended to a distance-preserving map on whole $\spc{S}$. \end{thm} \begin{thm}{Exercise}\label{ex:sc-urysohn} Show that ($d$-) Urysohn space is simply connected. \end{thm} \section{Uniqueness and homogeneity} \begin{thm}{Theorem}\label{thm:urysohn-unique} Suppose $\spc{F}\subset \spc{U}$ and $\spc{F}'\subset \spc{U}'$ be finite isometric subspaces in a pair of ($d$-)Urysohn spaces $\spc{U}$ and $\spc{U}'$. Then any isometry $\iota\:\spc{F}\leftrightarrow \spc{F}'$ can be extended to an isometry $\spc{U}\leftrightarrow \spc{U}'$. In particular, ($d$-)Urysohn space is unique up to isometry. \end{thm} While \ref{prop:sep-in-urys} implies that there are distance-preserving maps $\spc{U}\to \spc{U}'$ and $\spc{U}'\to \spc{U}$, it does not solely imply the existence of an isometry $\spc{U}\leftrightarrow \spc{U}'$. Its construction uses the idea of \ref{prop:sep-in-urys}, but it is applied {}\emph{back-and-forth} to ensure that the obtained distance-preserving map is onto. \parit{Proof.} Choose dense sequences $a_1,a_2,\dots\in \spc{U}$ and $b'_1,b'_2,\dots\in \spc{U}'$. We can assume that $\spc{F}=\{a_1,\dots,a_n\}$, $\spc{F}'=\{b_1',\dots,b_n'\}$ and $\iota(a_i)=b_i$ for $i\le n$. The required isometry $\spc{U}\leftrightarrow \spc{U}'$ will be denoted by $u \leftrightarrow u'$. Set $a'_i=b'_i$ if $i\le n$. Let us define recursively $a_{n+1}',b_{n+1}, a_{n+2}', b_{n+2},\dots$ --- on the odd step we define the images of $a_{n+1},a_{n+2},\dots$ and on the even steps we define inverse images of $b'_{n+1},b'_{n+2},\dots$. The same argument as in the proof of \ref{prop:sep-in-urys} shows that we can construct two sequences $a_1',a_2',\dots\in \spc{U}'$ and $b_1,b_2,\dots\in \spc{U}$ such that \begin{align*} \dist{a_i}{a_j}{\spc{U}}&=\dist{a_i'}{a_j'}{\spc{U}'} \\ \dist{a_i}{b_j}{\spc{U}}&=\dist{a_i'}{b_j'}{\spc{U}'} \\ \dist{b_i}{b_j}{\spc{U}}&=\dist{b_i'}{b_j'}{\spc{U}'} \end{align*} for all $i$ and $j$. It remains to observe that the constructed distance-preserving bijection defined by $a_i\leftrightarrow a_i'$ and $b_i\leftrightarrow b_i'$ extends continuously to an isometry $\spc{U}\leftrightarrow \spc{U}'$. \qeds Observe that \ref{thm:urysohn-unique} implies that the Urysohn space (as well as the $d$-Urysohn space) is \index{homogeneous}\emph{finite-set homogeneous}; that is, \begin{itemize} \item any distance-preserving map from a finite subset to the whole space can be extended to an isometry. \end{itemize} \begin{thm}{Open question} Is there a {}\emph{noncomplete} finite-set homogeneous metric space that meets the extension property? \end{thm} This is a question of Pavel Urysohn; it appeared already in \cite[§$2(6)$]{urysohn} and reappeared in \cite[p. 83]{gromov-2007} with a missing keyword. In fact, I do not see an example of a 1-point homogeneous space that meets the extension property. Recall that $S_r(p)_{\spc{X}}$ denotes the sphere of radius $r$ centered at $p$ in a metric space $\spc{X}$; that is, $$S_r(p)_{\spc{X}}=\set{x\in \spc{X}}{\dist{p}{x}{\spc{X}}=r}.$$ \begin{thm}{Exercise}\label{ex:sphere-in-urysohn} Choose $d\in [0,\infty]$. Denote by $\spc{U}_d$ the $d$-Urysohn space, so $\spc{U}_\infty$ is the Urysohn space. \begin{subthm}{ex:sphere-in-urysohn:sphere} Assume that $L=S_r(p)_{\spc{U}_d}\ne \emptyset$. Show that $L$ is isometric to $\spc{U}_{\ell}$; find $\ell$ in terms of $r$ and $d$. \end{subthm} \begin{subthm}{ex:sphere-in-urysohn:midpoint} Let $\ell=\dist{p}{q}{\spc{U}_d}$. Show that the subset $M\subset\spc{U}_d$ of midpoints between $p$ and $q$ is isometric to $\spc{U}_\ell$. \end{subthm} \begin{subthm}{ex:sphere-in-urysohn:homogeneous} Show that $\spc{U}_d$ is \emph{not} countable-set homogeneous; that is, there is a distance-preserving map from a countable subset of $\spc{U}_d$ to $\spc{U}_d$ that cannot be extended to an isometry of $\spc{U}_d$. \end{subthm} \end{thm} In fact, the Urysohn space is compact-set homogeneous; more precisely the following theorem holds. \begin{thm}{Theorem}\label{thm:compact-homogeneous} Let $K$ be a compact set in a ($d$-)Urysohn space $\spc{U}$. Then any distance-preserving map $K\to \spc{U}$ can be extended to an isometry of $\spc{U}$. \end{thm} A proof can be obtained by modifying the proofs of \ref{prop:completion-univeral} and \ref{thm:urysohn-unique} the same way as it is done in \ref{ex:compact-extension}. \begin{thm}{Exercise}\label{ex:homogeneous} Which of the following metric spaces are 1-point set homogeneous, finite set homogeneous, compact set homogeneous, countable homogeneous? \begin{subthm}{ex:homogeneous:euclidean} Euclidean plane, \end{subthm} \begin{subthm}{ex:homogeneous:hilbert} Hilbert space $\ell^2$, \end{subthm} \begin{subthm}{ex:homogeneous:ell-infty} $\ell^\infty$, \end{subthm} \begin{subthm}{ex:homogeneous:ell-1} $\ell^1$ \end{subthm} \end{thm} \section{Remarks} The statement in \ref{ex:frechet} was proved by Maurice René Fréchet in the paper where he first defined metric spaces \cite{frechet}; its extension \ref{lem:kuratowski} was given by Kazimierz Kuratowski~\cite{kuratowski}. The question about the existence of a separable universal space was posted by Maurice René Fréchet and answered by Pavel Urysohn~\cite{urysohn}. The idea of Urysohn's construction was reused in graph theory; it produces the so-called {}\emph{Rado graph}, also known as {}\emph{Erd\H{o}s–R\'enyi graph} or \emph{random graph}; a good survey on the subject is given by Peter Cameron~\cite{cameron}.
{ "timestamp": "2021-04-13T02:21:25", "yymm": "2007", "arxiv_id": "2007.09846", "language": "en", "url": "https://arxiv.org/abs/2007.09846", "abstract": "We discuss domestic affairs of metric spaces, keeping away from any extra structure. Topics include universal spaces, injective spaces, Hausdorff and Gromov--Hausdorff convergences, and ultralimits.", "subjects": "Metric Geometry (math.MG)", "title": "Pure metric geometry: introductory lectures", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9915543741819548, "lm_q2_score": 0.8267117962054049, "lm_q1q2_score": 0.81972969771529 }
https://arxiv.org/abs/math/0601638
Upper bounds for edge-antipodal and subequilateral polytopes
A polytope in a finite-dimensional normed space is subequilateral if the length in the norm of each of its edges equals its diameter. Subequilateral polytopes occur in the study of two unrelated subjects: surface energy minimizing cones and edge-antipodal polytopes. We show that the number of vertices of a subequilateral polytope in any d-dimensional normed space is bounded above by (d/2+1)^d for any d >= 2. The same upper bound then follows for the number of vertices of the edge-antipodal polytopes introduced by I.Talata (Period. Math. Hungar. 38 (1999), 231--246). This is a constructive improvement to the result of A.Pór (to appear) that for each dimension d there exists an upper bound f(d) for the number of vertices of an edge-antipodal d-polytopes. We also show that in d-dimensional Euclidean space the only subequilateral polytopes are equilateral simplices.
\section{Notation} Denote the $d$-dimensional real linear space by $\numbersystem{R}^d$, a norm on $\numbersystem{R}^d$ by $\norm{\cdot}$, its unit ball by $B$, and the ball with centre $x$ and radius $r$ by $B(x,r)$. Denote the diameter of a set $C\subseteq \numbersystem{R}^d$ by $\diam(C)$, and (if it is measurable) its volume (or $d$-dimensional Lebesgue measure) by $\vol(C)$. The \emph{dual norm} $\norm{\cdot}^\ast$ is defined by $\norm{x}^\ast:=\sup\{\ipr{x}{y} : \norm{y}\leq 1\}$, where $\ipr{\cdot}{\cdot}$ is the inner product on $\numbersystem{R}^d$. Denote the number of elements of a finite set $S$ by $\card{S}$. The \emph{difference body} of a set $S\subseteq\numbersystem{R}^d$ is $S-S:=\{x-y:x,y\in S\}$. A \emph{polytope} is the convex hull of finitely many points in some $\numbersystem{R}^d$. A \emph{$d$-polytope} is a polytope of dimension $d$. A \emph{convex body} $C$ is a compact convex subset of $\numbersystem{R}^d$ with nonempty interior. The boundary of $C$ is denoted by $\bd C$. Given any convex body $C$ we define the \emph{relative norm $\norm{\cdot}_C$ determined by $C$} to be the norm with unit ball $C-C$, or equivalently, \[ \norm{x}_C := \sup\{\lambda>0: a+\lambda x\in C \text{ for some } a\in C\}.\] See \cite{Grunbaum, Ba, MR97f:52001} for background on polytopes, convexity, and finite-dimensional normed spaces. \section{Introduction} \subsection{Antipodal and edge-antipodal polytopes} A $d$-polytope $P$ is \emph{antipodal} if for any two vertices $x$ and $y$ of $P$ there exist two parallel hyperplanes, one through $x$ and one through $y$, such that $P$ is contained in the closed slab bounded by the two hyperplanes. Klee \cite{Klee} posed the problem of finding an upper bound for the number of vertices of an antipodal $d$-polytope in terms of $d$. Danzer and Gr\"unbaum \cite{MR25:1488} proved the sharp upper bound of $2^d$. See \cite{MS} for a recent survey. A $d$-polytope $P$ is \emph{edge-antipodal} if for any two vertices $x$ and $y$ joined by an edge there exist two parallel hyperplanes, one through $x$ and one through $y$, such that $P$ is contained in the closed slab bounded by the two hyperplanes. This notion was introduced by Talata \cite{Talata}, who conjectured that the number of vertices of an edge-antipodal $3$-polytope is bounded above by a constant. Csik\'os \cite{Csikos} proved an upper bound of $12$, and K.~Bezdek, Bisztriczky and B\"or\"oczky \cite{BBB} gave the sharp upper bound of $8$. P\'or \cite{Por} proved that the number of vertices of an edge-antipodal $d$-polytope is bounded above by a function of $d$. However, his proof is existential, with no information on the size of the upper bound. Our main result is an explicit bound. \begin{theorem}\label{th1} Let $d\geq 2$. Then the number of vertices of an edge-antipodal $d$-polytope is bounded above by $(\frac{d}{2}+1)^d$. \end{theorem} In the plane, an edge-antipodal polytope is clearly antipodal, and in this case the above theorem is sharp. The bound given is not sharp for $d\geq 3$ (since the bound in Theorem~\ref{th2} below is not sharp). In \cite{BBB} it is stated without proof that all edge-antipodal $3$-polytopes are antipodal. On the other hand, Talata has an example of an edge-antipodal $d$-polytope that is not antipodal for each $d\geq 4$ (see \cite{Csikos} and Section~\ref{s4} below). Most likely the largest number of vertices of an edge-antipodal $d$-polytope has an upper bound exponential in $d$, perhaps even $2^d$. We also mention the paper by Bisztriczky and B{\"o}r{\"o}czky \cite{BB} discussing edge-antipodal $3$-polytopes. Theorem~\ref{th1} is proved by considering a metric relative of edge-antipodal polytopes, discussed next. \subsection{Equilateral and subequilateral polytopes} A polytope $P$ is \emph{equilateral} with respect to a norm $\norm{\cdot}$ on $\numbersystem{R}^d$ if its vertex set is an \emph{equidistant set}, i.e., the distance between any two vertices is a constant. This notion was first considered by Petty \cite{MR43:1051}, who showed that equilateral polytopes are antipodal, hence have at most $2^d$ vertices. We now introduce the following natural weakening of this notion, analogous to the weakening from antipodal to edge-antipodal. We say that a $d$-polytope $P$ is \emph{subequilateral} with respect to a norm $\norm{\cdot}$ on $\numbersystem{R}^d$ if the length of each of its edges equals its diameter. Although not explicitly given a name, the vertex sets of subequilateral polytopes appear in the study of surface energy minimizing cones by Lawlor and Morgan \cite{MR95i:58051}; see Section~\ref{s4} for a discussion. It is well-known and easy to prove that an edge-antipodal polytope $P$ is subequilateral with diameter $1$ in the relative norm $\norm{\cdot}_P$ determined by $P$ \cite{Talata, Csikos}. It is also easy to see that any subequilateral polytope is edge-antipodal. In order to prove Theorem~\ref{th1} it is therefore sufficient to bound the number of vertices of a subequilateral $d$-polytope. \begin{theorem}\label{th2} Let $d\geq 2$. Then the number of vertices of a subequilateral $d$-polytope with respect to some norm $\norm{\cdot}$ is bounded above by $(\frac{d}{2}+1)^d$. \end{theorem} The proof is in Section~\ref{s2}. In two-dimensional normed spaces subequilateral polytopes are always equilateral. Therefore, the above theorem is sharp for $d=2$. By analyzing equality in the proof of Theorem~\ref{th2}, it can be seen that the bound is not sharp for $d\geq 3$. Since edge-antipodal $3$-polytopes have at most $8$ vertices, with equality only for parallelepipeds \cite{BBB}, it follows that subequilateral $3$-polytopes with respect to any norm has size at most $8$, with equality only if the unit ball of the norm is a parallelepiped homothetic to the polytope. We finally mention that in Euclidean $d$-space $\mathbb{E}^d$ the only subequilateral polytopes are equilateral simplices, and give a proof. In the proof we have to consider subequilateral polytopes in spherical spaces, making it possible to formulate a more general theorem for spaces of constant curvature. Note that if we restrict ourselves to a hemisphere of the $d$-sphere $\mathbb{S}^d$ in $\mathbb{E}^{d+1}$, the notion of a polytope can be defined without ambiguity. The definition of a subequilateral polytope then still makes sense in in a hemisphere of $\mathbb{S}^d$, as well as in hyperbolic $d$-space $\mathbb{H}^d$. \begin{theorem}\label{th3} Let $P$ be a subequilateral $d$-polytope in either $\mathbb{E}^d$, $\mathbb{H}^d$, or a hemisphere of $\mathbb{S}^d$. Then $P$ is an equilateral $d$-simplex. \end{theorem} \begin{proof} The proof is by induction on $d\geq 1$, with $d=1$ trivial and $d=2$ easy. Suppose now $d\geq 3$. Let $P$ be a subequilateral $d$-polytope in any of the three spaces. By induction all facets of $P$ are equilateral simplices. In particular, $P$ is simplicial. Since $d\geq 3$, it is sufficient to show that $P$ is simple (see section~4.5 and exercise~4.8.11 of \cite{Grunbaum}). Consider any vertex $v$ with neighbours $v_1,\dots,v_k$, $k\geq d$. Then $v_1,\dots,v_k$ are contained in an open hemisphere $S$ of the $(d-1)$-sphere of radius $\diam(P)$ and centre $v$. (This sphere will be isometric to some sphere in $\mathbb{E}^{d}$, not necessarily of radius $\diam(P)$.) Consider the $(d-1)$-polytope $P'$ in $S$ generated by $v_1,\dots,v_k$ and any facet of $P'$ with vertex set $F\subset\{v_1,\dots,v_k\}$. There exists a great sphere $C$ of $S$ passing through $F$ with $P'$ in one of the closed hemispheres determined by $C$. It follows that the hyperplane $H$ generated by $C$ and $v$ passes through $F\cup\{v\}$, and $P$ is contained in one of the closed half spaces bounded by $H$. Therefore, $F\cup\{v\}$ is the vertex set of a facet of $P$. Similarly, it follows that for any vertex set $F$ of a facet of $P$ containing $v$, $F\setminus\{v\}$ is the vertex set of a facet of $P'$. Therefore, any edge $v_iv_j$ of $P'$ is an edge of $P$, hence of length the diameter of $P$. It follows that the distance between $v_i$ and $v_j$ in $H$ is the diameter of $P'$ as measured in $H$. This shows that $P'$ is subequilateral in $H$, and so by induction is an equilateral $(d-1)$-simplex. Therefore, $k=d$, giving that $P$ is a simple polytope, which finishes the proof. \end{proof} \section{A measure of non-equidistance}\label{s2} The key to the proof of Theorem~\ref{th2} is a lower bound for the distance between two nonadjacent vertices of a subequilateral polytope. For any finite set of points $V$ we define \[\lambda(V;\norm{\cdot})=\diam(V)/\min_{x,y\in V, x\neq y}\norm{x-y}.\] Since $\lambda(V;\norm{\cdot})\geq 1$, with equality if and only if $V$ is equidistant in the norm $\norm{\cdot}$, this functional measures how far $V$ is from being equidistant. The next lemma generalizes the theorem of Petty \cite{MR43:1051} and Soltan \cite{MR52:4127} that the number of points in an equidistant set is bounded above by $2^d$. In \cite{MR93d:52009} a proof of the $2^d$-upper bound was given using the isodiametric inequality for finite-dimensional normed spaces due to Busemann (equation~(2.2) on p.~241 of \cite{B}; see also Mel'nikov \cite{MR27:6191}). However, since the isodiametric inequality has a quick proof using the Brunn-Minkowski inequality \cite{BZ}, it is not surprising that the latter inequality occurs in the following proof. \begin{lemma}\label{l2} Let $V$ be a finite set in a $d$-dimensional normed space. Then $\card{V}\leq(\lambda(V;\norm{\cdot})+1)^d$. \end{lemma} \begin{proof} Let $\lambda=\lambda(V;\norm{\cdot})$. By scaling we may assume that $\diam(V)=\lambda$. Then $\norm{x-y}\geq 1$ for all $x,y\in V$, $x\neq y$, hence the balls $B(v,1/2)$, $v\in V$, have disjoint interiors. Define $C=\bigcup_{v\in V}B(v,1/2)$. Then $\vol(C)=\card{V}(1/2)^d\vol(B)$ and $\diam(C)\leq 1+\lambda$. By the Brunn-Minkowski inequality \cite{BZ} we obtain $\vol(C-C)^{1/d}\geq\vol(C)^{1/d}+\vol(-C)^{1/d}$. Noting that $C-C\subseteq(1+\lambda)B$, the result follows. \end{proof} In order to find an upper bound on the number of vertices of a subequilateral polytope with vertex set $V$, it remains to bound $\lambda(V;\norm{\cdot})$ from above. \begin{lemma}\label{l3} Let $d\geq 2$ and let $V$ be the vertex set of a subequilateral $d$-polytope. Then $\lambda(V;\norm{\cdot})\leq d/2$. \end{lemma} \begin{proof Let $P$ be a subequilateral $d$-polytope of diameter $1$, and let $V$ be its vertex set. We have to show that $\norm{x-y}\geq2/d$ for any distinct $x,y\in V$. Since this follows from the definition if $xy$ is an edge of $P$, we assume without loss that $xy$ is not an edge of $P$. Then $xy$ intersects the convex hull $P'$ of $V\setminus\{x,y\}$ in a (possibly degenerate) segment, say $x'y'$, with $x$, $x'$, $y'$, $y$ in this order on $xy$. Let $F_x$ and $F_y$ be facets of $P'$ containing $x'$ and $y'$, respectively. We show that $\norm{x-x'}\geq1/d$. For each vertex $z$ of $F_x$, $xz$ is an edge of $P$, hence $\norm{x-z}=1$. By Carath\'eodory's theorem \cite[(2.2)]{Ba}, there exist $d$ vertices $z_1,\dots,z_d$ of the $(d-1)$-polytope $F_x$ and real numbers $\lambda_1,\dots,\lambda_d$ such that \[ x'=\sum_{i=1}^d\lambda_i z_i,\quad \lambda_i\geq 0,\quad \sum_{i=1}^d\lambda_i=1.\] Suppose without loss that $\lambda_d=\max_i\lambda_i$. Then $\lambda_d\geq 1/d$. By the triangle inequality we obtain \begin{align*} \norm{x'-z_d} & = \norm{\sum_{i=1}^{d-1}\lambda_i(z_i-z_d)}\leq \sum_{i=1}^{d-1}\lambda_i\norm{z_i-z_d}\\ &\leq \sum_{i=1}^{d-1}\lambda_i=1-\lambda_d\leq 1-\frac{1}{d}, \end{align*} and \begin{align*} \norm{x-x'} & \geq\norm{x-z_d}-\norm{x'-z_d}\\ & \geq 1-(1-\frac{1}{d})=\frac{1}{d}. \end{align*} Similarly, $\norm{y-y'}\geq1/d$, and we obtain $\norm{x-y}\geq 2/d$. \end{proof} Lemmas~\ref{l2} and \ref{l3} now imply Theorem~\ref{th2}.\qed \section{Concluding remarks}\label{s4} \subsection{Sharpness of Lemma~\ref{l3}} The following example shows that Lemma~\ref{l3} cannot be improved in general. Consider the subspace $X=\{(x_1,\dots,x_{d+1}):\sum_{i=1}^dx_i=0\}$ of $\numbersystem{R}^{d+1}$ with the $\ell_1$ norm $\norm{(x_1,\dots,x_{d+1})}_1:=\sum_{i=1}^{d+1}\abs{x_i}$. Let the standard unit vector basis of $\numbersystem{R}^{d+1}$ be $e_1,\dots,e_{d+1}$. Let $c=\sum_{i=1}^d e_i$. Then $V=\{de_i-c:i=1,\dots,d\}\cup\{\pm 2e_{d+1}\}$ is the vertex set of a $d$-polytope $P$ in $X$, with all intervertex distances equal to $2d$, except for the distance between $\pm 2e_{d+1}$, which is $4$. It follows that $P$ is subequilateral and $\lambda(V;\norm{\cdot})=d/2$. However, the above polytope $P$ is in fact antipodal, and so it is equilateral in $\norm{\cdot}_P$, which gives $\lambda(V;\norm{\cdot}_P)=1$. It is easy to see that for any polytope $P$ subequilateral with respect to some norm $\norm{\cdot}$, and with vertex set $V$, we have $\lambda(V,\norm{\cdot})\leq\lambda(V,\norm{\cdot}_P)$. One may therefore hope that for the norm $\norm{\cdot}_P$ the upper bound in Lemma~\ref{l3} may be improved, thus giving a better bound in Theorem~\ref{th1}. The following example shows that any such improved upper bound will still have to be at least $(d-1)/2$, indicating that essentially new ideas will be needed to improve the upper bounds in Theorems~\ref{th1} and \ref{th2}. We consider Talata's example \cite{Csikos} of an edge-antipodal polytope that is not antipodal. Let $d\geq 4$, $e_1,\dots,e_{d}$ be the standard basis of $\numbersystem{R}^d$, $p=\frac{2}{d-1}\sum_{i=1}^{d-1}e_i$, and $\lambda=(d-1)/2-\varepsilon>1$ for some small $\varepsilon>0$. Then the polytope $P$ with vertex set $V=\{o,e_1,\dots,e_d,p,e_d+\lambda p\}$ is edge-antipodal but not antipodal. In fact, $\diam(V)\leq 1$ by definition of $\norm{\cdot}_P$, and since $\norm{e_d-o}_P=1$ and $\norm{p-o}_P=1/\lambda$, we obtain $\lambda(V,\norm{\cdot}_P)\geq\lambda$, which is arbitrarily close to $(d-1)/2$. \subsection{Subequilateral polytopes in the work of Lawlor and Morgan} Define the \emph{$\norm{\cdot}$-energy} of a hypersurface $S$ in $\numbersystem{R}^d$ to be $\norm{S}:=\int_S\norm{n(x)}dx$, where $n(x)$ is the Euclidean unit normal at $x\in S$. In \cite{MR95i:58051} a sufficient condition is given to obtain an energy minimizing hypersurface partitioning a convex body. We restate a special case of the ``General Norms Theorem I'' in \cite[pp.~66--67]{MR95i:58051} in terms of subequilateral polytopes. (In the notation of \cite{MR95i:58051} we take all the norms $\Phi_{ij}$ to be the same. Then the points $p_1,\dots,p_m$ in the hypothesis form an equidistant set with respect to the dual norm. The weakening of the hypothesis in the last sentence of the General Norms Theorem I is easily seen to be equivalent to the requirement that $p_1,\dots,p_m$ is the vertex set of a subequilateral polytope.) We refer to \cite{MR95i:58051} for the simple and enlightening proof using the divergence theorem. \begin{lmtheorem} Let $\norm{\cdot}$ be a norm on $\numbersystem{R}^n$, and let $p_1,\dots,p_m\in\numbersystem{R}^n$ be the vertex set of a subequilateral polytope of $\norm{\cdot}$-diameter $1$. Let $\Sigma=\bigcup H_{ij}\subset C$ be a hypersurface which partitions some convex body $C$ into regions $R_1,\dots,R_m$ with $R_i$ and $R_j$ separated by a piece $H_{ij}$ of a hyperplane such that the parallel hyperplane passing through $p_i-p_j$ supports the unit ball $B$ at $p_i-p_j$. Then for any hypersurface $M=\bigcup M_{ij}$ which also separates the $R_i\cap\bd C$ from each other in $C$, with the regions touching $R_i\cap\bd C$ and $R_j\cap\bd C$ facing each other across $M_{ij}$, we have $\norm{\Sigma}^\ast\leq\norm{M}^\ast$, i.e.\ $\Sigma$ minimizes $\norm{\cdot}^\ast$-energy, where $\norm{\cdot}^\ast$ is the norm dual to $\norm{\cdot}$. \end{lmtheorem}
{ "timestamp": "2006-01-26T13:51:21", "yymm": "0601", "arxiv_id": "math/0601638", "language": "en", "url": "https://arxiv.org/abs/math/0601638", "abstract": "A polytope in a finite-dimensional normed space is subequilateral if the length in the norm of each of its edges equals its diameter. Subequilateral polytopes occur in the study of two unrelated subjects: surface energy minimizing cones and edge-antipodal polytopes. We show that the number of vertices of a subequilateral polytope in any d-dimensional normed space is bounded above by (d/2+1)^d for any d >= 2. The same upper bound then follows for the number of vertices of the edge-antipodal polytopes introduced by I.Talata (Period. Math. Hungar. 38 (1999), 231--246). This is a constructive improvement to the result of A.Pór (to appear) that for each dimension d there exists an upper bound f(d) for the number of vertices of an edge-antipodal d-polytopes. We also show that in d-dimensional Euclidean space the only subequilateral polytopes are equilateral simplices.", "subjects": "Metric Geometry (math.MG)", "title": "Upper bounds for edge-antipodal and subequilateral polytopes", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9915543723101532, "lm_q2_score": 0.8267117876664789, "lm_q1q2_score": 0.8197296877010402 }
https://arxiv.org/abs/0912.5031
On two and three periodic Lyness difference equations
We describe the sequences {x_n}_n given by the non-autonomous second order Lyness difference equations x_{n+2}=(a_n+x_{n+1})/x_n, where {a_n}_n is either a 2-periodic or a 3-periodic sequence of positive values and the initial conditions x_1,x_2 are as well positive. We also show an interesting phenomenon of the discrete dynamical systems associated to some of these difference equations: the existence of one oscillation of their associated rotation number functions. This behavior does not appear for the autonomous Lyness difference equations.
\section{Introduction and main result} This paper fully describes the sequences given by the non-autonomous second order Lyness difference equations \begin{equation}\label{eq} x_{n+2}\,=\,\frac{a_n+x_{n+1}}{x_n}, \end{equation} where $\{a_n\}_n$ is a $k$-periodic sequence taking positive values, $k=2,3,$ and the initial conditions $x_1,x_2$ are as well positive. This question is proposed in \cite[Sec. 5.43]{CL}. Recall that non-autonomous recurrences appear for instance as population models with a variable structure affected by some seasonality \cite{ES1,ES2}, where $k$ is the number of seasons. Some dynamical issues of similar type of equations have been studied in several recent papers \cite{BHS,CJK,deA,JKN,KN,FJL,GKL}. Recall that when $k=1,$ that is $a_n=a>0,$ for all $n\in\mathbb{N}$, then (\ref{eq}) is the famous Lyness recurrence which is well understood, see for instance \cite{BR,Z}. The cases $k=2,3$ have been already studied and some partial results are established. For both cases it is known that the solutions are persistent near a given $k$-periodic solution, which is stable. This is proved by using some known invariants, see \cite{KN, FJL, GKL}. Recall that in our context it is said that a solution $\{x_n\}_n$ is persistent if there exist two real positive constants $c$ and $C,$ which depend on the initial conditions, such that for all $n\ge1, 0<c<x_n<C<\infty.$ We prove: \begin{teo}\label{main} Let $\{x_n\}_n$ be any sequence defined by (\ref{eq}) and $k\in\{2,3\}$. Then it is persistent. Furthermore, either \begin{enumerate} \item [(a)] the sequence $\{x_n\}_n$ is periodic, with period a multiple of $k$; or \item [(b)] the sequence $\{x_n\}_n$ densely fills one or two (resp. one, two or three) disjoint intervals of $\mathbb{R}^+$ when $\{a_n\}_n$ is 2-periodic (resp. 3-periodic). Moreover it is possible by algebraic tools to distinguish which is the situation. \end{enumerate} \end{teo} Our approach to describe the sequences $\{x_n\}_n$ is based on the study of the natural dynamical system associated to (\ref{eq}) and on the results of \cite{CGM}. The main tool that allows to distinguish the number of intervals for the adherence of the sequences $\{x_n\}_n$ is the computation of several resultants, see Section~\ref{provaA}. It is worth to comment that Theorem~\ref{main} is an extension of what happens in the classical case $k=1$. There, the same result holds but in statement (b) only appears one interval. Our second main result will prove that there are other more significative differences between the case $k=1$ and the cases $k=2,3.$ These differences are related with the lack of monotonicity of certain rotation number functions associated to the dynamical systems given by the Lyness recurrences, see Theorem~\ref{teonomonotonia}. The behaviors of these rotation number functions are important for the understanding of the recurrences, because they give the possible periods for them, see \cite{BR,BC,Z}. On the other hand in \cite{deA,GKL} it is proved that, at least for some values of $\{a_n\}_n,$ the behaviour of $\{x_n\}_n$ for the case $k=5$ is totally different. In particular unbounded positive solutions appear. In the forthcoming paper \cite{CGM3} we explore in more detail the differences between the cases $k=1,2,3$ and $k\ge4.$ This paper is organized as follows: Section~\ref{ds} presents the difference equations that we are studying as discrete dynamical systems and we state our main results on them, see Theorems~\ref{rotacions} and \ref{teonomonotonia}. Section~\ref{pt2} is devoted to the proof of Theorem~\ref{rotacions}. By using it, in Section~\ref{provaA}, we prove Theorem~\ref{main} and we give some examples of how to apply it to determine the number of closed intervals of the adherence of $\{x_n\}_n$. In Section~\ref{someproperties} we demonstrate Theorem~\ref{teonomonotonia} and we also present some examples where we study in more detail the rotation number function of the dynamical systems associated to \eqref{eq}. \section{Main results from the dynamical systems point of view}\label{ds} In this section we reduce the study of the sequence $\{x_n\}_n$ to the study of some discrete dynamical systems and we state our main results on them. First we introduce some notations. When $k=2,$ set \begin{equation}\label{k=2}a_n\,=\,\left\{\begin{array}{lllr} a&\mbox{for}&n=2\ell+1,\\ b&\mbox{for}&\,n=2\ell, \end{array}\right.\end{equation} and when $k=3,$ set \begin{equation}\label{k=3}a_n\,=\,\left\{\begin{array}{lllr} a&\mbox{for}&n=3\ell+1,\\ b&\mbox{for}&n=3\ell+2,\\ c&\mbox{for}&n=3\ell, \end{array}\right.\end{equation} where $\ell\in\mathbb{N}$ and $a>0,b>0$ and $c>0.$ We also consider the maps $F_\alpha(x,y)$, with $\alpha\in\{a,b,c\},$ as $$F_{\alpha}(x,y)=\left(y,\frac{\alpha+y}{x}\right),$$ defined on the open invariant set ${Q^+}:=\{(x,y):x>0,y>0\}\subset{\mathbb R}^2.$ Consider for instance $k=2.$ The sequence given by (\ref{eq}), \begin{equation}\label{seqq} x_1,x_2,x_3,x_4,x_5,x_6,x_7,\ldots, \end{equation} can be seen as \[ (x_1,x_2)\xrightarrow{F_a}(x_2,x_3)\xrightarrow{F_b}(x_3,x_4) \xrightarrow{F_a}(x_4,x_5)\xrightarrow{F_b}(x_5,x_6)\xrightarrow{F_a}\cdots. \] Hence the behavior of (\ref{seqq}) can be obtained from the study of the dynamical system defined in ${Q^+}$ by the map: $$F_{b,a}(x,y):=F_b\circ F_a (x,y)=\left(\frac {a+y}{x},\frac {a+b x+ y}{x y}\right). $$ Similarly, for $k=3$ we can consider the map: $$ F_{c,b,a}(x,y):=F_c \circ F_b\circ F_a(x,y)=\left(\frac {a+bx+y}{xy},\frac {a+bx+y+cxy}{y \left( a+y \right)} \right). $$ Notice that both maps have an only fixed point in ${Q^+},$ which depends on $a,b$ (and $c$), that for short we denote by $\bf{p}.$ It is easy to interpret the invariants for (\ref{eq}) and $k=2,3,$ given in \cite{JKN,KN}, in terms of first integrals of the above maps, see also Lemma~\ref{elemental}. We have that $$V_{b,a}(x,y):= {\frac {ax^2y+bxy^2+bx^2+ay^2+(b^2+a)x+(b+a^2)y+ab}{xy}}, $$ is a first integral for $F_{b,a}$ and $$V_{c,b,a}(x,y):= {\frac {c{x}^{2}y+ax{y}^{2}+b{x}^{2}+b{y}^{2}+(a+bc)x+(c+ab)y+ac}{xy}},$$ is a first integral for $F_{c,b,a}$. The topology of the level sets of these integrals in ${Q^+}$ as well as the dynamics of the maps restricted to them is described by the following result, that will be proved in Section~\ref{pt2}. \begin{teo}\label{rotacions} \begin{itemize} \item[(i)] The level sets of $V_{b,a}$ (resp. $V_{c,b,a}$) in $Q^+\setminus\{\bf{p}\}$ are diffeomorphic to circles surrounding $\bf p$, which is the unique fixed point of $F_{b,a}$ (resp. $F_{c,b,a}$). \item[(ii)] The action of $F_{b,a}$ (resp. $F_{c,b,a}$) on each level set of $V_{b,a}$ (resp. $V_{c,b,a}$) contained in $Q^+\setminus\{\bf{p}\}$ is conjugated to a rotation of the circle. \end{itemize} \end{teo} Once a result like the above one is established the study of the possible periods of the sequences $\{x_n\}_n$ given by (\ref{eq}) is quite standard. It suffices, first to get the rotation interval, which is the open interval formed by all the rotation numbers given by the above theorem, varying the level sets of the first integrals. Afterwards, it suffices to find which are the denominators of all the irreducible rational numbers that belong to the corresponding interval, see \cite{BC,CGM2,Z}. The study of the rotation number of these kind of rational maps is not an easy task, see again \cite{BR,BC,CGM2,Z}. In particular, in \cite{BR} was proved that the rotation number function parameterized by the energy levels of the Lyness map $F_a, a\ne1,$ is always monotonous, solving a conjecture of Zeeman given in \cite{Z}, see also \cite{ER}. As far as we know, in this paper we give the first simple example for which this rotation number function is neither constant nor monotonous. We prove: \begin{teo}\label{teonomonotonia} There are positive values of $a$ and $b$, such that the rotation number function $\rho_{b,a}(h)$ of $F_{b,a}$ associated to the closed ovals of $\{V_{b,a}=h\}\subset{Q^+}$ has a local maximum. \end{teo} Hence, apart from the known behaviors for the autonomous Lyness maps, that is global periodicity or monotonicity of the rotation number function, which trivially holds for $F_{b,a},$ taking for instance $a=b=1$ or $a=b\ne1,$ respectively, there appear more complicated behaviors for the rotation number function. Our proof of this result relies on the study of lower and upper bounds for the rotation number of $F_{b,a}$ on a given oval of a level set of $V_{b,a}$ given for some $(a,b)\in(\mathbb{Q}^+)^2$ and $\{V_{b,a}(x,y)=V_{b,a}(x_0,y_0)\},$ for $(x_0,y_0)\in(\mathbb{Q}^+)^2$. This can be done because the map on this oval is conjugated to a rotation and it is possible to use an algebraic manipulator to follow and to order a finite number iterates on it, which are also given by points with rational coordinates. So, only exact arithmetic is used. A similar study could be done for $F_{c,b,a}$. \section{Proof of Theorem~\ref{rotacions}}\label{pt2} {\rec {\it Proof of (i) of Theorem \ref{rotacions}.}} The orbits of $F_{b,a}$ and $F_{c,b,a}$ lie on the level sets $V_{b,a}=h$ and $V_{c,b,a}=h$ respectively. These level sets can be seen as the algebraic curves given by $$ C_2:=\{c_2(x,y)=ax^2y+bxy^2+bx^2-hxy+ay^2+(b^2+a)x+(b+a^2)y+ab=0\} $$ and $$ C_3:=\{c_3(x,y)=cx^2y+axy^2+bx^2-hxy+by^2+(a+bc)x+(c+ab)y+ac=0\}, $$ respectively. Taking homogeneous coordinates on the projective plane $P{\mathbb R}^2$ both curves $C_2$ and $C_3$ have the form $$ C:=\{Sx^2y+Txy^2+Ux^2z+Vxyz+Wy^2z+Lxz^2+Myz^2+Nz^3=0\}.$$ In order to find the branches of them tending to infinity, we examine the directions of approach to infinity ($z=0$) in the local charts determined by $x=1$ and $y=1$ respectively. In the local chart given by $x=1$, the curve $C$ writes as $$ Sy+Ty^2+Uz+Vyz+Wy^2z+Lz^2+Myz^2+Nz^3=0$$ and it meets the straight line at infinity $z=0$ when $y(S+Ty)=0.$ Since for both curves $C_2$ and $C_3$ the coefficients $S$ and $T$ are positive, the only intersection point that could give points in ${Q^+}$ is $(y,z)=(0,0).$ The algebraic curve $C$ arrives to $(y,z)=(0,0)$ tangentially to the line $Sy+Uz=0.$ Since for both curves, $C_2$ and $C_3,$ the coefficients $S$ and $U$ are also positive, we have that the branches of the level sets tending to infinity are not included in~$Q^+.$ An analogous study can be made in the chart given by $y=1,$ obtaining the same conclusions. Moreover, it can be easily checked that in the affine plane both curves $C_2$ and $C_3$ do not intersect the part of the axes $x=0$ and $y=0$ which is in the boundary of ${Q^+}$. In summary, there are no branches of the curves $C_2$ and $C_3$ tending to infinity or crossing the axes $x=0$ and $y=0$ in $Q^+$, and therefore the connected components of $C_i\cap Q^+$ for $i=2,3$ are bounded. Notice that this result in particular already implies the persistence of the sequences given by \eqref{eq}. Consider $k=2$. We claim the following facts: \begin{enumerate}[(a)] \item In ${Q^+}$, the set of fixed points of $F_{b,a}$ and the set of singular points of $C_1$ coincide and they only contain the point ${\bf p}=(\bar x_,\bar y)$. \item The function $V_{b,a}(x,y),$ has a local minimum at $\bf p$. \end{enumerate} We remark that item (b) is already known. We present a new simple proof for the sake of completeness. From the above claims and the fact that the connected components of the level sets of $V_{b,a}$ in ${Q^+}$ are bounded it follows that the level sets of $V_{b,a}$ in ${Q^+}\setminus\{\bf{p}\}$ are diffeomorphic to circles. Let us prove the above claims. The fixed points of $F_{b,a}$ are given by $$ \begin{cases} x=\frac{a+y}{x}, \\ y=\frac{a+bx+y}{xy}, \end{cases} \Leftrightarrow \begin{cases} x^2=a+y, \\ x(y^2-b)=a+y, \end{cases} $$ and so $x^2=x(y^2-b).$ Hence in $Q^+,$ we have that $x=y^2-b$ and the above system is equivalent to $$ \begin{cases} x=y^2-b, \\ xy^2-bx-y-a=0, \end{cases} \Leftrightarrow \begin{cases} x=y^2-b, \\ P(y):=y^4-2by^2-y+b^2-a=0. \end{cases} $$ It is not difficult to check that the last system of equations is precisely the one that gives the critical points of the curves $V_{a,b}=h.$ Moreover, from the first equation it is necessary that $x=y^2-b>0$ and hence $y>\sqrt{b}.$ Since $P(y)$ has only one real root in $(\sqrt{b},\infty)$ the uniqueness of the critical point holds. Let us prove that this critical point corresponds with a local minimum of $V_{b,a}.$ We will check the usual sufficient conditions given by the Hessian of $V_{b,a}$ at $\bf p$. Firstly, $$ \frac{\partial^2 }{\partial x^2}V_{b,a}(y^2-b,y)=2\,{\frac { \left( y+a \right) \left( ay+b \right) }{ \left({y}^{ 2}-b \right) ^{3}y}}>0\,\mbox{ for }y>\sqrt{b}.$$ Secondly, the determinant of the Hessian matrix at the points $(y^2-b,y)$ is $$ h(y)=\frac{f(y)}{(b-y^2)^4y^4}, $$ where $$f(y):=(by^2+a-b^2)(-b{y}^{6}+ 3\left( a+{b}^{2} \right) {y}^{4}+ 4\left( {a}^{2}+b \right) {y}^{3}+ 3b\left( 2a-{b}^{2} \right) {y}^{2}+b^2({b}^{2}-a ).$$ A tedious computation shows that $ f(y)=q(y)P(y)+r(y), $ with \begin{align*} r(y)=&\left(4{a}^{2}{b}^{2}+ 4{a}^{3}+4{b}^{3}+6ab \right) {y}^{3}+ \left(18{a}^{2}b+ 8{b}^{3}a+3{b}^{2} \right) {y}^{2}\\ &+ \left(-4{b}^{3}{a}^{2}+4{a}^{3}b-4{b}^{4} +12 a{b}^{2}+3{a}^{2} \right) y-8{b}^{4}a+5{a}^{2}{b}^{2}+3{a}^{3}. \end{align*} Observe that if $\bar y$ is the positive root of $P(y)$, then ${\rm sign}(h(\bar y ))={\rm sign}(r(\bar y ))$. Taking into account that $P(\bar y )=0$ implies that $a=\bar y ^4-2b\bar y ^2-\bar y +b^2$ we have that $$r(\bar y )= {\bar y}^{2} \left( 4\,{\bar y}^{3}-4\,b\bar y -1 \right) \left( b\bar y +1-{\bar y}^{3} \right) ^{2} \left( b-{\bar y}^{2} \right) ^{2}.$$ So, $ {\rm sign}\left( 4\,{\bar y}^{3}-4b\bar y -1\right)={\rm sign}\left(P'(\bar y )\right). $ Since $P(\sqrt{b})=-\sqrt{b}-a<0,$ $\lim\limits_{y\to\infty} P(y)=+\infty$ and, on this interval, there is only one critical point of $P(y),$ which is simple, we get that $P'(\bar y )>0$ and so $h(\bar y )>0.$ Hence ${\bf p}$ is a local minimum of $V_{b,a}(x,y)$, as we wanted to prove. The same kind of arguments work to end the proof for the case $k=3,$ but the computations are extremely more tedious. We only make some comments. The fixed points of $F_{c,b,a}$ in $Q^+$ are given by: \begin{equation*}\label{sistemafcba} \begin{cases} P(x):=x^5+ax^4-2x^3-(2a+bc)x^2+(1-b^2-c^2)x+a-bc=0,\, \\ y=Q(x):=\dfrac{(xb+a)}{(x-1)(x+1)}. \end{cases} \end{equation*} It can be proved again that they coincide with the singular points of $V_{c,b,a}$ in $Q^+.$ This fact follows from the computation of several suitable resultants between ${\partial V_{c,b,a}}/{\partial x},$ ${\partial V_{c,b,a}}/{\partial y}$ and $V_{c,b,a}.$ The uniqueness of the fixed point $\bf p$ in ${Q^+}$ can be shown as follows: since $Q(x)>0$ implies that $x>1$, we only need to search solutions of $P_5(x)=0$ in $(1,+\infty)$. With the new variable $z=x-1,$ \begin{equation*}\label{quantessolucions} \widetilde{P}(z):=P(z+1):=z^5+(c+5)z^4+(8+4c)z^3+(4-ab+4c)z^2-(a+b)z-(a+b)^2=0, \end{equation*} Since $\widetilde{P}(0)<0$; $\lim\limits_{z\to +\infty} \widetilde{P}(z)=+\infty$; and the Descarte's rule, we know that there is only one positive solution, as we wanted to see. Finally it can be proved that $\bf p$ is a non-degenerated local minimum of $V_{c,b,a}$. These computations are complicated, and they have been performed in a very smart way in \cite{KN}, so we skip them and we refer the reader to this last reference.\qed \subsection{Proof of (ii) of Theorem \ref{rotacions}} In \cite{CGM} it is proved a result that characterizes the dynamics of integrable diffeomorphisms having a \textsl{Lie Symmetry}, that is a vector field $X$ such that $ X(F(p))=(DF(p))\,X(p)$. Next theorem states it, particularized to the case we are interested. \begin{teo}[\cite{CGM}]\label{teor} Let ${\cal{U}}\subset \mathbb{R}^2$ be an open set and let $\Phi:{\cal{U}}\rightarrow {\cal{U}}$ be a diffeomorphism such that: \begin{enumerate} \item[(a)] It has a smooth regular first integral $V:{\cal{U}}\rightarrow {\mathbb R},$ having its level sets $\Gamma_h:=\{z=(x,y)\in{\cal{U}}\,:\, V(z)=h\}$ as simple closed curves. \item[(b)] There exists a smooth function $\mu:{\cal{U}}\rightarrow {\mathbb R}^+$ such that for any $z\in {\cal{U}},$ \begin{equation*}\label{mu} \mu(\Phi(z))=\det(D\Phi(z))\,\mu(z). \end{equation*} \end{enumerate} Then the map $\Phi$ restricted to each $\Gamma_h$ is conjugated to a rotation with rotation number $\tau(h)/T(h)$, where $T(h)$ is the period of $\Gamma_h$ as a periodic orbit of the planar differential equation \[ \dot z=\mu(z)\left(-\frac{\partial V(z)}{\partial y},\frac{\partial V(z)}{\partial x}\right) \] and $\tau(h)$ is the time needed by the flow of this equation for going from any $w\in\Gamma_h$ to $\Phi(w)\in\Gamma_h.$ \end{teo} Next lemma is one of the key points for finding a Lie symmetry for families of periodic maps, like the $2$ and $3$--periodic Lyness maps. \begin{lem}\label{mus} Let $\{G_a\}_{a\in A}$ be a family of diffeomorphisms of ${\cal{U}}\subset \mathbb{R}^2$. Suppose that there exists a smooth map $\mu:{\cal{U}}\to {\mathbb R}$ such that for any $a\in A$ and any $z\in {\cal{U}},$ the equation $ \mu(G_a(z))=\det(DG_a(z))\,\mu(z)$ is satisfied. Then, for every choice $a_1,\ldots,a_k\in A,$ we have $$ \mu(G_{[k]}(z))=\det(DG_{[k]}(z))\,\mu(z),$$ where $G_{[k]}=G_{a_k}\circ\cdots\circ G_{a_{2}}\circ G_{a_1}.$ \end{lem} \rec {\it Proof. }} % DEMOSTRACI\'{O It is only necessary to prove the result for $k=2$ because the general case follows easily by induction. Consider $a_1,a_2\in A$ then \begin{align*} \mu(G_{a_2,a_1}(z))& =\mu(G_{a_2}\circ G_{a_1}(z))=\det(DG_{a_2}(G_{a_1}(z)))\,\mu(G_{a_1}(z))= \\ &= \det(DG_{a_2}(G_{a_1}(z)))\,\det(DG_{a_1}(z))\,\mu(z)= \det(D (G_{a_2}\circ G_{a_1}(z)))\,\mu(z)=\\ &= \det(DG_{a_2,a_1}(z))\mu(z), \end{align*} and the lemma follows.\qed \medskip {\rec {\it Proof of (ii) of Theorem \ref{rotacions}.}} From part (i) of the theorem we know that the level sets of $V_{b,a}$ and $V_{c,b,a}$ in ${Q^+}\setminus\{{\bf p}\}$ are diffeomorphic to circles. Moreover these functions are first integrals of $F_{b,a}$ and $F_{c,b,a}$, respectively. Notice also that for any $a,$ the Lyness map $F_a(x,y)=(y,\frac{a+y}{x})$ satisfies $$\mu(F_{a}(x,y))=\det(DF_{a}(x,y))\mu(x,y),$$ with $\mu(x,y)=xy.$ Hence, by Lemma \ref{mus}, \[\mu(F_{b,a}(x,y))=\det(DF_{b,a}(x,y))\mu(x,y)\quad\mbox{and}\quad \mu(F_{c,b,a}(x,y))=\det(DF_{c,b,a}(x,y))\mu(x,y).\] Thus, from Theorem~\ref{teor}, the result follows.\qed \medskip It is worth to comment that once part (i) of the theorem is proved it is also possible to prove that the dynamics of $F_{b,a}$ (resp. $F_{c,b,a}$) restricted to the level sets of $V_{b,a}$ (resp. $V_{c,b,a}$) is conjugated to a rotation by using that they are given by cubic curves and that the map is birational, see \cite{JRV}. We prefer our approach because it provides a dynamical interpretation of the rotation number together with its analytic characterization. \section{Proof of Theorem \ref{main}}\label{provaA} In order to prove Theorem~\ref{main} we need a preliminary result. Consider the maps $F_{b,a}$ and $F_{a,b},$ jointly with their corresponding first integrals $V_{b,a}$ and $V_{a,b}.$ In a similar way consider $F_{c,b,a}\,,\,F_{a,c,b}$ and $F_{b,a,c}$ with $V_{c,b,a}\,,\,V_{a,c,b}$ and $V_{b,a,c}.$ Some simple computations prove the following elementary but useful lemma. Notice that it can be interpreted as the relation between the first integrals and the non-autonomous invariants. \begin{lem}\label{elemental} With the above notations: \begin{itemize} \item[(i)] $V_{b,a}(x,y)=V_{a,b}(F_a(x,y)).$ \item[(ii)] $V_{c,b,a}(x,y)=V_{a,c,b}(F_a(x,y))=V_{b,a,c}(F_b(F_a(x,y))).$ \end{itemize} \end{lem} {\rec {\it Proof of Theorem \ref{main}.}} We split the proof in two steps. For $k=2,3$ we first prove that there are only two types of behaviors for $\{x_n\}_n$, either this set of points is formed by $kp$ points for some positive integer $p,$ or it has infinitely many points whose adherence is given by at most $k$ intervals. Secondly, in this later case, we provide an algebraic way for studying the actual number of intervals. \noindent {\bf First step:} We start with the case $k=2.$ With the notation introduced in \eqref{k=2}, it holds that \[ F_{b,a}(x_{2n-1},x_{2n})=(x_{2n+1},x_{2n+2}), \quad F_{a,b}(x_{2n},x_{2n+1})=(x_{2n+2},x_{2n+3}), \] where $(x_1,x_2)\in{Q^+}$ and $n\ge 1$. So the odd terms of the sequence $\{x_n\}_n$ are contained in the projection on the $x$-axis of the oval of $\{V_{b,a}(x,y)=V_{b,a}(x_1,x_2)=h\}$ and the even terms in the corresponding projection of $\{V_{a,b}(x,y)=V_{a,b}(F_{a}(x_1,x_2))=h\}$, where notice that we have used Lemma~\ref{elemental}. Recall that the ovals of $V_{b,a}$ are invariant by $F_{b,a}$ and the ovals of $V_{a,b}$ are invariant by $F_{a,b}$. Notice also that the trivial equality $F_a\circ F_{b,a}=F_{a,b}\circ F_a$ implies that the action of $F_{b,a}$ on $\{V_{b,a}(x,y)=h\}$ is conjugated to the action of $F_{a,b}$ on $\{V_{a,b}(x,y)=h\}$ via $F_a.$ From Theorem \ref{rotacions} we know that $F_{b,a}$ on the corresponding oval is conjugated to a rotation of the circle. Hence, if the corresponding rotation number is rational, then the orbit starting at $(x_1,x_2)$ is periodic, of period say $q,$ then the sequence $\{x_n\}_n$ is $2q$-periodic. On the other hand if the rotation number is irrational, then the orbit of $(x_1,x_2)$ generated by $F_{b,a}$ fulfills densely the oval of $\{V_{b,a}(x,y)=h\}$ in ${Q^+}$ and hence the subsequence of odd terms also fulfills densely the projection of $\{V_{b,a}(x,y)=h\}$ in the $x$-axis. Clearly, the sequence of even terms do the same with the projection of the oval of $\{V_{a,b}(x,y)=h\}.$ Similarly when $k=3$ the equalities \begin{align*} F_{c,b,a}(x_{3n-2},x_{3n-1})&=(x_{3n+1},x_{3n+2}), \\ F_{a,c,b}(x_{3n-1},x_{3n})&=(x_{3n+2},x_{3n+3}),\\ F_{b,a,c}(x_{3n},x_{3n+1})&=(x_{3n+3},x_{3n+4}), \end{align*} where $n\ge1$, allow to conclude that each term $x_m,$ of the sequence $\{x_n\}_n$ where we use the notation (\ref{k=3}), is contained in one of the projections on the $x$-axis of the ovals $\{V_{c,b,a}(x,y)= V_{c,b,a}(x_1,x_2)=:h\}$ and $\{V_{a,c,b}(x,y)=h\}$ and $\{V_{b,a,c}(x,y)=h\}$, according with the remainder of $m$ after dividing it by 3. The rest of the proof in this case follows as in the case $k=2.$ So the first step is done. \noindent {\bf Second step:} From the above results it is clear that the problem of knowing the number of connected components of the adherence of $\{x_n\}_n$ is equivalent to the control of the projections of several invariant ovals on the $x$-axis. The strategy for $k=3$, and analogously for $k=2$, is the following. Consider the ovals contained in the level sets given by $\{V_{c,b,a}(x,y)=h\}$, $\{V_{a,c,b}(x,y)=h\}$ and $\{V_{b,a,c}(x,y)=h\}$ and denote by $I=I(a,b,c,h), J=J(a,b,c,h)$ and $K=K(a,b,c,h)$ the corresponding closed intervals of $(0,\infty)$ given by their projections on the $x$-axis. We want to detect the values of $h$ for which two of the intervals, among $I,J$ and $K,$ have exactly one common point. First we seek their boundaries. Since the level sets are given by cubic curves, that are quadratic with respect the $y$-variable, these points will correspond with values of $x$ for which the discriminant of the quadratic equation with respect to $y$ is zero. So, we compute \begin{align*} R_1(x,h,a,b,c)&:=\mbox{dis}\,(xyV_{c,b,a}(x,y)-hxy,y)=0,\\ R_2(x,h,a,b,c)&:=\mbox{dis}\,(xyV_{a,c,b}(x,y)-hxy,y)=0,\\ R_3(x,h,a,b,c)&:=\mbox{dis}\,(xyV_{b,a,c}(x,y)-hxy,y)=0. \end{align*} Now we have to search for relations among $a,b,c$ and $h$ for which two of these three functions have some common solution, $x.$ These relations can be obtained by computing some suitable resultants. Taking the resultants of $R_1$ and $R_2$; $R_2$ and $R_3$; and $R_1$ and $R_3$ with respect $x$ we obtain three polynomial equations $R_4(h,a,b,c)=0$, $R_5(h,a,b,c)=0$ and $R_6(h,a,b,c)=0.$ In short, once $a,b$ and $c$ are fixed we have obtained three polynomials in $h$ such that a subset of their zeroes give the bifurcation values which separate the number of intervals of the adherence of $\{x_n\}_n$. See the results of Proposition~\ref{ex2} and Example~\ref{ex3} for concrete applications of the method. Before ending the proof we want to comment that for most values of $a$, $b$ and $c,$ varying $h$ there appear the three possibilities, namely $1$, $2$ or $3$ different intervals. The last case appears for values of $h$ near $h_c:=V_{c,b,a}({\bf p})$, because the first coordinates of the three points $\bf p,$ $F_a({\bf p})$ and $F_b(F_a({\bf p}))$ almost never coincide. The other situations can be obtained by increasing $h.$ \qed \begin{propo}\label{ex2} Consider the recurrence \eqref{eq} with $k=2$ and $\{a_n\}_n$ as in \eqref{k=2} taking the values $a=3$ and $b=1/2.$ Define $h_c={(12z^3-33z+7)}/{(2(z^2-3))}\simeq 17.0394, $ where $z\simeq 2.1513$ is the biggest positive real root of \, $2z^4-12z^2-2z+17,$ and $h^*\simeq 17.1198,$ is the smallest positive root of \[p_4(h):=112900h^4-2548088h^3-48390204h^2+564028596h+7613699255.\] Then, \begin{enumerate}[(i)] \item The initial condition $(x_1,x_2)=(z,z^2-3)$ gives a two periodic recurrence $\{x_n\}_n$. Moreover $V_{1/2,3}(z,3-z^2)=h_c.$ \item Let $(x_1,x_2)$ be any positive initial conditions, different from $(z,z^2-3),$ and set $h=V_{1/2,3}(x_1,x_2).$ Let $\rho(h)$ denote the rotation number of $F_{1/2,3}$ restricted to the oval of $\{V_{1/2,3}(x,y)=h\}.$ Then \begin{enumerate}[(I)] \item If $\rho(h)=p /q\in\mathbb{Q},$ with $\gcd(p,q)=1,$ then the sequence $\{x_n\}_n$ is $2q$-periodic. \item If $\rho(h)\not\in\mathbb{Q}$ and $h\in (h_c,h^*)$ then the adherence of the sequence $\{x_n\}_n$ is formed by two disjoint closed intervals. \item If $\rho(h)\not\in\mathbb{Q}$ and $h\in [h^*,\infty)$ then the adherence of the sequence $\{x_n\}_n$ is one closed interval. \end{enumerate} \end{enumerate} \end{propo} We want to remark that, from a computational point of view, the case (I) almost never is detected. Indeed, taking $a$ and $b$ rational numbers and starting with rational initial conditions, by using Mazur's theorem it can be seen that almost never the rotation number will be rational, see the proof of \cite[Prop. 1]{BR}. Therefore, in numeric simulations only situations (II) and (III) appear, and the value $h=h^*$ gives the boundary between them. In general, for $k=2,$ the value $h^*$ is always the root of a polynomial of degree four, which is constructed from the values of $a$ and $b$. \noindent{\it Proof of Proposition~\ref{ex2}.} Clearly $(z,3-z^2)$ is the fixed point of $F_{b,a}$ in ${Q^+}.$ Some computations give the compact expression of $h_c:=V_{a,b}(z,z^2-3).$ To obtain the values $h^*$ we proceed as in the proof of Theorem~\ref{main}. In general, \begin{align*} R_1(x,h,a,b)&:=\mbox{dis}\,(xyV_{b,a}(x,y)-hxy,y)\\ &\phantom{:}=(ax^2-hx+a^2+b)^2-4(bx+a)(bx^2+b^2x+ax+ab),\\ R_2(x,h,a,b)&:=\mbox{dis}\,(xyV_{a,b}(x,y)-hxy,y)\\ &\phantom{:}=(bx^2-hx+a+b^2)^2-4(ax+b)(ax^2+a^2x+bx+ab). \end{align*} Then we have to compute the resultant of the above polynomials with respect to $x$. It always decomposes as the product of two quartic polynomials in $h.$ Its expression is very large, so we only give it when $a=3$ and $b=1/2$. It writes as \[ \frac{625}{65536}\left(4h^4-1176h^3+308h^2+287380h+1816975\right)p_4(h). \] It has four real roots, two for each polynomial. Some further work proves that the one that interests us is the smallest one of $p_4.$ \qed We also give an example when $k=3,$ but skipping all the details. \begin{example}\label{ex3} Consider the recurrence \eqref{eq} with $k=3$ and $\{a_n\}_n$ as in \eqref{k=3} taking the values $a=1/2, b=2$ and $c=3.$ Then for any positive initial conditions $x_1$ and $x_2$, $V_{c,b,a}(x_1,x_2)=h\ge V_{c,b,a}({\bf p})=h_c\simeq 15.9283.$ Moreover if the rotation number of $F_{c,b,a}$ associated to the oval $\{V_{c,b,a}(x,y)=h\}$ is irrational then the adherence of $\{x_n\}_n$ is given by: \begin{itemize} \item Three intervals when $h\in (h_c,h^*),$ where $h^*\simeq 15.9614;$ \item Two intervals when $h\in [h^*,h^{**}),$ where $h^{**}\simeq 16.0015;$ \item One interval when $h\in[h^{**},\infty).$ \end{itemize} The values $h^*$ and $h^{**}$ are roots of two polynomials of degree 8 with integer coefficients that can be explicitly given. \end{example} \section{Some properties of the rotation number function }\label{someproperties} From Theorem~\ref{rotacions} it is natural to introduce the {\it rotation number function} for $F_{b,a}$: \[ \rho_{b,a}:[h_c,\infty)\longrightarrow (0,1), \] where $h_c:=V_{b,a}({\bf p})$, as the map that associates to each invariant oval $\{V_{b,a}(x,y)=h\}$, the rotation number $\rho_{b,a}(h)$ of the function $F_{b,a}$ restricted to it. The following properties hold: \begin{enumerate}[(i)] \item The function $\rho_{b,a}(h)$ is analytic for $a>0,b>0$, $h>h_c$ and it is continuous at $h=h_c.$ This can be proved from the tools introduced in \cite[Sec. 4]{CGM2}. \item The value $\rho_{b ,a}(h_c)$ is given by the argument over $2\pi$ of the eigenvalues (which have modulus one due to the integrability of $F_{b,a}$) of the differential of $F_{b,a}$ at ${\bf p}$. \item $\rho_{b,a}(h)=\rho_{a,b}(h).$ \item $\rho_{a,a}(h)=2\rho_a(h) \mod 1,$ where $\rho_a$ is the rotation number\footnote{Notice that given a map of the circle there is an ambiguity between $\rho$ and $1-\rho$ when one considers its rotation number. So, while for us the rotation number of the classical Lyness map for $a=1$ is $4/5$, in other papers it is computed as $1/5.$} function associated to the classical Lyness map. Then, from the results of \cite{BC} we know that $\rho_{1,1}(h)\equiv3/5,$ that for $a\ne1,$ positive, $\rho_{a,a}(h)$ is monotonous and $\lim_{h\to\infty}\rho_{a,a}(h)=3/5.$ \end{enumerate} Note that item (iii) follows because $F_{a,b}$ is conjugated with $F_{b,a}$ via $\psi=F_a$ which is a diffeomorphism of ${Q^+}$, because $ \psi^{-1} F_{a,b} \psi=F_a^{-1}F_a F_b F_a=F_b F_a=F_{b,a}. $ Since $\psi$ preserves the orientation, the rotation number functions of $F_{a,b}$ and $F_{b,a}$ restricted to the corresponding ovals must coincide. Similar results to the ones given above hold for $F_{c,b,a}$ and its corresponding rotation number function. Some obvious differences are: \begin{align*} &\rho_{c,b,a}(h)=\rho_{b,a,c}(h)=\rho_{a,c,b}(h)\, &&\rho_{a,a,a}(h)=3\rho(h)\,\mod1,\\ &\rho_{1,1,1}(h)=2/5, &&\lim_{h\to\infty}\rho_{a,a,a}(h)=2/5. \end{align*} We are convinced that when $a>0$ and $b>0,$ \[\lim_{h\to\infty}\rho_{b,a}(h)=3/5\quad\mbox{and}\quad\lim_{h\to\infty}\rho_{c,b,a}(h)=2/5, \] but we have not been able to prove these equalities. If they were true, by combining them with the values of the rotation number function at $h=h_c$ this would give a very useful information to decide if, apart from the trivial cases $a=b=1 (c=1),$ there are other cases for which the rotation number function is constant. Notice that in these situations the maps $F_{b,a}$ or $F_{c,b,a}$ would be globally periodic in ${Q^+}.$ This information, together with the values at $h_c$, also would be useful to know the regions where the corresponding functions could be increasing or decreasing. Finally notice that this rotation number at infinity is not continuous when we approach to $a=0$ or $b=0$, where the recurrence and the first integral are also well defined on ${Q^+}.$ For instance $\rho_{0,0}(\rho)\equiv 2/3$ and the numerical experiments of next subsection seem to indicate that for $a>0$ or $b>0,$ \[ \lim_{h\to\infty}\rho_{0,a}(h)=\lim_{h\to\infty}\rho_{b,0}(h)=5/8. \] Before proving Theorem~\ref{teonomonotonia} we introduce with an example the algorithm that we will use along this section to compute lower and upper bounds for the rotation number. We have implemented it in an algebraic manipulator. Notice also that when we apply it taking rational values of $a$ and $b$ and rational initial conditions, it can be used as a method to achieve proofs, see next example or the proof of Theorem~\ref{teonomonotonia}. Fix $a=3,$ $b=2$ and $(x_0,y_0)=(1,1).$ Then $h=V_{2,3}(1,1)=34.$ Compute for instance the 27 points of the orbit starting at $(1,1),$ \[ (x_1,y_1)=(4,6),\quad (x_2,y_2)=\left(\frac9 4, \frac{17}{24}\right),\quad (x_3,y_3)=\left(\frac{89}{54}, \frac{788}{153}\right),\ldots \] and consider them as points on the oval $\{V_{2,3}(x,y)=34\},$ see Figure~\ref{fig1}. \begin{figure}[h] \begin{center} \includegraphics[scale=0.35]{fig1 \end{center} \caption{Oval of $\{V_{2,3}(x,y)=34\}$ with 27 iterates of $F_{2,3}$. The label $0$ indicates the initial condition $(1,1)$, and the label $k, k=1,\ldots,26,$ corresponds with the $k$-th point of the orbit.}\label{fig1} \end{figure} \begin{figure}[h] \begin{center} \includegraphics[height=5 cm,width=12 cm]{fig2 \end{center} \caption{Lower and upper bounds for $\rho_{2,3}(34)$ obtained after following some points of the orbit starting at $(1,1)$.}\label{fig2} \end{figure} We already know that the restriction of $F_{2,3}$ to the given oval is conjugated to a rotation, with rotation number $\rho:=\rho_{2,3}(34)$ that we want to estimate. This can be done by counting the number of turns that give the points $(x_j,y_j)$, after fixing some orientation in the closed curve. We orientate the curve with the counterclockwise sense. So, for instance we know that the second point $(x_2,y_2)$ has given more that one turn and less than two, giving that $1<2\rho<2,$ and hence that $\rho\in(1/2,1).$ Doing the same reasoning with all the points computed we obtain, \begin{align*} &4<7\rho<5&\Rightarrow&&& \rho\in\left(\frac47,\frac57\right),\\ &8<14\rho<9&\Rightarrow&&& \rho\in\left(\frac8{14},\frac9{14}\right),\\ &10<19\rho<11&\Rightarrow&&& \rho\in\left(\frac{10}{19},\frac{11}{19}\right),\\ &14<26\rho<15&\Rightarrow&&& \rho\in\left(\frac{14}{26},\frac{15}{26}\right), \end{align*} where we have only written the more relevant informacions obtained, which are given by the points of the orbit closer to the initial condition. So, we have shown that \[ 0.5714\simeq\frac{4}{7}<\rho_{2,3}(34)<\frac{15}{26}\simeq 0.5769. \] In Figure~\ref{fig2} we represent several successive lower and upper approximations obtained while the orbit is turning around the oval. We plot around six hundred steps, after skipping the first fifty ones. By taking 1000 points we get \[ 0.5761246\simeq\frac{338}{578}<\rho_{2,3}(34)<\frac{473}{821}\simeq 0.5761267, \] and after 3000 points, \[ 0.57612457\simeq\frac{338}{578}<\rho_{2,3}(34)<\frac{1472}{2555}\simeq 0.57612524. \] In fact when we say that $\rho_{2,3}(34)\in(\rho_{\mathrm{low}},\rho_{\mathrm{upp}})$, the value $\rho_{\mathrm{low}}$ is the upper lower bound obtained by following all the considered points of the orbit, and $\rho_{\mathrm{upp}}$ is the lowest upper bound. Notice that taking 1000 or 3000 points we have obtained the same lower bound for $\rho_{2,3}(34).$ Let us prove Theorem~\ref{teonomonotonia} by using the above approach. \bigskip \noindent{\it Proof of Theorem \ref{teonomonotonia}.} Consider $a=1/2,$ $b=3/2$ and the three points \[ \mathbf{p}^1=\left(\displaystyle{\frac{149}{100}},\displaystyle{\frac{173}{100}} \right),\quad\mathbf{p}^2=\left(\displaystyle{\frac{3}{40}},\displaystyle{\frac{173}{100}} \right),\quad\mathbf{p}^3=\left(\displaystyle{\frac{1}{1000}},\displaystyle{\frac{173}{100}} \right). \] Notice that \begin{align*} &h_1:=V_{3/2,1/2}(\mathbf{p}^1)=\frac{10655559}{1288850}\simeq8.27,\\ &h_2:=V_{3/2,1/2}(\mathbf{p}^2)=\frac{9328327}{207600}\simeq44.93,\\ &h_3:=V_{3/2,1/2}(\mathbf{p}^3)=\frac{1056238343}{346000}\simeq3052.71. \end{align*} Hence $h_c<h_1<h_2<h_3.$ By applying the algorithm described above, using 100 points of each orbit starting at each ${\bf p}^j, j=1,2,3,$ we obtain that \[ \rho_{3/2,1/2}(h_1),\rho_{3/2,1/2}(h_3)\in\left(\frac 35,\frac{59}{98}\right)\quad\mbox{and}\quad \rho_{3/2,1/2}(h_2)\in\left(\frac{56}{93},\frac{53}{88}\right). \] Since $59/98<56/93$ we have proved that the function $\rho_{3/2,1/2}(h)$ has at least a local maximum in $(h_1,h_3).$ From the continuity of the rotation number function, with respect $a,b$ and $h$, we notice that this result also holds for all values of $a$ and $b$ in a neighborhood of $ a=1/2,b=3/2.$ \qed \bigskip We believe that with the same method it can be proved that a similar result to the one given in Theorem~\ref{teonomonotonia} holds for some maps $F_{c,b,a},$ but we have decided do not perform this study. \subsection{Some numerical explorations for $k=2.$}\label{seccionumeric} We start by studying with more detail the rotation number function $\rho_{3/2,1/2}(h)$, that we have considered to prove Theorem~\ref{teonomonotonia}. In this case the fixed point is ${\bf p}\simeq(1.493363282,1.730133891)$ and $h_c=V_{b,a}({\mathbf{p}})=8.267483381.$ Moreover $\rho_{b,a}({h}_c)\simeq0.6006847931$. By applying our algorithm for approximating the rotation number, with $5000$ points on each orbit, we obtain the results presented in Table~1. In Figure~\ref{fig3} we also plot the upper and lower bounds of $\rho_{3/2,1/2}(h)$ that we have obtained by using a wide range of values of $h.$ \begin{center} \vglue 0.2cm \begin{tabular}{|r|r|r|r|} \hline Init. cond. $(x,\bar y)$ & Energy level $h$ & $\rho_{\mathrm{low}}(h)\qquad$ & $\rho_{\mathrm{upp}}(h) \qquad$\\ \hline $\bar x$&$h_c\simeq 8.2675$& $\simeq 0.6006848$ & $\simeq 0.6006848$ \\ 1.3&$8.3068$& $\frac{173}{288}\simeq 0.6006944$ & $\frac{2938}{4891}\simeq 0.6006951$ \\ 0.75&$9.2747$& $\frac{1435}{2388}\simeq 0.6009213$ & $\frac{2087}{3473}\simeq 0.6009214$ \\ 0.3&$14.7566$& $\frac{1548}{2573}\simeq 0.6016323$ & $\frac{2285}{3798}\simeq 0.6016324$ \\ 0.075&$44.9347$& $\frac{657}{1091}\simeq 0.6021998$ & $\frac{2354}{3909}\simeq 0.6022001$ \\ 0.001&$3052.75$& $\frac{2927}{4867}\simeq 0.6013972$ & $\frac{86}{143}\simeq 0.6013986$ \\ $5\cdot 10^{-6}$&$609716.07$& $\frac{1832}{3049}\simeq 0.6008527$ & $\frac{1409}{2345}\simeq 0.6008529$ \\ $5\cdot 10^{-256}$&$6.097\cdot 10^{255}$& $\frac{3}{5}= 0.6$ & $\frac{2999}{4998}\simeq 0.6000400$ \\ \hline \end{tabular} \vglue 1 cm Table 1: Lower and upper bounds of the rotation number $\rho_{3/2,1/2}(h)$, for some orbits of $F_{3/2,1/2}$ starting at $(x,\bar y),$ where ${\bf p}=(\bar x,\bar y).$ \end{center} \begin{figure}[h] \begin{center} \includegraphics[height=7 cm,width=14 cm]{fig3 \end{center} \caption{Lower and upper bounds for $\rho_{3/2,1/2}(h)$. On the horizontal axis we represent $-\log_{10}(h)$ and, on the vertical axis, the value of the rotation number. Notice that for values of $-\log_{10}(h)$ smaller that 70 both values are indistinguishable in the Figure.}\label{fig3} \end{figure} For other values of $a$ and $b$ we obtain different behaviors. All the experiments are performed by starting at the fixed point ${\bf p}=(\bar x,\bar y),$ and increasing the energy level by taking initial conditions of the form $(x,\overline{y})$, by decreasing $x$ to $0$. With this process we take orbits approaching to the boundary of ${Q^+}$, that is lying on level sets of $V_{b,a}$ with increasing energy. The step in the decrease of $x$ (and therefore in the increase of $h$) is not uniform, and it has been manually tuned making it smaller in those regions where a possible non monotonous behavior could appear. Consider the set of parameters $\Gamma=\{(a,b),\in [0,\infty)^2\}$, where notice that we also consider the boundaries $a=0$ or $b=0,$ where the map $F_{b,a}$ is well defined. We already know that the rotation number function behaves equal at $(a,b)$ and $(b,a).$ Moreover we know perfectly its behavior on the diagonal $(a,a)$ (when $a<1$ it is monotonous decreasing and when $a>1$ it is monotonous increasing) and that $\rho_{1,1}(h)\equiv 4/5$ and $\rho_{0,0}(h)\equiv 2/3.$ Hence a good strategy for an numerical exploration can be to produce sequences of experiments using our algorithm by fixing some $a\ge0$ and varying $b$. For instance we obtain: \begin{itemize} \item Case $a=1/2.$ For all the values of $b>0$ considered, the rotation number function seems to tend to $3/5$ when $h$ goes to infinity. Moreover it seems \begin{itemize} \item monotone decreasing for $b\in\{1/4,1\};$ \item to have a unique maximum when $b\in\{7/5,3/2\}$; \item monotone increasing for $b\in\{2,3\}.$ \end{itemize} \item Case $a=0.$ For all the values of $b>0$ considered, the rotation number function seems to tend to $5/8$ when $h$ goes to infinity. Moreover it seems \begin{itemize} \item monotone decreasing for $b\in\{1/10,3/10,1/2\}$; \item to have a unique maximum when $b\in\{7/10,3/4\}$; \item monotone decreasing for $b\in\{1,5\}$. \end{itemize} \end{itemize} The above results, together with some other experiments for other values of $a$ and $b$, not detailed in this paper, indicate the existence of a subset of positive measure in $\Gamma$ where the corresponding rotation number functions seem to present an unique maximum. This subset probably separates two other subsets of~$\Gamma$, one where $\rho_{b,a}(h)$ is monotonically decreasing to $3/5$ , and another one where $\rho_{b,a}(h)$ increases monotonically to the same value. The ``oscillatory subset'' seems to shrink to $(a,b)=(1,1)$ when it approaches to the line $a=b$ and seems to finish in one interval on each of the borders $\{a=0\}$ and $\{b=0\}$. Further analysis must be done in this direction in order to have a more accurate knowledge of the bifurcation diagram associated to the behavior of $\rho_{b,a}$ on $\Gamma$.
{ "timestamp": "2009-12-26T18:31:35", "yymm": "0912", "arxiv_id": "0912.5031", "language": "en", "url": "https://arxiv.org/abs/0912.5031", "abstract": "We describe the sequences {x_n}_n given by the non-autonomous second order Lyness difference equations x_{n+2}=(a_n+x_{n+1})/x_n, where {a_n}_n is either a 2-periodic or a 3-periodic sequence of positive values and the initial conditions x_1,x_2 are as well positive. We also show an interesting phenomenon of the discrete dynamical systems associated to some of these difference equations: the existence of one oscillation of their associated rotation number functions. This behavior does not appear for the autonomous Lyness difference equations.", "subjects": "Dynamical Systems (math.DS)", "title": "On two and three periodic Lyness difference equations", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9861513930404762, "lm_q2_score": 0.8311430499496095, "lm_q1q2_score": 0.8196328765237175 }
https://arxiv.org/abs/1107.0326
The Monty Hall Problem in the Game Theory Class
The basic Monty Hall problem is explored to introduce into the fundamental concepts of the game theory and to give a complete Bayesian and a (noncooperative) game-theoretic analysis of the situation. Simple combinatorial arguments are used to exclude the holding action and to find minimax solutions.
\section{Introduction} \begin{itemize} \item[] {\it Suppose you're on a game show, and you're given the choice of three doors: Behind one door is a car; behind the others, goats. You pick a door, say No. 1, and the host, who knows what's behind the doors, opens another door, say No. 3, which has a goat. He then says to you, ``Do you want to pick door No. 2?" Is it to your advantage to switch your choice?} \end{itemize} With these famous words the {\it Parade Magazine} columnist vos Savant opened an exciting chapter in mathematical didactics. The puzzle, which has become known as the {\it Monty Hall Problem} (MHP), has broken the records of popularity among the probability paradoxes. The book by Rosenhouse \cite{Rosenhouse} and the Wikipedia entry on the MHP present the history and variations of the problem. In the basic version of the MHP the rules of the game specify that the host must always reveal one of the unchosen doors to show that there is no prize there. Two remaining unrevealed doors may hide the prize, creating the illusion of symmetry and suggesting that the action does not matter. However, the symmetry is fallacious, and switching is a better action, doubling the probability of winning. There are two main explanations of the paradox. One of them, simplistic, amounts to just counting the mutually exclusive cases: either you win with switching or with holding the first choice. A more sophisticated argument, included in the textbooks as an exercise on the Bayes theorem, calculates the conditional probability of winning in the situation described in vos Savant's wording of the problem. In \cite{Gill} a critical analysis of the conventional approaches to the MHP has been done, with the advocated viewpoint that the whole situation of making decision, viewed as a multistep process, is a challenging instance of mathematical modelling, very much amenable to the analysis within the game-theoretic framework. The textbook by Haggstr\"om \cite{Haggstrom} puts the zero-sum game in matrix form and presents a minimax solution. Further steps in this direction were done in \cite{Doors} and \cite{Dominance}, where it was argued that the game-theoretic concept of dominance allows to analyse the problem under fairly general assumptions on the prior information of the decision-maker, including the very interesting case of nouniform distribution, only occasionally included in exercise sections of probability textbooks \cite{Grinstead, Tijms}. One elementary new observation we make is that for every contestant's strategy of playing the game after the initial door has been chosen, there is always (at least) one `unlucky' door, the same for every admissible algorithm for revealing the doors by the host. The prize is never found behind the `unlucky' door, hence the contestant will always lose in at least one case out of three. This leads, rather straightforwardly, to the worst-case winning probability 2/3. We prefer, however, for the sake of instruction and in anticipation of future generalizations to give here a full Bayesian analysis, with the host biased in any possible way. These notes, expanding upon the cited literature are intended to show that the MHP is indeed an excellent occasion to expose the undergraduate students to basic ideas of the game theory and to decision-making under various uncertainty scenarios. Due to its remarkable symmetry features the zero-sum version of the game will undoubtedly enter the Hall of Fame of the classical games like {\it matching pennies} and {\it paper-scissors-stone}. \section{The MHP as a sequential decision process} The game involves two actors which we christen Monte and Conie. In the basic scenario to follow, the prize is hidden behind one of three doors by Monte, then Conie picks a door which is kept unrevealed, then one of the unchosen doors is revealed as not containing the prize, and then an offer is made to switch the choice from the initial guess to another unrevealed door. Conie wants the prize and she wins it if her final choice falls on the door where the prize was hidden. To state the rules more rigorously and to introduce some notation for the admissible actions we number the doors 1,2,3. The game consists of four moves: \begin{itemize} \item[(i)] Monte chooses a door number $\theta$ from doors 1,2,3 to hide the prize. He keeps $\theta$ in secret. \item[(ii)] Conie picks door number $x$ out of 1,2,3. Monte observes Conie's choice. \item[(iii)] Monte reveals a door distinct from $x$ as not hiding the prize, and offers to Conie the possibility of choosing between door $x$ and another unrevealed door $y$. \item[(iv)] Conie finally chooses door $z$ from doors $x$ and $y$. Conie wins the prize if $z=\theta$, otherwise she wins nothing. \end{itemize} Conie's ignorance about the location of the prize means in the mathematical language that her actions cannot depend on $\theta$ explicitly. There is also another, more subtle, indefinite factor of which Conie is ignorant: this is the way Monte chooses between two doors to reveal in the case $x=\theta$. These two indefinite factors, which are under the control of her opponent, comprise Conie's decision-making environment. We say that there is a match if $\theta=x$, in which case $y$ on step (iii) can be one of two numbers distinct from $\theta$; whereas if there is a mismatch, $\theta\neq x$, the rules force $y=\theta$. On step (iv), we say that Conie takes action `hold' (of sticking with the initial choice), denoted $\tm$, if $z=x$; and that she takes action `switch' (from the initial choice), denoted $\ts$, if $z=y$. \par To illustrate, a sequence of admissible moves $\theta,x,y,z$ could be 2212, which means that Monte hides the prize behind door 2, Conie picks door 2 (so there is a match), then Monte offers to switch to door 1 (by revealing door 3 as not containing the prize), and finally Conie plays $\tm$ by sticking with her initial choice 2. Since $z=\theta~$ Conie wins the prize. Positions in the game represent all substantial information available for the move to follow. These are represented by vertices in the game tree in Figure \ref{TheBigGameTree}. An edge connects a position to another position achievable in one move. The play starts at the root vertex with Monte's move leading to a position $\theta\in\{1,2,3\}$, then the moves of actors alternate. A path in the tree from the root to a terminal vertex is determined by the actions of {\it both} actors. Each path directed away from the root ends in a leaf node, with Conie's winning positions $(\theta,x,y,z)$ being those with $z=\theta$. There is one important feature of the game which we indicate by coloring positions in the figure. Conie does not know the winning door $\theta$. The information of Conie on her second move can be specified by partitioning the collection of relevant positions in {\it information sets}. For the first Conie's move $x$ there is just one information set $\{1,2,3\}$. On her second move, Conie cannot distinguish e.g. between positions 121 and 221, since for $x=2$ and $y=1$ (when the revealed door is 3) the prize can be behind any of the doors 1 and 2; thus $\{121, 221\}$ is one information set which we denote $*21$, with $*$ staying for the unknown admissible value of $\theta$ (1 or 2 in this case). The complete list of information sets for the second Conie's move is $*12, *13, *21, *23, *31, *32$. Possible moves are labelled by actions $\{\tm,\ts\}$, and the action depends on the set. Thus if the action is $\tm$ from position $112$, then it must be $\tm$ from position $212$. The game tree with partition of positions into information sets is sometimes called the {\it Kuhn tree}. Monte always knows the current position in the game exactly, so his information sets are singletons. \begin{figure} $ { \resizebox*{10cm}{18cm} { \newcommand{\XX}[1] \Tr{\psframebox{\rule{0pt}{9pt}#1}}} \psset{angleB=90,angleA=90} \pstree [treemode=R] {\XX{start}} { \pstree{{\XX{1~~~~}}} { \pstree{{\XX{11~~~}}}{{\pstree{{\XX{\color{red}112~~}}}{{\XX{1121}~{\large\Aries}}\taput{\m}{\XX{1122}}}}\tbput{\s} \pstree{{\XX{\color{orange}113~~}}} {{\XX{1131}~{\large\Aries}}\taput{\m}{\XX{1133}}}}\tbput{\s} \pstree{{\XX{12~~~}}}{\pstree{{\XX{\color{yellow}121~~}}}{{\XX{1211}~{\large\Aries}}\taput{\s} {\XX{1212}}}}\tbput{\m} \pstree{{\XX{13~~~}}}{\pstree{{\XX{\color{green}131~~}}}{{\XX{1311}~{\large\Aries}}\taput{\s}{\XX{1313}}}}\tbput{\m} } \pstree{{\XX{2~~~~}}} { \pstree{{\XX{21~~~}}}{\pstree{{\XX{\color{red}212~~}}}{{\XX{2121}}\taput{\m}{\XX{2122}~{\large\Aries}}}}\tbput{\s} \pstree{{\XX{22~~~}}} {\pstree{{\XX{\color{yellow}221~~}}}{{\XX{2211}}\taput{\s} {\XX{2212}~{\large\Aries}}}\tbput{\m} \pstree{{\XX{\color{blue}223~~}}} {{\XX{2232}~{\large\Aries}}\taput{\m} {\XX{2233}}}}\tbput{\s} \pstree{{\XX{23~~~}}}{\pstree{{\XX{\color{violet}232~~}}}{{\XX{2322}~{\large\Aries}}\taput{\s} {\XX{2323}}}}\tbput{\m} } \pstree{{\XX{3~~~~ }}} { \pstree{{\XX{31~~~}}}{\pstree{{\XX{\color{orange}313~~}}}{{\XX{3131}}\taput{\m} {\XX{3133}~{\large\Aries}}}} \tbput{\s} \pstree{{\XX{32~~~}}}{\pstree{{\XX{\color{blue}323~~}}}{{\XX{3232}}\taput{\m} {\XX{3233}~{\large\Aries}}}} \tbput{\s} \pstree{{\XX{33~~~}}}{\pstree{{\XX{\color{green}331~~}}}{{\XX{3311}}\taput{\s} {\XX{3313}~{\large\Aries}}}\tbput{\m} \pstree{{\XX{\color{violet}332~~}}}{{\XX{3322}}\taput{\s}{\XX{3323}~{\large\Aries}}}}\tbput{\m} } } } } $ \caption{The game tree with succession of moves $\theta,x,y,z$, and Conie's winning terminal positions marked \Aries.} \label{TheBigGameTree} \end{figure} \section{Strategies and the payoff matrix} A {\it strategy} of Monte is a rule which for each position, where Monte is on the move, specifies a follower exactly. For his first move from the starting position a strategy specifies a value of $\theta$. For his second move a strategy specifies the value of $y$ for each position $(\theta,x)$, so we can consider $y$ as a function $y= d_\theta(x)$, where \begin{eqnarray*} d_\theta(x)=\theta ~~~{\rm if ~~~} x\neq\theta,\\ d_\theta(x)\neq \theta ~~~{\rm if~~~} \theta=x. \end{eqnarray*} Simply put, Monte's strategy can be encoded in a pair of door numbers like $12$, where $\theta=1$ is the door hiding the prize, and $d_1(1)=2$ is the door to which the switch is offered in the case of match. This indeed determines the function $d_1(\cdot)$ uniquely because $d_1(2)=d_1(3)=1$. With this notation, the complete list of Monte's strategies is $${\mathcal M}= \{12,~ 13,~ 21,~ 22,~ 31,~ 32\}.$$ A {\it strategy} of Conie is a rule which for each position where Conie is on the move specifies a follower, in a way consistent with partition in information sets. Thus, her strategy must specify the value of $x$. Furthermore, for each initial choice $x$ and the door offered for switch $y\neq x$ her strategy must specify an action from the set $\{\tm,\ts\}$, which is a function $a_x(\cdot)$, so that the second Conie's move is \begin{eqnarray*} z=x ~~~{\rm if ~~~} a_x(y)=\tm,\\ z=y ~~~{\rm if ~~~} a_x(y)=\ts. \end{eqnarray*} When $x$ is fixed $a_x(y)$ must be specified for two possible values of $y$. We can write therefore Conie's strategy as a triple like $2\ts\tm$ which specifies $x=2$ and $a_2(1)=\ts, a_2(3)=\tm$. The second entry $\ts$ of $2\ts\tm$ encodes the action for smaller door number $y$, and the third entry $\tm$ for larger. With similar conventions, the complete list of twelve strategies of Conie is \\ $${\mathcal C}= \{1\ts\ts,~ 1\ts\tm, ~1\tm\ts, ~1\tm\tm, ~2\ts\ts, ~2\ts\tm, ~2\tm\ts, ~2\tm\tm, ~3\ts\ts, ~3\ts\tm, ~3\tm\ts, ~3\tm\tm\}.$$ \noindent Every strategy of the kind $x\tm\tm$ or $x\ts\ts$ will be called {\it constant-action} strategy, and every strategy of the kind $x\ts\ts$ will be called {\it always-switching} stratgey. Strategies $x\ts\tm$ and $x\tm\ts$ are {\it context-dependent} strategies. A strategy profile of the actors is a strategy of Monte and a strategy of Conie. When a strategy profile is fixed, the course of the game, represented by a path in the game tree, is fully inambiguous. For instance, when the profile (12,2\ts\tm) is played by the actors, $\theta,x,y,z$ is 1211, because Monte offers a switch to door 1, and since 1 is the smaller door number of possible values of $y\in\{1,3\}$ Conie reacts with $\ts$, hence winning the prize. The second entry 2 of Monte's strategy 12 is immaterial for this outcome since there is a mismatch $\theta\neq x$. We adopt the convention that Conie receives payoff 1 when she wins the prize and 0 otherwise. The matrix $\CC$ in Figure \ref{ConPayoff} represents the correspondence between the strategy profiles and Conie's payoffs. \begin{figure} \begin{center} \begin{tabular}{c|cccccc} $\theta,y=$ & 12 & 13 & 21 & 23 & 31 & 32\\ \hline 1\swsw &0 & 0 & 1 & 1 & 1 &1\\ 1\masw &1 & 0 & 0 & 0 & 1 &1\\ 1\swma &0 & 1 & 1 & 1 & 0 &0\\ 1\mama &1 & 1 & 0 & 0 & 0 &0\\ & & & & & & \\ 2\swsw &1 & 1 & 0 & 0 & 1 &1\\ 2\masw &0 & 0 & 1 & 0 & 1 &1\\ 2\swma &1 & 1 & 0 & 1 & 0 &0\\ 2\mama &0 & 0 & 1 & 1 & 0 &0\\ & & & & & & \\ 3\swsw &1 & 1 & 1 & 1 & 0 &0\\ 3\masw &0 & 0 & 1 & 1 & 1 &0\\ 3\swma &1 & 1 & 0 & 0 & 0 &1\\ 3\mama &0 & 0 & 0 & 0 & 1 &1\\ \end{tabular} \end{center} \caption{Conie's payoff matrix $\CC$} \label{ConPayoff} \end{figure} Unlike the game tree, the simplified matrix representation ignores many substantial attributes like moves, positions and information sets. The game in this matrix form amounts to a very simple procedure: Monte picks a column and Conie picks a row of matrix $\CC$. Conie's payoff is then the entry of $\CC$ that stays on the intersection of the selected row and column. An advantage of the matrix representation is that it unifies and simplifies the analysis. In particular, we can compare different Conie's strategies under different assumptions on the decision-making environments, that is Monte's behaviors. For instance, we can think of two Conies which simultaneously play, say, 1\masw~ and 3\swsw~ (respectively) against the same strategy of Monte. Note that `simultaneous' in this context refers to a logical comparison of the outcomes, rather than to a real play. A quick inspection of the matrix $\CC$ shows that the problem has a property called {\it weak dominance}: \begin{itemize} \item[] { for every Conie's strategy $A$ which in some situation employs action $\tm$ (so $A$ is of the kind $x\tm\tm$, $x\tm\ts$ or $x\ts\tm$) there exists an always-switching strategy $B$ such that if $A$ wins against $S$, then $B$ wins against $S$ as well, whichever strategy of Monte $S$}. \end{itemize} For instance, strategy 1\tm\ts~ which does not switch to door $y=2$, is dominated by always-switching strategy 2\ts\ts. Amazingly, $1\tm\ts$ is beaten in the situation $\theta=1, y=3$ where the strategy uses the {\it switch} action! The never-switching strategy 1\tm\tm~ is dominated by both 2\ts\ts~and~ 3\ts\ts. The dominance is a very strong argument for excluding the strategies which may employ the action $\tm$. The dominance feature has a nice interpretation in terms of `unlucky' door. Suppose for a time-being that the prize-hiding is a move of nature (or quiz-team), out of control of Monte. All Monte can do is to choose door $y$ to reveal when he can. \begin{itemize} \item[] {The `unlucky' door theorem says: whichever Conie's strategy $A$, there exists a door $u$ (depending on $A$) such that under $A$ the final choice $z$ does not fall on $u$ when $\theta=u$, whichever Monte's way of revealing doors when he can. Strategy $u\ts\ts$~then weakly dominates $A$.} \end{itemize} Door $u=u(A)$ is marked in $\CC$ by two zeroes in positions $u*$ of the row $A$, for instance $u=2$ for $A=1\tm\ts$, and $u=2$ or $3$ for $A=1\tm\tm$. So the existence of the `unlucky' door means that Conie never wins for $\theta=u$, no matter how Monty reveals the doors. Therefore, \begin{itemize} \item[] {If $\theta$ is chosen uniformly at random Conie's winning probability cannot exceed $2/3$.} \end{itemize} \section{The Bayesian games}\label{Bayes} We described above the sets of {\it pure} strategies $\mathcal M$ and $\mathcal C$, selecting from which a particular profile determines inambiguously a succession of moves. A {\it mixed} strategy of an actor is an assignment of probability to each pure strategy of the actor. Playing mixed strategies means using chance devices to choose from $\mathcal M$ and $\mathcal C$. A mixed strategy of Monte can be written as a row vector $\QQ$ with six components corresponding to her pure strategies. A mixed strategy of Conie can be written as a row vector $\PP$ of length twelve. Conie's generic mixed strategy has 11 parameters, as probabilities add to one. A general theorem due to Kuhn says that if the actors do not forget their private history (perfect recall), like in our game, it is enough to consider a smaller class of {\it behavioral} strategies. A behavioral strategy specifies probability distribution on the set of admissible moves for every information set of an actor. Thus behavioral strategy of Conie is described by 8 parameters, 2 numbers for distribution of $x$ and 6 biases for coins tossed for each information set. This subtle distinction of the general and behavioral strategies is only mentioned here for the sake of instruction, but the distinction becomes important in some games where an actor may forget some elements of her private history. The {\it support} of a mixed strategies are the pure strategies with nonzero probability. For instance, the support of $(\dots,0,1,0,\dots)$ is one pure strategy, thus we simply identify pure strategies with mixed strategies of this kind. This identification allows us to view a mixed strategy as a convex combination (also called mixture) of the pure strategies constituting the support. A mixed strategy $\PP$ (respectively $\QQ$) is called {\it fully supported} if the support is $\cal C$ (respectively $\cal M$). When strategy profile $(\PP,\QQ)$ is played by the actors, the expected payoff for Conie, equal to her probability of winning the prize, is computed by the matrix multiplication as $\PP\CC\QQ^T$, where $^T$ denotes transposition. This way of computation presumes that the actors' choices of pure strategies are independent random variables with values in $\mathcal M$ and $\mathcal C$ (respectively), which may be simulated by the actors' private randomization devices. The independence of individual strategies is a feature required by the idea of {\it noncooperative} game, in which there is no communication of actors to play a joint strategy. In the {\it Bayesian setting} of the decision problem Monte is supposed to play some fixed strategy $\QQ$, known to Conie. The probability textbooks consider the Bayesian setting for the MHP, with $\QQ$ being the uniform distribution over $\mathcal M$, which is equivalent to the assumption that Monte picks $\theta$ from $\{1,2,3\}$ by rolling a three-sided die, and in the event of match picks a door $y$ from two possibilities by tossing a fair coin. The general Bayesian formulation, which may be called a game against the nature, models the situation where Conie deals with unconcious random algorithm which neither has own goals nor can take advantage of Conie's mistakes. Conie's optimal behavior is then a {\it Bayesian strategy} $\PP'=\PP'(\QQ)$ which maximizes the probability of winning the prize: $$\PP'\CC\QQ^T= \max_{\PP} \PP\CC\QQ^T.$$ Bayesian strategy always exists by continuity of $\PP\CC\QQ^T$ as function of $\PP$, and compactness of the set of mixed strategies. Moreover, by linearity of $\PP\CC\QQ^T$ and convexity \begin{itemize} \item[] The general Bayesian strategy is an arbitrary mixture of the pure Bayesian strategies. \end{itemize} Now suppose $\QQ$ is fully supported, that is every pure strategy from $\cal M$ is played with positive probability, and let $A$ and $B$ be two distinct pure strategies of Conie such that $B$ weakly dominates $A$. Since all rows of $\cal C$ are distinct the latter means that the collection of columns in which row $A$ has a 1 is a proper subset of the collection of columns in which $B$ has a 1. But then the winning probability is strictly larger for $B$ than for $A$. Since every strategy admitting action $\tm$ is weakly dominated by some always-switching strategy we conclude that \begin{itemize} \item[] If $\QQ$ is fully supported then only always-switching pure strategies can be Bayesian, hence every Bayesian strategy is a mixture of always-switching strategies. \end{itemize} Thus in the `generic' case the Bayesian principle of optimality excludes strategies which may use $\tm$. Having retained only always-switching strategies, it is easy to find all Bayesian strategies explicitly. \begin{itemize} \item[] Suppose that according to fully supported $\QQ$ the probability to hide the prize behind door $\theta\in\{1,2,3\}$ is $\pi_\theta$. Let $\pi_1\geq\pi_2\geq\pi_3>0$ (otherwise re-label the doors). Then the only pure Bayesian strategies are \begin{itemize} \item[\rm(1)] 3\ts\ts~ if $\pi_1\geq \pi_2>\pi_3$, \item[\rm(2)] 2\ts\ts~ and 3\ts\ts~ if $\pi_1>\pi_2=\pi_3$, \item[\rm(3)] 1\ts\ts,~ 2\ts\ts~ and 3\ts\ts~ if $\pi_1=\pi_2=\pi_3=1/3$. \end{itemize} \end{itemize} If $\QQ$ is not fully supported then the listed always-switching strategies are still Bayesian, but some other strategies may be Bayesian too. For instance, a never-switching strategy $x\tm\tm$ is Bayesian if (and only if) Monte always hides the prize behind door $x$, although even then every always-switching strategy $x'\ts\ts$ with $x'\neq x$ is Bayesian also. More interestingly, suppose $\QQ=(1/3,0,1/3,0,1/3,0)$. This is a model for a `crawl' behavior of the host \cite{Rosenthal}, who is eager to reveal the door with higher number when he has a choice. In this case the Bayesian pure strategies are $1\ts\ts,~ 2\ts\ts,~3\ts\ts$ and $1\tm\ts,~ 2\tm\ts,~3\tm\ts$. In general, the rule to determine all Bayesian pure strategies by excluding the dominated strategies is the following: \begin{itemize} \item[] If strategy $\theta,y$ with $y=\min\{1,2,3\}\setminus\{\theta\}$ enters $\QQ$ with nonzero probability then $\theta\ts\tm$ is not Bayesian. \item[] If strategy $\theta,y$ with $y=\max\{1,2,3\}\setminus\{x\}$ enters $\QQ$ with nonzero probability then $\theta\tm\ts$ is not Bayesian. \end{itemize} For arbitrary $\QQ$~ Conie's strategy `point at the door which is the least likely to hide the prize, then always switch' is Bayesian, giving the probability of win $1-\min_{\theta} \pi_{\theta}$, where $\pi_{\theta}$ is the probability of $\theta=1,2,3$. Remarkably, all what Conie needs to know to play optimally in a Bayesian game is the number of the least likely door. She tips at this door and switches all the time, winning the complementary probability no matter which are the biases for revealing the doors in case of match. In the special case $\pi_1=\pi_2=\pi_3=1/3$ every always-switching strategy is Bayesian no matter how $y$ is chosen in the event of match $x=\theta$; a conclusion usually shown in the literature with the help of conditional probabilities. In fact, for an optimal choice of $x$, the {\it conditional} probability of winning with switching in every position $xy$ is at least $1/2$. This is implied by the overall optimality (in the Bayesian sense), and is a simple instance of the general Bellman's dynamic programming principle. \section{The zero-sum game} The zero-sum game is a model for interaction of actors with antagonistic goals. Conie wins the prize when Monte loses it. The essense of the zero-sum game is the {\it worst-case analysis}. What Conie can guarantee in the game, when the behavior of Monte can be arbitrary (but agreeing with the game rules)? What is the worst behavior of Monte? We can write Monte's payoffs as another $12\times 6$ matrix $\HH$, but this is not necessary as $\HH=-\CC$, so $\CC$ contains all information about the payoffs. In this context, when Monty plays $\QQ$, Conies' Bayesian strategy $\PP'=\PP'(\QQ)$ is called a {\it best response} to $\QQ$. Reciprocally, when Conie plays a particular $\PP$ Monte's best response strategy $\QQ'=\QQ'(\PP)$ is the one for which Conie's winning probability is minimal, $$\PP\CC\QQ'^T= \min_{\PP} \PP\CC\QQ^T.$$ A profile of actors' mixed strategies $(\PP^*,\QQ^*)$ is said to be a {\it minimax solution} (or saddle point) if the strategies are the best responses to each other, $$ \PP^*\CC\QQ^{*T}= \max_{\PP} \PP\CC\QQ^{*T} = \min_{\QQ} \PP^*\CC\QQ^T.$$ Such a solution exists for arbitrary matrix game with finite sets of strategies, by the minimax theorem due to von Neumann. The number $V:=\PP^*\CC\QQ^{*T}$ is called the {\it value of the game}, and it does not depend on particular minimax profile due to the fundamental relation $$V=\max_{\PP}\min_{\QQ} \PP\CC\QQ^{T} = \min_{\QQ}\max_{\PP} \PP\CC\QQ^T$$ involved in the minimax theorem. Strategy $\PP^*$ of Conie is minimax if it guarantees the winning probability at least $V$ no matter what Monte does. Strategy $\QQ^*$ of Monte is minimax if it incures the winning probability at most $V$ no matter what Conie does. Recalling our discussion around the dominance it is really easy to see that $V=2/3$. If Monte chooses $\theta$ uniformly at random, Conie cannot winning probability higher than $2/3$. On the other hand, if $x$ is chosen uniformly at random and $x\ts\ts$ is played, Conie guarantees $2/3$ no matter what Monte does. So a solution is found. We wish to approach the solution more formally, manipulaing the payoff matrix. The principle of the elimination of dominated strategies says: \begin{itemize} \item[] the value of the game is not altered if the game matrix is (repeatedly) reduced by eliminating a (weakly) dominated row or column. \end{itemize} Explicitly, strategy $1\swsw$~ dominates 2\masw~ and 2\mama~ \begin{center} \begin{tabular}{c|cccccc} 1\swsw &0 & 0 & 1 & 1 & 1 &1\\ 2\masw &0 & 0 & 1 & 0 & 1 &1\\ 2\mama &0 & 0 & 1 & 1 & 0 &0\\ \end{tabular} \end{center} and 3\swsw~ dominates 2\swma~ \begin{center} \begin{tabular}{c|cccccc} 3\swsw &1 & 1 & 1 & 1 & 0 &0\\ 2\swma &1 & 1 & 0 & 1 & 0 &0\\ \end{tabular} \end{center} Continuing the elimination we reduce to \begin{center} \begin{tabular}{c|cccccc} 1\swsw &0 & 0 & 1 & 1 & 1 &1\\ 2\swsw &1 & 1 & 0 & 0 & 1 &1\\ 3\swsw &1 & 1 & 1 & 1 & 0 &0\\ \end{tabular} \end{center} Finally, discarding repeated columns the matrix is reduced to the square matrix $\Cc$: \vskip0.3cm \begin{center} \begin{tabular}{c|ccc} $\theta,y=$ & 12 & 21 & 32 \\ \hline 1\swsw &0 & 1 & 1 \\ 2\swsw &1 & 0 & 1 \\ 3\swsw &1 & 1 & 0 \\ \end{tabular} \end{center} \vskip0.3cm The reduced game $\Cc$ has a clear interpretation. Monte and Conie choose door numbers $\theta$ and $x$, respectively, from $\{1,2,3\}$. If the choices {\it mismatch} ($\theta\neq x$) Conie wins, otherwise there is no payoff. Let us find an {\it equalizing} strategy $\Pp^*$ for which ${\Pp}^*\Cc{\Qq}^T$ is the same no matter which counter-strategy $\Qq$ Monte plays. Taking for $\Qq$ pure strategies we arrive at the system of equations $$p_2+p_3=p_1+p_2=p_1+p_3,$$ which taken together with $p_1+p_2+p_3=1$ is complete. Solving the system we see that with $\Pp^*=(1/3,1/3,1/3)$ Conie wins with probability 2/3 no matter what Monte does. Similarly, when Monte plays $\Qq^*=(1/3,1/3,1/3)$ the winning probability is always 2/3 no matter what Conie does. Thus $(\Pp^*,\Qq^*)$ is a solution of the game and the value $V=2/3$ is confirmed. Yet another way to arrive at $(\Pp^*,\Qq^*)$ is to use symmetry of matrix $\Cc$ induced by permutations of the door numbers (see \cite{Ferguson}, Theorem 3.4). A related game with diagonal matrix $$\left( \begin{array}{ccc} -1 & 0 & 0 \\ 0 & -1 & 0 \\ 0 & 0 & -1 \\ \end{array}\right) $$ is obtained by subtracting a constant matrix from $\Cc$. In this diagonal game {\it Monte} wins if the choices match, otherwise there is no payoff. Going back to the original matrix $\CC$, we conclude that the profile \begin{eqnarray*} \PP^*=\left({1\over 3},0,0,0,{1\over 3},0,0,0,{1\over 3},0,0,0\right),~~ \QQ^*_{1,1,1}=\left({1\over 3},0,{1\over 3},0,{1\over 3},0\right) \end{eqnarray*} is a solution to the game. Strategy $\PP^*$ is equalizing, as $\PP^*\CC \QQ$ =2/3 for all $\QQ$. According to the solution $(\PP^*, \QQ^*_{1,1,1})$ Monte plays the `crawl' strategy: he hides the prize uniformly at random, and he always reveals the door with higher number, when there is a freedom for the second action. Conie selects door $x$ uniformly at random and always plays $\ts$. The generic Monte's strategy assigning probability $1/3$ to every value of $\theta$ is of the form $$\QQ^*_{{\lambda_1},{\lambda_2},{\lambda_3}}=\left({\lambda_1\over 3}\,,\,{1-\lambda_1\over 3}\,,\, {\lambda_2\over 3}\,,\,{1-\lambda_2\over 3}\,,\,{\lambda_3\over 3}\,,\,{1-\lambda_3\over 3}\right)$$ where $0\leq\lambda_\theta\leq 1$, $\theta\in \{1,2,3\}$. Parameter $\lambda_\theta$ is the conditional probability that Monte will offer switching to door with smaller number in the case of match. If Conie plays best response to $\QQ^*_{{\lambda_1},{\lambda_2},{\lambda_3}}$ she wins with probability $2/3$, as we have seen when considering the Bayesian setting. Therefore each strategy $\QQ^*_{{\lambda_1},{\lambda_2},{\lambda_3}}$ is minimax. On the other hand, if the values of $\theta$ have probabilities $\pi_\theta$ a best response of Conie yields winning probability $1-\min(\pi_1,\pi_2,\pi_3)$, which is minimized for $\pi_1=\pi_2=\pi_3=1/3$. We see that \begin{itemize} \item[] The strategies $\QQ^*_{{\lambda_1},{\lambda_2},{\lambda_3}}$ and only they are the minimax strategies of Monte. \end{itemize} Furthermore, if $\QQ^*_{{\lambda_1},{\lambda_2},{\lambda_3}}$ is fully supported, which is the case when $\lambda_{\theta}\in (0,1)$ for $\theta\in \{1,2,3\}$, then the unique best response is $\PP^*$, thus \begin{itemize} \item[] Conie's strategy $\PP^*$ of choosing a door uniformly at random, then always switching is the unique minimax strategy. \end{itemize} The subclass of Monte's strategies with the second action independent of the first given match consists of strategies with equal probabilities $\lambda_1=\lambda_2=\lambda_3$; these were discussed e.g. in \cite{Rosenhouse} (Version Five of the MHP). We note that the general theory does not preclude weakly dominated strategies from being minimax (see \cite{Ferguson}, Section 2.6, Exercise 9). This does not occur in the MHP because there exist fully supported minimax strategies of Monte. \section{The general-sum games} The logical way to go beyond the zero-sum game is a non-zero sum game. In such model there is a $12\times 6$ payoff matrix of Monte $\HH$ which need not be the negative of $\CC$. A best response to Monte's strategy $\QQ$ is defined as before, but best response to Conie's strategy $\PP$ is now a strategy which maximizes the expected payoff $\PP\HH\QQ^T$. A central solution concept for the general-sum game game is a {\it Nash equilibrium}, defined as a profile of mixed strategies $(\PP',\QQ')$ which are best responses to each other, $$\PP'\CC\QQ'^{T}=\max_{\PP} \PP\CC\QQ'^{T} ~~~{\rm and~~~} \PP'\HH\QQ'^{T}=\max_{\QQ} \PP'\HH\QQ^{T}.$$ That is to say, Nash equilibrium is a profile $(\PP',\QQ')$ such that a unilateral change of the strategy by one of the actors cannot improve the private payoff of the actor. A general theorem due to John Nash ensures that at least one such Nash equilibrium exists. Nash equilibrium is a concept of {\it noncooperative} game theory. The players cannot make binding agreements on a joint choice of strategy unless the agreements are self-enforced. This is formalized by the independence presumed in the product formulas for computing the payoffs. In every Nash equilibrium Conie will have winning probability not less than her minimax value $V=2/3$, which is her {\it safety level}. Higher probability in some Nash equilibria might be possible, since the game is no longer antagonistic. Our analysis of the Bayesian strategies can be applied to the general-sum games as well. Suppose in a Nash equilibrium $(\PP',\QQ')$ strategy $\QQ'$ is fully supported. Then $\PP'$, being a best response to $\QQ'$, is a mixture of always-switching strategies. Let $\pi_1\geq \pi_2\geq \pi_3>0$ be probabilities of the values $\theta=1,2,3$ under $\QQ'$. Then we have \begin{itemize} \item[] {\rm A profile $(\PP',\QQ')$ is a Nash equilibrium with fully supported $\QQ'$ if and only if there exists a probability vector $(p_1,p_2,p_3)$ such that the mixture with weights $p_1, p_2, p_3$ of rows $1\ts\ts, 2\ts\ts, 3\ts\ts$ of matrix $\HH$ is a row vector with equal entries. There are three possibilities \vskip0.1cm \noindent \begin{itemize} \item[\rm(1)] $p_3=1$, $\pi_2>\pi_3$ and the row $3\ts\ts$ of $\HH$ is a constant row, \item[\rm(2)] $p_1=0$, $\pi_1> \pi_2=\pi_3$ and a mixture of the rows $2\ts\ts$ and $3\ts\ts$ of $\HH$ is a constant row, \item[\rm(3)] $p_1p_2p_3>0,~ \pi_1=\pi_2=\pi_3=1/3$, and the arithmetical average of rows $1\ts\ts$, $2\ts\ts$ and $3\ts\ts$ of $\HH$ is a constant row \end{itemize} } \end{itemize} In the case (3) Conie's winning probability is her safety level 2/3 but, unlike the zero-sum game, the equilibrium strategy $\PP'$ need not give the same probability to every always-switching strategy. Thus if the equilibrium has a property of nondegeneracy, when every pure strategy of Monte and every always-switching strategy of Conie is played with positive probability, then for Conie the game brings the same as the game against antagonistic Monte. If none of the mixtures of the $x\ts\ts$-rows is constant, there is no fully supported Nash equilibrium. What could be plausible assumptions on Monte's payoff $\HH$? If Monte is only concerned about the fate of the prize, and not where and how the prize is won, his essentially distinct payoff structures are $\HH=-\CC$ (antagonistic Monte) $\HH=\CC$ (sympathetic Monte) and $\HH={\bf 0}$ (indifferent Monte). The first case is zero-sum, in the second case every entry 1 of $\CC$ corresponds to a pure Nash equilibrium, and in the third case every pair $(\QQ, \PP')$ with best-response $\PP'=\PP'(\QQ)$ is a Nash equilibrium. It is not hard to design further exotic examples of payoffs $\HH$ for which context-dependent strategies enter some Nash equilibrium profile. If there is a moral of that it is perhaps this: context-dependent strategies {\it may} be a rational kind of behavior under certain intensions of Monte. The source for this phenomenon is twofold. Firstly, this is the very idea of equilibrium: if Conie steps away from the context-dependent equilibrium strategy, Monty is not forced to play old strategy and may change biases in a way unfavorable for Conie, typically pushing her down to the safety level $2/3$. Secondly, the domination is only {\it weak}, thus can have no effect if support is not full. However, such context-dependent equilibria are highly unstable, and minor perturbations of $\HH$ will destroy them. Practically speaking, if Conie has any doubts about Monte's intensions it is safe to stay with always-switching strategies. As for the famous question, the noncooperative game theory gives more weight to vos Savant's solution by adding \begin{itemize} \item[] {\it Yes, you should switch. You knew the rules of the game. If your decision were to pick door 1 and hold when a switch to door 2 offered, then I could beat your strategy by picking door 2 and switching whichever happens. My strategy will be even strictly better than yours if the prize can ever be hidden behind door 2.} \end{itemize} If the game is {\it cooperative}, for instance if Monte and Conie want happily drive their new cadillac to Nice, they could just favor door 1. But this is a completely different story.
{ "timestamp": "2011-07-05T02:00:12", "yymm": "1107", "arxiv_id": "1107.0326", "language": "en", "url": "https://arxiv.org/abs/1107.0326", "abstract": "The basic Monty Hall problem is explored to introduce into the fundamental concepts of the game theory and to give a complete Bayesian and a (noncooperative) game-theoretic analysis of the situation. Simple combinatorial arguments are used to exclude the holding action and to find minimax solutions.", "subjects": "History and Overview (math.HO); Computer Science and Game Theory (cs.GT)", "title": "The Monty Hall Problem in the Game Theory Class", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9914225130656699, "lm_q2_score": 0.8267117983401363, "lm_q1q2_score": 0.8196206886914172 }
https://arxiv.org/abs/math/0503745
Pseudo-random graphs
Random graphs have proven to be one of the most important and fruitful concepts in modern Combinatorics and Theoretical Computer Science. Besides being a fascinating study subject for their own sake, they serve as essential instruments in proving an enormous number of combinatorial statements, making their role quite hard to overestimate. Their tremendous success serves as a natural motivation for the following very general and deep informal questions: what are the essential properties of random graphs? How can one tell when a given graph behaves like a random graph? How to create deterministically graphs that look random-like? This leads us to a concept of pseudo-random graphs and the aim of this survey is to provide a systematic treatment of this concept.
\section{Introduction} Random graphs have proven to be one of the most important and fruitful concepts in modern Combinatorics and Theoretical Computer Science. Besides being a fascinating study subject for their own sake, they serve as essential instruments in proving an enormous number of combinatorial statements, making their role quite hard to overestimate. Their tremendous success serves as a natural motivation for the following very general and deep informal questions: what are the essential properties of random graphs? How can one tell when a given graph behaves like a random graph? How to create deterministically graphs that look random-like? This leads us to a concept of {\em pseudo-random graphs}. Speaking very informally, a pseudo-random graph $G=(V,E)$ is a graph that behaves like a truly random graph $G(|V|,p)$ of the same edge density $p=|E|\left/{{|V|}\choose 2}\right.$. Although the last sentence gives some initial idea about this concept, it is not very informative, as first of all it does not say in which aspect the pseudo-random graph behavior is similar to that of the corresponding random graph, and secondly it does not supply any quantitative measure of this similarity. There are quite a few possible graph parameters that can potentially serve for comparing pseudo-random and random graphs (and in fact quite a few of them are equivalent in certain, very natural sense, as we will see later), but probably the most important characteristics of a truly random graph is its {\em edge distribution}. We can thus make a significant step forward and say that a pseudo-random graph is a graph with edge distribution resembling the one of a truly random graph with the same edge density. Still, the quantitative measure of this resemblance remains to be introduced. Although first examples and applications of pseudo-random graphs appeared very long time ago, it was Andrew Thomason who launched systematic research on this subject with his two papers \cite{Tho87a}, \cite{Tho87b} in the mid-eighties. Thomason introduced the notion of jumbled graphs, enabling to measure in quantitative terms the similarity between the edge distributions of pseudo-random and truly random graphs. He also supplied several examples of pseudo-random graphs and discussed many of their properties. Thomason's papers undoubtedly defined directions of future research for many years. Another cornerstone contribution belongs to Chung, Graham and Wilson \cite{ChuGraWil89} who in 1989 showed that many properties of different nature are in certain sense equivalent to the notion of pseudo-randomness, defined using the edge distribution. This fundamental result opened many new horizons by showing additional facets of pseudo-randomness. Last years brought many new and striking results on pseudo-randomness by various researchers. There are two clear trends in recent research on pseudo-random graphs. The first is to apply very diverse methods from different fields (algebraic, linear algebraic, combinatorial, probabilistic etc.) to construct and study pseudo-random graphs. The second and equally encouraging is to find applications, in many cases quite surprising, of pseudo-random graphs to problems in Graph Theory, Computer Science and other disciplines. This mutually enriching interplay has greatly contributed to significant progress in research on pseudo-randomness achieved lately. The aim of this survey is to provide a systematic treatment of the concept of pseudo-random graphs, probably the first since the two seminal contributions of Thomason \cite{Tho87a}, \cite{Tho87b}. Research in pseudo-random graphs has developed tremendously since then, making it impossible to provide full coverage of this subject in a single paper. We are thus forced to omit quite a few directions, approaches, theorem proofs from our discussion. Nevertheless we will attempt to provide the reader with a rather detailed and illustrative account of the current state of research in pseudo-random graphs. Although, as we will discuss later, there are several possible formal approaches to pseudo-randomness, we will mostly emphasize the approach based on graph eigenvalues. We find this approach, combining linear algebraic and combinatorial tools in a very elegant way, probably the most appealing, convenient and yet quite powerful. This survey is structured as follows. In the next section we will discuss various formal definitions of the notion of pseudo-randomness, from the so called jumbled graphs of Thomason to the $(n,d,\lambda)$-graphs defined by Alon, where pseudo-randomness is connected to the eigenvalue gap. We then describe several known constructions of pseudo-random graphs, serving both as illustrative examples for the notion of pseudo-randomness, and also as test cases for many of the theorems to be presented afterwards. The strength of every abstract concept is best tested by properties it enables to derive. Pseudo-random graphs are certainly not an exception here, so in Section 4 we discuss various properties of pseudo-random graphs. Section 5, the final section of the paper, is devoted to concluding remarks. \section{Definitions of pseudo-random graphs} Pseudo-random graphs are much more of a general concept describing some graph theoretic phenomenon than of a rigid well defined notion -- the fact reflected already in the plural form of the title of this section! Here we describe various formal approaches to the concept of pseudo-randomness. We start with stating known facts on the edge distribution of random graphs, that will serve later as a benchmark for all other definitions. Then we discuss the notion of jumbled graphs introduced by Thomason in the mid-eighties. Then we pass on to the discussion of graph properties, equivalent in a weak (qualitative) sense to the pseudo-random edge distribution, as revealed by Chung, Graham and Wilson in \cite{ChuGraWil89}. Our next item in this section is the definition of pseudo-randomness based on graph eigenvalues -- the approach most frequently used in this survey. Finally, we discuss the related notion of strongly regular graphs, their eigenvalues and their relation to pseudo-randomness. \subsection{Random graphs} As we have already indicated in the Introduction, pseudo-random graphs are modeled after truly random graphs, and therefore mastering the edge distribution in random graphs can provide the most useful insight on what can be expected from pseudo-random graphs. The aim of this subsection is to state all necessary definitions and results on random graphs. We certainly do not intend to be comprehensive here, instead referring the reader to two monographs on random graphs \cite{Bol01}, \cite{JanLucRuc00}, devoted entirely to the subject and presenting a very detailed picture of the current research in this area. A {\em random graph} $G(n,p)$ is a probability space of all labeled graphs on $n$ vertices $\{1,\ldots,n\}$, where for each pair $1\le i<j\le n$, $(i,j)$ is an edge of $G(n,p)$ with probability $p=p(n)$, independently of any other edges. Equivalently, the probability of a graph $G=(V,E)$ with $V=\{1,\ldots,n\}$ in $G(n,p)$ is $Pr[G]=p^{|E(G)|}(1-p)^{{n\choose 2}-|E(G)|}$. We will occasionally mention also the probability space $G_{n,d}$, this is the probability space of all $d$-regular graphs on $n$ vertices endowed with the uniform measure, see the survey of Wormald \cite{Wor99} for more background. We also say that a graph property ${\cal A}$ holds {\em almost surely}, or a.s. for brevity, in $G(n,p)~(G_{n,d})$ if the probability that $G(n,p)~(G_{n,d})$ has ${\cal A}$ tends to one as the number of vertices $n$ tends to infinity. From our point of view the most important parameter of random graph $G(n,p)$ is its edge distribution. This characteristics can be easily handled due to the fact that $G(n,p)$ is a product probability space with independent appearances of different edges. Below we cite known results on the edge distribution in $G(n,p)$. \begin{theo}\label{dist1} Let $p=p(n)\le 0.99$. Then almost surely $G\in G(n,p)$ is such that if $U$ is any set of $u$ vertices, then $$ \left|e(U)-p{u\choose 2}\right| = O\left(u^{3/2}p^{1/2}\log^{1/2}(2n/u)\right)\ . $$ \end{theo} \begin{theo}\label{dist2} Let $p=p(n)\le 0.99$. Then almost surely $G\in G(n,p)$ is such that if $U,W$ are disjoint sets of vertices satisfying $u=|U|\le w=|W|$, then $$ \left|e(U,W)-puw\right|=O\left(u^{1/2}wp^{1/2}\log^{1/2}(2n/w)\right)\ . $$ \end{theo} The proof of the above two statements is rather straightforward. Notice that both quantities $e(U)$ and $e(U,W)$ are binomially distributed random variables with parameters ${u\choose 2}$ and $p$, and $uw$ and $p$, respectively. Applying standard Chernoff-type estimates on the tails of the binomial distribution (see, e.g., Appendix A of \cite{AloSpe00}) and then the union bound, one gets the desired inequalities. It is very instructive to notice that we get less and less control over the edge distribution as the set size becomes smaller. For example, in the probability space $G(n,1/2)$ every subset is expected to contain half of its potential edges. While this is what happens almost surely for large enough sets due to Theorem \ref{dist1}, there will be almost surely sets of size about $2\log_2n$ containing all possible edges (i.e. cliques), and there will be almost surely sets of about the same size, containing no edges at all (i.e. independent sets). For future comparison we formulate the above two theorems in the following unified form: \begin{coro}\label{dist3} Let $p=p(n)\le 0.99$. Then almost surely in $G(n,p)$ for every two (not necessarily) disjoint subsets of vertices $U,W\subset V$ of cardinalities $|U|=u, |W|=w$, the number $e(U,W)$ of edges of $G$ with one endpoint in $U$ and the other one in $W$ satisfies: \begin{eqnarray}\label{dist} |e(U,W)-puw|= O(\sqrt{uwnp})\ . \end{eqnarray} \end{coro} (A notational agreement here and later in the paper: if an edge $e$ belongs to the intersection $U\cap W$, then $e$ is counted twice in $e(U,W)$.) Similar bounds for edge distribution hold also in the space $G_{n,d}$ of $d$-regular graphs, although they are significantly harder to derive there. Inequality (\ref{dist}) provides us with a quantitative benchmark, according to which we will later measure the uniformity of edge distribution in pseudo-random graphs on $n$ vertices with edge density $p=|E(G)|\left/{n\choose 2}\right.$ It is interesting to draw comparisons between research in random graphs and in pseudo-random graphs. In general, many properties of random graphs are much easier to study than the corresponding properties of pseudo-random graphs, mainly due to the fact that along with the almost uniform edge distribution described in Corollary \ref{dist3}, random graphs possess as well many other nice features, first and foremost of them being that they are in fact very simply defined product probability spaces. Certain graph properties can be easily shown to hold almost surely in $G(n,p)$ while they are not necessarily valid in pseudo-random graphs of the same edge density. We will see quite a few such examples in the next section. A general line of research appears to be not to use pseudo-random methods to get new results for random graphs, but rather to try to adapt techniques developed for random graphs to the case of pseudo-random graphs, or alternatively to develop original techniques and methods. \subsection{Thomason's jumbled graphs} In two fundamental papers \cite{Tho87a}, \cite{Tho87b} published in 1987 Andrew Thomason introduced the first formal quantitative definition of pseudo-random graphs. It appears quite safe to attribute the launch of the systematic study of pseudo-randomness to Thomason's papers. Thomason used the term "jumbled" graphs in his papers. A graph $G=(V,E)$ is said to be $(p,\alpha)$-{\em jumbled} if $p,\alpha$ are real numbers satisfying $0<p<1\le \alpha$ if every subset of vertices $U\subset V$ satisfies: \begin{eqnarray}\label{jumbled} \left|e(U)- p{{|U|}\choose 2}\right|\le \alpha |U|\ . \end{eqnarray} The parameter $p$ can be thought of as the density of $G$, while $\alpha$ controls the deviation from the ideal distribution. According to Thomason, the word "jumbled" is intended to convey the fact that the edges are evenly spread throughout the graph. The motivation for the above definition can be clearly traced to the attempt to compare the edge distribution in a graph $G$ to that of a truly random graph $G(n,p)$. Applying it indeed to $G(n,p)$ and recalling (\ref{dist}) we conclude that the random graph $G(n,p)$ is almost surely $O(\sqrt{np})$-jumbled. Thomason's definition has several trivial yet very nice features. Observe for example that if $G$ is $(p,\alpha)$-jumbled then the complement $\bar{G}$ is $(1-p,\alpha)$-jumbled. Also, the definition is hereditary -- if $G$ is $(p,\alpha)$-jumbled, then so is every induced subgraph $H$ of $G$. Note that being $(p,\Theta(np))$-jumbled for a graph $G$ on $n$ vertices and ${n\choose 2}p$ edges does not say too much about the edge distribution of $G$ as the number of edges in linear sized sets can deviate by a percentage from their expected value. However as we shall see very soon if $G$ is known to be $(p,o(np))$-jumbled, quite a lot can be said about its properties. Of course, the smaller is the value of $\alpha$, the more uniform or jumbled is the edge distribution of $G$. A natural question is then how small can be the parameter $\alpha=\alpha(n,p)$ for a graph $G=(V,E)$ on $|V|=n$ vertices with edge density $p=|E|\left/{n\choose 2}\right.$? Erd\H os and Spencer proved in \cite{ErdSpe72} that $\alpha$ satisfies $\alpha=\Omega(\sqrt{n})$ for a constant $p$; their method can be extended to show $\alpha=\Omega(\sqrt{np})$ for all values of $p=p(n)$. We thus may think about $(p,O(\sqrt{np}))$-jumbled graphs on $n$ vertices as in a sense best possible pseudo-random graphs. Although the fact that $G$ is $(p,\alpha)$-jumbled carries in it a lot of diverse information on the graph, it says almost nothing (directly at least) about small subgraphs, i.e. those spanned by subsets $U$ of size $|U|=o(\alpha/p)$. Therefore in principle a $(p,\alpha)$-jumbled graph can have subsets of size $|U|=O(\alpha/p)$ spanning by a constant factor less or more edges then predicted by the uniform distribution. In many cases however quite a meaningful local information (such as the presence of subgraphs of fixed size) can still be salvaged from global considerations as we will see later. Condition (\ref{jumbled}) has obviously a global nature as it applies to {\em all} subsets of $G$, and there are exponentially many of them. Therefore the following result of Thomason, providing a sufficient condition for pseudo-randomness based on degrees and co-degrees only, carries a certain element of surprise in it. \begin{theo}\label{jumlocal}\cite{Tho87a} Let $G$ be a graph on $n$ vertices with minimum degree $np$. If no pair of vertices of $G$ has more than $np^2+l$ common neighbors, then $G$ is $(p,\sqrt{(p+l)n})$-jumbled. \end{theo} The above theorem shows how the pseudo-randomness condition of (\ref{jumbled}) can be ensured/checked by testing only a polynomial number of easily accessible conditions. It is very useful for showing that specific constructions are jumbled. Also, it can find algorithmic applications, for example, a very similar approach has been used by Alon, Duke, Lefmann, R\"odl and Yuster in their Algorithmic Regularity Lemma \cite{AloDukLefRodYus94}. As observed by Thomason, the minimum degree condition of Theorem \ref{jumlocal} can be dropped if we require that every pair of vertices has $(1+o(1))np^2$ common neighbors. One cannot however weaken the conditions of the theorem so as to only require that every {\em edge} is in at most $np^2+l$ triangles. Another sufficient condition for pseudo-randomness, this time of global nature, has also been provided in \cite{Tho87a}, \cite{Tho87b}: \begin{theo}\label{jumglobal}\cite{Tho87a} Let $G$ be a graph of order $n$, let $\eta n$ be an integer between 2 and $n-2$, and let $\omega>1$ be a real number. Suppose that each induced subgraph $H$ of order $\eta n$ satisfies $|e(H)-p{{\eta n}\choose 2}|\le \eta n\alpha$. Then $G$ is $(p,7\sqrt{n\alpha/\eta}/(1-\eta))$-jumbled. Moreover $G$ contains a subset $U\subseteq V(G)$ of size $|U|\ge \left(1-\frac{380}{n(1-\eta)^2w}\right)n$ such that the induced subgraph $G[U]$ is $(p,\omega\alpha)$-jumbled. \end{theo} Thomason also describes in \cite{Tho87a}, \cite{Tho87b} several properties of jumbled graphs. We will not discuss these results in details here as we will mostly adopt a different approach to pseudo-randomness. Occasionally however we will compare some of later results to those obtained by Thomason. \subsection{Equivalent definitions of weak pseudo-randomness} Let us go back to the jumbledness condition (\ref{jumbled}) of Thomason. As we have already noted it becomes non-trivial only when the error term in (\ref{jumbled}) is $o(n^2p)$. Thus the latter condition can be considered as the weakest possible condition for pseudo-randomness. Guided by the above observation we now define the notion of weak pseudo-randomness as follows. Let $(G_n)$ be a sequence of graphs, where $G_n$ has $n$ vertices. Let also $p=p(n)$ is a parameter ($p(n)$ is a typical density of graphs in the sequence). We say that the sequence $(G_n)$ is {\em weakly pseudo-random} if the following condition holds: \begin{eqnarray}\label{weakpr} \mbox{For all subsets $U\subseteq V(G_n)$,}\quad\quad \left|e(U)-p{{|U|}\choose 2}\right|=o(n^2p)\ . \end{eqnarray} For notational convenience we will frequently write $G=G_n$, tacitly assuming that $(G)$ is in fact a sequence of graphs. Notice that the error term in the above condition of weak pseudo-randomness does not depend on the size of the subset $U$. Therefore it applies essentially only to subsets $U$ of linear size, ignoring subsets $U$ of size $o(n)$. Hence (\ref{weakpr}) is potentially much weaker than Thomason's jumbledness condition (\ref{jumbled}). Corollary \ref{dist3} supplies us with the first example of weakly pseudo-random graphs -- a random graph $G(n,p)$ is weakly pseudo-random as long as $p(n)$ satisfies $np\rightarrow\infty$. We can thus say that if a graph $G$ on $n$ vertices is weakly pseudo-random for a parameter $p$, then the edge distribution of $G$ is close to that of $G(n,p)$. In the previous subsection we have already seen examples of conditions implying pseudo-randomness. In general one can expect that conditions of various kinds that hold almost surely in $G(n,p)$ may imply or be equivalent to weak pseudo-randomness of graphs with edge density $p$. Let us first consider the case of the constant edge density $p$. This case has been treated extensively in the celebrated paper of Chung, Graham and Wilson from 1989 \cite{ChuGraWil89}, where they formulated several equivalent conditions for weak pseudo-randomness. In order to state their important result we need to introduce some notation. Let $G=(V,E)$ be a graph on $n$ vertices. For a graph $L$ we denote by $N^*_G(L)$ the number of labeled induced copies of $L$ in $G$, and by $N_G(L)$ the number of labeled not necessarily induced copies of $L$ in $G$. For a pair of vertices $x,y\in V(G)$, we set $s(x,y)$ to be the number of vertices of $G$ joined to $x$ and $y$ the same way: either to both or to none. Also, $codeg(x,y)$ is the number of common neighbors of $x$ and $y$ in $G$. Finally, we order the eigenvalues $\lambda_i$ of the adjacency matrix $A(G)$ so that $|\lambda_1|\ge |\lambda_2|\ge\ldots\ge |\lambda_n|$. \begin{theo}\label{CGW}\cite{ChuGraWil89} Let $p\in (0,1)$ be fixed. For any graph sequence $(G_n)$ the following properties are equivalent: \begin{description} \item[$P_1(l)$:\quad] For a fixed $l\ge 4$ for all graphs $L$ on $l$ vertices, $$ N_G^*(L)=(1+o(1))n^lp^{|E(L)|}(1-p)^{{l\choose 2}-|E(L)|}\ . $$ \item[$P_2(t)$:\quad] Let $C_t$ denote the cycle of length $t$. Let $t\ge 4$ be even, $$ e(G_n)=\frac{n^2p}{2}+o(n^2)\quad\mbox{and}\quad N_G(C_t)\le (np)^t+o(n^t)\ . $$ \item[$P_3$:\quad] $ e(G_n)\ge \frac{n^2p}{2}+o(n^2)\quad\mbox{and}\quad \lambda_1=(1+o(1))np,~~ \lambda_2=o(n)\ . $ \item[$P_4$:\quad] For each subset $U\subset V(G)$,\quad $e(U)=\frac{p}{2}|U|^2+o(n^2)$\ . \item[$P_5$:\quad] For each subset $U\subset V(G)$ with $|U|=\lfloor\frac{n}{2}\rfloor$,\quad we have\quad $e(U)=\left(\frac{p}{8}+o(1)\right)n^2\ .$ \item[$P_6$:\quad] $\sum_{x,y\in V} |s(x,y)-(p^2+(1-p)^2)n|=o(n^3)$\ . \item[$P_7$:\quad] $\sum_{x,y\in V} |codeg(x,y)-p^2n|=o(n^3)$\ . \end{description} \end{theo} Note that condition $P_4$ of this remarkable theorem is in fact identical to our condition (\ref{weakpr}) of weak pseudo-randomness. Thus according to the theorem all conditions $P_1$--$P_3$, $P_5-P_7$ are in fact equivalent to weak pseudo-randomness! As noted by Chung et al. probably the most surprising fact (although possibly less surprising for the reader in view of Theorem \ref{jumlocal}) is that apparently the weak condition $P_2(4)$ is strong enough to imply weak pseudo-randomness. It is quite easy to add another condition to the equivalence list of the above theorem: for all $U,W\subset V$, $e(U,W)=p|U||W|+o(n^2)$. A condition of a very different type, related to the celebrated Szemer\'edi Regularity Lemma has been added to the above list by Simonovits and S\'os in \cite{SimSos91}. They showed that if a graph $G$ possesses a Szemer\'edi partition in which almost all pairs have density $p$, then $G$ is weakly pseudo-random, and conversely if $G$ is weakly pseudo-random then in every Szemer\'edi partition all pairs are regular with density $p$. An extensive background on the Szemer\'edi Regularity Lemma, containing in particular the definitions of the above used notions, can be found in a survey paper of Koml\'os and Simonovits \cite{KomSim96}. The reader may have gotten the feeling that basically every property of random graphs $G(n,p)$ ensures weak pseudo-randomness. This feeling is quite misleading, and one should be careful while formulating properties equivalent to pseudo-randomness. Here is an example provided by Chung et al. Let $G$ be a graph with vertex set $\{1,\ldots,4n\}$ defined as follows: the subgraph of $G$ spanned by the first $2n$ vertices is a complete bipartite graph $K_{n,n}$, the subgraph spanned by the last $2n$ vertices is the complement of $K_{n,n}$, and for every pair $(i,j),1\le i\le 2n, 2n+1\le j\le 4n$, the edge $(i,j)$ is present in $G$ independently with probability $0.5$. Then $G$ is almost surely a graph on $4n$ vertices with edge density $0.5$. One can verify that $G$ has properties $P_1(3)$ and $P_2(2t+1)$ for every $t\ge 1$, but is obviously very far from being pseudo-random (contains a clique and an independent set of one quarter of its size). Hence $P_1(3)$ and $P_2(2t+1)$ are not pseudo-random properties. This example shows also the real difference between even and odd cycles in this context -- recall that Property $P_2(2t)$ does imply pseudo-randomness. A possible explanation to the above described somewhat disturbing phenomenon has been suggested by Simonovits and S\'os in \cite{SimSos97}. They noticed that the above discussed properties are not hereditary in the sense that the fact that the whole graph $G$ possesses one of these properties does not imply that large induced subgraphs of $G$ also have it. A property is called {\em hereditary} in this context if it is assumed to hold for all sufficiently large subgraphs $F$ of our graph $G$ with the same error term as for $G$. Simonovits and S\'os proved that adding this hereditary condition gives significant extra strength to many properties making them pseudo-random. \begin{theo}\cite{SimSos97}\label{SSher} Let $L$ be a fixed graph on $l$ vertices, and let $p\in (0,1)$ be fixed. Let $(G_n)$ be a sequence of graphs. If for every induced subgraph $H\subseteq G$ on $h$ vertices, $$ N_H(L)=p^{|E(L)|}h^l+o(n^l)\,, $$ then $(G_n)$ is weakly pseudo-random, i.e. property $P_4$ holds. \end{theo} Two main distinctive features of the last result compared to Theorem \ref{CGW} are: (a) $P_1(3)$ assumed hereditarily implies pseudo-randomness; and (b) requiring the right number of copies of a {\em single} graph $L$ on $l$ vertices is enough, compared to Condition $P_1(l)$ required to hold for {\em all} graphs on $l$ vertices simultaneously. Let us switch now to the case of vanishing edge density $p(n)=o(1)$. This case has been treated in two very recent papers of Chung and Graham \cite{ChuGra02} and of Kohayakawa, R\"odl and Sissokho \cite{KohRodSis02}. Here the picture becomes significantly more complicated compared to the dense case. In particular, there exist graphs with very balanced edge distribution not containing a single copy of some fixed subgraphs (see the Erd\H os-R\'enyi graph and the Alon graph in the next section (Examples 6, 9, resp.)). In an attempt to find properties equivalent to weak pseudo-randomness in the sparse case, Chung and Graham define the following properties in \cite{ChuGra02} : \noindent{\bf CIRCUIT($t$):} The number of closed walks $w_0,w_1,\ldots,w_t=w_0$ of length $t$ in $G$ is $(1+o(1))(np)^t$; \noindent{\bf CYCLE($t$):} The number of labeled $t$-cycles in $G$ is $(1+o(1))(np)^t$; \noindent{\bf EIG:} The eigenvalues $\lambda_i$, $|\lambda_1|\ge|\lambda_2|\ge\ldots |\lambda_n|$, of the adjacency matrix of $G$ satisfy: \begin{eqnarray*} \lambda_1&=&(1+o(1))np\,,\\ |\lambda_i|&=&o(np), i>1\,. \end{eqnarray*} \noindent{\bf DISC:} For all $X,Y\subset V(G)$, $$ |e(X,Y)-p|X||Y||=o(pn^2)\ . $$ (DISC here is in fact DICS(1) in \cite{ChuGra02}). \begin{theo}\label{CG1}\cite{ChuGra02} Let $(G=G_n: n\rightarrow \infty)$ be a sequence of graphs with $e(G_n)=(1+o(1))p{n\choose 2}$. Then the following implications hold for all $t\ge 1$: $$ CIRCUIT(2t)\Rightarrow EIG\Rightarrow DISC\ . $$ \end{theo} \noindent{\bf Proof.\quad} To prove the first implication, let $A$ be the adjacency matrix of $G$, and consider the trace $Tr(A^{2t})$. The $(i,i)$-entry of $A^{2t}$ is equal to the number of closed walks of length $2t$ starting and ending at $i$, and hence $Tr(A^{2t})=(1+o(1))(np)^{2t}$. On the other hand, since $A$ is symmetric it is similar to the diagonal matrix $D=diag(\lambda_1,\lambda_2,\ldots,\lambda_n)$, and therefore $Tr(A^{2t})=\sum_{i=1}^{2t}\lambda_i^{2t}$. We obtain: $$ \sum_{i=1}^n\lambda_i^{2t}=(1+o(1))(np)^{2t}\ . $$ Since the first eigenvalue of $G$ is easily shown to be as large as its average degree, it follows that $\lambda_1\ge 2|E(G)|/|V(G)|=(1+o(1))np$. Combining these two facts we derive that $\lambda_1=(1+o(1))np$ and $|\lambda_i|=o(np)$ as required. The second implication will be proven in the next subsection. \hfill $\Box$ \medskip Both reverse implications are false in general. To see why $DISC\not\Rightarrow EIG$ take a graph $G_0$ on $n-1$ vertices with all degrees equal to $(1+o(1))n^{0.1}$ and having property $DISC$ (see next section for examples of such graphs). Now add to $G_0$ a vertex $v^*$ and connect it to any set of size $n^{0.8}$ in $G_0$, let $G$ be the obtained graph. Since $G$ is obtained from $G_0$ by adding $o(|E(G_0|)$ edges, $G$ still satisfies $DISC$. On the other hand, $G$ contains a star $S$ of size $n^{0.8}$ with a center at $v^*$, and hence $\lambda_1(G)\ge\lambda_1(S)=\sqrt{n^{0.8}-1}\gg |E(G)|/n$ (see, e.g. Chapter 11 of \cite{Lov93} for the relevant proofs). This solves an open question from \cite{ChuGra02}. The Erd\H os-R\'enyi graph from the next section is easily seen to satisfy $EIG$, but fails to satisfy $CIRCUIT(4)$. Chung and Graham provide an alternative example in \cite{ChuGra02} (Example 1). The above discussion indicates that one probably needs to impose some additional condition on the graph $G$ to glue all these pieces together and to make the above stated properties equivalent. One such condition has been suggested by Chung and Graham who defined: \medskip \noindent{\bf U($t$):} For some absolute constant $c$, all degrees in $G$ satisfy: $d(v)< cnp$, and for every pair of vertices $x,y\in G$ the number $e_{t-1}(x,y)$ of walks of length $t-1$ from $x$ to $y$ satisfies: $e_{t-1}(x,y)\le cn^{t-2}p^{t-1}$. \medskip Notice that $U(t)$ can only hold for $p>c'n^{-1+1/(t-1)}$, where $c'$ depends on $c$. Also, every dense graph ($p=\Theta(1)$) satisfies $U(t)$. As it turns out adding property $U(t)$ makes all the above defined properties equivalent and thus equivalent to the notion of weak pseudo-randomness (that can be identified with property $DISC$): \begin{theo}\label{CG2}\cite{ChuGra02} Suppose for some constant $c>0$, $p(n)>cn^{-1+1/(t-1)}$, where $t\ge 2$. For any family of graphs $G_n$, $|E(G_n)|=(1+o(1))p{n\choose 2}$, satisfying $U(t)$, the following properties are all equivalent: $CIRCUIT(2t), CYCLE(2t), EIG$ and $DISC$. \end{theo} Theorem \ref{CG2} can be viewed as a sparse analog of Theorem \ref{CGW} as it also provides a list of conditions equivalent to weak pseudo-randomness. Further properties implying or equivalent to pseudo-randomness, including local statistics conditions, are given in \cite{KohRodSis02}. \subsection{Eigenvalues and pseudo-random graphs} In this subsection we describe an approach to pseudo-randomness based on graph eigenvalues -- the approach most frequently used in this survey. Although the eigenvalue-based condition is not as general as the jumbledness condition of Thomason or some other properties described in the previous subsection, its power and convenience are so appealing that they certainly constitute a good enough reason to prefer this approach. Below we first provide a necessary background on graph spectra and then derive quantitative estimates connecting the eigenvalue gap and edge distribution. Recall that the {\em adjacency matrix} of a graph $G=(V,E)$ with vertex set $V=\{1,\ldots,n\}$ is an $n$-by-$n$ matrix whose entry $a_{ij}$ is 1 if $(i,j)\in E(G)$, and is 0 otherwise. Thus $A$ is a $0,1$ symmetric matrix with zeroes along the main diagonal, and we can apply the standard machinery of eigenvalues and eigenvectors of real symmetric matrices. It follows that all eigenvalues of $A$ (usually also called the eigenvalues of the graph $G$ itself) are real, and we denote them by $\lambda_1\ge\lambda_2\ge\ldots\ge \lambda_n$. Also, there is an orthonormal basis $B=\{x_1,\ldots,x_n\}$ of the euclidean space $R^n$ composed of eigenvectors of $A$: $Ax_i=\lambda_i x_i$, $x_i^tx_i=1$, $i=1,\ldots,n$. The matrix $A$ can be decomposed then as: $A=\sum_{i=1}^n\lambda_ix_ix_i^t$ -- the so called spectral decomposition of $A$. (Notice that the product $xx^t$, $x\in R^n$, is an $n$-by-$n$ matrix of rank 1; if $x,y,z\in R^n$ then $y^t(xx^t)z=(y^tx)(x^tz)$). Every vector $y\in R^n$ can be easily represented in basis $B$: $y=\sum_{i=1}^n(y^tx_i)x_i$. Therefore, for $y,z\in R^n$, $y^tz=\sum_{i=1}^n (y^tx_i)(z^tx_i)$ and $\|y\|^2=y^ty=\sum_{i=1}^n (y^tx_i)^2$. All the above applies in fact to all real symmetric matrices. Since the adjacency matrix $A$ of a graph $G$ is a matrix with non-negative entries, one can derive some important extra features of $A$, most notably the Perron-Frobenius Theorem, that reads in the graph context as follows: if $G$ is connected then the multiplicity of $\lambda_1$ is one, all coordinates of the first eigenvector $x_1$ can be assumed to be strictly positive, and $|\lambda_i|\le \lambda_1$ for all $i\ge 2$. Thus, graph spectrum lies entirely in the interval $[-\lambda_1,\lambda_1]$. For the most important special case of regular graphs Perron-Frobenius implies the following corollary: \begin{prop}\label{PF} Let $G$ be a $d$-regular graph on $n$ vertices. Let $\lambda_1\ge\lambda_2\ge\ldots\ge \lambda_n$ be the eigenvalues of $G$. Then $\lambda_1=d$ and $-d\le \lambda_i \le d$ for all $1\le i\le n$. Moreover, if $G$ is connected then the first eigenvector $x_1$ is proportional to the all one vector $(1,\ldots,1)^t\in R^n$, and $\lambda_i<d$ for all $i\ge 2$. \end{prop} To derive the above claim from the Perron-Frobenius Theorem observe that $e=(1,\ldots,1)$ is immediately seen to be an eigenvector of $A(G)$ corresponding to the eigenvalue $d$: $Ae=de$. The positivity of the coordinates of $e$ implies then that $e$ is not orthogonal to the first eigenvector, and hence is in fact proportional to $x_1$ of $A(G)$. Proposition \ref{PF} can be also proved directly without relying on the Perron-Frobenius Theorem. We remark that $\lambda_n=-d$ is possible, in fact it holds if and only if the graph $G$ is bipartite. All this background information, presented above in a somewhat condensed form, can be found in many textbooks in Linear Algebra. Readers more inclined to consult combinatorial books can find it for example in a recent monograph of Godsil and Royle on Algebraic Graph Theory \cite{GodRoy01}. We now prove a well known theorem (see its variant, e.g., in Chapter 9, \cite{AloSpe00}) bridging between graph spectra and edge distribution. \begin{theo}\label{eigen} Let $G$ be a $d$-regular graph on $n$ vertices. Let $d=\lambda_1\ge\lambda_2\ge\ldots\lambda_n$ be the eigenvalues of $G$. Denote $$ \lambda=max_{2\le i\le n}|\lambda_i|\,. $$ Then for every two subsets $U,W\subset V$, \begin{equation}\label{eig} \left|e(U,W)-\frac{d|U||W|}{n}\right|\le \lambda\sqrt{|U||W|\left(1-\frac{|U|}{n}\right) \left(1-\frac{|W|}{n}\right)}\ . \end{equation} \end{theo} \noindent{\bf Proof.\ } Let $B=\{x_1,\ldots,x_n\}$ be an orthonormal basis of $R^n$ composed from eigenvectors of $A$: $Ax_i=\lambda_ix_i$, $1\le i\le n$. We represent $A=\sum_{i=1}^n\lambda_ix_ix_i^t$. Denote \begin{eqnarray*} A_1&=&\lambda_1x_1x_1^t\,,\\ {\cal E}&=&\sum_{i=2}^n\lambda_ix_ix_i^t\,, \end{eqnarray*} then $A=A_1+{\cal E}$. Let $u=|U|$, $w=|W|$ be the cardinalities of $U,W$, respectively. We denote the characteristic vector of $U$ by $\chi_U\in R^n$, i.e. $\chi_U(i)=1$ if $i\in U$, and $\chi_U(i)=0$ otherwise. Similarly, let $\chi_W\in R^n$ be the characteristic vector of $W$. We represent $\chi_U$, $\chi_W$ according to $B$: \begin{eqnarray*} \chi_U &=& \sum_{i=1}^n \alpha_ix_i,\quad \alpha_i=\chi_U^tx_i, \quad \sum_{i=1}^n\alpha_i^2=\|\chi_U\|^2=u\,,\\ \chi_W &=& \sum_{i=1}^n \beta_ix_i, \quad \beta_i=\chi_W^tx_i, \quad \sum_{i=1}^n\beta_i^2=\|\chi_W\|^2=w\ . \end{eqnarray*} It follows easily from the definitions of $A$, $\chi_U$ and $\chi_W$ that the product $\chi_U^tA\chi_W$ counts exactly the number of edges of $G$ with one endpoint in $U$ and the other one in $W$, i.e. $$ e(U,W)=\chi_U^tA\chi_W\ =\chi_U^tA_1\chi_W+\chi_U^t{\cal E}\chi_W\ . $$ Now we estimate the last two summands separately, the first of them will be the main term for $e(U,W)$, the second one will be the error term. Substituting the expressions for $\chi_U$, $\chi_W$ and recalling the orthonormality of $B$, we get: \begin{equation}\label{fir} \chi_U^tA_1\chi_W=\left(\sum_{i=1}^n\alpha_ix_i\right)^t (\lambda_1x_1x_1^t) \left(\sum_{j=1}^n\beta_jx_j\right)= \sum_{i=1}^n\sum_{j=1}^n \alpha_i\lambda_1\beta_j (x_i^tx_1)(x_1^tx_j)=\alpha_1\beta_1\lambda_1\ . \end{equation} Similarly, \begin{equation}\label{rest} \chi_U^t{\cal E}\chi_W=\left(\sum_{i=1}^n\alpha_ix_i\right)^t \left(\sum_{j=2}^n\lambda_jx_jx_j^t\right) \left(\sum_{k=1}^n\beta_kx_k\right)= \sum_{i=2}^n\alpha_i\beta_i\lambda_i\ . \end{equation} Recall now that $G$ is $d$-regular. Then according to Proposition \ref{PF}, $\lambda_1=d$ and $x_1=\frac{1}{\sqrt{n}}(1,\ldots,1)^t$. We thus get: $\alpha_1=\chi_U^tx_1=u/\sqrt{n}$ and $\beta_1=\chi_W^tx_1=w/\sqrt{n}$. Hence it follows from (\ref{fir}) that $\chi_U^tA_1\chi_W=duw/n$. Now we estimate the absolute value of the error term $\chi_U^t{\cal E}\chi_W$. Recalling (\ref{rest}), the definition of $\lambda$ and the obtained values of $\alpha_1$, $\beta_1$, we derive, applying Cauchy-Schwartz: \begin{eqnarray*} |\chi_U^t{\cal E}\chi_W|&=&|\sum_{i=2}^n\alpha_i\beta_i\lambda_i| \le\lambda|\sum_{i=2}^n\alpha_i\beta_i|\le \lambda\sqrt{\sum_{i=2}^n\alpha_i^2\sum_{i=2}^n\beta_i^2}\\ &=&\lambda\sqrt{(\|\chi_U\|^2-\alpha_1^2)(\|\chi_W\|^2-\beta_1^2)}= \lambda\sqrt{\left(u-\frac{u^2}{n}\right)\left(w-\frac{w^2}{n}\right)} \ . \end{eqnarray*} The theorem follows.\hfill $\Box$ \bigskip The above proof can be extended to the irregular (general) case. Since the obtained quantitative bounds on edge distribution turn out to be somewhat cumbersome, we will just indicate how they can be obtained. Let $G=(V,E)$ be a graph on $n$ vertices with {\em average} degree $d$. Assume that the eigenvalues of $G$ satisfy $\lambda<d$, with $\lambda$ as defined in the theorem. Denote $$ K=\sum_{v\in V}(d(v)-d)^2\ . $$ The parameter $K$ is a measure of irregularity of $G$. Clearly $K=0$ if and only if $G$ is $d$-regular. Let $e=\frac{1}{\sqrt{n}}(1,\ldots,1)^t$. We represent $e$ in the basis $B=\{x_1,\ldots,x_n\}$ of the eigenvectors of $A(G)$: $$ e=\sum_{i=1}^n\gamma_ix_i,\quad \gamma_i=e^tx_i,\quad \sum_{i=1}^n\gamma_i^2=\|e\|^2=1\ . $$ Denote $z=\frac{1}{\sqrt{n}}(d(v_1)-d,\ldots,d(v_n)-d)^t$, then $\|z\|^2=K/n$. Notice that $Ae=\frac{1}{\sqrt{n}}(d(v_1),\ldots,d(v_n))^t$ $=de+z$, and therefore $z=Ae-de=\sum_{i=1}^n\gamma_i(\lambda_i-d)x_i$. This implies: \begin{eqnarray*} \frac{K}{n}&=&\|z\|^2= \sum_{i=1}^n\gamma_i^2(\lambda_i-d)^2\ge \sum_{i=2}^n\gamma_i^2(\lambda_i-d)^2\\ &\ge& (d-\lambda)^2\sum_{i=2}^n\gamma_i^2\ . \end{eqnarray*} Hence $\sum_{i=2}^n\gamma_i^2\le \frac{K}{n(d-\lambda)^2}$. It follows that $\gamma_1^2=1-\sum_{i=2}^n\gamma_i^2\ge 1-\frac{K}{n(d-\lambda)^2}$ and $$ \gamma_1\ge \gamma_1^2 \ge 1-\frac{K}{n(d-\lambda)^2}\ . $$ Now we estimate the distance between the vectors $e$ and $x_1$ and show that they are close given that the parameter $K$ is small. \begin{eqnarray*} \|e-x_1\|^2&=& (e-x_1)^t(e-x_1)=e^te+x_1^tx_1-2e^tx_1=1+1-2\gamma_1 =2-2\gamma_1\\ &\le& \frac{2K}{n(d-\lambda)^2}\ . \end{eqnarray*} We now return to expressions (\ref{fir}) and (\ref{rest}) from the proof of Theorem \ref{eigen}. In order to estimate the main term $\chi_U^tA_1\chi_W$, we bound the coefficients $\alpha_1$, $\beta_1$ and $\lambda_1$ as follows: $$ \alpha_1=\chi_U^tx_1=\chi_U^te+\chi_U^t(x_1-e)= \frac{u}{\sqrt{n}}+\chi_U^t(x_1-e)\ , $$ and therefore \begin{equation}\label{alpha1} \left|\alpha_1-\frac{u}{\sqrt{n}}\right|=|\chi_U^t(x_1-e)| \le \|\chi_U||\cdot\|x_1-e\|\le \frac{\sqrt{\frac{2Ku}{n}}}{d-\lambda} \ . \end{equation} In a similar way one gets: \begin{equation}\label{beta1} \left|\beta_1-\frac{w}{\sqrt{n}}\right|\le \frac{\sqrt{\frac{2Kw}{n}}}{d-\lambda} \ . \end{equation} Finally, to estimate from above the absolute value of the difference between $\lambda_1$ and $d$ we argue as follows: $$ \frac{K}{n}=\|z\|^2=\sum_{i=1}^n\gamma_i^2(\lambda_i-d)^2\ge \gamma_1^2(\lambda_1-d)^2\,, $$ and therefore \begin{equation}\label{lambda1} |\lambda_1-d|\le \frac{1}{\gamma_1}\sqrt{\frac{K}{n}} \le \frac{n(d-\lambda)^2}{n(d-\lambda)^2-K} \sqrt{\frac{K}{n}}\ . \end{equation} Summarizing, we see from (\ref{alpha1}), (\ref{beta1}) and (\ref{lambda1}) that the main term in the product $\chi_U^tA_1\chi_W$ is equal to $\frac{duw}{n}$, just as in the regular case, and the error term is governed by the parameter $K$. In order to estimate the error term $\chi_U^t{\cal E} \chi_W$ we use (\ref{rest}) to get: \begin{eqnarray*} \hspace{3.1cm} |\chi_U^t{\cal E}\chi_W|&=& \left|\sum_{i=2}^n\alpha_i\beta_i\lambda_i\right| \le \lambda\left|\sum_{i=2}^n\alpha_i\beta_i\right| \le \lambda \sqrt{\sum_{i=2}^n\alpha_i^2\sum_{i=2}^n\beta_i^2}\\ &\le& \lambda \sqrt{\sum_{i=1}^n\alpha_i^2\sum_{i=1}^n\beta_i^2} = \lambda \|\chi_U\|\,\|\chi_W\|=\lambda\sqrt{uw}. \hspace{3.1cm} \Box \end{eqnarray*} Applying the above developed techniques we can prove now the second implication of Theorem \ref{CG1}. Let us prove first that $EIG$ implies $K=o(nd^2)$, where $d=(1+o(1))np$ is as before the average degree of $G$. Indeed, for every vector $v\in R^n$ we have $\|Av\|\le \lambda_1\|v\|$, and therefore $$ \lambda_1^2n=\lambda_1^2e^te\ge (Ae)^t(Ae)=\sum_{v\in V}d^2(v)\ . $$ Hence from $EIG$ we get: $\sum_{v\in V}d^2(v)\le (1+o(1))nd^2$. As $\sum_{v}d(v)=nd$, it follows that: $$ K=\sum_{v\in V}(d(v)-d)^2=\sum_{v\in V}d^2(v)-2d\sum_{v\in V}d(v)+nd^2 = (1+o(1))nd^2-2nd^2+nd^2=o(nd^2)\,, $$ as promised. Substituting this into estimates (\ref{alpha1}), (\ref{beta1}), (\ref{lambda1}) and using $\lambda=o(d)$ of $EIG$ we get: \begin{eqnarray*} \alpha_1 &=& \frac{u}{\sqrt{n}}+o(\sqrt{u})\,,\\ \beta_1 &=& \frac{w}{\sqrt{n}}+o(\sqrt{w})\,,\\ \lambda_1 &=& (1+o(1))d\,, \end{eqnarray*} and therefore $$ \chi_U^tA_1\chi_W=\frac{duw}{n}+o(dn)\ . $$ Also, according to $EIG$, $\lambda=o(d)$, which implies: $$ \chi_U^t{\cal E}\chi_w=o(d\sqrt{uw})=o(dn)\,, $$ and the claim follows. \hfill $\Box$ Theorem \ref{eigen} is a truly remarkable result. Not only it connects between two seemingly unrelated graph characteristics -- edge distribution and spectrum, it also provides a very good quantitative handle for the uniformity of edge distribution, based on easily computable, both theoretically and practically, graph parameters -- graph eigenvalues. According to the bound (\ref{eig}), a polynomial number of parameters can control quite well the number of edges in exponentially many subsets of vertices. The parameter $\lambda$ in the formulation of Theorem \ref{eigen} is usually called the {\em second eigenvalue} of the $d$-regular graph $G$ (the first and the trivial one being $\lambda_1=d$). There is certain inaccuracy though in this term, as in fact $\lambda=\max\{\lambda_2,-\lambda_n\}$. Later we will call, following Alon, a $d$-regular graph $G$ on $n$ vertices in which all eigenvalues, but the first one, are at most $\lambda$ in their absolute values, an {\em $(n,d,\lambda)$-graph}. Comparing (\ref{eig}) with the definition of jumbled graphs by Thomason we see that an $(n,d,\lambda)$-graph $G$ is $(d/n,\lambda)$-jumbled. Hence the parameter $\lambda$ (or in other words, the so called {\em spectral gap} -- the difference between $d$ and $\lambda$) is responsible for pseudo-random properties of such a graph. The smaller the value of $\lambda$ compared to $d$, the more close is the edge distribution of $G$ to the ideal uniform distribution. A natural question is then: how small can be $\lambda$? It is easy to see that as long as $d\le (1-\epsilon)n$, $\lambda=\Omega(\sqrt{d})$. Indeed, the trace of $A^2$ satisfies: $$ nd=2|E(G)|=Tr(A^2)=\sum_{i=1}^n\lambda_i^2\le d^2+(n-1)\lambda_2 \le (1-\epsilon)nd+(n-1)\lambda^2\,, $$ and $\lambda=\Omega(\sqrt{d})$ as claimed. More accurate bounds are known for smaller values of $d$ (see, e.g. \cite{Nil91}). Based on these estimates we can say that an $(n,d,\lambda)$-graph $G$, for which $\lambda=\Theta(\sqrt{d})$, is a very good pseudo-random graph. We will see several examples of such graphs in the next section. \subsection{Strongly regular graphs} A {\em strongly regular graph} $srg(n,d,\eta,\mu)$ is a $d$-regular graph on $n$ vertices in which every pair of adjacent vertices has exactly $\eta$ common neighbors and every pair of non-adjacent vertices has exactly $\mu$ common neighbors. (We changed the very standard notation in the above definition so as to avoid interference with other notational conventions throughout this paper and to make it more coherent, usually the parameters are denoted $(v,k,\lambda,\mu)$). Two simple examples of strongly regular graph are the pentagon $C_5$ that has parameters $(5,2,0,1)$, and the Petersen graph whose parameters are $(10,3,0,1)$. Strongly regular graphs were introduced by Bose in 1963 \cite{Bos63} who also pointed out their tight connections with finite geometries. As follows from the definition, strongly regular graphs are highly regular structures, and one can safely predict that algebraic methods are extremely useful in their study. We do not intend to provide any systematic coverage of this fascinating concept here, addressing the reader to the vast literature on the subject instead (see, e.g., \cite{BrovLi84}). Our aim here is to calculate the eigenvalues of strongly regular graphs and then to connect them with pseudo-randomness, relying on results from the previous subsection. \begin{prop}\label{srg} Let $G$ be a connected strongly regular graph with parameters $(n,d,\eta,\mu)$. Then the eigenvalues of $G$ are: $\lambda_1=d$ with multiplicity $s_1=1$, $$ \lambda_2=\frac{1}{2}\left(\eta-\mu+\sqrt{(\eta-\mu)^2+4(d-\mu)}\right) $$ and $$ \lambda_3=\frac{1}{2}\left(\eta-\mu-\sqrt{(\eta-\mu)^2+4(d-\mu)}\right) \,, $$ with multiplicities $$ s_2=\frac{1}{2}\left(n-1+\frac{(n-1)(\mu-\eta)-2d} {\sqrt{(\mu-\eta)^2+4(d-\mu)}}\right) $$ and $$ s_3=\frac{1}{2}\left(n-1-\frac{(n-1)(\mu-\eta)-2d} {\sqrt{(\mu-\eta)^2+4(d-\mu)}}\right)\,, $$ respectively. \end{prop} \noindent{\bf Proof.\ } Let $A$ be the adjacency matrix of $A$. By the definition of $A$ and the fact that $A$ is symmetric with zeroes on the main diagonal, the $(i,j)$-entry of the square $A^2$ counts the number of common neighbors of $v_i$ and $v_j$ in $G$ if $i\ne j$, and is equal to the degree $d(v_i)$ in case $i=j$. The statement that $G$ is $srg(n,d,\eta,\mu)$ is equivalent then to: \begin{equation}\label{srg1} AJ=dJ,\quad\quad A^2=(d-\mu)I+\mu J+(\eta-\mu)A\ , \end{equation} where $J$ is the $n$-by-$n$ all-one matrix and $I$ is the $n$-by-$n$ identity matrix. Since $G$ is $d$-regular and connected, we obtain from the Perron-Frobenius Theorem that $\lambda_1=d$ is an eigenvalue of $G$ with multiplicity 1 and with $e=(1,\ldots,1)^t$ as the corresponding eigenvector. Let $\lambda\ne d$ be another eigenvalue of $G$, and let $x\in R^n$ be a corresponding eigenvector. Then $x$ is orthogonal to $e$, and therefore $Jx=0$. Applying both sides of the second identity in (\ref{srg1}) to $x$ we get the equation: $\lambda^2x=(d-\mu)x+(\eta-\mu)\lambda x$, which results in the following quadratic equation for $\lambda$: $$ \lambda^2+(\mu-\eta)\lambda+(\mu-d)=0\ . $$ This equation has two solutions $\lambda_2$ and $\lambda_3$ as defined in the proposition formulation. If we denote by $s_2$ and $s_3$ the respective multiplicities of $\lambda_2$ and $\lambda_3$ as eigenvalues of $A$, we get: $$ 1+s_2+s_3=n,\quad\quad Tr(A)=d+s_2\lambda_2+s_3\lambda_3=0\ . $$ Solving the above system of linear equations for $s_2$ and $s_3$ we obtain the assertion of the proposition.\hfill $\Box$ \medskip Using the bound (\ref{eig}) we can derive from the above proposition that if the parameters of a strongly regular graph $G$ satisfy $\eta\approx \mu$ then $G$ has a large eigenvalue gap and is therefore a good pseudo-random graph. We will exhibit several examples of such graphs in the next section. \section{Examples}\label{examples} Here we present some examples of pseudo-random graphs. Many of them are well known and already appeared, e.g., in \cite{Tho87a} and \cite{Tho87b}, but there also some which have been discovered only recently. Since in the rest of the paper we will mostly discuss properties of $(n,d,\lambda)$-graphs, in our examples we emphasize the spectral properties of the constructed graphs. We will also use most of these constructions later to illustrate particular points and to test the strength of the theorems. \noindent {\bf Random graphs} \begin{enumerate} \item Let $G=G(n,p)$ be a random graph with edge probability $p$. If $p$ satisfies $pn/\log n \rightarrow \infty$ and $(1-p)n\log n \rightarrow \infty$, then almost surely all the degrees of $G$ are equal to $(1+o(1))np$. Moreover it was proved by F\"uredi and Koml\'os \cite{FK} that the largest eigenvalue of $G$ is a.s. $(1+o(1))np$ and that $\lambda(G) \leq (2+o(1))\sqrt{p(1-p)n}$. They stated this result only for constant $p$ but their proof shows that $\lambda(G) \leq O(\sqrt{np})$ also when $p\geq poly \log n/n$. \item For a positive integer-valued function $d=d(n)$ we define the model $G_{n,d}$ of random regular graphs consisting of all regular graphs on $n$ vertices of degree $d$ with the uniform probability distribution. This definition of a random regular graph is conceptually simple, but it is not easy to use. Fortunately, for small $d$ there is an efficient way to generate $G_{n,d}$ which is useful for theoretical studies. This is the so called {\em configuration model}. For more details about this model, and random regular graphs in general we refer the interested reader to two excellent monographs \cite{Bol01} and \cite{JanLucRuc00}, or to a survey \cite{Wor99}. As it turns out, sparse random regular graphs have quite different properties from those of the binomial random graph $G(n,p), p=d/n$. For example, they are almost surely connected. The spectrum of $G_{n,d}$ for a fixed $d$ was studied in \cite{FKS} by Friedman, Kahn and Szemer\'edi. Friedman \cite{FRI} proved that for constant $d$ the second largest eigenvalue of a random $d$-regular graph is $\lambda = (1+o(1))2\sqrt{d-1}$. The approach of Kahn and Szemer\'edi gives only $O(\sqrt{d})$ bound on $\lambda$ but continues to work also when $d$ is small power of $n$. The case $d \gg n^{1/2}$ was recently studied by Krivelevich, Sudakov, Vu and Wormald \cite{KSVW}. They proved that in this case for any two vertices $u,v \in G_{n,d}$ almost surely $$\big|codeg(u,v)-d^2/n\big| < Cd^3/n^2 + 6d\sqrt{ \log n} /\sqrt{n},$$ where $C$ is some constant and $codeg(u,v)$ is the number of common neighbors of $u,v$. Moreover if $d \geq n/\log n$, then $C$ can be defined to be zero. Using this it is easy to show that for $d \gg n^{1/2}$, the second largest eigenvalue of a random $d$-regular graph is $o(d)$. The true bound for the second largest eigenvalue of $G_{n,d}$ should be probably $(1+o(1))2\sqrt{d-1}$ for all values of $d$, but we are still far from proving it. \noindent \hspace{-0.95cm}{\bf Strongly regular graphs} \item Let $q=p^{\alpha}$ be a prime power which is congruent to $1$ modulo $4$ so that $-1$ is a square in the finite field $GF(q)$. Let $P_q$ be the graph whose vertices are all elements of $GF(q)$ and two vertices are adjacent if and only if their difference is a quadratic residue in $GF(q)$. This graph is usually called the {\em Paley graph}. It is easy to see that $P_q$ is $(q-1)/2$-regular. In addition one can easily compute the number of common neighbors of two vertices in $P_q$. Let $\chi$ be the {\em quadratic residue character} on $GF(q)$, i.e., $\chi(0)=0$, $\chi(x)=1$ if $x\not= 0$ and is a square in $GF(q)$ and $\chi(x)=-1$ otherwise. By definition, $\sum_x\chi(x)=0$ and the number of common neighbors of two vertices $a$ and $b$ equals $$\sum_{x\not=a,b}\left(\frac{1+\chi(a-x)}{2}\right) \left(\frac{1+\chi(b-x)}{2}\right)= \frac{q-2}{4}-\frac{\chi(a-b)}{2}+\frac{1}{4}\sum_{x\not=a,b} \chi(a-x)\chi(b-x).$$ Using that for $x \not =b$, $\chi(b-x)=\chi\big((b-x)^{-1}\big)$, the last term can be rewritten as $$\sum_{x\not=a,b}\chi(a-x)\chi\big((b-x)^{-1}\big)= \sum_{x\not=a,b} \chi\Big(\frac{a-x}{b-x}\Big)= \sum_{x\not=a,b}\chi\Big(1+\frac{a-b}{b-x}\Big)=\sum_{x\not=0,1}\chi(x)=-1.$$ Thus the number of common neighbors of $a$ and $b$ is $(q-3)/4-\chi(a-b)/2$. This equals $(q-5)/4$ if $a$ and $b$ are adjacent and $(q-1)/4$ otherwise. This implies that the Paley graph is a strongly regular graph with parameters $\big(q, (q-1)/2, (q-5)/4, (q-1)/4\big)$ and therefore its second largest eigenvalue equals $(\sqrt{q}+1)/2$. \item For any odd integer $k$ let $H_k$ denote the graph whose $n_k=2^{k-1}-1$ vertices are all binary vectors of length $k$ with an odd number of ones except the all one vector, in which two distinct vertices are adjacent iff the inner product of the corresponding vectors is $1$ modulo $2$. Using elementary linear algebra it is easy to check that this graph is $(2^{k-2}-2)$-regular. Also every two nonadjacent vertices vertices in it have $2^{k-3}-1$ common neighbors and every two adjacent vertices vertices have $2^{k-3}-3$ common neighbors. Thus $H_k$ is a strongly regular graph with parameters $\big(2^{k-1}-1, 2^{k-2}-2, 2^{k-3}-3, 2^{k-3}-1\big)$ and with the second largest eigenvalue $\lambda(H_k)=1+2^{\frac{k-3}{2}}$. \item Let $q$ be a prime power an let $V(G)$ be the elements of the two dimensional vector space over $GF(q)$, so $G$ has $q^2$ vertices. Partition the $q+1$ lines through the origin of the space into two sets $P$ and $N$, where $|P|=k$. Two vertices $x$ and $y$ of the graph $G$ are adjacent if $x-y$ is parallel to a line in $P$. This example is due to Delsarte and Goethals and to Turyn (see \cite {Sei}). It is easy to check that $G$ is strongly regular with parameters $\big(k(q-1),(k-1)(k-2)+q-2,k(k-1)\big)$. Therefore its eigenvalues, besides the trivial one are $-k$ and $q-k$. Thus if $k$ is sufficiently large we obtain that $G$ is $d=k(q-1)$-regular graph whose second largest eigenvalue is much smaller than $d$. \noindent \hspace{-0.95cm}{\bf Graphs arising from finite geometries} \item For any integer $t \geq 2$ and for any power $q=p^{\alpha}$ of prime $p$ let $PG(q,t)$ denote the projective geometry of dimension $t$ over the finite field $GF(q)$. The interesting case for our purposes here is that of large $q$ and fixed $t$. The vertices of $PG(q,t)$ correspond to the equivalence classes of the set of all non-zero vectors ${\bf x}=(x_0, \ldots, x_t)$ of length $t+1$ over $GF(q)$, where two vectors are equivalent if one is a multiple of the other by an element of the field. Let $G$ denote the graph whose vertices are the points of $PG(q,t)$ and two (not necessarily distinct) vertices ${\bf x}$ and ${\bf y}$ are adjacent if and only if $x_0y_0+\ldots+x_ty_t=0$. This construction is well known. In particular, in case $t=2$ this graph is often called the Erd\H os-R\'enyi graph and it contains no cycles of length $4$. It is easy to see that the number of vertices of $G$ is $n_{q,t}=\big(q^{t+1}-1\big)/\big(q-1\big)=\big(1+o(1)\big)q^t$ and that it is $d_{q,t}$-regular for $d_{q,t}=\big(q^t-1\big)/\big(q-1\big)=\big(1+o(1)\big)q^{t-1}$, where $o(1)$ tends to zero as $q$ tends to infinity. It is easy to see that the number of vertices of $G$ with loops is bounded by $2\big(q^{t}-1\big)/\big(q-1\big)=\big(2+o(1)\big)q^{t-1}$, since for every possible value of $x_0, \ldots, x_{t-1}$ we have at most two possible choices of $x_t$. Actually using more complicated computation, which we omit, one can determine the exact number of vertices with loops. The eigenvalues of $G$ are easy to compute (see \cite{AK}). Indeed, let $A$ be the adjacency matrix of $G$. Then, by the properties of $PG(q,t)$, $A^2=AA^T=\mu J+(d_{q,t}-\mu)I$, where $\mu=\big(q^{t-1}-1\big)/\big(q-1\big)$, $J$ is the all one matrix and $I$ is the identity matrix, both of size $n_{q,t} \times n_{q,t}$. Therefore the largest eigenvalue of $A$ is $d_{q,t}$ and the absolute value of all other eigenvalues is $\sqrt{d_{q,t}-\mu}=q^{(t-1)/2}$. \item The generalized polygons are incidence structures consisting of points $\cal P$ and lines $\cal L$. For our purposes we restrict our attention to those in which every point is incident to $q+1$ lines and every line is incident to $q+1$ points. A generalized $m$-gon defines a bipartite graph $G$ with bipartition $({\cal P},{\cal L})$ that satisfies the following conditions. The diameter of $G$ is $m$ and for every vertex $v \in G$ there is a vertex $u \in G$ such that the shortest path from $u$ to $v$ has length $m$. Also for every $r< m$ and for every two vertices $u, v$ at distance $r$ there exists a unique path of length $r$ connecting them. This immediately implies that every cycle in $G$ has length at least $2m$. For $q \geq 2$, it was proved by Feit and Higman \cite{FH} that $(q+1)$-regular generalized $m$-gons exist only for $m=3,4,6$. A {\em polarity} of $G$ is a bijection $\pi: {\cal P} \cup {\cal L} \rightarrow {\cal P} \cup {\cal L}$ such that $\pi({\cal P})={\cal L}$, $\pi({\cal L})={\cal P}$ and $\pi^2$ is the identity map. Also for every $p \in {\cal P}, l \in {\cal L}$, $\pi(p)$ is adjacent to $\pi(l)$ if and only if $p$ and $l$ are adjacent. Given $\pi$ we define a polarity graph $G^{\pi}$ to be the graph whose vertices are point in $\cal P$ and two (not necessarily distinct) points $p_1, p_2$ are adjacent iff $p_1$ was adjacent to $\pi(p_2)$ in $G$. Some properties of $G^{\pi}$ can be easily deduced from the corresponding properties of $G$. In particular, $G^{\pi}$ is $(q+1)$-regular and also contains no even cycles of length less than $2m$. For every $q$ which is an odd power of $2$, the incidence graph of the generalized $4$-gon has a polarity. The corresponding polarity graph is a $(q+1)$-regular graph with $q^3+q^2+q+1$ vertices. See \cite{BCN}, \cite{LUW} for more details. This graph contains no cycle of length $6$ and it is not difficult to compute its eigenvalues (they can be derived, for example, from the eigenvalues of the corresponding bipartite incidence graph, given in \cite{Ta}). Indeed, all the eigenvalues, besides the trivial one (which is $q+1$) are either $0$ or $\sqrt {2q}$ or $-\sqrt {2q}$. Similarly, for every $q$ which is an odd power of $3$, the incidence graph of the generalized $6$-gon has a polarity. The corresponding polarity graph is a $(q+1)$-regular graph with $q^5+q^4+ \cdots +q+1$ vertices ( see again \cite{BCN}, \cite{LUW}). This graph contains no cycle of length $10$ and its eigenvalues can be derived using the same technique as in case of the $4$-gon. All these eigenvalues, besides the trivial one are either $\sqrt {3q}$ or $-\sqrt{3q}$ or $\sqrt {q}$ or $-\sqrt {q}$. \noindent \hspace{-0.95cm}{\bf Cayley graphs} \item Let $G$ be a finite group and let $S$ be a set of non-identity elements of $G$ such that $S=S^{-1}$, i.e., for every $s \in S$, $s^{-1}$ also belongs to $S$. The {\em Cayley graph} $\Gamma(G,S)$ of this group with respect to the generating set $S$ is the graph whose set of vertices is $G$ and where two vertices $g$ and $g'$ are adjacent if and only if $g'g^{-1} \in S$. Clearly, $\Gamma(G,S)$ is $|S|$-regular and it is connected iff $S$ is a set of generators of the group. If $G$ is abelian then the eigenvalues of the Cayley graph can be computed in terms of the characters of $G$. Indeed, let $\chi: G \rightarrow C$ be a character of $G$ and let $A$ be the adjacency matrix of $\Gamma(G,S)$ whose rows and columns are indexed by the elements of $G$. Consider the vector ${\bf v}$ defined by ${\bf v}(g)=\chi(g)$. Then it is easy to check that $A{\bf v}=\alpha {\bf v}$ with $\alpha=\sum_{s\in S}\chi(s)$. In addition all eigenvalues can be obtained in this way, since every abelian group has exactly $|G|$ different character which are orthogonal to each other. Using this fact, one can often give estimates on the eigenvalues of $\Gamma(G,S)$ for abelian groups. One example of a Cayley graph that has already been described earlier is $P_q$. In that case the group is the additive group of the finite field $GF(q)$ and $S$ is the set of all quadratic residues modulo $q$. Next we present a slightly more general construction. Let $q=2kr+1$ be a prime power and let $\Gamma$ be a Cayley graph whose group is the additive group of $GF(q)$ and whose generating set is $S=\big\{x=y^k~|~\mbox{for some}~y \in GF(q)\big\}$. By definition, $\Gamma$ is $(q-1)/k$-regular. On the other hand, this graph is not strongly regular unless $k=2$, when it is the Paley graph. Let $\chi$ be a nontrivial additive character of $GF(q)$ and consider the Gauss sum $\sum_{y \in GF(q)} \chi(y^k)$. Using the classical bound $|\sum_{y \in GF(q)} \chi(y^k)|\leq (k-1)q^{1/2}$ (see e.g. \cite{LN}) and the above connection between characters and eigenvalues we can conclude that the second largest eigenvalue of our graph $\Gamma$ is bounded by $O(q^{1/2})$. \item Next we present a surprising construction obtained by Alon \cite{A94} of a very dense pseudo-random graph that on the other hand is triangle-free. For a positive integer $k$, consider the finite field $GF(2^k)$, whose elements are represented by binary vectors of length $k$. If $a, b, c$ are three such vectors, denote by $(a,b,c)$ the binary vector of length $3k$ whose coordinates are those of $a$, followed by coordinates of $b$ and then $c$. Suppose that $k$ is not divisible by $3$. Let $W_0$ be the set of all nonzero elements $\alpha \in GF(2^k)$ so that the leftmost bit in the binary representation of $\alpha^7$ is $0$, and let $W_1$ be the set of all nonzero elements $\alpha \in GF(2^k)$ for which the leftmost bit of $\alpha^7$ is $1$. Since $3$ does not divide $k$, $7$ does not divide $2^k-1$ and hence $|W_0|=2^{k-1}-1$ and $|W_1|=2^{k-1}$, as when $\alpha$ ranges over all nonzero elements of the field so does $\alpha^7$. Let $G_n$ be the graph whose vertices are all $n=2^{3k}$ binary vectors of length $3k$, where two vectors ${\bf v}$ and ${\bf v}'$ are adjacent if and only if there exist $w_0\in W_0$ and $w_1\in W_1$ so that ${\bf v}-{\bf v}'=(w_0,w_0^3,w_0^5)+(w_1,w_1^3,w_1^5)$, where here powers are computed in the field $GF(2^k)$ and the addition is addition modulo $2$. Note that $G_n$ is the Cayley graph of the additive group ${\bf Z}_2^{3k}$ with respect to the generating set $S=U_0+U_1$, where $U_0=\big\{(w_0,w_0^3,w_0^5)~|~w_0\in W_0\big\}$ and $U_1$ is defined similarly. A well known fact from Coding Theory (see e.g., \cite{MS}), which can be proved using the Vandermonde determinant, is that every set of six distinct vectors in $U_0 \cup U_1$ is linearly independent over $GF(2)$. In particular all the vectors in $U_0+U_1$ are distinct, $S=|U_0||U_1|$ and hence $G_n$ is $|S|=2^{k-1}(2^{k-1}-1)$-regular. The statement that $G_n$ is triangle free is clearly equivalent to the fact that the sum modulo $2$ of any set of $3$ nonzero elements of $S$ is not a zero-vector. Let $u_0+u_1, u'_0+u'_1$ and $u''_0+u''_1$ be three distinct element of $S$, where $u_0,u'_0,u''_0 \in U_0$ and $u_1,u'_1,u''_1 \in U_1$. By the above discussion, if the sum of these six vectors is zero, then every vector must appear an even number of times in the sequence $(u_0,u'_0,u''_0,u_1,u'_1,u''_1)$. However, since $U_0$ and $U_1$ are disjoint, this is clearly impossible. Finally, as we already mentioned, the eigenvalues of $G_n$ can be computed in terms of characters of ${\bf Z}_2^{3k}$. Using this fact together with the Carlitz-Uchiyama bound on the characters of ${\bf Z}_2^{3k}$ it was proved in \cite{A94} that the second eigenvalue of $G_n$ is bounded by $\lambda \leq 9\cdot 2^k+3\cdot 2^{k/2}+1/4$. \item The construction above can be extended in the obvious way as mentioned in \cite{ALONK}. Let $h\geq 1$ and suppose that $k$ is an integer such that $2^k-1$ is not divisible by $4h+3$. Let $W_0$ be the set of all nonzero elements $\alpha \in GF(2^k)$ so that the leftmost bit in the binary representation of $\alpha^{4h+3}$ is $0$, and let $W_1$ be the set of all nonzero elements $\alpha \in GF(2^k)$ for which the leftmost bit of $\alpha^{4h+3}$ is $1$. Since $4h+3$ does not divide $2^k-1$ we have that $|W_0|=2^{k-1}-1$ and $|W_1|=2^{k-1}$, as when $\alpha$ ranges over all nonzero elements of the field so does $\alpha^{4h+3}$. Define $G$ to be the Cayley graph of the additive group ${\bf Z}_2^{(2h+1)k}$ with respect to the generating set $S=U_0+U_1$, where $U_0=\big\{(w_0,w_0^3,\ldots,w_0^{4h+1})~|~w_0\in W_0\big\}$ and $U_1$ is defined similarly. Clearly, $G$ is a $2^{k-1}(2^{k-1}-1)$-regular graph on $2^{(2h+1)k}$ vertices. Using methods from \cite{A94}, one can show that $G$ contains no odd cycle of length $\leq 2h+1$ and that the second eigenvalue of $G$ is bounded by $O(2^k)$. \item Now we describe the celebrated expander graphs constructed by Lubotzky, Phillips and Sarnak \cite{LPS} and independently by Margulis \cite{Margulis}. Let $p$ and $q$ be unequal primes, both congruent to $1$ modulo $4$ and such that $p$ is a quadratic residue modulo $q$. As usual denote by $PSL(2,q)$ the factor group of the group of two by two matrices over $GF(q)$ with determinant $1$ modulo its normal subgroup consisting of the two scalar matrices $\bigg(\begin{array}{cc}1&0\\0&1\end{array}\bigg)$ and $\bigg(\begin{array}{cc}-1&0\\0&-1\end{array}\bigg)$. The graphs we describe are Cayley graphs of $PSL(2,q)$. A well known theorem of Jacobi asserts that the number of ways to represent a positive integer $n$ as a sum of $4$ squares is $8\sum_{4 \not \, |\, d, d|n}d$. This easily implies that there are precisely $p+1$ vectors ${\bf a}=(a_0, a_1, a_2, a_3)$, where $a_0$ is an odd positive integer, $a_1, a_2, a_3$ are even integers and $a_0^2+ a_1^2+a_2^2+a_3^2=p$. From each such vector construct the matrix $M_a$ in $PSL(2,q)$ where $M_a=\frac{1}{\sqrt{p}} \bigg(\begin{array}{cc}a_0+ia_1&a_2+ia_3\\-a_2+ia_3&a_0-ia_1\end{array}\bigg)$ and $i$ is an integer satisfying $i^2=-1(\mbox{mod}~ q)$. Note that, indeed, the determinant of $M_a$ is $1$ and that the square root of $p$ modulo $q$ does exist. Let $G^{p,q}$ denote the Cayley graph of $PSL(2,q)$ with respect to these $p+1$ matrices. In \cite{LPS} it was proved that if $q >2\sqrt{p}$ then $G^{p,q}$ is a connected $(p+1)$-regular graph on $n=q(q^2-1)/2$ vertices. Its girth is at least $2\log_p q$ and all the eigenvalues of its adjacency matrix, besides the trivial one $\lambda_1=p+1$, are at most $2 \sqrt{p}$ in absolute value. The bound on the eigenvalues was obtained by applying deep results of Eichler and Igusa concerning the Ramanujan conjecture. The graphs $G^{p,q}$ have very good expansion properties and have numerous applications in Combinatorics and Theoretical Computer Science. \item The {\em projective norm graphs} $NG_{p,t}$ have been constructed in \cite{ARS}, modifying an earlier construction given in \cite{KRS}. These graphs {\bf are not} Cayley graphs, but as one will immediately see, their construction has a similar flavor. The construction is the following. Let $t>2$ be an integer, let $p$ be a prime, let $GF(p)^*$ be the multiplicative group of the field with $p$ elements and let $GF(p^{t-1})$ be the field with $p^{t-1}$ elements. The set of vertices of the graph $NG_{p,t}$ is the set $V=GF(p^{t-1})\times GF(p)^*$. Two distinct vertices $(X,a)$ and $(Y,b)\in V$ are adjacent if and only if $N(X+Y)=ab$, where the norm $N$ is understood over $GF(p)$, that is, $N(X)=X^{1+p+\cdots+p^{t-2}}.$ Note that $|V|=p^t-p^{t-1}$. If $(X,a)$ and $(Y,b)$ are adjacent, then $(X,a)$ and $Y\neq -X$ determine $b$. Thus $NG_{p,t}$ is a regular graph of degree $p^{t-1}-1$. In addition, it was proved in \cite{ARS}, that $NG_{p,t}$ contains no complete bipartite graphs $K_{t,(t-1)!+1}$. These graphs can be also defined in the same manner starting with a prime power instead of the prime $p$. It is also not difficult to compute the eigenvalues of this graph. Indeed, put $q=p^{t-1}$ and let $A$ be the adjacency matrix of $NG_{p,t}$. The rows and columns of this matrix are indexed by the ordered pairs of the set $GF(q) \times GF(p)^*$. Let $\psi$ be a character of the additive group of $GF(q)$, and let $\chi$ be a character of the multiplicative group of $GF(p)$. Consider the vector ${\bf v}: GF(q) \times GF(p)^* \mapsto C$ defined by ${\bf v}(X,a)=\psi(X) \chi(a)$. Now one can check (see \cite{AR}, \cite{Sz} for more details) that the vector ${\bf v}$ is an eigenvector of $A^2$ with eigenvalue $\big| \sum_{Z\in GF(q), Z\not=0} \psi(Z)\chi(N(Z))\big|^2$ and that all eigenvalues of $A^2$ have this form. Set $\chi'(Z)=\chi(N(Z))$ for all nonzero $Z$ in $GF(q)$. Note that as the norm is multiplicative, $\chi'$ is a multiplicative character of the large field. Hence the above expression is a square of the absolute value of the Gauss sum and it is well known (see e.g. \cite{Da}, \cite{Bol01}) that the value of each such square, besides the trivial one (that is, when either $\psi$ or $\chi'$ are trivial), is $q$. This implies that the second largest eigenvalue of $NG_{p,t}$ is $\sqrt{q}=p^{(t-1)/2}$. \end{enumerate} \section{Properties of pseudo-random graphs} We now examine closely properties of pseudo-random graphs, with a special emphasis on $(n,d,\lambda)$-graphs. The majority of them are obtained using the estimate (\ref{eig}) of Theorem \ref{eigen}, showing again the extreme importance and applicability of the latter result. It is instructive to compare the properties of pseudo-random graphs, considered below, with the analogous properties of random graphs, usually shown to hold by completely different methods. The set of properties we chose to treat here is not meant to be comprehensive or systematic, but quite a few rather diverse graph parameters will be covered. \subsection{Connectivity and perfect matchings} The {\em vertex-connectivity} of a graph $G$ is the minimum number of vertices that we need to delete to make $G$ disconnected. We denote this parameter by $\kappa(G)$. For random graphs it is well known (see, e.g., \cite{Bol01}) that the vertex-connectivity is almost surely the same as the minimum degree. Recently it was also proved (see \cite{KSVW} and \cite{CFR}) that random $d$-regular graphs are $d$-vertex-connected. For $(n,d,\lambda)$-graphs it is easy to show the following. \begin{theo} \label{connectivity} Let $G$ be an $(n,d,\lambda)$-graph with $d \leq n/2$. Then the vertex-connectivity of $G$ satisfies: $$\kappa(G) \geq d-36\lambda^2/d.$$ \end{theo} \noindent {\bf Proof.}\, We can assume that $\lambda \leq d/6$, since otherwise there is nothing to prove. Suppose that there is a subset $S \subset V$ of size less than $d-36\lambda^2/d$ such that the induced graph $G[V-S]$ is disconnected. Denote by $U$ the set of vertices of the smallest connected component of $G[V-S]$ and set $W=V-(S \cup U)$. Then $|W| \geq (n-d)/2\geq n/4$ and there is no edge between $U$ and $W$. Also $|U|+|S|>d$, since all the neighbors of a vertex from $U$ are contained in $S \cup U$. Therefore $|U|\geq 36\lambda^2/d$. Since there are no edges between $U$ and $W$, by Theorem \ref{eigen}, we have that $d|U||W|/n <\lambda \sqrt{|U||W|}$. This implies that $$|U|<\frac{\lambda^2n^2}{d^2|W|}= \frac{\lambda}{d}\frac{n}{|W|} \frac{\lambda n}{d} \leq \frac{1}{6} \cdot 4 \cdot\frac{\lambda n}{d}< \frac{\lambda n}{d}.$$ Next note that, by Theorem \ref{eigen}, the number of edges spanned by $U$ is at most $$e(U) \leq \frac{d|U|^2}{2n}+\frac{\lambda|U|}{2}< \frac{\lambda n}{d}\frac{d|U|}{2n}+\frac{\lambda|U|}{2}= \frac{\lambda|U|}{2}+\frac{\lambda|U|}{2}= \lambda|U|.$$ As the degree of every vertex in $U$ is $d$, it follows that $$e(U,S)\geq d|U|-2e(U)> (d-2\lambda)|U|\geq 2d|U|/3.$$ On the other hand using again Theorem \ref{eigen} together with the facts that $|U|\geq 36\lambda^2/d$, $|S|<d$ and $d \leq n/2$ we conclude that \begin{eqnarray*} e(U,S)&\leq& \frac{d|U||S|}{n}+\lambda\sqrt{|U||S|} <\frac{d}{n}d|U|+\lambda\sqrt{d|U|} \leq \frac{d|U|}{2}+ \frac{\lambda\sqrt{d}|U|}{\sqrt{|U|}}\\ &\leq&\frac{d|U|}{2}+\frac{\lambda\sqrt{d}|U|}{6\lambda/\sqrt{d}}= \frac{d|U|}{2}+\frac{d|U|}{6}=\frac{2d|U|}{3}. \end{eqnarray*} This contradiction completes the proof.\hfill $\Box$ \medskip \noindent The constants in this theorem can be easily improved and we make no attempt to optimize them. Note that, in particular, for an $(n,d,\lambda)$-graph $G$ with $\lambda=O(\sqrt{d})$ we have that $\kappa(G)= d-\Theta(1)$. Next we present an example which shows that the assertion of Theorem \ref{connectivity} is tight up to a constant factor. Let $G$ be any $(n,d,\lambda)$-graph with $\lambda=\Theta(\sqrt{d})$. We already constructed several such graphs in the previous section. For an integer $k$, consider a new graph $G_k$, which is obtained by replacing each vertex of $G$ by the complete graph of order $k$ and by connecting two vertices of $G_k$ by an edge if and only if the corresponding vertices of $G$ are connected by an edge. Then it follows immediately from the definition that $G_k$ has $n'=nk$ vertices and is $d'$-regular graph with $d'=dk+k-1$. Let $\lambda'$ be the second eigenvalue of $G_k$. To estimate $\lambda'$ note that the adjacency matrix of $G_k$ equals to $A_G\otimes J_k+I_n\otimes A_{K_k}$. Here $A_G$ is the adjacency matrix of $G$, $J_k$ is the all one matrix of size $k\times k$, $I_n$ is the identity matrix of size $n \times n$ and $A_{K_k}$ is the adjacency matrix of the complete graph of order $k$. Also the tensor product of the $m\times n$ dimensional matrix $A=(a_{ij})$ and the $s\times t$-dimensional matrix $B=(b_{kl})$ is the $ms\times nt$-dimensional matrix $A\otimes B$, whose entry labeled $((i,k)(j,l))$ is $a_{ij}b_{kl}$. In case $A$ and $B$ are symmetric matrices with spectrums $\{\lambda_1, \ldots , \lambda_n\}$, $\{ \mu_1,\ldots , \mu_t\}$ respectively, it is a simple consequence of the definition that the spectrum of $A\otimes B$ is $\{ \lambda_i\mu_k: i=1, \ldots ,n, k=1,\ldots , t\}$ (see, e.g. \cite{Lov93}). Therefore the second eigenvalue of $A_G\otimes J_k$ is $k\lambda$. On the other hand $I_n\otimes A_{K_k}$ is the adjacency matrix of the disjoint union of $k$-cliques and therefore the absolute value of all its eigenvalues is at most $k-1$. Using these two facts we conclude that $\lambda'\leq \lambda k+k-1$ and that $G_k$ is $(n'=nk,d'=dk+k-1,\lambda'=\lambda k+k-1)$-graph. Also it is easy to see that the set of vertices of $G_k$ that corresponds to a vertex in $G$ has exactly $dk$ neighbors outside this set. By deleting these neighbors we can disconnect the graph $G_k$ and thus $$\kappa(G_k) \leq dk=d'-(k-1)=d'-\Omega\big((\lambda')^2/d'\big).$$ Sometimes we can improve the result of Theorem \ref{connectivity} using the information about co-degrees of vertices in our graph. Such result was used in \cite{KSVW} to determine the vertex-connectivity of dense random $d$-regular graphs. \begin{prop} \label{p42}\cite{KSVW} Let $G=(V,E)$ be a $d$-regular graph on $n$ vertices such that $ \sqrt{n}\log n < d \leq 3n/4$ and the number of common neighbors for every two distinct vertices in $G$ is $(1+o(1))d^2/n$. Then the graph $G$ is $d$-vertex-connected. \end{prop} Similarly to vertex-connectivity, define the {\em edge-connectivity} of a graph $G$ to be the minimum number of edges that we need to delete to make $G$ disconnected. We denote this parameter by $\kappa'(G)$. Clearly the edge-connectivity is always at most the minimum degree of a graph. We also say that $G$ has a {\em perfect matching} if there is a set of disjoint edges that covers all the vertices of $G$. Next we show that $(n,d,\lambda)$-graphs even with a very weak spectral gap are $d$-edge-connected and have a perfect matching (if the number of vertices is even). \begin{theo} \label{edge-connectivity} Let $G$ be an $(n,d,\lambda)$-graph with $d-\lambda\geq 2$. Then $G$ is $d$-edge-connected. When $n$ is even, it has a perfect matching. \end{theo} \noindent {\bf Proof.}\, Let $U$ be a subset of vertices of $G$ of size at most $n/2$. To prove that $G$ is $d$-edge-connected we need to show that there are always at least $d$ edges between $U$ and $V(G)-U$. If $1 \leq |U|\leq d$, then every vertex in $U$ has at least $d-(|U|-1)$ neighbors outside $U$ and therefore $e(U,V(G)-U) \geq |U|\big(d-|U|+1\big) \geq d$. On the other hand if $d \leq |U|\leq n/2$, then using that $d-\lambda\geq 2$ together with Theorem \ref{eigen} we obtain that \begin{eqnarray*} e\big(U,V(G)-U\big) &\geq& \frac{d|U|(n-|U|)}{n}-\lambda\sqrt{|U|(n-|U|)\left(1-\frac{|U|}{n}\right) \left(1-\frac{n-|U|}{n}\right)}\\ &=&(d-\lambda)\frac{(n-|U|)}{n}|U| \geq 2\cdot\frac{1}{2}\cdot|U|=|U|\geq d, \end{eqnarray*} and therefore $\kappa'(G)=d$. To show that $G$ contains a perfect matching we apply the celebrated Tutte's condition. Since $n$ is even, we need to prove that for every nonempty set of vertices $S$, the induced graph $G[V-S]$ has at most $|S|$ connected components of odd size. Since $G$ is $d$-edge-connected we have that there are at least $d$ edges from every connected component of $G[V-S]$ to $S$. On the other hand there are at most $d|S|$ edges incident with vertices in $S$. Therefore $G[V-S]$ has at most $|S|$ connected components and hence $G$ contains a perfect matching. \hfill $\Box$ \subsection{Maximum cut} Let $G=(V,E)$ be a graph and let $S$ be a nonempty proper subset of $V$. Denote by $(S,V-S)$ the cut of $G$ consisting of all edges with one end in $S$ and another one in $V-S$. The {\em size} of the cut is the number of edges in it. The MAX CUT problem is the problem of finding a cut of maximum size in $G$. Let $f(G)$ be the size of the maximum cut in $G$. MAX CUT is one of the most natural combinatorial optimization problems. It is well known that this problem is NP-hard \cite{GJ}. Therefore it is useful to have bounds on $f(G)$ based on other parameters of the graph, that can be computed efficiently. Here we describe two such folklore results. First, consider a random partition $V=V_1\cup V_2$, obtained by assigning each vertex $v\in V$ to $V_1$ or $V_2$ with probability $1/2$ independently. It is easy to see that each edge of $G$ has probability $1/2$ to cross between $V_1$ and $V_2$. Therefore the expected number of edges in the cut $(V_1,V_2)$ is $m/2$, where $m$ is the number of edges in $G$. This implies that for every graph $f(G)\geq m/2$. The example of a complete graph shows that this lower bound is asymptotically optimal. The second result provides an upper bound for $f(G)$, for a regular graph $G$, in terms of the smallest eigenvalue of its adjacency matrix. \begin{prop} \label{max-cut} Let $G$ be a $d$-regular graph (which may have loops) of order $n$ with $m=dn/2$ edges and let $\lambda_1\geq \lambda_2 \geq \ldots \geq \lambda_n$ be the eigenvalues of the adjacency matrix of $G$. Then $$f(G) \leq \frac{m}{2}-\frac{\lambda_n n}{4}.$$ In particular if $G$ is an $(n,d,\lambda)$-graph then $f(G) \leq (d+\lambda)n/4$. \end{prop} \noindent {\bf Proof.}\, Let $A=(a_{ij})$ be the adjacency matrix of $G=(V,E)$ and let $V=\{1, \ldots, n\}$. Let ${\bf x}=(x_1, \ldots, x_n)$ be any vector with coordinates $\pm 1$. Since the graph $G$ is $d$-regular we have $$\sum_{(i,j)\in E} (x_i-x_j)^2=d\sum_{i=1}^n x_i^2- \sum_{i,j}a_{ij}x_ix_j=dn-{\bf x}^tA{\bf x}.$$ By the variational definition of the eigenvalues of $A$, for any vector $z \in R^n$, $z^tAz \geq \lambda_n\|z\|^2$. Therefore \begin{equation} \label{a} \sum_{(i,j)\in E} (x_i-x_j)^2=dn-{\bf x}^tA{\bf x} \leq dn-\lambda_n\|{\bf x}\|^2=dn-\lambda_nn. \end{equation} Let $V=V_1\cup V_2$ be an arbitrary partition of $V$ into two disjoint subsets and let $e(V_1,V_2)$ be the number of edges in the bipartite subgraph of $G$ with bipartition $(V_1,V_2)$. For every vertex $v \in V(G)$ define $x_v=1$ if $v \in V_1$ and $x_v=-1$ if $v \in V_2$. Note that for every edge $(i,j)$ of $G$, $(x_i-x_j)^2=4$ if this edge has its ends in the distinct parts of the above partition and is zero otherwise. Now using (\ref{a}), we conclude that $$\hspace{3.2cm} e(V_1,V_2)=\frac{1}{4} \sum_{(i,j)\in E} (x_i-x_j)^2 \leq \frac{1}{4}(dn-\lambda_nn)=\frac{m}{2}-\frac{\lambda_nn}{4}. \hspace{3.2cm}\Box$$ This upper bound is often used to show that some particular results about maximum cuts are tight. For example this approach was used in \cite{ALON} and \cite{ABKS}. In these papers the authors proved that for every graph $G$ with $m$ edges and girth at least $r\geq 4$, $f(G) \geq m/2 +\Omega\big(m^{\frac{r}{r+1}}\big)$. They also show, using Proposition \ref{max-cut} and Examples 9, 6 from Section 3, that this bound is tight for $r=4,5$. \subsection{Independent sets and the chromatic number} The {\em independence number} $\alpha(G)$ of a graph $G$ is the maximum cardinality of a set of vertices of $G$ no two of which are adjacent. Using Theorem \ref{eigen} we can immediately establish an upper bound on the size of a maximum independent set of pseudo-random graphs. \begin{prop} \label{ind-set} Let $G$ be an $(n,d,\lambda)$-graph, then $$\alpha(G)\leq \frac{\lambda n}{d+\lambda}.$$ \end{prop} \noindent {\bf Proof.}\, Let $U$ be an independent set in $G$, then $e(U)=0$ and by Theorem \ref{eigen} we have that $d|U|^2/n\leq \lambda |U|(1-|U|/n)$. This implies that $|U|\leq \lambda n/(d+\lambda)$. \hfill $\Box$ \vspace{0.15cm} \noindent Note that even when $\lambda=O(\sqrt{d})$ this bound only has order of magnitude $O(n/\sqrt{d})$. This contrasts sharply with the behavior of random graphs where it is known (see \cite{Bol01} and \cite{JanLucRuc00}) that the independence number of random graph $G(n,p)$ is only $\Theta\big(\frac{n}{d}\log d\big)$ where $d=(1+o(1))np$. More strikingly there are graphs for which the bound in Proposition \ref{ind-set} cannot be improved. One such graph is the Paley graph $P_q$ with $q=p^2$ (Example 3 in the previous section). Indeed it is easy to see that in this case all elements of the subfield $GF(p)\subset GF(p^2)$ are quadratic residues in $GF(p^2)$. This implies that for every quadratic non-residue $\beta \in GF(p^2)$ all elements of any multiplicative coset $\beta GF(p)$ form an independent set of size $p$. As we already mentioned, $P_q$ is an $(n,d,\lambda)$-graph with $n=p^2,d=(p^2-1)/2$ and $\lambda=(p+1)/2$. Hence for this graph we get $\alpha(P_q)=\lambda n/(d+\lambda)$. Next we obtain a lower bound on the independence number of pseudo-random graphs. We present a slightly more general result by Alon et al. \cite{AKS} which we will need later. \begin{prop} \label{ind-set1}\cite{AKS} Let $G$ be an $(n,d,\lambda)$-graph such that $\lambda<d\leq 0.9n$. Then the induced subgraph $G[U]$ of $G$ on any subset $U, |U|=m$, contains an independent set of size at least $$\alpha(G[U]) \geq \frac{n}{2(d-\lambda)} \ln \left(\frac{m(d-\lambda)}{n(\lambda+1)} +1 \right).$$ In particular, $$\alpha(G) \geq \frac{n}{2(d-\lambda)} \ln \left(\frac{(d-\lambda)}{(\lambda+1)} +1 \right).$$ \end{prop} \noindent {\bf Sketch of proof.}\, First using Theorem \ref{eigen} it is easy to show that if $U$ is a set of $bn$ vertices of $G$, then the minimum degree in the induced subgraph $G[U]$ is at most $db+\lambda(1-b)=(d-\lambda)b+\lambda$. Construct an independent set $I$ in the induced subgraph $G[U]$ of $G$ by the following greedy procedure. Repeatedly choose a vertex of minimum degree in $G[U]$ , add it to the independent set $I$ and delete it and its neighbors from $U$, stopping when the remaining set of vertices is empty. Let $a_i, i\geq 0$ be the sequence of numbers defined by the following recurrence formula: $$a_0=m,~~ a_{i+1}=a_i-\left(d\frac{a_i}{n}+\lambda(1-\frac{a_i}{n})+1\right)= \left(1-\frac{d-\lambda}{n}\right)a_i-(\lambda+1),~\forall i \geq 0.$$ By the above discussion, it is easy to see that the size of the remaining set of vertices after $i$ iterations is at least $a_i$. Therefore the size of the resulting independent set $I$ is at least the smallest index $i$ such that $a_i\leq 0$. By solving the recurrence equation we obtain that this index satisfies: $$\hspace{5cm}i \geq \frac{n}{2(d-\lambda)} \ln \left(\frac{m(d-\lambda)}{n(\lambda+1)} +1 \right)\,. \hspace{5cm} \Box$$ \vspace{0.15cm} \noindent For an $(n,d,\lambda)$-graph $G$ with $\lambda \leq d^{1-\delta}, \delta>0$, this proposition implies that $\alpha(G) \geq \Omega\big(\frac{n}{d}\log d\big)$. This shows that the independence number of a pseudo-random graph with a sufficiently small second eigenvalue is up to a constant factor at least as large as $\alpha(G(n,p))$ with $p=d/n$. On the other hand the graph $H_k$ (Example 4, Section 3) shows that even when $\lambda \leq O(\sqrt{d})$ the independence number of $(n,d,\lambda)$-graph can be smaller than $\alpha(G(n,p))$ with $p=d/n$. This graph has $n=2^{k-1}-1$ vertices, degree $d=(1+o(1))n/2$ and $\lambda= \Theta(\sqrt{d})$. Also it is easy to see that every independent set in $H_k$ corresponds to a family of orthogonal vectors in ${\bf Z}_2^k$ and thus has size at most $k=(1+o(1))\log_2 n$. This is only half of the size of a maximum independent set in the corresponding random graph $G(n,1/2)$. A {\em vertex-coloring} of a graph $G$ is an assignment of a color to each of its vertices. The coloring is {\em proper} if no two adjacent vertices get the same color. The {\em chromatic number} $\chi(G)$ of $G$ is the minimum number of colors used in a proper coloring of it. Since every color class in the proper coloring of $G$ forms an independent set we can immediately obtain that $\chi(G)\geq |V(G)|/\alpha(G)$. This together with Proposition \ref{ind-set} implies the following result of Hoffman \cite{Ho}. \begin{coro} \label{hofman} Let $G$ be an $(n,d,\lambda)$-graph. Then the chromatic number of $G$ is at least $1+d/\lambda$. \end{coro} On the other hand, using Proposition \ref{ind-set1}, one can obtain the following upper bound on the chromatic number of pseudo-random graphs. \begin{theo} \label{chromatic}\cite{AKS} Let $G$ be an $(n,d,\lambda)$-graph such that $\lambda<d\leq 0.9n$. Then the chromatic number of $G$ satisfies $$ \chi(G) \leq \frac{6(d-\lambda)}{\ln\big (\frac{d-\lambda}{\lambda+1} +1\big)}\, . $$ \end{theo} \noindent {\bf Sketch of proof.}\, Color the graph $G$ as follows. As long as the remaining set of vertices $U$ contains at least $n/\ln (\frac{d-\lambda}{\lambda+1}+1)$ vertices, by Proposition \ref{ind-set1} we can find an independent set of vertices in the induced subgraph $G[U]$ of size at least $$\frac{n}{2(d-\lambda)} \ln \left( \frac{|U|(d-\lambda)}{n(\lambda+1)}+1\right) \geq \frac{n}{4(d-\lambda)} \ln \left( \frac{d-\lambda}{\lambda+1}+1\right).$$ Color all the members of such a set by a new color, delete them from the graph and continue. When this process terminates, the remaining set of vertices $U$ is of size at most $n/\ln (\frac{d-\lambda}{\lambda+1}+1)$ and we used at most $4(d-\lambda)/\ln(\frac{d-\lambda}{\lambda+1}+1)$ colors so far. As we already mentioned above, for every subset $U'\subset U$ the induced subgraph $G[U']$ contains a vertex of degree at most $$(d-\lambda)\frac{|U'|}{n}+\lambda\leq (d-\lambda)\frac{|U|}{n}+\lambda\leq \frac{d-\lambda}{\ln (\frac{d-\lambda}{\lambda+1}+1)}+\lambda \leq \frac{2(d-\lambda)}{\ln (\frac{d-\lambda}{\lambda+1}+1)}-1.$$ Thus we can complete the coloring of $G$ by coloring $G[U]$ using at most $2(d-\lambda)/\ln (\frac{d-\lambda}{\lambda+1}+1)$ additional colors. The total number of colors used is at most $6(d-\lambda)/\ln (\frac{d-\lambda}{\lambda+1}+1)$. \hfill$\Box$ \vspace{0.15cm} \noindent For an $(n,d,\lambda)$-graph $G$ with $\lambda \leq d^{1-\delta}, \delta>0$ this proposition implies that $\chi(G) \leq O\big(\frac{d}{\log d}\big)$. This shows that the chromatic number of a pseudo-random graph with a sufficiently small second eigenvalue is up to a constant factor at least as small as $\chi(G(n,p))$ with $p=d/n$. On the other hand, the Paley graph $P_q, q=p^2$, shows that sometimes the chromatic number of a pseudo-random graph can be much smaller than the above bound, even in the case $\lambda=\Theta(\sqrt{d})$. Indeed, as we already mentioned above, all elements of the subfield $GF(p)\subset GF(p^2)$ are quadratic residues in $GF(p^2)$. This implies that for every quadratic non-residue $\beta \in GF(p^2)$ all elements of a multiplicative coset $\beta GF(p)$ form an independent set of size $p$. Also all additive cosets of $\beta GF(p)$ are independent sets in $P_q$. This implies that $\chi(P_q)\leq \sqrt{q}=p$. In fact $P_q$ contains a clique of size $p$ (all elements of a subfield $GF(p)$), showing that $\chi(P_q)=\sqrt{q}\ll q/\log q$. Therefore the bound in Corollary \ref{hofman} is best possible. A more complicated quantity related to the chromatic number is the {\em list-chromatic number} $\chi_l(G)$ of $G$, introduced in \cite{ERT} and \cite{Vi}. This is the minimum integer $k$ such that for every assignment of a set $S(v)$ of $k$ colors to every vertex $v$ of $G$, there is a proper coloring of $G$ that assigns to each vertex $v$ a color from $S(v)$. The study of this parameter received a considerable amount of attention in recent years, see, e.g., \cite{Al2}, \cite{KTV} for two surveys. Note that from the definition it follows immediately that $\chi_l(G) \geq \chi(G)$ and it is known that the gap between these two parameters can be arbitrarily large. The list-chromatic number of pseudo-random graphs was studied by Alon, Krivelevich and Sudakov \cite{AKS} and independently by Vu \cite{Vu}. In \cite{AKS} and \cite{Vu} the authors mainly considered graphs with all degrees $(1+o(1))np$ and all co-degrees $(1+o(1))np^2$. Here we use ideas from these two papers to obtain an upper bound on the list-chromatic number of an $(n,d,\lambda)$-graphs. This bound has the same order of magnitude as the list chromatic number of the truly random graph $G(n,p)$ with $p=d/n$ (for more details see \cite{AKS}, \cite{Vu}). \begin{theo} \label{choice} Suppose that $0<\delta<1$ and let $G$ be an $(n,d,\lambda)$-graph satisfying $\lambda \leq d^{1-\delta}$, $d\le 0.9n$. Then the list-chromatic number of $G$ is bounded by $$\chi_l(G)\leq O\left(\frac{d}{\delta\log d}\right).$$ \end{theo} \noindent {\bf Proof.}\, Suppose that $d$ is sufficiently large and consider first the case when $d \leq n^{1-\delta/4}$. Then by Theorem \ref{eigen} the neighbors of every vertex in $G$ span at most $d^3/n+\lambda d \leq O(d^{2-\delta/4})$ edges. Now we can apply the result of Vu \cite{Vu} which says that if the neighbors of every vertex in a graph $G$ with maximum degree $d$ span at most $O(d^{2-\delta/4})$ edges then $\chi_l(G)\leq O\big(d/(\delta\log d)\big).$ Now consider the case when $d \geq n^{1-\delta/4}$. For every vertex $v \in V$, let $S(v)$ be a list of at least $\frac{7d}{\delta \log n}$ colors. Our objective is to prove that there is a proper coloring of $G$ assigning to each vertex a color from its list. As long as there is a set $C$ of at least $n^{1-\delta/2} $ vertices containing the same color $c$ in their lists we can, by Proposition \ref{ind-set1}, find an independent set of at least $\frac{\delta n}{6d} \log n$ vertices in $C$, color them all by $c$, omit them from the graph and omit the color $c$ from all lists. The total number of colors that can be deleted in this process cannot exceed $\frac{6d}{\delta \log n}$ (since in each such deletion at least $\frac{\delta n}{6d}\log n$ vertices are deleted from the graph). When this process terminates, no color appears in more than $n^{1-\delta/2}$ lists, and each list still contains at least $\frac{d}{\delta \log n} > n^{1-\delta/2}$ colors. Therefore, by Hall's theorem, we can assign to each of the remaining vertices a color from its list so that no color is being assigned to more than one vertex, thus completing the coloring and the proof.\hfill $\Box$ \subsection{Small subgraphs} We now examine small subgraphs of pseudo-random graphs. Let $H$ be a fixed graph of order $s$ with $r$ edges and with automorphism group $Aut(H)$. Using the second moment method it is not difficult to show that for every constant $p$ the random graph $G(n,p)$ contains $$(1+o(1))p^r(1-p)^{{s\choose 2}-r}\frac{n^s}{|Aut(H)|}$$ induced copies of $H$. Thomason extended this result to jumbled graphs. He showed in \cite{Tho87a} that if a graph $G$ is $(p,\alpha)$-jumbled and $p^sn\gg 42\alpha s^2$ then the number of induced subgraphs of $G$ which are isomorphic to $H$ is $(1+o(1))p^s(1-p)^{{s\choose 2}-r}n^s/|Aut(H)|$. Here we present a result of Noga Alon \cite{Alon} that proves that every large subset of the set of vertices of $(n,d,\lambda)$-graph contains the "correct" number of copies of any fixed sparse graph. An additional advantage of this result is that its assertion depends not on the number of vertices $s$ in $H$ but only on its maximum degree $\Delta$ which can be smaller than $s$. Special cases of this result have appeared in various papers including \cite{AK}, \cite{AP} and probably other papers as well. The approach here is similar to the one in \cite{AP}. \begin{theo} \label{number-subgraphs} \cite{Alon} Let $H$ be a fixed graph with $r$ edges, $s$ vertices and maximum degree $\Delta$, and let $G=(V,E)$ be an $(n,d,\lambda)$-graph, where, say, $d \leq 0.9n$. Let $m <n$ satisfy $m \gg \lambda (\frac{n}{d})^{\Delta}$. Then, for every subset $V' \subset V$ of cardinality $m$, the number of (not necessarily induced) copies of $H$ in $V'$ is $$ \big(1+o(1)\big)\frac{m^s}{|Aut(H)|} \left(\frac{d}{n}\right)^r. $$ \end{theo} Note that this implies that a similar result holds for the number of induced copies of $H$. Indeed, if $ n \gg d$ and $m \gg \lambda (\frac{n}{d})^{\Delta+1}$ then the number of copies of each graph obtained from $H$ by adding to it at least one edge is, by the above Theorem, negligible compared to the number of copies of $H$, and hence almost all copies of $H$ in $V'$ are induced. If $d=\Theta(n)$ then, by inclusion-exclusion, the number of induced copies of $H$ in $V'$ as above is also roughly the "correct" number. A special case of the above theorem implies that if $\lambda =O(\sqrt d)$ and $d \gg n^{2/3}$, then any $(n,d, \lambda)$-graph contains many triangles. As shown in Example 9, Section 3, this is not true when $d=(\frac{1}{4}+o(1)) n^{2/3}$, showing that the assertion of the theorem is not far from being best possible. \noindent {\bf Proof of Theorem \ref{number-subgraphs}.}\, To prove the theorem, consider a random one-to-one mapping of the set of vertices of $H$ into the set of vertices $V'$. Denote by $A(H)$ the event that every edge of $H$ is mapped on an edge of $G$. In such a case we say that the mapping is an embedding of $H$. Note that it suffices to prove that \begin{equation} \label{e1} Pr(A(H))=(1+o(1))\left(\frac{d}{n}\right)^r. \end{equation} We prove (\ref{e1}) by induction on the number of edges $r$. The base case $(r=0)$ is trivial. Suppose that (\ref{e1}) holds for all graphs with less than $r$ edges, and let $uv$ be an edge of $H$. Let $H_{uv}$ be the graph obtained from $H$ by removing the edge $uv$ (and keeping all vertices). Let $H_u$ and $H_v$ be the induced subgraphs of $H$ on the sets of vertices $V(H)\setminus \{ v\}$ and $V(H)\setminus \{ u\}$, respectively, and let $H'$ be the induced subgraph of $H$ on the set of vertices $V(H)\setminus \{ u,v\}$. Let $r'$ be the number of edges of $H'$ and note that $r-r' \leq 2(\Delta-1)+1=2\Delta -1. $ Clearly $Pr(A(H_{uv}))=Pr(A(H_{uv})|A(H'))\cdot Pr(A(H'))$. Thus, by the induction hypothesis applied to $H_{uv}$ and to $H'$: \[ Pr(A(H_{uv})|A(H'))=(1+o(1)) \left(\frac{d}{n}\right)^{r-1-r'}. \] For an embedding $f'$ of $H'$, let $\nu(u,f')$ be the number of extensions of $f'$ to an embedding of $H_u$ in $V'$; $\nu(v,f')$ denotes the same for $v$. Clearly, the number of extensions of $f'$ to an embedding of $H_{uv}$ in $V'$ is at least $\nu(u,f')\nu(v,f')-\min(\nu(u,f'),\nu(v,f'))$ and at most $\nu(u,f')\nu(v,f')$. Thus we have \[ \frac{\nu(u,f')\nu(v,f')-\min(\nu(u,f'),\nu(v,f'))}{(m-s+2)(m-s+1)}\leq Pr\big(A(H_{uv})|f'\big)\leq \frac{\nu(u,f')\nu(v,f')}{(m-s+2)(m-s+1)}. \] Taking expectation over all embeddings $f'$ the middle term becomes $Pr(A(H_{uv})|A(H'))$, which is $(1+o(1)) (\frac{d}{n})^{r-1-r'}$. Note that by our choice of the parameters and the well known fact that $\lambda =\Omega(\sqrt d)$, the expectation of the term $\min(\nu(u,f'),\nu(v,f'))~(~ \leq m)$ is negligible and we get \[ E_{f'}\big(\nu(u,f')\nu(v,f')|\ A(H')\big)=(1+o(1)) m^2 \left(\frac{d}{n}\right)^{r-1-r'}. \] Now let $f$ be a random one-to-one mapping of $V(H)$ into $V'$. Let $f'$ be a fixed embedding of $H'$. Then \[ Pr_f\big(A(H)|\ f|_{V(H)\setminus \{ u,v\}}=f'\big)= \left(\frac{d}{n}\right)\frac{\nu(u,f')\nu(v,f')}{(m-s+2)(m-s+1)}+\delta, \] where $|\delta|\leq \lambda\frac{\sqrt{\nu(u,f')\nu(v,f')}}{(m-s+2)(m-s+1)}$. This follows from Theorem \ref{eigen}, where we take the possible images of $u$ as the set $U$ and the possible images of $v$ as the set $W$. Averaging over embeddings $f'$ we get $Pr(A(H)|A(H'))$ on the left hand side. On the right hand side we get $(1+o(1)) (\frac{d}{n})^{r-r'}$ from the first term plus the expectation of the error term $\delta$. By Jensen's inequality, the absolute value of this expectation is bounded by \[ \lambda\frac{\sqrt{E(\nu(u,f')\nu(v,f'))}}{(m-s+2)(m-s+1)} =(1+o(1)) \frac{\lambda}{m} \left(\frac{d}{n}\right)^{(r-r'-1)/2}. \] Our assumptions on the parameters imply that this is negligible with respect to the main term. Therefore $ Pr(A(H))=Pr(A(H)|A(H')) \cdot Pr(A(H'))=(1+o(1)) \left(\frac{d}{n}\right)^r$, completing the proof of Theorem \ref{number-subgraphs}. \hfill $\Box$ If we are only interested in the existence of one copy of $H$ then one can sometimes improve the conditions on $d$ and $\lambda$ in Theorem \ref{number-subgraphs}. For example if $H$ is a complete graph of order $r$ then the following result was proved in \cite{AK}. \begin{prop} \label{cliques}\cite{AK} Let $G$ be an $(n,d,\lambda)$-graph. Then for every integer $r \geq 2$ every set of vertices of $G$ of size more than $$\frac{(\lambda+1)n}{d}\left(1+\frac{n}{d}+\ldots+\Big(\frac{n}{d}\Big)^{r-2}\right)$$ contains a copy of a complete graph $K_r$. \end{prop} \noindent In particular, when $d\geq\Omega(n^{2/3})$ and $\lambda \leq O(\sqrt{d})$ then any $(n,d,\lambda)$-graph contains a triangle and as shows Example 9 in Section 3 this is tight. Unfortunately we do not know if this bound is also tight for $r \geq 4$. It would be interesting to construct examples of $(n,d,\lambda)$-graphs with $d=\Theta\big(n^{1-1/(2r-3)}\big)$ and $\lambda \leq O(\sqrt{d})$ which contain no copy of $K_r$. Finally we present one additional result about the existence of odd cycles in pseudo-random graphs. \begin{prop} \label{cycles} Let $k\geq 1$ be an integer and let $G$ be an $(n,d,\lambda)$-graph such that $d^{2k}/n\gg \lambda^{2k-1}$. Then $G$ contains a cycle of length $2k+1$. \end{prop} \noindent {\bf Proof.}\, Suppose that $G$ contains no cycle of length $2k+1$. For every two vertices $u,v$ of $G$ denote by $d(u,v)$ the length of a shortest path from $u$ to $v$. For every $i \geq 1$ let $N_i(v)=\{u~|~d(u,v)=i\}$ be the set of all vertices in $G$ which are at distance exactly $i$ from $v$. In \cite{EFRS} Erd\H{o}s et al. proved that if $G$ contains no cycle of length $2k+1$ then for any $1 \leq i \leq k$ the induced graph $G[N_i(v)]$ contains an independent set of size $|N_i(v)|/(2k-1)$. This result together with Proposition \ref{ind-set} implies that for every vertex $v$ and for every $1 \leq i \leq k$, $|N_i(v)| \leq (2k-1)\lambda n/d$. Since $d^{2k}/n\gg \lambda^{2k-1}$ we have that $\lambda=o(d)$. Therefore by Theorem \ref{eigen} $$e\big(N_i(v)\big) \leq \frac{d}{2n}|N_i(v)|^2+\lambda |N_i(v)| \leq \frac{d}{n} \frac{(2k-1)\lambda n}{2d}|N_i(v)|+\lambda |N_i(v)| <2k\lambda|N_i(v)|=o\big(d|N_i(v)|\big).$$ Next we prove by induction that for every $1\le i \leq k$, $\frac{|N_{i+1}(v)|}{|N_i(v)|}\geq (1-o(1))d^2/\lambda^2$. By the above discussion the number of edges spanned by $N_1(v)$ is $o(d^2)$ and therefore $e\big(N_1(v),N_2(v)\big)=d^2-o(d^2)=(1-o(1))d^2$. On the other hand, by Theorem \ref{eigen} \begin{eqnarray*} e\big(N_1(v),N_2(v)\big) &\leq& \frac{d}{n}|N_1(v)||N_2(v)|+\lambda\sqrt{|N_1(v)||N_2(v)|} \leq \frac{d}{n}\,d\,\frac{(2k-1)\lambda n}{d} +\lambda\sqrt{d|N_2(v)|}\\ &=&\lambda d\sqrt{\frac{|N_2(v)|}{d}}+O(\lambda d)= \lambda d\sqrt{\frac{|N_2(v)|}{|N_1(v)|}}+o(d^2). \end{eqnarray*} Therefore $\frac{|N_2(v)|}{|N_1(v)|}\geq (1-o(1))d^2/\lambda^2$. Now assume that $\frac{|N_{i}(v)|}{|N_{i-1}(v)|}\geq (1-o(1))d^2/\lambda^2$. Since the number of edges spanned by $N_{i}(v)$ is $o\big(d|N_i(v)|\big)$ we obtain \begin{eqnarray*} e\big(N_i(v),N_{i+1}(v)\big)&=&d|N_i(v)|-2e\big(N_i(v)\big)-e\big(N_{i-1}(v),N_i(v)\big) \\ &\geq& d|N_i(v)|-o\big(d|N_i(v)|\big)-d|N_{i-1}(v)|\\ &\geq& (1-o(1))d|N_i(v)|-(1+o(1))d(\lambda^2/d^2)|N_i(v)|\\ &=&(1-o(1))d|N_i(v)|-o\big(d|N_i(v)|\big)= (1-o(1))d|N_i(v)|. \end{eqnarray*} On the other hand, by Theorem \ref{eigen} \begin{eqnarray*} e\big(N_i(v),N_{i+1}(v)\big) &\leq& \frac{d}{n}|N_i(v)||N_{i+1}(v)|+\lambda\sqrt{|N_i(v)||N_{i+1}(v)|}\\ &\leq& \frac{d}{n} \,\frac{(2k-1)\lambda n}{d}\,|N_i(v)| +\lambda\sqrt{|N_i(v)||N_{i+1}(v)|}\\ &=&O(\lambda |N_i(v)|)+\lambda |N_i(v)|\sqrt{\frac{|N_{i+1}(v)|}{|N_i(v)|}}= \lambda |N_i(v)|\sqrt{\frac{|N_{i+1}(v)|}{|N_i(v)|}}+o\big(d|N_i(v)|\big). \end{eqnarray*} Therefore $\frac{|N_{i+1}(v)|}{|N_i(v)|}\geq (1-o(1))d^2/\lambda^2$ and we proved the induction step. Finally note that $$|N_k(v)| = d \prod_{i=1}^{k-1}\frac{|N_{i+1}(v)|}{|N_i(v)|}\geq (1+o(1))d\left(\frac{d^2}{\lambda^2}\right)^{k-1}= (1+o(1)) \frac{d^{2k-1}}{\lambda^{2k-2}}\gg (2k-1)\frac{\lambda n}{d}.$$ This contradiction completes the proof. \hfill $\Box$ \vspace{0.15cm} \noindent This result implies that when $d\gg n^{\frac{2}{2k+1}}$ and $\lambda \leq O(\sqrt{d})$ then any $(n,d,\lambda)$-graph contains a cycle of length $2k+1$. As shown by Example 10 of the previous section this result is tight. It is worth mentioning here that it follows from the result of Bondy and Simonovits \cite{BS} that any $d$-regular graph with $d \gg n^{1/k}$ contains a cycle of length $2k$. Here we do not need to make any assumption about the second eigenvalue $\lambda$. This bound is known to be tight for $k=2,3,5$ (see Examples 6,7, Section 3). \subsection{Extremal properties} Tur\'an's theorem \cite{T} is one of the fundamental results in Extremal Graph Theory. It states that among $n$-vertex graphs not containing a clique of size $t$ the complete $(t-1)$-partite graph with (almost) equal parts has the maximum number of edges. For two graphs $G$ and $H$ we define the Tur\'an number $ex(G,H)$ of $H$ in $G$, as the largest integer $e$, such that there is an $H$-free subgraph of $G$ with $e$ edges. Obviously $ex (G,H) \leq |E(G)|$, where $E(G)$ denotes the edge set of $G$. Tur\'an's theorem, in an asymptotic form, can be restated as $$ex(K_n, K_t) = \left(\frac{t-2}{t-1}+o(1)\right){n\choose 2},$$ that is the largest $K_t$-free subgraph of $K_n$ contains approximately $\frac{t-2}{t-1}$-fraction of its edges. Here we would like to describe an extension of this result to $(n,d,\lambda)$-graphs. For an arbitrary graph $G$ on $n$ vertices it is easy to give a lower bound on $ex (G,K_t)$ following Tur\'an's construction. One can partition the vertex set of $G$ into $t-1$ parts such that the degree of each vertex within its own part is at most $\frac{1}{t-1}$-times its degree in $G$. Thus the subgraph consisting of the edges of $G$ connecting two different parts has at least a $\frac{t-2}{t-1}$-fraction of the edges of $G$ and is clearly $K_t$-free. We say that a graph (or rather a family of graphs) is {\em $t$-Tur\'an} if this trivial lower bound is essentially an upper bound as well. More precisely, $G$ is $t$-Tur\'an if $ex(G,K_t) = \big(\frac{t-2}{t-1} +o(1)\big) |E(G)|$. It has been shown that for any fixed $t$, there is a number $m(t,n)$ such that almost all graphs on $n$ vertices with $m \ge m(t,n) $ edges are $t$-Tur\'an (see \cite{SV}, \cite{KohRodS} for the most recent estimate for $m(t,n)$). However, these results are about random graphs and do not provide a deterministic sufficient condition for a graph to be $t$-Tur\'an. It appears that such a condition can be obtained by a simple assumption about the spectrum of the graph. This was proved by Sudakov, Szab\'o and Vu in \cite{SSV}. They obtained the following result. \begin{theo} \label{t1} \cite{SSV} Let $t\geq 3$ be an integer and let $G=(V,E)$ be an $(n,d,\lambda)$-graph. If $\lambda=o(d^{t-1}/n^{t-2})$ then $$ex(G,K_t)=\left(\frac{t-2}{t-1}+o(1)\right)|E(G)|.$$ \end{theo} Note that this theorem generalizes Tur\'an's theorem, as the second eigenvalue of the complete graph $K_n$ is 1. Let us briefly discuss the sharpness of Theorem \ref{t1}. For $t=3$, one can show that its condition involving $n,d$ and $\lambda$ is asymptotically tight. Indeed, in this case the above theorem states that if $d^2/n\gg \lambda$, then one needs to delete about half of the edges of $G$ to destroy all the triangles. On the other hand, by taking the example of Alon (Section \ref{examples}, Example 9) whose parameters are: $d=\Theta(n^{2/3})$, $\lambda=\Theta(n^{1/3})$, and blowing it up (which means replacing each vertex by an independent set of size $k$ and connecting two vertices in the new graph if and only if the corresponding vertices of $G$ are connected by an edge) we get a graph $G(k)$ with the following properties: \begin{center} $|V(G(k))|=n_k=nk$;\,\, $G(k)$ is $d_k=dk$-regular;\,\, $G(k)$ is triangle-free;\,\, $\lambda(G(k))=k\lambda$\, and \,$\lambda(G(k))=\Omega\big(d_k^2/n_k\big)$. \end{center} \noindent The above bound for the second eigenvalue of $G(k)$ can be obtained by using well known results on the eigenvalues of the tensor product of two matrices, see \cite{KriSudSza02} for more details. This construction implies that for $t=3$ and any sensible degree $d$ the condition in Theorem \ref{t1} is not far from being best possible. \subsection{Factors and fractional factors} Let $H$ be a fixed graph on $n$ vertices. We say that a graph $G$ on $n$ vertices has an {\em $H$-factor} if $G$ contains $n/h$ vertex disjoint copies of $H$. Of course, a trivial necessary condition for the existence of an $H$-factor in $G$ is that $h$ divides $n$. For example, if $H$ is just an edge $H=K_2$, then an $H$-factor is a perfect matching in $G$. One of the most important classes of graph embedding problems is to find sufficient conditions for the existence of an $H$-factor in a graph $G$, usually assuming that $H$ is fixed while the order $n$ of $G$ grows. In many cases such conditions are formulated in terms of the minimum degree of $G$. For example, the classical result of Hajnal and Szemer\'edi \cite{HajSze70} asserts that if the minimum degree $\delta(G)$ satisfies $\delta(G)\ge (1-\frac{1}{r})n$, then $G$ contains $\lfloor n/r\rfloor$ vertex disjoint copies of $K_r$. The statement of this theorem is easily seen to be tight. It turns our that pseudo-randomness allows in many cases to significantly weaken sufficient conditions for $H$-factors and to obtain results which fail to hold for general graphs of the same edge density. Consider first the case of a constant edge density $p$. In this case the celebrated Blow-up Lemma of Koml\'os, S\'ark\"ozy and Szemer\'edi \cite{KomSarSze97} can be used to show the existence of $H$-factors. In order to formulate the Blow-up Lemma we need to introduce the notion of a super-regular pair. Given $\epsilon>0$ and $0<p<1$, a bipartite graph $G$ with bipartition $(V_1,V_2)$, $|V_1|=|V_2|=n$, is called {\em super $(p,\epsilon)$-regular} if \begin{enumerate} \item For all vertices $v\in V(G)$, $$ (p-\epsilon)n\le d(v)\le (p+\epsilon)n\ ; $$ \item For every pair of sets $(U,W)$, $U\subset V_1$, $W\subset V_2$, $|U|,|W|\ge \epsilon n$, $$ \left|\frac{e(U,W)}{|U||W|}-\frac{|E(G)|}{n^2}\right|\le \epsilon\ . $$ \end{enumerate} \begin{theo}\label{KSS}\cite{KomSarSze97} For every choice of integers $r$ and $\Delta$ and a real $0<p<1$ there exist an $\epsilon>0$ and an integer $n_0(\epsilon)$ such that the following is true. Consider an $r$-partite graph $G$ with all partition sets $V_1,\ldots,V_r$ of order $n>n_0$ and all ${r\choose 2}$ bipartite subgraphs $G[V_i,V_j]$ super $(p,\epsilon)$-regular. Then for every $r$-partite graph $H$ with maximum degree $\Delta(H)\le \Delta$ and all partition sets $X_1,\ldots,X_r$ of order $n$, there exists an embedding $f$ of $H$ into $G$ with each set $X_i$ mapped onto $V_i$, $i=1,\ldots,r$. \end{theo} (The above version of the Blow-up Lemma, due to R\"odl and Ruci\'nski \cite{RodRuc99}, is somewhat different from and yet equivalent to the original formulation of Koml\'os et al. We use it here as it is somewhat closer in spirit to the notion of pseudo-randomness). The Blow-up Lemma is a very powerful embedding tool. Combined with another "big cannon", the Szemer\'edi Regularity Lemma, it can be used to obtain approximate versions of many of the most famous embedding conjectures. We suggest the reader to consult a survey of Koml\'os \cite{Kom99} for more details and discussions. It is easy to show that if $G$ is an $(n,d,\lambda)$-graph with $d=\Theta(n)$ and $\lambda=o(n)$, and $h$ divides $n$, then a random partition of $V(G)$ into $h$ equal parts $V_1,\ldots,V_h$ produces almost surely ${h\choose 2}$ super $(d/n,\epsilon)$-regular pairs. Thus the Blow-up Lemma can be applied to the obtained $h$-partite subgraph of $G$ and we get: \begin{coro}\label{Hdense} Let $G$ be an $(n,d,\lambda)$-graph with $d=\Theta(n)$, $\lambda=o(n)$. If $h$ divides $n$, then $G$ contains an $H$-factor, for every fixed graph $H$ on $h$ vertices. \end{coro} The case of a vanishing edge density $p=o(1)$ is as usual significantly more complicated. Here a sufficient condition for the existence of an $H$-factor should depend heavily on the graph $H$, as there may exist quite dense pseudo-random graphs without a single copy of $H$, see, for example, the Alon graph (Example 9 of Section \ref{examples}). When $H=K_2$, already a very weak pseudo-randomness condition suffices to guarantee an $H$-factor, or a perfect matching, as provided by Theorem \ref{edge-connectivity}. We thus consider the case $H=K_3$, the task here is to guarantee a {\em triangle factor}, i.e. a collection of $n/3$ vertex disjoint triangles. This problem has been treated by Krivelevich, Sudakov and Szab\'o \cite{KriSudSza02} who obtained the following result: \begin{theo}\cite{KriSudSza02}\label{KrSS} Let $G$ be an $(n,d,\lambda)$-graph. If $n$ is divisible by 3 and $$ \lambda=o\left(\frac{d^3}{n^2\log n}\right)\,, $$ then $G$ has a triangle factor. \end{theo} For best pseudo-random graphs with $\lambda=\Theta(\sqrt{d})$ the condition of the above theorem is fulfilled when $d\gg n^{4/5}\log^{2/5}n$. To prove Theorem \ref{KrSS} Krivelevich et al. first partition the vertex set $V(G)$ into three parts $V_1,V_2,V_3$ of equal cardinality at random. Then they choose a perfect matching $M$ between $V_1$ an $V_2$ at random and form an auxiliary bipartite graph $\Gamma$ whose parts are $M$ and $V_3$, and whose edges are formed by connecting $e\in M$ and $v\in V_3$ if both endpoints of $e$ are connected by edges to $v$ in $G$. The existence of a perfect matching in $\Gamma$ is equivalent to the existence of a triangle factor in $G$. The authors of \cite{KriSudSza02} then proceed to show that if $M$ is chosen at random then the Hall condition is satisfied for $\Gamma$ with positive probability. The result of Theorem \ref{KrSS} is probably not tight. In fact, the following conjecture is stated in \cite{KriSudSza02}: \begin{conj}\cite{KriSudSza02}\label{KrSScon} There exists an absolute constant $c>0$ so that every $d$-regular graph $G$ on $3n$ vertices, satisfying $\lambda(G)\le cd^2/n$, has a triangle factor. \end{conj} If true the above conjecture would be best possible, up to a constant multiplicative factor. This is shown by taking the example of Alon (Section \ref{examples}, Example 9) and blowing each of its vertices by an independent set of size $k$. As we already discussed in the previous section (see also \cite{KriSudSza02}), this gives a triangle-free $d_k$-regular graph $G(k)$ on $n_k$ vertices which satisfies $\lambda(G(k))=\Omega\big(d_k^2/n_k\big)$. Krivelevich, Sudakov and Szab\'o considered in \cite{KriSudSza02} also the fractional version of the triangle factor problem. Given a graph $G=(V,E)$, denote by $T=T(G)$ the set of all triangles of $G$. A function $f:T\rightarrow {\mathbb R}_+$ is called a {\em fractional triangle factor} if for every $v\in V(G)$ one has $\sum_{v\in t}f(t)=1$. If $G$ contains a triangle factor $T_0$, then assigning values $f(t)=1$ for all $t\in T_0$, and $f(t)=0$ for all other $t\in T$ produces a fractional triangle factor. This simple argument shows that the existence of a triangle factor in $G$ implies the existence of a fractional triangle factor. The converse statement is easily seen to be invalid in general. The fact that a fractional triangle factor $f$ can take non-integer values, as opposed to the characteristic vector of a "usual" (i.e. integer) triangle factor, enables to invoke the powerful machinery of Linear Programming to prove a much better result than Theorem \ref{KrSS}. \begin{theo}\label{KrSSfr}\cite{KriSudSza02} Let $G=(V,E)$ be a $(n,d,\lambda)$-graph If $\lambda \le 0.1d^2/n$ then $G$ has a fractional triangle factor. \end{theo} This statement is optimal up to a constant factor -- see the discussion following Conjecture \ref{KrSScon}. Already for the next case $H=K_4$ analogs of Theorem \ref{KrSS} and \ref{KrSSfr} are not known. In fact, even an analog of Conjecture \ref{KrSScon} is not available either, mainly due to the fact that we do not know the weakest possible spectral condition guaranteeing a single copy of $K_4$, or $K_r$ in general, for $r\ge 4$. Finally it would be interesting to show that for every integer $\Delta$ there exist a real $M$ and an integer $n_0$ so that the following is true. If $n\ge n_0$ and $G$ is an $(n,d,\lambda)$-graph for which $\lambda \le d\big(d/n\big)^M,$ then $G$ contains a copy of any graph $H$ on at most $n$ vertices with maximum degree $\Delta(H)\le \Delta$. This can be considered as a sparse analog of the Blow-up Lemma. \subsection{Hamiltonicity} A {\em Hamilton cycle} in a graph is a cycle passing through all the vertices of this graph. A graph is called {\em Hamiltonian} if it has at least one Hamilton cycle. For background information on Hamiltonian cycles the reader can consult a survey of Chv\'atal \cite{Chv85}. The notion of Hamilton cycles is one of the most central in modern Graph Theory, and many efforts have been devoted to obtain sufficient conditions for Hamiltonicity. The absolute majority of such known conditions (for example, the famous theorem of Dirac asserting that a graph on $n$ vertices with minimal degree at least $n/2$ is Hamiltonian) deal with graphs which are fairly dense. Apparently there are very few sufficient conditions for the existence of a Hamilton cycle in sparse graphs. As it turns out spectral properties of graphs can supply rather powerful sufficient conditions for Hamiltonicity. Here is one such result, quite general and yet very simple to prove, given our knowledge of properties of pseudo-random graphs. \begin{prop}\label{Ham1} Let $G$ be an $(n,d,\lambda)$-graph. If $$ d-36\frac{\lambda^2}{d} \ge \frac{\lambda n}{d+\lambda}\,, $$ then $G$ is Hamiltonian. \end{prop} \noindent{\bf Proof.\ } According to Theorem \ref{connectivity} $G$ is $(d-36\lambda^2/d)$-vertex-connected. Also, $\alpha(G)\le \lambda n/(d+\lambda)$, as stated in Proposition \ref{ind-set}. Finally, a theorem of Chv\'atal and Erd\H os \cite{ChvErd72} asserts that if the vertex-connectivity of a graph $G$ is at least as large as its independence number, then $G$ is Hamiltonian.\hfill $\Box$ \medskip The Chv\'atal-Erd\H os Theorem has also been used by Thomason in \cite{Tho87a}, who proved that a $(p,\alpha)$-jumbled graph $G$ with minimal degree $\delta(G)=\Omega(\alpha/p)$ is Hamiltonian. His proof is quite similar in spirit to that of the above proposition. Assuming that $\lambda=o(d)$ and $d\rightarrow\infty$, the condition of Proposition \ref{Ham1} reads then as: $\lambda\le (1-o(1))d^2/n$. For best possible pseudo-random graphs, where $\lambda=\Theta(\sqrt{d})$, this condition starts working when $d=\Omega(n^{2/3})$. One can however prove a much stronger asymptotical result, using more sophisticated tools for assuring Hamiltonicity. The authors prove such a result in \cite{KriSud02}: \begin{theo}\label{Ham2}\cite{KriSud02} Let $G$ be an $(n,d,\lambda)$-graph. If $n$ is large enough and $$ \lambda \leq \frac{(\log\log n)^2}{1000\log n(\log\log\log n)}\,d\,, $$ then $G$ is Hamiltonian. \end{theo} The proof of Theorem \ref{Ham2} is quite involved technically. Its main instrument is the famous rotation-extension technique of Posa \cite{Pos76}, or rather a version of it developed by Koml\'os and Szemer\'edi in \cite{KomSze83} to obtain the exact threshold for the appearance of a Hamilton cycle in the random graph $G(n,p)$. We omit the proof details here, referring the reader to \cite{KriSud02}. For reasonably good pseudo-random graphs, in which $\lambda\le d^{1-\epsilon}$ for some $\epsilon>0$, Theorem \ref{Ham2} starts working already when the degree $d$ is only polylogarithmic in $n$ -- quite a progress compared to the easy Proposition \ref{Ham1}! It is possible though that an even stronger result is true as given by the following conjecture: \begin{conj}\label{cHam}\cite{KriSud02} There exists a positive constant $C$ such that for large enough $n$, any $(n,d,\lambda)$-graph that satisfies $d/ \lambda>C$ contains a Hamilton cycle. \end{conj} This conjecture is closely related to another well known problem on Hamiltonicity. The {\em toughness} $t(G)$ of a graph $G$ is the largest real $t$ so that for every positive integer $x \geq 2$ one should delete at least $tx$ vertices from $G$ in order to get an induced subgraph of it with at least $x$ connected components. $G$ is $t$-tough if $t(G) \geq t$. This parameter was introduced by Chv\'atal in \cite{Chv73}, where he observed that Hamiltonian graphs are $1$-tough and conjectured that $t$-tough graphs are Hamiltonian for large enough $t$. Alon showed in \cite{Alo95} that if $G$ is an $(n,d,\lambda)$-graph, then the toughness of $G$ satisfies $t(G)>\Omega(d/\lambda)$. Therefore the conjecture of Chv\'atal implies the above conjecture. Krivelevich and Sudakov used Theorem \ref{Ham2} in \cite{KriSud02} to derive Hamiltonicity of sparse random Cayley graphs. Given a group $G$ of order $n$, choose a set $S$ of $s$ non-identity elements uniformly at random and form a Cayley graph $\Gamma(G,S\cup S^{-1})$ (see Example 8 in Section 3 for the definition of a Cayley graph). The question is how large should be the value of $t=t(n)$ so as to guarantee the almost sure Hamiltonicity of the random Cayley graph no matter which group $G$ we started with. \begin{theo}\label{Ham3}\cite{KriSud02} Let $G$ be a group of order $n$. Then for every $c>0$ and large enough $n$ a Cayley graph $X(G,S\cup S^{-1})$, formed by choosing a set $S$ of $c\log^5 n$ random generators in $G$, is almost surely Hamiltonian. \end{theo} \noindent{\bf Sketch of proof.\ } Let $\lambda$ be the second largest by absolute value eigenvalue of $X(G,S)$. Note that the Cayley graph $X(G,S)$ is $d$-regular for $d \geq c\log^5 n$. Therefore to prove Hamiltonicity of $X(G,S)$, by Theorem \ref{Ham2} it is enough to show that almost surely $\lambda/d \leq O(\log n)$. This can be done by applying an approach of Alon and Roichman \cite{AloRoi94} for bounding the second eigenvalue of a random Cayley graph.\hfill$\Box$ \medskip We note that a well known conjecture claims that every connected Cayley graph is Hamiltonian. If true the conjecture would easily imply that as few as $O(\log n)$ random generators are enough to give almost sure connectivity and thus Hamiltonicity. \subsection{Random subgraphs of pseudo-random graphs} There is a clear tendency in recent years to study random graphs different from the classical by now model $G(n,p)$ of binomial random graphs. One of the most natural models for random graphs, directly generalizing $G(n,p)$, is defined as follows. Let $G=(V,E)$ be a graph and let $0<p<1$. The {\em random subgraph} $G_p$ if formed by choosing every edge of $G$ independently and with probability $p$. Thus, when $G$ is the complete graph $K_n$ we get back the probability space $G(n,p)$. In many cases the obtained random graph $G_p$ has many interesting and peculiar features, sometimes reminiscent of those of $G(n,p)$, and sometimes inherited from those of the host graph $G$. In this subsection we report on various results obtained on random subgraphs of pseudo-random graphs. While studying this subject, we study in fact not a single probability space, but rather a family of probability spaces, having many common features, guaranteed by those of pseudo-random graphs. Although several results have already been achieved in this direction, overall it is much less developed than the study of binomial random graphs $G(n,p)$, and one can certainly expect many new results on this topic to appear in the future. We start with Hamiltonicity of random subgraphs of pseudo-random graphs. As we learned in the previous section spectral condition are in many cases sufficient to guarantee Hamiltonicity. Suppose then that a host graph $G$ is a Hamiltonian $(n,d,\lambda)$-graph. How small can the edge probability $p=p(n)$ be chosen so as to guarantee almost sure Hamiltonicity of the random subgraph $G_p$? This question has been studied by Frieze and the first author in \cite{FriKri02}. They obtained the following result. \begin{theo}\label{FK}\cite{FriKri02} Let $G$ be an $(n,d,\lambda)$-graph. Assume that $\lambda=o\left(\frac{d^{5/2}}{n^{3/2}(\log n)^{3/2}}\right)$. Form a random subgraph $G_p$ of $G$ by choosing each edge of $G$ independently with probability $p$. Then for any function $\omega(n)$ tending to infinity arbitrarily slowly: \begin{enumerate} \item if $p(n)=\frac{1}{d}(\log n+\log\log n-\omega(n))$, then $G_p$ is almost surely not Hamiltonian; \item if $p(n)=\frac{1}{d}(\log n+\log\log n+\omega(n))$, then $G_p$ is almost surely Hamiltonian. \end{enumerate} \end{theo} Just as in the case of $G(n,p)$ (see, e.g. \cite{Bol01}) it is quite easy to predict the critical probability for the appearance of a Hamilton cycle in $G_p$. An obvious obstacle for its existence is a vertex of degree at most one. If such a vertex almost surely exists in $G_p$, then $G_p$ is almost surely non-Hamiltonian. It is a straightforward exercise to show that the smaller probability in the statement of Theorem \ref{FK} gives the almost sure existence of such a vertex. The larger probability can be shown to be sufficient to eliminate almost surely all vertices of degree at most one in $G_p$. Proving that this is sufficient for almost sure Hamiltonicity is much harder. Again as in the case of $G(n,p)$ the rotation-extension technique of Posa \cite{Pos76} comes to our rescue. We omit technical details of the proof of Theorem \ref{FK}, referring the reader to \cite{FriKri02}. One of the most important events in the study of random graphs was the discovery of the sudden appearance of the giant component by Erd\H os and R\'enyi \cite{ErdRen60}. They proved that all connected components of $G(n,c/n)$ with $0<c<1$ are almost surely trees or unicyclic and have size $O(\log n)$. On the other hand, if $c>1$, then $G(n,c/n)$ contains almost surely a unique component of size linear in $n$ (the so called {\em giant component}), while all other components are at most logarithmic in size. Thus, the random graph $G(n,p)$ experiences the so called {\em phase transition} at $p=1/n$. Very recently Frieze, Krivelevich and Martin showed \cite{FriKriMar02} that a very similar behavior holds for random subgraphs of many pseudo-random graphs. To formulate their result, for $\alpha>1$ we define $\bar{\alpha}<1$ to be the unique solution (other than $\alpha$) of the equation $xe^{-x}=\alpha e^{-\alpha}$. \begin{theo}\label{FKM}\cite{FriKriMar02} Let $G$ be an $(n,d,\lambda)$-graph. Assume that $ \lambda=o(d). $ Consider the random subgraph $G_{\alpha/d}$, formed by choosing each edge of $G$ independently and with probability $p=\alpha/d$. Then: \begin{itemize} \item[(a)] If $\alpha<1$ then almost surely the maximum component size is $O(\log n)$. \item[(b)] If $\alpha>1$ then almost surely there is a unique giant component of asymptotic size $\left(1-\frac{\bar{\alpha}}{\alpha}\right)n$ and the remaining components are of size $O(\log n)$. \end{itemize} \end{theo} Let us outline briefly the proof of Theorem \ref{FKM}. First, bound (\ref{eig}) and known estimates on the number of $k$-vertex trees in $d$-regular graphs are used to get estimates on the expectation of the number of connected components of size $k$ in $G_{p}$, for various values of $k$. Using these estimates it is proved then that almost surely $G_p$ has no connected components of size between $(1/\alpha\gamma)\log n$ and $\gamma n$ for a properly chosen $\gamma=\gamma(\alpha)$. Define $f(\alpha)$ to be 1 for all $\alpha\le 1$, and to be $\bar{\alpha}/\alpha$ for $\alpha>1$. One can show then that almost surely in $G_{\alpha/d}$ the number of vertices in components of size between 1 and $d^{1/3}$ is equal to $nf(\alpha)$ up to the error term which is $O(n^{5/6}\log n)$. This is done by first calculating the expectation of the last quantity, which is asymptotically equal to $nf(\alpha)$, and then by applying the Azuma-Hoeffding martingale inequality. Given the above, the proof of Theorem \ref{FKM} is straightforward. For the case $\alpha<1$ we have $nf(\alpha)=n$ and therefore all but at most $n^{5/6}\log n$ vertices lie in components of size at most $(1/\alpha\gamma)\log n$. The remaining vertices should be in components of size at least $\gamma n$, but there is no room for such components. If $\alpha>1$, then $(\bar{\alpha}/\alpha)n+O(n^{5/6}\log n)$ vertices belong to components of size at most $(1/\alpha\gamma)\log n$, and all remaining vertices are in components of size at least $\gamma n$. These components are easily shown to merge quickly into one giant component of a linear size. The detail can be found in \cite{FriKriMar02} (see also \cite{ABS} for some related results). One of the recent most popular subjects in the study of random graphs is proving sharpness of thresholds for various combinatorial properties. This direction of research was spurred by a powerful theorem of Friedgut-Bourgain \cite{Fri99}, providing a sufficient condition for the sharpness of a threshold. The authors together with Vu apply this theorem in \cite{KriSudVu02} to show sharpness of graph connectivity, sometimes also called {\em network reliability}, in random subgraphs of a wide class of graphs. Here are the relevant definitions. For a connected graph $G$ and edge probability $p$ denote by $f(p)=f(G,p)$ the probability that a random subgraph $G_p$ is connected. The function $f(p)$ can be easily shown to be strictly monotone. For a fixed positive constant $x \leq 1$ and a graph $G$, let $p_{x} $ denote the (unique) value of $p$ where $f(G, p_{x})= x$. We say that a family $(G_i)_{i=1}^{\infty}$ of graphs satisfies the {\em sharp threshold} property if for any fixed positive $\epsilon \le 1/2$ $$ \lim_{i \rightarrow \infty} \frac{ p_{\epsilon} (G_i)} { p_{1-\epsilon} (G_i)} \rightarrow 1. $$ Thus the threshold for connectivity is sharp if the width of the transition interval is negligible compared to the critical probability. Krivelevich, Sudakov and Vu proved in \cite{KriSudVu02} the following theorem. \begin{theo}\cite{KriSudVu02}\label{KSV} Let $(G_i)_{i=1}^{\infty}$ be a family of distinct graphs, where $G_i$ has $n_i$ vertices, maximum degree $d_i$ and it is $k_i$-edge-connected. If $$ \lim_{i \rightarrow \infty} \frac{k_i \ln n_i}{d_i}=\infty, $$ then the family $(G_i)_{i=1}^{\infty}$ has a sharp connectivity threshold. \end{theo} The above theorem extends a celebrated result of Margulis \cite{Mar74} on network reliability (Margulis' result applies to the case where the critical probability is a constant). Since $(n,d,\lambda)$ graphs are $d(1-o(1))$-connected as long as $\lambda=o(d)$ by Theorem \ref{connectivity}, we immediately get the following result on the sharpness of the connectivity threshold for pseudo-random graphs. \begin{coro}\label{KSVpr} Let $G$ be an $(n,d,\lambda)$-graph. If $\lambda=o(d)$, then the threshold for connectivity in the random subgraph $G_p$ is sharp. \end{coro} Thus already weak connectivity is sufficient to guarantee sharpness of the threshold. This result has potential practical applications as discussed in \cite{KriSudVu02}. Finally we consider a different probability space created from a graph $G=(V,E)$. This space is obtained by putting random weights on the edges of $G$ independently. One can then ask about the behavior of optimal solutions for various combinatorial optimization problems. Beveridge, Frieze and McDiarmid treated in \cite{BevFriMcD98} the problem of estimating the weight of a random minimum length spanning tree in regular graphs. For each edge $e$ of a connected graph $G=(V,E)$ define the length $X_e$ of $e$ to be a random variable uniformly distributed in the interval $(0,1)$, where all $X_e$ are independent. Let $mst(G,\bf{X})$ denote the minimum length of a spanning tree in such a graph, and let $mst(G)$ be the expected value of $mst(G,{\bf X})$. Of course, the value of $mst(G)$ depends on the connectivity structure of the graph $G$. Beveridge et al. were able to prove however that if the graph $G$ is assumed to be almost regular and has a modest edge expansion, then $mst(G)$ can be calculated asymptotically: \begin{theo}\cite{BevFriMcD98}\label{BFM} Let $\alpha=\alpha(d)=O(d^{-1/3})$ and let $\rho(d)$ and $\omega(d)$ tend to infinity with $d$. Suppose that the graph $G=(V,E)$ satisfies $$ d\le d(v)\le (1+\alpha)d\quad\mbox{for all $v\in V(G)$}\,, $$ and $$ \frac{e(S,V\setminus S)}{|S|}\ge \omega d^{2/3}\log d \quad\mbox{for all $S\subset V$ with $d/2<|S|\le min\{\rho d,|V|/2\}$}\ . $$ Then $$ mst(G)=(1+o(1))\frac{|V|}{d}\zeta(3)\ , $$ where the $o(1)$ term tends to 0 as $d\rightarrow\infty$, and $\zeta(3)=\sum_{i=1}^{\infty}i^{-3}=1.202...$. \end{theo} The above theorem extends a celebrated result of Frieze \cite{Fri85}, who proved it in the case of the complete graph $G=K_n$. Pseudo-random graphs supply easily the degree of edge expansion required by Theorem \ref{BFM}. We thus get: \begin{coro}\label{BFMpr} Let $G$ be an $(n,d,\lambda)$-graph. If $\lambda=o(d)$ then $$ mst(G)=(1+o(1))\frac{n}{d}\zeta(3)\ . $$ \end{coro} Beveridge, Frieze and McDiarmid also proved that the random variable $mst(G,{\bf X})$ is sharply concentrated around its mean given by Theorem \ref{BFM}. Comparing between the very well developed research of binomial random graphs $G(n,p)$ and few currently available results on random subgraphs of pseudo-random graphs, we can say that many interesting problems in the latter subject are yet to be addressed, such as the asymptotic behavior of the independence number and the chromatic number, connectivity, existence of matchings and factors, spectral properties, to mention just a few. \subsection{Enumerative aspects} Pseudo-random graphs on $n$ vertices with edge density $p$ are quite similar in many aspects to the random graph $G(n,p)$. One can thus expect that counting statistics in pseudo-random graphs will be close to those in truly random graphs of the same density. As the random graph $G(n,p)$ is a product probability space in which each edge behaves independently, computing the expected number of most subgraphs in $G(n,p)$ is straightforward. Here are just a few examples: \begin{itemize} \item The expected number of perfect matchings in $G(n,p)$ is $\frac{n!}{(n/2)!2^{n/2}}p^{n/2}$ (assuming of course that $n$ is even); \item The expected number of spanning trees in $G(n,p)$ is $n^{n-2}p^{n-1}$; \item The expected number of Hamilton cycles in $G(n,p)$ is $\frac{(n-1)!}{2}p^n$. \end{itemize} In certain cases it is possible to prove that the actual number of subgraphs in a pseudo-random graph on $n$ vertices with edge density $p=p(n)$ is close to the corresponding expected value in the binomial random graph $G(n,p)$. Frieze in \cite{Fri00} gave estimates on the number of perfect matchings and Hamilton cycles in what he calls super $\epsilon$-regular graphs. Let $G=(V,E)$ be a graph on $n$ vertices with ${n\choose 2}p$ edges, where $0<p<1$ is a constant. Then $G$ is called {\em super $(p,\epsilon)$-regular}, for a constant $\epsilon>0$, if \begin{enumerate} \item For all vertices $v\in V(G)$, $$ (p-\epsilon)n \le d(v)\le (p+\epsilon)n\,; $$ \item For all $U,W\subset V$, $U\cap W=\emptyset$, $|U|, |W|\ge \epsilon n$, $$ \left|\frac{e(U,W)}{|U||W|}-p\right|\le \epsilon\ . $$ \end{enumerate} Thus, a super $(p,\epsilon)$-regular graph $G$ can be considered a non-bipartite analog of the notion of a super-regular pair defined above. In our terminology, $G$ is a weakly pseudo-random graph of constant density $p$, in which {\em all} degrees are asymptotically equal to $pn$. Assume that $n=2\nu$ is even. Let $m(G)$ denote the number of perfect matchings in $G$ and let $h(G)$ denote the number of Hamilton cycles in $G$, and let $t(G)$ denote the number of spanning trees in $G$. \begin{theo}\label{F}\cite{Fri00} If $\epsilon$ is sufficiently small and $n$ is sufficiently large then \begin{description} \item[{\bf (a)}] $$ (p-2\epsilon)^{\nu}\frac{n!}{\nu!2^{\nu}}\le m(G) \le (p+2\epsilon)^{\nu}\frac{n!}{\nu!2^{\nu}}\ ; $$ \item [{\bf (b)}] $$ (p-2\epsilon)^nn!\le h(G)\le (p+2\epsilon)^nn!\ ; $$ \end{description} \end{theo} Theorem \ref{F} thus implies that the numbers of perfect matchings and of Hamilton cycles in super $\epsilon$-regular graphs are quite close asymptotically to the expected values of the corresponding quantities in the random graph $G(n,p)$. Part (b) of Theorem \ref{F} improves significantly Corollary 2.9 of Thomason \cite{Tho87a} which estimates from below the number of Hamilton cycles in jumbled graphs. Here is a very brief sketch of the proof of Theorem \ref{F}. To estimate the number of perfect matchings in $G$, Frieze takes a random partition of the vertices of $G$ into two equal parts $A$ and $B$ and estimates the number of perfect matchings in the bipartite subgraph of $G$ between $A$ and $B$. This bipartite graph is almost surely super $2\epsilon$-regular, which allows to apply bounds previously obtained by Alon, R\"odl and Ruci\'nski \cite{AloRodRuc98} for such graphs. Since each Hamilton cycle is a union of two perfect matchings, it follows immediately that $h(G)\le m^2(G)/2$, establishing the desired upper bound on $h(G)$. In order to prove a lower bound, let $f_k$ be the number of 2-factors in $G$ containing exactly $k$ cycles, so that $f_1=h(G)$. Let also $A$ be the number of ordered pairs of edge disjoint perfect matchings in $G$. Then \begin{equation}\label{A} A= \sum_{i=1}^{\lfloor n/3\rfloor}2^kf_k\ . \end{equation} For a perfect matching $M$ in $G$ let $a_M$ be the number of perfect matchings of $G$ disjoint from $M$. Since deleting $M$ disturbs $\epsilon$-regularity of $G$ only marginally, one can use part (a) of the theorem to get $a_M\ge (p-2\epsilon)^{\nu}\frac{n!}{\nu!2^{\nu}}$. Thus \begin{equation}\label{A1} A=\sum_{M\in G}a_M\ge \left((p-2\epsilon)^{\nu}\frac{n!}{\nu!2^{\nu}} \right)^2 \ge (p-2\epsilon)^nn!\cdot \frac{1}{3n^{1/2}}\ . \end{equation} Next Frieze shows that the ratio $f_{k+1}/f_k$ can be bounded by a polynomial in $n$ for all $1\le k\le k_1=O(p^{-2})$, $f_k\le 5^{-(k-k_0)/2}\max\{f_{k_0+1},f_{k_0}\}$ for all $k\ge k_0+2, k_0=\Theta(p^{-3}\log n)$ and that the ratio $(f_{k_1+1}+\ldots+f_{\lfloor n/3\rfloor})/f_{k_1}$ is also bounded by a polynomial in $n$. Then from (\ref{A}), $A\le O_p(1)\sum_{k=1}^{k_0+1}f_k$ and thus $A\le n^{O(1)}f_1$. Plugging (\ref{A1}) we get the desired lower bound. One can also show (see \cite {Noga}) that the number of spanning trees $t(G)$ in super $(p,\epsilon)$-regular graphs satisfies: $$ (p-2\epsilon)^{n-1}n^{n-2}\le t(G)\le (p+2\epsilon)^{n-1}n^{n-2}\ , $$ for small enough $\epsilon>0$ and large enough $n$. In order to estimate from below the number of spanning trees in $G$, consider a random mapping $f:V(G)\rightarrow V(G)$, defined by choosing for each $v\in V$ its neighbor $f(v)$ at random. Each such $f$ defines a digraph $D_f=(V,A_f)$, $A_f=\{(v,f(v)): v\in V\}$. Each component of $D_f$ consists of cycle $C$ with a rooted forest whose roots are all in $C$. Suppose that $D_f$ has $k_f$ components. Then a spanning tree of $G$ can be obtained by deleting the lexicographically first edge of each cycle in $D_f$, and then extending the $k_f$ components to a spanning tree. Showing that $D_f$ has typically $O(\sqrt{n})$ components implies that most of the mappings $f$ create a digraph close to a spanning tree of $G$, and therefore: $$ t(G)\ge n^{-O(\sqrt{n})}|f:V\rightarrow V|\ge n^{-O(\sqrt{n})}(p-\epsilon)n^n\ . $$ For the upper bound on $t(G)$ let $\Omega^*=\{f:V\rightarrow V: (v,f(v))\in E(G)$ for $v\not =1$ and $f(1)=1 \}$. Then $$ t(G)\le |\Omega^*| \le \big((p+\epsilon)n\big)^{n-1}\le (p+2\epsilon)^{n-1}n^{n-2}\ . $$ To see this consider the following injection from the spanning trees of $G$ into $\Omega^*$: orient each edge of a tree $T$ towards vertex 1 and set $f(1)=1$. Note that this proof does not use the fact that the graph is pseudo-random. Surprisingly it shows that all nearly regular connected graphs with the same density have approximately the same number of spanning trees. For sparse pseudo-random graphs one can use Theorem \ref{FK} to estimate the number of Hamilton cycles. Let $G$ be an $(n,d,\lambda)$-graph satisfying the conditions of Theorem \ref{FK}. Consider the random subgraph $G_p$ of $G$, where $p=(\log n+2\log\log n)/d$. Let $X$ be the random variable counting the number of Hamilton cycles in $G_p$. According to Theorem \ref{FK}, $G_p$ has almost surely a Hamilton cycle, and therefore $E[X]\ge 1-o(1)$. On the other hand, the probability that a given Hamilton cycle of $G$ appears in $G_p$ is exactly $p^n$. Therefore the linearity of expectation implies $E[X]=h(G)p^n$. Combining the above two estimates we derive: $$ h(G)\ge \frac{1-o(1)}{p^n}=\left(\frac{d}{(1+o(1))\log n}\right)^n \ . $$ We thus get the following corollary: \begin{coro}\cite{FriKri02}\label{FKc} Let $G$ be an $(n,d,\lambda)$-graph with $\lambda=o(d^{5/2}/(n^{3/2}(\log n)^{3/2}))$. Then $G$ contains at least $\left(\frac{d}{(1+o(1))\log n}\right)^n$ Hamilton cycles. \end{coro} Note that the number of Hamilton cycles in any $d$-regular graph on $n$ vertices obviously does not exceed $d^n$. Thus for graphs satisfying the conditions of Theorem \ref{FK} the above corollary provides an asymptotically tight estimate on the exponent of the number of Hamilton cycles. \section{Conclusion} Although we have made an effort to provide a systematic coverage of the current research in pseudo-random graphs, there are certainly quite a few subjects that were left outside this survey, due to the limitations of space and time (and of the authors' energy). Probably the most notable omission is a discussion of diverse applications of pseudo-random graphs to questions from other fields, mostly Extremal Graph Theory, where pseudo-random graphs provide the best known bounds for an amazing array of problems. We hope to cover this direction in one of our future papers. Still, we would like to believe that this survey can be helpful in mastering various results and techniques pertaining to this field. Undoubtedly many more of them are bound to appear in the future and will make this fascinating subject even more deep, diverse and appealing. \medskip \noindent{\bf Acknowledgment.\ } The authors would like to thank Noga Alon for many illuminating discussions and for kindly granting us his permission to present his Theorem \ref{number-subgraphs} here. The proofs of Theorems \ref{connectivity}, \ref{edge-connectivity} were obtained in discussions with him.
{ "timestamp": "2005-03-31T18:59:47", "yymm": "0503", "arxiv_id": "math/0503745", "language": "en", "url": "https://arxiv.org/abs/math/0503745", "abstract": "Random graphs have proven to be one of the most important and fruitful concepts in modern Combinatorics and Theoretical Computer Science. Besides being a fascinating study subject for their own sake, they serve as essential instruments in proving an enormous number of combinatorial statements, making their role quite hard to overestimate. Their tremendous success serves as a natural motivation for the following very general and deep informal questions: what are the essential properties of random graphs? How can one tell when a given graph behaves like a random graph? How to create deterministically graphs that look random-like? This leads us to a concept of pseudo-random graphs and the aim of this survey is to provide a systematic treatment of this concept.", "subjects": "Combinatorics (math.CO)", "title": "Pseudo-random graphs", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.978384668497878, "lm_q2_score": 0.8376199572530448, "lm_q1q2_score": 0.819514524204227 }
https://arxiv.org/abs/1306.2385
Linear groups - Malcev's theorem and Selberg's lemma
An account of two fundamental facts concerning finitely generated linear groups: Malcev's theorem on residual finiteness, and Selberg's lemma on virtual torsion-freeness.
\section*{Introduction} A group is \textbf{linear} if it is (isomorphic to) a subgroup of $\mathrm{GL}_n(K)$, where $K$ is a field. If we want to specify the field, we say that the group is linear over $K$. The following theorems are fundamental, at least from the perspective of combinatorial group theory. \begin{thm*}[Malcev 1940] A finitely generated linear group is residually finite. \end{thm*} \begin{thm*}[Selberg 1960] A finitely generated linear group over a field of zero characteristic is virtually torsion-free. \end{thm*} A group is \textbf{residually finite} if its elements are distinguished by the finite quotients of the group, i.e., if each non-trivial element of the group remains non-trivial in a finite quotient. A group is \textbf{virtually torsion-free} if some finite-index subgroup is torsion-free. As a matter of further terminology, Selberg's theorem is usually referred to as Selberg's lemma, and Malcev is alternatively transliterated as Mal'cev or Maltsev. \begin{content} The main body of this text has three sections. In the first one we discuss residual finiteness and virtual torsion-freeness, with emphasis on their relation to a third property - roughly speaking, a ``$p$-adic'' refinement of residual finiteness. The main theorem we are actually interested in, due to Platonov (1968), gives such refined residual properties for finitely generated linear groups. Both Malcev's theorem and Selberg's lemma are consequences of this more powerful, but lesser known, theorem of Platonov. The second section is devoted to $\mathrm{SL}_n(\mathbb{Z})$. This example is too important, too interesting, too much fun to receive anything less than a scenic analysis. In the last section we return to a proof of Platonov's theorem. \end{content} \begin{comments} Besides the Russian original \cite{P}, the only other source in the literature for Platonov's theorem that I am aware of is the account of Wehrfritz in \cite[Chapter 4]{W}. The proof presented herein is, I think, considerably simpler. It is mainly influenced by the discussion of Malcev's theorem in the lecture notes by Stallings \cite{S}, and it is quite similar to Platonov's own arguments in \cite{P}. An alternative road to Selberg's lemma is to use valuations. This is the approach taken by Cassels in \emph{Local fields} (Cambridge University Press 1986), and by Ratcliffe in \emph{Foundations of hyperbolic manifolds} (2nd edition, Springer 2006). I thank, in chronological order, Andy Putman for a useful answer via \url{mathoverflow.net}, Jean-Fran\c{c}ois Planchat for a careful reading and constructive comments, and Vadim Alekseev for a translation of Platonov's article. \end{comments} \begin{convention} In this text, rings are commutative and unital. \end{convention} \section{Virtual and residual properties of groups}\label{VR} Virtual torsion-freeness and residual finiteness are instances of the following terminology. Let $\mathcal{P}$ be a group-theoretic property. A group is \textbf{virtually $\mathcal{P}$} if it has a finite-index subgroup enjoying $\mathcal{P}$. A group is \textbf{residually $\mathcal{P}$} if each non-trivial element of the group remains non-trivial in some quotient group enjoying $\mathcal{P}$. The virtually $\mathcal{P}$ groups and the residually $\mathcal{P}$ groups contain the $\mathcal{P}$ groups. It may certainly happen that a property is virtually stable (e.g., finiteness) or residually stable (e.g., torsion-freeness). Besides virtual torsion-freeness and residual finiteness, we are interested in the hybrid notion of \textbf{virtual residual $p$-finiteness} where $p$ is a prime. This is obtained by residualizing the property of being a finite $p$-group, followed by the virtual extension. The notion of virtual residual $p$-finiteness has, in fact, the leading role in this account for it relates both to residual finiteness and to virtual torsion-freeness. \begin{lem}[``Going down''] If $\mathcal{P}$ is inherited by subgroups, then both virtually $\mathcal{P}$ and residually $\mathcal{P}$ are inherited by subgroups. In particular, virtual torsion-freeness, residual finiteness, and virtual residual $p$-finiteness are inherited by subgroups. \end{lem} \begin{lem}[``Going up'']\label{going up} Virtually $\mathcal{P}$ passes to finite-index supergroups. In particular, both virtual torsion-freeness and virtual residual $p$-finiteness pass to finite-index supergroups. Residual finiteness passes to finite-index supergroups. \end{lem} Observe that residual $p$-finiteness, just like torsion-freeness, is not virtually stable. Residual finiteness does pass to finite-index supergroups because of the following equivalent description: a group is residually finite if and only if every non-trivial element lies outside of a finite-index subgroup. Residual $p$-finiteness trivially implies residual finiteness. Going up, we obtain: \begin{prop}\label{one p suffices} Virtual residual $p$-finiteness for some prime $p$ implies residual finiteness. \end{prop} On the other hand, residual $p$-finiteness imposes torsion restrictions. Namely, in a residually $p$-finite group, the order of a torsion element must be a $p$-th power. Hence, if a group is residually $p$-finite and residually $q$-finite for two different primes $p$ and $q$, then it is torsion-free. Virtualizing this statement, we obtain: \begin{prop}\label{p and q} Virtual residual $p$-finiteness and virtual residual $q$-finiteness for two different primes $p$ and $q$ imply virtual torsion-freeness. \end{prop} In light of Propositions~\ref{one p suffices} and ~\ref{p and q}, we see that Malcev's theorem and Selberg's lemma are consequences of the following: \begin{thm*}[Platonov 1968] Let $G$ be a finitely generated linear group over a field $K$. If $\mathrm{char}\:K=0$, then $G$ is virtually residually $p$-finite for all but finitely many primes $p$. If $\mathrm{char}\:K=p$, then $G$ is virtually residually $p$-finite. \end{thm*} \section{Essellennzee}\label{SL} In this section we examine $\mathrm{SL}_n(\mathbb{Z})$, where $n\geq 2$. We start with a most familiar fact. \begin{prop}\label{SLn f.g.} $\mathrm{SL}_n(\mathbb{Z})$ is generated by the elementary matrices $\{1_n+e_{ij}: i\neq j\}$. \end{prop} \begin{proof} In general, if $A$ is a euclidean domain then $\mathrm{SL}_n(A)$ is generated by the elementary matrices $\{1_n+a\cdot e_{ij}: a\in A, i\neq j\}$. The first step is to turn any matrix in $\mathrm{SL}_n(A)$ into a diagonal one via elementary operations. Using division with remainder, and the euclidean map on $A$ as a way of measuring the decrease in complexity, we can insure that a single non-zero entry, say $a$, remains in the first row. Column swapping brings $a$ in position $(1,1)$, and row reductions using the invertible $a$ turn all other entries in the first column to $0$. Now repeat this procedure for the remaining $(n-1)\times (n-1)$ block. The second step is to bring a diagonal matrix of determinant $1$ to the identity matrix $1_n$ via elementary operations. This is done using the transition \begin{align*} \begin{pmatrix} a & 0 \\ 0 & b \end{pmatrix} \leadsto \begin{pmatrix} a & a \\ 0 & b \end{pmatrix} \leadsto \begin{pmatrix} 1 & a \\ (a^{-1}-1)b & b \end{pmatrix} \leadsto \begin{pmatrix} 1 & a \\ 0 & ab \end{pmatrix} \leadsto \begin{pmatrix} 1 & 0 \\ 0 & ab \end{pmatrix}.\end{align*} Finally, if the additive group of $A$ is generated by $\{a_1,\dots, a_k\}$, then the corresponding matrices $\{1_n+a_1\cdot e_{ij},\dots, 1_n+a_k\cdot e_{ij}: i\neq j\}$ generate all the elementary matrices. \end{proof} Let $N$ be a positive integer. Reduction modulo $N$ defines a group homomorphism $\mathrm{SL}_n(\mathbb{Z})\to \mathrm{SL}_n(\mathbb{Z}/N)$, which enjoys the following remarkable property: \begin{lem} The congruence homomorphism $\mathrm{SL}_n(\mathbb{Z})\to \mathrm{SL}_n(\mathbb{Z}/N)$ is onto. \end{lem} \begin{proof} Since $\mathrm{SL}_n(\mathbb{Z})$ is generated by the elementary matrices, and the elementary matrices are mapped to the elementary matrices by the congruence homomorphism, its surjectivity is equivalent to the elementary generation of $\mathrm{SL}_n(\mathbb{Z}/N)$. The Chinese Remainder Theorem provides a decomposition $\mathbb{Z}/N\simeq \prod \mathbb{Z}/{p_i^{s_i}}$ into local rings, mirroring the decomposition $N=\prod p_i^{s_i}$ into primes. Direct products preserve elementary generation, and we now show that it holds over local rings. This is actually easier than elementary generation for euclidean domains. Let $A$ be a local ring and pick a matrix in $\mathrm{SL}_n(A)$. Some first-row entry is not in $\pi$, the maximal ideal of $A$, so it is invertible in $A$. Column swapping brings this element in the $(1,1)$-position, and then the first row and the first column can be cleared. The rest goes as in the proof of Proposition~\ref{SLn f.g.}. \end{proof} The kernel of the congruence homomorphism \begin{align*}\Gamma(N):=\ker \big( \mathrm{SL}_n(\mathbb{Z})\to \mathrm{SL}_n(\mathbb{Z}/N)\big)=\big\{X\in \mathrm{SL}_n(\mathbb{Z}): X \equiv 1_n \;\textrm{mod}\: N\big\}\end{align*} is the \textbf{principal congruence subgroup of level $N$}. In particular, $\Gamma(1)=\mathrm{SL}_n(\mathbb{Z})$. The principal congruence subgroups are normal, finite-index subgroups of $\mathrm{SL}_n(\mathbb{Z})$. The following lemma provides the formula for their index. \begin{lem}\label{computing the index} The index of $\Gamma(N)$ is given by \begin{align*} [\Gamma(1):\Gamma(N)]=|\mathrm{SL}_n(\mathbb{Z}/N)|=N^{n^2-1}\prod_{p|N}\Big(\prod_{i=2}^{n}(1-p^{-i})\Big). \end{align*} \end{lem} \begin{proof} First, recall that \begin{align*}|\mathrm{SL}_n(\mathbb{Z}/p)|=\frac{1}{p-1}|\mathrm{GL}_n(\mathbb{Z}/p)|=\frac{1}{p-1}\prod_{i=0}^{n-1}(p^n-p^i)=p^{n^2-1}\prod_{i=2}^{n}(1-p^{-i}). \end{align*} Next, we find the size of $\mathrm{SL}_n(\mathbb{Z}/{p^k})$. Consider again the general case of a local ring $A$ with maximal ideal $\pi$. The congruence map $\mathrm{GL}_n(A)\to \mathrm{GL}_n(A/\pi)$ is onto: \emph{any} lift to $A$ of a matrix in $\mathrm{GL}_n(A/\pi)$ has determinant not in $\pi$, i.e., invertible in $A$. Furthermore, the kernel of the congruence map $\mathrm{GL}_n(A)\to \mathrm{GL}_n(A/\pi)$ is $1_n+\mathrm{M}_n(\pi)$ since a matrix congruent to $1_n$ modulo $\pi$ has determinant in $1+\pi$, hence invertible in $A$. Thus, if $A$ is also finite, then \begin{align*}|\mathrm{GL}_n(A)|=|\pi|^{n^2}\cdot |\mathrm{GL}_n(A/\pi)|.\end{align*} Now $|\mathrm{GL}_n|=|\mathrm{GL}_1|\cdot |\mathrm{SL}_n|$ over any ring, and $|\mathrm{GL}_1(A)|=|\pi|\cdot |\mathrm{GL}_1(A/\pi)|$, so \begin{align*} |\mathrm{SL}_n(A)|=|\pi|^{n^2-1}\cdot |\mathrm{SL}_n(A/\pi)|. \end{align*} Returning to the particular case we are interested in, we obtain \begin{align*} |\mathrm{SL}_n(\mathbb{Z}/{p^k})|=p^{(k-1)(n^2-1)} |\mathrm{SL}_n(\mathbb{Z}/{p})|=(p^k)^{n^2-1}\prod_{i=2}^{n}(1-p^{-i}). \end{align*} Finally, the size of $\mathrm{SL}_n(\mathbb{Z}/N)$ is obtained by multiplying the above formula over the prime decomposition of $N$. \end{proof} Bass - Lazard - Serre (Bull. Amer. Math. Soc. 1964) and Mennicke (Ann. Math. 1965) have shown that, for $n\geq 3$, every finite-index subgroup of $\mathrm{SL}_n(\mathbb{Z})$ contains some principal congruence subgroup. This Congruence Subgroup Property is definitely not true for $n=2$. The first one to state this failure was Klein (1880), and proofs were subsequently provided by Fricke (1886) and Pick (1886). It is in fact known by now that it is an exceptional feature for a finite-index subgroup of $\mathrm{SL}_2(\mathbb{Z})$ to contain a principal congruence subgroup. The principal congruence subgroups are organized according to the divisibility of their levels: $\Gamma(M)\supseteq \Gamma(N) \Leftrightarrow M|N$, that is, ``to contain is to divide''. This puts the emphasis on the prime stratum $\{\Gamma(p) : p \textrm{ prime}\}$, and on the descending chains $\{\Gamma(p^k) : k\geq 1\}$ corresponding to each prime $p$. Observe that $\cap_{p}\: \Gamma(p)=\{1_n\}$, and that $\cap_{k}\: \Gamma(p^k)=\{1_n\}$ for each prime $p$, meaning that the elements of $\mathrm{SL}_n(\mathbb{Z})$ can be distinguished both along the prime stratum, as well as along each descending $p$-chain. Thus: \begin{prop} $\mathrm{SL}_n(\mathbb{Z})$ is residually finite. \end{prop} Clearly $\mathrm{SL}_n(\mathbb{Z})$ is not torsion-free. For example, \begin{align*} \begin{pmatrix} 0 & -1\\ 1 & 0\end{pmatrix}, \qquad \begin{pmatrix} 0 & -1\\ 1 & 1\end{pmatrix}\end{align*} are elements of order $4$, respectively $6$, in $\mathrm{SL}_2(\mathbb{Z})$. However, we have: \begin{prop} $\mathrm{SL}_n(\mathbb{Z})$ is virtually torsion-free. \end{prop} This is an immediate consequence of the following fact, due to Minkowski (1887): \begin{lem} $\Gamma(N)$ is torsion-free provided $N\geq 3$. \end{lem} \begin{proof} It suffices to show that $\Gamma (4)$ and $\Gamma(p)$, where $p\geq 3$ is a prime, are torsion-free. Let $p$ be any prime, and assume that $X\in \Gamma(p)$ is a non-trivial element having finite order. Up to replacing $X$ by a power of itself, we may assume that $X^q=1_n$ for a prime $q$. Then \begin{align*} -q (X-1_n)=\sum_{i\geq 2}^q \binom{q}{i} (X-1_n)^i. \end{align*} Let $p^s$, where $s\geq 1$, be the highest power of $p$ dividing all the entries of $X-1_n$. The left hand side of the displayed identity is divisible by at most $p^s$ if $q\neq p$, and by at most $p^{s+1}$ if $q=p$. The right hand side is divisible by $p^{2s}$, and even by $p^{2s+1}$ if $q=p\geq 3$. Hence $q=p=2$ and $s=1$. The conclusion that $p=2$ and $s=1$ means that $\Gamma(2)$ is the only one in the prime stratum which harbours torsion, and that $\Gamma(4)$, its successor in the descending $2$-chain, is free of torsion. The conclusion that $q=2$ means that torsion elements in $\Gamma(2)$ have order a power of $2$. As $X^2\in \Gamma(4)$ whenever $X\in \Gamma(2)$, and $\Gamma(4)$ is torsion-free, it follows that non-trivial torsion elements in $\Gamma(2)$ have order $2$. \end{proof} This lemma can be used to control the torsion spectrum - that is, the possible orders of torsion elements - in $\mathrm{SL}_n(\mathbb{Z})$. Let us illustrate the basic idea in the simplest case, when $n=2$. Given a group homomorphism, the torsion spectra of its domain, kernel, and range are trivially related by $\tau(\mathrm{dom})\subseteq \tau(\mathrm{ker})\cdot \tau(\mathrm{ran})$. In the case of a congruence homomorphism, this reads as $\tau(\mathrm{SL}_2(\mathbb{Z}))\subseteq \tau(\Gamma(N))\cdot \tau(\mathrm{SL}_2(\mathbb{Z}/N))$. If $N=3$ then $\tau(\Gamma(3))=\{1\}$, and it can be checked that $\tau(\mathrm{SL}_2(\mathbb{Z}/3))=\{1,2,3,4,6\}$. Somewhat easier, in fact, is to let $N=2$: then $\tau(\Gamma(2))=\{1,2\}$, and it is immediate that $\tau(\mathrm{SL}_2(\mathbb{Z}/2))=\{1,2,3\}$. We conclude that $\tau(\mathrm{SL}_2(\mathbb{Z}))\subseteq \{1,2,3,4,6\}$. Equality holds, actually, since there are elements of order $4$ and $6$ in $\mathrm{SL}_2(\mathbb{Z})$. The presence of a torsion element with composite order, namely $6$, implies that $\mathrm{SL}_n(\mathbb{Z})$ is not residually $p$-finite for any prime $p$. As with torsion-freeness, this is easily remedied by passing to a finite-index subgroup: \begin{prop} $\mathrm{SL}_n(\mathbb{Z})$ is virtually residually $p$-finite for each prime $p$. \end{prop} More precisely, we show: \begin{lem}\label{residual p-finiteness of PCS} $\Gamma(N)$ is residually $p$-finite for each prime $p$ dividing $N$. \end{lem} \begin{proof} It suffices to prove that, for each prime $p$, $\Gamma(p)$ is residually $p$-finite. To that end, we claim that each successive quotient $\Gamma(p^{k})/\Gamma(p^{k+1})$ in the descending chain $\{\Gamma(p^k) : k\geq 1\}$ is a $p$-group. This is seen most directly by observing that each element in $\Gamma(p^{k})/\Gamma(p^{k+1})$ has order $p$: for any matrix $1_n+p^kX\in \Gamma(p^k)$ we have \begin{align*}(1_n+p^kX)^p=1_n+\sum _{i=1}^p \binom{p}{i} p^{ki}X^i\in \Gamma(p^{k+1}).\end{align*} Another way is to use the formula for the index (Lemma~\ref{computing the index}), which yields that $\Gamma(p^{k})/\Gamma(p^{k+1})$ has size $p^{n^2-1}$. A third, more involved argument shows that each successive quotient $\Gamma(p^{k})/\Gamma(p^{k+1})$ is isomorphic to $(\mathbb{Z}/p,+)^{n^2-1}$. Start by noting that, for any matrix $1_n+p^kX\in \Gamma(p^k)$, we have $1=\det (1_n+p^kX)=1+p^k\:\mathrm{tr}(X)\textrm{ mod } p^{2k}$; in particular, $p$ divides $\mathrm{tr} (X)$. Let $\mathfrak{sl}_n(\mathbb{Z}/p)$ denote the additive group of traceless $n\times n$ matrices over $\mathbb{Z}/p$. Then the map \begin{align*} \phi_k: \Gamma(p^k)\to \mathfrak{sl}_n(\mathbb{Z}/p), \quad \phi_k(1_n+p^kX)= X \textrm{ mod } p\end{align*} is well-defined. Firstly, $\phi_k$ is a homomorphism: for $1_n+p^kX$ and $1_n+p^kY$ in $\Gamma(p^k)$ we have $\phi_k\big((1_n+p^kX)(1_n+p^kY)\big)= X+Y+p^kXY \textrm{ mod } p=X+Y \textrm{ mod } p$. Secondly, the kernel of $\phi_k$ is $\Gamma(p^{k+1})$. Thirdly, we claim that $\phi_k$ is onto. The target group $\mathfrak{sl}_n(\mathbb{Z}/p)$ is generated by the $n^2-n$ off-diagonal matrix units $\{e_{ij}: 1\leq i\neq j \leq n\}$ together with the $n-1$ diagonal differences $\{e_{ii}-e_{i+1\: i+1}: 1\leq i\leq n-1\}$. It is immediate that the off-diagonal matrix units are in the image of $\phi_k$, as $\phi_k(1_n+p^ke_{ij})=e_{ij}$ for $i\neq j$. To obtain the diagonal differences, consider an $n\times n$ matrix having the $2\times 2$-block \begin{align*} \begin{pmatrix} 1+p^k& p^k\\-p^k &1-p^k \end{pmatrix} \end{align*} on the diagonal, all the other non-zero entries being $1$'s along the remaining diagonal slots. This is a matrix in $\Gamma(p^k)$ which is mapped by $\phi_k$ to $e_{ii}-e_{i+1\: i+1} + e_{i\: i+1}-e_{i+1\: i}$. As $e_{i\: i+1}$ and $e_{i+1\: i}$ are in the image of $\phi_k$, the same is true for $e_{ii}-e_{i+1\: i+1}$. \end{proof} \begin{rem}\label{what lurks} Scratch most properties of $\mathrm{SL}_n(\mathbb{Z})$ and you will find a great discrepancy between $\mathrm{SL}_2(\mathbb{Z})$ and $\mathrm{SL}_{n\geq 3}(\mathbb{Z})$ lurking underneath. For the discussion at hand, the difference turns out to be the following: $\mathrm{SL}_2(\mathbb{Z})$ admits finite-index subgroups which are residually $p$-finite for all primes $p$, whereas in $\mathrm{SL}_{n\geq 3}(\mathbb{Z})$ every finite-index subgroup is residually $p$-finite for only finitely many primes $p$. The question which clarifies and sharpens this contrast is whether principal congruence subgroups can be residually $p$-finite for a prime $p$ not dividing the level. In $\mathrm{SL}_2(\mathbb{Z})$, the answer is that $\Gamma(2)$ is residually $p$-finite for $p=2$ only, but $\Gamma(N)$ with $N\geq 3$ is residually $p$-finite for all primes $p$. The exceptional case is due to the $2$-torsion in $\Gamma(2)$. In the higher level case there is no torsion. Now a torsion-free subgroup of $\mathrm{SL}_2(\mathbb{Z})$ is free, since $\mathrm{SL}_2(\mathbb{Z})$ acts on a tree with finite vertex stabilizers and without inversion (see I.\S 4 of Serre's \emph{Trees}, Springer 1980). Thus $\Gamma(N)$ with $N\geq 3$ is free. We may then use mutual abstract embeddings to conclude that $\Gamma(N)$ with $N\geq 3$, in fact every free group, is residually $p$-finite for all primes $p$. In $\mathrm{SL}_{n\geq 3}(\mathbb{Z})$, the answer is that $\Gamma(N)$ is residually $p$-finite if and only if $p$ divides $N$. Once we know this, the Congruence Subgroup Property will imply that each finite-index subgroup of $\mathrm{SL}_{n\geq 3}(\mathbb{Z})$ is residually $p$-finite for only finitely many primes $p$. Now let us justify the answer, specifically the forward implication. The proof hinges on computing the abelianization of $\Gamma (N)$, and this is essentially due to Lee and Szczarba (Invent. Math. 1976). As in the proof of Lemma~\ref{residual p-finiteness of PCS}, there is a well-defined homomorphism \begin{align*} \Gamma(N)\to \mathfrak{sl}_n(\mathbb{Z}/N), \quad 1_n+NX \mapsto X \textrm{ mod } N\end{align*} which is furthermore onto. Thus $\Gamma(N)/\Gamma(N^2)\simeq \mathfrak{sl}_n(\mathbb{Z}/N)\simeq (\mathbb{Z}/N, +)^{n^2-1}$, and the commutator subgroup $[\Gamma(N),\Gamma(N)]$ is contained in $\Gamma(N^2)$. On the other hand, we have $1_n+N^2e_{ik}=[1_n+Ne_{ij},1_n+Ne_{jk}]\in [\Gamma(N),\Gamma(N)]$ for distinct $i,j, k$. At this point we use the fact that the principal congruence subgroup of level $M$ is normally generated by $\{1_n+Me_{ij}: i\neq j\}$, the $M$-th powers of the elementary matrices. This is what Mennicke actually proved in his approach to the Congruence Subgroup Property. As pointed out soon after by Bass - Milnor - Serre (Publ. Math. IHES 1967), this fact is equivalent to the Congruence Subgroup Property. It follows that $\Gamma(N^2)$ is contained in $[\Gamma(N),\Gamma(N)]$, by the normality of $\Gamma(N)$. Summarizing, $[\Gamma(N),\Gamma(N)]=\Gamma(N^2)$, so that the abelianization of $\Gamma (N)$ is $(\mathbb{Z}/N, +)^{n^2-1}$. Finally, if $\Gamma(N)$ maps onto a non-trivial finite $p$-group then the abelianization of $\Gamma(N)$ maps onto the corresponding abelianization, which is a non-trivial $p$-group, and we conclude that $p$ divides $N$. \end{rem} \section{Proof of Platonov's theorem}\label{P} Let $G$ be a finitely generated linear group over a field $K$, say $G\leq \mathrm{GL}_n(K)$. In $K$, consider the subring $A$ generated by the multiplicative identity $1$ and the matrix entries of a finite, symmetric set of generators for $G$. Thus $A$ is a finitely generated domain, and $G\leq \mathrm{GL}_n(A)$. Platonov's theorem is then a consequence of the following: \begin{thm}\label{what we're actually proving} Let $A$ be a finitely generated domain. If $\mathrm{char}\:A=0$, then $\mathrm{GL}_n(A)$ is virtually residually $p$-finite for all but finitely many primes $p$. If $\mathrm{char}\:A=p$, then $\mathrm{GL}_n(A)$ is virtually residually $p$-finite. \end{thm} The proof of Theorem~\ref{what we're actually proving} is a straightforward variation on the example of $\mathrm{SL}_n(\mathbb{Z})$, as soon as we know the following facts: \begin{lem}\label{technical lemma} Let $A$ be a finitely generated domain. Then the following hold: \begin{itemize} \item[i.] $A$ is noetherian. \item[ii.] $\cap_{k}\: I^k =0$ for any ideal $I\neq A$. \item[iii.] If $A$ is a field, then $A$ is finite. \item[iv.] The intersection of maximal ideals of $A$ is $0$. \item[v.] If $\mathrm{char}\:A=0$, then only finitely many primes $p=p\cdot 1$ are invertible in $A$. \end{itemize} \end{lem} Let us postpone the proof of Lemma~\ref{technical lemma} for the moment, and see how to obtain Theorem~\ref{what we're actually proving}. The principal congruence subgroup of $\mathrm{GL}_n(A)$ corresponding to an ideal $I$ of $A$ is defined by \begin{align*}\Gamma(I)=\ker \big( \mathrm{GL}_n(A)\to \mathrm{GL}_n(A/I)\big).\end{align*} If $\pi$ is a maximal ideal then $A/\pi$ is a finite field, by \ref{technical lemma} iii, so $\Gamma(\pi)$ has finite index in $\mathrm{GL}_n(A)$. Also $\cap_{\pi}\: \Gamma(\pi)=\{1_n\}$ as $\pi$ runs over the maximal ideals of $A$, by \ref{technical lemma} iv. This shows that $\mathrm{GL}_n(A)$ is residually finite, thereby proving Malcev's theorem. We claim that $\pi^k/\pi^{k+1}$ is finite for each $k\geq 1$. In general, if $M$ is an $R$-module which is annihilated by an ideal $I$ -- in the sense that $IM=0$ -- then $M$ is also an $R/I$-module in a natural way: namely, define $\overline{r}\cdot m:=r\cdot m$ for $r\in R$ and $m\in M$. Furthermore, if $M$ is finitely generated as an $R$-module then $M$ is finitely generated as an $R/I$-module. In the case at hand, the $A$-module $\pi^k$ is finitely generated since $A$ is noetherian, so the $A$-module $\pi^k/\pi^{k+1}$ is also finitely generated. Therefore $\pi^k/\pi^{k+1}$ is finite dimensional as an $A/\pi$-module. As $A/\pi$ is finite, $\pi^k/\pi^{k+1}$ is finite as well. The ring $A/\pi^k$ is finite for each $k\geq 1$, so each $\Gamma(\pi^k)$ has finite index in $\mathrm{GL}_n(A)$. Furthermore, $\cap_{k}\: \Gamma(\pi^k)=\{1_n\}$ by \ref{technical lemma} ii. (This shows once again that $\mathrm{GL}_n(A)$ is residually finite.) Let $p$ be the characteristic of $A/\pi$, so $p\in \pi$. Then $\Gamma(\pi^{k})/\Gamma(\pi^{k+1})$ is a $p$-group: for $X\in \Gamma(\pi^k)$ we have \begin{align*}X^p=1_n+\sum _{i=1}^p \binom{p}{i} (X-1_n)^i\in \Gamma(\pi^{k+1}).\end{align*} To conclude, $\mathrm{GL}_n(A)$ is virtually residually $p$-finite for each prime $p$ not invertible in $A$. By \ref{technical lemma} v, this happens for all but finitely many primes $p$ in the zero characteristic case. In characteristic $p$, there is only such prime, namely $p$ itself. Theorem~\ref{what we're actually proving} is proved. We now return to the proof of the lemma. \begin{proof}[Proof of Lemma~\ref{technical lemma}] The first two points are standard: i) follows from the Hilbert Basis Theorem, and ii) is the Krull Intersection Theorem for domains. iii) We use the following fact: \begin{quotation} Let $F\subseteq F(u)$ be a field extension with $F(u)$ finitely generated as a ring. Then $F\subseteq F(u)$ is a finite extension and $F$ is finitely generated as a ring. \end{quotation} Here is how we use this fact. Let $F$ be the prime field of $A$ and let $a_1,\dots,a_N$ be generators of $A$ as a ring. Thus $A=F(a_1,\dots,a_N)$. Going down the chain \begin{align*} A=F(a_1,\dots,a_N)\supseteq F(a_1,\dots,a_{N-1})\supseteq \ldots\supseteq F \end{align*} we obtain that $F\subseteq A$ is a finite extension, and that $F$ is finitely generated as a ring. Then $F$ is a finite field, as $\mathbb{Q}$ is not finitely generated as a ring, and so $A$ is finite. Now here is how we prove the fact. Assume that $u$ is transcendental over $F$, i.e., $F(u)$ is the field of rational functions in $u$. Let $P_1/Q_1, \dots, P_N/Q_N$ generate $F(u)$ as a ring, where $P_i, Q_i\in F[u]$. The multiplicative inverse of $1+u\cdot \prod Q_i$ is a polynomial expression in the $P_i/Q_i$'s, which can be written as $R/\prod Q_i^{s_i}$. Therefore $\prod Q_i^{s_i}=(1+u\cdot \prod Q_i)R$ in $F[u]$. But this is impossible, since $\prod Q_i^{s_i}$ is relatively prime to $1+u\cdot \prod Q_i$. Thus $u$ is algebraic over $F$. Let $X^d+\alpha_1X^{d-1}+\dots+\alpha_d$ be the minimal polynomial of $u$ over $F$. Let also $a_1,\dots,a_N$ be ring generators of $F(u)=F[u]$. We may write each $a_i$ as $\sum_{0\leq m\leq d-1} \beta_{i,m} \: u^m$, with $\beta_{i,m} \in F$. We claim that the $\alpha_j$'s and the $\beta_{i,m}$'s are ring generators of $F$. Let $c\in F$. Then $c$ is a polynomial in $a_1,\dots,a_N$ over $F$, hence a polynomial in $u$ over the subring of $F$ generated by the $\beta_{i,m}$'s, hence a polynomial in $u$ of degree less than $d$ over the subring of $F$ generated by the $\alpha_j$'s and the $\beta_{i,m}$'s. By the linear independence of $\{1,u,\dots,u^{d-1}\}$, the latter polynomial is actually of degree $0$. Hence $c$ ends up in the subring of $F$ generated by the $\alpha_j$'s and the $\beta_{i,m}$'s. iv) Let $a\neq 0$ in $A$. To find a maximal ideal of $A$ not containing $a$, we rely on the basic avoidance: maximal ideals do not contain invertible elements. Consider the localization $A'=A[1/a]$. Let $\pi'$ be a maximal ideal in $A'$, so $a\notin \pi'$. The restriction $\pi=\pi'\cap A$ is an ideal in $A$, and $a\notin \pi$. We show that $\pi$ is maximal. The embedding $A\hookrightarrow A'$ induces an embedding $A/\pi\hookrightarrow A'/\pi'$. As $A'/\pi'$ is a field which is finitely generated as a ring, in follows from iii) that $A'/\pi'$ is finite field. Therefore the subring $A/\pi$ is a finite domain, hence a field as well. v) We shall use Noether's Normalization Theorem, which says the following. \begin{quotation} Let $R$ be a finitely generated algebra over a field $F\subseteq R$. Then there are elements $x_1,\dots, x_N\in R$ algebraically independent over $F$ such that $R$ is integral over $F[x_1,\dots,x_N]$. \end{quotation} In our case, $\mathbb{Z}$ is a subring of $A$, and $A$ is an integral domain which is finitely generated as a $\mathbb{Z}$-algebra. Extending to rational scalars, we have that $A_\mathbb{Q}=\mathbb{Q}\otimes_\mathbb{Z} A$ is a finitely generated $\mathbb{Q}$-algebra. By the Normalization Theorem, there exist elements $x_1,\dots, x_N$ in $A_\mathbb{Q}$ which are algebraically independent over $\mathbb{Q}$, and such that $A_\mathbb{Q}$ is integral over $\mathbb{Q}[x_1,\dots,x_N]$. Up to replacing each $x_i$ by an integral multiple of itself, we may assume that $x_1,\dots, x_N$ are in $A$. There is some positive $m\in \mathbb{Z}$ such that each ring generator of $A$ is integral over $\mathbb{Z}[1/m][x_1,\dots, x_N]$. Thus $A[1/m]$ is integral over the subring $\mathbb{Z}[1/m][x_1,\dots, x_N]$. If a prime $p$ is invertible in $A$, then it is also invertible in $A[1/m]$ while at the same time $p\in \mathbb{Z}[1/m][x_1,\dots, x_N]$. Now we use the following general fact. Let $R$ be a ring which is integral over a subring $S$. If $s\in S$ is invertible in $R$, then $s$ is already invertible in $S$. The proof is easy. Let $r\in R$ with $rs=1$. We have $r^d+s_{1}r^{d-1}+\dots+s_{d-1}r+s_d=0$ for some $s_i\in S$, since $r$ is integral over $S$. Multiplying through by $s^{d-1}$ yields $r\in S$. Returning to our proof, we infer that $p$ is invertible in $\mathbb{Z}[1/m][x_1,\dots, x_N]$. By the algebraic independence of $x_1,\dots, x_N$, it follows that $p$ is actually invertible in $\mathbb{Z}[1/m]$. But only finitely many primes have this property, namely the prime factors of $m$. \end{proof} \begin{rem} Let $A$ be an infinite, finitely generated domain with $\mathrm{char}\:A=p>0$. If $n\geq 2$ then the $p$-torsion group $(A,+)$ embeds in $\mathrm{GL}_n(A)$, and this prevents $\mathrm{GL}_n(A)$ from being virtually residually $\ell$-finite for any prime $\ell\neq p$. So we cannot do any better in the positive characteristic case of Theorem~\ref{what we're actually proving}. Selberg's lemma fails in positive characteristic for a similar reason. The elementary group $\mathrm{E}_n(A)=\langle 1_n+a\cdot e_{ij} : a\in A, i\neq j\rangle$ is linear over the fraction field of $A$, and it fails to be virtually torsion-free since it contains copies of the infinite torsion group $(A,+)$. Furthermore, if $n\geq 3$ then $\mathrm{E}_n(A)$ is finitely generated. This is due to the commutator relations $[1_n+a\cdot e_{ij}, 1_n+b\cdot e_{jk}]=1_n+ab\cdot e_{ik}$ for distinct $i,j,k$, which imply that $\mathrm{E}_n(A)$ is generated by $\{1_n+a_1\cdot e_{ij},\dots, 1_n+a_N\cdot e_{ij} : i\neq j\}$ whenever $a_1, \dots, a_N$ are ring generators for $A$. For a concrete example, take $A$ to be the polynomial ring $\mathbb{F}_p[t]$, in which case $\mathrm{E}_n(\mathbb{F}_p[t])=\mathrm{SL}_n(\mathbb{F}_p[t])$ since $\mathbb{F}_p[t]$ is a euclidean domain. \end{rem} \begin{rem} Among finitely generated groups, we have the following implications: \begin{align*} \textrm{linear}\;\Rightarrow\;\textrm{virtually residually $p$-finite for some prime $p$}\;\Rightarrow\;\textrm{residually finite} \end{align*} The first implication, a ``$p$-adic'' refinement of Malcev's theorem, is an immediate consequence of Platonov's theorem. The second implication is Proposition~\ref{one p suffices}. Neither implication can be reversed, as witnessed by the following examples. According to the previous remark, $\mathrm{SL}_n(\mathbb{F}_p[t])$ for $n\geq 3$ is finitely generated and virtually residually $\ell$-finite for $\ell=p$ only. Therefore $\mathrm{SL}_n(\mathbb{F}_{p}[t])\times \mathrm{SL}_n(\mathbb{F}_{q}[t])$, where $p$ and $q$ are different primes, is finitely generated, residually finite but not virtually residually $\ell$-finite for any prime $\ell$. The automorphism group of the free group on $n$ generators, $\mathrm{Aut} (F_{n})$, is virtually residually $p$-finite for all primes $p$. Indeed, as we have seen in Remark~\ref{what lurks}, free groups are residually $p$-finite for all primes $p$. Now a theorem of Lubotzky (J. Algebra 1980) says that $\mathrm{Aut} (G)$ is virtually residually $p$-finite whenever the finitely generated group $G$ is virtually residually $p$-finite. This is the ``$p$-adic'' analogue of an older, simpler, and better known theorem of G. Baumslag (J. London Math Soc. 1963) saying that $\mathrm{Aut} (G)$ is residually finite whenever the finitely generated group $G$ is residually finite. On the other hand, Formanek and Procesi (J. Algebra 1992) have shown that $\mathrm{Aut} (F_n)$ is not linear for $n\geq 3$. \end{rem}
{ "timestamp": "2013-06-12T02:00:49", "yymm": "1306", "arxiv_id": "1306.2385", "language": "en", "url": "https://arxiv.org/abs/1306.2385", "abstract": "An account of two fundamental facts concerning finitely generated linear groups: Malcev's theorem on residual finiteness, and Selberg's lemma on virtual torsion-freeness.", "subjects": "Group Theory (math.GR)", "title": "Linear groups - Malcev's theorem and Selberg's lemma", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9808759610129464, "lm_q2_score": 0.8354835452961425, "lm_q1q2_score": 0.8195057254028574 }
https://arxiv.org/abs/1409.1065
Inequalities via s-convexity and log-convexity
In this paper, we obtain some new inequalities for functions whose second derivatives' absolute value is s-convex and log-convex. Also, we give some applications for numerical integration.
\section{INTRODUCTION} We start with the well-known definition of convex functions: a function f:I\rightarrow \mathbb{R} ,$ $\emptyset \neq I\subset \mathbb{R} ,$ is said to be convex on $I$ if inequalit \begin{equation*} f(tx+(1-t)y)\leq tf(x)+(1-t)f(y) \end{equation* holds for all $x,y\in I$ and $t\in \left[ 0,1\right] .$ In the paper \cite{hudzik}, authors gave the class of functions which are s- $convex in the second sense by the following way. A function $f:[0,\infty )\rightarrow \mathbb{R} $ is said to be $s-$convex in the second sence i \begin{equation*} f(tx+(1-t)y)\leq t^{s}f(x)+(1-t)^{s}f(y) \end{equation* holds for all $x,y\in \lbrack 0,\infty ),t\in \left[ 0,1\right] $ and for some fixed $s\in (0,1].$ The class of $s-$convex functions in the second sense is usually denoted with $K_{s}^{2}.$ Besides in \cite{hudzik}, Hudzik and Maligranda proved that if $s\in \left( 0,1\right) $ $f\in K_{s}^{2}$ implies $f([0,\infty ))\subseteq \lbrack 0,\infty ),$ i.e., they proved that all functions from $K_{s}^{2},$ $s\in \left( 0,1\right) ,$ are nonnegative. \begin{example} (\cite{hudzik}) Let $s\in \left( 0,1\right) $ and $a,b,c\in \mathbb{R} .$ We define function $f:[0,\infty )\rightarrow \mathbb{R} $ as \begin{equation*} f(t)=\left\{ \begin{array}{cc} a, & t=0, \\ bt^{s}+c, & t>0 \end{array \right. \end{equation* It can be easily checked that (i) If $b\geq 0$ and $0\leq c\leq a,$ then $f\in K_{s}^{2},$ (ii) If $b>0$ and $c<0,$ then $f\notin K_{s}^{2}.$ \end{example} Several researchers studied on $s-$convex functions, some of them can be found in \cite{hudzik}-\cite{zeki}. Another kind of convexity is $\log -$convexity that is mentioned in \cite{6} by Niculescu as following. A positive function $f$ is called $\log -$convex on a real interval $I=\left[ a,b\right] $, if for all $x,y\in \left[ a,b\right] $ and $\lambda \in \left[ 0,1\right] $ \begin{equation*} f\left( \lambda x+\left( 1-\lambda \right) y\right) \leq f^{\lambda }\left( x\right) f^{1-\lambda }\left( y\right) . \end{equation*} For recent results for $\log -$convex functions, we refer to readers \cite{1 -\cite{18}. Now, we give a motivated inequality for convex functions: Let $f:I\subset \mathbb{R} \rightarrow \mathbb{R} $ be a convex function on the interval $I$ of real numbers and $a,b\in I$ with $a<b$. The inequalit \begin{equation*} \frac{1}{b-a}\int_{a}^{b}f\left( x\right) dx\leq \frac{1}{2}\left[ f\left( \frac{a+b}{2}\right) +\frac{f\left( a\right) +f\left( b\right) }{2}\right] \end{equation* is known as Bullen's inequality for convex functions \cite{17}, p. 39. We also consider the following useful inequality: Let $f:I\subset \left[ 0,\infty \right] \rightarrow \mathbb{R} $ be a differentiable mapping on $I^{\circ }$, the interior of the interval I$, such that $f^{\prime }\in L\left[ a,b\right] $ where $a,b\in I$ with a<b $. If $\left\vert f^{\prime }\left( x\right) \right\vert \leq M$, then the following inequality holds (see \cite{100}). \begin{equation} \left\vert f(x)-\frac{1}{b-a}\int_{a}^{b}f(u)du\right\vert \leq \frac{M}{b-a \left[ \frac{\left( x-a\right) ^{2}+\left( b-x\right) ^{2}}{2}\right] \label{h.1.1} \end{equation} This inequality is well known in the literature as the Ostrowski inequalit \textit{.}\textbf{\ } The main aim of this paper is to prove some new integral inequalities for s- $convex and $\log -$convex functions by using the integral identity that is obtained by Sar\i kaya and Set in \cite{0}. We also give some applications to our results in numerical integration. Some of our results are similar to the Ostrowski inequality and for special selections of the parameters, we proved some new inequalities of Bullen's type. \section{inequalities for $s-$convex functions} We need the following Lemma which is obtained by Sar\i kaya and Set in \cit {0}, so as to prove our results: \begin{lemma} \label{lem 2.1} Let $f:[a,b]\rightarrow \mathbb{R} $ be an absolutely continuous mapping. Denote by $K(x,.):[a,b]\rightarrow \mathbb{R} $ the kernel given b \begin{equation*} K(x,t)=\left\{ \begin{array}{ccc} \frac{\alpha }{\alpha +\beta }\frac{\left( t-a\right) \left( x-t\right) }{x- }, & & t\in \lbrack a,x] \\ & & \\ -\frac{\beta }{\alpha +\beta }\frac{\left( b-t\right) \left( x-t\right) }{b- }, & & t\in \lbrack x,b \end{array \right. \end{equation* where $\alpha ,\beta \in \mathbb{R} $ nonnegative and not both zero, then the identit \begin{eqnarray*} &&\int_{a}^{b}K(x,t)f^{\prime \prime }(t)dt \\ &=&f(x)+\frac{\alpha f(a)+\beta f(b)}{\alpha +\beta }-\frac{2}{\alpha +\beta }\left[ \frac{\alpha }{x-a}\int_{a}^{x}f(t)dt+\frac{\beta }{b-x \int_{x}^{b}f(t)dt\right] \end{eqnarray* holds. \end{lemma} \begin{theorem} \label{teo 2.1} Let $f:[a,b]\rightarrow \mathbb{R} $ be an absolutely continuous mapping such that $f^{\prime \prime }\in L[a,b].$ If $\left\vert f^{\prime \prime }\right\vert $ is $s-$ convex in the second sense on $[a,b]$ for some fixed $s\in (0,1],$ the \begin{eqnarray*} &&\left\vert f(x)+\frac{\alpha f(a)+\beta f(b)}{\alpha +\beta }-\frac{2} \alpha +\beta }\left[ \frac{\alpha }{x-a}\int_{a}^{x}f(t)dt+\frac{\beta }{b- }\int_{x}^{b}f(t)dt\right] \right\vert \\ &\leq &\frac{\alpha }{\alpha +\beta }\frac{\left( x-a\right) ^{2}}{\left( s+2\right) \left( s+3\right) }\left[ \left\vert f^{\prime \prime }(x)\right\vert +\left\vert f^{\prime \prime }(a)\right\vert \right] \\ &&+\frac{\beta }{\alpha +\beta }\frac{\left( b-x\right) ^{2}}{\left( s+2\right) \left( s+3\right) }\left[ \left\vert f^{\prime \prime }(x)\right\vert +\left\vert f^{\prime \prime }(b)\right\vert \right] \end{eqnarray* holds where $\alpha ,\beta \in \mathbb{R} $ nonnegative and not both zero. \end{theorem} \begin{proof} From Lemma \ref{lem 2.1}, using the property of the modulus and $s-$ convexity of $\left\vert f^{\prime \prime }\right\vert ,$ we can writ \begin{eqnarray*} &&\left\vert f(x)+\frac{\alpha f(a)+\beta f(b)}{\alpha +\beta }-\frac{2} \alpha +\beta }\left[ \frac{\alpha }{x-a}\int_{a}^{x}f(t)dt+\frac{\beta }{b- }\int_{x}^{b}f(t)dt\right] \right\vert \\ &\leq &\int_{a}^{b}\left\vert K(x,t)\right\vert \left\vert f^{\prime \prime }(t)\right\vert dt \\ &\leq &\int_{a}^{x}\frac{\alpha }{\alpha +\beta }\frac{1}{x-a}\left\vert t-a\right\vert \left\vert x-t\right\vert \left\vert f^{\prime \prime }(t)\right\vert dt \\ &&+\int_{x}^{b}\frac{\beta }{\alpha +\beta }\frac{1}{b-x}\left\vert b-t\right\vert \left\vert x-t\right\vert \left\vert f^{\prime \prime }(t)\right\vert dt \\ &=&\frac{\alpha }{\left( \alpha +\beta \right) \left( x-a\right) \int_{a}^{x}\left( t-a\right) \left( x-t\right) \left\vert f^{\prime \prime }\left( \frac{t-a}{x-a}x+\frac{x-t}{x-a}a\right) \right\vert dt \\ &&+\frac{\beta }{\left( \alpha +\beta \right) \left( b-x\right) \int_{x}^{b}\left( b-t\right) \left( t-x\right) \left\vert f^{\prime \prime }\left( \frac{t-x}{b-x}b+\frac{b-t}{b-x}x\right) \right\vert dt \\ &\leq &\frac{\alpha }{\left( \alpha +\beta \right) \left( x-a\right) \int_{a}^{x}\left( t-a\right) \left( x-t\right) \left[ \left( \frac{t-a}{x-a \right) ^{s}\left\vert f^{\prime \prime }(x)\right\vert +\left( \frac{x-t} x-a}\right) ^{s}\left\vert f^{\prime \prime }(a)\right\vert \right] dt \\ &&+\frac{\beta }{\left( \alpha +\beta \right) \left( b-x\right) \int_{x}^{b}\left( b-t\right) \left( t-x\right) \left[ \left( \frac{t-x}{b-x \right) ^{s}\left\vert f^{\prime \prime }(b)\right\vert +\left( \frac{b-t} b-x}\right) ^{s}\left\vert f^{\prime \prime }(x)\right\vert \right] dt \\ &=&\frac{\alpha }{\alpha +\beta }\frac{\left( x-a\right) ^{2}}{\left( s+2\right) \left( s+3\right) }\left[ \left\vert f^{\prime \prime }(x)\right\vert +\left\vert f^{\prime \prime }(a)\right\vert \right] +\frac \beta }{\alpha +\beta }\frac{\left( b-x\right) ^{2}}{\left( s+2\right) \left( s+3\right) }\left[ \left\vert f^{\prime \prime }(x)\right\vert +\left\vert f^{\prime \prime }(b)\right\vert \right] \end{eqnarray* where we use the fact tha \begin{equation*} \int_{a}^{x}\left( t-a\right) ^{s+1}\left( x-t\right) dt=\int_{a}^{x}\left( t-a\right) \left( x-t\right) ^{s+1}dt=\frac{\left( x-a\right) ^{s+3}}{\left( s+2\right) \left( s+3\right) } \end{equation* an \begin{equation*} \int_{x}^{b}\left( b-t\right) \left( t-x\right) ^{s+1}dt=\int_{x}^{b}\left( b-t\right) ^{s+1}\left( t-x\right) dt=\frac{\left( b-x\right) ^{s+3}}{\left( s+2\right) \left( s+3\right) }. \end{equation* The proof is completed. \end{proof} \begin{corollary} \label{co 2.1} Suppose that all the assumptions of Theorem \ref{teo 2.1} are satisfied with $\left\vert f^{\prime \prime }\right\vert \leq M.$ Then we have \begin{eqnarray*} &&\left\vert f(x)+\frac{\alpha f(a)+\beta f(b)}{\alpha +\beta }-\frac{2} \alpha +\beta }\left[ \frac{\alpha }{x-a}\int_{a}^{x}f(t)dt+\frac{\beta }{b- }\int_{x}^{b}f(t)dt\right] \right\vert \\ &\leq &\frac{2M}{\left( s+2\right) \left( s+3\right) }\left[ \frac{\alpha \left( x-a\right) ^{2}+\beta \left( b-x\right) ^{2}}{\alpha +\beta }\right] . \end{eqnarray*} \end{corollary} \begin{corollary} \label{co 2.2} In Theorem \ref{teo 2.1}, if we choose $\alpha =\beta =1,$ we obtai \begin{eqnarray*} &&\left\vert f(x)+\frac{f(a)+f(b)}{2}-\left[ \frac{1}{x-a}\int_{a}^{x}f(t)dt \frac{1}{b-x}\int_{x}^{b}f(t)dt\right] \right\vert \\ &\leq &\frac{\left( x-a\right) ^{2}+\left( b-x\right) ^{2}}{2\left( s+2\right) \left( s+3\right) }\left\vert f^{\prime \prime }(x)\right\vert \frac{1}{2\left( s+2\right) \left( s+3\right) }\left[ \left( x-a\right) ^{2}\left\vert f^{\prime \prime }(a)\right\vert +\left( b-x\right) ^{2}\left\vert f^{\prime \prime }(b)\right\vert \right] . \end{eqnarray*} \end{corollary} \begin{corollary} \label{co 2.3} In Theorem \ref{teo 2.1}, if we choose $\alpha =\beta =\frac{ }{2}$ and $x=\frac{a+b}{2},$ we obtain the following Bullen type inequality \begin{eqnarray*} &&\left\vert \frac{1}{2}\left[ f\left( \frac{a+b}{2}\right) +\frac{f(a)+f(b }{2}\right] -\frac{1}{b-a}\int_{a}^{b}f(t)dt\right\vert \\ &\leq &\frac{\left( b-a\right) ^{2}}{8\left( s+2\right) \left( s+3\right) \left[ \left\vert f^{\prime \prime }\left( \frac{a+b}{2}\right) \right\vert \frac{\left\vert f^{\prime \prime }(a)\right\vert +\left\vert f^{\prime \prime }(b)\right\vert }{2}\right] . \end{eqnarray*} \end{corollary} \begin{theorem} \label{teo 2.2} Let $f:[a,b]\rightarrow \mathbb{R} $ be an absolutely continuous mapping such that $f^{\prime \prime }\in L[a,b].$ If $\left\vert f^{\prime \prime }\right\vert ^{q}$ is $s-$ convex in the second sense on $[a,b]$ for some fixed $s\in (0,1]$ and $q>1$ with \frac{1}{p}+\frac{1}{q}=1,$ the \begin{eqnarray*} &&\left\vert f(x)+\frac{\alpha f(a)+\beta f(b)}{\alpha +\beta }-\frac{2} \alpha +\beta }\left[ \frac{\alpha }{x-a}\int_{a}^{x}f(t)dt+\frac{\beta }{b- }\int_{x}^{b}f(t)dt\right] \right\vert \\ &\leq &\left( \frac{\alpha }{\alpha +\beta }\right) ^{p}\frac{\left( x-a\right) ^{1+\frac{1}{q}}}{\left( s+1\right) ^{\frac{1}{q}}}\left( \beta \left( p+1,p+1\right) \right) ^{\frac{1}{p}}\left[ \left\vert f^{\prime \prime }(x)\right\vert ^{q}+\left\vert f^{\prime \prime }(a)\right\vert ^{q \right] ^{\frac{1}{q}} \\ &&+\left( \frac{\beta }{\alpha +\beta }\right) ^{p}\frac{\left( b-x\right) ^{1+\frac{1}{q}}}{\left( s+1\right) ^{\frac{1}{q}}}\left( \beta \left( p+1,p+1\right) \right) ^{\frac{1}{p}}\left[ \left\vert f^{\prime \prime }(b)\right\vert ^{q}+\left\vert f^{\prime \prime }(x)\right\vert ^{q}\right] ^{\frac{1}{q}} \end{eqnarray* where $\beta \left( x,y\right) =\int_{0}^{1}t^{x-1}\left( 1-t\right) ^{y-1}dt,$ $x,y>0$ is the Euler Beta function, $\alpha ,\beta \in \mathbb{R} $ nonnegative and not both zero. \end{theorem} \begin{proof} From Lemma \ref{lem 2.1}, using the property of the modulus, H\"{o}lder inequality and $s-$convexity of $\left\vert f^{\prime \prime }\right\vert ^{q},$ we can writ \begin{eqnarray*} &&\left\vert f(x)+\frac{\alpha f(a)+\beta f(b)}{\alpha +\beta }-\frac{2} \alpha +\beta }\left[ \frac{\alpha }{x-a}\int_{a}^{x}f(t)dt+\frac{\beta }{b- }\int_{x}^{b}f(t)dt\right] \right\vert \\ &\leq &\left( \int_{a}^{x}\left( \frac{\alpha }{\alpha +\beta }\frac{\left( t-a\right) \left( x-t\right) }{x-a}\right) ^{p}dt\right) ^{\frac{1}{p }\left( \int_{a}^{x}\left\vert f^{\prime \prime }\left( \frac{t-a}{x-a}x \frac{x-t}{x-a}a\right) \right\vert ^{q}dt\right) ^{\frac{1}{q}} \\ &&+\left( \int_{x}^{b}\left( \frac{\beta }{\alpha +\beta }\frac{\left( b-t\right) \left( x-t\right) }{b-x}\right) ^{p}dt\right) ^{\frac{1}{p }\left( \int_{x}^{b}\left\vert f^{\prime \prime }\left( \frac{t-x}{b-x}b \frac{b-t}{b-x}x\right) \right\vert ^{q}dt\right) ^{\frac{1}{q}} \\ &\leq &\frac{\alpha }{\alpha +\beta }\left( x-a\right) \left( \int_{a}^{x \frac{\left( t-a\right) ^{p}\left( x-t\right) ^{p}}{\left( x-a\right) ^{p}\left( x-a\right) ^{p}}dt\right) ^{\frac{1}{p}}\left( \int_{a}^{x}\left\vert f^{\prime \prime }\left( \frac{t-a}{x-a}x+\frac{x-t} x-a}a\right) \right\vert ^{q}dt\right) ^{\frac{1}{q}} \\ &&+\frac{\beta }{\alpha +\beta }\left( b-x\right) \left( \int_{x}^{b}\frac \left( b-t\right) ^{p}\left( x-t\right) ^{p}}{\left( b-x\right) ^{p}\left( b-x\right) ^{p}}dt\right) ^{\frac{1}{p}}\left( \int_{x}^{b}\left\vert f^{\prime \prime }\left( \frac{t-x}{b-x}b+\frac{b-t}{b-x}x\right) \right\vert ^{q}dt\right) ^{\frac{1}{q}} \\ &\leq &\frac{\alpha }{\alpha +\beta }\left( x-a\right) \left( \beta \left( p+1,p+1\right) \right) ^{\frac{1}{p}}\left[ \int_{a}^{x}\left( \frac{t-a}{x- }\right) ^{s}\left\vert f^{\prime \prime }(x)\right\vert ^{q}+\left( \frac x-t}{x-a}\right) ^{s}\left\vert f^{\prime \prime }(a)\right\vert ^{q}\right] ^{\frac{1}{q}} \\ &&+\frac{\beta }{\alpha +\beta }\left( b-x\right) \left( \beta \left( p+1,p+1\right) \right) ^{\frac{1}{p}}\left[ \int_{x}^{b}\left( \frac{t-x}{b- }\right) ^{s}\left\vert f^{\prime \prime }(b)\right\vert ^{q}+\left( \frac b-t}{b-x}\right) ^{s}\left\vert f^{\prime \prime }(x)\right\vert ^{q}\right] ^{\frac{1}{q}}. \end{eqnarray* We get the desired result by making use of the necessary computation. \end{proof} \begin{theorem} \label{teo 2.3} Under the assumptions of Theorem \ref{teo 2.2}, the following inequality \begin{eqnarray*} &&\left\vert f(x)+\frac{\alpha f(a)+\beta f(b)}{\alpha +\beta }-\frac{2} \alpha +\beta }\left[ \frac{\alpha }{x-a}\int_{a}^{x}f(t)dt+\frac{\beta }{b- }\int_{x}^{b}f(t)dt\right] \right\vert \\ &\leq &\beta \left( p+1,p+1\right) ^{\frac{1}{p}}\left( \left( \frac{\alpha }{\alpha +\beta }\right) ^{p}\left( x-a\right) ^{p}+\left( \frac{\beta } \alpha +\beta }\right) ^{p}\left( b-x\right) ^{p}\right) ^{\frac{1}{p}} \\ &&\times \left( \frac{b-a}{s+1}\right) ^{\frac{1}{q}}\left( \left\vert f^{\prime \prime }(a)\right\vert ^{q}+\left\vert f^{\prime \prime }(b)\right\vert ^{q}\right) ^{\frac{1}{q}} \end{eqnarray* holds where $\beta \left( x,y\right) $ is the Euler Beta function. \end{theorem} \begin{proof} From Lemma \ref{lem 2.1}, using the property of the modulus, H\"{o}lder inequality and $s-$convexity of $\left\vert f^{\prime \prime }\right\vert ^{q},$ we can writ \begin{eqnarray*} &&\left\vert f(x)+\frac{\alpha f(a)+\beta f(b)}{\alpha +\beta }-\frac{2} \alpha +\beta }\left[ \frac{\alpha }{x-a}\int_{a}^{x}f(t)dt+\frac{\beta }{b- }\int_{x}^{b}f(t)dt\right] \right\vert \\ &\leq &\left( \int_{a}^{b}\left\vert K(x,t)\right\vert ^{p}dt\right) ^{\frac 1}{p}}\left( \int_{a}^{b}\left\vert f^{\prime \prime }(t)\right\vert ^{q}dt\right) ^{\frac{1}{q}} \\ &=&\left( \int_{a}^{x}\left( \frac{\alpha }{\alpha +\beta }\frac{\left( t-a\right) \left( x-t\right) }{\left( x-a\right) \left( x-a\right) }\left( x-a\right) \right) ^{p}dt+\int_{x}^{b}\left( \frac{\beta }{\alpha +\beta \frac{\left( b-t\right) \left( x-t\right) }{\left( b-x\right) \left( b-x\right) }\left( b-x\right) \right) ^{p}dt\right) ^{\frac{1}{p}} \\ &&\times \left( \int_{a}^{b}\left\vert f^{\prime \prime }\left( \frac{t-a} b-a}b+\frac{b-t}{b-a}a\right) \right\vert ^{q}dt\right) ^{\frac{1}{q}}. \end{eqnarray* We get the desired result by making use of the necessary computation. \end{proof} The next result is obtained by using the well-known power-mean integral ineqaulity: \begin{theorem} \label{teo 2.4} Let $f:[a,b]\rightarrow \mathbb{R} $ be an absolutely continuous mapping such that $f^{\prime \prime }\in L[a,b].$ If $\left\vert f^{\prime \prime }\right\vert ^{q}$ is $s-$ convex in the second sense on $[a,b]$ for some fixed $s\in (0,1]$ and $q\geq 1,$ then \begin{eqnarray*} &&\left\vert f(x)+\frac{\alpha f(a)+\beta f(b)}{\alpha +\beta }-\frac{2} \alpha +\beta }\left[ \frac{\alpha }{x-a}\int_{a}^{x}f(t)dt+\frac{\beta }{b- }\int_{x}^{b}f(t)dt\right] \right\vert \\ &\leq &\frac{\alpha }{\alpha +\beta }\frac{\left( x-a\right) ^{2}}{6^{1-^ \frac{1}{q}}}\left[ \left( s+2\right) \left( s+3\right) \right] ^{\frac{1}{q }}\left[ \left\vert f^{\prime \prime }(x)\right\vert ^{q}+\left\vert f^{\prime \prime }(a)\right\vert ^{q}\right] ^{\frac{1}{q}} \\ &&+\frac{\beta }{\alpha +\beta }\frac{\left( b-x\right) ^{2}}{6^{1-^{\frac{ }{q}}}\left[ \left( s+2\right) \left( s+3\right) \right] ^{\frac{1}{q}} \left[ \left\vert f^{\prime \prime }(b)\right\vert ^{q}+\left\vert f^{\prime \prime }(x)\right\vert ^{q}\right] ^{\frac{1}{q}} \end{eqnarray* holds where $\alpha ,\beta \in \mathbb{R} $ nonnegative and not both zero. \end{theorem} \begin{proof} From Lemma \ref{lem 2.1}, using the property of the modulus, power-mean integral inequality and $s-$convexity of $\left\vert f^{\prime \prime }\right\vert ^{q},$ we can writ \begin{eqnarray*} &&\left\vert f(x)+\frac{\alpha f(a)+\beta f(b)}{\alpha +\beta }-\frac{2} \alpha +\beta }\left[ \frac{\alpha }{x-a}\int_{a}^{x}f(t)dt+\frac{\beta }{b- }\int_{x}^{b}f(t)dt\right] \right\vert \\ &\leq &\frac{\alpha }{\left( \alpha +\beta \right) \left( x-a\right) }\left( \int_{a}^{x}\left( t-a\right) \left( x-t\right) dt\right) ^{1-\frac{1}{q}} \\ &&\times \left( \int_{a}^{x}\left( t-a\right) \left( x-t\right) \left( \left( \frac{t-a}{x-a}\right) ^{s}\left\vert f^{\prime \prime }(x)\right\vert ^{q}+\left( \frac{x-t}{x-a}\right) ^{s}\left\vert f^{\prime \prime }(a)\right\vert ^{q}\right) dt\right) ^{\frac{1}{q}} \\ &&+\frac{\beta }{\left( \alpha +\beta \right) \left( b-x\right) }\left( \int_{x}^{b}\left( b-t\right) \left( t-x\right) dt\right) ^{1-\frac{1}{q}} \\ &&\times \left( \int_{x}^{b}\left( t-b\right) \left( t-x\right) \left( \left( \frac{t-x}{b-x}\right) ^{s}\left\vert f^{\prime \prime }(b)\right\vert ^{q}+\left( \frac{b-t}{b-x}\right) ^{s}\left\vert f^{\prime \prime }(x)\right\vert ^{q}\right) dt\right) ^{\frac{1}{q}} \\ &=&\frac{\alpha }{\left( \alpha +\beta \right) \left( x-a\right) }\left( \frac{\left( x-a\right) ^{3}}{6}\right) ^{1-\frac{1}{q}}\left( \frac{\left( x-a\right) ^{3}}{\left( s+2\right) \left( s+3\right) }\left( \left\vert f^{\prime \prime }(x)\right\vert ^{q}+\left\vert f^{\prime \prime }(a)\right\vert ^{q}\right) \right) ^{\frac{1}{q}} \\ &&+\frac{\beta }{\left( \alpha +\beta \right) \left( b-x\right) }\left( \frac{\left( b-x\right) ^{3}}{6}\right) ^{1-\frac{1}{q}}\left( \frac{\left( b-x\right) ^{3}}{\left( s+2\right) \left( s+3\right) }\left( \left\vert f^{\prime \prime }(b)\right\vert ^{q}+\left\vert f^{\prime \prime }(x)\right\vert ^{q}\right) \right) ^{\frac{1}{q}}. \end{eqnarray* The proof is completed. \end{proof} \begin{remark} \label{rem 2.1} In Theorem \ref{teo 2.4}, if we choose $q=1$ Theorem \re {teo 2.4} reduces to Theorem \ref{teo 2.1}. \end{remark} \begin{remark} \label{rem 2.2} If we choose $s=1$ for all the results, we obtain new results for convex functions. \end{remark} \section{inequalities for $\log -$convex functions} In this section, we will give some results for $\log -$convex functions. For the simplicity, we will use the following notations \begin{eqnarray*} \kappa &=&\left( \frac{\left\vert f^{\prime \prime }(x)\right\vert } \left\vert f^{\prime \prime }(a)\right\vert }\right) ^{\frac{1}{x-a}} \\ \tau &=&\left( \frac{\left\vert f^{\prime \prime }(b)\right\vert } \left\vert f^{\prime \prime }(x)\right\vert }\right) ^{\frac{1}{b-x}}. \end{eqnarray*} \begin{theorem} \label{teo 3.1} Let $f:[a,b]\rightarrow \mathbb{R} $ be an absolutely continuous mapping such that $f^{\prime \prime }\in L[a,b].$ If $\left\vert f^{\prime \prime }\right\vert $ is $\log -$ convex function on $[a,b]$ and $\kappa \neq 1,\tau \neq 1,$ the \begin{eqnarray*} &&\left\vert f(x)+\frac{\alpha f(a)+\beta f(b)}{\alpha +\beta }-\frac{2} \alpha +\beta }\left[ \frac{\alpha }{x-a}\int_{a}^{x}f(t)dt+\frac{\beta }{b- }\int_{x}^{b}f(t)dt\right] \right\vert \\ &\leq &\frac{\alpha }{\left( \alpha +\beta \right) \left( x-a\right) }\left( \frac{\left\vert f^{\prime \prime }(a)\right\vert ^{x}}{\left\vert f^{\prime \prime }(x)\right\vert ^{a}}\right) ^{\frac{1}{x-a}}\left( \frac{2\left( \kappa ^{a}-\kappa ^{x}\right) -(a-x)\left( \kappa ^{a}+\kappa ^{x}\right) \log \kappa }{\log ^{3}\kappa }\right) \\ &&+\frac{\beta }{\left( \alpha +\beta \right) \left( b-x\right) }\left( \frac{\left\vert f^{\prime \prime }(x)\right\vert ^{b}}{\left\vert f^{\prime \prime }(b)\right\vert ^{x}}\right) ^{\frac{1}{b-x}}\left( \frac{2\tau ^{x}-2\tau ^{b}+(b-x)\left( \tau ^{b}+\tau ^{x}\right) \log \tau }{\log ^{3}\tau }\right) \end{eqnarray* holds where $\kappa \neq 1,\tau \neq 1,$ $\alpha ,\beta \in \mathbb{R} $ nonnegative and not both zero. \end{theorem} \begin{proof} From Lemma \ref{lem 2.1} and by using the $\log -$ convexity of $\left\vert f^{\prime \prime }\right\vert ,$ we can writ \begin{eqnarray*} &&\left\vert f(x)+\frac{\alpha f(a)+\beta f(b)}{\alpha +\beta }-\frac{2} \alpha +\beta }\left[ \frac{\alpha }{x-a}\int_{a}^{x}f(t)dt+\frac{\beta }{b- }\int_{x}^{b}f(t)dt\right] \right\vert \\ &\leq &\int_{a}^{x}\frac{\alpha }{\alpha +\beta }\frac{1}{x-a}\left\vert t-a\right\vert \left\vert x-t\right\vert \left\vert f^{\prime \prime }(t)\right\vert dt \\ &&+\int_{x}^{b}\frac{\beta }{\alpha +\beta }\frac{1}{b-x}\left\vert b-t\right\vert \left\vert x-t\right\vert \left\vert f^{\prime \prime }(t)\right\vert dt \\ &=&\frac{\alpha }{\left( \alpha +\beta \right) \left( x-a\right) \int_{a}^{x}\left( t-a\right) \left( x-t\right) \left\vert f^{\prime \prime }\left( \frac{t-a}{x-a}x+\frac{x-t}{x-a}a\right) \right\vert dt \\ &&+\frac{\beta }{\left( \alpha +\beta \right) \left( b-x\right) \int_{x}^{b}\left( b-t\right) \left( t-x\right) \left\vert f^{\prime \prime }\left( \frac{t-x}{b-x}b+\frac{b-t}{b-x}x\right) \right\vert dt \\ &\leq &\frac{\alpha }{\left( \alpha +\beta \right) \left( x-a\right) \int_{a}^{x}\left( t-a\right) \left( x-t\right) \left[ \left\vert f^{\prime \prime }(x)\right\vert ^{\frac{t-a}{x-a}}\left\vert f^{\prime \prime }(a)\right\vert ^{\frac{x-t}{x-a}}\right] dt \\ &&+\frac{\beta }{\left( \alpha +\beta \right) \left( b-x\right) \int_{x}^{b}\left( b-t\right) \left( t-x\right) \left[ \left\vert f^{\prime \prime }(b)\right\vert ^{\frac{t-x}{b-x}}\left\vert f^{\prime \prime }(x)\right\vert ^{\frac{b-t}{b-x}}\right] dt \\ &=&\frac{\alpha }{\left( \alpha +\beta \right) \left( x-a\right) }\left( \frac{\left\vert f^{\prime \prime }(a)\right\vert ^{x}}{\left\vert f^{\prime \prime }(x)\right\vert ^{a}}\right) ^{\frac{1}{x-a}}\int_{a}^{x}\left( t-a\right) \left( x-t\right) \kappa ^{t}dt \\ &&+\frac{\beta }{\left( \alpha +\beta \right) \left( b-x\right) }\left( \frac{\left\vert f^{\prime \prime }(x)\right\vert ^{b}}{\left\vert f^{\prime \prime }(b)\right\vert ^{x}}\right) ^{\frac{1}{b-x}}\int_{x}^{b}\left( b-t\right) \left( t-x\right) \tau ^{t}dt. \end{eqnarray* By a simple computation, we get the result. \end{proof} \begin{corollary} \label{co 3.1} In Theorem \ref{teo 3.1}, if we choose $\alpha =\beta =1,$ we obtain \begin{eqnarray*} &&\left\vert f(x)+\frac{f(a)+f(b)}{2}-\left[ \frac{1}{x-a}\int_{a}^{x}f(t)dt \frac{1}{b-x}\int_{x}^{b}f(t)dt\right] \right\vert \\ &\leq &\frac{1}{2\left( x-a\right) }\left( \frac{\left\vert f^{\prime \prime }(a)\right\vert ^{x}}{\left\vert f^{\prime \prime }(x)\right\vert ^{a} \right) ^{\frac{1}{x-a}}\left( \frac{2\left( \kappa ^{a}-\kappa ^{x}\right) -(a-x)\left( \kappa ^{a}+\kappa ^{x}\right) \log \kappa }{\log ^{3}\kappa \right) \\ &&+\frac{1}{2\left( b-x\right) }\left( \frac{\left\vert f^{\prime \prime }(x)\right\vert ^{b}}{\left\vert f^{\prime \prime }(b)\right\vert ^{x} \right) ^{\frac{1}{b-x}}\left( \frac{2\tau ^{x}-2\tau ^{b}+(b-x)\left( \tau ^{b}+\tau ^{x}\right) \log \tau }{\log ^{3}\tau }\right) . \end{eqnarray*} \end{corollary} \begin{corollary} \label{co 3.2} In Theorem \ref{teo 3.1}, if we choose $\alpha =\beta =\frac{ }{2}$ and $x=\frac{a+b}{2},$ we obtain the following Bullen type inequality \begin{eqnarray*} &&\left\vert \frac{1}{2}\left[ f\left( \frac{a+b}{2}\right) +\frac{f(a)+f(b }{2}\right] -\frac{1}{b-a}\int_{a}^{b}f(t)dt\right\vert \\ &\leq &\left( \frac{\left\vert f^{\prime \prime }(a)\right\vert ^{\frac{a+b} 2}}}{\left\vert f^{\prime \prime }\left( \frac{a+b}{2}\right) \right\vert ^{a}}\right) ^{\frac{2}{b-a}}\left( \frac{\kappa _{1}^{a}-\kappa _{1}^{\frac a+b}{2}}+\left( \frac{b-a}{4}\right) \left( \kappa _{1}^{a}+\kappa _{1}^ \frac{a+b}{2}}\right) \log \kappa _{1}}{\left( b-a\right) \log ^{3}\kappa _{1}}\right) \\ &&+\left( \frac{\left\vert f^{\prime \prime }\left( \frac{a+b}{2}\right) \right\vert ^{b}}{\left\vert f^{\prime \prime }(b)\right\vert ^{\frac{a+b}{2 }}\right) ^{\frac{2}{b-a}}\left( \frac{\tau _{1}^{\frac{a+b}{2}}-\tau _{1}^{b}+\left( \frac{b-a}{4}\right) \left( \tau _{1}^{b}+\tau _{1}^{\frac a+b}{2}}\right) \log \tau _{1}}{\left( b-a\right) \log ^{3}\tau _{1}}\right) \end{eqnarray* where \begin{eqnarray*} \kappa _{1} &=&\left( \frac{\left\vert f^{\prime \prime }\left( \frac{a+b}{2 \right) \right\vert }{\left\vert f^{\prime \prime }(a)\right\vert }\right) ^ \frac{2}{b-a}} \\ \tau _{1} &=&\left( \frac{\left\vert f^{\prime \prime }(b)\right\vert } \left\vert f^{\prime \prime }\left( \frac{a+b}{2}\right) \right\vert \right) ^{\frac{2}{b-a}}. \end{eqnarray*} \end{corollary} \begin{theorem} \label{teo 3.2} Let $f:[a,b]\rightarrow \mathbb{R} $ be an absolutely continuous mapping such that $f^{\prime \prime }\in L[a,b].$ If $\left\vert f^{\prime \prime }\right\vert ^{q}$ is $\log -$ convex function on $[a,b]$ and $q>1$ with $\frac{1}{p}+\frac{1}{q}=1,$ the \begin{eqnarray*} &&\left\vert f(x)+\frac{\alpha f(a)+\beta f(b)}{\alpha +\beta }-\frac{2} \alpha +\beta }\left[ \frac{\alpha }{x-a}\int_{a}^{x}f(t)dt+\frac{\beta }{b- }\int_{x}^{b}f(t)dt\right] \right\vert \\ &\leq &\frac{\alpha }{\alpha +\beta }\left( x-a\right) \left( \beta \left( p+1,p+1\right) \right) ^{\frac{1}{p}}\left( \frac{\left\vert f^{\prime \prime }(a)\right\vert ^{x}}{\left\vert f^{\prime \prime }(x)\right\vert ^{a }\right) ^{\frac{1}{x-a}}\left( \frac{\kappa ^{\frac{qx}{x-a}}-\kappa ^ \frac{qa}{x-a}}}{\log \kappa ^{\frac{q}{x-a}}}\right) ^{\frac{1}{q}} \\ &&+\frac{\beta }{\alpha +\beta }\left( b-x\right) \left( \beta \left( p+1,p+1\right) \right) ^{\frac{1}{p}}\left( \frac{\left\vert f^{\prime \prime }(x)\right\vert ^{b}}{\left\vert f^{\prime \prime }(b)\right\vert ^{x }\right) ^{\frac{1}{b-x}}\left( \frac{\tau ^{\frac{qb}{b-x}}-\tau ^{\frac{q }{b-x}}}{\log \tau ^{\frac{q}{b-x}}}\right) ^{\frac{1}{q}}. \end{eqnarray* where $\beta \left( x,y\right) =\int_{0}^{1}t^{x-1}\left( 1-t\right) ^{y-1}dt,$ $x,y>0$ is the Euler Beta function and $\kappa \neq 1,\tau \neq 1 , $\alpha ,\beta \in \mathbb{R} $ nonnegative and not both zero. \end{theorem} \begin{proof} From Lemma \ref{lem 2.1}, by using $\log -$convexity of $\left\vert f^{\prime \prime }\right\vert ^{q}$ and by applying H\"{o}lder inequality$,$ we ge \begin{eqnarray*} &&\left\vert f(x)+\frac{\alpha f(a)+\beta f(b)}{\alpha +\beta }-\frac{2} \alpha +\beta }\left[ \frac{\alpha }{x-a}\int_{a}^{x}f(t)dt+\frac{\beta }{b- }\int_{x}^{b}f(t)dt\right] \right\vert \\ &\leq &\left( \int_{a}^{x}\left( \frac{\alpha }{\alpha +\beta }\frac{\left( t-a\right) \left( x-t\right) }{x-a}\right) ^{p}dt\right) ^{\frac{1}{p }\left( \int_{a}^{x}\left\vert f^{\prime \prime }\left( \frac{t-a}{x-a}x \frac{x-t}{x-a}a\right) \right\vert ^{q}dt\right) ^{\frac{1}{q}} \\ &&+\left( \int_{x}^{b}\left( \frac{\beta }{\alpha +\beta }\frac{\left( b-t\right) \left( x-t\right) }{b-x}\right) ^{p}dt\right) ^{\frac{1}{p }\left( \int_{x}^{b}\left\vert f^{\prime \prime }\left( \frac{t-x}{b-x}b \frac{b-t}{b-x}x\right) \right\vert ^{q}dt\right) ^{\frac{1}{q}} \\ &\leq &\frac{\alpha }{\alpha +\beta }\left( x-a\right) \left( \int_{a}^{x \frac{\left( t-a\right) ^{p}\left( x-t\right) ^{p}}{\left( x-a\right) ^{p}\left( x-a\right) ^{p}}dt\right) ^{\frac{1}{p}}\left( \int_{a}^{x}\left\vert f^{\prime \prime }\left( \frac{t-a}{x-a}x+\frac{x-t} x-a}a\right) \right\vert ^{q}dt\right) ^{\frac{1}{q}} \\ &&+\frac{\beta }{\alpha +\beta }\left( b-x\right) \left( \int_{x}^{b}\frac \left( b-t\right) ^{p}\left( x-t\right) ^{p}}{\left( b-x\right) ^{p}\left( b-x\right) ^{p}}dt\right) ^{\frac{1}{p}}\left( \int_{x}^{b}\left\vert f^{\prime \prime }\left( \frac{t-x}{b-x}b+\frac{b-t}{b-x}x\right) \right\vert ^{q}dt\right) ^{\frac{1}{q}} \\ &\leq &\frac{\alpha }{\alpha +\beta }\left( x-a\right) \left( \beta \left( p+1,p+1\right) \right) ^{\frac{1}{p}}\left[ \left( \frac{\left\vert f^{\prime \prime }(a)\right\vert ^{x}}{\left\vert f^{\prime \prime }(x)\right\vert ^{a}}\right) ^{\frac{q}{x-a}}\int_{a}^{x}\left( \frac \left\vert f^{\prime \prime }(x)\right\vert ^{\frac{q}{x-a}}}{\left\vert f^{\prime \prime }(a)\right\vert ^{\frac{q}{x-a}}}\right) ^{t}dt\right] ^ \frac{1}{q}} \\ &&+\frac{\beta }{\alpha +\beta }\left( b-x\right) \left( \beta \left( p+1,p+1\right) \right) ^{\frac{1}{p}}\left[ \left( \frac{\left\vert f^{\prime \prime }(x)\right\vert ^{b}}{\left\vert f^{\prime \prime }(b)\right\vert ^{x}}\right) ^{\frac{q}{b-x}}\int_{x}^{b}\left( \frac \left\vert f^{\prime \prime }(b)\right\vert ^{\frac{q}{b-x}}}{\left\vert f^{\prime \prime }(x)\right\vert ^{\frac{q}{b-x}}}\right) ^{t}dt\right] ^ \frac{1}{q}}. \end{eqnarray* By computing the above integrals, we get the desired result. \end{proof} \begin{theorem} \label{teo 3.3} Let $f:[a,b]\rightarrow \mathbb{R} $ be an absolutely continuous mapping such that $f^{\prime \prime }\in L[a,b].$ If $\left\vert f^{\prime \prime }\right\vert ^{q}$ is $\log - convex function on $[a,b]$ and $q\geq 1,$ then \begin{eqnarray*} &&\left\vert f(x)+\frac{\alpha f(a)+\beta f(b)}{\alpha +\beta }-\frac{2} \alpha +\beta }\left[ \frac{\alpha }{x-a}\int_{a}^{x}f(t)dt+\frac{\beta }{b- }\int_{x}^{b}f(t)dt\right] \right\vert \\ &\leq &\frac{\alpha \left( x-a\right) ^{2-\frac{3}{q}}}{6^{1-\frac{1}{q }\left( \alpha +\beta \right) }\left( \frac{\left\vert f^{\prime \prime }(a)\right\vert ^{x}}{\left\vert f^{\prime \prime }(x)\right\vert ^{a} \right) ^{\frac{1}{x-a}}\left( \frac{2\left( \kappa ^{qa}-\kappa ^{qx}\right) -q(a-x)\left( \kappa ^{qa}+\kappa ^{qx}\right) \log \kappa } \log ^{3}\kappa ^{q}}\right) ^{\frac{1}{q}} \\ &&+\frac{\beta \left( b-x\right) ^{2-\frac{3}{q}}}{6^{1-\frac{1}{q}}\left( \alpha +\beta \right) }\left( \frac{\left\vert f^{\prime \prime }(x)\right\vert ^{b}}{\left\vert f^{\prime \prime }(b)\right\vert ^{x} \right) ^{\frac{1}{b-x}}\left( \frac{2\tau ^{qx}-2\tau ^{qb}+q(b-x)\left( \tau ^{qb}+\tau ^{qx}\right) \log \tau }{\log ^{3}\tau ^{q}}\right) ^{\frac{ }{q}} \end{eqnarray* holds where $\kappa ^{q}\neq 1,\tau ^{q}\neq 1$, $\alpha ,\beta \in \mathbb{R} $ nonnegative and not both zero. \end{theorem} \begin{proof} From Lemma \ref{lem 2.1}, by using the well-known power-mean integral inequality and $\log -$convexity of $\left\vert f^{\prime \prime }\right\vert ^{q},$ we hav \begin{eqnarray*} &&\left\vert f(x)+\frac{\alpha f(a)+\beta f(b)}{\alpha +\beta }-\frac{2} \alpha +\beta }\left[ \frac{\alpha }{x-a}\int_{a}^{x}f(t)dt+\frac{\beta }{b- }\int_{x}^{b}f(t)dt\right] \right\vert \\ &\leq &\frac{\alpha }{\left( \alpha +\beta \right) \left( x-a\right) }\left( \int_{a}^{x}\left( t-a\right) \left( x-t\right) dt\right) ^{1-\frac{1}{q }\left( \int_{a}^{x}\left( t-a\right) \left( x-t\right) \left( \left\vert f^{\prime \prime }(x)\right\vert ^{q\frac{t-a}{x-a}}\left\vert f^{\prime \prime }(a)\right\vert ^{q\frac{x-t}{x-a}}\right) dt\right) ^{\frac{1}{q}} \\ &&+\frac{\beta }{\left( \alpha +\beta \right) \left( b-x\right) }\left( \int_{x}^{b}\left( b-t\right) \left( t-x\right) dt\right) ^{1-\frac{1}{q }\left( \int_{x}^{b}\left( t-b\right) \left( t-x\right) \left( \left\vert f^{\prime \prime }(b)\right\vert ^{q\frac{t-x}{b-x}}\left\vert f^{\prime \prime }(x)\right\vert ^{q\frac{b-t}{b-x}}\right) dt\right) ^{\frac{1}{q}} \\ &=&\frac{\alpha }{\left( \alpha +\beta \right) \left( x-a\right) }\left( \frac{\left( x-a\right) ^{3}}{6}\right) ^{1-\frac{1}{q}}\left( \frac \left\vert f^{\prime \prime }(a)\right\vert ^{x}}{\left\vert f^{\prime \prime }(x)\right\vert ^{a}}\right) ^{\frac{1}{x-a}}\left( \frac{2\left( \kappa ^{qa}-\kappa ^{qx}\right) -q(a-x)\left( \kappa ^{qa}+\kappa ^{qx}\right) \log \kappa }{\log ^{3}\kappa ^{q}}\right) ^{\frac{1}{q}} \\ &&+\frac{\beta }{\left( \alpha +\beta \right) \left( b-x\right) }\left( \frac{\left( b-x\right) ^{3}}{6}\right) ^{1-\frac{1}{q}}\left( \frac \left\vert f^{\prime \prime }(x)\right\vert ^{b}}{\left\vert f^{\prime \prime }(b)\right\vert ^{x}}\right) ^{\frac{1}{b-x}}\left( \frac{2\tau ^{qx}-2\tau ^{qb}+q(b-x)\left( \tau ^{qb}+\tau ^{qx}\right) \log \tau }{\log ^{3}\tau ^{q}}\right) ^{\frac{1}{q}}. \end{eqnarray* Which completes the proof. \end{proof} \begin{remark} \label{rem 3.1} In Theorem \ref{teo 3.3}, if we choose $q=1$ Theorem \re {teo 3.3} reduces to Theorem \ref{teo 3.1}. \end{remark} \begin{corollary} For the particular selections of the parameters $\alpha ,\beta $ and the variable $x,$ one can obtain several new inequalities for $\log -$convex functions, we omit the details. \end{corollary} \section{APPLICATIONS FOR NUMERICAL INTEGRATION} Suppose that $d=\left\{ a=x_{0}<x_{1}<...<x_{n}=b\right\} $ is a partition of the interval $\left[ a,b\right] ,$ $h_{i}=x_{i+1}-x_{i},$ for i=0,1,2,...,n-1$ and consider the averaged midpoint-trapezoid quadrature formul \begin{equation*} \int_{a}^{b}f\left( x\right) dx=A_{MT}\left( d,f\right) +R_{MT}\left( d,f\right) , \end{equation* where \begin{equation*} A_{MT}\left( \pi ,f\right) =\frac{1}{4}\sum_{i=0}^{n-1}h_{i}\left[ f\left( x_{i}\right) +2f\left( \frac{x_{i}+x_{i+1}}{2}\right) +f\left( x_{i+1}\right) \right] \end{equation* Here, the term $R_{MT}\left( d,f\right) $ denotes the associated approximation error. (See \cite{19}) \begin{proposition} Let $f:[a,b]\rightarrow \mathbb{R} $ be an absolutely continuous mapping such that $f^{\prime \prime }\in L[a,b].$ If $\left\vert f^{\prime \prime }\right\vert $ is $\log -$ convex function on $[a,b]$ and $\kappa _{1}\neq 1,\tau _{1}\neq 1,$ then for the partition $d,$ following inequality hold \begin{eqnarray*} &&\left\vert R_{MT}\left( d,f\right) \right\vert \\ &\leq &\left( \frac{\left\vert f^{\prime \prime }(x_{i})\right\vert ^{\frac x_{i}+x_{i+1}}{2}}}{\left\vert f^{\prime \prime }\left( \frac{x_{i}+x_{i+1}} 2}\right) \right\vert ^{x_{i}}}\right) ^{\frac{2}{h_{i}}}\left( \frac{\kappa _{i}^{x_{i}}-\kappa _{i}^{\frac{x_{i}+x_{i+1}}{2}}+\left( \frac{h_{i}}{4 \right) \left( \kappa _{i}^{x_{i}}+\kappa _{i}^{\frac{x_{i}+x_{i+1}}{2 }\right) \log \kappa _{i}}{h_{i}\log ^{3}\kappa _{i}}\right) \\ &&+\left( \frac{\left\vert f^{\prime \prime }\left( \frac{x_{i}+x_{i+1}}{2 \right) \right\vert ^{x_{i+1}}}{\left\vert f^{\prime \prime }(x_{i+1})\right\vert ^{\frac{x_{i}+x_{i+1}}{2}}}\right) ^{\frac{2}{h_{i} }\left( \frac{\tau _{i}^{\frac{x_{i}+x_{i+1}}{2}}-\tau _{i}^{x_{i+1}}+\left( \frac{h_{i}}{4}\right) \left( \tau _{i}^{x_{i+1}}+\tau _{i}^{\frac x_{i}+x_{i+1}}{2}}\right) \log \tau _{i}}{h_{i}\log ^{3}\tau _{i}}\right) . \end{eqnarray* where $\kappa _{i}\neq 1,\tau _{i}\neq 1$ and defined a \begin{eqnarray*} \kappa _{i} &=&\left( \frac{\left\vert f^{\prime \prime }\left( \frac x_{i}+x_{i+1}}{2}\right) \right\vert }{\left\vert f^{\prime \prime }(x_{i})\right\vert }\right) ^{\frac{2}{h_{i}}} \\ \tau _{i} &=&\left( \frac{\left\vert f^{\prime \prime }(x_{i+1})\right\vert }{\left\vert f^{\prime \prime }\left( \frac{x_{i}+x_{i+1}}{2}\right) \right\vert }\right) ^{\frac{2}{h_{i}}}. \end{eqnarray*} \end{proposition} \begin{proof} By applying Corollary \ref{co 3.2} to the subintervals $\left[ x_{i},x_{i+1 \right] $ of $d,$ ($i=0,1,...,n-1)$ and by summation. We obtain the desired result. \end{proof} \begin{proposition} Let $f:[a,b]\rightarrow \mathbb{R} $ be an absolutely continuous mapping such that $f^{\prime \prime }\in L[a,b].$ If $\left\vert f^{\prime \prime }\right\vert $ is $s-$ convex in the second sense on $[a,b]$ for some fixed $s\in (0,1],$ then for partition d$ of $[a,b]$ the following inequality holds \begin{eqnarray*} &&\left\vert R_{MT}\left( d,f\right) \right\vert \\ &\leq &\frac{h_{i}^{2}}{8\left( s+2\right) \left( s+3\right) }\left[ \left\vert f^{\prime \prime }\left( \frac{x_{i}+x_{i+1}}{2}\right) \right\vert +\frac{\left\vert f^{\prime \prime }(x_{i})\right\vert +\left\vert f^{\prime \prime }(x_{i+1})\right\vert }{2}\right] . \end{eqnarray*} \end{proposition} \begin{proof} By applying Corollary \ref{co 2.3} to the\ subintervals $\left[ x_{i},x_{i+1 \right] $ of $d,$ ($i=0,1,...,n-1)$ and by summation. we get the result. \end{proof}
{ "timestamp": "2014-09-04T02:10:26", "yymm": "1409", "arxiv_id": "1409.1065", "language": "en", "url": "https://arxiv.org/abs/1409.1065", "abstract": "In this paper, we obtain some new inequalities for functions whose second derivatives' absolute value is s-convex and log-convex. Also, we give some applications for numerical integration.", "subjects": "Classical Analysis and ODEs (math.CA)", "title": "Inequalities via s-convexity and log-convexity", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9808759610129464, "lm_q2_score": 0.8354835411997897, "lm_q1q2_score": 0.8195057213848433 }
https://arxiv.org/abs/2005.14125
Notes on ridge functions and neural networks
These notes are about ridge functions. Recent years have witnessed a flurry of interest in these functions. Ridge functions appear in various fields and under various guises. They appear in fields as diverse as partial differential equations (where they are called plane waves), computerized tomography and statistics. These functions are also the underpinnings of many central models in neural networks.We are interested in ridge functions from the point of view of approximation theory. The basic goal in approximation theory is to approximate complicated objects by simpler objects. Among many classes of multivariate functions, linear combinations of ridge functions are a class of simpler functions. These notes study some problems of approximation of multivariate functions by linear combinations of ridge functions. We present here various properties of these functions. The questions we ask are as follows. When can a multivariate function be expressed as a linear combination of ridge functions from a certain class? When do such linear combinations represent each multivariate function? If a precise representation is not possible, can one approximate arbitrarily well? If well approximation fails, how can one compute/estimate the error of approximation, know that a best approximation exists? How can one characterize and construct best approximations? If a smooth function is a sum of arbitrarily behaved ridge functions, can it be expressed as a sum of smooth ridge functions? We also study properties of generalized ridge functions, which are very much related to linear superpositions and Kolmogorov's famous superposition theorem. These notes end with a few applications of ridge functions to the problem of approximation by single and two hidden layer neural networks with a restricted set of weights.We hope that these notes will be useful and interesting to both researchers and students.
\chapter*{Preface} These notes are about \textit{ridge functions}. Recent years have witnessed a flurry of interest in these functions. Ridge functions appear in various fields and under various guises. They appear in fields as diverse as partial differential equations (where they are called \textit{plane waves}), computerized tomography and statistics. These functions are also the underpinnings of many central models in neural networks. We are interested in ridge functions from the point of view of approximation theory. The basic goal in approximation theory is to approximate complicated objects by simpler objects. Among many classes of multivariate functions, linear combinations of ridge functions are a class of simpler functions. These notes study some problems of approximation of multivariate functions by linear combinations of ridge functions. We present here various properties of these functions. The questions we ask are as follows. When can a multivariate function be expressed as a linear combination of ridge functions from a certain class? When do such linear combinations represent each multivariate function? If a precise representation is not possible, can one approximate arbitrarily well? If well approximation fails, how can one compute/estimate the error of approximation, know that a best approximation exists? How can one characterize and construct best approximations? If a smooth function is a sum of arbitrarily behaved ridge functions, is it true that it can be expressed as a sum of smooth ridge functions? We also study properties of generalized ridge functions, which are very much related to linear superpositions and Kolmogorov's famous superposition theorem. These notes end with a few applications of ridge functions to the problem of approximation by single and two hidden layer neural networks with a restricted set of weights. We hope that these notes will be useful and interesting to both researchers and graduate students. \newpage \tableofcontents \newpage \chapter*{Introduction} \addcontentsline{toc}{chapter}{Introduction} Recent years have seen a growing interest in the study of special multivariate functions called ridge functions. A \textit{ridge function}, in its simplest format, is a multivariate function of the form $g\left( \mathbf{% a}\cdot \mathbf{x}\right) $, where $g:\mathbb{R}\rightarrow \mathbb{R}$, $% \mathbf{a}=\left( a_{1},...,a_{d}\right) $ is a fixed vector (direction) in $% \mathbb{R}^{d}\backslash \left\{ \mathbf{0}\right\} $, $\mathbf{x}=\left( x_{1},...,x_{d}\right) $ is the variable and $\mathbf{a}\cdot \mathbf{x}$ is the standard inner product. In other words, a ridge function is a multivariate function constant on the parallel hyperplanes $\mathbf{a}\cdot \mathbf{x}=c$, $c\in \mathbb{R}$. These functions arise naturally in various fields. They arise in computerized tomography (see, e.g., \cite% {72,73,74,97,106,111}), statistics (see, e.g., \cite{13,14,27,33,42}) and neural networks (see, e.g., \cite{22,Is,58,94,100,119,123}). These functions are also used in modern approximation theory as an effective and convenient tool for approximating complicated multivariate functions (see, e.g., \cite{38,66,57,Kr,101,114,118,137}). It should be remarked that long before the appearance of the name \textquotedblleft ridge", these functions were used in PDE theory under the name of \textit{plane waves}. For example, see the book by F. John \cite{69}. In general, sums of ridge functions with fixed directions occur in the study of hyperbolic constant coefficient partial differential equations. As an example, assume that $(\alpha _{i},\beta _{i}),~i=1,...,r,$ are pairwise linearly independent vectors in $\mathbb{R}% ^{2}$. Then the general solution to the homogeneous partial differential equation \begin{equation*} \prod\limits_{i=1}^{r}\left( \alpha {_{i}{\frac{\partial }{\partial {x}}}% +\beta _{i}{\frac{\partial }{\partial {y}}}}\right) {u}\left( {x,y}\right) =0 \end{equation*}% are all functions of the form \begin{equation*} u(x,y)=\sum\limits_{i=1}^{r}g_{i}\left( \beta {_{i}x-\alpha _{i}y}\right) \end{equation*}% for arbitrary continuous univariate functions $g_{i}$, $i=1,...,r$. Here the derivatives are understood in the sense of distributions. The term \textquotedblleft ridge function" was coined by Logan and Shepp in their seminal paper \cite{97} devoted to the basic mathematical problem of computerized tomography. This problem consists of reconstructing a given multivariate function from values of its integrals along certain straight lines in the plane. The integrals along parallel lines can be considered as a ridge function. Thus, the problem is to reconstruct $f$ from some set of ridge functions generated by the function $f$ itself. In practice, one can consider only a finite number of directions along which the above integrals are taken. Obviously, reconstruction from such data needs some additional conditions to be unique, since there are many functions $g$ having the same integrals. For uniqueness, Logan and Shepp \cite{97} used the criterion of minimizing the $L_{2}$ norm of $g$. That is, they found a function $g(x,y)$ with the minimum $L_{2}$ norm among all functions, which has the same integrals as $f$. More precisely, let $D$ be the unit disk in the plane and an unknown function $f(x,y)$ be square integrable and supported on $D.$ We are given projections $P_{f}(t,\theta )$ (integrals of $f$ along the lines $% x\cos \theta +y\sin \theta =t$) and looking for a function $g=g(x,y)$ of minimum $L_{2}$ norm, which has the same projections as $f:$ $P_{g}(t,\theta _{j})=P_{f}(t,\theta _{j}),$ $j=0,1,...,n-1$, where the angles $\theta _{j}$ generate equally spaced directions, i.e. $\theta _{j}=\frac{j\pi }{n},$ $% j=0,1,...,n-1.$ The authors of \cite{97} showed that this problem of tomography is equivalent to the problem of $L_{2}$-approximation of the function $f$ by sums of ridge functions with the equally spaced directions $% (\cos \theta _{j},\sin \theta _{j})$, $j=0,1,...,n-1.$ They gave a closed-form expression for the unique function $g(x,y)$ and showed that the unique polynomial $P(x,y)$ of degree $n-1$ which best approximates $f$ in $% L_{2}(D)$ is determined from the above $n$ projections of $f$ and can be represented as a sum of $n$ ridge functions. Kazantsev \cite{72} solved the above problem of tomography without requiring that the considered directions are equally spaced. Marr \cite{106} considered the problem of finding a polynomial of degree $n-2$, whose projections along lines joining each pair of $n$ equally spaced points on the circumference of $D$ best matches the given projections of $f$ in the sense of minimizing the sum of squares of the differences. Thus we see that the problems of tomography give rise to an independent study of approximation theoretic properties of the following set of linear combinations of ridge functions: \begin{equation*} \mathcal{R}\left( \mathbf{a}^{1},...,\mathbf{a}^{r}\right) =\left\{ \sum\limits_{i=1}^{r}g_{i}\left( \mathbf{a}^{i}\cdot \mathbf{x}\right) :g_{i}:\mathbb{R}\rightarrow \mathbb{R},i=1,...,r\right\} , \end{equation*}% where directions $\mathbf{a}^{1},...,\mathbf{a}^{r}$ are fixed and belong to the $d$-dimensional Euclidean space. Note that the set $\mathcal{R}\left( \mathbf{a}^{1},...,\mathbf{a}^{r}\right) $ is a linear space. Ridge function approximation also appears in statistics in \textit{% Projection Pursuit}. This term was introduced by Friedman and Tukey \cite{32} to name a technique for the explanatory analysis of large and multivariate data sets. This technique seeks out ``interesting" linear projections of the multivariate data onto a line or a plane. Projection Pursuit algorithms approximate a multivariate function $f$ by sums of ridge functions with variable directions, that is, by functions from the set \begin{equation*} \mathcal{R}_{r}=\left\{ \sum\limits_{i=1}^{r}g_{i}\left( \mathbf{a}^{i}\cdot \mathbf{x}\right) :\mathbf{a}^{i}\in \mathbb{R}^{d}\setminus \{\mathbf{0}% \},\ g_{i}:\mathbb{R}\rightarrow \mathbb{R},i=1,...,r\right\} . \end{equation*}% Here $r$ is the only fixed parameter, directions $\mathbf{a}^{1},...,\mathbf{% a}^{r}$ and functions $g_{1},...,g_{r}$ are free to choose. The first method of such approximation was developed by Friedman and Stuetzle \cite{33}. Their approximation process called \textit{Projection Pursuit Regression} (PPR) operates in a stepwise and greedy fashion. The process does not find a best approximation from $\mathcal{R}_{r}$, it algorithmically constructs functions $g_{r}\in \mathcal{R}_{r},$ such that $\left\Vert g_{r}-f\right\Vert _{L_{2}}\rightarrow 0,$ as $r\rightarrow \infty $. At stage $m$, PPR looks for a univariate function $g_{m}$ and direction $% \mathbf{a}^{m}$ such that the ridge function $g_{m}\left( \mathbf{a}% ^{m}\cdot \mathbf{x}\right) $ best approximates the residual $% f(x)-\sum\limits_{j=1}^{m-1}g_{j}\left( \mathbf{a}^{j}\cdot \mathbf{x}% \right) $. Projection pursuit regression has been proposed as an approach to bypass the curse of dimensionality and now is applied to prediction in applied sciences. In \cite{13,14}, Candes developed a new approach based not on stepwise construction of approximation but on a new transform called the \textit{ridgelet transform}. The ridgelet transform represents general functions as integrals of \textit{ridgelets} -- specifically chosen ridge functions. The significance of approximation by ridge functions is well understood from its role in the theory of \textit{neural networks}. Ridge functions appear in the definitions of many central neural network models. It is a broad knowledge that neural networks are being successfully applied across an extraordinary range of problem domains, in fields as diverse as finance, medicine, engineering, geology and physics. Generally speaking, neural networks are being introduced anywhere that there are problems of prediction, classification or control. Thus not surprisingly, there is a great interest to this powerful and very popular area of research (see, e.g., \cite{119} and a great deal of references therein). An artificial neural network is a way to perform computations using networks of interconnected computational units vaguely analogous to neurons simulating how our brain solves them. An artificial neuron, which forms the basis for designing neural networks, is a device with $d$ real inputs and an output. This output is generally a ridge function of the given inputs. In mathematical terms, a neuron may be described as \begin{equation*} y=\sigma (\mathbf{w\cdot x}-\theta ), \end{equation*}% where $\mathbf{x=(x}_{1},...,x_{d})\in \mathbb{R}^{d}$ are the input signals, $w=(w_{1},...,w_{d})\in \mathbb{R}^{d}$ are the synaptic weights, $% \theta \in \mathbb{R}$ is the bias, $\sigma $ is the activation function and $y$ is the output signal of the neuron. In a layered neural network the neurons are organized in the form of layers. We have at least two layers: an input and an output layer. The layers between the input and the output layers (if any) are called hidden layers, whose computation nodes are correspondingly called hidden neurons or hidden units. The output signals of the first layer are used as inputs to the second layer, the output signals of the second layer are used as inputs to the third layer, and so on for the rest of the network. Neural networks with this kind of architecture is called a \textit{Multilayer Feedforward Perceptron} (MLP). This is the most popular model among other neural network models. In this model, a neural network with a single hidden layer and one output represents a function of the form \begin{equation*} \sum_{i=1}^{r}c_{i}\sigma (\mathbf{w}^{i}\mathbf{\cdot x}-\theta _{i}). \end{equation*} Here the weights $\mathbf{w}^{i}$ are vectors in $\mathbb{R}^{d}$, the thresholds $\theta _{i}$ and the coefficients $c_{i}$ are real numbers and \ the activation function $\sigma $ is a univariate function. We fix only $% \sigma $ and $r$. Note that the functions $\sigma (\mathbf{w}^{i}\mathbf{% \cdot x}-\theta _{i})$ are ridge functions. Thus it is not surprising that some approximation theoretic problems related to neural networks have strong association with the corresponding problems of approximation by ridge functions. It is clear that in the special case, linear combinations of ridge functions turn into sums of univariate functions. This is also the simplest case. The simplicity of the approximation apparatus itself guarantees its utility in applications where multivariate functions are constant obstacles. In mathematics, this type of approximation has arisen, for example, in connection with the classical functional equations \cite{11}, the numerical solution of certain PDE boundary value problems \cite{9}, dimension theory \cite{132,133}, etc. In computer science, it arises in connection with the efficient storage of data in computer databases (see, e.g., \cite{140}). There is an interesting interconnection between the theory of approximation by univariate functions and problems of equilibrium construction in economics (see \cite{136}). Linear combinations of ridge functions with fixed directions allow a natural generalization to functions of the form $g(\alpha _{1}(x_{1})+\cdot \cdot \cdot +\alpha _{d}(x_{d}))$, where $\alpha _{i}(x_{i})$, $i=\overline{1,d},$ are real univariate functions. Such a generalization has a strong association with linear superpositions. A linear superposition is a function expressed as the sum% \begin{equation*} \sum\limits_{i=1}^{r}g_{i}(h_{i}(x)), \; x \in X, \end{equation*}% where $X$ is any set (in particular, a subset of $\mathbb{R}^{d}$), $% h_{i}:X\rightarrow {{\mathbb{R}}},~i=1,...,r,$ are arbitrarily fixed functions, and $g_{i}:\mathbb{R}\rightarrow \mathbb{R},~i=1,...,r.$ Note that here we deal with more complicated composition than the composition of a univariate function with the inner product. A starting point in the study of linear superpositions was the well known superposition theorem of Kolmogorov \cite{83} (see also the paper on Kolmogorov's works by Tikhomirov \cite{139}). This theorem states that for the unit cube $\mathbb{I}^{d},~% \mathbb{I}=[0,1],~d\geq 2,$ there exist $2d+1$ functions $% \{s_{q}\}_{q=1}^{2d+1}\subset C(\mathbb{I}^{d})$ of the form \begin{equation*} s_{q}(x_{1},...,x_{d})=\sum_{p=1}^{d}\varphi _{pq}(x_{p}),~\varphi _{pq}\in C(\mathbb{I}),~p=1,...,d,~q=1,...,2d+1 \end{equation*}% such that each function $f\in C(\mathbb{I}^{d})$ admits the representation \begin{equation*} f(x)=\sum_{q=1}^{2d+1}g_{q}(s_{q}(x)),~x=(x_{1},...,x_{d})\in \mathbb{I}% ^{d},~g_{q}\in C({{\mathbb{R)}}}. \end{equation*}% Thus, any continuous function on the unit cube can be represented as a linear superposition with the fixed inner functions $s_{1},...,s_{2d+1}$. In literature, these functions are called universal functions or the Kolmogorov functions. Note that all the functions $g_{q}(s_{q}(x))$ in the Kolmogorov superposition formula are generalized ridge functions, since each $s_{q}$ is a sum of univariate functions. In these notes, we consider some problems of approximation and/or representation of multivariate functions by linear combinations of ridge functions, generalized ridge functions and feedforward neural networks. The notes consist of five chapters. Chapter 1 is devoted to the approximation from some sets of ridge functions with arbitrarily fixed directions in $C$ and $L_{2}$ metrics. First, we study problems of representation of multivariate functions by linear combinations of ridge functions. Then, in case of two fixed directions and under suitable conditions, we give complete solutions to three basic problems of uniform approximation, namely, problems on existence, characterization, and construction of a best approximation. We also study problems of well approximation (approximation with arbitrary accuracy) and representation of continuous multivariate functions by sums of two continuous ridge functions. The reader will see the main difficulties and remained open problems in the uniform approximation by sums of more than two ridge functions. For $L_{2}$ approximation, a number of summands does not play such an essential role as it plays in the uniform approximation. In this case, it is known that a best approximation always exists and unique. For some special domains in $\mathbb{R}^{d}$, we characterize and then construct the best approximation. We also give an explicit formula for the approximation error. Chapter 2 explores the following open problem raised in Buhmann and Pinkus \cite{12}, and Pinkus \cite[p. 14]{117}. Assume we are given a function $f(\mathbf{x})=f(x_{1},...,x_{n})$ of the form \begin{equation*} f(\mathbf{x})=\sum_{i=1}^{k}f_{i}(\mathbf{a}^{i}\cdot \mathbf{x}), \end{equation*}% where the $\mathbf{a}^{i},$ $i=1,...,k,$ are pairwise linearly independent vectors (directions) in $\mathbb{R}^{d}$, $f_{i}$ are arbitrarily behaved univariate functions and $\mathbf{a}^{i}\cdot \mathbf{x}$ are standard inner products. Assume, in addition, that $f$ is of a certain smoothness class, that is, $f\in C^{s}(\mathbb{R}^{d})$, where $s\geq 0$ (with the convention that $C^{0}(\mathbb{R}^{d})=C(\mathbb{R}^{d})$). Is it true that there will always exist $g_{i}\in C^{s}(\mathbb{R})$ such that \begin{equation*} f(\mathbf{x})=\sum_{i=1}^{k}g_{i}(\mathbf{a}^{i}\cdot \mathbf{x})\text{ ?}% \end{equation*} In this chapter, we solve this problem up to some multivariate polynomial. We find various conditions on the directions $% \mathbf{a}^{i}$ allowing to express this polynomial as a sum of smooth ridge functions with these directions. We also consider the question of constructing $g_{i}$ using the information about the known functions $f_{i}$. Chapter 3 is devoted to the simplest type of ridge functions -- univariate functions. Note that a ridge function depends only on one variable if its direction coincides with the coordinate direction. Thus, in case of coincidence of all given directions with the coordinate directions, the problem of ridge function approximation turns into the problem of approximation of multivariate functions by sums of univariate functions. In this chapter, we first consider the approximation of a bivariate function $% f(x,y)$ by sums $\varphi (x)+\psi (y)$ on a rectangular domain $R$. We construct special classes of continuous functions depending on a numerical parameter and characterize each class in terms of the approximation error calculation formulas. This parameter will show which points of $R$ the calculation formula involves. We will also construct a best approximating sum $\varphi _{0}(x)+\psi _{0}(y)$ to a function from constructed classes. Then we develop a method for obtaining explicit formulas for the error of approximation of bivariate functions, defined on a union of rectangles, by sums of univariate functions. It should be remarked that formulas of such type were known only for functions defined on a rectangle with sides parallel to the coordinate axes. Our method, based on a maximization process over certain objects, called \textquotedblleft closed bolts", allows the consideration of functions defined on hexagons, octagons and stairlike polygons with sides parallel to the coordinate axes. At the end of this chapter we discuss one important result from Golomb's paper \cite{37}. This paper, published in 1959, made a start of a systematic study of approximation of multivariate functions by various compositions, including sums of univariate functions. In \cite{37}, along with many other results, Golomb obtained a duality formula for the error of approximation to a multivariate function from the set of sums of univariate functions. Unfortunately, his proof had a gap, which was 24 years later pointed out by Marshall and O'Farrell \cite{107}. But the question if Golomb's formula was correct, remained unsolved. In Chapter 3, we show that Golomb's formula is correct, and moreover it holds in a stronger form. Chapter 4 tells us about some problems concerning generalized ridge functions $% g(\alpha _{1}(x_{1})+\cdot \cdot \cdot +\alpha _{d}(x_{d}))$ and linear superpositions. We consider the problem of representation of general functions by linear superpositions. We show that if some representation by linear superpositions, in particular by linear combinations of generalized ridge functions, holds for continuous functions, then it holds for all functions. This leads us to extensions of many superpositions theorems (such as the well-known Kolmogorov superposition theorem, Ostrand's superposition theorem, etc.) from continuous to arbitrarily behaved multivariate functions. Concerning generalized ridge functions, we see that every multivariate function can be written as a generalized ridge function or as a sum of finitely many such functions. We also study the uniqueness of representation of functions by linear superpositions. Chapter 5 is about neural network approximation. The analysis in this chapter is based on properties of ordinary and generalized ridge functions. We consider a single and two hidden layer feedforward neural network models with a restricted set of weights. Such network models are important from the point of view of practical applications. We study approximation properties of single hidden layer neural networks with weights varying on a finite set of directions and straight lines. We give several necessary and sufficient conditions for well approximation by such networks. For a set of weights consisting of two directions (and two straight lines), we show that there is a geometrically explicit solution to the problem. Regarding two hidden layer feedforward neural networks, we prove that two hidden layer neural networks with $d$ inputs, $d$ neurons in the first hidden layer, $2d+2 $ neurons in the second hidden layer and with a specifically constructed sigmoidal, infinitely differentiable and almost monotone activation function can approximate any continuous multivariate function with arbitrary precision. We show that for this approximation only a finite number of fixed weights (precisely, $d$ fixed weights) suffice. There are topics related to ridge functions that are not presented here. The glaring omission is that of interpolation at points and on straight lines by ridge functions. We also do not address, for example, questions of linear independence and spanning by linear combinations of ridge monomials in the spaces of homogeneous and algebraic polynomials of a fixed degree, integral representations of functions where the kernel is a ridge function, approximation algorithms for finding best approximations from spaces of linear combinations of ridge functions. These and similar topics may be found in the monograph by Pinkus \cite{117}. The reader may also consult the survey articles \cite{57,87,118}. \newpage \chapter{Properties of linear combinations of ridge functions} In this chapter, we consider approximation-theoretic problems arising in ridge function approximation. First we briefly review some results on approximation by sums of ridge functions with both fixed and variable directions. Then we analyze the problem of representability of an arbitrary multivariate function by linear combinations of ridge functions with fixed directions. In the special case of two fixed directions, we characterize a best uniform approximation from the set of sums of ridge functions with these directions. For a class of bivariate functions we use this result to construct explicitly a best approximation. Questions on existence of a best approximation are also studied. We also study problems of well approximation (approximation with arbitrary accuracy) and representation of continuous multivariate functions by sums of two continuous ridge functions. The reader will see the main difficulties and remained open problems in the uniform approximation by sums of more than two ridge functions. For $L_{2}$ approximation, a number of summands does not play such an essential role as it plays in the uniform approximation. In this case, it is known that a best approximation always exists and unique. For some special domains in $% \mathbb{R}^{d}$, we characterize and then construct the best approximation. We also give an explicit formula for the approximation error. \bigskip \section{A brief excursion into the approximation theory of ridge functions} In this section we briefly review some results on approximation properties of the sets $% \mathcal{R}\left( \mathbf{a}^{1},...,\mathbf{a}^{r}\right) $ and $\mathcal{R}% _{r}$. These results are presented without proofs but with discussions and complete references. We hope this section will whet the reader's appetite for the rest of these notes, where a more comprehensive study of concrete mathematical problems is provided. \subsection{$\mathcal{R}\left( \mathbf{a}^{1},...,\mathbf{a}^{r}\right) $ -- ridge functions with fixed directions} It is clear that well approximation of a multivariate function $f:X\rightarrow \mathbb{R}$ from some normed space by using elements of the set $\mathcal{R}\left( \mathbf{a}% ^{1},...,\mathbf{a}^{r}\right)$ is not always possible. The value of the approximation error depends not only on the approximated function $f$ but also on geometrical structure of the given set $X$. This poses challenging research problems on computing the error of approximation and constructing best approximations from $\mathcal{R}\left( \mathbf{a}^{1},...,\mathbf{a}^{r}\right)$. Serious difficulties arise when one attempts to solve these problems in continuous function spaces endowed with the uniform norm. For example, let us consider the algorithm for finding best approximations, called the \textit{Diliberto-Straus algorithm} (see \cite{18}). The essence of this algorithm is as follows. Let $X$ be a compact subset of $\mathbb{R}^{d}$ and $A_{i}$ be a best approximation operator from the space of continuous functions $C(X)$ to the subspace of ridge functions $G_{i}=\{g_{i}\left( \mathbf{a}^{i}\cdot \mathbf{x}\right) :~g_{i}\in C(\mathbb{R)},~\mathbf{x}% \in X\}$, $i=1,...,r.$ That is, for each function $f$ $\in C(X)$, the function $A_{i}f$ is a best approximation to $f$ from $G_{i}.$ Set \begin{equation*} Tf=(I-A_{r})(I-A_{r-1})\cdot \cdot \cdot (I-A_{1})f, \end{equation*}% where $I$ is the identity operator. It is clear that \begin{equation*} Tf=f-g_{1}-g_{2}-\cdot \cdot \cdot -g_{r}, \end{equation*}% where $g_{k}$ is a best approximation from $G_{k}$ to the function $% f-g_{1}-g_{2}-\cdot \cdot \cdot -g_{k-1}$, $k=1,...,r.$ Consider powers of the operator $T$: $T^{2},T^{3}$ and so on. Is the sequence $% \{T^{n}f\}_{n=1}^{\infty }$ convergent? In case of an affirmative answer, which function is the limit of $T^{n}f,$ as $n\rightarrow \infty$? One may expect that the sequence $\{T^{n}f\}_{n=1}^{\infty }$ converges to $% f-g^{\ast },$ where $g^{\ast }$ is a best approximation from $\mathcal{R}% \left( \mathbf{a}^{1},...,\mathbf{a}^{r}\right) $ to $f$. This conjecture was first stated by Diliberto and Straus \cite{26} in 1951 for the uniform approximation of a multivariate function, defined on the unit cube, by sums of univariate functions (that is, sums of ridge functions with the coordinate directions). But later it was shown by Aumann \cite{4} that the sequence generated by this algorithm may not converge if $r>2$. For $r=2$ and certain convex compact sets $X$, the sequence $\{\|T^{n}f\|\}_{n=1}^{\infty }$ converges to the approximation error $\|f-g_{0}\|$, where $g_{0}$ is a best approximation from $\mathcal{R}% \left( \mathbf{a}^{1},\mathbf{a}^{2}\right)$ (see \cite{61,117}). However, it is not yet clear whether $\|T^{n}f-(f-g_{0})\|$ converges to zero as $n\rightarrow \infty$. In the case $r>2$ no efficient algorithm is known for finding a best uniform approximation from $\mathcal{R}\left( \mathbf{a}^{1},...,\mathbf{a}% ^{r}\right)$. Note that in the $L_{2}$ metric, the Diliberto-Straus algorithm converges as desired for an arbitrary number of distinct directions. This also holds in the $L_{p}$ space setting, provided that $p>1$ and $\mathcal{R}\left( \mathbf{a}^{1},...,\mathbf{a}^{r}\right)$ is closed (see \cite{118}). But in the $L_{1}$ space setting, the alternating algorithm does not work even in the case of two directions (see \cite{117}). One of the basic problems concerning the approximation by sums of ridge functions with fixed directions is the problem of verifying if a given function $f$ belongs to the space $\mathcal{R}\left( \mathbf{a}^{1},...,% \mathbf{a}^{r}\right) $. This problem has a simple solution if the space dimension $d=2$ and a given function $f(x,y)$ has partial derivatives up to $% r$-th order. For the representation of $f(x,y)$ in the form% \begin{equation*} f(x,y)=\sum_{i=1}^{r}g_{i}(a_{i}x+b_{i}y), \end{equation*}% it is necessary and sufficient that \begin{equation*} \prod\limits_{i=1}^{r}\left( b_{i}\frac{\partial }{\partial x}-a_{i}\frac{% \partial }{\partial y}\right) f=0.\eqno(1.1) \end{equation*}% This recipe is also valid for continuous bivariate functions provided that the derivatives are understood in the sense of distributions. Unfortunately such a simple characterization does not carry over to the case of more than two variables. Below we provide two results concerning the general case of arbitrarily many variables. \bigskip \textbf{Proposition 1.1 }(Diaconis, Shahshahani \cite{25}). \textit{Let $% \mathbf{a}^{1},...,\mathbf{a}^{r}$ be pairwise linearly independent vectors in $\mathbb{R}^{d}. $ Let for $i=1,2,...,r$, $H^{i}$ denote the hyperplane $% \{\mathbf{c}\in \mathbb{R}^{d}$: $\mathbf{c\cdot a}^{i}=0\}.$ Then a function $f\in C^{r}(\mathbb{R}^{d})$ can be represented in the form} \begin{equation*} f(\mathbf{x})=\sum\limits_{i=1}^{r}g_{i}\left( \mathbf{a}^{i}\cdot \mathbf{x}% \right) +P(\mathbf{x}), \end{equation*}% \textit{where $P(\mathbf{x})$ is a polynomial of degree not more than $r$, if and only if} \begin{equation*} \prod\limits_{i=1}^{r}\sum_{s=1}^{d}c_{s}^{i}\frac{\partial f}{\partial x_{s}% }=0, \end{equation*}% \textit{for all vectors $\mathbf{c}^{i}=(c_{1}^{i},c_{2}^{i},...,c_{d}^{i})% \in H^{i}, $ $i=1,2,...,r.$} \bigskip There are examples showing that one cannot simply dispense with the polynomial $P(\mathbf{x})$ in the above proposition (see \cite{25}). In fact, a polynomial term appears in the sufficiency part of the proof of this proposition. Lin and Pinkus \cite{95} obtained more general result on the representation by sums of ridge functions with fixed directions. We need some notation to present their result. Each polynomial $p(x_{1},...,x_{d})$ generates the differential operator $p(\frac{% \partial }{\partial x_{1}},...,\frac{\partial }{\partial x_{d}}).$ Let $P(% \mathbf{a}^{1},...,\mathbf{a}^{r})$ denote the set of polynomials which vanish on all the lines $\{\lambda \mathbf{a}^{i},\lambda \in \mathbb{R}\},$ $i=1,...,r.$ Obviously, this is an ideal in the ring of all polynomials. Let $Q$ be the set of polynomials $q=q(x_{1},...,x_{d})$ such that $p(\frac{% \partial }{\partial x_{1}},...,\frac{\partial }{\partial x_{d}})q=0$, for all $p(x_{1},...,x_{d})\in P(\mathbf{a}^{1},...,\mathbf{a}^{r}).$ \bigskip \textbf{Proposition 1.2 }(Lin, Pinkus \cite{95}). \textit{Let $\mathbf{a}% ^{1},...,\mathbf{a}^{r}$ be pairwise linearly independent vectors in $% \mathbb{R}^{d}.$ A function $f\in C(\mathbb{R}^{d})$ can be expressed in the form} \begin{equation*} f(\mathbf{x})=\sum\limits_{i=1}^{r}g_{i}(\mathbf{a}^{i}\cdot \mathbf{x)}, \end{equation*}% \textit{if and only if $f$ belongs to the closure of the linear span of $Q.$} \bigskip In \cite{120}, A.Pinkus considered the problems of smoothness and uniqueness in ridge function representation. For a given function $f$ $\in $ $\mathcal{R% }\left( \mathbf{a}^{1},...,\mathbf{a}^{r}\right) $, he posed and answered the following questions. If $f$ belongs to some smoothness class, what can we say about the smoothness of the functions $g_{i}$? How many different ways can we write $f$ as a linear combination of ridge functions? These and similar problems will be extensively discussed in Chapter 2. The above problem of representation of fixed functions by sums of ridge functions gives rise to the problem of representation of some classes of functions by such sums. For example, one may consider the following problem. Let $X$ be a subset of the $d$-dimensional Euclidean space. Let $% C(X),$ $B(X),$ $T(X)$ denote the set of continuous, bounded and all real functions defined on $X$, respectively. In the first case, we additionally suppose that $X$ is a compact set. Let $\mathcal{R}_{c}\left( \mathbf{a}% ^{1},...,\mathbf{a}^{r}\right) $ and $\mathcal{R}_{b}\left( \mathbf{a}% ^{1},...,\mathbf{a}^{r}\right) $ denote the subspaces of $\mathcal{R}\left( \mathbf{a}^{1},...,\mathbf{a}^{r}\right) $ comprising only sums with continuous and bounded terms $g_{i}\left( \mathbf{a}^{i}\cdot \mathbf{x}% \right) $, $i=1,...,r$, respectively. The following questions naturally arise: For which sets $X$, $(1)$ $C(X)=\mathcal{R}_{c}\left( \mathbf{a}^{1},...,\mathbf{a}^{r}\right)$? $(2)$ $B(X)=\mathcal{R}_{b}\left( \mathbf{a}^{1},...,\mathbf{a}^{r}\right)$? $(3)$ $T(X)=\mathcal{R}\left( \mathbf{a}^{1},...,\mathbf{a}^{r}\right)$? The first two questions in a more general setting were answered in Sternfeld \cite{132,131}. The third question will be answered in the next section. Let us briefly discuss some results of Sternfeld concerning ridge function representation. These results have been mostly overlooked in the corresponding ridge function literature, as they have to do with more general superpositions of functions and do not directly mention ridge functions. Assume we are given directions $% \mathbf{a}^{1},...,\mathbf{a}^{r}\in \mathbb{R}^{d}\backslash \{\mathbf{0}\}$ and a set $X\subseteq \mathbb{R}^{d}.$ Following Sternfeld, we say that a family $F=\{\mathbf{a}^{1},...,% \mathbf{a}^{r}\}$ \textit{uniformly separates points} of $X$ if there exists a number $0<\lambda \leq 1$ such that for each pair $\{\mathbf{x}% _{j}\}_{j=1}^{m}$, $\{\mathbf{z}_{j}\}_{j=1}^{m}$ of disjoint finite sequences in $X$, there exists some direction $\mathbf{a}^{k}\in F$ so that if from the two sequences $\{\mathbf{a}^{k}\cdot \mathbf{x}_{j}\}_{j=1}^{m}$ and $\{\mathbf{a}^{k}\cdot \mathbf{z}_{j}\}_{j=1}^{m}$ we remove a maximal number of pairs of points $\mathbf{a}^{k}\cdot \mathbf{x}_{j_{1}}$ and $% \mathbf{a}^{k}\cdot \mathbf{z}_{j_{2}}$ with $\mathbf{a}^{k}\cdot \mathbf{x}% _{j_{1}}=\mathbf{a}^{k}\cdot \mathbf{z}_{j_{2}},$ then there remains at least $\lambda m$ points in each sequence (or, equivalently, at most $% (1-\lambda )m$ pairs can be removed). Sternfeld \cite{132}, in particular, proved that a family of directions $F=\{\mathbf{a}^{1},...,\mathbf{a}^{r}\}$ uniformly separates points of $X$ if and only if $\mathcal{R}_{b}\left( \mathbf{a}^{1},...,\mathbf{a}^{r}\right) =B(X)$. In \cite{132}, he also obtained a practically convenient sufficient condition for the equality $% \mathcal{R}_{b}\left( \mathbf{a}^{1},...,\mathbf{a}^{r}\right) =B(X).$ To describe his condition, define the set functions \begin{equation*} \tau _{i}(Z)=\{\mathbf{x}\in Z:~|p_{i}^{-1}(p_{i}(\mathbf{x}))\bigcap Z|\geq 2\}, \end{equation*}% where $Z\subset X,~p_{i}(\mathbf{x})=\mathbf{a}^{i}\cdot \mathbf{x}$, \ $% i=1,\ldots ,r,$ and $|Y|$ denotes the cardinality of a set $Y$. Define $\tau (Z)$ to be $\bigcap_{i=1}^{k}\tau _{i}(Z)$ and define $\tau ^{2}(Z)=\tau (\tau (Z))$, $\tau ^{3}(Z)=\tau (\tau ^{2}(Z))$ and so on inductively. \bigskip \textbf{Proposition 1.3 }(Sternfeld \cite{132}). \textit{If $\tau ^{n}(X)=\emptyset $ for some $n$, then $\mathcal{R}_{b}\left( \mathbf{a}% ^{1},...,\mathbf{a}^{r}\right) =B(X)$. If $X$ is a compact subset of $% \mathbb{R}^{d}$, and $\tau ^{n}(X)=\emptyset $ for some $n$, then $\mathcal{R% }_{c}\left( \mathbf{a}^{1},...,\mathbf{a}^{r}\right) =C(X)$.} \bigskip If $r=2$, the sufficient condition \textquotedblleft $\tau ^{n}(X)=\emptyset $ for some $n$" turns out to be also necessary. In this case, the equality $\mathcal{R}_{b}\left( \mathbf{a}^{1},\mathbf{a}^{2}\right) =B(X)$ is equivalent to the equality $\mathcal{R}_{c}\left( \mathbf{a}^{1},% \mathbf{a}^{2}\right) =C(X)$. In another work \cite{131}, \ Sternfeld obtained a measure-theoretic necessary and sufficient condition for the equality $\mathcal{R}_{c}\left( \mathbf{a}^{1},...,\mathbf{a}^{r}\right) =C(X)$. Let $p_{i}(\mathbf{x})=\mathbf{a}^{i}\cdot \mathbf{x}$, \ $% i=1,\ldots ,r$, $X$ be a compact set in $\mathbb{R}^{d}$ and $M(X)$ be a class of measures defined on some field of subsets of $X$. Following Sternfeld, we say that a family $F=\{% \mathbf{a}^{1},...,\mathbf{a}^{r}\}$ \textit{uniformly separates measures} of the class $M(X)$ if there exists a number $0<\lambda \leq 1$ such that for each measure $\mu $ in $M(X)$ the equality $\left\Vert \mu \circ p_{k}^{-1}\right\Vert \geq \lambda \left\Vert \mu \right\Vert $ holds for some direction $\mathbf{a}^{k}\in F$. Sternfeld \cite{134,131}, in particular, proved that the equality $\mathcal{R}_{c}\left( \mathbf{a}% ^{1},...,\mathbf{a}^{r}\right) =C(X)$ holds if and only if the family of directions $\{\mathbf{a}^{1},...,\mathbf{a}^{r}\}$ uniformly separates measures of the class $C(X)^{\ast }$ (that is, the class of regular Borel measures). In addition, he proved that $\mathcal{R}_{b}\left( \mathbf{a}^{1},...,% \mathbf{a}^{r}\right) =B(X)$ if and only if the family of directions $\{% \mathbf{a}^{1},...,\mathbf{a}^{r}\}$ uniformly separates measures of the class $l_{1}(X)$ (that is, the class of finite measures defined on countable subsets of $X$). Since $l_{1}(X)\subset C(X)^{\ast },$ the first equality $% \mathcal{R}_{c}\left( \mathbf{a}^{1},...,\mathbf{a}^{r}\right) =C(X)$ implies the second equality $\mathcal{R}_{b}\left( \mathbf{a}^{1},...,% \mathbf{a}^{r}\right) =B(X).$ The inverse is not true (see \cite{131}). We emphasize again that the above results of Sternfeld were obtained for more general functions, than linear combinations of ridge functions, namely for functions of the form $\sum_{i=1}^{r}g_{i}(h_{i}(x))$, where $h_{i}$ arbitrarily fixed functions (bounded or continuous) defined on $X.$ Such functions will be discussed in Chapter 4. \bigskip \subsection{$\mathcal{R}_{r}$ -- ridge functions with variable directions} Obviously, the\ set $\mathcal{R}_{c}\left( \mathbf{a}^{1},...,\mathbf{a}% ^{r}\right) $ is not dense in $C(\mathbb{R}^{d})$ in the topology of uniform convergence on compact subsets of $\mathbb{R}^{d}.$ Density here does not hold because the number of considered directions is finite. If consider all the possible directions, then the set $\mathcal{R}=span\{g(\mathbf{a}\cdot \mathbf{x)}:~g\in C(\mathbb{R)},~\mathbf{a}\in \mathbb{R}^{d}\backslash \{% \mathbf{0}\}\}$ will certainly be dense in the space $C(\mathbb{R}^{d})$ in the above mentioned topology. In order to be sure, it is enough to consider only the functions $e^{\mathbf{a}\cdot \mathbf{x}}\in \mathcal{R}$, the linear span of which is dense in $C(\mathbb{R}^{d})$ by the Stone-Weierstrass theorem. In fact, for density it is not necessary to comprise all directions. The following theorem shows how many directions in totality satisfy the density requirements. \bigskip \textbf{Proposition 1.4 }(Vostrecov and Kreines \cite{142}, Lin and Pinkus \cite{95}). \textit{For density of the set} \begin{equation*} \mathcal{R(A)}=span\{g(\mathbf{a}\cdot \mathbf{x)}:~g\in C(\mathbb{R)},~% \mathbf{a}\in \mathcal{A}\subset \mathbb{R}^{d}\} \end{equation*}% \textit{in $C(\mathbb{R}^{d})$ (in the topology of uniform convergence on compact sets) it is necessary and sufficient that the only homogeneous polynomial which vanishes identically on $\mathcal{A}$ is the zero polynomial.} \bigskip Since in the definition of $\mathcal{R(A)}$ we vary over all univariate functions$~g,$ allowing one direction $\mathbf{a}$ is equivalent to allowing all directions $k\mathbf{a}$ for every real $k$. Thus it is sufficient to consider only the set $\mathcal{A}$ of directions normalized to the unit sphere $S^{n-1}.$ For example, if $\mathcal{A}$ is a subset of the sphere $% S^{n-1},$ which contains an interior point (interior point with respect to the induced topology on $S^{n-1}$), then $\mathcal{R(A)}$ is dense in the space $C(\mathbb{R}^{d}).$ The proof of Proposition 1.4 highlights an important fact that the set $\mathcal{R(A)}$ is dense in $C(\mathbb{R}^{d})$ in the topology of uniform convergence on compact subsets if and only if $% \mathcal{R(A)}$ contains all the polynomials (see \cite{95}). Representability of polynomials by sums of ridge functions is a building block for many results. In many works (see, e.g., \cite{119}), the following fact is fundamental: Every multivariate polynomial $h(\mathbf{x}% )=h(x_{1},...,x_{d})$ of degree $k $ can be represented in the form \begin{equation*} h(\mathbf{x})=\sum\limits_{i=1}^{l}p_{i}(\mathbf{a}^{i}\cdot \mathbf{x),} \end{equation*}% where $p_{i}$ is a univariate polynomial, $\mathbf{a}^{i}\in \mathbb{R}^{d}$% , and $l=$ $\binom{d-1+k}{k}$. For example, for the representation of a bivariate polynomial of degree $k$, it is needed $k+1$ univariate polynomials and $k+1$ directions (see \cite{97}% ). The proof of this fact is organized so that the directions $\mathbf{a}^{i} $, $i=1,...,k+1$, are chosen once for all multivariate polynomials of $k$-th degree. At one of the seminars in the Technion -- Israel Institute of Technology in 2007, A. Pinkus posed two problems: 1) Can every multivariate polynomial of degree $k$ be represented by less than $l$ ridge functions? 2) How large is the set of polynomials represented by $l-1,$ $l-2,...$ ridge functions? Note that for bivariate polynomials the 1-st problem is solved positively, that is, the number $l=k+1$ can be reduced. Indeed, for a bivariate polynomial $P(x,y)$ of $k$-th degree, there exist many combinations of real numbers $c_{0},...,c_{k}$ such that \begin{equation*} \sum_{i=0}^{k}c_{i}\frac{\partial ^{k}}{\partial x^{i}\partial y^{k-i}}% P(x,y)=0. \end{equation*}% Further the numbers $c_{i}$, $i=0,...,k$, can be selected to enjoy the property that the polynomial $\sum_{i=0}^{k}c_{i}t^{i}$ has distinct real zeros. Then it is not difficult to verify that the differential operator $% \sum_{i=0}^{k}c_{i}\frac{\partial ^{k}}{\partial x^{i}\partial y^{k-i}}$ can be written in the form \begin{equation*} \prod\limits_{i=1}^{k}\left( b_{i}\frac{\partial }{\partial x}-a_{i}\frac{% \partial }{\partial y}\right) , \end{equation*}% for some pairwise linearly independent vectors $(a_{i},b_{i})$, $i=1,...,k$. Now from the above criterion (1.1) we obtain that the polynomial $P(x,y)$ can be represented as a sum of $k$ ridge functions. Note that the problem of representation of a multivariate algebraic polynomial $P(\mathbf{x})$ in the form $\sum_{i=0}^{r}g_{i}(\mathbf{a}^{i}\cdot \mathbf{x})$ with minimal $r$ was extensively studied in the monograph by Pinkus \cite{117}. In connection with the 2-nd problem of Pinkus, V. Maiorov \cite{103} studied certain geometrical properties of the manifold $\mathcal{R}_{r}$. Namely, he estimated the $\varepsilon $-entropy numbers in terms of smaller $% \varepsilon $-covering numbers of the compact class formed by the intersection of the class $\mathcal{R}_{r}$ with the unit ball in the space of polynomials of degree at most $s$ on $\mathbb{R}^{d}$. Let $E$ be a Banach space and let for $x\in E$ and $\delta >0,$ $S(x,\delta )$ denote the ball of radius $\delta $ centered at the point $x$. For any positive number $% \varepsilon $, the $\varepsilon $-covering number of a set $F$ in the space $% E$ represents the quantity \begin{equation*} L_{\varepsilon }(F,E)=\min \left\{ N:~\exists x_{1},...,x_{N}\in F\text{ such that }F\subset \bigcup_{i=1}^{N}S(x_{i},\varepsilon )\text{ }\right\}. \end{equation*} The $\varepsilon $-entropy of $F$ is defined as the number $H_{\varepsilon }(F,E)\overset{def}{=}$ $\log _{2}L_{\varepsilon }(F,E)$. The notion of $% \varepsilon $-entropy has been devised by A.N.Kolmogorov (see \cite{84,85}) to classify compact metric sets according to their massivity. In order to formulate Maiorov's result, let $\mathcal{P}_{s}^{d}$ be the space of all polynomials of degree at most $s$ on $\mathbb{R}^{d}$, $% L_{q}=L_{q}(I)$, $1\leq q\leq \infty $, be the space of $q$-integrable functions on the unit cube $I=[0,1]^{d}$ with the norm $\left\Vert f\right\Vert _{q}=\left( \int_{I}\left\vert f(x)\right\vert ^{q}dx\right) ^{1/q}$, $BL_{q}$ be the unit ball in the space $L_{q},$ and $B_{q}\mathcal{P% }_{s}^{d}=$ $BL_{q}\cap $ $\mathcal{P}_{s}^{d}$ be the unit ball in the space $\mathcal{P}_{s}^{d}$ equipped with the $L_{q}$ metric. \bigskip \textbf{Proposition 1.5 }(Maiorov \cite{103}). \textit{Let $r,s\in \mathbb{N} $, $1\leq q\leq \infty $, $0 < \varepsilon < 1$. The $\varepsilon $-entropy of the class $B_{q}\mathcal{P}_{s}^{d}\cap \mathcal{R}_{r}$ in the space $% L_{q}$ satisfies the inequalities} 1) \begin{equation*} c_{1}rs\leq \frac{H_{\varepsilon }(B_{q}\mathcal{P}_{s}^{d}\cap \mathcal{R}% _{r},L_{q})}{\log _{2}\frac{1}{\varepsilon }}\leq c_{2}rs\log _{2}\frac{% 2es^{d-1}}{r}, \end{equation*} \textit{for $r\leq s^{d-1}.$} 2) \begin{equation*} c_{1}^{^{\prime }}s^{d}\leq \frac{H_{\varepsilon }(B_{q}\mathcal{P}% _{s}^{d}\cap \mathcal{R}_{r},L_{q})}{\log _{2}\frac{1}{\varepsilon }}\leq c_{2}^{^{\prime }}s^{d}, \end{equation*} \textit{for $r>s^{d-1}.$ In these inequalities $c_{1},c_{2},c_{1}^{^{\prime }},c_{2}^{^{\prime }}$ are constants depending only on $d$.} \bigskip Let us consider $\mathcal{R}_{r}$ as a subspace of some normed linear space $% X$ endowed with the norm $\left\Vert \cdot \right\Vert _{X}.$ The error of approximation of a given function $f\in X$ by functions $g\in \mathcal{R}% _{r} $ is defined as follows \begin{equation*} E(f,\mathcal{R}_{r},X)\overset{def}{=}\underset{g\in \mathcal{R}_{r}}{\inf }% \left\Vert f-g\right\Vert _{X}. \end{equation*} Let $B^{d}$ denote the unit ball in the space $\mathbb{R}^{d}.$ Besides, let $\mathbb{Z}_{+}^{d}$ denote the lattice of nonnegative multi-integers in $% \mathbb{R}^{d}.$ For $k=(k_{1},...,k_{d})\in \mathbb{Z}_{+}^{d},$ set $% \left\vert k\right\vert =k_{1}+\cdot \cdot \cdot +k_{d}$, $\mathbf{x}^{% \mathbf{k}}=x_{1}^{k_{1}}\cdot \cdot \cdot x_{d}^{k_{d}}$ and \begin{equation*} D^{\mathbf{k}}=\frac{\partial ^{\left\vert k\right\vert }}{\partial ^{k_{1}}x_{1}\cdot \cdot \cdot \partial ^{k_{d}}x_{d}} \end{equation*} The Sobolev space $W_{p}^{m}(B^{d})$ is the space of functions defined on $% B^{d}$ with the norm \begin{equation*} \left\Vert f\right\Vert _{m,p}=\left\{ \begin{array}{c} \left( \sum_{0\leq \left\vert \mathbf{k}\right\vert \leq m}\left\Vert D^{% \mathbf{k}}f\right\Vert _{p}^{p}\right) ^{1/p},\text{ if }1\leq p<\infty \\ \max_{0\leq \left\vert \mathbf{k}\right\vert \leq m}\left\Vert D^{\mathbf{k}% }f\right\Vert _{\infty },\text{ if }p=\infty .% \end{array}% \right. \end{equation*} Here \begin{equation*} \left\Vert h(\mathbf{x})\right\Vert _{p}=\left\{ \begin{array}{c} \left( \int_{B^{n}}\left\vert h(\mathbf{x})\right\vert ^{p}d\mathbf{x}% \right) ^{1/p},\text{ if }1\leq p<\infty \\ ess\sup_{\mathbf{x}\in B^{d}}\left\vert h(\mathbf{x})\right\vert ,\text{ if }% p=\infty .% \end{array}% \right. \end{equation*} Let $S_{p}^{m}(B^{d})$ be the unit ball in $W_{p}^{m}(B^{d})$: \begin{equation*} S_{p}^{m}(B^{d})=\{f\in W_{p}^{m}(B^{d}):\left\Vert f\right\Vert _{m,p}\leq 1~\}. \end{equation*} In 1999, Maiorov \cite{102} proved the following result \bigskip \textbf{Proposition 1.6 }(Maiorov \cite{102}). \textit{Assume $m\geq 1$ and $% d\geq 2$. Then for each $r\in \mathbb{N}$ there exists a function $f\in S_{2}^{m}(B^{d})$ such that} \begin{equation*} E(f,\mathcal{R}_{r},L_{2})\geq Cr^{-m/(d-1)},\eqno(1.2) \end{equation*} \textit{where $C$ is a constant independent of $f$ and $r.$} \bigskip For $d=2,$ this inequality was proved by Oskolkov \cite{114}. In \cite{102}, Maiorov also proved that for each function $f\in S_{2}^{m}(B^{d})$ \begin{equation*} E(f,\mathcal{R}_{r},L_{2})\leq Cr^{-m/(d-1)}.\eqno(1.3) \end{equation*} Thus he established the following order for the error of approximation to functions in $S_{2}^{m}(B^{d})$ from the class $\mathcal{R}_{r}$: \begin{equation*} E(S_{2}^{m}(B^{d}),\mathcal{R}_{r},L_{2})\overset{def}{=}\sup_{f\in S_{2}^{m}(B^{d})}E(f,\mathcal{R}_{r},L_{2})\asymp r^{-m/(d-1)}. \end{equation*} Pinkus \cite{119} revealed that the upper bound (1.3) is also valid in the $L_{p} $ metric ($1\leq p\leq \infty $). In other words, for every function $f\in S_{p}^{m}(B^{d})$ \begin{equation*} E(f,\mathcal{R}_{r},L_{p})\leq Cr^{-m/(d-1)}. \end{equation*} These inequalities were successfully applied to some problems of approximation of multivariate functions by neural networks with a single hidden layer. Recall that such networks are given by the formula $% \sum_{i=1}^{r}c_{i}\sigma (\mathbf{w}^{i}\mathbf{\cdot x}-\theta _{i}).$ By $% \mathcal{M}_{r}(\sigma )$ let us denote the set of all single hidden layer networks with the activation function $\sigma $. That is, \begin{equation*} \mathcal{M}_{r}(\sigma )=\left\{ \sum_{i=1}^{r}c_{i}\sigma (\mathbf{w}^{i}% \mathbf{\cdot x}-\theta _{i}):~c_{i},\theta _{i}\in \mathbb{R},~\mathbf{w}% ^{i}\in \mathbb{R}^{d}\right\} . \end{equation*} The above results on ridge approximation from $\mathcal{R}_{r}$ enable us to estimate the rate with which the approximation error $E(f,\mathcal{M}% _{r}(\sigma ),L_{2})$ tends to zero. First note that $\mathcal{M}_{r}(\sigma )\subset \mathcal{R}_{r},$ since each function of the form $\sigma (\mathbf{% w\cdot x}-\theta )$ is a ridge function with the direction $\mathbf{w}$. Thus the lower bound (1.2) holds also for the set $\mathcal{M}_{r}(\sigma )$% : there exists a function $f\in S_{2}^{m}(B^{d})$ for which \begin{equation*} E(f,\mathcal{M}_{r}(\sigma ),L_{2})\geq Cr^{-m/(d-1)}. \end{equation*} It remains to see whether the upper bound (1.3) is valid for $\mathcal{M}% _{r}(\sigma )$. Clearly, it cannot be valid if $\sigma $ is an arbitrary continuous function. Here we are dealing with the question if there exists a function $\sigma ^{\ast }\in C(\mathbb{R})$, for which \begin{equation*} E(f,\mathcal{M}_{r}(\sigma ^{\ast }),L_{2})\leq Cr^{-m/(d-1)}. \end{equation*} This question is answered affirmatively by the following result. \bigskip \textbf{Proposition 1.7 }(Maiorov, Pinkus \cite{99}). \textit{There exists a function $\sigma ^{\ast }\in C(\mathbb{R})$ with the following properties} 1) \textit{$\sigma ^{\ast }$ is infinitely differentiable and strictly increasing;} 2) \textit{$\lim_{t\rightarrow \infty }\sigma^{\ast } (t)=1$ and $\lim_{t\rightarrow -\infty }\sigma^{\ast } (t)=0;$} 3) \textit{for every $g\in \mathcal{R}_{r}$ and $\varepsilon >0$ there exist $c _{i},\theta _{i}\in \mathbb{R}$ and $\mathbf{w}^{i}\in \mathbb{R}^{d}$ satisfying} \begin{equation*} \sup_{\mathbf{x}\in B^{d}}\left\vert g(\mathbf{x})-\sum_{i=1}^{r+d+1}c_{i}% \sigma^{\ast } (\mathbf{w}^{i}\mathbf{\cdot x}-\theta _{i})\right\vert <\varepsilon . \end{equation*} \bigskip Temlyakov \cite{138} considered the approximation from some certain subclass of $\mathcal{R}_{r}$ in $L_{2}$ metric. More precisely, he considered the approximation of a function $f\in L_{2}(D),$ where $D$ is the unit disk in $% \mathbb{R}^{2}$, by functions $\sum\limits_{i=1}^{r}g_{i}(\mathbf{a}% ^{i}\cdot \mathbf{x)}\in \mathcal{R}_{r}\cap L_{2}(D)$, which satisfy the additional condition $\left\Vert g_{i}(\mathbf{a}^{i}\cdot \mathbf{x)}% \right\Vert _{2}\leq B\left\Vert f\right\Vert _{2},$ $i=1,...,r$ ($B$ is a given positive number). Let $\sigma _{r}^{B}(f)$ be the error of this approximation. For this approximation error, the author of \cite{138} obtained upper and lower bounds. Let, for $\alpha >0,$ $H^{\alpha }(D)$ denote the set of all functions $f\in L_{2}(D)$, which can be represented in the form \begin{equation*} f=\sum_{n=1}^{\infty }P_{n}, \end{equation*}% where $P_{n}$ are bivariate algebraic polynomials of total degree $2^{n}-1$ satisfying the inequalities \begin{equation*} \left\Vert P_{n}\right\Vert _{2}\leq 2^{-\alpha n},\text{ }n=1,2,... \end{equation*} \bigskip \textbf{Proposition 1.8 }(Temlyakov \cite{138}). \textit{1) For every $f\in H^{\alpha }(D) $, we have} \begin{equation*} \sigma _{r}^{1}(f)\leq C(\alpha )r^{-\alpha }. \end{equation*} \textit{2) For any given $\alpha >0$, $B>0$, $r>1$, there exists a function $% f\in H^{\alpha }(D)$ such that} \begin{equation*} \sigma _{r}^{B}(f)\geq C(\alpha ,B)(r\ln r)^{-\alpha }. \end{equation*} \bigskip Petrushev \cite{116} proved the following interesting result: Let $X_{k}$ be the $k$ dimensional linear space of univariate functions in $L_{2}[-1,1],$ $% k=1,2,...$. Besides, let $B^{d}$ and $S^{d-1}$ denote correspondingly the unit ball and unit sphere in the space $\mathbb{R}^{d}$. If $X_{k}$ provides order of approximation $O(k^{-m})$ for univariate functions with $m$ derivatives in$\ L_{2}[-1,1]$ and $\Omega _{k}$ are appropriately chosen finite sets of directions distributed on $S^{d-1}$, then the space $% Y_{k}=span\{p_{k}(\mathbf{a}\cdot \mathbf{x}):~p_{k}\in X_{k},~\mathbf{a}\in \Omega _{k}\}$ will provide approximation of order $O(k^{-m-d/2+1/2})$ for every function $f\in L_{2}(B^{d})$ with smoothness of order $m+d/2-1/2$. Thus, Petrushev showed that the above form of ridge approximation has the same efficiency of approximation as the traditional multivariate polynomial approximation. Many other results concerning the approximation of multivariate functions by functions from the set $\mathcal{R}_{r}$ and their applications in neural network theory may be found in \cite{43,94,99,119,123}. \bigskip \section{Representation of multivariate functions by linear combinations of ridge functions} In this section we develop a technique for verifying if a multivariate function can be expressed as a sum of ridge functions with given directions. We also obtain a necessary and sufficient condition for the representation of all multivariate functions on a subset $X$ of $\mathbb{R}^{d}$ by sums of ridge functions with fixed directions. \subsection{Two representation problems} Let $X$ be a subset of ${{\mathbb{R}}}^{d}$ and $\{\mathbf{a}% ^{i}\}_{i=1}^{r} $ be arbitrarily fixed nonzero directions (vectors) in ${{% \mathbb{R}}}^{d}$. Consider the following set of linear combinations of ridge functions. \begin{equation*} \mathcal{R}(\mathbf{a}^{1},...,\mathbf{a}^{r};X)=\left\{ \sum\limits_{i=1}^{r}g_{i}(\mathbf{a}^{i}\cdot \mathbf{x}),~\mathbf{x}\in X,~g_{i}:\mathbb{R}\rightarrow \mathbb{R},~i=1,...,r\right\} \end{equation*}% In this section, we are going to deal with the following two problems: \bigskip \textbf{Problem 1.} \textit{What conditions imposed on $f:X\rightarrow \mathbb{R}$ are necessary and sufficient for the inclusion $f\in \mathcal{R}(% \mathbf{a}^{1},...,\mathbf{a}^{r};X)$?} \bigskip \textbf{Problem 2.} \textit{What conditions imposed on $X$ are necessary and sufficient that every function defined on $X$ belongs to the space $\mathcal{% R}(\mathbf{a}^{1},...,\mathbf{a}^{r};X)$?} \bigskip As noticed in Section 1.1, Problem 1 was considered for continuous functions in \cite{95} and a theoretical result was obtained. It was also noticed there that the similar problem of representation of $f$ in the form $% \sum_{i=1}^{r}g_{i}(\mathbf{a}^{i}\cdot \mathbf{x})+P(\mathbf{x})$ with polynomial $P(\mathbf{x})$ was solved for continuously differentiable functions in \cite{25}. Problem 2 was solved in \cite{10} for finite subsets $X$ of ${{\mathbb{R}}}^{d}$ and in \cite{81} for the case when $r=d$ and $% \mathbf{a}^{i}$ are the coordinate directions. Here we consider both Problem 1 and Problem 2 without imposing on $X$, $f$ and $r$ any conditions. In fact, we solve these problems for more general, than $\mathcal{R}(\mathbf{a}^{1},...,\mathbf{a}^{r};X)$, set of functions. Namely, we solve them for the set \begin{equation*} \mathcal{B}(X)=\mathcal{B}(h_{1},...,h_{r};X)=\left\{ \sum\limits_{i=1}^{r}g_{i}(h_{i}(x)),~x\in X,~g_{i}:\mathbb{R}\rightarrow \mathbb{R},~i=1,...,r\right\} , \end{equation*}% where $h_{i}:X\rightarrow {{\mathbb{R}}},~i=1,...,r,$ are arbitrarily fixed functions. In particular, the functions $h_{i},~i=1,...,r$, may be equal to scalar products of the variable $\mathbf{x}$ with some vectors $\mathbf{a}% ^{i}$, $i=1,...,r$. Only in this special case, we have $\mathcal{B}% (h_{1},...,h_{r};X)=\mathcal{R}(\mathbf{a}^{1},...,\mathbf{a}^{r};X).$ \bigskip \subsection{Cycles} The main idea leading to solutions of the above problems is in using new objects called \textit{cycles} with respect to $r$ functions $% h_{i}:X\rightarrow \mathbb{R},~i=1,...,r$ (and in particular, with respect to $r$ directions $\mathbf{a}^{1},...,\mathbf{a}^{r}$). In the sequel, by $% \delta _{A}$ we will denote the characteristic function of a set $\ A\subset \mathbb{R}.$ That is, \begin{equation*} \delta _{A}(y)=\left\{ \begin{array}{c} 1,~if~y\in A \\ 0,~if~y\notin A.% \end{array}% \right. \end{equation*} \bigskip \textbf{Definition 1.1.} \textit{Given a subset $X\subset \mathbb{R}^{d}$\ and functions $h_{i}:X\rightarrow \mathbb{R},~i=1,...,r$. A set of points $% \{x_{1},...,x_{n}\}\subset X$ is called a cycle with respect to the functions $h_{1},...,h_{r}$ (or, concisely, a cycle if there is no confusion), if there exists a vector $\lambda =(\lambda _{1},...,\lambda _{n})$ with the nonzero real coordinates $\lambda _{i},~i=1,...,n,$ such that% } \begin{equation*} \sum_{j=1}^{n}\lambda _{j}\delta _{h_{i}(x_{j})}=0,~i=1,...,r.\eqno(1.4) \end{equation*} \bigskip If $h_{i}=\mathbf{a}^{i}\cdot \mathbf{x}$, $i=1,...,r$, where $\mathbf{a}% ^{1},...,\mathbf{a}^{r}$ are some directions in $\mathbb{R}^{d}$, a cycle, with respect to the functions $h_{1},...,h_{r}$, is called a cycle with respect to the directions\textit{\ }$\mathbf{a}^{1},...,\mathbf{a}^{r}.$ Let for $i=1,...,r,$ the set $\{h_{i}(x_{j}),~j=1,...,n\}$ have $k_{i}$ different values. Then it is not difficult to see that Eq. (1.4) stands for a system of $\sum_{i=1}^{r}k_{i}$ homogeneous linear equations in unknowns $% \lambda _{1},...,\lambda _{n}.$ If this system has any solution with the nonzero components, then the given set $\{x_{1},...,x_{n}\}$ is a cycle. In the last case, the system has also a solution $m=(m_{1},...,m_{n})$ with the nonzero integer components $m_{i},~i=1,...,n.$ Thus, in Definition 1.1, the vector $\lambda =(\lambda _{1},...,\lambda _{n})$ can be replaced with a vector $m=(m_{1},...,m_{n})$ with $m_{i}\in \mathbb{Z}\backslash \{0\}.$ For example, the set $l=\{(0,0,0),~(0,0,1),~(0,1,0),~(1,0,0),~(1,1,1)\}$ is a cycle in $\mathbb{R}^{3}$ with respect to the functions $% h_{i}(z_{1},z_{2},z_{3})=z_{i},~i=1,2,3.$ The vector $\lambda $ in Definition 1.1 can be taken as $(2,1,1,1,-1).$ In case $r=2,$ the picture of cycles becomes more clear. Let, for example, $% h_{1}$ and $h_{2}$ be the coordinate functions on $\mathbb{R}^{2}.$ In this case, a cycle is the union of some sets $A_{k}$ with the property: each $% A_{k}$ consists of vertices of a closed broken line with the sides parallel to the coordinate axis. These objects (sets $A_{k}$) have been exploited in practically all works devoted to the approximation of bivariate functions by univariate functions, although under various different names (see ``bolt of lightning" in Section 1.3). If the functions $h_{1}$ and $h_{2}$ are arbitrary, the sets $A_{k}$ can be described as a trace of some point traveling alternatively in the level sets of $h_{1}$ and $h_{2},$ and then returning to its primary position. It should be remarked that in the case $% r>2,$ cycles do not admit such a simple geometric description. We refer the reader to Braess and Pinkus \cite{10} for the description of cycles when $r=3 $ and $h_{i}(\mathbf{x})=\mathbf{a}^{i}\cdot \mathbf{x},$ $\mathbf{x}\in \mathbb{R}^{2},~\mathbf{a}^{i}\in \mathbb{R}^{2}\backslash \{\mathbf{0}% \},~i=1,2,3.$ Let $T(X)$ denote the set of all functions on $X.$ With each pair $% \left\langle p,\lambda \right\rangle ,$ where $p=\{x_{1},...,x_{n}\}$ is a cycle in $X$ and $\lambda =(\lambda _{1},...,\lambda _{n})$ is a vector known from Definition 1.1, we associate the functional \begin{equation*} G_{p,\lambda }:T(X)\rightarrow \mathbb{R},~~G_{p,\lambda }(f)=\sum_{j=1}^{n}\lambda _{j}f(x_{j}). \end{equation*}% In the following, such pairs $\left\langle p,\lambda \right\rangle $ will be called \textit{cycle-vector pairs} of $X.$ It is clear that the functional $% G_{p,\lambda }$ is linear and $G_{p,\lambda }(g)=0$ for all functions $g\in \mathcal{B}(h_{1},...,h_{r};X).$ \bigskip \textbf{Lemma 1.1.} \textit{Let $X$ have cycles and $h_{i}(X)\cap h_{j}(X)=\varnothing ,$ for all $i,j\in \{1,...,r\},~i\neq j.$ Then a function $f:X\rightarrow \mathbb{R}$ belongs to the set $\mathcal{B}% (h_{1},...,h_{r};X)$ if and only if $G_{p,\lambda }(f)=0$ for any cycle-vector pair $\left\langle p,\lambda \right\rangle $ of $X.$} \bigskip \begin{proof} The necessity is obvious, since the functional $G_{p,\lambda }$ annihilates all members of $\mathcal{B}(h_{1},...,h_{r};X)$. Let us prove the sufficiency. Introduce the notation \begin{eqnarray*} Y_{i} &=&h_{i}(X),~i=1,...,r; \\ \Omega &=&Y_{1}\cup ...\cup Y_{r}. \end{eqnarray*} Consider the following set. \begin{equation*} \mathcal{L}=\{Y=\{y_{1},...,y_{r}\}:\text{if there exists }x\in X\text{ such that }h_{i}(x)=y_{i},~i=1,...,r\}\eqno(1.5) \end{equation*} Note that $\mathcal{L}$ is not a subset of $\Omega $. It is a set of some certain subsets of $\Omega .$ Each element of $\mathcal{L}$ is a set $Y=\{y_{1},...,y_{r}\}\subset \Omega $ with the property that there exists $% x\in X$ such that $h_{i}(x)=y_{i},~i=1,...,r.$ In what follows, all the points $x$ associated with $Y$ by (1.5) will be called $(\ast )$-points of $Y.$ It is clear that the number of such points depends on $Y$ as well as on the functions $h_{1},...,h_{r}$, and may be greater than 1. But note that if any two points $x_{1}$ and $x_{2}$ are $% (\ast )$-points of $Y$, then the set $\{x_{1}$, $x_{2}\}$ necessarily forms a cycle with the associated vector $\lambda _{0}=(1;-1).$ Indeed, if $x_{1}$ and $x_{2}$ are $(\ast )$-points of $Y$, then $h_{i}(x_{1})=h_{i}(x_{2})$, $% i=1,...,r,$ whence \begin{equation*} 1\cdot \delta _{h_{i}(x_{1})}+(-1)\cdot \delta _{h_{i}(x_{2})}\equiv 0,~i=1,...,r. \end{equation*} The last identity means that the set $p_{0}=\{x_{1},$ $x_{2}\}$ forms a cycle and $\lambda _{0}=(1;-1)$ is an associated vector. Then by the the sufficiency condition, $G_{p_{0},\lambda _{0}}(f)=0$, whcih yields that $f(x_{1})=f(x_{2})$. Let now $Y^{\ast }$ be the set of all $(\ast )$-points of $Y.$ Since we have already known that $f(Y^{\ast })$ is a single number, we can define the function \begin{equation*} t:\mathcal{L}\rightarrow \mathbb{R},~t(Y)=f(Y^{\ast }). \end{equation*}% Or, equivalently, $t(Y)=f(x),$ where $x$ is an arbitrary $(\ast )$-point of $% Y$. Consider now a class $\mathcal{S}$ of functions of the form $% \sum_{j=1}^{k}r_{j}\delta _{D_{j}},$ where $k$ is a positive integer, $r_{j}$ are real numbers and $D_{j}$ are elements of $\mathcal{L},~j=1,...,k.$ We fix neither the numbers $\ k,~r_{j},$ nor the sets $D_{j}.$ Clearly, $% \mathcal{S\ }$is a linear space. Over $\mathcal{S}$, we define the functional \begin{equation*} F:\mathcal{S}\rightarrow \mathbb{R},~F\left( \sum_{j=1}^{k}r_{j}\delta _{D_{j}}\right) =\sum_{j=1}^{k}r_{j}t(D_{j}). \end{equation*} First of all, we must show that this functional is well defined. That is, the equality \begin{equation*} \sum_{j=1}^{k_{1}}r_{j}^{\prime }\delta _{D_{j}^{\prime }}=\sum_{j=1}^{k_{2}}r_{j}^{\prime \prime }\delta _{D_{j}^{\prime \prime }} \end{equation*}% always implies the equality \begin{equation*} \sum_{j=1}^{k_{1}}r_{j}^{\prime }t(D_{j}^{\prime })=\sum_{j=1}^{k_{2}}r_{j}^{\prime \prime }t(D_{j}^{\prime \prime }). \end{equation*}% In fact, this is equivalent to the implication \begin{equation*} \sum_{j=1}^{k}r_{j}\delta _{D_{j}}=0\Longrightarrow \sum_{j=1}^{k}r_{j}t(D_{j})=0,~\text{for all }k\in \mathbb{N}\text{, }% r_{j}\in \mathbb{R}\text{, }D_{j}\subset \mathcal{L}\text{.}\eqno(1.6) \end{equation*} Suppose that the left-hand side of the implication (1.6) be satisfied. Each set $D_{j}$ consists of $r$ real numbers $y_{1}^{j},...,y_{r}^{j}$, $% j=1,...,k.$ By the hypothesis of the lemma, all these numbers are different. Therefore, \begin{equation*} \delta _{D_{j}}=\sum_{i=1}^{r}\delta _{y_{i}^{j}},~j=1,...,k.\eqno(1.7) \end{equation*}% Eq. (1.7) together with the left-hand side of (1.6) gives \begin{equation*} \sum_{i=1}^{r}\sum_{j=1}^{k}r_{j}\delta _{y_{i}^{j}}=0.\eqno(1.8) \end{equation*}% Since the sets $\{y_{i}^{1},y_{i}^{2},...,y_{i}^{k}\}$, $i=1,...,r,$ are pairwise disjoint, we obtain from (1.8) that \begin{equation*} \sum_{j=1}^{k}r_{j}\delta _{y_{i}^{j}}=0,\text{ }i=1,...,r.\eqno(1.9) \end{equation*} Let now $x_{1},...,x_{k}$ be some $(\ast )$-points of the sets $% D_{1},...,D_{k}$ respectively. Since by (1.5), $y_{i}^{j}=h_{i}(x_{j})$, for $i=1,...,r$ and $j=1,...,k,$ it follows from (1.9) that the set $% \{x_{1},...,x_{k}\}$ is a cycle. Then by the condition of the sufficiency, $% \sum_{j=1}^{k}r_{j}f(x_{j})=0.$ Hence $\sum_{j=1}^{k}r_{j}t(D_{j})=0.$ We have proved the implication (1.6) and hence the functional $F$ is well defined. Note that the functional $F$ is linear (this can be easily seen from its definition). Consider now the following space: \begin{equation*} \mathcal{S}^{\prime }=\left\{ \sum_{j=1}^{k}r_{j}\delta _{\omega _{j}}\right\} , \end{equation*}% where $k\in \mathbb{N}$, $r_{j}\in \mathbb{R}$, $\omega _{j}\subset \Omega .$ As above, we do not fix the parameters $k$, $r_{j}$ and $\omega _{j}.$ Clearly, the space $\mathcal{S}^{\prime }$ is larger than $\mathcal{S}$. Let us prove that the functional $F$ can be linearly extended to the space $% \mathcal{S}^{\prime }$. So, we must prove that there exists a linear functional $F^{\prime }:\mathcal{S}^{\prime }\rightarrow \mathbb{R}$ such that $F^{\prime }(x)=F(x)$, for all $x\in \mathcal{S}$. Let $H$ denote the set of all linear extensions of $F$ to subspaces of $\mathcal{S}^{\prime }$ containing $\mathcal{S}$. The set $H$ is not empty, since it contains a functional $F.$ For each functional $v\in H$, let $dom(v)$ denote the domain of $v$. Consider the following partial order in $H$: $v_{1}\leq v_{2}$, if $% v_{2}$ is a linear extension of $v_{1}$ from the space $dom(v_{1})$ to the space $dom(v_{2}).$ Let now $P$ be any chain (linearly ordered subset) in $H$% . Consider the following functional $u$ defined on the union of domains of all functionals $p\in P$: \begin{equation*} u:\bigcup\limits_{p\in P}dom(p)\rightarrow \mathbb{R},~u(x)=p(x),\text{ if }% x\in dom(p) \end{equation*} Obviously, this functional is well defined and linear. Besides, the functional $u$ provides an upper bound for $P.$ We see that the arbitrarily chosen chain $P$ has an upper bound. Then by Zorn's lemma, there is a maximal element $F^{\prime }\in H$. We claim that the functional $F^{\prime } $ must be defined on the whole space $\mathcal{S}^{\prime }$. Indeed, if $% F^{\prime }$ is defined on a proper subspace $\mathcal{D\subset }$ $\mathcal{% S}^{\prime }$, then it can be linearly extended to a space larger than $% \mathcal{D}$ by the following way: take any point $x\in \mathcal{S}^{\prime }\backslash \mathcal{D}$ and consider the linear space $\mathcal{D}^{\prime }=\{\mathcal{D}+\alpha x\}$, where $\alpha $ runs through all real numbers. For an arbitrary point $y+\alpha x\in \mathcal{D}^{\prime }$, set $% F^{^{\prime \prime }}(y+\alpha x)=F^{\prime }(y)+\alpha b$, where $b$ is any real number considered as the value of $F^{^{\prime \prime }}$ at $x$. Thus, we constructed a linear functional $F^{^{\prime \prime }}\in H$ satisfying $% F^{\prime }\leq F^{^{\prime \prime }}.$ The last contradicts the maximality of $F^{\prime }.$ This means that the functional $F^{\prime }$ is defined on the whole $\mathcal{S}^{\prime }$ and $F\leq F^{\prime }$ ($F^{\prime }$ is a linear extension of $F$). Define the following functions by means of the functional $F^{\prime }$: \begin{equation*} g_{i}:Y_{i}\rightarrow \mathbb{R},\text{ }g_{i}(y_{i})\overset{def}{=}% F^{\prime }(\delta _{y_{i}}),\text{ }i=1,...,r. \end{equation*}% Let $x$ be an arbitrary point in $X.$ Obviously, $x$ is a $(\ast )$-point of some set $Y=\{y_{1},...,y_{r}\}\subset \mathcal{L}.$ Thus, \begin{eqnarray*} f(x) &=&t(Y)=F(\delta _{Y})=F\left( \sum_{i=1}^{r}\delta _{y_{i}}\right) =F^{\prime }\left( \sum_{i=1}^{r}\delta _{y_{i}}\right) = \\ \sum_{i=1}^{r}F^{\prime }(\delta _{y_{i}}) &=&\sum_{i=1}^{r}g_{i}(y_{i})=\sum_{i=1}^{r}g_{i}(h_{i}(x)). \end{eqnarray*}% \end{proof} \bigskip \subsection{Minimal cycles and the main results} \textbf{Definition 1.2.} \textit{A cycle $p=\{x_{1},...,x_{n}\}$ is said to be minimal if $p$ does not contain any cycle as its proper subset.} \bigskip For example, the set $l=\{(0,0,0),~(0,0,1),~(0,1,0),~(1,0,0),~(1,1,1)\}$ considered above is a minimal cycle with respect to the functions $% h_{i}(z_{1},z_{2},z_{3})=z_{i},~i=1,2,3.$ Adding the point $(0,1,1)$ to $l$, we will have a cycle, but not minimal. The vector $\lambda $ associated with $l\cup \{(0,1,1)\}$ can be taken as $(3,-1,-1,-2,2,-1).$ A minimal cycle $p=\{x_{1},...,x_{n}\}$ has the following obvious properties: \begin{description} \item[(a)] \textit{The vector $\lambda $ associated with $p$ through Eq. (1.4) is unique up to multiplication by a constant;} \item[(b)] \textit{If in (1.4), $\sum_{j=1}^{n}\left\vert \lambda _{j}\right\vert =1,$ then all the numbers $\lambda _{j},~j=1,...,n,$ are rational.} \end{description} Thus, a minimal cycle $p$ uniquely (up to a sign) defines the functional \begin{equation*} ~G_{p}(f)=\sum_{j=1}^{n}\lambda _{j}f(x_{j}),\text{ \ }\sum_{j=1}^{n}\left% \vert \lambda _{j}\right\vert =1. \end{equation*} \bigskip \textbf{Lemma 1.2.} \textit{The functional $G_{p,\lambda }$ is a linear combination of functionals $G_{p_{1}},...,G_{p_{k}},$ where $p_{1},...,p_{k}$ are minimal cycles in $p.$} \bigskip \begin{proof} Let $\left\langle p,\lambda \right\rangle $ be a cycle-vector pair of $X$, where $p=\{x_{1},...,x_{n}\}$ and $\lambda =(\lambda _{1},...,\lambda _{n})$. Let $p_{1}=$ $\{y_{1}^{1},...,y_{s_{1}}^{1}\},$ $% s_{1}<n$, be a minimal cycle in $p$ and \begin{equation*} G_{p_{1}}(f)=\sum_{j=1}^{s_{1}}\nu _{j}^{1}f(y_{j}^{1}),\text{ }% \sum_{j=1}^{s_{1}}\left\vert \nu _{j}^{1}\right\vert =1. \end{equation*} Without loss of generality, we may assume that $y_{1}^{1}=x_{1}.$ Put \begin{equation*} t_{1}=\frac{\lambda _{1}}{\nu _{1}^{1}}. \end{equation*}% Then the functional $G_{p,\lambda }-t_{1}G_{p_{1}}$ has the form \begin{equation*} G_{p,\lambda }-t_{1}G_{p_{1}}=\sum_{j=1}^{n_{1}}\lambda _{j}^{1}f(x_{j}^{1}), \end{equation*}% where $x_{j}^{1}\in p$, $\lambda _{j}^{1}\neq 0$, $j=1,...,n_{1}$. Clearly, the set $l_{1}=\{x_{1}^{1},...,x_{n_{1}}^{1}\}$ is a cycle in $p$ with the associated vector $\lambda ^{1}=(\lambda _{1}^{1},...,\lambda _{n_{1}}^{1})$% . Besides, $x_{1}\notin l_{1}$. Thus, $n_{1}<n$ and $G_{l_{1},\lambda ^{1}}=$ $G_{p,\lambda }-t_{1}G_{p_{1}}$. If $l_{1}$ is minimal, then the proof is completed. Assume $l_{1}$ is not minimal. Let $p_{1}=$ $% \{y_{1}^{2},...,y_{s_{2}}^{2}\},$ $s_{2}<n_{1},$ be a minimal cycle in $% l_{1} $ and \begin{equation*} G_{p_{2}}(f)=\sum_{j=1}^{s_{2}}\nu _{j}^{2}f(y_{j}^{2}),\text{ }% \sum_{j=1}^{s_{2}}\left\vert \nu _{j}^{2}\right\vert =1. \end{equation*} Without loss of generality, we may assume that $y_{1}^{2}=x_{1}^{1}.$ Put \begin{equation*} t_{2}=\frac{\lambda _{1}^{1}}{\nu _{1}^{2}}. \end{equation*}% Then the functional $G_{l_{1},\lambda ^{1}}-t_{2}G_{p_{2}}$ has the form \begin{equation*} G_{l_{1},\lambda ^{1}}-t_{2}G_{p_{2}}=\sum_{j=1}^{n_{2}}\lambda _{j}^{2}f(x_{j}^{2}), \end{equation*}% where $x_{j}^{2}\in l_{1}$, $\lambda _{j}^{2}\neq 0$, $j=1,...,n_{2}$. Clearly, the set $l_{2}=\{x_{1}^{2},...,x_{n_{2}}^{2}\}$ is a cycle in $% l_{1} $ with the associated vector $\lambda ^{2}=(\lambda _{1}^{2},...,\lambda _{n_{2}}^{2})$. Besides, $x_{1}^{1}\notin l_{2}$. Thus, $n_{2}<n_{1}$ and $G_{l_{2},\lambda ^{2}}=$ $G_{l_{1},\lambda ^{1}}-t_{2}G_{p_{2}}.$ If $l_{2}$ is minimal, then the proof is completed. Let $l_{2}$ be not minimal. Repeating the above process for $l_{2}$, then for $l_{3}$, etc., after some $k-1$ steps we will come to a minimal cycle $% l_{k-1}$ and the functional \begin{equation*} G_{l_{k-1},\lambda ^{k-1}}=G_{l_{k-2},\lambda ^{k-2}}-t_{k-1}G_{p_{k-1}}=\sum_{j=1}^{n_{k-1}}\lambda _{j}^{k-1}f(x_{j}^{k-1}). \end{equation*}% Since the cycle $l_{k-1}$ is minimal, \begin{equation*} G_{l_{k-1},\lambda ^{k-1}}=t_{k}G_{l_{k-1}},\text{ \ where }% t_{k}=\sum_{j=1}^{n_{k-1}}\left\vert \lambda _{j}^{k-1}\right\vert . \end{equation*}% Now putting $p_{k}=l_{k-1}$ and considering the above chain relations between the functionals $G_{l_{i},\lambda ^{i}}$, $i=1,...,k-1,$ we obtain that \begin{equation*} G_{p,\lambda }=\sum_{i=1}^{k}t_{i}G_{p_{i}}. \end{equation*} \end{proof} \textbf{Theorem 1.1.} \textit{Assume $X\subset \mathbb{R}^{d}$ and $% h_{1},...,h_{r}$ are arbitrarily fixed real functions on $X.$ The following assertions are valid.} \textit{1) Let $X$ have cycles with respect to the functions $% h_{1},...,h_{r} $. A function $f:X\rightarrow \mathbb{R}$ belongs to the space $\mathcal{B}(h_{1},...,h_{r};X)$ if and only if $G_{p}(f)=0$ for any minimal cycle $p\subset X$.} \textit{2) Let $X$ have no cycles. Then $\mathcal{B}% (h_{1},...,h_{r};X)=T(X). $} \bigskip \begin{proof} 1) The necessity is clear. Let us prove the sufficiency. On the strength of Lemma 1.2, it is enough to prove that if $G_{p,\lambda }(f)=0$ for any cycle-vector pair $\left\langle p,\lambda \right\rangle $ of $X$, then $f\in \mathcal{B}(X).$ Consider a system of intervals $\{(a_{i},b_{i})\subset \mathbb{R}% \}_{i=1}^{r} $ such that $(a_{i},b_{i})\cap (a_{j},b_{j})=\varnothing $ for all the indices $i,j\in \{1,...,r\}$, $~i\neq j.$ For $i=1,...,r$, let $\tau _{i}$ be one-to-one mappings of $\mathbb{R}$\ onto $(a_{i},b_{i}).$ Introduce the following functions on $X$: \begin{equation*} h_{i}^{^{\prime }}(x)=\tau _{i}(h_{i}(x)),\text{ }i=1,...,r. \end{equation*} It is clear that any cycle with respect to the functions $h_{1},...,h_{r}$ is also a cycle with respect to the functions $h_{1}^{^{\prime }},...,h_{r}^{^{\prime }}$, and vice versa. Besides, $h_{i}^{\prime }(X)\cap h_{j}^{\prime }(X)=\varnothing ,$ for all $i,j\in \{1,...,r\},~i\neq j.$ Then by Lemma 1.1, \begin{equation*} f(x)=g_{1}^{\prime }(h_{1}^{\prime }(x))+\cdots +g_{r}^{\prime }(h_{r}^{\prime }(x)), \end{equation*}% where $g_{1}^{\prime },...,g_{r}^{\prime }$ are univariate functions depending on $f$. From the last equality we obtain that \begin{equation*} f(x)=g_{1}^{\prime }(\tau _{1}(h_{1}(x)))+\cdots +g_{r}^{\prime }(\tau _{r}(h_{r}(x)))=g_{1}(h_{1}(x))+\cdots +g_{r}(h_{r}(x)). \end{equation*}% That is, $f\in \mathcal{B}(X)$. 2) Let $f:X\rightarrow \mathbb{R}$ be an arbitrary function. First suppose that $h_{i}(X)\cap h_{j}(X)=\varnothing ,$ for all $i,j\in \{1,...,r\}$,$% ~i\neq j.$ In this case, the proof is similar to and even simpler than that of Lemma 1.1. Indeed, the set of all $(\ast )$-points of $Y$ consists of a single point, since otherwise we would have a cycle with two points, which contradicts the hypothesis of the 2-nd part of the theorem. Further, well definition of the functional $F$ becomes obvious, since the left-hand side of (1.6) also contradicts the nonexistence of cycles. Thus, as in the proof of Lemma 1.1, we can extend $F$ to the space $\mathcal{S}^{\prime }$ and then obtain the desired representation for the function $f$. Since $f$ is arbitrary, $T(X)=\mathcal{B}(X).$ Using the techniques from the proof of the 1-st part of the theorem, one can easily generalize the above argument to the case when the functions $% h_{1},...,h_{r}$ have arbitrary ranges. \end{proof} \bigskip \textbf{Theorem 1.2.} \textit{$\mathcal{B}(h_{1},...,h_{r};X)=T(X)$ if and only if $X$ has no cycles with respect to the functions }$h_{1},...,h_{r}$% \textit{.} \bigskip \begin{proof} The sufficiency immediately follows from Theorem 1.1. To prove the necessity, assume that $X$ has a cycle $p=\{x_{1},...,x_{n}\}$. Let $% \lambda =(\lambda _{1},...,\lambda _{n})$ be a vector associated with $p$ by Eq. (1.4). Consider a function $f_{0}$ on $X$ with the property: $% f_{0}(x_{i})=1,$ for indices $i$ such that $\lambda _{i}\,>0$ and $% f_{0}(x_{i})=-1,$ for indices $i$ such that $\lambda _{i}\,<0$. For this function, $G_{p,\lambda }(f_{0})\neq 0$. Then by Theorem 1.1, $f_{0}\notin \mathcal{B}(X)$. Hence $\mathcal{B}(X)\neq T(X)$. The contradiction shows that $X$ does not admit cycles. \end{proof} \subsection{Corollaries} From Theorems 1.1 and 1.2 we obtain the following corollaries for the ridge function representation. \bigskip \textbf{Corollary 1.1.} \textit{Assume $X\subset \mathbb{R}^{d}$ and $\mathbf{a}^{1},...,\mathbf{a}^{r}\in \mathbb{R}^{d}\backslash \{\mathbf{0}% \}$. The following assertions are valid.} \textit{1) Let $X$ have cycles with respect to the directions $% \mathbf{a}^{1},...,\mathbf{a}^{r}$. A function $f:X\rightarrow \mathbb{R}$ belongs to the space $\mathcal{R}(\mathbf{a}^{1},...,% \mathbf{a}^{r};X)$ if and only if $G_{p}(f)=0$ for any minimal cycle $p\subset X$.} \textit{2) Let $X$ have no cycles. Then every function $% f:X\rightarrow \mathbb{R}$ belongs to the space $\mathcal{R}(% \mathbf{a}^{1},...,\mathbf{a}^{r};X)$.} \bigskip \textbf{Corollary 1.2.} \textit{$\mathcal{R}(\mathbf{a}^{1},...,\mathbf{a}^{r};X)=T(X)$ if and only if $X$ has no cycles with respect to the directions $\mathbf{a}^{1},...,\mathbf{a}^{r}$.} \bigskip Note that solutions to Problems 1 and 2 are given by Corollaries 1.1 and 1.2, correspondingly. Although it is not always easy to find all cycles of a given set $X$ and even to know if $X$ possesses a single cycle, Corollaries 1.1 and 1.2 are of more practical than theoretical character. Particular cases of Problems 1 and 2 evidence in favor of our opinion. For example, for the problem of representation by sums of two ridge functions, the picture of cycles is completely describable (see the beginning of this section). The interpretation of cycles with respect to three directions in the plane can be found in Braess and Pinkus \cite{10}. A geometric description of cycles with respect to 4 and more directions is quite complicated and requires deep techniques from geometry and graph theory. This is not within the aim of our study. From the last corollary, it follows that if representation by sums of ridge functions with fixed directions $\mathbf{a}^{1},...,\mathbf{a}^{r}$ is valid in the class of continuous functions (or in the class of bounded functions), then such representation is valid in the class of all functions. For a rigid mathematical formulation of this result, let us introduce the notation: \begin{equation*} \mathcal{R}_{c}(\mathbf{a}^{1},...,\mathbf{a}^{r};X)=\left\{ \sum\limits_{i=1}^{r}g_{i}(\mathbf{a}^{i}\cdot \mathbf{x}),~\mathbf{x}\in X,~g_{i}(\mathbf{a}^{i}\cdot \mathbf{x})\in C(X\mathbb{)},~i=1,...,r\right\} \end{equation*} and \begin{equation*} \mathcal{R}_{b}(\mathbf{a}^{1},...,\mathbf{a}^{r};X)=\left\{ \sum\limits_{i=1}^{r}g_{i}(\mathbf{a}^{i}\cdot \mathbf{x}),~\mathbf{x}\in X,~g_{i}(\mathbf{a}^{i}\cdot \mathbf{x})\in B(X\mathbb{)},~i=1,...,r\right\} \end{equation*} Here $C(X)$ and $B(X)$ denote the spaces of continuous and bounded functions defined on $X\subset \mathbb{R}^{d}$ correspondingly (for the first space, the set $X$ is supposed to be compact). As we know (see Section 1.1) from the results of Sternfeld it follows that the equality $\mathcal{R}_{c}(% \mathbf{a}^{1},...,\mathbf{a}^{r};X)=C(X)$ implies the equality $\mathcal{R}% _{b}(\mathbf{a}^{1},...,\mathbf{a}^{r};X)=B(X).$ In other words, if every continuous function is represented by sums of ridge functions (with fixed directions!), then every bounded function also obeys such representation (with bounded summands). Corollaries 1.1 and 1.2 allow us to obtain the following result. \bigskip \textbf{Corollary 1.3.} \textit{Let $X$ be a compact subset of $\mathbb{R}% ^{d}$ and $\mathbf{a}^{1},...,\mathbf{a}^{r}$ be given directions in $% \mathbb{R}^{d}\backslash \{\mathbf{0}\}$. If $\mathcal{R}_{c}(\mathbf{a}% ^{1},...,\mathbf{a}^{r};X)=C(X),$ then $\mathcal{R}(\mathbf{a}^{1},...,% \mathbf{a}^{r};X)=T(X).$} \bigskip \begin{proof} If every continuous function defined on $X\subset \mathbb{R}^{d}$ is represented by sums of ridge functions with the directions $\mathbf{a}% ^{1},...,\mathbf{a}^{r}$, then it can be shown by applying the same idea (as in the proof of Theorem 1.2) that the set $X$ has no cycles with respect to the given directions. Only, because of continuity, Urysohn's great lemma should be taken into account. That is, it should be taken into account that, by assuming the existence of a cycle $p_{0}=\{x_{1},...,x_{n}\}$ with an associated vector $\lambda _{0}=(\lambda _{1},...,\lambda _{n})$, we can deduce from Urysohn's great lemma the existence of a continuous function $u:X\rightarrow \mathbb{R}$ satisfying 1) $u(x_{i})=1,$ for indices $i$ such that $\lambda _{i}\,>0$ 2) $u(x_{j})=-1,$ for indices $j$ such that $\lambda _{j}\,<0$, 3) $-1<u(x)<1,$ for all $x\in X\backslash p_{0}.$ These properties mean that $G_{p_{0},\lambda _{0}}(u)\neq 0\Longrightarrow u\notin \mathcal{R}_{c}(\mathbf{a}^{1},...,\mathbf{a}% ^{r};X)\Longrightarrow \mathcal{R}_{c}(\mathbf{a}^{1},...,\mathbf{a}% ^{r};X)\neq C(X).$ But if $X$ has no cycles with respect to the directions $\mathbf{a}^{1},...,% \mathbf{a}^{r}$, then by Corollary 1.2, $\mathcal{R}(\mathbf{a}^{1},...,% \mathbf{a}^{r};X)=T(X).$ \end{proof} Let us now give some examples of sets over which the representation by linear combinations of ridge functions is possible. \begin{description} \item[(1)] Let $r=2$ and $X$ be the union of two parallel lines not perpendicular to the directions $\mathbf{a}^{1}$ and $\mathbf{a}^{2}$. Then $X$ has no cycles with respect to $\{\mathbf{a}^{1},\mathbf{a}^{2}\}$. Therefore, by Corollary 1.2, $\mathcal{R}\left( \mathbf{a}^{1},\mathbf{a}% ^{2};X\right) =T(X).$ \item[(2)] Let $r=2,$ $\mathbf{a}^{1}=(1,1)$, $\mathbf{a}^{2}=(1,-1)$ and $X$ be the graph of the function $y=\arcsin (\sin x)$. Then $X$ has no cycles and hence $\mathcal{R}\left( \mathbf{a}^{1},\mathbf{a}^{2};X\right) =T(X).$ \item[(3)] Assume now we are given $r$ directions $\{\mathbf{a}^{j}\}_{j=1}^{r}$ and $% r+1$ points $\{\mathbf{x}^{i}\}_{i=1}^{r+1}\subset \mathbb{R}^{d}$ such that% \begin{eqnarray*} \mathbf{a}^{1}\cdot \mathbf{x}^{i} &=&\mathbf{a}^{1}\cdot \mathbf{x}^{j}\neq \mathbf{a}^{1}\cdot \mathbf{x}^{2}\text{, \ for }1\leq i,j\leq r+1\text{, }% i,j\neq 2 \\ \mathbf{a}^{2}\cdot \mathbf{x}^{i} &=&\mathbf{a}^{2}\cdot \mathbf{x}^{j}\neq \mathbf{a}^{2}\cdot \mathbf{x}^{3}\text{, \ for }1\leq i,j\leq r+1\text{, }% i,j\neq 3 \\ &&\mathbf{......................................} \\ \mathbf{a}^{r}\cdot \mathbf{x}^{i} &=&\mathbf{a}^{r}\cdot \mathbf{x}^{j}\neq \mathbf{a}^{r}\cdot \mathbf{x}^{r+1}\text{, \ for }1\leq i,j\leq r. \end{eqnarray*}% The simplest data realizing these equations are the basis directions in $% \mathbb{R}^{d}$ and the points $(0,0,...,0)$, $(1,0,...,0)$, $(0,1,...,0)$% ,..., $(0,0,...,1)$. From the first equation we obtain that $\mathbf{x}^{2}$ cannot be a point of any cycle in $X=\{\mathbf{x}^{1},...,\mathbf{x}^{r+1}\}$% . Sequentially, from the second, third, ..., $r$-th equations it follows that the points $\mathbf{x}^{3},\mathbf{x}^{4},...,\mathbf{x}^{r+1}$ also cannot be points of cycles in $X$, respectively. Thus the set $X$ does not contain cycles at all. By Corollary 1.2, $\mathcal{R}\left( \mathbf{a}% ^{1},...,\mathbf{a}^{r};X\right) =T(X).$ \item[(4)] Assume we are given directions $\{\mathbf{a}^{j}\}_{j=1}^{r}$ and a curve $% \gamma $ in $\mathbb{R}^{d}$ such that for any $c\in \mathbb{R}$, $\gamma $ has at most one common point with at least one of the hyperplanes $\mathbf{a}% ^{j}\cdot \mathbf{x}=c$, $j=1,...,r.$ Clearly, the curve $\gamma $ has no cycles and hence $\mathcal{R}\left( \mathbf{a}^{1},...,\mathbf{a}^{r};\gamma \right) =T(\gamma ).$ \end{description} Braess and Pinkus \cite{10} considered the partial case of Problem 2: characterize a set of points $\left( \mathbf{x}^{1},...,\mathbf{x}% ^{k}\right) \subset \mathbb{R}^{d}$ such that for any data $\{\alpha _{1},...,\alpha _{k}\}\subset \mathbb{R}$ there exists a function $g\in \mathcal{R}\left( \mathbf{a}^{1},...,\mathbf{a}^{r};\mathbb{R}^{d}\right) $ satisfying $g(\mathbf{x}^{i})=\alpha _{i},$ $i=1,...,k$. In connection with this problem, they introduced the notion of the \textit{NI}-property (non interpolation property) and \textit{MNI}-property (minimal non interpolation property) of a finite set of points as follows: Given directions $\{\mathbf{a}^{j}\}_{j=1}^{r}\subset \mathbb{R}% ^{d}\backslash \{\mathbf{0}\}$, we say that a set of points $\{\mathbf{x}% ^{i}\}_{i=1}^{k}\subset \mathbb{R}^{d}$ has the \textit{NI}-property with respect to $\{\mathbf{a}^{j}\}_{j=1}^{r}$, if there exists $\{\alpha _{i}\}_{i=1}^{k}\subset \mathbb{R}$ such that we cannot find a function $% g\in \mathcal{R}\left( \mathbf{a}^{1},...,\mathbf{a}^{r};\mathbb{R}% ^{d}\right) $ satisfying $g(\mathbf{x}^{i})=\alpha _{i},$ $i=1,...,k$. We say that \ the set $\{\mathbf{x}^{i}\}_{i=1}^{k}\subset \mathbb{R}^{d}$ has the \textit{MNI}-property with respect to $\{\mathbf{a}^{j}\}_{j=1}^{r}$, if $\{\mathbf{x}^{i}\}_{i=1}^{k}$ but no proper subset thereof has the \textit{% NI}-property. It follows from Corollary 1.2 that a set $\{\mathbf{x}^{i}\}_{i=1}^{k}$ has the \textit{NI}-property if and only if $\{\mathbf{x}^{i}\}_{i=1}^{k}$ contains a cycle with respect to the functions $h_{i}=\mathbf{a}^{i}\cdot \mathbf{x},$ $i=1,...,r$ (or, simply, to the directions $\mathbf{a}^{i},$ $% i=1,...,r$) and the \textit{MNI}-property if and only if the set $\{\mathbf{x% }^{i}\}_{i=1}^{k}$ itself is a minimal cycle with respect to the given directions. Taking into account this argument and Definitions 1.1 and 1.2, we obtain that the set $\{\mathbf{x}^{i}\}_{i=1}^{k}$ has the \textit{NI}% -property if and only if there is a vector $\mathbf{m}=(m_{1},...,m_{k})\in \mathbb{Z}^{k}\backslash \{\mathbf{0}\}$ such that \begin{equation*} \sum_{j=1}^{k}m_{j}g(\mathbf{a}^{i}\cdot \mathbf{x}^{j})=0, \end{equation*}% for $i=1,...,r$ and all functions $g:\mathbb{R\rightarrow R}$. This set has the \textit{MNI}-property if and only if the vector $\mathbf{m}$ has the additional properties: it is unique up to multiplication by a constant and all its components are different from zero. This special consequence of Corollary 1.2 was proved in \cite{10}. \bigskip \section{Characterization of an extremal sum of ridge functions} The approximation problem considered in this section is to approximate a continuous multivariate function $f\left( \mathbf{x}\right) =f\left( {% x_{1},...,x_{d}}\right) $ by sums of two ridge functions in the uniform norm. We give a necessary and sufficient condition for a sum of two ridge functions to be a best approximation to $f\left( \mathbf{x}\right) .$ This main result is next used in a special case to obtain an explicit formula for the approximation error and to construct a best approximation. The problem of well approximation by such sums is also considered. \subsection{Exposition of the problem} Consider the following set of sums of ridge functions \begin{equation*} \mathcal{R}=\mathcal{R}\left( \mathbf{a}^{1},\mathbf{a}^{2}\right) ={\left\{ {% g_{1}\left( \mathbf{a}^{1}{\cdot }\mathbf{x}\right) +g_{2}\left( \mathbf{a}^{2}{% \cdot }\mathbf{x}\right) :g}_{i}{\in C\left( {\mathbb{R}}\right) ,i=1,2}% \right\} }. \end{equation*} That is, we fix directions $\mathbf{a}^{1}$ and $\mathbf{a}^{2}$ and consider linear combinations of ridge functions with these directions. Assume $f\left( \mathbf{x}\right) $ is a continuous function on a compact subset $Q$ of $\mathbb{R}^{d}$. We want to find conditions that are necessary and sufficient for a function $g_{_{0}}\in \mathcal{R}\left( \mathbf{a}^{1},\mathbf{a}^{2}\right) $ to be an extremal element (or a best approximation) to $f$. In other words, we want to characterize such sums $% g_{0}\left( \mathbf{x}\right) =g_{1}\left( \mathbf{a}^{1}{\cdot }\mathbf{x}% \right) +g_{2}\left( \mathbf{a}^{2}{\cdot }\mathbf{x}\right) $ of ridge functions that \begin{equation*} {\left\Vert {f-g_{0}}\right\Vert }={\max\limits_{{\mathbf{x}\in Q}}}{% \left\vert {f\left( \mathbf{x}\right) -g}_{{0}}{\left( \mathbf{x}\right) }% \right\vert }=E\left( {f}\right) , \end{equation*}% where \begin{equation*} E\left( {f}\right) =E(f,\mathcal{R})\overset{def}{=}{\inf_{g \in \mathcal{R}\left( \mathbf{a}^{1},\mathbf{a}^{2}\right)}}{\left\Vert {f-g}% \right\Vert } \end{equation*}% is the error in approximating from $\mathcal{R}\left( \mathbf{a}^{1},\mathbf{a}^{2}% \right) .$ The other related problem is how to construct these sums of ridge functions. We also want to know if we can approximate well, i.e. for which compact sets $Q,$ $\mathcal{R}\left( \mathbf{a}^{1},\mathbf{a}^{2}\right) $ is dense in $C\left( {Q}\right) $ in the topology of uniform convergence. It should be remarked that solutions to these problems may be useful in connection with the study of partial differential equations. For example, assume that $\left( {a_{1},b_{1}}\right) $ and $\left( {a_{2},b_{2}}\right) $ are linearly independent vectors in $\mathbb{R}^{2}.$ Then the general solution to the homogeneous partial differential equation \begin{equation*} \left( {a_{1}{\frac{\partial }{\partial {x}}}+b_{1}{\frac{\partial }{% \partial {y}}}}\right) \left( {a_{2}{\frac{\partial }{\partial {x}}}+b_{2}{% \frac{\partial }{\partial {y}}}}\right) {u}\left( {x,y}\right) =0\eqno(1.10) \end{equation*}% are all functions of the form \begin{equation*} u\left( {x,y}\right) =g_{1}\left( {b_{1}x-a_{1}y}\right) +g_{_{2}}\left( {% b_{2}x-a_{2}y}\right) \eqno(1.11) \end{equation*}% for arbitrary $g_{1}$ and $g_{2}.$ In \cite{36}, Golitschek and Light described an algorithm that computes the error of approximation of a continuous function $f\left( {x,y}\right) $ by solutions of equation (1.10), provided that $a_{1}=b_{2}=1$, $a_{2}=b_{1}=0.$ Using our result (see Theorem 1.3), one can characterize those solutions (1.11) that are extremal to a given function $f(x,y)$. For a certain class of functions $f(x,y) $, one can also easily calculate the approximation error and construct an extremal solution (see Theorems 1.5 and 1.6 below). The problem of approximating by functions from the set $\mathcal{R}\left( \mathbf{a}^{1},\mathbf{a}^{2}\right) $ arises in other contexts too. Buck \cite{11} studied the classical functional equation: given $\beta (t)\in C[0,1]$, $% 0\leq \beta (t)\leq 1$, for which $u\in C[0,1]$ does there exist $\varphi \in C[0,1]$ such that \begin{equation*} \varphi (t)=\varphi \left( \beta (t)\right) +u(t)? \end{equation*}% He proved that the set of all $u$ satisfying this condition is dense in the set \begin{equation*} \{v\in C[0,1]:\ v(t)=0\ \mbox{whenever}\ \beta (t)=t\} \end{equation*}% if and only if $\mathcal{R}\left( \mathbf{a}^{1},\mathbf{a}^{2}\right) $ with the unit directions $\mathbf{a}^{1}=(1;0)$ and $\mathbf{a}^{2}=(0,1)$ is dense in $C(K)$% , where $K=\{(x,y):y=x\ \mbox{or}\ y=\beta (x),\ 0\leq x\leq 1\}$. Although there are enough reasons to consider approximation problems associated with the set $\mathcal{R}\left( \mathbf{a}^{1},\mathbf{a}^{2}\right) $ in an independent way, one may ask why sums of only two ridge functions are considered instead of sums with an arbitrary number of terms. We will try to answer this fair question in Section 1.3.4. \ \subsection{The characterization theorem} Let $Q$ be a compact subset of $\mathbb{R}^{d}$ and $\mathbf{a}^{1},\mathbf{a}^{2}% \in \mathbb{R}^{d}\backslash {\left\{ \mathbf{0}\right\} }.$ \bigskip \textbf{Definition 1.3.} \textit{A finite or infinite ordered set $p=\left( \mathbf{p}{_{1},\mathbf{p}_{2},...}\right) \subset Q$ with $\mathbf{p}% _{i}\neq \mathbf{p}_{i+1},$ and either $\mathbf{a}^{1}\cdot \mathbf{p}_{1}=% \mathbf{a}^{1}\cdot \mathbf{p}_{2},\mathbf{a}^{2}\cdot \mathbf{p}_{2}=\mathbf{a}^{2}% \cdot \mathbf{p}_{3},\mathbf{a}^{1}\cdot \mathbf{p}_{3}=\mathbf{a}^{1}\cdot \mathbf{p% }_{4},...$ or $\mathbf{a}^{2}\cdot \mathbf{p}_{1}=\mathbf{a}^{2}\cdot \mathbf{p}% _{2},~\mathbf{a}^{1}\cdot \mathbf{p}_{2}=\mathbf{a}^{1}\cdot \mathbf{p}_{3}, \mathbf{a}^{2}\cdot \mathbf{p}_{3}=\mathbf{a}^{2}\cdot \mathbf{p}_{4},...$is called a path with respect to the directions $\mathbf{a}^{1}$ and $\mathbf{a}^{2}$.} \bigskip This notion (in the two-dimensional case) was introduced by Braess and Pinkus \cite{10}. They showed that paths give geometric means of deciding if a set of points ${\left\{ {\mathbf{x}}^{i}\right\} }_{i=1}^{m}\subset \mathbb{R}% ^{2}$ has the \textit{NI} property (see Section 1.2.4). Ismailov and Pinkus \cite{63} used these objects to study the problem of interpolation on straight lines by linear combinations of a finite number of ridge functions with fixed directions. In \cite{51,53,44} paths were generalized to those with respect to two functions. The last objects turned out to be useful in problems of approximation and representation by sums of compositions of fixed multivariate functions with univariate functions. If $\mathbf{a}^{1}$ and $\mathbf{a}^{2}$ are the coordinate vectors in $\mathbb{R}% ^{2}$, then Definition 1.3 defines a \textit{bolt of lightning}. The idea of bolts was first introduced in Diliberto and Straus \cite{26}, where these objects are called \textit{permissible lines}. They appeared further in a number of papers, although under several different names (see, e.g., \cite% {29,34,36,55,56,79,78,76,82,93,108,107,113}). Note that the term \textquotedblleft bolt of lightning" is due to Arnold \cite{3}. For the sake of brevity, we use the term ``path" instead of the long expression ``path with respect to the directions $\mathbf{a}^{1}$ and $\mathbf{a}^{2}$". The length of a path is the number of its points. A single point is a path of the unit length. A finite path $\left( \mathbf{p}_{1} ,\mathbf{p}_{2} ,...,\mathbf{p}_{2n} \right)$ is said to be closed if $\left(\mathbf{p}_{1} ,% \mathbf{p}_{2} ,...,\mathbf{p}_{2n}, \mathbf{p}_{1}\right)$ is a path. We associate each closed path $p=\left(\mathbf{p}_{1}, \mathbf{p}_{2} ,...,% \mathbf{p}_{2n} \right) $ with the functional \begin{equation*} G_{p} (f)=\frac{1}{2n} \sum\limits_{k=1}^{2n}(-1)^{k+1} f(\mathbf{p}_{k}). \end{equation*} This functional has the following obvious properties: (a) If $g\in \mathcal{R}\left( \mathbf{a}^{1},\mathbf{a}^{2}\right)$, then $G_{p} (g)=0$. (b) $\left\| G_{p} \right\| \leq 1$ and if $\mathbf{p}_{i} \neq \mathbf{p}_{j}$ for all $% i\neq j,$ $1\leq i,j\leq 2n$ , then $\left\| G_{p} \right\| =1$. \bigskip \textbf{Lemma 1.3.} \textit{Let a compact set $Q$ have closed paths. Then \begin{equation*} \sup\limits_{p\subset Q}\left\vert G_{p}(f)\right\vert \leq E\left( f\right) ,\eqno(1.12) \end{equation*}% where the sup is taken over all closed paths. Moreover, inequality (1.12) is sharp, i.e. there exist functions for which (1.12) turns into equality.} \bigskip \begin{proof} Let $p$ be a closed path in $Q$ and $g$ be any function from $% \mathcal{R}\left( \mathbf{a}^{1},\mathbf{a}^{2}\right) $. By the linearity of $G_{p}$ and properties (a) and (b), \begin{equation*} \left\vert G_{p}(f)\right\vert =\left\vert G_{p}(f-g)\right\vert \leq \left\Vert f-g\right\Vert .\eqno(1.13) \end{equation*}% Since the left-hand and the right-hand sides of (1.13) do not depend on $g$ and $p$ respectively, it follows from (1.13) that \begin{equation*} \sup_{p\subset Q}\left\vert G_{p}(f)\right\vert \leq \inf_{g \in \mathcal{R}\left( \mathbf{a}^{1},\mathbf{a}^{2}\right) }\left\Vert f-g\right\Vert .\eqno(1.14) \end{equation*} Now we prove the sharpness of (1.12). By assumption $Q$ has closed paths. Then $Q$ has a closed path $p^{\prime }=\left( \mathbf{p}_{1}^{\prime },...,% \mathbf{p}_{2m}^{\prime }\right)$ with distinct points $\mathbf{p}_{1}^{\prime },...,% \mathbf{p}_{2m}^{\prime}$. In fact, such a special path can be obtained from any closed path $p=\left( \mathbf{p}_{1},...,\mathbf{p}_{2n}\right) $ by the following simple algorithm: if the points of the path $p$ are not all distinct, let $i$ and $k>0$ be the minimal indices such that $\mathbf{p}_{i}=% \mathbf{p}_{i+2k}$; delete from $p$ the subsequence $\mathbf{p}_{i+1},...,% \mathbf{p}_{i+2k}$ and call $p$ the obtained path; repeat the above step until all points of $p$ are all distinct; set $p^{\prime }:=p$. On the other hand there exist continuous functions $h=h(\mathbf{x})$ on $Q$ such that $h(% \mathbf{p}_{i}^{\prime })=1$, $i=1,3,...,2m-1$, $h(\mathbf{p}_{i}^{\prime })=-1$, $i=2,4,...,2m$ and $-1<h(\mathbf{x})<1$ elsewhere. For such functions we have \begin{equation*} G_{p^{\prime }}(h)=\Vert h\Vert =1\eqno(1.15) \end{equation*}% and \begin{equation*} E(h)\leq \Vert h\Vert ,\eqno(1.16) \end{equation*}% where the last inequality follows from the fact that $0\in \mathcal{R}\left( \mathbf{a}^{1},\mathbf{a}^{2}\right) .$ From (1.14)-(1.16) it follows that \begin{equation*} \sup_{p\subset Q}\left\vert G_{p}(h)\right\vert =E\left( h\right) . \end{equation*}% \end{proof} \textbf{Lemma 1.4.} \textit{Let $Q$ be a convex compact subset of $% \mathbb{R}^{d}$ and $f \in C(Q)$. For a vector $\mathbf{e}\in \mathbb{R}^{d}\backslash \mathbf{\{0\}}$ and a real number $t$ set \begin{equation*} Q_{t}=\left\{ {\mathbf{x}}\in Q:\mathbf{e}\cdot \mathbf{x}=t\right\} ,\ \ \ T_{h}=\left\{ t\in \mathbb{R}:Q_{t}\neq \emptyset \right\} . \end{equation*}% The functions \begin{equation*} g_{1} (t)=\max_{\mathbf{x} \in Q_t} f(\mathbf{x} ),\ \ t\in T_{h}\ \ % \mbox{and}\ \ g_{2} (t)=\min\limits_{\mathbf{x} \in Q_t} f(\mathbf{x} ),\ \ \ t\in T_{h} \end{equation*} are defined and continuous on $T_{h} $.} \bigskip The proof of this lemma is not difficult and can be obtained by the well-known elementary methods of mathematical analysis. \bigskip \textbf{Definition 1.4.} \textit{A finite or infinite path $(\mathbf{p}_{1},% \mathbf{p}_{2},...)$ is said to be extremal for a function $u \in C(Q)$ if $u(\mathbf{p}_{i})=(-1)^{i}\left\Vert u\right\Vert ,i=1,2,...$ or $% u(\mathbf{p}_{i})=(-1)^{i+1}\left\Vert u\right\Vert ,$ $i=1,2,...$} \bigskip \textbf{Theorem 1.3.} \textit{Let $Q\subset \mathbb{R}^{d}$ be a convex compact set satisfying the following condition} \textit{Condition (A): For any path $q=(\mathbf{q}_{1},\mathbf{q}_{2},...,% \mathbf{q}_{n})\subset Q$ there exist points $\mathbf{q}_{n+1},\mathbf{q}% _{n+2},...,\mathbf{q}_{n+s}\in Q$ such that $(\mathbf{q}_{1},\mathbf{q}% _{2},...,\mathbf{q}_{n+s})$ is a closed path and $s$ is not more than some positive integer $N_{0}$ independent of $q$.} \textit{Then a necessary and sufficient condition for a function $g_{0}\in \mathcal{R}\left( \mathbf{a}^{1},\mathbf{a}^{2}\right) $ to be an extremal element to the given function $f \in C(Q)$ is the existence of a closed or infinite path $l=(\mathbf{p}_{1},\mathbf{p}_{2},...)$ extremal for the function $% f_{1}=f-g_{0}$.} \bigskip It should be remarked that the above condition (A) strongly depends on the fixed directions $\mathbf{a}^{1}$ and $\mathbf{a}^{2}$. For example, in the familiar case of a square $S\subset \mathbb{R}^2$ there are many directions which are not allowed. If it is possible to reach a corner of $S$ with not more than one of the two directions orthogonal to $\mathbf{a}^{1} $ and $\mathbf{a}^{2}$, respectively (we don't differentiate between directions $% \mathbf{c}$ and $-\mathbf{c}$), the triple $(S,\mathbf{a}^{1}, \mathbf{a}^{2})$ does not satisfy condition (A) of the theorem. Here are simple examples: Let $% S=[0;1]^2$, $\mathbf{a}^{1}=(1;0)$, $\mathbf{a}^{2}=(1;1)$. Then the ordered set $% \{(0;1), (1;0), (1;1)\}$ is a path in $S$ which can not be made closed. In this case, $(1;1)$ is not reached with the direction orthogonal to $\mathbf{b% }$. Let now $\mathbf{a}^{1}=\left(1;\frac{1}{2}\right)$, $\mathbf{a}^{2}=(1;1)$. Then the corner $(1;1)$ is reached with none of the directions orthogonal to $\mathbf{a}^{1}$ and $\mathbf{a}^{2}$ respectively. In this case, for any positive integer $N_0$ and any point $\mathbf{q}_0$ in $S$ one can chose a point $% \mathbf{q}_1\in S$ from a sufficiently small neighborhood of the corner $% (1;1)$ so that any path containing $\mathbf{q}_0$ and $\mathbf{q}_1$ has the length more than $N_0$. These examples and a little geometry show that if a convex compact set $Q\subset\mathbb{R}^2$ satisfies condition (A) of Theorem 1.3, then any point in the boundary of $Q$ must be reached with each of the two directions orthogonal to $\mathbf{a}^{1}$ and $\mathbf{a}^{2}$ respectively. If $% Q\subset \mathbb{R}^d, \mathbf{a}^{1}, \mathbf{a}^{2}\in \mathbb{R}^d\backslash\{% \mathbf{0}\}$, $d>2$, there are many directions orthogonal to $\mathbf{a}^{1}$ and $\mathbf{a}^{2} $. In this case, condition (A) requires that any point in the boundary of $Q$ should be reached with at least two directions orthogonal to $\mathbf{a}^{1}$ and $\mathbf{a}^{2}$, respectively. \begin{proof} \textit{Necessity.} Let $g_{0}(\mathbf{x}) =g_{1,0} \left( \mathbf{a}^{1}{\cdot }\mathbf{x}\right) +g_{2,0} \left(\mathbf{a}^{2}{\cdot }\mathbf{x}\right)$ be an extremal element from $\mathcal{R}\left( \mathbf{a}^{1},\mathbf{a}^{2} \right)$ to $f$. We must show that if there is not a closed path extremal for $f_{1} $, then there exists a path extremal for $f_{1} $ with the infinite length (number of points). Suppose the contrary. Suppose that there exists a positive integer $N$ such that the length of each path extremal for $f_{1} $ is not more than $N$. Set the following functions: \begin{equation*} f_{n} =f_{n-1} -g_{1,n-1} -g_{2,n-1} ,\ \ n=2, 3, ..., \end{equation*} where \begin{equation*} g_{1,n-1} =g_{1,n-1} \left(\mathbf{a}^{1}{\cdot }\mathbf{x}\right)=\frac{1}{2} \left(\max\limits_{\substack{ \mathbf{y}\in Q \\ \mathbf{a}^{1}{\cdot }\mathbf{y% }=\mathbf{a}^{1}{\cdot }\mathbf{x}}} f_{n-1}(\mathbf{y}) +\min\limits_{\substack{ \mathbf{y}\in Q \\ \mathbf{a}^{1}{\cdot }\mathbf{y}=\mathbf{a}^{1}{\cdot }\mathbf{x} }} f_{n-1}(\mathbf{y})\right) \end{equation*} \begin{equation*} g_{2,n-1} =g_{2,n-1} (\mathbf{a}^{2}{\cdot }\mathbf{x})=\frac{1}{2} \left( \max _{\substack{ \mathbf{y}\in Q \\ \mathbf{a}^{2}{\cdot }\mathbf{y}=\mathbf{a}^{2}{% \cdot }\mathbf{x}}} \left( f_{n-1} (\mathbf{y})-g_{1,n-1}(\mathbf{a}^{1}{\cdot }% \mathbf{y})\right)\right. \end{equation*} \begin{equation*} \left.+\min\limits_{\substack{ \mathbf{y}\in Q \\ \mathbf{a}^{2}{\cdot }\mathbf{% y}=\mathbf{a}^{2}{\cdot }\mathbf{x}}} \left( f_{n-1}(\mathbf{y}) -g_{1,n-1}(% \mathbf{a}^{1}{\cdot }\mathbf{y}) \right) \right). \end{equation*} By Lemma 1.4, all the functions $f_{n}(\mathbf{x}),$ $n=2,3,...,$ are continuous on $Q$. By assumption $g_{0}$ is a best approximation to $f$. Hence $\left\Vert f_{1}\right\Vert =E\left( f\right) $. Now let us show that $% \left\Vert f_{2}\right\Vert =E\left( f\right) $. Indeed, for any $\mathbf{x}% \in Q$ \begin{equation*} f_{1}(\mathbf{x})-g_{1,1}(\mathbf{a}^{1}{\cdot }\mathbf{x})\leq \frac{1}{2}% \left( \max\limits_{\substack{ \mathbf{y}\in Q \\ \mathbf{a}^{1}{\cdot }\mathbf{% y}=\mathbf{a}^{1}{\cdot }\mathbf{x}}}f_{1}(\mathbf{y})-\min\limits_{\substack{ \mathbf{y}\in Q \\ \mathbf{a}^{1}{\cdot }\mathbf{y}=\mathbf{a}^{1}{\cdot }\mathbf{x} }}f_{1}(\mathbf{y})\right) \leq E(f)\eqno(1.17) \end{equation*}% and \begin{equation*} f_{1}(\mathbf{x})-g_{1,1}(\mathbf{a}^{1}{\cdot }\mathbf{x})\geq \frac{1}{2}% \left( \min\limits_{\substack{ \mathbf{y}\in Q \\ \mathbf{a}^{1}{\cdot }\mathbf{% y}=\mathbf{a}^{1}{\cdot }\mathbf{x}}}f_{1}(\mathbf{y})-\max\limits_{\substack{ \mathbf{y}\in Q \\ \mathbf{a}^{1}{\cdot }\mathbf{y}=\mathbf{a}^{1}{\cdot }\mathbf{x} }}f_{1}(\mathbf{y})\right) \geq -E(f).\eqno(1.18) \end{equation*} Using the definition of $g_{2,1}(\mathbf{a}^{2}\cdot \mathbf{x})$, for any $% \mathbf{x}\in Q$ we have \begin{equation*} f_{1}(\mathbf{x})-g_{1,1}(\mathbf{a}^{1}\cdot \mathbf{x})-g_{2,1}(\mathbf{a}^{2}% \cdot \mathbf{x}) \end{equation*} \begin{equation*} \leq \frac{1}{2}\left( \max\limits_{\substack{ \mathbf{y}\in Q \\ \mathbf{a}^{2}% \cdot \mathbf{y}=\mathbf{a}^{2}\cdot \mathbf{x}}}\left( f_{1}(\mathbf{y}% )-g_{1,1}(\mathbf{a}^{1}\cdot \mathbf{y})\right) -\min\limits_{ _{\substack{ \mathbf{y}\in Q \\ \mathbf{a}^{2}\cdot \mathbf{y}=\mathbf{a}^{2}\cdot \mathbf{x}}}% }\left( f_{1}(\mathbf{y})-g_{1,1}(\mathbf{a}^{1}\cdot \mathbf{y})\right) \right) \end{equation*} and \begin{equation*} f_{1}(\mathbf{x})-g_{1,1}(\mathbf{a}^{1}\cdot \mathbf{x})-g_{2,1}(\mathbf{a}^{2}% \cdot \mathbf{x}) \end{equation*} \begin{equation*} \leq \frac{1}{2}\left( \min\limits_{_{\substack{ \mathbf{y}\in Q \\ \mathbf{% a}^{2}\cdot \mathbf{y}=\mathbf{a}^{2}\cdot \mathbf{x}}}}\left( f_{1}(\mathbf{y}% )-g_{1,1}(\mathbf{a}^{1}\cdot \mathbf{y})\right) -\max\limits_{\substack{ \mathbf{y}\in Q \\ \mathbf{a}^{2}\cdot \mathbf{y}=\mathbf{a}^{2}\cdot \mathbf{x}}}% \left( f_{1}(\mathbf{y})-g_{1,1}(\mathbf{a}^{1}\cdot \mathbf{y})\right) \right). \end{equation*} Using (1.17) and (1.18) in the last two inequalities, we obtain that for any $\mathbf{x}\in Q$ \begin{equation*} -E(f)\leq f_{2}(\mathbf{x})=f_{1}(\mathbf{x})-g_{1,1}(\mathbf{a}^{1}{\cdot }% \mathbf{x})-g_{2,1}(\mathbf{a}^{2}{\cdot }\mathbf{x})\leq E(f). \end{equation*} Therefore, \begin{equation*} \left\Vert f_{2}\right\Vert \leq E(f).\eqno(1.19) \end{equation*} Since $f_{2}(\mathbf{x})-f(\mathbf{x})$ belongs to $\mathcal{R}\left( \mathbf{a}^{1},\mathbf{a}^{2}\right) $, we deduce from (1.19) that \begin{equation*} \left\Vert f_{2}\right\Vert =E(f). \end{equation*} By the same way, one can show that $\|f_3\|=E(f)$, $\|f_4\|=E(f)$, and so on. Thus we can write \begin{equation*} \left\| f_{n} \right\| =E(f), \ \mbox{for any}\ n . \end{equation*} Let us now prove the implications \begin{equation*} f_{1}(\mathbf{p}_{0})<E(f)\Rightarrow f_{2}(\mathbf{p}_{0})<E(f)\eqno(1.20) \end{equation*}% and \begin{equation*} f_{1}(\mathbf{p}_{0})>-E(f)\Rightarrow f_{2}(\mathbf{p}_{0})>-E(f),\eqno% (1.21) \end{equation*}% where $\mathbf{p}_{0}\in Q$. First, we are going to prove the implication \begin{equation*} f_{1}(\mathbf{p}_{0})<E(f)\Rightarrow f_{1}(\mathbf{p}_{0})-g_{1,1}(\mathbf{a% }\cdot \mathbf{p}_{0})<E(f).\eqno(1.22) \end{equation*} There are two possible cases. 1) $\max\limits_{\substack{ \mathbf{y}\in Q \\ \mathbf{a}^{1}{\cdot }\mathbf{y}=% \mathbf{a}^{1}{\cdot }\mathbf{p}_0}} f_{1} (\mathbf{y} )=E(f)$ and $\min\limits _{\substack{ \mathbf{y}\in Q \\ \mathbf{a}^{1}{\cdot }\mathbf{y}=\mathbf{a}^{1}{% \cdot }\mathbf{p}_0}} f_{1} (\mathbf{y}) = -E(f). $ In this case, $g_{1,1}(% \mathbf{a}^{1}\cdot \mathbf{p}_0)=0$. Hence \begin{equation*} f_1(\mathbf{p}_0)-g_{1,1}(\mathbf{a}^{1}\cdot \mathbf{p}_0)< E(f). \end{equation*} 2) $\max\limits_{\substack{ \mathbf{y}\in Q \\ \mathbf{a}^{1}{\cdot }\mathbf{y}=% \mathbf{a}^{1}{\cdot }\mathbf{p}_{0}}}f_{1}(\mathbf{y})=E(f)-\varepsilon _{1}$ and $\min\limits_{\substack{ \mathbf{y}\in Q \\ \mathbf{a}^{1}{\cdot }\mathbf{y}% =\mathbf{a}^{1}{\cdot }\mathbf{p}_{0}}}f_{1}(\mathbf{y})=-E(f)+\varepsilon _{2}$,% \newline where $\varepsilon _{1}$, $\varepsilon _{2}$ are nonnegative real numbers with the sum $\varepsilon _{1}+\varepsilon _{2}\not=0$. In this case, \begin{eqnarray*} f_{1}(\mathbf{p}_{0})-g_{1,1}(\mathbf{a}^{1}{\cdot }\mathbf{p}_{0}) &\leq &\max\limits_{\substack{ \mathbf{y}\in Q \\ \mathbf{a}^{1}{\cdot }\mathbf{y}=% \mathbf{a}^{1}{\cdot }\mathbf{p}_{0}}}f_{1}(\mathbf{y})-g_{1,1}(\mathbf{a}^{1}{\cdot }\mathbf{p}_{0})= \\ &=&\frac{1}{2}\left( \max\limits_{\substack{ \mathbf{y}\in Q \\ \mathbf{a}^{1}{% \cdot }\mathbf{y}=\mathbf{a}^{1}{\cdot }\mathbf{p}_{0}}}f_{1}(\mathbf{y}% )-\min\limits_{\substack{ \mathbf{y}\in Q \\ \mathbf{a}^{1}{\cdot }\mathbf{y}=% \mathbf{a}^{1}{\cdot }\mathbf{p}_{0}}}f_{1}(\mathbf{y})\right) = \end{eqnarray*}% \begin{equation*} =E(f)-\frac{\varepsilon _{1}+\varepsilon _{2}}{2}<E(f). \end{equation*} Thus we have proved (1.22). Using this method, we can also prove that \begin{equation*} f_{1}(\mathbf{p}_{0})-g_{1,1}(\mathbf{a}^{1}{\cdot }\mathbf{p}% _{0})<E(f)\Rightarrow f_{1}(\mathbf{p}_{0})-g_{1,1}(\mathbf{a}^{1}{\cdot }% \mathbf{p}_{0})-g_{2,1}(\mathbf{a}^{2}{\cdot }\mathbf{p}_{0})<E(f).\eqno(1.23) \end{equation*}% Now (1.20) follows from (1.22) and (1.23). By the same way we can prove (1.21). It follows from implications (1.20) and (1.21) that if $f_{2}(% \mathbf{p}_{0})=E(f)$, then $f_{1}(\mathbf{p}_{0})=E(f)$ and if $f_{2}(% \mathbf{p}_{0})=-E(f)$, then $f_{1}(\mathbf{p}_{0})=-E(f)$. This simply means that each path extremal for $f_{2}$ will be extremal for $f_{1}$. Now we show that if any path extremal for $f_{1} $ has the length not more than $N$, then any path extremal for $f_{2} $ has the length not more than $% N-1$. Suppose the contrary. Suppose that there is a path extremal for $f_{2} $ with the length equal to $N$. Denote it by $q=(\mathbf{q}_{1} ,\mathbf{q}% _{2} ,...,\mathbf{q}_{N} )$. Without loss of generality we may assume that $% \mathbf{a}^{2}\cdot \mathbf{q}_{N-1} =\mathbf{a}^{2}\cdot \mathbf{q}_{N}$. As it has been shown above, the path $q$ is also extremal for $f_{1}$. Assume that $% f_{1} (\mathbf{q}_{N} )=E(f)$. Then there is not any $\mathbf{q}_{0} \in Q$ such that $\mathbf{q}_{0} \neq \mathbf{q}_{N} $, $\mathbf{a}^{1}\cdot \mathbf{q}% _{0} =\mathbf{a}^{1}\cdot \mathbf{q}_{N}$ and $f_{1} (\mathbf{q}_{0} )=-E(f)$. Indeed, if there was such $\mathbf{q}_{0}$ and $\mathbf{q}_{0} \not\in q$, the path $(\mathbf{q}_{1} ,\mathbf{q}_{2} ,...,\mathbf{q}_{N} ,\mathbf{q}% _{0} )$ would be extremal for $f_{1} $. But this would contradict our assumption that any path extremal for $f_{1} $ has the length not more than $% N$. Besides, if there was such $\mathbf{q}_{0} $ and $\mathbf{q}_{0} \in q$, we could form some closed path extremal for $f_{1} $. This also would contradict our assumption that there does not exist a closed path extremal for $f_{1} $. Hence \begin{equation*} \max\limits_{\substack{ \mathbf{y}\in Q \\ \mathbf{a}^{1}{\cdot }\mathbf{y}=% \mathbf{a}^{1}{\cdot }\mathbf{q}_N}} f_{1} (\mathbf{y} )=E(f),\ \ \min\limits _{\substack{ \mathbf{y}\in Q \\ \mathbf{a}^{1}{\cdot }\mathbf{y}=\mathbf{a}^{1}{% \cdot }\mathbf{q}_N}} f_{1} (\mathbf{y})>-E(f). \end{equation*} Therefore, \begin{equation*} \left| f_{1} (\mathbf{q}_{N} )-g_{1,1} (\mathbf{a}^{1}{\cdot }\mathbf{q}% _N)\right| <E(f). \end{equation*} From the last inequality it is easy to obtain that (see the proof of implications (1.20) and (1.21)) \begin{equation*} \left\vert f_{2}(\mathbf{q}_{N})\right\vert <E(f). \end{equation*}% This means, on the contrary to our assumption, that the path $(\mathbf{q}% _{1},\mathbf{q}_{2},...,\mathbf{q}_{N})$ can not be extremal for $f_{2}$. Hence any path extremal for $f_{2}$ has the length not more than $N-1$. By the same way, it can be shown that any path extremal for $f_{3}$ has the length not more than $N-2$, any path extremal for $f_{4}$ has the length not more than $N-3$ and so on. Finally, we will obtain that there is not a path extremal for $f_{N+1}$. Hence there is not a point $\mathbf{p}_{0}\in Q$ such that $|f_{N+1}(\mathbf{p}_{0})|=\Vert f_{N+1}\Vert $. But by Lemma 1.4, all the functions $f_{2}$, $f_{3},...,f_{N+1}$ are continuous on the compact set $Q$; hence the norm $\Vert f_{N+1}\Vert $ must be attained. This contradiction means that there exists a path extremal for $f_{1}$ with the infinite length. \bigskip \textit{Sufficiency.} Let a path $p=(\mathbf{p}_{1},\mathbf{p}_{2},...,% \mathbf{p}_{2n})$ be closed and extremal for $f_{1}$. Then \begin{equation*} \left\vert G_{p}(f)\right\vert =\left\Vert f-g_{0}\right\Vert .\eqno(1.24) \end{equation*} By Lemma 1.3, \begin{equation*} \left\vert G_{p}(f)\right\vert \leq E(f).\eqno(1.25) \end{equation*} It follows from (1.24) and (1.25) that $g_{0}$ is a best approximation. Let now a path $p=(\mathbf{p}_{1},\mathbf{p}_{2},...,\mathbf{p}_{n},...)$ be infinite and extremal for $f_{1}$. Consider the sequence $p_{n}=(\mathbf{p}% _{1},\mathbf{p}_{2},...,\mathbf{p}_{n})$, $n=1,2,...,$ of finite paths. By condition (A) of the theorem, for each $p_{n}$ there exists a closed path $p_{n}^{m_{n}}=(\mathbf{p}_{1},\mathbf{p}_{2},...,% \mathbf{p}_{n},\mathbf{q}_{n+1},...,\mathbf{q}_{n+m_{n}})$, where $m_{n}\leq N_{0}$. Then for any positive integer $n$, \begin{equation*} \left\vert G_{p_{n}^{m_{n}}}(f)\right\vert =\left\vert G_{p_{n}^{m_{n}}}(f-g_{0})\right\vert \leq \frac{n\left\Vert f-g_{0}\right\Vert +m_{n}\left\Vert f-g_{0}\right\Vert }{n+m_{n}}=\left\Vert f-g_{0}\right\Vert \end{equation*}% and \begin{equation*} \left\vert G_{p_{n}^{m_{n}}}(f)\right\vert \geq \frac{n\left\Vert f-g_{0}\right\Vert -m_{n}\left\Vert f-g_{0}\right\Vert }{n+m_{n}}=\frac{% n-m_{n}}{n+m_{n}}\left\Vert f-g_{0}\right\Vert . \end{equation*} It follows from the above two inequalities for $\left\vert G_{p_{n}^{m_{n}}}(f)\right\vert$ that \begin{equation*} \sup_{p_{n}^{m_{n}}}\left\vert G_{p_{n}^{m_{n}}}(f)\right\vert =\left\Vert f-g_{0}\right\Vert. \end{equation*} This together with Lemma 1.3 give that \begin{equation*} \Vert f-g_{0}\Vert \leq E(f). \end{equation*}% Hence $g_{0}$ is a best approximation. \end{proof} \bigskip Theorem 1.3 has been proved by using only methods of classical analysis. By implementing more deep techniques from functional analysis we will see below that condition (A) and the convexity assumption on a compact set $Q$ can be dropped. \bigskip \textbf{Theorem 1.4.}\ \textit{\ Assume $Q$ is a compact subset of $\mathbb{R% }^{d}$. A function $g_{0}\in \mathcal{R}$ is a best approximation to a function $f\in C(Q)$ if and only if there exists a closed or infinite path $% p=(\mathbf{p}_{1},\mathbf{p}_{2},...)$ extremal for the function $f-g_{0}$.} \bigskip \begin{proof} \textit{Sufficiency.} There are two possible cases. The first case happens when there exists a closed path $(\mathbf{p}_{1},...,\mathbf{p}% _{2n}) $ extremal for the function $f-g_{0}.$ Let us check that in this case, $f-g_{0}$ is a best approximation. Indeed, on the one hand, the following equalities are valid \begin{equation*} \left\vert \sum_{i=1}^{2n}(-1)^{i}f(\mathbf{p}_{i})\right\vert =\left\vert \sum_{i=1}^{2n}(-1)^{i}\left[ f-g_{0}\right] (\mathbf{p}_{i})\right\vert =2n\left\Vert f-g_{0}\right\Vert . \end{equation*}% On the other hand, for any function $g\in \mathcal{R}$, we have \begin{equation*} \left\vert \sum_{i=1}^{2n}(-1)^{i}f(\mathbf{p}_{i})\right\vert =\left\vert \sum_{i=1}^{2n}(-1)^{i}\left[ f-g\right] (\mathbf{p}_{i})\right\vert \leq 2n\left\Vert f-g\right\Vert . \end{equation*}% Therefore, $\left\Vert f-g_{0}\right\Vert \leq \left\Vert f-g\right\Vert $ for any $g\in \mathcal{R}$. That is, $g_{0}$ is a best approximation. The second case happens when we do not have closed paths extremal for $% f-g_{0}$, but there exists an infinite path $(\mathbf{p}_{1},\mathbf{p}% _{2},...)$ extremal for $f-g_{0}$. To analyze this case, consider the following linear functional \begin{equation*} L_{q}:C(Q)\rightarrow \mathbb{R}\text{, \ }L_{q}(F)=\frac{1}{n}% \sum_{i=1}^{n}(-1)^{i}F(\mathbf{q}_{i}), \end{equation*}% where $q=\{\mathbf{q}_{1},...,\mathbf{q}_{n}\}$ is a finite path in $Q$. It is easy to see that the norm $\left\Vert L_{q}\right\Vert \leq 1$ and $% \left\Vert L_{q}\right\Vert =1$ if and only if the set of points of $q$ with odd indices $O=\{\mathbf{q}_{i}\in q:$ $i$ \textit{is an odd number}$\}$ do not intersect with the set of points of $q$ with even indices $E=\{\mathbf{q}% _{i}\in q:$ $i$ \textit{is an even number}$\}$. Indeed, from the definition of $L_{q}$ it follows that $\left\vert L_{q}(F)\right\vert \leq \left\Vert F\right\Vert $ for all functions $F\in C(Q)$, whence $\left\Vert L_{q}\right\Vert \leq 1.$ If $O\cap E=\varnothing $, then for a function $% F_{0}$ with the property $F_{0}(\mathbf{q}_{i})=-1$ if $i$ is odd, $F_{0}(% \mathbf{q}_{i})=1$ if $i$ is even and $-1<F_{0}(x)<1$ elsewhere on $Q,$ we have $\left\vert L_{q}(F_{0})\right\vert =\left\Vert F_{0}\right\Vert .$ Hence, $\left\Vert L_{q}\right\Vert =1$. Recall that such a function $F_{0}$ exists on the basis of Urysohn's great lemma. Note that if $q$ is a closed path, then $L_{q}$ annihilates all members of the class $\mathcal{R}$. But in general, when $q$ is not closed, we do not have the equality $L_{q}(g)=0,$ for all members $g\in \mathcal{R}$. Nonetheless, this functional has the important property that \begin{equation*} \left\vert L_{q}(g_{1}+g_{2})\right\vert \leq \frac{2}{n}(\left\Vert g_{1}\right\Vert +\left\Vert g_{2}\right\Vert ),\eqno(1.26) \end{equation*}% where $g_{1}$ and $g_{2}$ are ridge functions with the directions $\mathbf{a}% _{1}$ and $\mathbf{a}_{2}$, respectively, that is, $g_{1}=g_{1}(\mathbf{a}% _{1}\cdot \mathbf{x})$ and $g_{2}=g_{2}(\mathbf{a}_{2}\cdot \mathbf{x}).$ This property is important in the sense that if $n$ is sufficiently large, then the functional $L_{q}$ is close to an annihilating functional. To prove (1.26), note that $\left\vert L_{q}(g_{1})\right\vert \leq \frac{2}{n}% \left\Vert g_{1}\right\Vert $ and $\left\vert L_{q}(g_{2})\right\vert \leq \frac{2}{n}\left\Vert g_{2}\right\Vert $. These estimates become obvious if consider the chain of equalities $g_{1}(\mathbf{a}_{1}\cdot \mathbf{q}% _{1})=g_{1}(\mathbf{a}_{1}\cdot \mathbf{q}_{2}),$ $g_{1}(\mathbf{a}_{1}\cdot \mathbf{q}_{3})=g_{1}(\mathbf{a}_{1}\cdot \mathbf{q}_{4}),...$(or $g_{1}(% \mathbf{a}_{1}\cdot \mathbf{q}_{2})=g_{1}(\mathbf{a}_{1}\cdot \mathbf{q}% _{3}),$ $g_{1}(\mathbf{a}_{1}\cdot \mathbf{q}_{4})=g_{1}(\mathbf{a}_{1}\cdot \mathbf{q}_{5}),...$) for $g_{1}(\mathbf{a}_{1}\cdot \mathbf{x})$ and the corresponding chain of equalities for $g_{2}(\mathbf{a}_{2}\cdot \mathbf{x})$% . Now consider the infinite path $p=(\mathbf{p}_{1},\mathbf{p}_{2},...)$ and form the finite paths $p_{k}=(\mathbf{p}_{1},...,\mathbf{p}_{k}),$ $% k=1,2,... $. For ease of notation, let us set $L_{k}=L_{p_{k}}.$ The sequence $\{L_{_{k}}\}_{k=1}^{\infty }$ is a subset of the unit ball of the conjugate space $C^{\ast }(Q).$ By the Banach-Alaoglu theorem, the unit ball is weak$^{\text{*}}$ compact in the weak$^{\text{*}}$ topology of $C^{\ast }(Q)$ (see \cite[p.68]{122}). It follows from this theorem that the sequence $\{L_{_{k}}\}_{k=1}^{\infty }$ must have weak$^{\text{*}}$ cluster points. Suppose $L^{\ast }$ denotes one of them. Without loss of generality we may assume that $L_{k}\overset{weak^{\ast }}{\longrightarrow }% L^{\ast },$ as $k\rightarrow \infty .$ From (1.26) it follows that $L^{\ast }(g_{1}+g_{2})=0.$ That is, $L^{\ast }\in \mathcal{R}^{\bot },$ where the symbol $\mathcal{R}^{\bot }$ stands for the annihilator of $\mathcal{R}$. Since in addition $\left\Vert L^{\ast }\right\Vert \leq 1,$ we can write that \begin{equation*} \left\vert L^{\ast }(f)\right\vert =\left\vert L^{\ast }(f-g)\right\vert \leq \left\Vert f-g\right\Vert ,\eqno(1.27) \end{equation*}% for all functions $g\in \mathcal{R}.$ On the other hand, since the infinite bolt $p$ is extremal for $f-g_{0}$ \begin{equation*} \left\vert L_{k}(f-g_{0})\right\vert =\left\Vert f-g_{0}\right\Vert ,\text{ }% k=1,2,... \end{equation*}% Therefore, \begin{equation*} \left\vert L^{\ast }(f)\right\vert =\left\vert L^{\ast }(f-g_{0})\right\vert =\left\Vert f-g_{0}\right\Vert .\eqno(1.28) \end{equation*}% From (1.27) and (1.28) we conclude that \begin{equation*} \left\Vert f-g_{0}\right\Vert \leq \left\Vert f-g\right\Vert , \end{equation*}% for all $g\in \mathcal{R}.$ In other words, $g_{0}$ is a best approximation to $f$. We proved the sufficiency of the theorem. \textit{Necessity.} The proof of this part is mainly based on the following result of Singer \cite{S}: Let $X$ be a compact space, $U$ be a linear subspace of $C(X)$, $f\in C(X)\backslash U$ and $u_{0}\in U.$ Then $u_{0}$ is a best approximation to $f$ if and only if there exists a regular Borel measure $\mu $ on $X$ such that (1) The total variation $\left\Vert \mu \right\Vert =1$; (2) $\mu $ is orthogonal to the subspace $U$, that is, $\int_{X}ud\mu =0$ for all $u\in U$; (3) For the Jordan decomposition $\mu =\mu ^{+}-\mu ^{-}$, \begin{equation*} f(x)-u_{0}(x)=\left\{ \begin{array}{c} \left\Vert f-u_{0}\right\Vert \text{ for }x\in S^{+}, \\ -\left\Vert f-u_{0}\right\Vert \text{ for }x\in S^{-},% \end{array}% \right. \end{equation*}% where $S^{+}$ and $S^{-}$ are closed supports of the positive measures $\mu ^{+}$ and $\mu ^{-}$, respectively. Let us show how we use this theorem in the proof of necessity part of our theorem. Assume $g_{0}\in \mathcal{R}$ is a best approximation. For the subspace $\mathcal{R},$ the existence of a measure $\mu $ satisfying the conditions (1)-(3) is a direct consequence of Singer's result. Let $\mathbf{x}% _{0}$ be any point in $S^{+}.$\ Consider the point $y_{0}=\mathbf{a}% _{1}\cdot \mathbf{x}_{0}$ and a $\delta $-neighborhood of $y_{0}$. That is, choose an arbitrary $\delta >0$ and consider the set $I_{\delta }=(y_{0}-\delta ,y_{0}+\delta )\cap \mathbf{a}_{1}\cdot Q.$ Here, $\mathbf{a}% _{1}\cdot Q=\{\mathbf{a}_{1}\cdot \mathbf{x}:$ $\mathbf{x}\in Q\}.$ For any subset $E\subset \mathbb{R}$, put \begin{equation*} E^{i}=\{\mathbf{x}\in Q:\mathbf{a}_{i}\cdot \mathbf{x}\in E\},\text{ }i=1,2.% \text{ } \end{equation*} Clearly, for some sets $E,$ one or both the sets $E^{i}$ may be empty. Since $I_{\delta }^{1}\cap S^{+}$ is not empty (note that $\mathbf{x}_{0}\in I_{\delta }^{1}$), it follows that $\mu ^{+}(I_{\delta }^{1})>0.$ At the same time $\mu (I_{\delta }^{1})=0,$ since $\mu $ is orthogonal to all functions $g_{1}(\mathbf{a}_{1}\cdot \mathbf{x}).$ Therefore, $\mu ^{-}(I_{\delta }^{1})>0.$ We conclude that $I_{\delta }^{1}\cap S^{-}$ is not empty. Denote this intersection by $A_{\delta }.$ Tending $\delta $ to $% 0,$ we obtain a set $A$ which is a subset of $S^{-}$ and has the property that for each $\mathbf{x}\in A,$ we have $\mathbf{a}_{1}\cdot \mathbf{x}=% \mathbf{a}_{1}\cdot \mathbf{x}_{0}.$ Fix any point $\mathbf{x}_{1}\in A$. Changing $\mathbf{a}_{1}$, $\mu ^{+}$, $S^{+}$ to $\mathbf{a}_{2}$, $\mu ^{-} $ and $S^{-}$ correspondingly, repeat the above process with the point $% y_{1}=\mathbf{a}_{2}\cdot \mathbf{x}_{1}$ and a $\delta $-neighborhood of $% y_{1}$. Then we obtain a point $\mathbf{x}_{2}\in S^{+}$ such that $\mathbf{a% }_{2}\cdot \mathbf{x}_{2}=\mathbf{a}_{2}\cdot \mathbf{x}_{1}.$ Continuing this process, one can construct points $\mathbf{x}_{3}$, $\mathbf{x}_{4}$, and so on. Note that \ the set of all constructed points $\mathbf{x}_{i}$, $% i=0,1,...,$ forms a path. By Singer's above result, this path is extremal for the function $f-g_{0}$. We have proved the necessity and hence Theorem 1.4. \end{proof} Theorem 1.4, in a more general setting, was proven in Pinkus \cite[p.99]{117} under additional assumption that $Q$ is convex. Convexity assumption was made to guarantee continuity of the following functions \begin{equation*} g_{1,i}(t)=\max_{\substack{ \mathbf{x}\in Q \\ \mathbf{a}_{i}\cdot \mathbf{x% }=t}}F(\mathbf{x})\ \ \text{and }\ g_{2,i}(t)=\min\limits_{\substack{ \mathbf{x}\in Q \\ \mathbf{a}_{i}\cdot \mathbf{x}=t}}F(\mathbf{x}),\text{ }% i=1,2, \end{equation*}% where $F$ is an arbitrary continuous function on $Q$. Note that in the proof of Theorem 1.4 we did not need continuity of these functions. \bigskip It is well known that characterization theorems of this type are very essential in approximation theory. Chebyshev was the first to prove a similar result for polynomial approximation. Khavinson \cite{79} characterized extremal elements in the special case of the problem considered here. His case allows the approximation of a continuous bivariate function $f\left( {x,y}\right) $ by functions of the form $\varphi \left( {x}% \right) +\psi \left( {y}\right)$. \bigskip \subsection{Construction of an extremal element} In 1951, Diliberto and Straus \cite{26} established a formula for the error of approximation of a bivariate function by sums of univariate functions. Their formula contains the supremum over all closed bolts (see Section 3.3.1). Although the mentioned formula is valid for all continuous functions, it is not easily calculable. Therefore, it cannot give a desired effect if one is interested in the precise value of the approximation error. After this general result some authors started to seek easily calculable formulas for the approximation error by considering not the whole space of continuous functions, but some subsets thereof (see, for example, \cite{4,7,55,56,79,121}). These subsets were chosen so that they could provide precise and easy computation of the approximation error. Since the set of ridge functions contains univariate functions, one may ask for explicit formulas for the error of approximation of a multivariate function by sums of ridge functions. In this section, we see how with the use of Theorem 1.3 (or 1.4) it is possible to find the approximation error and construct an extremal element in the problem of approximation by sums of ridge functions. We restrict ourselves to $% \mathbb{R}^{2}.$ To make the problem more precise, let $\Omega $ be a compact set in $\mathbb{R}^{2},$ $f \in C\left( {% \Omega }\right)$, $\mathbf{a}=\left( {a_{1},a_{2}}\right)$ and $\mathbf{b}% =\left( {b_{1},b_{2}}\right)$ be linearly independent vectors. Consider the approximation of $f$ by functions from $\mathcal{R}=\mathcal{R}\left( \mathbf{a},\mathbf{b}\right)$. We want, under some suitable conditions on $f\;$and $\Omega $, to establish a formula for an easy and direct computation of the approximation error $E\left(f,\mathcal{R}\right)$. \bigskip \textbf{Theorem 1.5.} \textit{Let \begin{equation*} \Omega =\left\{ \mathbf{x}\in \mathbb{R}^{2}:c_{1}\leq \mathbf{a}\cdot \mathbf{x}\leq d_{1},\ \ c_{2}\leq \mathbf{b}\cdot \mathbf{x}\leq d_{2}\right\} , \end{equation*}% where $c_{1}<d_{1}$ and $c_{2}<d_{2}$. Let a function $f(\mathbf{x})\in C(\Omega )$ have the continuous partial derivatives $\frac{\partial ^{2}f}{% \partial x_{1}^{2}},\frac{\partial ^{2}f}{\partial x_{1}\partial x_{2}},% \frac{\partial ^{2}f}{\partial x_{2}^{2}}$ and for any $\mathbf{x}\in \Omega $ \begin{equation*} \frac{\partial ^{2}f}{\partial x_{1}\partial x_{2}}\left( a_{1}b_{2}+a_{2}b_{1}\right) -\frac{\partial ^{2}f}{\partial x_{1}^{2}}% a_{2}b_{2}-\frac{\partial ^{2}f}{\partial x_{2}^{2}}a_{1}b_{1}\geq 0. \end{equation*}% Then \begin{equation*} E\left(f,\mathcal{R}\right)=\frac{1}{4}\left( f_{1}(c_{1},c_{2})+f_{1}(d_{1},d_{2})-f_{1}(c_{1},d_{2})-f_{1}(d_{1},c_{2})% \right) , \end{equation*}% where \begin{equation*} f_{1}(y_{1},y_{2})=f\left( \frac{y_{1}b_{2}-y_{2}a_{2}}{a_{1}b_{2}-a_{2}b_{1}% },\frac{y_{2}a_{1}-y_{1}b_{1}}{a_{1}b_{2}-a_{2}b_{1}}\right) .\eqno(1.29) \end{equation*}% } \bigskip \begin{proof} Introduce the new variables \begin{equation*} y_{1}=a_{1}x_{1}+a_{2}x_{2},\ \ y_{2}=b_{1}x_{1}+b_{2}x_{2}.\eqno(1.30) \end{equation*} Since the vectors $(a_{1},a_{2})$ and $(b_{1},b_{2})$ are linearly independent, for any $(y_{1},y_{2})\in Y$, where $Y=[c_{1},d_{1}]\times \lbrack c_{2},d_{2}]$, there exists only one solution $(x_{1},x_{2})\in \Omega $ of the system (1.30). The coordinates of this solution are \begin{equation*} x_{1}=\frac{y_{1}b_{2}-y_{2}a_{2}}{a_{1}b_{2}-a_{2}b_{1}},\qquad \ x_{2}=% \frac{y_{2}a_{1}-y_{1}b_{1}}{a_{1}b_{2}-a_{2}b_{1}}.\eqno(1.31) \end{equation*} The linear transformation (1.31) transforms the function $f(x_{1},x_{2})$ to the function $f_{1}(y_{1},y_{2})$. Consider the approximation of $% f_{1}(y_{1},y_{2})$ from the set \begin{equation*} \mathcal{Z}=\left\{ z_{1}(y_{1})+z_{2}(y_{2}):z_{i}\in C(\mathbb{R}),\ i=1,2\right\} . \end{equation*} It is easy to see that \begin{equation*} E\left( f,\mathcal{R}\right) =E\left( f_{1},\mathcal{Z}\right) .\eqno(1.32) \end{equation*} With each rectangle $S=\lbrack u_{1} ,v_{1} ]\times \lbrack u_{2} ,v_{2} ]\subset Y$ we associate the functional \begin{equation*} L\left(h, S\right) =\frac{1}{4} \left(h(u_{1} ,u_{2} )+h(v_{1} ,v_{2} )-h(u_{1} ,v_{2} )-h (v_{1} ,u_{2})\right),\ \ h\in C(Y). \end{equation*} This functional has the following obvious properties: (i) $L(z,S)=0$ for any $z\in \mathcal{Z}$ and $S\subset Y$. (ii) For any point $(y_{1} ,y_{2} )\in Y$, $L(f_1, Y)=\sum\limits_{i=1}^{4}L(f_1, S_{i} ) $, where $S_{1} =[c_{1} ,y_{1}]\times [c_{2} ,y_{2}],$ $S_{2} =[y_{1} ,d_{1}]\times [y_{2} ,d_{2}],$ $S_{3} =[c_{1} ,y_{1}]\times [y_{2} ,d_{2}],$ $S_{4} =[y_{1} ,d_{1}]\times [c_{2} ,y_{2}]$. By the conditions of the theorem, it is not difficult to verify that \begin{equation*} \frac{\partial ^{2} f_1}{\partial y_{1} \partial y_{2} } \geq 0\ \ \mbox{for any}\ \ (y_{1} ,y_{2} )\in Y. \end{equation*} Integrating both sides of the last inequality over arbitrary rectangle $% S=[u_{1},v_{1}]\times \lbrack u_{2},v_{2}]\subset Y$, we obtain that \begin{equation*} L\left( f_{1},S\right) \geq 0.\eqno(1.33) \end{equation*}% Set the function \begin{equation*} f_{2}(y_{1},y_{2})=L\left( f_{1},S_{1}\right) +L\left( f_{1},S_{2}\right) -L\left( f_{1},S_{3}\right) -L\left( f_{1},S_{4}\right) .\eqno(1.34) \end{equation*} It is not difficult to verify that the function $f_{1}-f_{2}$ belongs to $% \mathcal{Z}$. Hence \begin{equation*} E\left( f_{1},\mathcal{Z}\right) =E\left( f_{2},\mathcal{Z}\right) .\eqno% (1.35) \end{equation*} Calculate the norm $\left\Vert f_{2}\right\Vert $. From the property (ii), it follows that \begin{equation*} f_{2}(y_{1},y_{2})=L(f_{1},Y)-2(L(f_{1},S_{3})+L(f_{1},S_{4})) \end{equation*}% and \begin{equation*} f_{2}(y_{1},y_{2})=2\left( L\left( f_{1},S_{1}\right) +L\left( f_{1},S_{2}\right) \right) -L\left( f_{1},Y\right) . \end{equation*}% From the last equalities and (1.33), we obtain that \begin{equation*} \left\vert f_{2}(y_{1},y_{2})\right\vert \leq L\left( f_{1},Y\right) ,\ % \mbox{for any}\ (y_{1},y_{2})\in Y. \end{equation*}% On the other hand, one can check that \begin{equation*} f_{2}(c_{1},c_{2})=f_{2}(d_{1},d_{2})=L\left( f_{1},Y\right) \eqno(1.36) \end{equation*}% and \begin{equation*} f_{2}(c_{1},d_{2})=f_{2}(d_{1},c_{2})=-L\left( f_{1},Y\right) .\eqno(1.37) \end{equation*}% Therefore, \begin{equation*} \left\Vert f_{2}\right\Vert =L\left( f_{1},Y\right) .\eqno(1.38) \end{equation*}% Note that the points $% (c_{1},c_{2}),(c_{1},d_{2}),(d_{1},d_{2}),(d_{1},c_{2}) $ in the given order form a closed path with respect to the directions $(0;1) $ and $(1;0)$. We conclude from (1.36)-(1.38) that this path is extremal for $f_{2}$. By Theorem 1.3, $z_{0}=0$ is a best approximation to $f_{2}$. Hence \begin{equation*} E\left( f_{2},\mathcal{Z}\right) =L\left( f_{1},Y\right) .\eqno(1.39) \end{equation*} Now from (1.32),(1.35) and (1.39) we finally conclude that \begin{equation*} E\left( f,\mathcal{R}\right) =L\left( f_{1},Y\right) =\frac{1}{4}\left( f_{1}(c_{1},c_{2})+f_{1}(d_{1},d_{2})-f_{1}(c_{1},d_{2})-f_{1}(d_{1},c_{2})% \right) , \end{equation*}% which is the desired result. \end{proof} \textbf{Corollary 1.4.}\ \textit{Let all the conditions of Theorem 1.5 hold and $f_{1}(y_{1},y_{2})$ is the function defined in (1.29). Then the function $g_{0}(y_{1},y_{2})=g_{1,0}(y_{1})+g_{2,0}(y_{2})$, where \begin{equation*} g_{1,0}(y_{1})=\frac{1}{2}f_{1}(y_{1},c_{2})+\frac{1}{2}f_{1}(y_{1},d_{2})-% \frac{1}{4}f_{1}(c_{1},c_{2})-\frac{1}{4}f_{1}(d_{1},d_{2}), \end{equation*}% \begin{equation*} g_{2,0}(y_{2})=\frac{1}{2}f_{1}(c_{1},y_{2})+\frac{1}{2}f_{1}(d_{1},y_{2})-% \frac{1}{4}f_{1}(c_{1},d_{2})-\frac{1}{4}f_{1}(d_{1},c_{2}) \end{equation*}% and $y_{1}=a_{1}x_{1}+a_{2}x_{2}$, $y_{2}=b_{1}x_{1}+b_{2}x_{2}$, is a best approximation from the set $\mathcal{R}(a,b)$ to the function $f$.} \bigskip \begin{proof} It is not difficult to verify that the function $% f_{2}(y_{1},y_{2})$ defined in (1.34) has the form \begin{equation*} f_{2}(y_{1},y_{2})=f_{1}(y_{1},y_{2})-g_{1,0}(y_{1})-g_{2,0}(y_{2}). \end{equation*} On the other hand, we know from the proof of Theorem 1.5 that \begin{equation*} E(f_{1},\mathcal{Z})=\left\Vert f_{2}\right\Vert . \end{equation*}% Therefore, the function $g_{1,0}(y_{1})+g_{2,0}(y_{2})$ is a best approximation to $f_{1}$. Then the function $g_{1,0}(\mathbf{a}\cdot \mathbf{% x})+g_{2,0}(\mathbf{b}\cdot \mathbf{x})$ is an extremal element from $% \mathcal{R}(\mathbf{a},\mathbf{b})$ to $f(\mathbf{x})$. \end{proof} \textbf{Remark 1.1.} Rivlin and Sibner \cite{121}, and Babaev \cite{7} proved Theorem 1.5 for the case in which $\mathbf{a}${\ and\ }$\mathbf{b}$ are the coordinate vectors. Our proof of Theorem 1.5 is different, short and elementary. Moreover, it has turned out to be useful in constructing an extremal element (see the proof of Corollary 1.4). \bigskip \subsection{Density of ridge functions and some problems} Let $\mathbf{a}^{1}${\ and }$\mathbf{a}^{2}$ be nonzero directions in $% \mathbb{R}^{d}$. One may ask the following question: are there cases in which the set $\mathcal{R}\left( \mathbf{a}^{1},\mathbf{a}^{2}\right) $ is dense in the space of all continuous functions? Undoubtedly, a positive answer depends on the geometrical structure of compact sets over which all the considered functions are defined. This problem may be interesting in the theory of partial differential equations. Take, for example, equation (1.10). A positive answer to the problem means that for any continuous function $f$ there exist solutions of the given equation uniformly converging to $f$. It should be remarked that our problem is a special case of the problem considered by Marshall and O'Farrell. In \cite{107}, they obtained a necessary and sufficient condition for a sum $A_{1}+A_{2}$ of two subalgebras to be dense in $C(U)$, where $C(U)$ denotes the space of real-valued continuous functions on a compact Hausdorff space $U$. Below we describe Marshall and O' Farrell's result for sums of ridge functions. Let $X$ be a compact subset of $\mathbb{R}^{d}.$ The relation on $X$, defined by setting $\mathbf{x}\approx \mathbf{y}$ if $\mathbf{x}$ and $% \mathbf{y}$ belong to some path in $X$, is an equivalence relation. The equivalence classes we call orbits. \bigskip \textbf{Theorem 1.6.} \textit{Let $X$ be a compact subset of $\mathbb{R}^{d}$ with all its orbits closed. The set $\mathcal{\ R}\left( \mathbf{a}^{1},% \mathbf{a}^{2}\right) $ is dense in $C(X)$ if and only if $X$ contains no closed path with respect to the directions $\mathbf{a}^{1}${\ and }$\mathbf{a% }^{2}$.} \bigskip The proof immediately follows from proposition 2 in \cite{108} established for the sum of two algebras. Since that proposition was given without proof, for completeness of the exposition we give the proof of Theorem 1.6. \begin{proof} \textit{Necessity}. If $X$ has closed paths, then $X$ has a closed path $% p^{\prime }=\left( \mathbf{p}_{1}^{\prime },...,\mathbf{p}_{2m}^{\prime }\right) $ such that all points $\mathbf{p}_{1}^{\prime },...,\mathbf{p}% _{2m}^{\prime }$ are distinct. In fact, such a special path can be obtained from any closed path $p=\left( \mathbf{p}_{1},...,\mathbf{p}_{2n}\right) $ by the following simple algorithm: if the points of the path $p$ are not all distinct, let $i$ and $k>0$ be the minimal indices such that $\mathbf{p}_{i}=% \mathbf{p}_{i+2k}$; delete from $p$ the subsequence $\mathbf{p}_{i+1},...,% \mathbf{p}_{i+2k}$ and call $p$ the obtained path; repeat the above step until all points of $p$ are all distinct; set $p^{\prime }:=p$. By Urysohn's great lemma, there exist continuous functions $h=h(\mathbf{x})$ on $X$ such that $h(\mathbf{p}_{i}^{\prime })=1$, $i=1,3,...,2m-1$, $h(\mathbf{p}% _{i}^{\prime })=-1$, $i=2,4,...,2m$ and $-1<h(\mathbf{x})<1$ elsewhere. Consider the measure \begin{equation*} \mu _{p^{\prime }}=\frac{1}{2m}\sum_{i=1}^{2m}(-1)^{i-1}\delta _{\mathbf{p}% _{i}^{\prime }}\text{ ,} \end{equation*}% where $\delta _{\mathbf{p}_{i}^{\prime }}$ is a point mass at $\mathbf{p}% _{i}^{\prime }$. For this measure, $\int\limits_{X}hd\mu _{p^{\prime }}=1$ and $\int\limits_{X}gd\mu _{p^{\prime }}=0$ for all functions $g\in \mathcal{% R}\left( \mathbf{a}^{1},\mathbf{a}^{2}\right) $. Thus the set $\mathcal{R}% \left( \mathbf{a}^{1},\mathbf{a}^{2}\right) $ cannot be dense in\textit{\ }% $C(X)$. \textit{Sufficiency}. We are going to prove that the only annihilating regular Borel measure for $\mathcal{R}\left( \mathbf{a}^{1},\mathbf{a}% ^{2}\right) $ is the zero measure. Suppose, contrary to this assumption, there exists a nonzero annihilating measure on $X$ for $\mathcal{R}\left( \mathbf{a}^{1},\mathbf{a}^{2}\right) $. The class of such measures with total variation not more than $1$ we denote by $S.$ Clearly, $S$ is weak-* compact and convex. By the Krein-Milman theorem, there exists an extreme measure $\mu $ in $S.$ Since the orbits are closed, $\mu $ must be supported on a single orbit. Denote this orbit by $T.$ For $i=1,2,$ let $X_{i}$ be the quotient space of $X$ obtained by identifying the points $\mathbf{y}$ and $\mathbf{z}$ whenever $\mathbf{a}^{i}% {\cdot }\mathbf{y}=\mathbf{a}^{i}{\cdot }\mathbf{z}$. Let $\pi _{i}$ be the natural projection of $X$ onto $X_{i}$. For a fixed point $t\in X$ set $% T_{1}=\{t\}$, $T_{2}=\pi _{1}^{-1}\left( \pi _{1}T_{1}\right) $, $T_{3}=\pi _{2}^{-1}\left( \pi _{2}T_{2}\right) $, $T_{4}=\pi _{1}^{-1}\left( \pi _{1}T_{3}\right) $, $...$ Obviously, $T_{1}\subset T_{2}\subset T_{3}\subset \cdot \cdot \cdot $ . Therefore, for some $k\in \mathbb{N}$, $\left\vert \mu \right\vert (T_{2k})>0$, where $\left\vert \mu \right\vert $ is a total variation measure of $\mu$. Since $\mu $ is orthogonal to every continuous function of the form ${g\left( \mathbf{a}^{1}{\cdot }\mathbf{x}\right) }$, $% \mu (T_{2k})=0$. From the Haar decomposition $\mu (T_{2k})=\mu ^{+}(T_{2k})-\mu ^{-}(T_{2k})$ it follows that $\mu ^{+}(T_{2k})=\mu ^{-}(T_{2k})>0$. Fix a Borel subset $S_{0}\subset T_{2k}$ such that $\mu ^{+}(S_{0})>0$ and $\mu ^{-}(S_{0})=0$. Since $\mu $ is orthogonal to every continuous function of the form ${g\left( \mathbf{a}^{2}{\cdot }\mathbf{x}% \right) }$, $\mu (\pi _{2}^{-1}\left( \pi _{2}S_{0}\right) )=0.$ Therefore, one can chose a Borel set $S_{1}$ such that $S_{1}\subset \pi _{2}^{-1}\left( \pi _{2}S_{0}\right) \subset T_{2k+1}$, $S_{1}\cap S_{0}=\varnothing $, $\mu ^{+}(S_{1})=0$, $\mu ^{-}(S_{1})\geqslant \mu ^{+}(S_{0})$. By the same way one can chose a Borel set $S_{2}$ such that $% S_{2}\subset \pi _{1}^{-1}\left( \pi _{1}S_{1}\right) \subset T_{2k+2}$, $% S_{2}\cap S_{1}=\varnothing $, $\mu ^{-}(S_{2})=0$, $\mu ^{+}(S_{2})\geqslant \mu ^{-}(S_{1})$, and so on. The sets $S_{0},S_{1},S_{2},...$are pairwise disjoint. For otherwise, there would exist positive integers $n$ and $m,$ with $n<m$ and a path $% (y_{n},y_{n+1},...,y_{m})$ such that $y_{i}\in S_{i}$ for $i=n,...,m$ and $% y_{m}\in S_{m}\cap S_{n}$. But then there would exist paths $% (z_{1},z_{2},...,z_{n-1},y_{n})$ and $(z_{1},z_{2}^{^{\prime }},...,z_{n-1}^{^{\prime }},y_{m})$ with $z_{i}$ and $z_{i}^{^{\prime }}$ in $T_{i}$ for $i=2,...,n-1.$ Hence, the set \begin{equation*} \{z_{1},z_{2},...,z_{n-1},y_{n},y_{n+1},...,y_{m},z_{n-1}^{^{\prime }},...,z_{2}^{^{\prime }},z_{1}\} \end{equation*}% would contain a closed path. This would contradict our assumption on $X.$ Now, since the sets $S_{0},S_{1},S_{2},...,$ are pairwise disjoint and $% \left\vert \mu \right\vert (S_{i})\geqslant \mu ^{+}(S_{0})>0$ for each $% i=1,2,...,$ it follows that the total variation of $\mu $ is infinite. This contradiction completes the proof. \end{proof} The following corollary concerns the problem considered by Colitschek and Light \cite{36}. \bigskip \textbf{Corollary 1.5.} \textit{Let $D$ be a compact subset of \ $% \mathbb{R}^{2}$ with all its orbits closed. Let $W$ denote the set of all solutions of the wave equation \begin{equation*} \frac{\partial ^{2}w}{\partial s\partial t}(s,t)=0,\;\ \ \ \ (s,t)\in D. \end{equation*}% Then \begin{equation*} \inf\limits_{w\in W}\left\Vert f-w\right\Vert =0 \end{equation*}% for any continuous function $f(s,t)$ on $D$ if and only if $D$ contains no closed bolt of lightning.} \bigskip \begin{proof} Let $\pi _{1}$ and $\pi _{2}$ denote the usual coordinate projections, viz: $\pi _{1}(s,t)=s$ and $\pi _{2}(s,t)=t$, $% (s,t)\in \mathbb{R}^{2}$. Set $S=\pi _{1}(D)$ and $T=\pi _{2}(D)$. It is easy to see that \begin{equation*} W=\left\{ w\in C(D):w(s,t)=x(s)+y(t),\;\ \ x\in C^{2}(S),\;\ y\in C^{2}(T)\right\} . \end{equation*} Set \begin{equation*} \widetilde{W}=\left\{ w\in C(D):w(s,t)=x(s)+y(t),\;\ \ x\in C(S),\;\ y\in C(T)\right\} . \end{equation*} Since the set $W$ is dense in $\widetilde{W},$ \begin{equation*} \inf\limits_{w\in W}\left\Vert f-w\right\Vert =\inf\limits_{w\in \widetilde{W% }}\left\Vert f-w\right\Vert. \end{equation*} But by Theorem 1.6, the equality \begin{equation*} \inf\limits_{w\in \widetilde{W}}\left\Vert f-w\right\Vert =0 \end{equation*}% holds for any $f\in C(D)$ if and only if $D$ contains no closed bolt of lightning. \end{proof} Let us discuss some difficulties that arise when studying sums of more than two ridge functions. Consider the set \begin{equation*} \mathcal{R}\left( \mathbf{a}{^{1},...,}\mathbf{a}{^{r}}\right) ={\left\{ \sum\limits_{i=1}^{r}{g}_{i}{{{\left( \mathbf{a}{^{i}\cdot }\mathbf{x}% \right) ,g}}}_{{i}}{{{\ \in C\left( \mathbb{R}\right) ,i=1,...,r}}}\right\} }% , \end{equation*}% where $\mathbf{a}{^{1},...,}\mathbf{a}{^{r}}$ are pairwise linearly independent vectors in $\mathbb{R}^{d}\backslash \{\mathbf{0}\}$. Let $r\geq 3$. How can one define a path in this general case? Recall that in the case when $r=2$, a path is an ordered set of points $\left( \mathbf{p}_{1},\mathbf{p}_{2},...,\mathbf{p}% _{n}\right) $ in $\mathbb{R}^{d}$ with edges $\mathbf{p}_{i}\mathbf{p}_{i+1}$ in alternating hyperplanes. The first, the third, the fifth,... hyperplanes (also the second, the fourth, the sixth,... hyperplanes) are parallel. If not differentiate between parallel hyperplanes, the path $\left( \mathbf{p}% _{1},\mathbf{p}_{2},...,\mathbf{p}_{n}\right) $ can be considered as a trace of some point traveling in two alternating hyperplanes. In this case, if the point starts and stops at the same location (i.e., if $\mathbf{p}_{n}=% \mathbf{p}_{1})$ and $n$ is an odd number, then the path functional \begin{equation*} G(f)=\frac{1}{n-1}\sum\limits_{i=1}^{n-1}(-1)^{i+1}f(\mathbf{p}_{i}), \end{equation*}% annihilates sums of ridge functions with the corresponding two fixed directions. The picture becomes quite different and more complicated when the number of directions more than two. The simple generalization of the above-mentioned arguments demands a point traveling in three or more alternating hyperplanes. But in this case the appropriate generalization of the functional $G$ does not annihilate functions from $\mathcal{R}\left( \mathbf{a}{^{1},...,}\mathbf{a}{^{r}}% \right) $. There were several attempts to fill this gap in the special case when $r=d$ and $\mathbf{a}{^{1},...,}\mathbf{a}{^{r}}$ are the coordinate vectors. Unfortunately, all these attempts failed (see, for example, the attempts in \cite{26,37} and the refutations in \cite{4,109}). At the end of this subsection we want to draw the readers attention to the following problems. All these problems are open and cannot be solved by the methods presented here. Let $Q$ be a compact subset of $\mathbb{R}^{d}$. Consider the approximation of a continuous function defined on $Q$ by functions from $\mathcal{R}\left( \mathbf{a}{^{1},...,}\mathbf{a}{^{r}}\right)$. Let $r\ge 3$. \textbf{Problem 3.} \textit{Characterize those functions from $\mathcal{R}% \left( \mathbf{a}{^{1},...,}\mathbf{a}{^{r}}\right) $ that are extremal to a given continuous function.} \textbf{Problem 4.} \textit{Establish explicit formulas for the error in approximating from $\mathcal{R}\left( \mathbf{a}{^{1},...,}\mathbf{a}{^{r}}% \right) $ \ and construct a best approximation.} \textbf{Problem 5.} \textit{Find necessary and sufficient geometrical conditions for the set $\mathcal{R}\left( \mathbf{a}{^{1},...,}\mathbf{a}{% ^{r}}\right) $ to be dense in $C(Q)$.} It should be remarked that in \cite{108}, Problem 5 was set up for the sum of $r $ subalgebras of $C(Q)$. Lin and Pinkus \cite{95} proved that the set $% \mathcal{R}\left( \mathbf{a}{^{1},...,}\mathbf{a}{^{r}}\right) $ ($r$ may be very large) is not dense in $C(\mathbb{R}^{d})$ in the topology of uniform convergence on compact subsets of $\mathbb{R}^{d}$. That is, there are compact sets $Q\subset \mathbb{R}^{d}$ such that $\mathcal{R}\left( \mathbf{a% }{^{1},...,}\mathbf{a}{^{r}}\right) $ is not dense in $C(Q)$. In the case $% r=2$, Theorem 1.6 complements this result, by describing compact sets $% Q\subset \mathbb{R}^{2}$, for which $\mathcal{R}\left( \mathbf{a}{^{1},}% \mathbf{a}{^{2}}\right) $ is dense in $C(Q)$. \bigskip \section{Sums of continuous ridge functions} In this section, we find geometric means of deciding if any continuous multivariate function can be represented by a sum of two continuous ridge functions. \subsection{Exposition of the problem} In this section, we will consider the following representation problem associated with the set $\mathcal{R}\left( \mathbf{a}^{1},...,\mathbf{a}% ^{r}\right) .$ \bigskip \textbf{Problem 6.}\ \textit{Let $X$ be a compact subset of \ $\mathbb{R}% ^{d}.$ Give geometrical conditions that are necessary and sufficient for \begin{equation*} \mathcal{R}\left( \mathbf{a}^{1},...,\mathbf{a}^{r}\right)=C\left( X\right) , \end{equation*}% where $C\left( X\right) $ is the space of continuous functions on $X$ furnished with the uniform norm.} \bigskip We solve this problem for $r=2$. Problem 6, like Problems 3--5 from the previous section, is open in the case $r\geq 3$. Geometrical characterization of compact sets $X \subset \mathbb{R}^{d}$ with the property $\mathcal{R}\left(\mathbf{a}^{1},...,\mathbf{a}^{r}\right)=C\left(X\right)$, $r\geq 3$, seems to be beyond the scope of the methods discussed herein. Nevertheless, recall that this problem in a quite abstract form, which involves regular Borel measures on $X$, was solved by Sternfeld (see Section 1.1.1). In the sequel, we will use the notation \begin{equation*} H_{1}=H_{1}\left( X\right) =\left\{ g_{1}\left( \mathbf{a}^{1}\cdot \mathbf{x% }\right) :g_{1}\in C\left( \mathbb{R}\right) \right\} , \end{equation*} \begin{equation*} H_{2}=H_{2}\left( X\right) =\left\{ g_{2}\left( \mathbf{a}^{2}\cdot \mathbf{x% }\right) :g_{2}\in C\left( \mathbb{R}\right) \right\} . \end{equation*} Note that by this notation, $\mathcal{R}\left( \mathbf{a}^{1},\mathbf{a}% ^{2}\right) =H_{1}+H_{2}.$ At the end of this section, we generalize the obtained result from $% H_{1}+H_{2}$ to the set of sums $g_{1}\left( h_{1}\left( \mathbf{x}\right) \right) +g_{2}\left( h_{2}\left( \mathbf{x}\right) \right)$, where $h_{1}, h_{2}$ are fixed continuous functions on $X$. \bigskip \subsection{The representation theorem} \textbf{Theorem 1.7.} \textit{Let $X$ be a compact subset of $\mathbb{R}% ^{d} $. The equality \begin{equation*} H_{1}\left( X\right) +H_{2}\left( X\right) =C\left( X\right) \end{equation*}% holds if and only if $X$ contains no closed path and there exists a positive integer $n_{0}$ such that the lengths of paths in $X$ are bounded by $n_{0}$.% } \bigskip \begin{proof} \textit{Necessity.} Let $H_{1}+H_{2}=C\left( X\right) $. Consider the linear operator \begin{equation*} A:H_{1}\times H_{2}\rightarrow C\left( X\right), ~~~ A\left[ \left( g_{1},g_{2}\right) \right] =g_{1}+g_{2}, \end{equation*}% where $g_{1}\in H_{1},g_{2}\in H_{2}.$ The norm on $H_{1}\times H_{2}$ we define as \begin{equation*} \left\Vert \left( g_{1},g_{2}\right) \right\Vert =\left\Vert g_{1}\right\Vert +\left\Vert g_{2}\right\Vert . \end{equation*}% It is obvious that the operator $A$ is continuous with respect to this norm. Besides, since $C\left( X\right) =H_{1}+H_{2},$ $A$ is a surjection. Consider the conjugate operator \begin{equation*} A^*:C\left( X\right) ^{\ast }\rightarrow \left[ H_{1}\times H_{2}\right] ^{\ast }, ~~~ A^{\ast }\left[ G\right] =\left( G_{1},G_{2}\right) , \end{equation*}% where the functionals $G_{1}$ and $G_{2}$ are defined as follows \begin{equation*} G_{1}\left( g_{1}\right) =G\left( g_{1}\right) ,g_{1}\in H_{1}; ~~~ G_{2}\left( g_{2}\right) =G\left( g_{2}\right) ,g_{2}\in H_{2}. \end{equation*} An element $\left( G_{1},G_{2}\right) $ from $\left[ H_{1}\times H_{2}\right] ^{\ast }$ has the norm \begin{equation*} \left\Vert \left( G_{1},G_{2}\right) \right\Vert =\max \left\{ \left\Vert G_{1}\right\Vert ,\left\Vert G_{2}\right\Vert \right\} .\eqno(1.40) \end{equation*} Let now $p=\left( p_{1},...,p_{m}\right) $ be any path with different points: $p_{i}\neq p_{j}$ for any $i\neq j$, $1\leq i,~j\leq m$. We associate with $p$ the following functional over $C\left( X\right) $ \begin{equation*} L\left[ f\right] =\frac{1}{m}\sum\limits_{i=1}^{m}\left( -1\right) ^{i-1}f\left( p_{i}\right) . \end{equation*}% Since $\left\vert L(f)\right\vert \leq \left\Vert f\right\Vert $ and $% \left\vert L(g)\right\vert =\left\Vert g\right\Vert $ for a continuous function $g(\mathbf{x})$ such that $g(p_{i})=1,\ $for odd indices $i,\ g(p_{j})=-1,$ for even\ indices$\ j\ $and $-1<g(\mathbf{x})<1$ elsewhere, we obtain that $\left\Vert L\right\Vert =1$. Let $A^{\ast }\left[ L\right] =\left( L_{1},L_{2}\right) $. One can easily verify that \begin{equation*} \left\Vert L_{i}\right\Vert \leq \frac{2}{m},i=1,2. \end{equation*}% Therefore, from (1.40) we obtain that \begin{equation*} \left\Vert A^{\ast }\left[ L\right] \right\Vert \leq \frac{2}{m}.\eqno(1.41) \end{equation*}% Since $A$ is a surjection, there exists $\delta >0$ such that \begin{equation*} \left\Vert A^{\ast }\left[ G\right] \right\Vert \geq \delta \left\Vert G\right\Vert ~~~~\;\mbox{for any functional}\;\ G\in C\left( X\right) ^{\ast } \end{equation*}% Hence \begin{equation*} \left\Vert A^{\ast }\left[ L\right] \right\Vert \geq \delta .\eqno(1.42) \end{equation*}% Now from (1.41) and (1.42) we conclude that \begin{equation*} m\leq \frac{2}{\delta }. \end{equation*} This means that for a path with different points, $n_{0}$ can be chosen as $% \left[ \frac{2}{\delta }\right] +1$. Let now $p=\left( p_{1},...,p_{m}\right) $ be a path with at least two coinciding points. Then we can form a closed path with different points. This may be done by the following way: let $i\ $and $j\ $be indices such that $p_{i}=\ p_{j}\ $and $j-i\ $takes its minimal value. Note that in this case all the points $p_{i},p_{i+1},...,p_{j-1}\ $are distinct. Now if $j-i\ $% is an even number, then the path $(p_{i},p_{i+1},...,p_{j-1})\ $, and if $\ j-i\ $is an odd number, then the path $(p_{i+1},...,p_{j-1})$ is a closed path with different points. It remains to show that $X$ can not possess closed paths with different points. Indeed, if $q=\left( q_{1},...,q_{2k}\right) $ is a path of this type, then the functional $L,$ associated with $q,$ annihilates all functions from $H_{1}+H_{2}$. On the other hand, $L\left[ f\right] =1$ for a continuous function $f$ on $X$ satisfying the conditions $f\left( t\right) =1$ if $t\in \left\{ q_{1},q_{3},...,q_{2k-1}\right\} ;$ $f\left( t\right) =-1$ if $t\in \left\{ q_{2},q_{4},...,q_{2k}\right\} ;$ $f\left( t\right) \in \left( -1;1\right) $ if $t\in X\backslash q$ . This implies on the contrary to our assumption that $H_{1}+H_{2}\neq C\left( X\right) $. The necessity has been proved. \textit{Sufficiency.} Let $X$ contains no closed path and the lengths of all paths are bounded by some positive integer $n_{0}$. We may suppose that any path has different points. Indeed, in other case we can form a closed path, which contradicts our assumption. For $i=1,2,$ let $X_{i}$ be the quotient space of $X$ obtained by identifying the points $a$ and $b$ whenever $g\left( a\right) =g\left( b\right) $ for each $g$ in $H_{i}$. Let $\pi _{i}$ be the natural projection of $X$ onto $X_{i}$. For a point $t\in X$ set $T_{1}=\pi _{1}^{-1}\left( \pi _{1}t\right) ,T_{2}=\pi _{2}^{-1}\left( \pi _{2}T_{1}\right) ,\ldots .$ By $% O\left( t\right) $ denote the orbit of $X$ containing $t.$ Since the length of any path in $X$ is not more than $n_{0}$, we conclude that $O\left( t\right) =T_{n_{0}}$. Since $X\ $is compact, the sets $% T_{1},T_{2},...,T_{n_{0}},\ $hence $O(t),$ are compact. By Theorem 1.6, $\overline{H_{1}+H_{2}}=C\left( X\right)$. Now let us show that $H_{1}+H_{2}$ is closed in $C\left( X\right) $. Set \begin{equation*} H_{3}=H_{1}\cap H_{2}. \end{equation*} Let $X_{3}$ and $\pi _{3}$ be the associated quotient space and projection. Fix some $a\in X_{3}$. Show, within conditions of our theorem, that if $t\in \pi _{3}^{-1}\left( a\right) ,$ then $O\left( t\right) =\pi _{3}^{-1}\left( a\right) $. The inclusion $O\left( t\right) \subset \pi _{3}^{-1}\left( a\right) $ is obvious. Suppose that there exists a point $t_{1}\in \pi _{3}^{-1}\left( a\right) $ such that $t_{1}\notin O\left( t\right) $. Then $O\left( t\right) \cap O\left( t_{1}\right) =\emptyset $. By $X|O$ denote the factor space generated by orbits of $X$. $X|O$ is a normal topological space with its natural factor topology. Hence we can construct a continuous function $u\in C\left( X|O\right) $ such that $u\left( O\left( t\right) \right) =0,$ $u\left( O\left( t_{1}\right) \right) =1$. The function $\upsilon \left( x\right) =u\left( O\left( x\right) \right) ,\;\ x\in X,$ is continuous on $X$ and belongs to $H_{3}$ as a function being constant on each orbit. But, since $O\left( t\right) \subset \pi _{3}^{-1}\left( a\right) $ and $O\left( t_{1}\right) \subset \pi _{3}^{-1}\left( a\right) $, the function $\upsilon \left( x\right) $ can not take different values on $O\left( t\right) $ and $O\left( t_{1}\right) $. This contradiction means that there is not a point $t_{1}\in \pi _{3}^{-1}\left( a\right) $ such that $t_{1}\notin O\left( t\right) $. Thus, \begin{equation*} O\left( t\right) =\pi _{3}^{-1}\left( a\right) \eqno(1.43) \end{equation*}% for any $a\in X_{3}$ and $t\in \pi _{3}^{-1}\left( a\right) $. Now prove that there exists a positive real number $c$ such that \begin{equation*} \sup\limits_{z\in X_{3}}\underset{\pi _{3}^{-1}\left( z\right) }{var}f\leq c\sup\limits_{y\in X_{2}}\underset{\pi _{2}^{-1}\left( y\right) }{var}f\eqno% (1.44) \end{equation*}% for all $f$ in $H_{1}$. Note that for $Y\subset X,\ \;\underset{Y}{var}f$ is the variation of $f$ on the set $Y.$ That is, $\;$% \begin{equation*} \underset{Y}{var}f=\sup\limits_{x,y\in Y}\left\vert f\left( x\right) -f\left( y\right) \right\vert . \end{equation*} Due to (1.43), inequality (1.44) can be written in the following form \begin{equation*} \sup_{t\in X}\underset{O\left( t\right) }{var}f\leq c\sup_{t\in X}\underset{% \pi _{2}^{-1}\left( \pi _{2}\left( t\right) \right) }{var}f\eqno(1.45) \end{equation*}% for all $f\in H_{1}$. Let $t\in X$ and $t_{1},t_{2}$ be arbitrary points of $O\left( t\right) $. Then there is a path $\left( b_{1},b_{2},...,b_{m}\right) $ with $% b_{1}=t_{1} $ and $b_{m}=t_{2}$. Besides, by the condition, $m\leq n_{0}$ . Let first $\mathbf{a}^{2}\cdot b_{1}=\mathbf{a}^{2}\cdot b_{2},$ $\mathbf{a}% ^{1}\cdot b_{2}=\mathbf{a}^{1}\cdot b_{3},...,\mathbf{a}^{2}\cdot b_{m-1}=% \mathbf{a}^{2}\cdot b_{m}$. Then for any function $f\in H_{1}$ \begin{equation*} \left\vert f\left( t_{1}\right) -f\left( t_{2}\right) \right\vert =\left\vert f\left( b_{1}\right) -f\left( b_{2}\right) +...-f\left( b_{m}\right) \right\vert \leq \end{equation*}% \begin{equation*} \leq \left\vert f\left( b_{1}\right) -f\left( b_{2}\right) \right\vert +...+\left\vert f\left( b_{m-1}\right) -f\left( b_{m}\right) \right\vert \leq \frac{n_{o}}{2}\sup_{t\in X}\underset{\pi _{2}^{-1}\left( \pi _{2}\left( t\right) \right) }{var}f.\eqno(1.46) \end{equation*} It is not difficult to verify that inequality (1.46) holds in all other possible cases of the path $\left( b_{1},...,b_{m}\right) $. Now from (1.46) we obtain (1.45), hence (1.44), where $c=\frac{n_{0}}{2}$. In \cite{108}, Marshall and O'Farrell proved the following result (see \cite[Proposition 4]{108}): Let $A_{1}\ $and $A_{2}\ $be closed subalgebras of $C(X)\ $that contain the constants. Let $(X_{1},\pi _{1}),\ (X_{2},\pi _{2})\ $and $% (X_{3},\pi _{3})\ $be the quotient spaces and projections associated with the algebras $A_{1},$ $A_{2}\ $and $A_{3}=A_{1}\cap A_{2}\ $respectively. Then $A_{1}+A_{2}\ $is closed in $C(X)\ $if and only if there exists a positive real number $c$ such that \begin{equation*} \sup\limits_{z\in X_{3}}\underset{\pi _{3}^{-1}\left( z\right) }{var}f\leq c\sup\limits_{y\in X_{2}}\underset{\pi _{2}^{-1}\left( y\right) }{var}f \end{equation*}% for all $f\ $in $A_{1}.$ By this proposition, (1.44) implies that $H_{1}+H_{2}$ is closed in $C\left( X\right) $. Thus we finally obtain that $H_{1}+H_{2}=C\left( X\right) $. \end{proof} Paths with respect to two directions are explicit objects and give geometric means of deciding if $H_{1}+H_{2}=C\left( X\right) $. Let us show this in the example of the bivariate ridge functions $g_{1}=x_{1}+x_{2}\ $and $% g_{2}=x_{1}-x_{2}.$ If $X$ is the union of two parallel line segments in $% \mathbb{R}^{2},$ not parallel to any of the lines $x_{1}+x_{2}=0$ and $% x_{1}-x_{2}=0,\ $then Theorem 1.7 holds. If $X$ is any bounded part of the graph of the function $x_{2}=\arcsin (\sin x_{1}),$ then Theorem 1.7 also holds. Let now $X\ $be the set \begin{equation*} \begin{array}{c} \{(0,0),(1,-1),(0,-2),(-1\frac{1}{2},-\frac{1}{2}),(0,1),(\frac{3}{4},\frac{1% }{4}),(0,-\frac{1}{2}), \\ (-\frac{3}{8},-\frac{1}{8}),(0,\frac{1}{4}),(\frac{3}{16},\frac{1}{16}% ),...\}.% \end{array}% \end{equation*} In this case, there is no positive integer bounding lengths of all paths. Thus Theorem 1.7 fails. Note that since orbits of all paths are closed, Theorem 1.6 from the previous section shows $H_{1}+H_{2}$ is dense in $% C\left( X\right) .$ If $X$ is any set with interior points, then both Theorem 1.6 and Theorem 1.7 fail, since any such set contains the vertices of some parallelogram with sides parallel to the directions $\mathbf{a}^{1}$ and $\mathbf{a}^{2}$, that is a closed path. Theorem 1.7 admits a direct generalization to the representation by sums $% g_{1}\left( h_{1}\left( \mathbf{x}\right) \right) +g_{2}\left( h_{2}\left( \mathbf{x}\right) \right) $, where $h_{1}\left( \mathbf{x}\right) $ and $% h_{2}\left( \mathbf{x}\right) $ are fixed continuous functions on $X$. This generalization needs consideration of new objects -- paths with respect to two continuous functions. \bigskip \textbf{Definition 1.5.} \textit{Let $X$ be a compact set in $\mathbb{R}% ^{d}$ and $h_{1},h_{2}\in C\left( X\right)$. A finite ordered subset $% \left( p_{1},p_{2},...,p_{m}\right)$ of $X$ with $p_{i}\neq p_{i+1}\left( i=1,...,m-1\right)$, and either $h_{1}\left( p_{1}\right) =h_{1}\left( p_{2}\right)$, $h_{2}\left( p_{2}\right) =h_{2}\left( p_{3}\right)$, $% h_{1}\left( p_{3}\right) =h_{1}\left( p_{4}\right),...,$ or $h_{2}\left( p_{1}\right) =h_{2}\left( p_{2}\right)$, $h_{1}\left( p_{2}\right) =h_{1}\left( p_{3}\right)$, $h_{2}\left( p_{3}\right) =h_{2}\left( p_{4}\right),...,$ is called a path with respect to the functions $h_{1}$ and $h_{2}$ or, shortly, an $h_{1}$-$h_{2}$ path.} \bigskip \textbf{Theorem 1.8.} \textit{Let $X$ be a compact subset of $\mathbb{R}^{d}$% . All functions $f \in C(X)$ admit a representation \begin{equation*} f(\mathbf{x})=g_{1}\left( h_{1}\left( \mathbf{x}\right) \right) +g_{2}\left( h_{2}\left( \mathbf{x}\right) \right) ,~g_{1},g_{2}\in C(\mathit{\mathbb{R}}) \end{equation*}% if and only if the set $X$ contains no closed $h_{1}$-$h_{2}$ path and there exists a positive integer $n_{0}$ such that the lengths of $h_{1}$-$h_{2}$ paths in $X$ are bounded by $n_{0}$.} \bigskip The proof can be carried out by the same arguments as above. It should be noted that Theorem 1.8 was first proved by Khavinson in his monograph \cite{76}. Khavinson's proof (see \cite[p.87]{76}) used theorems of Sternfeld \cite{132} and Medvedev \cite[Theorem 2.2]{76}, whereas our proof, which generalizes the ideas of Khavinson, was based on the above proposition of Marshall and O'Farrell. \bigskip \section{On the proximinality of ridge functions} In this section, using two results of Garkavi, Medvedev and Khavinson \cite% {35}, we give sufficient conditions for proximinality of sums of two ridge functions with bounded and continuous summands in the spaces of bounded and continuous multivariate functions, respectively. In the first case, we give an example which shows that the corresponding sufficient condition cannot be made weaker for certain subsets of $\mathbb{R}^{n}$. In the second case, we obtain also a necessary condition for proximinality. All the results are furnished with plenty of examples. The results, examples and following discussions naturally lead us to a conjecture on the proximinality of the considered class of ridge functions. \subsection{Problem statement} Let $E$ be a normed linear space and $F$ be its subspace. We say that $F$ is proximinal in $E$ if for any element $e\in E$ there exists at least one element $f_{0}\in F$ such that \begin{equation*} \left\Vert e-f_{0}\right\Vert =\inf_{f\in F}\left\Vert e-f\right\Vert . \end{equation*} In this case, the element $f_{0}$ is said to be extremal to $e$. We are interested in the problem of proximinality of the set of linear combinations of ridge functions in the spaces of bounded and continuous functions respectively. This problem will be considered in the simplest case when the class of approximating functions is the set \begin{equation*} \mathcal{R}=\mathcal{R}\left( \mathbf{a}^{1},\mathbf{a}^{2}\right) ={\ \left\{ {g_{1}\left( \mathbf{a}^{1}{\cdot }\mathbf{x}\right) +g_{2}\left( \mathbf{a}^{2}{\cdot }\mathbf{x}\right) :g}_{i}:{{\mathbb{R\rightarrow R}} ,i=1,2}\right\} }. \end{equation*} Here $\mathbf{a}^{1}${and }$\mathbf{a}^{2}$ are fixed directions and we vary over ${g}_{i}$. It is clear that this is a linear space. Consider the following three subspaces of $\mathcal{R}$. The first is obtained by taking only bounded sums ${g_{1}\left( \mathbf{a}^{1}{\cdot }\mathbf{x}\right) +g_{2}\left( \mathbf{a}^{2}{\cdot }\mathbf{x}\right) }$ over some set $X$ in $\mathbb{R}^{n}.$ We denote this subspace by $\mathcal{R}_{a}(X)$. The second and the third are subspaces of $\mathcal{R}$ with bounded and continuous summands $g_{i}\left( \mathbf{a}^{i}\cdot \mathbf{x}\right) ,~i=1,2,$ on $X$ respectively. These subspaces will be denoted by $\mathcal{% R }_{b}(X)$ and $\mathcal{R}_{c}(X).$ In the case of $\mathcal{R}_{c}(X),$ the set $X$ is considered to be compact. Let $B(X)$ and $C(X)$ be the spaces of bounded and continuous multivariate functions over $X$ respectively. What conditions must one impose on $X$ in order that the sets $\mathcal{R}_{a}(X)$ and $\mathcal{R}_{b}(X)$ be proximinal in $B(X)$ and the set $\mathcal{R}_{c}(X)$ be proximinal in $C(X)$% ? We are also interested in necessary conditions for proximinality. It follows from one result of Garkavi, Medvedev and Khavinson (see \cite[Theorem 1]{35}) that $\mathcal{R}_{a}(X)$ is proximinal in $B(X)$ for all subsets $X$ of $\mathbb{R}^{n}$. There is also an answer (see \cite[Theorem 2]{35}) for proximinality of $\mathcal{R}_{b}(X)$ in $B(X)$. This will be discussed in Section 1.5.2. Is the set $\mathcal{R}_{b}(X)$ always proximinal in $B(X)$% ? There is an an example of a set $X\subset \mathbb{R}^{n}$ and a bounded function $f$ on $X$ for which there does not exist an extremal element in $% \mathcal{R}_{b}(X)$. In Section 1.5.3, we will obtain sufficient conditions for the existence of extremal elements from $\mathcal{R}_{c}(X)$ to an arbitrary function $f$ $% \in $ $C(X)$. Based on one result of Marshall and O'Farrell \cite{108}, we will also give a necessary condition for proximinality of $\mathcal{R}_{c}(X) $ in $C(X)$. All the theorems, following discussions and examples of the paper will lead us naturally to a conjecture on the proximinality of the subspaces $\mathcal{R}_{b}(X)$ and $\mathcal{R}_{c}(X)$ in the spaces $B(X)$ and $C(X)$ respectively. The reader may also be interested in the more general case with the set $% \mathcal{R}=\mathcal{R}\left( \mathbf{a}^{1},...,\mathbf{a}^{r}\right)$. In this case, the corresponding sets $\mathcal{R}_{a}(X)$, $\mathcal{R}_{b}(X)$ and $\mathcal{R}_{c}(X)$ are defined similarly. Using the results of \cite% {35}, one can obtain sufficient (but not necessary) conditions for proximinality of these sets. This needs, besides paths, the consideration of some additional and more complicated relations between points of $X$. Here we will not consider the case $r\geq 3$, since our main purpose is to draw the reader's attention to the arisen problems of proximinality in the simplest case of approximation. For the existing open problems connected with the set $\mathcal{R}\left( \mathbf{a}^{1},...,\mathbf{a}^{r}\right) $, where $r\geq 3$, see \cite{53} and \cite{118}. \bigskip \subsection{Proximinality of $\mathcal{R}_{b}(X)$ in $B(X)$} Let $\mathbf{a}^{1}$ and $\mathbf{a}^{2}$ be two different directions in $% \mathbb{R}^{n}$. In the sequel, we will use paths with respect to the directions $\mathbf{a}^{1}$ and $\mathbf{a}^{2}$. Recall that a length of a path is the number of its points and can be equal to $\infty $ if the path is infinite. A singleton is a path of the unit length. We say that a path $% \left( \mathbf{x}^{1},...,\mathbf{x}^{m}\right) $ belonging to some subset $X $ of $\mathbb{R}^{n}$ is irreducible if there is not another path $\left( \mathbf{y}^{1},...,\mathbf{y}^{l}\right) \subset X$ with $\mathbf{y}^{1}=% \mathbf{x}^{1},~\mathbf{y}^{l}=\mathbf{x}^{m}$ and $l<m$. The following theorem follows from \cite[Theorem 2]{35}. \bigskip \textbf{Theorem 1.9.} \textit{Let $X\subset $ $\mathbb{R}^{n}$ and the lengths of all irreducible paths in $X$ be uniformly bounded by some positive integer. Then each function in $B(X)$ has an extremal element in $% \mathcal{R}_{b}(X)$.} \bigskip There are a large number of sets in $\mathbb{R}^{n}$ satisfying the hypothesis of this theorem. For example, if a set $X$ has a cross section according to one of the directions $\mathbf{a}^{1}$ or $\mathbf{a}^{2}$, then the set $X$ satisfies the hypothesis of Theorem 1.9. By a cross section according to the direction $\mathbf{a}^{1}$ we mean any set $X_{\mathbf{a}% ^{1}}=\{x\in X:\ \mathbf{a}^{1}\cdot \mathbf{x}=c\}$, $c\in \mathbb{R}$, with the property: for any $\mathbf{y}\in X$ there exists a point $\mathbf{y}% ^{1}\in X_{\mathbf{a}^{1}}$ such that $\mathbf{a}^{2}\cdot \mathbf{y}=% \mathbf{a}^{2}\cdot \mathbf{y}^{1}$. By the similar way, one can define a cross section according to the direction $\mathbf{a}^{2}$. For more on cross sections in problems of proximinality of sums of univariate functions see \cite{34,77}. Regarding Theorem 1.9 one may ask if the condition of the theorem is necessary for proximinality of $\mathcal{R}_{b}(X)$ in $B(X)$. While we do not know a complete answer to this question, we are going to give an example of a set $X $ for which Theorem 1.9 fails. Let $\mathbf{a}% ^{1}=(1;-1),\ \mathbf{a}^{2}=(1;1).$ Consider the set \begin{eqnarray*} X &=&\{(2;\frac{2}{3}),(\frac{2}{3};-\frac{2}{3}),(0;0),(1;1),(1+\frac{1}{2}% ;1-\frac{1}{2}),(1+\frac{1}{2}+\frac{1}{4};1-\frac{1}{2}+\frac{1}{4}), \\ &&(1+\frac{1}{2}+\frac{1}{4}+\frac{1}{8};1-\frac{1}{2}+\frac{1}{4}-\frac{1}{8% }),...\}. \end{eqnarray*}% In what follows, the elements of $X$ in the given order will be denoted by $% \mathbf{x}^{0},\mathbf{x}^{1},\mathbf{x}^{2},...$ . It is clear that $X$ is a path of the infinite length and $\mathbf{x}^{n}\rightarrow \mathbf{x}^{0}$ as $n\rightarrow \infty $. Let $\sum_{n=1}^{\infty }c_{n}$ be any divergent series with the terms $c_{n}>0$ and $c_{n}\rightarrow 0$ as $% n\rightarrow \infty $. Besides let $f_{0}$ be a function vanishing at the points $\mathbf{x}^{0},\mathbf{x}^{2},\mathbf{x}^{4},...,$ and taking values $c_{1},c_{2},c_{3},...$ at the points $\mathbf{x}^{1},\mathbf{x}^{3},\mathbf{% \ x}^{5},...$, respectively. It is obvious that $f_{0}$ is continuous on $X$. The set $X$ is compact and satisfies all the conditions of Theorem 1.6. By that theorem, $\overline{\mathcal{R}_{c}(X)}=C(X).$ Therefore, for any continuous function on $X$, thus for $f_{0}$, \begin{equation*} \inf_{g\in \mathcal{R}_{c}(X)}\left\Vert f_{0}-g\right\Vert _{C(X)}=0.\eqno% (1.47) \end{equation*} Since $\mathcal{R}_{c}(X)\subset \mathcal{R}_{b}(X),$ we obtain from (1.47) that \begin{equation*} \inf_{g\in \mathcal{R}_{b}(X)}\left\Vert f_{0}-g\right\Vert _{B(X)}=0.\eqno% (1.48) \end{equation*} Suppose that $f_{0}$ has an extremal element ${g_{1}^{0}\left( \mathbf{a}^{1}% {\cdot }\mathbf{x}\right) +g_{2}^{0}\left( \mathbf{a}^{2}{\ \cdot }\mathbf{x}% \right) }$ in $\mathcal{R}_{b}(X).$ By the definition of $\mathcal{R}_{b}(X)$% , the ridge functions ${g_{i}^{0},i=1,2}$, are bounded on $X.$ From (1.48)\ it follows that $f_{0}={g_{1}^{0}\left( \mathbf{a}^{1}{\ \cdot }\mathbf{x}% \right) +g_{2}^{0}\left( \mathbf{a}^{2}{\cdot }\mathbf{x}\right) .}$ Since $% \mathbf{a}^{1}\cdot \mathbf{x}^{2n}=\mathbf{a}^{1}\cdot \mathbf{x}^{2n+1}$ and $\mathbf{a}^{2}\cdot \mathbf{x}^{2n+1}=\mathbf{a}^{2}\cdot \mathbf{x}% ^{2n+2},$ for $n=0,1,...,$ we can write that \begin{equation*} \sum_{n=0}^{k}c_{n+1}=\sum_{n=0}^{k}\left[ f(\mathbf{x}^{2n+1})-f(\mathbf{x}% ^{2n})\right] \end{equation*} \begin{equation*} =\sum_{n=0}^{k}\left[ {g_{2}^{0}}(\mathbf{x}^{2n+1})-{g_{2}^{0}}(\mathbf{x}% ^{2n})\right] ={g_{2}^{0}(}\mathbf{a}^{2}\cdot \mathbf{x}^{2k+1})-{g_{2}^{0}(% }\mathbf{a}^{2}\cdot \mathbf{x}^{0}).\eqno(1.49) \end{equation*} Since $\sum_{n=1}^{\infty }c_{n}=\infty ,$ we deduce from (1.49)\ that the function ${g_{2}^{0}\left( \mathbf{a}^{2}{\cdot }\mathbf{x}\right) }$ is not bounded on $X.$ This contradiction means that the function $f_{0}$ does not have an extremal element in $\mathcal{R}_{b}(X).$ Therefore, the space $% \mathcal{R}_{b}(X)$ is not proximinal in $B(X).$ \bigskip \subsection{Proximinality of $\mathcal{R}_{c}(X)$ in $C(X)$} In this section, we give a sufficient condition and also a necessary condition for proximinality of $\mathcal{R}_{c}(X)$ in $C(X)$. \bigskip \textbf{Theorem 1.10.} \textit{Let the system of linearly independent vectors $% \mathbf{a}^{1}$ and $\mathbf{a}^{2}$ have a complement to a basis $\{\mathbf{a% }^{1},...,\mathbf{a}^{n}\}$ in $\mathbb{R}^{n}$ with the property: for any point $\mathbf{x}^{0}\in X$ and any positive real number $\delta $ there exist a number $\delta _{0}\in (0,\delta ]$ and a point $\mathbf{x}^{\sigma } $ in the set} \begin{equation*} \sigma =\{\mathbf{x}\in X:\mathbf{a}^{2}\cdot \mathbf{x}^{0}-\delta _{0}\leq \mathbf{a}^{2}\cdot \mathbf{x}\leq \mathbf{a}^{2}\cdot \mathbf{x}^{0}+\delta _{0}\}, \end{equation*} \textit{such that the system} \begin{equation*} \left\{ \begin{array}{c} \mathbf{a}^{2}\cdot \mathbf{x}^{\prime }=\mathbf{a}^{2}\cdot \mathbf{x}% ^{\sigma } \\ \mathbf{a}^{1}\cdot \mathbf{x}^{\prime }=\mathbf{a}^{1}\cdot \mathbf{x} \\ \sum_{i=3}^{n}\left\vert \mathbf{a}^{i}\cdot \mathbf{x}^{\prime }-\mathbf{a}% ^{i}\cdot \mathbf{x}\right\vert <\delta% \end{array}% \right. \eqno(1.50) \end{equation*}% \textit{has a solution $\mathbf{x}^{\prime }\in \sigma $ for all points $% \mathbf{x}\in \sigma .$Then the space $\mathcal{R}_{c}(X)$ is proximinal in $% C(X).$} \bigskip \begin{proof} Introduce the following mappings and sets: \begin{equation*} \pi _{i}:X\rightarrow \mathbb{R}\text{, }\pi _{i}(\mathbf{x)=a}^{i}\cdot \mathbf{x}\text{, }Y_{i}=\pi _{i}(X\mathbf{)}\text{, }i=1,...,n. \end{equation*} Since the system of vectors $\{\mathbf{a}^{1},...,\mathbf{a}^{n}\}$ is linearly independent, the mapping $\pi =(\pi _{1},...\pi _{n})$ is an injection from $X$ into the Cartesian product $Y_{1} \times ...\times Y_{n} $ . Besides, $\pi $ is linear and continuous. By the open mapping theorem, the inverse mapping $\pi ^{-1}$ is continuous from $Y=\pi (X)$ onto $X.$ Let $f$ be a continuous function on $X$. Then the composition $f\circ \pi ^{-1}(y_{1},...y_{n})$ will be continuous on $Y,$ where $y_{i}=\pi _{i}( \mathbf{x),}\ i=1,...,n,$ are the coordinate functions. Consider the approximation of the function $f\circ \pi ^{-1}$ by elements from \begin{equation*} G_{0}=\{g_{1}(y_{1})+g_{2}(y_{2}):\ g_{i}\in C(Y_{i}),\ i=1,2\} \end{equation*} over the compact set $Y$. Then one may observe that the function $f$ has an extremal element in $\mathcal{R}_{c}(X)$ if and only if the function $f\circ \pi ^{-1}$ has an extremal element in $G_{0}$. Thus the problem of proximinality of $\mathcal{R}_{c}(X)$ in $C(X)$ is reduced to the problem of proximinality of $G_{0}$ in $C(Y).$ Let $T,T_{1},...,T_{m+1}$ be metric compact spaces and $T\subset $ $% T_{1}\times ...\times T_{m+1}.$ For $i=1,...,m,$ let $\varphi _{i}$ be the continuous mappings from $T$ onto $T_{i}.$ In \cite{35}, the authors obtained sufficient conditions for proximinality of the set \begin{equation*} C_{0}=\{\sum_{i=1}^{n}g_{i}\circ \varphi _{i}:\ g_{i}\in C(T_{i}),\ i=1,...m\} \end{equation*}% in the space $C(T)$ of continuous functions on $T.$ Since $Y\subset $ $% Y_{1}\times Y_{2}\times Z_{3},$ where $Z_{3}=Y_{3}\times ...\times Y_{n},$ we can use this result in our case, for the approximation of the function $% f\circ \pi ^{-1}$ by elements from $G_{0}$. By this theorem, the set $G_{0}$ is proximinal in $C(Y)$ if for any $y_{2}^{0}\in Y_{2}$ and $\delta >0$ there exists a number $\delta _{0}\in (0,$ $\delta )$ such that the set $% \sigma (y_{2}^{0},\delta _{0})=[y_{2}^{0}-\delta _{0},y_{2}^{0}+\delta _{0}]\cap Y_{2}$ has $(2,\delta )$ maximal cross section. The last means that there exists a point $y_{2}^{\sigma }\in \sigma (y_{2}^{0},\delta _{0})$ with the property: for any point $(y_{1},y_{2},z_{3})\in Y,$ with the second coordinate $y_{2}$ from the set $\sigma (y_{2}^{0},\delta _{0}),$ there exists a point $(y_{1}^{\prime },y_{2}^{\sigma },z_{3}^{\prime })\in Y$ such that $y_{1}=y_{1}^{\prime }$ and $\rho (z_{3},z_{3}^{\prime })<\delta ,$ where $\rho $ is a metrics in $Z_{3}.$ Since these conditions are equivalent to the conditions of Theorem 1.10, the space $G_{0}$ is proximinal in the space $C(Y).$ Then by the above conclusion, the space $\mathcal{R}_{c}(X)$ is proximinal in $C(X).$ \end{proof} Let us give some simple examples of compact sets satisfying the hypothesis of Theorem 1.10. For the sake of brevity, we restrict ourselves to the case $% n=3.$ \begin{enumerate} \item[(a)] Assume $X$ is a closed ball in $\mathbb{R}^{3}$ and $\mathbf{a}^{1}$, $\mathbf{a}^{2}$ are orthogonal directions. Then Theorem 1.10 holds. Note that in this case, we can take $\delta _{0}=\delta $ and $\mathbf{a}^{3}$ as an orthogonal vector to both the vectors $\mathbf{a}^{1}$ and $\mathbf{a}^{2}.$ \item[(b)] Let $X$ be the unite cube, $\mathbf{a}^{1}=(1;1;0),\ a^{2}=(1;-1;0).$ Then Theorem 1.10 also holds. In this case, we can take $\delta _{0}=\delta $ and $\mathbf{a}^{3}=(0;0;1).$ Note that the unit cube does not satisfy the hypothesis of the theorem for many directions (take, for example, $\mathbf{a}^{1}=(1;2;0)$ and $\mathbf{a}^{2}=(2;-1;0)$). \end{enumerate} In the following example, one can not always chose $\delta _{0}$ as equal to $\delta $. \begin{enumerate} \item[(c)] Let $X=\{(x_{1},x_{2},x_{3}):\ (x_{1},x_{2})\in Q,\ 0\leq x_{3}\leq 1\},$ where $Q$ is the union of two triangles $A_{1}B_{1}C_{1}$ and $A_{2}B_{2}C_{2}$ with the vertices $A_{1}=(0;0),\ B_{1}=(1;2),\ C_{1}=(2;0),\ A_{2}=(1\frac{1}{2};1),\ B_{2}=(2\frac{1}{2};-1),\ C_{2}=(3% \frac{1}{2};1).$ Let $\mathbf{a}^{1}=(0;1;0)$ and $\mathbf{a}^{2}=(1;0;0).$ Then it is easy to see that Theorem 1.10 holds (the vector $\mathbf{a}^{3}$ can be chosen as $(0;0;1)$). In this case, $\delta _{0}$ can not be always chosen as equal to $\delta $. Take, for example, $\mathbf{x}^{0}=(1\frac{3}{4};0;0)$ and $\delta =1\frac{3% }{4}.$ If $\delta _{0}=\delta ,$ then the second equation of the system (1.50) has not a solution for a point $(1;2;0)$ or a point $(2\frac{1}{2}% ;-1;0).$ But if we take $\delta _{0}$ not more than $\frac{1}{4}$, then for $% \mathbf{x}^{\sigma }=\mathbf{x}^{0}$ the system has a solution. Note that the last inequality $\left\vert \mathbf{a}^{3}\cdot \mathbf{x}^{\prime }-% \mathbf{a}^{3}\cdot \mathbf{x}\right\vert <\delta $ of the system can be satisfied with the equality $\mathbf{a}^{3}\cdot \mathbf{x}^{\prime }=% \mathbf{a}^{3}\cdot \mathbf{x}$ if $\mathbf{a}^{3}=(0;0;1).$ \end{enumerate} It should be remarked that the results of \cite{35} tell nothing about necessary conditions for proximinality of the spaces considered there. To fill this gap in our case, we want to give a necessary condition for proximinality of $\mathcal{R}_{c}(X)$ in $C(X)$. First, let us introduce some notation. By $\mathcal{R}_{c}^{i},\ i=1,2,$ we will denote the set of continuous ridge functions $g\left( \mathbf{a}^{i}\cdot \mathbf{x}\right) $ on the given compact set $X\subset \mathbb{R}^{n}.$ Note that $\mathcal{R}% _{c}=\mathcal{R}_{c}^{1}+\mathcal{R}_{c}^{2}.$ Besides, let $\mathcal{R}% _{c}^{3}=\mathcal{R}_{c}^{1}\cap \mathcal{R}_{c}^{2}.$ For $i=1,2,3,$ let $% X_{i}$ be the quotient space obtained by identifying points $y_{1}$ and $% y_{2}$ in $X$ whenever $f(y_{1})=f(y_{2})$ for each $f$ in $\mathcal{R}% _{c}^{i}.$ By $\pi _{i}$ denote the natural projection of $X$ onto $X_{i},$ $% i=1,2,3.$ Note that we have already dealt with the quotient spaces $X_{1}$, $% X_{2}$ and the projections $\pi _{1},\pi _{2}$ in the previous section. Recall that the relation on $X$, defined by setting $\ y_{1}\approx y_{2}$ if $y_{1}$ and $y_{2}$ belong to some path, is an equivalence relation and the equivalence classes are called orbits. By $O(t)$ denote the orbit of $X$ containing $t.$ For $Y\subset X,$ let $var_{Y}\ f$ be the variation of a function $f$ on the set $Y.$ That is, \begin{equation*} \underset{Y}{var}f=\sup\limits_{x,y\in Y}\left\vert f\left( x\right) -f\left( y\right) \right\vert . \end{equation*} The following theorem is valid. \bigskip \textbf{Theorem 1.11.} \textit{Suppose that the space $\mathcal{R}_{c}(X)$ is proximinal in $C(X).$Then there exists a positive real number c such that} \begin{equation*} \sup_{t\in X}\underset{O\left( t\right) }{var}\mathit{f\leq c}\sup_{t\in X}% \underset{\pi _{2}^{-1}\left( \pi _{2}\left( t\right) \right) }{var}\mathit{f% }\eqno(1.51) \end{equation*}% \textit{for all $f$ in $\mathcal{R}_{c}^{1}.$} \bigskip \begin{proof} The proof is based on the following result of Marshall and O'Farrell (see \cite[Proposition 4]{108}): Let $A_{1}\ $and $A_{2}\ $be closed subalgebras of $C(X)\ $that contain the constants. Let $% (X_{1},\pi _{1}),\ (X_{2},\pi _{2})\ $and $(X_{3},\pi _{3})\ $be the quotient spaces and projections associated with the algebras $A_{1},$ $% A_{2}\ $and $A_{3}=A_{1}\cap A_{2}\ $respectively. Then $A_{1}+A_{2}\ $is closed in $C(X)\ $if and only if there exists a positive real number $c$ such that \begin{equation*} \sup\limits_{z\in X_{3}}\underset{\pi _{3}^{-1}\left( z\right) }{var}f\leq c\sup\limits_{y\in X_{2}}\underset{\pi _{2}^{-1}\left( y\right) }{var}f\eqno% (1.52) \end{equation*}% for all $f\ $in $A_{1}.$ If $\mathcal{R}_{c}(X)$ is proximinal in $C(X),$ then it is necessarily closed and therefore, by the above proposition, (1.52) holds for the algebras $A_{1}^{i}=\mathcal{R}_{c}^{i},\ i=1,2,3.$ The right-hand side of (1.52) is equal to the right-hand side of (1.51). Let $t$ be some point in $% X $ and $z=\pi _{3}(t).$ Since each function $\ f\in \mathcal{R}_{c}^{3}$ is constant on the orbit of $t$ (note that $f$ is both of the form ${\ g_{1}\left( \mathbf{a}^{1}{\cdot }\mathbf{x}\right) }$ and of the form ${\ g_{2}\left( \mathbf{a}^{2}{\cdot }\mathbf{x}\right) }$), $O(t)\subset \pi _{3}^{-1}(z).$ Hence, \begin{equation*} \sup_{t\in X}\underset{O\left( t\right) }{var}f\leq c\sup\limits_{z\in X_{3}}% \underset{\pi _{3}^{-1}\left( z\right) }{var}f\eqno(1.53) \end{equation*} From (1.52) and (1.53) we obtain (1.51). \end{proof} Note that the inequality (1.52) provides not worse but less practicable necessary condition for proximinality than the inequality (1.51) does. On the other hand, there are many cases in which both the inequalities are equivalent. For example, assume the lengths of irreducible paths of $X$ are bounded by some positive integer $n_{0}$. In this case, it can be shown that the inequality (1.52), hence (1.51), holds with the constant $c=\frac{n_{0}}{% 2}$ and moreover $O(t)=\pi _{3}^{-1}(z)$ for all $t\in X$, where $z=\pi _{3}(t)$ (see the proof of \cite[Theorem 5]{53}). Therefore, the inequalities (1.51) and (1.52) are equivalent for the considered class of sets $X.$ The last argument shows that all the compact sets $X\subset $ $% \mathbb{R}^{n}$ over which $\mathcal{R}_{c}(X)$ is not proximinal in $C(X)$ should be sought in the class of sets having irreducible paths consisting of sufficiently many points. For example, let $I=[0;1]^{2}$ be the unit square, $\mathbf{a}^{1}=(1;1)$, $\mathbf{a}^{2}=(1;\frac{1}{2}).$ Consider the path \begin{equation*} l_{k}=\{(1;0),(0;1),(\frac{1}{2};0),(0;\frac{1}{2}),(\frac{1}{4};0),...,(0; \frac{1}{2^{k}})\}. \end{equation*} It is clear that $l_{k}$ is an irreducible path with the length $2k+2 $, where $k$ may be very large. Let $g_{k}$ be a continuous univariate function on $\mathbb{R}$ satisfying the conditions: $g_{k}(\frac{1}{2^{k-i}} )=i,\ i=0,...,k,$ $g_{k}(t)=0$ if $t<\frac{1}{2^{k}},\ i-1\leq g_{k}(t)\leq i $ if $t\in (\frac{1}{2^{k-i+1}},\frac{1}{2^{k-i}}),\ i=1,...,k,$ and $g_{k}(t)=k$ if $t>1.$ Then it can be easily verified that \begin{equation*} \sup_{t\in X}\underset{\pi _{2}^{-1}\left( \pi _{2}\left( t\right) \right) }{% var}g_{k}(\mathbf{a}^{1}{\cdot }\mathbf{x})\leq 1.\eqno(1.54) \end{equation*} Since $\max_{\mathbf{x}\in I}g_{k}(\mathbf{a}^{1}{\cdot }\mathbf{x} )=k,$ $% \min_{\mathbf{x}\in I}g_{k}(\mathbf{a}^{1}{\cdot }\mathbf{x})=0$ and $var_{% \mathbf{x}\in O\left( t_{1}\right) }g_{k}(\mathbf{a}^{1}{\cdot }\mathbf{\ x}% )=k $ for $t_{1}=(1;0),$ we obtain that \begin{equation*} \sup_{t\in X}\underset{O\left( t\right) }{var}g_{k}(\mathbf{a}^{1}{\cdot }% \mathbf{x})=k.\eqno(1.55) \end{equation*} Since $k$ may be very large, from (1.54) and (1.55) it follows that the inequality (1.51) cannot hold for the function $g_{k}(\mathbf{a}^{1}{\ \cdot }\mathbf{x})\in \mathcal{R}_{c}^{1}.$ Thus the space $\mathcal{R}_{c}(I)$ with the directions $\mathbf{a}^{1}=(1;1)$ and $\mathbf{a}^{2}=(1;\frac{1}{2})$ is not proximinal in $C(I)$. It should be remarked that if a compact set $X\subset $ $\mathbb{R}^{n}$ satisfies the hypothesis of Theorem 1.10, then the length of all irreducible paths are uniformly bounded (see the proof of Theorem 1.10 and lemma in \cite% {35}). We have already seen that if the last condition does not hold, then the proximinality of both $\mathcal{R}_{c}(X)$ in $C(X)$ and $\mathcal{R}% _{b}(X)$ in $B(X)$ fail for some sets $X$. In addition to the examples given above and in Section 1.5.2, one can easily construct many other examples of such sets. All these examples, Theorems 1.9--1.11 and the subsequent remarks justify the statement of the following conjecture: \bigskip \textbf{Conjecture. }\textit{Let $X$ be some subset of $\mathbb{R} ^{n}.$ The space $\mathcal{R}_{b}(X)$ is proximinal in $B(X)$ and the space $% \mathcal{R}_{c}(X)$ is proximinal in $C(X)$ (in this case, $X$ is considered to be compact) if and only if the lengths of all irreducible paths of $X$ are uniformly bounded.} \bigskip \textbf{Remark 1.2.} Medvedev's result (see \cite[p.58]{76}), which later came to our attention, in particular, says that the set $R_{c}(X)$ is closed in $C(X)$ if and only if the lengths of all irreducible paths of $X$ are uniformly bounded. Thus, in the case of $C(X)$, the necessity of the above conjecture was proved by Medvedev. \bigskip \textbf{Remark 1.3.} Note that there are situations in which a continuous function (a specific function on a specially constructed set) has an extremal element in $\mathcal{R}_{b}(X)$, but not in $\mathcal{R}_{c}(X)$ (see \cite[p.73]{76}). One subsection of \cite{76} (see p.68 there) was devoted to the proximinality of sums of two univariate functions with continuous and bounded summands in the spaces of continuous and bounded bivariate functions, respectively. If $X\subset \mathbb{R}^{2}$ and $\mathbf{% a}^{1},\mathbf{a}^{2} $ be linearly independent directions in $\mathbb{R}^{2} $, then the linear transformation $y_{1}=$ $\mathbf{a}^{1}\cdot \mathbf{x\,}$% , $y_{2}=$ $\mathbf{a}^{2}\cdot \mathbf{x}$ reduces the problems of proximinality of $\mathcal{R}_{b}(X)$ in $B(X)$ and $\mathcal{R}_{c}(X)$ in $% C(X)$ to the problems considered in that subsection. But in general, when $% X\subset \mathbb{R}^{n}$ and $n>2$, they cannot be reduced to those in \cite{76}. \bigskip \section{On the approximation by weighted ridge functions} In this section, we characterize the best $L_{2}$-approximation to a multivariate function by linear combinations of ridge functions multiplied by some fixed weight functions. In the special case, when the weight functions are constants, we obtain explicit formulas for both the best approximation and approximation error. \subsection{Problem statement} Ridge approximation in $L_{2}$ started to be actively studied in the late 90's by K.I. Oskolkov \cite{114}, V.E. Maiorov \cite{102}, A. Pinkus \cite{118}, V.N. Temlyakov \cite{138}, P. Petrushev \cite{116} and other researchers. Let $D$ be the unit disk in $\mathbb{R}^{2}$. In \cite{97}, Logan and Shepp along with other results gave a closed-form expression for the best $L_{2}$-approximation to a function $f \in L_{2}\left( D\right)$ from the set $\mathcal{R}\left( \mathbf{a}^{1},...,\mathbf{a}^{r}\right)$. Their solution requires that the directions $\mathbf{a}^{1},...,\mathbf{a}% ^{r}$ be equally-spaced and involves finite sums of convolutions with explicit kernels. In the $n$-dimensional case, we obtained an expression of simpler form for the best $L_{2}$-approximation to square-integrable multivariate functions over a certain domain, provided that $r=n$ and the directions $\mathbf{a}^{1},...,\mathbf{a}^{r}$ are linearly independent (see \cite{52}). In this section, we consider the approximation by functions from the following more general set \begin{equation*} \mathcal{R}\left( \mathbf{a}^{1},...,\mathbf{a}^{r};~w_{1},...,w_{r}\right) =\left\{ \sum\limits_{i=1}^{r}w_{i}(\mathbf{x})g_{i}\left( \mathbf{a}% ^{i}\cdot \mathbf{x}\right) :g_{i}:\mathbb{R}\rightarrow \mathbb{R}% ,~i=1,...,r\right\} , \end{equation*}% where $w_{1},...,w_{r}$ are fixed multivariate functions. We characterize the best $L_{2}$-approximation from this set in the case $% r\leq n.$ Then, in the special case when the weight functions $% w_{1},...,w_{r}$ are constants, we will prove two theorems on explicit formulas for the best approximation and the approximation error, respectively. At present, we do not yet know how to approach these problems in other possible cases of $r.$ \bigskip \subsection{Characterization of the best approximation} Let $X$ be a subset of $\mathbb{R}^{n}$ with a finite Lebesgue measure. Consider the approximation of a function $f\left( \mathbf{x}\right) =f\left( x_{1},...,x_{n}\right) $ in $L_{2}\left( X\right) $ by functions from the manifold $% \mathcal{R}\left( \mathbf{a}^{1},...,\mathbf{a}^{r};~w_{1},...,w_{r}\right) $% , where $r\leq n.$ We suppose that the functions $w_{i}(\mathbf{x})$ and the products $w_{i}(\mathbf{x})\cdot g_{i}\left( \mathbf{a}^{i}\cdot \mathbf{x}% \right) ,~i=1,...,r$, belong to the space $L_{2}\left( X\right) .$ Besides, we assume that the vectors $\mathbf{a}^{1},...,\mathbf{a}^{r}$ are linearly independent. We say that a function $g_{w}^{0}=\sum\limits_{i=1}^{r}w_{i}(% \mathbf{x})g_{i}^{0}\left( \mathbf{a}^{i}\cdot \mathbf{x}\right) $ in $% \mathcal{R}\left( \mathbf{a}^{1},...,\mathbf{a}^{r};~w_{1},...,w_{r}\right) $ is the best approximation (or extremal) to $f$ if \begin{equation*} \left\Vert f-g_{w}^{0}\right\Vert _{L_{2}\left( X\right) }=\inf\limits_{g\in \mathcal{R}\left( \mathbf{a}^{1},...,\mathbf{a}^{r};~w_{1},...,w_{r}\right) }\left\Vert f-g\right\Vert _{L_{2}\left( X\right) }. \end{equation*} Let the system of vectors $\{\mathbf{a}^{1},...,\mathbf{a}^{r},\mathbf{a% }^{r+1},...,\mathbf{a}^{n}\}$ be a completion of the system $\{\mathbf{a}% ^{1},...,\mathbf{a}^{r}\}$ to a basis in $\mathbb{R}^{n}.$\ Let $% J:X\rightarrow \mathbb{R}^{n}$ be the linear transformation given by the formulas \begin{equation*} y_{i}=\mathbf{a}^{i}\cdot \mathbf{x,}\quad \,i=1,...,n.\eqno(1.56) \end{equation*}% Since the vectors $\mathbf{a}^{i},$ $i=1,...,n$, are linearly independent, it is an injection. The Jacobian $\det J$ of this transformation is a constant different from zero. Let the formulas \begin{equation*} x_{i}=\mathbf{b}^{i}\cdot \mathbf{y},\;\;i=1,...,n, \end{equation*}% stand for the solution of linear equations (1.56) with respect to $% x_{i},\;i=1,...,n.$ Introduce the notation \begin{equation*} Y=J\left( X\right) \end{equation*}% and \begin{equation*} Y_{i}=\left\{ y_{i}\in \mathbb{R}:\;\;y_{i}=\mathbf{a}^{i}\cdot \mathbf{x}% ,\;\;\mathbf{x}\in X\right\} ,\,i=1,...,n. \end{equation*} For any function $u\in L_{2}\left( X\right) ,$ put \begin{equation*} u^{\ast }=u^{\ast }\left( \mathbf{y}\right) \overset{def}{=}u\left( \mathbf{b% }^{1}\cdot \mathbf{y},...,\mathbf{b}^{n}\cdot \mathbf{y}\right) . \end{equation*}% It is obvious that $u^{\ast }\in L_{2}\left( Y\right) .$ Besides, \begin{equation*} \int\limits_{Y}u^{\ast }\left( \mathbf{y}\right) d\mathbf{y}=\left\vert \det J\right\vert \cdot \int\limits_{X}u\left( \mathbf{x}\right) d\mathbf{x}\eqno% (1.57) \end{equation*}% and \begin{equation*} \left\Vert u^{\ast }\right\Vert _{L_{2}\left( Y\right) }=\left\vert \det J\right\vert ^{1/2}\cdot \left\Vert u\right\Vert _{L_{2}\left( X\right) }.% \eqno(1.58) \end{equation*}% Set \begin{equation*} L_{2}^{i}=\{w_{i}^{\ast }(\mathbf{y})g\left( y_{i}\right) \in L_{2}(Y)\},~i=1,...,r. \end{equation*} We need the following auxiliary lemmas. \bigskip \textbf{Lemma 1.5.} \textit{Let $f\left( \mathbf{x}\right) \in L_{2}\left( X\right) $. A function $\sum\limits_{i=1}^{r}w_{i}(\mathbf{x}% )g_{i}^{0}\left( \mathbf{a}^{i}\cdot \mathbf{x}\right) $ is extremal to the function $f\left( \mathbf{x}\right) $ if and only if \ $\sum% \limits_{i=1}^{r}w_{i}^{\ast }(\mathbf{y})g_{i}^{0}\left( y_{i}\right) $ is extremal from the space $L_{2}^{1}\mathit{\oplus }...\oplus L_{2}^{r}$ to the function $f^{\ast }\left( \mathbf{y}\right) $.} \bigskip Due to (1.58) the proof of this lemma is obvious. \bigskip \textbf{Lemma 1.6.} \textit{Let $f\left( \mathbf{x}\right) \in L_{2}\left( X\right) $. A function $\sum\limits_{i=1}^{r}w_{i}(\mathbf{x}% )g_{i}^{0}\left( \mathbf{a}^{i}\cdot \mathbf{x}\right) $ is extremal to the function $f\left( \mathbf{x}\right) $ if and only if} \begin{equation*} \int\limits_{X}\left( f\left( \mathbf{x}\right) -\sum\limits_{i=1}^{r}w_{i}(% \mathbf{x})g_{i}^{0}\left( \mathbf{a}^{i}\cdot \mathbf{x}\right) \right) w_{j}(\mathbf{x})h\left( \mathbf{a}^{j}\cdot \mathbf{x}\right) d\mathbf{x}% =0\ \end{equation*}% \textit{for any ridge function $h\left( \mathbf{a}^{j}\cdot \mathbf{x}% \right) $ such that $\mathit{w}_{j}\mathit{(x)h}\left( \mathbf{a}^{j}\cdot \mathbf{x}\right) $$\in L_{2}\left( X\right)$, $j=1,...,r$.} \bigskip \textbf{Lemma 1.7.} \textit{The following formula is valid for the error of approximation to a function $f\left( \mathbf{x}\right) $ in $L_{2}\left( X\right) $ from $\mathcal{R}\left( \mathbf{a}^{1},...,\mathbf{a}% ^{r};~w_{1},...,w_{r}\right) $:} \begin{equation*} E\left( f\right) =\left( \left\Vert f\left( \mathbf{x}\right) \right\Vert _{L_{2}\left( X\right) }^{2}-\left\Vert \sum\limits_{i=1}^{r}w_{i}(\mathbf{x}% )g_{i}^{0}\left( \mathbf{a}^{i}\cdot \mathbf{x}\right) \right\Vert _{L_{2}\left( X\right) }^{2}\right) ^{\frac{1}{2}}, \end{equation*}% \textit{where $\sum\limits_{i=1}^{r}w_{i}(\mathbf{x})g_{i}^{0}\left( \mathbf{% a}^{i}\cdot \mathbf{x}\right) $ is the best approximation to $f\left( \mathbf{x}\right) $.} \bigskip Lemmas 1.6 and 1.7 follow from the well-known facts of functional analysis that the best approximation of an element $x$ in a Hilbert space $H$ from a linear subspace $Z$ of $H$ must be the image of $x$ via the orthogonal projection onto $Z$ and the sum of squares of norms of orthogonal vectors is equal to the square of the norm of their sum. We say that $Y$ is an \textit{$r$-set} if it can be represented as $Y_{1}\times ...\times Y_{r}\times Y_{0},$ where $Y_{0}$ is some set from the space $% \mathbb{R}^{n-r}.$ In a special case, $Y_{0}$ may be equal to $Y_{r+1}\times ...\times Y_{n},$ but it is not necessary. By $Y^{\left( i\right) },$ we denote the Cartesian product of the sets $Y_{1},...,Y_{r},Y_{0}$ except for $% Y_{i},\;i=1,...,r$. That is, $Y^{\left( i\right) }=Y_{1}\times ...\times Y_{i-1}\times Y_{i+1}\times ...\times Y_{r}\times Y_{0},\,\ i=1,...,r$. \bigskip \textbf{Theorem 1.12.} \textit{Let $Y$ be an $r$-set. A function $\sum\limits_{i=1}^{r}w_{i}(\mathbf{x})g_{i}^{0}\left( \mathbf{a}^{i}\cdot \mathbf{x}\right) $ is the best approximation to $f(\mathbf{x)}$ if and only if} \begin{equation*} g_{j}^{0}\left( y_{j}\right) =\frac{1}{\int\limits_{Y^{\left( j\right) }}w_{j}^{\ast 2}(\mathbf{y})d\mathbf{y}^{\left( j\right) }}% \int\limits_{Y^{\left( j\right) }}\left( f^{\ast }\left( \mathbf{y}\right) -\sum\limits_{\substack{ i=1 \\ i\neq j}}^{r}w_{i}^{\ast }(\mathbf{y}% )g_{i}^{0}\left( y_{i}\right) \right) w_{j}^{\ast }(\mathbf{y})d\mathbf{y}% ^{\left( j\right) },\eqno(1.59) \end{equation*}% \textit{for $j=1,...,r$.} \bigskip \begin{proof} \textit{Necessity.} Let a function $\sum\limits_{i=1}^{r}w_{i}(% \mathbf{x})g_{i}^{0}\left( \mathbf{a}^{i}\cdot \mathbf{x}\right) $ be extremal to $f$. Then by Lemma 1.5, the function $\sum% \limits_{i=1}^{r}w_{i}^{\ast }(\mathbf{y})g_{i}^{0}\left( y_{i}\right) $ in $% L_{2}^{1}\oplus ...\oplus L_{2}^{r}$ is extremal to $f^{\ast }$. By Lemma 1.6 and equality (1.57), \begin{equation*} \int\limits_{Y}f^{\ast }\left( \mathbf{y}\right) w_{j}^{\ast }(\mathbf{y}% )h\left( y_{j}\right) d\mathbf{y}=\int\limits_{Y}w_{j}^{\ast }(\mathbf{y}% )h\left( y_{j}\right) \sum\limits_{i=1}^{r}w_{i}^{\ast }(\mathbf{y}% )g_{i}^{0}\left( y_{i}\right) d\mathbf{y}\eqno(1.60) \end{equation*}% for any product $w_{j}^{\ast }(\mathbf{y})h\left( y_{j}\right) $ in $% L_{2}^{j},\;\;j=1,...,r$. Applying Fubini's theorem to the integrals in (1.60), we obtain that \begin{eqnarray*} &&\int\limits_{Y_{j}}h\left( y_{j}\right) \left[ \int\limits_{Y^{\left( j\right) }}f^{\ast }\left( \mathbf{y}\right) w_{j}^{\ast }(\mathbf{y})d% \mathbf{y}^{\left( j\right) }\right] dy_{j} \\ &=&\int\limits_{Y_{j}}h\left( y_{j}\right) \left[ \int\limits_{Y^{\left( j\right) }}w_{j}^{\ast }(\mathbf{y})\sum\limits_{i=1}^{r}w_{i}^{\ast }(% \mathbf{y})g_{i}^{0}\left( y_{i}\right) d\mathbf{y}^{\left( j\right) }\right] dy_{j}. \end{eqnarray*}% Since $h\left( y_{j}\right) $ is an arbitrary function such that $% w_{j}^{\ast }(\mathbf{y})h\left( y_{j}\right) \in L_{2}^{j}$, \begin{equation*} \int\limits_{Y^{\left( j\right) }}f^{\ast }\left( \mathbf{y}\right) w_{j}^{\ast }(\mathbf{y})d\mathbf{y}^{(j)}=\int\limits_{Y^{\left( j\right) }}w_{j}^{\ast }(\mathbf{y})\sum\limits_{i=1}^{r}w_{i}^{\ast }(\mathbf{y}% )g_{i}^{0}\left( y_{i}\right) d\mathbf{y}^{\left( j\right) },\;\;j=1,...,r. \end{equation*}% Therefore, \begin{equation*} \int\limits_{Y^{\left( j\right) }}w_{j}^{\ast 2}(\mathbf{y})g_{j}^{0}\left( {% y_{j}}\right) d\mathbf{y}^{\left( j\right) }=\int\limits_{Y^{\left( j\right) }}\left( f^{\ast }\left( \mathbf{y}\right) -\sum\limits_{\substack{ i=1 \\ % i\neq j}}^{r}w_{i}^{\ast }(\mathbf{y})g_{i}^{0}\left( y_{i}\right) \right) w_{j}^{\ast }(\mathbf{y})d\mathbf{y}^{\left( j\right) }, \end{equation*}% for $j=1,...,r.$ Now, since $y_{j}\notin Y^{\left( j\right) }$, we obtain (1.59). \textit{Sufficiency.} Note that all the equalities in the proof of the necessity can be obtained in the reverse order. Thus, (1.60) can be obtained from (1.59). Then by (1.57) and Lemma 1.6, we finally conclude that the function $% \sum\limits_{i=1}^{r}w_{i}(\mathbf{x})g_{i}^{0}\left( \mathbf{a}^{i}\cdot \mathbf{x}\right) $ is extremal to $f\left( \mathbf{x}\right) $. \end{proof} In the following, $\left\vert Q\right\vert $ will denote the Lebesgue measure of a measurable set $Q.$ The following corollary is obvious. \bigskip \textbf{Corollary 1.6.} \textit{Let $Y$ be an }$r$\textit{-set. A function $\sum\limits_{i=1}^{r}g_{i}^{0}\left( \mathbf{a}^{i}\cdot \mathbf{x}% \right) $ in $\mathcal{R}\left( \mathbf{a}^{1},...,\mathbf{a}^{r}\right) $ is the best approximation to $f(\mathbf{x)}$ if and only if} \begin{equation*} g_{j}^{0}\left( y_{j}\right) =\frac{1}{\left\vert Y^{\left( j\right) }\right\vert }\int\limits_{Y^{\left( j\right) }}\left( f^{\ast }\left( \mathbf{y}\right) -\sum\limits_{\substack{ i=1 \\ i\neq j}}% ^{r}g_{i}^{0}\left( y_{i}\right) \right) d\mathbf{y}^{\left( j\right) },\;\;j=1,...,r. \end{equation*}% \bigskip In \cite{52}, this corollary was proven for the case $r=n.$ \bigskip \subsection{Formulas for the best approximation and approximation error} In this section, we establish explicit formulas for both the best approximation and approximation error, provided that the weight functions are constants. In this case, since we vary over $g_{i},$ the set $% \mathcal{R}\left( \mathbf{a}^{1},...,\mathbf{a}^{r};~w_{1},...,w_{r}\right) $ coincides with $\mathcal{R}\left( \mathbf{a}^{1},...,\mathbf{a}^{r}\right) .$ Thus, without loss of generality, we may assume that $w_{i}(\mathbf{x})=1$ for $i=1,...,r.$ For brevity of the further exposition, introduce the notation \begin{equation*} A=\int\limits_{Y}f^{\ast }\left( \mathbf{y}\right) d\mathbf{y}\text{ and \ }% f_{i}^{\ast }=f_{i}^{\ast }(y_{i})=\int\limits_{Y^{\left( i\right) }}f^{\ast }\left( \mathbf{y}\right) d\mathbf{y}^{\left( i\right) },~i=1,...,r. \end{equation*} The following theorem is a generalization of the main result of \cite% {52} from the case $r=n$ to the cases $r<n.$ \bigskip \textbf{Theorem 1.13.} \textit{Let $Y$ be an }$r$\textit{-set. Set the functions} \begin{equation*} g_{1}^{0}\left( y_{1}\right) =\frac{1}{\left\vert Y^{\left( 1\right) }\right\vert }f_{1}^{\ast }-\left( r-1\right) \frac{A}{\left\vert Y\right\vert } \end{equation*}% \textit{and} \begin{equation*} g_{j}^{0}\left( y_{j}\right) =\frac{1}{\left\vert Y^{\left( j\right) }\right\vert }f_{j}^{\ast },\;j=2,...,r. \end{equation*}% \textit{Then the function $\sum\limits_{i=1}^{r}g_{i}^{0}\left( \mathbf{a}% ^{i}\cdot \mathbf{x}\right) $ is the best approximation from $\mathcal{R}% \left( \mathbf{a}^{1},...,\mathbf{a}^{r}\right) $ to $f\left( \mathbf{x}% \right) $.} \bigskip The proof is simple. It is sufficient to verify that the functions $g_{j}^{0}\left( y_{j}\right) ,\;j=1,...,r$, satisfy the conditions of Corollary 1.6. This becomes obvious if note that \begin{equation*} \sum\limits_{\underset{i\neq j}{i=1}}^{r}\frac{1}{\left\vert Y^{\left( j\right) }\right\vert }\frac{1}{\left\vert Y^{\left( i\right) }\right\vert }% \int\limits_{Y^{\left( j\right) }}\left[ \int\limits_{Y^{\left( i\right) }}f^{\ast }\left( \mathbf{y}\right) d\mathbf{y}^{\left( i\right) }\right] d% \mathbf{y}^{\left( j\right) }=\left( r-1\right) \frac{1}{\left\vert Y\right\vert }\int\limits_{Y}f^{\ast }\left( \mathbf{y}\right) d\mathbf{y} \end{equation*}% for $j=1,...,r$. \bigskip \textbf{Theorem 1.14.} \textit{Let $Y$ be an $r$-set. Then the error of approximation to a function $f(x)$ from the set $\mathcal{R}\left( \mathbf{a}^{1},...,\mathbf{a}^{r}\right) $ can be calculated by the formula} \begin{equation*} E(f)=\left\vert \det J\right\vert ^{-1/2}\left( \left\Vert f^{\ast }\right\Vert _{L_{2}(Y)}^{2}-\sum_{i=1}^{r}\frac{1}{\left\vert Y^{\left( i\right) }\right\vert ^{2}}\left\Vert f_{i}^{\ast }\right\Vert _{L_{2}(Y)}^{2}+(r-1)\frac{A^{2}}{\left\vert Y\right\vert }\right) ^{1/2}. \end{equation*} \bigskip \begin{proof} From Eq. (1.58), Lemma 1.7 and Theorem 1.13, it follows that \begin{equation*} E(f)=\left\vert \det J\right\vert ^{-1/2}\left( \left\Vert f^{\ast }\right\Vert _{L_{2}(Y)}^{2}-I\right) ^{1/2},\eqno(1.61) \end{equation*}% where \begin{equation*} I=\left\Vert \sum_{i=1}^{r}\frac{1}{\left\vert Y^{\left( i\right) }\right\vert }f_{i}^{\ast }-(r-1)\frac{A}{\left\vert Y\right\vert }% \right\Vert _{L_{2}(Y)}^{2}. \end{equation*}% The integral $I$ can be written as a sum of the following four integrals: \begin{eqnarray*} I_{1} &=&\sum_{i=1}^{r}\frac{1}{\left\vert Y^{\left( i\right) }\right\vert ^{2}}\left\Vert f_{i}^{\ast }\right\Vert _{L_{2}(Y)}^{2},~I_{2}=\sum_{i=1}^{r}\sum\limits_{\substack{ j=1 \\ j\neq i }}^{r}\frac{1}{\left\vert Y^{\left( i\right) }\right\vert }\frac{1}{% \left\vert Y^{\left( j\right) }\right\vert }\int\limits_{Y}f_{i}^{\ast }f_{j}^{\ast }d\mathbf{y,} \\ I_{3} &=&-2(r-1)\frac{1}{\left\vert Y\right\vert }A\sum_{i=1}^{r}\frac{1}{% \left\vert Y^{\left( i\right) }\right\vert }\int\limits_{Y}f_{i}^{\ast }d% \mathbf{y,}~I_{4}=(r-1)^{2}\frac{A^{2}}{\left\vert Y\right\vert }. \end{eqnarray*}% \qquad \qquad It is not difficult to verify that \begin{equation*} \int\limits_{Y}f_{i}^{\ast }f_{j}^{\ast }d\mathbf{y=}\left\vert Y_{0}\times \prod\limits_{\substack{ k=1 \\ k\neq i,j}}^{r}Y_{k}\right\vert A^{2},\text{ for }i,j=1,...,r,~i\neq j,\eqno(1.62) \end{equation*}% and \begin{equation*} \int\limits_{Y}f_{i}^{\ast }d\mathbf{y}=\left\vert Y_{0}\times \prod\limits _{\substack{ k=1 \\ k\neq i}}^{r}Y_{k}\right\vert A,\text{ for }i=1,...,r.% \eqno(1.63) \end{equation*} Considering (1.62) and (1.63) in the expressions of $I_{2}$ and $I_{3}$ respectively, we obtain that \begin{equation*} I_{2}=r(r-1)\frac{A^{2}}{\left\vert Y\right\vert }\text{ and }I_{3}=-2r(r-1)% \frac{A^{2}}{\left\vert Y\right\vert }. \end{equation*}% Therefore, \begin{equation*} I=I_{1}+I_{2}+I_{3}+I_{4}=\sum_{i=1}^{r}\frac{1}{\left\vert Y^{\left( i\right) }\right\vert ^{2}}\left\Vert f_{i}^{\ast }\right\Vert _{L_{2}(Y)}^{2}-(r-1)\frac{A^{2}}{\left\vert Y\right\vert }. \end{equation*}% Now the last equality together with (1.61) complete the proof. \end{proof} \textbf{Example.} Consider the following set \begin{equation*} X=\{\mathbf{x}\in \mathbb{R}^{4}:y_{i}=y_{i}(\mathbf{x})\in \lbrack 0;1],~i=1,...,4\}, \end{equation*}% where \begin{equation*} \left\{ \begin{array}{c} y_{1}=x_{1}+x_{2}+x_{3}-x_{4} \\ y_{2}=x_{1}+x_{2}-x_{3}+x_{4} \\ y_{3}=x_{1}-x_{2}+x_{3}+x_{4} \\ y_{4}=-x_{1}+x_{2}+x_{3}+x_{4}% \end{array}% \right. \eqno(1.64) \end{equation*}% Let the function \begin{equation*} f=8x_{1}x_{2}x_{3}x_{4}-\sum_{i=1}^{4}x_{i}^{4}+2\sum_{i=1}^{3}% \sum_{j=i+1}^{4}x_{i}^{2}x_{j}^{2} \end{equation*}% be given on $X.$ Consider the approximation of this function by functions from $% \mathcal{R}\left( \mathbf{a}^{1},\mathbf{a}^{2},\mathbf{a}^{3}\right) ,% \mathcal{\ }$where $\mathbf{a}^{1}=(1;1;1;-1),~\mathbf{a}^{2}=(1;1;-1;1),~% \mathbf{a}^{3}=(1;-1;1;1).$ Putting $\mathbf{a}^{4}=(-1;1;1;1),$ we complete the system of vectors $\mathbf{a}^{1},\mathbf{a}^{2},\mathbf{a}^{3}$ to the basis $\{\mathbf{a}^{1},\mathbf{a}^{2},\mathbf{a}^{3},\mathbf{a}^{4}\}$ in $% \mathbb{R}^{4}.$ The linear transformation $J$ defined by (1.64) maps the set $X$ onto the set $Y=[0;1]^{4}.$ The inverse transformation is given by the formulas \begin{equation*} \left\{ \begin{array}{c} x_{1}=\frac{1}{4}y_{1}+\frac{1}{4}y_{2}+\frac{1}{4}y_{3}-\frac{1}{4}y_{4} \\ x_{2}=\frac{1}{4}y_{1}+\frac{1}{4}y_{2}-\frac{1}{4}y_{3}+\frac{1}{4}y_{4} \\ x_{3}=\frac{1}{4}y_{1}-\frac{1}{4}y_{2}+\frac{1}{4}y_{3}+\frac{1}{4}y_{4} \\ x_{4}=-\frac{1}{4}y_{1}+\frac{1}{4}y_{2}+\frac{1}{4}y_{3}+\frac{1}{4}y_{4}% \end{array}% \right. \end{equation*}% It can be easily verified that $f^{\ast }=y_{1}y_{2}y_{3}y_{4}$ and $Y$ is a $3$-set with $Y_{i}=[0;1],$ $i=1,2,3.$ Besides, $Y_{0}=[0;1].$ After easy calculations we obtain that $A=\allowbreak \frac{1}{16};~$\ $% f_{i}^{\ast }=\allowbreak \frac{1}{8}y_{i}$ for $i=1,2,3;$ $\det J=-16;$ $% \left\Vert f^{\ast }\right\Vert _{L_{2}(Y)}^{2}=\frac{1}{81};$ $\left\Vert f_{i}^{\ast }\right\Vert _{L_{2}(Y)}^{2}=\frac{1}{192},$ $i=1,2,3.$ Now from Theorems 1.13 and 1.14 it follows that the function $\frac{1}{8}% \sum_{i=1}^{3}\left( \mathbf{a}^{i}\cdot \mathbf{x}\right) -\allowbreak \frac{1}{8}$ is the best approximation from $\mathcal{R}\left( \mathbf{a}^{1},% \mathbf{a}^{2},\mathbf{a}^{3}\right) $ to $f$ and $E(f)=\frac{1}{576}\sqrt{2}% \sqrt{47}.$ \bigskip \textbf{Remark 1.4.} Most of the material in this chapter is to be found in \cite{52,53,54,49,50,47,66,64}. \newpage \chapter{The smoothness problem in ridge function representation} This chapter discusses the following open problem raised in Buhmann and Pinkus \cite{12}, and Pinkus \cite[p. 14]{117}. Assume we are given a function $f(\mathbf{x})=f(x_{1},...,x_{n})$ of the form \begin{equation*} f(\mathbf{x})=\sum_{i=1}^{k}f_{i}(\mathbf{a}^{i}\cdot \mathbf{x}),\eqno(2.1) \end{equation*}% where the $\mathbf{a}^{i},$ $i=1,...,k,$ are pairwise linearly independent vectors (directions) in $\mathbb{R}^{n}$, $f_{i}$ are arbitrarily behaved univariate functions and $\mathbf{a}^{i}\cdot \mathbf{x}$ are standard inner products. Assume, in addition, that $f$ is of a certain smoothness class, that is, $f\in C^{s}(\mathbb{R}^{n})$, where $s\geq 0$ (with the convention that $C^{0}(\mathbb{R}^{n})=C(\mathbb{R}^{n})$). Is it true that there will always exist $g_{i}\in C^{s}(\mathbb{R})$ such that \begin{equation*} f(\mathbf{x})=\sum_{i=1}^{k}g_{i}(\mathbf{a}^{i}\cdot \mathbf{x})\text{ ?}% \eqno(2.2) \end{equation*} In this chapter, we solve this problem up to some multivariate polynomial. In the special case $n=2$, we see that this multivariate polynomial can be written as a sum of polynomial ridge functions with the given directions $% \mathbf{a}^{i}$. In addition, we find various conditions on the directions $% \mathbf{a}^{i}$ guaranteeing a positive solution to the problem. We also consider the question on constructing $g_{i}$ using the information about the known functions $f_{i}$. Most of the material of this chapter may be found in \cite{2,1,A2,A1,120}. \bigskip \section{A solution to the problem up to a multivariate polynomial} In this section, we solve the above problem up to a multivariate polynomial. That is, we show that if (2.1) holds for $f\in C^{s}(\mathbb{R}^{n})$ and arbitrarily behaved $f_{i}$, then there exist $g_{i}\in C^{s}(\mathbb{R})$ such that \begin{equation*} f(\mathbf{x})=\sum_{i=1}^{k}g_{i}(\mathbf{a}^{i}\cdot \mathbf{x})+P(\mathbf{x% }), \end{equation*}% where $P(\mathbf{x})$ is a polynomial of degree at most $k-1$. In the special case $n=2$, we see that this multivariate polynomial can be written as a sum of polynomial ridge functions with the given directions $\mathbf{a}% ^{i}$ and thus (2.2) holds with $g_{i}\in C^{s}(\mathbb{R})$. \subsection{A brief overview of some results} We start this subsection with the simple observation that for $k=1$ and $k=2$ the smoothness problem is easily solved. Indeed for $k=1$ by choosing $% \mathbf{c}\in \mathbb{R}^{n}$ satisfying $\mathbf{a}^{1}\cdot \mathbf{c}=1$, we have that $f_{1}(t)=f(t\mathbf{c)}$ is in $C^{s}(\mathbb{R})$. The same argument can be carried out for the case $k=2.$ In this case, since the vectors $\mathbf{a}^{1}$ and $\mathbf{a}^{2}$ are linearly independent, there exists a vector $\mathbf{c}\in \mathbb{R}^{n}$ satisfying $\mathbf{a}% ^{1}\cdot \mathbf{c}=1$ and $\mathbf{a}^{2}\cdot \mathbf{c}=0.$ Therefore, we obtain that the function $f_{1}(t)=f(t\mathbf{c)}-f_{2}(0)$ is in the class $C^{s}(\mathbb{R})$. Similarly, one can verify that $f_{2}\in C^{s}(% \mathbb{R})$. The above cases with one and two ridge functions in (2.1) show that the functions $f_{i}$ inherit smoothness properties of the given $f$. The picture is absolutely different if the number of directions $k\geq 3$. For $% k=3$, there are ultimately smooth functions which decompose into sums of very badly behaved ridge functions. This phenomena comes from the classical Cauchy Functional Equation (CFE). This equation,% \begin{equation*} h(x+y)=h(x)+h(y),\text{ }h:\mathbb{R\rightarrow R},\eqno(2.3) \end{equation*}% looks very simple and has a class of simple solutions $h(x)=cx,$ $c\in \mathbb{R}$. However, it easily follows from Hamel basis theory that CFE also has a large class of wild solutions. These solutions are called \textquotedblleft wild" because they are extremely pathological. They are, for example, not continuous at a point, not monotone on an interval, not bounded on any set of positive measure (see, e.g., \cite{A}). Let $h_{1}$ be any wild solution of the equation (2.3). Then the zero function can be represented as% \begin{equation*} 0=h_{1}(x)+h_{1}(y)-h_{1}(x+y).\eqno(2.4) \end{equation*}% Note that the functions involved in (2.4) are bivariate ridge functions with the directions $(1,0)$, $(0,1)$ and $(1,1)$, respectively. This example shows that for $k\geq 3$ the functions $f_{i}$ in (2.1) may not inherit smoothness properties of the function $f$, which in the case of (2.4) is the identically zero function. Thus the above problem arises naturally. However, it was shown by some authors that, additional conditions on $f_{i}$ or the directions $\mathbf{a}^{i}$ guarantee smoothness of the representation (2.1). It was first proved by Buhmann and Pinkus \cite{12} that if in (2.1) $f\in C^{s}(\mathbb{R}^{n})$, $s\geq k-1$ and $f_{i}\in L_{loc}^{1}(\mathbb{R)}$ for each $i$, then $f_{i}\in C^{s}(\mathbb{R)}$ for $i=1,...,k.$ Later Pinkus \cite{120} found a strong relationship between CFE and the problem of smoothness in ridge function representation. He generalized extensively the previous result of Buhmann and Pinkus \cite{12}. He showed that the solution is quite simple and natural if the functions $% f_{i}$ are taken from a certain class $\mathcal{B}$ of real-valued functions defined on $\mathbb{R}$. $\mathcal{B}$ includes, for example, the set of continuous functions, the set of bounded functions, the set of Lebesgue measurable functions (for the precise definition of $\mathcal{B}$ see the next subsection). The result of Pinkus \cite{120} states that if in (1.1) $% f\in C^{s}(\mathbb{R}^{n})$ and each $f_{i}\in \mathcal{B}$, then necessarily $f_{i}\in C^{s}(\mathbb{R)}$ for $i=1,...,k$. Note that severe restrictions on the directions $\mathbf{a}^{i}$ also guarantee smoothness of the representation (2.1). For example, in (2.1) the inclusions $f_{i}\in C^{s}(\mathbb{R})$, $i=1,...,k,$ are automatically valid if the directions $\mathbf{a}^{i}$ are linearly independent and if these directions are not linearly independent, then there exists $f\in C^{s}(% \mathbb{R}^{n})$ of the form (2.1) such that the $f_{i}\notin C^{s}(\mathbb{R% }),$ $i=1,...,k$ (see \cite{86}). Indeed, if the directions $\mathbf{a}^{i}$ are linearly independent, then for each $i=1,...,k,$ we can choose a vector $% \mathbf{b}^{i}$ such that $\mathbf{b}^{i}\cdot \mathbf{a}^{i}=1,$ but at the same time $\mathbf{b}^{i}\cdot \mathbf{a}^{j}=0,$ for all $j=1,...,k,$ $% j\neq i$. Putting $\mathbf{x}=\mathbf{b}^{i}t$ in (2.1) yields that \begin{equation*} f(\mathbf{b}^{i}t)=f_{i}(t)+\sum_{j=1,j\neq i}^{k}f_{j}(0),\text{ }i=1,...,k. \end{equation*}% This shows that all the functions $f_{i}$ and $f$ belong to the same smoothness class. If the directions $\mathbf{a}^{i}$ are not linearly independent, then there exist numbers $\lambda _{1},...,\lambda _{k}$ such that $\sum_{i=1}^{k}\left\vert \lambda _{i}\right\vert >0$ and $% \sum_{i=1}^{k}\lambda _{i}\mathbf{a}^{i}=\mathbf{0}$. Let $h$ be any wild solution of CFE. Then it is not difficult to verify that \begin{equation*} 0=\sum_{i=1}^{k}h_{i}(\mathbf{a}^{i}\cdot \mathbf{x}), \end{equation*}% where $h_{i}(t)=h(\lambda _{i}t),$ $i=1,...,k.$ Note that in the last representation, the zero function is an ultimately smooth function, while all the functions $h_{i}$ are highly nonsmooth. The above result of Pinkus was a starting point for further research on continuous and smooth sums of ridge functions. Much work in this direction was done by Konyagin and Kuleshov \cite{86,K2}, and Kuleshov \cite{K4}. They mainly analyze the continuity of $f_{i}$, that is, the question of if and when continuity of $f$ guarantees the continuity of $f_{i}$. There are also other results concerning different properties, rather than continuity, of $% f_{i}$. Most results in \cite{86,K2,K4} involve certain subsets (convex open sets, convex bodies, etc.) of $\mathbb{R}^{n}$ instead of only $\mathbb{R}% ^{n}$ itself. In \cite{1}, Aliev and Ismailov gave a partial solution to the smoothness problem. Their solution comprises the cases in which $s\geq 2$ and $k-1$ directions of the given $k$ directions are linearly independent. Kuleshov \cite{88} generalized Aliev and Ismailov's result \cite[Theorem 2.3]% {1} to all possible cases of $s$. That is, he proved that if a function $% f\in C^{s}(\mathbb{R}^{n})$, where $s\geq 0$, is of the form (2.1) and $% (k-1) $-tuple of the given set of $k$ directions $\mathbf{a}^{i}$ forms a linearly independent system, then there exist $g_{i}\in C^{s}(\mathbb{R})$, $% i=1,...,k $, such that (2.2) holds (see \cite[Theorem 3]{88}). In Section 2.2 we give a new constructive proof of Kuleshov's result. \bigskip \subsection{A result of A. Pinkus} In \cite{120}, A. Pinkus considered the smoothness problem in ridge function representation. For a given function $f$ $:\mathbb{R}^{n}\rightarrow \mathbb{% R}$, he posed and partially answered the following question. If $f$ belongs to some smoothness class and (2.1) holds, what can we say about the smoothness of the functions $f_{i}$? He proved that for a large class of representing functions $f_{i}$, these $f_{i}$ are smooth. That is, if apriori we assume that in the representation (2.1) the functions $f_{i}$ is of a certain class of \textquotedblleft reasonably well behaved functions", then they have the same degree of smoothness as the function $f.$ As the mentioned class of \textquotedblleft reasonably well behaved functions" one may take, e.g., the set of functions that are continuous at a point, the set of Lebesgue measurable functions, etc. All these classes arise from the class $\mathcal{B}$ considered by Pinkus \cite{120} and the classical theory of CFE. In \cite{120}, $\mathcal{B}$ denotes any linear space of real-valued functions $u$ defined on $\mathbb{R}$, closed under translation, such that if there is a function $v\in C(\mathbb{R)}$ for which $u-v$ satisfies CFE, then $u-v$ is necessarily linear, i.e. $u(x)-v(x)=cx,$ for some constant $% c\in \mathbb{R}$. Such a definition of $\mathcal{B}$ is required in the proof of the following theorem. \bigskip \textbf{Theorem 2.1} (Pinkus \cite{120}). \textit{Assume $f\in C^{s}(\mathbb{% R}^{n})$ is of the form (2.1). Assume, in addition, that each $f_{i}\in \mathcal{B}$. Then necessarily $f_{i}\in C^{s}(\mathbb{R)}$ for $i=1,...,k.$} \bigskip \begin{proof} We prove this theorem by induction on $k.$ The result is valid when $k=1$. Indeed, taking any direction $\mathbf{c}$ such that $% \mathbf{a}^{1}\cdot \mathbf{c}=1$ and putting $x=\mathbf{c}t$ in (2.1), we obtain that $f_{1}(t)=f(\mathbf{c}t)\in C^{s}(\mathbb{R})$. Assume that the result is valid for $k-1.$ Let us show that it is valid for $k$. Chose any vector $\mathbf{e}\in \mathbb{R}^{n}$ satisfying $\mathbf{e\cdot a}% ^{k}=0$ and $\mathbf{e\cdot a}^{i}=b_{i}\neq 0$, for $i=1,...,k-1.$ Clearly, there exists a vector with this property. The property of $\mathbf{e}$ enables us to write that \begin{equation*} f(\mathbf{x+e}t)-f(\mathbf{x)=}\sum_{i=1}^{k-1}f_{i}(\mathbf{a}^{i}\cdot \mathbf{x}+b_{i}t)-f_{i}(\mathbf{a}^{i}\cdot \mathbf{x}). \end{equation*}% Thus% \begin{equation*} F(\mathbf{x}):=f(\mathbf{x+e}t)-f(\mathbf{x})=\sum_{i=1}^{k-1}h_{i}(\mathbf{a% }^{i}\cdot \mathbf{x}), \end{equation*}% where% \begin{equation*} h_{i}(y)=f_{i}(y+b_{i}t)-f_{i}(y)\text{, }i=1,...,k-1. \end{equation*}% Since $f_{i}\in \mathcal{B}$ and $\mathcal{B}$ is translation invariant, $% h_{i}\in \mathcal{B}$. In addition, since $F\in C^{s}(\mathbb{R}^{n})$, it follows by our induction assumption that $h_{i}\in C^{s}(\mathbb{R})$. Note that this inclusion is valid for all $t\in \mathbb{R}$. In \cite{B1}, de Bruijn proved that if for any $c\in \mathbb{R}$ the difference $u(y+c)-u(y)$ ($u$ is any real function on $\mathbb{R}$) belongs to the class $C^{s}(\mathbb{R})$, then $u$ is necessarily of the form $u=v+r$% , where $v\in $ $C^{s}(\mathbb{R})$ and $r$ satisfies CFE. Thus each function $f_{i}$ is of the form $f_{i}=v_{i}+r_{i}$, where $v_{i}\in $ $% C^{s}(\mathbb{R})$ and $r_{i}$ satisfies CFE. By our assumption, each $f_{i}$ is in $\mathcal{B}$, and from the definition of $\mathcal{B}$ it follows that $r_{i}=f_{i}-v_{i}$ is a linear function. Thus $f_{i}=v_{i}+r_{i}$, where both $v_{i},r_{i}\in $ $C^{s}(\mathbb{R})$, implying that $f_{i}\in C^{s}(\mathbb{R})$. This is valid for $i=1,...,k-1$, and hence also for $i=k$. \end{proof} \textbf{Remark 2.1. }In de Bruijn \cite{B1,B2}, there are delineated various classes of real-valued functions $\mathcal{D}$ with the property that if $% \bigtriangleup _{t}f=f(\cdot +t)-f(\cdot )\in \mathcal{D}$ for all $t\in \mathbb{R}$, then $f-s\in \mathcal{D}$, for some $s$ satisfying CFE (for such classes see the next subsection). Some translation invariant classes among them are $C^{\infty }(\mathbb{R})$ functions; analytic functions; algebraic polynomials; trigonometric polynomials. Theorem 2.1 can be suitably restated for any of these classes. \bigskip \subsection{Polynomial functions of $k$-th order} Given $h_{1},...,h_{k}\in \mathbb{R}$, we define inductively the difference operator $\Delta _{h_{1}...h_{k}}$ as follows \begin{eqnarray*} \Delta _{h_{1}}f(x) &:&=f(x+h_{1})-f(x), \\ \Delta _{h_{1}...h_{k}}f &:&=\Delta _{h_{k}}(\Delta _{h_{1}...h_{k-1}}f),% \text{ }f:\mathbb{R\rightarrow R}. \end{eqnarray*}% If $h_{1}=\cdots=h_{k}=h,$ then we write briefly $\Delta _{h}^{k}f$ instead of $\Delta _{\underset{n\text{ times}}{\underbrace{h...h}}}f$. For various properties of difference operators see \cite[Section 15.1]{Kuc}. \bigskip \textbf{Definition 2.1 }(see \cite{Kuc}). \textit{A function $f:\mathbb{R\rightarrow R}$ is called a polynomial function of order $k$ ($k\in \mathbb{N}$% ) if for every $x\in \mathbb{R}$ and $h\in \mathbb{R}$ we have} \begin{equation*} \Delta _{h}^{k+1}f(x)=0. \end{equation*} It can be shown that if $\Delta _{h}^{k+1}f=0$ for any $h\in \mathbb{R}$, then $\Delta _{h_{1}...h_{k+1}}f=0$ for any $h_{1},...,h_{k+1}\in \mathbb{R}$ (see \cite[Theorem 15.3.3]{Kuc}). A polynomial of degree at most $k$ is a polynomial function of order $k$ (see \cite[Theorem 15.9.4]{Kuc}). The polynomial functions generalize ordinary polynomials, and reduce to the latter under mild regularity assumptions. For example, if a polynomial function is continuous at one point, or bounded on a set of positive measure, then it continuous at all points (see \cite{Cies, Kurepa}), and therefore is a polynomial of degree $k$ (see \cite[Theorem 15.9.4]{Kuc}). Basic results concerning polynomial functions are due to S. Mazur-W. Orlicz \cite{Maz}, McKiernan \cite{Mc}, Djokovi\'{c} \cite{Djok}. The following theorem, which we will use in the sequel, yield implicitly the general construction of polynomial functions. \bigskip \textbf{Theorem 2.2 }(see \cite[Theorems 15.9.1 and 15.9.2]{Kuc}). \textit{A function $f:\mathbb{R\rightarrow R}$ is a polynomial function of order $k$ if and only if it admits a representation} \begin{equation*} f=f_{0}+f_{1}+...+f_{k}, \end{equation*}% \textit{where $f_{0}$ is a constant and $f_{j}:\mathbb{R\rightarrow R}$, $% j=1,...,k$, are diagonalizations of $j$-additive symmetric functions $F_{j}:% \mathbb{R}^{j}\mathbb{\rightarrow R}$, i.e.,} \begin{equation*} f_{j}(x)=F_{j}(x,...,x). \end{equation*} \bigskip Note that a function $F_{p}:\mathbb{R}^{p}\mathbb{\rightarrow R}$ is called $% p$-additive if for every $j,$ $1\leq j\leq p,$ and for every $% x_{1},...,x_{p},y_{j}\in \mathbb{R}$ \begin{equation*} F(x_{1},...,x_{j}+y_{j},...,x_{p})=F(x_{1},...,x_{p})+F(x_{1},...,x_{j-1},y_{j},x_{j+1},...,x_{p}), \end{equation*}% i.e., $F$ is additive in each of its variables $x_{j}$ (see \cite[p.363]% {Kuc}). A simple example of a $p$-additive function is given by the product \begin{equation*} f_{1}(x_{1})\times\cdots\times f_{p}(x_{p}), \end{equation*}% where the univariate functions $f_{j},$ $j=1,...,p$, are additive. Following de Bruijn, we say that a class $\mathcal{D}$ of real functions has the \textit{difference property} if any function $f:\mathbb{R\rightarrow R}$ such that $\bigtriangleup _{h}f\in \mathcal{D}$ for all $h\in \mathbb{R}$, admits a decomposition $f=g+S$, where $g\in \mathcal{D}$ and $S$ satisfies the Cauchy Functional Equation (2.3). Several classes with the difference property are investigated in de Bruijn \cite{B1,B2}. Some of these classes are: \smallskip 1) $C(\mathbb{R)}$, continuous functions; 2) $C^{s}(\mathbb{R)}$, functions with continuous derivatives up to order $s$% ; 3) $C^{\infty }(\mathbb{R)}$, infinitely differentiable functions; 4) analytic functions; 5) functions which are absolutely continuous on any finite interval; 6) functions having bounded variation over any finite interval; 7) algebraic polynomials; 8) trigonometric polynomials; 9) Riemann integrable functions. \smallskip A natural generalization of classes with the difference property are classes of functions with the difference property of $k$-th order. \bigskip \textbf{Definition 2.2 }(see \cite{Gajda}). \textit{A class $\mathcal{F}$ is said to have the difference property of $k$-th order if any function $f:\mathbb{R\rightarrow R}$ such that $\bigtriangleup _{h}^{k}f\in \mathcal{F}$ for all $h\in \mathbb{R}$, admits a decomposition $f=g+H$, where $g\in \mathcal{F}$ and $H$ is a polynomial function of $k$-th order.} \bigskip It is not difficult to see that the class $\mathcal{F}$ has the difference property of first order if and only if it has the difference property in de Bruijn's sense. There arises a natural question: which of the above classes have difference properties of higher orders? Gajda \cite{Gajda} considered this question in its general form, for functions defined on a locally compact Abelian group and showed that for any $k\in \mathbb{N}$, continuous functions have the difference property of $k$-th order (see \cite[Theorem 4]% {Gajda}). The proof of this result is based on several lemmas, in particular, on the following lemma, which we will also use in the sequel. \bigskip \textbf{Lemma 2.1.} (see \cite[Lemma 5]{Gajda}). \textit{For each $k\in \mathbb{N}$ the class of all continuous functions defined on $\mathbb{R}$ has the difference property of $k$-th order.} \bigskip In fact, Gajda \cite{Gajda} proved this lemma for Banach space valued functions, but the simplest case with the space $\mathbb{R}$ has all the difficulties. Unfortunately, the proof of the lemma has an essential gap. The author of \cite{Gajda} tried to reduce the proof to $\mod1$ periodic functions, but made a mistake in proving the continuity of the difference $% \Delta _{h_{1}...h_{k-1}}(f-f^{\ast })$. Here $f^{\ast }:\mathbb{% R\rightarrow R}$ is a $\mod1$ periodic function defined on the interval $% [0,1)$ as $f^{\ast }(x)=f(x)$ and extended to the whole $\mathbb{R}$ with the period $1$. That is, $f^{\ast }(x)=f(x)$ for $x\in \lbrack 0,1)$ and $% f^{\ast }(x+1)=f^{\ast }(x)$ for $x\in \mathbb{R}$. In the proof, the author of \cite{Gajda} takes a point $x\in \lbrack m,m+1)$ and writes that \begin{equation*} \Delta _{h_{1}...h_{k-1}}(f-f^{\ast })(x)=\Delta _{h_{1}...h_{k-1}}(f(x)-f(x-m)), \end{equation*}% which is not valid. Even though $f^{\ast }(x)=f(x-m)$ for any $x\in \lbrack m,m+1)$, the differences $\Delta _{h_{1}...h_{k-1}}f^{\ast }(x)$ and $\Delta _{h_{1}...h_{k-1}}f(x-m)$ are completely different, since the latter may involve values of $f$ at points outside $[0,1)$, which have no relationship with the definition of $f^{\ast }$. In the next section, we give a new proof for Lemma 2.1 (see Theorem 2.3 below). We hope that our proof is free from mathematical errors and thus the above lemma itself is valid. \bigskip \subsection{Some auxiliary results on polynomial functions} In this section, we do further research on polynomial functions and prove some auxiliary results. \bigskip \textbf{Lemma 2.2.} \textit{If $f:\mathbb{R\rightarrow R}$ is a polynomial function of order $k$, then for any $p\in $ $\mathbb{N}$ and any fixed $\xi _{1},...,\xi _{p}\in \mathbb{R}$, the function}% \begin{equation*} g(x_{1},...,x_{p})=f(\xi _{1}x_{1}+\cdots +\xi _{p}x_{p}), \end{equation*}% \textit{considered on the $p$ dimensional space $\mathbb{Q}^{p}$ of rational vectors, is an ordinary polynomial of degree at most $k$.} \bigskip \begin{proof} By Theorem 2.2, \begin{equation*} f=\sum_{m=0}^{k}f_{m},\eqno(2.5) \end{equation*}% where $f_{0}$ is a constant and $f_{m}:\mathbb{R\rightarrow R}$, $1,...,m$, are diagonalizations of $m$-additive symmetric functions $F_{m}:\mathbb{R}% ^{m}\mathbb{\rightarrow R}$, i.e., \begin{equation*} f_{m}(x)=F_{m}(x,...,x). \end{equation*}% For a $m$-additive function $F_{m}$ the equality \begin{equation*} F_{m}(\xi _{1},...,\xi _{i-1},r\xi _{i},\xi _{i+1},...,\xi _{m})=rF_{m}(\xi _{1},...,\xi _{m}) \end{equation*}% holds for all $i=1,...,m$ and any $r\in \mathbb{Q}$, $\xi _{i}\in $ $\mathbb{% R}$, $i=1,...,m$ (see \cite[Theorem 13.4.1]{Kuc}). Using this, it is not difficult to verify that for any $(x_{1},...,x_{p})\in \mathbb{Q}^{p}$, \begin{eqnarray*} f_{m}(\xi _{1}x_{1}+\cdots +\xi _{p}x_{p}) &=&F_{m}(\xi _{1}x_{1}+\cdots +\xi _{p}x_{p},...,\xi _{1}x_{1}+\cdots +\xi _{p}x_{p}) \\ &=&\sum_{\substack{ 0\leq s_{i}\leq m,~\overline{i=1,p} \\ s_{1}+\cdots +s_{p}=m}}A_{s_{1}...s_{p}}F_{m}(\underset{s_{1}}{\underbrace{\xi _{1},...,\xi _{1}}},...,\underset{s_{p}}{\underbrace{\xi _{p},...,\xi _{p}}}% )x_{1}^{s_{1}}...x_{p}^{s_{p}}. \end{eqnarray*}% Here $A_{s_{1}...s_{p}}$ are some coefficients, namely $% A_{s_{1}...s_{p}}=m!/(s_{1}!...s_{p}!).$ Considering the last formula in (2.5), we conclude that the function $g(x_{1},...,x_{p})$, restricted to $% \mathbb{Q}^{p}$, is a polynomial of degree at most $k$. \end{proof} \textbf{Lemma 2.3.} \textit{Assume $f$ is a polynomial function of order $k$% . Then there exists a polynomial function $H$ of order $k+1$ such that $% H(0)=0$ and} \begin{equation*} f(x)=H(x+1)-H(x).\eqno(2.6) \end{equation*} \bigskip \begin{proof} Consider the function \begin{equation*} H(x):=xf(x)+\sum_{i=1}^{k}(-1)^{i}\frac{x(x+1)...(x+i)}{(i+1)!}\Delta _{1}^{i}f(x).\eqno(2.7) \end{equation*}% Clearly, $H(0)=0.$ We are going to prove that $H$ is a polynomial function of order $k+1$ and satisfies (2.6). Let us first show that for any polynomial function $g$ of order $m$ the function $G_{1}(x)=xg(x)$ is a polynomial function of order $m+1.$ Indeed, for any $h_{1},...,h_{m+2}\in \mathbb{R}$ we can write that \begin{equation*} \Delta _{h_{1}...h_{m+2}}G_{1}(x)=(x+h_{1}+\cdots +h_{m+2})\Delta _{h_{1}...h_{m+2}}g(x) \end{equation*}% \begin{equation*} +\sum_{i=1}^{m+2}h_{i}\Delta _{h_{1}...h_{i-1}h_{i+1...}h_{m+2}}g(x).\eqno(2.8) \end{equation*}% The last formula is verified directly by using the known product property of differences, that is, the equality \begin{equation*} \Delta _{h}(g_{1}g_{2})=g_{1}\Delta _{h}g_{2}+g_{2}\Delta _{h}g_{1}+\Delta _{h}g_{1}\Delta _{h}g_{2}.\eqno(2.9) \end{equation*}% Now since $g$ is a polynomial function of order $m$, all summands in (2.8) is equal to zero; hence we obtain that $G_{1}(x)$ is a polynomial function of order $m+1$. By induction, we can prove that the function $% G_{p}(x)=x^{p}g(x)$ is a polynomial function of order $m+p.$ Since $\Delta _{1}^{i}f(x)$ in (2.7) is a polynomial function of order $k-i$, it follows that all summands in (2.7) are polynomial functions of order $k+1$. Therefore, $H(x)$ is a polynomial function of order $k+1$. Now let us prove (2.6). Considering the property (2.9) in (2.7) we can write that% \begin{equation*} \Delta _{1}H(x)=\left[ f(x)+(x+1)\Delta _{1}f(x)\right] \end{equation*}% \begin{equation*} +\sum_{i=1}^{k}(-1)^{i}\left[ \frac{(x+1)...(x+i+1)}{(i+1)!}\Delta _{1}^{i+1}f(x)+\Delta _{1}\left( \frac{x(x+1)...(x+i)}{(i+1)!}\right) \Delta _{1}^{i}f(x)\right] .\eqno(2.10) \end{equation*} Note that in (2.10) \begin{equation*} \Delta _{1}\left( \frac{x(x+1)...(x+i)}{(i+1)!}\right) =\frac{(x+1)...(x+i)}{% i!}. \end{equation*}% Considering this and the assumption $\Delta _{1}^{k+1}f(x)=0$, it follows from (2.10) that \begin{equation*} \Delta _{1}H(x)=f(x), \end{equation*}% that is, (2.6) holds. \end{proof} \bigskip The next lemma is due to Gajda \cite{Gajda}. \bigskip \textbf{Lemma 2.4 }(see \cite[Corollary 1]{Gajda}). \textit{Let $f:$ $% \mathbb{R\rightarrow R}$ be a $\mod1$ periodic function such that, for any $% h_{1},...,h_{k}\in \mathbb{R}$, $\Delta _{h_{1}...h_{k}}f$ is continuous. Then there exist a continuous function $g:$ $\mathbb{R\rightarrow R}$ and a polynomial function $H$ of $k$-th order such that $f=g+H$.} \bigskip The following theorem generalizes de Bruijn's theorem (see \cite[Theorem 1.1]% {B1}) on the difference property of continuous functions and shows that Gajda's above lemma (see Lemma 2.1) is valid. Note that the main result of \cite{Gajda} also uses this theorem. \bigskip \textbf{Theorem 2.3.} \textit{Assume for any $h_{1},...,h_{k}\in \mathbb{R}$% , the difference $\Delta _{h_{1}...h_{k}}f(x)$ is a continuous function of the variable $x$. Then there exist a function $g\in C(\mathbb{R})$ and a polynomial function $H$ of $k$-th order with the property $H(0)=0$ such that} \begin{equation*} f=g+H. \end{equation*} \bigskip \begin{proof} We prove this theorem by induction. For $k=1$, the theorem is the result of de Bruijn: if $f$ is such that, for each $h$, $\Delta _{h}f(x)$ is a continuous function of $x$, then it can be written in the form $g+H$, where $g$ is continuous and $H$ is additive (that is, satisfies the Cauchy Functional Equation). Assume that the theorem is valid for $k-1.$ Let us prove it for $k$. Without loss of generality we may assume that $% f(0)=f(1)$. Otherwise, we can prove the theorem for $f_{0}(x)=f(x)-\left[ f(1)-f(0)\right] x$ and then automatically obtain its validity for $f$. Consider the function \begin{equation*} F_{1}(x)=f(x+1)-f(x)\text{, }x\in \mathbb{R}.\eqno(2.11) \end{equation*}% Since for\ any $h_{1},...,h_{k}\in \mathbb{R}$, $\Delta _{h_{1}...h_{k}}f(x)$ is a continuous function of $x$ and $\Delta _{h_{1}...h_{k-1}}F_{1}=\Delta _{h_{1}...h_{k-1}1}f$, the difference $\Delta _{h_{1}...h_{k-1}}F_{1}(x)$ will be a continuous function of $x$, as well. By assumption, there exist a function $g_{1}\in C(\mathbb{R})$ and a polynomial function $H_{1}$ of $% (k-1) $-th order with the property $H_{1}(0)=0$ such that \begin{equation*} F_{1}=g_{1}+H_{1}.\eqno(2.12) \end{equation*}% It follows from Lemma 2.3 that there exists a polynomial function $H_{2}$ of order $k$ such that $H_{2}(0)=0$ and \begin{equation*} H_{1}(x)=H_{2}(x+1)-H_{2}(x).\eqno(2.13) \end{equation*}% Substituting (2.13) in (2.12) we obtain that \begin{equation*} F_{1}(x)=g_{1}(x)+H_{2}(x+1)-H_{2}(x).\eqno(2.14) \end{equation*}% It follows from (2.11) and (2.14) that \begin{equation*} g_{1}(x)=\left[ f(x+1)-H_{2}(x+1)\right] -\left[ f(x)-H_{2}(x)\right] .\eqno% (2.15) \end{equation*} Consider the function \begin{equation*} F_{2}=f-H_{2}.\eqno(2.16) \end{equation*}% Since $H_{2}$ is a polynomial function of order $k$ and for any $% h_{1},...,h_{k}\in \mathbb{R}$ the difference $\Delta _{h_{1}...h_{k}}f(x)$ is a continuous function of $x$, we obtain that $\Delta _{h_{1}...h_{k}}F_{2}(x)$ is also a continuous function of $x$. In addition, since $f(0)=f(1)$ and $H_{2}(0)=H_{2}(1)=0$, it follows from (2.16) that $% F_{2}(0)=F_{2}(1)$. We will use these properties of $F_{2}$ below. Let us write (2.15) in the form \begin{equation*} g_{1}(x)=F_{2}(x+1)-F_{2}(x),\eqno(2.17) \end{equation*}% and define the following $\mod1$ periodic function \begin{eqnarray*} F^{\ast }(x) &=&F_{2}(x)\text{ for }x\in \lbrack 0,1), \\ F^{\ast }(x+1) &=&F^{\ast }(x)\text{ for }x\in \mathbb{R}. \end{eqnarray*} Consider the function \begin{equation*} F=F_{2}-F^{\ast }.\eqno(2.18) \end{equation*}% Let us show that $F\in C(\mathbb{R)}$. Indeed since $F(x)=0$ for $x\in \lbrack 0,1)$, $F$ is continuous on $(0,1)$. Consider now the interval $% [1,2) $. For any $x\in \lbrack 1,2)$ by the definition of $F^{\ast }$ and (2.17) we can write that \begin{equation*} F(x)=F_{2}(x)-F_{2}(x-1)=g_{1}(x-1).\eqno(2.19) \end{equation*}% Since $g_{1}\in C(\mathbb{R)}$, it follows from (2.19) that $F$ is continuous on $(1,2)$. Note that by (2.17) $g_{1}(0)=0$; hence $% F(1)=g_{1}(0)=0$. Since $F\equiv 0$ on $[0,1)$, $F(1)=0$ and $F\in C(1,2),$ we obtain that $F$ is continuous on $(0,2)$. Consider the interval $[2,3)$. For any $x\in \lbrack 2,3)$ we can write that \begin{equation*} F(x)=F_{2}(x)-F_{2}(x-2)=g_{1}(x-1)+g_{1}(x-2).\eqno(2.20) \end{equation*}% Since $g_{1}\in C(\mathbb{R)}$, $F$ is continuous on $(2,3)$. Note that by (2.19) $\lim_{x\rightarrow 2-}F(x)=g_{1}(1)$ and by (2.20) $F(2)=g_{1}(1).$ We obtain from these arguments that $F$ is continuous on $(0,3)$. In the same way, we can prove that $F$ is continuous on $(0,m)$ for any $m\in \mathbb{N}$. Similar arguments can be used to prove the continuity of $F$ on $(-m,0)$ for any $m\in \mathbb{N}$. We show it for the first interval $[-1,0)$. For any $% x\in \lbrack -1,0)$ by the definition of $F^{\ast }$ and (2.17) we can write that \begin{equation*} F(x)=F_{2}(x)-F_{2}(x+1)=-g_{1}(x). \end{equation*}% Since $g_{1}\in C(\mathbb{R)}$, it follows that $F$ is continuous on $(-1,0)$% . Besides, \newline $\lim_{x\rightarrow 0-}F(x)=-g_{1}(0)=0.$ This shows that $F$ is continuous on $(-1,1)$, since $F\equiv 0$ on $[0,1).$ Combining all the above arguments we conclude that $F\in C(\mathbb{R)}$. Since $F\in C(\mathbb{R)}$ and $\Delta _{h_{1}...h_{k}}F_{2}(x)$ is a continuous function of $x$, we obtain from (2.18) that $\Delta _{h_{1}...h_{k}}F^{\ast }(x)$ is also a continuous function of $x.$ By Lemma 2.4, there exist a function $g_{2}\in C(\mathbb{R)}$ and a polynomial function $H_{3}$ of order $k$ such that \begin{equation*} F^{\ast }=g_{2}+H_{3}.\eqno(2.21) \end{equation*}% It follows from (2.16), (2.18) and (2.21) that \begin{equation*} f=F+g_{2}+H_{2}+H_{3}.\eqno(2.22) \end{equation*} Introduce the notation \begin{eqnarray*} H(x) &=&H_{2}(x)+H_{3}(x)-H_{3}(0), \\ g(x) &=&F(x)+g_{2}(x)+H_{3}(0). \end{eqnarray*}% Obviously, $g\in C(\mathbb{R)}$ and $H(0)=0$. It follows from (2.22) and the above notation that \begin{equation*} f=g+H. \end{equation*}% This completes the proof of the theorem. \end{proof} \bigskip \subsection{Main results} We start this subsection with the following lemma. \bigskip \textbf{Lemma 2.5.} \textit{Assume we are given pairwise linearly independent vectors $\mathbf{a}^{i},$ $i=1,...,k,$ and a function $f\in C(% \mathbb{R}^{n})$ of the form (2.1) with arbitrarily behaved univariate functions $f_{i}$. Then for any $h_{1},...,h_{k-1}\in \mathbb{R}$, and all indices $i=1,...,k$, $\Delta _{h_{1}...h_{k-1}}f_{i}\in C(\mathbb{R})$.} \bigskip \begin{proof} We prove this lemma for the function $f_{k}.$ It can be proven for the other functions $f_{i}$ in the same way. Let $% h_{1},...,h_{k-1}\in \mathbb{R}$ be given. Since the vectors $\mathbf{a}^{i}$ are pairwise linearly independent, for each $j=1,...,k-1,$ there is a vector $\mathbf{b}^{j}$ such that $\mathbf{b}^{j}\cdot \mathbf{a}^{j}=0$ and $% \mathbf{b}^{j}\cdot \mathbf{a}^{k}\neq 0$. It is not difficult to see that for any $\lambda \in \mathbb{R}$, $\Delta _{\lambda \mathbf{b}^{j}}f_{j}(% \mathbf{a}^{j}\cdot \mathbf{x})=0.$ Therefore, for any $\lambda _{1},...,\lambda _{k-1}\in \mathbb{R}$, we obtain from (2.1) that \begin{equation*} \Delta _{\lambda _{1}\mathbf{b}^{1}...\lambda _{k-1}\mathbf{b}^{k-1}}f(% \mathbf{x})=\Delta _{\lambda _{1}\mathbf{b}^{1}...\lambda _{k-1}\mathbf{b}% ^{k-1}}f_{k}(\mathbf{a}^{k}\cdot \mathbf{x}).\eqno(2.23) \end{equation*}% Note that in multivariate setting the difference operator $\Delta _{\mathbf{h% }^{1}...\mathbf{h}^{k}}f(\mathbf{x})$ is defined similarly as in the previous section. If in (2.23) we take \begin{eqnarray*} \mathbf{x} &\mathbf{=}&\frac{\mathbf{a}^{k}}{\left\Vert \mathbf{a}% ^{k}\right\Vert ^{2}}t\text{, }t\in \mathbb{R}, \\ \lambda _{j} &=&\frac{h_{j}}{\mathbf{a}^{k}\cdot \mathbf{b}^{j}}\text{, }% j=1,...,k-1, \end{eqnarray*}% we will obtain that $\Delta _{h_{1}...h_{k-1}}f_{k}\in C(\mathbb{R})$. \end{proof} The following theorem is valid. \bigskip \textbf{Theorem 2.4.} \textit{Assume a function $f\in C(\mathbb{R}^{n})$ is of the form (2.1). Then there exist continuous functions $g_{i}:\mathbb{% R\rightarrow R}$, $i=1,...,k$, and a polynomial $P(\mathbf{x})$ of degree at most $k-1$ such that} \begin{equation*} f(\mathbf{x})=\sum_{i=1}^{k}g_{i}(\mathbf{a}^{i}\cdot \mathbf{x})+P(\mathbf{x% }).\eqno(2.24) \end{equation*} \bigskip \begin{proof} By Lemma 2.5 and Theorem 2.3, for each $i=1,...,k$, there exists a function $g_{i}\in C(\mathbb{R})$ and a polynomial function $H_{i}$ of $(k-1)$-th order with the property $H_{i}(0)=0$ such that \begin{equation*} f_{i}=g_{i}+H_{i}.\eqno(2.25) \end{equation*} Consider the function \begin{equation*} F(\mathbf{x})=f(\mathbf{x})-\sum_{i=1}^{k}g_{i}(\mathbf{a}^{i}\cdot \mathbf{x% }).\eqno(2.26) \end{equation*}% It follows from (2.1), (2.25) and (2.26) that \begin{equation*} F(\mathbf{x})=\sum_{i=1}^{k}H_{i}(\mathbf{a}^{i}\cdot \mathbf{x}).\eqno(2.27) \end{equation*} Denote the restrictions of the multivariate functions $H_{i}(\mathbf{a}% ^{i}\cdot \mathbf{x})$ to the space $\mathbb{Q}^{n}$ by $P_{i}(\mathbf{x})$, respectively. By Lemma 2.2, the functions $P_{i}(\mathbf{x})$ are ordinary polynomials of degree at most $k-1$. Since the space $\mathbb{Q}^{n}$ is dense in $\mathbb{R}^{n}$, and the functions $F(\mathbf{x})$, $P_{i}(\mathbf{% x})$, $i=1,...,k$, are continuous on $\mathbb{R}^{n}$, and the equality \begin{equation*} F(\mathbf{x})=\sum_{i=1}^{k}P_{i}(\mathbf{x}),\eqno(2.28) \end{equation*}% holds for all $\mathbf{x}\in \mathbb{Q}^{n}$, we obtain that (2.28) holds also for all $\mathbf{x}\in \mathbb{R}^{n}$. Now (2.24) follows from (2.26) and (2.28) by putting $P=\sum_{i=1}^{k}P_{i}$. \end{proof} Now we generalize Theorem 2.4 from $C(\mathbb{R}^{n})$ to any space $C^{s}(% \mathbb{R}^{n})$ of $s$-th order continuously differentiable functions. \bigskip \textbf{Theorem 2.5.} \textit{Assume $f\in C^{s}(\mathbb{R}^{n})$ is of the form (2.1). Then there exist functions $g_{i}\in C^{s}(\mathbb{R})$, $% i=1,...,k$, and a polynomial $P(\mathbf{x})$ of degree at most $k-1$ such that (2.24) holds.} \bigskip The proof is based on Theorems 2.1 and 2.4. On the one hand, it follows from Theorem 2.4 that the $s$-th order continuously differentiable function $f-P$ can be expressed as $\sum_{i=1}^{k}g_{i}$ with continuous $g_{i}$. On the other hand, since the class $\mathcal{B}$ in Theorem 2.1, in particular, can be taken as $C(\mathbb{R}),$ it follows that $g_{i}\in C^{s}(\mathbb{R})$. \bigskip Note that Theorem 2.5 solves the problem posed in Buhmann and Pinkus \cite% {12} and Pinkus \cite[p.14]{117} up to a polynomial. The following theorem shows that in the two dimensional setting $n=2$ it solves the problem completely. \bigskip \textbf{Theorem 2.6.} \textit{Assume a function $f\in C^{s}(\mathbb{R}^{2})$ is of the form} \begin{equation*} f(x,y)=\sum_{i=1}^{k}f_{i}(a_{i}x+b_{i}y), \end{equation*}% \textit{where $(a_{i},b_{i})$ are pairwise linearly independent vectors in $% \mathbb{R}^{2}$ and $f_{i}$ are arbitrary univariate functions. Then there exist functions $g_{i}\in C^{s}(\mathbb{R})$, $i=1,...,k$, such that} \begin{equation*} f(x,y)=\sum_{i=1}^{k}g_{i}(a_{i}x+b_{i}y).\eqno(2.29) \end{equation*} \bigskip The proof of this theorem is not difficult. First we apply Theorem 2.5 and obtain that \begin{equation*} f(x,y)=\sum_{i=1}^{k}\overline{g}_{i}(a_{i}x+b_{i}y)+P(x,y),\eqno(2.30) \end{equation*}% where $\overline{g}_{i}\in C^{s}(\mathbb{R})$ and $P(x,y)$ is a bivariate polynomial of degree at most $k-1$. Then we use the known fact that a bivariate polynomial $P(x,y)$ of degree $k-1$ is decomposed into a sum of ridge polynomials with any given $k$ pairwise linearly independent directions $(a_{i},b_{i}),$ $i=1,...,k$ (see e.g. \cite{97}). That is, \begin{equation*} P(x,y)=\sum_{i=1}^{k}p_{i}(a_{i}x+b_{i}y), \end{equation*}% where $p_{i}$ are univariate polynomials of degree at most $k-1$. Considering this in (2.30) gives the desired representation (2.29). \bigskip \textbf{Remark 2.2.} Theorem 2.5 can be restated also for the classes $% C^{\infty }(\mathbb{R})$ of infinitely differentiable functions and $D(% \mathbb{R})$ of analytic functions. That is, if under the conditions of Theorem 2.5, we have $f\in C^{\infty }(\mathbb{R}^{n})$ (or $f\in D(\mathbb{R% }^{n})$), then this function can be represented also in the form (2.24) with $g_{i}\in C^{\infty }(\mathbb{R})$ (or $g_{i}\in D(\mathbb{R})$). This follows, similarly to the case $C^{s}(\mathbb{R})$ above, from Theorem 2.4 and Remark 2.1. These arguments are also valid for Theorem 2.6. \bigskip \section{A solution to the smoothness problem under certain conditions} Assume we are given a function $f\in C^{s}(\mathbb{R}^{n})$ of the form (2.1). In this section, we discuss various conditions on the directions $% \mathbf{a}^{i}$ guaranteeing the validity of (2.2) with $g_{i}\in C^{s}(% \mathbb{R})$. \subsection{Directions with only rational components} The following theorems, in particular, show that if directions of ridge functions have only rational coordinates then no polynomial term appears in Theorems 2.4 and 2.5. \bigskip \textbf{Theorem 2.7.} \textit{Assume a function $f\in C(\mathbb{R}^{n})$ is of the form (2.1) and there is a nonsingular linear transformation $T:$ $% \mathbb{R}^{n}\rightarrow \mathbb{R}^{n}$ such that $T\mathbf{a}^{i}\in \mathbb{Q}\mathit{^{n}},$ $i=1,...,k$. Then there exist continuous functions $g_{i}:\mathbb{R\rightarrow R}$, $i=1,...,k$, such that (2.2) holds.} \bigskip \begin{proof} Applying the coordinate change $\mathbf{x\rightarrow y}$, given by the formula $\mathbf{x}=T\mathbf{y}$, to both sides of (2.1) we obtain that \begin{equation*} \tilde{f}(\mathbf{y})=\sum_{i=1}^{k}f_{i}(\mathbf{b}^{i}\cdot \mathbf{y}), \end{equation*}% where $\tilde{f}(\mathbf{y})=f(T\mathbf{y})$ and $\mathbf{b}^{i}=T\mathbf{a}% ^{i},$ $i=1,...,k.$ Let us repeat the proof of Theorem 2.4 for the function $% \tilde{f}$. Since the vectors $\mathbf{b}^{i}$, $i=1,...,k,$ have rational coordinates, it is not difficult to see that the restrictions of the functions $H_{i}$ to $\mathbb{Q}$ are univariate polynomials. Indeed, for each $\mathbf{b}^{i}$ we can choose a vector $\mathbf{c}^{i}$ with rational coordinates such that $\mathbf{b}^{i}\cdot \mathbf{c}^{i}=1$. If in the equality $H_{i}(\mathbf{b}^{i}\cdot \mathbf{x})=P_{i}(\mathbf{x}),$ $\mathbf{% x}\in \mathbb{Q}^{n}$, we take $\mathbf{x=c}^{i}t$ with $t\in \mathbb{Q}$, we obtain that $H_{i}(t)=P_{i}(\mathbf{c}^{i}t)$ for all $t\in \mathbb{Q}$. Now since $P_{i}$ is a multivariate polynomial on $\mathbb{Q}^{n}$, $H_{i}$ is a univariate polynomial on $\mathbb{Q}$. Denote this univariate polynomial by $L_{i}$. Thus the formula \begin{equation*} P_{i}(\mathbf{x})=L_{i}(\mathbf{b}^{i}\cdot \mathbf{x})\eqno(2.31) \end{equation*}% holds for each $i=1,...,k$, and all $\mathbf{x}\in \mathbb{Q}^{n}$. Since $% \mathbb{Q}^{n}$ is dense in $\mathbb{R}^{n}$, we see that (2.31) holds, in fact, for all $\mathbf{x}\in \mathbb{R}^{n}$. Thus the polynomial $P(\mathbf{% x})$ in (2.24) can be expressed as $\sum_{i=1}^{k}L_{i}(\mathbf{b}^{i}\cdot \mathbf{x})$. Considering this in Theorem 2.4, we obtain that \begin{equation*} \tilde{f}(\mathbf{y})=\sum_{i=1}^{k}g_{i}(\mathbf{b}^{i}\cdot \mathbf{y}),% \eqno(2.32) \end{equation*}% where $g_{i}$ are continuous functions. Using the inverse transformation $% \mathbf{y}=T^{-1}\mathbf{x}$ in (2.32) we arrive at (2.2). \end{proof} \textbf{Theorem 2.8.} \textit{Assume a function $f\in C^{s}(\mathbb{R}^{n})$ is of the form (2.1) and there is a nonsingular linear transformation $T:$ $% \mathbb{R}^{n}\rightarrow \mathbb{R}^{n}$ such that $T\mathbf{a}^{i}\in \mathbb{Q}\mathit{^{n}},$ $i=1,...,k$. Then there exist functions $g_{i}\in C^{s}(\mathbb{R})$, $i=1,...,k$, such that (2.2) holds.} \bigskip The proof of this theorem easily follows from Theorem 2.7 and Theorem 2.5. \bigskip \subsection{Linear independence of $k-1$ directions} We already know that if the given directions $\mathbf{a}^{i}$ form a linearly independent set, then the smoothness problem has a positive solution (see Section 2.1.1). What can we say if all $\mathbf{a}^{i}$ are not linearly independent? In the sequel, we show that if $k-1$ of the directions $\mathbf{a}^{i}$, $i=1,...,k,$ are linearly independent, then in (2.1) $f_{i}$ can be replaced with $g_{i}\in C^{s}(\mathbb{R})$. We will also estimate the modulus of continuity of $g_{i}$ in terms of the modulus of continuity of a function generated from $f$ under a linear transformation. Let $F:\mathbb{R}^{n}\rightarrow \mathbb{R}$, $n\geq 1,$ be any function and $\Omega \subset \mathbb{R}^{n}$. The function \begin{equation*} \omega (F;\delta ;\Omega )=\sup \left\{ \left\vert F(\mathbf{x})-F(\mathbf{y}% )\right\vert :\mathbf{x},\mathbf{y}\in \Omega ,\text{ }\left\vert \mathbf{x}-% \mathbf{y}\right\vert \leq \delta \right\} ,\text{ }0\leq \delta \leq diam\Omega , \end{equation*}% is called the modulus of continuity of the function $F(\mathbf{x}% )=F(x_{1},...,x_{n})$ on the set $\Omega .$ We will also use the notation $% \omega _{\mathbb{Q}}(F;\delta ;\Omega )$, which stands for the function $% \omega (F;\delta ;\Omega \cap \mathbb{Q}^{n})$. Here $\mathbb{Q}$ denotes the set of rational numbers. Note that $\omega _{\mathbb{Q}}(F;\delta ;\Omega )$ makes sense if the set $\Omega \cap \mathbb{Q}^{n}$ is not empty. Clearly, $\omega _{\mathbb{Q}}(F;\delta ;\Omega )\leq \omega (F;\delta ;\Omega )$. The equality $\omega _{\mathbb{Q}}(F;\delta ;\Omega )=\omega (F;\delta ;\Omega )$ holds for continuous $F$ and certain sets $\Omega $. For example, it holds if for any $\mathbf{x},\mathbf{y}\in \Omega $ with $% \left\vert \mathbf{x}-\mathbf{y}\right\vert \leq \delta $ there exist sequences $\left\{ \mathbf{x}_{m}\right\} ,\left\{ \mathbf{y}_{m}\right\} \subset \Omega \cap \mathbb{Q}^{n}$ such that $\mathbf{x}_{m}\rightarrow \mathbf{x}$, $\mathbf{y}_{m}\rightarrow \mathbf{y}$ and $\left\vert \mathbf{x% }_{m}-\mathbf{y}_{m}\right\vert \leq \delta ,$ for all $m$. There are many sets $\Omega $, which satisfy this property. \bigskip The following lemma is valid. \bigskip \textbf{Lemma 2.6.} \textit{Assume a function $G\in C(\mathbb{R}^{n})$ has the form} \begin{equation*} G(x_{1},...,x_{n})=\sum_{i=1}^{n}g(x_{i})-g(x_{1}+\cdot \cdot \cdot +x_{n}),% \eqno(2.33) \end{equation*}% \textit{where $g$ is an arbitrarily behaved function. Then the following inequality holds} \begin{equation*} \omega _{\mathbb{Q}}(g;\delta ;[-M,M])\leq 2\delta \left\vert g(1)-g(0)\right\vert +3\omega \left( G;\delta ;[-M,M]^{n}\right) ,\eqno(2.34) \end{equation*}% \textit{where $\delta \in \left( 0,\frac{1}{2}\right) \cap \mathbb{Q}$ and $% M\geq 1$.} \bigskip \begin{proof} Consider the function $f(t)=g(t)-g(0)$ and write (2.33) in the form \begin{equation*} F(x_{1},...,x_{n})=\sum_{i=1}^{n}f(x_{i})-f(x_{1}+\cdot \cdot \cdot +x_{n}),% \eqno(2.35) \end{equation*}% where \begin{equation*} F(x_{1},...,x_{n})=G(x_{1},...,x_{n})-(n-1)g(0). \end{equation*}% Note that the functions $f$ and $g$, as well as the functions $F$ and $G,$ have a common modulus of continuity. Thus we prove the lemma if we prove it for the pair $\left\langle F,f\right\rangle .$ Since $f(0)=0,$ it follows from (2.35) that \begin{equation*} F(x_{1},0,...,0)=F(0,x_{2},0,...,0)=\cdot \cdot \cdot =F(0,0,...,x_{n})=0.% \eqno(2.36) \end{equation*}% For the sake of brevity, introduce the notation $\mathcal{F}(x_{1},x_{2})=$ $% F(x_{1},x_{2},0,...,0)$. Obviously, for any real number $x,$ \begin{eqnarray*} \mathcal{F}(x,x) &=&2f(x)-f(2x); \\ \mathcal{F}(x,2x) &=&f(x)+f(2x)-f(3x); \\ &&\cdot \cdot \cdot \\ \mathcal{F}(x,(k-1)x) &=&f(x)+f((k-1)x)-f(kx). \end{eqnarray*} We obtain from the above equalities that \begin{eqnarray*} f(2x) &=&2f(x)-\mathcal{F}(x,x), \\ f(3x) &=&3f(x)-\mathcal{F}(x,x)-\mathcal{F}(x,2x), \\ &&\cdot \cdot \cdot \\ f(kx) &=&kf(x)-\mathcal{F}(x,x)-\mathcal{F}(x,2x)-\cdot \cdot \cdot -% \mathcal{F}(x,(k-1)x). \end{eqnarray*}% Thus for any nonnegative integer $k$, \begin{equation*} f(x)=\frac{1}{k}f(kx)+\frac{1}{k}\left[ \mathcal{F}(x,x)+\mathcal{F}% (x,2x)+\cdot \cdot \cdot +\mathcal{F}(x,(k-1)x)\right] .\eqno(2.37) \end{equation*} Consider now the simple fraction $\frac{p}{m}\in (0,\frac{1}{2})$ and set $% m_{0}=\left[ \frac{m}{p}\right] .$ Here $[r]$ denotes the whole number part of $r$. Clearly, $m_{0}\geq 2$ and the remainder $p_{1}=m-m_{0}p<p.$ Taking $% x=\frac{p}{m}$ and $k=m_{0}$ in (2.37) gives us the following equality \begin{equation*} f\left( \frac{p}{m}\right) =\frac{1}{m_{0}}f\left( 1-\frac{p_{1}}{m}\right) \end{equation*}% \begin{equation*} +\frac{1}{m_{0}}\left[ \mathcal{F}\left( \frac{p}{m},\frac{p}{m}\right) +% \mathcal{F}\left( \frac{p}{m},\frac{2p}{m}\right) +\cdot \cdot \cdot +% \mathcal{F}\left( \frac{p}{m},(m_{0}-1)\frac{p}{m}\right) \right] .\eqno% (2.38) \end{equation*}% On the other hand, since \begin{equation*} \mathcal{F}\left( \frac{p_{1}}{m},1-\frac{p_{1}}{m}\right) =f\left( \frac{% p_{1}}{m}\right) +f\left( 1-\frac{p_{1}}{m}\right) -f(1), \end{equation*}% it follows from (2.38) that \begin{equation*} f\left( \frac{p}{m}\right) =\frac{f(1)}{m_{0}} \end{equation*}% \begin{equation*} +\frac{1}{m_{0}}\left[ \mathcal{F}\left( \frac{p}{m},\frac{p}{m}\right) +\cdot \cdot \cdot +% \mathcal{F}\left( \frac{p}{m},(m_{0}-1)\frac{p}{m}\right) +\mathcal{F}\left( \frac{p_{1}}{m},1-\frac{p_{1}}{m}\right) \right] \end{equation*} \begin{equation*} -\frac{1}{m_{0}}f\left( \frac{p_{1}}{m}\right) .\eqno(2.39) \end{equation*} Put $m_{1}=\left[ \frac{m}{p_{1}}\right] $, $p_{2}=m-m_{1}p_{1}.$ Clearly, $% 0\leq p_{2}<p_{1}$. Similar to (2.39), we can write that \begin{equation*} f\left( \frac{p_{1}}{m}\right) =\frac{f(1)}{m_{1}} \end{equation*}% \begin{equation*} +\frac{1}{m_{1}}\left[ \mathcal{F}\left( \frac{p_{1}}{m},\frac{p_{1}}{m}\right) +\cdot \cdot \cdot +% \mathcal{F}\left( \frac{p_{1}}{m},(m_{1}-1)\frac{p_{1}}{m}\right) +\mathcal{F% }\left( \frac{p_{2}}{m},1-\frac{p_{2}}{m}\right) \right] \end{equation*} \begin{equation*} -\frac{1}{m_{1}}f\left( \frac{p_{2}}{m}\right) .\eqno(2.40) \end{equation*} Let us make a convention that (2.39) is the $1$-st and (2.40) is the $2$-nd formula. One can continue this process by defining the chain of pairs $% (m_{2},p_{3}),$ $(m_{3},p_{4})$ until the pair $(m_{k-1},p_{k})$ with $% p_{k}=0$ and writing out the corresponding formulas for each pair. For example, the last $k$-th formula will be of the form \begin{equation*} f\left( \frac{p_{k-1}}{m}\right) =\frac{f(1)}{m_{k-1}} \end{equation*}% \begin{equation*} +\frac{1}{m_{k-1}}\left[ \mathcal{F}\left( \frac{p_{k-1}}{m},\frac{p_{k-1}}{m% }\right) +\cdot \cdot \cdot +\mathcal{F}\left( \frac{p_{k-1}}{m},(m_{k-1}-1)% \frac{p_{k-1}}{m}\right) +\mathcal{F}\left( \frac{p_{k}}{m},1-\frac{p_{k}}{m}% \right) \right] \end{equation*}% \begin{equation*} -\frac{1}{m_{k-1}}f\left( \frac{p_{k}}{m}\right) .\eqno(2.41) \end{equation*}% Note that in (2.41), $f\left( \frac{p_{k}}{m}\right) =0$ and $\mathcal{F}% \left( \frac{p_{k}}{m},1-\frac{p_{k}}{m}\right) =0$. Considering now the $k$% -th formula in the $(k-1)$-th formula, then the obtained formula in the $% (k-2)$-th formula, and so forth, we will finally arrive at the equality \begin{equation*} f\left( \frac{p}{m}\right) =f(1)\left[ \frac{1}{m_{0}}-\frac{1}{m_{0}m_{1}}% +\cdot \cdot \cdot +\frac{(-1)^{k-1}}{m_{0}m_{1}\cdot \cdot \cdot m_{k-1}}% \right] \end{equation*} \begin{equation*} +\frac{1}{m_{0}}\left[ \mathcal{F}\left( \frac{p}{m},\frac{p}{m}\right) +\cdot \cdot \cdot +\mathcal{F}\left( \frac{p}{m},(m_{0}-1)\frac{p}{m}% \right) +\mathcal{F}\left( \frac{p_{1}}{m},1-\frac{p_{1}}{m}\right) \right] \end{equation*} \begin{equation*} -\frac{1}{m_{0}m_{1}}\left[ \mathcal{F}\left( \frac{p_{1}}{m},\frac{p_{1}}{m}% \right) +\cdot \cdot \cdot +\mathcal{F}\left( \frac{p_{1}}{m},(m_{1}-1)\frac{% p_{1}}{m}\right) +\mathcal{F}\left( \frac{p_{2}}{m},1-\frac{p_{2}}{m}\right) % \right] \end{equation*} \begin{equation*} +\cdot \cdot \cdot + \end{equation*}% \begin{equation*} \frac{(-1)^{k-1}}{m_{0}m_{1}\cdot \cdot \cdot m_{k-1}}\left[ \mathcal{F}% \left( \frac{p_{k-1}}{m},\frac{p_{k-1}}{m}\right) +\cdot \cdot \cdot +% \mathcal{F}\left( \frac{p_{k-1}}{m},(m_{k-1}-1)\frac{p_{k-1}}{m}\right) % \right] .\eqno(2.42) \end{equation*}% Taking into account (2.36) and the definition of $\mathcal{F}$, for any point of the form $\left( \frac{p_{i}}{m},c\right) $, $i=0,1,...,k-1,$ $% p_{0}=p$, $c\in \lbrack 0,1]$, we can write that \begin{equation*} \left\vert \mathcal{F}\left( \frac{p_{i}}{m},c\right) \right\vert =\left\vert \mathcal{F}\left( \frac{p_{i}}{m},c\right) -\mathcal{F}\left( 0,c\right) \right\vert \leq \omega \left( F;\frac{p_{i}}{m};[0,1]^{n}\right) \leq \omega \left( F;\frac{p}{m};[0,1]^{n}\right) . \end{equation*}% Applying this inequality to each term $\mathcal{F}\left( \frac{p_{i}}{m}% ,\cdot \right) $ in (2.42), we obtain that \begin{equation*} \left\vert f\left( \frac{p}{m}\right) \right\vert \leq \left[ \frac{1}{m_{0}}% -\frac{1}{m_{0}m_{1}}+\cdot \cdot \cdot +\frac{(-1)^{k-1}}{m_{0}m_{1}\cdot \cdot \cdot m_{k-1}}\right] \left\vert f(1)\right\vert \end{equation*} \begin{equation*} +\left[ 1+\frac{1}{m_{0}}+\cdot \cdot \cdot +\frac{1}{m_{0}\cdot \cdot \cdot m_{k-2}}\right] \omega \left( F;\frac{p}{m};[0,1]^{n}\right) .\eqno(2.43) \end{equation*} Since $m_{0}\leq m_{1}\leq \cdot \cdot \cdot \leq m_{k-1},$ it is not difficult to see that in (2.43) \begin{equation*} \frac{1}{m_{0}}-\frac{1}{m_{0}m_{1}}+\cdot \cdot \cdot +\frac{(-1)^{k-1}}{% m_{0}m_{1}\cdot \cdot \cdot m_{k-1}}\leq \frac{1}{m_{0}} \end{equation*}% and \begin{equation*} 1+\frac{1}{m_{0}}+\cdot \cdot \cdot +\frac{1}{m_{0}\cdot \cdot \cdot m_{k-2}}% \leq \frac{m_{0}}{m_{0}-1}. \end{equation*}% Considering the above two inequalities in (2.43) we obtain that \begin{equation*} \left\vert f\left( \frac{p}{m}\right) \right\vert \leq \frac{\left\vert f(1)\right\vert }{m_{0}}+\frac{m_{0}}{m_{0}-1}\omega \left( F;\frac{p}{m}% ;[0,1]^{n}\right) .\eqno(2.44) \end{equation*}% Since $m_{0}=\left[ \frac{m}{p}\right] \geq 2,$ it follows from (2.44) that \begin{equation*} \left\vert f\left( \frac{p}{m}\right) \right\vert \leq \frac{2p\left\vert f(1)\right\vert }{m}+2\omega \left( F;\frac{p}{m};[0,1]^{n}\right) .\eqno% (2.45) \end{equation*} Let now $\delta \in \left( 0,\frac{1}{2}\right) \cap \mathbb{Q}$ be a rational increment, $M\geq 1$ and $x,x+\delta $ be two points in $\left[ -M,M% \right] \cap \mathbb{Q}.$ By (2.45) we can write that \begin{equation*} \left\vert f(x+\delta )-f(x)\right\vert \leq \left\vert f(\delta )\right\vert +\left\vert F(x,\delta ,0,...,0)\right\vert \leq 2\delta \left\vert f(1)\right\vert +3\omega \left( F;\delta ;[-M,M]^{n}\right) .\eqno% (2.46) \end{equation*}% Now (2.34) follows from (2.46) and the definitions of $f$ and $F$. \end{proof} \textbf{Remark 2.3.} The above lemma shows that the restriction of $g$ to the set of rational numbers $\mathbb{Q}$ is uniformly continuous on any interval $[-M,M]\cap \mathbb{Q}$. \bigskip To prove the main result of this section we need the following lemma. \bigskip \textbf{Lemma 2.7.} \textit{Assume a function $G\in C(\mathbb{R}^{n})$ has the form} \begin{equation*} G(x_{1},...,x_{n})=\sum_{i=1}^{n}g(x_{i})-g(x_{1}+\cdot \cdot \cdot +x_{n}), \end{equation*}% \textit{where $g$ is an arbitrary function. Then there exists a function $% F\in C(\mathbb{R})$ such that} \begin{equation*} G(x_{1},...,x_{n})=\sum_{i=1}^{n}F(x_{i})-F(x_{1}+\cdot \cdot \cdot +x_{n})% \eqno(2.47) \end{equation*}% \textit{and the following inequality holds} \begin{equation*} \omega (F;\delta ;[-M,M])\leq 3\omega \left( G;\delta ;[-M,M]^{n}\right) ,% \eqno(2.48) \end{equation*}% \textit{where $0\leq \delta \leq \frac{1}{2}$ and $M\geq 1$.} \bigskip \begin{proof} Consider the function \begin{equation*} u(t)=g(t)-\left[ g(1)-g(0)\right] t. \end{equation*}% Obviously, $u(1)=u(0)$ and \begin{equation*} G(x_{1},...,x_{n})=\sum_{i=1}^{n}u(x_{i})-u(x_{1}+\cdot \cdot \cdot +x_{n}).% \eqno(2.49) \end{equation*}% By Lemma 2.6, the restriction of $u$ to $\mathbb{Q}$ is continuous and uniformly continuous on every interval $[-M,M]\cap \mathbb{Q}$. Denote this restriction by $v$. Let $y$ be any real number and $\{y_{k}\}_{k=1}^{\infty }$ be any sequence of rational numbers converging to $y$. We can choose $M>0$ so that $y_{k}\in \lbrack -M,M]$ for any $k\in \mathbb{N}$. It follows from the uniform continuity of $v$ on $[-M,M]\cap \mathbb{Q}$ that the sequence $% \{v(y_{k})\}_{k=1}^{\infty }$ is Cauchy. Thus there exits a finite limit $% \lim_{k\rightarrow \infty }v(y_{k})$. It is not difficult to see that this limit does not depend on the choice of $\{y_{k}\}_{k=1}^{\infty }$. Let $F$ denote the following extension of $v$ to the set of real numbers. \begin{equation*} F(y)=\left\{ \begin{array}{c} v(y),\text{ if }y\in \mathbb{Q}\text{;} \\ \lim_{k\rightarrow \infty }v(y_{k}),\text{ if }y\in \mathbb{R}\backslash \mathbb{Q}\text{ and }\{y_{k}\}\text{ is a sequence in }\mathbb{Q}\text{ tending to }y.% \end{array}% \right. \end{equation*}% In view of the above arguments, $F$ is well defined on the whole real line. Let us prove that for this function (2.47) is valid. Consider an arbitrary point $(x_{1},...,x_{n})\in \mathbb{R}^{n}$ and sequences of rationale numbers $\{y_{k}^{i}\}_{k=1}^{\infty },i=1,...,n,$ tending to $x_{1},...,x_{n},$ respectively. Taking into account (2.49), we can write that \begin{equation*} G(y_{k}^{1},...,y_{k}^{n})=\sum_{i=1}^{n}v(y_{k}^{i})-v(y_{k}^{1}+\cdot \cdot \cdot +y_{k}^{n}),\text{ for all }k=1,2,...,\eqno(2.50) \end{equation*}% since $v$ is the restriction of $u$ to $\mathbb{Q}$. Tending $k\rightarrow \infty $ in both sides of (2.50) we obtain (2.47). Let us now prove that $F\in C(\mathbb{R})$ and (2.48) holds. Since $% v(1)=v(0) $ we obtain from (2.49) and (2.34) that for $\delta \in \left( 0,% \frac{1}{2}\right) \cap \mathbb{Q}$, $M\geq 1$ and any numbers $a,b\in \lbrack -M,M]\cap \mathbb{Q}$, $\left\vert a-b\right\vert \leq \delta ,$ the following inequality holds \begin{equation*} \left\vert v(a)-v(b)\right\vert \leq 3\omega \left( G;\delta ;[-M,M]^{n}\right) .\eqno(2.51) \end{equation*}% Consider any real numbers $r_{1}$ and $r_{2}$ satisfying $r_{1},r_{2}\in \lbrack -M,M]$, $\left\vert r_{1}-r_{2}\right\vert \leq \delta $ and take sequences $\{a_{k}\}_{k=1}^{\infty }\subset \lbrack -M,M]\cap \mathbb{Q}$, $% \{b_{k}\}_{k=1}^{\infty }\subset \lbrack -M,M]\cap \mathbb{Q}$ with the property $\left\vert a_{k}-b_{k}\right\vert \leq \delta ,$ $k=1,2,...,$ and tending to $r_{1}$ and $r_{2}$, respectively. By (2.51), \begin{equation*} \left\vert v(a_{k})-v(b_{k})\right\vert \leq 3\omega \left( G;\delta ;[-M,M]^{n}\right) . \end{equation*}% If we take limits\ on both sides of the above inequality, we obtain that \begin{equation*} \left\vert F(r_{1})-F(r_{2})\right\vert \leq 3\omega \left( G;\delta ;[-M,M]^{n}\right) , \end{equation*}% which means that $F$ is uniformly continuous on $[-M,M]$ and \begin{equation*} \omega \left( F;\delta ;[-M,M]\right) \leq 3\omega \left( G;\delta ;[-M,M]^{n}\right) . \end{equation*}% Note that in the last inequality $\delta $ is a rational number from the interval $\left( 0,\frac{1}{2}\right) .$ It is well known that the modulus of continuity $\omega (f;\delta ;\Omega )$ of a continuous function $f$ is continuous from the right for any compact set $\Omega \subset \mathbb{R}^{n}$ and it is continuous from the left for certain compact sets $\Omega $, in particular for rectangular sets (see \cite{Kol}). It follows immediately that (2.48) is valid for all $\delta \in \lbrack 0,\frac{1}{2}].$ \end{proof} The following theorem was first obtained by Kuleshov \cite{88}. Below, we prove this using completely different ideas. Our proof, which is taken from \cite{2}, contains a theoretical method for constructing the functions $% g_{i}\in C^{s}(\mathbb{R})$ in (2.2). Using this method, we will also estimate the modulus of continuity of $g_{i}$ in terms of the modulus of continuity of $f$ (see Remark 2.4 below). \bigskip \textbf{Theorem 2.9.} \textit{Assume we are given $k$ directions $\mathbf{a}% ^{i}$, $i=1,...,k$, in $\mathbb{R}^{n}\backslash \{\mathbf{0}\}$ and $k-1$ of them are linearly independent.\ Assume that a function $f\in C(\mathbb{R}% ^{n})$ is of the form (2.1). Then $f$ can be represented also in the form (2.2) with $g_{i}\in C(\mathbb{R})$, $i=1,...,k$.} \bigskip \begin{proof} Without loss of generality, we may assume that the first $% k-1 $ vectors $\mathbf{a}^{1},\mathbf{...},\mathbf{a}^{k-1}$ are linearly independent. Thus there exist numbers $\lambda _{1},...,\lambda _{k-1}\in \mathbb{R}$ such that $\mathbf{a}^{k}=\lambda _{1}\mathbf{a}^{1}+\cdot \cdot \cdot +\lambda _{k-1}\mathbf{a}^{k-1}$. We may also assume that the first $p$ numbers $\lambda _{1},...,\lambda _{p}$, $1\leq p\leq k-1$, are nonzero and the remaining $\lambda _{j}$s are zero. Indeed, if necessary, we can rearrange the vectors $\mathbf{a}^{1},\mathbf{...},\mathbf{a}^{k-1}$ so that this assumption holds. Complete the system $\{\mathbf{a}^{1},...,\mathbf{a}% ^{k-1}\}$ to a basis $\{\mathbf{a}^{1},...,\mathbf{a}^{k-1},\mathbf{b}% ^{k},...,\mathbf{b}^{n}\}$ and consider the linear transformation $\mathbf{y}% =A\mathbf{x,}$ where $\mathbf{x}=(x_{1},...,x_{n})^{T},$ $\mathbf{y}% =(y_{1},...,y_{n})^{T}$ and $A$ is the matrix, rows of which are formed by the coordinates of the vectors $\mathbf{a}^{1},...,\mathbf{a}^{k-1},\mathbf{b% }^{k},...,\mathbf{b}^{n}.$ Using this transformation, we can write (2.1) in the form \begin{equation*} f(A^{-1}\mathbf{y})=f_{1}(y_{1})+\cdot \cdot \cdot +f_{k-1}(y_{k-1})+f_{k}(\lambda _{1}y_{1}+\cdot \cdot \cdot +\lambda _{p}y_{p}).\eqno(2.52) \end{equation*} For the brevity of exposition in the sequel, we put $l=k-1$ and use the notation \begin{equation*} w=f_{l+1},\text{ }\Phi (y_{1},...,y_{l})=f(A^{-1}\mathbf{y})\text{ and }% Y_{j}=(y_{1},...,y_{j-1},y_{j+1},...,y_{l}), j=1,...,l. \end{equation*}% Using this notation, we can write (2.52) in the form \begin{equation*} \Phi (y_{1},...,y_{l})=f_{1}(y_{1})+\cdot \cdot \cdot +f_{l}(y_{l})+w(\lambda _{1}y_{1}+\cdot \cdot \cdot +\lambda _{p}y_{p}).\eqno% (2.53) \end{equation*}% In (2.53), taking sequentially $Y_{1}=0,$ $Y_{2}=0,$..., $Y_{l}=0$ we obtain that \begin{equation*} f_{j}(y_{j})=\Phi (y_{1},...,y_{l})|_{Y_{j}=\mathbf{0}}-w(\lambda _{j}y_{j})-\sum_{\substack{ i=1 \\ i\neq j}}^{l}f_{i}(0),\text{ }j=1,...,l.% \eqno(2.54) \end{equation*}% Substituting (2.54) in (2.53), we obtain the equality \begin{equation*} \left. \begin{array}{c} w(\lambda _{1}y_{1}+\cdot \cdot \cdot +\lambda _{p}y_{p})-\sum_{j=1}^{p}w(\lambda _{j}y_{j})-(l-p)w(0)= \\ =\Phi (y_{1},...,y_{l})-\sum_{j=1}^{l}\Phi (y_{1},...,y_{l})|_{Y_{j}=\mathbf{% 0}}+(l-1)\sum_{j=1}^{l}f_{j}(0).% \end{array}% \right. \eqno(2.55) \end{equation*}% We see that the right hand side of (2.55) depends only on the variables $% y_{1},y_{2},...,y_{p}.$ Denote the right hand side of (2.55) by $% H(y_{1},...,y_{p}).$ That is, set% \begin{equation*} H(y_{1},...,y_{p})\overset{def}{=}\Phi (y_{1},...,y_{l})-\sum_{j=1}^{l}\Phi (y_{1},...,y_{l})|_{Y_{j}=\mathbf{0}}+(l-1)\sum_{j=1}^{l}f_{j}(0).\eqno(2.56) \end{equation*}% We will use the following identity, which follows from (2.55) and (2.56)% \begin{equation*} H(y_{1},...,y_{p})=w(\lambda _{1}y_{1}+\cdot \cdot \cdot +\lambda _{p}y_{p})-\sum_{j=1}^{p}w(\lambda _{j}y_{j})-(l-p)w(0).\eqno(2.57) \end{equation*} It follows from (2.56) and the continuity of $f$ that the function $H$ is continuous on $\mathbb{R}^{p}$. Then, defining the function \begin{equation*} G(y_{1},...,y_{p})=H(\frac{y_{1}}{\lambda _{1}},...,\frac{y_{p}}{\lambda _{p}% })+(l-p)w(0)\eqno(2.58) \end{equation*}% and applying Lemma 2.7, we obtain that there exists a function $F\in C(% \mathbb{R})$ such that \begin{equation*} G(y_{1},...,y_{p})=F(y_{1}+\cdot \cdot \cdot +y_{p})-\sum_{j=1}^{p}F(y_{p}).% \eqno(2.59) \end{equation*}% It follows from the formulas (2.57)-(2.59) that \begin{equation*} w(\lambda _{1}y_{1}+\cdot \cdot \cdot +\lambda _{p}y_{p})-\sum_{j=1}^{p}w(\lambda _{j}y_{j})=F(\lambda _{1}y_{1}+\cdot \cdot \cdot +\lambda _{p}y_{p})-\sum_{j=1}^{p}F(\lambda _{j}y_{j}).\eqno% (2.60) \end{equation*} Let us introduce the following functions \begin{equation*} \left. \begin{array}{c} g_{j}(y_{j})=\Phi (y_{1},...,y_{l})|_{Y_{j}=\mathbf{0}}-F(\lambda _{j}y_{j})-\sum_{\substack{ i=1 \\ i\neq j}}^{l}f_{i}(0),\text{ }j=1,...,p, \\ g_{j}(y_{j})=\Phi (y_{1},...,y_{l})|_{Y_{j}=\mathbf{0}}-\sum_{\substack{ i=1 \\ i\neq j}}^{l}f_{i}(0)-w(0),\text{ }j=p+1,...,l.% \end{array}% \right. \eqno(2.61) \end{equation*} Note that $g_{j}\in C(\mathbb{R}),$ $j=1,...,l.$ Considering (2.55), (2.60) and (2.61) it is not difficult to verify that \begin{equation*} \Phi (y_{1},...,y_{l})=g_{1}(y_{1})+\cdot \cdot \cdot +g_{l}(y_{l})+F(\lambda _{1}y_{1}+\cdot \cdot \cdot +\lambda _{p}y_{p}).\eqno% (2.62) \end{equation*}% In (2.62), denoting $F=g_{k}$, recalling the definition of $\Phi $ and going back to the variable $\mathbf{x}=(x_{1},...,x_{n})$ by using again the linear transformation $\mathbf{y}=A\mathbf{x}$, we finally obtain (2.2). \end{proof} \textbf{Remark 2.4.} Using Theorem 2.9 and Lemma 2.7, one can estimate the modulus of continuity of the functions $g_{i}$ in representation (2.2) in terms of the modulus of continuity of $\Phi $. To show how one can do this, assume $C>0,$ $M\geq 1,$ $0\leq \delta \leq \frac{1}{2\max \{\left\vert \lambda _{j}\right\vert \}}$ and introduce the following sets \begin{equation*} \mathcal{M}_{j}=\left\{ (x_{1},...,x_{l})\in \mathbb{R}^{l}:x_{j}\in \lbrack -M/\left\vert \lambda _{j}\right\vert ,M/\left\vert \lambda _{j}\right\vert ]% \text{, }x_{i}=0\text{ for }i\neq j\right\}, \end{equation*}% \begin{equation*} j=1,...,p; \end{equation*}% \begin{equation*} \mathcal{C}_{j}=\left\{ (x_{1},...,x_{l})\in \mathbb{R}^{l}:x_{j}\in \lbrack -C,C]\text{, }x_{i}=0\text{ for }i\neq j\right\} ,\text{ }j=p+1,...,l. \end{equation*}% It can be easily obtained from (2.61) that \begin{equation*} \omega (g_{j};\delta ;[-M/\left\vert \lambda _{j}\right\vert ,M/\left\vert \lambda _{j}\right\vert ])\leq \omega \left( \Phi ;\delta ;\mathcal{M}% _{j}\right) +\omega (F;\delta _{1};[-M,M]),\text{ }j=1,...,p,\eqno(2.63) \end{equation*}% \begin{equation*} \omega (g_{j};\delta ;[-C,C])\leq \omega \left( \Phi ;\delta ;\mathcal{C}% _{j}\right) ,\text{ }j=p+1,...,l,\eqno(2.64) \end{equation*}% where $\delta _{1}=\delta \cdot \max \{\left\vert \lambda _{j}\right\vert \}. $ To estimate $\omega (F;\delta _{1};[-M,M])$ in (2.63), we refer to Lemma 2.7. Applying Lemma 2.7 to the function $G$ in (2.58) we obtain that in addition to (2.59) the following inequality holds. \begin{equation*} \omega (F;\delta _{1};[-M,M])\leq 3\omega \left( G;\delta _{1};[-M,M]^{p}\right) .\eqno(2.65) \end{equation*}% Note that here $0\leq \delta _{1}\leq \frac{1}{2}$ as in Lemma 2.7. It follows from (2.58) and (2.65) that \begin{equation*} \omega (F;\delta _{1};[-M,M])\leq 3\omega \left( H;\delta _{2};\mathcal{K}% \right) ,\eqno(2.66) \end{equation*}% where $\mathcal{K}=[-M/\left\vert \lambda _{1}\right\vert ,M/\left\vert \lambda _{1}\right\vert ]\times \cdot \cdot \cdot \times \lbrack -M/\left\vert \lambda _{p}\right\vert ,M/\left\vert \lambda _{p}\right\vert ] $ and \newline $\delta _{2}=\delta _{1}/\min \{\left\vert \lambda _{j}\right\vert \} $. Further, (2.66) and (2.56) together yield that \begin{equation*} \omega (F;\delta _{1};[-M,M])\leq (3l+3)\omega \left( \Phi ;\delta _{2};% \mathcal{S}\right) ,\eqno(2.67) \end{equation*}% where $\mathcal{S}=\left\{ (x_{1},...,x_{l})\in \mathbb{R}% ^{l}:(x_{1},...,x_{p})\in \mathcal{K}\text{, }x_{i}=0\text{ for }i>p\right\} $. Now it follows from (2.63) and (2.67) that \begin{equation*} \omega (g_{j};\delta ;[-M/\left\vert \lambda _{j}\right\vert ,M/\left\vert \lambda _{j}\right\vert ])\leq \omega \left( \Phi ;\delta ;\mathcal{M}% _{j}\right) +(3l+3)\omega \left( \Phi ;\delta _{2};\mathcal{S}\right) ,\text{ }j=1,...,p.\eqno(2.68) \end{equation*}% Formulas (2.64), (2.67) and (2.68) provide us with upper estimates for the modulus of continuity of the functions $g_{j},$ $j=1,...,k,$ in terms of the modulus of continuity of $\Phi $. Recall that in these estimates $l=k-1$, $% F=g_{k}$ and $\lambda _{j}$ are coefficients\textit{\ }in the expression $% \mathbf{a}^{k}=\lambda _{1}\mathbf{a}^{1}+\cdot \cdot \cdot +\lambda _{p}% \mathbf{a}^{p}$. \bigskip Theorems 2.1 and 2.9 together give the following result. \bigskip \textbf{Theorem 2.10. }\textit{Assume we are given $k$ directions $\mathbf{a}% ^{i}$, $i=1,...,k$, in $\mathbb{R}^{n}\backslash \{\mathbf{0}\}$ and $k-1$ of them are linearly independent.\ Assume that a function $f\in C^{s}(% \mathbb{R}^{n})$ is of the form (2.1). Then $f$ can be represented also in the form (2.2), where the functions $g_{i}\in C^{s}(\mathbb{R})$, $i=1,...,k$% .} \bigskip Indeed, on the one hand, it follows from Theorem 2.9 that $f$ can be expressed as (2.2) with continuous $g_{i}$. On the other hand, since the class $\mathcal{B}$ in Theorem 2.1, in particular, can be taken as $C(% \mathbb{R}),$ it follows that $g_{i}\in C^{s}(\mathbb{R})$. \bigskip \textbf{Remark 2.5.} In addition to the above $C^{s}(\mathbb{R})$, Theorems 2.9 can be restated also for some other subclasses of the space of continuous functions. These are $C^{\infty }(\mathbb{R})$ functions; analytic functions; algebraic polynomials; trigonometric polynomials. More precisely, assume $\mathcal{H}(\mathbb{R})$ is any of these subclasses and $% \mathcal{H}(\mathbb{R}^{n})$ is the $n$-variable analog of the $\mathcal{H}(% \mathbb{R})$. If under the conditions of Theorem 2.9, we have $f\in \mathcal{% H}(\mathbb{R}^{n})$, then this function can be represented in the form (2.2) with $g_{i}\in \mathcal{H}(\mathbb{R}).$ This follows, similarly to the case $C^{s}(\mathbb{R})$ above, from Theorem 2.9 and Remark 2.1. \bigskip \section{A constructive analysis of the smoothness problem} Note that Theorems 2.4-2.10 are generally existence results. They tell about existence of smooth ridge functions $g_{i}$ in the corresponding representation formula (2.2) or (2.24). They are uninformative if we want to construct explicitly these functions. In this section, we give two theorems which do not only address the smoothness problem, but also are useful in constructing the mentioned $g_{i}$% . \subsection{Bivariate case} We start with the constructive analysis of the smoothness problem for bivariate functions. We show that if a bivariate function of a certain smoothness class is represented by a sum of finitely many, arbitrarily behaved ridge functions, then, under suitable conditions, it also can be represented by a sum of ridge functions of the same smoothness class and these ridge functions can be constructed explicitely. \bigskip \textbf{Theorem 2.11.} \textit{Assume $(a_{i},b_{i})$, $i=1,...,n$ \ are pairwise linearly independent vectors in $\mathbb{R}^{2}$. Assume that a function $f\in C^{s}(\mathbb{R}^{2})$ has the form} \begin{equation*} f(x,y)=\sum_{i=1}^{n}f_{i}(a_{i}x+b_{i}y), \end{equation*}% \textit{where $f_{i}$ are arbitrary univariate functions and $s\geq n-2.$ Then $f$ can be represented also in the form} \begin{equation*} f(x,y)=\sum_{i=1}^{n}g_{i}(a_{i}x+b_{i}y),\eqno(2.69) \end{equation*}% \textit{where the functions $g_{i}\in C^{s}(\mathbb{R})$, $i=1,...,n$. In (2.69), the functions $g_{i}$, $i=1,...,n,$ can be constructed by the formulas} \begin{eqnarray*} g_{p} &=&\varphi _{p,n-p-1},\text{ }p=1,...,n-2; \\ g_{n-1} &=&h_{1,n-1};\text{ }g_{n}=h_{2,n-1}. \end{eqnarray*}% \textit{Here all the involved functions $\varphi _{p,n-p-1}$\textit{, }$% h_{1,n-1}$ and $h_{2,n-1}$ can be found inductively as follows}% \begin{eqnarray*} h_{1,1}(t) &=&\frac{\partial ^{n-2}}{\partial l_{1}\cdot \cdot \cdot \partial l_{n-2}}f^{\ast }(t,0),~ \\ h_{2,1}(t) &=&\frac{\partial ^{n-2}}{\partial l_{1}\cdot \cdot \cdot \partial l_{n-2}}f^{\ast }(0,t)-\frac{\partial ^{n-2}}{\partial l_{1}\cdot \cdot \cdot \partial l_{n-2}}f^{\ast }(0,0); \\ h_{1,k+1}(t) &=&\frac{1}{e_{1}\cdot l_{k}}\int_{0}^{t}h_{1,k}(z)dz,\text{ }% k=1,...,n-2; \\ h_{2,k+1}(t) &=&\frac{1}{e_{2}\cdot l_{k}}\int_{0}^{t}h_{2,k}(z)dz,\text{ }% k=1,...,n-2; \end{eqnarray*}% \textit{and} \begin{equation*} \varphi _{p,1}(t)=\frac{\partial ^{n-p-2}f^{\ast }}{\partial l_{p+1}\cdot \cdot \cdot \partial l_{n-2}}\left( \frac{\widetilde{a}_{p}t}{\widetilde{a}% _{p}^{2}+\widetilde{b}_{p}^{2}},\frac{\widetilde{b}_{p}t}{\widetilde{a}% _{p}^{2}+\widetilde{b}_{p}^{2}}\right) -h_{1,p+1}\left( \frac{\widetilde{a}% _{p}t}{\widetilde{a}_{p}^{2}+\widetilde{b}_{p}^{2}}\right) \end{equation*}% \begin{equation*} -h_{2,p+1}\left( \frac{\widetilde{b}_{p}t}{\widetilde{a}_{p}^{2}+\widetilde{b% }_{p}^{2}}\right)-\sum_{j=1}^{p-1}\varphi _{j,p-j+1}\left( \frac{\widetilde{a% }_{j}\widetilde{a}_{p}+\widetilde{b}_{j}\widetilde{b}_{p}}{\widetilde{a}% _{p}^{2}+\widetilde{b}_{p}^{2}}t\right), \end{equation*} \begin{equation*} p=1,...,n-2\left( \text{for }p=n-2\text{, }\frac{\partial ^{n-p-2}f^{\ast }}{% \partial l_{p+1}\cdot \cdot \cdot \partial l_{n-2}}:=f^{\ast }\right) ; \end{equation*} \begin{equation*} \varphi _{p,k+1}(t)=\frac{1}{(\widetilde{a}_{p},\widetilde{b}_{p})\cdot l_{k+p}}\int_{0}^{t}\varphi _{p,k}(z)dz,\text{ }p=1,...,n-3,\text{ }% k=1,...,n-p-2. \end{equation*} \textit{In the above formulas} \begin{equation*} \widetilde{a}_{p}=\frac{a_{p}b_{n}-a_{n}b_{p}}{a_{n-1}b_{n}-a_{n}b_{n-1}};~% \widetilde{b}_{p}=\frac{a_{n-1}b_{p}-a_{p}b_{n-1}}{a_{n-1}b_{n}-a_{n}b_{n-1}}% ,~p=1,...,n-2, \end{equation*} \begin{equation*} l_{p}=\left( \frac{\widetilde{b}_{p}}{\sqrt{\widetilde{a}_{p}^{2}+\widetilde{% b}_{p}^{2}}},\frac{-\widetilde{a}_{p}}{\sqrt{\widetilde{a}_{p}^{2}+% \widetilde{b}_{p}^{2}}}\right) ,\text{ }p=1,...,n-2. \end{equation*} \begin{equation*} f^{\ast }(x,y)=f\left( \frac{b_{n}x-b_{n-1}y}{a_{n-1}b_{n}-a_{n}b_{n-1}},% \frac{a_{n}x-a_{n-1}y}{a_{n}b_{n-1}-a_{n-1}b_{n}}\right) . \end{equation*} \bigskip \begin{proof} Since the vectors $(a_{n-1},b_{n-1})$ and $(a_{n},b_{n})$ are linearly independent, there is a nonsingular linear transformation $% S:(x,y)\rightarrow (x^{^{\prime }},y^{^{\prime }})$ such that $% S:(a_{n-1},b_{n-1})\rightarrow (1,0)$ and $S:(a_{n},b_{n})\rightarrow (0,1).$ Thus, without loss of generality we may assume that the vectors $% (a_{n-1},b_{n-1})$ and $(a_{n},b_{n})$ coincide with the coordinate vectors $% e_{1}=(1,0)$ and $e_{2}=(0,1)$ respectively. Therefore, to prove the first part of the theorem it is enough to show that if a function $f\in C^{s}(% \mathbb{R}^{2})$ is expressed in the form \begin{equation*} f(x,y)=\sum_{i=1}^{n-2}f_{i}(a_{i}x+b_{i}y)+f_{n-1}(x)+f_{n}(y), \end{equation*}% with arbitrary $f_{i}$, then there exist functions $g_{i}$ $\in C^{s}(% \mathbb{R})$, $i=1,...,n$, such that $f$ is also expressed in the form \begin{equation*} f(x,y)=\sum_{i=1}^{n-2}g_{i}(a_{i}x+b_{i}y)+g_{n-1}(x)+g_{n}(y).\eqno(2.70) \end{equation*} By $\Delta _{l}^{(\delta )}F$ we denote the increment of a function $F$ in a direction $l=(l^{\prime },l^{\prime \prime }).$ That is, \begin{equation*} \Delta _{l}^{(\delta )}F(x,y)=F(x+l^{\prime }\delta ,y+l^{\prime \prime }\delta )-F(x,y). \end{equation*}% We also use the notation $\frac{\partial F}{\partial l}$ which denotes the derivative of $F$ in the direction $l$. It is easy to check that the increment of a ridge function $g(ax+by)$ in a direction perpendicular to $(a,b)$ is zero. Let $l_{1},...,l_{n-2}$ be unit vectors perpendicular to the vectors $(a_{1},b_{1}),...,(a_{n-2},b_{n-2})$ correspondingly. Then for any set of numbers $\delta _{1},...,\delta _{n-2}\in \mathbb{R}$ we have \begin{equation*} \Delta _{l_{1}}^{(\delta _{1})}\cdot \cdot \cdot \Delta _{l_{n-2}}^{(\delta _{n-2})}f(x,y)=\Delta _{l_{1}}^{(\delta _{1})}\cdot \cdot \cdot \Delta _{l_{n-2}}^{(\delta _{n-2})}\left[ f_{n-1}(x)+f_{n}(y)\right] .\eqno(2.71) \end{equation*} Denote the left hand side of (2.71) by $S(x,y).$ That is, set% \begin{equation*} S(x,y)\overset{def}{=}\Delta _{l_{1}}^{(\delta _{1})}\cdot \cdot \cdot \Delta _{l_{n-2}}^{(\delta _{n-2})}f(x,y). \end{equation*}% Then from (2.71) it follows that for any real numbers $\delta _{n-1}$and $% \delta _{n}$, \begin{equation*} \Delta _{e_{1}}^{(\delta _{n-1})}\Delta _{e_{2}}^{(\delta _{n})}S(x,y)=0, \end{equation*}% or in expanded form, \begin{equation*} S(x+\delta _{n-1},y+\delta _{n})-S(x,y+\delta _{n})-S(x+\delta _{n-1},y)+S(x,y)=0. \end{equation*}% Putting in the last equality $\delta _{n-1}=-x,$ $\delta _{n}=-y$, we obtain that \begin{equation*} S(x,y)=S(x,0)+S(0,y)-S(0,0). \end{equation*}% This means that \begin{equation*} \Delta _{l_{1}}^{(\delta _{1})}\cdot \cdot \cdot \Delta _{l_{n-2}}^{(\delta _{n-2})}f(x,y) \end{equation*} \begin{equation*} =\Delta _{l_{1}}^{(\delta _{1})}\cdot \cdot \cdot \Delta _{l_{n-2}}^{(\delta _{n-2})}f(x,0)+\Delta _{l_{1}}^{(\delta _{1})}\cdot \cdot \cdot \Delta _{l_{n-2}}^{(\delta _{n-2})}f(0,y)-\Delta _{l_{1}}^{(\delta _{1})}\cdot \cdot \cdot \Delta _{l_{n-2}}^{(\delta _{n-2})}f(0,0). \end{equation*} By the hypothesis of the theorem, the derivative $\frac{\partial ^{n-2}}{% \partial l_{1}\cdot \cdot \cdot \partial l_{n-2}}f(x,y)$ exists at any point $(x,y)\in $ $\mathbb{R}^{2}$. Thus, it follows from the above formula that \begin{equation*} \frac{\partial ^{n-2}f}{\partial l_{1}\cdot \cdot \cdot \partial l_{n-2}}% (x,y)=h_{1,1}(x)+h_{2,1}(y),\eqno(2.72) \end{equation*}% where $h_{1,1}(x)=\frac{\partial ^{n-2}}{\partial l_{1}\cdot \cdot \cdot \partial l_{n-2}}f(x,0)$ and $h_{2,1}(y)=\frac{\partial ^{n-2}}{\partial l_{1}\cdot \cdot \cdot \partial l_{n-2}}f(0,y)-\frac{\partial ^{n-2}}{% \partial l_{1}\cdot \cdot \cdot \partial l_{n-2}}f(0,0)$. Note that $h_{1,1}$ and $h_{2,1}$ belong to the class $C^{s-n+2}(\mathbb{R}).$ By $h_{1,2}$ and $h_{2,2}$ denote the antiderivatives of $h_{1,1}$ and $% h_{2,1}$ satisfying the condition $h_{1,2}(0)=h_{2,2}(0)=0$ and multiplied by the numbers $1/(e_{1}\cdot l_{1})$ and $1/(e_{2}\cdot l_{1})$ correspondingly. That is, \begin{eqnarray*} h_{1,2}(x) &=&\frac{1}{e_{1}\cdot l_{1}}\int_{0}^{x}h_{1,1}(z)dz; \\ h_{2,2}(y) &=&\frac{1}{e_{2}\cdot l_{1}}\int_{0}^{y}h_{2,1}(z)dz. \end{eqnarray*}% Here $e\cdot l$ denotes the scalar product between vectors $e$ and $l$. Obviously, the function \begin{equation*} F_{1}(x,y)=h_{1,2}(x)+h_{2,2}(y) \end{equation*}% obeys the equality \begin{equation*} \frac{\partial F_{1}}{\partial l_{1}}(x,y)=h_{1,1}(x)+h_{2,1}(y).\eqno(2.73) \end{equation*}% From (2.72) and (2.73) we obtain that \begin{equation*} \frac{\partial }{\partial l_{1}}\left[ \frac{\partial ^{n-3}f}{\partial l_{2}\cdot \cdot \cdot \partial l_{n-2}}-F_{1}\right] =0. \end{equation*}% Hence, for some ridge function $\varphi _{1,1}(a_{1}x+b_{1}y),$ \begin{equation*} \frac{\partial ^{n-3}f}{\partial l_{2}\cdot \cdot \cdot \partial l_{n-2}}% (x,y)=h_{1,2}(x)+h_{2,2}(y)+\varphi _{1,1}(a_{1}x+b_{1}y).\eqno(2.74) \end{equation*}% Here all the functions $h_{2,1},h_{2,2}(y),\varphi _{1,1}\in C^{s-n+3}(% \mathbb{R}).$ Set the following functions \begin{eqnarray*} h_{1,3}(x) &=&\frac{1}{e_{1}\cdot l_{2}}\int_{0}^{x}h_{1,2}(z)dz; \\ h_{2,3}(y) &=&\frac{1}{e_{2}\cdot l_{2}}\int_{0}^{y}h_{2,2}(z)dz; \\ \varphi _{1,2}(t) &=&\frac{1}{(a_{1},b_{1})\cdot l_{2}}\int_{0}^{t}\varphi _{1,1}(z)dz. \end{eqnarray*}% Note that the function \begin{equation*} F_{2}(x,y)=h_{1,3}(x)+h_{2,3}(y)+\varphi _{1,2}(a_{1}x+b_{1}y) \end{equation*}% obeys the equality \begin{equation*} \frac{\partial F_{2}}{\partial l_{2}}(x,y)=h_{1,2}(x)+h_{2,2}(y)+\varphi _{1,1}(a_{1}x+b_{1}y).\eqno(2.75) \end{equation*}% From (2.74) and (2.75) it follows that \begin{equation*} \frac{\partial }{\partial l_{2}}\left[ \frac{\partial ^{n-4}f}{\partial l_{3}\cdot \cdot \cdot \partial l_{n-2}}-F_{2}\right] =0. \end{equation*}% The last equality means that for some ridge function $\varphi _{2,1}(a_{2}x+b_{2}y),$ \begin{equation*} \frac{\partial ^{n-4}f}{\partial l_{3}\cdot \cdot \cdot \partial l_{n-2}}% (x,y)=h_{1,3}(x)+h_{2,3}(y)+\varphi _{1,2}(a_{1}x+b_{1}y)+\varphi _{2,1}(a_{2}x+b_{2}y).\eqno(2.76) \end{equation*}% Here all the functions $h_{1,3},$ $h_{2,3},$ $\varphi _{1,2},$ $\varphi _{2,1}\in C^{s-n+4}(\mathbb{R}).$ Note that in the left hand sides of (2.72), (2.74) and (2.76) we have the mixed directional derivatives of $f$ and the order of these derivatives is decreased by one in each consecutive step. Continuing the above process, until it reaches the function $f$, we obtain the desired representation (2.70). The formulas for $g_{i}$ are obtained in the process of the above proof. These formulas involve certain functions which can be found inductively as described in the proof. The validity of the formulas for the functions $% h_{1,k}$ and $h_{2,k}$, $k=1,...,n-1,$ is obvious. The formulas for $\varphi _{p,1}$ and $\varphi _{p,k+1}$ can be obtained from (2.74), (2.76) and the subsequent (assumed but not written) equations if we put $x=\widetilde{a}% _{p}t/(\widetilde{a}_{p}^{2}+\widetilde{b}_{p}^{2})$ and $y=\widetilde{b}% _{p}t/(\widetilde{a}_{p}^{2}+\widetilde{b}_{p}^{2})$. Note that $(\widetilde{% a}_{p},\widetilde{b}_{p}),$ $p=1,...,n-2,$ are the images of vectors $% (a_{p},b_{p})$ under the linear transformation $S$ which takes the vectors $% (a_{n-1},b_{n-1})$ and $(a_{n},b_{n})$ to the coordinate vectors $e_{1}=(1,0)$ and $e_{2}=(0,1),$ respectively. Besides, note that for $p=1,...,n-2,$ the vectors $l_{p}$ are perpendicular to the vectors $(\widetilde{a}_{p},% \widetilde{b}_{p})$, respectively and $f^{\ast }$ is the function generated from $f$ by the above liner transformation. \end{proof} Theorem 2.11 can be applied to some higher order partial differential equations in two variables, e.g., to the following homogeneous equation \begin{equation*} \prod\limits_{i=1}^{r}\left( \alpha _{i}\frac{\partial }{\partial x}+\beta _{i}\frac{\partial }{\partial y}\right) u(x,y)=0,\eqno(2.77) \end{equation*}% where $(\alpha _{i},\beta _{i}),~i=1,...,r,$ are pairwise linearly independent vectors in $\mathbb{R}^{2}$. Clearly, the general solution to this equation are all functions of the form \begin{equation*} u(x,y)=\sum\limits_{i=1}^{r}v_{i}(\beta _{i}x-\alpha _{i}y),\eqno(2.78) \end{equation*}% where $v_{i}\in C^{r}(\mathbb{R})$, $i=1,...,r$. Based on Theorem 2.11, for the general solution, one can demand only smoothness of the sum $u$ and dispense with smoothness of the summands $v_{i}$. More precisely, the following corollary is valid. \bigskip \textbf{Corollary 2.1.} \textit{Assume a function $u\in C^{r}(\mathbb{R}% ^{2}) $ is of the form (2.78) with arbitrarily behaved $v_{i}$. Then $u$ is a solution to Equation (2.77).} \bigskip \textbf{Remark 2.6.} If in Theorem 2.11 $s\geq n-1,$ then the functions $% g_{i}$, $i=1,...,n,$ can be constructed (up to polynomials) by the method discussed in Buhmann and Pinkus \cite{12}. This method is based on the fact that for a direction $\mathbf{c}=(c_{1},...,c_{m})$ orthogonal to a given direction $\mathbf{a}\in \mathbb{R}^{m}\backslash \{\mathbf{0}\},$ the operator \begin{equation*} D_{\mathbf{c}}=\sum_{k=1}^{m}c_{k}\frac{\partial }{\partial x_{k}} \end{equation*}% acts on $m$-variable ridge functions $g(\mathbf{a}\cdot \mathbf{x})$ as follows \begin{equation*} D_{\mathbf{c}}g(\mathbf{a}\cdot \mathbf{x})=\left( \mathbf{c}\cdot \mathbf{a}% \right) g^{\prime }(\mathbf{a}\cdot \mathbf{x}). \end{equation*}% Thus, if in our case for fixed $r\in \{1,...,n\},$ vectors $l_{k},$ $k\in \{1,...,n\}$, $k\neq r$, are perpendicular to the vectors $(a_{k},b_{k})$, then \begin{equation*} \prod\limits_{\substack{ k=1 \\ k\neq r}}^{n}D_{l_{k}}f(x,y)=\prod\limits _{\substack{ k=1 \\ k\neq r}}^{n}D_{l_{k}}\sum_{i=1}^{n}g_{i}(a_{i}x+b_{i}y) \end{equation*}% \begin{equation*} =\sum_{i=1}^{n}\left( \prod\limits _{\substack{ k=1 \\ k\neq r}}^{n}\left( (a_{i},b_{i})\cdot l_{k}\right) \right) g_{i}^{(n-1)}(a_{i}x+b_{i}y)=\prod\limits_{\substack{ k=1 \\ k\neq r}}% ^{n}\left( (a_{r},b_{r})\cdot l_{k}\right) g_{r}^{(n-1)}(a_{r}x+b_{r}y). \end{equation*}% Now $g_{r}$ can be easily constructed from the above formula (up to a polynomial of degree at most $n-2$). Note that this method is not feasible if in Theorem 2.11 the function $f$ is of the class $C^{n-2}(\mathbb{R}^{2})$% . \bigskip \subsection{Multivariate case} In this subsection, we generalize ideas from the previous subsection to prove constructively that if a multivariate function of a certain smoothness class is represented by a sum of $k$ arbitrarily behaved ridge functions, then, under suitable conditions, it can be represented by a sum of ridge functions of the same smoothness class and some polynomial of a certain degree. The appearance of a polynomial term is mainly related to the fact that in $\mathbb{R}^{n}$ ($n\geq 3)$ there are many directions orthogonal to a given direction. Such a result was proved nonconstructively in Section 2.1.5 (see Theorem 2.5), but here under a mild hypothesis on the degree of smoothness, we give a new proof for this theorem, which will provide us with a recipe for constructing the functions $g_{i}$ in (2.24). The following theorem is valid. \bigskip \textbf{Theorem 2.12.} \textit{Assume $f\in C^{s}(\mathbb{R}^{n})$ is of the form (2.1). Let $s\geq k-p+1,$ where $p$ is the number of vectors $\mathbf{a}% ^{i}$ forming a maximal linearly independent system. Then there exist functions $g_{i}\in C^{s}(\mathbb{R})$ and a polynomial $P(\mathbf{x})$ of total degree at most $k-p+1$ such that (2.24) holds and $g_{i}$ can be constructed algorithmically.} \bigskip \begin{proof} We start the proof by choosing a maximal linearly independent system in $\{\mathbf{a}^{1},....,\mathbf{a}^{k}\}$. The case when the system $\{\mathbf{a}^{1},....,\mathbf{a}^{k}\}$ itself is linearly independent is obvious (see Section 2.1.1). Thus we omit this special case here. Without loss of generality we may assume that the first $p$ vectors $% \mathbf{a}^{1},....,\mathbf{a}^{p}$, $p<k$, are linearly independent. Thus, the vectors $\mathbf{a}^{j},$ $j=p+1,...,k,$ can be expressed as linear combinations $\lambda _{1}^{j}\mathbf{a}^{1}+\cdot \cdot \cdot +\lambda _{p}^{j}\mathbf{a}^{p}$, where $\lambda _{1}^{j},...,\lambda _{p}^{j}$ are real numbers. In addition, we can always apply a nonsingular linear transformation $S$ of the coordinates such that $S:\mathbf{a}^{i}\rightarrow \mathbf{e}_{i},$ $i=1,...,p,$ where $\mathbf{e}_{i}$ denotes the $i$-th unit vector. This reduces the initial representation (2.1) to the following simpler form \begin{equation*} f(\mathbf{x})=f_{1}(x_{1})+\cdot \cdot \cdot +f_{p}(x_{p})+\sum_{i=1}^{m}f_{p+i}(\mathbf{a}^{i}\cdot \mathbf{x}).\eqno% (2.79) \end{equation*}% Note that we keep the notation of (2.1), but here $\mathbf{x}% =(x_{1},...,x_{p}),$ $\mathbf{a}^{i}=(\lambda _{1}^{i}\mathbf{,...,}\lambda _{p}^{i})\in \mathbb{R}^{p}$ and $m=k-p.$ Obviously, we prove Theorem 2.12 if we prove it for the representation (2.79). Thus, in the sequel, we prove that if $f\in C^{s}(\mathbb{R}^{n})$ is of the form (2.79) and $s\geq m+1,$ then there exist functions $g_{i}\in C^{s}(\mathbb{R})$ and a polynomial $P(% \mathbf{x})$ of total degree at most $m+1$ such that \begin{equation*} f(\mathbf{x})=g_{1}(x_{1})+\cdot \cdot \cdot +g_{p}(x_{p})+\sum_{i=1}^{m}g_{p+i}(\mathbf{a}^{i}\cdot \mathbf{x})+P(% \mathbf{x}). \end{equation*} In the process of the proof, we also see how these $g_{i}$ are constructed. For each $i=1,...,m,$ let $\{\mathbf{e}_{1}^{(i)},...,\mathbf{e}% _{p-1}^{(i)}\}$ denote an orthonormal basis in the hyperplane perpendicular to $\mathbf{a}^{i}.$ By $\Delta _{\mathbf{e}}^{(\delta )}F$ we denote the increment of a function $F$ in a direction $\mathbf{e}$ of length $\delta .$ That is, \begin{equation*} \Delta _{\mathbf{e}}^{(\delta )}F(\mathbf{x})=F(\mathbf{x}+\delta \mathbf{e}% )-F(\mathbf{x}). \end{equation*}% We also use the notation $\frac{\partial F}{\partial \mathbf{e}}$ to denote the derivative of $F$ in a direction $\mathbf{e}$. It is easy to check that the increment of a ridge function $g(\mathbf{a\cdot x})$ in any direction perpendicular to $\mathbf{a}$ is zero. For example, \begin{equation*} \Delta _{\mathbf{e}_{j}^{(i)}}^{(\delta )}g(\mathbf{a}^{i}\mathbf{\cdot x)}% =0, \end{equation*}% for all $i=1,...,m,$ $j=1,...,p-1.$ Therefore, for any indices $% i_{1},...,i_{m}\in \{1,...,p-1\}$, $q\in \{1,...,p\}$ and numbers $\delta _{1},...,\delta _{m},\delta \in \mathbb{R}$ we have the formula \begin{equation*} \Delta _{\mathbf{e}_{i_{1}}^{(1)}}^{(\delta _{1})}\Delta _{\mathbf{e}% _{i_{2}}^{(2)}}^{(\delta _{2})}\cdot \cdot \cdot \Delta _{\mathbf{e}% _{i_{m}}^{(m)}}^{(\delta _{m})}\Delta _{\mathbf{e}_{q}}^{(\delta )}f(\mathbf{% x})=\Delta _{\mathbf{e}_{i_{1}}^{(1)}}^{(\delta _{1})}\Delta _{\mathbf{e}% _{i_{2}}^{(2)}}^{(\delta _{2})}\cdot \cdot \cdot \Delta _{\mathbf{e}% _{i_{m}}^{(m)}}^{(\delta _{m})}\Delta _{\mathbf{e}_{q}}^{(\delta )}f_{q}(x_{q}), \end{equation*}% where $\mathbf{e}_{q}$ denotes the $q$-th unit vector. This means that for each $q=1,...,p,$ the mixed directional derivative \begin{equation*} \frac{\partial ^{m+1}f}{\partial \mathbf{e}_{i_{1}}^{(1)}\cdot \cdot \cdot \partial \mathbf{e}_{i_{m}}^{(m)}\partial x_{q}}(\mathbf{x}) \end{equation*}% depends only on the variable $x_{q}.$ Denote this derivative by $% h_{i_{1},...,i_{m}}^{0,q}(x_{q})$: \begin{equation*} h_{i_{1},...,i_{m}}^{0,q}(x_{q})=\frac{\partial ^{m+1}f}{\partial \mathbf{e}% _{i_{1}}^{(1)}\cdot \cdot \cdot \partial \mathbf{e}_{i_{m}}^{(m)}\partial x_{q}}(\mathbf{x}),\text{ }q=1,...,p.\eqno(2.80) \end{equation*}% Since $f\in C^{s}(\mathbb{R}^{p}),$ we obtain that $% h_{i_{1},...,i_{m}}^{0,q}\in C^{s-m-1}(\mathbb{R}).$ It follows from (2.80) that \begin{equation*} d\left( \frac{\partial ^{m}f}{\partial \mathbf{e}_{i_{1}}^{(1)}\cdot \cdot \cdot \partial \mathbf{e}_{i_{m}}^{(m)}}\right) =h_{i_{1},...,i_{m}}^{0,1}(x_{1})dx_{1}+\cdot \cdot \cdot +h_{i_{1},...,i_{m}}^{0,p}(x_{p})dx_{p}.\eqno(2.81) \end{equation*}% We conclude from (2.81) that \begin{equation*} \frac{\partial ^{m}f}{\partial \mathbf{e}_{i_{1}}^{(1)}\cdot \cdot \cdot \partial \mathbf{e}_{i_{m}}^{(m)}}(\mathbf{x}% )=h_{i_{1},...,i_{m}}^{1,1}(x_{1})+\cdot \cdot \cdot +h_{i_{1},...,i_{m}}^{1,p}(x_{p})+c_{i_{1},...,i_{m}},\eqno(2.82) \end{equation*}% where the functions $h_{i_{1},...,i_{m}}^{1,q}(x_{q})$, $q=1,...,p,$ are antiderivatives of $h_{i_{1},...,i_{m}}^{0,q}(x_{q})$ satisfying the condition $h_{i_{1},...,i_{m}}^{1,q}(0)=0$ and $c_{i_{1},...,i_{m}}$ is a constant. Note that $h_{i_{1},...,i_{m}}^{1,q}\in C^{s-m}(\mathbb{R}),$ $% q=1,...,p.$ Obviously, for any pair $k,t\in \{1,...,p-1\}$, \begin{equation*} \frac{\partial ^{m+1}f}{\partial \mathbf{e}_{i_{1}}^{(1)}\cdot \cdot \cdot \partial \mathbf{e}_{i_{m-1}}^{(m-1)}\partial \mathbf{e}_{k}^{(m)}\partial \mathbf{e}_{t}^{(m)}}=\frac{\partial ^{m+1}f}{\partial \mathbf{e}% _{i_{1}}^{(1)}\cdot \cdot \cdot \partial \mathbf{e}_{i_{m-1}}^{(m-1)}% \partial \mathbf{e}_{t}^{(m)}\partial \mathbf{e}_{k}^{(m)}}\eqno(2.83) \end{equation*}% It follows from (2.82) and (2.83) that \begin{equation*} (\mathbf{e}_{q}\cdot \mathbf{e}_{k}^{(m)})\left( h_{i_{1},...,i_{m-1},t}^{1,q}\right) ^{^{\prime }}(x_{q})=(\mathbf{e}% _{q}\cdot \mathbf{e}_{t}^{(m)})\left( h_{i_{1},...,i_{m-1},k}^{1,q}\right) ^{^{\prime }}(x_{q})+c, \end{equation*}% where $c$ is a constant depending on the parameters $i_{1},...,i_{m-1},k,t$ and $q.$ Recall that by construction, $h_{i_{1},...,i_{m}}^{1,q}(0)=0.$ Hence \begin{equation*} (\mathbf{e}_{q}\cdot \mathbf{e}% _{k}^{(m)})h_{i_{1},...,i_{m-1},t}^{1,q}(x_{q})=(\mathbf{e}_{q}\cdot \mathbf{% e}_{t}^{(m)})h_{i_{1},...,i_{m-1},k}^{1,q}(x_{q})+cx_{q}.\eqno(2.84) \end{equation*} Since for each $q=1,...,p,$ the vectors $\mathbf{e}_{q}$ and $\mathbf{a}^{m}$ are linearly independent, there exists an index $i_{m}(q)\in \{1,...,p-1\}$ such that the vector $\mathbf{e}_{i_{m}(q)}^{(m)}$ is not orthogonal to $% \mathbf{e}_{q}.$ That is, $\mathbf{e}_{q}\cdot \mathbf{e}_{i_{m}(q)}^{(m)}$ $% \neq 0.$ For each $q=1,...,p,$ fix the index $i_{m}(q)$ and define the following functions \begin{equation*} h_{i_{1},...,i_{m-1}}^{2,q}(x_{q})=\frac{1}{\mathbf{e}_{q}\cdot \mathbf{e}% _{i_{m}(q)}^{(m)}}\int_{0}^{x_{q}}h_{i_{1},...,i_{m-1},i_{m}(q)}^{1,q}(z)dz.% \eqno(2.85) \end{equation*}% and \begin{equation*} F_{i_{1},...,i_{m-1}}(\mathbf{x})=h_{i_{1},...,i_{m-1}}^{2,1}(x_{1})+\cdot \cdot \cdot +h_{i_{1},...,i_{m-1}}^{2,p}(x_{p}). \end{equation*} \bigskip It is easy to obtain from (2.84) and (2.85) that for any $i_{m}\in \{1,...,p-1\},$% \begin{equation*} \frac{\partial F_{i_{1},...,i_{m-1}}}{\partial \mathbf{e}_{i_{m}}^{(m)}}(% \mathbf{x})=h_{i_{1},...,i_{m}}^{1,1}(x_{1})+\cdot \cdot \cdot +h_{i_{1},...,i_{m}}^{1,p}(x_{p})+P_{i_{1},...,i_{m}}^{(1)},\eqno(2.86) \end{equation*}% where $P_{i_{1},...,i_{m}}^{(1)}$ is a polynomial of total degree not greater than $1$. It follows from (2.82) and (2.86) that \begin{equation*} \frac{\partial }{\partial \mathbf{e}_{i_{m}}^{(m)}}\left[ \frac{\partial ^{m-1}f}{\partial \mathbf{e}_{i_{1}}^{(1)}\cdot \cdot \cdot \partial \mathbf{% e}_{i_{m-1}}^{(m-1)}}-F_{i_{1},...,i_{m-1}}\right] (\mathbf{x}% )=c_{i_{1},...,i_{m}}-P_{i_{1},...,i_{m}}^{(1)}(\mathbf{x}).\eqno(2.87) \end{equation*}% Note that the last equality is valid for all vectors $\mathbf{e}% _{i_{m}}^{(m)},$ which form a basis in the hyperplane orthogonal to $\mathbf{% a}^{m}$. Thus from (2.87) we conclude that the following expansion is valid \begin{equation*} \left. \begin{array}{c} \frac{\partial ^{m-1}f}{\partial \mathbf{e}_{i_{1}}^{(1)}\cdot \cdot \cdot \partial \mathbf{e}_{i_{m-1}}^{(m-1)}}(\mathbf{x}% )=h_{i_{1},...,i_{m-1}}^{2,1}(x_{1})+\cdot \cdot \cdot +h_{i_{1},...,i_{m-1}}^{2,p}(x_{p}) \\ +\varphi _{i_{1},...,i_{m-1}}^{2,1}(\mathbf{a}^{m}\cdot \mathbf{x}% )+P_{i_{1},...,i_{m-1}}^{(2)}(\mathbf{x}).% \end{array}% \right. \eqno(2.88) \end{equation*}% Here all the functions $% h_{i_{1},...,i_{m-1}}^{2,1},...,h_{i_{1},...,i_{m-1}}^{2,p}(x_{p}),\varphi _{i_{1},...,i_{m-1}}^{2,1}\in C^{s-m+1}(\mathbb{R})$ and $% P_{i_{1},...,i_{m-1}}^{(2)}$ is a polynomial of total degree not greater than $2.$ Since for each $q=1,...,p,$ the vector $\mathbf{e}_{q}$ is not collinear to $% \mathbf{a}^{m-1},$ there is an index $i_{m-1}(q)\in \{1,...,p-1\}$ such that $\mathbf{e}_{i_{m-1}(q)}^{(m-1)}$ is not orthogonal to $\mathbf{e}_{q}$. Similarly, since $\mathbf{a}^{m-1}$ is not collinear to $\mathbf{a}^{m},$ there is an index $i_{m-1}(m)\in \{1,...,p-1\}$ such that $\mathbf{e}% _{i_{m-1}(m)}^{(m-1)}$ is not orthogonal to $\mathbf{a}^{m}$. Fix the indices $i_{m-1}(q)$, $i_{m-1}(m)$ and consider the following functions \begin{equation*} h_{i_{1},...,i_{m-2}}^{3,q}(x_{q})=\frac{1}{\mathbf{e}_{q}\cdot \mathbf{e}% _{i_{m-1}(q)}^{(m-1)}}% \int_{0}^{x_{q}}h_{i_{1},...,i_{m-2},i_{m-1}(q)}^{2,q}(z)dz,\text{ }% q=1,...,p,\eqno(2.89) \end{equation*} \begin{equation*} \varphi _{i_{1},...,i_{m-2}}^{3,1}(t)=\frac{1}{\mathbf{a}^{m}\cdot \mathbf{e}% _{i_{m-1}(m)}^{(m-1)}}\int_{0}^{t}\varphi _{i_{1},...,i_{m-2},i_{m-1}(m)}^{2,1}(z)dz,\eqno(2.90) \end{equation*}% and \begin{equation*} F_{i_{1},...,i_{m-2}}(\mathbf{x})=h_{i_{1},...,i_{m-2}}^{3,1}(x_{1})+\cdot \cdot \cdot +h_{i_{1},...,i_{m-2}}^{3,p}(x_{p})+\varphi _{i_{1},...,i_{m-2}}^{3,1}(\mathbf{a}^{m}\cdot \mathbf{x}).\eqno(2.91) \end{equation*} Similar to (2.84), one can easily verify that for any pair $k,t\in \{1,...,p-1\}$ and for all $q=1,...,p,$ the following equalities are valid. \begin{equation*} \left. \begin{array}{c} (\mathbf{e}_{q}\cdot \mathbf{e}% _{k}^{(m-1)})h_{i_{1},...,i_{m-2},t}^{2,q}(x_{q})=(\mathbf{e}_{q}\cdot \mathbf{e}_{t}^{(m-1)})h_{i_{1},...,i_{m-2},k}^{2,q}(x_{q})+H_{q}(x_{q}), \\ (\mathbf{a}^{m}\cdot \mathbf{e}_{k}^{(m-1)})\varphi _{i_{1},...,i_{m-2},t}^{2,1}(\mathbf{a}^{m}\cdot \mathbf{x})=(\mathbf{a}% ^{m}\cdot \mathbf{e}_{t}^{(m-1)})\varphi _{i_{1},...,i_{m-2},k}^{2,1}(% \mathbf{a}^{m}\cdot \mathbf{x})+\Phi (\mathbf{x}),% \end{array}% \right. \eqno(2.92) \end{equation*}% where $H_{q}$ and $\Phi $ are univariate and $n$-variable polynomials of degree not greater than $2.$ Indeed, applying the Schwarz formula \begin{equation*} \frac{\partial ^{m+1}f}{\partial \mathbf{e}_{i_{1}}^{(1)}\cdot \cdot \cdot \partial \mathbf{e}_{i_{m-2}}^{(m-2)}\partial \mathbf{e}_{k}^{(m-1)}\partial \mathbf{e}_{t}^{(m-1)}\partial \mathbf{e}_{i_{m}(q)}^{(m)}}=\frac{\partial ^{m+1}f}{\partial \mathbf{e}_{i_{1}}^{(1)}\cdot \cdot \cdot \partial \mathbf{% e}_{i_{m-2}}^{(m-2)}\partial \mathbf{e}_{t}^{(m-1)}\partial \mathbf{e}% _{k}^{(m-1)}\partial \mathbf{e}_{i_{m}(q)}^{(m)}} \end{equation*}% on the symmetry of derivatives, it follows from (2.82) that for any $k,t\in \{1,...,p-1\}$ \begin{equation*} (\mathbf{e}_{q}\cdot \mathbf{e}_{k}^{(m-1)})\left( h_{i_{1},...,i_{m-2},t,i_{m}(q)}^{1,q}\right) ^{^{\prime }}(x_{q})=(\mathbf{e% }_{q}\cdot \mathbf{e}_{t}^{(m-1)})\left( h_{i_{1},...,i_{m-2},k,i_{m}(q)}^{1,q}\right) ^{^{\prime }}(x_{q})+d, \end{equation*}% where $d$ is a constant depending on the parameters $i_{1},...,i_{m-2},k,t$ and $i_{m}(q).$ Since, by construction, $h_{i_{1},...,i_{m}}^{1,q}(0)=0,$ we obtain that \begin{equation*} (\mathbf{e}_{q}\cdot \mathbf{e}% _{k}^{(m-1)})h_{i_{1},...,i_{m-2},t,i_{m}(q)}^{1,q}(x_{q})=(\mathbf{e}% _{q}\cdot \mathbf{e}% _{t}^{(m-1)})h_{i_{1},...,i_{m-2},k,i_{m}(q)}^{1,q}(x_{q})+dx_{q}. \end{equation*}% The last equality together with (2.85) yield that \begin{equation*} (\mathbf{e}_{q}\cdot \mathbf{e}_{k}^{(m-1)})\left( h_{i_{1},...,i_{m-2},t}^{2,q}\right) ^{^{\prime }}(x_{q})=(\mathbf{e}% _{q}\cdot \mathbf{e}_{t}^{(m-1)})\left( h_{i_{1},...,i_{m-2},k}^{2,q}\right) ^{^{\prime }}(x_{q})+dx_{q}. \end{equation*}% Therefore, the first equality in (2.92) holds. Considering this and applying the corresponding Schwarz formula to (2.88) we obtain the second equality in (2.92). Taking into account the definitions (2.89), (2.90) and the relations (2.92), we obtain from (2.91) that for any $i_{m-1}\in \{1,...,p-1\},$% \begin{equation*} \frac{\partial F_{i_{1},...,i_{m-2}}}{\partial \mathbf{e}_{i_{m-1}}^{(m-1)}}(% \mathbf{x}) \end{equation*} \begin{equation*} =h_{i_{1},...,i_{m-1}}^{2,1}(x_{1})+\cdot \cdot \cdot +h_{i_{1},...,i_{m-1}}^{2,p}(x_{p})+\varphi _{i_{1},...,i_{m-1}}^{2,1}(% \mathbf{a}^{m}\cdot \mathbf{x})+\widetilde{P}_{i_{1},...,i_{m-1}}^{(2)}(% \mathbf{x}),\eqno(2.93) \end{equation*}% where $\widetilde{P}_{i_{1},...,i_{m-1}}^{(2)}$ is a polynomial of degree not greater than $2.$ It follows from (2.88) and (2.93) that \begin{equation*} \frac{\partial }{\partial \mathbf{e}_{i_{m-1}}^{(m-1)}}\left[ \frac{\partial ^{m-2}f}{\partial \mathbf{e}_{i_{1}}^{(1)}\cdot \cdot \cdot \partial \mathbf{% e}_{i_{m-2}}^{(m-2)}}-F_{i_{1},...,i_{m-2}}\right] (\mathbf{x}% )=P_{i_{1},...,i_{m-1}}^{(2)}(\mathbf{x})-\widetilde{P}% _{i_{1},...,i_{m-1}}^{(2)}(\mathbf{x}).\eqno(2.94) \end{equation*}% Note that the last equality is valid for all vectors $\mathbf{e}% _{i_{m-1}}^{(m-1)},$ which form a basis in the hyperplane orthogonal to $% \mathbf{a}^{m-1}$. Considering this, from (2.94) we derive the following representation \begin{equation*} \left. \begin{array}{c} \frac{\partial ^{m-2}f}{\partial \mathbf{e}_{i_{1}}^{(1)}\cdot \cdot \cdot \partial \mathbf{e}_{i_{m-2}}^{(m-2)}}(\mathbf{x}% )=h_{i_{1},...,i_{m-2}}^{3,1}(x_{1})+\cdot \cdot \cdot +h_{i_{1},...,i_{m-2}}^{3,p}(x_{p}) \\ +\varphi _{i_{1},...,i_{m-2}}^{3,1}(\mathbf{a}^{m}\cdot \mathbf{x})+\varphi _{i_{1},...,i_{m-2}}^{3,2}(\mathbf{a}^{m-1}\cdot \mathbf{x}% )+P_{i_{1},...,i_{m-2}}^{(3)}(\mathbf{x}).% \end{array}% \right. \eqno(2.95) \end{equation*}% Here all the functions \newline $% h_{i_{1},...,i_{m-2}}^{3,1},...,h_{i_{1},...,i_{m-2}}^{3,p}(x_{p}),$ $% \varphi _{i_{1},...,i_{m-2}}^{3,1},$ $\varphi _{i_{1},...,i_{m-2}}^{3,2}\in C^{s-m+2}(\mathbb{R})$ and $P_{i_{1},...,i_{m-2}}^{(3)}$ is a polynomial of total degree not greater than $3.$ Note that in the left hand sides of (2.82), (2.88) and (2.95) we have the mixed directional derivatives of $f$ and the order of these derivatives is decreased by one at each consecutive step. Continuing the above process, until it reaches the function $f$, we obtain the desired representation. Note that the above proof gives a recipe for constructing the smooth ridge functions $g_{i}$. Writing out explicit recurrent formulas for $g_{i}$, as in Theorem 2.11, is technically cumbersome here and hence is avoided. \end{proof} \textbf{Remark 2.7.} Note that using Theorem 2.12, the degree of polynomial $% P(\mathbf{x})$ in Theorem 2.5 can be reduced. Indeed, it follows from (2.27) and (2.28) that the the above polynomial $P(\mathbf{x})$ is of the form (2.1). On the other hand, by Theorem 2.12 there exist functions $g_{i}^{\ast }\in C^{s}(\mathbb{R})$, $i=1,...,k$, and a polynomial $G(\mathbf{x})$ of degree at most $k-p+1$ such that \begin{equation*} P(\mathbf{x})=\sum_{i=1}^{k}g_{i}^{\ast }(\mathbf{a}^{i}\cdot \mathbf{x})+G(% \mathbf{x}). \end{equation*}% Now considering this in (2.24) we see that our assertion is true. \bigskip At the end of this chapter, we want to draw the reader's attention to the following uniqueness question. Assume we are given pairwise linearly independent vectors $\mathbf{a}^{i},$ $i=1,...,k,$ in $\mathbb{R}^{n}$ and a function $f:\mathbb{R}^{n}\rightarrow \mathbb{R}$ of the form (2.1). How many different ways can $f$ be written as a sum of ridge functions with the directions $\mathbf{a}^{i}$? Clearly, representation (2.1) is not unique, since we can always add some constants $c_{i}$ to $f_{i}$ without changing the resulting sum in (2.1) provided that $\sum_{i=1}^{k}c_{i}=0$. It turns out that under minimal requirements representation (2.1) is unique up to polynomials of degree at most $k-2$. More precisely, if, in addition to (2.1), $% f$ also has the form (2.2) and $% f_{i},g_{i}\in \mathcal{B}$, $i=1,...,k$, then the functions $f_{i}-g_{i}$ are univariate polynomials of degree at most $k-2$. This result is due to Pinkus \cite[Theorem 3.1]{117}. It follows immediately from this result that in Theorems 2.6--2.10 the functions $g_{i}$ is unique up to a univariate polynomial. This is also valid for $g_{i}$ in Theorems 2.4 and 2.5, but in this case for the proof we must apply a slightly different result of Pinkus \cite[Corollary 3.2]{117}: Assume a multivariate polynomial $f$ of degree $m$ is of the form (2.1) and $f_{i}\in \mathcal{B}$ for $i=1,...,k$. Then $f_{i}$ are univariate polynomials of degree at most $% l=\max \left\{ m,k-2\right\} $. A different uniqueness problem, in a more general setting, will be analyzed in Chapter 4. In that problem we will look for sets $Q\subset \mathbb{R}^{n}$ for which representation (2.1), considered on $Q$, is unique. \newpage \chapter{Approximation of multivariate functions by sums of univariate functions} It is clear that in the special case, when directions of ridge functions coincide with the coordinate directions, the problem of approximation by linear combinations of these functions turn into the problem of approximation by sums of univariate functions. This is also the simplest case in ridge function approximation. The simplicity of the approximation guarantees its practicability in application areas, where complicated multivariate functions are main obstacles. In mathematics, this type of approximation has arisen, for example, in connection with the classical functional equations \cite{11}, the numerical solution of certain PDE boundary value problems \cite{9}, dimension theory \cite{133,132}, etc. In this chapter, we obtain some results concerning the problem of best approximation by sums of univariate functions. Most of the material of this chapter is taken from \cite{59,55,56,48}. \bigskip \section{Characterization of some bivariate function classes by formulas for the error of approximation} This section is devoted to calculation formulas for the error of approximation of bivariate functions by sums of univariate functions. Certain classes of bivariate functions depending on some numerical parameter are constructed and characterized in terms of the approximation error calculation formulas. \subsection{Exposition of the problem} The approximation problem considered here is to approximate a continuous and real-valued function of two variables by sums of two continuous functions of one variable. To make the problem precise, let $Q$ be a compact set in the $% xOy$ plane. Consider the approximation of a continuous function $f \in C(Q)$ by functions from the manifold $D=\left\{ \varphi (x)+\psi (y)\right\} ,$ where $\varphi (x),\psi (y)$ are defined and continuous on the projections of $Q$ into the coordinate axes $x$ and $y$, respectively. The approximation error is defined as the distance from $f$ to $D:$ \begin{equation*} E(f)=dist(f,D)=\inf\limits_{D}\left\Vert f-\varphi -\psi \right\Vert _{C(Q)}= \end{equation*}% \begin{equation*} =\inf\limits_{D}\max\limits_{(x,y)\in Q}\left\vert f(x,y)-\varphi (x)-\psi (y)\right\vert. \end{equation*}% A function $\varphi _{0}(x)+\psi _{0}(y)$ from $D$, if it exists, is called an extremal element or a best approximating sum if \begin{equation*} E(f)=\left\Vert f-\varphi _{0}-\psi _{0}\right\Vert _{C(Q)}. \end{equation*}% To show that $E(f)$ depends also on $Q$, in some cases to avoid confusion, we will write $E(f,Q)$ instead of $E(f)$. In this section we deal with calculation formulas for $E(f)$. In 1951 Diliberto and Straus published a paper \cite{26}, in which along with other results they established a formula for $E(f,R)$, where $R$ here and throughout this section is a rectangle with sides parallel to the coordinate axes, containing supremum over all closed lightning bolts. Later the same formula was established by other authors differently, in cases of both rectangle (see \cite{113}) and more general sets (see \cite{79,107}). Although the formula was valid for all continuous functions, it was not easily calculable. Some authors started to seek easily calculable formulas for the approximation error for some subsets of continuous functions. Rivlin and Sibner \cite{121} proved a result, which allow one to find the exact value of $E(f,R)$ for a function $f(x,y)$ having the continuous and nonnegative derivative $\frac{\partial ^{2}f}{\partial x\partial y}$. This result in a more general case (for functions of $n$ variables) was proved by Flatto \cite{30}. Babaev \cite{6} generalized Rivlin and Sibner's result (as well as Flatto's result, see \cite{7}). More precisely, he considered the class $M(R)$ of continuous functions $f(x,y) $ with the property \begin{equation*} \Delta _{h_{1},h_{1}}f=f(x,y)+f(x+h_{1},y+h_{2})-f(x,y+h_{2})-f(x+h_{1},y)\geq 0 \end{equation*}% for each rectangle $\left[ x,x+h_{1}\right] \times \left[ y,y+h_{2}\right] \subset R$, and proved that if $f(x,y)$ belongs to $M(R)$, where $R=\left[ a_{1},b_{1}\right] \times \left[ a_{2},b_{2}\right] $, then \begin{equation*} E(f,R)=\frac{1}{4}\left[ f(a_{1},a_{2})+f(b_{1},b_{2})-f(a_{1},b_{2})-f(b_{1},a_{2})\right] . \end{equation*}% As seen from this formula, to calculate $E(f)$ it is sufficient to find only values of $f(x,y)$ at the vertices of $R$. One can see that the formula also gives a sufficient condition for membership in the class $M(R)$, i.e. if \begin{equation*} E(f,S)=\frac{1}{4}\left[ f(x_{1},y_{1})+f(x_{2},y_{2})-f(x_{1},y_{2})-f(x_{2},y_{1})\right] , \end{equation*}% for a given $f$ and for each $S=\left[ x_{1},x_{2}\right] \times \left[ y_{1},y_{2}\right] \subset R$, then the function $f(x,y)$ is from $M(R)$. Our purpose is to construct new classes of continuous functions, which will depend on a numerical parameter, and characterize each class in terms of the approximation error calculation formulas. The mentioned parameter will show which points of $R$ the calculation formula involves. We will also construct a best approximating sum $\varphi _{0}+\psi _{0}$ to a function from constructed classes. \bigskip \subsection{Definition of the main classes} Let throughout this section $R=\left[ a_{1},b_{1}\right] \times \left[ a_{2},b_{2}\right] $ be a rectangle and $c\in (a_{1},b_{1}]$. Denote $R_{1}=% \left[ a_{1},c\right] \times \left[ a_{2},b_{2}\right] $ and $R_{2}=\left[ c,b_{1}\right] \times \left[ a_{2},b_{2}\right] $. It is clear that $% R=R_{1}\cup R_{2}$ and if $c=b_{1}$, then $R=R_{1}$. We associate each rectangle $S=\left[ x_{1},x_{2}\right] \times \left[ y_{1},y_{2}\right] $ lying in $R$ with the following functional: \begin{equation*} L(f,S)=\frac{1}{4}\left[ f(x_{1},y_{1})+f(x_{2},y_{2})-f(x_{1},y_{2})-f(x_{2},y_{1})\right] . \end{equation*} \bigskip \textbf{Definition 3.1.} \textit{We say that a continuous function $f(x,y)$ belongs to the class $V_{c}(R)$ if } \textit{1) $L(f,S)\geq 0$, for each $S\subset R_{1}$; } \textit{2) $L(f,S)\leq 0$, for each $S\subset R_{2}$; } \textit{3) $L(f,S)\geq 0$, for each $S=\left[ a_{1},b_{1}\right] \times % \left[ y_{1},y_{2}\right] ,~\ S\subset R$.} \bigskip It can be shown that for any $c\in (a_{1},b_{1}]$ the class $V_{c}(R)$ is not empty. Indeed, one can easily verify that the function \begin{equation*} v _{c}(x,y)=\left\{ \begin{array}{c} w(x,y)-w(c,y),\;\ (x,y)\in R_{1} \\ w(c,y)-w(x,y),\;\ (x,y)\in R_{2}% \end{array}% \right. \end{equation*} where $w(x,y)=\left( \frac{x-a_{1}}{b_{1}-a_{1}}\right) ^{\frac{1}{n}}\cdot y $ and $n\geq \log _{2}\frac{b_{1}-a_{1}}{c-a_{1}}$, satisfies conditions 1)-3) and therefore belongs to $V_{c}(R)$. The class $V_{c}(R)$ has the following obvious properties:\newline a) For given functions $f_{1},f_{2}\in V_{c}(R)$ and numbers $\alpha _{1},\alpha _{2}\geq 0$, $\alpha _{1}f_{1}+\alpha _{2}f_{2}\in V_{c}(R)$. $% V_{c}(R)$ is a closed subset of the space of continuous functions.\newline b) $V_{b_{1}}(R)=M(R)$.\newline c) If $f$ is a common element of $V_{c_{1}}(R)$ and $V_{c_{2}}(R)$, $% a_{1}<c_{1}<c_{2}\leq b_{1}$ then $f(x,y)=\varphi (x)+\psi (y)$ on the rectangle $\left[ c_{1},c_{2}\right] \times \left[ a_{2},b_{2}\right] $. The properties a) and b) are clear. The property c) also becomes clear if note that according to the definition of the classes $V_{c_{1}}(R)$ and $% V_{c_{2}}(R)$, for each rectangle \begin{equation*} S\subset \left[ c_{1},c_{2}\right] \times \left[ a_{2},b_{2}\right] \end{equation*} we have \begin{equation*} L(f,S)\leq 0\;\; \mbox{and}\;\; L(f,S)\geq 0, \end{equation*} respectively. Hence \begin{equation*} L(f,S)=0\;\; \mbox{for each}\;\; S\subset \left[ c_{1},c_{2}\right] \times % \left[ a_{2},b_{2}\right]. \end{equation*} Thus it is not difficult to understand that $f$ is of the form $\varphi (x)+\psi (y)$ on the rectangle $\left[ c_{1},c_{2}\right] \times \left[ a_{2},b_{2}\right] $. \bigskip \textbf{Lemma 3.1.} \textit{Assume a function $f(x,y)$ has the continuous derivative $\frac{\partial ^{2}f}{\partial x\partial y}$ on the rectangle $R$ and satisfies the following conditions} \textit{1) $\frac{\partial ^{2}f}{\partial x\partial y}\geq 0$, for all $% (x,y)\in R_{1}$; } \textit{2) $\frac{\partial ^{2}f}{\partial x\partial y}\leq 0$, for all $% (x,y)\in R_{2}$; } \textit{3) $\frac{df(a_{1},y)}{dy}\leq \frac{df(b_{1},y)}{dy}$, for all $% y\in \left[ a_{2},b_{2}\right]$.} \textit{Then $f(x,y)$ belongs to $V_{c}(R)$.} \bigskip The proof of this lemma is very simple and can be obtained by integrating both sides of inequalities in conditions 1)-3) through sets $\left[ x_{1},x_{2}\right] \times \left[ y_{1},y_{2}\right] \subset R_{1}$, $\left[ x_{1},x_{2}\right] \times \left[ y_{1},y_{2}\right] \subset R_{2}$ and $% \left[ y_{1},y_{2}\right] \subset \left[ a_{2},b_{2}\right]$, respectively. \bigskip \textbf{Example 3.1.} Consider the function $f(x,y)=y\sin \pi x$ on the unit square $K=\left[ 0,1\right] \times \left[ 0,1\right] $ and rectangles $K_{1}=% \left[ 0,\frac{1}{2}\right] \times \left[ 0,1\right] ,K_{2}=\left[ \frac{1}{2% },1\right] \times \left[ 0,1\right] $. It is not difficult to verify that this function satisfies all conditions of the lemma and therefore belongs to $V_{\frac{1}{2}}(K)$. \bigskip \subsection{Construction of an extremal element} The following theorem is valid. \bigskip \textbf{Theorem 3.1.}\ \textit{The approximation error of a function $f(x,y)$ from the class $V_{c}(R)$ can be calculated by the formula \begin{equation*} E(f,R)=L(f,R_{1})=\frac{1}{4}\left[ f(a_{1},a_{2})+f(c,b_{2})-f(a_{1},b_{2})-f(c,a_{2})\right] . \end{equation*}% Let $y_{0}$ be any solution from $\left[ a_{2},b_{2}\right] $ of the equation \begin{equation*} L(f,Y)=\frac{1}{2}L(f,R_{1}),\;\;Y=\left[ a_{1},c\right] \times \left[ a_{2},y\right] . \end{equation*}% Then the function $\varphi _{0}(x)+\psi _{0}(y)$, where \begin{equation*} \varphi _{0}(x)=f(x,y_{0}), \end{equation*}% \begin{equation*} \psi _{0}(y)=\frac{1}{2}\left[ f(a_{1},y)+f(c,y)-f(a_{1},y_{0})-f(c,y_{0})% \right] \end{equation*}% is a best approximating sum from the manifold $D$ to $f$.} \bigskip To prove this theorem we need the following lemma. \bigskip \textbf{Lemma 3.2.}\ \textit{Let $f(x,y)$ be a function from $V_{c}(R)$ and $% X=\left[ a_{1},x\right] \times \left[ y_{1},y_{2}\right] $ be a rectangle with fixed $y_{1},y_{2}\in \left[ a_{2},b_{2}\right] $. Then the function $% h(x)=L(f,X)$ has the properties: } \textit{1) $h(x)\geq 0$, for any $x\in \left[ a_{1},b_{1}\right] $; } \textit{2) $\max\limits_{\left[ a_{1},b_{1}\right] }h(x)=h(c)$ and $% \min\limits_{\left[ a_{1},b_{1}\right] }h(x)=h(a_{1})=0$.} \bigskip \textbf{Proof.} If $X\subset R_{1}$, then the validity of $h(x)\geq 0$ follows from the definition of $V_{c}(R)$. If $X$ is from $R$ but not lying in $R_{1}$, then by denoting $X^{\prime }=\left[ x,b_{1}\right] \times \left[ y_{1},y_{2}\right] ,S=X\cup X^{\prime }$ and using the obvious equality \begin{equation*} L(f,S)=L(f,X)+L(f,X^{\prime }) \end{equation*} we deduce from the definition of $V_{c}(R)$ that $h(x)\geq 0$. To prove the second part of the lemma, it is enough to show that $h(x)$ increases on the interval $\left[ a_{1},c\right] $ and decreases on the interval $\left[ c,b_{1}\right] $. Indeed, if $a_{1}\leq x_{1}\leq x_{2}\leq c$, then \begin{equation*} h(x_{2})=L(f,X_{2})=L(f,X_{1})+L(f,X_{12}),\eqno(3.1) \end{equation*}% where $X_{1}=\left[ a_{1},x_{1}\right] \times \left[ y_{1},y_{2}\right] ,$ $% X_{2}=\left[ a_{1},x_{2}\right] \times \left[ y_{1},y_{2}\right] ,$ $X_{12}=% \left[ x_{1},x_{2}\right] \times \left[ y_{1},y_{2}\right] $. Taking into consideration that $L(f,X_{1})=h(x_{1})$ and $X_{12}$ lies in $R_{1}$ we obtain from (3.1) that $h(x_{2})\geq h(x_{1})$. If $c\leq x_{1}\leq x_{2}\leq b_{1}$, then $X_{12}$ lies in $R_{2}$ and we obtain from (3.1) that $h(x_{2})\leq h(x_{1})$. \bigskip \textbf{Proof of Theorem 3.1.} It is obvious that $L(f,R_{1})=L(f-\varphi -\psi ,R_{1})$ for each sum $\varphi (x)+\psi (y)$. Hence \begin{equation*} L(f,R_{1})\leq \left\Vert f-\varphi -\psi \right\Vert _{C(R_{1})}\leq \left\Vert f-\varphi -\psi \right\Vert _{C(R)}. \end{equation*}% Since a sum $\varphi (x)+\psi (y)$ is arbitrary, $L(f,R_{1})\leq E(f,R)$. To complete the proof it is sufficient to construct a sum $\varphi _{0}(x)+\psi _{0}(y)$ for which the equality \begin{equation*} \left\Vert f-\varphi _{0}-\psi _{0}\right\Vert _{C(R)}=L(f,R_{1})\eqno(3.2) \end{equation*}% holds. Consider the function \begin{equation*} g(x,y)=f(x,y)-f(x,a_{2})-f(a_{1},y)+f(a_{1},a_{2}). \end{equation*}% This function has the following obvious properties 1) $g(x,a_{2})=g(a_{1},y)=0$; 2) $L(f,R_{1})=L(g,R_{1})=\frac{1}{4}g(c,b_{2})$; 3) $E(f,R)=E(g,R)$; 4) The function of one variable $g(c,y)$ increases on the interval $\left[ a_{2},b_{2}\right] $. The last property of $g$ allows us to write that \begin{equation*} 0=g(c,a_{2})\leq \frac{1}{2}g(c,b_{2})\leq g(c,b_{2}). \end{equation*}% Since $g(x,y)$ is continuous, there exists at least one solution $y=y_{0}$ of the equation \begin{equation*} g(c,y)=\frac{1}{2}g(c,b_{2}) \end{equation*}% or, in other notation, of the equation \begin{equation*} L(f,Y)=\frac{1}{2}L(f,R_{1}),\;\;\text{where}\;\;Y=\left[ a_{1},c\right] \times \left[ a_{2},y\right] , \end{equation*}% Introduce the functions \begin{equation*} \varphi _{1}(x)=g(x,y_{0}), \end{equation*}% \begin{equation*} \psi _{1}(y)=\frac{1}{2}\left( g(c,y)-g(c,y_{0})\right) , \end{equation*}% \begin{equation*} G(x,y)=g(x,y)-\varphi _{1}(x)-\psi _{1}(y). \end{equation*}% Calculate the norm of $G(x,y)$ on $R$. Consider the rectangles $R^{\prime }=% \left[ a_{1},b_{1}\right] \times \left[ y_{0},b_{2}\right] $ and $R^{\prime \prime }=\left[ a_{1},b_{1}\right] \times \left[ a_{2},y_{0}\right] $. It is clear that \begin{equation*} \left\Vert G\right\Vert _{C(R)}=\max \left\{ \left\Vert G\right\Vert _{C(R^{\prime })},\left\Vert G\right\Vert _{C(R^{\prime \prime })}\right\} . \end{equation*}% First calculate the norm $\left\Vert G\right\Vert _{C(R^{\prime })}$: \begin{equation*} \left\Vert G\right\Vert _{C(R^{\prime })}=\max\limits_{(x,y)\in R^{\prime }}\left\vert G(x,y)\right\vert =\max\limits_{y\in \left[ y_{0},b_{2}\right] }\max\limits_{x\in \left[ a_{1},b_{1}\right] }\left\vert G(x,y)\right\vert .% \eqno(3.3) \end{equation*}% For a fixed point $y$ (we keep it fixed until (3.6)) from the interval $% \left[ y_{0},b_{2}\right] $ we can write that \begin{equation*} \max\limits_{x\in \left[ a_{1},b_{1}\right] }G(x,y)=\max\limits_{x\in \left[ a_{1},b_{1}\right] }\left( g(x,y)-g(x,y_{0})\right) -\psi _{1}(y)\eqno(3.4) \end{equation*}% and \begin{equation*} \min\limits_{x\in \left[ a_{1},b_{1}\right] }G(x,y)=\min\limits_{x\in \left[ a_{1},b_{1}\right] }\left( g(x,y)-g(x,y_{0})\right) -\psi _{1}(y).\eqno(3.5) \end{equation*}% By Lemma 3.2, the function \begin{equation*} h_{1}(x)=4L(f,X)=g(x,y)-g(x,y_{0}),\;\;\mbox{where}\;\;X=\left[ a_{1},x% \right] \times \left[ y_{0},y\right] , \end{equation*}% reaches its maximum on $x=c$ and minimum on $x=a_{1}$: \begin{equation*} \max\limits_{x\in \left[ a_{1},b_{1}\right] }h_{1}(x)=g(c,y)-g(c,y_{0}) \end{equation*}% \begin{equation*} \min\limits_{x\in \left[ a_{1},b_{1}\right] }h_{1}(x)=g(a_{1},y)-g(a_{1},y_{0})=0. \end{equation*}% Considering these facts in (3.4) and (3.5) we obtain that \begin{equation*} \max\limits_{x\in \left[ a_{1},b_{1}\right] }G(x,y)=g(c,y)-g(c,y_{0})-\psi _{1}(y)=\frac{1}{2}\left( g(c,y)-g(c,y_{0})\right) , \end{equation*}% \begin{equation*} \min\limits_{x\in \left[ a_{1},b_{1}\right] }G(x,y)=-\psi _{1}(y)=-\frac{1}{2% }\left( g(c,y)-g(c,y_{0})\right) . \end{equation*}% Consequently, \begin{equation*} \max\limits_{x\in \left[ a_{1},b_{1}\right] }\left\vert G(x,y)\right\vert =% \frac{1}{2}\left( g(c,y)-g(c,y_{0})\right) .\eqno(3.6) \end{equation*}% Taking (3.6) and the $4$-th property of $g$ into account in (3.3) yields \begin{equation*} \left\Vert G\right\Vert _{C(R^{\prime })}=\frac{1}{2}\left( g(c,b_{2})-g(c,y_{0})\right) =\frac{1}{4}g(c,b_{2}). \end{equation*}% Similarly it can be shown that \begin{equation*} \left\Vert G\right\Vert _{C(R^{\prime \prime })}=\frac{1}{4}g(c,b_{2}). \end{equation*}% Hence \begin{equation*} \left\Vert G\right\Vert _{C(R)}=\frac{1}{4}g(c,b_{2})=L(f,R_{1}). \end{equation*}% But by the definition of $G$, \begin{equation*} G(x,y)=g(x,y)-\varphi _{1}(x)-\psi _{1}(y)=f(x,y)-\varphi _{0}(x)-\psi _{0}(y), \end{equation*}% where \begin{equation*} \varphi _{0}(x)=\varphi _{1}(x)+f(x,a_{2})-f(a_{1},a_{2})+f(a_{1},y_{0})=f(x,y_{0}), \end{equation*}% \begin{equation*} \psi _{0}(y)=\psi _{1}(y)+f(a_{1},y)-f(a_{1},y_{0})= \end{equation*}% \begin{equation*} =\frac{1}{2}\left( f(a_{1},y)+f(c,y)-f(a_{1},y_{0})-f(c,y_{0})\right) . \end{equation*}% Therefore, \begin{equation*} \left\Vert f-\varphi _{0}-\psi _{0}\right\Vert _{C(R)}=L(f,R_{1}). \end{equation*}% We proved (3.2) and hence Theorem 3.1. Note that the function $\varphi _{0}(x)+\psi _{0}(y)$ is a best approximating sum from the manifold ${D}$ to $f$. \bigskip \textbf{Remark 3.1.} In the special case $c=b_{1}$, Theorem 3.1 turns into Babaev's result from \cite{6}. \bigskip \textbf{Corollary 3.1.} \textit{Let a function $f(x,y)$ have the continuous derivative $\frac{\partial ^{2}f}{\partial x\partial y}$ on the rectangle $R$ and satisfy the following conditions } \textit{1) $\frac{\partial ^{2}f}{\partial x\partial y}\geq 0$, for all $% (x,y)\in R_{1}$; } \textit{2) $\frac{\partial ^{2}f}{\partial x\partial y}\leq 0$, for all $% (x,y)\in R_{2}$; } \textit{3) $\frac{df(a_{1},y)}{dy}\leq \frac{df(b_{1},y)}{dy}$, for all $% y\in \left[ a_{2},b_{2}\right] $. } \textit{Then} \begin{equation*} E(f,R)=L(f,R_{1})=\frac{1}{4}\left[ f(a_{1},a_{2})+f(c,b_{2})-f(a_{1},b_{2})-f(c,a_{2})\right] . \end{equation*} The proof of this corollary can be obtained directly from Lemma 3.1 and Theorem 3.1. \bigskip \textbf{Remark 3.2.} Rivlin and Sibner \cite{121} proved Corollary 3.1 in the special case $c=b_{1}$. \bigskip \textbf{Example 3.2.} As we know (see Example 3.1) the function $f=y\sin \pi x$ belongs to $V_{\frac{1}{2}}(K)$, where $K=\left[ 0,1\right] \times \left[ 0,1\right] $. By Theorem 3.1, $E(f,K)=\frac{1}{4}$ and the function $\frac{1% }{2}\sin \pi x+\frac{1}{2}y-\frac{1}{4}$ is a best approximating sum. \bigskip The following theorem shows that in some cases the approximation error formula in Theorem 3.1 is valid for more general sets than rectangles with sides parallel to the coordinate axes. \bigskip \textbf{Theorem 3.2.} \textit{Let $f(x,y)$ be a function from $V_{c}(R)$ and $Q\subset R$ be a compact set which contains all vertices of $R_{1}$ (points $(a_{1},a_{2}),(a_{1},b_{2}),(c,a_{2}),(c,b_{2})$). Then} \begin{equation*} E(f,Q)=L(f,R_{1})=\frac{1}{4}\left[ f(a_{1},a_{2})+f(c,b_{2})-f(a_{1},b_{2})-f(c,a_{2})\right] . \end{equation*} \textbf{Proof.} Since $Q\subset R,$ $E(f,Q)\leq E(f,R)$. On the other hand by Theorem 3.1, $E(f,R)=L(f,R_{1})$. Hence $E(f,Q)\leq L(f,R_{1})$. It can be shown, as it has been shown in the proof of Theorem 3.1, that $% L(f,R_{1})\leq E(f,Q)$. But then automatically $E(f,Q)=L(f,R_{1})$. \bigskip \textbf{Example 3.3.} Calculate the approximation error of the function $% f(x,y)=-(x-2)^{2n}y^{m}$ ($n$ and $m$ are positive integers) on the domain \begin{equation*} Q=\left\{ (x,y):0\leq x\leq 2,0\leq y\leq (x-1)^{2}+1\right\} . \end{equation*}% It can be easily verified that $f \in V_{2}(R)$, where $R=\left[ 0,4\right] \times \left[ 0,2\right] $. Besides, $Q$ contains all vertices of $R_{1}=% \left[ 0,2\right] \times \left[ 0,2\right] $. Consequently, by Theorem 3.2, $% E(f,Q)=L(f,R_{1})=2^{2(n-1)+m}$. \bigskip \subsection{Characterization of $V_{c}(R)$} The following theorem characterizes the class $V_{c}(R)$ in terms of the approximation error calculation formulas. \bigskip \textbf{Theorem 3.3.} \textit{The following conditions are necessary and sufficient for a continuous function $f(x,y)$ belong to $V_{c}(R):$ } \textit{1) $E(f,S)=L(f,S)$, for each rectangle $S=\left[ x_{1},x_{2}\right] \times \left[ y_{1},y_{2}\right] ,S\subset R_{1}$; } \textit{2) $E(f,S)=-L(f,S)$, for each rectangle $S=\left[ x_{1},x_{2}\right] \times \left[ y_{1},y_{2}\right] ,S\subset R_{2}$; } \textit{3) $E(f,S)=L(f,S_{1})$, for each rectangle $S=\left[ a_{1},b_{1}% \right] \times \left[ y_{1},y_{2}\right] ,S\subset R$ and $S_{1}=\left[ a_{1},c\right] \times \left[ y_{1},y_{2}\right] $.} \bigskip \textbf{Proof.} The necessity easily follows from the definition of $V_{c}(R) $, Babaev's above-mentioned result (see Section 3.1.1) and Theorem 3.1. The sufficiency is clear if pay attention to the fact that $E(f,S)\geq 0 $. \bigskip \subsection{Classes $V_{c}^{-}(R),U(R)$ and $U_{c}^{-}(R)$} By $V_{c}^{-}(R)$ we denote the class of functions $f(x,y)$ such that $-f \in V_{c}(R)$. It is clear that $E(f,R)=-L(f,R_{1})$ for each $f\in V_{c}^{-}(R)$. We define $U_{c}(R),a_{1}\leq c< b_{1}$, as a class of continuous functions $% f(x,y)$ with the properties 1) $L(f,S)\leq 0$, for each rectangle $S=\left[ x_{1},x_{2}\right] \times % \left[ y_{1},y_{2}\right] ,\;S\subset R_{1};$ 2) $L(f,S)\geq 0$, for each rectangle $S=\left[ x_{1},x_{2}\right] \times % \left[ y_{1},y_{2}\right] ,\;S\subset R_{2};$ 3) $L(f,S)\geq 0$, for each rectangle $S=\left[ a_{1},b_{1}\right] \times % \left[ y_{1},y_{2}\right] ,\;S\subset R.$ Using the same techniques in the proof of Theorem 3.1 it can be shown that the following theorem is valid: \bigskip \textbf{Theorem 3.4.} \textit{The approximation error of a function $f(x,y)$ from the class $U_{c}(R)$ can be calculated by the formula \begin{equation*} E(f,R)=L(f,R_{2})=\frac{1}{4}\left[ f(c,a_{2})+f(b_{1},b_{2})-f(c,b_{2})-f(b_{1},a_{2})\right] . \end{equation*}% Let $y_{0}$ be any solution from $\left[ a_{2},b_{2}\right] $ of the equation \begin{equation*} L(f,Y)=\frac{1}{2}L(f,R_{2}),\qquad Y=\left[ c,b_{1}\right] \times \left[ a_{2},y\right] . \end{equation*}% Then the function $\varphi _{0}(x)+\psi _{0}(y)$, where \begin{equation*} \varphi _{0}(x)=f(x,y_{0}),\quad \psi _{0}(y)=\frac{1}{2}\left[ f(c,y)+f(b_{1},y)-f(c,y_{0})-f(b_{1},y_{0})\right] , \end{equation*}% is a best approximating sum from the manifold $D$ to $f$. } \bigskip By $U_{c}^{-}(R)$ denote the class of functions $f(x,y)$ such that $-f \in U_{c}(R)$. It is clear that $E(f,R)=-L(f,R_{2})$ for each $f\in U_{c}^{-}(R)$% . \bigskip \textbf{Remark 3.3.} The correspondingly modified versions of Theorems 2.2, 2.3 and Corollary 3.1 are valid for the classes $V_{c}^{-}(R),U_{c}(R)$ and $% U_{c}^{-}(R)$. \bigskip \textbf{Example 3.4.} Consider the function $f(x,y)=\left( x-\frac{1}{2}% \right) ^{2}y$ on the unit square $K=\left[ 0,1\right] \times \left[ 0,1% \right] $. It can be easily verified that $f\in U_{\frac{1}{2}}(K)$. Hence, by Theorem 3.4, $E(f,K)=\frac{1}{16}$ and the function $\frac{1}{2}\left( x-% \frac{1}{2}\right) ^{2}+\frac{1}{8}y-\frac{1}{16}$ is a best approximating function. \bigskip \section{Approximation by sums of univariate functions on certain domains} The purpose of this section is to develop a method for obtaining explicit formulas for the error of approximation of bivariate functions by sums of univariate functions. It should be remarked that formulas of this type were known only for functions defined on a rectangle with sides parallel to the coordinate axes. Our method, based on a maximization process over closed bolts, allows the consideration of functions defined on hexagons, octagons and stairlike polygons with sides parallel to the coordinate axes. \subsection{Problem statement} Let $Q$ be a compact set in $\mathbb{R}^2$. Consider the approximation of a continuous function $f \in C(Q)$ by functions from the set $D=\left\{ \varphi (x)+\psi (y)\right\} ,$ where $\varphi (x),\psi (y)$ are defined and continuous on the projections of $Q$ into the coordinate axes $x$ and $y$, respectively. The approximation error is defined as follows \begin{equation*} E(f,Q)=\inf\limits_{\varphi+\psi \in D}\left\Vert f-\varphi -\psi \right\Vert _{C(Q)}. \end{equation*}% Our purpose is to develop a method for obtaining explicit formulas providing precise and easy computation of $E(f,Q)$ for polygons $Q$ with sides parallel to the coordinate axes. This method will be based on the herein developed \textit{closed bolts maximization process} and can be used in alternative proofs of the known results from \cite{6}, \cite{59} and \cite% {121}. First, we show efficiency of the method in the example of a hexagon with sides parallel to the coordinate axes. Then we formulate an analogous theorem for staircase polygons and two theorems for octagons, which can be proved in a similar way, and touch some aspects of the question about the case of an arbitrary polygon with sides parallel to the coordinate axes. The condition posed on sides of polygons (being parallel to the coordinate axes) is essential for our method. This has several reasons, which get clear through the proof of Theorem 3.5. Here we are able to explain one of these reasons: by \cite[Theorem 3]{34}, a continuous function $f(x,y)$ defined on a polygon with sides parallel to the coordinate axes has an extremal element, the existence of which is required in our method. Now let $K$ be a rectangle (not speaking about polygons) with sides not parallel to the coordinate axes. Does any function $f\in C(K)$ have an extremal element? No one knows (see \cite{34}). In the sequel, all the considered polygons are supposed to have sides parallel to the coordinate axes. \bigskip \subsection{The maximization process} Let $H$ be a closed hexagon. It is clear that $H$ can be uniquely represented in the form \begin{equation*} H=R_{1}\cup R_{2},\eqno(3.7) \end{equation*}% where $R_{1},R_{2}$ are rectangles and there does not exist any rectangle $R$ such that $R_{1}\subset R\subset H$ or $R_{2}\subset R\subset H$. We associate each closed bolt $p=\left\{ p_{1},p_{2},\cdots p_{2n}\right\} $ with the following functional \begin{equation*} l(f,p)=\frac{1}{2n}\sum\limits_{k=1}^{2n}(-1)^{k-1}f(p_{k}). \end{equation*} Denote by $M(H)$ the class of bivariate continuous functions $f$ on $H$ satisfying the condition \begin{equation*} f(x_{1},y_{1})+f(x_{2},y_{2})-f(x_{1},y_{2})-f(x_{2},y_{1})\geq 0 \end{equation*} for any rectangle $\left[ x_{1},x_{2}\right] \times \left[ y_{1},y_{2}\right] \subset H.$\newline \textbf{Theorem 3.5.}\ \textit{Let $H$ be a hexagon and (3.7) be its representation. Let $f \in M(H)$. Then \begin{equation*} E(f,H)=\max \left\{ \left\vert l(f,h)\right\vert ,\left\vert l(f,r_{1})\right\vert ,\left\vert l(f,r_{2})\right\vert \right\} ,\eqno(3.8) \end{equation*}% where $h,r_{1},r_{2}$ are closed bolts formed by vertices of the polygons $% H,R_{1},R_{2}$ respectively. } \bigskip \begin{proof} Without loss of generality, we may assume that the rectangles $% R_{1}$ and $R_{2}$ are of the following form \begin{equation*} R_{1}=\left[ a_{1},a_{2}\right] \times \left[ b_{1},b_{3}\right] ,~\ R_{2}=% \left[ a_{1},a_{3}\right] \times \left[ b_{1},b_{2}\right] ,~\ a_{1}<a_{2}<a_{3},\;b_{1}<b_{2}<b_{3}. \end{equation*}% Introduce the notation \begin{equation*} \begin{array}{c} f_{11}=f\left( a_{1},b_{1}\right) ,~\ f_{12}=-f\left( a_{1},b_{2}\right) ,\;f_{13}=-f\left( a_{1},b_{3}\right) ; \\ f_{21}=-f\left( a_{2},b_{1}\right) ,\;f_{22}=-f\left( a_{2},b_{2}\right) ,~\ f_{23}=f\left( a_{2},b_{3}\right) ; \\ f_{31}=-f\left( a_{3},b_{1}\right) ,\;f_{32}=f\left( a_{3},b_{2}\right) .% \end{array}% \eqno(3.9) \end{equation*}% It is clear that \begin{equation*} \begin{array}{c} \left\vert l(f,r_{1})\right\vert =\dfrac{1}{4}\left( f_{11}+f_{13}+f_{23}+f_{21}\right) , \\ \left\vert l(f,r_{2})\right\vert =\dfrac{1}{4}\left( f_{11}+f_{12}+f_{32}+f_{31}\right) , \\ \left\vert l(f,h)\right\vert =\dfrac{1}{6}\left( f_{11}+f_{13}+f_{23}+f_{22}+f_{32}+f_{31}\right) .% \end{array}% \eqno(3.10) \end{equation*} Let $p=\left\{ p_{1},p_{2},\cdots p_{2n}\right\} $ be any closed bolt. We group the points $p_{1},p_{2},\cdots p_{2n}$ by putting \begin{equation*} p_{+}=\left\{ p_{1},p_{3},\cdots p_{2n-1}\right\} ,\;p_{-}=\left\{ p_{2},p_{4},\cdots p_{2n}\right\}. \end{equation*} First, assume that $l(f,p)\geq 0$. We apply the following algorithm, which we call \textit{the maximization process over closed bolts}, to $p$. \textbf{Step 1.} Consider sequentially the units $p_ip_{i+1}$ $\left(i=\overline{% 1,2n}, p_{2n+1}=p_1\right)$ with the vertices $p_{i}\left( x_{i},y_{i}\right) ,~\ p_{i+1}\left( x_{i+1},y_{i+1}\right) $ having equal abscissae: $x_{i}=x_{i+1}$. Four cases are possible. 1) $p_{i}\in p_{+}$ and $y_{i+1}>y_{i}$. In this case, replace the unit $% p_{i}p_{i+1}$ by a new unit $q_{i}q_{i+1}$ with the vertices $% q_{i}=(a_{1},y_{i}),\;\ q_{i+1}=(a_{1},y_{i+1})$. 2) $p_{i}\in p_{+}$ and $y_{i+1}<y_{i}$. In this case, replace the unit $% p_{i}p_{i+1}$ by a new unit $q_{i}q_{i+1}$ with the vertices $q_{i}=\left( a_{2},y_{i}\right) ,\;q_{i+1}=(a_{2},y_{i+1})$ if $b_{2}<y_{i}\leq b_{3}$ or with the vertices $q_{i}=\left( a_{3},y_{i}\right), q_{i+1}=(a_{3},y_{i+1})$ if $b_{1}\leq y_{i}\leq b_{2}$. 3) $p_{i}\in p_{-}$ and $y_{i+1}<y_{i}$. In this case, replace $p_{i}p_{i+1}$ by a new unit $q_{i}q_{i+1}$ with the vertices $% q_{i}=(a_{1},y_{i}),~q_{i+1}=(a_{1},y_{i+1})$. 4) $p_{i}\in p_{-}$ and $y_{i+1}>y_{i}$. In this case, replace $p_{i}p_{i+1}$ by a new unit $q_{i}q_{i+1}$ with the vertices $% q_{i}=(a_{2},y_{i}),~q_{i+1}=(a_{2},y_{i+1})$\ if\ $b_{2}<y_{i+1}\leq b_{3}$ or with the vertices $q_{i}=(a_{3},y_{i}),~q_{i+1}=(a_{3},y_{i+1})$ if $% b_{1}\leq y_{i+1}\leq b_{2}$. Since $f\in M(H)$, it is not difficult to verify that \begin{equation*} \begin{array}{c} f(p_{i})-f(p_{i+1})\leq f(q_{i})-f(q_{i+1})\ \ \mbox{for cases 1) and 2)}, \\ -f(p_{i})+f(p_{i+1})\leq -f(q_{i})+f(q_{i+1})\ \ \mbox{for cases 3) and 4)}% \end{array}% \eqno(3.11) \end{equation*} It is clear that after Step 1 the bolt $p$ will be replaced by the ordered set $% q=\left\{ q_{1},q_{2},\cdots ,q_{2n}\right\} $. We do not say a bolt but an ordered set because of a possibility of coincidence of some successive points $q_{i}, q_{i+1}$ (this, for example, may happen if the 1-st case takes place for the units $p_{i-1}p_{i}$ and $p_{i+1}p_{i+2}$). Let us exclude simultaneously successive and coincident points from $q$. Then we obtain some closed bolt, which we denote by $q^{\prime }=\left\{ q_{1}^{\prime },q_{2}^{\prime },\cdots ,q_{2m}^{\prime }\right\}$. It is not difficult to understand that all points of the bolt $q^{\prime}$ are located on straight lines $x=a_{1},~x=a_{2},~x=a_{3}$. From inequalities (3.11) and the fact that $2m\leq 2n,$ we deduce that \begin{equation*} l(f,p)\leq l(f,q^{\prime }).\eqno(3.12) \end{equation*} \textbf{Step 2.} Consider sequentially units $q_{i}^{\prime }q_{i+1}^{\prime }\;\left( i=\overline{1,2m},q_{2m+1}^{\prime }=q_{1}^{\prime }\right) $ with the vertices $q_{i}^{\prime }=\left( x_{i}^{\prime },y_{i}^{\prime }\right) ,~\ q_{i+1}^{\prime }\left( x_{i+1}^{\prime },y_{i+1}^{\prime }\right) $ having equal ordinates: $y_{i}^{\prime }=y_{i+1}^{\prime }$. The following four cases are possible. 1) $q_{i}^{\prime }\in q_{+}^{\prime }$ and $x_{i+1}^{\prime }>x_{i}^{\prime }$. In this case, replace the unit $q_{i}^{\prime }q_{i+1}^{\prime }$ by a new unit $p_{i}^{\prime }p_{i+1}^{\prime }$ with the vertices $p_{i}^{\prime }=\left( x_{i}^{\prime },b_{1}\right) ,~\ p_{i+1}^{\prime }=\left( x_{i+1}^{\prime },b_{1}\right) $. 2) $q_{i}^{\prime }\in q_{+}^{\prime }$ and $x_{i+1}^{\prime }<x_{i}^{\prime }$. In this case, replace the unit $q_{i}^{\prime }q_{i+1}^{\prime }$ by a new unit $p_{i}^{\prime }p_{i+1}^{\prime }$ with the vertices $p_{i}^{\prime }=\left( x_{i}^{\prime },b_{2}\right) ,$ $\ p_{i+1}^{\prime }=\left( x_{i+1}^{\prime },b_{2}\right) $ if $x_{i}^{\prime }=a_{3}$ and with the vertices $p_{i}^{\prime }=\left( x_{i}^{\prime },b_{3}\right) ,$ $% p_{i+1}^{\prime }=\left( x_{i+1}^{\prime },b_{3}\right) $ if $x_{i}^{\prime }=a_{2}$. 3) $q_{i}^{\prime }\in q_{-}^{\prime }$ and $x_{i+1}^{\prime }<x_{i}^{\prime }$. In this case, replace $q_{i}^{\prime }q_{i+1}^{\prime }$ by a new unit $% p_{i}^{\prime }p_{i+1}^{\prime }$ with the vertices $p_{i}^{\prime }=\left( x_{i}^{\prime },b_{1}\right) ,~\ p_{i+1}^{\prime }=\left( x_{i+1}^{\prime },b_{1}\right)$. 4) $q_{i}^{\prime }\in q_{-}^{\prime }$ and $x_{i+1}^{\prime }>x_{i}^{\prime }$. In this case, replace $q_{i}^{\prime }q_{i+1}^{\prime }$ by a new unit $% p_{i}^{\prime }p_{i+1}^{\prime }$ with the vertices $p_{i}^{\prime }=\left( x_{i}^{\prime },b_{2}\right) ,~\ p_{i+1}^{\prime }=\left( x_{i+1}^{\prime },b_{2}\right)$ if $x_{i+1}^{\prime }=a_{3}$ and with the vertices $% p_{i}^{\prime }=\left( x_{i}^{\prime },b_{3}\right) ,~\ p_{i+1}^{\prime }=\left( x_{i+1}^{\prime },b_{3}\right) $ if $x_{i+1}^{\prime }=a_{2}$. It is easy to see that after Step 2 the bolt $q^{\prime }$ will be replaced by the bolt $p^{\prime }=\left\{ p_{1}^{\prime },p_{2}^{\prime },\cdots p_{2m}^{\prime }\right\} $ and \begin{equation*} l(f,q^{\prime })\leq l(f,p^{\prime }).\eqno(3.13) \end{equation*} From (3.12) and (3.13) we obtain that \begin{equation*} l(f,p)\leq l(f,p^{\prime }).\eqno(3.14) \end{equation*} It is clear that each point of the set $p_{+}^{\prime }$ coincides with one of the points $\left( a_{1},b_{1}\right) ,~\left( a_{2},b_{3}\right) ,$ $% \left( a_{3},b_{2}\right) $ and each point of the set $p_{-}^{\prime }$ coincides with one of the points $\left( a_{1},b_{2}\right) ,~\left( a_{1},b_{3}\right) ,$ $~\left( a_{2},b_{1}\right) ,~\left( a_{2},b_{2}\right) ,~\left( a_{3},b_{1}\right) .$ Denote by $m_{ij}$ the number of points of the bolt $p^{\prime }$ coinciding with the point $\left( a_{i},b_{j}\right) ,~i,j=\overline{1,3},~i+j\neq 6$. By (3.9), we can write that \begin{equation*} l(f,p^{\prime })=\frac{1}{2m}\sum\limits_{\substack{ i,j=\overline{1,3} \\ % i+j\leq 5}}m_{ij}f_{ij}.\eqno(3.15) \end{equation*} On the straight line $x=a_{i}~\ $or $\ y=b_{i},~i=\overline{1,3}$, the number of points of the set $p_{+}^{\prime }$ is equal to the number of points of the set $p_{-}^{\prime }$. Hence \begin{equation*} m_{11}=m_{12}+m_{13}=m_{21}+m_{31};\ m_{23}=m_{22}+m_{21}=m_{13};\ m_{32}=m_{31}=m_{12}+m_{22}. \end{equation*}% From these equalities we deduce that \begin{equation*} m_{11}=m_{12}+m_{21}+m_{22};\ m_{13}=m_{21}+m_{22};\ m_{23}=m_{21}+m_{22};\ m_{31}=m_{12}+m_{22}.\eqno(3.16) \end{equation*}% Consequently, \begin{equation*} 2m=\sum\limits_{\substack{ i,j=\overline{1,3} \\ i+j\leq 5}}% m_{ij}=4m_{12}+4m_{21}+6m_{22}.\eqno(3.17) \end{equation*}% Considering (3.16) and (3.17) in (3.15) and taking (3.10) into account, we obtain that \begin{equation*} l(f,p^{\prime })=\dfrac{4m_{12}\left\vert l(f,r_{2})\right\vert +4m_{21}\left\vert l(f,r_{1})\right\vert +6m_{22}\left\vert l(f,h)\right\vert }{4m_{12}+4m_{21}+6m_{22}} \end{equation*}% \begin{equation*} \leq \max \left\{ \left\vert l(f,r_{1})\right\vert ,\left\vert l(f,r_{2})\right\vert ,\left\vert l(f,h)\right\vert \right\} . \end{equation*}% Therefore, due to (3.14), \begin{equation*} l(f,p)\leq \max \left\{ \left\vert l(f,r_{1})\right\vert ,\left\vert l(f,r_{2})\right\vert ,\left\vert l(f,h)\right\vert \right\} .\eqno(3.18) \end{equation*} Note that in the beginning of the proof the bolt $p$ has been chosen so that $l(f,p)\geq 0$. Let now $p=\left\{ p_{1},p_{2},\cdots p_{2n}\right\} $ be any closed bolt such that $l(f,p)\leq 0$. Since $l(f,p^{\prime \prime })=$ $-l(f,p)\geq 0$ for the bolt $p^{\prime \prime }=\left\{ p_{2},p_{3},\cdots ,p_{2n},p_{1}\right\} $,we obtain from (3.18) that \begin{equation*} -l(f,p)\leq \max \left\{ \left\vert l(f,r_{1})\right\vert ,\left\vert l(f,r_{2})\right\vert ,\left\vert l(f,h)\right\vert \right\} .\eqno(3.19) \end{equation*}% From (3.18) and (3.19) we deduce on the strength of arbitrariness of $p$ that \begin{equation*} \sup\limits_{p\subset H}\left\{ \left\vert l(f,p)\right\vert \right\} =\max \left\{ \left\vert l(f,r_{1})\right\vert ,\left\vert l(f,r_{2})\right\vert ,\left\vert l(f,h)\right\vert \right\} ,\eqno(3.20) \end{equation*}% where the $sup$ is taken over all closed bolts of the hexagon $H$. The hexagon $H$ satisfies the conditions of Theorem 1.10 on the existence of a best approximation. By \cite[Theorem 2]{79} (see Section 3.3), we obtain that \begin{equation*} E(f,H)=\sup\limits_{p\subset H}\left\{ \left\vert l(f,p)\right\vert \right\} .\eqno(3.21) \end{equation*}% From (3.20) and (3.21) we finally conclude that \begin{equation*} E(f,H)=\max \left\{ \left\vert l(f,r_{1})\right\vert ,\left\vert l(f,r_{2})\right\vert ,\left\vert l(f,h)\right\vert \right\} . \end{equation*}% \end{proof} \textbf{Corollary 3.2.}\ \textit{Let a function $f(x,y)$ have the continuous nonnegative derivative $\dfrac{\partial ^{2}f}{\partial x\partial y}$ on $H$% . Then the formula (3.8) is valid.} \bigskip The proof is very simple and can be obtained by integrating the inequality \linebreak $\dfrac{\partial ^{2}f}{\partial x\partial y}\!\geq \!0$ over an arbitrary rectangle $\left[ x_{1},x_{2}\right] \times \left[ y_{1},y_{2}% \right] \subset H$ and applying Theorem 3.5. The method used in the proof of Theorem 3.5 can be generalized to obtain similar results for stairlike polygons. For example, let $S$ be a closed polygon of the following form \begin{equation*} S=\bigcup\limits_{i=1}^{N-1}P_{i}, \end{equation*}% where $N\geq 2,$ $P_{i}=\left[ a_{i},a_{i+1}\right] \times \left[ b_{1},b_{N+1-i}\right] ,$ $i=\overline{1,N-1},$ $a_{1}<a_{2}<\dots <a_{N},$ $% b_{1}<b_{2}<\dots <b_{N}$. Such polygons will be called \textit{stairlike polygons} (see \cite{56}). A closed $2m$-gon $F$ with sides parallel to the coordinate axes is called a maximal $2m$-gon of the polygon $S$ if $F\subset S$ and there is no another $% 2m$-gon $F^{\prime }$ such that $F\subset F^{\prime }\subset S$. Clearly, if $F$ is a maximal $2m$-gon of the polygon $S$, then $m\leq N.$ A closed bolt formed by the vertices of a maximal polygon $F$ is called a maximal bolt of $% S$. By $S^{B}$ denote the set of all maximal bolts of the stairlike polygon $% S.$ \bigskip \textbf{Theorem 3.6.} \textit{Let $S$ be a stairlike polygon. The approximation error of a function $f \in M(S)$ can be computed by the formula% } \begin{equation*} E\left( f,S\right) =\max \left\{ \left\vert r(f,h)\right\vert ,\;h\in S^{B}\right\} . \end{equation*} \bigskip For the proof of this theorem see \cite{56}. \bigskip \subsection{$E$-bolts} The main idea in the proof of Theorem 3.5 can be successfully used in obtaining formulas of type (3.8) for functions $f(x,y)$ defined on another simple polygons. The following two theorems include cases of some octagons and can be proved in a similar way. \bigskip \textbf{Theorem 3.7.}\ \textit{Let $a_{1}<a_{2}<a_{3}<a_{4},$ $% b_{1}<b_{2}<b_{3}$ and $Q$ be an octagon of the following form \begin{equation*} Q=\bigcup\limits_{i=1}^{4}R_{i},\ \ \ where \end{equation*}% $R_{1}=\left[ a_{1},a_{2}\right] \times \left[ b_{1},b_{2}\right] ,R_{2}=% \left[ a_{2},a_{3}\right] \times \left[ b_{1},b_{2}\right] ,R_{3}=\left[ a_{3},a_{4}\right] \times \left[ b_{1},b_{2}\right] ,R_{4}=\left[ a_{2},a_{3}% \right] \times \left[ b_{2},b_{3}\right] $.} \textit{Let $f\in M(Q)$. Then the following formula holds \begin{equation*} E(f,Q)=\max \left\{ \left| l(f,q)\right| ,\left| l(f,r_{123} )\right| ,\left| l(f,r_{124} )\right| ,\left| l(f,r_{234} )\right| ,\left| l(f,r_{24} )\right| \right\}, \end{equation*} where $q,$ $r_{123},$ $r_{124},$ $r_{234},$ $r_{24} $ are closed bolts formed by the vertices of the polygons $Q,$ $R_{1} \cup R_{2}\cup R_{3} ,R_{1}\cup R_{2}\cup R_{4} ,R_{2}\cup R_{3}\cup R_{4} $ and $R_{2}\cup R_{4} $, respectively.} \bigskip \textbf{Theorem 3.8.}\ \textit{Let $a_{1}<a_{2}<a_{3}<a_{4},\ b_{1}<b_{2}<b_{3}$ and $Q$ be an octagon of the following form \begin{equation*} Q=\bigcup_{i=1}^{3}R_{i}, \end{equation*}% where $R_{1}=\left[ a_{1},a_{4}\right] \times \left[ b_{1},b_{2}\right] ,R_{2}=\left[ a_{1},a_{2}\right] \times \left[ b_{2},b_{3}\right] ,R_{3}=% \left[ a_{3},a_{4}\right] \times \left[ b_{2},b_{3}\right] $.} \textit{Let $% f\in M(Q)$. Then \begin{equation*} E(f,Q)=\max \left\{ \left| l(f,r)\right| ,\left| l(f,r_{12} )\right| ,\left| l(f,r_{13} )\right| \right\}, \end{equation*} where $r,r_{12} ,r_{13} $ are closed bolts formed by the vertices of the polygons $R=\left[ a_{1} ,a_{4} \right] \times \left[ b_{1} ,b_{3} \right],$ $R_{1}\cup R_{2} ,R_{1}\cup R_{3}$, respectively.} \bigskip Although the closed bolts maximization process can be applied to bolts of an arbitrary polygon, some combinatorial difficulties arise when grouping values at points of maximized bolts (bolts obtained after the maximization process, see (3.15)-(3.18)). While we do not know a complete answer to this problem, we can describe points of a polygon $F$ with which points of maximized bolts coincide and state a conjecture concerning the approximation error. Let $F=A_{1}A_{2}...A_{2n}$ be any polygon with sides parallel to the coordinate axes. The vertices $A_{1},$ $A_{2},$ $...,$ $A_{2n}$ in the given order form a closed bolt, which we denote by $r_{F}$. By $\left[ r_{F}\right] $ denote the length of $r_{F}$. In our case, $\left[ r_{F}\right] =2n$. \bigskip \textbf{Definition 3.2.}\ \textit{Let $F$ and $S$ be polygons with sides parallel to the coordinate axes. We say that the closed bolt $r_{F}$ is an $% e $-bolt (extended bolt) of $S$ if\ $r_{F}\subset S$\ and there does not exist any polygon $F^{^{\prime }}$ such that $F\subset F^{^{\prime }},\ \ r_{F^{^{\prime }}}\subset S,\ \ \left[ r_{F^{^{\prime }}}\right] \leq \left[ r_{F}\right] .$} \bigskip For example, in Theorem 3.8 the octagon $Q$ has $3$ $e$-bolts. They are $% r,r_{12}$ and $r_{13}$. In Theorem 3.7, the octagon $Q$ has $5$ $e$-bolts, which are $q,r_{123},r_{124},r_{234}$ and $r_{24}$ . The polygon $% S_{2n}=\bigcup\limits_{i=1}^{n-1}R_{i}$, where $R_{i}=\left[ a_{i},a_{i+1}% \right] \times \left[ b_{1},b_{n+1-i}\right] ,i=\overline{1,n-1}% ,a_{1}<a_{2}<...<a_{n},b_{1}<b_{2}<...<b_{n}$ has exactly $2^{n-1}-1$ $e$% -bolts. It is not difficult to observe that the set of points of a closed bolt obtained after the maximization process is a subset of the set of points of all $e$-bolts. This condition and Theorems 2.5-2.8 justify the statement of the following conjecture: \textit{Let $S$ be any polygon with sides parallel to the coordinate axes and $f \in M(S)$. Then \begin{equation*} E(f,S)=\max_{h\in S^{E}}\left\{ \left\vert l(f,h)\right\vert \right\}, \end{equation*}% where $S^{E}$ is a set of all $e$-bolts of the polygon $S$.} \bigskip \subsection{Error estimates} Theorem 3.5 allows us to consider classes wider than $M(H)$ and establish sharp estimates for the approximation error. \bigskip \textbf{Theorem 3.9.}\ \textit{Let $H$ be a hexagon and (3.7) be its representation. The following sharp estimates are valid for a function $% f(x,y)$ having the continuous derivative $\dfrac{\partial ^{2}f}{\partial x\partial y}$ on $H$: \begin{equation*} A\leq E(f,H)\leq BC+\frac{3}{2}\left( B\left\vert l(g,h)\right\vert -\left\vert l(f,h)\right\vert \right) ,\eqno(3.22) \end{equation*}% where \begin{equation*} B=\max_{(x,y)\in H}\left\vert \frac{\partial ^{2}f(x,y)}{\partial x\partial y% }\right\vert ,\ \ \ g=g(x,y)=x\cdot y, \end{equation*}% \begin{equation*} A=\max \left\{ \left\vert l(f,h)\right\vert ,\ \left\vert l(f,r_{1})\right\vert ,\left\vert l(f,r_{2})\right\vert \right\} ,\ C=\max \left\{ \left\vert l(g,h)\right\vert ,\left\vert l(g,r_{1})\right\vert ,\ \left\vert l(g,r_{2})\right\vert \right\}, \end{equation*}% where $h,r_{1},r_{2}$ are closed bolts formed by vertices of the polygons $% H,R_{1}$ and $R_{2}$, respectively.} \bigskip \textbf{Remark 3.4.} Inequalities similar to (3.22) were established in Babaev \cite{8} for the approximation of a function $f(x)=f(x_{1},...,x_{n})$% , defined on a parallelepiped with sides parallel to the coordinate axes, by sums $\sum\limits_{i=1}^{n}\varphi _{i}(x\backslash x_{i})$. For the approximation of bivariate functions, Babaev's result contains only rectangular case. \bigskip \textbf{Remark 3.5.} Estimates (3.22) are easily calculable in contrast to those established in \cite{5} for continuous functions defined on certain domains, which are different from polygons. \bigskip To prove Theorem 3.9 we need the following lemmas. \bigskip \textbf{Lemma 3.3.}\ \textit{Let $X$ be a normed space, $F$ be a subspace of $X$. The following inequality is valid for an element $x=x_{1}+x_{2}$ from $X $: \begin{equation*} \left\vert E(x_{1})-E(x_{2})\right\vert \leq E(x)\leq E(x_{1})+E(x_{2}), \end{equation*}% where \begin{equation*} E(x)=E(x,F)=\inf_{y\in F}\left\Vert x-y\right\Vert . \end{equation*}% } \bigskip \textbf{Lemma 3.4.}\ \textit{If $f\in M(H)$, then \begin{equation*} \left\vert l(f,r_{i})\right\vert \leq \frac{3}{2}\left\vert l(f,h)\right\vert ,i=1,2. \end{equation*}% } Lemma 3.3 is obvious. To prove Lemma 3.4, note that for any $f\in M(H)$ \begin{equation*} 6\left\vert l(f,h)\right\vert =4\left\vert l(f,r_{i})\right\vert +4\left\vert l(f,r_{3})\right\vert ,\ \ i=1,2, \end{equation*}% where $r_{3}$ is a closed bolt formed by the vertices of the rectangle $% R_{3}=H\backslash R_{i}.$ Now let us prove Theorem 3.9. \begin{proof} It is not difficult to verify that if $\frac{\partial ^{2}u}{% \partial x\partial y}\geq 0$ on $H$ for some $u(x,y),$ $\frac{\partial ^{2}u(x,y)}{\partial x\partial y}\in C(H)$, then $u\in M(H)$ (see the proof of Corollary 3.2). Set $f_{1}=f+Bg$. Since $\frac{\partial ^{2}f_{1}}{% \partial x\partial y}\geq 0$ on $H$, $f_{1}\in M(H)$. By Lemma 3.4, \begin{equation*} \left\vert l(f_{1},r_{i})\right\vert \leq \frac{3}{2}\left\vert l(f_{1},h)\right\vert ,i=1,2.\eqno(3.23) \end{equation*} Theorem 3.5 implies that \begin{equation*} E(f_{1},H)=\max \left\{ \left\vert l(f_{1},h)\right\vert ,\left\vert l(f_{1},r_{1})\right\vert ,\left\vert l\left( f_{1},r_{2}\right) \right\vert \right\} .\eqno(3.24) \end{equation*} We deduce from (3.23) and (3.24) that \begin{equation*} E(f_{1},H)\leq \frac{3}{2}\left\vert l(f_{1},h)\right\vert . \end{equation*} First, let the closed bolt $h$ start at the point $(a_{1},b_{1})$. Then it is clear that \begin{equation*} E(f_{1},H)\leq \frac{3}{2}l(f_{1},h).\eqno(3.25) \end{equation*} By Lemma 3.3, \begin{equation*} E(f,H)-E(Bg,H)\leq E(f_{1},H).\eqno(3.26) \end{equation*} Inequalities (3.25) and (3.26) yield \begin{equation*} E(f,H)\leq BE(g,H)+\frac{3}{2}l(f_{1},h).\eqno(3.27) \end{equation*} Since the functional $l(f,h)$ is linear, \begin{equation*} l(f_{1},h)=l(f,h)+Bl(g,h). \end{equation*}% Considering this expression of $l(f_{1},h)$ in (3.27), we obtain that \begin{equation*} E(f,H)\leq BE(g,H)+\frac{3}{2}Bl(g,h)+\frac{3}{2}l(f,h).\eqno(3.28) \end{equation*} Now consider the function $f_{2}=Bg-f$. Obviously, $\frac{\partial ^{2}f_{2}}{\partial x\partial y}% \geq 0$ on $H$. It can be shown, in the same way as (3.28) has been obtained, that \begin{equation*} E(f,H)\leq BE(g,H)+\frac{3}{2}Bl(g,h)-\frac{3}{2}l(f,h).\eqno(3.29) \end{equation*} From (3.28) and (3.29) it follows that \begin{equation*} E(f,H)\leq BE(g,H)+\frac{3}{2}Bl(g,h)-\frac{3}{2}\left\vert l(f,h)\right\vert .\eqno(3.30) \end{equation*}% Since $g\in M(H)$ and $h$ starts at the point $(a_{1},b_{1}),$ we have $% l(g,h)\geq 0$. Let now $h$ start at a point such that $l(u,h)\leq 0$ for any $u\in M(H)$. Then in a similar way as above we can prove that \begin{equation*} E(f,H)\leq BE(g,H)-\frac{3}{2}Bl(g,h)-\frac{3}{2}\left\vert l(f,h)\right\vert ,\eqno(3.31) \end{equation*}% where $l(g,h)\leq 0$. From (3.30), (3.31) and the fact that $E(g,H)=C$ (in view of Theorem 3.5), it follows that \begin{equation*} E(f,H)\leq BC+\frac{3}{2}\left( B\left\vert l(g,h)\right\vert -\left\vert l(f,h)\right\vert \right) . \end{equation*}% The upper bound in (3.22) has been established. Note that it is attained by $f=g=xy$. The proof of the lower bound in (3.22) is simple. One of the obvious properties of the functional $l(f,p)$ is that $\left\vert l(f,p)\right\vert \leq E(f,H)$ for any continuous function $f$ on $H$ and a closed bolt $p$. Hence, \begin{equation*} A=\max \left\{ \left\vert l(f,h)\right\vert ,\left\vert l(f,r_{1})\right\vert ,\left\vert l(f,r_{2})\right\vert \right\} \leq E(f,H). \end{equation*} Note that by Theorem 3.5 the lower bound in (3.22) is attained by an arbitrary function from $M(H)$. \end{proof} \textbf{Remark 3.6.} Using Theorems 2.7 and 2.8 one can obtain sharp estimates of type (3.22) for bivariate functions defined on the corresponding simple polygons with sides parallel to the coordinate axes. \bigskip \section{On the theorem of M. Golomb} Let $X_{1},...,X_{n}$ be compact spaces and $X=X_{1}\times \cdots \times X_{n}.$ Consider the approximation of a function $f\in C(X)$ by sums $% g_{1}(x_{1})+\cdots +g_{n}(x_{n}),$ where $g_{i}\in C(X_{i}),$ $i=1,...,n.$ In \cite{37}, M.Golomb obtained a formula for the error of this approximation in terms of measures constructed on special points of $X$, called ``projection cycles". However, his proof had a gap, which was pointed out later by Marshall and O'Farrell \cite{107}. But the question if the formula was correct, remained open. The purpose of this section is to prove that Golomb's formula is valid, and moreover it holds in a stronger form. \subsection{History of Golomb's formula} Let $X_{i},i=1,...,n,$ be compact Hausdorff spaces. Consider the approximation to a continuous function $f$, defined on $X=X_{1}\times \cdots \times X_{n}$, from the manifold \begin{equation*} M=\left\{ \sum_{i=1}^{n}g_{i}(x_{i}):g_{i}\in C(X_{i}),~~i=1,...,n\right\} . \end{equation*}% The approximation error is defined as the distance from $f$ to $M$: \begin{equation*} E(f)\overset{def}{=}dist(f,M)=\underset{g\in M}{\inf }\left\Vert f-g\right\Vert _{C(X)}. \end{equation*} The well-known duality relation says that \begin{equation*} E(f)=\underset{\left\Vert \mu \right\Vert \leq 1}{\underset{\mu \in M^{\bot }% }{\sup }}\left\vert \int\limits_{X}fd\mu \right\vert ,\eqno(3.32) \end{equation*}% where $M^{\bot }$ is the space of regular Borel measures annihilating all functions in $M$ and $\left\Vert \mu \right\Vert $ stands for the total variation of a measure $\mu $. It should be noted that the $\sup $ in (3.32) is attained by some measure $\mu ^{\ast }$ with total variation $\left\Vert \mu ^{\ast }\right\Vert =1.$ We are interested in the problem: is it possible to replace in (3.32) the class $M^{\bot }$ by some subclass of it consisting of measures of simple structure? For the case $n=2,$ this problem was first considered by Diliberto and Straus \cite{26}. They showed that the measures generated by closed bolts are sufficient for the equality (3.32). In case of general topological spaces, a lightning bolt is defined similarly to the case $\mathbb{R}^2$. Let $X=X_{1}\times X_{2}$ and $\pi _{i}$ be the projections of $X$ onto $X_{i},$ $i=1,2.$ A lightning bolt (or, simply, a bolt) is a finite ordered set $\{a_{1},...,a_{k}\}$ contained in $X$, such that $a_{i}\neq a_{i+1}$, for $i=1,2,...,k-1$, and either $\pi _{1}(a_{1})=\pi _{1}(a_{2}),$ $\pi _{2}(a_{2})=\pi _{2}(a_{3})$, $\pi _{1}(a_{3})=\pi _{1}(a_{4}),...,$ or $\pi _{2}(a_{1})=\pi _{2}(a_{2}),$ $\pi _{1}(a_{2})=\pi _{1}(a_{3})$, $\pi _{2}(a_{3})=\pi _{2}(a_{4}),...$ A bolt $% \{a_{1},...,a_{k}\}$ is said to be closed if $k$ is an even number and the set $\{a_{2},...,a_{k},a_{1}\}$ is also a bolt. Let $l=\{a_{1},...,a_{2k}\}$ be a closed bolt. Consider a measure $\mu _{l}$ having atoms $\pm \frac{1}{2k}$ with alternating signs at the vertices of $l$% . That is, \begin{equation*} \mu _{l}=\frac{1}{2k}\sum_{i=1}^{2k}(-1)^{i-1}\delta _{a_{i}}\text{ \ or \ }% \mu _{l}=\frac{1}{2k}\sum_{i=1}^{2k}(-1)^{i}\delta _{a_{i}}, \end{equation*}% where $\delta _{a_{i}}$ is a point mass at $a_{i}.$ It is clear that $\mu _{l}\in M^{\bot }$ and $\left\Vert \mu _{l}\right\Vert \leq 1$. $\left\Vert \mu _{l}\right\Vert =1$ if and only if the set of vertices of the bolt $l$ having even indices does not intersect with that having odd indices. The following duality relation was first established by Diliberto and Straus \cite{26} \begin{equation*} E(f)=\underset{l\subset X}{\sup }\left\vert \int\limits_{X}fd\mu _{l}\right\vert ,\eqno(3.33) \end{equation*}% where $X=X_{1}\times X_{2}$ and the $\sup $ is taken over all closed bolts of $X$. In fact, Diliberto and Straus obtained the formula (3.33) for the case when $X$ is a rectangle in $\mathbb{R}^{2}$ with sides parallel to the coordinate axis. The same result was independently proved by Smolyak (see \cite{113}). Yet another proof of (3.33), in the case when $X$ is a Cartesian product of two compact Hausdorff spaces, was given by Light and Cheney \cite{93}. For $X$'s other than a rectangle in $\mathbb{R}^{2}$, the theorem under some additional assumptions appeared in the works \cite% {62,79,107}. But we shall not discuss these works here. Golomb's paper \cite{37} made a start to a systematic study of approximation of multivariate functions by various compositions, including sums of univariate functions. Golomb generalized the notion of a closed bolt to the $% n$-dimensional case and obtained the analogue of formula (3.33) for the error of approximation from the manifold $M$. The objects introduced in \cite% {37} were called \textit{projection cycles} and they are defined as sets of the form \begin{equation*} p=\{b_{1},...,b_{k};~c_{1},...,c_{k}\}\subset X,\eqno(3.34) \end{equation*}% with the property that $b_{i}\neq c_{j}$, $i,j=1,...,k$ and for all $\nu =1,...,n,$ the group of the $\nu $-th coordinates of $c_{1},...,c_{k}$ is a permutation of that of the $\nu $-th coordinates of $b_{1},...,b_{k}.$ Some points in the $b$-part $\left( b_{1},...,b_{k}\right) $ or $c$-part $\left( c_{1},...,c_{k}\right) $ of $p$ may coincide. The measure associated with $p$ is \begin{equation*} \mu _{p}=\frac{1}{2k}\left( \sum_{i=1}^{k}\delta _{b_{i}}-\sum_{i=1}^{k}\delta _{c_{i}}\right). \end{equation*} It is clear that $\mu _{p}\in M^{\bot }$ and $\left\Vert \mu _{p}\right\Vert =1.$ Besides, if $n=2,$ then a projection cycle is the union of closed bolts after some suitable permutation of its points. Golomb's result states that \begin{equation*} E(f)=\underset{p\subset X}{\sup }\left\vert \int\limits_{X}fd\mu _{p}\right\vert ,\eqno(3.35) \end{equation*}% where $X=X_{1}\times \cdots \times X_{n}$ and the $\sup $ is taken over all projection cycles of $X$. It can be proved that in the case $n=2,$ the formulas (3.33) and (3.35) are equivalent. Unfortunately, the proof of (3.35) had a gap, which was pointed out many years later by Marshall and O'Farrell \cite{107}. But the question if the formula (3.35) was correct, remained unsolved (see also the monograph by Khavinson \cite{76}% ). Note that Golomb's result was used and cited in the literature, for example, in works \cite{75,126}. In the following subsection, we will construct families of normalized measures (that is, measures with the total variation equal to $1$) on projection cycles. Each measure $\mu _{p}$ defined above will be a member of some family. We will also consider minimal projection cycles and measures constructed on them. By properties of these measures, we show that Golomb's formula (3.35) is valid in a stronger form. \bigskip \subsection{Measures supported on projection cycles} Let us give an equivalent definition of a projection cycle. This will be useful in constructing of certain measures having simple structure and capability of approximating arbitrary measures in $M^{\bot }$. In the sequel, $\chi _{a}$ will denote the characteristic function of a single point set $\{a\}\subset \mathbb{R}$. \bigskip \textbf{Definition 3.3.} \textit{Let $X=X_{1}\times \cdots \times X_{n}$ and $\pi _{i}$ be the projections of $X$ onto the sets $X_{i},$ $i=1,...,n.$ We say that a set $p=\{x_{1},...,x_{m}\}\subset X$ is a projection cycle if there exists a vector $\lambda =(\lambda _{1},...,\lambda _{m})$ with nonzero real coordinates such that} \begin{equation*} \sum_{j=1}^{m}\lambda _{j}\chi _{\pi _{i}(x_{j})}=0,\text{ \ }i=1,...,n.\eqno% (3.36) \end{equation*} \bigskip Let us give some explanatory remarks concerning Definition 3.3. Fix the subscript $i.$ Let the set $\{\pi _{i}(x_{j})$, $j=1,...,m\}$ have $s_{i}$ different values, which we denote by $\gamma _{1}^{i},\gamma _{2}^{i},...,\gamma _{s_{i}}^{i}.$ Then (3.36) implies that \begin{equation*} \sum_{j}\lambda _{j}=0, \end{equation*}% where the sum is taken over all $j$ such that $\pi _{i}(x_{j})=\gamma _{k}^{i},$ $k=1,...,s_{i}.$ Thus for fixed $i$, we have $s_{i}$ homogeneous linear equations in $\lambda _{1},...,\lambda _{m}.$ The coefficients of these equations are the integers $0$ and $1.$ By varying $i$, we obtain $% s=\sum_{i=1}^{n}s_{i}$ such equations. Hence (3.36), in its expanded form, stands for the system of these equations. One can observe that if this system has a solution $(\lambda _{1},...,\lambda _{m})$ with nonzero real components $\lambda _{i},$ then it also has a solution $(n_{1},...,n_{m})$ with nonzero integer components $n_{i},$ $i=1,...,m.$ This means that in Definition 3.3, we can replace the vector $\lambda $ by the vector $% n=(n_{1},...,n_{m})\,$, where $n_{i}\in \mathbb{Z}\backslash \{0\},$ $% i=1,...,m.$ Thus, Definition 3.3 is equivalent to the following definition. \bigskip \textbf{Definition 3.4.} \textit{A set $p=\{x_{1},...,x_{m}\}\subset X$ is called a projection cycle if there exist nonzero integers $n_{1},...,n_{m}$ such that} \begin{equation*} \sum_{j=1}^{m}n_{j}\chi _{\pi _{i}(x_{j})}=0,\text{ \ }i=1,...,n.\eqno(3.37) \end{equation*} \bigskip \textbf{Lemma 3.5.} \textit{Definition 3.4 is equivalent to Golomb's definition of a projection cycle.} \bigskip \begin{proof} Let $p=\{x_{1},...,x_{m}\}$ be a projection cycle with respect to Definition 3.4. By $b$ and $c$ denote the set of all points $x_{i}$ such that the integers $n_{i}$ associated with them in (3.37) are positive and negative correspondingly. Write out each point $x_{i}$ $n_{i}$ times if $% n_{i}>0$ and $-n_{i}$ times if $n_{i}<0.$ Then the set $\{b;c\}$ is a projection cycle with respect to Golomb's definition. The inverse is also true. Let a set $p_{1}=\{b_{1},...,b_{k};~c_{1},...,c_{k}\}$ be a projection cycle with respect to Golomb's definition. Here, some points $b_{i}$ or $c_{i}$ may be repeated. Let $p=\{x_{1},...,x_{m}\}$ stand for the set $p_{1}$, but with no repetition of its points. Let $n_{i}$ show how many times $x_{i}$ appear in $p_{1}.$ We take $n_{i}$ positive if $x_{i}$ appears in the $b$-part of $p_{1}$ and negative if it appears in the $c$% -part of $p_{1}.$ Clearly, the set $\{x_{1},...,x_{m}\}$ is a projection cycle with respect to Definition 3.4, since the integers $n_{i},$ $% i=1,...,m, $ satisfy (3.37). \end{proof} In the sequel, we will use Definition 3.3. A pair $\left\langle p,\lambda \right\rangle ,$ where $p$ is a projection cycle in $X$ and $\lambda $ is a vector associated with $p$ by (3.36), will be called a ``projection cycle-vector pair" of $X.$ To each such pair $\left\langle p,\lambda \right\rangle $ with $p=\{x_{1},...,x_{m}\}$ and $\lambda =(\lambda _{1},...,\lambda _{m})$, we correspond the measure \begin{equation*} \mu _{p,\lambda }=\frac{1}{\sum_{j=1}^{m}\left\vert \lambda _{j}\right\vert }% \sum_{j=1}^{m}\lambda _{j}\delta _{x_{j}}.\eqno(3.38) \end{equation*} Clearly, $\mu _{p,\lambda }\in M^{\bot }$ and $\left\Vert \mu _{p,\lambda }\right\Vert =1$. We will also deal with measures supported on some certain subsets of projection cycles called \textit{minimal projection cycles}. A projection cycle is said to be minimal if it does not contain any projection cycle as its proper subset. For example, the set $p=% \{(0,0,0),~(0,0,1),~(0,1,0),~(1,0,0),~(1,1,1)\}$ is a minimal projection cycle in $\mathbb{R}^{3},$ since the vector $\lambda =(2,-1,-1,-1,1)$ satisfies Eq. (3.36) and there is no such vector for any other subset of $p$% . Adding one point $(0,1,1)$ from the right to $p$, we will also have a projection cycle, but not minimal. Note that in this case, $\lambda $ can be taken as $(3,-1,-1,-2,2,-1).$ \bigskip \textbf{Remark 3.7.} A minimal projection cycle under the name of a \textit{% loop} was introduced and used in the works of Klopotowski, Nadkarni, Rao \cite{81,80}. \bigskip To prove our main result we need some auxiliary facts. \bigskip \textbf{Lemma 3.6.} (1)\textit{\ The vector $\lambda =(\lambda _{1},...,\lambda _{m})$ associated with a minimal projection cycle $% p=(x_{1},...,x_{m})$ is unique up to multiplication by a constant.} (2)\textit{\ If in (1), $\sum_{j=1}^{m}\left\vert \lambda _{j}\right\vert =1, $ then all the numbers $\lambda _{j}$, $j=1,...,m,$ are rational.} \bigskip \begin{proof} Let $\lambda ^{1}=(\lambda _{1}^{1},...,\lambda _{m}^{1})$ and $% \lambda ^{2}=(\lambda _{1}^{2},...,\lambda _{m}^{2})$ be any two vectors associated with $p.$ That is, \begin{equation*} \sum_{j=1}^{m}\lambda _{j}^{1}\chi _{\pi _{i}(x_{j})}=0\text{ and }% \sum_{j=1}^{m}\lambda _{j}^{2}\chi _{\pi _{i}(x_{j})}=0,\text{ \ }i=1,...,n. \end{equation*}% After multiplying the second equality by $c=\frac{\lambda _{1}^{1}}{\lambda _{1}^{2}}$ and subtracting from the first, we obtain that \begin{equation*} \sum_{j=2}^{m}(\lambda _{j}^{1}-c\lambda _{j}^{2})\chi _{\pi _{i}(x_{j})}=0% \text{, \ }i=1,...,n. \end{equation*}% Now since the cycle $p$ is minimal, $\lambda _{j}^{1}=c\lambda _{j}^{2},$ for all $j=1,...,m.$ The second part of the lemma is a consequence of the first part. Indeed, let $n=(n_{1},...,n_{m})$ be a vector with the nonzero integer coordinates associated with $p.$ Then the vector $\lambda ^{^{\prime }}=(\lambda _{1}^{^{\prime }},...,\lambda _{m}^{^{\prime }}),$ where $\lambda _{j}^{^{\prime }}=\frac{n_{j}}{\sum_{j=1}^{m}\left\vert n_{j}\right\vert },$ $j=1,...,m,$ is also associated with $p.$ All coordinates of $\lambda ^{^{\prime }}$ are rational and therefore by the first part of the lemma, it is the unique vector satisfying $\sum_{j=1}^{m}\left\vert \lambda _{j}^{^{\prime }}\right\vert =1.$ \end{proof} By this lemma, a minimal projection cycle $p$ uniquely (up to a sign) defines the measure \begin{equation*} ~\mu _{p}=\sum_{j=1}^{m}\lambda _{j}\delta _{x_{j}},\text{ \ }% \sum_{j=1}^{m}\left\vert \lambda _{j}\right\vert =1. \end{equation*} \bigskip \textbf{Lemma 3.7}. \textit{Let $\mu $ be a normalized orthogonal measure on a projection cycle $l\subset X$. Then it is a convex combination of normalized orthogonal measures on minimal projection cycles of $l$. That is,} \begin{equation*} \mu =\sum_{i=1}^{s}t_{i}\mu _{l_{i}},\text{ }\sum_{i=1}^{s}t_{i}=1,~t_{i}>0, \end{equation*} \textit{where $l_{i},$ $i=1,...,s,$ are minimal projection cycles in $l.$} \bigskip This lemma follows from the result of Navada (see \cite[Theorem 2]{112}): Let $S\subset X_{1}\times \cdots \times X_{n}$ be a finite set. Then any extreme point of the convex set of measures $\mu $ on $S$, $\mu \in M^{\bot } $, $\left\Vert \mu \right\Vert \leq 1$, has its support on a minimal projection cycle contained in $S$. \bigskip \textbf{Remark 3.8.} In the case $n=2$, Lemma 3.7 was proved by Medvedev (see \cite[p.77]{76}). \bigskip \textbf{Lemma 3.8} (see \cite[p.73]{76}). \textit{Let $X=X_{1}\times \cdots \times X_{n}$ and $\pi _{i}$ be the projections of $X$ onto the sets $X_{i},$ $i=1,...,n.$ In order that a measure $\mu \in C(X)^{\ast }$ be orthogonal to the subspace $M$, it is necessary and sufficient that} \begin{equation*} \mu \circ \pi _{i}^{-1}=0,\text{ }i=1,...,n. \end{equation*} \bigskip \textbf{Lemma 3.9} (see \cite[p.75]{76}). \textit{Let $\mu \in M^{\bot }$ and $\left\Vert \mu \right\Vert =1.$ Then there exist a net of measures $% \{\mu _{\alpha }\}\subset M^{\bot }$ weak$^{\text{*}}$ converging in $% C(X)^{\ast }$ to $\mu $ and satisfying the following properties:} 1) $\left\Vert \mu _{\alpha }\right\Vert =1;$ 2) \textit{The closed support of each $\mu _{\alpha }$ is a finite set.} \bigskip Our main result is the following theorem. \bigskip \textbf{Theorem 3.10.} \textit{The error of approximation from the manifold $% M$ obeys the equality} \begin{equation*} E(f)=\underset{l\subset X}{\sup }\left\vert \int\limits_{X}fd\mu _{l}\right\vert , \end{equation*}% \textit{where the $\sup $ is taken over all minimal projection cycles of $X.$% } \bigskip \begin{proof} Let $\overset{\sim }{\mu }$ be a measure with finite support $% \{x_{1},...,x_{m}\}$ and orthogonal to the space $M.$ Put $\lambda _{j}=% \overset{\sim }{\mu }(x_{j}),$ $j=1,...m.$ By Lemma 3.8, $\overset{\sim }{% \mu }(\pi _{i}^{-1}(\pi _{i}(x_{j})))=0,$ for all $i=1,...,n,$ $j=1,...,m.$ Fix the indices $i$ and $j.$ Then we have the equation $\sum_{k}\lambda _{k}=0,$ where the sum is taken over all indices $k$ such that $\pi _{i}(x_{k})=\pi _{i}(x_{j}).$ Varying $i$ and $j,$ we obtain a system of such equations, which concisely can be written as \begin{equation*} \sum_{k=1}^{m}\lambda _{k}\chi _{\pi _{i}(x_{k})}=0,\text{ \ }i=1,...,n. \end{equation*}% This means that the finite support of $\overset{\sim }{\mu }$ forms a projection cycle. Therefore, a net of measures approximating the given measure $\mu $ in Lemma 3.9 are all of the form (3.38). Let now $\mu _{p,\lambda }$ be any measure of the form (3.38). Since $\mu _{p,\lambda }\in M^{\bot }$ and $\left\Vert \mu _{p,\lambda }\right\Vert =1,$ we can write \begin{equation*} \left\vert \int\limits_{X}fd\mu _{p,\lambda }\right\vert =\left\vert \int\limits_{X}(f-g)d\mu _{p,\lambda }\right\vert \leq \left\Vert f-g\right\Vert ,\eqno(3.39) \end{equation*}% where $g$ is an arbitrary function in $M$. It follows from (3.39) that \begin{equation*} \underset{\left\langle p,\lambda \right\rangle }{\sup }\left\vert \int\limits_{X}fd\mu _{p,\lambda }\right\vert \leq E(f),\eqno(3.40) \end{equation*}% where the $\sup $ is taken over all projection cycle-vector pairs of $X.$ Consider the general duality relation (3.32). Let $\mu _{0}$ be a measure attaining the supremum in (3.32) and $\left\{ \mu _{p,\lambda }\right\} $ be a net of measures of the form (3.38) approximating $\mu _{0}$ in the weak$^{% \text{*}}$ topology of $C(X)^{\ast }.$ We already know that this is possible. For any $\varepsilon >0,$ there exists a measure $\mu _{p_{0},\lambda _{0}}$ in $\left\{ \mu _{p,\lambda }\right\} $ such that \begin{equation*} \left\vert \int\limits_{X}fd\mu _{0}-\int\limits_{X}fd\mu _{p_{0},\lambda _{0}}\right\vert <\varepsilon . \end{equation*}% From the last inequality we obtain that \begin{equation*} \left\vert \int\limits_{X}fd\mu _{p_{0},\lambda _{0}}\right\vert >\left\vert \int\limits_{X}fd\mu _{0}\right\vert -\varepsilon =E(f)-\varepsilon . \end{equation*}% Hence, \begin{equation*} \underset{\left\langle p,\lambda \right\rangle }{\sup }\left\vert \int\limits_{X}fd\mu _{p,\lambda }\right\vert \geq E(f).\eqno(3.41) \end{equation*}% From (3.40) and (3.41) it follows that \begin{equation*} \underset{\left\langle p,\lambda \right\rangle }{\sup }\left\vert \int\limits_{X}fd\mu _{p,\lambda }\right\vert =E(f).\eqno(3.42) \end{equation*} By Lemma 3.7, \begin{equation*} \mu _{p,\lambda }=\sum_{i=1}^{s}t_{i}\mu _{l_{i}}, \end{equation*}% where $l_{i}$, $i=1,...,s,$ are minimal projection cycles in $p$ and $% \sum_{i=1}^{s}t_{i}=1,~t_{i}>0.$ Let $k$ be an index in the set $\{1,...,s\} $ such that \begin{equation*} \left\vert \int\limits_{X}fd\mu _{l_{k}}\right\vert =\max \left\{ \left\vert \int\limits_{X}fd\mu _{l_{i}}\right\vert ,\text{ }i=1,...,s\right\} . \end{equation*}% Then \begin{equation*} \left\vert \int\limits_{X}fd\mu _{p,\lambda }\right\vert \leq \left\vert \int\limits_{X}fd\mu _{l_{k}}\right\vert .\eqno(3.43) \end{equation*}% Now since \begin{equation*} \left\vert \int\limits_{X}fd\mu _{l}\right\vert \leq E(f), \end{equation*}% for any minimal cycle $l,$ from (3.42) and (3.43) we obtain the assertion of the theorem. \end{proof} \textbf{Remark 3.9.} Theorem 3.10 not only proves Golomb's formula, but also improves it. Indeed, based on Lemma 3.5, one can easily observe that the formula (3.35) is equivalent to the formula \begin{equation*} E(f)=\underset{\left\langle p,\lambda \right\rangle }{\sup }\left\vert \int\limits_{X}fd\mu _{p,\lambda }\right\vert , \end{equation*}% where the $\sup $ is taken over all projection cycle-vector pairs $% \left\langle p,\lambda \right\rangle $ of $X$ provided that all the numbers $% \lambda _{i}\diagup \sum_{j=1}^{m}\left\vert \lambda _{j}\right\vert $, $% i=1,...,m,$ are rational. But by Lemma 3.6, minimal projection cycles enjoy this property. \newpage \chapter{Generalized ridge functions and linear superpositions} A ridge function $g(\mathbf{a}\cdot \mathbf{x})$ with a direction $\mathbf{a}% \in \mathbb{R}^{d}\backslash \{\mathbf{0}\}$ admits a natural generalization to a multivariate function of the form $g(\alpha _{1}(x_{1})+\cdot \cdot \cdot +\alpha _{d}(x_{d}))$, where $\alpha _{i}(x_{i})$, $i=\overline{1,d},$ are real, presumably well behaved, fixed univariate functions. We know from Chapter 1 that finitely many directions $\mathbf{a}^{j}$ are not enough for sums $\sum g_{j}\left( \mathbf{a}^{j}\cdot \mathbf{x}\right) $ to approximate multivariate functions. However, we will see in this chapter that sums of the form $\sum g_{j}(\alpha _{1}^{j}(x_{1})+\cdot \cdot \cdot +\alpha _{d}^{j}(x_{d}))$ with finitely many $\alpha _{i}^{j}(x_{i})$ is capable not only approximating multivariate functions but also precisely representing them. First we study the problem of representation of a function $f:X\rightarrow \mathbb{R}$, where $X$ is any set, as a linear superposition $\sum_{j}g_{j}(h_{j}(x))$ with arbitrary but fixed functions $% h_{j}:X\rightarrow {{\mathbb{R}}}$. Then we apply the obtained result and the famous Kolmogorov superposition theorem to prove representability of an arbitrarily behaved multivariate function in the form of a generalized ridge function $\sum g_{j}(\alpha _{1}^{j}(x_{1})+\cdot \cdot \cdot +\alpha _{d}^{j}(x_{d}))$. We also study the uniqueness of representation of functions by linear superpositions. The material of this chapter is taken from \cite{49,Ism}. \bigskip \section{Representation theorems} In this section, we study some problems of representation of real functions by linear superpositions and linear combinations of generalized ridge functions. \subsection{Problem statement and historical notes} Let $X$ be any set and $h_{i}:X\rightarrow {{\mathbb{R}}},~i=1,...,r,$ be arbitrarily fixed functions. Consider the set \begin{equation*} \mathcal{B}(X)=\mathcal{B}(h_{1},...,h_{r};X)=\left\{ \sum\limits_{i=1}^{r}g_{i}(h_{i}(x)),~x\in X,~g_{i}:\mathbb{R}\rightarrow \mathbb{R},~i=1,...,r\right\} \eqno(4.1) \end{equation*}% Members of this set will be called linear superpositions with respect to the functions $h_{1},...,h_{r}$ (see \cite{141}). For a detailed study of linear superpositions and their approximation-theoretic properties we refer the reader to the monograph by Khavinson \cite{76}. Note that sums of generalized ridge functions $\sum g_{j}(\alpha _{1}^{j}(x_{1})+\cdot \cdot \cdot +\alpha _{d}^{j}(x_{d}))$ with fixed $\alpha _{i}^{j}(x_{i})$ are a special case of linear superpositions. In Section 1.2, we considered linear superpositions defined on a subset of the $d$% -dimensional Euclidean space, while here $X$ is a set of arbitrary nature. As in Section 1.2, we are interested in the question: what conditions on $X$ guarantee that each function on $X$ will be in the set $\mathcal{B}(X)$? The simplest case $X\subset \mathbb{R}^{d},~r=d$ and $h_{i}$ are the coordinate functions was solved in \cite{81}. See also \cite[p.57]{76} for the case $% r=2.$ By $\mathcal{B}_{c}(X)$ and $\mathcal{B}_{b}(X)$ denote the right hand side of (4.1) with continuous and bounded $g_{i}:\mathbb{R}\rightarrow \mathbb{R}% ,~i=1,...,r,$ respectively. Our starting point is the well-known superposition theorem of Kolmogorov \cite{83}. It states that for the unit cube $\mathbb{I}^{d},~\mathbb{I}=[0,1],~d\geq 2,$ there exists $2d+1$ functions $\{s_{q}\}_{q=1}^{2d+1}\subset C(\mathbb{I}^{d})$ of the form \begin{equation*} s_{q}(x_{1},...,x_{d})=\sum_{p=1}^{d}\varphi _{pq}(x_{p}),~\varphi _{pq}\in C(\mathbb{I}),~p=1,...,d,~q=1,...,2d+1\eqno(4.2) \end{equation*}% such that each function $f\in C(\mathbb{I}^{d})$ admits the representation \begin{equation*} f(x)=\sum_{q=1}^{2d+1}g_{q}(s_{q}(x)),~x=(x_{1},...,x_{d})\in \mathbb{I}% ^{d},~g_{q}\in C({{\mathbb{R)}}}.\eqno(4.3) \end{equation*} Note that the functions $g_{q}(s_{q}(x))$, involved in the right hand side of (4.3), are generalized ridge functions. In our notation, (4.3) means that $% \mathcal{B}_{c}(s_{1},...,s_{2d+1};\mathbb{I}^{d})=C(\mathbb{I}^{d}).$ This surprising and deep result, which solved (negatively) Hilbert's 13-th problem, was improved and generalized in several directions. It was first observed by Lorentz \cite{98} that the functions $g_{q}$ can be replaced by a single continuous function $g.$ Sprecher \cite{128} showed that the theorem can be proven with constant multiples of a single function $\varphi $ and translations. Specifically, $\varphi _{pq}$ in (4.2) can be chosen as $% \lambda ^{p}\varphi (x_{p}+\varepsilon q),$ where $\varepsilon $ and $% \lambda $ are some positive constants. Fridman \cite{31} succeeded in showing that the functions $\varphi _{pq}$ can be constructed to belong to the class $Lip(1).$ Vitushkin and Henkin \cite{141} showed that $\varphi _{pq}$ cannot be taken to be continuously differentiable. Ostrand \cite{115} extended the Kolmogorov theorem to general compact metric spaces. In particular, he proved that for each compact $d$-dimensional metric space $X$ there exist continuous real functions $\{\alpha _{i}\}_{i=1}^{2d+1}\subset C(X)$ such that $\mathcal{B}_{c}(\alpha _{1},...,\alpha _{2d+1};X)=C(X).$ Sternfeld \cite{130} showed that the number $2d+1$ cannot be reduced for any $d$-dimensional space $X.$ Thus the number of terms in the Kolmogorov superposition theorem is the best possible. Some papers of Sternfeld were devoted to the representation of continuous and bounded functions by linear superpositions. Let $C(X)$ and $B(X)$ denote the space of continuous and bounded functions on some set $X$ respectively (in the first case, $X$ is supposed to be a compact metric space). Let $% F=\{h\}$ be a family of functions on $X.$ $F$ is called a uniformly separating family (\textit{u.s.f.}) if there exists a number $0<\lambda \leq 1$ such that for each pair $\{x_{j}\}_{j=1}^{m}$, $\{z_{j}\}_{j=1}^{m}$ of disjoint finite sequences in $X$, there exists some $h\in F$ so that if from the two sequences $\{h(x_{j})\}_{j=1}^{m}$and $\{h(z_{j})\}_{j=1}^{m}$ in $% h(X)$ we remove a maximal number of pairs of points $h(x_{j_{1}})$ and $% h(z_{j_{2}})$ with $h(x_{j_{1}})=h(z_{j_{2}}),$ there remains at least $% \lambda m$ points in each sequence (or , equivalently, at most $(1-\lambda )m $ pairs can be removed). Sternfeld \cite{132} proved that for a finite family $F=\{h_{1},...,h_{r}\}$ of functions on $X$, being a \textit{u.s.f.} is equivalent to the equality $\mathcal{B}_{b}(h_{1},...,h_{r};X)=B(X),$ and that in the case where $X$ is a compact metric space and the elements of $F$ are continuous functions on $X$, the equality $\mathcal{B}% _{c}(h_{1},...,h_{r};X)=C(X)$ implies that $F$ is a \textit{u.s.f.} Thus, in particular, Sternfeld obtained that the formula (4.3) is valid for all bounded functions, where $g_{q}$ are bounded functions depending on $f$ (see also \cite[p.21]{76}). Let $X$ be a compact metric space. The family $F=\{h\}\subset C(X)$ is said to be a measure separating family (\textit{m.s.f.}) if there exists a number $0<\lambda \leq 1$ such that for any measure $\mu $ in $\ C(X)^{\ast },$ the inequality $\left\Vert \mu \circ h^{-1}\right\Vert \geq \lambda \left\Vert \mu \right\Vert $ holds for some $h\in F.$ Sternfeld \cite{131} proved that $% \mathcal{B}_{c}(h_{1},...,h_{r};X)=C(X)$ if and only if the family $% \{h_{1},...,h_{r}\}$ is a \textit{m.s.f.} In \cite{132}, it was also shown that if $r=2,$ then the properties \textit{u.s.f.} and \textit{m.s.f.} are equivalent. Therefore, the equality $\mathcal{B}_{b}(h_{1},h_{2};X)=B(X)$ is equivalent to $\mathcal{B}_{c}(h_{1},h_{2};X)=C(X).$ But for $r\,>2$, these two properties are no longer equivalent. That is, $\mathcal{B}% _{b}(h_{1},...,h_{r};X)=B(X)$ does not always imply $\mathcal{B}% _{c}(h_{1},...,h_{r};X)=C(X)$ (see \cite{131}). Our purpose is to consider the above mentioned problem of representation by linear superpositions without involving any topology (that of continuity or boundedness). We start with characterization of those sets $X$ for which $% \mathcal{B}(h_{1},...,h_{r};X)=T(X),$ where $T(X)$ is the space of all functions on $X.$ As in Section 1.2, this will be done in terms of cycles. We claim that nonexistence of cycles in $X$ is equivalent to the equality $% \mathcal{B}(X)=T(X)$ for an arbitrary set $X$. In particular, we show that $% \mathcal{B}_{c}(X)=C(X)$ always implies $\mathcal{B}(X)=T(X).$ This implication will enable us to obtain some new results, namely extensions of the previously known theorems from continuous to discontinuous multivariate functions. For example, we will prove that the formula (4.3) is valid for all discontinuous multivariate functions $f$ defined on the unite cube $% \mathbb{I}^{d},$ where $g_{q}$ are univariate functions depending on $f.$ \bigskip \subsection{Extension of Kolmogorov's superposition theorem} In this subsection, we show that if some representation by linear superpositions holds for continuous functions, then it holds for all functions. This will lead us to natural extensions of some known superposition theorems (such as Kolmogorov's superposition theorem, Ostrand's superposition theorem, etc) from continuous to discontinuous functions. In the sequel, by $\chi _{A}$ we will denote the characteristic function of a set $\ A\subset \mathbb{R}.$ That is, \begin{equation*} \chi _{A}(y)=\left\{ \begin{array}{c} 1,~if~y\in A \\ 0,~if~y\notin A.% \end{array}% \right. \end{equation*} The following definition is a generalized version of Definition 1.1 from Section 1.2, where in connection with ridge functions only subsets of $% \mathbb{R}^{d}$ were considered. \bigskip \textbf{Definition 4.1.} \textit{Given an arbitrary set $X$ and functions $% h_{i}:X\rightarrow \mathbb{R},~i=1,...,r$. A set of points $% \{x_{1},...,x_{n}\}\subset X$ is called to be a cycle with respect to the functions $h_{1},...,h_{r}$ (or, concisely, a cycle if there is no confusion), if there exists a vector $\lambda =(\lambda _{1},...,\lambda _{n})$ with the nonzero real coordinates $\lambda _{i},~i=1,...,n,$ such that } \begin{equation*} \sum_{j=1}^{n}\lambda _{j}\chi _{h_{i}(x_{j})}=0,~i=1,...,r.\eqno(4.4) \end{equation*} \textit{A cycle $p=\{x_{1},...,x_{n}\}$ is said to be minimal if $p$ does not contain any cycle as its proper subset.} \bigskip Note that in this definition the vector $\lambda =(\lambda _{1},\ldots ,\lambda _{n})$ can be chosen so that it has only integer components. Indeed, let for $i=1,...,r,$ the set $\{h_{i}(x_{j}),~j=1,...,n\}$ have $% k_{i}$ different values. Then it is not difficult to see that Eq. (4.4) stands for a system of $\sum_{i=1}^{r}k_{i}$ homogeneous linear equations in unknowns $\lambda _{1},...,\lambda _{n}.$ This system can be written in the matrix form $(\lambda _{1},\ldots ,\lambda _{n})\times C=0,$ where $C$ is an $n$ by $\sum_{i=1}^{r}k_{i}$ matrix. The basic property of this matrix is that all of its entries are 0's and 1's and no row or column of $C$ is identically zero. Since Eq. (4.4) has a nontrivial solution $(\lambda _{1}^{^{\prime }},\ldots ,\lambda _{n}^{^{\prime }})\in \mathbf{R}^{n}$ and all entries of $C$ are integers, by applying the Gauss elimination method we can see that there always exists a nontrivial solution $(\lambda _{1},\ldots ,\lambda _{n})$ with the integer components $\lambda _{i}$, $i=1,...,n$. For a number of simple examples, see Section 1.2. Let $T(X)$ denote the set of all functions on $X.$ With each pair $% \left\langle p,\lambda \right\rangle ,$ where $p=\{x_{1},...,x_{n}\}$ is a cycle in $X$ and $\lambda =(\lambda _{1},...,\lambda _{n})$ is a vector known from Definition 4.1, we associate the functional \begin{equation*} G_{p,\lambda }:T(X)\rightarrow \mathbb{R},~~G_{p,\lambda }(f)=\sum_{j=1}^{n}\lambda _{j}f(x_{j}). \end{equation*}% In the following, such pairs $\left\langle p,\lambda \right\rangle $ will be called \textit{cycle-vector pairs} of $X.$ It is clear that the functional $% G_{p,\lambda }$ is linear. Besides, $G_{p,\lambda }(g)=0$ for all functions $% g\in \mathcal{B}(h_{1},...,h_{r};X).$ Indeed, assume that (4.4) holds. Given $i\leq r$, let $z=h_{i}(x_{j})$ for some $j$. Hence, $% \sum_{j~(h_{i}(x_{j})=z)}\lambda _{j}=0$ and $\sum_{j~(h_{i}(x_{j})=z)}% \lambda _{j}g_{i}(h_{i}(x_{j}))=0$. A summation yields $G_{p,\lambda }(g_{i}\circ h_{i})=0$. Since $G_{p,\lambda }$ is linear, we obtain that $% G_{p,\lambda }(\sum_{i=1}^{r}g_{i}\circ h_{i})=0$. A minimal cycle $p=\{x_{1},...,x_{n}\}$ has the following obvious properties: \begin{description} \item[(a)] \textit{The vector $\lambda $ associated with $p$ by Eq. (4.4) is unique up to multiplication by a constant;} \item[(b)] \textit{If in (4.4), $\sum_{j=1}^{n}\left\vert \lambda _{j}\right\vert =1,$ then all the numbers $\lambda _{j},~j=1,...,n,$ are rational.} \end{description} Thus, a minimal cycle $p$ uniquely (up to a sign) defines the functional \begin{equation*} ~G_{p}(f)=\sum_{j=1}^{n}\lambda _{j}f(x_{j}),\text{ \ }\sum_{j=1}^{n}\left% \vert \lambda _{j}\right\vert =1. \end{equation*} \bigskip \textbf{Proposition 4.1.} \textit{1) Let $X$ have cycles. A function $% f:X\rightarrow \mathbb{R}$ belongs to the space $\mathcal{B}% (h_{1},...,h_{r};X)$ if and only if $G_{p}(f)=0$ for any minimal cycle $% p\subset X$ with respect to the functions $h_{1},...,h_{r}$.} \textit{2) Let $X$ has no cycles. Then $\mathcal{B}(h_{1},...,h_{r};X)=T(X).$% } \bigskip \textbf{Proposition 4.2.} \textit{$\mathcal{B}(h_{1},...,h_{r};X)=T(X)$ if and only if $X$ has no cycles.} \bigskip These propositions are proved by the same way as Theorems 1.1 and 1.2. We use these propositions to obtain our main result (see Theorem 4.1 below). The condition whether $X$ have cycles or not, depends both on $X$ and the functions $h_{1},...,h_{r}$. In the following, we see that if $% h_{1},...,h_{r}$ are ``nice" functions (smooth functions with the simple structure. For example, ridge functions) and $X\subset \mathbb{R}^{d}$ is a ``rich" set (for example, the set with interior points), then $X$ has always cycles. Thus the representability by linear combinations of univariate functions with the fixed ``nice" multivariate functions requires at least that $X$ should not possess interior points. The picture is quite different when the functions $h_{1},...,h_{r}$ are not ``nice". Even in the case when they are continuous, we will see that many sets of $\mathbb{R}^{d}$ (the unite cube, any compact subset of that, or even the whole space $\mathbb{R}% ^{d}$ itself) may have no cycles. If disregard the continuity, there exists even one function $h$ such that every multivariate function is representable as $g\circ h$ over any subset of $\mathbb{R}^{d}$. First, let us introduce the following definition. \bigskip \textbf{Definition 4.2.} \textit{Let $X$ be a set and $h_{i}:X\rightarrow \mathbb{R}, $ $i=1,...,r,$ be arbitrarily fixed functions. A class $A(X)$ of functions on $X$ will be called a ``permissible function class" if for any minimal cycle $p\subset X$ with respect to the functions $h_{1},...,h_{r}$ (if it exists), there is a function $f_{0}$ in $A(X)$ such that $% G_{p}(f_{0})\neq 0. $} \bigskip Clearly, $C(X)$ and $B(X)$ are both permissible function classes (in case of $C(X),$ $X$ is considered to be a normal topological space). \bigskip \textbf{Theorem 4.1.} \textit{Let $A(X)$ be a permissible function class. If $A(X) \subset \mathcal{B}(h_{1},...,h_{r};X)$, then $\mathcal{B}% (h_{1},...,h_{r};X)=T(X).$} \bigskip The proof is simple and based on Propositions 4.1 and 4.2. Assume for a moment that $X$ admits a cycle $p$. By Proposition 4.1, the functional $G_{p} $ annihilates all members of the set $B(h_{1},...,h_{r};X).$ By Definition 4.2 of permissible function classes, $A(X)\ $contains a function $f_{0}$ such that $G_{p}(f_{0})\neq 0.$ Therefore, $f_{0}\notin B(h_{1},...,h_{r};X)$% . We see that the embedding $A(X) \subset B(h_{1},...,h_{r};X)$ is impossible if $X$ has a cycle. Thus $X$ has no cycles. Then by Proposition 4.2, $\mathcal{B}(h_{1},...,h_{r};X)=T(X).$ In the ``if part" of Theorem 4.1, instead of $\mathcal{B}(h_{1},...,h_{r};X)$ and $A(X)$ one can take $\mathcal{B}_{c}(h_{1},...,h_{r};X)$ and $C(X)$ (or $% \mathcal{B}_{b}(h_{1},...,h_{r};X)$ and $B(X)$) respectively. That is, the following corollaries are valid. \bigskip \textbf{Corollary 4.1.} \textit{Let $X$ be a set and $h_{i}:X\rightarrow \mathbb{R},$ $i=1,...,r,$ be arbitrarily fixed bounded functions\textit{. If $\mathcal{B}_{b}(h_{1},...,h_{r};X)=B(X)$, then $\mathcal{B}% (h_{1},...,h_{r};X)=T(X).$}} \bigskip \textbf{Corollary 4.2.} \textit{Let $X$ be a normal topological space and $h_{i}:X\rightarrow \mathbb{R},$ $i=1,...,r,$ be arbitrarily fixed continuous functions\textit{. If $\mathcal{B}_{c}(h_{1},...,h_{r};X)=C(X)$, then $\mathcal{B}(h_{1},...,h_{r};X)=T(X).$}} \bigskip The main advantage of Theorem 4.1 is that we need not check directly if the set $X$ has no cycles, which in many cases may turn out to be very tedious task. Using this theorem, we can extend free-of-charge\ the existing superposition theorems from the classes $B(X)$ or $C(X)$ (or some other permissible function classes) to all functions defined on $X.$ For example, this theorem allows us to extend the Kolmogorov superposition theorem from continuous to all multivariate functions. \bigskip \textbf{Theorem 4.2.} \textit{Let $d\geq 2$, $\mathbb{I}=[-1;1]$, and $% ~\varphi _{pq}, ~p=1,...,d, ~q=1,...,2d+1$, be the universal continuous functions in (4.2). Then each multivariate function $f:\mathbb{I}% ^{d}\rightarrow \mathbb{R}$ can be represented in the form} \begin{equation*} f(x)=\sum_{q=1}^{2d+1}g_{q}(\sum_{p=1}^{d}\varphi _{pq}(x_{p})),~x=(x_{1},...,x_{d})\in \mathbb{I}^{d}. \end{equation*}% \textit{where $g_{q}$ are univariate functions depending on $f.$} \bigskip It should be remarked that Sternfeld \cite{132}, in particular, obtained that the formula (4.3) is valid for functions $f\in B(\mathbb{I}^{d})$ provided that $g_{q}$ are bounded functions depending on $f$ (see \cite[% Chapter 1]{76} for more detailed information and interesting discussions). Let $X$ be a compact metric space and $h_{i}\in C(X)$, $i=1,...,r.$ The result of Sternfeld (see Section 4.1) and Corollary 4.1 give us the implications \begin{equation*} \mathcal{B}_{c}(h_{1},...,h_{r};X)=C(X)\Rightarrow \mathcal{B}% _{b}(h_{1},...,h_{r};X)=B(X) \end{equation*} \begin{equation*} \Rightarrow \mathcal{B}(h_{1},...,h_{r};X)=T(X). \end{equation*} The first implication is invertible when $r=2$ (see \cite{132}). We want to show that the second is not invertible even in the case $r=2.$ The following interesting example is due to Khavinson \cite[p.67]{76}. Let $X\subset \mathbb{R}^{2}$ consist of a broken line whose sides are parallel to the coordinate axis and whose vertices are \begin{equation*} (0;0),(1;0),(1;1),(1+\frac{1}{2^{2}};1),(1+\frac{1}{2^{2}};1+\frac{1}{2^{2}} ),(1+\frac{1}{2^{2}}+\frac{1}{3^{2}};1+\frac{1}{2^{2}}),... \end{equation*} We add to this line the limit point of the vertices $(\frac{\pi ^{2}}{6},% \frac{\pi ^{2}}{6})$. Let $r=2$ and $h_{1},h_{2}$ be the coordinate functions. Then the set $X$ has no cycles with respect to $h_{1}$ and $% h_{2}. $ By Proposition 4.1, every function $f$ on $X$ is of the form $% g_{1}(x_{1})+g_{2}(x_{2})$, $(x_{1},x_{2})\in X$. Now construct a function $% f_{0}$ on $X$ as follows. On the link joining $(0;0)$ to $(1;0)$ $% f_{0}(x_{1},x_{2})$ continuously increases from $0$ to $1$; on the link from $(1;0)$ to $(1;1)$ it continuously decreases from $1$ to $0$; on the link from $(1;1)$ to $(1+\frac{1}{2^{2}};1)$ it increases from $0$ to $\frac{1}{2} $; on the link from $(1+\frac{1}{2^{2}};1)$ to $(1+\frac{1}{2^{2}};1+\frac{1% }{2^{2}})$ it decreases from $\frac{1}{2}$ to $0$; on the next link it increases from $0$ to $\frac{1}{3}$, etc. At the point $(\frac{\pi ^{2}}{6},% \frac{\pi ^{2}}{6})$ set the value of $f_{0}$ equal to $0.$ Obviously, $% f_{0} $ is a continuous functions and by the above argument, $% f_{0}(x_{1},x_{2})=g_{1}(x_{1})+g_{2}(x_{2}).$ But $g_{1}$ and $g_{2}$ cannot be chosen as continuous functions, since they get unbounded as $x_{1}$ and $x_{2}$ tends to $\frac{\pi ^{2}}{6}$. Thus, $\mathcal{B}% (h_{1},h_{2};X)=T(X)$, but at the same time $\mathcal{B}_{c}(h_{1},h_{2};X)% \neq C(X)$ (or, equivalently, $\mathcal{B}_{b}(h_{1},h_{2};X)\neq B(X)$). \bigskip \subsection{Some other superposition theorems} We have seen in the previous subsection that the unit cube in $\mathbb{R}% ^{d} $ has no cycles with respect to some $2d+1$ continuous functions (namely, the Kolmogorov functions $s_{q}$ (4.2)). From the result of Ostrand \cite{115} (see Section 4.1) and Corollary 4.2 it follows that compact sets $% X$ of finite dimension also lack cycles with respect to a certain family of finitely many continuous functions on $X$. Namely, the following generalization of Ostrand's theorem is valid. \bigskip \textbf{Theorem 4.3.} \textit{For $p=1,2,...,m$ let $X_{p}$ be a compact metric space of finite dimension $d_{p}$ and let $n=\sum_{p=1}^{n}d_{p}.$ There exist continuous functions $\alpha _{pq}:X_{p}\rightarrow \lbrack 0,1], $ $p=1,...,m,$ $q=1,...,2n+1,$ such that every real function $f$ defined on $\Pi _{p=1}^{m}X_{p}$ is representable in the form} \begin{equation*} f(x_{1},...,x_{m})=\sum_{q=1}^{2n+1}g_{q}(\sum_{p=1}^{m}\alpha _{pq}(x_{p})).% \eqno(4.5) \end{equation*}% \textit{where $g_{q}$ are real functions depending on $f$. If $f$ is continuous, then the functions $g_{q}$ can be chosen continuous.} \bigskip Note that Ostrand proved ``if $f$ is continuous..." part of Theorem 4.3, while we prove the validity of (4.5) for discontinuous $f$. One may ask if there exists a finite family of functions $\{h_{i}:\mathbb{R}% ^{d}\rightarrow \mathbb{R}\}_{i=1}^{n}$ such that any subset of $\mathbb{R}% ^{d}$ does not admit cycles with respect to this family? The answer is positive. This follows from the result of Demko \cite{23}: there exist $2d+1$ continuous functions $\varphi _{1},...,\varphi _{2d+1}$ defined on $\mathbb{R% }^{d}$ such that every bounded continuous function on $\mathbb{R}^{d}$ is expressible in the form $\sum_{i=1}^{2d+1}g\circ \varphi _{i}$ for some $% g\in C(\mathbb{R})$. This theorem together with Corollary 4.1 yield that every function on $\mathbb{R}^{d}$ is expressible in the form $% \sum_{i=1}^{2d+1}g_{i}\circ \varphi _{i}$ for some $g_{i}:\mathbb{R}% \rightarrow \mathbb{R},~i=1,...,2d+1$. We do not yet know if $g_{i}$ here can be replaced by a single univariate function. We also don't know if the number $2d+1$ can be reduced so that the whole space of $\mathbb{R}^{d}$ (or any $d$-dimensional compact subset of that, or at least the unit cube $% \mathbb{I}^{d}$) has no cycles with respect to some continuous functions $% \varphi _{1},...,\varphi _{k}:\mathbb{R}^{d}\rightarrow \mathbb{R}$, where $% k<2d+1$. One of the basic results of Sternfeld \cite{130} says that the dimension of a compact metric space $X$ equals $d$ if and only if there exist functions $\varphi _{1},...,\varphi _{2d+1}\in C(X)$ such that $% \mathcal{B}_{c}(\varphi _{1},...,\varphi _{2d+1};X)=C(X)$ and for any fmily $% \{\psi _{i}\}_{i=1}^{k}\subset C(X),$ $k<2d+1$, we have $\mathcal{B}% _{c}(\psi _{1},...,\psi _{k};X)\neq C(X).$ In particular, from this result it follows that the number of terms in the Kolmogorov superposition theorem cannot be reduced. But since the equalities $\mathcal{B}_{c}(X)=C(X)$ and $% \mathcal{B}(X)=T(X)$ are not equivalent, the above question on the nonexistence of cycles in $\mathbb{R}^{d}$ with respect to less than $2d+1$ continuous functions is far from trivial. If disregard the continuity, one can construct even one function $\varphi :% \mathbb{R}^{d}\rightarrow \mathbb{R}$ such that the whole space $\mathbb{R}% ^{d}$ will not possess cycles with respect to $\varphi $ and therefore, every function $f:\mathbb{R}^{d}\rightarrow \mathbb{R}$ will admit the representation $f=g\circ \varphi $ with some univariate $g$ depending on $f$% . Our argument easily follows from Corollary 4.2 and the result of Sprecher \cite{127}: for any natural number $d$, $d\geq 2$, there exist functions $% h_{p}:\mathbb{I}\rightarrow \mathbb{R}$, $p=1,...,d,$ such that every function $f\in C(\mathbb{I}^{d})$ can be represented in the form \begin{equation*} f(x_{1},...,x_{d})=g\left( \sum_{p=1}^{d}h_{p}(x_{p})\right) ,\eqno(4.6) \end{equation*}% where $g$ is a univariate (generally discontinuous) function depending on $f$% . Note that the function involved in the right hand side of (4.6) is a generalized ridge function. Thus, the result of Sprecher together with our result means that every multivariate function $f$ is representable as a generalized ridge function $g\left( \cdot \right) $ and if $f$ is continuous, then $g$ can be chosen continuous as well. \bigskip \textbf{Remark 4.1.} Concerning ordinary ridge functions $g(\mathbf{a}\cdot \mathbf{x})$, representation of every multivariate function by linear combinations of such functions may not be possible over many sets in $% \mathbb{R}^{d}$. For example, this is not possible for sets having interior points. More precisely, assume we are given finitely many nonzero directions $\mathbf{a}^{1},...,\mathbf{a}^{r}$ in $\mathbb{R}^{d}$. Then $\mathcal{R}% \left( \mathbf{a}^{1},...,\mathbf{a}^{r};X\right) \neq T(X)$ for any set $% X\subset \mathbb{R}^{d}$ with a nonempty interior. Indeed, let $\mathbf{y}$ be a point in the interior of $X$. Consider vectors $\mathbf{b}^{i}$, $% i=1,...,r,$ with sufficiently small coordinates such that $\mathbf{a}% ^{i}\cdot \mathbf{b}^{i}=0$, $i=1,...,r$. Note that the vectors $\mathbf{b}% ^{i}$, $i=1,...,r,$ can be chosen pairwise linearly independent. With each vector $\mathbf{\varepsilon }=(\varepsilon _{1},...,\varepsilon _{r})$, $% \varepsilon _{i}\in \{0,1\}$, $i=1,...,r,$ we associate the point \begin{equation*} \mathbf{x}_{\mathbf{\varepsilon }}=\mathbf{y+}\sum_{i=1}^{r}\varepsilon _{i}% \mathbf{b}^{i}. \end{equation*}% Since the coordinates of $\mathbf{b}^{i}$ are sufficiently small, we may assume that all the points $\mathbf{x}_{\mathbf{\varepsilon }}$ are in the interior of $X$. We correspond each point $\mathbf{x}_{\mathbf{\varepsilon }% } $ to the number $(-1)^{\left\vert \mathbf{\varepsilon }\right\vert }$, where $\left\vert \mathbf{\varepsilon }\right\vert =\varepsilon _{1}+\cdots +\varepsilon _{r}.$ One may easily verify that the pair $\left\langle \{% \mathbf{x}_{\mathbf{\varepsilon }}\},\{(-1)^{\left\vert \mathbf{\varepsilon }% \right\vert }\}\right\rangle $ is a cycle-vector pair of $X$. Therefore, by Proposition 4.2, $\mathcal{R}\left( \mathbf{a}^{1},...,\mathbf{a}% ^{r};X\right) \neq T(X).$ Note that the above method of construction of the set $\{\mathbf{x}_{\mathbf{% \varepsilon }}\}$ is due to Lin and Pinkus \cite{95}. \bigskip \textbf{Remark 4.2.} A different generalization of ridge functions was considered in Lin and Pinkus \cite{95}. This generalization involves multivariate functions of the form $g(A\mathbf{x})$, where $\mathbf{x}\in \mathbb{R}^{d}$ is the variable, $A$ is a fixed $d\times n$ matrix, $1\leq n<d$, and $g$ is a real-valued function defined on $\mathbb{R}^{n}$. For $% n=1,$ this reduces to a ridge function. \bigskip \section{Uniqueness theorems} Let $Q$ be a set such that every function on $Q$ can be represented by linear superpositions. This representation is generally not unique. But for some sets it may be unique provided that initial values of the representing functions are prescribed at some point of $Q$. In this section, we are going to study properties of such sets. All the obtained results are valid, in particular, for linear combinations of generalized ridge functions. \subsection{Formulation of the problem} Assume $X$ is an arbitrary set, $h_{i}:X\rightarrow \mathbb{R}$, $i=1,\ldots ,r$, are fixed functions and $\mathcal{B}(X)$ is the set defined in (4.1). Let $T(X)$ denote the set of all real functions on $X$. {Obviously, }$% \mathcal{B}(X)${\ is a linear subspace of $T(X)$. For a set }$Q\subset X$, let $T(Q)$ and $\mathcal{B}(Q)$ denote the restrictions of $T(X)$ and $% \mathcal{B}(X)$ to $Q$, respectively. Sets $Q$ with the property $\mathcal{B}% (Q)=T(Q)$ will be called \textit{representation sets}. Recall that Proposition 4.2 gives a complete characterization of such sets. For a representation set $Q$, we will also use the notation $Q\in RS.$ Here, $RS$ stands for the set of all representation sets in $X$. Let $Q\in RS.$ Clearly for a function $f$ defined on $Q$ the representation \begin{equation*} f(x)=\sum_{i=1}^{r}g_{i}(h_{i}(x)),~x\in Q\eqno(4.7) \end{equation*}% is not unique. We are interested in the uniqueness of such representation under some reasonable restrictions on the functions $g_{i}\circ h_{i}$. These restrictions may be various, but in this section, we require that the values of $g_{i}\circ h_{i}$ are prescribed at some point $x_{0}\in Q$. That is, we require that \begin{equation*} g_{i}(h_{i}(x_{0}))=a_{i},~i=1,...,r-1,\eqno(4.8) \end{equation*}% where $a_{i}$ are arbitrarily fixed real numbers. Is representation (4.7) subject to initial conditions (4.8) always unique? Obviously, not. We are going to identify those representation sets $Q$ for which representation (4.7) subject to conditions (4.8) is unique for all functions $% f:Q\rightarrow \mathbb{R}$. In the sequel, such sets $Q$ will be called \textit{unicity sets}. \bigskip \subsection{Complete representation sets} From Proposition 4.2 it is easy to obtain the following set-theoretic properties of representation sets: \bigskip (1) $Q\in RS$ $\Longleftrightarrow $ $A\in RS$ for every finite set $% A\subset Q$; (2) The union of any linearly ordered (under inclusion) system of representation sets is also a representation set (3) For any representation set $Q$ there is a maximal representation set, that is, a set $M\in RS$ such that $Q\subset M$ and for any $P\supset M$, $% P\in RS$ we have $P=M$. (4) If $M\subset X$ is a maximal representation set, then $h_{i}(M)=h_{i}(X)$% , $i=1,...,r$. \bigskip Properties (1) and (2) are obvious, since any cycle is a finite set. The (3)-rd property follows from (2) and Zorn's lemma. To prove property (4) note that if $x_{0}\in X$ and $h_{i}(x_{0})\notin h_{i}(M)$ for some $i$, one can construct the representation set $M\cup \{x_{0}\}$, which is bigger than $M$. But this is impossible, since $M$ is maximal. \bigskip \textbf{Definition 4.3.} \textit{A set $Q\subset X$ is called a complete representation set if $Q$ itself is a representation set and there is no other representation set $P$ such that $Q\subset P$ and $h_{i}(P)=h_{i}(Q)$, $i=1,...,r$.} \bigskip The set of all complete representation sets of $X$ will be denoted by $CRS$. Obviously, every representation set is contained in a complete representation set. That is, if $A\in RS$, then there exists $B\in CRS$ such that $h_{i}(B)=h_{i}(A),$ $i=1,...,r.$ It turns out that for the functions $% h_{1},...,h_{r}$, complete representation sets entirely characterize unicity sets. To prove this fact we need some auxiliary lemmas. \bigskip \textbf{Lemma 4.1.} \textit{Let $Q\subset X$ be a representation set and for some point $x_{0}\in Q$ the zero function representation} \begin{equation*} 0=\sum_{i=1}^{r}g_{i}(h_{i}(x)),\text{ \ }x\in Q, \end{equation*}% \textit{is unique, provided that $g_{i}(h_{i}(x_{0}))=0,$ $i=1,...,r-1$. That is, all the functions $g_{i}\equiv 0$ on the sets $h_{i}(Q)$, $% i=1,...,r.$ Then $Q\in CRS.$} \bigskip \begin{proof} Assume that $Q\notin CRS$. Then there exists a point $p\in X$ such that $p\notin Q$, $h_{i}(p)\in h_{i}(Q)$, for all $i=1,...,r,$ and $% Q^{^{\prime }}=Q\cup \{p\}$ is also a representation set. Consider a function $f_{0}:Q^{^{\prime }}\rightarrow \mathbb{R}$ such that $f_{0}(q)=0$% , for any $q\in Q$ and $f_{0}(p)=1.$ Since $Q^{^{\prime }}\in RS$, \begin{equation*} f_{0}(x)=\sum_{i=1}^{r}s_{i}(h_{i}(x)),\text{ \ }x\in Q^{^{\prime }}. \end{equation*}% Then \begin{equation*} f_{0}(x)=\sum_{i=1}^{r}g_{i}(h_{i}(x)),\text{ \ }x\in Q^{^{\prime }},\eqno% (4.9) \end{equation*}% where \begin{equation*} g_{i}(h_{i}(x))=s_{i}(h_{i}(x))-s_{i}(h_{i}(x_{0})),\text{ }i=1,...,r-1 \end{equation*}% and \begin{equation*} g_{r}(h_{r}(x))=s_{r}(h_{r}(x))+\sum_{i=1}^{r-1}s_{i}(h_{i}(x_{0})). \end{equation*}% \qquad A restriction of representation (4.9) to the set $Q$ gives the equality \begin{equation*} \sum_{i=1}^{r}g_{i}(h_{i}(x))=0,\text{ for all }x\in Q.\eqno(4.10) \end{equation*}% Note that $g_{i}(h_{i}(x_{0}))=0,$ $i=1,...,r-1.$ It follows from the hypothesis of the lemma that representation (4.10) is unique. Hence, $% g_{i}(h_{i}(x))=0,$ for all $x\in Q$ and $i=1,...,r.$ But from (4.9) it follows that \begin{equation*} \sum_{i=1}^{r}g_{i}(h_{i}(p))=f_{0}(p)=1. \end{equation*}% Since $h_{i}(p)\in h_{i}(Q)$ for all $i=1,...,r,$ the above relation contradicts that the functions $g_{i}$ are identically zero on the sets $% h_{i}(Q)$, $i=1,...,r.$ This means that our assumption is not true and $Q\in CRS.$ \end{proof} The following lemma is a strengthened version of Lemma 4.1. \bigskip \textbf{Lemma 4.2.} \textit{Let $Q\in RS$ and for some point $x_{0}\in Q$, numbers $c_{1},c_{2},...,c_{r-1}\in \mathbb{R}$ and a function $v\in T(Q)$ the representation} \begin{equation*} v(x)=\sum_{i=1}^{r}v_{i}(h_{i}(x)) \end{equation*}% \textit{is unique under the initial conditions $v_{i}(h_{i}(x_{0}))=c_{i},$ $% i=1,...,r-1$. Then for any numbers $b_{1},b_{2}...,b_{r-1}\in \mathbb{R}$ and an arbitrary function $f\in T(Q)$ the representation} \begin{equation*} f(x)=\sum_{i=1}^{r}f_{i}(h_{i}(x)) \end{equation*}% \textit{is also unique, provided that $f_{i}(h_{i}(x_{0}))=b_{i},$ $% i=1,...,r-1$. Besides, $Q\in CRS.$} \bigskip \begin{proof} Assume the contrary. Assume that there is a function $f\in T(Q)$ having two different representations subject to the same initial conditions. That is, \begin{equation*} f(x)=\sum_{i=1}^{r}f_{i}(h_{i}(x))=\sum_{i=1}^{r}f_{i}^{^{\prime }}(h_{i}(x)) \end{equation*}% with $f_{i}(h_{i}(x_{0}))=f_{i}^{^{\prime }}(h_{i}(x_{0}))=b_{i},$ $% i=1,...,r-1$ and $f_{i}\neq f_{i}^{^{\prime }}$ for some indice $i\in \{1,...,r\}.$ In this case, the function $v(x)$ will possess the following two different representations \begin{equation*} v(x)=\sum_{i=1}^{r}v_{i}(h_{i}(x))=\sum_{i=1}^{r}\left[ v_{i}(h_{i}(x))+f_{i}(h_{i}(x))-f_{i}^{^{\prime }}(h_{i}(x))\right] . \end{equation*}% both satisfying the initial conditions. The obtained contradiction and above Lemma 4.1 complete the proof. \end{proof} In the sequel, we will assume that for any points $t_{i}\in h_{i}(X),$ $% i=1,...,r,$ the system of equations $h_{i}(x)=t_{i}$, $i=1,...,r,$ has at least one solution. \bigskip \textbf{Lemma 4.3.} \textit{Let $Q\in CRS.$ Then for any point $x_{0}\in Q$ the representation} \begin{equation*} 0=\sum_{i=1}^{r}g_{i}(h_{i}(x)),\text{ }x\in Q,\eqno(4.11) \end{equation*}% \textit{subject to the conditions} \begin{equation*} g_{i}(h_{i}(x_{0}))=0,\text{ }i=1,...,r-1,\eqno(4.12) \end{equation*}% \textit{is unique. That is, $g_{i}\equiv 0$ on the sets $h_{i}(Q)$, $% i=1,...,r.$} \bigskip \begin{proof} Assume the contrary. Assume that representation (4.11) subject to (4.12) is not unique, or in other words, not all of $g_{i}$ are identically zero. Without loss of generality, we may suppose that $g_{r}(h_{r}(y))\neq 0, $ for some $y\in Q.$ Let $\xi \in X$ be a solution of the system of equations $h_{i}(x)=h_{i}(x_{0}),$ $i=1,...,r-1,$ and $h_{r}(x)=h_{r}(y)$. Therefore, $g_{i}(h_{i}(\xi ))=0,$ $i=1,...,r-1,$ and $g_{r}(h_{r}(\xi ))\neq 0.$ Obviously, $\xi \notin Q.$ Otherwise, we may have $% g_{r}(h_{r}(\xi ))=0.$ We are going to prove that $Q^{\prime }=Q\cup \{\xi \}$ is a representation set. For this purpose, consider an arbitrary function $f:Q^{\prime }\rightarrow \mathbb{R}$. The restriction of $f$ to the set $Q$ admits a decomposition \begin{equation*} f(x)=\sum_{i=1}^{r}t_{i}(h_{i}(x)),\text{ }x\in Q. \end{equation*} One is allowed to fix the values $t_{i}(h_{i}(x_{0}))=0,$ $i=1,...,r-1.$ Note that then $t_{i}(h_{i}(\xi ))=0,$ $i=1,...,r-1.$ Consider now the functions \begin{equation*} v_{i}(h_{i}(x))=t_{i}(h_{i}(x))+\frac{f(\xi )-t_{r}(h_{r}(\xi ))}{% g_{r}(h_{r}(\xi ))}g_{i}(h_{i}(x)),\text{ }x\in Q^{\prime },\text{ }% i=1,...,r. \end{equation*} It can be easily verified that \begin{equation*} f(x)=\sum_{i=1}^{r}v_{i}(h_{i}(x)),\text{ }x\in Q^{\prime }. \end{equation*}% Since $f$ is arbitrary, we obtain that $Q^{\prime }\in RS,$ where $Q^{\prime }\supset Q$ and $h_{i}(Q^{\prime })=h_{i}(Q),$ $i=1,...,r.$ But this contradicts the hypothesis of the lemma that $Q\in CRS$. \end{proof} The following theorem is valid. \bigskip \textbf{Theorem 4.4.} \textit{$Q\in CRS$ if and only if for any $x_{0}\in Q,$ any $f\in T(Q)$ and any $a_{1},...,a_{r-1}\in \mathbb{R}$ the representation} \begin{equation*} f(x)=\sum_{i=1}^{r}g_{i}(h_{i}(x)),\text{ }x\in Q, \end{equation*}% \textit{subject to the conditions $g_{i}(h_{i}(x_{0}))=a_{i},$ $i=1,...,r-1,$ is unique. Equivalently, a set $Q\in CRS$ if and only if it is a unicity set.% } \bigskip Theorem 4.4 is an obvious consequence of Lemmas 4.2 and 4.3. \bigskip \textbf{Remark 4.3.} In Theorem 4.4, all the words "any" can be replaced with the word "some". \bigskip \textbf{Remark 4.4. }For the case $X=X_{1}\times \cdot \cdot \cdot \times X_{n}$, the possibility and uniqueness of the representation by sums $% \sum_{i=1}^{n}u_{i}(x_{i})$, \thinspace $u_{i}:X_{i}\rightarrow \mathbb{R}$, $i=1,...,n$, were investigated in \cite{81,80}. \bigskip \textbf{Examples.} Let $r=2,$ $X=\mathbb{R}^{2},$ $% h_{1}(x_{1},x_{2})=x_{1}+x_{2},$ $h_{2}(x_{1},x_{2})=x_{1}-x_{2},$ $Q$ be the graph of the function $x_{2}=\arcsin (\sin x_{1}).$ The set $Q$ has no cycles with respect to the functions $h_{1}$ and $h_{2}.$ Therefore, by Proposition 4.2, $Q\in RS.$ By adding a point $p\notin Q$, we obtain the set $Q\cup \{p\},$ which contains a cycle and hence is not a representation set. Thus, $Q\in CRS$ and hence $Q$ is a unicity set. Let now $r=2,$ $X=\mathbb{R}^{2},$ $h_{1}(x_{1},x_{2})=x_{1},$ $% h_{2}(x_{1},x_{2})=x_{2},$ and $Q$ be the graph of the function $x_{2}=x_{1}. $ Clearly, $Q\in RS$ and $Q\notin CRS.$ By the definition of complete representation sets, there is a set $P\supset Q$ such that $P\in RS$ and for any $T\supset P$, $T$ is not a representation set. There are many sets $P$ with this property. One of them can be obtained by adding to $Q$ any straight line $l$ parallel to one of the coordinate axes. Indeed, if $% y\notin Q\cup l,$ then the set $Q_{1}=Q\cup l\cup \{y\}$ contains a four-point cycle (with one vertex as $y$, two vertices lying on $l$ and one vertex lying on $Q$). This means that $Q_{1}\notin RS$ and hence $Q\cup l\in CRS.$ \bigskip The following corollary can be easily obtained from Theorem 4.4 and Lemma 4.2. \bigskip \textbf{Corollary 4.3.} \textit{$Q\in CRS$ if and only if $Q\in RS$ and in the representation} \begin{equation*} 0=\sum_{i=1}^{r}g_{i}(h_{i}(x)),\text{ }x\in Q, \end{equation*}% \textit{all the functions $g_{i},$ $i=1,...,r,$ are constants.} \bigskip We have seen that complete representation sets enjoy the unicity property. Let us study some other properties of these sets. The following properties are valid. \bigskip (a) If $Q_{1},Q_{2}\in CRS,$ $Q_{1}\cap Q_{2}\neq \emptyset $ and $Q_{1}\cup Q_{2}\in RS$, then $Q_{1}\cup Q_{2}\in CRS.$ (b) Let $\{Q_{\alpha }\},$ $\alpha \in \Phi ,$ be a family of complete representation sets such that $\cap _{\alpha \in \Phi }Q_{\alpha }\neq \emptyset $ and $\cup _{\alpha \in \Phi }Q_{\alpha }\in RS.$ Then $\cup _{\alpha \in \Phi }Q_{\alpha }\in CRS.$ \bigskip The above two properties follow from Corollary 4.3. Note that (b) is a generalization of (a). The following property is a consequence of (b) and property (2) of representation sets. \bigskip (c) Let $\{Q_{\alpha }\},$ $\alpha \in \Phi ,$ be a totally ordered (under inclusion) family of complete representation sets. Then $\cup _{\alpha \in \Phi }Q_{\alpha }\in CRS.$ \bigskip We know that every representation set $A$ is contained in a complete representation set $Q$ such that $h_{i}(A)=h_{i}(Q),$ $i=1,...,r.$ What can we say about the set $Q\backslash A$? Clearly, $Q\backslash A\in RS.$ But can we chose $Q$ so that $Q\backslash A\in CRS$? The following theorem answers this question. \bigskip \textbf{Theorem 4.5.} \textit{Let $A\in RS$ and $A\notin CRS.$ Then there exists a set $B\in CRS$ such that $A\subset B,$ $h_{i}(A)=h_{i}(B),$ $% i=1,...,r,$ and $B\backslash A\in CRS.$} \bigskip \begin{proof} Since the representation set $A$ is not complete, there exists a point $p\notin A$ such that $h_{i}(p)\in h_{i}(A),$ $i=1,...,r,$ and $% A^{\prime }=A\cup \{p\}\in RS$. By $\mathcal{M}$ denote the collection of sets $M$ such that 1) $A\subset M$ and $M\in RS$; 2) $h_{i}(M)=h_{i}(A)$ for all $i=1,...,r$; 3) $M\backslash A\in CRS.$ Obviously, $\mathcal{M}$ is not empty. It contains the above set $A^{\prime } $. Consider the partial order on $\mathcal{M}$ defined by inclusion. Let $% \{M_{\beta }\},$ $\beta \in \Gamma $, be any chain in $\mathcal{M}$. The set $\cup _{\beta \in \Gamma }M_{\beta }$ is an upper bound for this chain. To see this, let us check that $\cup _{\beta \in \Gamma }M_{\beta }$ belongs to $\mathcal{M}$. That is, all the above conditions 1)-3) are satisfied. Indeed, 1) $A\subset \cup _{\beta \in \Gamma }M_{\beta }$ and $\cup _{\beta \in \Gamma }M_{\beta }\in RS.$ This follows from property (2) of representation sets; 2) $h_{i}(\cup _{\beta \in \Gamma }M_{\beta })=\cup _{\beta \in \Gamma }h_{i}(M_{\beta })=\cup _{\beta \in \Gamma }h_{i}(A)=h_{i}(A),$ $i=1,...,r$; 3) $\cup _{\beta \in \Gamma }M_{\beta }\backslash A\in CRS$. This follows from property (c) of complete representation sets and the facts that $% M_{\beta }\backslash A\in CRS$ for any $\beta \in \Gamma $ and the system $% \{M_{\beta }\backslash A\}$, $\beta \in \Gamma $, is totally ordered under inclusion. Thus we see that any chain in $\mathcal{M}$ has an upper bound. By Zorn's lemma, there are maximal sets in $\mathcal{M}$. Assume $B$ is one of such sets. Let us now prove that $B\in CRS$. Assume on the contrary that $B\notin CRS$. Then by Lemma 4.2, for any point $x_{0}\in B$ the representation \begin{equation*} 0=\sum_{i=1}^{r}g_{i}(h_{i}(x)),\text{ }x\in B,\eqno(4.13) \end{equation*}% subject to the conditions $g_{i}(h_{i}(x_{0}))=0,$ $i=1,...,r-1,$ is not unique. That is, there is a point $y\in B$ such that for some index $i,$ $% g_{i}(h_{i}(y))\neq 0.$ Without loss of generality we may assume that $% g_{r}(h_{r}(y))\neq 0$. Clearly, $y$ cannot belong to $B\backslash A$, since $B\backslash A\in CRS$ and over complete representation sets, the zero function has a trivial representation provided that conditions (4.12) hold. Thus, $y\in A$. Let $\xi \in X$ be a point such that $h_{i}(\xi )=h_{i}(x_{0}),$ $i=1,...,r-1$, and $h_{r}(\xi )=h_{r}(y).$ The point $\xi \notin B,$ otherwise from (4.13) we would obtain that $% g_{r}(h_{r}(y))=g_{r}(h_{r}(\xi ))=0$. Following the techniques in the proof of Lemma 4.3, it can be shown that $B_{1}=B\cup \{\xi \}\in RS$. Now we prove that $B_{1}\backslash A\in CRS$. Consider the representation \begin{equation*} 0=\sum_{i=1}^{r}g_{i}^{\prime }(h_{i}(x)),\text{ }x\in B_{1}\backslash A,% \eqno(4.14) \end{equation*}% subject to the conditions $g_{i}^{\prime }(h_{i}(x_{0}))=0,$ $i=1,...,r-1,$ where \ $x_{0}$ is some point in $B\backslash A.$ Such representation holds uniquely on $B\backslash A,$ since $B\backslash A\in CRS$. That is, all the functions $g_{i}^{\prime }$ are identically zero on $h_{i}(B\backslash A),$ $% i=1,...,r$. On the other hand, since $g_{i}^{\prime }(h_{i}(\xi ))=g_{i}^{\prime }(h_{i}(x_{0}))=0$, for all $i=1,...,r-1$, we obtain that $% g_{r}^{\prime }(h_{r}(\xi ))=0.$ This means that representation (4.14) subject to the conditions $g_{i}^{\prime }(h_{i}(x_{0}))=0,$ $i=1,...,r-1,$ is unique on $B_{1}\backslash A.$ That is, all the functions $g_{i}^{\prime } $ in (4.14) are zero functions on $h_{i}(B_{1}\backslash A),$ $i=1,...,r.$ Hence by Lemma 4.1, $B_{1}\backslash A\in CRS$. Thus, $B_{1}\in \mathcal{M}$% . But the set $B$ was chosen as a maximal set in $\mathcal{M}$. We see that our assumption $B\notin CRS$ leads to the contradiction that there is a set $% B_{1}\in \mathcal{M}$ bigger than the maximal set $B$. Thus, in fact, $B\in CRS$. \end{proof} \bigskip \subsection{$C$-orbits and $C$-trips} Let $A$ be a representation set. The relation on $A$ defined by setting $% x\sim y$ if there is a finite complete representation subset of $A$ containing both $x$ and $y$, is an equivalence relation. Indeed, it is reflexive and symmetric. It is transitive by property (a) of complete representation sets. The equivalence classes we call $C$\textit{-orbits}. In the case $r=2$, $C$-orbits turn into classical orbits considered by Marshall and O'Farrell \cite{108,107}, which have a very nice geometric interpretation in terms of paths (see Section 1.3). A classical orbit consists of all possible traces of an arbitrary point in it traveling alternatively in the level sets of $h_{1}$ and $h_{2}.$ In the general setting, one partial case of $C$-orbits were introduced by Klopotowski, Nadkarni, Rao \cite{80} under the name of \textit{% related components}. The case considered in \cite{80} requires that $A\subset X=X_{1}\times \cdot \cdot \cdot \times X_{n}$ and $h_{i}$ be the canonical projections of $X$ onto $X_{i},$ $% i=1,...,r,$ respectively. Finite complete representation sets containing $x$ and $y$ will be called $C$% \textit{-trips} connecting $x$ and $y$. A $C$-trip of the smallest cardinality connecting $x$ and $y$ will be called a \textit{minimal }$C$% \textit{-trip}. \bigskip \textbf{Theorem 4.6.} \textit{Let $A$ be a representation set and $x$ and $y$ be any two points of some $C$-orbit in $A$. Then there is only one minimal $C $-trip connecting them.} \bigskip \begin{proof} Assume that $L_{1}$ and $L_{2}$ are two minimal $C$-trips connecting $% x$ and $y.$ By the definition, $L_{1}$ and $L_{2}$ are complete representation sets. Note that $L_{1}\cup L_{2}$ is also complete. Let us prove that the set $L_{1}\cap L_{2}$ is complete. Clearly, $L_{1}\cap L_{2}\in RS.$ Let $x_{0}\in L_{1}\cap L_{2}$. In particular, $x_{0}$ can be one of the points $x$ and $y$. Consider the representation \begin{equation*} 0=\sum_{i=1}^{r}g_{i}(h_{i}(x)),\text{ }x\in L_{1}\cap L_{2},\eqno(4.15) \end{equation*}% subject to $g_{i}(h_{i}(x_{0}))=0,$ $i=1,...,r-1$. On the strength of Lemma 4.1, it is enough to prove that this representation is unique. For $i=1,...,r $, let $g_{i}^{\prime }$ be any extension of $g_{i}$ from the set $% h_{i}(L_{1}\cap L_{2})$ to the set $h_{i}(L_{1})$. Construct the function \begin{equation*} f^{\prime }(x)=\sum_{i=1}^{r}g_{i}^{\prime }(h_{i}(x)),\text{ }x\in L_{1}.% \eqno(4.16) \end{equation*}% Since $f^{\prime }(x)=0$ on $L_{1}\cap L_{2}$, the following function is well defined \begin{equation*} f(x)=\left\{ \begin{array}{c} f^{\prime }(x),\text{ }x\in L_{1}, \\ 0,\text{ }x\in L_{2}.% \end{array}% \right. \end{equation*}% Since $L_{1}\cup L_{2}\in CRS$, the representation \begin{equation*} f(x)=\sum_{i=1}^{r}w_{i}(h_{i}(x)),\text{ }x\in L_{1}\cup L_{2}.\eqno(4.17) \end{equation*}% subject to \begin{equation*} w_{i}(h_{i}(x_{0}))=0,\text{ }i=1,...,r-1.\eqno(4.18) \end{equation*}% is unique. Besides, since $L_{1}\in CRS$ and $g_{i}^{\prime }(h_{i}(x_{0}))=g_{i}(h_{i}(x_{0}))=0,$ $i=1,...,r-1$, representation (4.16) is unique. This means that for each function $g_{i}$, there is only one extension $g^{\prime }$. Note that \begin{equation*} f(x)=f^{\prime }(x)=\sum_{i=1}^{r}w_{i}(h_{i}(x)),\text{ }x\in L_{1}. \end{equation*}% Now from the uniqueness of representation (4.16) we obtain that \begin{equation*} w_{i}(h_{i}(x))=g_{i}^{\prime }(h_{i}(x)),\text{ }i=1,...,r,\text{ }x\in L_{1}.\eqno(4.19) \end{equation*} A restriction of formula (4.17) to the set $L_{2}$ gives \begin{equation*} 0=\sum_{i=1}^{r}w_{i}(h_{i}(x)),\text{ }x\in L_{2}.\eqno(4.20) \end{equation*}% Since $L_{2}\in CRS$, representation (4.20) subject to conditions (4.18) is unique, whence \begin{equation*} w_{i}(h_{i}(x))=0,\text{ }i=1,...,r\text{, \ }x\in L_{2}.\eqno(4.21) \end{equation*}% From (4.19) and (4.21) it follows that \begin{equation*} g_{i}(h_{i}(x))=g_{i}^{\prime }(h_{i}(x))=0,\text{ }i=1,...,r,\text{ }x\in L_{1}\cap L_{2}. \end{equation*}% Thus, we see that representation (4.15) subject to the conditions $% g_{i}(h_{i}(x_{0}))=0,$ $i=1,...,r-1$ is unique on the intersection $% L_{1}\cap L_{2}.$ Therefore by Lemma 4.1, $L_{1}\cap L_{2}\in CRS.$ Let the cardinalities of $L_{1}$ and $L_{2}$ be equal to $n.$ Since $x,y\in L_{1}\cap L_{2}$ and $L_{1}\cap L_{2}\in CRS$, we obtain from the definition of minimal $C$-trips that the cardinality of $L_{1}\cap L_{2}$ is also $n.$ Hence, $L_{1}\cap L_{2}=L_{1}=L_{2}.$ \end{proof} Let $Q$ be a representation set. That is, each function $f:Q\rightarrow \mathbb{R}$ enjoys representation (4.7). Can we construct the functions $% g_{i},$ $i=1,...,r,$ for a given $f$? There is a procedure for constructing one certain collection of $g_{i}$, provided that $Q$ consists of a single $C$% -orbit, that is, any two points of $Q$ can be connected by a $C$-trip. To describe this procedure, take a point $x_{0}\in Q$ and fix it. We are going to find $g_{i}$ from (4.7) and conditions (4.8). Let $y$ be any point in $Q$% . To find the values of $g_{i}$ at the points $h_{i}(y),$ $i=1,...,r,$ connect $x_{0}$ and $y$ by a minimal $C$-trip $S=\{x_{1},...,x_{n}\},$ where $x_{1}=x_{0}$ and $x_{n}=y.$ Since $S$ is a complete representation set, equation (4.7) subject to (4.8) has a unique solution on $S$. That is, we can find $g_{i}(h_{i}(y)),$ $i=1,...,r,$ by solving the system of linear equations \begin{equation*} \sum_{i=1}^{r}g_{i}(h_{i}(x_{j}))=f(x_{j}),\text{ }j=1,...,n\text{.} \end{equation*} We see that each minimal $C$-trip containing $x_{0}$ generates a system of linear equations, which is uniquely solvable. Since any point in $Q$ can be connected with $x_{0}$ by such a trip, we can find $g_{i}(t)$ at each point $% t\in h_{i}(Q),$ $i=1,...,r.$ The above procedure can still be effective for some particular representation sets $Q$ consisting of many $C$-orbits. Let $\{C_{\alpha }\},$ $\alpha \in \Lambda ,$ denote the set of all $C$-orbits of $Q$. Fix some points $x_{\alpha }\in C_{\alpha },$ $\alpha \in \Lambda $, one in each orbit. Let $y_{\alpha }$ be any points of $C_{\alpha },$ $\alpha \in \Lambda ,$ respectively. We can apply the above procedure of finding the values of $% g_{i}$ at each $y_{\alpha }$ if $h_{i}(y_{\alpha })\neq h_{i}(y_{\beta })$ for all $i$ and $\alpha \neq \beta $. For $h_{i}(y_{\alpha })=h_{i}(y_{\beta }),$ one cannot guarantee that after solving the corresponding systems of linear equations (associated with $y_{\alpha }$ and $y_{\beta }$), the solutions $g_{i}(h_{i}(y_{\alpha })$ and $g_{i}(h_{i}(y_{\beta }))$ will be equal. That is, for the case $h_{i}(y_{\alpha })=h_{i}(y_{\beta })$, the constructed functions $g_{i}$ may not be well defined. \bigskip \textbf{Remark 4.5.} All the results in this section are valid, in particular, for linear combinations of generalized ridge functions. \newpage \chapter{Applications to neural networks} Neural networks have increasingly been used in many areas of applied sciences. Most of the applications employ neural networks to approximate complex nonlinear functional dependencies on a high dimensional data set. The theoretical justification for such applications is that any continuous function can be approximated within an arbitrary precision by carefully selecting parameters in the network. The most commonly used model of neural networks is the \textit{multilayer feedforward perceptron} (MLP) model. This model consists of a finite number of successive layers. The first and the last layers are called the input and the output layers, respectively. The intermediate layers are called hidden layers. MLP models are usually classified not by their number of layers, but by their number of hidden layers. In this chapter, we study approximation properties of the single and two hidden layer feedforward perceptron models. Our analysis is based on ridge functions and the Kolmogorov superposition theorem. The material of this chapter may be found in \cite{39,46,43,65}. \bigskip \section{Single hidden layer neural networks} In this section, we consider single hidden layer neural networks with a set of weights consisting of a finite number of directions or straight lines. For certain activation functions, we characterize compact sets $X$ in the $d$-dimensional space such that the corresponding neural network can approximate any continuous function on $X$. \subsection{Problem statement} Approximation capabilities of neural networks have been investigated in a great deal of works over the last 30 years (see, e.g., \cite% {Alm,An,B,15,17,Ch,19,20,21,24,GI,39,40,41,67,68,71,70,89,104,110,119,123,135}). In this section, we are interested in questions of density of a single hidden layer perceptron model. A typical density result shows that this model can approximate an arbitrary function in a given class with any degree of accuracy. \textit{A single hidden layer perceptron model} with $r$ units in the hidden layer and input $\mathbf{x}=(x_{1},...,x_{d})$ evaluates a function of the form \begin{equation*} \sum_{i=1}^{r}c_{i}\sigma (\mathbf{w}^{i}\mathbf{\cdot x}-\theta _{i}),\eqno% (5.1) \end{equation*}% where the \textit{weights} $\mathbf{w}^{i}$ are vectors in $\mathbb{R}^{d}$, the \textit{thresholds} $\theta _{i}$ and the coefficients $c_{i}$ are real numbers and \ the \textit{activation function} $\sigma $ is a univariate function, which is considered to be continuous here. Note that in Eq (5.1) each function $% \sigma (\mathbf{w}^{i}\mathbf{\cdot x}-\theta _{i})$ is a ridge function with the direction $\mathbf{w}^{i}$. For various activation functions $\sigma $, it has been proved in a number of papers that one can approximate arbitrarily well a given continuous function by functions of the form (5.1) ($r$ is not fixed!) over any compact subset of $\mathbb{R}^{d}$. In other words, the set \begin{equation*} \mathcal{M}(\sigma )=span\text{\ }\{\sigma (\mathbf{w\cdot x}-\theta ):\ \theta \in \mathbb{R}\text{, }\mathbf{w\in }\mathbb{R}^{d}\} \end{equation*}% is dense in the space $C(\mathbb{R}^{d})$ in the topology of uniform convergence on compact sets (see, e.g., \cite{17,21,41,67,68}). The most general result of this type belongs to Leshno, Lin, Pinkus and Schocken \cite% {89}. They proved that a necessary and sufficient condition for a continuous activation function to have the density property is that it not be a polynomial. This result shows the efficacy of the single hidden layer perceptron model within all possible choices of the activation function $% \sigma $, provided that $\sigma $ is continuous. In fact, density of the set $\mathcal{M}(\sigma )$ also holds for some reasonable sets of weights and thresholds. (see\cite{119}). Some authors showed that a single hidden layer perceptron with a suitably restricted set of weights can also have the density property (or, in neural network terminology, the \textit{universal approximation property}). For example, White and Stinchcombe \cite{135} proved that a single layer network with a polygonal, polynomial spline or analytic activation function and a bounded set of weights has the density property. Ito \cite{68} investigated this property of networks using monotone sigmoidal functions (tending to $0$ at minus infinity and $1$ at infinity), with only weights located on the unit sphere. We see that weights required for the density property are not necessary to be of an arbitrarily large magnitude. But what if they are too restricted. How can one learn approximation properties of networks with an arbitrarily restricted set of weights? This problem is too difficult to be solved completely in this general formulation. But there are some cases that deserve a special attention. The most interesting case is, of course, neural networks with weights varying on a finite set of directions or lines. To the best of our knowledge, approximation capabilities of such networks have not been studied yet. More precisely, let $W$ be a set of weights consisting of a finite number of vectors (or straight lines) in $\mathbb{R}^{d}$. It is clear that if $w$ varies only in $W$, the set $\mathcal{M}(\sigma )$ can not be dense in $C(\mathbb{R}^{d})$ in the topology of uniform convergence on compacta (compact sets). In this case, one may want to determine boundaries of efficacy of the model. Over which compact sets $X\subset \mathbb{R}^{d}$ does the model preserve its general propensity to approximate arbitrarily well every continuous multivariate function? In Section 5.1.2, we will consider this problem and give both sufficient and necessary conditions for well approximation (approximation with arbitrary precision) by networks with weights from a finite set of directions or lines. For a set $W$ of weights consisting of two vectors, we show that there is a geometrically explicit solution to the problem. In Section 5.1.3, we discuss some aspects of the exact representation by neural networks with weights varying on finitely many straight lines. \bigskip \subsection{Density results} In this subsection we give a sufficient and also a necessary conditions for approximation by neural networks with finitely many weights and with weights varying on a finite set of straight lines (through the origin). Let $X$ be a compact subset of $\mathbb{R}^{d}$. Consider the following set functions \begin{equation*} \tau _{i}(Z)=\{\mathbf{x}\in Z:~|p_{i}^{-1}(p_{i}(\mathbf{x}))\bigcap Z|\geq 2\},\quad Z\subset X,~i=1,\ldots ,k, \end{equation*}% where $p_{i}(\mathbf{x})=\mathbf{a}^{i}\cdot \mathbf{x}$, $|Y|$ denotes the cardinality of a considered set $Y$. Define $\tau (Z)$ to be $% \bigcap_{i=1}^{k}\tau _{i}(Z)$ and define $\tau ^{2}(Z)=\tau (\tau (Z))$, $% \tau ^{3}(Z)=\tau (\tau ^{2}(Z))$ and so on inductively. These functions first appeared in the work \cite{132} by Sternfeld, where he investigated problems of representation by linear superpositions. Clearly, $\tau (Z)\supseteq \tau ^{2}(Z)\supseteq \tau ^{3}(Z)\supseteq ...$It is possible that for some $n$, $\tau ^{n}(Z)=\emptyset .$ In this case, one can see that $Z$ does not contain a cycle. In general, if some set $Z\subset X$ forms a cycle, then $\tau ^{n}(Z)=Z.$ But the reverse is not true. Indeed, let $% Z=X=\{(0,0,\frac{1}{2}),(0,0,1),(0,1,0),(1,0,1),(1,1,0),(\frac{1}{2},\frac{1% }{2},0),(\frac{1}{2},\frac{1}{2},\frac{1}{2})\}$, $\mathbf{a}^{i},i=1,2,3,$ are the coordinate directions in $\mathbb{R}^{3}$. It is not difficult to verify that $X$ does not possess cycles with respect to these directions and at the same time $\tau (X)=X$ (and so $\tau ^{n}(X)=X$ for every $n)$. Consider the linear combinations of ridge functions with fixed directions $% \mathbf{a}^{1},...,\mathbf{a}^{k}$ \begin{equation*} \mathcal{R}\left( \mathbf{a}^{1},...,\mathbf{a}^{k}\right) =\left\{ \sum\limits_{i=1}^{k}g_{i}\left( \mathbf{a}^{i}\cdot \mathbf{x}\right) :g_{i}\in C(\mathbb{R)},~i=1,...,k\right\} .\eqno(5.2) \end{equation*} Let $K$ be a family of functions defined on $\mathbb{R}^{d}$ and $X$ be a subset of $\mathbb{R}^{d}.$ By $K_{X}$ we will denote the restriction of this family to $X.$ Thus $\mathcal{R}_{X}\left( \mathbf{a}^{1},...,\mathbf{a}% ^{k}\right) $ stands for the set of sums of ridge functions in (5.2) defined on $X$. The following theorem is a particular case of the known general result of Sproston and Strauss \cite{129} established for the sum of subalgebras of $% C(X)$. \bigskip \textbf{Theorem 5.1. }\textit{Let $X$ be a compact subset of $\mathbb{R}^{d}$% . If $\cap _{n=1,2,...}\tau ^{n}(X)=\emptyset $, then the set $\mathcal{R}% _{X}\left( \mathbf{a}^{1},...,\mathbf{a}^{k}\right) $ is dense in $C(X)$.} \bigskip In our analysis, we need the following lemma. \bigskip \textbf{Lemma 5.1. } \textit{If $\mathcal{R}_{X}\left( \mathbf{a}^{1},...,% \mathbf{a}^{k}\right) $ is dense in $C(X),$ then the set $X$ does not contain a cycle with respect to the directions $\mathbf{a}^{1},...,\mathbf{a}% ^{k}$.} \bigskip \begin{proof} Suppose the contrary. Suppose that the set $X$ contains cycles. Each cycle $% l=(x_{1},\ldots ,x_{n})$ and the associated vector $\lambda =(\lambda _{1},\ldots ,\lambda _{n})$ generate the functional \begin{equation*} G_{l,\lambda }(f)=\sum_{j=1}^{n}\lambda _{j}f(x_{j}),\quad f\in C(X). \end{equation*} Clearly, $G_{l,\lambda }$ is linear and continuous with the norm $% \sum_{j=1}^{n}|\lambda _{j}|.$It is not difficult to verify that $% G_{l,\lambda }(g)=0$ for all functions $g\in \mathcal{R}\left( \mathbf{a}% ^{1},...,\mathbf{a}^{k}\right) .$ Let $f_{0}$ be a continuous function such that $f_{0}(x_{j})=1$ if $\lambda _{j}>0$ and $f_{0}(x_{j})=-1$ if $\lambda _{j}<0$, $j=1,\ldots ,n$. For this function, $G_{l,\lambda }(f_{0})\neq 0$. Thus, we have constructed a nonzero linear functional which belongs to the annihilator of the manifold $\mathcal{R}_{X}\left( \mathbf{a}^{1},...,% \mathbf{a}^{k}\right) $. This means that $\mathcal{R}_{X}\left( \mathbf{a}% ^{1},...,\mathbf{a}^{k}\right) $ is not dense in $C(X)$. The obtained contradiction proves the lemma. \end{proof} Now we are ready to step forward from ridge function approximation to neural networks. Let $\sigma \in C(\mathbb{R)}$ be a continuous activation function. For a subset $W\subset \mathbb{R}^{d},$ let $\mathcal{M}(\sigma ;W,% \mathbb{R})$ stand for the set of neural networks with weights from $W.$ That is, \begin{equation*} \mathcal{M}(\sigma ;W,\mathbb{R})=span\{\sigma (\mathbf{w}\cdot \mathbf{x}% -\theta ):~\mathbf{w}\in W,~\theta \in \mathbb{R}\}. \end{equation*} \bigskip \textbf{Theorem 5.2.} \textit{Let $\sigma \in C(\mathbb{R})\cap L_{p}(% \mathbb{R)}$, where $1\leq p<\infty $, or $\sigma $ be a continuous, bounded, nonconstant function, which has a limit at infinity (or minus infinity). Let $W=\{\mathbf{a}^{1},...,\mathbf{a}^{k}\} \subset \mathbb{R}^{d}$ be the given set of weights and $X$ be a compact subset of $\mathbb{R}^{d}$. The following assertions are valid:} \textit{(1) if $\cap _{n=1,2,...}\tau ^{n}(X)=\emptyset $, then the set $% \mathcal{M}_{X}(\sigma ;W,\mathbb{R})$ is dense in the space of all continuous functions on $X$.} \textit{(2) if $\mathcal{M}_{X}(\sigma ;W,\mathbb{R})$ is dense in $C(X)$, then the set $X$ does not contain cycles.} \bigskip \begin{proof} Part (1). Let $X$ be a compact subset of $\mathbb{R}^{d}$ for which \textit{% \ }$\cap _{n=1,2,...}\tau ^{n}(X)=\emptyset $. By Theorem 5.1, the set $% \mathcal{R}_{X}\left( \mathbf{a}^{1},...,\mathbf{a}^{k}\right) $ is dense in $C(X)$. This means that for any positive real number $\varepsilon $ there exist continuous univariate functions $g_{i},$ $i=1,...,k$ such that \begin{equation*} \left\vert f(\mathbf{x})-\sum_{i=1}^{k}{g_{i}\left( \mathbf{a}^{i}{\cdot }% \mathbf{x}\right) }\right\vert <\frac{\varepsilon }{k+1}\eqno(5.3) \end{equation*}% for all $\mathbf{x}\in X$. Since $X$ is compact, the sets $Y_{i}=\{\mathbf{a}% ^{i}{\cdot }\mathbf{x:\ x}\in X\},\ i=1,2,...,k$ are also compacts. In 1947, Schwartz \cite{124} proved that continuous and $p$-th degree Lebesgue integrable univariate functions or continuous, bounded, nonconstant functions having a limit at infinity (or minus infinity) are not mean-periodic. Note that a function $f\in C(\mathbb{R}^{d})$ is called mean periodic if the set $span$\ $\{f(\mathbf{x}-\mathbf{b}):\ \mathbf{b}\in \mathbb{R}^{d}\}$ is not dense in $C(\mathbb{R}^{d})$ in the topology of uniform convergence on compacta (see \cite{124}). Thus, Schwartz proved that the set \begin{equation*} span\text{\ }\{\sigma (y-\theta ):\ \theta \in \mathbb{R}\} \end{equation*}% is dense in $C(\mathbb{R)}$ in the topology of uniform convergence. We learned about this result from Pinkus \cite[p.162]{119}. This density result means that for the given $\varepsilon $ there exist numbers $c_{ij},\theta _{ij}\in \mathbb{R}$, $i=1,2,...,k$, $j=1,...,m_{i}$ such that% \begin{equation*} \left\vert g_{i}(y)-\sum_{j=1}^{m_{i}}c_{ij}\sigma (y-\theta _{ij})\right\vert \,<\frac{\varepsilon }{k+1}\eqno(5.4) \end{equation*}% for all $y\in Y_{i},\ i=1,2,...,k.$ From (5.3) and (5.4) we obtain that \begin{equation*} \left\Vert f(\mathbf{x})-\sum_{i=1}^{k}\sum_{j=1}^{m_{i}}c_{ij}\sigma (% \mathbf{a}^{i}{\cdot }\mathbf{x}-\theta _{ij})\right\Vert _{C(X)}<\varepsilon .\eqno(5.5) \end{equation*}% Hence $\overline{\mathcal{M}_{X}(\sigma ;W,\mathbb{R})}=C(X).$ \bigskip Part (2). Let $X$ be a compact subset of $\mathbb{R}^{d}$ and the set $% \mathcal{M}_{X}(\sigma ;W,\mathbb{R})$ be dense in $C(X).$ Then for an arbitrary positive real number $\varepsilon $, inequality (5.5) holds with some coefficients $c_{ij},\theta _{ij},\ i=1,2,\ j=1,...,m_{i}.$ Since for each $i=1,2,...,k$, the function $\sum_{j=1}^{m_{i}}c_{ij}\sigma (\mathbf{a}% ^{i}{\cdot }\mathbf{x}-\theta _{ij})$ is a function of the form $g_{i}(% \mathbf{a}^{i}{\cdot }\mathbf{x}),$ the subspace $\mathcal{R}_{X}\left( \mathbf{a}^{1},...,\mathbf{a}^{k}\right) $ is dense in $C(X)$. Then by Lemma 5.1, the set $X$ contains no cycles. \end{proof} The above theorem still holds if the set of weights $W=\{\mathbf{a}^{1},...,\mathbf{a}^{k}\}$ is replaced by the set $W_{1}=\{t_{1}\mathbf{a}^{1},...,t_{k}\mathbf{a}^{k}:\ t_{1},...,t_{k}\in \mathbb{R}\}$. In fact, for $W_{1}$, the above restrictions on the activation function $\sigma $ may be weakened. \bigskip \textbf{Theorem 5.3.} \textit{Assume $\sigma \in C(\mathbb{R})$ is not a polynomial. Let $W_{1}=\{t_{1}\mathbf{a}^{1},...,t_{k}\mathbf{a}^{k}:\ t_{1},...,t_{k}\in \mathbb{R}\}$ be the given set of weights and $X$ be a compact subset of $\mathbb{R}^{d}$. The following assertions are valid:} \textit{(1) if $\cap _{n=1,2,...}\tau ^{n}(X)=\emptyset $, then the set $% \mathcal{M}_{X}(\sigma ;W_{1},\mathbb{R})$ is dense in the space of all continuous functions on $X$.} \textit{(2) if $\mathcal{M}_{X}(\sigma ;W_{1},\mathbb{R})$ is dense in $C(X)$% , then the set $X$ does not contain cycles.} \bigskip The proof of this theorem is similar to that of Theorem 5.2 and based on the following result of Leshno, Lin, Pinkus and Schocken \cite{89}: if $\sigma $ is not a polynomial, then the set \begin{equation*} span\text{\ }\{\sigma (ty-\theta ):\ t,\theta \in \mathbb{R}\} \end{equation*}% is dense in $C(\mathbb{R)}$ in the topology of uniform convergence on compacta. The above example with the set \begin{equation*} \{(0,0,\frac{1}{2}),(0,0,1),(0,1,0),(1,0,1),(1,1,0),(\frac{1}{2},\frac{1}{2}% ,0),(\frac{1}{2},\frac{1}{2},\frac{1}{2})\} \end{equation*}% shows that the sufficient condition in part (1) of Theorem 5.2 is not necessary. The necessary condition in part (2), in general, is not sufficient. But it is not easily seen. Here, is the nontrivial example showing that nonexistence of cycles is not sufficient for the density $% \overline{\mathcal{M}_{X}(\sigma ;W,\mathbb{R})}=C(X).$ For the sake of simplicity, we restrict ourselves to $\mathbb{R}^{2}.$ Let $\mathbf{a}% ^{1}=(1;1),$ $\mathbf{a}^{2}=(1;-1)$ and the set of weights $W=\{\mathbf{a}% ^{1},\mathbf{a}^{2}\}.$ The set $X$ can be constructed as follows. Let $% X_{1} $ be the union of the four line segments $[(-3;0),(-1;0)],$ $% [(-1;2),(1;2)],$ $[(1;0),(3;0)]$ and $[(-1;-2),(1;-2)].$ Rotate one segment in $X_{1}$ $90^{\circ }$ about its center and remove the middle one-third from each line segment. The obtained set denote by $X_{2}$. By the same way, one can construct $X_{3},X_{4},$ and so on. It is clear that the set $X_{i}$ has $2^{i+1}$ line segments. Let $X$ be a limit of the sets $X_{i}$, $% i=1,2,...$. Note that there are no cycles. By $S_{i}$, $i=\overline{1,4},$ denote the closed discs with the unit radius and centered at the points $(-2;0),$ $(0;2),$ $(2;0)$ and $(0;-2)$ respectively. Consider a continuous function $f_{0}$ such that $f_{0}(% \mathbf{x})=1$ for $\mathbf{x}\in (S_{1}\cup S_{3})\cap X$, $f_{0}(\mathbf{x}% )=-1$ for $\mathbf{x}\in (S_{2}\cup S_{4})\cap X$, and $-1<f_{0}(\mathbf{x}% )<1$ elsewhere on $\mathbb{R}^{2}$. Let $p=(\mathbf{y}^{1},\mathbf{y}% ^{2},...)$ be any infinite path in $X.$ Note that the points $\mathbf{y}% ^{i}, $ $i=1,2,...,$ are alternatively in the sets $(S_{1}\cup S_{3})\cap X$ and $(S_{2}\cup S_{4})\cap X$. Obviously, \begin{equation*} E(f_{0},X)\overset{def}{=}\inf_{g\in \mathcal{R}_{X}\left( \mathbf{a}^{1},% \mathbf{a}^{2}\right) }\left\Vert f_{0}-g\right\Vert _{C(X)}\leq \left\Vert f_{0}\right\Vert _{C(X)}=1.\eqno(5.6) \end{equation*} For each positive integer $k=1,2,...$, set $p_{k}=(\mathbf{y}^{1},...,% \mathbf{y}^{k})$ and consider the path functionals \begin{equation*} G_{p_{k}}(f)=\frac{1}{k}\sum_{i=1}^{k}(-1)^{i-1}f(\mathbf{y}^{i}). \end{equation*} $G_{p_{k}}$ is a continuous linear functional obeying the following obvious properties: \begin{enumerate} \item[(1)] $\left\Vert G_{p_{k}}\right\Vert =G_{p_{k}}(f_{0})=1;$ \item[(2)] $G_{p_{k}}(g_{1}+g_{2})\leq \frac{2}{k}(\left\Vert g_{1}\right\Vert +\left\Vert g_{2}\right\Vert )$ for ridge functions $g_{1}={% g_{1}\left( \mathbf{a}^{1}{\cdot }\mathbf{x}\right) }$ and $g_{2}={% g_{2}\left( \mathbf{a}^{2}{\cdot }\mathbf{x}\right) .}$ \end{enumerate} By property (1), the sequence $\{G_{p_{k}}\}_{k=1}^{\infty }$ has a weak$^{% \text{*}}$ cluster point. This point will be denoted by $G.$ By property (2), $G\in \mathcal{R} _{X}\left( \mathbf{a}^{1},\mathbf{a}^{2}\right) ^{\bot}.$ Therefore, \begin{equation*} 1=G(f_{0})=G(f_{0}-g)\leq \left\Vert f_{0}-g\right\Vert _{C(X)}\text{ \ for any }g\in \mathcal{R}_{X}\left( \mathbf{a}^{1},\mathbf{a}^{2}\right) . \end{equation*} Taking $\inf$ over $g$ in the right-hand side of the last inequality, we obtain that $1\leq E(f_{0},X).$ Now it follows from (5.6) that $% E(f_{0},X)=1. $ Recall that $\mathcal{M}_{X}(\sigma ;W,\mathbb{R})\subset \mathcal{R}_{X}\left( \mathbf{a}^{1},\mathbf{a}^{2}\right) .$ Thus% \begin{equation*} \inf_{h\in \mathcal{M}_{X}(\sigma ;W,\mathbb{R})}\left\Vert f-h\right\Vert _{C(X)}\geq 1. \end{equation*} The last inequality finally shows that $\overline{\mathcal{M}_{X}(\sigma ;W,% \mathbb{R})}\neq C(X).$ \bigskip For neural networks with weights consisting of only two vectors (or directions) the problem of density becomes more clear. In this case, under some minor restrictions on $X,$ the necessary condition in part (2) of Theorem 5.2 (nonexistence of cycles) is also sufficient for the density of $% \mathcal{M}_{X}(\sigma ;W,\mathbb{R})$ in $C(X)$. These restrictions are imposed on the following equivalent classes of $X$ induced by paths. The relation $\mathbf{x}\thicksim \mathbf{y}$ when $\mathbf{x}$ and $\mathbf{y}$ belong to some path in a given compact set $X\subset \mathbb{R}^{d}$ defines an equivalence relation. Recall that the equivalence classes are called orbits (see Section 1.3.4). \bigskip \textbf{Theorem 5.4.} \textit{Let $\sigma \in C(\mathbb{R})\cap L_{p}(% \mathbb{R)}$, where $1\leq p<\infty $, or $\sigma $ be a continuous, bounded, nonconstant function, which has a limit at infinity (or minus infinity). Let $W=\{\mathbf{a}^{1},\mathbf{a}^{2}\} \subset \mathbb{R}^{d}$ be the given set of weights and $X$ be a compact subset of $\mathbb{R}% ^{d}$ with all its orbits closed. Then $\mathcal{M}_{X}(\sigma ;W,\mathbb{R}% ) $ is dense in the space of all continuous functions on $X$ if and only if $X$ contains no closed paths with respect to the directions $\mathbf{a}^{1}$ and $\mathbf{a}^{2}$.} \bigskip \begin{proof} \textit{Sufficiency.} Let $X$ be a compact subset of $\mathbb{R}% ^{d}$ with all its orbits closed. Besides, let $X$ contain no closed paths. By Theorem 1.6 (see Section 1.3.4), the set $\mathcal{R}_{X}\left( \mathbf{a}% ^{1},\mathbf{a}^{2}\right) $ is dense in $C(X)$. This means that for any positive real number $\varepsilon $ there exist continuous univariate functions $g_{1}$ and $g_{2}$ such that \begin{equation*} \left\vert f(\mathbf{x})-{g_{1}\left( \mathbf{a}^{1}{\cdot }\mathbf{x}% \right) -g_{2}\left( \mathbf{a}^{2}{\cdot }\mathbf{x}\right) }\right\vert <% \frac{\varepsilon }{3}\eqno(5.7) \end{equation*}% for all $\mathbf{x}\in X$. Since $X$ is compact, the sets $Y_{i}=\{\mathbf{a}% ^{i}{\cdot }\mathbf{x:\ x}\in X\},\ i=1,2,$ are also compacts. As mentioned above, Schwartz \cite{124} proved that continuous and $p$-th degree Lebesgue integrable univariate functions or continuous, bounded, nonconstant functions having a limit at infinity (or minus infinity) are not mean-periodic. Thus, the set \begin{equation*} span\text{\ }\{\sigma (y-\theta ):\ \theta \in \mathbb{R}\} \end{equation*}% is dense in $C(\mathbb{R)}$ in the topology of uniform convergence. This density result means that for the given $\varepsilon $ there exist numbers $% c_{ij},\theta _{ij}\in \mathbb{R}$, $i=1,2,$ $j=1,\dots ,m_{i}$ such that% \begin{equation*} \left\vert g_{i}(y)-\sum_{j=1}^{m_{i}}c_{ij}\sigma (y-\theta _{ij})\right\vert \,<\frac{\varepsilon }{3}\eqno(5.8) \end{equation*}% for all $y\in Y_{i},\ i=1,2.$ From (5.7) and (5.8) we obtain that \begin{equation*} \left\Vert f(\mathbf{x})-\sum_{i=1}^{2}\sum_{j=1}^{m_{i}}c_{ij}\sigma (% \mathbf{a}^{i}{\cdot }\mathbf{x}-\theta _{ij})\right\Vert _{C(X)}<\varepsilon .\eqno(5.9) \end{equation*}% Hence $\overline{\mathcal{M}_{X}(\sigma ;W,\mathbb{R})}=C(X).$ \bigskip \textit{Necessity.} Let $X$ be a compact subset of $\mathbb{R}^{n}$ with all its orbits closed and the set $\mathcal{M}_{X}(\sigma ;W,\mathbb{R})$ be dense in $C(X).$ Then for an arbitrary positive real number $\varepsilon $, inequality (5.9) holds with some coefficients $c_{ij},\theta _{ij},\ i=1,2,\ j=1,\dots ,m_{i}.$ Since for $i=1,2,$ $\sum_{j=1}^{m_{i}}c_{ij}\sigma (% \mathbf{a}^{i}{\cdot }\mathbf{x}-\theta _{ij})$ is a function of the form $% g_{i}(\mathbf{a}^{i}{\cdot }\mathbf{x}),$ the subspace $\mathcal{R}% _{X}\left( \mathbf{a}^{1},\mathbf{a}^{2}\right) $ is dense in $C(X)$. Then by Theorem 1.6, the set $X$ contains no closed paths. \end{proof} \bigskip \textbf{Remark 5.1.} It can be shown that the necessity of the theorem is valid without any restriction on orbits of $X$. Indeed if $X$ contains a closed path, then it contains a closed path $p=(\mathbf{x}^{1},\dots ,% \mathbf{x}^{2m})$ with different points. The functional $G_{p}=% \sum_{i=1}^{2m}(-1)^{i-1}f(\mathbf{x}^{i})$ belongs to the annihilator of the subspace $\mathcal{R}_{X}\left( \mathbf{a}^{1},\mathbf{a}^{2}\right) .$ There exist nontrivial continuous functions $f_{0}$ on $X$ such that $% G_{p}(f_{0})\neq 0$ (take, for example, any continuous function $f_{0}$ taking values $+1$ at $\{\mathbf{x}^{1},\mathbf{x}^{3},\dots ,\mathbf{x}% ^{2m-1}\}$, $-1$ at $\{\mathbf{x}^{2},\mathbf{x}^{4},\dots ,\mathbf{x}% ^{2m}\} $ and $-1<f_{0}(\mathbf{x})<1$ elsewhere). This shows that the subspace $\mathcal{R}_{X}\left( \mathbf{a}^{1},\mathbf{a}^{2}\right) $ is not dense in $C(X)$. But in this case, the set $\mathcal{M}_{X}(\sigma ;W,% \mathbb{R})$ cannot be dense in $C(X)$. The obtained contradiction means that our assumption is not true and $X$ contains no closed paths. \bigskip Theorem 5.4 remains valid if the set of weights $\ W=\{\mathbf{a}^{1},% \mathbf{a}^{2}\}$ is replaced by the set $W_{1}=\{t_{1}\mathbf{a}^{1},t_{2}% \mathbf{a}^{2}:\ t_{1},t_{2}\in \mathbb{R\}}$. In fact, for the set $W_{1}$, the required conditions on $\sigma $ may be weakened. As in Theorem 5.3, the activation function $\sigma $ can be taken only non-polynomial. \bigskip \textbf{Theorem 5.5.} \textit{Assume $\sigma \in C(\mathbb{R})$ is not a polynomial. Let $\mathbf{a}^{1}$ and $\mathbf{a}^{2}$ be fixed vectors and $W_{1}=\{t_{1}\mathbf{a}^{1},t_{2}% \mathbf{a}^{2}:\ t_{1},t_{2}\in \mathbb{R\}}$ be the set of weights. Let $X$ be a compact subset of $\mathbb{R% }^{d}$ with all its orbits closed. Then $\mathcal{M}_{X}(\sigma ;W_{1},% \mathbb{R})$ is dense in the space of all continuous functions on $X$ if and only if $X$ contains no closed paths with respect to the directions $\mathbf{a}^{1}$ and $\mathbf{a}^{2}$.} \bigskip The proof is analogous to that of Theorem 5.4 and based on the above mentioned result of Leshno, Lin, Pinkus and Schocken \cite{89}. \bigskip \textbf{Examples:} \begin{description} \item[(a)] Let $\mathbf{a}^{1}$ and $\mathbf{a}^{2}$ be two noncollinear vectors in $\mathbb{R}^{2}.$ Let $B=B_{1}...B_{k}$ be a broken line with the sides $B_{i}B_{i+1},\ i=1,...,k-1,$ alternatively perpendicular to $\mathbf{a% }^{1}$ and $\mathbf{a}^{2}$. Besides, let $B$ does not contain vertices of any parallelogram with sides perpendicular to these vectors. Then the set $% \mathcal{M}_{B}(\sigma ;\{\mathbf{a}^{1},\mathbf{a}^{2}\},\mathbb{R})$ is dense in $C(B).$ \item[(b)] Let $\mathbf{a}^{1}$ and $\mathbf{a}^{2}$ be two noncollinear vectors in $\mathbb{R}^{2}.$ If $X$ is the union of two parallel line segments, not perpendicular to any of the vectors $\mathbf{a}^{1}$ and $% \mathbf{a}^{2}$, then the set $\mathcal{M}_{X}(\sigma ;\{\mathbf{a}^{1}, \mathbf{a}^{2}\},\mathbb{R})$ is dense in $C(X).$ \item[(c)] Let now $\mathbf{a}^{1}$ and $\mathbf{a}^{2}$ be two collinear vectors in $\mathbb{R}^{2}.$ Note that in this case any path consisting of two points is automatically closed. Thus the set $\mathcal{M}_{X}(\sigma ;\{% \mathbf{a}^{1},\mathbf{a}^{2}\},\mathbb{R})$ is dense in $C(X)$ if and only if $X$ contains no path different from a singleton. A simple example is a line segment not perpendicular to the given direction. \item[(d)] Let $X$ be a compact set with an interior point. Then Theorem 5.4 fails, since any such set contains vertices of some parallelogram with sides perpendicular to the given directions $\mathbf{a}^{1}$ and $\mathbf{a}^{2}$, that is a closed path. \end{description} \bigskip \subsection{A necessary condition for the representation by neural networks} In this subsection we give a necessary condition for the representation of functions by neural networks with weights from a finitely many straight lines. Before formulating our result, we introduce new objects, namely \textit{semicycles} with respect to directions $\mathbf{a}^{1},...,% \mathbf{a}^{k}\in \mathbb{R}^{d}\backslash \{\mathbf{0}\}$. \bigskip \textbf{Definition 5.1.} \textit{A set of points $l=(\mathbf{x}^{1},\ldots ,% \mathbf{x}^{n})\subset \mathbb{R}^{d}$ is called a semicycle with respect to the directions $\mathbf{a}^{1},...,\mathbf{a}^{k}$ if there exists a vector $\lambda =(\lambda _{1},\ldots ,\lambda _{n})\in \mathbf{Z}% ^{n}\setminus \{\mathbf{0}\}$ such that for any $i=1,\ldots ,k,$ we have} \begin{equation*} \sum_{j=1}^{n}\lambda _{j}\delta _{\mathbf{a}^{i}\cdot \mathbf{x}% ^{j}}=\sum_{s=1}^{r_{i}}\lambda _{i_{s}}\delta _{\mathbf{a}^{i}\cdot \mathbf{% x}^{i_{s}}},\quad where~r_{i}\leq k.\eqno(5.10) \end{equation*}% Here $\delta _{a}$ is the characteristic function of the single point set $% \{a\}$. Note that for $i=1,\ldots ,k$, the set $\{\lambda _{i_{s}},~s=1,...,r_{i}\}$ is a subset of the set $\{\lambda _{j},~j=1,...,n\}$. Thus, Eq. (5.10) means that for each $i$, we actually have at most $k$ terms in the sum $\sum_{j=1}^{n}\lambda _{j}\delta _{% \mathbf{a}^{i}\cdot \mathbf{x}^{j}}$. Recall that if in (5.10) for any $i=1,\ldots ,k$, we have \begin{equation*} \sum_{j=1}^{n}\lambda _{j}\delta _{\mathbf{a}^{i}\cdot \mathbf{x}^{j}}=0, \end{equation*}% then the set $l=(\mathbf{x}^{1},\ldots ,\mathbf{x}^{n})$ is a cycle with respect to the directions $\mathbf{a}^{1},...,\mathbf{a}^{k}$ (see Section 1.2). Thus a cycle is a special case of a semicycle. Let us give a simple example of a semicycle. Assume $k=2$ and $\mathbf{a}^{1}\cdot \mathbf{x}^{1}=\mathbf{a}% ^{1}\cdot \mathbf{x}^{2}$, $\mathbf{a}^{2}\cdot \mathbf{x}^{2}=\mathbf{a}% ^{2}\cdot \mathbf{x}^{3}$, $\mathbf{a}^{1}\cdot \mathbf{x}^{3}=\mathbf{a}% ^{1}\cdot \mathbf{x}^{4}$,..., $\mathbf{a}^{2}\cdot \mathbf{x}^{n-1}=\mathbf{% a}^{2}\cdot \mathbf{x}^{n}$. Then it is not difficult to see that for a vector $\lambda =(\lambda _{1},\ldots ,\lambda _{n})$ with the components $\lambda _{j}=(-1)^{j},$ the following equalities hold: \begin{eqnarray*} \sum_{j=1}^{n}\lambda _{j}\delta _{\mathbf{a}^{1}\cdot \mathbf{x}^{j}} &=&\lambda _{n}\delta _{\mathbf{a}^{1}\cdot \mathbf{x}^{n}}, \\ \sum_{j=1}^{n}\lambda _{j}\delta _{\mathbf{a}^{2}\cdot \mathbf{x}^{j}} &=&\lambda _{1}\delta _{\mathbf{a}^{2}\cdot \mathbf{x}^{1}}. \end{eqnarray*}% Thus, by Definition 5.1, the set $l=\{\mathbf{x}^{1},\ldots ,\mathbf{x}% ^{n}\}$ is a semicycle with respect to the directions $\mathbf{a}^{1}$ and $\mathbf{a}^{2}$. Note that this set, in the given order of its points, forms a path with respect to the directions $\mathbf{a}^{1}$ and $\mathbf{a}^{2}$ (see Section 1.3). It is not difficult to see that any path with respect to $\mathbf{a}^{1}$ and $\mathbf{a}^{2}$ is a semicycle with respect to these directions. But semicycles may also involve some union of paths. Note that one can construct many semicycles by adding not more than $k$ arbitrary points to a cycle with respect to the directions $\mathbf{a}^{1},% \mathbf{a}^{2},...,\mathbf{a}^{k}$. A cycle (or semicycle) $l$ is called a \textit{$q$-cycle} (\textit{$q$% -semicycle}) if the vector $\lambda $ associated with $l$ can be chosen so that $\left\vert \lambda _{i}\right\vert \leq q,$ $i=1,...,n,$ and $q$ is the minimal number with this property. The semicycle considered above is a $1$-semicycle. If in that example, $% \mathbf{a}^{2}\cdot \mathbf{x}^{n-1}=\mathbf{a}^{2}\cdot \mathbf{x}^{1}$, then the set $\{x_{1},x_{2},...,x_{n-1}\}$ is a $1$-cycle. Let us give a simple example of a $2$-cycle with respect to the directions $\mathbf{a}% ^{1}=(1,0)$ and $\mathbf{a}^{2}=(0,1)$. Consider the union \begin{equation*} \{0,1\}^{2}\cup \{0,2\}^{2}=\{(0,0),(1,1),(2,2),(0,1),(1,0),(0,2),(2,0)\}. \end{equation*} It is easy to see that this set is a $2$-cycle with the associated vector $% (2,1,1,-1,-1,-1,-1).$ Similarly, one can construct a $q$-cycle or $q$% -semicycle for any positive integer $q$. \bigskip \textbf{Theorem 5.6.} \textit{Assume $W=\{t_{1}\mathbf{a}^{1},...,t_{k}\mathbf{a}^{k}:\ t_{1},...,t_{k}\in \mathbb{R}\}$ is the given set of weights. If $\mathcal{M}_{X}(\sigma ;W,\mathbb{R})=C(X)$, then $X$ contains no cycles and the lengths (number of points) of all $q$-semicycles in $X$ are bounded by some positive integer.} \bigskip \begin{proof} Let $\mathcal{M}_{X}(\sigma ;W,\mathbb{R})=C(X).$\textit{\ Then} $\mathcal{R}_{1}+\mathcal{R}_{2}+...+\mathcal{R}_{k}=C\left( X\right) $, where \begin{equation*} \mathcal{R}_{i}=\{g_{i}(\mathbf{a}^{i}\cdot \mathbf{x):~}g_{i}\in C(\mathbb{% R)}\},~i=1,2,...,k. \end{equation*} Consider the linear space% \begin{equation*} \mathcal{U}=\prod_{i=1}^{k}\mathcal{R}_{i}=\{(g_{1},\ldots ,g_{k}):~g_{i}\in \mathcal{R}_{i},~i=1,\ldots ,k\} \end{equation*}% endowed with the norm \begin{equation*} \Vert (g_{1},\ldots ,g_{k})\Vert =\Vert g_{1}\Vert +\cdots +\Vert g_{k}\Vert . \end{equation*} By $\mathcal{U}^{\ast}$ denote the dual space of $\mathcal{U}$. Each functional $% F\in \mathcal{U}^{\ast }$ can be written as \begin{equation*} F=F_{1}+\cdots +F_{k}, \end{equation*}% where the functionals $F_{i}\in \mathcal{R}_{i}^{\ast }$ and \begin{equation*} F_{i}(g_{i})=F[(0,\ldots ,g_{i},\ldots ,0)],\quad i=1,\ldots ,k. \end{equation*}% We see that the functional $F$ determines the collection $% (F_{1},\ldots ,F_{k})$. Conversely, every collection $(F_{1},\ldots ,F_{k})$ of continuous linear functionals $F_{i}\in \mathcal{R}_{i}^{\ast }$, $% i=1,\ldots ,k$, determines the functional $F_{1}+\cdots +F_{k},$ on $% \mathcal{U}$. Considering this, in what follows, elements of $\mathcal{U}% ^{\ast }$ will be denoted by $(F_{1},\ldots ,F_{k})$. It is not difficult to verify that \begin{equation*} \Vert (F_{1},\ldots ,F_{k})\Vert =\max \{\Vert F_{1}\Vert ,\ldots ,\Vert F_{k}\Vert \}.\eqno(5.11) \end{equation*} Let $l=(\mathbf{x}^{1},\ldots ,\mathbf{x}^{n})$ be any $q$-semicycle (with respect to the directions $\mathbf{a}^{1}$,...,$\mathbf{a}% ^{k}$) in $X$ and $\lambda =(\lambda _{1},\ldots ,\lambda _{n})$ be a vector associated with it. Consider the following functional \begin{equation*} G_{l,\lambda }(f)=\sum_{j=1}^{n}\lambda _{j}f(\mathbf{x}^{j}),\quad f\in C(X). \end{equation*} Since $l$ satisfies (5.10), for each function $g_{i}\in \mathcal{R}_{i}$, $% i=1,\ldots ,k$, we have \begin{equation*} G_{l,\lambda }(g_{i})=\sum_{j=1}^{n}\lambda _{j}g_{i}(\mathbf{a}^{i}\cdot \mathbf{x}^{j})=\sum_{s=1}^{r_{i}}\lambda _{i_{s}}g_{i}(\mathbf{a}^{i}\cdot \mathbf{x}^{i_{s}}),\eqno(5.12) \end{equation*}% where $r_{i}\leq k$. That is, for each set $\mathcal{R}_{i}$, $G_{l,\lambda } $ can be reduced to a functional defined with the help of not more than $k$ points of the semicycle $l$. Consider the operator \begin{equation*} A:\mathcal{U}\rightarrow C(X),\quad A[(g_{1},\ldots ,g_{k})]=g_{1}+\cdots +g_{k}. \end{equation*}% Clearly, $A$ is a linear continuous operator with the norm $\Vert A\Vert =1$% . Besides, since $\mathcal{R}_{1}+\mathcal{R}_{2}+...+\mathcal{R}_{k}=C(X)$, $A$ is a surjection. Consider also the conjugate operator \begin{equation*} A^{\ast }:C(X)^{\ast }\rightarrow \mathcal{U}^{\ast },~A^{\ast }[H]=(F_{1},\ldots ,F_{k}), \end{equation*}% where $F_{i}(g_{i})=H(g_{i})$, for any $g_{i}\in \mathcal{R}_{i}$, $% i=1,\ldots ,k$. Set $A^{\ast }[G_{l,\lambda }]=(G_{1},\ldots ,G_{k})$. From (5.12) it follows that \begin{equation*} |G_{i}(g_{i})|=|G_{l,\lambda }(g_{i})|\leq \Vert g_{i}\Vert \sum_{s=1}^{r_{i}}|\lambda _{i_{s}}|\leq kq\Vert g_{i}\Vert ,\quad i=1,\ldots ,k, \end{equation*}% Therefore, \begin{equation*} \Vert G_{i}\Vert \leq kq,\quad i=1,\ldots ,k. \end{equation*}% From (5.11) we obtain that \begin{equation*} \Vert A^{\ast }[G_{l,\lambda }]\Vert =\Vert (G_{1},\ldots ,G_{k})\Vert \leq kq.\eqno(5.13) \end{equation*}% Since $A$ is a surjection, there exists a positive real number $\delta $ such that \begin{equation*} \Vert A^{\ast }[H]\Vert >\delta \Vert H\Vert \end{equation*}% for any functional $H\in C(X)^{\ast }$(see \cite[p.100]{122}). Taking into account that $% \Vert G_{l,\lambda }\Vert =\sum_{j=1}^{n}|\lambda _{j}|$, for the functional $G_{l,\lambda }$ we have \begin{equation*} \Vert A^{\ast }[G_{l,\lambda }]\Vert >\delta \sum_{j=1}^{n}|\lambda _{j}|.% \eqno(5.14) \end{equation*}% It follows from (5.13) and (5.14) that \begin{equation*} \delta <\frac{kq}{\sum_{j=1}^{n}|\lambda _{j}|}. \end{equation*}% The last inequality shows that $n$ (the length of the arbitrarily chosen $q$% -semicycle $l$) cannot be as great as possible, otherwise $\delta =0$. This simply means that there must be some positive integer bounding the lengths of all $q$-semicycles in $X$. It remains to show that there are no cycles in $X$. Indeed, if $l=(\mathbf{x}% ^{1},\ldots ,\mathbf{x}^{n})$ is a cycle in $X$ and $\lambda =(\lambda _{1},\ldots ,\lambda _{n})$ is a vector associated with it, then the above functional $G_{l,\lambda }$ annihilates all functions from $\mathcal{R}_{1}+% \mathcal{R}_{2}+...+\mathcal{R}_{k}$. On the other hand, $G_{l,\lambda }(f)=\sum_{j=1}^{n}|\lambda _{j}|\neq 0$ for a continuous function $f$ on $X$ satisfying the conditions $f(\mathbf{x}^{j})=1$ if $\lambda _{j}>0$ and $f(% \mathbf{x}^{j})=-1$ if $\lambda _{j}<0$, $j=1,\ldots ,n$. This implies that $% \mathcal{R}_{1}+\mathcal{R}_{2}+...+\mathcal{R}_{k}\neq C\left( X\right) $. Since $\mathcal{M}_{X}(\sigma ;W,\mathbb{R})\subseteq \mathcal{R}_{1}+% \mathcal{R}_{2}+...+\mathcal{R}_{k}$, we obtain that $\mathcal{M}_{X}(\sigma ;W,\mathbb{R})\neq C\left( X\right) $ on the contrary to our assumption. \end{proof} \bigskip \textbf{Remark 5.2.} Assume $\mathcal{M}_{X}(\sigma ;W,\mathbb{R})$ is dense in $C(X).$ Is it necessarily closed? Theorem 5.6 may describe cases when it is not. For example, let $\mathbf{a}^{1}=(1;-1),\ \mathbf{a}^{2}=(1;1),$ $% W=\{\mathbf{a}^{1},\mathbf{a}^{2}\}$ and $\sigma $ be any continuous, bounded and nonconstant function, which has a limit at infinity. Consider the set \begin{eqnarray*} X &=&\{(2;\frac{2}{3}),(\frac{2}{3};\frac{2}{3}),(0;0),(1;1),(1+\frac{1}{2}% ;1-\frac{1}{2}),(1+\frac{1}{2}+\frac{1}{4};1-\frac{1}{2}+\frac{1}{4}), \\ &&(1+\frac{1}{2}+\frac{1}{4}+\frac{1}{8};1-\frac{1}{2}+\frac{1}{4}-\frac{1}{8% }),...\}. \end{eqnarray*} It is clear that $X$ is a compact set with all its orbits closed. (In fact, there is only one orbit, which coincides with $X$). Hence, by Theorem 5.4, $% \overline{\mathcal{M}_{X}(\sigma ;W,\mathbb{R})}=C(X).$ But by Theorem 5.6, $% \mathcal{M}_{X}(\sigma ;W,\mathbb{R})\neq C(X).$ Therefore, the set $% \mathcal{M}_{X}(\sigma ;W,\mathbb{R})$ is not closed in $C(X).$ \bigskip \section{Two hidden layer neural networks} A single hidden layer perceptron is able to approximate a given data with any degree of accuracy. But in applications it is necessary to define how many neurons one should take in a hidden layer. The more the number of neurons, the more the probability of the network to give precise results. Unfortunately, practicality decreases with the increase of the number of neurons in the hidden layer. In other words, single hidden layer perceptrons are not always effective if the number of neurons in the hidden layer is prescribed. In this section, we show that this phenomenon is no longer true for perceptrons with two hidden layers. We prove that a two hidden layer neural network with $d$ inputs, $d$ neurons in the first hidden layer, $2d+2$ neurons in the second hidden layer and with a specifically constructed sigmoidal and infinitely differentiable activation function can approximate any continuous multivariate function with arbitrary accuracy. \subsection{Relation of the Kolmogorov superposition theorem to two hidden layer neural networks} Note that if $r$ is fixed in (5.1), then the set \begin{equation*} \mathcal{M}_{r}(\sigma )=\left\{ \sum_{i=1}^{r}c_{i}\sigma (\mathbf{w}^{i}% \mathbf{\cdot x}-\theta _{i}):~c_{i},\theta _{i}\in \mathbb{R},\mathbf{w\in }% \mathbb{R}^{d}\right\} \end{equation*}% is no longer dense in in the space $C(\mathbb{R}^{d})$ (in the topology of uniform convergence on compact sets) for any activation function $\sigma $. The set $\mathcal{M}_{r}(\sigma )$ will not be dense even if we variate over all univariate continuous functions $\sigma $ (see \cite[Theorem 5.1]% {95}). In the following, we will see that this property of single hidden layer neural networks does not carry over to networks with more than one hidden layer. A two hidden layer network is defined by iteration of the single hidden layer neural network model. The output of two hidden layer perceptron with $r$ units in the first layer, $s$ units in the second layer and the input $x=(x_{1},...,x_{d})$ is \begin{equation*} \sum_{i=1}^{s}d_{i}\sigma \left( \sum_{j=1}^{r}c_{ij}\sigma (\mathbf{w}% ^{ij}\cdot \mathbf{x-}\theta _{ij})-\gamma _{i}\right) . \end{equation*}% Here $d_{i},c_{ij},\theta _{ij},\gamma _{i}$ are real numbers, $\mathbf{w}% ^{ij}$ are vectors of $\mathbb{R}^{d}$ and $\sigma $ is a fixed univariate function. In many applications, it is convenient to take the activation function $% \sigma $ as a \textit{sigmoidal function} which is defined as \begin{equation*} \lim_{t\rightarrow -\infty }\sigma (t)=0\quad \text{ and }\quad \lim_{t\rightarrow +\infty }\sigma (t)=1. \end{equation*}% The literature on neural networks abounds with the use of such functions and their superpositions. The following are typical examples of sigmoidal functions: \begin{align*} \sigma (t)& =\frac{1}{1+e^{-t}} & & \text{(the squashing function),} \\ \sigma (t)& =% \begin{cases} 0, & t\leq -1, \\ \dfrac{t+1}{2}, & -1\leq t\leq 1, \\ 1, & t\geq 1% \end{cases} & & \text{(the piecewise linear function),} \\ \sigma (t)& =\frac{1}{\pi }\arctan t+\frac{1}{2} & & \text{(the arctan sigmoid function),} \\ \sigma (t)& =\frac{1}{\sqrt{2\pi }}\int\limits_{-\infty }^{t}e^{-x^{2}/2}dx & & \text{(the Gaussian function).} \end{align*} In this section, we prove that there exists a two hidden layer neural network model with $d$ units in the first layer and $2d+2$ units in the second layer such that it has the ability to approximate any $d$-variable continuous function with arbitrary accuracy. As an activation function for this model we take a specific sigmoidal function. The idea behind the proof of this result is very much connected to the Kolmogorov superposition theorem (see Section 4.1). This theorem has been much discussed in neural network literature (see, e.g., \cite{119}). In our opinion, the most remarkable application of the Kolmogorov superposition theorem to neural networks was given by Maiorov and Pinkus \cite{99}. They showed that there exists a sigmoidal, strictly increasing, analytic activation function, for which a fixed number of units in both hidden layers are sufficient to approximate arbitrarily well any continuous multivariate function. Namely, the authors of \cite{99} proved the following theorem. \bigskip \textbf{Theorem 5.7 (Maiorov and Pinkus \cite{99}).} \textit{There exists an activation function $\sigma $ which is analytic, strictly increasing and sigmoidal and has the following property: For any $f\in C[0,1]^{d}$ and $% \varepsilon >0,$ there exist constants $d_{i},$ $c_{ij},$ $\theta _{ij},$ $% \gamma _{i}$, and vectors $\mathbf{w}^{ij}\in \mathbb{R}^{d}$ for which} \begin{equation*} \left\vert f(\mathbf{x})-\sum_{i=1}^{6d+3}d_{i}\sigma \left( \sum_{j=1}^{3d}c_{ij}\sigma (\mathbf{w}^{ij}\cdot \mathbf{x-}\theta _{ij})-\gamma _{i}\right) \right\vert <\varepsilon \eqno(5.15) \end{equation*}% \textit{for all $\mathbf{x}=(x_{1},...,x_{d})\in \lbrack 0,1]^{d}.$} \bigskip This theorem is based on the following version of the Kolmogorov superposition theorem given by Lorentz \cite{98} and Sprecher \cite{128}. \bigskip \textbf{Theorem 5.8 (Kolmogorov's superposition theorem).} \textit{For the unit cube $\mathbb{I}^{d},~\mathbb{I}=[0,1],~d\geq 2,$ there exists constants $\lambda _{q}>0,$ $q=1,...,d,$ $\sum_{q=1}^{d}\lambda _{q}=1,$ and nondecreasing continuous functions $\phi _{p}:[0,1]\rightarrow \lbrack 0,1],$ $p=1,...,2d+1,$ such that every continuous function $f:\mathbb{I}% ^{d}\rightarrow \mathbb{R}$ admits the representation} \begin{equation*} f(x_{1},...x_{d})=\sum_{p=1}^{2d+1}g\left( \sum_{q=1}^{d}\lambda _{q}\phi _{p}(x_{q})\right) \eqno(5.16) \end{equation*}% \textit{for some $g\in C[0,1]$ depending on $f.$} \bigskip In the next subsection, using the general ideas developed in \cite{99}, we show that the bounds of units in hidden layers in (5.15) may be chosen even equal to the bounds in the Kolmogorov superposition theorem. More precisely, these bounds can be taken as $2d+2$ and $d$ instead of $6d+3$ and $3d$. To attain this purpose, we change the ``analyticity" of $\sigma $ to ``infinite differentiability". In addition, near infinity we assume that $\sigma $ is ``$% \lambda $-strictly increasing" instead of being ``strictly increasing". \bigskip \subsection{The main result} \smallskip We begin this subsection with a definition of a \textit{$\lambda$-monotone function}. Let $\lambda $ be any nonnegative number. A real function $f$ defined on $(a,b)$ is called \textit{$\lambda $-increasing} (\textit{$\lambda $-decreasing}) if there exists an increasing (decreasing) function $u:(a,b)\rightarrow \mathbb{R}$ such that $\left\vert f(x)-u(x)\right\vert \leq \lambda ,$ for all $x\in (a,b)$. If $u$ is strictly increasing (or strictly decreasing), then the above function $f$ is called a \textit{$\lambda $-strictly increasing} (or \textit{$% \lambda $-strictly decreasing}) function. Clearly, $0$-monotonicity coincides with the usual concept of monotonicity and a $\lambda _{1}$-monotone function is $\lambda _{2}$-monotone if $\lambda _{1}\leq \lambda _{2}$. It is also clear from the definition that a $\lambda $-monotone function behaves like a usual monotone function as $\lambda $ gets very small. Our purpose is to prove the following theorem. \bigskip \textbf{Theorem 5.9.} \textit{For any positive numbers $\alpha $ and $% \lambda $,\textit{\ }there exists a $C^{\infty }(\mathbb{R}),$ sigmoidal activation function $\sigma :$ $\mathbb{R\rightarrow R}$ which is strictly increasing on $(-\infty ,\alpha )$, $\lambda$-strictly increasing on $[\alpha ,+\infty )$, and satisfies the following property: For any $f\in C[0,1]^{d}$ and $\varepsilon >0,$ there exist constants $% d_{p}, $ $c_{pq},$ $\theta _{pq},$ $\gamma _{p}$, and vectors $\mathbf{w}% ^{pq}\in \mathbb{R}^{d}$ for which} \begin{equation*} \left\vert f(\mathbf{x})-\sum_{p=1}^{2d+2}d_{p}\sigma \left( \sum_{q=1}^{d}c_{pq}\sigma (\mathbf{w}^{pq}\cdot \mathbf{x-}\theta _{pq})-\gamma _{p}\right) \right\vert <\varepsilon \eqno(5.17) \end{equation*}% \textit{for all $\mathbf{x}=(x_{1},...,x_{d})\in \lbrack 0,1]^{d}.$} \bigskip \begin{proof} Let $\alpha $ be any positive number. Divide the interval $% [\alpha ,+\infty )$ into the segments $[\alpha ,2\alpha ],$ $[2\alpha ,3\alpha ],...$. Let $h(t)$ be any strictly increasing, infinitely differentiable function on $[\alpha ,+\infty )$ with the properties \bigskip 1) $0<h(t)<1$ for all $t\in \lbrack \alpha ,+\infty )$; 2) $1-h(\alpha )\leq \lambda ;$ 3) $h(t)\rightarrow 1,$ as $t\rightarrow +\infty .$ \bigskip The existence of a strictly increasing smooth function satisfying these properties is easy to verify. Note that from conditions (1)-(3) it follows that any function $f(t)$ satisfying the inequality $h(t)<f(t)<1$ for all $% t\in \lbrack \alpha ,+\infty ),$ is $\lambda $-strictly increasing and $% f(t)\rightarrow 1,$ as $t\rightarrow +\infty .$ We are going to construct $\sigma $ obeying the required properties in stages. Let $\{u_{n}(t)\}_{n=1}^{\infty }$ be the sequence of all polynomials with rational coefficients defined on $[0,1].$ First, we define $% \sigma $ on the closed intervals $[(2m-1)\alpha ,2m\alpha ],$ $m=1,2,...$, as the function \begin{equation*} \sigma (t)=a_{m}+b_{m}u_{m}(\frac{t}{\alpha }-2m+1),\text{ }t\in \lbrack (2m-1)\alpha ,2m\alpha ],\eqno(5.18) \end{equation*}% or equivalently, \begin{equation*} \sigma (\alpha t+(2m-1)\alpha )=a_{m}+b_{m}u_{m}(t),\text{ }t\in \lbrack 0,1],\eqno(5.19) \end{equation*}% where $a_{m}$ and $b_{m}\neq 0$ are appropriately chosen constants. These constants are determined from the condition \begin{equation*} h(t)<\sigma (t)<1,\eqno(5.20) \end{equation*}% \bigskip for all $t\in \lbrack (2m-1)\alpha ,2m\alpha ].$ There is a simple procedure for determining a suitable pair of $a_{m}$ and $b_{m}$. Indeed, let \begin{equation*} M=\max h(t)\text{, }A_{1}=\min u_{m}(\frac{t}{\alpha }-2m+1)\text{, }% A_{2}=\max u_{m}(\frac{t}{\alpha }-2m+1), \end{equation*}% where in all the above $\max $ and $\min $, the variable $t$ runs over the closed interval $[(2m-1)\alpha ,2m\alpha ].$ Note that $M<1$. If $% A_{1}=A_{2} $ (that is, if the function $u_{m}$ is constant on $[0,1]$), then we can set $\sigma (t)=(1+M)/2$ and easily find a suitable pair of $% a_{m}$ and $b_{m}$ from (5.18). Let now $A_{1}\neq A_{2}$ and $y=a+bx,$ $% b\neq 0,$ be a linear function mapping the segment $[A_{1},A_{2}]$ into $% (M,1).$ Then it is enough to take $a_{m}=a$ and $b_{m}=b.$ At the second stage we define $\sigma $ on the intervals $[2m\alpha ,(2m+1)\alpha ],$ $m=1,2,...,$ so that it is in $C^{\infty }(\mathbb{R})$ and satisfies the inequality (5.20). Finally, in all of $(-\infty ,\alpha )$ we define $\sigma $ while maintaining the $C^{\infty }$ strict monotonicity property, and also in such a way that $\lim_{t\rightarrow -\infty }\sigma (t)=0.$ We obtain from the properties of $h$ and the condition (5.20) that $% \sigma (t)$ is a $\lambda $-strictly increasing function on the interval $% [\alpha ,+\infty )$ and $\sigma (t)\rightarrow 1$, as $t\rightarrow +\infty . $ From the above construction of $\sigma $, that is, from (5.19) it follows that for each $m=1,2,...,$ there exists numbers $A_{m}$,$\ B_{m}$ and $r_{m}$ such that \begin{equation*} u_{m}(t)=A_{m}\sigma (\alpha t-r_{m})-B_{m},\eqno(5.21) \end{equation*}% where $A_{m}\neq 0.$ Let $f$ be any continuous function on the unit cube $[0,1]^{d}.$ By the Kolmogorov superposition theorem the expansion (5.16) is valid for $f.$ For the exterior continuous univariate function $g(t)$ in (5.16) and for any $% \varepsilon >0$ there exists a polynomial $u_{m}(t)$ of the above form such that \begin{equation*} \left\vert g(t)-u_{m}(t)\right\vert <\frac{\varepsilon }{2(2d+1)}, \end{equation*}% for all $t\in \lbrack 0,1].$ This together with (5.21) means that \begin{equation*} \left\vert g(t)-[a\sigma (\alpha t-r)-b]\right\vert <\frac{\varepsilon }{% 2(2d+1)},\eqno(5.22) \end{equation*}% for some $a,b,r\in \mathbb{R}$ and all $t\in \lbrack 0,1].$ Substituting (5.22) in (5.16) we obtain that \begin{equation*} \left\vert f(x_{1},...,x_{d})-\sum_{p=1}^{2d+1}\left( a\sigma \left( \alpha \cdot \sum_{q=1}^{d}\lambda _{q}\phi _{p}(x_{q})-r\right) -b\right) \right\vert <\frac{\varepsilon }{2}\eqno(5.23) \end{equation*}% for all $(x_{1},...,x_{d})\in \lbrack 0,1]^{d}.$ For each $p\in \{1,2,...,2d+1\}$ and $\delta >0$ there exist constants $% a_{p},b_{p}$ and $r_{p}$ such that \begin{equation*} \left\vert \phi _{p}(x_{q})-[a_{p}\sigma (\alpha x_{q}-r_{p})-b_{p}]\right\vert <\delta ,\eqno(5.24) \end{equation*}% for all $x_{q}\in \lbrack 0,1].$ Since $\lambda _{q}>0,$ $q=1,...,d,$ $% \sum_{q=1}^{d}\lambda _{q}=1,$ it follows from (5.24) that \begin{equation*} \left\vert \sum_{q=1}^{d}\lambda _{q}\phi _{p}(x_{q})-\left[ \sum_{q=1}^{d}\lambda _{q}a_{p}\sigma (\alpha x_{q}-r_{p})-b_{p}\right] \right\vert <\delta ,\eqno(5.25) \end{equation*}% for all $(x_{1},...,x_{d})\in \lbrack 0,1]^{d}.$ Now since the function $a\sigma (\alpha t-r)$ is uniformly continuous on every closed interval, we can choose $\delta $ sufficiently small and obtain from (5.25) that \[ \left\vert \sum_{p=1}^{2d+1}a\sigma \left( \alpha \sum_{q=1}^{d}\lambda _{q}\phi _{p}(x_{q})-r\right) \ -\sum_{p=1}^{2d+1}a\sigma \left( \alpha \left[ \sum_{q=1}^{d}\lambda _{q}a_{p}\sigma (\alpha x_{q}-r_{p})-b_{p}\right] -r\right) \right\vert \] \[ < \frac{\varepsilon }{2}. \] This inequality may be rewritten as \begin{equation*} \left\vert \sum_{p=1}^{2d+1}a\sigma \left( \alpha \sum_{q=1}^{d}\lambda _{q}\phi _{p}(x_{q})-r\right) -\sum_{p=1}^{2d+1}d_{p}\sigma \left( \sum_{q=1}^{d}c_{pq}\sigma (\mathbf{w}% ^{pq}\cdot \mathbf{x}-\theta _{pq})-\gamma _{p}\right) \right\vert <\frac{% \varepsilon }{2}.\eqno(5.26) \end{equation*}% From (5.23) and (5.26) it follows that \begin{equation*} \left\vert f(\mathbf{x})-\left[ \sum_{p=1}^{2d+1}d_{p}\sigma \left( \sum_{q=1}^{d}c_{pq}\sigma (\mathbf{w}^{pq}\cdot \mathbf{x-}\theta _{pq})-\gamma _{p}\right) -s\right] \right\vert <\varepsilon ,\eqno(5.27) \end{equation*}% where $s=(2d+1)b$. Since the constant $s$ can be written in the form \begin{equation*} s=d\sigma \left( \sum_{q=1}^{d}c_{q}\sigma (\mathbf{w}^{q}\cdot \mathbf{x-}% \theta _{q})-\gamma \right) , \end{equation*}% from (5.27) we finally obtain the validity of (5.17). \end{proof} \textbf{Remark 5.3.} It is easily seen in the proof of Theorem 5.9 that all the weights $\mathbf{w}^{ij}$ are fixed (see (5.26)). Namely, $\mathbf{w}% ^{ij}=\alpha \mathbf{e}^{j},$ for all $i=1,...,2d+2,$ $j=1,...,d,$ where $% \mathbf{e}^{j}$ is the $j$-th coordinate vector of the space $\mathbb{R}^{d}$% . \bigskip The next theorem follows from Theorem 5.9 easily, since the Kolmogorov superposition theorem is valid for all compact sets of $\mathbb{R}^{d}$. \bigskip \textbf{Theorem 5.10.} \textit{Let $Q$ be a compact set in $\mathbb{R}^{d}.$ For any numbers $\alpha \in \mathbb{R}$ and $\lambda >0,$ there exists a \textit{$C^{\infty }(\mathbb{R}),$ sigmoidal activation function }$\sigma :$ $\mathbb{R\rightarrow R}$ which is strictly increasing on $(-\infty ,\alpha ) $, $\lambda$-strictly increasing on $[\alpha ,+\infty )$, and satisfies the following property: For any $f\in C(Q)$ and $\varepsilon >0$ there exist real numbers $d_{i},$ $c_{ij},$ $\theta _{ij}$, $\gamma _{i},$ and vectors $\mathbf{w}^{ij}\in \mathbb{R}^{d}$ for which} \begin{equation*} \left\vert f(\mathbf{x})-\sum_{i=1}^{2d+2}d_{i}\sigma \left( \sum_{j=1}^{d}c_{ij}\sigma (\mathbf{w}^{ij}\cdot \mathbf{x-}\theta _{ij})-\gamma _{i}\right) \right\vert <\varepsilon \end{equation*}% \textit{for all $\mathbf{x}=(x_{1},...,x_{d})\in Q.$} \bigskip \textbf{Remark 5.4.} In some literature, a single hidden layer perceptron is defined as the function \begin{equation*} \sum_{i=1}^{r}c_{i}\sigma (\mathbf{w}^{i}\mathbf{\cdot x}-\theta _{i})-c_{0}. \end{equation*}% A two hidden layer network then takes the form \begin{equation*} \sum_{i=1}^{s}d_{i}\sigma \left( \sum_{j=1}^{r}c_{ij}\sigma (\mathbf{w}% ^{ij}\cdot \mathbf{x-}\theta _{ij})-\gamma _{i}\right) -d_{0}.\eqno(5.28) \end{equation*}% The proof of Theorem 5.9 shows that for networks of type (5.28) the theorem is valid if we take $2d+1$ neurons in the second hidden layer (instead of $% 2d+2$ neurons as above). That is, there exist networks of type (5.28) having the universal approximation property and for which the number of units in the hidden layers is equal to the number of summands in the Kolmogorov superposition theorem. \bigskip \textbf{Remark 5.5.} It is known that the $2d+1$ in the Kolmogorov superposition theorem is minimal (see Sternfeld \cite{130}). Thus it is doubtful if the number of neurons in Theorems 5.9 and 5.10 can be reduced. \bigskip \textbf{Remark 5.6.} Inequality (5.22) shows that single hidden layer neural networks of the form (5.28) with the activation function $\sigma$ and with only one neuron in the hidden layer can approximate any continuous function on the interval $[0,1]$ with arbitrary precision. Since the number $b$ in (5.22) can always be written as $b=a_1\sigma (0 \cdot t-r_1)$ for some $a_1$ and $r_1$, we see that two neurons in the hidden layer are sufficient for traditional single hidden layer neural networks with the activation function $\sigma$ to approximate continuous functions on $[0,1]$. Applying the linear transformation $x=a+(b-a)t$ it can be proven that the same argument holds for any interval $[a,b]$. \bigskip \section{Construction of a universal sigmoidal function} In the preceding section, we considered two theorems (Theorem 5.7 of Maiorov and Pinkus, and Theorem 5.9) on the approximation capabilities of the MLP model of neural networks with a prescribed number of hidden neurons. Note that both results are more theoretical than practical, as they indicate only the existence of the corresponding activation functions. In this section, we construct algorithmically a smooth, sigmoidal, almost monotone activation function $\sigma $ providing approximation to an arbitrary continuous function within any degree of accuracy. This algorithm is implemented in a computer program, which computes the value of $\sigma $ at any reasonable point of the real axis. \bigskip \subsection{A construction algorithm} In this subsection, we construct algorithmically a sigmoidal function $% \sigma $ which we use in our results in Section 5.3.3. To start with the construction of $\sigma$, assume that we are given a closed interval $[a, b]$ and a sufficiently small real number $\lambda$. We construct $\sigma$ algorithmically, based on two numbers, namely $\lambda$ and $d := b - a$. The following steps describe the algorithm. \textit{Step 1.} Introduce the function \begin{equation*} h(x) := 1 - \frac{\min\{1/2, \lambda\}}{1 + \log(x - d + 1)}. \end{equation*} Note that this function is strictly increasing on the real line and satisfies the following properties: \begin{enumerate} \item $0 < h(x) < 1$ for all $x \in [d, +\infty)$; \item $1 - h(d) \le \lambda$; \item $h(x) \to 1$, as $x \to +\infty$. \end{enumerate} We want to construct $\sigma $ satisfying the inequalities \begin{equation*} h(x)<\sigma (x)<1\eqno(5.29) \end{equation*}% for $x\in \lbrack d,+\infty )$. Then our $\sigma $ will tend to $1$ as $x$ tends to $+\infty $ and obey the inequality \begin{equation*} |\sigma (x)-h(x)|\leq \lambda , \end{equation*}% i.e., it will be a $\lambda $-increasing function. \textit{Step 2.} Before proceeding to the construction of $\sigma$, we need to enumerate the monic polynomials with rational coefficients. Let $q_n$ be the Calkin--Wilf sequence (see~\cite{CW00}). Then we can enumerate all the rational numbers by setting \begin{equation*} r_0 := 0, \quad r_{2n} := q_n, \quad r_{2n-1} := -q_n, \ n = 1, 2, \dots. \end{equation*} Note that each monic polynomial with rational coefficients can uniquely be written as $r_{k_0} + r_{k_1} x + \ldots + r_{k_{l-1}} x^{l-1} + x^l$, and each positive rational number determines a unique finite continued fraction \begin{equation*} [m_0; m_1, \ldots, m_l] := m_0 + \dfrac1{m_1 + \dfrac1{m_2 + \dfrac1{\ddots + \dfrac1{m_l}}}} \end{equation*} with $m_0 \ge 0$, $m_1, \ldots, m_{l-1} \ge 1$ and $m_l \ge 2$. We now construct a bijection between the set of all monic polynomials with rational coefficients and the set of all positive rational numbers as follows. To the only zeroth-degree monic polynomial 1 we associate the rational number 1, to each first-degree monic polynomial of the form $r_{k_0} + x$ we associate the rational number $k_0 + 2$, to each second-degree monic polynomial of the form $r_{k_0} + r_{k_1} x + x^2$ we associate the rational number $[k_0; k_1 + 2] = k_0 + 1 / (k_1 + 2)$, and to each monic polynomial \begin{equation*} r_{k_0} + r_{k_1} x + \ldots + r_{k_{l-2}} x^{l-2} + r_{k_{l-1}} x^{l-1} + x^l \end{equation*} of degree $l \ge 3$ we associate the rational number $[k_0; k_1 + 1, \ldots, k_{l-2} + 1, k_{l-1} + 2]$. In other words, we define $u_1(x) := 1$, \begin{equation*} u_n(x) := r_{q_n-2} + x \end{equation*} if $q_n \in \mathbb{Z}$, \begin{equation*} u_n(x) := r_{m_0} + r_{m_1-2} x + x^2 \end{equation*} if $q_n = [m_0; m_1]$, and \begin{equation*} u_n(x) := r_{m_0} + r_{m_1-1} x + \ldots + r_{m_{l-2}-1} x^{l-2} + r_{m_{l-1}-2} x^{l-1} + x^l \end{equation*} if $q_n = [m_0; m_1, \ldots, m_{l-2}, m_{l-1}]$ with $l \ge 3$. For example, the first few elements of this sequence are \begin{equation*} 1, \quad x^2, \quad x, \quad x^2 - x, \quad x^2 - 1, \quad x^3, \quad x - 1, \quad x^2 + x, \quad \ldots. \end{equation*} \textit{Step 3.} We start with constructing $\sigma$ on the intervals $% [(2n-1)d, 2nd]$, $n = 1, 2, \ldots$. For each monic polynomial $u_n(x) = \alpha_0 + \alpha_1 x + \ldots + \alpha_{l-1} x^{l-1} + x^l$, set \begin{equation*} B_1 := \alpha_0 + \frac{\alpha_1-|\alpha_1|}{2} + \ldots + \frac{% \alpha_{l-1} - |\alpha_{l-1}|}{2} \end{equation*} and \begin{equation*} B_2 := \alpha_0 + \frac{\alpha_1+|\alpha_1|}{2} + \ldots + \frac{% \alpha_{l-1} + |\alpha_{l-1}|}{2} + 1. \end{equation*} Note that the numbers $B_1$ and $B_2$ depend on $n$. To avoid complication of symbols, we do not indicate this in the notation. Introduce the sequence \begin{equation*} M_n := h((2n+1)d), \qquad n = 1, 2, \ldots. \end{equation*} Clearly, this sequence is strictly increasing and converges to $1$. Now we define $\sigma $ as the function \begin{equation*} \sigma (x):=a_{n}+b_{n}u_{n}\left( \frac{x}{d}-2n+1\right) ,\quad x\in \lbrack (2n-1)d,2nd],\eqno(5.30) \end{equation*}% where \begin{equation*} a_{1}:=\frac{1}{2},\qquad b_{1}:=\frac{h(3d)}{2},\eqno(5.31) \end{equation*}% and \begin{equation*} a_{n}:=\frac{(1+2M_{n})B_{2}-(2+M_{n})B_{1}}{3(B_{2}-B_{1})},\qquad b_{n}:=% \frac{1-M_{n}}{3(B_{2}-B_{1})},\qquad n=2,3,\ldots .\eqno(5.32) \end{equation*} It is not difficult to notice that for $n>2$ the numbers $a_{n}$, $b_{n}$ are the coefficients of the linear function $y=a_{n}+b_{n}x$ mapping the closed interval $[B_{1},B_{2}]$ onto the closed interval $% [(1+2M_{n})/3,(2+M_{n})/3]$. Besides, for $n=1$, i.e. on the interval $% [d,2d] $, \begin{equation*} \sigma (x)=\frac{1+M_{1}}{2}. \end{equation*}% Therefore, we obtain that \begin{equation*} h(x)<M_{n}<\frac{1+2M_{n}}{3}\leq \sigma (x)\leq \frac{2+M_{n}}{3}<1,\eqno% (5.33) \end{equation*}% for all $x\in \lbrack (2n-1)d,2nd]$, $n=1$, $2$, $\ldots $. \textit{Step 4.} In this step, we construct $\sigma $ on the intervals $% [2nd,(2n+1)d]$, $n=1,2,\ldots $. For this purpose we use the \textit{smooth transition function }% \begin{equation*} \beta _{a,b}(x):=\frac{\widehat{\beta }(b-x)}{\widehat{\beta }(b-x)+\widehat{% \beta }(x-a)}, \end{equation*}% where \begin{equation*} \widehat{\beta }(x):=% \begin{cases} e^{-1/x}, & x>0, \\ 0, & x\leq 0.% \end{cases}% \end{equation*}% Obviously, $\beta _{a,b}(x)=1$ for $x\leq a$, $\beta _{a,b}(x)=0$ for $x\geq b$, and $0<\beta _{a,b}(x)<1$ for $a<x<b$. Set \begin{equation*} K_n := \frac{\sigma(2nd) + \sigma((2n+1)d)}{2}, \qquad n = 1, 2, \ldots. \end{equation*} Note that the numbers $\sigma(2nd)$ and $\sigma((2n+1)d)$ have already been defined in the previous step. Since both the numbers $\sigma(2nd)$ and $% \sigma((2n+1)d)$ lie in the interval $(M_n, 1)$, it follows that $K_n \in (M_n, 1)$. First we extend $\sigma $ smoothly to the interval $[2nd,2nd+d/2]$. Take $% \varepsilon :=(1-M_{n})/6$ and choose $\delta \leq d/2$ such that \begin{equation*} \left\vert a_{n}+b_{n}u_{n}\left( \frac{x}{d}-2n+1\right) -\left( a_{n}+b_{n}u_{n}(1)\right) \right\vert \leq \varepsilon ,\quad x\in \lbrack 2nd,2nd+\delta ].\eqno(5.34) \end{equation*}% One can choose this $\delta $ as \begin{equation*} \delta :=\min \left\{ \frac{\varepsilon d}{b_{n}C},\frac{d}{2}\right\} , \end{equation*}% where $C>0$ is a number satisfying $|u_{n}^{\prime }(x)|\leq C$ for $x\in (1,1.5)$. For example, for $n=1$, $\delta $ can be chosen as $d/2$. Now define $\sigma $ on the first half of the interval $[2nd,(2n+1)d]$ as the function \begin{equation*} \sigma(x) :=K_{n}-\beta _{2nd,2nd+\delta }(x) \end{equation*} \begin{equation*} \times \left( K_{n}-a_{n}-b_{n}u_{n}\left( \frac{x}{d}-2n+1\right) \right) , x\in \left[ 2nd,2nd+\frac{d}{2}\right].\eqno(5.35) \end{equation*} Let us prove that $\sigma (x)$ satisfies the condition~(5.29). Indeed, if $% 2nd+\delta \leq x\leq 2nd+d/2$, then there is nothing to prove, since $% \sigma (x)=K_{n}\in (M_{n},1)$. If $2nd\leq x<2nd+\delta $, then $0<\beta _{2nd,2nd+\delta }(x)\leq 1$ and hence from~(5.35) it follows that for each $% x\in \lbrack 2nd,2nd+\delta )$, $\sigma (x)$ is between the numbers $K_{n}$ and $A_{n}(x):=a_{n}+b_{n}u_{n}\left( \frac{x}{d}-2n+1\right) $. On the other hand, from~(5.34) we obtain that \begin{equation*} a_{n}+b_{n}u_{n}(1)-\varepsilon \leq A_{n}(x)\leq a_{n}+b_{n}u_{n}(1)+\varepsilon , \end{equation*}% which together with~(5.30) and~(5.33) yields $A_{n}(x)\in \left[ \frac{% 1+2M_{n}}{3}-\varepsilon ,\frac{2+M_{n}}{3}+\varepsilon \right] $ for $x\in \lbrack 2nd,2nd+\delta )$. Since $\varepsilon =(1-M_{n})/6$, the inclusion $% A_{n}(x)\in (M_{n},1)$ is valid. Now since both $K_{n}$ and $A_{n}(x)$ belong to $(M_{n},1)$, we finally conclude that \begin{equation*} h(x)<M_{n}<\sigma (x)<1,\quad \text{for }x\in \left[ 2nd,2nd+\frac{d}{2}% \right] . \end{equation*} We define $\sigma $ on the second half of the interval in a similar way: \begin{equation*} \begin{split} \sigma (x)& :=K_{n}-(1-\beta _{(2n+1)d-\overline{\delta },(2n+1)d}(x)) \\ & \times \left( K_{n}-a_{n+1}-b_{n+1}u_{n+1}\left( \frac{x}{d}-2n-1\right) \right) ,\quad x\in \left[ 2nd+\frac{d}{2},(2n+1)d\right] , \end{split}% \end{equation*}% where \begin{equation*} \overline{\delta }:=\min \left\{ \frac{\overline{\varepsilon }d}{b_{n+1}% \overline{C}},\frac{d}{2}\right\} ,\qquad \overline{\varepsilon }:=\frac{% 1-M_{n+1}}{6},\qquad \overline{C}\geq \sup_{\lbrack -0.5,0]}|u_{n+1}^{\prime }(x)|. \end{equation*}% One can easily verify, as above, that the constructed $\sigma (x)$ satisfies the condition~(5.29) on $[2nd+d/2,2nd+d]$ and \begin{equation*} \sigma \left( 2nd+\frac{d}{2}\right) =K_{n},\qquad \sigma ^{(i)}\left( 2nd+% \frac{d}{2}\right) =0,\quad i=1,2,\ldots . \end{equation*} Steps 3 and 4 construct $\sigma$ on the interval $[d, +\infty)$. \textit{Step 5.} On the remaining interval $(-\infty ,d)$, we define $\sigma $ as \begin{equation*} \sigma (x):=\left( 1-\widehat{\beta }(d-x)\right) \frac{1+M_{1}}{2},\quad x\in (-\infty ,d). \end{equation*}% It is not difficult to verify that $\sigma $ is a strictly increasing, smooth function on $(-\infty ,d)$. Note also that $\sigma (x)\rightarrow \sigma (d)=(1+M_{1})/2$, as $x$ tends to $d$ from the left and $\sigma ^{(i)}(d)=0$ for $i=1$, $2$, $\ldots $. This final step completes the construction of $\sigma $ on the whole real line. \bigskip \subsection{Properties of the constructed sigmoidal function} It should be noted that the above algorithm allows one to compute the constructed $\sigma$ at any point of the real axis instantly. The code of this algorithm is available at % \url{http://sites.google.com/site/njguliyev/papers/monic-sigmoidal}. As a practical example, we give here the graph of $\sigma$ (see Figure~\ref% {fig:sigma}) and a numerical table (see Table~\ref{tbl:sigma}) containing several computed values of this function on the interval $[0, 20]$. Figure~% \ref{fig:sigma100} shows how the graph of $\lambda$-increasing function $% \sigma$ changes on the interval $[0,100]$ as the parameter $\lambda$ decreases. The above $\sigma$ obeys the following properties: \begin{enumerate} \item $\sigma$ is sigmoidal; \item $\sigma \in C^{\infty}(\mathbb{R})$; \item $\sigma$ is strictly increasing on $(-\infty, d)$ and $\lambda$% -strictly increasing on $[d, +\infty)$; \item $\sigma$ is easily computable in practice. \end{enumerate} All these properties are easily seen from the above exposition. But the essential property of our sigmoidal function is its ability to approximate an arbitrary continuous function using only a fixed number of translations and scalings of $\sigma$. More precisely, only two translations and scalings are sufficient. We formulate this important property as a theorem in the next section. \begin{figure}[tbp] \includegraphics[width=1.0\textwidth]{monic-0-20} \caption{The graph of $\protect\sigma$ on $[0, 20]$ ($d = 2$, $\protect% \lambda = 1/4$)} \label{fig:sigma} \end{figure} \begin{table}[tbp] \caption{Some computed values of $\protect\sigma$ ($d = 2$, $\protect\lambda % = 1/4$)} \label{tbl:sigma}% \begin{tabular}{|c|c|c|c|c|c|c|c|c|c|} \hline $t$ & $\sigma$ & $t$ & $\sigma$ & $t$ & $\sigma$ & $t$ & $\sigma$ & $t$ & $% \sigma$ \\ \hline $0.0$ & $0.37462$ & $4.0$ & $0.95210$ & $8.0$ & $0.97394$ & $12.0$ & $% 0.97662 $ & $16.0$ & $0.96739$ \\ \hline $0.4$ & $0.44248$ & $4.4$ & $0.95146$ & $8.4$ & $0.96359$ & $12.4$ & $% 0.97848 $ & $16.4$ & $0.96309$ \\ \hline $0.8$ & $0.53832$ & $4.8$ & $0.95003$ & $8.8$ & $0.96359$ & $12.8$ & $% 0.97233 $ & $16.8$ & $0.96309$ \\ \hline $1.2$ & $0.67932$ & $5.2$ & $0.95003$ & $9.2$ & $0.96314$ & $13.2$ & $% 0.97204 $ & $17.2$ & $0.96307$ \\ \hline $1.6$ & $0.87394$ & $5.6$ & $0.94924$ & $9.6$ & $0.95312$ & $13.6$ & $% 0.97061 $ & $17.6$ & $0.96067$ \\ \hline $2.0$ & $0.95210$ & $6.0$ & $0.94787$ & $10.0$ & $0.95325$ & $14.0$ & $% 0.96739$ & $18.0$ & $0.95879$ \\ \hline $2.4$ & $0.95210$ & $6.4$ & $0.94891$ & $10.4$ & $0.95792$ & $14.4$ & $% 0.96565$ & $18.4$ & $0.95962$ \\ \hline $2.8$ & $0.95210$ & $6.8$ & $0.95204$ & $10.8$ & $0.96260$ & $14.8$ & $% 0.96478$ & $18.8$ & $0.96209$ \\ \hline $3.2$ & $0.95210$ & $7.2$ & $0.95725$ & $11.2$ & $0.96727$ & $15.2$ & $% 0.96478$ & $19.2$ & $0.96621$ \\ \hline $3.6$ & $0.95210$ & $7.6$ & $0.96455$ & $11.6$ & $0.97195$ & $15.6$ & $% 0.96565$ & $19.6$ & $0.97198$ \\ \hline \end{tabular}% \end{table} \begin{figure}[tbp] \includegraphics[width=0.75\textwidth]{monic-0-100} \caption{The graph of $\protect\sigma$ on $[0, 100]$ ($d = 2$)} \label{fig:sigma100} \end{figure} \bigskip \subsection{Theoretical results} The following theorems are valid. \bigskip \textbf{Theorem 5.11.} \textit{Assume that $f$ is a continuous function on a finite segment $[a,b]$ of $\mathbb{R}$ and $\sigma $ is the sigmoidal function constructed in Section~5.3.1. Then for any sufficiently small $% \varepsilon >0 $ there exist constants $c_{1}$, $c_{2}$, $\theta _{1}$ and $% \theta _{2}$ such that} \begin{equation*} |f(x)-c_{1}\sigma (x-\theta _{1})-c_{2}\sigma (x-\theta _{2})|<\varepsilon \end{equation*}% \textit{for all $x\in \lbrack a,b]$.} \bigskip \begin{proof} Set $d:=b-a$ and divide the interval $[d,+\infty )$ into the segments $[d,2d]$, $[2d,3d]$, $\ldots $. It follows from~(5.30) that \begin{equation*} \sigma (dx+(2n-1)d)=a_{n}+b_{n}u_{n}(x),\quad x\in \lbrack 0,1]\eqno(5.36) \end{equation*}% for $n=1$, $2$, $\ldots $. Here $a_{n}$ and $b_{n}$ are computed by~(5.31) and~(5.32) for $n=1$ and $n>1$, respectively. From~(5.36) it follows that for each $n=1$, $2$, $\ldots $, \begin{equation*} u_{n}(x)=\frac{1}{b_{n}}\sigma (dx+(2n-1)d)-\frac{a_{n}}{b_{n}}.\eqno(5.37) \end{equation*} Let now $g$ be any continuous function on the unit interval $[0,1]$. By the density of polynomials with rational coefficients in the space of continuous functions on any compact subset of $\mathbb{R}$, for any $\varepsilon >0$ there exists a polynomial $p(x)$ of the above form such that \begin{equation*} |g(x)-p(x)|<\varepsilon \end{equation*}% for all $x\in \lbrack 0,1]$. Denote by $p_{0}$ the leading coefficient of $p$% . If $p_{0}\neq 0$ (i.e., $p\not\equiv 0$) then we define $u_{n}$ as $% u_{n}(x):=p(x)/p_{0}$, otherwise we just set $u_{n}(x):=1$. In both cases \begin{equation*} |g(x)-p_{0}u_{n}(x)|<\varepsilon ,\qquad x\in \lbrack 0,1]. \end{equation*}% This together with~(5.37) means that \begin{equation*} |g(x)-c_{1}\sigma (dx-s_{1})-c_{0}|<\varepsilon \eqno(5.38) \end{equation*}% for some $c_{0}$, $c_{1}$, $s_{1}\in \mathbb{R}$ and all $x\in \lbrack 0,1]$% . Namely, $c_{1}=p_{0}/b_{n}$, $s_{1}=d-2nd$ and $c_{0}=p_{0}a_{n}/b_{n}$. On the other hand, we can write $c_{0}=c_{2}\sigma (dx-s_{2})$, where $% c_{2}:=2c_{0}/(1+h(3d))$ and $s_{2}:=-d$. Hence, \begin{equation*} |g(x)-c_{1}\sigma (dx-s_{1})-c_{2}\sigma (dx-s_{2})|<\varepsilon .\eqno(5.39) \end{equation*}% Note that~(5.39) is valid for the unit interval $[0,1]$. Using linear transformation it is not difficult to go from $[0,1]$ to the interval $[a,b]$% . Indeed, let $f\in C[a,b]$, $\sigma $ be constructed as above, and $% \varepsilon $ be an arbitrarily small positive number. The transformed function $g(x)=f(a+(b-a)x)$ is well defined on $[0,1]$ and we can apply the inequality~(5.39). Now using the inverse transformation $x=(t-a)/(b-a)$, we can write \begin{equation*} |f(t)-c_{1}\sigma (t-\theta _{1})-c_{2}\sigma (t-\theta _{2})|<\varepsilon \end{equation*}% for all $t\in \lbrack a,b]$, where $\theta _{1}=a+s_{1}$ and $\theta _{2}=a+s_{2}$. The last inequality completes the proof. \end{proof} Since any compact subset of the real line is contained in a segment $[a,b]$, the following generalization of Theorem 5.11 holds. \bigskip \textbf{Theorem 5.12.} \textit{Let $Q$ be a compact subset of the real line and $d$ be its diameter. Let $\lambda $ be any positive number. Then one can algorithmically construct a computable sigmoidal activation function $\sigma \colon \mathbb{R}\rightarrow \mathbb{R}$, which is infinitely differentiable, strictly increasing on $(-\infty ,d)$, $\lambda $-strictly increasing on $[d,+\infty )$, and satisfies the following property: For any $% f\in C(Q)$ and $\varepsilon >0$ there exist numbers $c_{1}$, $c_{2}$, $% \theta _{1}$ and $\theta _{2}$ such that} \begin{equation*} |f(x)-c_{1}\sigma (x-\theta _{1})-c_{2}\sigma (x-\theta _{2})|<\varepsilon \end{equation*}% \textit{for all $x\in Q$.} \bigskip \textbf{Remark 5.7.} Theorems 5.11 and 5.12 show that single hidden layer neural networks with the constructed sigmoidal activation function $\sigma $ and only two neurons in the hidden layer can approximate any continuous univariate function. Moreover, in this case, one can fix the weights equal to $1$. For the approximation of continuous multivariate functions two hidden layer neural networks with $3d+2$ hidden neurons can be taken. Namely, Theorem 5.9 (and hence Theorem 5.10) is valid with the constructed in Section 5.3.1 activation function $\sigma$. Indeed, the proof of this theorem shows that any activation function with the property (5.22) suffices. But the activation function constructed in Section 5.3.1 satisfies this property (see (5.38)). \bigskip \subsection{Numerical results} We prove in Theorem 5.11 that any continuous function on $[a,b]$ can be approximated arbitrarily well by single hidden layer neural networks with the fixed weight $1$ and with only two neurons in the hidden layer. An activation function $\sigma $ for such a network is constructed in Section 5.3.1. We have seen from the proof that our approach is totally constructive. One can evaluate the value of $\sigma $ at any point of the real axis and draw its graph instantly, using the programming interface at the URL shown at the beginning of Section 5.3.2. In the current subsection, we demonstrate our result in various examples. For different error bounds we find the parameters $c_{1}$, $c_{2}$, $\theta _{1}$ and $\theta _{2}$ in Theorem 5.11. All computations were done in SageMath~\cite{Sage}. For computations, we use the following algorithm, which works well for analytic functions. Assume $f$ is a function, whose Taylor series around the point $% (a+b)/2$ converges uniformly to $f$ on $[a,b]$, and $\varepsilon >0$. \begin{enumerate} \item Consider the function $g(t) := f(a + (b - a) t)$, which is well-defined on $[0, 1]$; \item Find $k$ such that the $k$-th Taylor polynomial \begin{equation*} T_k(x) := \sum_{i=0}^k \frac{g^{(i)}(1/2)}{i!} \left( x - \frac{1}{2} \right)^i \end{equation*} satisfies the inequality $|T_k(x) - g(x)| \le \varepsilon / 2$ for all $x \in [0, 1]$; \item Find a polynomial $p$ with rational coefficients such that \begin{equation*} |p(x) - T_k(x)| \le \frac{\varepsilon}{2}, \qquad x \in [0, 1], \end{equation*} and denote by $p_0$ the leading coefficient of this polynomial; \item If $p_0 \ne 0$, then find $n$ such that $u_n(x) = p(x) / p_0$. Otherwise, set $n := 1$; \item For $n=1$ and $n>1$ evaluate $a_{n}$ and $b_{n}$ by~(5.31) and~(5.32), respectively; \item Calculate the parameters of the network as \begin{equation*} c_1 := \frac{p_0}{b_n}, \qquad c_2 := \frac{2 p_0 a_n}{b_n (1 + h(3d))}, \qquad \theta_1 := b - 2 n (b - a), \qquad \theta_2 := 2 a - b; \end{equation*} \item Construct the network $\mathcal{N}=c_{1}\sigma (x-\theta _{1})+c_{2}\sigma (x-\theta _{2}).$ Then $\mathcal{N}$ gives an $\varepsilon $-approximation to $f.$ \end{enumerate} In the sequel, we give four practical examples. To be able to make comparisons between these examples, all the considered functions are given on the same interval $[-1,1]$. First we select the polynomial function $% f(x)=x^{3}+x^{2}-5x+3$ as a target function. We investigate the sigmoidal neural network approximation to $f(x)$. This function was considered in \cite{HH04} as well. Note that the authors of \cite{HH04} chose the sigmoidal function as \begin{equation*} \sigma (x)=% \begin{cases} 1, & \text{if }x\geq 0, \\ 0, & \text{if }x<0,% \end{cases}% \end{equation*}% and obtained the numerical results (see Table~\ref{tbl:Hahm}) for single hidden layer neural networks with $8$, $32$, $128$, $532$ neurons in the hidden layer (see also~\cite{CX10} for an additional constructive result concerning the error of approximation in this example). \begin{table}[tbp] \caption{The Heaviside function as a sigmoidal function} \label{tbl:Hahm}% \begin{tabular}{|c|c|c|} \hline $N$ & Number of neurons ($2N^2$) & Maximum error \\ \hline $2$ & $8$ & $0.666016$ \\ \hline $4$ & $32$ & $0.165262$ \\ \hline $8$ & $128$ & $0.041331$ \\ \hline $16$ & $512$ & $0.010333$ \\ \hline \end{tabular}% \end{table} As it is seen from the table, the number of neurons in the hidden layer increases as the error bound decreases in value. This phenomenon is no longer true for our sigmoidal function. Using Theorem 5.11, we can construct explicitly a single hidden layer neural network model with only two neurons in the hidden layer, which approximates the above polynomial with arbitrarily given precision. Here by \textit{explicit construction} we mean that all the network parameters can be computed directly. Namely, the calculated values of these parameters are as follows: $c_{1}\approx 2059.373597$, $c_{2}\approx -2120.974727$, $\theta _{1}=-467$, and $\theta _{2}=-3$. It turns out that for the above polynomial we have an exact representation. That is, on the interval $[-1,1]$ we have the identity \begin{equation*} x^{3}+x^{2}-5x+3\equiv c_{1}\sigma (x-\theta _{1})+c_{2}\sigma (x-\theta _{2}). \end{equation*} Let us now consider the other polynomial function \begin{equation*} f(x) = 1 + x + \frac{x^2}{2} + \frac{x^3}{6} + \frac{x^4}{24} + \frac{x^5}{% 120} + \frac{x^6}{720}. \end{equation*} For this function we do not have an exact representation as above. Nevertheless, one can easily construct a $\varepsilon$-approximating network with two neurons in the hidden layer for any sufficiently small approximation error $\varepsilon$. Table~\ref{tbl:polynomial} displays numerical computations of the network parameters for six different approximation errors. \begin{table}[tbp] \caption{Several $\protect\varepsilon$-approximators of the function $1 + x + x^2 / 2 + x^3 / 6 + x^4 / 24 + x^5 / 120 + x^6 / 720$} \label{tbl:polynomial}% \begin{tabular}{|c|c|c|l|c|c|} \hline Number of & \multicolumn{4}{|c|}{Parameters of the network} & Maximum \\ \cline{2-5} neurons & $c_1$ & $c_2$ & \multicolumn{1}{c|}{$\theta_1$} & $\theta_2$ & error \\ \hline $2$ & $2.0619 \times 10^{2}$ & $2.1131 \times 10^{2}$ & $-1979$ & $-3$ & $% 0.95$ \\ \hline $2$ & $5.9326 \times 10^{2}$ & $6.1734 \times 10^{2}$ & $-1.4260 \times 10^{8}$ & $-3$ & $0.60$ \\ \hline $2$ & $1.4853 \times 10^{3}$ & $1.5546 \times 10^{3}$ & $-4.0140 \times 10^{22}$ & $-3$ & $0.35$ \\ \hline $2$ & $5.1231 \times 10^{2}$ & $5.3283 \times 10^{2}$ & $-3.2505 \times 10^{7}$ & $-3$ & $0.10$ \\ \hline $2$ & $4.2386 \times 10^{3}$ & $4.4466 \times 10^{3}$ & $-2.0403 \times 10^{65}$ & $-3$ & $0.04$ \\ \hline $2$ & $2.8744 \times 10^{4}$ & $3.0184 \times 10^{4}$ & $-1.7353 \times 10^{442}$ & $-3$ & $0.01$ \\ \hline \end{tabular}% \end{table} \begin{figure}[tbp] \includegraphics[width=1.0\textwidth]{monic-f1} \caption{The graphs of $f(x) = 1 + x + x^2 / 2 + x^3 / 6 + x^4 / 24 + x^5 / 120 + x^6 / 720$ and some of its approximators ($\protect\lambda = 1/4$)} \label{fig:polynomial} \end{figure} At the end we consider the nonpolynomial functions $f(x) = 4x / (4 + x^2)$ and $f(x) = \sin x - x \cos(x + 1)$. Tables~\ref{tbl:nonpolynomial} and \ref% {tbl:nonpolynomial2} display all the parameters of the $\varepsilon$% -approximating neural networks for the above six approximation error bounds. As it is seen from the tables, these bounds do not alter the number of hidden neurons. Figures~\ref{fig:polynomial}, \ref{fig:nonpolynomial} and % \ref{fig:nonpolynomial2} show how graphs of some constructed networks $% \mathcal{N}$ approximate the corresponding target functions $f$. \begin{table}[tbp] \caption{Several $\protect\varepsilon$-approximators of the function $4x / (4 + x^2)$} \label{tbl:nonpolynomial}% \begin{tabular}{|@{\hspace{4pt}}c|@{\hspace{4pt}}c|c|l|c|c|} \hline Number of & \multicolumn{4}{|c|}{Parameters of the network} & Maximum \\ \cline{2-5} neurons & $c_1$ & $c_2$ & \multicolumn{1}{c|}{$\theta_1$} & $\theta_2$ & error \\ \hline $2$ & $\phantom{-}1.5965 \times 10^{2}$ & $\phantom{-}1.6454 \times 10^{2}$ & $-283$ & $-3$ & $0.95$ \\ \hline $2$ & $\phantom{-}1.5965 \times 10^{2}$ & $\phantom{-}1.6454 \times 10^{2}$ & $-283$ & $-3$ & $0.60$ \\ \hline $2$ & $-1.8579 \times 10^{3}$ & $-1.9428 \times 10^{3}$ & $-6.1840 \times 10^{11}$ & $-3$ & $0.35$ \\ \hline $2$ & $\phantom{-}1.1293 \times 10^{4}$ & $\phantom{-}1.1842 \times 10^{4}$ & $-4.6730 \times 10^{34}$ & $-3$ & $0.10$ \\ \hline $2$ & $\phantom{-}2.6746 \times 10^{4}$ & $\phantom{-}2.8074 \times 10^{4}$ & $-6.8296 \times 10^{82}$ & $-3$ & $0.04$ \\ \hline $2$ & $-3.4218 \times 10^{6}$ & $-3.5939 \times 10^{6}$ & $-2.9305 \times 10^{4885}$ & $-3$ & $0.01$ \\ \hline \end{tabular}% \end{table} \begin{table}[tbp] \caption{Several $\protect\varepsilon$-approximators of the function $\sin x - x \cos(x + 1)$} \label{tbl:nonpolynomial2}% \begin{tabular}{|c|c|c|l|c|c|} \hline Number of & \multicolumn{4}{|c|}{Parameters of the network} & Maximum \\ \cline{2-5} neurons & $c_1$ & $c_2$ & \multicolumn{1}{c|}{$\theta_1$} & $\theta_2$ & error \\ \hline $2$ & $\phantom{-}8.950 \times 10^{3}$ & $\phantom{-}9.390 \times 10^{3}$ & $% -3.591 \times 10^{53}$ & $-3$ & $0.95$ \\ \hline $2$ & $\phantom{-}3.145 \times 10^{3}$ & $\phantom{-}3.295 \times 10^{3}$ & $% -3.397 \times 10^{23}$ & $-3$ & $0.60$ \\ \hline $2$ & $\phantom{-}1.649 \times 10^{5}$ & $\phantom{-}1.732 \times 10^{5}$ & $% -9.532 \times 10^{1264}$ & $-3$ & $0.35$ \\ \hline $2$ & $-4.756 \times 10^{7}$ & $-4.995 \times 10^{7}$ & $-1.308 \times 10^{180281}$ & $-3$ & $0.10$ \\ \hline $2$ & $-1.241 \times 10^{7}$ & $-1.303 \times 10^{7}$ & $-5.813 \times 10^{61963}$ & $-3$ & $0.04$ \\ \hline $2$ & $\phantom{-}1.083 \times 10^{9}$ & $\phantom{-}1.138 \times 10^{9}$ & $% -2.620 \times 10^{5556115}$ & $-3$ & $0.01$ \\ \hline \end{tabular}% \end{table} \begin{figure}[tbp] \includegraphics[width=1.0\textwidth]{monic-f2} \caption{The graphs of $f(x) = 4x / (4 + x^2)$ and some of its approximators ($\protect\lambda = 1/4$)} \label{fig:nonpolynomial} \end{figure} \begin{figure}[tbp] \includegraphics[width=1.0\textwidth]{monic-f3} \caption{The graphs of $f(x)=\sin x-x\cos (x+1)$ and some of its approximators ($\protect\lambda =1/4$)} \label{fig:nonpolynomial2} \end{figure} \newpage \addcontentsline{toc}{chapter}{Bibliography}
{ "timestamp": "2020-09-01T02:06:22", "yymm": "2005", "arxiv_id": "2005.14125", "language": "en", "url": "https://arxiv.org/abs/2005.14125", "abstract": "These notes are about ridge functions. Recent years have witnessed a flurry of interest in these functions. Ridge functions appear in various fields and under various guises. They appear in fields as diverse as partial differential equations (where they are called plane waves), computerized tomography and statistics. These functions are also the underpinnings of many central models in neural networks.We are interested in ridge functions from the point of view of approximation theory. The basic goal in approximation theory is to approximate complicated objects by simpler objects. Among many classes of multivariate functions, linear combinations of ridge functions are a class of simpler functions. These notes study some problems of approximation of multivariate functions by linear combinations of ridge functions. We present here various properties of these functions. The questions we ask are as follows. When can a multivariate function be expressed as a linear combination of ridge functions from a certain class? When do such linear combinations represent each multivariate function? If a precise representation is not possible, can one approximate arbitrarily well? If well approximation fails, how can one compute/estimate the error of approximation, know that a best approximation exists? How can one characterize and construct best approximations? If a smooth function is a sum of arbitrarily behaved ridge functions, can it be expressed as a sum of smooth ridge functions? We also study properties of generalized ridge functions, which are very much related to linear superpositions and Kolmogorov's famous superposition theorem. These notes end with a few applications of ridge functions to the problem of approximation by single and two hidden layer neural networks with a restricted set of weights.We hope that these notes will be useful and interesting to both researchers and students.", "subjects": "Classical Analysis and ODEs (math.CA); Functional Analysis (math.FA)", "title": "Notes on ridge functions and neural networks", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9808759638081522, "lm_q2_score": 0.8354835371034369, "lm_q1q2_score": 0.8195057197021778 }
https://arxiv.org/abs/1805.12119
A combinatorial characterization of finite groups of prime exponent
The power graph of a group $G$ is a simple and undirected graph with vertex set $G$ and two distinct vertices are adjacent if one is a power of the other. In this article, we characterize (non-cyclic) finite groups of prime exponent and finite elementary abelian $2$-groups (of rank at least $2$) in terms of their power graphs.
\section{Introduction} Let $G$ be a group. The \emph{power graph} $\mathcal{G}(G)$ of $G$ is a simple and undirected graph whose vertex set is $G$ and distinct vertices $u$ and $v$ are adjacent if $v=u^n$ for some $n \in \mathbb{N}$ or $u=v^m$ for some $m \in \mathbb{N}$. The notion of (directed) power graph of a group was introduced by Kelarev and Quinn \cite{kelarev2000combinatorial, kelarevDirectedSemigr}. Afterwards, Chakrabarty et al. \cite{GhoshSensemigroups} gave the above definition of power graph of a group. Recently, many interesting results on power graphs have been obtained, for instance, see \cite{curtin2014edge,feng2015,ma2018,moghaddamfar2013} and the references therein. In the study of graphs constructed from groups, a useful topic is the extent to which structure of a group is reflected by the corresponding graph. By considering it for power graphs, Cameron \cite{Cameron} proved that if two finite groups $G_1$ and $G_2$ have isomorphic power graphs, they have same number of elements of each order (in particular, have the same spectra). Additionally, if $G_1$ and $G_2$ are abelian, it was shown by Cameron and Ghosh \cite{Ghosh} that $G_1$ and $G_2$ are isomorphic. Mirzargar et al. \cite{mirzargar2012power} proved that finite simple groups and finite cyclic groups are uniquely determined by their power graphs. In this article, we characterize non-cyclic finite groups of prime exponent (cf. Theorem \ref{min.pg.all2}) and finite elementary abelian $2$-groups of rank at least $2$ (cf. Theorem \ref{min.vercon}) in terms of connectedness of their power graphs. The terminology used throughout are standard; for instance, we can refer to \cite{Dummit} for groups and \cite{west1996graph} for graphs. For any graph $\Gamma$, its (vertex) connectivity, edge-connectivity and minimum degree are denoted by $\kappa(\Gamma)$, $\kappa'(\Gamma)$ and $\delta(\Gamma)$, respectively. Let $\Gamma$ be a non-trivial and connected graph. Then it is said to be \emph{minimally connected} if $\kappa(\Gamma-\varepsilon) = \kappa(\Gamma)-1$ for every edge $\varepsilon$ of $\Gamma$. Analogously, $\Gamma$ is said to be \emph{minimally edge-connected} if $\kappa'(\Gamma-\varepsilon)= \kappa'(\Gamma)-1$ for every edge $\varepsilon$ of $\Gamma$. These graphs have been explored in various problems in extremal and structural graph theory (cf. \cite{bollobas2004graph, kriesell2013minimal}). We employ these notions on power graphs to study the corresponding finite groups. Now we state our main results. \begin{theorem}\label{min.pg.all2} A finite group $G$ is a non-cyclic group of prime exponent if and only if $\mathcal{G}(G)$ is non-complete and minimally edge-connected. \end{theorem} \begin{theorem}\label{min.vercon} A finite group $G$ is an elementary abelian $2$-group of rank at least $2$ if and only if $\mathcal{G}(G)$ is non-complete and minimally connected. \end{theorem} It is trivial to see that a finite group $G$ is a cyclic group of prime exponent $p$ if and only if $\mathcal{G}(G)$ is a complete graph on $p$ vertices. A similar statement can be made about a finite elementary abelian $2$-group of rank $1$. \section{Preliminaries} \label{prelim} Let $G$ be a group with an element $x$. The order of $x$ in $G$ and the degree of $x$ in $\mathcal{G}(G)$ are denoted by $\mathrm{o}(x)$ and $\deg(x)$, respectively. We denote by $[x]$ the set of generators of the cyclic subgroup $\langle x \rangle$. Since an element of $G$ is a vertex of $\mathcal{G}(G)$ and vice versa, we use the terms element and vertex interchangeably. Let $\Gamma$ be a graph with vertex set $V(\Gamma)$ and edge set $E(\Gamma)$. For any subgraph $\Gamma'$ and a set of edges $S$ of $\Gamma$, we denote $\Gamma' - \{S \cap E(\Gamma')\}$ simply by $\Gamma' - S$. We next recall some necessary results on power graphs of finite groups. \begin{lemma}[\cite{GhoshSensemigroups,power2017conn,power2017mindeg}]\label{lemma1} Suppose that $G$ is a finite group. \begin{enumerate}[\rm(i)] \item $\mathcal{G}(G)$ is complete if and only if $G$ is a cyclic group of order one or prime power. \item For any induced subgraph $\Gamma$ of $\mathcal{G}(G)$ with identity vertex, $\kappa'(\Gamma)=\delta(\Gamma)$. \item Any minimum separating set of $\mathcal{G}(G)$ is of the form $\cup_{i=1}^s [x_i]$ for $x_1,\ldots, x_s$ in $G$. \end{enumerate} \end{lemma} We use Lemma \ref{lemma1}(i) without referring to it explicitly. Given any group $G$, we denote by $\mathcal{G}^*(G)$, the subgraph of $\mathcal{G}(G)$ obtained by deleting the identity vertex. \begin{lemma}[{\cite[Theorem 4]{doostabadi2015power}}]\label{p.exponent} For any finite group $G$, $\mathcal{G}^*(G)$ is regular if and only if $G$ is a cyclic group of prime power order or $G$ is of prime exponent. \end{lemma} \begin{lemma}[{\cite[Proposition 3.1]{power2017conn}}]\label{lemma3} Suppose that $G$ is a finite $p$-group for some prime $p$. If $x$ is an element of order $p$ in $G$, then $x$ is adjacent to all other vertices of the component of $\mathcal{G}^*(G)$ that contains $x$. \end{lemma} \section{Proofs of the main results} \label{proof.main.results} In this section $G$ is a non-trivial group with identity element $e$. We first prove Theorem \ref{min.pg.all2} and then Theorem \ref{min.vercon}. We begin with an essential lemma. \begin{lemma}\label{min.necessary} Let $G$ be a finite group such that $\mathcal{G}(G)$ is minimally edge-connected. \begin{enumerate}[\rm(i)] \item If $x \in G$ and $\mathrm{o}(x) > 2$, then $\deg(x) = \delta(\mathcal{G}(G))$. \item If $\langle y \rangle$ is a maximal cyclic subgroup of $G$ and $\mathrm{o}(y) > 2$, then $\mathrm{o}(y) = \delta(\mathcal{G}(G))+1$. \end{enumerate} \end{lemma} \begin{proof} (i) Note that $[x]$ has two or more elements. Let $\varepsilon$ be the edge in $\mathcal{G}(G)$ with endpoints $x$ and $x' \in [x]-\{x\}$. Then by Lemma \ref{lemma1}(ii) and the fact that $\mathcal{G}(G)$ is minimally edge-connected, we have $\delta(\mathcal{G}(G)-\varepsilon) = \delta(\mathcal{G}(G))-1$. Thus at least one endpoint of $\varepsilon$ has the minimum degree in $\mathcal{G}(G)$. Furthermore, as $\deg(x) = \deg(x')$, we get $ \deg(x) = \delta(\mathcal{G}(G))$. \smallskip \noindent (ii) Observe that $\deg(y) = \mathrm{o}(y)-1$. Hence by (i), the proof follows. \end{proof} \begin{lemma}\label{no.maximal} Let $G$ be a finite group with no maximal subgroup of order two. If $\mathcal{G}(G)$ is minimally edge-connected, then $G$ is of prime power order. \end{lemma} \begin{proof} It is known that every element of $G$ is in some of its maximal cyclic subgroups. Thus it follows from Lemma \ref{min.necessary}(ii) that the exponent of $G$ is $\delta(\mathcal{G}(G))+1$. The following arguments are similar to that of the proof of Lemma \ref{p.exponent}. We denote $d = \delta(\mathcal{G}(G))$. If possible, suppose $G$ is not of prime power order. Let $p_1 < p_2 < \ldots < p_r$, $r \geq 2$, be the prime factors of $d+1$. Then by Lemma \ref{min.necessary}(ii), $G$ has an element $x$ of order $\frac{d+1}{p_1}$. The degree of $x$ in $\mathcal{G}(G)$ is $\frac{d+1}{p_1}-1+m\phi(d+1)$, where $m$ is the number of maximal cyclic subgroups of $G$ containing $x$ and $\phi$ is the Euler's totient function. However, since $\mathrm{o}(x) > 2$, it follows from Lemma \ref{min.necessary}(i) that $\deg(x) = d$. As a result, $\frac{d+1}{p_1}+m\phi(d+1) = d+1$, which yields $m(p_2-1)(p_3-1)\ldots(p_r-1) = p_2p_3\ldots p_r$. This implies that if $q$ is a positive divisor of $p_2-1$, then $q=p_i$ for some $2 \leq i \leq r$, which is a contradiction. Hence $G$ is a group of prime power order. \end{proof} Now we provide the proof of Theorem \ref{min.pg.all2}. \begin{proof}[Proof of Theorem \ref{min.pg.all2}] Let $\mathcal{G}(G)$ be non-complete and minimally edge-connected. If $G$ has a maximal subgroup of order two, then $\kappa'(\mathcal{G}(G)) = 1$. Thus $\mathcal{G}(G)$ is a tree. Consequently, all elements of $G$ have order two, so that $G$ is an elementary abelian $2$-group of rank at least $2$. Now suppose $G$ has no maximal subgroup of order two. Then by Lemma \ref{no.maximal}, $G$ is a finite $p$-group for some prime $p$. If $p > 2$, then it follows from Lemma \ref{min.necessary}(i) that $\mathcal{G}^*(G)$ is regular. We next take $p=2$. If $\mathcal{G}(G)$ has more than one block, then the only vertex common to any two distinct blocks is $e$. Furthermore, by Lemma \ref{lemma3}, every block of $\mathcal{G}(G)$ has exactly one vertex that has order two in $G$. Note that for any $x$ in $G$, all elements of $\langle x \rangle$ are vertices in the same block of $\mathcal{G}(G)$. If possible, suppose $\langle x_1 \rangle$ and $\langle x_2 \rangle$ are distinct maximal cyclic subgroups of $G$ whose elements are vertices in the same block, say $\Gamma$, of $\mathcal{G}(G)$. Let $y$ be the vertex in $\Gamma$ having order two in $G$ and $\varepsilon$ be the edge with endpoints $e$ and $y$. Then by assumption, $\langle y \rangle$ is not a maximal subgroup in $G$. Thus $\mathcal{G}(G)-\varepsilon$ and in particular, $\Gamma-\varepsilon$ is connected. We continue to write $d =\delta(\mathcal{G}(G))$. By applying Lemma \ref{lemma3} and Lemma \ref{min.necessary}(i), we have $d \geq 3$. Suppose $S$ is a minimum disconnecting set of $\mathcal{G}(G)-\varepsilon$. In view of Lemma \ref{lemma1}(ii), $|S| = d-1$. If $\Gamma$ is the only block of $\mathcal{G}(G)$, then $\Gamma = \mathcal{G}(G)$. We observe that the minimum degree (hence the edge-connectivity) of any block of $\mathcal{G}(G)$ is $d$. Accordingly, if $\mathcal{G}(G)$ has any block $\Gamma'$ different from $\Gamma$, then $\Gamma'-S$ is connected. Thus we deduce that $(\Gamma-\varepsilon)-S$ is disconnected. Moreover, as $\kappa'(\Gamma-\varepsilon) = d-1$, all elements of $S$ are therefore edges in $\Gamma$. Since all elements of $\langle x_1 \rangle$ and $\langle x_2 \rangle$ are vertices in $\Gamma$, by applying Lemma \ref{min.necessary}(ii), we get $|V(\Gamma)| \geq d+2$ (in fact, $|V(\Gamma)| \geq d+3$). So considering Lemma \ref{lemma3}, $e$ is connected to $y$ by a path of length two in $(\Gamma-\varepsilon)-S$. Because $(\Gamma-\varepsilon)-S$ is disconnected, there exists a vertex $z$ in $\Gamma$, different from $e$ and $y$, that is not connected to $e$ by any path in $(\Gamma-\varepsilon)-S$. As a result, $S$ contains the edge with endpoints $e$ and $z$ as well as the edge with endpoints $y$ and $z$. Since $\deg(z) = d$ and $|S|=d-1$, there exists a path of length two between $e$ and $z$ in $(\Gamma-\varepsilon)-S$. However, this contradicts the initial assumption about $z$. We thus conclude that every block of $\mathcal{G}(G)$ is a clique induced by a maximal cyclic subgroup of $G$. Consequently, in view of Lemma \ref{min.necessary}(ii), $\mathcal{G}^*(G)$ is regular. Additionally, $\mathcal{G}(G)$ is non-complete. Hence it follows from Lemma \ref{p.exponent} that $G$ is a non-cyclic group of prime exponent. Conversely, let $G$ be a non-cyclic group of prime exponent $p$. Then two non-identity vertices $u$ and $v$ are adjacent in $\mathcal{G}(G)$ if and only if $\langle u \rangle = \langle v \rangle$. Thus $\mathcal{G}(G)$ is union of finitely many maximal cliques of size $p$ with any two distinct maximal cliques having $e$ as the only common vertex. Therefore, $\mathcal{G}(G)$ is non-complete and minimally edge-connected. \end{proof} We next give a simple lemma and then prove Theorem \ref{min.vercon}. \begin{lemma}\label{min.sepset.endpts} Let $\Gamma$ be a graph with an edge $\varepsilon$ such that $\Gamma-\varepsilon$ is connected. If $\kappa(\Gamma - \varepsilon) = \kappa(\Gamma)-1$, then no minimum separating set of $\Gamma - \varepsilon$ contains endpoints of $\varepsilon$. \end{lemma} \begin{proof} Suppose $u$ and $v$ are the endpoints of $\varepsilon$. If possible, let $S$ be a minimum separating set of $\Gamma - \varepsilon$ containing at least one of $u$ or $v$. We have $(\Gamma - \varepsilon)-S = \Gamma-S$, so that $S$ is a separating set of $\Gamma$. This implies $\kappa(\Gamma) \leq \kappa(\Gamma - \varepsilon)$, which contradicts the given condition. This proves the lemma. \end{proof} \begin{proof}[Proof of Theorem \ref{min.vercon}] Let $\mathcal{G}(G)$ be non-complete and minimally connected. We notice that $|G| \geq 4$. Let $x$ be a non-identity element of $G$ and if possible, let there exist $y \neq x$ such that $\langle x \rangle = \langle y \rangle$. Suppose $\varepsilon$ is the edge with endpoints $x$ and $y$. Since $x$ and $y$ are connected by the path $x,e,y$, $\mathcal{G}(G)-\varepsilon$ is connected. By using the fact that $\mathcal{G}(G)$ is minimally connected and $\kappa(\mathcal{G}(G)) \leq |G|-2$, we have $\kappa(\mathcal{G}(G)-\varepsilon) \leq |G|-3$. Let $S$ be a minimum separating set of $\mathcal{G}(G)-\varepsilon$. Then $\mathcal{G}(G)-S$ is a connected graph with three or more vertices, and by Lemma \ref{min.sepset.endpts}, $x,y \notin S$. We thus deduce that there is no path in $\mathcal{G}(G)-S$ that has vertices $x$ and $y$, and does not have the edge $\varepsilon$. Let $z$ be a vertex different from $x$ and $y$ in $\mathcal{G}(G)-S$. Then in $\mathcal{G}(G)-S$, $z$ is connected to exactly one of $x$ and $y$ by a path that does not have $\varepsilon$. Without loss of generality, let $z$ be connected to $x$ by a path that does not have $\varepsilon$ in $\mathcal{G}(G)-S$. Then $S \cup \{x\}$ is a separating set of $\mathcal{G}(G)$. Since $\kappa(\mathcal{G}(G)-\varepsilon)=\kappa(\mathcal{G}(G))-1$, we conclude that $S \cup \{x\}$ is a minimum separating set of $\mathcal{G}(G)$. Then by Lemma \ref{lemma1}(iii), $[x] \subseteq S \cup \{x\}$, which contradicts the fact that $y \notin S$. Accordingly, we get $[x]=\{x\}$. Because $|[x]|=\phi(\mathrm{o}(x))$ and $x \neq e$, we have $\mathrm{o}(x)=2$. Hence $G$ is an elementary abelian $2$-group. Additionally, as $\mathcal{G}(G)$ is non-complete, the rank of $G$ is at least $2$. For converse, suppose $G$ is an elementary abelian $2$-group of rank at least $2$. Then $\mathcal{G}(G)$ is a star on three or more vertices. As a result, $\kappa(\mathcal{G}(G))=1$ and $\mathcal{G}(G)-\varepsilon$ is disconnected for every edge $\varepsilon$ of $\mathcal{G}(G)$. Therefore we conclude that $\mathcal{G}(G)$ is non-complete and minimally connected. \end{proof} \section*{Acknowledgement} The author is thankful to Dr. K. V. Krishna for his constructive suggestions. This research work was partially supported by the fellowship of IIT Guwahati, India.
{ "timestamp": "2019-03-20T01:26:12", "yymm": "1805", "arxiv_id": "1805.12119", "language": "en", "url": "https://arxiv.org/abs/1805.12119", "abstract": "The power graph of a group $G$ is a simple and undirected graph with vertex set $G$ and two distinct vertices are adjacent if one is a power of the other. In this article, we characterize (non-cyclic) finite groups of prime exponent and finite elementary abelian $2$-groups (of rank at least $2$) in terms of their power graphs.", "subjects": "Combinatorics (math.CO)", "title": "A combinatorial characterization of finite groups of prime exponent", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9833429624315188, "lm_q2_score": 0.8333246015211009, "lm_q1q2_score": 0.8194438823268242 }
https://arxiv.org/abs/1911.08067
Sets in $\mathbb{R}^d$ determining $k$ taxicab distances
We address an analog of a problem introduced by Erdős and Fishburn, itself an inverse formulation of the famous Erdős distance problem, in which the usual Euclidean distance is replaced with the metric induced by the $\ell^1$-norm, commonly referred to as the $\textit{taxicab metric}$. Specifically, we investigate the following question: given $d,k\in \mathbb{N}$, what is the maximum size of a subset of $\mathbb{R}^d$ that determines at most $k$ distinct taxicab distances, and can all such optimal arrangements be classified? We completely resolve the question in dimension $d=2$, as well as the $k=1$ case in dimension $d=3$, and we also provide a full resolution in the general case under an additional hypothesis.
\section{Introduction} In 1946, Erd\H{o}s \cite{Erdos} asked a now famous question: given $n\in \mathbb{N}$, what is the minimum number of distinct distances determined by $n$ points in a plane? Denoting this minimum by $f(n)$, he proved via an elementary counting argument that $f(n)=\Omega(\sqrt{n})$, and he conjectured that the correct order of growth is $n/\sqrt{\log n}$, as attained by a $\sqrt{n}\times\sqrt{n}$ integer grid. After decades of incremental progress, this conjecture was effectively resolved in a celebrated result of Guth and Katz \cite{GuthKatz}, who established that $f(n)=\Omega(n/\log n)$. 50 years after Erd\H{o}s's original paper, Erd\H{o}s and Fishburn \cite{EF} addressed the same question from the inverse perspective, and aspired to precise results in fixed cases rather than general asymptotic results. Specifically, they investigated the following: given $k\in \mathbb{N}$, what is the maximum number of points in a plane that determine at most $k$ distinct distances, and can such optimal arrangements be classified? This question, which we refer to as the \textit{Erd\H{o}s-Fishburn problem}, was fully resolved by Erd\H{o}s and Fishburn for $1\leq k \leq 4$, then by Shinahara \cite{Shin} for $k=5$, and Wei \cite{Wei} for $k=6$, while it remains open for $k\geq 7.$ By convention, in the quoted results and throughout this paper, $0$ is not counted as a distance determined by a set of points. These questions can also be adapted to higher dimensions, and to alternative notions of distance. Here we focus on a particular, well-known alternative metric. \begin{definition} For $d\in \mathbb{N}$ and $x=(x_1,\dots,x_d)\in \R^d$, we define the \textit{$\ell^1$-norm} of $x$ by $$\norm{x}_1=|x_1|+\cdots+|x_d|,$$ which in particular satisfies the \textit{triangle inequality} $\norm{x+y}_1\leq \norm{x}_1+\norm{y}_1$. Like every norm, the $\ell^1$-norm induces a metric on $\R^d$ by defining $\norm{x-y}_1$ to be the \textit{$\ell^1$-distance} between $x,y\in \R^d.$ \end{definition} The metric induced by the $\ell^1$-norm is commonly referred to as the \textit{taxicab metric}, because it measures the length of the shortest path between two points in space, under the restriction that one can only travel in directions parallel to the coordinate axes, as if in a taxicab on a grid of city streets. For example, if two people at city intersections are separated by 3 blocks horizontally and 4 blocks vertically, then, as the crow flies, they are 5 blocks apart by the Pythagorean theorem. However, to actually make the journey without cutting through buildings, they must walk 7 blocks, which is the $\ell^1$-distance. As noted in Chapters 0 and 1 of \cite{ErdosBook}, one can show that the minimum number of $\ell^1$-distances determined by $n$ points in $\R^d$ is $\Omega(n^{1/d})$, and this order of growth is attained by $\{1,2,3,\dots,\lceil n^{1/d}\rceil\}^d$. Therefore, in the case of the taxicab metric, the big picture asymptotic question is immediately resolved, which begs the question of whether this case can be analyzed more precisely. To begin this journey, we first consider the Erd\H{o}s-Fishburn problem in the plane with $k=1$. We fix two points $U,V\in \R^2$, say $U=(-1,0)$ and $V=(1,0)$. With the usual notion of distance, if any additional points can be added without determining an additional distance, those points necessarily lie on the circles of radius $2$ centered at $U$ and $V$, respectively. Those two circles intersect in only two points, and we find that they are $Q=(0,\sqrt{3})$ and $R=(0,-\sqrt{3})$. Since the distance between $Q$ and $R$ is greater than $2$, only one of the two can be added to $\{U,V\}$ while maintaining only a single distance. In summary, a set $P\subseteq R^2$ determining a single distance satisfies $|P|\leq 3$, and equality holds if and only if $P$ is the set of vertices of an equilateral triangle. However, even in this simplest case, the taxicab metric case diverges from that of the usual distance. With the taxicab metric, the ``circle" (which we refer to as an \textit{$\ell^1$-circle}) of radius $2$ centered at $U$ is in fact a square, rotated $45^{\circ}$ from axis-parallel, with the four sides connecting the points $(-3,0)$, $(-1,2)$, $(1,0)$, and $(-1,-2)$. Similarly, the $\ell^1$-circle of radius $2$ centered at $V$ is a square with sides connecting $(3,0)$, $(1,-2)$, $(-1,0)$, and $(1,2)$. Like the usual distance case, these two circles intersect in exactly two points, this time $Q=(0,1)$ and $R=(0,-1)$. The difference is that here $Q$ and $R$ are indeed separated by $\ell^1$-distance $2$, and hence the four-point configuration $\{U,V,Q,R\}$ determines a single $\ell^1$-distance. \section{Main Definition and Results} \label{mainres} Inspired by the four-point construction above, as well as additional trial and error, we define the following family of sets, which serve as our candidates for resolving the Erd\H{o}s-Fishburn problem for the taxicab metric. \begin{definition}For integers $d>0$ and $k\geq 0$, we define $$\Lambda_d(k)=\left\{ n=(n_1,n_2,\dots,n_d)\in \mathbb{Z}^d: \norm{n}_1\leq k, \ n_1+\cdots+n_d \equiv k \ (\text{mod }2)\right\}. $$ \end{definition} \noindent $\Lambda_d(k)$ is the union of the integer lattice points lying on the $\ell^1$-spheres (which in dimension $d$ are $2^d$-faced polytopes) centered at the origin of \textit{every other} integer radius, starting with either $0$ or $1$ depending on the parity of $k$. The four-point configuration discussed in the introduction is $\Lambda_2(1)$, and some additional examples are pictured below. \begin{figure}[H] \centering \begin{subfigure}{.5\textwidth} \centering \includegraphics[width=.5\linewidth]{Lambda22.pdf} \caption{$\Lambda_2(2)$: $9$ points in $\R^2$ determining two $\ell^1$-distances} \label{fig:sub1} \end{subfigure}% \begin{subfigure}{.5\textwidth} \centering \includegraphics[width=.5\linewidth]{Lambda32.pdf} \caption{$\Lambda_3(2)$: $19$ points in $\R^3$ determining two $\ell^1$-distances} \label{fig:sub2} \end{subfigure} \caption{}\label{fig:test} \end{figure} \noindent In Section \ref{propsec}, we establish the following properties of $\Lambda_d(k)$, including the crucial fact that it determines exactly $k$ distinct $\ell^1$-distances, the primary motivation for its definition. \begin{theorem} \label{prop} The following hold for all $d,k\in \mathbb{N}$: \begin{enumerate}[(i)] \item \label{2k} $\Lambda_d(k)$ determines exactly $k$ distinct $\ell^1$-distances, specifically $2,4,\dots, 2k$ \\ \item \label{start} $|\Lambda_1(k)|=k+1$, $|\Lambda_d(1)|=2d$ \\ \item \label{recur} $\displaystyle{|\Lambda_{d+1}(k)|=|\Lambda_{d}(k)|+2\sum_{j=0}^{k-1} |\Lambda_{d}(j)|}$ \end{enumerate} \end{theorem} \noindent Parts (\ref{start}) and (\ref{recur}) of Theorem \ref{prop}, combined with known formulas for sums of powers, allow one to determine explicit formulas for $|\Lambda_d(k)|$ for any fixed $d\in \mathbb{N}$. We include the first few examples in the following table: \begin{center} \begin{table}[H] \caption{Explicit Formulas for $|\Lambda_d(k)|$} \renewcommand{\arraystretch}{1.3} \begin{tabular}{||c | c ||} \hline $d$ & $|\Lambda_d(k)|$ \\ \hline\hline $2$ & $(k+1)^2$\\ \hline $3$ & $\frac{2}{3}(k+1)^3+\frac{1}{3}(k+1)$\\ \hline $4$ & $\frac{1}{3}(k+1)^4+\frac{2}{3}(k+1)^2$ \\ \hline $5$ & $\frac{2}{15}(k+1)^5+\frac{2}{3}(k+1)^3+\frac{1}{5}(k+1) $ \\ \hline $6$ & $ \frac{2}{45}(k+1)^6+\frac{4}{9}(k+1)^4+\frac{23}{45}(k+1)^2 $ \\ \hline $7$ & $ \frac{4}{315}(k+1)^7+\frac{2}{9}(k+1)^5 + \frac{28}{45}(k+1)^3 + \frac{1}{7}(k+1) $\\ \hline \end{tabular} \label{one} \end{table} \end{center} \noindent Some of the patterns observed in Table \ref{one} can be generalized using Faulhaber's Formula for sums of powers, as seen in the following formulation, which we also prove in Section \ref{propsec}. \begin{theorem}\label{ffthm} For each $d\in \mathbb{N}$ and each integer $k\geq 0$, we have the formula \begin{equation*}|\Lambda_d(k)|=\sum_{i=0}^{\lceil d/2 \rceil-1} a_{d,i}(k+1)^{d-2i},\end{equation*} where the coefficients $a_{d,i}$ satisfy the recursive formula \begin{equation*} a_{d,i}=2\sum_{\ell=0}^i \frac{a_{d-1,\ell}}{d-2\ell} {d-2\ell \choose 2(i-\ell)}B_{2(i-\ell)}, \end{equation*} where $B_i$ is the $i$-th Bernoulli number. In particular, we have the explicit formulas $a_{d,0}=2^{d-1}/d! $ for all $d\in \mathbb{N}$ and $a_{d,1}=2^{d-3}/(3(d-3)!)$ for all $d\geq 3$. \end{theorem} \noindent Detailed analysis of $\Lambda_d(k)$ is perhaps of independent interest, but to make headway toward our goal, we need to address the important questions: does $\Lambda_d(k)$ have maximal size amongst subsets of $\R^d$ determining at most $k$ distinct $\ell^1$-distances? If so, is $\Lambda_d(k)$ the only such optimal arrangement? In anticipation of the latter question, we observe that any optimal arrangement can undergo any scaling, or any transformation that preserves the $\ell^1$-norm, and remain optimal, leading to the following definition. \begin{definition} \label{simdef} For $d\in \mathbb{N}$, we say that two subsets of $\R^d$ are \textit{$\ell^1$-similar} if one can be mapped to the other via a composition of translations, reflections about coordinate hyperplanes, dilations, and coordinate permutations, as these transformations either preserve or uniformly scale collections of $\ell^1$-distances. \end{definition} \noindent We note that the list of transformations in Definition \ref{simdef} does not include rotations, because, unlike the usual Euclidean metric, the taxicab metric is \textit{not} invariant under rotation, unless the rotation can alternatively be obtained through reflection about coordinate hyperplanes and permutation of coordinates. This fact rears its head in our exploration of the taxicab metric in higher dimensions, and plays a key role in our discussions in Section \ref{hdsec}. For now, though, the following result established in Section \ref{2dsec} completely resolves the taxicab analog of the Erd\H{o}s-Fishburn problem in the plane. \begin{theorem} \label{opt} If $k\in \mathbb{N}$ and $P\subseteq \R^2$ determines at most $k$ distinct $\ell^1$-distances, then $|P|\leq (k+1)^2$. Further, $|P|=(k+1)^2$ if and only if $P$ is $\ell^1$-similar to $\Lambda_2(k)$. \end{theorem} \noindent As we discuss in Section \ref{2dsec}, the $d=2$ case is simplified by the fact that, for the purposes of analyzing distance sets, the $\ell^1$-norm in $\R^2$ is effectively the same as the $\ell^{\infty}$-norm defined by $\norm{(x,y)}_{\infty}=\max\{|x|,|y|\}$. However, this equivalence does not persist in dimension $d\geq 3$, and for this reason, our proof strategy does not immediately generalize to higher dimensions. (Although, for the interested reader, the proof does generalize to show that if $P\subseteq \R^d$ determines at most $k$ distinct $\ell^{\infty}$-distances, then $|P|\leq (k+1)^d$, and equality holds if and only if $P$ is $\ell^1$-similar to $\{0,1,2,\dots,k\}^d$.) \noindent With considerable additional effort, we successfully get our foot into the higher-dimensional door in Section \ref{3dsec}, which assures us that the unique optimality of $\Lambda_d(k)$ is not completely dependent on a connection to the $\ell^{\infty}$-norm. \begin{theorem}\label{singdist} If $P\subseteq \R^3$ determines a single $\ell^1$-distance, then $|P|\leq 6$. Further, $|P|=6$ if and only if $P$ is $\ell^1$-similar to $\Lambda_3(1)$. \end{theorem} \noindent \textbf{Remark on previous work for $k=1$.} After the initial posting of this paper to the arxiv server, we were alerted to previous work done in the $k=1$ case (referred to as \textit{equilateral sets}) in a variety of metric spaces, including $\R^d$ with the taxicab metric (referred to as \textit{rectilinear space}). Specifically, Theorem \ref{singdist} above follows from Corollary 4.2 of \cite{Band}, due to Bandelt, Chepoi, and Laurent, while Koolen, Laurent, and Schrijver \cite{Koolen} showed that if $P\subseteq \R^4$ determines a single $\ell^1$-distance, then $|P|\leq 8=|\Lambda_4(1)|$. This partially settles a question of Kusner (see Problem 0 in \cite{Guy}), who asked if $|P|\leq 2d=|\Lambda_d(1)|$ holds for subsets of $\R^d$ determining a single $\ell^1$-distance, and this remains open for $d\geq 5$. Conjecture \ref{optcon} below can be thought of as a precise, multi-distance generalization of Kusner's question. While the conclusion of Theorem \ref{singdist} was known previously, we believe our alternative, elementary proof given in Section \ref{3dsec} remains of interest. \noindent In Section \ref{hdsec}, we explore the question of what additional hypotheses are required to prove the optimality of $\Lambda_3(k)$ for all $k\in \mathbb{N}$, or even $\Lambda_d(k)$ in full generality. We find that the proof of Theorem \ref{opt} can be fully adapted with a seemingly mild additional assumption, leading us to make the following general conjecture. \begin{conjecture} \label{optcon} If $d,k\in \mathbb{N}$ and $P\subseteq \R^d$ determines at most $k$ distinct $\ell^1$-distances, then $|P|\leq |\Lambda_d(k)|$. Further, $|P|=|\Lambda_d(k)|$ if and only if $P$ is $\ell^1$-similar to $\Lambda_d(k)$. \end{conjecture} \section{Properties of $\Lambda_d(k)$: Proof of Theorems \ref{prop} and \ref{ffthm}} \label{propsec} We begin this section by proving the essential properties of $\Lambda_d(k)$ that make it a worthy candidate for resolving the Erd\H{o}s-Fishburn problem for the taxicab metric. \subsection{Proof of Theorem \ref{prop}} Fix $k\in \mathbb{N}$. For (\ref{2k}), fix $d\in \mathbb{N}$, note that by definition of $\Lambda_d(k)$, we have $\norm{n}_1\leq k$ for all $n\in \Lambda_d(k)$. In particular, for any $n,m\in \Lambda_d(k)$, we have by the triangle inequality that $$\norm{n-m}_1 \leq \norm{n}_1+\norm{m}_1 \leq k+k=2k. $$ \noindent Further, $\norm{n-m}_1=|n_1-m_1|+\cdots+|n_d-m_d|$ is certainly an integer, and by definition of $\Lambda_d(k)$, and the fact that an integer is congruent to its absolute value modulo 2, we have \begin{align*} |n_1-m_1|+\cdots+|n_d-m_d| & \equiv n_1-m_1+\cdots +n_d-m_d \\ & \equiv (n_1+\cdots+n_d)-(m_1+\cdots+m_d) \\ & \equiv k-k \\& \equiv 0 \ (\text{mod }2). \end{align*} Therefore, the only possible nonzero values of $\norm{n-m}_1$ are $2,4,\dots, 2k$, and for each $1\leq j \leq k$ the distance $2j$ is attained between the points $(j,0,\dots,0)$ and $(-j,0,\dots,0)$ if $j\equiv k \ (\text{mod }2)$, or between $(j,1,\dots,0)$ and $(-j,1,\dots,0)$ if $j \not \equiv k \ (\text{mod }2)$. \noindent For (\ref{start}), we first see that $$\Lambda_1(k)=\begin{cases}\{-k,-k+2,\dots,-1,1,\dots,k-2,k\} & \text{ if }k \text{ is odd} \\ \{-k, -k+2,\dots, -2, 0, 2,\dots, k-2, k\} & \text{ if }k \text{ is even}\end{cases}.$$ In particular, $|\Lambda_1(k)|=2\lceil k/2 \rceil=k+1$ if $k$ is odd and $|\Lambda_1(k)|=2(k/2)+1=k+1$ if $k$ is even. Secondly, we see that $\Lambda_d(1)$ is precisely $\{\pm e_i: 1\leq i \leq d\}$, where $\{e_i\}$ is the standard basis for $\R^d$. \noindent For (\ref{recur}), we see that the possible values of the final coordinate for elements of $\Lambda_{d+1}(k)$ are integers satisfying $-k\leq x_{d+1} \leq k$. Further, for a fixed value $x_{d+1}=c$, the intersection of this hyperplane with $\Lambda_{d+1}(k)$ is $$\left\{(n_1,\dots,n_{d},c)\in \mathbb{Z}^{d+1}: |n_1|+\cdots+|n_{d}|\leq k-|c|, n_1+\cdots+n_{d}\equiv k-c\equiv k-|c| \ (\text{mod }2) \right\}, $$ which is in natural bijection with $\Lambda_{d}(k-|c|)$. Therefore, $$|\Lambda_{d+1}(k)|=\sum_{c=-k}^k |\Lambda_{d}(k-|c|)|=|\Lambda_{d}(k)|+2\sum_{j=0}^{k-1} |\Lambda_{d}(j)|. $$\qed \noindent We continue by establishing a detailed formula for $|\Lambda_d(k)|$, which in particular guarantees that it has the correct order of magnitude $\Omega(k^d)$. \subsection{Proof of Theorem \ref{ffthm}} We first note that by Theorem \ref{prop}(ii), we have $|\Lambda_1(k)|=k+1$ for all $k\geq 0$. We now fix $d\geq 2$, let $h=\lceil d/2\rceil -1$, and make the inductive hypothesis that \begin{equation} \label{indhyp} |\Lambda_{d-1}(k)|=a_{d-1,0}(k+1)^{d-1}+a_{d-1,1}(k+1)^{d-3}+\cdots+a_{d-1,h}(k+1)^{d-1-2h} \end{equation} for all $k\geq 0$. Faulhaber's formula gives \begin{equation}\label{ff} F_{p}(n)=\sum_{j=1}^n j^p=\frac{n^{p+1}}{p+1}+\frac{n^p}{2}+\frac{1}{p+1}\sum_{i=0}^{p-1}{p+1 \choose i}B_{p+1-i} n^i, \end{equation} for all $n,p\in \mathbb{N}$, where $B_i$ is the $i$-th Bernoulli number. By Theorem \ref{prop}(iii), we have \begin{equation} \label{recur2}|\Lambda_d(k)|=2\sum_{j=0}^{k-1}|\Lambda_{d-1}(j)|+|\Lambda_{d-1}(k)|=2\sum_{j=0}^{k}|\Lambda_{d-1}(j)|-|\Lambda_{d-1}(k)|, \end{equation} which combines with (\ref{indhyp}) to yield \begin{align*} |\Lambda_d(k)| &=2\left(a_{d-1,0}\sum_{j=0}^k (j+1)^{d-1}+\cdots+a_{d-1,h}\sum_{j=0}^k (j+1)^{d-1-2h}\right)-|\Lambda_{d-1}(k)| \\ &=2\left(a_{d-1,0}\sum_{j=1}^{k+1} j^{d-1}+\cdots+a_{d-1,h}\sum_{j=1}^{k+1} j^{d-1-2h}\right)-|\Lambda_{d-1}(k)| \\ &=2\left(a_{d-1,0}F_{d-1}(k+1)+\cdots+a_{d-1,h}F_{d-1-2h}(k+1)\right)-|\Lambda_{d-1}(k)|.\end{align*} This tells us that we can indeed write $|\Lambda_d(k)|$ as a polynomial in $k+1$, but we wish to establish the claimed explicit and recursive formulas for the coefficients, as well as the fact that every other coefficient is zero. First we consider the $(k+1)^d$ coefficient, which only arises from the term $2a_{d-1,0}F_{d-1}(k+1)$. Since the $n^{p+1}$ coefficient of $F_{p}(n)$ is $1/(p+1)$, we have $a_{d,0}=2a_{d-1,0}/d$. Using the base case $a_{1,0}=1$, we have by induction that $a_{d,0}=2^{d-1}/d!$, as claimed. \noindent Next we consider the $(k+1)^{d-1}$ coefficient, which arises from two sources: the $(k+1)^{d-1}$ coefficients of $2a_{d-1,0}F_{d-1}(k+1)$ and $-|\Lambda_{d-1}(k)|$, respectively. The former is $2a_{d-1,0}(1/2)=a_{d-1,0}$, while the latter is $-a_{d-1,0}$, which means that the $(k+1)^{d-1}$ coefficient of $|\Lambda_d(k)|$ is indeed $0$. More generally, for other coefficients corresponding to terms of the form $(k+1)^{d-1-2i}$, we use the following three facts: the $(k+1)^{d-1-2i}$ coefficient on $2a_{d-1,i}F_{d-1-2i}(k+1)=a_{d-1,i}$ by the same logic as above, the $(k+1)^{d-1-2i}$ coefficient of $-|\Lambda_{d-1}(k)|$ is $-a_{d-1,i}$, and the $(k+1)^{d-1-2i}$ coefficient of $F_{d-1-2\ell}(k+1)$ is $0$ for all $\ell<i$, because $B_n=0$ for all odd $n\geq 3$. Therefore, all $(k+1)^{d-1-2i}$ coefficients of $|\Lambda_d(k)|$ are $0$. \noindent For the the $(k+1)^{d-2}$ coefficient, we begin by noting that a direct calculation using (\ref{ff}) and (\ref{recur2}) yields $|\Lambda_3(k)|=\frac{2}{3}(k+1)^3+\frac{1}{3}(k+1)$, hence $a_{3,1}=1/3$, which serves as the base case for another induction. Fixing $d\geq 4$ and assuming the claimed formula $a_{d-1,1}=2^{d-4}/(3(d-4)!)$ holds, the $(k+1)^{d-2}$ coefficient of $|\Lambda_d(k)|$ is formed by two contributions, from $2a_{d-1,0}F_{d-1}(k+1)$ and $2a_{d-1,1}F_{d-3}(k+1)$, respectively. \noindent The former is given by $$2a_{d-1,0}\left(\frac{1}{d}\right){d \choose d-2}B_2=2\cdot\frac{2^{d-2}}{(d-1)!}\cdot\frac{1}{d}\cdot\frac{d(d-1)}{2}\cdot\frac{1}{6}=\frac{2^{d-3}}{3(d-2)!},$$ while the latter is given by $$2a_{d-1,1}\cdot\frac{1}{d-2}=2\cdot\frac{2^{d-4}}{3(d-4)!}\cdot\frac{1}{d-2}=\frac{2^{d-3}}{3(d-4)!(d-2)}.$$ Therefore, we have $$a_{d,1}=\frac{2^{d-3}}{3(d-2)!}+\frac{2^{d-3}}{3(d-4)!(d-2)}=\frac{2^{d-3}+(d-3)2^{d-3}}{3(d-2)!}=\frac{2^{d-3}(d-2)}{3(d-2)!}=\frac{2^{d-3}}{3(d-3)!},$$ as claimed. \noindent More generally, by (\ref{ff}) and (\ref{recur2}), we see that the $(k+1)^{d-2i}$ coefficient of $|\Lambda_d(k)|$ receives a contribution from $2a_{d-1,\ell}F_{d-1-2\ell}(k+1)$ for each $0\leq \ell \leq i$. Specifically, that contribution is $$2a_{d-1,\ell}\cdot \frac{1}{d-2\ell} \cdot {d-2\ell \choose d-2i} \cdot B_{2i-2\ell}=\frac{2a_{d-1,\ell}}{d-2\ell}{d-2\ell \choose 2(i-\ell)} B_{2(i-\ell)}, $$ and the recursive formula for $a_{d,i}$ follows. \qed \section{Optimality in Two Dimensions: Proof of Theorem \ref{opt}} \label{2dsec} In this section, we prove the unique optimality of $\Lambda_2(k)$, in that it is the unique subset of $\R^2$, up to $\ell^1$-similarity, of maximal size amongst sets determining at most $k$ distinct $\ell^1$-distances. As referenced in Section \ref{mainres}, the proof is in part enabled by an equivalence between the $\ell^1$-norm and the $\ell^{\infty}$-norm on $\R^2$. We frame our discussion entirely in the context of the $\ell^1$-norm, but the connection is implicit in our proof, particularly the following lemma. \begin{lemma} \label{linfty} Let $v_1=(1,1)$ and $v_2=(-1,1)$. If $x \in \R^2$ with $x=c_1v_1+c_2v_2$, then $$\norm{x}_1=2\max\{|c_1|, |c_2|\}. $$ \end{lemma} \begin{proof} Let $v_1=(1,1)$, $v_2=(-1,1)$, fix $x\in \R^2$, and write $x$ uniquely as $x=c_1v_1+c_2v_2=(c_1-c_2,c_1+c_2)$. By potentially reflecting over the diagonal $x_1=x_2$ and/or replacing $x$ by $-x$, both of which preserve the $\ell_1$-norm, we can assume without loss of generality that $|c_1|\geq |c_2|$ and $c_1 \geq 0$. In this case, $$\norm{x}_1=|c_1-c_2|+|c_1+c_2|=c_1-c_2+c_1+c_2=2c_1=2\max\{|c_1|,|c_2|\}.$$\end{proof} \noindent \textbf{Remark.} As an alternative approach to Section 4, one could treat the connection between the $\ell^1$ and $\ell^{\infty}$ norms on $\R^2$ in a more explicit way. Namely, Lemma \ref{linfty} can be reframed as the statement that the map $f: (\R^2,\norm{\cdot}_1)\to (\R^2,\norm{\cdot}_{\infty})$ defined by $f(x,y)=(x+y,x-y)$ is a linear isomorphism satisfying $\norm{(x,y)}_1=\norm{f(x,y)}_{\infty}$. Therefore, any results results related to the $\ell^{\infty}$-norm can be immediately transferred to the $\ell^1$-norm via this isomorphism. In particular, the proofs that follow could be rewritten in a slightly cleaner way in the $\ell^{\infty}$ context. However, in order to maintain our hands-on approach with the taxicab metric, we have chosen to leave the proofs in their $\ell^1$ form.\\ \noindent Our main strategy for proving Theorem \ref{opt} is inspired by Erd\H{o}s and Fishburn \cite{EF}. Specifically, we suppose that $P \subseteq \R^2$ determines at most $k$ distinct $\ell^1$-distances, and we seek an upper bound on the number of points we must remove from $P$ in order to eliminate the largest $\ell^1$-distance, hence reducing to the case of $k-1$ distinct $\ell^1$-distances and allowing us to invoke an inductive hypothesis. The following sequence of lemmas formalizes this strategy. Here we define an \textit{$\ell^1$-ball} in the expected way, as the region bounded by an $\ell^1$-sphere, which for $d=2$ is an $\ell^1$-circle. \ \begin{lemma} \label{SD} Suppose $P\subseteq \R^2$ is finite. If $D$ is the largest $\ell^1$-distance determined by $P$, then $P$ is contained in a closed $\ell^1$-ball of diameter $D$. \end{lemma} \begin{proof} Suppose $P \subseteq \R^2$ is finite. Let $v_1 =(1,1)$ and $v_2 = (-1,1)$. Since $\{v_1,v_2\}$ forms a basis for $\R^2$, every $x \in P$ can be written uniquely as $x= c_1 v_1 + c_2 v_2$. Choose $x_1,x_2,x_3,x_4\in P$ such that $x_1$ maximizes $c_1$, $x_2$ minimizes $c_1$, $x_3$ maximizes $c_2$, and $x_4$ minimizes $c_2$. Call these values $c_{1,\max}$, $c_{1,\min}$, $c_{2,\max}$, and $c_{2,\min}$, respectively. These choices contain $P$ inside of a rectangle $R$, rotated $45^{\circ}$ from axis parallel, determined by the inequalities $c_{1,\min}\leq c_1 \leq c_{1, \max}$ and $c_{2,\min} \leq c_2 \leq c_{2,\max}$. \noindent Let $w_1 = c_{1, \max} - c_{1, \min}$ and $w_2 = c_{2, \max} - c_{2, \min}$, and assume without loss of generality that $w_1 \geq w_2$. By Lemma \ref{linfty}, we have that $\norm{x_1-x_2}_1=2w_1$ and $\norm{p_1-p_2}_1 \leq 2w_1$ for all $p_1,p_2\in R$, so $D=2w_1$ is the largest $\ell_1$-distance determined by $P$. Let $c_{2,\text{new}} = c_{2, \max} - w_1\leq c_{2,\min}$, and let $B\supseteq R \supseteq P$ be defined by the inequalities $c_{1,\min}\leq c_1 \leq c_{1, \max}$ and $c_{2,\text{new}} \leq c_2 \leq c_{2,\max}$. $B$ is a square rotated $45^{\circ}$ from axis parallel, or in other words a closed $\ell^1$-ball, of diameter $D$, as required. \end{proof} \ \begin{lemma} \label{remove} If $P\subseteq \R^2$ is contained in a closed $\ell^1$-ball $B$ of diameter $D$, then the $\ell^1$-distance $D$ can be eliminated from $P$ by removing the points of P contained in any two adjacent sides of the boundary of $B$. \end{lemma} \begin{proof} Suppose $P\subseteq \R^2$ is contained in a closed $\ell^1$-ball $B$ of diameter $D$. \noindent Let $a_1, a_2$ be the left and right vertices of $B$, respectively, so in particular $\norm{a_1-a_2}_1=D$. Let $U$ denote the closed (including $a_1,a_2$) upper $\ell^1$-semicircle connecting $a_1$ and $a_2$, and let $L$ denote the open (not including $a_1,a_2$) lower $\ell^1$-semicircle connecting $a_1$ and $a_2$. Since the $\ell^1$-norm is invariant under $90^{\circ}$ rotation, it suffices to establish the conclusion of the lemma for removing the points of $P$ lying in $U$. Suppose $x_1,x_2\in P\setminus U$. \noindent \textbf{Case 1:} At least one of $x_1,x_2$ lies in $B\setminus(U\cup L)$, which is an open $\ell^1$-ball of radius $D/2$. \noindent Assume without loss of generality that $x_1\in B\setminus(U\cup L)$, and let $c$ be the center of $B$. Therefore, $\norm{x_1-c}_1 < D/2$ and $\norm{x_2-c}_1\leq D/2$. By the triangle inequality, $\norm{x_1-x_2}_1\leq \norm{x_1-c}_1+\norm{c-x_2}_1 < D/2+D/2=D.$ \noindent \textbf{Case 2:} $x_1,x_2\in L$. After possibly reflecting, assume without loss of generality that $x_1$ is to the left of $x_2$ and $\norm{x_1-a_1}_1\leq \norm{x_2-a_2}_1$, so $x_1$ is positioned at least as high as $x_2$, By replacing $x_1$ with $a_1$, we move up and to the left, so both the horizontal and vertical components of the $\ell^1$-distance to $x_2$ get larger, hence $$\norm{x_1-x_2}_1<\norm{a_1-x_2}_1=D. $$ \noindent In both cases, all distances amongst points in $P\setminus U$ are strictly less than $D$, and the lemma follows. \end{proof} \ \begin{lemma}\label{line} If $P\subseteq \R^d$ is contained in a line and determines at most $k$ distinct $\ell^1$-distances, then $|P|\leq k+1$. Further, if $|P|=k+1$, then $P$ is an arithmetic progression, meaning the $\ell^1$-distances are $\lambda,2\lambda,\dots, k\lambda$ for some $\lambda>0$. \end{lemma} \begin{proof} Since $\ell^1$-distance along a straight line in $\R^d$ is just a constant multiple, depending on the direction of the line, times the standard Euclidean distance, it suffices to establish the lemma with $d=1$, for which we induct on $k$. \noindent The base case $k=1$ is trivial, as three points $x_1<x_2<x_3$ in $\R$ automatically determine two distances $x_2-x_1<x_3-x_1$, and any two points form an arithmetic progression. \noindent Now, fix $k\geq 2$, and assume that if $Q\subseteq \R$ determines at most $k-1$ distances, then $|Q|\leq k$, and further, if $|Q|=k$, then $Q$ is an arithmetic progression. Now suppose $P\subseteq \R$ determines at most $k$ distances. \noindent Let $P=\{x_1<x_2<\cdots<x_n\}$. The $n-1$ distances $x_2-x_1<x_3-x_1<\cdots<x_n-x_1$ are all distinct, hence $n-1\leq k$, or in other words $n\leq k+1$. Further, suppose $n=k+1$. By removing $x_{k+1}$, we also remove the longest distance $x_{k+1}-x_1$, so the set $Q=\{x_1,\dots, x_k\}$ determines $k-1$ distances. By our inductive hypothesis, $Q$ must be an arithmetic progression, in other words $Q=\{x_1,x_1+\lambda,x_1+2\lambda, \dots x_1+(k-1)\lambda\}$. \noindent If $x_{k+1}<x_1+k\lambda$, then both $x_{k+1}-x_1>(k-1)\lambda$ and $x_{k+1}-x_k<\lambda$ are new distances not determined by $Q$. If $x_{k+1}>x_1+k\lambda$, then both $x_{k+1}-x_1>k\lambda$ and $x_{k+1}-x_2>(k-1)\lambda$ are new distances not determined by $Q$. In either case, $P$ determines at least $k+1$ distinct distances, contradicting the assumption that it determines at most $k$ distances. Therefore, $x_{k+1}$ must be $x_1+k\lambda$, and the lemma follows.\end{proof} \begin{lemma} \label{semicirc} If $S \subseteq \R^2$ is contained in the union of two adjacent sides of an $\ell^1$-circle and determines at most $k$ distinct $\ell^1$-distances, then $|S|\leq 2k+1$. Further, if $|S|=2k+1$, then the points of $S$ on each side form an arithmetic progression containing the shared vertex. \end{lemma} \begin{proof}Suppose $S \subseteq \R^2$ is contained in the union of two adjacent sides of an $\ell^1$-circle and determines at most $k$ distinct $\ell^1$-distances. Assume without loss of generality that the two adjacent sides are the closed upper semicircle. We know from Lemma \ref{line} that there are at most $k+1$ points on each of the two sides. \noindent Further, if $|S|\geq 2k+1$, then there are exactly $k+1$ points on one side, assume the left, and at least $k$ points on the right side. Let $x_1,\dots,x_{k+1}$ denote the points of $P$ on the left side, ordered left to right, and let $y$ be any point of $P$ on the right side. We note that $$\norm{x_1-x_2}_1<\norm{x_1-x_3}_1<\cdots<\norm{x_1-x_{k+1}}_1\leq \norm{x_1-y}_1, $$ and $\norm{x_1-x_{k+1}}_1= \norm{x_1-y}_1$ is only possible if $x_{k+1}$ is the vertex shared by the two sides. In particular, if the shared vertex is not included amongst the $k+1$ points on the left side, then at least $k+1$ distinct $\ell^1$-distances occur from the leftmost point, contradicting our assumption. \noindent Therefore, if $|S|\geq 2k+1$, it must be the case that there are exactly $k+1$ points on both the left and right sides, including the shared vertex, meaning in fact $|S|=2k+1$. Finally, by Lemma \ref{line}, we know that the $k+1$ points on each side must form an arithmetic progression. \end{proof} \noindent We are now fully armed to show the unique optimality of $\Lambda_2(k)$. \begin{proof}[Proof of Theorem \ref{opt}] We induct on $k$. For our base case, consider $k=0$. In order for a set to determine $0$ $\ell^1$-distances (as always, not including $0$), it can contain at most $1=(0+1)^2$ point, and if it contains a point, then it is trivially a translation of $\Lambda_2(0)=\{(0,0)\}$. \noindent Now, fix $k\in \mathbb{N}$, assume the conclusion of the theorem holds for $k-1$, and suppose $P\subseteq \R^2$ determines at most $k$ distinct $\ell^1$-distances. By Lemma \ref{SD}, $P$ is contained in a closed $\ell^1$-ball $B$ of diameter $D$, where $D$ is the largest $\ell^1$-distance determined by $P$. By Lemma \ref{remove}, we can remove the distance $D$ by removing the points of $P$ that lie on the closed upper $\ell^1$-semicircle $U$ on the boundary of $B$. Since $D$ has been removed as an $\ell^1$-distance, we know that $T=P\setminus U$ determines at most $k-1$ distinct $\ell^1$-distances. By our inductive hypothesis, $|T|\leq k^2$, and if $|T|=k^2$, then $T$ is $\ell^1$-similar to $\Lambda_2(k-1)$. \noindent Further, by Lemma \ref{semicirc}, we know that $S=P\cap U$ satisfies $|S|\leq 2k+1$, and if $|S|=2k+1$, then $S$ consists of two $(k+1)$-term arithmetic progressions, one on each side of $U$, which meet at the shared vertex. Therefore, $|P|\leq |T|+|S|\leq k^2+2k+1=(k+1)^2$, and $|P|=(k+1)^2$ if and only if $T$ is $\ell^1$-similar to $\Lambda_2(k-1)$ and $S$ is a union of two arithmetic progressions meeting at the shared vertex. Finally, the only way these two sets can be combined without creating additional $\ell^1$-distances is for $S\cup T$ to be $\ell^1$-similar to $\Lambda_2(k)$. \end{proof} \section{Single $\ell^1$-distance in three dimensions: Proof of Theorem \ref{singdist}} \label{3dsec} Without analogs of Lemmas \ref{linfty} and \ref{SD} in dimension $d\geq 3$, our strategy for proving Theorem \ref{opt} does not naturally generalize to higher dimensions. However, in the case of $k=1$, we make the observation that if $P\subseteq \R^d$ determines a single $\ell^1$-distance, then all but the ``southernmost" point (the point minimizing the last coordinate) of $P$ lie on a single closed upper $\ell^1$-hemisphere. The following sequence of lemmas provide a detailed investigation into how $\ell^1$-distance behaves between points on a single upper $\ell^1$-hemisphere in $\R^3$, which consists of four flat faces, one for each quadrant determined by the first two coordinates, intersecting at a single northernmost point. The three lemmas correspond to the cases where the points lie on the same face, opposite faces, or neighboring faces, respectively. \ \begin{lemma} \label{face} Suppose $V,W\in \R^3$ with $V=(x_1,y_1,z_1)$ and $W=(x_2,y_2,z_2)$. If $\norm{V}_1=\norm{W}_1$ and $x_1x_2,y_1y_2,z_1z_2 \geq 0$, then $$\norm{V-W}_1=2\max\{|x_1-x_2|,|y_1-y_2|,|z_1-z_2|\}. $$ \end{lemma} \ \begin{proof} Suppose $V,W\in \R^3$, $V=(x_1,y_1,z_1)$, $W=(x_2,y_2,z_2)$, $\norm{V}_1=\norm{W}_1=\lambda$, and $x_1x_2,y_1y_2,z_1z_2 \geq 0$. After reflections about coordinate planes, coordinate permutations, and relabeling $V$ and $W$ (which all preserve both sides of the equation in the conclusion of the lemma), we can assume without loss of generality that all coordinates are nonnegative and $x_1-x_2\geq |y_1-y_2| \geq |z_1-z_2|$. Since $$\norm{V}_1=x_1+y_1+z_1=\norm{W}_1=x_2+y_2+z_2=\lambda,$$ we have in particular that $(x_1-x_2)+(y_1-y_2)+(z_1-z_2)=0$. Since the largest coordinate distance is in the $x$-direction, and $x_1\geq x_2$, we must have $y_1\leq y_2$ and $z_1\leq z_2$. Therefore \begin{align*}\norm{V-W}_1&=(x_1-x_2)+(y_2-y_1)+(z_2-z_1) \\ &=x_1-x_2+y_2-y_1+(\lambda-x_2-y_2)-(\lambda-x_1-y_1) \\ &=2(x_1-x_2), \end{align*} and the lemma follows. \end{proof} \ \begin{lemma} \label{oppface} Suppose $V,W\in \R^3$ with $V=(x_1,y_1,z_1)$ and $W=(x_2,y_2,z_2)$. If $\norm{V}_1=\norm{W}_1=\lambda$, $x_1x_2\leq 0$, $y_1y_2\leq 0$, and $z_1,z_2\geq 0$, then $$\norm{V-W}_1=2(\lambda-\min\{z_1,z_2\}). $$ \end{lemma} \begin{proof} Suppose $V,W\in \R^3$ with $V=(x_1,y_1,z_1)$, $W=(x_2,y_2,z_2)$, $\norm{V}_1=\norm{W}_1=\lambda$, $x_1x_2\leq 0$, $y_1y_2\leq 0$, and $z_1,z_2\geq 0$. After reflections about coordinate planes and relabeling $V$ and $W$, we can assume without loss of generality that $x_1,y_1\geq 0$, $x_2,y_2\leq 0$, and $z_1\leq z_2$. \noindent Therefore, $x_1+y_1=\lambda-z_1$ while $-x_2-y_2=\lambda-z_2$, hence \begin{align*} \norm{V-W}_1&=(x_1-x_2)+(y_1-y_2)+(z_2-z_1) \\ &= \lambda-z_1+\lambda-z_2+z_2-z_1 \\ &=2(\lambda-z_1), \end{align*} and the lemma follows. \end{proof} \ \begin{lemma} \label{neighbor} Suppose $V,W\in \R^3$ with $V=(x_1,y_1,z_1)$, $W=(-x_2,y_2,z_2)$, $\norm{V}_1=\norm{W}_1=\lambda$, and $x_1x_2, y_1y_2, z_1z_2 \geq 0$. If $\norm{V-W}_1=\lambda$, then $|x_1|\leq \lambda/2$. \end{lemma} \begin{proof} Suppose $V,W\in \R^3$ with $V=(x_1,y_1,z_1)$, $W=(-x_2,y_2,z_2)$, $\norm{V}_1=\norm{W}_1=\lambda$, $x_1x_2, y_1y_2, z_1z_2 \geq 0$. After reflecting about coordinate planes and scaling, we can assume $x_1,x_2,y_1,y_2,z_1,z_2\geq 0$, and $\lambda=2$. If $\norm{V-W}_1=2$, then the largest possible value of $y_2+z_2$ is $y_1+z_1+2-(x_1+x_2)$. However, since $\norm{W}_1=2$, we must have $y_2+z_2=2-x_2$, hence $2-x_2\leq y_1+z_1+2-(x_1+x_2)$, which rearranges to $x_1\leq y_1+z_2=2-x_1$, hence $x_1\leq 1$, as required. \end{proof} \ \noindent We now establish the unique optimality of $\Lambda_3(1)$ by conducting a case analysis based on the concentration of the points of $P$, apart from the southernmost point, on the four faces of a single closed upper $\ell^1$-hemisphere. \ \begin{proof}[Proof of Theorem \ref{singdist}] Suppose $P\subseteq \R^3$ determines a single $\ell^1$-distance $\lambda$, and choose a point $c\in P$ that minimizes the $z$-coordinate. By translating and dilating, we can assume without loss of generality that $c=(0,0,0)$ and $\lambda=2$, and hence the remaining elements of $P$ are all contained in the closed upper $\ell^1$-hemisphere $H$ of radius $2$ centered at $(0,0,0)$. We note that the southernmost point of $\Lambda_3(1)$ is $(0,0,-1)$, so our end goal in this proof is to show that $|P|<6$ unless $P$ is $\Lambda_3(1)$ shifted up by $1$. \noindent We consider the different ways that $P$ can be concentrated on the faces of $H$. To this end, we define $H_{++}=\{(x,y,z)\in H: x,y\geq 0\}$ and $H_{+-}=\{(x,y,z)\in H: x\geq 0, y\leq 0\}$, with analogous definitions for $H_{-+}$ and $H_{--}$. We refer to the pair $H_{++}$, $H_{--}$ as \textit{opposite} faces, and likewise for $H_{+-}$, $H_{-+}$. The three lemmas proven at the beginning of this section allow us to make the following assertions: \begin{enumerate}[(i)] \item \label{oppitem} For any pair of distinct points $U=(x_1,y_1,z_1),V=(x_2,y_2,z_2)\in P\cap H$, with $U$ and $V$ lying on opposite faces, we have by Lemma \ref{oppface} that $\min\{z_1,z_2\}=1$. \\ \item \label{sameitem} For any pair of distinct points $U=(x_1,y_1,z_1),V=(x_2,y_2,z_2)\in P\cap H$, with $U$ and $V$ lying on the same face, we have by Lemma \ref{face} that $\max\{ |x_1-x_2|,|y_1-y_2|,|z_1-z_2| \}=1$. \\ \item \label{nitem} For distinct points $U=(x_1,y_1,z_1),V=(x_2,y_2,z_2)\in P\cap H$, with $U\in H_{++}$ and $V\in H_{-+}$, we have by Lemma \ref{neighbor} that $x_1\leq 1$. Similarly, by permuting coordinates, if $U\in H_{++}$ and $V\in H_{+-}$, then $y_1\leq 1$. \end{enumerate} \noindent If $|P|\geq 6$, then at least five points of $P$ lie on $H$, and in particular the sizes of the four intersections of $P$ with the respective faces of $H$ must add to at least five. Therefore, the only possible arrangements of $P\cap H$ include either three points on a single face, or two points on one face and a point on the opposite face. Further, the proof is greatly simplified in the case that the ``north pole" $(0,0,2)\in P$, so we divide the argument into the following three cases: \begin{itemize} \item{\textbf{Case 1:}} $(0,0,2)\in P$. \\ \item{\textbf{Case 2:}} $(0,0,2)\notin P$, and $P$ contains three points $U,V,W\in H$ such that $U$ and $V$ lie on the same face, and $W$ lies on the opposite face. \\ \item{\textbf{Case 3:}} $(0,0,2)\notin P$, and there exists a face of $H$ containing at least three points of $P$. \end{itemize} \noindent \textbf{Proof for Case 1:} Let $V=(0,0,2)$. For $Q=(x,y,z)\in (P\cap H)\setminus \{V\}$, we have by (\ref{oppitem}) that $z=1$. In particular, the elements of $P$ other than $(0,0,0)$ and $(0,0,2)$ take the form $(x,y,1)$ with $|x|+|y|=1$, and all pairs are separated by $\ell^1$-distance $2$. By Theorem \ref{opt}, there can be at most four such elements, and the only choice of four that works is $(1,0,1)$, $(-1,0,1)$, $(0,1,1)$, and $(0,-1,1)$. The resulting arrangement is $\Lambda_3(1)$ translated up by $1$, which establishes Theorem \ref{singdist} in this case. \noindent \textbf{Proof for Case 2:} After reflecting about coordinate planes, we can assume that $U,V\in H_{++}$ and $W\in H_{--}$. Letting $U=(x_0,y_0,z_0)$, $V=(x_1,y_1,z_1)$, and $W=(x_2,y_2,z_2)$, we have by (\ref{oppitem}) and (\ref{sameitem}) that $$\max\{|x_0-x_1|,|y_0-y_1|,|z_0-z_1|\}=\min\{z_0,z_2\}=\min\{z_1,z_2\}=1.$$ In particular, all three $z$ coordinates are at least $1$, and since $(0,0,2)\notin P$, we have $|z_0-z_1|<1$. Therefore, we simultaneously know that $0\leq x_0,x_1,y_0,y_1 \leq 1$ and $\max\{|x_0-y_0|,|x_1-y_1|\}=1$. \noindent This implies that (after potentially relabeling) either $U=(1,0,1)$ and $V=(0,y,2-y)$ for some $0<y\leq 1$ or $U=(x,0,2-x)$ for some $0<x\leq 1$ and $V=(0,1,1)$. In either case, $U\in H_{++}\cap H_{+-}$, and $V\in H_{++}\cap H_{-+}$, so $P$ contains at least one element on every face of $H$. Therefore, by (\ref{oppitem}), all points of $P$ lying on $H$ have $z$-coordinate at least $1$. Further, by the same reasoning as above, there are at most two points of $P$ on each face, and the only way two points of $P$ can lie on the same face is if they lie on opposite sides of the boundary, as with $U$ and $V$. In particular, at most four points of $P$ lie on $H$, and hence $P$ contains at most five points in total. \noindent \textbf{Proof for Case 3:} This case gets a bit stickier, because, as some trial and error reveals, there are a variety of possible arrangements of three points on a single face of $H$ that are all separated by $\ell^1$-distance $2$. \noindent Focusing on $H_{++}$ for the sake of exposition, we see that our desired configuration of $\{(1,0,1), (0,1,1), (0,0,2)\}$ is merely one member of a family of arrangements obtained from the following process: \begin{itemize} \item Choose $x_0,y_0,z_0\geq 0$ with $x_0+y_0+z_0\leq 1$, and let $\alpha=1-(x_0+y_0+z_0)$ \\ \item Starting from $(x_0,y_0,z_0)$, construct a point by adding $1$ to one coordinate and $\alpha$ to another coordinate (so the coordinates add to $2$), then produce two additional points in a similar way by rotating the original choice of coordinates. For example, the initial choice of $U=(x_0+1,y_0+\alpha,z_0)$ uniquely determines the two additional points $V=(x_0,y_0+1,z_0+\alpha)$ and $W=(x_0+\alpha,y_0,z_0+1)$. All of these points lie on $H_{++}$, and by (\ref{sameitem}) they are all separated by $\ell^1$-distance $2$. In fact, the only other possible set of three points yielded by this process (up to labeling) is $U=(x_0+1,y_0,z_0+\alpha)$, $V=(x_0+\alpha, y_0+1, z_0)$, and $W=(x_0,y_0+\alpha, z_0+1)$. For additional clarity, a specific example is $x_0=0.1$, $y_0=0.3$, $z_0=0.4$, hence $\alpha=0.2$, which could yield the three-point arrangements $\{(1.1,0.5,0.4),(0.1,1.3,0.6), (0.3,0.3,1.4)\}$ or $\{(1.1,0.3,0.6),(0.3,1.3,0.4),(0.1,0.5,1.4)\}$. \end{itemize} \noindent We hope to demystify the situation by arguing that the arrangements discussed above are in fact the only possible arrangements. To this end, after reflections, we can assume $P$ contains three points $U,V,W\in H_{++}$, and we settle Case 3 with the following steps: \begin{itemize} \item \textbf{Step 1 :} Show that $U,V,W$ take the form discussed above. In particular, after specifying the minimum values of each coordinate and a single point, the second and third points are uniquely determined, hence there cannot be a fourth point in $P\cap H_{++}$. \\ \item \textbf{Step 2:} Show that $P$ can contain at most one point in $(H_{+-}\cup H_{-+})\setminus H_{++}$ before necessarily reducing to Case 2. This means that any hypothetical fifth point of $P\cap H$ necessarily lies on $H_{--}$, which itself reduces the argument back to Case 2.\\ \end{itemize} \noindent \textbf{Step 1:} Let $x_0$, $y_0$, and $z_0$ be the minimum $x$, $y$, and $z$-coordinates, respectively, attained by $U$, $V$, and $W$. In what follows, we repeatedly appeal to (\ref{sameitem}), which tells us that for every pair of points in $\{U,V,W\}$, the maximum coordinate distance is exactly $1$. In particular, the maximum $x$, $y$, and $z$-coordinates attained by $U$,$V$, and $W$ are at most $x_0+1$, $y_0+1$, and $z_0+1$, respectively, and we begin by arguing that this inequality must be equality in all three coordinates. \noindent Suppose that this inequality is strict in at least one coordinate. By permuting coordinates and relabeling points we may assume that $U=(x_0,y,z)$, and neither of $V$ and $W$ has $x$-coordinate $x_0+1$. Therefore, the maximum coordinate distance of $1$ required by (\ref{sameitem}) must occur in either the $y$ or $z$-coordinates, and since $x_0$ is the minimum $x$-coordinate, $V$ and $W$ must both take one of the following forms: $(x_0+\alpha,y-1, z+(1-\alpha))$ for some $0\leq\alpha< 1$, or $(x_0+\beta, y+(1-\beta),z-1)$ for some $0\leq \beta <1$. However, no combination of these choices for $V$ and $W$ have a maximum coordinate distance of $1$ from each other, so this arrangement is impossible. Therefore, all the maxima $x_0+1$, $y_0+1$, and $z_0+1$ are indeed achieved. For the remainder of the proof, we will refer to the respective coordinate values $x_0$, $y_0$, and $z_0$ as \textit{minimum coordinates}, and we will similarly refer to the respective coordinate values $x_0+1$, $y_0+1$, $z_0+1$ as \textit{maximum coordinates}. We complete step one by considering the following three subcases. \begin{itemize} \item \textbf{Subcase A:} Two maximum coordinates appear simultaneously in a single point. \\ \noindent Since all the points have $\ell^1$-norm $2$, this subcase necessitates that $x_0=y_0=z_0=0$, and we assume without loss of generality that $U=(1,1,0)$. Since the minimum $x$ and $y$ coordinates of $0$ must be attained, $\{V,W\}$ contains points of the form $(x,0,2-x)$ and $(0,y,2-y)$, respectively, for some $0\leq x,y\leq 1$. However, since the maximum $z$-coordinate is $1$, the only admissible choices are $x=y=1$, hence the three points are $(1,1,0)$, $(0,1,1)$, and $(1,0,1)$, which take the required form with $\alpha=1$. \\ \item \textbf{Subcase B:} Two minimum coordinates appear simultaneously in a single point.\\ \noindent Assume without loss of generality that $U=(x_0,y_0,z)$. Since $x_0$ and $y_0$ are minimum coordinates, each of $V$ and $W$ must take the form $(x_0+\alpha, y_0+\beta, z-(\alpha+\beta))$ for some $\alpha,\beta\geq 0$, and by (\ref{sameitem}) we must have $\alpha+\beta=1$. Further, since the $z$-coordinate of both $V$ and $W$ is $z-1$ (which is hence the minimum coordinate $z_0$), the maximum coordinate distance of $1$ must occur in the first two coordinates, meaning that $\{V,W\}=\{(x_0+1,y_0,z_0),(x_0,y_0+1,z_0)\}$. In particular, the arrangement takes the required form with $\alpha=0$. \\ \item \textbf{Subcase C:} Exactly one minimum coordinate and one maximum coordinate occurs in each point.\\ \noindent After permuting coordinates and relabeling points we assume $U=(x_0+1,y_0+\alpha,z_0)$ where $0<\alpha=1-(x_0+y_0+z_0)<1$. In order to meet the subcase conditions, have a maximum coordinate distance of $1$ from $U$, and have $\ell^1$-norm $2$, the options for $V$ and $W$ are $(x_0, y_0+1, z_0+\alpha)$, $(x_0, y_0+\alpha, z_0+1)$, and $(x_0+\alpha, y_0, z_0+1)$. Of these three possibilities, there is only one pair that are separated by $\ell^1$-distance $2$ from each other, hence $\{V,W\}=\{(x_0, y_0+1, z_0+\alpha),(x_0+\alpha, y_0, z_0+1)\}$, as required. \end{itemize} \noindent \textbf{Step 2:} Suppose $P$ contains a point $Q\in H_{-+}\setminus H_{++}$ (the argument is completely analogous for $Q\in H_{+-}\setminus H_{++}$). By (\ref{nitem}), in order for $Q$ to be separated from $U=(x_0+1,y_0+\alpha, z_0)$ by $\ell^1$-distance $2$, we must have $x_0+1\leq 1$, and hence $x_0=0$. In particular, $V=(0,y_0+1, z_0+\alpha) \in P \cap (H_{-+}\cap H_{++})$, so $P$ contains at least two points on $H_{-+}$. This means that, in order to avoid reducing to Case 2, $P$ cannot contain any elements of $H_{+-}$. \noindent If instead $P$ contains a second point $R\in H_{-+}\setminus H_{++}$, hence a third point in $H_{-+}$, then we fall back to our previous analysis of three points on a single face, adapted by taking negatives of all $x$-coordinates. In particular, because $V$ has $x$-coordinate $0$, which minimizes the $x$-coordinate in absolute value among the points in $P\cap H_{-+}$, either $Q$ or $R$ must maximize the $x$-coordinate in absolute value and have $x$-coordinate $-1$. Assuming $Q$ has $x$-coordinate $-1$, in order for $Q$ to be separated from $U=(1,y_0+\alpha,z_0)$ by $\ell^1$-distance $2$, we must have $Q=(-1,y_0+\alpha,z_0)$. In order for $\{V,Q,R\}$ to meet the required form for three points of $P$ on $H_{-+}$ established in Step 1, we must have $R=(-\alpha,y_0,z_0+1)$. However, in this case we see that $\norm{U-R}_1=2+2\alpha=2$, hence $\alpha=0$, which contradicts the assumption that $R\notin H_{++}$. \end{proof} \section{Conditional Results in Higher Dimensions} \label{hdsec} In the remainder of our discussion, we use the terms \textit{$\ell^1$-sphere} and \textit{$\ell^1$-ball} as before, defined analogously to regular spheres and balls in $\R^d$, with the usual distance replaced by $\ell^1$-distance. In an effort to establish results in higher dimensions, we make the following observations, heavily inspired by our journey thus far: \begin{enumerate}[(a)] \item \label{capt} As noted at the beginning of Section \ref{3dsec}, our proof of Theorem \ref{opt} does not naturally generalize to higher dimensions, because in dimension $d\geq 3$, it is not necessarily the case that if the largest $\ell^1$-distance determined by a finite set $P\subseteq \R^d$ is $\lambda$, then $P$ is contained in a closed $\ell^1$-ball of diameter $\lambda$. However, the argument in Lemma \ref{remove} does generalize to all dimensions: the distance $\lambda$ can be removed from an $\ell^1$-ball of diameter $\lambda$ by removing the closed upper $\ell^1$-hemisphere. In particular, if we somehow could capture our set inside such a ball, then by mimicking the proof of Theorem \ref{opt}, the problem is reduced to determining maximal configurations of points arranged on single closed upper $\ell^1$-hemisphere, which would then facilitate an induction on the number of distinct distances. \\ \item \label{ballitem} Suppose $P \subseteq \R^d$ is a finite set determining at most $k$ distinct $\ell^1$-distances, with largest $\ell^1$-distance $\lambda$. By translating and scaling, we can assume that $\lambda=2k$ and the ``southernmost point" of $P$, minimizing the $x_d$ coordinate, is $-ke_d$, where $e_d$ is the $d$-th standard basis vector. An enticing observation, particularly in juxtaposition with (\ref{capt}), is the following: if $ke_d$ is also in $P$, then, since $2k$ is the largest $\ell^1$-distance, $P$ is contained in the intersection of the closed $\ell^1$-ball of radius $2k$ centered at $-ke_d$ and the closed $\ell^1$-ball of radius $2k$ centered at $ke_d$, which is conveniently the closed $\ell^1$-ball of radius $k$ centered at the origin. \\ \item \label{hyper} Inspired by the simplicity of Case 1 in the proof of Theorem \ref{singdist}, we see that if $U$ lies on an upper $\ell^1$-hemisphere $H\subseteq \R^d$, then the $\ell^1$-distance between $U$ and the ``north pole" of $H$ is determined entirely by the $x_d$-coordinate of $U$. More specifically, if $H$ is the closed upper $\ell^1$-hemisphere of radius $k$ centered at the origin and $U=(x_1,\dots,x_d)\in H$, then $$\norm{U-ke_d}_1=|x_1|+\cdots+|x_{d-1}|+k-x_d=2(k-x_d). $$ In particular, if $ke_d\in P$ and $P$ determines only $k$ distinct $\ell^1$-distances $\{\lambda_i\}_{i=1}^k$, then the points of $(P\cap H)\setminus ke_d$ are restricted to the hyperplanes $\{x_d=c_i\}$ for $1\leq i \leq k$., where $c_i=k-\lambda_i/2$. Further, the intersection of $H$ with the hyperplane $\{x_d=c_i\}$ is $$\{(x_1,\dots,x_{d-1},c_i): |x_1|+\cdots+|x_{d-1}|=k-c_i\},$$ which is a copy of the $\ell^1$-sphere of radius $k-c_i$ centered at the origin in $\R^{d-1}$. This would allow us to analyze $P\cap H$ by inducting on dimension, analogous to the invocation of Theorem \ref{opt} during Case 1 in the proof of Theorem \ref{singdist}. \end{enumerate} \noindent These three items combine to a clear aspirational reality: given a finite set $P\subseteq \R^d$ that determines at most $k$ distinct $\ell^1$-distances, the largest of which is $\lambda$, letting $H$ denote the closed upper $\ell^1$-hemisphere of radius $\lambda$ centered at the ``southernmost" point of $P$, we could fully adapt the proof of Theorem \ref{opt} and induct on both $d$ and $k$, if only we could assume that the ``north pole" of H is also in $P$. If we were considering the usual Euclidean distance, this would be no obstruction at all, as we could rotate our set and assume without loss of generality that the largest distance $\lambda$ occurs parallel to the $x_d$-axis. However, since $\ell^1$-distance is not invariant under rotation, we require an additional assumption to establish a conditional version of Conjecture \ref{optcon}. The following definition, conjecture, and theorem fully formalize this conditional result, after which we conclude our discussion. \begin{definition} Given $P\subseteq \R^d$ and an $\ell^1$-distance $\lambda>0$, we say that $\lambda$ \textit{occurs in an axis-parallel direction} if there exists $x\in P$ and $1\leq i \leq d$ such that $x+\lambda e_i\in P$, where $e_i$ is the $i$-th standard basis vector. Further, if $P$ is bounded, we say that $P$ is \textit{axis-parallel} if the largest $\ell^1$-distance determined by $P$ occurs in an axis-parallel direction. \end{definition} \begin{conjecture} \label{axisp} Suppose $d,k\in \mathbb{N}$. If $P$ is of maximal size amongst subsets of $\R^d$ determining at most $k$ distinct $\ell^1$-distances, then $P$ is axis-parallel. The same holds within the class of sets contained in an $\ell^1$-sphere in $\R^d$. \end{conjecture} \begin{theorem}\label{condinc} Conjecture \ref{axisp} implies Conjecture \ref{optcon}. \end{theorem} \begin{proof} We proceed via two inductions, one on the dimension $d$ and another on the number of distinct $\ell^1$-distances $k$. We streamline the argument by defining the following propositions for each $d\in \mathbb{N}$ and each nonnegative integer $k$: \begin{itemize} \item Opt$(d,k)$: $\Lambda_d(k)$ is the unique set, up to $\ell^1$-similarity, of maximal size amongst subsets of $\R^d$ determining at most $k$ distinct $\ell^1$-distances. Conjecture \ref{optcon} is precisely the statement that Opt$(d,k)$ holds for all $d,k\in \mathbb{N}$. \\ \item S-Opt$(d,k)$: $\Lambda_d(k)\setminus \Lambda_d(k-2)$ is the unique set, up to $\ell^1$ similarity, of maximal size amongst sets contained in an $\ell^1$-sphere in $\R^d$ determining at most $k$ distinct $\ell^1$-distances. For $k=0$ or $1$, we take the convention that $\Lambda_d(-1)=\Lambda_d(-2)=\emptyset$.\\ \item H-Opt$(d,k)$: Let $H$ denote the closed upper $\ell^1$-hemisphere of radius $k$ centered at the origin in $\R^d$. If $ke_d\in E\subseteq H$ and $E$ determines at most $k$ distinct $\ell^1$-distances, the largest of which is $2k$, then $|E|\leq |\Lambda_d(k)\cap H|$, and $|E|=|\Lambda_d(k)\cap H|$ if and only if $E=\Lambda_d(k)\cap H$.\\ \end{itemize} \noindent For the necessary base cases, we note that S-Opt$(1,k)$ and H-Opt$(1,k)$ trivially hold for all $k\in \mathbb{N}$ as $\ell^1$-spheres and $\ell^1$-hemispheres in $\R$ contain just two points and one point, respectively. Also, Opt$(d,0)$ holds trivially for all $d\in \mathbb{N}$ because a set determining no $\ell^1$-distances contains at most a single point. Under the assumption that Conjecture \ref{axisp} holds, we verify Conjecture \ref{optcon} by establishing the following implications: \begin{enumerate} \item \label{SH} S-Opt$(d-1,k)$ for all $k\in \mathbb{N} \implies$ H-Opt$(d,k)$ and S-Opt$(d,k)$ for all $k\in \mathbb{N}$, so S-Opt$(d,k)$ and H-Opt$(d,k)$ hold for all $d,k\in \mathbb{N}$. \\ \item \label{OI} Opt$(d,k-1)$ and H-Opt$(d,k) \implies$ Opt$(d,k)$, so Opt$(d,k)$ holds for all $d,k\in \mathbb{N}$, as required.\\ \end{enumerate} \noindent \textbf{Proof of (\ref{OI}):} Fix $d,k\in \mathbb{N}$, and suppose Opt$(d,k-1)$ and H-Opt$(d,k)$ hold. Suppose $P\subseteq \R^d$ determines at most $k$ distinct $\ell^1$-distances, and has maximal size amongst sets with this property. By scaling we can assume that the largest $\ell^1$-distance determined by $P$ is $2k$, which by Conjecture \ref{axisp} we know occurs in an axis-parallel direction. After permuting coordinates and translating, we can assume that $-ke_d,ke_d\in P$. As noted in (\ref{ballitem}), this implies that $P$ is contained the closed $\ell^1$-ball of radius $k$ centered at the origin. To verify this, suppose $U=(x_1,\dots,x_d)\in P$ with $\norm{U}_1 >k$. If $x_d\geq0$, then $\norm{U-(-ke_d)}_1=\norm{U}_1+k>2k$, while if $x_d\leq 0$, the same holds for the $\ell^1$-distance between $U$ and $ke_d$, contradicting the fact that $2k$ is the largest $\ell^1$-distance determined by $P$. \noindent As noted in (\ref{capt}), the $\ell^1$-distance $2k$ can be eliminated from $P$ by removing the points in $P\cap H$, where $H$ is the closed upper $\ell^1$-hemisphere of radius $k$ centered at the origin. This is because, within a closed $\ell^1$-ball of radius $k$, the $\ell^1$-distance $2k$ only occurs between pairs of points on opposite faces of the boundary. By H-Opt$(d,k)$, we know that $|P\cap H|\leq |\Lambda_d(k)\cap H|$, and further $|P\cap H|= |\Lambda_d(k)\cap H|$ if and only if $P\cap H=\Lambda_d(k)\cap H$, in which case the $\ell^1$-distances determined by $P$ are $2,4,\dots,2k$. \noindent Because the $\ell^1$-distance $2k$ does not occur in $P\setminus H$, we have that $P\setminus H$ determines at most $k-1$ distinct $\ell^1$-distances. By Opt$(d,k-1)$, we know that $|P\setminus H|\leq |\Lambda_d(k-1)|$, and further $|P\setminus H|= |\Lambda_d(k-1)|$ if and only if $P\setminus H$ is $\ell^1$-similar to $\Lambda_d(k-1)$. In order for both $P\cap H$ and $P\setminus H$ to attain their maximum possible sizes, the $\ell^1$-distances determined by $P\setminus H$ must be $2,4,\dots,2k-2$. In this case, since an $\ell^1$-similar copy of $\Lambda_d(k-1)$ is uniquely determined by its ``south pole" and its largest distance, we know that $P\setminus H$ must be $\Lambda_d(k-1)$ shifted down by $1$. In other words, \begin{align*} P\setminus H&= \{(x_1,\dots,x_{d}-1)\in \mathbb{Z}^d: |x_1|+\cdots+|x_d|\leq k-1, \ x_1+\dots+x_d\equiv k-1 \ (\text{mod }2)\} \\ &=\{(x_1,\dots,x_{d})\in \mathbb{Z}^d: |x_1|+\cdots+|x_d+1|\leq k-1, \ x_1+\dots+x_d\equiv k \ (\text{mod }2)\}. \end{align*} The latter description ensures that $P\setminus H \subseteq \Lambda_d(k) \setminus H$, and conversely, if $U=(x_1,\dots,x_d) \in \Lambda_d(k)\setminus H$, then either $x_d<0$ or $\norm{U}_1\leq k-2$. In either case $|x_1|+\cdots+|x_d+1|\leq k-1$, and hence $U\in P\setminus H$. Bringing everything together, we have that if $P\cap H$ and $P\setminus H$ both attain their maximum possible size, then $P\cap H= \Lambda_d(k)\cap H$ and $P\setminus H=\Lambda_d(k)\setminus H$, hence $P=\Lambda_d(k)$, so Opt$(d,k)$ holds. \noindent \textbf{Proof of (\ref{SH}):} \noindent Fix $d\geq 2$, suppose S-Opt$(d-1,k)$ holds for all $k\in \mathbb{N}$, and fix $k\in \mathbb{N}$. Suppose $P\subseteq \R^d$ is contained in the $\ell^1$-sphere $S$ of radius $k$ centered at the origin, and that $P$ has maximal size amongst all such sets determining at most $k$ distinct $\ell^1$-distances. To establish S-Opt$(d,k)$, we must show that $P=\Lambda_d(k)\setminus \Lambda_d(k-2)$. Thanks to our inductive hypothesis, we can assume $P$ determines exactly $k$ distinct $\ell^1$-distances, not fewer, and we denote those $\ell^1$-distances by $\lambda_1<\cdots<\lambda_k$. By the $\ell^1$-sphere component of Conjecture \ref{axisp}, we know that $\lambda_k$ occurs in an axis-parallel direction. By permuting coordinates, we assume that $\lambda_k$ occurs in the last coordinate direction, in other words $(x_1,\dots,x_d),(x_1,\dots,x_d+\lambda_k)\in P$ for some $x_1,\dots,x_d\in \R$ with $|x|+\cdots+|x_d|=|x_1|+\cdots+|x_d+\lambda_k|=k$, which in particular forces $x_d=-\lambda_k/2$ and $|x_1|+\cdots+|x_{d-1}|=k-\lambda_k/2$. This transformation is allowable because our end goal, $\Lambda_d(k)\setminus \Lambda_d(k-2)$, is invariant under coordinate permutation. \noindent We argue (informally for the moment) that the only reasonable choice is $\lambda_k=2k$ and $x_1=\cdots=x_{d-1}=0$, meaning $ke_d,-ke_d\in P$. This is because, since $\lambda_k$ is the largest $\ell^1$-distance, $P$ is contained in the intersection of the closed $\ell^1$-balls of radius $\lambda_k$ centered at $(x_1,\dots,x_{d-1},-\lambda_k/2)$ and $(x_1,\dots,x_{d-1},\lambda_k/2)$, respectively, and this intersection is the $\ell^1$-ball of radius $\lambda_k/2$ centered at $(x_1,\dots,x_{d-1},0)$. However, $P$ is also contained in $S$, so if it is not the case that $\lambda_k=2k$, then $P$ would in fact be contained in the intersection of an $\ell^1$-sphere with a closed $\ell^1$-ball of a smaller radius, which is at most a closed $\ell^1$-hemisphere. The idea that a maximal subset of an $\ell^1$-sphere determining at most $k$ distinct $\ell^1$-distances could actually be contained in a closed $\ell^1$-hemisphere is intuitively suspect, and we return to this issue near the end of the proof. For now, we assume $-ke_d,ke_d\in P$. \noindent We let $H$ denote the closed upper $\ell^1$-hemisphere of $S$, and we establish H-Opt$(d,k)$ along the way. As discussed in (\ref{hyper}), all the points of $(P\cap H)\setminus ke_d$ have $x_d$-coordinates in the list $c_1> c_2 > \cdots > c_k$, where $c_i=k-\lambda_i/2$. For each $c_i$, the points of $H$ with $x_d$-coordinate equal to $c_i$ take the form $(x_1,\dots,x_{d-1},c_i)$ where $|x_1|+\cdots+|x_{d-1}| = k-c_i$, and we refer to the set of such points as $S_i$. With regard to $\ell^1$-distances, $S_i$ is equivalent to an $\ell^1$-sphere in $\R^{d-1}$, centered at the origin with radius $k-c_i$. All $\ell^1$-distances determined by $P\cap S_i$ are at most $2(k-c_i)=\lambda_i$, so $P\cap S_i$ determines at most $i$ distinct $\ell^1$-distances. By our inductive hypothesis, $|P\cap S_i| \leq |\Lambda_{d-1}(i)|-|\Lambda_{d-1}(i-2)|$, with equality holding if and only if the projection $P\cap S_i$ onto the first $d-1$ coordinates is $\ell^1$-similar to $\Lambda_{d-1}(i)\setminus \Lambda_{d-1}(i-2)$, so in particular $\lambda_1,\dots,\lambda_i$ form an arithmetic progression. Since $k-z_i=\lambda_i/2$ and $\lambda_k=2k$, this equality holds for all $1\leq i \leq k$ if and only if $P\cap S_i=\{(x,k-i): x\in \Lambda_{d-1}(i)\setminus \Lambda_{d-1}(i-2)\}$ for all $1\leq i \leq k$. Here we note that if it were not the case that $ke_d,-ke_d\in P$ as previously assumed, then $P\cap S_i$ would be, at most, equivalent to an $\ell^1$-hemisphere in $\R^{d-1}$, in which case our inductive hypothesis would prohibit it from having $|\Lambda_{d-1}(i)|-|\Lambda_{d-1}(i-2)|$ elements. \noindent In summary, \begin{equation}\label{hemiub}|P\cap H| \leq \sum_{i=0}^{k} |\Lambda_{d-1}(i)|-|\Lambda_{d-1}(i-2)|,\end{equation} taking $\Lambda_{d-1}(-1)$ and $\Lambda_{d-1}(-2)$ to be empty, and equality holds if and only if $$P\cap H = \bigcup_{i=0}^k\left\{(x,k-i): x\in \Lambda_{d-1}(i)\setminus \Lambda_{d-1}(i-2)\right\}= \Lambda_d(k)\cap H,$$ which establishes H-Opt$(d,k)$. \noindent Further, $P\cap H$ can contain at most $\sum_{i=0}^{k-1} |\Lambda_{d-1}(i)|-|\Lambda_{d-1}(i-2)|$ points with $x_d>0$. Letting $H'$ denote the closed lower $\ell^1$-hemisphere of $S$, we employ the identical reasoning as above to yield the same upper bound (\ref{hemiub}) on $|P\cap H'|$, with equality holding if and only if $$P\cap H' = \bigcup_{i=0}^{k}\{(x,i-k): x\in \Lambda_{d-1}(i)\setminus \Lambda_{d-1}(i-2)\}=\Lambda_d(k)\cap H.$$ Further, $P\cap H'$ can contain at most $\sum_{i=0}^{k-1} |\Lambda_{d-1}(i)|-|\Lambda_{d-1}(i-2)|$ points with $x_d<0$. Putting all this together, we have \begin{align*}|P|&=|P\cap (H\setminus H')|+|P\cap (H' \setminus H)|+|P\cap H\cap H'| \\&\leq 2\left(\sum_{i=0}^{k-1} |\Lambda_{d-1}(i)|-|\Lambda_{d-1}(i-2)|\right)+\left(|\Lambda_{d-1}(k)|-|\Lambda_{d-1}(k-2)|\right), \end{align*} and equality holds if and only if $$ P= \bigcup_{i=-k}^k\left\{(x,i): x\in \Lambda_{d-1}(k-|i|)\setminus \Lambda_{d-1}(k-|i|-2)\right\} =\Lambda_d(k)\setminus \Lambda_d(k-2).$$ Therefore, S-Opt$(d,k)$ holds, and the induction on dimension is complete. \end{proof} \noindent \textbf{Remark.} If one is specifically interested in dimension $d=3$, then, because we have fully resolved the problem in dimension $d=2$, no inductive hypothesis is needed for dimension, just for the number of $\ell^1$-distances. In other words, if Opt$(3,k-1)$ holds, then $\Lambda_3(k)$ is the unique set, up to $\ell^1$-similarity, of maximal size amongst axis-parallel subsets of $\R^3$ determining at most $k$ distinct $\ell^1$-distances. In particular, because Theorem \ref{singdist} tells us that O$(3,1)$ holds, we know that $\Lambda_3(2)$, which contains $19$ points and is pictured in Figure \ref{fig:sub2}, is uniquely optimal amongst axis-parallel sets determining only two $\ell^1$-distances. However, we cannot make the analogous claim for $\Lambda_3(3)$, because we cannot exclude the possibility of a non-axis-parallel set determining two $\ell^1$-distances that contains more than $19$ points (or that contains exactly $19$ points but is not $\ell^1$-similar to $\Lambda_3(2)$). This possibility disables our bridge from two $\ell^1$-distances to three, and exemplifies the need in assuming Conjecture \ref{axisp} if we wish to glean additional information in dimension $d\geq 3$. \noindent \textbf{Acknowledgements:} This research was initiated during the Summer 2019 Kinnaird Institute Research Experience at Millsaps College. All authors were supported during the summer by the Kinnaird Endowment, gifted to the Millsaps College Department of Mathematics. At the time of submission, all authors except Alex Rice were Millsaps College undergraduate students. The authors would like to thank Alex Iosevich for his helpful references, and Tomasz Tkocz for alerting us to previous research done in the $k=1$ case. Finally, the authors would like to thank the anonymous referee for their encouraging comments and helpful recommendations, particular regarding Sections 5 and 6.
{ "timestamp": "2020-05-26T02:08:51", "yymm": "1911", "arxiv_id": "1911.08067", "language": "en", "url": "https://arxiv.org/abs/1911.08067", "abstract": "We address an analog of a problem introduced by Erdős and Fishburn, itself an inverse formulation of the famous Erdős distance problem, in which the usual Euclidean distance is replaced with the metric induced by the $\\ell^1$-norm, commonly referred to as the $\\textit{taxicab metric}$. Specifically, we investigate the following question: given $d,k\\in \\mathbb{N}$, what is the maximum size of a subset of $\\mathbb{R}^d$ that determines at most $k$ distinct taxicab distances, and can all such optimal arrangements be classified? We completely resolve the question in dimension $d=2$, as well as the $k=1$ case in dimension $d=3$, and we also provide a full resolution in the general case under an additional hypothesis.", "subjects": "Combinatorics (math.CO); Metric Geometry (math.MG)", "title": "Sets in $\\mathbb{R}^d$ determining $k$ taxicab distances", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9911526443943918, "lm_q2_score": 0.8267117962054048, "lm_q1q2_score": 0.8193975829610245 }
https://arxiv.org/abs/1901.06759
Submultiplicativity of the numerical radius of commuting matrices of order two
Denote by $w(T)$ the numerical radius of a matrix $T$. An elementary proof is given to the fact that $w(AB) \leq w(A)w(B)$ for a pair of commuting matrices of order two, and characterization is given for the matrix pairs that attain the quality.
\section{Introduction} Let $M_n$ be the set of $n\times n$ matrices. The numerical range and numerical radius of $A \in M_n$ are defined by $$W(A) = \{x^*Ax: x \in {\mathbb C}^n, x^*x = 1\} \qquad \hbox{ and } \qquad w(A) = \max\{|\mu|: \mu \in W(A)\},$$ respectively. The numerical range and numerical radius are useful tools in studying matrices and operators. There are strong connection between the algebraic properties of a matrix $A$ and the geometric properties of $W(A)$. For example, $W(A) = \{\mu I\}$ if and only if $A = \mu I$; $W(A) \subseteq {\mathbb R}$ if and only if $A = A^*$; $W(A) \subseteq [0, \infty)$ if and only if $A$ is positive semi-definite. The numerical radius is a norm on $M_n$, and has been used in the analysis of basic iterative solution methods \cite{Ax}. Researchers have obtained interesting inequalities related to the numerical radius; for example, see \cite{G,H,Hol,HolS,HJ1}. We mention a few of them in the following. Let $\|A\|$ be the operator norm of $A$. It is known that $$w(A) \le \|A\| \le 2w(A).$$ While the spectral norm is submultiplicative, i.e., $\|AB\| \le \|A\| \|B\|$ for all $A, B \in M_n$, the numerical radius is not. In general, $$w(AB) \le \xi w(A)w(B) \quad \hbox{ for all } A, B \in M_n$$ if and only if $\xi \ge 4$; e.g., see \cite{GW}. Despite the fact that the numerical radius is not submultiplicative, $$w(A^m) \le w(A)^m \qquad \hbox{ for all positive integers } m.$$ For a normal matrix $A\in M_n$, we have $w(A) = \|A\|$. Thus, for a normal matrix $A$ and any $B \in M_n$, $$w(AB) \le \|AB\| \le \|A\| \|B\| = w(A) \|B\| \le 2 w(A)w(B),$$ and also $$w(BA) \le \|BA\| \le \|B\|\|A\| = \|B\| w(A) \le 2w(B)w(A).$$ In case $A, B\in M_n$ are normal matrices, $$w(AB) \le \|AB\| \le \|A\| \|B\| = w(A) w(B).$$ Also, for any pairs of commuting matrices $A, B \in M_n$, $$w(AB) \le 2w(A)w(B).$$ To see this, we may assume $w(A) = w(B) = 1$, and observe that \begin{eqnarray*} 4w(AB) &=& w((A+B)^2 - (A-B)^2) \le w((A+B)^2) + w((A-B)^2) \\ &\le& w(A+B)^2 + w(A-B)^2 \le 8. \end{eqnarray*} The constant 2 is best (smallest) possible for matrices of order at least 4 because $w(AB) = 2w(A)w(B)$ if $A = E_{12} + E_{34}$ and $B = E_{13} + E_{24}$, where $ E_{ij} \in M_n$ has $1$ at the $(i,j)$ position and $0$ elsewhere; see \cite[Theorem 3.1]{GW}. In connection to the above discussion, there has been interest in studying the best (smallest) constant $\xi > 0$ such that $$w(AB) \le \xi w(A)w(B)$$ for all commuting matrices $A, B \in M_n$ with $n \le 3$. For $n = 2$, the best constant $\xi$ is one; the existing proofs of the $2\times 2$ case depend on deep theory on analytic functions, von Neumann inequality, and functional calculus on operators with numerical radius equal to one, etc.; for example, see \cite{Hol,HolS}. Researchers have been trying to find an elementary proof for this result in view of the fact that the numerical range of $A \in M_2$ is well understood, namely, $W(A)$is an elliptical disk with the eigenvalues $\lambda_1, \lambda_2$ as foci and the length of minor axis $\sqrt{({\mathrm tr}\, A^*A) - |\lambda_1|^2 - |\lambda_2|^2}$; for example, see \cite{L,M} and \cite[Theorem 1.3.6]{HJ1}. The purpose of this note is to provide such a proof. Our analysis is based on elementary theory in convex analysis, co-ordinate geometry, and inequalities. Using our approach, we readily give a characterization of commuting pairs of matrices $A, B \in M_2$ satisfying $w(AB) = w(A)w(B)$, which was done in \cite[Theorem 4.1]{GW} using yet another deep result of Ando \cite{An} that a matrix $ A$ has numerical radius bounded by one if and only if $A = (I-Z)^{1/2}C(A+Z)^{1/2}$ for some contractions $C$ and $Z$, where $Z = Z^*$. Here is our main result. \begin{theorem} Let $A, B \in M_2$ be nonzero matrices such that $AB = BA$. Then $w(AB) \le w(A)w(B)$. The equality holds if and only if one of the following holds. \begin{itemize} \item[{\rm (a)}] $A$ or $B$ is a scalar matrix, i.e. of the form $ \mu I_2$ for some $\mu \in{\mathbb C}$. \item[ {\rm (b)}] There is a unitary $U$ such that $U^*AU = {\rm diag}\,(a_1,a_2)$ and $U^*BU = {\rm diag}\,(b_1, b_2)$ with $|a_1| \ge |a_2|$ and $|b_1| \ge |b_2|$. \iffalse \item[ {\rm (c)}] There is a unitary $U$ such that $U^*AU = a R$ and $ U^*BU =b\overline R$ for some $a, b \in {\mathbb C}$ with $|a|= w(A)$, $|b| = w(B)$, $R = ic^2 \gamma I_2 + \sqrt{1- (c\gamma )^2} \begin{pmatrix} c & 2\sqrt{1-c^2} \cr 0 & -c \end{pmatrix}$ for some $\gamma, c \in (0,1]$. \fi \end{itemize} \end{theorem} One can associate the conditions (a) and (b) in the theorem with the geometry of the numerical range of $A$ and $B$ as follows. Condition (a) means that $W(A)$ or $W(B)$ is a single point; condition (b) means that $W(A)$, $W(B)$, $W(AB)$ are line segments with three sets of end points, $\{a_1, a_2\}, \{b_1, b_2\}, \{a_1b_1, a_2b_2\}$, respectively, such that $|a_1|\ge |a_2|$ and $|b_1| \ge |b_2|$. \iffalse condition (c) means that there are $a, b \in {\mathbb C}$ with $(|a|, |b|) = w(A), w(B))$ such that $W(A/a) = {\mathcal E}$ is an elliptical disk symmetric about the imaginary axis insider the unit circle at two points, and $W(B/b) = \overline{{\mathcal E}} = \{\bar\mu \in {\mathbb C}: \mu \in {\mathcal E}\}.$ \fi \section{Proof of Theorem 1} Let $A, B \in M_2$ be commuting matrices. We may replace $(A,B)$ by $(A/w(A),B/w(B))$ and assume that $w(A) = w(B) = 1$. We need to show that $w(AB)\le 1$. Since $AB = BA$, there is a unitary matrix $U \in M_2$ such that both $U^*AU$ and$U^*BU$ are in triangular form; for example, see \cite[Theorem 2.3.3]{HJ2}. We may replace $(A,B)$ by $(U^*AU,U^*BU)$ and assume that $A = \begin{pmatrix} a_1&a_3\cr 0 & a_2\cr \end{pmatrix}$, $B = \begin{pmatrix} b_1&b_3\cr 0 & b_2\cr \end{pmatrix}$ and $w(A) = w(B) = 1$. The result is clear if $A$ or $B$ is normal. So, we assume that $a_3, b_3 \ne 0$. Furthermore, comparing the $(1,2)$ entries on both sides of $AB = BA$, we see that $\displaystyle\frac{a_1-a_2}{a_3}= \displaystyle\frac{b_1-b_2}{b_3}$. Applying a diagonal unitary similarity to both $A$ and $B$, we may further assume that $\gamma =\displaystyle\frac{a_1-a_2}{a_3}\ge 0$. Let $r=\displaystyle\frac{1}{\sqrt{\gamma^2+1}}$. We have $0<r\le 1$. Then $A = z_1 I + s_1C$ and $B = z_2 I + s_2 C$ with \bigskip \centerline{ $z_1 = \displaystyle\frac{a_1+a_2}{2}, \quad z_2 =\displaystyle\frac{b_1+b_2}{2} , \quad s_1=\displaystyle\frac{a_3}{2r}, \quad s_2 =\displaystyle\frac{b_3}{2r}$, \quad and \quad $C = \begin{pmatrix} \sqrt{1-r^2} & 2r \cr 0 & -\sqrt{1-r^2}\end{pmatrix}$.} \medskip\noindent Note that $W(C)$ is the elliptical disk with boundary $$\{\cos\theta + i r\sin\theta: \theta \in [0, 2\pi]\};$$ see \cite {L} and \cite[Theorem 1.3.6]{HJ1}. Replacing $(A, B)$ with $(e^{it_1} A, e^{it_2 } B)$ for suitable $t_1, t_2\in [0,2\pi ]$, if necessary, we may assume that ${\rm Re}\,z_1,\ {\rm Re}\,z_2 \ge 0 $ and $s_1,s_2$ are real. Suppose $z_1 = \alpha_1 + i \alpha_2$ with $\alpha_1\ge 0$ and the boundary of $W(A)$ touches the unit circle at the point $\cos\phi_1+i\sin\phi_1$ with $\phi_1 \in [-\pi/2, \pi/2]$. Then $W(A)$ has boundary $$\{\alpha_1 +| s_1|\cos\theta + i (\alpha_2 + |s_1|r\sin\theta): \theta \in [0, 2\pi]\} .$$ \noindent We {\bf claim} that the matrix $A$ is a convex combination of $A_0 = e^{i\phi_1} I$ and another matrix $A_1$ of the form $A_1 = i(1-r^2)\sin \phi_1I + \xi C$ for some $\xi \in {\mathbb R}$ such that $w(A_1) \le 1$. To prove our claim, we first determine $\theta_1 \in [-\pi/2, \pi/2]$ satisfying $$\cos\phi_1+i\sin\phi_1=(\alpha_1 + |s_1| \cos\theta_1) + i (\alpha_2 + |s_1| r\sin\theta_1).$$ Since the boundary of $W(A)$ touches the unit circle at the point $\cos\phi_1+i\sin\phi_1$, using the parametric equation \begin{equation}\label{para} x+iy = (\alpha_1 + |s_1| \cos\theta) + i (\alpha_2 + |s_1| r\sin\theta), \end{equation} of the boundary of $W(A)$, we see that the direction of the tangent at the intersection point $\cos\phi_1+i\sin\phi_1$ is $- \sin \theta_1 + i r \cos \theta_1$, which agrees with $-\sin\phi_1+i \cos \phi_1 $, the direction of the tangent line of the unit circle at the same point. As a result, we have $$(\cos \theta_1, \sin \theta_1) = \displaystyle\frac{(\cos \phi_1, r\sin\phi_1)}{\sqrt{\cos^2\phi_1 + r^2\sin^2\phi_1}}.$$ Furthermore, since $\cos\phi_1+i\sin\phi_1=(\alpha_1 + |s_1| \cos\theta_1) + i (\alpha_2 + |s_1| r\sin\theta_1)$, we have $$ \alpha_1=\cos\phi_1 - \displaystyle\frac{|s_1|\cos\phi_1}{\sqrt{\cos^2\phi_1 + r^2\sin^2\phi_1}}\ge 0 \quad \hbox{ and } \quad \alpha_2=\sin \phi_1-\displaystyle\frac{|s_1|r^2\sin\phi_1}{\sqrt{\cos^2\phi_1 + r^2\sin^2\phi_1}}. $$ \noindent \bf Assertion. \it If $\hat s_1 = \sqrt{\cos^2\phi_1+r^2 \sin^2 \phi_1}$, then $|s_1|\le \hat s_1$. \rm If $\cos \phi_1>0$, then $\alpha_1=\(1 - \displaystyle\frac{|s_1| }{\hat s_1}\)\cos\phi_1\ge 0$, and hence $|s_1|\le \hat s_1$. If $\cos\phi_1=0$, then $\sin\phi_1=\pm 1$, $\hat s_1=r$ and $(\alpha_1,\alpha_2) = (0,\sin\phi_1 (1-|s_1|r))$ so that the parametric equation of the boundary of $W(A)$ in (\ref{para}) becomes $$x+iy = |s_1| \cos\theta + i ( \sin\phi_1 (1-|s_1|r)\ + |s_1| r\sin\theta)\,.$$ Since $w(A)=1$ and $\sin\phi_1=\pm 1$, for all $\theta\in [0,2\pi)$ , we have \begin{eqnarray*} 0 &\le& 1 - \[(|s_1| \cos\theta)^2+(\sin\phi_1 (1-|s_1|r) + |s_1| r\sin\theta)^2 \]\\ &= &1- \[ |s_1| (1-\sin^2\theta) +(\pm (1-|s_1|r) + |s_1| r\sin\theta)^2 \]\\ &= &1- \[ |s_1|^2 (1-(\pm 1\mp(1\mp\sin\theta))^2) +( 1-|s_1|r(1\mp \sin\theta))^2 \]\\ &= &1- \[ |s_1|^2 ( 2(1\mp\sin\theta) - (1\mp\sin\theta)^2) + 1-2|s_1|r(1\mp \sin\theta)+ |s_1|^2r^2(1\mp \sin\theta)^2 \]\\ &= & 2|s_1|(r-|s_1|)(1\mp\sin\theta)+(1-r^2)|s_1|^2(1\mp \sin\theta)^2 . \end{eqnarray*} Therefore, $(r-|s_1|)\ge 0$, which gives $|s_1|\le r=\hat s_1$. \medskip Now, we show that our {\bf claim} holds with \begin{equation}\label{a0a1} A_0=e^{i\phi_1}I \qquad \hbox{ and } \qquad A_1 = i(1-r^2)\sin\phi_1I+\nu_1 \hat s_1 C, \end{equation} where $\nu_1 =1$ if $s_1\ge 0$ and $\nu_1 =-1$ if $s_1< 0$. Note that $W(A_1)$ is the elliptical disk with boundary $\{ \hat s_1 \cos \theta + i[(1-r^2) \sin\phi_1+ \hat s_1r\sin\theta): \theta \in [0, 2\pi)\}$, and for every $\theta\in [0, 2\pi]$, we have \begin{eqnarray*} &&(\hat s_1\cos \theta)^2+((1-r^2)\sin\phi_1+\hat s_1r\sin \theta)^2\\ &=&\hat s_1^2(1-\sin^2 \theta)+(1-r^2)^2\sin^2\phi_1+\hat s_1^2r^2\sin^2 \theta+2\hat s_1r(1-r^2)\sin\phi_1\sin \theta\\ &=&\hat s_1^2 +(1-r^2)^2\sin^2\phi_1+(1-r^2)r^2\sin^2\phi_1 -(1-r^2)\(\hat s_1^2\sin^2 \theta-2\hat s_1r \sin\phi_1\sin \theta+r^2\sin^2\phi_1\)\\ &=&(\cos^2\phi_1+r^2 \sin^2 \phi_1) +(1-r^2)^2\sin^2\phi_1+(1-r^2)r^2\sin^2\phi_1-(1-r^2)(\hat s_1\sin \theta-r\sin\phi_1)^2 \\ &=&1-(1-r^2)(\hat s_1\sin \theta-r\sin\phi_1)^2\\ &\le& 1. \end{eqnarray*} Therefore, $w(A_1)\le 1$. By the Assertion, $|s_1| \le \hat s_1$. Hence $A=\(1-\displaystyle\frac{|s_1|}{\hat s_1}\)A_0+\displaystyle\frac{|s_1|}{\hat s_1} A_1$ is a convex combination of $ A_0$ and $A_1$. \medskip Similarly, if $W(B)$ touches the unit circle at $e^{i\phi_2}$ with $\phi_2\in[-\pi/2,\pi/2]$, then $B$ is a convex combination of \begin{equation}\label{b0b1} B_0=e^{i\phi_2}I \qquad \hbox{ and } \qquad B_1 = i(1-r^2)\sin\phi_2I+\nu_2\hat s_2 C \end{equation} with $\hat s_2 = \sqrt{\cos^2\phi_2+r^2 \sin^2 \phi_2}$ and $\nu_2\in\{1,-1\}$. Let $U = \begin{pmatrix}-r& \sqrt{1-r^2} \cr \sqrt{1-r^2}& r\end{pmatrix}$. Then $U^*CU=-C$. If $\nu_2 = -1$, we may replace $(A,B)$ by $(U^*AU,U^*BU)$ so that $(\nu_1,\nu_2)$ will change to $(-\nu_1, -\nu_2)$. So, we may further assume that $\nu_2=1$. \medskip By the above analysis, $AB$ is a convex combination of $A_0B_0, A_0B_1, A_1B_0$ and $A_1B_1$. Since $w(e^{it}T)=w(T)$ for all $t\in{\mathbb R}$ and $T\in M_n$, the first three matrices have numerical radius 1. We will prove that \begin{equation} \label{a1b1} w(A_1 B_1) < 1. \end{equation} It will then follow that $w(AB) \le 1$, where the equality holds only when $A=A_0$ or $B=B_0$. \medskip For simplicity of notation, let $w_1=\sin\phi_1$ and $w_2=\sin\phi_2$. Then \begin{equation}\label{hats}\hat s_i=\sqrt{1-(1-r^2)w_i^2}\ \mbox{ for }\ i=1,2. \end{equation} Recall from (\ref{a0a1}) and (\ref{b0b1}) that $A_1 = i(1-r^2)w_1 I +\nu_1 \hat s_1 C$ and $B_1 = i(1-r^2)w_2 I + \hat s_2 C$ because $\nu_2=1$. Since $C^2 = (1-r^2)I_2$, we have $$A_1B_1 = (1-r^2)(u I_2 + iv C),$$ where $$ u = \nu_1 \hat s_1\hat s_2-w_1w_2(1-r^2) \quad\hbox{ and } \quad v = w_1 \hat s_2 +\nu_1w_2\hat s_1. $$ If $r=1$, then $A_1B_1 =0$. Assume that $0<r<1$. We need to show that $$ \frac{1}{1-r^2}w(A_1B_1) = w(uI+ivC) < \displaystyle\frac{1}{(1-r^2)}. $$ Because $W(uI+ivC)$ is an elliptical disk with boundary $\{u+iv(\cos\theta+ir\sin\theta): \theta \in [0, 2\pi]\}$, it suffices to show that $$f(\theta) = |u + iv(\cos\theta + i r\sin \theta)|^2 < \frac{1}{(1-r^2)^2} \quad \hbox{ for all } \ \theta \in [0, 2\pi].$$ Note that \begin{eqnarray*} f(\theta) &=& (u-rv\sin\theta)^2 + (v\cos\theta)^2 \\ &=&u^2-2ruv\sin\theta+r^2v^2\sin^2\theta+v^2(1-\sin^2\theta) \\ &=&\displaystyle\frac{u^2}{1-r^2}+v^2-\(\sqrt{1-r^2}v\sin\theta+\displaystyle\frac{ru}{\sqrt{1-r^2}}\)^2\\ &\le&\displaystyle\frac{u^2}{1-r^2}+v^2\\ &=&\displaystyle\frac{1}{(1-r^2)}\[u^2+(1-r^2)v^2\]\\ &=&\displaystyle\frac{1}{(1-r^2)}\[ (\nu_1 \hat s_1\hat s_2-w_1w_2(1-r^2) )^2+(1-r^2)(w_1 \hat s_2+\nu_1w_2\hat s_1)^2 \]\\ &=&\displaystyle\frac{1}{(1-r^2)}\[ \hat s_1^2\hat s_2^2+w_1^2w_2^2(1-r^2)^2 +(1-r^2) (w_1^2 \hat s_2^2+ w_2^2\hat s_1^2)\]\quad \mbox{because }\nu_1=\pm 1 \\ &=&\displaystyle\frac{1}{(1-r^2)}\[ (\hat s_1^2+ (1-r^2) w_1^2)( \hat s_2^2 +(1-r^2) w_2^2)\]\\ &=&\displaystyle\frac{1}{(1-r^2)}\hskip 3.25in \ \mbox{ by (\ref{hats})} \\ &< &\displaystyle\frac{1}{(1-r^2)^2}\hskip 3.25in \ \mbox{because }0<r<1. \end{eqnarray*} Consequently, we have $w(A_1B_1)<1$ as asserted in (\ref{a1b1}). Moreover, by the comment after (\ref{a1b1}), if $w(AB) = w(A)w(B)$, then $A = A_0$ or $B = B_0$. Conversely, if $A = A_0$ or $B_0$, then we clearly have $W(AB) = w(A)w(B)$. The proof of the theorem is complete. \hfill $\Box$\medskip \bigskip\noindent {\bf Acknowledgment} We would like to thank Professor Pei Yuan Wu, Professor Hwa-Long Gau, and the referee for some helpful comments. Li is an affiliate member of the Institute for Quantum Computing, University of Waterloo, and is an honorary professor of the Shanghai University. His research was supported by USA NSF grant DMS 1331021, Simons Foundation Grant 351047, and NNSF of China Grant 11571220.
{ "timestamp": "2019-03-01T02:02:57", "yymm": "1901", "arxiv_id": "1901.06759", "language": "en", "url": "https://arxiv.org/abs/1901.06759", "abstract": "Denote by $w(T)$ the numerical radius of a matrix $T$. An elementary proof is given to the fact that $w(AB) \\leq w(A)w(B)$ for a pair of commuting matrices of order two, and characterization is given for the matrix pairs that attain the quality.", "subjects": "Functional Analysis (math.FA)", "title": "Submultiplicativity of the numerical radius of commuting matrices of order two", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9910145731051029, "lm_q2_score": 0.8267117962054049, "lm_q1q2_score": 0.8192834377974522 }
https://arxiv.org/abs/2010.01539
A discussion on the approximate solutions of first order systems of non-linear ordinary equations
We develop a one step matrix method in order to obtain approximate solutions of first order systems and non-linear ordinary differential equations, reducible to first order systems. We find a sequence of such solutions that converge to the exact solution. We study the precision, in terms of the local error, of the method by applying it to different well known examples. The advantage of the method over others widely used lies on the simplicity of its implementation.
\section{Introduction} This paper pretends to be a contribution to methods to find the approximate solutions of nonlinear first order equations (or systems) with given initial values of the form \begin{equation}\label{1} {\mathbf y}'(t) = {\mathbf f}({\mathbf y}(t))\,,\qquad {\mathbf y}(t_0)={\mathbf y}_0\,, \end{equation} where the prime represents first derivative with respect to the variable $t$. As many higher order ordinary differential equations either linear or non-linear may be written as a system of the form \eqref{1}, our method to obtain approximate solutions will apply also to these kind of equations. Our approach is based in a generalization of the one-step matrix method developed by Demidovich and other authors \cite{DEM,KOT,HSU} some time ago, valid for systems of the form $\mathbf y'(t)=A(t)\mathbf y(t)$. One clear presentation of this method appears in the textbook by Farkas \cite{FAR} and it is interesting to compare it with some related procedures, see for instance \cite{NEP,NEP1}. It is quite important to remark that, while the Demidovich matrix method is applied to linear equations (with variable coefficients), our matrix method is a generalization to non-linear systems. The advantage that our method may have in comparison of other one step methods lies on its great algorithmic simplicity. In addition, our solutions have a reasonable level of accuracy in few steps and this is workable in a table computer. This method is quite easily programmable and is very suitable for its use with the package Mathematica. It may be also seen as an alternative to Runge-Kutta and Taylor methods due precisely to its simplicity and precision. The objetive of the present study is to obtain approximate solutions of all kind of systems described by non-linear systems of differential equations (although we may include those linear systems with variable coefficients), including those who appear from physics. In the derivation of the present approach, our motivation was rooted in the practice of operational numerical calculus. Nevertheless, we are mainly focused in the mathematical analysis of our method instead of a detailed analysis of the algorithm or CPU times. As for the case of one step Taylor polynomial method, ours shows a great conceptual simplicity. As a consequence, our proposal for obtaining approximate solutions of non-linear systems can be very easily implemented. The presentation of the method to obtain the approximate solutions is introduced in Section 2. Thus, we have a sequence of approximate solutions, which are defined on a given interval of the real line. This sequence converges uniformly to the exact solution on the given interval as is proven in Section 3, where we also discuss a question of order. It is important to remark that we do not impose any periodicity conditions, so that our results are valid either for periodic or for non-periodic solutions. We have applied our method to various examples of two or three dimensional examples. It is also necessary to test the applicability and accuracy of the method on widely used equations and/or systems. To this end, we have used the van der Pol \cite{VDP}, Duffing \cite{DUF}, Lorentz \cite{LOR} equations, a pseudo diffusive equation depending on a parameter, studied in the standard literature \cite{HAL}, an epidemic equation and the predator-prey Lotka-Volterra \cite{LOT,VOL} equation. We have compared the precision of our solutions with the exact solution, whenever this is known. If not, the comparison is based on Runge-Kutta solutions, for a modern presentation, see \cite{BEL,EGG,KAL}. Also with the widely used Taylor method. The analysis of a variety of examples suggest that our method may be more precise for two dimensional systems than for higher dimensional ones, although this is not always exact: we give two examples of three dimensional systems (Lorentz and epidemic equations), for which our method gives different precisions. And it looks like particularly efficient in the case of the pseudo diffusive equation we mentioned earlier, at least when compared with the standard perturbative method studied in \cite{MIC}. We usually obtain precisions in between of those obtained for the Taylor method of third and fourth order (although the precision depends also on the length of subintervals in which is divided the interval in which we are looking for solutions), which is reasonable when we use a table computer. We close this article with concluding remarks and a conjecture on the Li\'enard equation which is motivated by some of our results and confirmed through numerical experiments. \section{A matrix method.} We begin with an initial value problem as given in \eqref{1}. While the function ${\mathbf y}(t)$ is a $\mathbb R^n$ valued function with real variable $t$ of class $C^1$ on the neighbourhood $|t-t_0|\le a$, $\mathbf f(-)$ is a $\mathbb R^n$ valued function with variable in $\mathbb R^n$, which is continuous on $D\equiv \{ \mathbf y \in \mathbb R^n\;\;/\;\; ||\mathbf y - \mathbf y_0||\le d\}$ and satisfies a Lipschitz condition with respect to $\mathbf y$. Needless to say that $a,d$ are positive constants. In relation with the identity \eqref{1}, we shall use either the denomination of ``equation'' or ``system'' indistinctly. In any case, it is well known that the initial value problem \eqref{1} has one unique solution on the interval $|t-t_0|\le \inf(a,d/M)$ with $M:=\sup_D||\mathbf f||$. Our objective is to introduce a generalization of a method of solutions of \eqref{1} proposed in \cite{FAR}. This generalization is based in an iterative procedure of numerical integration for equations of the type $\mathbf f(\mathbf y) =A(\mathbf y) \cdot \mathbf y $, where $A(-)$ is an $n\times n$ matrix. With this choice for the function $\mathbf f(-)$, the differential equation in \eqref{1} has the form \begin{equation}\label{2} \mathbf y'(t) = A(\mathbf y(t))\cdot \mathbf y(t)\,. \end{equation} We assume that the entries of $A(\mathbf y)$ are continuous on $D$. Let us define a uniform partition of the interval $[0,t_N]$ into subintervals $I_k\equiv [t_k,t_{k+1}]$, where $t_k=hk$ with $k=0,1,2,\dots,N$, with $N$ natural and $h=t_N)/N$, $t_N<a$. We have chosen this form of the interval by simplicity, needless to say that if the original interval is somehow else, it always may be transformed into $[0,t_N]$ by a translation. Also, the equal spacing of all subintervals is not strictly necessary, although it simplifies our notation. Conventionally, we may call {\it nodes} to the points $\{t_k\}$. We proceed as follows: On each interval $I_k$, we approximate Equation \eqref{2} by \begin{equation}\label{3} \mathbf y'_{N,k}(t) = A(\mathbf y^*_{N,k}) \cdot \mathbf y_{N,k}(t)\,. \end{equation} At each node, $t_k$, we impose $\mathbf y_{N,k}(t_k) = \mathbf y_{N,k-1}(t_k)$, while $\mathbf y^*_{N,k} \in \mathbb R^n$ is a constant to be determined. This gives the segmentary solution, which has to be of the form $\mathbf y_N(t)\equiv \{\mathbf y_{N,k}(t)\;;\; k=0,1,\dots,N-1\}$. It satisfies \begin{equation}\label{4} \mathbf y'_N(t) = A(\mathbf y^*_N) \cdot \mathbf y_N(t)\,, \end{equation} where $\mathbf y^*_N$ coincides with $\mathbf y^*_{N,k}$ on each of the $k$-th intervals. These segmentary solutions give a sequence of functions $\{\mathbf y_N(t)\}$, $t\in [0,t_N]$, that are approximations to the solution of \eqref{2}. Here, we shall give a method to obtain each of the $\mathbf y_N(t)$ and, in next sections, we shall discuss the properties of the sequence. Then, we proceed to an iterative integration of \eqref{4} as follows: Take the first interval $I_0$ and fix an initial value $\mathbf y(t_0)$. This initial value gives the solution $\mathbf y_{N,0}(t)$ on $I_0$. Thus, we have the value $\mathbf y_{N,0}(t_1)= \mathbf y_{N,1}(t_1)$, which serves as the initial value for the solution on the interval $I_1$. Then, we repeat the procedure in an obvious manner for $I_2$ and so on. For each interval $I_k$ the matrix $A(\mathbf y^*_{N,k})$, which appears in \eqref{3}, is a constant matrix and, therefore, \eqref{3} is a system with constant coefficients. Therefore, the solution of \eqref{3} has the form \begin{equation}\label{5} \mathbf y_{N,k}(t) = \exp \{ A(\mathbf y^*_{N,k})(t-t_k) \} \cdot \mathbf y_{N,k}(t_k)\,,\qquad k=0,1,2,\dots, N-1\,, \end{equation} where we have used the notation $\mathbf y_{N,0}(t_0)= \mathbf y(t_0)$. We determine the numbers $\mathbf y^*_{N,k}$ through the following expression: \begin{equation}\label{6} \mathbf y^*_{N,k} = \mathbf y_{N,k}\left(t_k+\frac h2 \right) = \exp \left\{ A(\mathbf y_{N,k}(t_k))\,\frac h2 \right\} \cdot \mathbf y_{N,k}(t_k)\,, \end{equation} where $k=0,1,2,\dots N-1$. Then, the approximate solution $\mathbf y_N(t)$ gives at $t\in I_k$ and on each of the intervals $I_k$: \begin{equation}\label{7} \mathbf y_N(t) = \exp \left\{ A(\mathbf y^*_{N,k}) (t-t_k) \right\} \cdot \prod_{j=0}^{k-1} \exp\left\{ A(\mathbf y^*_{N,j})\,h \right\} \cdot \mathbf y_0\,. \end{equation} The determination of the exponential of a matrix may often be rather complicated for large matrices. Then, we may use the Putzer spectral formula. This establish that if $A$ is a constant matrix of order $n\times n$ with eigenvalues $\{\lambda_k\}_{k=1}^n$, then, its exponential verifies the following expression: \begin{equation}\label{8} \exp\{ A\,t \} = \sum_{k=1}^n r_k(t)\,P_{k-1}\,, \end{equation} with \begin{equation}\label{9} P_0\equiv I\,,\qquad P_k= \prod_{j=1}^k (A-\lambda_j I)\,, \qquad k=1,2,\dots,n-1\,, \end{equation} where $I$ is the $n\times n$ identity matrix and the coefficients $r_k(t)$ are to be determined through the following first order system of differential equations: \begin{eqnarray}\label{10} r'_1(t) = \lambda_1\,r_1(t)\,, \quad r_1(0)=1\,; \nonumber \\[2ex] r'_k(t) =\lambda_k\,r_k(t) + r_{k-1}(t)\,, \quad r_k(0)=0\,, \end{eqnarray} for $k=2,3,\dots,n$. For simplicity, let us consider the particular case, in which $A(\mathbf y^*_{N,j})$ are matrices of order $2\times 2$. Each of these matrices has two eigenvalues, $\lambda_{1,j}$ and $\lambda_{2,j}$, which may be either different or equal. Let us assume that $\lambda_{1,j} \ne \lambda_{2,j}$. Then, \begin{eqnarray}\label{11} \exp\left\{ A(\mathbf y^*_{N,j})\,h \right\} \nonumber\\[2ex]= \frac{1}{\lambda_{1,j} - \lambda_{2,j}} \left\{ (A(\mathbf y^*_{N,j}) - \lambda_{2,j}\,I ) \exp\{\lambda_{1,j}\,h \} - (A(\mathbf y^*_{N,j}) - \lambda_{1,j}\,I ) \exp\{\lambda_{2,j}\,h \} \right\}\,. \end{eqnarray} On the other hand, when $\lambda_{1,j}=\lambda_{2,j}=\lambda_j$, we have for the exponential \begin{equation}\label{12} \exp\left\{ A(\mathbf y^*_{N,j})\,h \right\} = \exp\{ \lambda_j\,h \}\, \{I + h(A(\mathbf y^*_{N,j}) -\lambda_j\,I)\}\,. \end{equation} We conclude here the crude description of the method. In the sequel, we shall show the convergence of the sequence, $\{\mathbf y_N(t)\}$, of approximate segmentary solutions and shall evaluate the precision of the method. \section{On the convergence of approximate solutions} In the previous Section, we have obtained a set of approximate solutions of the initial value problem on a compact interval of the real line. The question is now, assuming we have obtained by the previous method a sequence of solutions. Does this sequence converges to the exact solution in any reasonable sense as the length of the sub intervals, here called $h$, becomes arbitrarily small. To investigate this possibility is the goal of the present Section. Let us go back to \eqref{4} and rewrite it as \begin{equation}\label{13} \mathbf y'_N(t)= A(\mathbf y_N) \cdot \mathbf y_N(t) +\eta_N\,, \end{equation} so that, \begin{equation}\label{14} \eta_N(t) =(A(\mathbf y^*_N) - A(\mathbf y_N)) \cdot \mathbf y_N(t)\,. \end{equation} Let us add and subtract $A(\mathbf y^*_N) \cdot \mathbf y^*_N$ in the right hand side of \eqref{14}. Then, let us take the supremum norm on the interval $[t-t_0,t+t_0]$ and use the triangle inequality of the norm, so as to obtain the following inequality: \begin{equation}\label{15} ||\eta_N|| \le || (A(\mathbf y^*_N) \cdot \mathbf y^*_N -A(\mathbf y_N) \cdot \mathbf y_N)|| + ||A(\mathbf y^*_N)|| \,||\mathbf y^*_N - \mathbf y_N ||\,. \end{equation} Then, we apply in \eqref{15} the Lipschitz condition with respect to the variable $\mathbf y$ with constant $K>0$. Then, it comes that \begin{equation}\label{16} ||\eta_N|| \le (K +|| A(\mathbf y_N^*)||) \, ||\mathbf y_N^*-\mathbf y_N||\,. \end{equation} On each one of the intervals $I_k$, let us expand $\mathbf y_N(t)$ in Taylor series around $t_k$. We obtain the following inequality: \begin{equation}\label{17} ||\mathbf y_N^* - \mathbf y_N(t)|| \le \frac h2 \,||\mathbf y'_n(t_k)|| \le \frac h2\,\max_{t_0\le t\le t_N} ||\mathbf y'_N(t)||\,. \end{equation} Since $\mathbf y'_N(t)$ is continuous on the interval $t_0 \le t \le t_N$, the maximum in the right hand side of \eqref{17} exists. Furthermore, $A(\mathbf y)$ is continuous with respect to $\mathbf y$ on the neighborhood $||\mathbf y- \mathbf y_0||\le d$, so that there exists a constant $C>0$ such that $||A(\mathbf y)|| \le C$. Taking norms in \eqref{4}, we have that \begin{equation}\label{18} ||\mathbf y'_N|| \le ||A(\mathbf y^*_N)||\,||\mathbf y_N|| \le C\, ||\mathbf y_N||\,. \end{equation} Equation \eqref{7} implies that \begin{equation}\label{19} \max_{t_0 \le t \le t_N} ||\mathbf y_N(t)|| \le C' \exp\{C(t_N-t_0)\}\,, \end{equation} where $C'>0$ is a constant. After (\ref{18}-\ref{19}), we see that $||\mathbf y'(t)||$ is bounded for all $t$ in the interval $[t_0,t_N]$. Consequently after \eqref{16} and (\ref{18}-\ref{19}), we have that \begin{equation}\label{20} ||\eta_N|| \le \frac h2 \, (K+C)\, C' \, \exp\{C(t_N-t_0)\} =S\,h\,, \end{equation} where the meaning of the constant $S>0$ is obvious. Next, le us integrate \eqref{13} on the interval $[t_0,t]$. Since for all value of $N$, we use the same initial value $y(t_0)$, we have \begin{equation}\label{21} \mathbf y_N(t) = \mathbf y(t_0) + \int_{t_0}^t A(\mathbf y_N(s))\cdot \mathbf y_N(s)\,ds + \int_{t_0}^t \eta_N(s)\,ds\,. \end{equation} From \eqref{21}, we have that \begin{equation}\label{22} ||\mathbf y_{N+M}(t) -\mathbf y_N(t)|| \le \int_{t_0}^t ||A(\mathbf y_{N+M}(s))\cdot \mathbf y_{N+M}(s) - A(\mathbf y_N(s)) \cdot \mathbf y_N(s)||\,ds + \delta_N(t)\,, \end{equation} with \begin{equation}\label{23} \delta_N(t) = \int_{t_0}^t ||\eta_{N+M}(s)-\eta_N(s)||\,ds \le 2Sh (t_N-t_0)\,. \end{equation} Using the Lipschitz condition in \eqref{22}, we obtain \begin{equation}\label{24} ||\mathbf y_{N+M}(t) -\mathbf y_N(t)|| \le K \int_{t_0}^t ||\mathbf y_{N+M}(s) -\mathbf y_N(s)|| \, ds + 2Sh (t_N-t_0)\,. \end{equation} At this point, we use the Gronwall lema, which states the following \medskip {\bf Lemma (Gronwall)}.- Let $f(t): I\longmapsto \mathbb R$ an integrable function on the compact real interval $I$ such that there exists two positive constants $A$ and $B$, with \begin{equation}\label{25} 0 \le f(t) \le A + B \int_{t_0}^t f(s)\,ds\,, \qquad t_0 \in I\, \end{equation} for all $t \in I$. Then, \begin{equation}\label{26} f(t) \le A\, e^{B(t-t_0)}\,. \end{equation} \hfill $\blacksquare$ \bigskip Then, we use the Gronwall lema with \begin{equation}\label{27} f(t)\equiv ||\mathbf y_{N+M}(t) -\mathbf y_N(t)|| \,, \quad A\equiv Mh := 2S(t_N-t_0)h\,, \quad B\equiv K\,, \end{equation} to conclude that \begin{equation}\label{28} ||\mathbf y_{N+M}(t) -\mathbf y_N(t)|| \le Mh\, e^{K(t-t_0)} \le [M\,e^{K(t_n-t_0)}]\,h =K'\,h\,, \end{equation} With $K'>0$ a positive constant. Therefore, \begin{equation}\label{29} ||\mathbf y_{N+M}(t) -\mathbf y_N(t)|| \longmapsto 0\,, \end{equation} as $h\longmapsto 0$. Since the space $C^0[t_0,t_N]$ is complete\footnote{Note that $t_N$ is fixed and $N$ just denotes the number of intervals in the partition or equivalently, the length of $h$.}, \eqref{29} implies the existence of a continuous function $\mathbf z(t): [t_0,t_N] \longmapsto \mathbb R$, such that \begin{equation}\label{30} \mathbf z(t) := \lim_{N\to\infty} \mathbf y_N(t)\,, \end{equation} uniformly. Now, we claim that $\mathbf z(t)$ is differentiable on $(t_0,t_N)$. Furthermore, $\mathbf z(t)$ is a solution of the differential equation \eqref{2}. The proof goes as follows: The Lipschitz condition applied to our situation implies that \begin{equation}\label{31} ||A(\mathbf y_{N+M}(t)) \cdot \mathbf y_{N+M}(t) -A(\mathbf y_N(t)) \cdot \mathbf y_N(t)|| \le K\,||\mathbf y_{N+M}(t)-\mathbf y_N(t)||\,, \end{equation} so that $A(\mathbf y_N(t)) \cdot \mathbf y_N(t)$ converges uniformly to $A(\mathbf z(t))\cdot \mathbf z(t)$. In addition, after \eqref{20}, we have that \begin{equation}\label{32} ||\eta_N(t)|| \le S h \le S(t_N-t_0)\,. \end{equation} Recall that $t_N$ is fixed for whatever value of $N$. Then, taken limits in \eqref{21}, we have \begin{eqnarray}\label{33} \mathbf z(t) = \mathbf y(t_0) + \lim_{N\to\infty} \int_{t_0}^t A(\mathbf y_N(s)) \cdot \mathbf y_N(s)\,ds + \lim_{N\to\infty} \int_{t_0}^t \eta_N(s)\,ds \nonumber\\[2ex] = \mathbf y(t_0) + \int_{t_0}^t [ \lim_{N\to\infty} A(\mathbf y_N(s)) \cdot \mathbf y_N(s)]\,ds + \int_{t_0}^t \lim_{N\to\infty}[\eta_N(s)]\,ds\,. \end{eqnarray} In the second integral, we may interchange the limit and the integral due to the uniform convergence of the sequence under the integral to its limit. In the case of the second integral, we have used the Lebesgue convergence theorem, which can be applied here due to \eqref{32}. Since obviously $\lim_{N\to\infty}[\eta_N(s)]=0$, we finally conclude that \begin{equation}\label{34} \mathbf z(t)= \mathbf y(t_0) + \int_{t_0}^t A(\mathbf z(s))\cdot \mathbf z(s) \,ds\,. \end{equation} From \eqref{34}, we conclude the following: \medskip 1.- The function $\mathbf z(t)$ is differentiable in the considered interval. \smallskip 2.- The function $\mathbf z(t)$ is the solution of equation \eqref{2} with initial value $\mathbf z(t_0)=\mathbf y(t_0)$. \bigskip \subsection{A question of order} The expansion into Taylor series on a neighborhood of $t_k$ of the solutions of equations \eqref{2} and \eqref{3} have these forms, respectively: \begin{equation}\label{35} \mathbf y(t_{k+1}) = \mathbf y(t_k) +A(y_k)\,\mathbf y(t_k)\,h +\frac 12 (A^2(\mathbf y(t_k))\cdot \mathbf y(t_k))h^2 + \frac 12 \frac{d}{dt} [A(\mathbf y(t_k)) \cdot \mathbf y(t_k)] h^2 + O(h^3)\,. \end{equation} Taking into account that \begin{equation}\label{36} \frac{d}{dt}\,A(\mathbf y(t)) = \sum_{j=1}^n \frac{\partial}{\partial\,y_j}\, A(\mathbf y(t)) \,\frac{d}{dt}\, y_j(t)\,, \end{equation} where $y_j$ is the $j$-th component of $\mathbf y$, equation \eqref{35} becomes: \begin{eqnarray}\label{37} \mathbf y(t_{k+1}) = \mathbf y(t_k) +A(y_k)\,\mathbf y(t_k)\,h +\frac 12 (A^2(\mathbf y(t_k))\cdot \mathbf y(t_k))h^2 \nonumber\\[2ex] + \frac12 \left( \sum_{j=1}^n \frac{\partial}{\partial\,y_j}\, A(\mathbf y(t_k)) \,\frac{d}{dt}\, y_j(t_k) \right) \cdot \mathbf y(t_k) \,h^2 + o(h^3)\,. \end{eqnarray} On the $k$-th interval, equation \eqref{4} takes te formula \begin{equation}\label{38} \mathbf y_N(t_{k+1}) = \mathbf y_N(t_k) + A(\mathbf y^*_{N_k}) \cdot \mathbf y_N(t_k)\,h +\frac 12\, A^2(\mathbf y^*_{N_k}) \cdot \mathbf y_N(t_k)\,h^2 +o(h^3)\,. \end{equation} Then, we may proceed to expand into Taylor series the matrix $A(\mathbf y^*_{N_k})$ on a neighborhood of $\mathbf y_{N_k}$. Taking into account \eqref{5} and after some simple calculations, we obtain: \begin{equation}\label{39} A(\mathbf y^*_{N_k}) = A(\mathbf y_N(t_k) ) + \sum_{j=1}^n \frac{\partial}{\partial \,y_j}\, A(\mathbf y_N(t_k) ) \,(y_{N,j}(t_k+h/2)- y_{N,j}(t_k))\,, \end{equation} where $y_{N,j}(t)$ is the $j$-th component of $\mathbf y_N(t)$. A first order expansion on the last factor on the right hand side of \eqref{39} gives \begin{equation}\label{40} y_{N,j}(t_k+h/2)- y_{N,j}(t_k) = \frac 12\,\frac{d}{dt}\, y_{N,j}(t_k) \, h + o(h^2)\,, \end{equation} so that using \eqref{40} in \eqref{39}, we have \begin{equation}\label{41} A(\mathbf y^*_{N_k}) = A(\mathbf y_N(t_k) ) + \frac h2 \sum_{j=1}^n \frac{\partial}{\partial \,y_j}\, A(\mathbf y_N(t_k) ) \, \frac{d}{dt} \, y_{N,j}(t_k)\,. \end{equation} Then, we replace \eqref{41} into \eqref{38}, and performing some simple manipulations, taking into account that up to second order in $h$: \begin{equation}\label{42} \frac12 \, A^2(\mathbf y^*_{N_k}) \cdot \mathbf y_N(t_k)\,h^2 \approx \frac 12\,A^2(\mathbf y(t_k)) \cdot \mathbf y_N(t_k)\, h^2\,, \end{equation} we finally obtain that \begin{eqnarray}\label{43} \mathbf y(t_{k+1})= \mathbf y_N(t_k) + A(\mathbf y_N(t_k)) \cdot \mathbf y_N(t_k)\,h +\frac{h^2}{2}\sum_{j=1}^n \,\frac{\partial}{\partial\,y_j} A(\mathbf y_n(t_k))\,\frac{d}{dt}\, y_{N,j}(t_k) \nonumber \\[2ex] + \frac12 \,A^2(\mathbf y_N(t_k)) \cdot \mathbf y_N(t_k)\,h^2 +o(h^3)\,. \end{eqnarray} The advantage that \eqref{43} offers with respect to \eqref{37} is that in \eqref{43} the terms up to second order in $h$ are correctly shown. \subsubsection{Going beyond second order} In (3.1), we have found the solution up to second order in $h$. Would we obtain a better precision for third or higher order keeping at the same time the simplicity of the method? First of all our construction is based on equation \eqref{3}, which is not longer valid if we require an approximation of order higher than two. To fix ideas, let us take ${\mathbf y}(t)=(y_1(t),y_2(t))$ bidimensional for simplicity, a choice which does not affect to our argument. Let us expand $A({\mathbf y}(t))$ around ${\mathbf y}^*(t)=(y_1^*,y_2^*)$ (where we have omitted the sub-indices $N,k$ for simplicity) and take one more term in the Taylor span. The result is \begin{equation}\label{44} A({\mathbf y}(t)) = A({\mathbf y}^*(t)) + \frac{\partial}{\partial y_1}\, A({\mathbf y}^*(t)) (y_1(t)- y_1^*) + \frac{\partial}{\partial y_2}\, A({\mathbf y}^*(t)) (y_2(t)- y_2^*)\,. \end{equation} Then, instead of the approximation \eqref{3}, we have the following, where again we have suppressed the indices $N$ and $k$: \begin{equation}\label{45} {\mathbf y}'(t) = \left[ A({\mathbf y}^*(t)) + \frac{\partial}{\partial y_1}\, A({\mathbf y}^*(t)) (y_1(t)- y_1^*) + \frac{\partial}{\partial y_2}\, A({\mathbf y}^*(t)) (y_2(t)- y_2^*) \right] \cdot {\mathbf y}(t)\,. \end{equation} The solution to be determined is just ${\mathbf y}(t)=(y_1(t),y_2(t))$, which now is a part of the Ansatz \eqref{45}. One possibility to solve this contradiction is to proceed with the following span on each of the $I_k$ intervals: \begin{equation}\label{46} y_1(t)= y_1(t_k) + \frac{\partial y_1}{\partial t}(t_k) \,(t-t_k) +\dots\,, \end{equation} and same for $y_2(t)$. If we use these manipulations in \eqref{3}, we obtain an equation of the type: \begin{equation}\label{47} {\mathbf y'_{N,k}}(t) = G_{N,k}(t) \cdot {\mathbf y_{N,k}}(t) \,, \end{equation} with \begin{equation}\label{48} G_{N,k}(t) = A({\mathbf y^*_{N,k}}(t)) + \frac{\partial }{\partial y_1}\, A({\mathbf y^*_{N,k}}(t)) (y_1(t)-y_1^*) + \frac{\partial }{\partial y_2}\, A({\mathbf y^*_{N,k}}(t)) (y_2(t)-y_2^*) + \dots\,. \end{equation} Obviously, system \eqref{48} is non-autonomic. We have to use a new approximation of $G_{N,k}(t) $ by a constant on the interval $I_k$ and re-start again. As we see, advancing to just one higher order of accuracy destroys the simplicity of the method which is one of its more interesting added values. Therefore, we cannot consider going to higher orders an advantage. It may slightly improve the precision at the price of destroying the efficiency and simplicity of the method. \subsection{Some examples} \begin{itemize} \item{{\bf The van der Pol equation} The van der Pol equation \begin{equation}\label{49} y''(t) +\mu(y^2(t)-1)y'(t) + y(t)=0 \end{equation} is a particular case of the Li\'enard equation, \begin{equation}\label{50} y''(t)+f(y)\,y'(t) +g(y)=0\,. \end{equation} which will be discussed in the Appendix. In the van der Pol equation, we have obviously that $f(y)=\mu(y^2(t)-1)$ and $g(y)\equiv y$. This equation can be easily written in the matrix form \eqref{2} by writing $y_1(t)\equiv y(t)$ and $y_2(t)\equiv y'(t)$, as \begin{equation}\label{51} \left(\begin{array}{c} y'_1(t)\\[2ex] y'_2(t) \end{array} \right) = \left( \begin{array}{cc} 0 & 1 \\[2ex] -1 & \mu(1-y_1^2) \end{array} \right) \left(\begin{array}{c} y_1(t)\\[2ex] y_2(t) \end{array} \right)\,. \end{equation} Our goal is to compare the precision of our method with a reference solution. No explicit solutions to the van der Pol equation \eqref{49} are known, so that we use the Runge-Kutta solution of eight order, $y_{rk}(t)$, as reference solution (alternatively, one may consider a Taylor solution of eighth order, which has a comparable precision). We compare the precision of our method with the precision of the solutions as obtained by a third or fourth order Taylor method. As a measure of the error, we use \begin{equation}\label{52} e_h:= \frac 1N \sum_{j=0}^{N-1} (y_{rk}(t_j)-y(t_j))^2\,, \end{equation} where $y(t)$ is the solution obtained using our method or any other, like the Taylor method. Then, we need to use given values of the parameters and inicial conditions. In Table 1 below, we compre the errors produced by our method as compared with the error given with the use of third and fourth order Taylor method for different values of the interval width $h$ for $T=20$, $\mu=1/2$ and the initial conditions $y(0)=0$ and $y'(0)=2$. \vskip1cm \centerline{$ \begin{array} [c]{cccc} h & {\rm Matrix} & {\rm Taylor\, 3^{\rm rd}} & {\rm Taylor\, 4^{\rm rd}}\\[2ex] 10^{-4} & 2.63\, 10^{-11} & 8.60\,10^{-7} & 6.40 \,10^{-7} \\ 10^{-3} & 2.08\, 10^{-11} & 8.30\, 10^{-9} & 6.45\, 10^{-9} \\ 10^{-2} & 1.46\, 10^{-8} & 9.11\, 10^{-9} & 6.60\, 10^{-11} \\ 10^{-1} & 2.04\, 10^{-5} & 1.17\, 10^{-4} & 1.67\, 10^{-7} \\ 2\, 10^{-1} & 2.30\, 10^{-4} & 2.23\, 10^{-3} & 7.94\, 10^{-6} \\ 5\, 10^{-1} & 8.10\, 10^{-3} & 1.49\, 10^{-1} & 2.75\, 10^{-3} \end{array} $} \medskip TABLE 1.- Values for the error $e_h$ for our matrix method and the Taylor method of orders three and fourth for distinct values of $h$ for the van der Pol equation. \vskip 1cm It is clear that our matrix method has a precision in between those of the third and fourth orden Taylor method. These results are just an example of the results obtain in multiple numerical experiments, we have performed showing essentially the same result. However, Table 1 as well as other numerical experiments show and important tendency: the lower $h$ the better our precision as compared with the precision given by the Taylor method. The reason is clear, the smaller $h$ the bigger the number of operations needed to obtain the approximate solution. Our matricidal method requieres less operations than the Taylor method, so that our precision gets better as $h$ gets smaller. In Appendix B, we give our source code when our method is to be applied in the present case. } \item{{\bf The Duffing equation} This is another second order non linear equation, which has the following form: \begin{equation}\label{53} y''(t) + y(t) + y^3(t)=0\,, \end{equation} where we have omitted the term in the first derivative of the indeterminate, $y'(t)$. For the Duffing equation, there are explicit solutions in terms of the Jacobi elliptic functions. For instance, being given the initial conditions $y(0)=1$ and $y;(0)=0$, we have the following solution: \begin{equation}\label{54} y_e(t) = -i \sqrt{1+k} \,{\rm sn}\,(u;m)\,, \end{equation} where ${\rm sn}\,(u;m)$ denotes the elliptic sine. The arguments in \eqref{49} denote the following: \begin{eqnarray}\label{55} u= \frac 1{\sqrt 2} \, (t^2(1-k) + 2t(c_2-kc_1) + (1-k)c_2^2)\,, \nonumber\\[2ex] m= \frac{1+k}{1-k}\,,\qquad k=\sqrt{1+2c_1}\,. \end{eqnarray} The value of the constants included in \eqref{55} are $c_1= 1.5$ and $c_2= 1.1920055$. The Duffing equation may be written in matrix form, if we write again $y_1(t)\equiv y(t)$ and $y_2(t)\equiv y'(t)$, as \begin{equation}\label{56} \left(\begin{array}{c} y'_1(t)\\[2ex] y'_2(t) \end{array} \right) = \left( \begin{array}{cc} 0 & 1 \\[2ex] -(1+y_1^2) & 0 \end{array} \right) \left(\begin{array}{c} y_1(t)\\[2ex] y_2(t) \end{array} \right)\,. \end{equation} We define the error $e_h$ as in \eqref{52}, where we replace $y_{rk}(t)$ by the exact solution, $y_e(t)$, which does exist in the present case. Also $y(t)$ is the solution for which we want to compare its precision with the exact solution, in our case the solutions obtained by our matrix method as well as the third or fourth order Taylor solutions. The errors produced by each method are given on Table 2 below. \vskip1cm \centerline{$ \begin{array} [c]{cccc} h & {\rm Matrix} & {\rm Taylor\, 3^{\rm rd}} & {\rm Taylor\, 4^{\rm rd}}\\[2ex] 10^{-4} & 1.69\, 10^{-9} & 5.60\,10^{-5} & 4.13 \,10^{-5} \\ 10^{-3} & 1.03\, 10^{-10} & 5.06\, 10^{-7} & 4.24\, 10^{-7} \\ 10^{-2} & 9.61\, 10^{-9} & 2.47\, 10^{-8} & 4.24\, 10^{-9} \\ 10^{-1} & 1.16\, 10^{-5} & 4.92\, 10^{-4} & 4.12\, 10^{-8} \\ 2\, 10^{-1} & 1.18\, 10^{-4} & 6.78\, 10^{-3} & 7.56\, 10^{-6} \\ 5\, 10^{-1} & 82.00\, 10^{-3} & 1.04\, 10^{-1} & 7.73\, 10^{-3} \end{array} $} \medskip \medskip TABLE 2.- Values for the error $e_h$ for our matrix method and the Taylor method of orders three and fourth for distinct values of $h$ for the Duffing equation. \vskip1cm We see that the results are quite similar to those obtained with the van der Pol equation. Similarly, we have made some numerical experiments that confirm these results. } \item{ {\bf The Lorenz equation} The Lorenz equation, which is a model for the study of chaotic systems has been introduced in the study of atmospheric behaviour \cite{LOR}. This equations arises in many problems of physics, where chaoticity is present \cite{HAK,GOR,COU,MIS}. The Lorenz equation is usually written in matrix form and is three-dimensional: \begin{equation}\label{57} \left(\begin{array}{c} y'_1(t) \\[2ex] y'_2(t) \\[2ex] y'_3 (t) \end{array} \right) = \left(\begin{array}{ccc} -a & a & 0 \\[2ex] b-y_3(t) & -1 & 0 \\[2ex] y_2(t) & 0 & -c \end{array} \right) \left(\begin{array}{c} y_1(t) \\[2ex] y_2(t) \\[2ex] y_3 (t) \end{array} \right)\,, \end{equation} $a$, $b$ and $c$ being positive constants. This system is very sensitive to the particular choice of the parameters and of initial values, as may numerical experiments show. Based in these experiments, we have choosed the following values for the parameters, $a=10$, $b=9.996$ and $c=8/3$, the fixed point $(y_1,y_2,y_3) =(4.88808,4.88808,8.996)$ is an attractor. We proceed as in the previous case and compare solutions of the errors \eqref{52} resulting of the use of the Taylor method of order two and this Matrix method using as reference the solution obtained with the Ruge Kutta method of eighth order. We use $T=10$ and the solution with initial values $(y_1(0),y_2(0),y_3(0))= (1,0,1)$. We have obtained a table (Table 3) of errors given in terms of the interval width $h$ as \vskip1cm \centerline{$ \begin{array} [c]{ccccc} h & {\rm Matrix} & \text{Taylor second order} & \text{Taylor third order} & \text{Taylor fourth order}\\[2ex] 10^{-2} & 5.2\,10^{-6} & 6.73\,10^{-5} & 3.6\,10^{-8} & 3.1\, 10^{-10} \\ 10^{-1} & 5.3\,10^{-2} & 7.1 \, 10^{-1} & 1.8\, 10^{-1} & 5.2\, 10^{-2} \\ 2\, 10^{-1} & 3.7 \,10^{-1} & {\rm error} & {\rm error} & {\rm error} \end{array} $} \medskip \centerline{TABLE 3.- Values of the precision $e_h$ in terms of $h$ for $T=10$, $y_1(0)=1$, $y_2=0$ and $y_3(0)=1$.} \vskip1cm The word ``error'' written in two entries in Table 2 means that the error that appears if the interval width is of the order of 0.2, when we use the Taylor method, is incontrollable. We also see that the precision obtained with the matrix method clearly improves the precision by the Taylor method the larger the length of the subintervals. \item{{\bf Neutral damping equation}. This equation has been discussed in the literature \cite{HAL,MIC} and is \begin{equation}\label{58} x''(t) +\varepsilon\, ( x'(t))^2 +x(t)=0\,, \end{equation} where the tilde means derivative with respect to the variable $t$ and $\varepsilon$ is a real parameter. As in previous cases, let us define $y(t):= x'(t)$, so that \eqref{58} can be written in the following form: \begin{equation}\label{59} (-x+\varepsilon y^2)\,dx -y\,dy=0\,. \end{equation} This equation is integrable with integrating factor: \begin{equation}\label{60} \mu(x,y)\equiv e^{2\varepsilon x}\,. \end{equation} It is readily shown that equation \eqref{60} admits the following first integral: \begin{equation}\label{61} f(x,y) = \left[\frac 12 \,y^2 + \frac{1}{4\varepsilon^2}\,(2\varepsilon x -1) \right] e^{2\varepsilon x} \end{equation} If we write \eqref{58} in the standard matrix form as \begin{equation}\label{62} \left(\begin{array}{c} x' \\[2ex] y' \end{array} \right) = \left( \begin{array}{cc} 0 & 1 \\[2ex] -1 & -\varepsilon y \end{array} \right) \left( \begin{array}{c} x\\[2ex] y \end{array} \right)\,, \end{equation} we readily observe that the only fixed point is the origin. It is also the origin the point at which the minimum of the first integral \eqref{61} lies. All orbits are periodic around the origin. In Figure 1, we have depicted the periodic solutions in phase space. Note that, although equation \eqref{58} may look like dissipative, it is not as Figure 1 manifests. Now, let us repeat the comparison between errors given by our matrix method with the errors given by the Taylor method at some low orders. As always, we use \eqref{47} as the definition of the error and the numerical Runge-Kutta solution of eighth order as the reference solution. We obtained the following results, which appear in Table 4, where $h$ is, as always, the distance between two consecutive nodes: \vskip1cm \centerline{$ \begin{array}[c]{cccc} h & {\rm Matrix} & {\rm Taylor\, 3^{\rm rd}} & {\rm Taylor\, 4^{\rm rd}}\\[2ex] 10^{-4} & 2.2\, 10^{-14} & 4.2\, 10^{-14} & 3.2 \,10^{-11} \\ 10^{-3} & 1.9\,10^{-14} & 4.1\,10^{-11} & 3.1\, 10^{-11} \\ 10^{-2} & 9.0\, 10^{-14} & 4.6\, 10^{-13} & 4.0\,10^{-12} \\ 10^{-1} & 1.6\, 10^{-9} & 2.7\, 10^{-7} & 8.2\, 10^{-11} \\ 2.10^{-1} & 2.6\, 10^{-8} & 8.6\, 10^{-6} & 1.0\, 10^{-9} \\ 5.10^{-1} & 1.1 \, 10^{-6} & 7.5\, 10^{-4} & 6.1\, 10^{-6} \end{array} $} \medskip TABLE 4. Values of the error in terms of $h$ for $\varepsilon=0.1$, $T=2\pi$ (see \eqref{47}) and the initial values $x(0)=1$, $ x'(0)=0$. \vskip1cm We observe that the precision of the matrix method is much higher than the precision of the third and forth order Taylor method for a distance between nodes $h\le 0.01$. Moreover, we have to underline that our matrix one step method is much simpler to programming that the Taylor method as the reader can easily convince him/herself using any of these examples. There is another method based in the theory of perturbations in order to obtain approximate solutions to \eqref{53}, which receives the name of Lindstedt-Poincar\'e. It is described in \cite{MIC}. It consists in a series in terms of $\varepsilon$ of the form: \begin{equation}\label{63} x(t) = x_0(t) +\varepsilon\, x_1(t) + \varepsilon^2\,x_2(t) +\dots\,. \end{equation} Coefficients $x_i(t)$ may be obtain iteratively, once we have fixed the initial values. For instance, for $x(0)=1$ and $ x'(0)=0$, we obtain: \begin{eqnarray}\label{64} x_0(t)= \cos\omega t\,,\qquad x_1(t)=\frac 16 \left(-3+4 \cos \omega t- \cos 2\omega t \right)\,,\nonumber\\[2ex] x_2(t)= \frac 13 \left(-2 +\frac{61}{24}\, \cos\omega t -\frac 23 \,\cos 2\omega t + \frac 18\, \cos 3\omega t \right)\,, \nonumber\\[2ex] \omega= 1-\frac 16\,\varepsilon + o(\varepsilon^3)\,. \end{eqnarray} We may also evaluate the error for this perturbative method, which is independent of any division of the interval, in which the solution is considered, into subintervals. This error is $1.1\,10^{-4}$, which is obviously higher to those obtained using any of the numerical method considered. We finish the present example by proposing another approach to an approximate solution. By either method, matrix or Taylor, we obtain on each of the nodes $\{t_k\}$ a value, $x_k$, of the approximate solution. Let us interpolate each interval by means of cubic splines, so that we obtain a segmentary approximation by cubic polynomials \footnote{In \cite{LG}, we have already proposed segmentary cubic solutions and studied their properties, although in \cite{LG} they were not necessarily cubic splines.}. Let us assume that the cubit interpolating solution is $s(t)$, then the error of this solution on an interval $T$, with respect to the exact solution is given by (recall that cubic splines admit first and second continuous derivatives at the nodes) \begin{equation}\label{65} e=\frac 1T \int_0^T (s''(t)+\varepsilon\,[s'(t)]^2+s(t))^2\,dt\,. \end{equation} The form of the error for the Taylor method is also given by \eqref{60}. The resulting errors appear in the following table (Table 5): \vskip1cm \centerline{$ \begin{array}[c]{ccccc} h & {\rm Spline} & {\rm Taylor\, 2^{\rm rd}} & {\rm Taylor\, 3^{\rm rd}} & {\rm Taylor\, 4^{\rm rd}}\\[2ex] 10^{-4} & 7.2\, 10^{-16} & 4.7 \,10^{-16} & 6.7\,10^{-16} & 6.5\,10^{-16}\\ 10^{-3} & 1.8\,10^{-14} & 1.8\,10^{-14} & 1.8\,10^{-14} & 1.8\,10^{-14}\\ 10^{-2} & 1.9\,10^{-10} & 1.9\,10^{-11} & 1.1\,10^{-10} & 1.1\,10^{-10} \\ 10^{-1} & 4.6\,10^{-6} & 6.4\,10^{-6} & 4.6\,10^{-6} & 4.6\,10^{-6} \\ 2.10^{-1} & 6.1\,10^{-5} & 9.8\,10^{-5} & 6.0\,10^{-5} & 6.0\,10^{-5}\\ 5.10^{-1} & 4.8\,10^{-3} & 1.2\,10^{-3} & 4.6\,10^{-3} \ & 4.9\,10^{-3} \end{array} $} \medskip TABLE 5.- Comparison between errors by the cubic spline and Taylor methods. \vskip1cm Finally, for the perturbative method the error obtained is $7.4\,10^{-5}$. Concerning the conservation of the constant of motion, we measure its dispersion by means of the following parameter, $e_f$, defined as \begin{equation}\label{66} e_f:= \frac 1T \int_0^T(f(x_0,y_0)-f(x,y))^2\, dt\,, \end{equation} where $f(x,y)$ should be calculated using different approximations such as Taylor, matrix and the analytic approximate solution as obtained by the perturbative Lindstedt-Poincar\'e method mentioned earlier. The point $(x_0,y_0)$ gives the chosen initial conditions. In the latter case, we have obtained $e_f = 1.1\,10^{-4}$, in all others, we always got $e_f<10^{-8}$. \begin{figure} \centering \includegraphics[width=0.5\textwidth]{StreamPlot} \caption{\small Periodic orbits around the origin for the neutral damping equation. The horizontal coordinate represents the values of $x$, while vertical coordinate gives the values of $y=x'$. Note that these periodic orbits are represented in phase space. \label{Figure1}} \end{figure} } \item{{\bf An epidemic equation} A model for an epidemia has been proposed as early as in 1927 \cite{KK,STR,TC}. If $x(t)$, $y(t)$ and $z(t)$ are the number, at certain time $t$, of healthy, sick and dead persons, respectively, in some society, the model assumes that these functions satisfy the following non-linear system: \begin{eqnarray}\label{67} \dot x(t) &=& -a x(t)\,y(t)\,, \nonumber\\[2ex] \dot y(t) &=& a x(t)\,y(t) -b y(t)\,, \nonumber\\[2ex] \dot z(t) &=& by(t)\,, \end{eqnarray} where $a$ and $b$ are positive constants and the dot means derivation with respect to time $t$. The model assumes infection of healthy persons from sick persons. The latter died after some time. Note that, in the studied time interval, the sole cause of population dynamics is the epidemic. For obvious reasons, we consider only positive solutions. The fixed points for \eqref{67} have the form $(\alpha,0,\beta)$ with $\alpha,\beta>0$. Let us consider the following vector field, also called the {\it flux} of \eqref{62}, \begin{equation}\label{68} F(t):=(-a x(t)\,y(t),a x(t)\,y(t) -b y(t),by(t)) \,. \end{equation} Note that $F_x<0$. For $x(t)<b/a$, we have $F_y<0$, while $x>b/a$, it results that $F_y>0$. Thus, the fixed point is inestable if $\alpha >b/a$. This is a necessary condition for the existence of the epidemic. If $\alpha <b/a$, then the fixed points are asymptotically stable as depicted in Figure 2, where we have chosen $a=0.0005$ and $b=0.1$. Different curves in Figure 2 represent different initial conditions. \begin{figure} \centering \includegraphics[width=0.5\textwidth]{FIXEDPOINTS} \caption{\small Asymptotic stability of fixed points with $a=0.0005$ and $b=0.1$ in \eqref{67}. The horizontal line represent the scaled number of healthy people, while the vertical figures give the scaled number of sick people. The arrow means the direction of time. Different curves are obtained using different initial conditions. Observe the presence of a maximum at a point which is independent on the initial conditions. \label{Figure2}} \end{figure} Let us write \eqref{67} in matricial form as \begin{equation}\label{69} \dot X(t) =A(x,y,z)\cdot X(t)\,, \end{equation} where, \begin{equation}\label{70} X(t):= \left(\begin{array}{c} x(t)\\[2ex] y(t)\\[2ex] z(t) \end{array} \right)\,, \qquad A(x,y,z):= \left( \begin{array}{ccc} 0 & -ax & 0 \\[2ex] 0 & ax-b & 0 \\[2ex] 0 & b & 0 \end{array} \right)\,. \end{equation} In Figure 3, we obtain the curve giving the total number of infected people with time. After a maximum, the number of sick people decays quickly. We have used the values for the parameters $a=0.0005$ and $b=0.1$ and the initial values $x(0)=300 >b/a$, $y(0)=20$, $z(0)=0$. \begin{figure} \centering \includegraphics[width=0.5\textwidth]{SICK} \caption{\small Number of infected people in relation with time. Observe the existence of a maximum. After the maximum the curve decreases steeply. \label{Figure3}} \end{figure} The form of the error for the method is obtained using \eqref{65} again. Next in Table 6, we compare the errors for given values of $h$ of our Matrix Method as compared to second and third order Taylor. \vskip1cm \centerline{$ \begin{array}[c]{cccc} h & \text{Matrix Method} & {\rm Taylor\, 2^{\rm rd}} & {\rm Taylor\, 3^{\rm rd}} \\[2ex] 0.01 & 2.2\, 10^{-11} & 4.0 \, 10^{-11} & 1.9\, 10^{-11} \\ 0.1 & 2.0\, 10^{-8} & 7.3 \, 10^{-8} & 2.0 \, 10^{-11} \\ 1.0 & 2.1\, 10^{-4} & 7.2 \, 10^{-4} & 1.7\,10^{-7} \\ 2.0 & 3.6\, 10^{-3} & 1.1\, 10^{-2} & 1.2\, 10^{-5} \end{array} $} \medskip TABLE 6.- Comparison between the errors for the Matrix Method and the Taylor method in the case of the epidemic model. \vskip1cm We see that the level of error is similar, although our method keeps the advantage of needing much less arithmetics than any Taylor method. } \item{{\bf Lotka-Volterra equation.} Models in population dynamics, as for instance the predator-prey competition, were independently developed by the american biologist A.J. Lotka \cite{LOT} and the Italian mathematician V. Volterra \cite{VOL}, see modern references for the Lotka-Volterra equation in \cite{FAR,CI,MUR,LV}. The most general form of the Lotka-Volterra equation has the form \begin{eqnarray}\label{71} \dot x_1= x_1(\varepsilon_1 -a_{11}\,x_1 - a_{12}\, x_2)\,, \nonumber\\[2ex] \dot x_2 =x_2(\varepsilon_2 -a_{12}\,x_1 - a_{22}\, x_2)\,, \end{eqnarray} where $x_1$ and $x_2$ are functions of time $t$ and $\varepsilon_i$, $a_{ij}$, $i,j=1,2$ are constants. Following \cite{CI}, we consider here a simpler version, yet non-linear, of \eqref{71}, which is \begin{eqnarray}\label{72} \dot x= a\,x-b\,xy\,,\nonumber\\[2ex] \dot y = d\,xy - c\,y\,, \end{eqnarray} and the initial conditions $x(t_0)=x_0$, $y(t_0)= y_0$. We may consider the solutions, $x=x(t)$, $y=y(t)$ of \eqref{72} as the equations determining a parametric curve on the $x-y$ plane. Then, by elimination of $t$ and integration, we obtain: \begin{equation}\label{73} C(x,y)= x^cy^a \,\exp\{-(b+d)x\}\,. \end{equation} Note that $C(x,y)$ is a constant on each curve solution (constant of motion) and its value over each solution is determined via the initial conditions. Equations \eqref{71} have two fixed points, which are $(0,0)$ and $(c/d,a/b)$. For $x>0$ and $y>0$ all solutions are periodic. To apply the method proposed in the present article to this system, let us write it in matrix form as \begin{equation}\label{74} \dot X =A(x,y)\,X\,, \end{equation} where, \begin{equation}\label{75} X(t)\equiv \left(\begin{array}{c} x(t)\\[2ex] y(t) \end{array} \right)\,,\qquad A\equiv \left(\begin{array}{cc} a-by & 0 \\[2ex] 0 & dx-c \end{array} \right)\,. \end{equation} In order to estimate the precision of the method, we need to choose values for the initial values as well as for the parameters $a$, $b$, $c$ and $d$. We have use several choices and obtain in all of them similar values. to show a table comparing the precision of our method with some others, let us choose as the values of the parameters, $a=1.2$, $b=0.6$, $c=0.8$and $d=0.3$. As the initial values, let us choose, $x_0=y_0=3$. In addition, we have to take an integration time, in our case we took $T=20$, which approximately accounts for three periods. In order to determine the error produced by our matrix method is determined by the formula \eqref{65} above. We compare this error with those of second and third order Taylor method as compared to the numerical solution obtained by a forth order Runge-Kutta. These errors are shown in Table 7. \vskip1cm \centerline{$ \begin{array}[c]{cccc} h & \text{Matrix Method} & {\rm Taylor\, 2^{\rm rd}} & {\rm Taylor\, 3^{\rm rd}} \\[2ex] 0.01 & 4.1\, 10^{-8} & 3.5 \, 10^{-8} & 8.2\, 10^{-13} \\ 0.1 & 3.5\, 10^{-8} & 5.4 \, 10^{-8} & 2.3 \, 10^{-9} \\ 0.2 & 2.6\, 10^{-3} & 5.4 \, 10^{-3} & 1.0\,10^{-5} \end{array} $} \medskip TABLE 7.- Comparison between the errors for the Matrix Method and the Taylor method in the case of the Lotka-Volterra equation. \vskip1cm We see that the precision of our method is equivalent to those of a second order Taylor, with much less arithmetic operations. } \item{{\bf On the possibility of extending the method to PDE: The Burger's equation.} Can we extend this precedent discussion to partial differential equations admitting separation of variables? One possible example would have been the convection diffusion equation in one spatial dimension \cite{HWX}: \begin{equation}\label{76} \frac{\partial}{\partial t} u(x,t)= \frac{\partial}{\partial x} \left[ D(u)\,\frac{\partial}{\partial x}\, u(x,t) \right] +Q(x,u)\, \frac{\partial}{\partial x}\,u(x,t) + P(x,u)\,. \end{equation} Separation of variables for \eqref{76} is discussed in \cite{HWX}. In general, our method is not applicable here since the resulting equations after separation of variables are not of the form \eqref{1}. Nevertheless, another point of view is possible. Assume we want to obtain approximate solutions of an equation of the type \eqref{76} on the interval $[0,X]$ under the conditions $u(0,t)=0$, $u(X,t)=0$, $u(x,0)=h(x)$, $h(x)$ being a given smooth function and $u(0,0)=u(X,0)=0$. On the interval $[0,X]$, we define a uniform partition of width $h:=X/n$ and nodes $x_k=kh$, $k=0,1,\dots,n$. One may propose one discretization of the solution of the form $u_k(t):= u(x_k,t)$, $k=0,1,\dots,n$. Then, the second derivative in \eqref{76} could be approximated using finite differences \cite{KIN}: \begin{equation}\label{77} \frac{\partial^2}{\partial x^2}\,u(x_k,t) = \frac{u(x_k-1,t)-2u(x_k,t) + u(x_{k+1},t)}{h^2}\,, \end{equation} while for the first spatial derivative, we have \begin{equation}\label{78} \frac{\partial}{\partial x}\,u(x_k,t) = \frac{u(x_k+1,t) - u(x_{k-1},t)}{2h}\,. \end{equation} with $k=1,2,\dots,n-1$, so that equation \eqref{76} takes the form \begin{equation}\label{79} \frac{d}{dt}\,U(t) =F(U(t))\,, \end{equation} where $F$ is a square matrix of order $n-1$ and $U(t)= (u_1(t),u_2(t),\dots,u_{n-1}(t)$ with $u_k(t):= u(x_k,t)$, $k=1,2,\dots,n-1$ and initial values $U(0)=(u(x_1,0), u(x_2,0),\dots,u(x_{n-1},0))$ and initial value $u(x,0)=h(x)$. If it were possible to write equation \eqref{79} in the form: \begin{equation}\label{80} \frac{d}{dt}\,U(t) = A(U(t)) \cdot U(t)\,, \end{equation} then, we would be able to apply our method to find an approximate solution of \eqref{80} on a given finite interval. This property is not fulfilled by the general convection diffusion equation \eqref{76}. However, it is satisfied by a particular choice of this type of parabolic equations: {\it the non-linear Burger's diffusion equation}, which is \cite{BUR,BA}: \begin{equation}\label{81} \frac{\partial}{\partial t}\, u(x,t) = \frac{\partial^2}{\partial x^2}\, u(x,t) - u(x,t)\, \frac{\partial}{\partial x}\, u(x,t)\,. \end{equation} We choose $[0,1]$ as the integration interval for the coordinate variable, so that $X=1$ as above. For the time variable, we use $t\in [0,1]$. This choice is made just for simplicity and as an example to implement our numerical calculations. We use the finite difference method, where the second derivative and first derivatives are replaced as in \eqref{77} and \eqref{78}, respectively. Using \eqref{77} and \eqref{78} in \eqref{81}, we have for each of the nodes the following recurrence relation: \begin{eqnarray}\label{82} \frac{d}{dt}\, u(x_k,t) = \frac{1}{h^2} \left( \left( 1-\frac h2\, u(x_k,t) \right) u(x_{k-1},t) -2 u(x_k,t) + \left( 1 + \frac h2 \, u(x_k,t) \right) u(x_{k+1},t) \right)\,, \end{eqnarray} for $k=1,2,\dots,n-1$. When written expressions \eqref{82} in matrix form, we obtain a matrix equation of the form \eqref{6} and, then, suitable for applying our method. In our numerical realization, we have used a small number of nodes, to begin with, say $n=5$. Then, we integrate on the time variable, on the interval $t\in [0,1]$, using $h_t:=1/m$, where $m$ is a given integer, as the distance between time nodes. Then, we compare our solution with the solution of \eqref{81} given by the sentence NDSolve provided by the Mathematica software, solution that we denote as $v(x,t)$. Then, we can estimate the error of our solution as compared with $v(x,t)$. This is given by \begin{equation}\label{83} {\rm error}:= \frac 1m \sum_{j=0}^m \sum_{k=1}^{n-1} (u(x_k,t_j)-v(x_k,t_j))^2\,, \end{equation} where $t_j:=jh_t$ for all $j=0,1,2,\dots,m$. To estimate the error, we have to give values to $m$. For $m=10$, the error is $1.07\, 10^{-4}$. A similar result can be obtained with $m=20$ or even higher, so that $m=10$ gives already a reasonable approximation. The solution for $t=1$ is nearly zero, which one may have expected taking into account that equation \eqref{80} describes a dissipative model. Thus, our approximate solution may be considered satisfactory also in this case. A few more words on the comparison of our solution and the solution using the Euler method, performed through NDSolve. First of all, we use the explicit Euler method, for which the local error is $o(h^2_t)$, through the option ``Method $\to$ ``ExplicitEuler'', ``StartingStepSize'' $\to$ $1/100$'', since for $h_t>0.01$ the result is unstable. Once we have done the spatial discretization, we integrate \eqref{82} with respect to time. The instability often appears when one uses an explicit method of spatial and time discretization for parabolic PDE \cite{CLW}. This instability comes after the errors due to the arithmetic calculations and are amplified after time integration. Then, let us consider $n=5$, where $n$ is the number of spatial nodes, $0\le t \le 1$ and $m=100$, so that the time integration interval becomes $h_t=0.01$. Then, the error \eqref{83} is $3.40\, 10^{-4}$. We may improve time integration using Euler mid-point integration. In this case, one uses the option ``Method $\to$ ``ExplicitMidpoint'', ``StartingStepSize'' $\to$ 1/10''. Choosing $m=10$, we obtain an error of $1.76\, 10^{-4}$, so that $h_t=0.1$. This error is of the order given by our method. Just recalling that the local error given by the mid-point Euler method is of the order of $o(h_t^3)$. In addition, we may compare the precision of our method and both Euler methods mentioned above by the errors obtained using the third and forth order Adams method (see Chapter III in \cite{HNW}), which for $h_t=0.1$ are respectively given by $1.76\, 10^{-4}$ and $1.74\,10^{-4}$. This error has always been obtained using \eqref{83}, where $u(x,t)$ is our solution and $v(x,t)$ is the solution given by either Euler, Euler mid-point or Adams methods. Nevertheless, the Adams method is a multi-step method while ours is one step method, which means that ours is more easily programmable. } } \end{itemize} \section{Concluding remarks} In the present paper, we have generalized a one step integration method, which has been developed in the seventies of last century by Demidovich and some other authors \cite{DEM,KOT,HSU}. While Demidovich and others restrict themselves to the search for approximate solutions of linear systems of first order differential equations, albeit with variable coefficients, we propose a way to extend the ideas of the mentioned authors to non-linear systems. The solutions we have obtained have a similar degree of precision than those proposed in \cite{DEM,KOT,HSU}. In our method, we obtain a sequence of approximate solutions on a finite interval of ordinary differential equations, and we have proven that this sequence converges uniformly to the exact solution. In order to obtain each approximate solution, we have divided the integration interval into subintervals of length $h$. The sequence of approximate solutions can be obtained after successive refinements of $h$. For each approximate solution, characterized by a value of $h$, we have determined the local error up to $o(h^3)$. It is certainly true, that our one step matrix method does not improve the precision obtained with the fourth order Runge-Kutta method or other equivalent. However, the great advantage of the proposed method with any other is that its algorithm of construction, through the exponential matrix as described in Section 2, is much simpler than their competitors. Simplicity that is inherited from its antecesor the Demidovich method \cite{DEM,KOT,HSU}. This paper is somehow the continuation of previous research of the authors in the same field \cite{GL1,GLP}. We have added some examples of the application of the method, on where we have performed numerous numerical test on the precision of the method, which lies between second and third Taylor method. We have used the sofware Mathematica to implement these numerical tests. \section*{Acknowledgements} We acknowledge partial financial support to the Spanish MINECO, grant MTM2014-57129-C2-1-P, and the Junta de Castilla y Le\'on, grants VA137G18, BU229P18 and VA057U16. We are grateful to Prof L.M. Nieto (Valladolid) for some useful suggestions. \section{Appendix A: A conjecture relative to the Li\'enard equation.} As is well known, the Li\'enard equation has the following form: \begin{equation}\label{84} y''(t)+f(y)\,y'(t) +g(y)=0\,. \end{equation} Let us assume that the function $g(y)$ is a product of some function, that we also call $g(y)$ for simplicity, and $y$, so that equation \eqref{67} takes the form: \begin{equation}\label{85} y''(t)+f(y)\,y'(t) +g(y)\,y(t)=0\,. \end{equation} This second order equation may be easily transformed into a two dimensional first order equation ($y(t)=y(t)$, $z(t)=y'(t)$): \begin{equation}\label{86} \left( \begin{array}{c} y' \\[2ex] z' \end{array} \right) = \left( \begin{array}{cc} 0 & 1 \\[2ex] -g(y) & -f(y) \end{array} \right) \left( \begin{array}{c} y \\[2ex] z \end{array} \right) = A \left( \begin{array}{c} y \\[2ex] z \end{array} \right)\,, \end{equation} where the meaning of the matrix $A$ is obvious. Its eigenvalues are \begin{equation}\label{87} \lambda_\pm(y) = -\frac12 \left(f(y) \pm \sqrt{f^2(y)-4g(y)}\,\right)\,. \end{equation} The conjecture is the following: A sufficient condition for the solutions of \eqref{67} to be bounded for $t>0$ is that the following three properties hold simultaneously: i.) The discriminant in \eqref{60} be negative, i.e., $f^2(y)-4g(y)<0$; ii.) The function $f(y)$ be non-negative, $f(y)\ge 0$ and iii.) The function $g(y)$ is positive and smaller than one, $0< g(y)<1$. This conjecture is based in the form of equations \eqref{11} and \eqref{12} and we have tested it in several numerical experiments. Two more comments in relation to the Li\'enard equation: 1.- The vector field associated to the equation is given by $J\equiv(z,-g(y)y-f(y)z)$, for which the divergence is ${\rm div}\, (J)=-f(y) \le 0$ if $f(y)\ge 0$. Therefore, if $f(y)$ is non-negative the divergence is always negative, so that the origin $(0,0)$ is an attractor. We conjecture that this attractor is also global. 2.- If we take $f(y)\equiv 0$, then \eqref{72} represents a Hamiltonian flow with Hamiltonian given by \begin{equation}\label{88} H(y,z)=\frac 12 \,z^2+V(y)\,, \end{equation} with \begin{equation}\label{89} V(y)=\int_0^y ug(u)\,du \end{equation} Under the assumption $g(y)>0$, the derivative $V'(y)=yg(y)$ is positive for $y>0$ and negative for $y<0$. Thus the only critical point is $y^*=0$, at this point $V'(0)=0$ and $V''(0)=g(0)>0$, so that all the orbits are closed, hence periodic. \section{Appendix B: Source code to use our method in the van der Pole equation} \begin{figure} \centering \includegraphics[width=1.4\textwidth]{S_C.jpeg} \caption{\small Source Code used when applying the method to our example using the van der Pole equation. \label{Figure5}} \end{figure} \vfill\eject
{ "timestamp": "2021-03-12T02:07:33", "yymm": "2010", "arxiv_id": "2010.01539", "language": "en", "url": "https://arxiv.org/abs/2010.01539", "abstract": "We develop a one step matrix method in order to obtain approximate solutions of first order systems and non-linear ordinary differential equations, reducible to first order systems. We find a sequence of such solutions that converge to the exact solution. We study the precision, in terms of the local error, of the method by applying it to different well known examples. The advantage of the method over others widely used lies on the simplicity of its implementation.", "subjects": "Numerical Analysis (math.NA)", "title": "A discussion on the approximate solutions of first order systems of non-linear ordinary equations", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9857180677531123, "lm_q2_score": 0.8311430436757312, "lm_q1q2_score": 0.8192727150384824 }
https://arxiv.org/abs/1701.07963
Negative (and Positive) Circles in Signed Graphs: A Problem Collection
A signed graph is a graph whose edges are labelled positive or negative. The sign of a circle (cycle, circuit) is the product of the signs of its edges. Most of the essential properties of a signed graph depend on the signs of its circles. Here I describe several questions regarding negative circles and their cousins the positive circles. Topics include incidence between signed circles and edges or vertices, characterizing signed graphs with special circle properties, counting negative circles, signed-circle packing and covering, signed circles and eigenvalues, and directed cycles in signed digraphs. A few of the questions come with answers.
\section*{Introduction}\label{intro} A signed graph is a graph with a \emph{signature} that assigns to each edge a positive or negative sign. To me the most important thing about a signed graph is the signs of its circles,\footnote{A circle is a connected, 2-regular graph. The common name ``cycle'' has too many other meanings.} which are calculated by multiplying the signs of the edges in the circle. Thus a signature is essentially its list of negative circles, or (of course) its list of positive circles. I will describe some of the uses of and questions about circles of different signs in a signed graph. Both theorems and algorithms will be significant. The topic of this report is broad. Of necessity, I will be very selective and arbitrarily so, omitting many fine contributions. (Let no one take offense!) I chose this topic in part because it has many fine open problems, but especially in honor of our dear friend Dr.~B.~Devadas Acharya---``our'' because he was the dear friend of so many. Among Dr.~Acharya's wide combinatorial interests, I believe signed graphs were close to his heart, one of his---and his collaborator and wife's, Prof.~Mukti Acharya's---first and lasting areas of research. Circles (or ``cycles'') in signed graphs exemplify well Dr.~B.~D.~Acharya's approach to mathematics, that new ideas and new problems are its lifeblood. He himself was an enthusiastic and inspiring font of new ideas. I hope some of his spirit will be found in this survey. \section{Groundwork}\label{ground} \subsection{Signed graphs}\ A \emph{signed graph} $\S = (\G,\s) = (V,E,\s)$ is defined as an \emph{underlying graph} $\G=(V,E)$, also written $|\S|$, and a signature $\s: E \to \{+,-\}$ (or $\{+1,-1\}$), the sign group. The sets of positive and negative edges are $E^+(\S)$ and $E^-(\S)$. In the literature $\G$ may be assumed to be simple, or it may not (this is graph theory); I do not assume simplicity. Each circle and indeed each walk $W = e_1e_2\cdots e_l$ has a sign $\s(W) := \s(e_1)\s(e_2)\cdots\s(e_l)$. $\S$ is called \emph{balanced} if every circle is positive. Two important signatures are the all-positive one, denoted by $+\G=(\G,+)$, and the all-negative one, $-\G=(\G,-)$, where every edge has the same sign. In most ways an unsigned graph behaves like $+\G$, while $-\G$ acts rather like a generalization of a bipartite graph. In particular, in $+\G$ every circle is positive. In $-\G$ the even circles are positive while the odd ones are negative, so $-\G$ is balanced if and only if $\G$ is bipartite. Signed graphs and balance were introduced by Frank Harary\footnote{Signed graphs, like graphs, have been rediscovered many times; but Harary was certainly the first. K\"onig \cite[Chapter X]{Konig} had an equivalent idea but he missed the idea of labelling edges by the sign group, which leads to major generalizations; cf.\ \cite[Section 5]{BG1}.} in \cite{NB} with this fundamental theorem: \begin{thm}[Harary's Balance Theorem]\label{T:balance} A signed graph $\S$ is balanced if and only if there is a bipartition of its vertex set, $V = X \cup Y$, such that every positive edge is induced by $X$ or $Y$ while every negative edge has one endpoint in $X$ and one in $Y$. Also, if and only if for any two vertices $v,w$, every path between them has the same sign. \end{thm} A \emph{bipartition} of a set $V$ is any pair $\{X,Y\}$ of complementary subsets, including the possibility that one subset is empty (in which case the bipartition is not, technically, a partition). I call a bipartition of $V$ as in the Balance Theorem a \emph{Harary bipartition} of $V$. The Harary bipartition is unique if and only if $\S$ is connected; if $\S$ is also all positive (all edges are positive), then $X$ or $Y$ is empty. Harary later defined $\S$ to be \emph{antibalanced} if every even circle is positive and every odd circle is negative; equivalently, $-\S$ is balanced \cite{SDual}. (The negative of $\S$, $-\S$, has signature $-\s$.) A basic question about a signed graph is whether it is balanced; in terms of our theme, whether there exists a negative circle. If $\S$ is unbalanced, any negative circle provides a simple verification (a \emph{certificate}) that it is unbalanced, since computing the sign of a circle is easy. The Balance Theorem tells us how to provide a certificate that $\S$ is balanced, if in fact it is; namely, one presents the bipartition $\{X,Y\}$, since any mathematical person can easily verify that a given bipartition is, or is not, a Harary bipartition. What is hard about deciding whether $\S$ is balanced is to find a negative circle out of the (usually) exponential number of circles, or a Harary bipartition out of all $2^{n-1}$ possible bipartitions. Fortunately, there is a powerful technique that enables us to quickly find a certificate for imbalance. \emph{Switching} $\S$ consists in choosing a function $\zeta: V \to \{+,-\}$ and changing the signature $\s$ to $\s^\zeta$ defined by $\s^\zeta(e_{vw}) := \zeta(v)\s(e_{vw})\zeta(w)$. The resulting switched signed graph is $\S^\zeta := (|\S|,\s^\zeta)$. It is clear that switching does not change the signs of circles. Let us denote by $\cC(\S)$ the set of all circles of a signed graph (and similarly for an unsigned graph) and by $\cC^+(\S)$ or $\cC^-(\S)$ the set of all positive or, respectively, negative circles. Thus, $\cC^+(\S^\zeta) = \cC^+(\S)$. There is a converse due to Zaslavsky \cite{CSG} and, essentially, Soza\'nski \cite{Soz}. \begin{thm}\label{T:switching} Let $\S$ and $\S'$ be two signed graphs with the same underlying graph $\G$. Then $\cC^+(\S) = \cC^+(\S')$ if and only if $\S'$ is obtained by switching $\S$. In particular, $\S$ is balanced if and only if it switches to the all-positive signed graph $+\G$. \end{thm} \subsubsection*{Algorithmics of balance}\ How do we use this to determine balance or imbalance of $\S$? Assume $\S$ is connected, since we can treat each component separately. Find a spanning tree $T$ and choose a vertex $r$ to be its root. For each vertex $v$ there is a unique path $T_{rv}$ in $T$ from $r$ to $v$. Calculate $\zeta(v) = \s(T_{rv})$ (so, for instance, $\zeta(r)=+$) and switch $\S$ by $\zeta$. In $\S^\zeta$ every tree edge is positive. Every non-tree edge $e$ belongs to a unique circle $C_e$ in $T \cup e$ and $\s(C_e) = \s^\zeta(C_e) = \s^\zeta(e)$. If there is an edge $e$ that is negative in $\S^\zeta$, then there is a circle $C_e$ that is negative in $\S$ and $\S$ is unbalanced. If there is no such edge, then $\{X,Y\}$ with $X = \zeta\inv(+) \subseteq V$ and $Y = \zeta\inv(-)$ is a Harary bipartition of $\S$, confirming that $\S$ is balanced. Since $T$ can be found quickly by standard algorithms and it is obviously fast to find $\zeta$, this gives us a quick way of determining whether $\S$ is balanced or not. This simple algorithm was first published (in different terminology) independently by Hansen \cite{Hansen} and then by Harary and Kabell \cite{HK}. \subsubsection*{About circles}\ A \emph{chordless} or \emph{induced circle} is a circle $C$ that is an induced subgraph. Any extra induced edge besides $C$ itself is considered a \emph{chord} of $C$. An unsigned graph has girth, $g(\G) = \min_C |C|$, minimized over all circles $C$. It also has (though less frequently mentioned) even girth and odd girth, where $C$ varies over circles of even or odd length. These quantities are naturally signed-graphic. A signed graph has, besides its girth $g(\S)=g(\G)$, also \emph{positive girth} and \emph{negative girth}, $g_+(\S)$ and $g_-(\S)$, which are the minimum lengths of positive and negative circles; they reduce to even and odd girth when applied to $\S=-\G$. Girth is not explicit in any of my questions but signed girth may be worth keeping in mind. \subsubsection*{Contraction}\ Contracting an edge $e=vw$ with two distinct endpoints (a ``link'') in an ordinary graph means shrinking it to a point, i.e., identifying $v$ and $w$ to a single vertex and then deleting the edge $e$. In a signed graph $\S$, first $\S$ must be switched so that $e$ is positive. Then contraction is the same as it is without signs; the remaining edges retain the sign they have after switching. \subsubsection*{Balancing edges and vertices}\ A \emph{balancing vertex} is a vertex $v$ of an unbalanced signed graph $\S$ that lies in every negative circle; that is, $\S \setm v$ is balanced. A \emph{balancing edge} is an edge $e$ in an unbalanced signed graph such that $\S \setm e$ is balanced; that is, $e$ is in every negative circle. An endpoint of a balancing edge is a balancing vertex but there can (easily) be a balancing vertex without there being a balancing edge. A constructive characterization of balancing vertices is the next proposition. \emph{Contracting} a negative edge $vw$ that is not a loop means switching $w$ (so $vw$ becomes positive) and then identifying $v$ with $w$ and deleting the edge. \begin{prop}\label{P:balvert} Let $\S$ be a signed graph and $v$ a vertex in it. The following statements about $v$ are equivalent. \begin{enumerate}[{\rm (1)}] \item $v$ is a balancing vertex. \label{P:balvert:bv} \item $\S$ is obtained, up to switching, by adding a negative nonloop edge $vw$ to a signed graph with only positive edges and then contracting $vw$ to a vertex, which is the balancing vertex $v$. \label{P:balvert:contract} \item $\S$ can be switched so that all edges are positive except those incident with $v$, and at $v$ there is at least one edge of each sign. \label{P:balvert:vsigns} \end{enumerate} \end{prop} \begin{proof} The equivalence of \eqref{P:balvert:bv} with \eqref{P:balvert:contract} is from \cite{VLI}. The result of contraction in \eqref{P:balvert:contract} is precisely the description in \eqref{P:balvert:vsigns}. \end{proof} \subsubsection*{Blocks and necklaces}\ A \emph{cutpoint} is a vertex $v$ that has a pair of incident edges such that every path containing those edges passes through $v$. For instance, a vertex that supports a loop is a cutpoint unless the vertex is only incident with that loop and no other edge. A graph is called \emph{inseparable} if it is connected and has no cutpoints. A maximal inseparable subgraph of $\G$ is called a \emph{block of $\G$}; a graph that is inseparable is also called a \emph{block}. A \emph{block of $\S$} means just a block of $|\S|$. Blocks are important to signed graphs because every circle lies entirely within a block. An \emph{unbalanced necklace of balanced blocks} is an unbalanced signed graph constructed from balanced signed blocks $B_{1}, B_{2}, \ldots, B_{k}$ ($k\geq2$) and distinct vertices $v_i, w_i \in B_i$ by identifying $v_i$ with $w_{i-1}$ for $i=2,\ldots,k$ and $v_1$ with $w_k$. To make the necklace unbalanced, before the last step (identifying $v_1$ and $w_k$) make sure by switching that a path between them in $B_{1}\cup B_{2}\cup \cdots\cup B_{k}$ has negative sign. (All such paths have the same sign by the second half of Theorem \ref{T:balance}, because the union is balanced before the last identification.) An unbalanced necklace of balanced blocks is an unbalanced block in which each $v_i$ is a balancing vertex and there are no other balancing vertices. If a $B_i$ has only a single edge, that edge is a balancing edge. In fact, any signed block $\S$ with a nonloop balancing edge $e$ is an unbalanced necklace of balanced blocks: the balancing edge is one of the $B_i$'s, and the others are the blocks of $\S \setm e$. Unbalanced necklaces of balanced blocks are important in signed graphs; for instance, they require special treatment in matroid structure \cite{BG2}. If we allow $k=1$ in the definition of a necklace we can say that any signed block with a balancing vertex is an unbalanced necklace of balanced blocks. \subsection{Parity}\label{parity}\ There is a close connection between negative and positive circles in signed graphs on the one hand, and on the other hand odd and even circles in unsigned graphs---that is, parity of unsigned circles. First, parity is what one sees when all edges are negative, or (with switching) when the signature is antibalanced. There is considerable literature on parity problems that can be studied for possible generalization to signed graphs; I mention some of it in the following sections. The point of view here is that parity problems about circles are a special case of problems about signed circles. Some existing work on odd or even circles will generalize easily to negative or positive circles. For example, the computational difficulty of a signed-graph problem cannot be less than that of the specialization to antibalanced signatures---that is, the corresponding parity problem---and this may imply that the two problems have the same level of difficulty. \subsubsection*{Negative subdivision}\label{negsub}\ Second, there is \emph{negative subdivision}, which means replacing a positive edge by a path of two negative edges. Negatively subdividing every positive edge converts positive circles to even ones and negative circles to odd ones. Many problems on signed circles have the same answer after negative subdivision. The point of view here is that those signed-circle problems are a special case of parity-circle problems. Negative subdivision most obviously fails when connectivity is involved since the subdivided graph cannot be 3-connected. Another disadvantage is that contraction of edges makes sense only in signed graphs; a solution that involves contraction should be done in the signed framework. Denote by $\S^\sim$ the all-negative graph that results from negatively subdividing every positive edge. Let $\tilde e$ be the path of length 1 (if $\s(e)=-$) or 2 (if $\s(e)=+$) in $\S^\sim$ that corresponds to the edge $e \in E(\S)$, and for a negative $e$ let $v_e$ be the middle vertex of $\tilde e$; thus, $V(\S^\sim)=V(\S) \cup \{v_e : e \in E^+(\S)\}$. The essence of negative subdivision is the canonical sign-preserving bijection between the circles of $\S$ and those of $\S^\sim$, induced by mapping $e \in E(\S)$ to $\tilde e$ in $\S^\sim$. (There is such a bijection for every choice of positive edges to subdivide, even if that is not all positive edges.) \begin{prop}\label{T:negsub-bal} A signed graph $\S$ is balanced if and only if $|\S^\sim|$ is bipartite. \end{prop} \begin{proof} It follows from the sign-preserving circle bijection that $\S$ is balanced if and only if $\S^\sim$ is balanced. Since $\S^\sim$ is all negative, it is balanced if and only if its underlying graph is bipartite. \end{proof} \subsection{Weirdness}\label{terms}\ \subsubsection*{Groups or no group}\ Any two-element group will do instead of the sign group. Some people prefer to use the additive group $\bbZ_2$ of integers modulo $2$, which is the additive group of the two-element field $\bbF_2$. This is useful when the context favors a vector space over $\bbF_2$. Another variant notation is to define a signed graph as a pair $(\G,\S)$ where $\S \subseteq E(\G)$; the understanding is that the edges in $\S$ are negative and the others are positive. I do not use this notation. \subsubsection*{Terminologies}\ Switching has been called ``re-signing'' and other names. Stranger terminology exists. Several otherwise excellent works redefine the words ``even'', ``odd'', and ``bipartite'' to mean positive, negative, and balanced, all of which empty those words of their standard meanings and invite confusion. I say, ``That way madness lies'' \cite{Lear}.
{ "timestamp": "2018-01-17T02:01:47", "yymm": "1701", "arxiv_id": "1701.07963", "language": "en", "url": "https://arxiv.org/abs/1701.07963", "abstract": "A signed graph is a graph whose edges are labelled positive or negative. The sign of a circle (cycle, circuit) is the product of the signs of its edges. Most of the essential properties of a signed graph depend on the signs of its circles. Here I describe several questions regarding negative circles and their cousins the positive circles. Topics include incidence between signed circles and edges or vertices, characterizing signed graphs with special circle properties, counting negative circles, signed-circle packing and covering, signed circles and eigenvalues, and directed cycles in signed digraphs. A few of the questions come with answers.", "subjects": "Combinatorics (math.CO)", "title": "Negative (and Positive) Circles in Signed Graphs: A Problem Collection", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9830850872288504, "lm_q2_score": 0.8333246015211009, "lm_q1q2_score": 0.8192289885763184 }
https://arxiv.org/abs/1612.07904
On Garland's vanishing theorem for $\mathrm{SL}_n$
This is an expository paper on Garland's vanishing theorem specialized to the case when the linear algebraic group is $\mathrm{SL}_n$. Garland's theorem can be stated as a vanishing of the cohomology groups of certain finite simplicial complexes. The method of the proof is quite interesting on its own. It relates the vanishing of cohomology to the assertion that the minimal positive eigenvalue of a certain combinatorial laplacian is sufficiently large. Since the 1970's, this idea has found applications in a variety of problems in representation theory, group theory, and combinatorics, so the paper might be of interest to a wide audience. The paper is intended for non-specialists and graduate students.
\section{Introduction} \subsection{Statement of the theorem} This is an expository paper on Howard Garland's work \cite{Garland} specialized to the case when the linear algebraic group is $\mathrm{SL}_n$. Reading \cite{Garland} requires knowledge of the theory of buildings. On the other hand, the ideas in \cite{Garland}, in their essence, are combinatorial. Since the relevant Bruhat-Tits building in the $\mathrm{SL}_n$ case has a simple description in terms of lattices in a finite dimensional vector space, one can give a proof of Garland's vanishing theorem which requires from the reader only familiarity with linear algebra and some group theory. There is already an excellent Bourbaki Expos\'e by Borel \cite{Borel} on Garland's work. The main difference of our exposition, besides the more detailed proofs, is the absence of any references to the theory of buildings, which makes the article completely self-contained. Since Garland's result applies to quite general discrete subgroups of $p$-adic groups, to give a full account of his work one cannot avoid a discourse on the theory of buildings and representation theory. Other expositions of Garland's method and its generalizations can be found in \cite{BS} and \cite{ABM}; these papers also give nice applications of Garland's method to group theory and combinatorics, although they do not give a complete proof of Theorem \ref{thm1.1}. Let $K$ be a non-archimedean local field with residue field of order $q$; such a field is either isomorphic to a finite extension of the $p$-adic numbers $\mathbb{Q}_p$, or to the field of formal Laurent series $\mathbb{F}_q(\!(T)\!)$ over the finite field $\mathbb{F}_q$ with $q$ elements. Let $\Gamma$ be a discrete subgroup of the topological group $\mathrm{SL}_n(K)$, $n\geq 2$. There is an infinite contractible $(n-1)$-dimensional simplicial complex $\mathfrak B$, the \textit{Bruhat-Tits building of $\mathrm{SL}_{n}(K)$}, on which $\Gamma$ naturally acts. The complex $\mathfrak B$ can be described in terms of lattices in the vector space $K^n$. One way to formulate the main result of \cite{Garland} for $\mathrm{SL}_n$ is as follows: \begin{thm}\label{thm1.1} Assume the quotient $\mathfrak B/\Gamma$ is a finite complex. There is a constant $q(n)$ depending only on $n$ such that if $q\geq q(n)$ then for all $0<i<n-1$ the simplicial cohomology groups $H^i(\mathfrak B/\Gamma, \mathbb{R})$ are zero. \end{thm} Such vanishing theorems were originally conjectured by Serre \cite{Serre}. We will prove Theorem \ref{thm1.1} in Section \ref{sBTB} under a mild assumption on $\Gamma$. In the same section we also explain how central division algebras naturally give rise to such groups. We should mention that the restriction on $q$ being sufficiently large in Theorem \ref{thm1.1} was removed by Casselman \cite{Casselman}, who proved the vanishing of the middle cohomology groups by a completely different method, using representation theory of $p$-adic groups. Garland's vanishing theorem plays an important role in some problems in representation theory; for example, it puts strict restrictions on the continuous cohomology of topological groups with coefficients in infinite dimensional representations (cf. \cite{Casselman} and $\S$\ref{ss4.3}). An application of Garland's theorem in arithmetic geometry arises in the calculation of the cohomology groups of certain algebraic varieties possessing rigid-analytic uniformization. More precisely, $\mathfrak B$ can be realized as the skeleton of Drinfeld's symmetric space $\Omega^n$, so the cohomology groups of the algebraic variety uniformized as $\Omega^n/\Gamma$ are related to the cohomology groups of $\mathfrak B/\Gamma$ (cf. \cite{SS}). \subsection{Outline of the paper} Now we give an outline of the proof of Theorem \ref{thm1.1} and the contents of this paper. Let $X$ be a finite simplicial complex of dimension $n$. Let $w$ be a ``Riemannian metric'' on $X$, by which we mean, following \cite{Garland2}, a function from the non-oriented simplices of $X$ to the positive real numbers. Let $C^i(X)$ denote the $\mathbb{R}$-vector space of $i$-cochains of $X$ with values in $\mathbb{R}$. Define an inner product on $C^i(X)$: $$ (f, g)=\sum_\sigma w(\sigma) f(\hat{\sigma}) g(\hat{\sigma}), $$ where the sum is over all non-oriented $i$-simplices of $X$ and $\hat{\sigma}$ is an oriented simplex corresponding to $\sigma$. Let $d: C^i(X)\to C^{i+1}(X)$ denote the coboundary operator, and $\delta: C^i(X)\to C^{i-1}(X)$ denote the adjoint of $d$ with respect to $(\cdot ,\cdot )$. Let $\harm{i}(X)\subset C^i(X)$ be the subspace of \textit{harmonic cocycles}; by definition, these are the $i$-cochains annihilated by both $d$ and $\delta$. It is not hard to show that $H^i(X, \mathbb{R})\cong \harm{i}(X)$. This isomorphism is a consequence of the ``Hodge decomposition'' for $C^i(X)$. In Section \ref{SecGM}, after recalling some standard terminology related to simplicial complexes, we prove this well-known fact. Thus, to prove $H^i(X, \mathbb{R})=0$, it suffices to prove that there are no non-zero harmonic cocycles. Motivated by the work of Matsushima \cite{Matsushima}, who reduced the study of harmonic forms on real locally symmetric spaces to the computation of the minimal eigenvalues of certain curvature transformations, Garland reduces the study of $\harm{i}(X)$ to estimating the minimal non-zero eigenvalue $m^i(X)$ of the linear operator $\Delta=\delta d$ acting on $C^i(X)$. Section \ref{sFI} contains a key part of Garland's argument: it gives the precise relationship between the vanishing of $\harm{i}(X)$ and lower bounds on $m^i(X)$, and it also gives a method for estimating $m^i(X)$ inductively. In Section \ref{sBTB}, we describe the Bruhat-Tits building $\mathfrak B$ of $\mathrm{SL}_n(K)$ as a simplicial complex and explain how Theorem \ref{thm1.1} follows from the results in Section \ref{sFI}, assuming a certain lower bound on $m^i(\mathrm{Lk}(v))$ for vertices of $\mathfrak B$, where $\mathrm{Lk}(v)$ denotes the link of the vertex $v$. We relegate the proof of this lower bound, which is the most technical part of the paper, to Section \ref{sec3}. At the end of Section \ref{sBTB} we give a brief discussion of some of the more recent applications of Garland's method to producing examples of groups having Kazhdan's property (T). Based on numerical calculations, in $\S$\ref{ss4.1} we state a conjecture about the asymptotic behaviour of the eigenvalues of $\Delta$ acting on $C^i(\mathrm{Lk}(v))$ for vertices of $\mathfrak B$, and in $\S$\ref{ss4.3} we give some evidence for this conjecture. None of the results of this paper are original, except possibly those in $\S$\ref{ss4.3}. \section{Simplicial cohomology and harmonic cocycles}\label{SecGM} In this section we recall the basic definitions from the theory of simplicial cohomology and prove a combinatorial analogue of the Hodge decomposition theorem. This last theorem identifies simplicial cohomology groups with spaces of harmonic cocycles. Its importance for the proof of Theorem \ref{thm1.1} is that, instead of proving that $H^i(\mathfrak B/\Gamma, \mathbb{R})=0$ directly, we will actually show that there are no non-zero harmonic $i$-cocycles on $\mathfrak B/\Gamma$. \subsection{Basic concepts} An (abstract) \textit{simplicial complex} is a collection $X$ of finite nonempty sets, called simplices, such that if $s$ is an element of $X$, so is every nonempty subset of $s$. A nonempty subset of a simplex $s$ is called a \textit{face} of $s$. A \textit{simplex of dimension $i$}, or simply an \textit{$i$-simplex}, is a simplex with $i+1$ elements. The vertex set $\mathrm{Ver}(X)$ of $X$ is the union of its $0$-simplices. A subcollection of $X$ that is itself a complex is called a \textit{subcomplex} of $X$. The \textit{dimension} of $X$ is the largest dimension of one of its simplices, or is infinite if there is no such largest dimension. Let $s$ be a simplex of $X$. The \textit{star} of $s$ in $X$, denoted $\mathrm{St}(s)$, is the subcomplex of $X$ consisting of the union of all simplices of $X$ having $s$ as a face. The \textit{link} of $s$, denoted $\mathrm{Lk}(s)$, is the subcomplex of $\mathrm{St}(s)$ consisting of the simplices which are disjoint from $s$. If one thinks of $\mathrm{St}(s)$ as the ``unit ball'' around $s$ in $X$, then $\mathrm{Lk}(v)$ is the ``unit sphere'' around $s$. Let $X$ and $Y$ be simplicial complexes. The \textit{join} of $X$ and $Y$ is the simplicial complex $X\ast Y$ such that $s\in X\ast Y$ if either $s\in X$ or $s\in Y$, or $s=x\ast y:=\{x_0,\dots, x_i, y_0, \dots, y_j\}$, where $x=\{x_0,\dots, x_i\}\in X$ and $y=\{y_0,\dots, y_j\}\in Y$. It is clear that $X\ast Y$ is a simplicial complex and $\dim(X\ast Y)=\dim(X)+\dim(Y)+1$. Note that, as a special case of this construction, $\mathrm{St}(s)=s\ast \mathrm{Lk}(s)$. A specific ordering of the vertices of $s$, up to an even permutation, is called an \textit{orientation} of $s$. Each positive dimensional simplex has two orientations. Denote the set of $i$-simplices by $\widehat{S}_i(X)$, and the set of oriented $i$-simplices by $S_i(X)$. Note that $\widehat{S}_0(X)=S_0(X)=\mathrm{Ver}(X)$. For $s\in S_i(X)$, $\bar{s}\in S_i(X)$ denotes the same simplex but with opposite orientation. An $\mathbb{R}$-valued \textit{$i$-cochain} on $X$ is a function $f: S_i(X)\to \mathbb{R}$ which is alternating if $i\geq 1$, i.e., $f(s)=-f(\bar{s})$. (A $0$-cochain is just a function on $\mathrm{Ver}(X)$.) The $i$-cochains naturally form an $\mathbb{R}$-vector space which is denoted $C^i(X)$. If $i<0$ or $i>\dim(X)$, we set $C^i(X)=0$. The \textit{coboundary} operator is the linear transformation $d:C^i(X)\to C^{i+1}(X)$ defined by \begin{equation}\label{eq-d} df([v_0,\dots, v_{i+1}])=\sum_{j=0}^{i+1}(-1)^jf([v_0,\dots,\hat{v}_j,\dots, v_{i+1}]), \end{equation} where $[v_0,\dots,v_{i+1}]\in S_{i+1}(X)$ and the symbol $\hat{v}_j$ means that the vertex $v_j$ is to be deleted from the array. The kernel of $d:C^i(X)\to C^{i+1}(X)$ is called the subspace of \textit{$i$-cocycles} and denoted $Z^i(X)$. As one easily verifies $d\circ d = 0$ (cf. \cite[p. 30]{Munkres}), so $dC^{i-1}(X)\subset Z^i(X)$. The \textit{$i$-th cohomology group of $X$} (with real coefficients) is $$H^i(X):=Z^i(X)/dC^{i-1}(X).$$ Let $\mathbf{1}\in C^0(X)$ be the function defined by $\mathbf{1}(v)=1$ for all $v\in \mathrm{Ver}(X)$. The subspace $\mathbb{R} \mathbf{1}\subset C^0(X)$ spanned by $\mathbf{1}$ is the space of constant function. It is easy to see that $\mathbb{R} \mathbf{1} \subset Z^0(X)$. One defines the \textit{reduced $i$-th cohomology group} $\tilde{H}^i(X)$ of $X$ by setting $\tilde{H}^i(X)=H^i(X)$ if $i\geq 1$, and $\tilde{H}^0(X)=Z^0(X)/\mathbb{R} \mathbf{1}$. It is easy to show that the ``geometric realization'' of $X$ is connected if and only if $\tilde{H}^0(X)=0$; see \cite[p. 256]{Munkres}. Now assume $X$ is finite, i.e., has finitely many vertices. To each simplex $s$ of $X$ assign a positive real number $w(s)=w(\bar{s})$, which we call the \textit{weight} of $s$. Define an inner-product on $C^i(X)$ by \begin{equation}\label{eq-pairing} (f,g)=\sum_{{s} \in \widehat{S}_i(X)} w(s)f(s)g(s),\qquad f,g\in C^i(X), \end{equation} where in $w(s)f(s)g(s)$ we choose some orientation of $s$; this is well-defined since $f(s)g(s)=f(\bar{s})g(\bar{s})$. Let $s=[v_0,\dots, v_i]\in S_i(X)$ and $v\in \mathrm{Ver}(X)$. If the set $\{v,v_0,\dots, v_i\}$ is an $(i+1)$-simplex of $X$, then we denote by $[v, s]\in S_{i+1}(X)$ the oriented simplex $[v,v_0,\dots, v_i]$. Define a linear transformation $\delta: C^i(X)\to C^{i-1}(X)$ by \begin{equation}\label{eq-delta} \delta f(s)=\sum_{\substack{v\in \mathrm{Ver}(X)\\ [v,s]\in S_i(X)}} \frac{w([v,s])}{w(s)}f([v,s]). \end{equation} This operator is the adjoint of $d$ with respect to \eqref{eq-pairing}: \begin{lem}\label{v-prop1.12} If $f\in C^i(X)$ and $g\in C^{i+1}(X)$, then $(df,g)=(f,\delta g)$. \end{lem} \begin{proof} \begin{align*} &(df,g)\\ &=\sum_{\substack{s=[v_0,\dots,v_{i+1}] \in \widehat{S}_{i+1}(X)}} w(s)\sum_{j=0}^{i+1}f([v_0,\dots, \hat{v}_j,\dots, v_{i+1}])g([v_j,v_0,\dots,\hat{v}_j,\dots, v_{i+1}])\\ &=\sum_{\substack{\sigma \in \widehat{S}_{i}(X)}} w(\sigma)f(\sigma)\sum_{\substack{v\in \mathrm{Ver}(X) \\ [v,\sigma]\in S_{i+1}(X)}}\frac{w([v,\sigma])}{w(\sigma)}g([v, \sigma])=(f,\delta g). \end{align*} \end{proof} The kernel of $\delta: C^i(X)\to C^{i-1}(X)$ will be denoted by $\sZ^i(X)$. \subsection{Combinatorial Hodge decomposition} The intersection $$\harm{i}(X):=\sZ^i(X)\cap Z^i(X)$$ in $C^i(X)$ is the subspace of \textit{harmonic $i$-cocycles}. (This subspace depends on the choice of the inner-product \eqref{eq-pairing}.) \begin{rem} The term \textit{harmonic} comes from the fact that $f\in C^i(X)$ is in $\harm{i}(X)$ if and only if $f$ is in the kernel of the operator $d\delta+\delta d$, which is a combinatorial analogue of the Laplacian. \end{rem} \begin{thm}\label{lem1.11} Relative to the inner-product \eqref{eq-pairing}, we have the orthogonal direct sum decompositions \begin{align} \label{eq-ci} C^i(X) &=\harm{i}(X)\oplus dC^{i-1}(X)\oplus \delta C^{i+1}(X), \\ \label{eq-zi} Z^i(X) &=\harm{i}(X)\oplus dC^{i-1}(X), \\ \label{eq-hi} \sZ^i(X) &=\harm{i}(X)\oplus \delta C^{i+1}(X). \end{align} This implies \begin{equation}\label{eq-CohomHarm} H^i(X)=Z^i(X)/dC^{i-1}(X)\cong \sZ^i(X)/\delta C^{i+1}(X)\cong \harm{i}(X). \end{equation} \end{thm} \begin{proof} Let $f\in dC^{i-1}(X)$ and $g\in \delta C^{i+1}(X)$. We can write $f=df'$ and $g=\delta g'$ for some $f'\in C^{i-1}(X)$ and $g'\in C^{i+1}(X)$. Since $d^2=0$, using Lemma \ref{v-prop1.12} we get $$ (f, g)=(df', \delta g')=(d^2 f', g')=(0, g')=0. $$ Hence $dC^{i-1}(X)\perp \delta C^{i+1}(X)$. Now suppose $h\in C^i(X)$ is orthogonal to $dC^{i-1}(X)\oplus \delta C^{i+1}(X)$. Then $$ 0=(h, df')=(\delta h, f') $$ for all $f'\in C^{i-1}(X)$, which implies $\delta h=0$. Similarly, $0=(h, \delta g')$ implies that $dh=0$. Therefore, $h\in \harm{i}(X)$. In other words, the orthogonal complement of $dC^{i-1}(X)\oplus \delta C^{i+1}(X)$ in $C^i(X)$ is $\harm{i}(X)$. This proves \eqref{eq-ci}. A cochain $f\in C^i(X)$ is in the orthogonal complement of $ \delta C^{i+1}(X)$ if and only if $(f, \delta g')=(df, g')=0$ for all $g'\in C^{i+1}(X)$, which implies $df=0$. Thus \begin{equation}\label{eq-ci2} C^i(X) =Z^i(X)\oplus \delta C^{i+1}(X). \end{equation} A similar argument shows that the orthogonal complement of $dC^{i-1}(X)$ is $\sZ^i(X)$: \begin{equation}\label{eq-ci3} C^i(X) =\sZ^i(X)\oplus dC^{i-1}(X). \end{equation} Comparing \eqref{eq-ci2} and \eqref{eq-ci3} with \eqref{eq-ci}, we get \eqref{eq-zi} and \eqref{eq-hi}. Finally, \eqref{eq-zi} and \eqref{eq-hi} imply \eqref{eq-CohomHarm}. \end{proof} \begin{defn}\label{defn1.7} Following \cite{Garland}, we call the linear transformation $\Delta=\delta d$ on $C^i(X)$ the \textit{curvature transformation}. (What we denote by $\Delta$ in this paper is denoted by $\Delta^+$ in \cite{Garland} and \cite{Borel}.) \end{defn} By Lemma \ref{v-prop1.12}, for any $f, g\in C^i(X)$ we have $$ (\Delta f, g)= (\delta d f, g)= (d f, dg) = (f, \delta d g)= (f, \Delta g) $$ and \begin{equation}\label{eq-see} (\Delta f, f)= (df, df)\geq 0. \end{equation} Hence $\Delta$ is a self-adjoint positive operator on $C^i(X)$, which implies that $C^i(X)$ has an orthonormal basis consiting of eigenvectors of $\Delta$, and the eigenvalues of $\Delta$ are nonnegative. \begin{lem}\label{lem1.7} Let $i\geq 0$. \begin{enumerate} \item The subspace of $C^i(X)$ spanned by the eigenfunctions of $\Delta$ with positive eigenvalues is $\delta C^{i+1}(X)=Z^i(X)^\perp$. \item If $H^i(X)=0$, then $\delta C^{i+1}(X)=\sZ^i(X)$. \end{enumerate} \end{lem} \begin{proof} It is clear from \eqref{eq-see} that if $\Delta f=0$ then $df=0$. Hence $\ker(\Delta)\subseteq Z^i(X)$. Conversely, if $f\in Z^i(X)$ then $\Delta f=\delta df=\delta 0=0$. Therefore the subspace of $C^i(X)$ spanned by the eigenfunctions of $\Delta$ with positive eigenvalues is $Z^i(X)^\perp$. This latter subspace is $\delta C^{i+1}(X)$, as follows from \eqref{eq-ci2}. The second claim of the lemma follows from \eqref{eq-CohomHarm}. \end{proof} \section{Garland's method}\label{sFI} In this section we discuss what is nowadays is called \textit{Garland's method}. Let $X$ be a finite simplicial complex. Let $$\lambda_{\min}^i(X):=\min_{v\in \mathrm{Ver}(X)} m^i(\mathrm{Lk}(v)).$$ Garland's method shows that a strong enough lower bound on $\lambda_{\min}^{i-1}(X)$ implies that $\harm{i}(X)=0$ (hence also $H^i(X, \mathbb{R})=0$ by the Hodge decomposition discussed in the previous section). Moreover, the method gives a lower bound on $m^k(X)$ in terms of $\lambda_{\min}^{k-1}(X)$, hence allows to estimate $\lambda_{\min}^{i-1}(X)$ inductively in certain situations. (We will see an example of such an inductive estimate in Section \ref{sec3}.) The observation that Garland's ideas from \cite{Garland} apply to any finite simplicial complex satisfying a certain condition is due to Borel; in this section we partly follow \cite[$\S$1]{Borel}. The proof of the main results has two parts: It starts with a decomposition of $(\Delta f, f)$ into a sum $\sum_v (\Delta f_v, f_v)$ over the vertices of $X$, where $f_v$ is the restriction of $f$ to the ``unit ball'' around $v$; this is the content of $\S$\ref{ss3.1}. Then one bounds $(\Delta f_v, f_v)$ in terms of $m^{i-1}(\mathrm{Lk}(v))$ by studying the local version of the curvature transformation; this is the content of $\S$\ref{ss3.2}. We combine these two parts in $\S$\ref{ss3.3} to prove the main results. One of the subtleties is that to make this strategy work one has to choose an appropriate metric \eqref{eq-themetric}. In this section we assume that $X$ is a finite $n$-dimensional complex which satisfies the following property: \begin{center} $(\star)$ each simplex of $X$ is a face of some $n$-simplex. \end{center} For $s\in S_i(X)$, let \begin{equation}\label{eq-themetric} w(s) = \text{the number of (non-oriented) $n$-simplices containing $s$.} \end{equation} Note that, due to $(\star)$, $w(s)\neq 0$. \subsection{Decomposition of $(\Delta f,f)$ into local factors}\label{ss3.1} We start with a simple lemma: \begin{lem}\label{lem-w} Let $s\in \widehat{S}_i(X)$ be fixed. Then $$ \sum_{\substack{\sigma\in \widehat{S}_{i+1}(X)\\ s\subset \sigma}}w(\sigma)=(n-i)w(s). $$ \end{lem} \begin{proof} Given an $n$-simplex $t$ such that $s\subset t$ there are exactly $(n-i)$ simplices $\sigma$ of dimension $(i+1)$ such that $s\subset \sigma\subset t$. Hence in the sum of the lemma we count every $n$-simplex containing $s$ exactly $(n-i)$ times. \end{proof} \begin{comment} \begin{rem} The appearance of the weights $w(s)$ in \eqref{eq-pairing} might seem unmotivated at the moment, but for the arguments in the next subsection to go through one crucially needs exactly this weighted version of the pairing. \end{rem} \end{comment} For a fixed $v\in \mathrm{Ver}(X)$ define a linear transformation $\rho_v: C^i(X)\to C^i(X)$ by: $$ \rho_vf (s)=\left\{ \begin{array}{ll} f(s) & \hbox{if } v\in s; \\ 0 & \hbox{otherwise.} \end{array} \right. $$ It is clear that \begin{equation}\label{eq-rho^2} \rho_v\rho_v=\rho_v, \end{equation} and for $f,g\in C^i(X)$ \begin{equation}\label{eq-(rho)} (\rho_vf, g)= (\rho_vf, \rho_v g)=(f, \rho_vg). \end{equation} Moreover, since any $i$-simplex has $(i+1)$-vertices, for $f\in C^i(X)$ we have the equality \begin{equation}\label{eq-obv} \sum_{v\in \mathrm{Ver}(X)}\rho_vf=(i+1)f. \end{equation} \begin{comment} We also have the following obvious lemma: \begin{lem}\label{v-prop1.13} \hfill \begin{enumerate} \item $\rho_v\rho_v=\rho_v$; \item For $f \in C^i(X)$ and $g\in C^i(X)$, $(\rho_vf, g)=(f, \rho_vg)$. \end{enumerate} \end{lem} \end{comment} \begin{lem}\label{lem-borel} For $f\in C^i(X)$, we have $$ i\cdot (\Delta f, f)=\sum_{v\in \mathrm{Ver}(X)}(\Delta \rho_v f,\rho_v f) - (n-i)(f,f). $$ \end{lem} \begin{proof} First, to simplify the notation in our calculations we introduce new notation. Let $\sigma\in S_{i+1}(X)$ and $s\in S_i(X)$ be a face of $\sigma$. The orientation on $\sigma$ induces an orientation on $s$; we define $[\sigma:s]=\pm 1$ depending on whether this induces orientation is the original orientation of $s$ or its opposite. With this definition, for $f\in C^{i}(X)$ we have $$ df(\sigma) = \sum_{\substack{s\in \widehat{S}_i(X)\\ s\subset \sigma}}[\sigma:s]f(s), $$ where for each face $s$ of $\sigma$ we choose some orientation. (Note that $[\sigma:s]f(s)$ does not depend on the choice of the orientation of $s$.) Let $v\in \sigma$ be a fixed vertex and $s_0\in \widehat{S}_i(X)$ be the unique face of $\sigma$ not containing $v$. Then $$ df(\sigma) = d\rho_v f(\sigma)+[\sigma:s_0]f(s_0). $$ Hence $$ df(\sigma)^2 = d\rho_v f(\sigma)^2+2[\sigma:s_0]f(s_0)\sum_{v\in s\subset \sigma}[\sigma:s]f(s)+f(s_0)^2. $$ Summing both sides over all vertices of $\sigma$ we get $$ (i+2)df(\sigma)^2=\sum_{v\in \sigma}d\rho_v f(\sigma)^2+2\sum_{\substack{s, s'\subset \sigma\\ s\neq s'}}[\sigma:s][\sigma:s']f(s)f(s') +\sum_{s\subset \sigma} f(s)^2 $$ $$ =\sum_{v\in \sigma}d\rho_v f(\sigma)^2 + 2\sum_{\substack{s, s'\subset \sigma}}[\sigma:s][\sigma:s']f(s)f(s') -\sum_{s\subset \sigma} f(s)^2 $$ $$ =\sum_{v\in \sigma}d\rho_v f(\sigma)^2 + 2df(\sigma)^2 -\sum_{s\subset \sigma} f(s)^2. $$ Hence \begin{equation}\label{Borel-noproof} i\cdot df(\sigma)^2 = \sum_{v\in \sigma}d\rho_v f(\sigma)^2 -\sum_{s\subset \sigma} f(s)^2. \end{equation} Now \begin{align*} i\cdot (\Delta f, f) &\overset{\mathrm{Lem.} \ref{v-prop1.12}}{=} i\cdot (df, df) = i\sum_{\sigma\in \widehat{S}_{i+1}(X)} w(\sigma) df(\sigma)^2 \\ &\overset{\eqref{Borel-noproof}}{=}\sum_{\sigma\in \widehat{S}_{i+1}(X)} \sum_{v\in \sigma}w(\sigma) d\rho_v f(\sigma)^2- \sum_{\sigma\in \widehat{S}_{i+1}(X)}\sum_{s\subset \sigma} w(\sigma)f(s)^2 \\ &=\sum_{v\in \mathrm{Ver}(X)} (d\rho_v f,d\rho_v f) - \sum_{s\in \widehat{S}_{i}(X)}f(s)^2 \sum_{\substack{\sigma\in \widehat{S}_{i+1}(X)\\ s\subset \sigma}} w(\sigma) \\ &\overset{\mathrm{Lem.} \ref{lem-w}}{=} \sum_{v\in \mathrm{Ver}(X)} (\Delta\rho_v f,\rho_v f)-(n-i)\sum_{s\in \widehat{S}_{i}(X)}w(s)f(s)^2 \\ &= \sum_{v\in \mathrm{Ver}(X)} (\Delta\rho_v f,\rho_v f)-(n-i)(f,f). \end{align*} \end{proof} \subsection{Local curvature transformations}\label{ss3.2} \begin{comment} \begin{cor}\label{cor1.14} Let $f\in C^i(X)$. Suppose there is a positive real number $\lambda$ such that $(\Delta \rho_v f, \rho_v f)\geq \lambda\cdot (\rho_v f, f)$ for all $v\in \mathrm{Ver}(X)$. Then $$i\cdot (\Delta f, f)\geq \left(\lambda\cdot (i+1)-(n-i)\right)(f,f).$$ Similarly, if there is a positive real number $\Lambda$ such that $(\Delta \rho_v f, \rho_v f)\leq \Lambda\cdot (\rho_v f, f)$ for all $v\in \mathrm{Ver}(X)$, then $$i\cdot (\Delta f, f)\leq \left(\Lambda\cdot (i+1)-(n-i)\right)(f,f).$$ \end{cor} \begin{proof} This follows from Lemma \ref{v-prop1.13}, Lemma \ref{lem-borel} and \eqref{obv}. \end{proof} \end{comment} For $f,g\in C^i(\mathrm{Lk}(v))$ define their inner-product by \begin{equation}\label{eq-locp} (f,g)_v=\sum_{s\in \widehat{S}_i(\mathrm{Lk}(v))}w_v(s)\cdot f(s)\cdot g(s), \end{equation} where $w_v(s)$ is the number of $(n-1)$-simplices in $\mathrm{Lk}(v)$ containing $s$. Note that $\mathrm{Lk}(v)$ is an $(n-1)$-dimensional complex satisfying $(\star)$. Another simple observation is that for a simplex $\sigma$ in $\mathrm{Lk}(v)$ there is a one-to-one correspondence between the $n$-simplices of $X$ containing $[v,\sigma]$ and the $(n-1)$-simplices of $\mathrm{Lk}(v)$ containing $\sigma$. Hence \begin{equation}\label{eq-w_vw} w_v(\sigma)=w([v,\sigma]) \quad \text{for any }\sigma\in \mathrm{Lk}(v). \end{equation} Let $$d_v: C^i(\mathrm{Lk}(v))\to C^{i+1}(\mathrm{Lk}(v))$$ be the coboundary operator acting on the cochains of the finite simplicial complex $\mathrm{Lk}(v)$. Let $\delta_v$ be the adjoint of $d_v$ with respect to \eqref{eq-locp}, and let $$\Delta_v:=\delta_vd_v.$$ Lemma \ref{lem-borel} essentially decomposes $\Delta$ into a sum of its restrictions $\Delta\rho_v$ to $\mathrm{St}(v)$ over all vertices. We want to relate $\Delta\rho_v$ to $\Delta_v$, and hence to relate the eigenvalues of $\Delta$ to the eigenvalues of its local version $\Delta_v$. For this we need to introduce one more linear operator: For $i\geq 1$, define \begin{align*} &\tau_v:C^{i}(X)\to C^{i-1}(X),\\ & \tau_vf(s)=\left\{ \begin{array}{ll} f([v,s]), & \hbox{if $s\in S_{i-1}(\mathrm{Lk}(v))$;}\\ 0, & \hbox{otherwise.} \end{array} \right. \end{align*} Given $f\in C^i(X)$, its restriction to $\mathrm{Lk}(v)$ defines a function in $C^i(\mathrm{Lk}(v))$, which by slight abuse of notation we denote by the same letter. With this convention, for $f\in C^i(X)$ we can consider $d_v f$ and $\delta_v f$, which are now functions on $\mathrm{Lk}(v)$. Similarly, we can compute the pairing \eqref{eq-locp} on $C^i(X)$. \begin{lem}\label{prop7.12} Assume $i\geq 1$. For $f, g\in C^i(X)$, we have $$(\tau_vf,\tau_vg)_v=(\rho_vf,\rho_vg).$$ \end{lem} \begin{proof} We have \begin{align*} (\tau_vf,\tau_vg)_v &=\sum_{\sigma\in \widehat{S}_{i-1}(\mathrm{Lk}(v))}w_v(\sigma)\cdot \tau_vf(\sigma)\cdot \tau_vg(\sigma)\\ &\overset{\eqref{eq-w_vw}}{=} \sum_{s\in \widehat{S}_{i}(\mathrm{St}(v))}w(s)\cdot \rho_vf(s)\cdot \rho_vg(s). \end{align*} Since $\rho_v f$ is zero away from $\mathrm{St}(v)$, the last sum can be extended to the whole $\widehat{S}_{i}(X)$, so the lemma follows. \end{proof} \begin{lem} Let $f\in C^i(X)$. We have \begin{align} \label{eq-taud} \tau_v d\rho_v f &=-d_v\tau_v f, \quad i\geq 1,\\ \label{eq-taudelta} \tau_v \delta f &=-\delta_v\tau_v f, \quad i\geq 2, \\ \label{eq-tauDelta} \tau_v \Delta \rho_v f &=\Delta_v\tau_v f, \quad i\geq 1. \end{align} \end{lem} \begin{proof} Let $s\in S_i(\mathrm{Lk}(v))$, $i\geq 1$. We have $$ \tau_v d\rho_v f(s)=d\rho_vf([v,s])=\rho_v(f(s)-f([v, ds]))=-f([v, ds]). $$ In the last term $ds$ denotes the image of $s$ under the boundary operator and $[\cdot]$ is extended linearly to $\mathbb{Z}[S_i(\mathrm{Lk}(v))]$. Since $d$ restricted to $\mathrm{Lk}(v)$ coincides with $d_v$, we have $f([v, ds])=d_v\tau_v f(s)$. This proves \eqref{eq-taud}. Now assume $i\geq 2$ and let $s\in S_{i-2}(\mathrm{Lk}(v))$. We have $$ \tau_v \delta f(s)=\delta f([v, s])= \sum_{\substack{x\in \mathrm{Ver}(X)\\ [x,v, s]\in S_{i}(X)}}\frac{w([x,v,s])}{w([v, s])}f([x,v,s]) $$ $$ \overset{\eqref{eq-w_vw}}{=}- \sum_{\substack{x\in \mathrm{Ver}(\mathrm{Lk}(v))\\ [x, s]\in S_{i-1}(\mathrm{Lk}(v))}}\frac{w_v([x, s])}{w_v(s)}\tau_v f([x,s])=-\delta_v\tau_v f(s). $$ This proves \eqref{eq-taudelta}. Finally, $$ \tau_v \Delta \rho_v f =\tau_v \delta d \rho_v f\overset{\eqref{eq-taudelta}}{=}-\delta_v \tau_v d\rho_v f \overset{\eqref{eq-taud}}{=} \delta_v d_v\tau_v f=\Delta_v\tau_v f. $$ This proves \eqref{eq-tauDelta}. \end{proof} \begin{lem}\label{prop7.14} Assume $i\geq 1$. For $f\in C^i(X)$, we have $$(\Delta \rho_v f,\rho_v f)=(\Delta_v\tau_vf,\tau_vf)_v.$$ \end{lem} \begin{proof} We have $$ (\Delta \rho_v f,\rho_vf)\overset{\eqref{eq-(rho)}}{=}(\rho_v\Delta \rho_v f,\rho_vf)\overset{\mathrm{Lem.} \ref{prop7.12}}{=}(\tau_v\Delta \rho_vf,\tau_vf)_v \overset{\eqref{eq-tauDelta}}{=}(\Delta_v\tau_vf,\tau_vf)_v. $$ \begin{comment} Now it is enough to show $\tau_v\Delta\rho_v f=\Delta_v\tau_vf$. For $s\in S_{i-1}(\mathrm{Lk}(v))$, we have $$ \tau_v\Delta\rho_v f(s)=\delta d \rho_vf([v,s])=\sum_{\substack{x\in \mathrm{Ver}(\mathrm{Lk}(v))\\ [x,s]\in S_{i}(\mathrm{Lk}(v))}}\frac{w([x,v,s])}{w([v,s])}d\rho_vf([x,v,s]) $$ $$ =\sum_{\substack{x\in \mathrm{Ver}(\mathrm{Lk}(v))\\ [x,s]\in S_{i}(\mathrm{Lk}(v))}}\frac{w_v([x,s])}{w_v(s)}(\rho_vf([v,s])-\rho_vf([x,s])+\rho_vf([x,v,ds])). $$ In the last term $ds$ denotes the image of $s$ under the boundary operator and $[\cdot]$ is extended linearly to $\mathbb{Z}[S_i(\mathrm{Lk}(v))]$. Continuing our calculation $$ =\sum_{\substack{x\in \mathrm{Ver}(\mathrm{Lk}(v))\\ [x,s]\in S_{i}(\mathrm{Lk}(v))}}\frac{w_v([x,s])}{w_v(s)}(f([v,s])-f([v,x,ds])) $$ $$ =\sum_{\substack{x\in \mathrm{Ver}(\mathrm{Lk}(v))\\ [x,s]\in S_{i}(\mathrm{Lk}(v))}}\frac{w_v([x,s])}{w_v(s)}(\tau_vf(s)-\tau_vf([x,ds])) $$ $$ =\sum_{\substack{x\in \mathrm{Ver}(\mathrm{Lk}(v))\\ [x,s]\in S_{i}(\mathrm{Lk}(v))}}\frac{w_v([x,s])}{w_v(s)}d_v\tau_vf([x,s])=\delta_v d_v\tau_v f(s)=\Delta_v\tau_vf(s). $$ \end{comment} \end{proof} \begin{notn} Given a simplicial complex $Y$, let $M^i(Y)$ and $m^i(Y)$ be the maximal and minimal non-zero eigenvalues of $\Delta$ acting on $C^i(Y)$, respectively. Denote $$ \lambda^i_{\max}(Y)= \max_{\substack{v\in \mathrm{Ver}(Y)}}M^i(\mathrm{Lk}(v)), $$ $$ \lambda^i_{\min}(Y)= \min_{\substack{v\in \mathrm{Ver}(Y)}}m^i(\mathrm{Lk}(v)). $$ \end{notn} \begin{lem}\label{lemDec5} Assume $i\geq 1$. For $f\in C^i(X)$, we have $$ (\Delta_v \tau_v f, \tau_v f)_v\leq M^{i-1}(\mathrm{Lk}(v))\cdot(\tau_v f, \tau_v f)_v. $$ If $\tilde{H}^{i-1}(\mathrm{Lk}(v))=0$, then for $f\in \sZ^i(X)$ we have $$ (\Delta_v \tau_v f, \tau_v f)_v\geq m^{i-1}(\mathrm{Lk}(v))\cdot(\tau_v f, \tau_v f)_v. $$ \end{lem} \begin{proof} We can choose an orthonormal basis $\{e_1, \dots, e_h\}$ of $C^{i-1}(\mathrm{Lk}(v))$ with respect to $(\cdot , \cdot)_v$ which consists of $\Delta_v$-eigenvectors. Let $\{\kappa_1,\dots, \kappa_h\}$ be the corresponding eigenvalues. We have $\kappa_j\geq 0$ for all $j$. Write $\tau_vf=\sum_{j=1}^h a_j e_j$. Then $$ (\Delta_v \tau_v f, \tau_v f)_v = (\sum_{j=1}^h a_j \kappa_j e_j, \sum_{j=1}^h a_j e_j)_v=\sum_{j=1}^h a_j^2 \kappa_j $$ $$ \leq M^{i-1}(\mathrm{Lk}(v)) \sum_{j=1}^h a_j^2 = M^{i-1}(\mathrm{Lk}(v)) (\tau_v f, \tau_v f)_v. $$ The second claim will follow from a similar argument if we show that $\tau_vf$ belongs to the subspace of $C^{i-1}(\mathrm{Lk}(v))$ spanned by $\Delta_v$-eigenfunctions with \textbf{positive} eigenvalues. First assume $i\geq 2$. In this case ${H}^{i-1}(\mathrm{Lk}(v))=\tilde{H}^{i-1}(\mathrm{Lk}(v))=0$. Hence, thanks to Lemma \ref{lem1.7}, it is enough to show that $\tau_vf \in\cH^{i-1}(\mathrm{Lk}(v))$. \begin{comment} For $s\in S_{i-2}(\mathrm{Lk}(v))$ we have $$ \delta_v\tau_v f(s)=\sum_{\substack{x\in \mathrm{Ver}(\mathrm{Lk}(v))\\ [x, s]\in S_{i}(\mathrm{Lk}(v))}}\frac{w_v([x, s])}{w_v(s)}\tau_v f([x,s]) $$ $$ = \sum_{\substack{x\in \mathrm{Ver}(X)\\ [x,v, s]\in S_{i+1}(X)}}\frac{w([x,v,s])}{w([v, s])}f([v, x,s]) = -\delta f([v, s])=-\tau_v \delta f(s). $$ \end{comment} Since by assumption $\delta f=0$, from \eqref{eq-taudelta} we get $\delta_v\tau_v f=-\tau_v \delta f=0$. Thus $\tau_v f\in \sZ^{i-1}(\mathrm{Lk}(v))$. Now assume $i=1$ and $\tilde{H}^0(\mathrm{Lk}(v))=0$. This last assumption is equivalent to $\mathrm{Lk}(v)$ being connected. In this case $Z^0(\mathrm{Lk}(v))$ is spanned by the function $\mathbf{1}\in C^0(\mathrm{Lk}(v))$ which assumes value $1$ on all vertices of $\mathrm{Lk}(v)$. By Lemma \ref{lem1.7}, we need to show that $\mathbf{1}$ is orthogonal to $\tau_v f$ with respect to the inner-product \eqref{eq-locp}. We compute \begin{align} \label{eq-1tuaf} (\mathbf{1}, \tau_v f)_v &=\sum_{x\in \mathrm{Ver}(\mathrm{Lk}(v))}{w}_v(x)\cdot \tau_vf(x) \overset{\eqref{eq-w_vw}}{=}\sum_{x\in \mathrm{Ver}(\mathrm{Lk}(v))}{w}([v,x]) f([v,x])\\ \nonumber &=-w(v)\delta f(v)=0. \end{align} \end{proof} \begin{comment} To see this let $[v, v']\in S_1(X)$ and $f\in Z^0(X)$. We have $$ 0=df([v, v'])=f(v')-f(v). $$ Hence $f$ assumes the same value on adjacent vertices. But $\tilde{H}^0(X)=0$ is equivalent to $X$ being connected, so $f$ must be a constant function. \end{comment} \begin{comment} \begin{lem}\label{cor7.15} Assume $i\geq 1$. Let $f\in C^i(X)$. We have $$ (\Delta \rho_vf,\rho_vf)\leq \lambda^{i-1}_{\max}(X)\cdot(\rho_vf,f). $$ If $\tilde{H}^{i-1}(\mathrm{Lk}(v))=0$ and $\delta f(s)=0$ for every $s\in S_{i-1}(X)$ containing $v$, then $$ (\Delta \rho_vf,\rho_vf)\geq \lambda^{i-1}_{\min}(X)\cdot(\rho_vf,f). $$ \end{lem} \begin{proof} By Lemma \ref{v-prop1.13} and Lemma \ref{prop7.12}, $$ (\tau_vf,\tau_vf)_v =(\rho_vf,\rho_vf)= (\rho_vf, f). $$ On the other hand, by Lemma \ref{prop7.14} and Lemma \ref{lemDec5}, $$ (\Delta \rho_vf,\rho_vf)=(\Delta_v\tau_vf,\tau_vf)_v\leq \lambda^{i-1}_{\max}(X)\cdot (\tau_vf,\tau_vf)_v, $$ and if $\tilde{H}^{i-1}(\mathrm{Lk}(v))=0$ and $\delta f(s)=0$ for every $s\in S_{i-1}(X)$ containing $v$, then $$ (\Delta \rho_vf,\rho_vf)\geq \lambda^{i-1}_{\min}(X)\cdot (\tau_vf,\tau_vf)_v. $$ \end{proof} \end{comment} \subsection{Fundamental inequalities}\label{ss3.3} Now we are ready to prove the main results of this section. \begin{thm}\label{thmFI} Assume $i\geq 1$. For $f\in C^i(X)$ we have $$ i\cdot (\Delta f, f)\leq \left((i+1)\cdot \lambda^{i-1}_{\max}(X)-(n-i)\right)(f,f). $$ If $\tilde{H}^{i-1}(\mathrm{Lk}(v))=0$ for every $v\in \mathrm{Ver}(X)$, then for $f\in \sZ^i(X)$ we have $$ i\cdot (\Delta f, f)\geq \left((i+1)\cdot \lambda^{i-1}_{\min}(X)-(n-i)\right)(f,f). $$ \end{thm} \begin{proof} Combining Lemma \ref{lem-borel} with Lemma \ref{prop7.14}, we get \begin{equation}\label{eq-locexp} i\cdot (\Delta f, f)=\sum_{v\in \mathrm{Ver}(X)}(\Delta_v \tau_v f,\tau_v f)_v - (n-i)(f,f). \end{equation} If $\tilde{H}^{i-1}(\mathrm{Lk}(v))=0$ for every $v\in \mathrm{Ver}(X)$ and $f\in \sZ^i(X)$, then by Lemma \ref{lemDec5} \begin{align*} \sum_{v\in \mathrm{Ver}(X)}(\Delta_v \tau_v f,\tau_v f)_v & \geq \sum_{v\in \mathrm{Ver}(X)}m^{i-1}(\mathrm{Lk}(v))(\tau_v f,\tau_v f)_v \\ & \geq \lambda^{i-1}_{\min}(X) \sum_{v\in \mathrm{Ver}(X)}(\tau_v f,\tau_v f)_v. \end{align*} On the other hand, $$ \sum_{v\in \mathrm{Ver}(X)}(\tau_v f,\tau_v f)_v \overset{\mathrm{Lem.} \ref{prop7.12}}{=} \sum_{v\in \mathrm{Ver}(X)}(\rho_v f,\rho_v f) \overset{\eqref{eq-(rho)}}{=} \sum_{v\in \mathrm{Ver}(X)}(\rho_v f, f) $$ $$ = (\sum_{v\in \mathrm{Ver}(X)}\rho_v f, f) \overset{\eqref{eq-obv}}{=} (i+1)(f,f). $$ Hence \begin{equation}\label{eq-suminequl} \sum_{v\in \mathrm{Ver}(X)}(\Delta_v \tau_v f,\tau_v f)_v\geq \lambda^{i-1}_{\min}(X) (i+1)(f,f). \end{equation} Substituting this inequality into \eqref{eq-locexp} we get the second claim of the theorem. The first claim follows from a similar argument. \end{proof} \begin{cor}\label{CorFI} Assume $i\geq 1$. If $\tilde{H}^{i-1}(\mathrm{Lk}(v))=0$ for every $v\in \mathrm{Ver}(X)$ and $\lambda^{i-1}_{\min}(X)>\frac{n-i}{i+1}$, then $\harm{i}(X)=0$. \end{cor} \begin{proof} Let $f\in \harm{i}(X)$. Then $df=\delta f=0$. We obviously have $(\Delta f, f)=(df, df)=0$. Under our current assumptions, Theorem \ref{thmFI} then implies $(f,f)\leq 0$, which implies $f=0$. \end{proof} \begin{thm}\label{thmFIeigen} Assume $i\geq 1$. We have $$ i\cdot M^i(X)\leq (i+1)\cdot \lambda^{i-1}_{\max}(X)-(n-i), $$ and $$ i\cdot m^i(X)\geq (i+1)\cdot \lambda^{i-1}_{\min}(X)-(n-i). $$ \end{thm} \begin{proof} Let $f\in C^i(X)$ be an eigenfunction of $\Delta$ with non-zero eigenvalue $c\neq 0$. Then, by Lemma \ref{lem1.7} (1), $f=\delta g$ for some $g\in C^{i+1}(X)$. By \eqref{eq-taudelta} $$\tau_v f=\tau_v \delta g=-\delta_v\tau_v g.$$ Hence, again by Lemma \ref{lem1.7} (1), $\tau_v f$ belongs to the subspace of $C^{i-1}(\mathrm{Lk}(v))$ spanned by the eigenfunctions of $\Delta_v$ with positive eigenvalues. As in the proof of Lemma \ref{lemDec5}, this implies $$(\Delta_v \tau_v f, \tau_v f)_v\geq m^{i-1}(\mathrm{Lk}(v)) (\tau_v f, \tau_v f)_v \geq \lambda^{i-1}_{\min}(X) (\tau_v f, \tau_v f)_v.$$ This inequality, as in the proof of Theorem \ref{thmFI}, implies \eqref{eq-suminequl}. Combining \eqref{eq-suminequl} with \eqref{eq-locexp} we get $$ ic(f, f)=i(\Delta f, f)\geq \left((i+1)\cdot \lambda^{i-1}_{\min}(X)-(n-i)\right)(f,f), $$ which implies the second inequality of the theorem. The first inequality can be proven by a similar argument. \end{proof} \begin{rem} Note that in Theorem \ref{thmFIeigen} we do not assume $\tilde{H}^{i-1}(\mathrm{Lk}(v))=0$. Also, since $\Delta$ is not the zero operator, we must have $M^i(X)>0$ for $i\leq n-1$, which implies $\lambda_{\max}^{i-1}(X)>\frac{n-i}{i+1}$. \end{rem} \begin{notn} For $m\geq 1$, let $I_m$ denote the $m\times m$ identity matrix and let $J_m$ denote the $m\times m$ matrix whose entries are all equal to $1$. The minimal polynomial of $J_m$ is $x(x-m)$. \end{notn} \begin{example}\label{Ex-P} Let $X$ be an $n$-simplex. We claim that the eigenvalues of $\Delta$ acting on $C^i(X)$ are $0$ and $(n+1)$ for any $0\leq i \leq n-1$. It is easy to see that $0$ is an eigenvalue, so we need to show that the only non-zero eigenvalue of $\Delta$ is $(n+1)$. It is enough to show that $m^i(X)=M^i(X)=n+1$. First, suppose $i=0$. Since for any simplex of $X$ there is a unique $n$-simplex containing it, one easily checks that $\Delta$ acts on $C^0(X)$ as the matrix $(n+1)I_{n+1}-J_{n+1}$. The only eigenvalues of this matrix are $0$ and $(n+1)$. Now let $i\geq 1$. The link of any vertex is an $(n-1)$-simplex, so by induction $\lambda_{\min}^{i-1}(X)=\lambda^{i-1}_{\max}(X)=n$. Theorem \ref{thmFIeigen} implies $$ i\cdot M^i(X)\leq (i+1)n-(n-i)=i(n+1) $$ and $$ i\cdot m^i(X)\geq (i+1)n-(n-i)=i(n+1). $$ Hence $(n+1)\leq m^i(X)\leq M^i(X)\leq (n+1)$, which implies the claim. (Of course, for this simple example it is possible, but not completely trivial, to compute the eigenvalues of $\Delta$ directly.) \end{example} \begin{notn} Let $0\leq j\leq n-1$. Given $s\in \widehat{S}_j(X)$, its link $\mathrm{Lk}(s)$ in $X$ has dimension $n-(j+1)$ and satisfies $(\star)$. For $0\leq i\leq n-(j+1)$, denote $$ \lambda^{i,j}_{\max}(X)= \max_{\substack{s\in \widehat{S}_j(X)}}M^i(\mathrm{Lk}(s)), $$ $$ \lambda^{i,j}_{\min}(X)= \min_{\substack{s\in \widehat{S}_j(X)}}m^i(\mathrm{Lk}(s)). $$ With our earlier notation, we have $\lambda^{i,0}_{\max}(X)=\lambda^{i}_{\max}(X)$ and $\lambda^{i,0}_{\min}(X)=\lambda^{i}_{\min}(X)$. \end{notn} \begin{cor}\label{corFIeigen} Let $0\leq j<i\leq n-1$. We have $$ (i-j)\cdot M^i(X)\leq (i+1)\cdot \lambda^{i-(j+1),j}_{\max}(X)-(j+1)(n-i). $$ and $$ (i-j)\cdot m^i(X)\geq (i+1)\cdot \lambda^{i-(j+1),j}_{\min}(X)-(j+1)(n-i). $$ \end{cor} \begin{proof} We will prove the second inequality. The first inequality can be proven by a similar argument. If $j=0$, then the claim is just Theorem \ref{thmFIeigen}. Now, given $j\geq 1$, assume that we proved the inequality for $j-1$: \begin{equation}\label{eqNN1} (i-(j-1))\cdot m^i(X)\geq (i+1)\cdot \lambda^{i-j,j-1}_{\min}(X)-j(n-i). \end{equation} Let $s\in \widehat{S}_{j-1}(X)$. The dimension of $\mathrm{Lk}(s)$ is $n-j$, so by Theorem \ref{thmFIeigen} we have $$ (i-j)m^{i-j}(\mathrm{Lk}(s))\geq (i-j+1)\lambda^{i-j-1,0}_{\min}(\mathrm{Lk}(s))-((n-j)-(i-j)). $$ On the other hand, the link of a vertex $v\in \mathrm{Lk}(s)$ in $\mathrm{Lk}(s)$ is the same as the link of the $j$-simplex $[v, s]$ in $X$. Thus, $$ \lambda^{i-j-1,0}_{\min}(\mathrm{Lk}(s))\geq \lambda^{i-j-1,j}_{\min}(X), $$ and $$ (i-j)m^{i-j}(\mathrm{Lk}(s))\geq (i-j+1)\lambda^{i-j-1,j}_{\min}(X)-(n-i). $$ Taking the minimum over all $s\in \widehat{S}_{j-1}(X)$, we get \begin{equation}\label{eqNN2} (i-j)\lambda^{i-j,j-1}_{\min}(X)\geq (i-j+1)\lambda^{i-j-1,j}_{\min}(X)-(n-i). \end{equation} Substituting \eqref{eqNN2} into \eqref{eqNN1}, gives the desired inequality for $j$. \end{proof} The argument in the previous proof can be easily adapted to prove the following inequality: $$ (i+1)\lambda^{i-1,0}_{\min}(X)-(n-i)\geq \frac{i}{i-j}\left((i+1)\lambda^{i-j-1,j}_{\min}(X)-(j+1)(n-i)\right). $$ Now substituting this inequality into the second inequality of Theorem \ref{thmFI}, we get: \begin{cor}\label{corNN1} Assume $0\leq j< i\leq n-1$. If $\tilde{H}^{i-1}(\mathrm{Lk}(v))=0$ for every $v\in \mathrm{Ver}(X)$, then for $f\in \sZ^i(X)$ we have $$ (i-j)\cdot (\Delta f, f)\geq \left((i+1)\cdot \lambda^{i-j-1,j}_{\min}(X)-(j+1)(n-i)\right)(f,f). $$ In particular, if $\lambda^{i-j-1,j}_{\min}(X)>\frac{(j+1)(n-i)}{(i+1)}$, then $\harm{i}(X)=0$. \end{cor} \begin{rem}\label{remAddedLast} Using \eqref{eqNN2} it is easy to check that $\lambda^{i-j-1,j}_{\min}(X)>\frac{(j+1)(n-i)}{(i+1)}$ implies $\lambda^{i-k-1,k}_{\min}(X)>\frac{(k+1)(n-i)}{(i+1)}$ for all $0\leq k\leq j$. Hence the strongest assumption for Corollary \ref{corNN1} is $\lambda^{0,i-1}_{\min}(X)>\frac{i(n-i)}{(i+1)}$. The advantage in trying to prove this last inequality, besides the fact that it implies all the others, is that the question about the vanishing of $H^i(X)$ reduces to estimating the minimal non-zero eigenvalues of laplacians on graphs. \end{rem} For the purposes of proving Theorem \ref{thm1.1} we will need a variant of Corollary \ref{CorFI}. Let $X$ be an $n$-dimensional simplicial complex satisfying $(\star)$ but which is not necessarily finite. Let $\Gamma$ be a group \textit{acting} on $X$. This means that $\Gamma$ acts on the vertices of $X$ and preserves the simplicial structure of $X$, i.e., whenever the vertices $\{v_0,\dots, v_i\}$ of $X$ form an $i$-simplex, $0\leq i\leq \dim(X)$, then for any $\gamma\in \Gamma$ the vertices $\{\gamma v_0,\dots,\gamma v_i\}$ also form an $i$-simplex. Consider the following condition on the action of $\Gamma$: \begin{center} $(\dag)\quad \mathrm{St}(v)\cap \mathrm{St}(\gamma v)=\varnothing$ for any $v\in \mathrm{Ver}(X)$ and any $1\neq \gamma\in \Gamma$. \end{center} In particular, this implies that the stabilizer of any simplex is trivial. \begin{defn} Let $X/\Gamma$ be the simplicial complex whose vertices $\mathrm{Ver}(X/\Gamma)$ are the orbits $\mathrm{Ver}(X)/\Gamma$ and a subset $\{\tilde{v}_0,\dots, \tilde{v}_i\}$ of $\mathrm{Ver}(X/\Gamma)$ forms an $i$-simplex, $0\leq i\leq \dim(X)$, if we can choose a representative $v_j\in \mathrm{Ver}(X)$ from the orbit $\tilde{v}_j$ for each $0\leq j\leq i$ so that that $\{v_0, \dots, v_i\}$ form an $i$-simplex in $X$. \end{defn} It is obvious that $X/\Gamma$ is a simplicial complex. Moreover, if $(\dag)$ holds then $S_i(X/\Gamma)$ is in bijection with the orbits $S_i(X)/\Gamma$ for any $i$. Indeed, as is easy to check, $(\dag)$ implies that if $\{v_0,\dots, v_{i}\}$ and $\{\gamma_0 v_0,\dots, \gamma_{i} v_{i}\}$ are in $\widehat{S}_i(X)$ for some $\gamma_0,\dots, \gamma_i\in \Gamma$, then $\gamma_0=\cdots=\gamma_i$. \begin{cor}\label{CorFII} Let $\Gamma$ be a group acting on $X$ so that $(\dag)$ is satisfied. Assume $X/\Gamma$ is finite and $i\geq 1$. If $\tilde{H}^{i-1}(\mathrm{Lk}(v))=0$ for every $v\in \mathrm{Ver}(X)$ and $\lambda^{i-1}_{\min}(X)>\frac{n-i}{i+1}$, then $H^i(X/\Gamma)=0$. \end{cor} \begin{proof} $X/\Gamma$ is a finite $n$-dimensional complex satisfying $(\star)$. Let $v\in \mathrm{Ver}(X)$ and let $\tilde{v}$ be the image of $v$ in $X/\Gamma$. Due to $(\dag)$, $\mathrm{Lk}(v)\cong \mathrm{Lk}(\tilde{v})$. Hence $\tilde{H}^{i-1}(\mathrm{Lk}(\tilde{v}))=0$ for every $\tilde{v}\in \mathrm{Ver}(X/\Gamma)$ and $\lambda^{i-1}_{\min}(X/\Gamma)>\frac{n-i}{i+1}$. By Corollary \ref{CorFI}, we have $\harm{i}(X/\Gamma)=0$. Therefore, by Theorem \ref{lem1.11}, $H^i(X/\Gamma)=0$. \end{proof} \section{The Bruhat-Tits building of $\mathrm{SL}_n(K)$}\label{sBTB} In this section we describe the Bruhat-Tits building of $\mathrm{SL}_n$, and the links of its vertices. Then, assuming a certain lower bound on the minimal non-zero eigenvalue of the curvature transformation acting on the links, we prove Theorem \ref{thm1.1}. This is an important application of the ideas developed in the previous section. (The required lower bound on the minimal non-zero eigenvalue will be proven in Section \ref{sec3}.) We conclude the section with a brief discussion of some applications of Garland's method to produce examples of groups having the so-called property (T). \subsection{The building} Let $n\geq 0$ be a non-negative integer. Let $K$ be a complete discrete valuation field. Let $\mathcal{O}$ be the ring of integers in $K$, $\pi$ be a uniformizer, $\mathcal{O}/\pi \mathcal{O}\cong \mathbb{F}_q$, where $\mathbb{F}_q$ denotes the finite field with $q$ elements. be the residue field, and $q$ be the order of $k$. Let $\mathcal{V}$ be an $(n+2)$-dimensional vector space over $K$. A subset $L\subset \mathcal{V}$ which has a structure of a free $\mathcal{O}$-module of rank $(n+2)$ such that $L\otimes_\mathcal{O} K=\mathcal{V}$ is called a \textit{lattice}. It is clear that if $L$ is a lattice then $xL:=\{x\cdot \ell\ |\ \ell\in L\}$, $x\in K^\times$, is also a lattice. We say that $L$ and $xL$ are \textit{similar}. Similarity defines an equivalence relation on the set of lattices in $\mathcal{V}$. We denote the equivalence class of $L$ by $[L]$. The \textit{Bruhat-Tits building of $\mathrm{SL}_{n+2}(K)$} is the simplicial complex $\mathfrak B_n$ with set of vertices $\{[L]\ |\ L \text{ is a lattice in }\mathcal{V}\}$ where $\{[L_0],\dots,[L_i]\}$ form an $i$-simplex if there is $L'_j\in [L_j]$ for each $j$ with $$ \pi L_i'\subsetneqq L_0'\subsetneqq L_1'\subsetneqq\cdots\subsetneqq L'_i. $$ To visualize $\mathfrak B_n$ in some way, fix a basis $\{e_1,\dots, e_{n+2}\}$ of $\mathcal{V}$. It is easy to see that the classes of lattices $$ \mathcal{O}\pi^{a_1} e_1\oplus\cdots\oplus \mathcal{O}\pi^{a_{n+2}} e_{n+2}, \qquad a_1,\dots, a_{n+2}\in \mathbb{Z}, $$ are in bijection with the elements of $\mathbb{Z}^{n+2}/\mathbb{Z}\cdot (1,1,\dots, 1)$. In particular, $\mathfrak B_n$ is infinite. Next, the vertices corresponding to $(a_1, \dots, a_{n+2})$ and $(b_1, \dots, b_{n+2})$ are adjacent in $\mathfrak B_n$ if and only if modulo $\mathbb{Z}\cdot (1,1,\dots, 1)$ we have $a_i\leq b_i\leq a_i+1$ for all $i$. For example, when $n=0$, these vertices form an infinite line as in Figure \ref{Fig1}. When $n=1$, these vertices give a triangulation of $\mathbb{R}^2$ part of which looks like Figure \ref{Fig2}. It is important to stress that the vertices that we considered above do not give all vertices of $\mathfrak B_n$, but only of a part of the building, called an \textit{apartment}. For example, it is not hard to see that $\mathfrak B_0$ is an infinite tree in which every vertex is adjacent to exactly $(q+1)$ other vertices. Similarly, $\mathfrak B_n$ is very symmetric in the sense that the simplicial complexes $\mathrm{St}(v)$, $v\in \mathrm{Ver}(\mathfrak B_n)$, are all isomorphic to each other. To see what this complex is take a lattice $\cL$ corresponding to $v$. Then $\cL/\pi\cL= \mathbb{F}_q^{n+2}=:V$, and the vertices of $\mathrm{St}(v)$ are in one-to-one correspondence with the positive dimensional linear subspaces of $V$ ($v$ itself corresponds to $V$). Let $\{V_0,\dots,V_i\}$ be the linear subspaces corresponding to the vertices $\{v_0,\dots,v_i\}$ of $\mathrm{St}(v)$. Then $\{v_0,\dots,v_i\}$ form an $i$-simplex if and only if the linear subspaces $\{V_0,\dots,V_i\}$ fit into an ascending sequence (possibly after reindexing): $$ \mathcal{F}: V_0\subset V_1\subset \cdots \subset V_i. $$ The $i$-simplices of $\mathrm{Lk}(v)$ correspond to those $\mathcal{F}$ for which $V_i\neq V$. Next, we consider more carefully $\mathrm{Lk}(v)$ as a simplicial complex. \begin{figure \begin{tikzpicture}[scale=1.5, semithick, inner sep=.5mm, vertex/.style={circle, fill=black}] \node (0) at (0, 0) {(0,0)}; \node (1) at (1, 0) {(1,0)}; \node (2) at (2, 0) {(2,0)}; \node (3) at (3, 0) {}; \node (-1) at (-1, 0) {(0,1)}; \node (-2) at (-2, 0) {(0,2)}; \node (-3) at (-3, 0) {}; \path[] (0) edge (1) (1) edge (2) (2) edge[dashed] (3) (0) edge (-1) (-1) edge (-2) (-2) edge[dashed] (-3); \end{tikzpicture} \caption{}\label{Fig1} \end{figure} \begin{figure \begin{tikzpicture}[scale=1, semithick, inner sep=.5mm, vertex/.style={circle, fill=black}] \node (00) at (0, 0) {(0,0,0)}; \node (20) at (2, 0) {(1,0,0)}; \node (11) at (1, 1) {(1,1,0)}; \node (-11) at (-1, 1) {(0,1,0)}; \node (-20) at (-2, 0) {(0,1,1)}; \node (-1-1) at (-1, -1) {(0,0,1)}; \node (1-1) at (1, -1) {(1,0,1)}; \path[] (00) edge (20) (00) edge (11) (00) edge (-11) (00) edge (-20) (00) edge (-1-1) (00) edge (1-1); \path[] (20) edge (11) (11) edge (-11) (-11) edge (-20) (-20) edge (-1-1) (-1-1) edge (1-1) (1-1) edge (20); \end{tikzpicture} \caption{}\label{Fig2} \end{figure} \subsection{Complexes of flags}\label{ssCF} Fix some $n\geq 0$. Let $V$ be a linear space of dimension $n+2$ over the finite field $\mathbb{F}_q$. A \textit{flag} in $V$ is an ascending sequence \begin{equation}\label{eq-flag} \mathcal{F}: F_0\subset F_1\subset \cdots \subset F_i \end{equation} of distinct linear subspaces $F_0,\dots, F_i$ of $V$ such that $F_0\neq 0$ and $F_i\neq V$. The \textit{length} of $\mathcal{F}$ is $i$. We will refer to a flag of length $i$ as $i$-flag. In particular, the $0$-flags are simply the proper non-zero linear subspaces of $V$. Given two flags $$ \mathcal{F}: F_0\subset F_1\subset \cdots \subset F_i\quad \text{and} \quad \mathcal{G}: G_0 \subset G_1\subset \cdots \subset G_j $$ we say that $\mathcal{G}$ \textit{refines} $\mathcal{F}$, and write $\mathcal{F}\prec \mathcal{G}$, if for every $0\leq k\leq i$ there is $0\leq t\leq j$ such that $F_k=G_t$. It is convenient also to have the empty flag $\varnothing$, which is the empty sequence of linear subspaces; we put $-1$ for the length of $\varnothing$. The refinement defines a partial ordering on the set of flags in $V$; the empty flag is refined by every other flag. Let $\mathcal{F}$ be a fixed flag of length $\ell$. Consider the following simplicial complex $X_\mathcal{F}$ (when $\mathcal{F}=\varnothing$ we will also denote this complex by $X_\varnothing^n$). The vertices of $X_\mathcal{F}$ are the $(\ell+1)$-flags refining $\mathcal{F}$. The vertices $v_0,\dots, v_h$ form an $h$-simplex if the corresponding flags are all refined by a single $(\ell+h+1)$-flag. It is easy to see that $X_\mathcal{F}$ is indeed a finite simplicial complex of dimension $n-1-\ell$. Since any flag can be refined into an $n$-flag, $X_\mathcal{F}$ satisfies $(\star)$. Note that the link of a vertex of the Bruhat-Tits building $\mathfrak B_n$ is isomorphic to $X_\varnothing^n$. Now assume $\mathcal{F}\neq \varnothing$. Let $\mathcal{F}: F_0\subset F_1\subset \cdots \subset F_{\ell}$, with $\ell\geq 0$. Consider the array of integers $(t_0, t_1,\dots, t_{\ell+1})$ defined by \begin{equation}\label{eq-t} t_j=\left\{ \begin{array}{ll} \dim(F_0), & \mbox{if }j=0; \\ \dim(F_j)-\dim(F_{j-1}), & \mbox{if }1\leq j\leq \ell; \\ \dim(V)-\dim(F_\ell), & \mbox{if }j=\ell+1. \end{array} \right. \end{equation} It is not hard to see that \begin{equation}\label{eqXFdecomp} X_\mathcal{F}\cong X_\varnothing^{t_0-2}\ast X_\varnothing^{t_1-2}\ast\cdots \ast X_\varnothing^{t_{\ell+1}-2}, \end{equation} where $X_\varnothing^{-1}$ denotes the empty complex. \begin{lem}\label{lemn2.1} If $\dim(X_\mathcal{F})>0$ then $X_\mathcal{F}$ is connected. \end{lem} \begin{proof} First we show that $X_\varnothing^n$ is connected if $n\geq 1$. Let $x$ and $y$ be two vertices of $X_\varnothing^n$, and let $W_1$ and $W_2$ be the corresponding subspaces of $V$. Choose a $1$-dimensional subspace $L_i$ of $W_i$, $i=1,2$. Consider the subspace $P:=L_1+L_2$ of $V$. Since $n\geq 1$, $P\neq V$, so $P$ gives a vertex of $X_\varnothing^n$, which we denote by the same letter. The vertex $P$ is adjacent to both $L_1$ and $L_2$, $L_1$ is adjacent to $x$, and $L_2$ is adjacent to $y$, so there is a path from $x$ to $y$. Now assume $\mathcal{F}\neq \varnothing$ and $X_\mathcal{F}$ is given by \eqref{eqXFdecomp}. If $\dim(X_\mathcal{F})\geq 1$, then either at least two of $X_\varnothing^{t_j-2}$'s are non-empty or at least one $X_\varnothing^{t_j-2}$ has dimension $1$. In either case $X_\mathcal{F}$ is clearly connected. \end{proof} \begin{thm}\label{thm-G} Assume $N:=\dim(X_\mathcal{F})\geq 1$. Then for any $0\leq i\leq N-1$ and $\varepsilon>0$ there is a constant $q(\varepsilon, n)$ depending only on $\varepsilon$ and $n$ such that $m^i(X_\mathcal{F})\geq N-i-\varepsilon$ once $q>q(\varepsilon, n)$. \end{thm} The proof of this theorem is quite complicated and will be given in Section \ref{sec3}. \begin{cor}\label{cor-vanishing} There is a constant $q(n)$ depending only of $n$ such that if $q>q(n)$, then $\tilde{H}^i(X_\mathcal{F})=0$ for all $0\leq i\leq N-1$. \end{cor} \begin{proof} We use induction on $N$ and $i$. When $N=1$ or $i=0$, the claim follows from Lemma \ref{lemn2.1}, since $\tilde{H}^0(X_\mathcal{F})=0$ is equivalent to $X_\mathcal{F}$ being connected. Now assume $N>1$ and $i\geq 1$. For any $v\in \mathrm{Ver}(X_\mathcal{F})$, we have $\mathrm{Lk}(v)\cong X_\mathcal{G}$ for some $\mathcal{G}\succ \mathcal{F}$. Since $\dim(X_\mathcal{G})=N-1$, by the induction assumption $\tilde{H}^{i-1}(X_\mathcal{G})=0$ for $q$ large enough. Next, choosing $\varepsilon$ small enough in Theorem \ref{thm-G}, we can make $$ \lambda^{i-1}_{\min}(X_\mathcal{F})> \frac{N-i}{i+1}. $$ Now the assumptions of Corollary \ref{CorFI} are satisfied, so $\tilde{H}^i(X_\mathcal{F})=0$. \end{proof} \subsection{Main theorem} Since $\mathrm{Lk}(v)$ is isomorphic to the simplicial complex $X^n_\varnothing$, the complex $\mathfrak B_n$ is $(n+1)$-dimensional and satisfies $(\star)$. \begin{thm}\label{thm3.1} Let $\Gamma$ be a group acting on $\mathfrak B_n$ so that $(\dag)$ is satisfied. Assume $\mathfrak B_n/\Gamma$ is finite. There is a constant $q(n)$ depending only on $n$ such that if $q>q(n)$ then $H^i(\mathfrak B_n/\Gamma)=0$ for $1\leq i\leq n$. \end{thm} \begin{proof} Since $\mathrm{Lk}(v)\cong X_\varnothing^n$ for any $v\in \mathrm{Ver}(\mathfrak B_n)$, Theorem \ref{thm-G} and Corollary \ref{cor-vanishing} imply that there is a constant $q(n)$ depending only on $n$ such that if $q>q(n)$ then for any $1\leq i\leq n$ we have $\tilde{H}^{i-1}(\mathrm{Lk}(v))=0$ and $\lambda_{\min}^{i-1}(\mathfrak B_n)=m^{i-1}(X_\varnothing^n)>\frac{n+1-i}{i+1}$. Now the claim follows from Corollary \ref{CorFII}. \end{proof} \begin{rem}\label{rem-ST} It is known that the cohomology groups $\tilde{H}^i(X_\mathcal{F})$ vanish for $0\leq i\leq N-1$, without any assumptions on $q$, by a general result of Solomon and Tits; see Appendix II in \cite{Garland} for a proof. Assuming this result, to prove Theorem \ref{thm3.1} we only need the bound $m^i(X^n_\varnothing)\geq n-i-\varepsilon$. On the other hand, the proof of Theorem \ref{thm-G} is inductive, and requires proving this bound for all $X_\mathcal{F}$. Another observation is that to prove Theorem \ref{thm3.1} it is enough to prove $m^0(X_\mathcal{F})\geq N-\varepsilon$ for all $\mathcal{F}$. Indeed, the link of any simplex in $\mathfrak B_n$ is isomorphic to some $X_\mathcal{F}$, so one can appeal to Corollary \ref{corNN1} to get the vanishing of the cohomology. \end{rem} \begin{rem}\label{rem4.2} In $\S$\ref{ss4.1} we will compute that $m^0(X^1_\varnothing)>1/2$ and $m^0(X^2_\varnothing)>1$. Hence for $n=1, 2$ Garland's method proves the vanishing of $H^1(\mathfrak B_n/\Gamma)$ for all $q$. On the other hand, $m^1(X^2_\varnothing)=1/3$ when $q=2$. To apply Corollary \ref{CorFII} to show that $H^2(\mathfrak B_2/\Gamma)=0$ we need $\lambda^1_{\min}(\mathfrak B_2/\Gamma)>1/3$, so we need to assume $q>2$. \end{rem} There is an abundance of groups $\Gamma$ satisfying $(\dag)$. The most important examples of such groups come from arithmetic. One possible construction proceeds as follows. Let $F=\mathbb{F}_q(T)$ be the field of rational functions in indeterminate $T$ with $\mathbb{F}_q$ coefficients. Fix a place $\infty=1/T$ of $F$. Let $A=\mathbb{F}_q[T]$ be the polynomial ring. Let $K=\mathbb{F}_q(\!(1/T)\!)$ be the completion of $F$ at $\infty$. Let $D$ be a central division algebra over $F$ of dimension $(n+2)^2$. Assume $D$ is split at $\infty$, i.e., $D\otimes_F K\cong \mathrm{Mat}_{n+2}(K)$. Let $\cD$ be a maximal $A$-order in $D$; see \cite{Reiner} for the definitions. Let $\cD^\times$ be the group of multiplicative units in $\cD$. The quotient $\cD^\times/\mathbb{F}_q^\times$ can be identified with a discrete, cocompact subgroup of $\mathrm{PGL}_{n+2}(K)$. Replacing $\cD^\times/\mathbb{F}_q^\times$ by a subgroup $\Gamma\subset \cD^\times/\mathbb{F}_q^\times$ of finite index if necessary, we get a group which naturally acts on $\mathfrak B_n$ and satisfies $(\dag)$. Moreover, the quotient $\mathfrak B_n/\Gamma$ is finite. For these facts we refer to \cite[p. 140]{Laumon}, \cite{Li}, \cite{LSV}, \cite{Serre}. Theorem \ref{thm3.1} implies $H^i(\mathfrak B_n/\Gamma)=0$ for all $1\leq i\leq n$. On the contrary, $H^{n+1}(\mathfrak B_n/\Gamma)$ is usually quite large. Its dimension approximately equals the volume of $\mathrm{PGL}_{n+2}(K)/\Gamma$ with respect to an appropriately normalized Haar measure on $\mathrm{PGL}_{n+2}(K)$; see \cite{Serre}. The simplicial complexes $\mathfrak B_n/\Gamma$ are often used in the construction of Ramanujan complexes; see \cite{Li}, \cite{LSV}. \subsection{Property (T)} Garland's method has been applied to prove that certain groups have Kazhdan's property (T). Let $\Gamma$ be a group generated by a finite set $S$. Let $\pi:\Gamma\to U(H_\pi)$ be a unitary representation. We say that $\pi$ almost has invariant vectors if for every $\varepsilon>0$ there exists a non-zero vector $u_\varepsilon$ in the Hilbert space $H_\pi$ such that $|\!|\pi(s)u_\varepsilon-u_\varepsilon|\!|\leq \varepsilon |\!|u_\varepsilon|\!|$ for every $s\in S$. The group $\Gamma$ is said to have \textit{property (T)} if every unitary representation of $\Gamma$ which almost has invariant vectors has a non-zero invariant vector. Property (T) has important applications to representation theory, ergodic theory, geometric group theory and the theory of networks. For example, Margulis used groups with property (T) to give the first explicit examples of expanding graphs and to solve the Banach-Ruziewicz problem that asks whether the Lebesgue measure is the only normalized rotationally invariant finitely additive measure on the $n$-dimensional sphere. We refer to Lubotzky's book \cite{Lubotzky} for a discussion of property (T) and its applications. It is known that a group $\Gamma$ has property (T) if and only if for any unitary representation $\pi$ of $\Gamma$, the first cohomology group $H^1(\Gamma, \pi)$ is zero. This suggests the following line of attack to prove that $\Gamma$ has property (T). Suppose that $\Gamma$ is the fundamental group of a finite simplicial complex $X$. By group cohomology, $H^1(\Gamma, \pi)=H^1(X, E_\pi)$, where $E_\pi$ is a local system on $X$ associated to $\pi$. Then one can try to prove the vanishing of $H^1(X, E_\pi)$ by a generalization of Garland's method. This approach in the case when $X$ is a $2$-dimensional finite simplicial complex was pursued independently by Ballmann and {\'S}wiatkowski \cite{BS}, Pansu \cite{Pansu}, and \.Zuk \cite{Zuk}. For example, in \cite{BS}, the authors prove the following theorem: Assume $X$ is a $2$-dimensional finite simplicial complex, $\mathrm{Lk}(v)$ is a connected graph for any vertex of $X$, and $\lambda^0_{\min}(X)>1/2$. Then $\Gamma=\pi_1(X)$ has property (T). Note that these assumptions are the same as in Corollary \ref{CorFI} for $n=2$. They are fulfilled when $X$ is a finite quotient of a 2-dimensional Bruhat-Tits building. These results gave new explicit examples of groups with property (T) which were significantly different from the earlier known examples. In \cite{DJ1}, \cite{DJ2}, Dymara and Januszkiewicz applied a generalization of Garland's method to groups acting on buildings of arbitrary type and dimension (e.g. hyperbolic buildings), and produced examples of groups having property (T), not coming from locally symmetric spaces or euclidean buildings. \section{Complexes of flags}\label{sec3} The main goal of this section is to prove Theorem \ref{thm-G}. The notation will be the same as in $\S$\ref{ssCF}. In particular, $V$ is a linear space of dimension $n+2$ over the finite field $\mathbb{F}_q$, and $\mathcal{F}$ is a (possibly empty) flag in $V$ of length $\ell$. We denote $$ N=\dim(X_\mathcal{F}). $$ The proof of Theorem \ref{thm-G} proceeds by induction on $N$ and $i$. The base case $N=1$ follows from a direct calculation. We will carry out this calculation in $\S$\ref{ss4.1}. In the same subsection we give some explicit examples which provide a sense of the complexity of the eigenvalues of $\Delta$ acting on $C^i(X_\mathcal{F})$. These examples suggest a remarkable asymptotic behaviour of the eigenvalues of $\Delta$ as $q\to \infty$, which we state as a conjecture. The inductive step, discussed in $\S$\ref{ss4.2}, has two parts. Assuming the claim holds for $i=0$ and all $N$, the proof of the general case quickly follows from the inequality in Theorem \ref{thmFIeigen}. On the other hand, the argument which proves the claim for $i=0$ and $N\geq 1$ is fairly intricate. The outline is approximately the following: We start with a $\Delta$-eigenfunction $f\in C^0(X_\mathcal{F})$ having eigenvalue $c>0$. The machinery developed in $\S$\ref{ss3.2} cannot be applied to this function, since we cannot apply the operator $\tau_v$ directly to $f$. Instead, we introduce a parameter $R\in \mathbb{R}$, and multiply the values of $f$ on an appropriate subset of $\mathrm{Ver}(X_\mathcal{F})$ by $R$. The resulting function $f_\alpha$ is no longer an eigenfunction of $\Delta$, but we get some flexibility because we can vary $R$. We apply the machinery of $\S$\ref{ss3.2} to $df_\alpha\in C^1(X_\mathcal{F})$. Choosing $R$ appropriately forces some miraculous cancellations, which in the end give the desired bound $c\geq N-\varepsilon$. In $\S$\ref{ss4.3}, we prove some auxiliary results about the eigenvalues of curvature transformations. These results are not used elsewhere in the paper, and are given as some evidence for the conjecture in $\S$\ref{ss4.1}. \subsection{The base case and explicit examples}\label{ss4.1} For $N=1$ we need to consider only $\Delta$ acting on $C^0(X_\mathcal{F})$, since $0\leq i\leq N-1$. \begin{lem}\label{lem-n2.2} If $N=1$, then $m^0(X_\mathcal{F})$ is equal either to $1$ or $1-\frac{\sqrt{q}}{q+1}$. \end{lem} \begin{proof} If $\dim(X_\mathcal{F})=1$ then the length of $\mathcal{F}$ is $\ell=n-2$. Let $(t_0,\dots, t_{n-1})$ be defined by \eqref{eq-t}. Since $t_i\geq 1$ and $\sum_{i=0}^{n-1} (t_i-1)=2$, either exactly two $t_i$, $t_j$, $i<j$, are equal to $2$ and all others are $1$, or exactly one $t_i$ is equal to $3$ and all others are $1$. In the first case $X_\mathcal{F}\cong X_\varnothing^0\ast X_\varnothing^0$, in the second case $X_\mathcal{F}\cong X_\varnothing^1$. In the first case $X_\mathcal{F}$ is a $(q+1)$-regular bipartite graph with $2(q+1)$ vertices. \begin{comment}More precisely, $\mathrm{Ver}(X_\mathcal{F})$ is a union of two disjoint subsets $A$ and $B$ of cardinality $m=(q+1)$ each. A vertex in $A$ is not adjacent to any other vertex in $A$ but is adjacent to every vertex in $B$, and similarly, every vertex in $B$ is not adjacent to a vertex in $B$ but is adjacent to every vertex in $A$. \end{comment} It is easy to check that $(q+1)\Delta$ acts on $C^0(X_\mathcal{F})$ as the matrix $$ (q+1)I_{2(q+1)}-\begin{pmatrix} 0 & J_{q+1} \\ J_{q+1} & 0\end{pmatrix}. $$ The minimal polynomial of this matrix is $x(x-(q+1))(x-2(q+1))$, so the eigenvalues of $\Delta$ are $0$, $1$, and $2$. In the second case, $X_\mathcal{F}$ is isomorphic to the graph whose vertices correspond to $1$ and $2$-dimensional subspaces of a $3$-dimensional vector space $V$ over $\mathbb{F}_q$, two vertices being adjacent if one of the corresponding subspaces is contained in the other. With a slight abuse of terminology, we will call $1$ and $2$ dimensional subspaces lines and planes, respectively. The number of lines and planes in $V$ is $m=q^2+q+1$ each. Let $A=(a_{ij})$ be the $m\times m$ matrix whose rows are enumerated by the lines in $V$ and columns by the planes, and $a_{ij}=-1$ if the $i$th line lies in the $j$th plane, and is $0$ otherwise. We can choose a basis of $C^0(X_\mathcal{F})$ so that $(q+1)\Delta$ acts as the matrix $$ (q+1)I_{2m}+\begin{pmatrix} 0 & A \\ A^t & 0\end{pmatrix}, $$ where $A^t$ denotes the transpose of $A$. Let $M=\begin{pmatrix} 0 & A \\ A^t & 0\end{pmatrix}$. Since any two distinct lines lie in a unique plane and any line lies in $(q+1)$ planes, $AA^t=qI_m+J_m$. By a similar argument, $A^tA=qI_m+J_m$. Hence $$ M^2= qI_{2m}+\begin{pmatrix} J_m & 0 \\ 0 & J_m\end{pmatrix}. $$ This implies that $(M^2-qI_{2m})(M^2-(q+1)^2I_{2m})=0$. Since $(q+1)\Delta - (q+1)I_{2m}=M$, we conclude that $(q+1)\Delta$ satisfies the polynomial equation $$ x(x-(2q+2))(x^2-(2q+2)x+(q^2+q+1))=0. $$ It is not hard to see that this is in fact the minimal polynomial of $(q+1)\Delta$. Hence the eigenvalues of $\Delta$ are $0$, $2$, and $1\pm\frac{\sqrt{q}}{q+1}$. \end{proof} Denote by $\mathrm{min.pol}^i_n(x)$ the minimal polynomial of $\Delta$ acting on $C^i(X^n_\varnothing)$. The proof of Lemma \ref{lem-n2.2} shows that $$ \mathrm{min.pol}^0_1(x)=x(x-2)\left(x^2-2x+\frac{q^2+q+1}{q^2+2q+1}\right). $$ Note that $$m^0(X^1_\varnothing)=1- \frac{\sqrt{q}}{q+1}$$ is always strictly larger than $1/2$ and tends to $1$ as $q\to \infty$. Moreover, the whole polynomial tends coefficientwise to the polynomial $x(x-2)(x-1)^2$. Now assume $n=2$. In this case it is considerably harder to compute the minimal polynomials. With the help of a computer, we deduced that \begin{align*} \mathrm{min.pol}^0_2(x)=&x(x-2)\left(x-3\right)\left(x-\frac{2q^2+3q+2}{q^2+q+1}\right)\\ &\times\left(x^2-\frac{4q^2+3q+4}{q^2+q+1}x+\frac{4q^2+4}{q^2+q+1}\right). \end{align*} This implies $$ m^0(X^2_\varnothing)= \frac{1}{2(q^2+q+1)}\left(4q^2+3q+4-\sqrt{8q^3+9q^2+8q}\right) $$ is at least $1.08$ and tends to $2$ from below as $q\to \infty$. The whole polynomial tends coefficientwise to the polynomial $x(x-3)(x-2)^4$ as $q\to \infty$. Next \begin{align*} \mathrm{min.pol}^1_2(x)=&x(x-1)(x-2)(x-3)\\ &\times\left(x^2-2x+\frac{q^2+1}{q^2+2q+1}\right) \left(x^2-3x+\frac{2q^2+2q+2}{q^2+2q+1}\right)\\ &\times\left(x^2-4x+\frac{4q^2+6q+4}{q^2+2q+1}\right). \end{align*} In this case $$m^1(X^2_\varnothing)=1-\frac{\sqrt{2q}}{q+1}.$$ It is easy to see that $1/3\leq m^1(X^2_\varnothing)<1$. Moreover, $m^1(X^2_\varnothing)$ is strictly larger than $1/3$ for $q>2$ and tends to $1$ as $q\to \infty$; the whole polynomial tends to $x(x-3)(x-2)^4(x-1)^4$. \begin{conj}\label{conj} The previous examples, combined with some calculations for $n=3$ which we do not list, suggest a remarkable property of the eigenvalues of $\Delta$ acting on $C^i(X^n_\varnothing)$, $0\leq i\leq n-1$: \begin{enumerate} \item The number of distinct eigenvalues of $\Delta$ depends only on $i$, i.e., does not depend on $q$, even though the eigenvalues themselves and the dimension of $C^i(X^n_\varnothing)$ depend on $q$. \item The positive eigenvalues of $\Delta$, which in general are neither rational nor integral, tend to the integers $$ n-i,\ n-i+1,\ \dots,\ n+1 $$ as $q\to \infty$. \end{enumerate} \end{conj} \subsection{Inductive step}\label{ss4.2} Since we proved Theorem \ref{thm-G} for $N=1$, we assume $N\geq 2$. Let $1\leq i\leq N-1$ be given. Assume for the moment that we proved the bound in Theorem \ref{thm-G} for $\Delta$ acting on $C^{i-1}(X_\mathcal{G})$, where $\mathcal{G}$ is any flag with $\dim(\mathcal{G})= N-1$. Since for any $v\in \mathrm{Ver}(X_\mathcal{F})$ its link $\mathrm{Lk}(v)$ is isomorphic to $X_\mathcal{G}$ for some $\mathcal{G}\succ\mathcal{F}$ with $\dim(X_\mathcal{G})=N-1$, we get \begin{align*} \lambda^{i-1}_{\min}(X_\mathcal{F}) &\geq (N-1)-(i-1)-\varepsilon'=N-i-\varepsilon', \end{align*} where $\varepsilon'=i\cdot\varepsilon/(i+1)$. Then, by Theorem \ref{thmFIeigen}, we have \begin{equation}\label{eqNNN2} m^i(X_\mathcal{F}) \geq \frac{(i+1)\cdot \lambda^{i-1}_{\min}(X_\mathcal{F})-(N-i)}{i} \geq N-i-\varepsilon. \end{equation} Therefore, to complete the proof of Theorem \ref{thm-G} it remains to show that \begin{equation}\label{eqNNN} m^0(X_\mathcal{F})\geq N-\varepsilon. \end{equation} This will occupy the rest of this subsection. \begin{rem} Instead of induction, one can deduce the lower bound \eqref{eqNNN2} directly from \eqref{eqNNN} using Corollary \ref{corFIeigen}. Indeed, the link of any $(i-1)$-simplex in $X_\mathcal{F}$ is isomorphic to some $X_\mathcal{G}$ with $\dim(X_\mathcal{G})=N-i$, so assuming $m^0(X_\mathcal{G})\geq (N-i)-\varepsilon'$, $\varepsilon'=\varepsilon/(i+1)$, Corollary \ref{corFIeigen} gives $$ m^i(X_\mathcal{F})\geq (i+1)(N-i-\varepsilon')-i(N-i)= N-i-\varepsilon $$ On the other hand, the proof of Corollary \ref{corFIeigen} uses similar inductive argument as above. \end{rem} We start by proving some preliminary lemmas. For an integer $m\geq 1$ we put $(m)_q=\prod_{k=1}^m(q^k-1)$, and we put $(0)_q=1$. The number of $d$-dimensional subspaces in an $m$-dimensional linear space over $\mathbb{F}_q$ is equal to $$ \gc{m}{d}:=\frac{(m)_q}{(d)_q(m-d)_q}. $$ With this notation it is easy to give a formula for the number of $n$-flags refining a given flag: \begin{lem}\label{lem-3.1} Let $s$ be a simplex in $X_\mathcal{F}$ corresponding to $\mathcal{G}\succ \mathcal{F}$. Let $(r_0,\dots, r_{j})$ be the integers defined for $\mathcal{G}$ by \eqref{eq-t}. The number of $N$-simplices in $X_\mathcal{F}$ containing $s$ is given by the formula $$ w(s)=\prod_{k=0}^j\prod_{z=1}^{r_k}\gc{z}{1}=\prod_{k=0}^{j}(r_k)_q/(1)^{r_k}_q. $$ \end{lem} Let $v\in \mathrm{Ver}(X_\mathcal{F})$ and let $\mathcal{G}$ be the corresponding $(\ell+1)$-flag. There is a unique subspace $G$ in the sequence of $\mathcal{G}$ which does not occur in $\mathcal{F}$. Let $$\mathrm{Type}(v):=\dim(G).$$ Denote the set of types of vertices of $X_\mathcal{F}$ by $\mathfrak T$. It is easy to see that the vertices of a simplex in $X_\mathcal{F}$ have distinct types. Moreover, $\# \mathfrak T=N+1$. \begin{lem}\label{eq-yet} Let $v\in \mathrm{Ver}(X_\mathcal{F})$. Assume $\alpha\in \mathfrak T$ is fixed and $\alpha\neq \mathrm{Type}(v)$. Then $$ \sum_{\substack{x\in \mathrm{Ver}(X_\mathcal{F})\\ [v, x]\in \widehat{S}_{1}(X_\mathcal{F}) \\ \mathrm{Type}(x)=\alpha}}w([v,x]) = w(v). $$ \end{lem} \begin{proof} Let $\mathcal{G}$ be the flag of length $i:=\ell+1$ in $V$ corresponding to $v$. Let $(t_0,\dots, t_{i+1})$ be the array \eqref{eq-t} of $\mathcal{G}$. Let $[v, x]\in \widehat{S}_{1}(X_\mathcal{F})$ and $\mathcal{G}'\succ \mathcal{G}$ be the corresponding $(i+1)$-flag. There is a unique $t_a$ such that the array of $\mathcal{G}'$ is $(t_0,\dots,t_a', t_a'',\cdots, t_{i+1})$ with $t_a'+t_a''=t_a$. Moreover, the type of $x$ uniquely determines $a$ and $t_a'$. The number of $[v, x]\in \widehat{S}_{1}(X_\mathcal{F})$ with $\mathrm{Type}(x)=\alpha$ is equal to $\gc{t_a}{t_a'}$. Using Lemma \ref{lem-3.1}, we compute $$ \sum_{\substack{x\in \mathrm{Ver}(X_\mathcal{F})\\ [v, x]\in \widehat{S}_{1}(X_\mathcal{F}) \\ \mathrm{Type}(x)=\alpha}}\frac{w([v,x])}{w(v)} =\gc{t_a}{t_a'}\frac{(1)_q^{t_a} (t_a')_q(t_a'')_q}{(t_a)_q(1)_q^{t_a'}(1)_q^{t_a''}}=1. $$ \end{proof} \begin{rem} Lemma \ref{eq-yet} is a refined version of Lemma \ref{lem-w}. Indeed, $$ \sum_{\substack{x\in \mathrm{Ver}(X_\mathcal{F})\\ [v, x]\in \widehat{S}_{1}(X_\mathcal{F}) }}w([v,x]) = \sum_{\substack{\alpha\in \mathfrak T\\ \alpha\neq \mathrm{Type}(v)}} \sum_{\substack{x\in \mathrm{Ver}(X_\mathcal{F})\\ [v, x]\in \widehat{S}_{1}(X_\mathcal{F}) \\ \mathrm{Type}(x)=\alpha}}w([v,x]) = \sum_{\substack{\alpha\in \mathfrak T\\ \alpha\neq \mathrm{Type}(v)}} w(v) = Nw(v). $$ \end{rem} Let $f\in C^0(X_\mathcal{F})$ and let $R\in \mathbb{R}$ be a fixed constant. For each $\alpha\in \mathfrak T$ define the function $f_\alpha\in C^0(X_\mathcal{F})$ by $$ f_\alpha(v) = \begin{cases} R\cdot f(v), & \text{if $\mathrm{Type}(v)= \alpha$};\\ f(v), & \text{if $\mathrm{Type}(v)\neq \alpha$}. \end{cases} $$ Also, for $i\geq 0$ define a linear transformation $\rho_\alpha: C^i(X_\mathcal{F})\to C^i(X_\mathcal{F})$ by $$ \rho_\alpha=\sum_{\substack{v\in \mathrm{Ver}(X_\mathcal{F})\\ \mathrm{Type}(v)=\alpha}}\rho_v. $$ \begin{lem} We have \begin{align} \label{eq-ny3} \sum_{\alpha\in \mathfrak T}(1-\rho_\alpha)df&=(N-1)df, \\ \label{eq-ny} \sum_{\substack{v\in \mathrm{Ver}(X_\mathcal{F})\\ \mathrm{Type}(v)=\alpha}} (\Delta\rho_v df_\alpha, \rho_vdf_\alpha) &=(\Delta \rho_\alpha df_\alpha, \rho_\alpha df_\alpha), \\ \label{lem2.3} (\rho_\alpha df_\alpha, df_\alpha) & =(df_\alpha, df_\alpha)-((1-\rho_\alpha)df, df),\\ \label{lem2.2} (\Delta\rho_\alpha df_\alpha, \rho_\alpha df_\alpha) &=((1-\rho_\alpha)df,df). \end{align} \end{lem} \begin{proof} Equation \eqref{eq-ny3} follows from a straightforward calculation: $$ \sum_{\alpha\in \mathfrak T}(1-\rho_\alpha)df=(N+1)df-\sum_{v\in \mathrm{Ver}(X_\mathcal{F})}\rho_v df \nonumber =(N+1)df-2df=(N-1)df. $$ To prove \eqref{eq-ny}, expand its right hand-side as $$ (d \rho_\alpha df_\alpha, d\rho_\alpha df_\alpha)=\sum_{\substack{v, v'\in \mathrm{Ver}(X_\mathcal{F})\\ \mathrm{Type}(v)=\mathrm{Type}(v)=\alpha}} (d\rho_v df_\alpha, d\rho_{v'} df_\alpha). $$ Now let $s=[x,y,z]\in S_2(X_\mathcal{F})$. Since the vertices of the same simplex have distinct types, only one of $x,y,z$ can be of type $\alpha$. Therefore, if $v\neq v'$ but $\mathrm{Type}(v)=\mathrm{Type}(v)=\alpha$, then $d\rho_v df_\alpha(s)\cdot d\rho_{v'} df_\alpha(s)=0$. This implies that in the above sum only the terms with $v=v'$ are possibly non-zero, so $$ \sum_{\substack{v, v'\in \mathrm{Ver}(X_\mathcal{F})\\ \mathrm{Type}(v)=\mathrm{Type}(v)=\alpha}} (d\rho_v df_\alpha, d\rho_{v'} df_\alpha) = \sum_{\substack{v \in \mathrm{Ver}(X_\mathcal{F})\\ \mathrm{Type}(v)=\alpha}} (d\rho_v df_\alpha, d\rho_v df_\alpha). $$ To prove \eqref{lem2.3}, note that if $s\in S_1(X_\mathcal{F})$ contains a vertex of type $\alpha$, then $(1-\rho_\alpha)g(s)=0$ for any $g\in C^1(X)$. On the other hand, if $s$ does not contain a vertex of type $\alpha$, then $(1-\rho_\alpha)df_\alpha(s)= df_\alpha(s)=df(s)=(1-\rho_\alpha)df(s)$. Hence $$((1-\rho_\alpha)df_\alpha, df_\alpha)=((1-\rho_\alpha)df, df).$$ Now $$ ((1-\rho_\alpha)df, df)=((1-\rho_\alpha)df_\alpha, df_\alpha)=(df_\alpha, df_\alpha)-(\rho_\alpha df_\alpha, df_\alpha). $$ Finally, to prove \eqref{lem2.2}, let $s=[x,y,z]\in S_2(X_\mathcal{F})$. If none of the vertices of $s$ has type $\alpha$ then $d\rho_\alpha d f_\alpha(s)=0$. If $s$ has a vertex of type $\alpha$, then such a vertex is unique. Without loss of generality, assume $\mathrm{Type}(x)=\alpha$. Then $$ d\rho_\alpha d f_\alpha([x,y,z])=f(y)-f(z)=-df([y,z]). $$ Hence $$ (d\rho_\alpha d f_\alpha, d\rho_\alpha d f_\alpha)=\sum_{\substack{v\in \mathrm{Ver}(X_\mathcal{F})\\ \mathrm{Type}(v)=\alpha}}\sum_{s\in \widehat{S}_1(\mathrm{Lk}(v))}w([v,s])df(s)^2 $$ $$ =\sum_{s\in \widehat{S}_1(X_\mathcal{F})} (1-\rho_\alpha)df(s)\cdot df(s) \sum_{\substack{v\in \mathrm{Ver}(\mathrm{Lk}(s))\\ \mathrm{Type}(v)=\alpha}}w([v,s])= ((1-\rho_\alpha)df, df), $$ where in the last equality we used Lemma \ref{eq-yet}. \end{proof} \begin{lem}\label{lemd31} Let $f\in C^0(X_\mathcal{F})$ and suppose $\Delta f=c\cdot f$. Then $$ \sum_{\alpha\in \mathfrak T}(\Delta f_\alpha, f_\alpha)=\left[(N-c)(R-1)^2+c(R^2+N)\right]\cdot (f,f). $$ \end{lem} \begin{proof} Fix some type $\alpha$ and let $g\in C^0(X_\mathcal{F})$ be a function such that $g(v)=0$ if $\mathrm{Type}(v)\neq \alpha$. Then $(\Delta g, g)=N\cdot (g,g)$. Indeed, \begin{align*} (\Delta g, g) &=(dg,dg)=\sum_{[x,v]\in \widehat{S}_1(X_\mathcal{F})}w([x,v])(g(v)-g(x))^2 \\ & =\sum_{\substack{v\in \mathrm{Ver}(X_\mathcal{F})\\ \mathrm{Type}(v)=\alpha}}g(v)^2\sum_{\substack{x\in \mathrm{Ver}(X_\mathcal{F})\\ [x,v]\in \widehat{S}_1(X_\mathcal{F})}}w([x,v]) \\ &\overset{\mathrm{Lem.} \ref{lem-w}}{=}N\sum_{\substack{v\in \mathrm{Ver}(X_\mathcal{F})\\ \mathrm{Type}(v)=\alpha}}w(v)\cdot g(v)^2=N\cdot (g,g). \end{align*} If we apply this to $g=f_\alpha-f$, then we get \begin{equation}\label{eq-d31} (\Delta f_\alpha, f_\alpha)=N\cdot (f_\alpha, f_\alpha)-2(N-c)(f_\alpha, f)+(N-c)(f,f). \end{equation} Since the cardinality of $\mathfrak T$ is $(N+1)$, $$ \sum_{\alpha\in \mathfrak T}f_\alpha= (N+R)\cdot f\quad \text{and} \quad \sum_{\alpha\in \mathfrak T}(f_\alpha, f_\alpha)=(N+R^2)\cdot (f,f). $$ Summing \eqref{eq-d31} over all types and using the previous two equalities, we get the claim. \end{proof} \begin{prop}\label{prop4.15-15} For any $\varepsilon>0$ there is a constant $q(\varepsilon, n)$ depending only on $\varepsilon$ and $n$, such that if $q>q(\varepsilon, n)$ then $m^0(X_\mathcal{F})\geq N-\varepsilon$. \end{prop} \begin{proof} Since Lemma \ref{lem-n2.2} implies this claim for $N=1$, we can assume from now on that $N\geq 2$. Let $f\in C^0(X_\mathcal{F})$ and suppose $\Delta f=c\cdot f$. If $\mathrm{Type}(v)=\alpha$, then \begin{align} \Delta f_\alpha(v) = \sum_{\substack{x\in \mathrm{Ver}(X_\mathcal{F})\\ [x,v]\in S_1(X_\mathcal{F})}}\frac{w([x,v])}{w(v)}(Rf(v)-f(x)) =NRf(v)-C, \end{align} where $C$ does not depend on $R$ since $\mathrm{Type}(x)\neq \mathrm{Type}(v)$ if $[x,v]\in S_1(X_\mathcal{F})$. If we take $R=1$, then $f_\alpha=f$, so $\Delta f_\alpha(v)=c\cdot f(v)$. We conclude that $C=(N-c)f(v)$, and $\Delta f_\alpha(v)=(NR-(N-c))f(v)$. From now on we assume that $R=(N-c)/N$. With this choice of $R$ our calculation implies \begin{equation}\label{eq-deltaalpha} \Delta f_\alpha(v)=0\quad \text{if } \mathrm{Type}(v)=\alpha. \end{equation} Let $v\in \mathrm{Ver}(X_\mathcal{F})$ be a vertex of type $\alpha$. By Lemma \ref{prop7.14}, $$ (\Delta \rho_v df_\alpha, \rho_vdf_\alpha)= (\Delta_v\tau_v df_\alpha, \tau_v df_\alpha)_v. $$ Since $$ (\mathbf{1}, \tau_v df_\alpha)_v \overset{\eqref{eq-1tuaf}}{=} -w(v)\delta df_\alpha(v)= -w(v)\Delta f_\alpha(v) \overset{\eqref{eq-deltaalpha}}{=} 0, $$ we can use the argument in the proof of Lemma \ref{lemDec5} to conclude $$ (\Delta_v\tau_v df_\alpha, \tau_v df_\alpha)_v\geq m^{0}(\mathrm{Lk}(v))(\tau_v df_\alpha, \tau_v df_\alpha)_v \geq \lambda^0_{\min}(X_\mathcal{F}) (\tau_v df_\alpha, \tau_v df_\alpha)_v. $$ (Note that $\mathrm{Lk}(v)$ is connected since $\mathrm{Lk}(v)\cong X_\mathcal{G}$ for some $\mathcal{G}$ with $\dim(X_\mathcal{G})=N-1\geq 1$.) Hence \begin{align*} (\Delta \rho_v df_\alpha, \rho_vdf_\alpha)\geq \lambda^0_{\min}(X_\mathcal{F}) (\tau_v df_\alpha, \tau_v df_\alpha)_v &\overset{\mathrm{Lem}. \ref{prop7.12}}{=} \lambda^0_{\min}(X_\mathcal{F}) (\rho_v df_\alpha, \rho_v df_\alpha) \\ & \overset{\eqref{eq-(rho)}}{=}\lambda^0_{\min}(X_\mathcal{F}) (\rho_v df_\alpha, df_\alpha). \end{align*} Summing these inequalities over all vertices of type $\alpha$ and using \eqref{eq-ny}, we get $$ (\Delta \rho_\alpha df_\alpha, \rho_\alpha df_\alpha)\geq \lambda^0_{\min}(X_\mathcal{F})\cdot (\rho_\alpha df_\alpha, df_\alpha). $$ Using \eqref{lem2.3} and \eqref{lem2.2}, we can rewrite this inequality as $$ (1+\lambda^0_{\min}(X_\mathcal{F}))((1-\rho_\alpha)df,df)\geq \lambda^0_{\min}(X_\mathcal{F})\cdot (df_\alpha, df_\alpha) $$ Summing these inequalities over all types and using \eqref{eq-ny3} and Lemma \ref{lemd31}, we get \begin{equation}\label{eq2.1} (1+\lambda^0_{\min}(X_\mathcal{F}))(N-1)c\geq \lambda^0_{\min}(X_\mathcal{F})\cdot \left[(N-c)(R-1)^2+c(R^2+N)\right]. \end{equation} Suppose $c=m^0(X_\mathcal{F})$. If $c\geq N$, then we are done. On the other hand, if $c<N$, then $(N-c)(R-1)^2$ is positive, so \eqref{eq2.1} implies $$ (1+\lambda^0_{\min}(X_\mathcal{F}))(N-1)c \geq \lambda^0_{\min}(X_\mathcal{F})c (R^2+N). $$ Dividing both sides by $c$ (recall that $c>0$), we get $$ N-1\geq (1+R^2)\lambda^0_{\min}(X_\mathcal{F}). $$ By induction on $N$, for any $\varepsilon>0$ there is a constant $q(\varepsilon, n)$ such that $\lambda^0_{\min}(X_\mathcal{F})\geq N-1-\varepsilon$ if $q\geq q(\varepsilon, n)$. Thus $$ \varepsilon\geq R^2(N-1-\varepsilon). $$ We see that $R^2\to 0$ as $\varepsilon\to 0$. Since $R=(N-c)/N$, this forces $c\to N$. \end{proof} \subsection{Auxiliary results about eigenvalues}\label{ss4.3} In this subsection we prove that $M^i(X_\mathcal{F})\leq N+1$ and $m^0(X_\mathcal{F})\leq N$. This implies that if we allow $q$ to vary, then the lower bound $m^0(X_\mathcal{F})\geq N-\varepsilon$ in Proposition \ref{prop4.15-15} is optimal; in other terms, $m^0(X_\mathcal{F})\to N$ as $q\to \infty$, which is consistent with Conjecture \ref{conj}. We also show that $M^0(X_\mathcal{F})=N+1$ and its multiplicity is $N$, so does not depend on $q$. \begin{prop}\label{prop_ny} For all $0\leq i\leq N-1$ we have $M^i(X_\mathcal{F})\leq N+1$. \end{prop} \begin{proof} The proof is again by induction on $N$. If $N=1$, then the calculations in the proof of Lemma \ref{lem-n2.2} show that $M^0(X_\mathcal{F})=2$. Now assume $N\geq 2$ and $i\geq 1$. Assume we proved that $M^{i-1}(X_\mathcal{G})\leq N$ for any $\mathcal{G}$ with $\dim(X_\mathcal{G})=N-1$. Then $\lambda_{\max}^{i-1}(X_\mathcal{F})\leq N$, so Theorem \ref{thmFIeigen} implies $M^i(X_\mathcal{F})\leq N+1$. It remains to prove that $M^0(X_\mathcal{F})\leq N+1$. Let $f\in C^0(X_\mathcal{F})$. By an argument very similar to the proof of Proposition \ref{prop4.15-15} we get \begin{align*} (\Delta \rho_v df_\alpha, \rho_vdf_\alpha) = (\Delta_v\tau_v df_\alpha, \tau_v df_\alpha)_v &\leq \lambda^0_{\max}(X_\mathcal{F})(\tau_v df_\alpha, \tau_v df_\alpha)_v \\ & = \lambda^0_{\max}(X_\mathcal{F})\cdot (\rho_v df_\alpha, df_\alpha), \end{align*} which leads to $$ (\Delta \rho_\alpha df_\alpha, \rho_\alpha df_\alpha)\leq \lambda^0_{\max}(X_\mathcal{F})\cdot (\rho_\alpha df_\alpha, df_\alpha). $$ By induction, $\lambda^0_{\max}(X_\mathcal{F})\leq N$, so using \eqref{lem2.3} and \eqref{lem2.2}, we can rewrite the previous inequality as $$ (1+N)\cdot ((1-\rho_\alpha)df,df)\leq N\cdot (df_\alpha, df_\alpha). $$ Assume $\Delta f=c\cdot f$ is an eigenfunction. Summing the above inequalities over all types and using \eqref{eq-ny3} and Lemma \ref{lemd31}, we get $$ (N+1)(N-1)c\leq N\cdot \left[(N-c)(R-1)^2+c(R^2+N)\right]. $$ If we put $R=(N-c)/N$, then this inequality forces $c\leq N+1$. In particular, $M^0(X_\mathcal{F})\leq N+1$. \end{proof} Let $Y$ be an $N$-simplex. Since $Y$ has a unique simplex of maximal dimension, the weights \eqref{eq-themetric} of the simplices of $Y$ are all equal to $1$. Then, relative to the inner-product \eqref{eq-pairing}, we have the orthogonal direct sum decomposition (cf. Lemma \ref{lem1.7}) $$ C^0(Y)=\mathbb{R}\mathbf{1}\oplus \delta C^1(Y), $$ where $\delta C^1(Y)$ can be explicitly described as the space of functions satisfying $$\sum_{v\in \mathrm{Ver}(Y)}g(v)=0.$$ It is easy to check that $\Delta g = 0$ if and only if $g\in \mathbb{R}\mathbf{1}$, and $\Delta g= (N+1)g$ if and only if $g\in \delta C^1(Y)$. Hence $0$ and $N+1$ are the only eigenvalues of $\Delta$ acting on $C^0(Y)$, and their multiplicities are $1$ and $N$, respectively. \begin{defn} We say that $f\in C^0(X_\mathcal{F})$ is \textit{type-constant} if $f(v)=f(v')$ for all $v, v'\in \mathrm{Ver}(X_\mathcal{F})$ of the same type. We denote the space of type-constant functions by $\mathscr{C}$. \end{defn} Label the vertices of $Y$ by the elements of $\mathfrak T$. Given a function $f\in \mathscr{C}$, define $\tilde{f}\in C^0(Y)$ by $\tilde{f}(x)=c_{\alpha}(f)$, $x\in \mathrm{Ver}(Y)$, where $\alpha=\mathrm{Type}(x)$ and $c_\alpha(f)$ is the value of $f$ on vertices of type $\alpha\in \mathfrak T$. It is clear that $\mathscr{C}\to C^0(Y)$, $f\mapsto \tilde{f}$, is an isomorphism of vector spaces which restricts to an isomorphism $\mathscr{C}_0 \xrightarrow{\sim} \delta C^1(Y)$, where $\mathscr{C}_0\subset \mathscr{C}$ is the subspace of functions $f\in \mathscr{C}$ satisfying $$ \sum_{\alpha\in \mathfrak T}c_\alpha(f)=0. $$ \begin{lem}\label{lemNNC} If $f\in \mathscr{C}$, then $\Delta f\in \mathscr{C}$. Moreover, $\widetilde{\Delta f} = \Delta \tilde{f}$. This implies that for $f\in \mathscr{C}_0$ we have $\Delta f= (N+1)f$. \end{lem} \begin{proof} For a fixed $v\in \mathrm{Ver}(X_\mathcal{F})$, we have $$ \Delta f(v)=\sum_{\substack{x\in \mathrm{Ver}(X_\mathcal{F})\\ [x,v]\in S_1(X_\mathcal{F})}}\frac{w([x,v])}{w(v)}(f(v)-f(x)) $$ $$ \overset{\mathrm{Lem.} \ref{lem-w}}{=}N f(v)- \sum_{\substack{x\in \mathrm{Ver}(X_\mathcal{F})\\ [x,v]\in S_1(X_\mathcal{F})}}\frac{w([x,v])}{w(v)}f(x) $$ $$ = N f(v)- \sum_{\substack{\alpha\in \mathfrak T\\ \alpha\neq \mathrm{Type}(v)}} \sum_{\substack{x\in \mathrm{Ver}(X_\mathcal{F})\\ \mathrm{Type}(x)=\alpha \\ [x,v]\in S_1(X_\mathcal{F})}}\frac{w([x,v])}{w(v)}f(x). $$ Now $$ \sum_{\substack{x\in \mathrm{Ver}(X_\mathcal{F})\\ \mathrm{Type}(x)=\alpha \\ [x,v]\in S_1(X_\mathcal{F})}}\frac{w([x,v])}{w(v)}f(x) = c_\alpha(f)\sum_{\substack{x\in \mathrm{Ver}(X_\mathcal{F})\\ \mathrm{Type}(x) =\alpha \\ [x,v]\in S_1(X_\mathcal{F})}}\frac{w([x,v])}{w(v)}\overset{\mathrm{Lem.}\ref{eq-yet}}{=} c_\alpha(f). $$ Thus, if we denote $\beta=\mathrm{Type}(v)$, $$ \Delta f(v) = N c_\beta(f) - \sum_{\substack{\alpha\in \mathfrak T\\ \alpha\neq \beta}} c_\alpha(f). $$ It is clear from this that $\Delta f\in \mathscr{C}$. Moreover, for any $z\in \mathrm{Ver}(Y)$ we have $$ \widetilde{\Delta f}(z)=N\tilde{f}(z)-\sum_{\substack{y\in \mathrm{Ver}(Y_\mathcal{F})\\ y\neq z}} \tilde{f}(y)=\Delta\tilde{f}(z). $$ \end{proof} \begin{lem}\label{lemNNC2} Let $f\in C^0(X_\mathcal{F})$ be a $\Delta$-eigenfunction with eigenvalue $c$. If $c=0$ or $N+1$, then $f\in \mathscr{C}$. \end{lem} \begin{proof} Using the fact that $X_\mathcal{F}$ is connected, it is easy to show that $\Delta f=0$ if and only if $f$ is constant. We prove that if $\Delta f=(N+1)f$, then $f$ is type-constant. First, assume $N=1$. Then either $X_\mathcal{F}=X_\varnothing^0\ast X_\varnothing^0$ or $X_\mathcal{F}=X_\varnothing^1$. In either case, the matrix of $m\Delta$ has the form $$ \begin{pmatrix} m I_m & -A \\ -A^t & m I_m \end{pmatrix}, $$ where $A$ gives the adjacency relations between vertices of type $\alpha$ and $\beta$ (there are only two types), and either $m=q+1$ or $m=q^2+q+1$. Let $\mathbf{x}=(x_1, x_2, \dots, x_{2m})^t$ be an eigenvector with eigenvalue $2m$. Let $\mathbf{x}_\alpha=(x_1, x_2, \dots, x_{m})^t$ and $\mathbf{x}_\beta=(x_{m+1}, x_2, \dots, x_{2m})^t$. Then \begin{align*} mI_m \mathbf{x}_\alpha - A\mathbf{x}_\beta &= 2m \mathbf{x}_\alpha \\ -A^t \mathbf{x}_\alpha + mI_m \mathbf{x}_\beta &= 2m \mathbf{x}_\beta. \end{align*} Hence $m \mathbf{x}_\alpha = -A\mathbf{x}_\beta$ and $m \mathbf{x}_\beta =-A^t \mathbf{x}_\alpha$. This implies $m^2 \mathbf{x}_\alpha = AA^t \mathbf{x}_\alpha$. In the first case, $AA^t=m J_m$, so $m \mathbf{x}_\alpha = J_m \mathbf{x}_\alpha$. This implies that $m x_j=\sum_{i=1}^m x_i$ for all $1\leq j\leq m$. Hence $x_1=x_2=\dots=x_m$. Similarly, one shows that $x_{m+1}=\dots=x_{2m}$. In the second case, $AA^t=qI_m+J_m$, so $$ m^2 \mathbf{x}_\alpha= q \mathbf{x}_\alpha + J_m \mathbf{x}_\alpha. $$ Hence $(m^2-q) x_j=\sum_{i=1}^m x_i$ for all $1\leq j\leq m$, which again implies $x_1=x_2=\dots=x_m$. Similarly, one shows that $x_{m+1}=\dots=x_{2m}$, since $A^t A=qI_m+J_m$. Now assume $N>1$ and that we proved the claim for all $X_\mathcal{F}$ of dimension less than $N$. Suppose $f$ is not type-constant. Then there are two vertices $x, y$ of the same type such that $f(x)\neq f(y)$. We claim that we can choose $x$ and $y$ so that there is a vertex $v\in \mathrm{Ver}(X_\mathcal{F})$ such that $x, y\in \mathrm{Lk}(v)$. We start with $X_\mathcal{F}=X_\varnothing^n$. In that case $x$ and $y$ correspond to subspaces $W_1$ and $W_2$ of $\mathbb{F}_q^{n+2}$ of the same dimension. By assumption $n>1$. If $\dim(W_i)=1$, then $v$ corresponding to $W_1+W_2$ is adjacent to both $x$ and $y$. (Note that $\dim(W_1+W_2)=2<n+2$.) If $r=\dim(W_i)>1$, choose a line $\ell_i\in W_i$. Let $W_3$ be a subspace of dimension $r$ which contains $\ell_1+\ell_2$. Let $z$ be the corresponding vertex. If $f(x)=f(z)$, then we replace $x$ by $z$ and take $v$ corresponding to $\ell_2$. If $f(x)\neq f(z)$, then we replace $y$ by $z$ and take $v$ corresponding to $\ell_1$. Now suppose $X_\mathcal{F}\cong X_\varnothing^{n_1}\ast \cdots \ast X_\varnothing^{n_s}$, $s\geq 2$. Our vertices are in the same $X_\varnothing^{n_i}$ since they have the same type, but then any vertex in another $X_\varnothing^{n_j}$ is adjacent to both $x$ and $y$. Let $x, y\in \mathrm{Lk}(v)$ be as in the previous paragraph. Let $\mathrm{Type}(v)=\alpha$. Consider the function $\tau_vdf_\alpha$. We have $$ \tau_vdf_\alpha(x)=df_\alpha([v,x])=f(x)-Rf(v)\neq f(y)-Rf(v) = \tau_vdf_\alpha(y). $$ Hence $\tau_vdf_\alpha\in C^0(\mathrm{Lk}(v))$ is not type-constant. By induction, $\tau_vdf_\alpha$ does not lie in the subspace of $C^0(\mathrm{Lk}(v))$ spanned by eigenfunctions with eigenvalue $N$. This implies (use the orthonormal decomposition of Lemma \ref{lemDec5} and Proposition \ref{prop_ny}) $$ (\Delta_v \tau_vdf_\alpha, \tau_vdf_\alpha)_v< N (\tau_vdf_\alpha, \tau_vdf_\alpha)_v. $$ This inequality implies, as in the proof of Proposition \ref{prop_ny}, that $$ (\Delta \rho_\alpha df_\alpha, \rho_\alpha df_\alpha) < N (\rho_\alpha df_\alpha, df_\alpha). $$ As in the proof of Proposition \ref{prop_ny}, this leads to $$ (N+1)(N-1)c< N[(N-c)(R-1)^2+(N+1)(R^2+N)], $$ where $c=N+1$ and $R=(N-c)/N$. But for these $c$ and $R$ both sides are equal, so the inequality cannot be strict. \end{proof} \begin{prop} A function $f\in C^0(X_\mathcal{F})$ is a $\Delta$-eigenfunction with eigenvalue $0$ if and only if $f$ is constant. A function $f\in C^0(X_\mathcal{F})$ is a $\Delta$-eigenfunction with eigenvalue $N+1$ if and only if $f\in \mathscr{C}_0$. This implies that $M^0(X_\mathcal{F})=N+1$ and its multiplicity as an eigenvalue of $\Delta$ is $N$. \end{prop} \begin{proof} It is easy to check that $\Delta f = 0\Leftrightarrow df=0\Leftrightarrow f$ is constant (since $X_\mathcal{F}$ is connected). By Lemma \ref{lemNNC}, if $f\in \mathscr{C}_0$, then $\Delta f=(N+1)f$. Conversely, suppose $\Delta f=(N+1)f$. By Lemma \ref{lemNNC2}, $f$ is type-constant, so by Lemma \ref{lemNNC}, $$ \widetilde{\Delta f} = \widetilde{(N+1) f} = (N+1)\tilde{f}=\Delta \tilde{f}. $$ This implies $f\in \mathscr{C}_0$. \end{proof} \begin{rem} In \cite{PapMM}, we proved a general result about finite buildings which implies that $M^i(X_\mathcal{F})=N+1$ for all $0\leq i\leq N-1$. \end{rem} \begin{prop}\label{thm-last} $m^0(X_\mathcal{F})\leq N$. \end{prop} \begin{proof} Denote $c:=m^0(X_\mathcal{F})$ and let $f$ be a $\Delta$-eigenfunction with eigenvalue $c$. First we claim that $c\neq N+1$. Indeed, $\Delta$ is a semi-simple operator and if $c=N+1$ then by Proposition \ref{prop_ny} it has only two distinct eigenvalues, namely $0$ and $N+1$. This implies that $\Delta^2=(N+1)\Delta$. In $X_\mathcal{F}$ we can find two vertices $x$ and $y$ which are not adjacent but such that there is another vertex $v$ which is adjacent to both $x$ and $y$. Let $g\in C^0(X_\mathcal{F})$ be a function such that $g(x)\neq 0$ but $g(x')=0$ if $x'\neq x$. Now $\Delta g(y)=0$ because this is a sum of the values of $g$ at $y$ and the vertices adjacent to $y$, and $x$ is not one of them. On the the other hand, $\Delta^2 g(y)\neq 0$ since this is a sum which involves $g(x)$ with a non-zero coefficient. This contradicts the equality $\Delta^2=(N+1)\Delta$. Define a function $h\in C^0(X_\mathcal{F})$ by $$h(v)=\sum_{\substack{x\in \mathrm{Ver}(X_\mathcal{F})\\ \mathrm{Type}(x)=\mathrm{Type}(v)}} f(x), \qquad \text{for any }v\in \mathrm{Ver}(X_\mathcal{F}). $$ It is clear that $h$ is type-constant, and because $f$ is a $\Delta$-eigenfunction, we have $\Delta h=ch$. Since $c\neq 0, N+1$, the function $h$ must be identically $0$. Therefore, $\sum_{\mathrm{Type}(v)=\beta}f(v)=0$ for any fixed $\beta\in \mathfrak T$. Obviously the same is also true for $f_\alpha$, i.e., $\sum_{\mathrm{Type}(v)=\beta}f_\alpha(v)=0$. Since $w(v)$ depends only on the type of $v$, we see that $f_\alpha$ is orthogonal to $\mathbf{1}$ in $C^0(X_\mathcal{F})$ with respect to the pairing \eqref{eq-pairing}. As in the proof of Lemma \ref{lemDec5}, this implies that $$ (\Delta f_\alpha, f_\alpha)\geq c\cdot (f_\alpha, f_\alpha). $$ Summing over all types, we get $$ \sum_{\alpha\in \mathfrak T} (\Delta f_\alpha, f_\alpha)\geq c(N+R^2)\cdot (f, f). $$ Comparing this inequality with the expression in Lemma \ref{lemd31}, we conclude that $(N-c)(R-1)^2\geq 0$. Since $R$ is arbitrary, we must have $c\leq N$. \end{proof} \subsection*{Acknowledgements} The author thanks Ori Parzanchevski and Farbod Shokrieh for useful comments on an earlier version of the paper. \def\polhk#1{\setbox0=\hbox{#1}{\ooalign{\hidewidth \lower1.5ex\hbox{`}\hidewidth\crcr\unhbox0}}} \providecommand{\bysame}{\leavevmode\hbox to3em{\hrulefill}\thinspace} \providecommand{\MR}{\relax\ifhmode\unskip\space\fi MR } \providecommand{\MRhref}[2]{% \href{http://www.ams.org/mathscinet-getitem?mr=#1}{#2} } \providecommand{\href}[2]{#2}
{ "timestamp": "2016-12-26T02:02:27", "yymm": "1612", "arxiv_id": "1612.07904", "language": "en", "url": "https://arxiv.org/abs/1612.07904", "abstract": "This is an expository paper on Garland's vanishing theorem specialized to the case when the linear algebraic group is $\\mathrm{SL}_n$. Garland's theorem can be stated as a vanishing of the cohomology groups of certain finite simplicial complexes. The method of the proof is quite interesting on its own. It relates the vanishing of cohomology to the assertion that the minimal positive eigenvalue of a certain combinatorial laplacian is sufficiently large. Since the 1970's, this idea has found applications in a variety of problems in representation theory, group theory, and combinatorics, so the paper might be of interest to a wide audience. The paper is intended for non-specialists and graduate students.", "subjects": "Combinatorics (math.CO); Representation Theory (math.RT)", "title": "On Garland's vanishing theorem for $\\mathrm{SL}_n$", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9755769085257165, "lm_q2_score": 0.839733963661418, "lm_q1q2_score": 0.8192250642528526 }
https://arxiv.org/abs/1903.10668
On the Weakly Prime-Additive Numbers with Length 4
In 1992, Erd$ő$s and Hegyv$á$ri showed that for any prime p, there exist infinitely many length 3 weakly prime-additive numbers divisible by p. In 2018, Fang and Chen showed that for any positive integer m, there exists infinitely many length 3 weakly prime-additive numbers divisible by m if and only if 8 does not divide m. Under the assumption (*) of existence of a prime in certain arithmetic progression with prescribed primitive root, which is true under the Generalized Riemann Hypothesis (GRH), we show for any positive integer m, there exists infinitely many length 4 weakly prime-additive numbers divisible by m. We also present another related result analogous to the length 3 case shown by Fang and Chen.
\section{Introduction} A number $n$ with at least 2 distinct prime divisors is called \textit{prime-additive} if $n=\sum_{p|n}p^{a_p}$ for some $a_p>0$. If additionally $p^{a_p}<n\leq p^{a_p+1}$ for all $p|n$, then $n$ is called \textit{strongly prime-additive}. In 1992, Erd\H{o}s and Hegyv\'{a}ri \cite{Erdos} stated a few examples and conjectured that there are infinitely many strongly prime-additive numbers. However, this problem was and is still far from being solved. For example, not even the infinitude of prime-additive numbers is known. Therefore they introduced the following weaker version of prime-additive numbers. \begin{defn} A positive integer $n$ is said to be \textit{weakly prime-additive} if $n$ has 2 distinct prime divisors, and there exists distinct prime divisors $p_1,...,p_t$ of $n$ and positive integers $a_1,...,a_t$ such that $n=p_1^{a_1}+\cdots+p_t^{a_t}$. The minimal value of such $t$ is defined to be the \textit{length} of $n$, denoted as $\kappa_n$. Note that if $n$ is a weakly prime-additive number, then $\kappa_n\geq3$. So we call a weakly prime-additive number with length 3 a \textit{shortest weakly prime-additive} number. \end{defn} Erd\H{o}s and Hegyv\'{a}ri \cite{Erdos} showed that for any prime $p$, there exist infinitely many weakly prime-additive numbers divisible by $p$. In fact, they showed that these weakly prime-additive numbers can be taken to be shortest weakly prime-additive in their proof. They also showed that the number of shortest weakly prime-additive numbers up to some integer $N$ is at least $c(\log N)^3$ for sufficiently small constant $c>0$. In 2018, Fang and Chen \cite{JHF} showed that for any positive integer $m$, there exists infinitely many shortest weakly prime-additive numbers divisible by $m$ if any only if $8$ does not divide $m$. This is Theorem \ref{shortest} stated in this paper. Also, they showed that for any positive integer $m$, there exists infinitely many weakly prime-additive numbers with length $\kappa_n\leq 5$ and divisible by $m$. In the same paper, Fang and Chen posted 4 open problems, the first one asking for any positive integer $m$, if there are infinitely many weakly prime-additive numbers $n$ with $m|n$ and $\kappa_n=4$. In Theorem 1 of this paper, we confirm this is true under the assumption $(\ast)$ of existence of a prime in certain arithmetic progression with prescribed primitive root, and such assumption is true under the Generalized Riemann Hypothesis (GRH). Finally, it was also shown in \cite{JHF} that for any distinct primes $p,q$, there exists a prime $r$ and infinitely many $a,b,c$ such that $pqr|p^a+q^b+r^c$. In Theorem 2, we showed an analogous result for 4 primes with a mild congruence conditions assuming $(\ast)$, the same assumption as above. \section{Main Results} \begin{assump} Let $1\leq a\leq f$ be positive integers with $(a,f)=1$ and $4|f$. Let $g$ be an odd prime dividing $f$ such that $\left(\frac{g}{a}\right)=-1$ with $\left(\frac{\cdot}{\cdot}\right)$ being the Kronecker symbol. Then there exists a prime $p$ such that $p\equiv a\Mod f$ and $g$ is a primitive root of $p$. \end{assump} It is known that ($\ast$) is a consequence of the Grand Riemann Hypothesis (GRH), see Corollary \ref{moree} in the next section for details. Under the assumption ($\ast$), we have the following. \begin{thm}\label{main} Assume ($\ast$). For any positive integer $m$, there exist infinitely many weakly prime-additive numbers $n$ with $m|n$ and $\kappa_n=4$. \end{thm} Note that if a positive integer $n=p^a+q^b+r^c+s^d$ for some distinct primes $p,q,r,s$, and positive integers $a,b,c,d$ such that $p,q,r,s|n$, then $p,q,r,s$ are all odd primes. We have the following partial converse and analog to Theorem 1.4 in \cite{JHF}. \begin{thm} Assume ($\ast$). For any distinct odd primes $p,q,r$ with one of them $\equiv 3$ or $5\Mod 8$, there exists infinitely many prime $s$, infinitely many positive integers $a,b,c,d$ such that $$pqrs|p^a+q^b+r^c+s^d.$$ \end{thm} \section{Preliminaries} \begin{lemma}\label{fermat} (The Fermat-Euler Theorem, Theorem 72, \cite{elementary}) Let $a,n$ be coprime positive integers, then $$a^{\phi(n)}\equiv1\Mod n$$ where $\phi$ is the Euler totient function. \end{lemma} We will need the following properties of the Kronecker Symbol $(\frac{\cdot}{\cdot})$, a generalization of the Legendre symbol. Whenever we write $(\frac{a}{b})$ for some integers $a,b$, it means the Kronecker symbol. \begin{lemma}\label{kronecker} For any nonzero integers $a,b,c$, and any odd primes $p,q$. Let $a'$, $b'$ be the odd part of $a$ and $b$ respectively, then we have:\\ \begin{align*} 1. & \left(\frac{ab}{c}\right)=\left(\frac{a}{c}\right)\left(\frac{b}{c}\right) \text{ unless c=-1}\\ 2. & \left(\frac{a}{b}\right)=(-1)^{\frac{a'-1}{2}\frac{b'-1}{2}}\left(\frac{b}{a}\right)\\ 3. & \left(\frac{-2}{p}\right)=\begin{cases}1&\text{ if } p\equiv 1,3\Mod 8\\ -1&\text{ if } p\equiv 5,7\Mod 8\end{cases}\\ 4. & \left(\frac{a}{p}\right)\equiv a^\frac{p-1}{2}\Mod p\\ 5. & \left(\frac{p}{q}\right)=\left(\frac{q}{p}\right) \text{ unless } p\equiv q\equiv 3\Mod 4.\\ & If p\equiv q\equiv 3\Mod 4, \left(\frac{p}{q}\right)=-\left(\frac{q}{p}\right) \end{align*} \end{lemma} \begin{proof} See \cite{kronecker}, p. 289-290. \end{proof} \begin{thm}\label{dirichlet} (Dirichlet's Theorem) Let $a,d$ be coprime positive integers, then there are infinitely many primes $p$ such that $p\equiv a\Mod d$. \end{thm} \begin{proof} See \cite{kronecker}, chapter 1. \end{proof} Under GRH, we have the following generalization. \begin{thm}\label{more} (Theorem 1.3, \cite{PM}) Let $1\leq a\leq f$ be positive integers with $(a,f)=1$. Let $g$ be an integer that is not equal to $-1$ or a square, and let $h\geq1$ be the largest integer such that $g$ is a $h$th power. Write $g=g_1g_2^2$ with $g_1$ square free, $g_1,g_2$ are integers. Let $$\beta=\frac{g_1}{(g_1,f)} \text{ and } \gamma_1=\begin{cases}(-1)^\frac{\beta-1}{2}(f,g_1) & \text{if }\beta \text{ is odd;}\\1 & \text{otherwise.}\end{cases}$$ Let $\pi_g(x;f,a)$ be the number of primes $p\leq x$ such that $p\equiv a\Mod{f}$ and $g$ is a primitive root $\Mod{p}$. Then, assuming GRH, we have $$\pi_g(x;f,a)=\delta(a,f,g)\frac{x}{\log x}+O_{f,g}\left(\frac{x\log\log x}{\log^2 x}\right)$$ where $$\delta(a,f,g)=\frac{A(a,f,h)}{\phi(f)}\left(1-\left(\frac{\gamma_1}{a}\right)\frac{\mu(|\beta|)}{\prod_{\substack{p|\beta\\ p|h}}(p-1)\prod_{\substack{p|\beta \\ p\not|h}}(p^2-p-1)}\right)$$ if $g_1\equiv1\Mod{4}$ or $(g_1\equiv 2\Mod{4}$ and $8|f)$ or $(g_1\equiv3\Mod{4}$ and $4|f$), and $$\delta(a,f,g)=\frac{A(a,f,h)}{\phi(f)}$$ otherwise. Here $\mu$ is the mobius function, $\left(\frac{\cdot}{\cdot}\right)$ is the Kronecker symbol, and $$A(a,f,h)=\prod_{p|(a-1,f)}\left(1-\frac{1}{p}\right)\prod_{\substack{p\not| f \\ p|h}}\left(1-\frac{1}{p-1}\right)\prod_{\substack{p\not| f\\ p\not| h}}\left(1-\frac{1}{p(p-1)}\right)$$ if $(a-1,f,h)=1$ and $A(a,f,h)=0$ otherwise. \end{thm} \begin{cor}\label{moree} Assume GRH. Let $a,f,g$ as above and $\left(\frac{g}{a}\right)=-1$. Then there exists a prime $p$ such that $p\equiv a\Mod f$ and $g$ is a primitive root of $p$, i.e. ($\ast$) is true under GRH. \end{cor} \begin{proof} This is a special case of Theorem \ref{more}, where with our conditions on $a,f,g$, $\beta=h=1$, $\gamma_1=g$. $$\delta(a,f,g)=\frac{2}{\phi(f)}\prod_{p|(a-1,f)}\left(1-\frac{1}{p}\right)\prod_{p\not| f}\left(1-\frac{1}{p(p-1)}\right)>0$$ \end{proof} \begin{remark} This shows that our result also follows from GRH, which is a much stronger assumption than ($\ast$). \end{remark} \begin{thm}\label{shortest} (Corollary 1.1, \cite{JHF}) Let $m$ be a positive integer, then there exists infinitely many shortest weakly prime-additive numbers $n$ with $m|n$ if and only if 8 does not divide $m$. \end{thm} \section{Proof of Theorem 1} We first prove the following weaker version of Theorem 1. \begin{thm}\label{main} Assume ($\ast$). For any positive integer $m$, there exist infinitely many weakly prime-additive numbers $n$ with $m|n$ and $\kappa_n\leq4$. \end{thm} \begin{proof} Let $m$ be a positive integer. Write $m=2^km_1$ with $(m_1,2)=1$ and $k\geq0\in\Z$. Without loss of generality, we assume $k\geq3$. We will construct a family of distinct primes $p,q,r,s$ and positive integers $a,b,c$ such that $m,p,q,r,s|n:=p^a+q^b+r^c+s$. Let $p$ be an odd prime such that $(p,m)=1$. By the Chinese Remainder Theorem and Theorem \ref{dirichlet}, there exists an odd prime $q$ such that $$q\equiv 1 \Mod{2^kp}\text{ and }q\equiv-1\Mod{m_1}$$ Again by the same two theorems, there exists an odd prime $r$ such that $$r\equiv 3\Mod{2^k}\text{ and }r\equiv1\Mod{pqm_1}$$ By the Chinese Remainder Theorem, let $s_0$ be the unique integer such that $1\leq s_0\leq pqrm$ and \begin{align*}\label{acong} &s_0\equiv -5\Mod{2^k}\\ &s_0\equiv -1\Mod{m_1}\\ &s_0\equiv -2\Mod{pqr} \end{align*} Then we can see that $(s_0,pqrm)=1$. Now note that as $k\geq3$, $r\equiv3\Mod{2^k}$, $s_0\equiv -5\mod{2^k}$ gives $r\equiv 3\Mod 8$ and $s_0\equiv 3\Mod 8$ respectively, so we have by Lemma \ref{kronecker}, \begin{align*} \left(\frac{r}{s_0}\right)=\left(\frac{s_0}{r}\right)(-1)^{\frac{s_0-1}{2}\frac{r-1}{2}}=-\left(\frac{s_0}{r}\right)=-\left(\frac{-2}{r}\right)=-1 \end{align*} where we used $s_0\equiv -2\Mod r$ as $s_0\equiv -2\Mod{pqr}$. Hence applying Corollary \ref{moree} with $a=s_0$, $f=pqrm$ and $g=r$, there exists an odd prime $s$ such that $s\equiv s_0\Mod{pqrm}$ and $r$ is a primitive root of $s$. So $s$ satisfies all of the above congruence relations satisfied by $s_0$, and there exists a positive integer $c_0$ s.t. $$r^{c_0}\equiv -2 \Mod s$$ Note by construction, $p,q,r,s$ are all distinct odd primes. Now for any positive integer $c'$, take $$c=(p-1)(q-1)(r-1)\phi(m)c'+c_0$$ and for any positive odd integer $b'$, take $$b=\frac{1}{4}(r-1)(s-1)b'$$ Note that since $r\equiv 3\Mod{2^k}$ and $s\equiv -5\Mod{2^k}$, we have $r,s\equiv 3\Mod 4$. So $b$ is odd. Finally, for any positive integer $a'$, take $$a=(q-1)(r-1)(s-1)\phi(m)a'$$ where $\phi$ is the euler $\phi$ function, and let $$n=p^a+q^b+r^c+s$$ Then we have the following congruence conditions:\\ 1. As $q\equiv r\equiv1\Mod p$, $s\equiv -2\Mod p$, we have $$n\equiv p^a+q^b+r^c+s\equiv0+1+1-2\equiv 0\Mod p$$ 2. As $q-1|a$, by Lemma \ref{fermat}, $p^a\equiv1\mod q$. So we have $$n\equiv p^a+q^b+r^c+s\equiv1+0+1-2\equiv0\Mod q$$ 3. Similarly, $p^a\equiv1\Mod r$ as $r-1|a$. Since $q\equiv1\Mod{2^kp}$, $q\equiv 1\Mod 8$. By Lemma \ref{kronecker}, with $r\equiv3\Mod 8$ and $r\equiv1\Mod q$, $$q^b\equiv (q^{\frac{1}{2}(r-1)})^{\frac{1}{2}(s-1)b'}\equiv\left(\frac{q}{r}\right)^{\frac{1}{2}(s-1)b'}\equiv\left(\frac{r}{q}\right)\equiv\left(\frac{1}{q}\right)\equiv1\Mod{r}$$ So we have $$n\equiv p^a+q^b+r^c+s\equiv1+1+0-2\equiv0\Mod r$$ 4. Similarly, $p^a\equiv q^b\equiv1\Mod s$. As $r^c\equiv-2\Mod s$, we have $$n\equiv p^a+q^b+r^c+s\equiv1+1-2+0\equiv 0\Mod s$$ 5. As $\phi(m)|a$, by Lemma \ref{fermat}, $p^a\equiv1\Mod m$. Since $b$ is odd and $q\equiv-1\Mod{m_1}$, we get $q^b\equiv -1\Mod{m_1}$. Together with $r\equiv1\Mod{m_1}$ and $s\equiv -1\Mod{m_1}$, we have $$n\equiv p^a+q^b+r^c+s\equiv1-1+1-1\equiv0\Mod{m_1}$$ 6. As $p^a\equiv1\Mod m$, $q\equiv 1\Mod{2^k}$, $r\equiv 3\Mod{2^k}$ and $s\equiv -5\Mod{2^k}$, we have $$n\equiv p^a+q^b+r^c+s\equiv1+1+3-5\equiv0\Mod{2^k}$$ Hence $n=p^a+q^b+r^c+s$ is weakly prime additive and is divisible by $m$. Since $a',c'$ can be any positive integer, $b'$ can be any positive odd integer and $p$ can be any arbitrary odd prime coprime to $m$, we have constructed infinitely many weakly prime-additive $n$ with length $\leq 4$. \end{proof} \begin{remark} In the above construction, $s$ can be raised to any $d$-th power for any positive integer $d\equiv 1\Mod{\phi(pqrm)}$. \end{remark} Together with Theorem \ref{shortest}, we can prove Theorem 1. \begin{proof}[Proof of Theorem 1] Let $m$ be a positive integer. Then by Theorem \ref{main}, there exists infinitely many weakly prime-additive numbers with length $\leq4$ such that they are divisible by $8m$. By Theorem \ref{shortest}, as $8|8m$, these numbers cannot be shortest weakly prime-additive, hence they are all weakly prime-additive numbers with length 4. \end{proof} \section{Proof of Theorem 2} Let $p,q,r$ be distinct odd primes with one of them, WLOG say, $r\equiv 3$ or $5\Mod 8$. Let $k\in\N$ such that $\frac{(p-1)(q-1)}{2^k}$ is odd, and let $A=\frac{(p-1)(q-1)}{2^k}$. Let $f=8Apqr$ and by the Chinese Remainder Theorem, let $s_0$ be the unique integer such that $1\leq s_0\leq f$ and \begin{align*} &s_0\equiv3\Mod{8}\\ &s_0\equiv -2\Mod{Apqr} \end{align*} Note that by Lemma \ref{kronecker}, using $r\equiv3$ or $5\Mod 8$, we have \begin{align*} \left(\frac{r}{s_0}\right)=(-1)^{\frac{r-1}{2}\frac{s_0-1}{2}}\left(\frac{s_0}{r}\right)=(-1)^\frac{r-1}{2}\left(\frac{-2}{r}\right)=-1 \end{align*} Apply Theorem 5 with the above $a, f$, and $g=r$, there exists an odd prime $s$ such that\\ $s\equiv s_0\Mod{8Apqr}$ and $r$ is a primitive root of $s$. So there exists $0<c_0<s-1$ such that $$r^{c_0}\equiv -2\Mod{s}.$$ Now note that as $s\equiv3\Mod{8}$, we have $\left(\frac{-2}{s}\right)=1$. So $c_0$ is even. Now by the Chinese Remainder Theorem, take any positive integer $c$ such that \begin{align*} &c\equiv c_0\Mod{\frac{s-1}{2}}\\ &c\equiv 0\Mod{(p-1)(q-1)} \end{align*} This is possible as $s\equiv3\Mod{8}$ and $s\equiv -2\Mod{A}$, so $(\frac{s-1}{2},(p-1)(q-1))=1$. Since $c_0$ is even and $\frac{s-1}{2}$ is odd, this makes $c\equiv c_0\Mod{s-1}$. So we have $r^c\equiv -2\Mod{s}$ and $r^c\equiv1\Mod{pq}$. Finally, for any positive integers $a,b,d$ such that $(q-1)(r-1)(s-1)|a$, $(p-1)(r-1)(s-1)|b$, $d\equiv1\Mod{(p-1)(q-1)(r-1)}$, we have the following:\begin{align*} p^a+q^b+r^c+s^d&\equiv0+1+1-2\equiv0\Mod p\\ p^a+q^b+r^c+s^d&\equiv1+0+1-2\equiv0\Mod q\\ p^a+q^b+r^c+s^d&\equiv1+1+0-2\equiv0\Mod r\\ p^a+q^b+r^c+s^d&\equiv1+1-2+0\equiv0\Mod s \end{align*} So we get for any positive integers $a,b,c,d$ as above, $$pqrs|p^a+q^b+r^c+s^d.$$ \qed
{ "timestamp": "2019-03-27T01:08:12", "yymm": "1903", "arxiv_id": "1903.10668", "language": "en", "url": "https://arxiv.org/abs/1903.10668", "abstract": "In 1992, Erd$ő$s and Hegyv$á$ri showed that for any prime p, there exist infinitely many length 3 weakly prime-additive numbers divisible by p. In 2018, Fang and Chen showed that for any positive integer m, there exists infinitely many length 3 weakly prime-additive numbers divisible by m if and only if 8 does not divide m. Under the assumption (*) of existence of a prime in certain arithmetic progression with prescribed primitive root, which is true under the Generalized Riemann Hypothesis (GRH), we show for any positive integer m, there exists infinitely many length 4 weakly prime-additive numbers divisible by m. We also present another related result analogous to the length 3 case shown by Fang and Chen.", "subjects": "Number Theory (math.NT)", "title": "On the Weakly Prime-Additive Numbers with Length 4", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9852713826904225, "lm_q2_score": 0.8311430520409023, "lm_q1q2_score": 0.8189014640978776 }
https://arxiv.org/abs/1906.02781
Tutte Polynomial Activities
Unlike Whitney's definition of the corank-nullity generating function $T(G;x+1,y+1)$, Tutte's definition of his now eponymous polynomial $T(G;x,y)$ requires a total order on the edges of which the polynomial is a posteriori independent. Tutte presented his definition in terms of internal and external activities of maximal spanning forests. Although Tutte's original definition may appear somewhat ad hoc upon first inspection, subsequent work by various researchers has demonstrated that activity is a deep combinatorial concept. In this survey, we provide an introduction to activities for graphs and matroids. Our primary goal is to survey several notions of activity for graphs which admit expansions of the Tutte polynomial. Additionally, we describe some fundamental structural theorems, and outline connections to the topological notion of shellability as well as several topics in algebraic combinatorics.
\chapter*{} \chapterauthor{Spencer Backman}{Einstein Institute for Mathematics\\ Edmond J. Safra Campus\\ The Hebrew University of Jerusalem\\ Givat Ram. Jerusalem, 9190401, Israel\\ \texttt{[email protected]}\\} \chapter{Tutte Polynomial Activities} \section{Synopsis} Activities are certain statistics associated to spanning forests and more general objects, which can be used for defining the Tutte polynomial. This chapter is intended to serve as an introduction to activities for graphs and matroids. We describe \index{activity} \begin{itemize} \item Tutte's original spanning forest activities \item Gordon and Traldi's subgraph activities \item Gessel and Sagan's depth-first search activities \item Bernardi's embedding activities \item Gordon and McMahon's generalized subgraph activities \item Las Vergnas' orientation activities \item Etienne and Las Vergnas' activity bipartition \item Crapo's activity interval decomposition \item Las Vergnas' active orders \item Shellability and the algebraic combinatorics of activities \end{itemize} \section{Introduction} \index{activity} \index{Tutte polynomial} \index{spanning forest} \index{matroid} \index{cut} \index{cycle} Unlike Whitney's definition of the corank-nullity generating function \index{corank-nullity generating function} $T(G;x+1,y+1)$, Tutte's definition of his now eponymous polynomial $T(G;x,y)$ \index{Tutte polynomial} requires a total order on the edges of which the polynomial is {\it a posteriori} independent. Tutte presented his definition in terms of \emph{internal and external activities} of maximal spanning forests. Although Tutte's original definition may appear somewhat ad hoc upon first inspection, subsequent work by various researchers has demonstrated that activity is a deep combinatorial concept. In this chapter, we provide an introduction to activities for graphs and matroids. Our primary goal is to survey several notions of activity for graphs which admit expansions of the Tutte polynomial. Additionally, we describe some fundamental structural theorems, and outline connections to the topological notion of shellability as well as several topics in algebraic combinatorics. We use the language of graphs except in sections \ref{active} and \ref{shell} where matroid terminology is employed, although sections \ref{spanningforest}, \ref{activebipartition}, \ref{activesubgraph}, and \ref{unified} apply equally well to matroids, and section \ref{orientation} applies to oriented matroids. \section{Activities for maximal spanning forests}\label{spanningforest} We recall Whitney's original definition of the the Tutte polynomial \cite{MR1562461}. The rank of a subset of edges $A$, written $r(A)$, is the maximum cardinality of a forest contained in $A$. \index{forest} \begin{definition}\label{EMM:def:ss} If $G=(V,E)$ is a graph, then the Tutte polynomial of $G$ is \begin{equation}\label{eq:rank_expansion} T(G; x, y) = \sum_{A\subseteq E} (x-1)^{r(E)-r(A)} (y-1)^{|A|-r(A)}. \end{equation} \end{definition} Let $G=(V,E)$ be a graph and $F$ be a maximal spanning forest of $G$. The maximality of $F$ means that if $G$ is connected then $F$ is a spanning tree, and if $G$ is not connected, restricting $F$ to any component of $G$ gives a spanning tree of that component. \index{activity} \index{Tutte polynomial} \index{spanning forest} \index{cut!fundamental} \index{cycle!fundamental} \begin{definition} Let $F$ be a maximal spanning forest of $G$, $f \in F$, and $e \in E \setminus F$. The \emph{fundamental cut associated with $F$ and $e$} \index{cut}\index{cycle} is \[ U_F(e) = \text{\{the edges of the unique cut in $(E \setminus F) \cup e$\}}. \] Similarly, the \emph{fundamental cycle associated with $F$ and $e$} is \[ Z_F(e) =\text{\{the edges of the unique cycle in $F\cup e$\}}. \] \end{definition} We describe Tutte's activities using fundamental cuts and cycles. \begin{definition} Let $G = (V,E)$ be a graph with a total order on $E$, and $F$ be a maximal spanning forest of $G$. We say that an edge $e \in E$ is \index{spanning forest} \begin{itemize} \item \emph{internally active} ($e \in \mathrm I(F)$) with respect to $F$ if $e\in F$ and it is the smallest edge in $U_F(e)$, \item \emph{externally active} ($e \in \mathrm E(F)$) with respect to $F$ if $e\not\in F$ and it is the smallest edge in $Z_F(e)$. \end{itemize} We note that all bridges are internally active and all loops are externally active. \end{definition} \begin{figure}[h] \begin{subfigure} \centering \includegraphics[scale=1]{activities2} \caption{(i) A graph with a total order on the edges and a spanning tree $T$ in red. (ii) Edge 1 is the only internally active edge and its fundamental cut is colored blue. (iii) Edge 2 is the only externally active edge and its fundamental cycle is colored green. } \end{subfigure} \end{figure} \index{activity} We can now give the \emph{spanning tree (maximal spanning forest) activities expansion} of the Tutte polynomial. \index{Tutte polynomial} \begin{definition}[Tutte \cite{MR0061366}]\label{EMM:d.intext} If $G=(V,E)$ is a graph with a fixed total order of $E$, then \begin{equation}\label{EMM:e2} T(G;x,y)=\sum\limits_{\substack{F }}x^{|\mathrm{I}(F)|} y^{|\mathrm{E}(F)|}, \end{equation} where the sum is over all maximal spanning forests of $G$. \end{definition} Tutte demonstrated that this polynomial is well defined, i.e., it is independent of the total order on the edges. \section{Activity bipartition}\label{activebipartition} \index{activity}\index{activity!bipartition} \index{Tutte polynomial} \index{spanning forest} \index{matroid}\index{flat}\index{convolution} Etienne and Las Vergnas showed that the activities of a maximal spanning forest induce a canonical bipartition of the edge set of a graph or more generally a matroid. First we recall the definition of a flat. \begin{definition} Let $\mathcal{F}\subset E$. If there exists no $e \in E \setminus \mathcal{F}$ such that $e$ is contained in a non loop cycle in $e \cup \mathcal{F}$, we say that $\mathcal{F}$ is a \emph{flat} of $G$. A flat is \emph{cyclic} if it is a union of cycles. \index{forest} \end{definition} \begin{theorem}(Etienne-Las Vergnas \cite{MR1489076}). Given a maximal spanning forest $F$ of $G$, there exists a unique cyclic flat $\mathcal{F} $ of G such that $\mathcal{F} \cap F$ is a maximal spanning forest of $G|_{\mathcal{F}}$ with no internal activity, and $\mathcal{F}^c \cap F$ is a maximal spanning forest of $G/\mathcal{F} $ with no external activity. \end{theorem} As an application of this decomposition, one may obtain a convolution formula for the Tutte polynomial which was independently discovered by Kook, Reiner, and Stanton via incidence algebra methods, and has since been substantially refined and generalized \cite{MR1676189, :ab, MR2718681, MR3678586, MR3887405}. \index{Tutte polynomial} \begin{theorem}[Etienne-Las Vergnas \cite{MR1489076}, Kook-Reiner-Stanton \cite{MR1699230}] Given a graph G, then $$\sum_{\mathcal{F}}T(G;x,y) = T(G/\mathcal{F};x,0)T(G|_{\mathcal{F}};0,y),$$ where the sum is over all cyclic flats of G. \end{theorem} \section{Activities for subgraphs}\label{activesubgraph} \index{subgraph}\index{activity!subgraph} Gordon and Traldi introduced a notion of activities for arbitrary subgraphs, and used this to provide a 4-variable expansion of the Tutte polynomial which naturally specializes to both Whitney's and Tutte's original expansions. \index{activity} \begin{definition} Let $G = (V,E)$ be a graph with a total order on $E$, and $S \subset E$, then we say that an edge $e$ is \begin{itemize} \item \emph{internally active present} with respect to $S$ $(e \in I(S)\cap S)$ if $e\in S$ and $e$ is the smallest edge in some cut in $S^c \cup e$, \item \emph{internally active absent} with respect to $S$ $(e \in I(S)\cap S^c)$ if $e\notin S$ and $e$ is the smallest edge in some cut in $S^c$, \item \emph{externally active present} with respect to $S$ $(e \in L(S)\cap S)$ if $e\in S$ and $e$ is the smallest edge in some cycle in $S$, \item \emph{externally active absent} with respect to $S$ $(e \in \mathrm L(S)\cap S^c)$ if $e\notin S$ and $e$ is the smallest edge in some cycle in $S \cup e$. \end{itemize} \end{definition} \begin{theorem}[Gordon-Traldi \cite{MR1080623}]\label{EMM:d.intext} \index{Tutte polynomial} If $G$ is a graph with a fixed total order of $E$, then \begin{equation}\label{EMM:e2} T(G;x+w,y +z)=\sum\limits_{\substack{S \subset E}} x^{|{I(S)\cap S}|}w^{| I(S)\cap S^c|} y^{| L(S)\cap S|}z^{|L(S)\cap S^c|}. \end{equation} \end{theorem} By setting $x= 1$ and $z=1$, we recover Whitney's definition, and by setting $w= 0$ and $y=0$, we recover Tutte's definition. While Gordon and Traldi's expansion is proven recursively via deletion-contraction, from which they obtain more general formulae, the 4-variable expansion is equivalent to an earlier theorem of Crapo. \index{forest} \begin{theorem}[Crapo \cite{MR0262095}] Let $P(E)$ be the Boolean lattice of subgraphs of $G$ ordered by containment. Given a spanning forest $F$, define an interval in this lattice $[ F \setminus I(F), F \cup E(F)]$. Then $$P(E) = \sqcup_{F} [ F \setminus I(F), F \cup E(F)]$$ where the disjoint union is over all maximal spanning forests. \end{theorem} \begin{figure}[h] \begin{subfigure} \centering \includegraphics[scale=.9]{activities3.pdf} \caption{A Crapo decomposition of the subgraphs of $K_3$. Each row collects the subgraphs corresponding to the spanning tree on the left.} \end{subfigure} \end{figure} In the following sections \ref{DFS}, \ref{combmaps}, and \ref{unified} we will describe other notions of activity for maximal spanning forests which employ input data different from a total order on the edges, but still allow for expansions of the Tutte polynomial and an analogue of Crapo's interval decomposition. \index{activity}\index{depth-first search}\index{Crapo interval}\index{spanning tree}\index{activity!depth-first search} \section{Depth-first search external activity}\label{DFS} Gessel and Sagan introduced a notion of external activity for maximal spanning forests based on depth-first search. For simplicity sake, we will assume that our graph is connected and has no parallel edges. In what follows, we view a tree rooted at a vertex $q$ to be oriented so that every vertex is reachable from $q$ by a directed path. \index{activity} \begin{definition}[Gessel-Sagan \cite{MR1392494}] Let $<$ be a total order on the vertices of $G$, and $F$ be a spanning tree of $G$ rooted at the smallest vertex $q$. Let $e = (u,v)$ be an edge of $G \setminus F$. We say that $e$ is {\emph{depth-first search externally active (DFS externally active)}}, and write $e\in E_{DFS}(F)$, if either $u=v$, or $(u,w)$ is an oriented edge in $F$ belonging to the unique cycle in $F \cup (u,v)$, and $w>v$. \end{definition} \begin{figure}[h] \begin{subfigure} \centering \includegraphics[scale=.8]{activities4.pdf} \caption{$K_4$ with a root $q$ and a total order on its vertices. A spanning tree in red and a DFS externally active edge in blue.} \end{subfigure} \end{figure} The name DFS externally active is justified by the following observation: given a spanning subgraph of $G$, we can produce a spanning forest $F$ by performing a DFS search which favors larger labeled vertices. Then $(u,v)$ is DFS externally active if when we apply DFS search to the graph $F \cup (u,v)$, we obtain $F$. Gessel and Sagan showed that DFS external activity when combined with Tutte's notion of internal activity allows for an expansion of the Tutte polynomial.\index{spanning forest}\index{Tutte polynomial} \begin{theorem}[Gessel-Sagan \cite{MR1392494}]\index{Tutte polynomial} \index{forest} If $G$ is a connected graph with a total order of its vertices, then \begin{equation}\label{GS} T(G;x,y)=\sum\limits_{\substack{T}} x^{|{I(T)}|}y^{|{E_{DFS}(T)}|}, \end{equation} where the sum is over all spanning trees $T$ of $G$. \end{theorem} In the same article, Gessel and Sagan produced a notion of Nearest Neighbor First activity, which will not be reviewed here. \index{activity!depth-first search}\index{depth-first search}\index{spanning tree}\index{map}\index{embedding}\index{half-edge}\index{activity!embedding}\index{spanning tree} \section{Activities via combinatorial maps}\label{combmaps} \index{activity} Bernardi proposed a notion of activity induced by a rooted combinatorial map, which is essentially an embedding of a graph $G=(V,E)$ with a distinguished half-edge $h$ into an orientable surface. In what follows, we assume that $G$ is connected and loopless, although these restrictions are not essential. Informally, given a spanning tree $T$ of $G$, we can use a rooted combinatorial map to tour the edges of $G$ by starting at $h$ and traveling counterclockwise around the outside of $T$. We then declare an edge $e \notin T$ to be externally active if it is the first edge in its associated fundamental cycle which we meet during the tour. We similarly say an edge $e \in T$ is internally active if it is the first edge in its associated fundamental cut which we meet during the tour. We now describe Cori's maps \cite{MR0404045} and Bernardi's activities more formally. We define a \emph{half-edge} to be an edge and an incident vertex, e.g. if $e \in E$ with $e = (u,v)$ such that $u,v \in V$, then $(e,u)$ and $(e,v)$ are the two associated half-edges. Let $\sigma$ be a permutation of the half-edges of $G$ such that $\sigma((e,u))=(e',u)$ some other half-edge incident to $u$, and for any two half-edges $(e,u)$ and $(e',u)$ with the same endpoint, there exists some $k >0$ such that $\sigma^k((e,u))=(e',u)$. Let $h$ be a distinguished half-edge. We define a \emph{rooted combinatorial map} to be a triple $(G,\sigma,h)$. Let $\alpha$ be the involution on the set of half-edges such that for all $e = (u,v) \in E$, $\alpha((e,u)) = (e,v)$. Given a half-edge $i$ and a spanning tree $T$, we define the \emph{motion operator} \[ t(i)= \begin{cases} \sigma(i) &\text{if } i \notin T\\ \sigma \circ \alpha(i) & \text{if } i \in T \end{cases} \] It is easy to check that the iterated motion operator defines a tour of the half-edges of $G$. This tour induces a total order $<_{T}$ on the edges given by the first time one of its half-edges is visited. We use this total order to define internally and externally active edges. \index{forest} \begin{figure}[!ht] \begin{subfigure \centering \includegraphics[scale=.35]{bernardi_tour1.pdf} \end{subfigure} \begin{subfigure \centering \includegraphics[scale=.35]{bernardi_tour2.pdf} \caption{Tours of two different spanning trees induced by the motion operator associated to the same rooted combinatorial map.} \label{bernardi} \end{subfigure} \end{figure} \index{activity} \begin{definition}[Bernardi \cite{MR2428901}] Let $(G,\sigma, h)$ be a rooted combinatorial map, $T$ a spanning tree of $G$, and $e$ an edge of $G$. \begin{itemize} \item The edge $e$ is \emph{embedding-internally active} $(e \in I_B(T))$ if $e \in T$ and $e$ is the minimum edge with respect to $<_{T}$ in its associated fundamental cut. \item The edge $e$ is \emph{embedding-externally active} $(e \in E_B(T))$ if $e \notin T$ and $e$ is the minimum edge with respect to $<_{T}$ in its associated fundamental cycle. \end{itemize} \end{definition} Bernardi showed that this definition of activity admits an expansion of the Tutte polynomial. \begin{theorem}[Bernardi \cite{MR2428901}]\index{Tutte polynomial} Let $(G,\sigma,h)$ be a rooted combinatorial map, then \begin{equation}\label{GS} T(G;x,y)=\sum\limits_{\substack{T}} x^{|{I_B(T)}|}y^{|{E_B(T)}|}, \end{equation} where the sum is over all spanning trees. \end{theorem} We remark that Courtiel recently introduced a different notion of activity via combinatorial maps which he calls the ``blossom activity" \cite{courtiel:hal-01088871}. \index{Tutte polynomial} \section{Unified activities for subgraphs via decision trees}\label{unified} The Gordon-Traldi activities were further generalized by Gordon-McMahon \cite{MR1425948} in a way which also applies to greedoids. The Gordon-McMahon notion of activity was rediscovered by Courtiel \cite{courtiel:hal-01088871} who proved that it allows for a unification of all of the aforementioned notions of activity. We describe these activities following Courtiel and using the language of decision trees. \index{activity}\index{activity!blossom}\index{activity!unified}\index{decision tree}\index{spanning tree}\index{binary tree} \begin{definition} \index{activity} A \emph{decision tree} $D$ for $G$ consists of a perfect binary tree, i.e., a rooted tree such that all non-leaf nodes have two descendants, and a labeling of each node of the tree by elements of $E$ such that the labels along any particular branch give a permutation of $E$. \end{definition} \begin{figure}[h] \begin{subfigure} \centering \includegraphics[scale=.9]{activities6.pdf} \caption{An example of a decision tree for a graph on 4 edges.} \end{subfigure} \end{figure} Given a decision tree $D$ and a subgraph $S \subset E$, we can use $D$ to partition $S$ into four sets: $I(S)$, $L(S)$, $S_E$, $S_I$. We describe the recursive algorithm for producing this partition informally and refer the reader to \cite{courtiel:hal-01088871} for a pseudocode presentation. \begin{algorithm}{\text Recursive generalized activities algorithm} \ Initialize with $X = E, I(S) = L(S) = S_I = S_E = \emptyset$, and $e$ corresponding to the label of the root of $D$. While $X \neq \emptyset$, do the following: \begin{itemize} \item If $e$ is a bridge, add $e$ to $I(S)$, contract $e$ in $X$, and move to the right descendant of $e$ in $D$. \item If $e \in S $ is neither a bridge nor a loop, add $e$ to $S_I$, contract $e$ in $X$, and move to the right descendant of $e$ in $D$. \item If $e$ is a loop, add $e$ to $L(S)$, delete $e$ in $X$, and move to the left descendant of $e$ in $D$. \item If $e \notin S$ is neither a bridge nor a loop, add $e$ to $S_E$, delete $e$ in $X$, and move to the left descendant of $e$ in $D$. After we move to a descendant, we update $e$ to be the label of the new node and recurse. \end{itemize} \end{algorithm} \index{activity} \begin{theorem}[Gordon-McMahon \cite{MR1425948}]\index{Tutte polynomial} Let $G$ be a graph, $D$ a decision tree for $G$, and $I(S)$ and $L(S)$ as above, then the Tutte polynomial has the following expansion $$T(G;x+w, y+z) = \sum_{S \subset E} x^{|S \cap I(S)|}w^{|S^c \cap I(S)|}y^{|S^c \cap L(S)|}z^{|S \cap L(S)|}.$$ \end{theorem} \index{activity!orientation}\index{algorithm}\index{algorithm!recursive}\index{Tutte polynomial}\index{orientation}\index{cycle!directed}\index{cut!directed} \section{Orientation activities}\label{orientation} \index{activity} A famous result of Stanley states that $T(G;2,0)$ (equivalently, the chromatic polynomial evaluated at $-1$) counts the number of acyclic orientations of $G$ \cite{MR0317988}. This result was generalized to hyperplane arrangements by Zaslavsky \cite{MR0357135}, and to oriented matroids by Las Vergnas \cite{MR586435}. Las Vergnas \cite{MR776814} introduced a notion of orientation activities, which parallels those of subsets, and allows for an orientation expansion of the Tutte polynomial which recovers Stanley's result. He later introduced refined orientation activities \cite{Vergnas:aa}, which we now describe. Similar to the way that Tutte's activities are defined in terms of fundamental cuts and cycles, Las Vergnas' orientation activities are defined in terms of directed cuts and cycles. \begin{definition} Let $\mathcal{O}$ be an orientation of the edges of $G$. Let $Z$ be a cycle in $G$, and $U = (X,X^c)$ be a cut in $G$. We say that $Z$ is a \emph{directed cycle} if we can walk around the cycle traveling in the direction of the edge orientations. We similarly define $U$ to be a \emph{directed cut} if all of its edges are, without loss of generality, oriented from $X$ to $X^c$. \end{definition} We now use directed cuts and cycles to introduce notions of orientations activities. \begin{definition}[Las Vergnas \cite{Vergnas:aa}]\label{oract} \index{activity}\index{orientation}\index{activity!orientation}\index{orientation!reference} Let $G = (V,E)$ be a graph with a total order on $E$, and $\mathcal{O}_{ref}$ a reference orientation of the edges of $G$. If $\mathcal{O}$ is an orientation of $G$ and $e \in E$, then we say $e$ is \begin{itemize} \item \emph{positive cut active} ($e \in I(\mathcal{O})^+$) if $e$ is the smallest edge in some directed cut and is oriented in agreement with $\mathcal{O}_{ref}$, \item \emph{negative cut active} ($e \in I(\mathcal{O})^-$) if $e$ is the smallest edge in some directed cut and is oriented in disagreement with $\mathcal{O}_{ref}$, \item \emph{positive cycle active} ($e \in E(\mathcal{O})^+$) if $e$ is the smallest edge in some directed cycle and is oriented in agreement with $\mathcal{O}_{ref}$, \item \emph{negative cycle active} ($e \in E(\mathcal{O})^-$) if $e$ is the smallest edge in some directed cycle and is oriented in disagreement with $\mathcal{O}_{ref}$. \end{itemize} \end{definition} \begin{figure}[h] \begin{subfigure} \centering \includegraphics[scale=1.2]{activities7} \caption{Left: A total order and reference orientation of the edges of a graph $G$. Right: An orientation $\mathcal{O}$ of $G$ with $I(\mathcal{O})^+ = \emptyset , I(\mathcal{O})^- = \{ 1\}, L(\mathcal{O})^+ =\{2\}, {\rm and \, } L(\mathcal{O})^- = \emptyset$.} \end{subfigure} \end{figure} \begin{theorem}[Las Vergnas \cite{Vergnas:aa}]\label{EMM:d.intext}\index{Tutte polynomial} Let $G$ be a graph with a fixed total order and reference orientation of $E$, and $I(\mathcal{O})^+, I(\mathcal{O})^-, L(\mathcal{O})^+,$ and $L(\mathcal{O})^-$ as in Definition \ref{oract}, then \begin{equation}\label{EMM:e2} T(G;x +w,y +z)=\sum\limits_{\substack{\mathcal{O} }} x^{|I(\mathcal{O})^+|}w^{|I(\mathcal{O})^-|} y^{|E(\mathcal{O})^+|}z^{|E(\mathcal{O})^-|}, \end{equation} where the sum is over all orientations of $E$. \end{theorem} Las Vergnas orientation expansion holds for all oriented matroids. By specializing variables $x=w = u/2$ and $y = z = v/2$, we recover Las Vergnas earlier expansion which does not make use of a reference orientation. \begin{corollary}[Las Vergnas \cite{MR776814}] Let $G$ be a graph with a fixed total order $E$, and let $I(\mathcal{O})$ and $L(\mathcal{O})$ to be the set of edges which are minimum is some directed cut or cycle, respectively, then \begin{equation}\label{EMM:e2} T(G;x,y)=\sum\limits_{\substack{\mathcal{O}}} \left({u \over 2}\right)^{|I(\mathcal{O})|}\left({v\over 2}\right)^{|L(\mathcal{O})|}, \end{equation} where the sum is over all orientations of $E$. \end{corollary} \index{activity}\index{orientation}\index{activity!orientation}\index{active bijection}\index{fourientation}\index{subgraph}\index{Tutte polynomial}\index{active orders}\index{pivoting}\index{matroid}\index{activity!fourientation} We remark that Berman \cite{MR0469810} was the first to propose an orientation expansion of the Tutte polynomial, although his definition was not correct. There are natural notions of orientation activity classes which parallel Crapo's subset intervals. The refined orientation expansion of the Tutte polynomial also follows as a direct consequence of ``the active bijection" of Gioan and Las Vergnas which gives a bijection between orientation activity classes and Crapo subset intervals, which respects the four different activities \cite{MR2552669, MR3895683}. See Gioan's chapter for further details. \index{activity} A \emph{fourientation} of a graph is a choice for each edge of the graph whether to orient that edge in either direction, leave it unoriented, or biorient it. One may naturally view fourientations as a mixture of orientations and subgraphs where absent and present edges correspond to unoriented and bioriented edges, respectively. Backman, Hopkins, and Traldi \cite{MR3707217} introduced notions of activities for fourientations which provide a common refinement of the Gordon and Traldi subgraph activities and Las Vergnas' orientation activities. \section{Active orders}\label{active} \index{activity} For deepening our understanding of activities, Las Vergnas introduced three \emph{active orders} on the bases of a matroid. We describe these partial orders here and Las Vergnas' key result that they induce lattice structures on bases. Let $B_1$ and $B_2$ be bases of a matroid $M$ with fixed order on its ground set. We say that $B_1$ is obtained from $B_2$ by an \emph{externally active pivoting} if $B_2 = B_1 \setminus e \cup f$, where $e$ is the minimum element in $Z_{B_1}(f)$, and we write $B_1 \leftarrow_M B_2$. Dually, we say that $B_1$ is obtained from $B_2$ by an \emph{internally active pivoting} if $B_1 = B_2 \setminus e \cup f$, where $e$ is the smallest element in $U_{B_1}(f)$, and we write $B_1 \leftarrow *_M B_2$. We let $<_{Ext}$ and $<_{Int}$ denote the partial orders on the bases obtained by taking the transitive closures of the relations $ \leftarrow_M$ and $ \leftarrow *_M$, respectively. We refer to $<_{Ext}$ as the \emph{external order}, and $<_{Int}$ as the \emph{internal order}. Las Vergnas also defined the following join of the external and internal orders. Let $B_1$ and $B_2$ be bases, then we say that $B_1 <_{Ext/Int} B_2$ if there exists bases $C_1, \dots , C_k$ such that $B_1 = C_1$, $B_2 = C_k$, and for each $i$ either $C_i \leftarrow_M C_{i+1}$ or $C_i \leftarrow *_M C_{i+1}$. Recall that a lattice is a poset such that every pair of elements have a unique join and meet. \begin{theorem}[Las Vergnas \cite{MR1845495}] Let $\mathcal{B} (M)$ be the set of bases of a matroid $M$ with a total order on its ground set, then the posets \mbox{$(\mathcal{B} (M) \cup \{\mathbf{0}\}, <_{Ext})$}, \mbox{$(\mathcal{B} (M) \cup \{\mathbf{1}\}, <_{Int})$}, and $(\mathcal{B} (M), <_{Ext/Int})$ are lattices. \end{theorem} \index{active orders}\index{pivoting}\index{matroid}\index{bases}\index{lattice}\index{lattice!atomistic}\index{lattice!supersolvable}\index{lattice!distributive}\index{shellability}\index{simplicial complex}\index{facet}\index{face}\index{complex!independence}\index{complex!no broken circuit}\index{$f$-polynomial}\index{$h$-polynomial}\index{$f$-vector}\index{$h$-vector}\index{Tutte polynomial} Las Vergnas observed that the lattice associated to the external order is not distributive, although it is atomistic. This appears to have been remedied in the recent PhD thesis of Gillespie \cite{:aa} where it is shown that by extending the external order to all independent sets, a supersolvable join-distributive lattice is obtained. \section{Shellability and activity}\label{shell} \index{activity} There are important connections between activity and combinatorial topology. We refer the reader to Bj\"orner \cite{MR1165544} for an excellent introduction to this topic. Let $[n]$ denote the finite set $\{1, \dots , n\}$. An \emph{abstract simplicial complex} (often just called a simplicial complex) $\Delta$ on $n$ elements is a collection of subsets (faces) of $[n]$ which is closed under taking subsets. Informally, a simplicial complex $\Delta$ is \emph{shellable} if there is an ordering of the maximal faces $(\emph{facets})$ of $\Delta$ so that each facet can be added to the previous ones by glueing along codimension 1 faces. The \emph{f-polynomial} of a simplicial complex $\Delta$ is $f_{\Delta}(x) = \sum_{i=0}^n f_ix^{d-i}$ where $f_i$ is the number of faces of $\Delta$ of size $i$. The \emph{h-polynomial} of $\Delta$ is the $h_{\Delta}(x) = f_{\Delta}(x-1)$. The $f$-vector and $h$-vector of $\Delta$ are the vectors whose entries are the coefficients of the $f$-polynomial and $h$-polynomial, respectively. A shellable complex is homotopy equivalent to a wedge of spheres, and its $h$-vector is nonnegative. The following complexes are shellable, and the proofs of shellability are related to activities. \begin{itemize} \item The \emph{independence complex} $(IN(M))$ whose faces are the independent sets of a matroid $M$ \cite{MR2627467, MR593648}. Its $h$-polynomial is $T(M;x,1)$. Matroids are characterized by the fact that the $IN(M)$ are the pure simplicial complexes which are lexicographically shellable with respect to any order on the ground set \cite{MR1165544}. \item The \emph{no broken circuit complex} $(NBC(M))$ whose faces are the independent sets of a matroid with no external activity in the sense of Tutte \cite{MR2627467, MR593648}. Its $h$-polynomial is $T(M;x,0)$. \item The \emph{external activity complex} defined on $E(M)\times E(M)$, whose facets are given by $B \cup L(B) \times B\cup (B \cup L(B))^c$, where $B$ ranges over the bases of $M$ \cite{MR3558045}. The shelling makes use of Las Vergnas' order $<_{Ext/Int}$. It has the same $h$-polynomial as $IN(M)$ \cite{MR3558045}. \item The \emph{order complex of IN(M)} modulo Las Vergnas' external active order $<_{Ext}$ \cite{:aa}. \item The \emph{order complex of the lattice of flats} \cite{MR570784}. It has Euler characteristic $T(M;1,0)$. \end{itemize} \index{Tutte polynomial} These complexes admit many interesting connections with algebraic and geometric combinatorics. We briefly mention a few. Orlik and Solomon \cite{MR558866} introduced a certain graded algebra which is isomorphic to the cohomology ring of the complement of a complex hyperplane arrangement, and showed that monomials corresponding to faces of $NBC(M)$ give a basis for this algebra; see Falk and Kung's chapter for an introduction to these algebras. \index{activity}\index{shellability}\index{simplicial complex}\index{facet}\index{face}\index{complex!independence}\index{complex!no broken circuit}\index{$f$-polynomial}\index{$h$-polynomial}\index{complex!order}\index{algebra!Orlik-Solomon}\index{tropical geometry}\index{Bergman fan}\index{log-concavity}\index{unimodular}\index{$h$-vector}\index{$f$-vector}\index{$O$-sequence}\index{flat}\index{lattice}\index{monomial} The external activity complex was introduced by Ardila and Boocher in their investigation of commutative algebraic aspects of the closure of a linear space in a product of projective lines \cite {MR3439307}. The ideals they consider are homogenizations of ones considered earlier by Proudfoot and Speyer \cite{MR2246531} and Terao\cite{MR1899865}. A slight variation of these objects play an important role in Huh and Wang's proof of the Dowling-Wilson conjecture for realizable matroids \cite{MR3733101}. In the field of tropical geometry, which is often referred to a as a piecewise linear version of algebraic geometry, Bergman fans are certain balanced polyhedral fans which provide a new and exciting framework for studying matroids. Ardila and Klivans \cite{MR2185977} showed that they are unimodularily triangulated by the fan over the order complex of the lattice of flats. This relationship has lead authors to uncover interesting connections between algebraic geometry and matroids, most notably the proof of the Heron-Rota-Welsh conjecture that the $f$-vector of $NBC(M)$ is log-concave by Adirprasito, Huh, and Katz \cite{MR3862944} building on earlier works of Huh \cite{MR2904577}, and Huh and Katz \cite{MR2983081}. Recently, Fink, Speyer, and Woo introduced an \emph{extended NBC complex} in order to shed some light on these different manifestations of the $f$-polynomial of $NBC(M)$ \cite{fink}. Finally, a major open question in combinatorial commutative algebra is Stanley's 1977 conjecture \cite{MR0572989} that the $h$-vector of $IN(M)$ is a pure $O$-sequence. Merino settled this conjecture in the case of cographic matroids by application of chip-firing \cite{MR1888777}; see Merino's chapter for further details. \section{Acknowledgements} Many thanks to the anonymous referee for providing very helpful feedback on an earlier draft of this chapter, and to Sam Hopkins for pointing out several typographical errors. Additional thanks to Matt Baker for generously sharing Figure \ref{bernardi}. \bibliographystyle{plain}
{ "timestamp": "2019-06-11T02:27:43", "yymm": "1906", "arxiv_id": "1906.02781", "language": "en", "url": "https://arxiv.org/abs/1906.02781", "abstract": "Unlike Whitney's definition of the corank-nullity generating function $T(G;x+1,y+1)$, Tutte's definition of his now eponymous polynomial $T(G;x,y)$ requires a total order on the edges of which the polynomial is a posteriori independent. Tutte presented his definition in terms of internal and external activities of maximal spanning forests. Although Tutte's original definition may appear somewhat ad hoc upon first inspection, subsequent work by various researchers has demonstrated that activity is a deep combinatorial concept. In this survey, we provide an introduction to activities for graphs and matroids. Our primary goal is to survey several notions of activity for graphs which admit expansions of the Tutte polynomial. Additionally, we describe some fundamental structural theorems, and outline connections to the topological notion of shellability as well as several topics in algebraic combinatorics.", "subjects": "Combinatorics (math.CO)", "title": "Tutte Polynomial Activities", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.985271387015241, "lm_q2_score": 0.8311430457670241, "lm_q1q2_score": 0.8189014615109478 }
https://arxiv.org/abs/0706.4112
Induced Ramsey-type theorems
We present a unified approach to proving Ramsey-type theorems for graphs with a forbidden induced subgraph which can be used to extend and improve the earlier results of Rodl, Erdos-Hajnal, Promel-Rodl, Nikiforov, Chung-Graham, and Luczak-Rodl. The proofs are based on a simple lemma (generalizing one by Graham, Rodl, and Rucinski) that can be used as a replacement for Szemeredi's regularity lemma, thereby giving much better bounds. The same approach can be also used to show that pseudo-random graphs have strong induced Ramsey properties. This leads to explicit constructions for upper bounds on various induced Ramsey numbers.
\section{Background and Introduction} Ramsey theory refers to a large body of deep results in mathematics concerning partitions of large structures. Its underlying philosophy is captured succinctly by the statement that ``In a large system, complete disorder is impossible.'' This is an area in which a great variety of techniques from many branches of mathematics are used and whose results are important not only to graph theory and combinatorics but also to logic, analysis, number theory, and geometry. Since the publication of the seminal paper of Ramsey \cite{Ra} in 1930, this subject has grown with increasing vitality, and is currently among the most active areas in combinatorics. For a graph $H$, the {\it Ramsey number} $r(H)$ is the least positive integer $n$ such that every two-coloring of the edges of the complete graph $K_n$ on $n$ vertices contains a monochromatic copy of $H$. Ramsey's theorem states that $r(H)$ exists for every graph $H$. A classical result of Erd\H{o}s and Szekeres~\cite{ErSz}, which is a quantitative version of Ramsey's theorem, implies that $r(K_k) \leq 2^{2k}$ for every positive integer $k$. Erd\H{o}s~\cite{Er} showed using probabilistic arguments that $r(K_k) > 2^{k/2}$ for $k > 2$. Over the last sixty years, there has been several improvements on the lower and upper bounds of $r(K_k)$, the most recent by Conlon \cite{Co}. However, despite efforts by various researchers, the constant factors in the exponents of these bounds remain the same. A subset of vertices of a graph is {\it homogeneous} if it is either an independent set (empty subgraph) or a clique (complete subgraph). For a graph $G$, denote by $\hom(G)$ the size of the largest homogeneous subset of vertices of $G$. A restatement of the Erd\H{o}s-Szekeres result is that every graph $G$ on $n$ vertices satisfies $\hom(G) \geq \frac{1}{2}\log n$, while the Erd\H{o}s result says that for each $n\geq 2$ there is a graph $G$ on $n$ vertices with $\hom(G) \leq 2\log n$. (Here, and throughout the paper, all logarithms are base $2$.) The only known proofs of the existence of {\it Ramsey graphs}, i.e., graphs for which $\hom(G)=O(\log n)$, come from various models of random graphs with edge density bounded away from $0$ and $1$. This supports the belief that any graph with small $\hom(G)$ looks `random' in one sense or another. There are now several results which show that Ramsey graphs have random-like properties. A graph $H$ is an {\it induced subgraph} of a graph $G$ if $V(H) \subset V(G)$ and two vertices of $H$ are adjacent if and only if they are adjacent in $G$. A graph is {\it $k$-universal} if it contains all graphs on at most $k$ vertices as induced subgraphs. A basic property of large random graphs is that they almost surely are $k$-universal. There is a general belief that graphs which are not $k$-universal are highly structured. In particular, they should contain a homogeneous subset which is much larger than that guaranteed by the Erd\H{o}s-Szekeres bound for general graphs. In the early 1970's, an important generalization of Ramsey's theorem, known as the Induced Ramsey Theorem, was discovered independently by Deuber \cite{De}, Erd\H{o}s, Hajnal, and Posa \cite{ErHaPo}, and R\"odl \cite{Ro1}. It states that for every graph $H$ there is a graph $G$ such that in every $2$-edge-coloring of $G$ there is an induced copy of $H$ whose edges are monochromatic. The least positive integer $n$ for which there is an $n$-vertex graph with this property is called the {\it induced Ramsey number} $r_{\textrm{ind}}(H)$. All of the early proofs of the Induced Ramsey Theorem give enormous upper bounds on $r_{\textrm{ind}}(H)$. It is still a major open problem to prove good bounds on induced Ramsey numbers. Ideally, we would like to understand conditions for a graph $G$ to have the property that in every two-coloring of the edges of $G$, there is an induced copy of graph $H$ that is monochromatic. In this paper, we present a unified approach to proving Ramsey-type theorems for graphs with a forbidden induced subgraph which can be used to extend and improve results of various researchers. The same approach is also used to prove new bounds on induced Ramsey numbers. In the few subsequent sections we present in full detail our theorems and compare them with previously obtained results. \subsection{Ramsey properties of $H$-free graphs}\label{hfree} As we already mentioned, there are several results (see, e.g., \cite{ErSzem,She,AlKrSu,BuSu}) which indicate that Ramsey graphs, graphs $G$ with relatively small $\hom(G)$, have random-like properties. The first advance in this area was made by Erd\H{o}s and Szemer\'edi \cite{ErSzem}, who showed that the Erd\H{o}s-Szekeres bound $\hom(G) \geq \frac{1}{2}\log n$ can be improved for graphs which are very sparse or very dense. The edge density of a graph $G$ is the fraction of pairs of distinct vertices of $G$ that are edges. The Erd\H{o}s-Szemer\'edi theorem states that there is an absolute positive constant $c$ such that $\hom(G) \geq \frac{c \log n}{\epsilon \log \frac{1}{\epsilon}}$ for every graph $G$ on $n$ vertices with edge density $\epsilon \in (0,1/2)$. This result shows that the Erd\H{o}s-Szekeres bound can be significantly improved for graphs that contain a large subset of vertices that is very sparse or very dense. R\"odl \cite{Ro} proved that if a graph is not $k$-universal with $k$ fixed, then it contains a linear-sized induced subgraph that is very sparse or very dense. A graph is called {\it $H$-free} if it does not contain $H$ as an induced subgraph. More precisely, R\"odl's theorem says that for each graph $H$ and $\epsilon \in (0,1/2)$, there is a positive constant $\delta(\epsilon,H)$ such that every $H$-free graph on $n$ vertices contains an induced subgraph on at least $\delta(\epsilon,H)n$ vertices with edge density either at most $\epsilon$ or at least $1-\epsilon$. Together with the theorem of Erd\H{o}s and Szemeredi, it shows that the Erd\H{o}s-Szekeres bound can be improved by any constant factor for any family of graphs that have a forbidden induced subgraph. R\"odl's proof uses Szemer\'edi's regularity lemma \cite{Sz}, a powerful tool in graph theory, which was introduced by Szemer\'edi in his celebrated proof of the Erd\H{o}s-Tur\'an conjecture on long arithmetic progressions in dense subsets of the integers. The regularity lemma roughly says that every large graph can be partitioned into a small number of parts such that the bipartite subgraph between almost every pair of parts is random-like. To properly state the regularity lemma requires some terminology. The edge density $d(X,Y)$ between two subsets of vertices of a graph $G$ is the fraction of pairs $(x,y) \in X \times Y$ that are edges of $G$, i.e., $d(X,Y)=\frac{e(X,Y)}{|X||Y|}$, where $e(X,Y)$ is the number of edges with one endpoint in $X$ and the other in $Y$. A pair $(X,Y)$ of vertex sets is called $\epsilon$-regular if for every $X' \subset X$ and $Y' \subset Y$ with $|X'| > \epsilon |X|$ and $|Y'| > \epsilon |Y|$ we have $|d(X',Y')-d(X,Y)|<\epsilon$. A partition $V=\bigcup_{i=1}^k V_i$ is called {\it equitable} if $\big||V_i|-|V_j|\big|\leq 1$ for all $i,j$. Szemer\'edi's regularity lemma \cite{Sz} states that for each $\epsilon>0$, there is a positive integer $M(\epsilon)$ such that the vertices of any graph $G$ can be equitably partitioned $V(G)=\bigcup_{i=1}^k V_i$ into $k$ subsets with $\epsilon^{-1} \leq k \leq M(\epsilon)$ satisfying that all but at most $\epsilon k^2$ of the pairs $(V_i,V_j)$ are $\epsilon$-regular. For more background on the regularity lemma, see the excellent survey by Koml\'os and Simonovits \cite{KoSi}. In the regularity lemma, $M(\epsilon)$ can be taken to be a tower of $2$'s of height proportional to $\epsilon^{-5}$. On the other hand, Gowers \cite{Go} proved a lower bound on $M(\epsilon)$ which is a tower of $2$'s of height proportional to $\epsilon^{-\frac{1}{16}}$. His result demonstrates that $M(\epsilon)$ is inherently large as a function of $\epsilon^{-1}$. Unfortunately, this implies that the bounds obtained by applications of the regularity lemma are often quite poor. In particular, this is a weakness of the bound on $\delta(\epsilon,H)$ given by R\"odl's proof of his theorem. It is therefore desirable to find a new proof of R\"odl's theorem that does not use the regularity lemma. The following theorem does just that, giving a much better bound on $\delta(\epsilon,H)$. Its proof works as well in a multicolor setting (see concluding remarks). \begin{theorem} \label{main} There is a constant $c$ such that for each $\epsilon \in (0,1/2)$ and graph $H$ on $k \geq 2$ vertices, every $H$-free graph on $n$ vertices contains an induced subgraph on at least $2^{-ck(\log \frac{1}{\epsilon})^2}n$ vertices with edge density either at most $\epsilon$ or at least $1-\epsilon$. \end{theorem} Nikiforov \cite{Ni} recently strengthened R\"odl's theorem by proving that for each $\epsilon>0$ and graph $H$ of order $k$, there are positive constants $\kappa=\kappa(\epsilon,H)$ and $C=C(\epsilon,H)$ such that for every graph $G=(V,E)$ that contains at most $\kappa|V|^{k}$ induced copies of $H$, there is an equitable partition $V=\bigcup_{i=i}^C V_i$ of the vertex set such that the edge density in each $V_i$ ($i \geq 1$) is at most $\epsilon$ or at least $1-\epsilon$. Using the same technique as the proof of Theorem \ref{main}, we give a new proof of this result without using the regularity lemma, thereby solving the main open problem posed in \cite{Ni}. Erd\H{o}s and Hajnal \cite{ErHa} gave a significant improvement on the Erd\H{o}s-Szekeres bound on the size of the largest homogeneous set in $H$-free graphs. They proved that for every graph $H$ there is a positive constant $c(H)$ such that $\hom(G) \geq 2^{c(H)\sqrt{\log n}}$ for all $H$-free graphs $G$ on $n$ vertices. Erd\H{o}s and Hajnal further conjectured that every such $G$ contains a complete or empty subgraph of order $n^{c(H)}$. This beautiful problem has received increasing attention by various researchers, and was also featured by Gowers \cite{Go1} in his list of problems at the turn of the century. For various partial results on the Erd\H{o}s-Hajnal conjecture see, e.g., \cite{AlPaSo, ErHaPa, FoSu, AlPaPiRaSh, FoPaTo2, LaMaPaTo, ChSa} and their references. Recall that a graph is $k$-universal if it contains all graphs on at most $k$ vertices as induced subgraphs. Note that the Erd\H{o}s-Hajnal bound, in particular, implies that, for every fixed $k$, sufficiently large Ramsey graphs are $k$-universal. This was extended further by Pr\"omel and R\"odl \cite{PrRo}, who obtained an asymptotically best possible result. They proved that if $\hom(G) \leq c_1 \log n$ then $G$ is $c_2\log n$-universal for some constant $c_2$ which depends on $c_1$. Let $\hom(n,k)$ be the largest positive integer such that every graph $G$ on $n$ vertices is $k$-universal or satisfies $\hom(G) \geq \hom(n,k)$. The Erd\H{o}s-Hajnal theorem and the Promel-R\"odl theorem both say that $\hom(n,k)$ is large for fixed or slowly growing $k$. Indeed, from the first theorem it follows that for fixed $k$ there is $c(k)>0$ such that $\hom(n,k) \geq 2^{c(k)\sqrt{\log n}}$, while the second theorem says that for each $c_1$ there is $c_2>0$ such that $\hom(n,c_2\log n) \geq c_1 \log n$. One would naturally like to have a general lower bound on $\hom(n,k)$ that implies both the Erd\H{o}s-Hajnal and Promel-R\"odl results. This is done in the following theorem. \begin{theorem}\label{combined} There are positive constants $c_3$ and $c_4$ such that for all $n,k$, every graph on $n$ vertices is $k$-universal or satisfies $\hom(G) \geq c_32^{c_4\sqrt{\frac{\log n}{k}}}\log n.$ \end{theorem} Theorem \ref{main} can be also used to answer a question of Chung and Graham \cite{ChGr}, which was motivated by the study of quasirandom graphs. Given a fixed graph $H$, it is well known that a typical graph on $n$ vertices contains many induced copies of $H$ as $n$ becomes large. Therefore if a large graph $G$ contains no induced copy of $H$, its edge distribution should deviate from ``typical" in a rather strong way. This intuition was made rigorous in \cite{ChGr}, where the authors proved that if a graph $G$ on $n$ vertices is not $k$-universal, then there is a subset $S$ of $\lfloor \frac{n}{2} \rfloor$ vertices of $G$ such that $|e(S)-\frac{1}{16}n^2|>2^{-2(k^2+27)}n^2$. For positive integers $k$ and $n$, let $D(k,n)$ denote the largest integer such that every graph $G$ on $n$ vertices that is not $k$-universal contains a subset $S$ of vertices of size $\lfloor \frac{n}{2} \rfloor$ with $|e(S)-\frac{1}{16}n^2|>D(k,n)$. Chung and Graham asked whether their lower bound on $D(k,n)$ can be substantially improved, e.g., replaced by $c^{-k}n^2$. Using Theorem \ref{main} this can be easily done as follows. A lemma of Erd\H{o}s, Goldberg, Pach, and Spencer \cite{ErGoPaSp} implies that if a graph on $n$ vertices has a subset $R$ that deviates by $D$ edges from having edge density $1/2$, then there is a subset $S$ of size $\lfloor n/2 \rfloor$ that deviates by at least a constant times $D$ edges from having edge density $1/2$. By Theorem \ref{main} with $\epsilon=1/4$, there is a positive constant $C$ such that every graph on $n$ vertices that is not $k$-universal has a subset $R$ of size at least $C^{-k}n$ with edge density at most $1/4$ or at least $3/4$. This $R$ deviates from having edge density $1/2$ by at least $$\frac{1}{4}{|R| \choose 2}\geq \frac{1}{16}|R|^2 \geq \frac{1}{16}C^{-2k}n^2$$ edges. Thus, the above mentioned lemma from \cite{ErGoPaSp} implies that there is an absolute constant $c$ such that every graph $G$ on $n$ vertices which is not $k$-universal contains a subset $S$ of size $\lfloor n/2 \rfloor$ with $|e(S)-\frac{n^2}{16}|>c^{-k}n^2$. Chung and Graham also ask for non-trivial upper bounds on $D(k,n)$. In this direction, we show that there are $K_k$-free graphs on $n$ vertices for which $|e(S)-\frac{1}{16}n^2|=O(2^{-k/4}n^2)$ holds for every subset $S$ of $\lfloor \frac{n}{2} \rfloor$ vertices of $G$. Together with the lower bound it determines the asymptotic behavior of $D(k,n)$ and shows that there are constants $c_1,c_2>1$ such that $c_1^{-k}n^2<D(k,n)<c_2^{-k}n^2$ holds for all positive integers $k$ and $n$. This completely answers the questions of Chung and Graham. Moreover, we can obtain a more precise result about the relation between the number of induced copies of a fixed graph $H$ in a large graph $G$ and the edge distribution of $G$. In their celebrated paper, Chung, Graham, and Wilson \cite{ChGrWi} introduced a large collection of equivalent graph properties shared by almost all graphs which are called {\it quasirandom}. For a graph $G=(V,E)$ on $n$ vertices, two of these properties are \vspace{0.1cm} ${\bf P_1}:\textrm{For each subset}~S \subset V,$ $$e(S)=\frac{1}{4}|S|^2+o(n^2).$$ ${\bf P}_2$: For every fixed graph $H$ with $k$ vertices, the number of labeled induced copies of $H$ in $G$ is $$(1+o(1))n^{k}2^{-{k \choose 2}}.$$ So one can ask naturally, by how much does a graph deviate from ${\bf P}_1$ assuming a deviation from ${\bf P}_2$? The following theorem answers this question. \begin{theorem}\label{dev} Let $H$ be a graph with $k$ vertices and $G=(V,E)$ be a graph with $n$ vertices and at most $(1-\epsilon)2^{-{k \choose 2}}n^k$ labeled induced copies of $H$. Then there is a subset $S \subset V$ with $|S|=\lfloor n/2 \rfloor$ and $|e(S)-\frac{n^2}{16}| \geq \epsilon c^{-k} n^2$, where $c$ is an absolute constant. \end{theorem} The proof of Theorem \ref{dev} can be easily adjusted if we replace the ``at most'' with ``at least'' and the $(1-\epsilon)$ factor by $(1+\epsilon)$. Note that this theorem answers the original question of Chung and Graham in a very strong sense. \subsection{Induced Ramsey numbers}\label{inducedsubsection} Recall that the induced Ramsey number $r_{\textrm{ind}}(H)$ is the minimum $n$ for which there is a graph $G$ with $n$ vertices such that for every $2$-edge-coloring of $G$, one can find an induced copy of $H$ in $G$ whose edges are monochromatic. One of the fundamental results in graph Ramsey theory (see chapter 9.3 of \cite{Di}), the Induced Ramsey Theorem, says that $r_{\textrm{ind}}(H)$ exists for every graph $H$. R\"odl \cite{Ro} noted that a relatively simple proof of the theorem follows from a simple application of his result discussed in the previous section. However, all of the early proofs of the Induced Ramsey Theorem give poor upper bounds on $r_{\textrm{ind}}(H)$. Since these early proofs, there has been a considerable amount of research on induced Ramsey numbers. Erd\H{o}s \cite{Er2} conjectured that there is a constant $c$ such that every graph $H$ on $k$ vertices satisfies $r_{\textrm{ind}}(H) \leq 2^{ck}$. Erd\H{o}s and Hajnal \cite{Er1} proved that $r_{\textrm{ind}}(H) \leq 2^{2^{k^{1+o(1)}}}$ holds for every graph $H$ on $k$ vertices. Kohayakawa, Pr\"omel, and R\"odl \cite{KoPrRo} improved this bound substantially and showed that if a graph $H$ has $k$ vertices and chromatic number $\chi$, then $r_{\textrm{ind}}(H) \leq k^{ck(\log \chi)},$ where $c$ is a universal constant. In particular, their result implies an upper bound of $2^{ck (\log k)^2}$ on the induced Ramsey number of any graph on $k$ vertices. In their proof, the graph $G$ which gives this bound is randomly constructed using projective planes. There are several known results that provide upper bounds on induced Ramsey numbers for sparse graphs. For example, Beck \cite{Be} studied the case when $H$ is a tree; Haxell, Kohayakawa, and \L uczak \cite{HaKoLu} proved that the cycle of length $k$ has induced Ramsey number linear in $k$; and, settling a conjecture of Trotter, \L uczak and R\"odl \cite{LuRo} showed that the induced Ramsey number of a graph with bounded degree is at most polynomial in the number of its vertices. More precisely, they proved that for every integer $d$, there is a constant $c_d$ such that every graph $H$ on $k$ vertices and maximum degree at most $d$ satisfies $r_{\textrm{ind}}(H) \leq k^{c_d}$. Their proof, which also uses random graphs, gives an upper bound on $c_d$ that is a tower of $2$'s of height proportional to $d^2$. As noted by Schaefer and Shah \cite{SchSh}, all known proofs of the Induced Ramsey Theorem either rely on taking $G$ to be an appropriately chosen random graph or give a poor upper bound on $r_{\textrm{ind}}(H)$. However, often in combinatorics, explicit constructions are desirable in addition to existence proofs given by the probabilistic method. For example, one of the most famous such problems was posed by Erd\H{o}s \cite{AlSp}, who asked for the explicit construction of a graph on $n$ vertices without a complete or empty subgraph of order $ c\log n$. Over the years, this intriguing problem and its bipartite variant has drawn a lot of attention by various researches (see, e.g., \cite{FrWi,Al2,BaKiShSuWi,Bo,BaRaShWi}), but, despite these efforts, it is still open. Similarly, one would like to have an explicit construction for the Induced Ramsey Theorem. We obtain such a construction using pseudo-random graphs. The {\it random graph} $G(n,p)$ is the probability space of all labeled graphs on $n$ vertices, where every edge appears randomly and independently with probability $p$. An important property of $G(n,p)$ is that, with high probability, between any two large subsets of vertices $A$ and $B$, the edge density $d(A,B)=\frac{e(A,B)}{|A||B|}$ is approximately $p$. This observation is one of the motivations for the following useful definition. A graph $G=(V,E)$ is {\it $(p,\lambda)$-pseudo-random} if the following inequality holds for all subsets $A,B \subset V$: $$|d(A,B)-p| \leq \frac{\lambda}{\sqrt{|A||B|}}.$$ It is easy to show that if $p<0.99$, then with high probability, the random graph $G(n,p)$ is $(p,\lambda)$-pseudo-random with $\lambda=O(\sqrt{pn})$. Moreover, there are also many explicit constructions of pseudo-random graphs which can be obtained using the following fact. Let $\lambda_1 \geq \lambda_2 \geq \ldots \geq \lambda_n$ be the eigenvalues of the adjacency matrix of a graph $G$. An {\it $(n,d,\lambda)$-graph} is a $d$-regular graph on $n$ vertices with $\lambda = \max_{i \geq 2} |\lambda_i|$. It was proved by Alon (see, e.g., \cite{AlSp}, \cite{KrSu}) that every $(n,d,\lambda)$-graph is in fact $(\frac{d}{n},\lambda)$-pseudo-random. Therefore to construct good pseudo-random graphs we need regular graphs with $\lambda \ll d$. For more details on pseudo-random graphs, including many constructions, we refer the interested reader to the recent survey \cite{KrSu}. A graph is {\it $d$-degenerate} if every subgraph of it has a vertex of degree at most $d$. The degeneracy number of a graph $H$ is the smallest $d$ such that $H$ is $d$-degenerate. This quantity, which is always bounded by the maximum degree of the graph, is a natural measure of its sparseness. In particular, in a $d$-degenerate graph every subset $X$ spans at most $d|X|$ edges. The chromatic number $\chi(H)$ of graph $H$ is the minimum number of colors needed to color vertices of $H$ such that adjacent vertices get different colors. Using a greedy coloring, it is easy to show that $d$-degenerate graphs have chromatic number at most $d+1$. The following theorem, which is special case of a more general result which we prove in Section 4, shows that any sufficiently pseudo-random graph of appropriate density has strong induced Ramsey properties. \begin{theorem}\label{quasicor1} There is an absolute constant $c$ such that for all integers $k,d,\chi \geq 2$, every $(\frac{1}{k},n^{0.9})$-pseudo-random graph $G$ on $n \geq k^{cd\log \chi}$ vertices satisfies that every $d$-degenerate graph on $k$ vertices with chromatic number at most $\chi$ occurs as an induced monochromatic copy in all $2$-edge-colorings of $G$. Moreover, all of these induced monochromatic copies can be found in the same color. \end{theorem} This theorem implies that, with high probability, $G(n,p)$ with $p=1/k$ and $n \geq k^{cd\log \chi}$ satisfies that every $d$-degenerate graph on $k$ vertices with chromatic number at most $\chi$ occurs as an induced monochromatic copy in all $2$-edge-colorings of $G$. It gives the first polynomial upper bound on the induced Ramsey numbers of $d$-degenerate graphs. In particular, for bounded degree graphs this is a significant improvement of the above mentioned \L uczak-R\"odl result. It shows that the exponent of the polynomial in their theorem can be taken to be $O(d\log d)$, instead of the previous bound of a tower of $2$'s of height proportional to $d^2$. \begin{corollary}\label{corollaryobvious} There is an absolute constant $c$ such that every $d$-degenerate graph $H$ on $k$ vertices with chromatic number $\chi \geq 2$ has induced Ramsey number $r_{\textrm{ind}}(H) \leq k^{cd \log \chi}$. \end{corollary} A significant additional benefit of Theorem \ref{quasicor1} is that it leads to explicit constructions for induced Ramsey numbers. One such example can be obtained from a construction of Delsarte and Goethals and also of Turyn (see \cite{KrSu}). Let $r$ be a prime power and let $G$ be a graph whose vertices are the elements of the two dimensional vector space over finite field $\mathbb{F}_r$, so $G$ has $r^2$ vertices. Partition the $r+1$ lines through the origin of the space into two sets $P$ and $N$, where $|P|=t$. Two vertices $x$ and $y$ of the graph $G$ are adjacent if $x-y$ is parallel to a line in $P$. This graph is known to be $t(r-1)$-regular with eigenvalues, besides the largest one, being either $-t$ or $r-t$. Taking $t= \frac{r^2}{k(r-1)}$, we obtain an $(n,d,\lambda)$-graph with $n=r^2$, $d = n/k$, and $\lambda=r-t\leq r \leq n^{1/2}$. This gives a $(p,\lambda)$-pseudo-random graph with $p =d/n=1/k$ and $\lambda \leq n^{1/2}$ which satisfies the assertion of Theorem \ref{quasicor1}. Another well-known explicit construction is the Paley graph $P_n$. Let $n$ be a prime power which is congruent to 1 modulo 4 so that $-1$ is a square in the finite field $\mathbb{F}_n$. The {\it Paley graph} $P_n$ has vertex set $\mathbb{F}_n$ and distinct elements $x,y \in \mathbb{F}_n$ are adjacent if $x-y$ is a square. It is well known and not difficult to prove that the Paley graph $P_n$ is $(1/2,\lambda)$-pseudo-random with $\lambda=\sqrt{n}$. This can be used together with the generalization of Theorem \ref{quasicor1}, which we discuss in Section 5, to prove the following result. \begin{corollary}\label{payley} There is an absolute constant $c$ such that for prime $n \geq 2^{ck\log^2 k}$, every graph on $k$ vertices occurs as an induced monochromatic copy in all $2$-edge-colorings of the Paley graph $P_n$. \end{corollary} This explicit construction matches the best known upper bound on induced Ramsey numbers of graphs on $k$ vertices obtained by Kohayakawa, Pr\"omel, and R\"odl \cite{KoPrRo}. Similarly, we can prove that there is a constant $c$ such that, with high probability, $G(n,1/2)$ with $n \geq 2^{ck\log^2 k}$ satisfies that every graph on $k$ vertices occurs as an induced monochromatic copy in all $2$-edge-colorings of $G$. Very little is known about lower bounds for induced Ramsey numbers beyond the fact that an induced Ramsey number is at least its corresponding Ramsey number. A well-known conjecture of Burr and Erd\H{o}s \cite{BuEr} from 1973 states that for each positive integer $d$ there is a constant $c(d)$ such that the Ramsey number $r(H)$ is at most $c(d)k$ for every $d$-degenerate graph $H$ on $k$ vertices. As mentioned earlier, Haxell et al. \cite{HaKoLu} proved that the induced Ramsey number for the cycle on $k$ vertices is linear in $k$. This implies that the induced Ramsey number for the path on $k$ vertices is also linear in $k$. Also, using a star with $2k-1$ edges, it is trivial to see that the induced Ramsey number of a star with $k$ edges is $2k$. It is natural to ask whether the Burr-Erd\H{o}s conjecture extends to induced Ramsey numbers. The following result shows that this fails already for trees, which are $1$-degenerate graphs. \begin{theorem} \label{tree} For every $c>0$ and sufficiently large integer $k$ there is a tree $T$ on $k$ vertices such that $r_{\textrm{ind}}(T) \geq ck$. \end{theorem} The tree $T$ in the above theorem can be taken to be any sufficiently large tree that contains a matching of linear size and a star of linear size as subgraphs. It is interesting that the induced Ramsey number for a path on $k$ vertices or a star on $k$ vertices is linear in $k$, but the induced Ramsey number for a tree which contains both a path on $k$ vertices and a star on $k$ vertices is superlinear in $k$. \vspace{0.3cm} \noindent {\bf Organization of the paper.}\,\, In the next section we give short proofs of Theorem \ref{main} and Theorem \ref{combined} which illustrate our methods. Section \ref{section12} contains the key lemma that is used as a replacement for Szemer\'edi's regularity lemma in the proofs of several results. We answer questions of Chung-Graham and Nikiforov on the edge distribution in graphs with a forbidden induced subgraph in Section \ref{sectionmulticolor}. In Section \ref{moreoninduced} we show that any sufficiently pseudo-random graph of appropriate density has strong induced Ramsey properties. Combined with known examples of pseudo-random graphs, this leads to explicit constructions which match and improve the best known estimates for induced Ramsey numbers. The proof of the result that there are trees whose induced Ramsey number is superlinear in the number of vertices is in Section \ref{superlinear}. The last section of this paper contains some concluding remarks together with a discussion of a few conjectures and open problems. Throughout the paper, we systematically omit floor and ceiling signs whenever they are not crucial for the sake of clarity of presentation. We also do not make any serious attempt to optimize absolute constants in our statements and proofs. \section{Ramsey-type results for $H$-free graphs}\label{section2} In this section, we prove Theorems \ref{main} and \ref{combined}. While we obtain more general results later in the paper, the purpose of this section is to illustrate on simple examples the main ideas and techniques that we will use in our proofs. Our theorems strengthen and generalize results from \cite{Ro} and \cite{PrRo} and the proofs we present here are shorter and simpler than the original ones. We start with the proof of Theorem \ref{main}, which uses the following lemma of Erd\H{o}s and Hajnal \cite{ErHa}. We prove a generalization of this lemma in Section \ref{sectionmulticolor}. \begin{lemma}\label{lemmaerdoshajnal} For each $\epsilon \in (0,1/2)$, graph $H$ on $k$ vertices, and $H$-free graph $G=(V,E)$ on $n \geq 2$ vertices, there are disjoint subsets $A$ and $B$ of $V$ with $|A|, |B| \geq \epsilon^{k-1}\frac{n}{k}$ such that either every vertex in $A$ has at most $\epsilon|B|$ neighbors in $B$, or every vertex in $A$ has at least $(1-\epsilon)|B|$ neighbors in $B$. \end{lemma} Actually, the statement of the lemma in \cite{ErHa} is a bit weaker than that of Lemma \ref{lemmaerdoshajnal} but it is easy to get the above statement by analyzing more carefully the proof of Erd\H{o}s and Hajnal. Lemma \ref{lemmaerdoshajnal} roughly says that every $H$-free graph contains two large disjoint vertex subsets such that the edge density between them is either very small or very large. However, to prove Theorem \ref{main}, we need to find a large induced subgraph with such edge density. Our next lemma shows how one can iterate the bipartite density result of Lemma \ref{lemmaerdoshajnal} in order to establish the complete density result of Theorem \ref{main}. For $\epsilon_1,\epsilon_2 \in (0,1)$ and a graph $H$, define $\delta(\epsilon_1,\epsilon_2,H)$ to be the largest $\delta$ (which may be 0) such that for each $H$-free graph on $n$ vertices, there is an induced subgraph on at least $\delta n$ vertices with edge density at most $\epsilon_1$ or at least $1-\epsilon_2$. Notice that for $2 \leq n_0 \leq n_1$, the edge-density of a graph on $n_1$ vertices is the average of the edge-densities of the induced subgraphs on $n_0$ vertices. Therefore, from definition of $\delta$, it follows that for every $2 \leq n_0 \leq \delta(\epsilon_1,\epsilon_2,H) n$ and $H$-free graph $G$ on $n$ vertices, $G$ contains an induced subgraph on exactly $n_0$ vertices with edge density at most $\epsilon_1$ or at least $1-\epsilon_2$. Recall that the edge-density $d(A)$ of a subset $A$ of $G$ equals $e(A)/{|A| \choose 2}$, where $e(A)$ is the number of edges spanned by $A$. \begin{lemma}\label{useful} Suppose $\epsilon_1,\epsilon_2 \in (0,1)$ with $\epsilon_1+\epsilon_2<1$ and $H$ is a graph on $k \geq 2$ vertices. Let $\epsilon=\min(\epsilon_1,\epsilon_2)$. We have $$\delta(\epsilon_1,\epsilon_2,H) \geq (\epsilon/4)^{k}k^{-1}\min\Big(\delta\big(3\epsilon_1/2,\epsilon_2,H\big), \delta\big(\epsilon_1, 3\epsilon_2/2,H\big)\Big).$$ \end{lemma} \begin{proof} Let $G$ be a $H$-free graph on $n \geq 2$ vertices. If $n<k$ then we may consider any two-vertex induced subgraph of $G$ which has always density either 0 or 1. Therefore, for $G$ of order less than $k$ we can take $\delta=2/k$, which is clearly larger than the right hand side of the inequality in the assertion of the lemma. Thus we can assume that $n \geq k$. Applying Lemma \ref{lemmaerdoshajnal} to $G$ with $\epsilon/4$ in place of $\epsilon$, we find two subsets $A$ and $B$ with $|A|,|B| \geq (\epsilon/4)^{k-1} n/k $, such that either every vertex in $A$ is adjacent to at most $\frac{\epsilon}{4}|B|$ vertices of $B$ or every vertex of $A$ is adjacent to at least $(1-\frac{\epsilon}{4})|B|$ vertices of $B$. Consider the first case in which every vertex in $A$ is adjacent to at most $\frac{\epsilon}{4}|B|$ vertices of $B$ (the other case can be treated similarly) and let $G[A]$ be the subgraph of $G$ induced by the set $A$. By definition of function $\delta$, $G[A]$ contains a subset $A'$ with $$|A'|= \delta(3\epsilon_1/2,\epsilon_2,H)\Big(\frac{\epsilon}{4}\Big)^{k}\frac{n}{k} \leq \delta(3\epsilon_1/2,\epsilon_2,H)|A|,$$ such that the subgraph induced by $A'$ has edge density at most $\frac{3}{2}\epsilon_1$ or at least $1-\epsilon_2$. If $A'$ has edge density at least $1-\epsilon_2$ we are done, since $G[A']$ is an induced subgraph of $G$ with at least $(\epsilon/4)^{k}k^{-1}\delta(3\epsilon_1/2,\epsilon_2,H)n$ vertices and edge density at least $1-\epsilon_2$. So we may assume that the edge density in $A'$ is at most $\frac{3}{2}\epsilon_1$. Let $B_1 \subset B$ be those vertices of $B$ that have at most $\frac{\epsilon}{2}|A'|$ neighbors in $A'$. Since $A' \subset A$, each vertex of $A'$ has at most $\frac{\epsilon}{4}|B|$ neighbors in $B$ and the number of edges $e(A',B) \leq \frac{\epsilon}{4}|A'||B|$. Therefore $B_1$ has at least $|B|/2$ vertices. Then, by definition of $\delta$, $B_1$ contains a subset $B'$ with $$|B'| = \delta(3\epsilon_1/2,\epsilon_2,H)\Big(\frac{\epsilon}{4}\Big)^{k}\frac{n}{k} \leq \delta(3\epsilon_1/2,\epsilon_2,H)|B_1|,$$ such that the induced subgraph $G[B']$ has edge density at most $\frac{3}{2}\epsilon_1$ or at least $1-\epsilon_2$. If it has edge density at least $1-\epsilon_2$ we are done, so we may assume that the edge density $d(B')$ is at most $\frac{3}{2}\epsilon_1$. Finally to complete the proof note that, since $|A'|=|B'|$, $|A' \cup B'|=2|A'|$, $d(A'),d(B') \leq \frac{3}{2}\epsilon_1$, and $d(A',B') \leq \frac{\epsilon_1}{2}$, we have that \begin{eqnarray*} e(A' \cup B')&=&e(A')+e(B')+e(A',B') \leq \frac{3}{2}\epsilon_1 {|A'| \choose 2}+ \frac{3}{2}\epsilon_1 {|B'| \choose 2}+ \frac{\epsilon_1}{2}|A'||B'|\\ &=&2\epsilon_1|A'|^2-3\epsilon_1 |A'|/2 \leq \epsilon_1{2|A'| \choose 2}. \end{eqnarray*} Therefore, $d(A' \cup B') \leq \epsilon_1$. \end{proof} From this lemma, the proof of our first result, that every $H$-free graph on $n$ vertices contains a subset of at least $2^{-ck(\log \frac{1}{\epsilon})^2}n$ vertices with edge density either $ \leq \epsilon$ or $\geq 1-\epsilon$, follows in a few lines. \noindent {\bf Proof of Theorem \ref{main}:} Notice that if $\epsilon_1+\epsilon_2 \geq 1$, then trivially $\delta(\epsilon_1,\epsilon_2,H) =1$. In particular, if $\epsilon_1\epsilon_2 \geq \frac{1}{4}$, then $\epsilon_1+\epsilon_2 \geq 1$ and $\delta(\epsilon_1,\epsilon_2,H) =1$. Therefore, by iterating Lemma \ref{useful} for $t=\log \frac{1}{\epsilon^2}/\log \frac{3}{2}$ iterations and using that $\epsilon \leq 1/2$, we obtain $$\delta(\epsilon,\epsilon,H) \geq \left(\frac{\epsilon^{k}}{4^{k}k}\right)^t \geq 2^{-\frac{2}{\log 3/2}\left(k(\log 1/\epsilon)^2+\left(2k+\log k \right)\log 1/\epsilon\right)} \geq 2^{-15k (\log 1/\epsilon)^2},$$ which, by definition of $\delta$, completes the proof of the theorem. \ifvmode\mbox{ }\else\unskip\fi\hskip 1em plus 10fill$\Box$ Recall the Erd\H{o}s-Szemer\'edi theorem, which states that there is an absolute constant $c$ such that every graph $G$ on $n$ vertices with edge density $\epsilon \in (0,1/2)$ has a homogeneous set of size at least $\frac{c \log n}{\epsilon \log \frac{1}{\epsilon}}$. Theorem \ref{combined} follows from a simple application of Theorem \ref{main} and the Erd\H{o}s-Szemer\'edi theorem. \noindent {\bf Proof of Theorem \ref{combined}:} Let $G$ be a graph on $n$ vertices which is not $k$-universal, i.e., it is $H$-free for some fixed graph $H$ on $k$ vertices. Fix $\epsilon=2^{-\frac{1}{5}\sqrt{\frac{\log n}{k}}}$ and apply Theorem \ref{main} to $G$. It implies that $G$ contains a subset $W \subset V(G)$ of size at least $2^{-15k(\log \frac{1}{\epsilon})^2}n=n^{2/5}$ such that the subgraph induced by $W$ has edge density at most $\epsilon$ or at least $1-\epsilon$. Applying the Erd\H{o}s-Szemer\'edi theorem to the induced subgraph $G[W]$ or its complement and using that $\epsilon \log 1/\epsilon \leq 4\epsilon^{1/2}$ for all $\epsilon \leq 1$, we obtain a homogeneous subset $W' \subset W$ with $$|W'| \geq \frac{c \log n^{2/5}}{\epsilon \log \frac{1}{\epsilon}} \geq \frac{c\log n}{10\epsilon^{1/2}} \geq \frac{c}{10}2^{\frac{1}{10}\sqrt{\frac{\log n}{k}}}\log n,$$ which completes the proof of Theorem \ref{combined}. \ifvmode\mbox{ }\else\unskip\fi\hskip 1em plus 10fill$\Box$ \section{Key Lemma}\label{section12} In this section we present our key lemma. We use it as a replacement for Szemer\'edi's regularity lemma in the proofs of several Ramsey-type results, thereby giving much better estimates. A very special case of this statement was essentially proved in Lemma \ref{useful} in the previous section. Our key lemma generalizes the result of Graham, R\"odl, and Rucinski \cite{GrRoRu} and has a simpler proof than the one in \cite{GrRoRu}. Roughly, our result says that if $(G_1,\ldots,G_r)$ is a sequence of graphs on the same vertex set $V$ with the property that every large subset of $V$ contains a pair of large disjoint sets with small edge density between them in at least one of the graphs $G_i$, then every large subset of $V$ contains a large set with small edge density in one of the $G_i$. To formalize this concept, we need a couple definitions. For a graph $G=(V,E)$ and disjoint subsets $W_1,\ldots,W_t \subset V$, the {\it density} $d_{G}(W_1,\ldots,W_t)$ between the $t \geq 2$ vertex subsets $W_1,\ldots,W_t$ is defined by $$d_G(W_1,\ldots,W_t)=\frac{\sum_{ i < j} e(W_i,W_j)}{\sum_{i < j } |W_i||W_j|}.$$ If $|W_1|=\ldots=|W_t|$, then $$d_G(W_1,\ldots,W_t)={t \choose 2}^{-1}\sum_{ i < j } d_G(W_i,W_j).$$ Also, in this section if $t=1$ we define the density to be zero. \begin{definition} \label{d31} For $\alpha,\rho,\epsilon \in [0,1]$ and positive integer $t$, a sequence $(G_1,\ldots,G_r)$ of graphs on the same vertex set $V$ is {\bf $(\alpha,\rho,\epsilon,t)$-sparse} if for all subsets $U \subset V$ with $|U| \geq \alpha |V|$, there are positive integers $t_1,\ldots,t_r$ such that $\prod_{i=1}^r t_i \geq t$ and for each $i \in [r]=\{1, \ldots, r\}$ there are disjoint subsets $W_{i,1},\ldots,W_{i,t_i} \subset U$ with $|W_{i,1}|=\ldots =|W_{i,t_i}|=\lceil \rho |U|\rceil$ and $d_{G_i}(W_{i,1},\ldots,W_{i,t_i})\leq \epsilon$. \end{definition} We call a graph $(\alpha,\rho,\epsilon,t)$-sparse if the one-term sequence consisting of that graph is $(\alpha,\rho,\epsilon,t)$-sparse. By averaging, if $\alpha'\geq \alpha$, $\rho' \leq \rho$, $\epsilon'\geq \epsilon$, $t' \leq t$, and $(G_1,\ldots,G_r)$ is $(\alpha,\rho,\epsilon,t)$-sparse, then $(G_1,\ldots,G_r)$ is also $(\alpha',\rho',\epsilon',t')$-sparse. The following is our main result in this section. \begin{lemma} \label{l32} If a sequence of graphs $(G_1,\ldots,G_r)$ with common vertex set $V$ is $(\frac{1}{2}\alpha\rho,\rho',\epsilon,t)$-sparse and $(\alpha,\rho,\epsilon/4,2)$-sparse, then $(G_1,\ldots,G_r)$ is also $(\alpha,\frac{1}{2}\rho\rho',\epsilon,2t)$-sparse. \end{lemma} \begin{proof} Since $(G_1,\ldots,G_r)$ is $(\alpha,\rho,\epsilon/4,2)$-sparse, then for each $U \subset V$ with $|U| \geq \alpha |V|$, there is $i \in [r]$ and disjoint subsets $X,Y \subset U$ with $|X|=|Y| = \rho|U|$ and $d_{G_{i}}(X,Y) \leq \epsilon/4$. Let $X_1$ be the set of vertices in $X$ that have at most $\frac{\epsilon}{2}|Y|$ neighbors in $Y$ in graph $G_i$. Then $e_{G_i}(X\setminus X_1,Y) \geq \epsilon|X\setminus X_1||Y|/2$ and we also have $e_{G_i}(X,Y) \leq \epsilon|X||Y|/4$. Therefore $|X_1| \geq |X|/2 \geq \frac{1}{2}\rho|U|$ and by removing extra vertices we assume that $|X_1|= \frac{1}{2}\rho|U|$. Since $(G_1,\ldots,G_r)$ is $(\frac{1}{2}\alpha\rho,\rho',\epsilon,t)$-sparse, then there are positive integers $t_{1},\ldots,t_{r}$ such that $\prod_{j=1}^r t_{j} \geq t$ and for each $j \in [r]$ there are disjoint subsets $X_{j,1},\ldots,X_{j,t_j} \subset X_1$ of size $|X_{j,1}|=\ldots=|X_{j,t_j}|=\rho' |X_1|$ with density $d_{G_j}(X_{j,1},\ldots,X_{j,t_j})\leq \epsilon$. Let $Y_1$ the set of vertices in $Y$ that have at most $\epsilon|X_{i,1} \cup \ldots \cup X_{i,t_i}|$ neighbors in $X_{i,1} \cup \ldots \cup X_{i,t_i}$ in graph $G_i$. Since every vertex of $X_1$ is adjacent to at most $\frac{\epsilon}{2}|Y|$ vertices of $Y$ and since $X_{i,1} \cup \ldots \cup X_{i,t_i}\subset X_1$ we have that $d_{G_i}(X_{i,1} \cup \ldots \cup X_{i,t_i},Y)\leq \epsilon/2$. On the other hand, $d_{G_i}(X_{i,1} \cup \ldots \cup X_{i,t_i},Y\setminus Y_1)\geq \epsilon$. Therefore $|Y_1|\geq |Y|/2$, so again we can assume that $|Y_1|=\frac{1}{2}\rho|U|=|X_1|$. Since $(G_1,\ldots,G_r)$ is $(\frac{1}{2}\alpha\rho,\rho',\epsilon,t)$-sparse, then there are positive integers $s_{1},\ldots,s_{r}$ such that $\prod_{j=1}^r s_{j} \geq t$ and for each $j \in [r]$ there are disjoint subsets $Y_{j,1},\ldots,Y_{j,s_j} \subset Y_1$ with $d_{G_j}(Y_{j,1},\ldots,Y_{j,s_j})\leq \epsilon$ and $|Y_{j,1}|=\ldots=|Y_{j,s_j}|=\rho' |Y_1|$. By the above construction, the edge density between $X_{i,1}\cup \ldots \cup X_{i,t_i}$ and $Y_{i,1}\cup \ldots \cup Y_{i,s_i}$ is bounded from above by $\epsilon$. We also have that both $d_{G_i}(X_{i,1},\ldots,X_{i,t_i})$ and $d_{G_i}(Y_{i,1},\ldots,Y_{i,s_i})$ are at most $\epsilon$ and these two sets have the same size. Therefore $d_{G_i}(X_{i,1},\ldots,X_{i,t_i},Y_{i,1},\ldots,Y_{i,s_i}) \leq \epsilon$, implying that $(G_1,\ldots,G_r)$ is $\left(\alpha,\frac{1}{2}\rho\rho',\epsilon,u\right)$-sparse with $u=(t_i+s_i)\prod_{j \in [r]\setminus \{i\}}\max(t_j,s_j)$ for some $i$. By the arithmetic mean-geometric mean inequality, we have $$t^2 \leq \prod_{j=1}^r t_j \prod_{j=1}^r s_j \leq t_is_i\left(\prod_{j \in [r]\setminus \{i\}} \max(t_j,s_j)\right)^2 =\frac{t_is_i}{(t_i+s_i)^2}u^2 \leq \frac{u^2}{4}.$$ Thus $u \geq 2t$. Altogether this shows that $(G_1,\ldots,G_r)$ is $\left(\alpha,\frac{1}{2}\rho\rho',\epsilon,2t\right)$-sparse, completing the proof. \end{proof} Rather than using this lemma directly, in applications we usually need the following two corollaries. The first one is obtained by simply applying Lemma \ref{l32} $h-1$ times. \begin{corollary}\label{secondcor} If $(G_1,\ldots,G_r)$ is $(\alpha,\rho,\epsilon/4,2)$-sparse and $h$ is a positive integer, then $(G_1,\ldots,G_r)$ is also $\left((\frac{2}{\rho})^{h-1}\alpha,2^{1-h}\rho^{h},\epsilon,2^h\right)$-sparse. \end{corollary} If we use the last statement with $h=r\log \frac{1}{\epsilon}$ and $\alpha=(\frac{\rho}{2})^{h-1}$, then we get that there is an index $i \in [r]$ and disjoint subsets $W_1,\ldots,W_{t} \subset V$ with $t \geq 2^{h/r}= \frac{1}{\epsilon}$, $|W_1|=\ldots=|W_t|= 2^{1-h}\rho^{h}|V|$, and $d_{G_i}(W_1,\ldots,W_t) \leq \epsilon$. Since ${|W_1| \choose 2} \leq \frac{\epsilon}{t}{t|W_1| \choose 2}$, even if every $W_i$ has edge density one, still the edge density in the set $W_1 \cup \ldots \cup W_t$ is at most $2\epsilon$. Therefore, (using $\epsilon/2$ instead of $\epsilon$) we have the following corollary. \begin{corollary}\label{corbip} If $(G_1,\ldots,G_r)$ is $((\frac{\rho}{2})^{h-1},\rho,\epsilon/8,2)$-sparse where $h=r\log \frac{2}{\epsilon}$, then there is $i \in [r]$ and an induced subgraph $G'$ of $G_i$ on $2\epsilon^{-1}2^{1-h}\rho^{h}|V|$ vertices that has edge density at most $\epsilon$. \end{corollary} The key lemma in the paper of Graham, R\"odl, and Rucinski \cite{GrRoRu} on the Ramsey number of graphs (their Lemma 1) is essentially the $r=1$ case of Corollary \ref{corbip}. \section{Edge distribution in $H$-free graphs}\label{sectionmulticolor} In this section, we obtain several results on the edge distribution of graphs with a forbidden induced subgraph which answer open questions by Nikiforov and Chung-Graham. We first prove a strengthening of R\"odl's theorem (mentioned in the introduction) without using the regularity lemma. Then we present a proof of Theorem \ref{dev} on the dependence of error terms in quasirandom properties. We conclude this section with an upper bound on the maximum edge discrepancy in subgraphs of $H$-free graphs. To obtain these results we need the following generalization of Lemma \ref{lemmaerdoshajnal}. \begin{lemma}\label{lemmaerdoshajnal4} Let $H$ be a $k$-vertex graph and let $G$ be a graph on $n \geq k^2$ vertices that contains less than $n^{k}(1-\frac{k^2}{2n})\prod_{i=1}^{k-1} (1-\delta_i)\epsilon_i^{k-i}$ labeled induced copies of $H$, where $\epsilon_0=1$ and $\epsilon_i,\delta_{i} \in (0,1)$ for all $1 \leq i \leq k-1$. Then there is an index $i\leq k-1$ and disjoint subsets $A$ and $B$ of $G$ with $|A|\geq \frac{\delta_in}{k(k-i)}\prod_{j<i} \epsilon_j$ and $|B| \geq \frac{n}{k}\prod_{j < i} \epsilon_j$ such that either every vertex of $A$ is adjacent to at most $\epsilon_i|B|$ vertices of $B$ or every vertex of $A$ is adjacent to at least $(1-\epsilon_i)|B|$ vertices of $B$. \end{lemma} \begin{proof} Let $M$ denote the number of labeled induced copies of $H$ in $G$, which by our assumption is at most \begin{eqnarray} \label{M-bound} M < n^{k}\left(1-\frac{k^2}{2n}\right)\prod_{i=1}^{k-1} (1-\delta_i)\epsilon_i^{k-i}. \end{eqnarray} We may assume that the vertex set of $H$ is $[k]$. Consider a random partition $V_1 \cup \ldots \cup V_k$ of the vertices of $G$ such that each $V_i$ has cardinality $n/k$. Note that for any such partition there are $(n/k)^k$ ordered $k$-tuples of vertices of $G$ with the property that the $i$-th vertex of the $k$-tuple is in $V_i$ for all $i \in [k]$. On the other hand the total number of ordered $k$-tuples of vertices is $n(n-1)\cdots(n-k+1)$ and each of these $k$-tuples has the above property with equal probability. This implies that for any given $k$-tuple the probability that its $i$-th vertex is in $V_i$ for all $i \in [k]$ equals $\prod_{i=1}^k \frac{n/k}{n-i+1}$. In particular, by linearity of expectation, the expected number of labeled induced copies of $H$ in $G$ for which the image of every vertex $i \in [k]$ is in $V_i$ is at most $M\cdot \prod_{i=1}^k \frac{n/k}{n-i+1}$. Using that $\prod (1-x_i) \geq 1-\sum x_i$ for any $0 \leq x_i \leq 1$ and that $n \geq k^2$, we obtain \begin{eqnarray*} \prod_{i=1}^k \frac{n/k}{n-i+1} &=& k^{-k}\prod_{i=0}^{k-1}(1-i/n)^{-1} \leq k^{-k}\left(1-\sum_{i=0}^{k-1} i/n\right)^{-1} = k^{-k}\left(1-{k \choose 2}/n\right)^{-1}\\ &<& \left(1-\frac{k^2}{2n}\right)^{-1}k^{-k}. \end{eqnarray*} This, together with (\ref{M-bound}), shows that there is a partition $V_1 \cup \ldots \cup V_k$ of $G$ into sets of cardinality $n/k$ such that the total number of labeled induced copies of $H$ in $G$ for which the image of every vertex $i\in [k]$ is in $V_i$ is less than \begin{equation} \label{upbound} M\left(1-\frac{k^2}{2n}\right)^{-1}k^{-k}<k^{-k}n^{k}\prod_{i=1}^{k-1} (1-\delta_i)\epsilon_i^{k-i}. \end{equation} We use this estimate to construct sets $A$ and $B$ which satisfy the assertion of the lemma. For a vertex $v \in V$, the {\it neighborhood} $N(v)$ is the set of vertices of $G$ that are adjacent to $v$. For $v \in V_i$ and a subset $S \subset V_j$ with $i \not = j$, let $\tilde{N}(v,S)=N(v) \cap S$ if $(i,j)$ is an edge of $H$ and $\tilde{N}(v,S)=S \setminus N(v)$ otherwise. We will try iteratively to build many induced copies of $H$. After $i$ steps, we will have vertices $v_1,\ldots,v_{i}$ with $v_j \in V_j$ for $j \leq i$ and subsets $V_{i+1,i}, V_{i+2,i},\ldots, V_{k,i}$ such that \begin{enumerate} \item $V_{\ell,i}$ is a subset of $V_{\ell}$ of size $|V_{\ell,i}| \geq \frac{n}{k}\prod_{j=1}^i\epsilon_j$ for all $i+1 \leq \ell \leq k$, \item for $1 \leq j<\ell \leq i$, $(v_j,v_{\ell})$ is an edge of $G$ if and only if $(j,\ell)$ is an edge of $H$, \item and if $j \leq i<\ell$ and $w \in V_{\ell,i}$, then $(v_j,w)$ is an edge of $G$ if and only if $(j,\ell)$ is an edge of $H$. \end{enumerate} In the first step, we call a vertex $v\in V_1$ {\it good} if $|\tilde{N}(v,V_i)| \geq \epsilon_1|V_i|$ for each $i >1$. If less than a fraction $1-\delta_1$ of the vertices in $V_1$ are good, then, by the pigeonhole principle, there is a subset $A \subset V_1$ with $|A|\geq \frac{\delta_1}{k-1}|V_1|=\frac{\delta_1}{k(k-1)}n$ and an index $j > 1$ such that $|\tilde{N}(v,V_j)| < \epsilon_1|V_j|$ for each $v \in A$. Letting $B=V_j$, one can easily check that $A$ and $B$ satisfy the assertion of the lemma. Hence, we may assume that at least a fraction $1-\delta_1$ of the vertices $v_1 \in V_1$ are good, choose any good $v_1$ and define $V_{i,1}=\tilde{N}(v_1,V_i)$ for $i>1$, completing the first step. Suppose that after step $i$ the properties 1-3 are satisfied. Then, in step $i+1$, we again call a vertex $v \in V_{i+1,i}$ {\it good} if $|\tilde{N}(v,V_{j,i})| \geq \epsilon_{i+1}|V_{j,i}|$ for each $j>i+1$. If less than a fraction $1-\delta_{i+1}$ vertices of $V_{i+1,i}$ are good, then, by the pigeonhole principle, there is a subset $A \subset V_{i+1,i}$ with $|A|\geq \frac{\delta_{i+1}}{k-i-1}|V_{i+1,i}|$ and index $j >i+1$ such that $|\tilde{N}(v,V_{j,i})|<\epsilon_{i+1}|V_{j,i}|$ for each $v \in A$. Letting, $B=V_{j,i}$, one can check using properties 1-3, that $A$ and $B$ satisfy the assertion of the lemma. Hence, we may assume that a fraction $1-\delta_{i+1}$ of the vertices $v_{i+1} \in V_{i+1,i}$ are good, choose any good $v_{i+1}$ and define $V_{j,i+1}=\tilde{N}(v_{i+1},V_{j,i})$ for $j>i+2$, completing step $i+1$. Notice that after step $i+1$, we have $|V_{j,i+1}|\geq \epsilon_{i+1}|V_{j,i}|$ for $j>i+1$, which guarantees that property 1 is satisfied. The remaining properties (2 and 3) follow from our construction of sets $V_{j,i+1}$. Thus if our process fails in one of the first $k-1$ steps we obtain desired sets $A$ and $B$. Suppose now that we successfully performed $k-1$ steps. Note that in step $i+1$, we had at least $(1-\delta_{i+1})|V_{i+1,i}| \geq \frac{n}{k}(1-\delta_{i+1})\prod_{j=1}^i \epsilon_j$ vertices to choose for vertex $v_{i+1}$. Also note that, by property 3, after step $k-1$ we can choose any vertex in the set $V_{k,k-1}$ to be $v_k$. Moreover, by the property 2, every choice of the vertices $v_1, \ldots, v_k$ form a labeled induced copy of $H$. Altogether, this gives at least \begin{eqnarray*} |V_{k,k-1}| \cdot \prod_{i=1}^{k-1}\bigg( \frac{n}{k}(1-\delta_{i})\prod_{0 \leq j<i} \epsilon_j\bigg) &\geq& \frac{n}{k}\prod_{j=1}^{k-1}\epsilon_j \, \cdot \, \prod_{i=1}^{k-1}\bigg( \frac{n}{k}(1-\delta_{i})\prod_{0 \leq j<i} \epsilon_j\bigg)\\ &=&(n/k)^k\prod_{i=1}^{k-1}(1-\delta_i)\epsilon_i^{k-i} \end{eqnarray*} labeled induced copies of $H$ for which the image of every vertex $i \in [k]$ is in $V_i$. This contradicts (\ref{upbound}) and completes the proof. \end{proof} Notice that the number of induced copies of $H$ in any induced subgraph of $G$ is at most the number of induced copies of $H$ in $G$. Let $\epsilon_i=\epsilon\leq 1/2$ and $\delta_i=\frac{1}{2}$ for $1 \leq i \leq k-1$ and let $\alpha \geq k^2/n$. Applying Lemma \ref{lemmaerdoshajnal4} with these $\epsilon_i, \delta_i$ to subsets of $G$ of size $\alpha n$ and using that $\frac{\epsilon^{i-1}}{k-i} \geq \epsilon^{k-1}, 1-\frac{k^2}{2\alpha n}\geq 1/2$ we obtain the following corollary. \begin{corollary}\label{last6} Let $H$ be a graph with $k$ vertices, $\alpha \geq k^2/n$, $\epsilon \leq 1/2$, and $G$ be a graph with at most $2^{-k}\epsilon^{k \choose 2}(\alpha n)^{k}$ induced copies of $H$. Then the pair $(G, \bar G)$ is $(\alpha,\frac{\epsilon^{k-1}}{2k},\epsilon,2)$-sparse. \end{corollary} The next statement strengthens Theorem \ref{main} by allowing for many induced copies of $H$. It follows from Corollary \ref{corbip} with $r=2, h=2\log (2/\epsilon), \rho=\frac{\epsilon^{k-1}}{2k}$, combined with the last statement in which we set $\alpha=(\rho/2)^{h-1}$. \begin{corollary}\label{maingeneralized7} There is a constant $c$ such that for each $\epsilon \in (0,1/2)$ and graph $H$ on $k$ vertices, every graph $G$ on $n$ vertices with less than $2^{-c(k \log \frac{1}{\epsilon})^2}n^k$ induced copies of $H$ contains an induced subgraph of size at least $2^{-ck (\log \frac{1}{\epsilon})^2}n$ with edge density at most $\epsilon$ or at least $1-\epsilon$. \end{corollary} This result demonstrates that for each $\epsilon \in (0,1/2)$ and graph $H$, there exist positive constants $\delta^*=\delta^*(\epsilon,H)$ and $\kappa^*=\kappa^*(\epsilon,H)$ such that every graph $G=(V,E)$ on $n$ vertices with less than $\kappa^* \,n^k$ induced copies of $H$ contains a subset $W \subset V$ of size at least $\delta^*\,n$ such that the edge density of $W$ is at most $\epsilon$ or at least $1-\epsilon$. Furthermore, there is a constant $c$ such that we can take $\delta^*(\epsilon,H)=2^{-ck (\log \frac{1}{\epsilon})^2}$ and $\kappa^*(\epsilon,H)=2^{-c(k\log \frac{1}{\epsilon})^2}$. Applying Corollary \ref{maingeneralized7} recursively one can obtain an equitable partition of $G$ into a small number of subsets each with low or high density. \begin{theorem}\label{weakversion} For each $\epsilon \in (0,1/2)$ and graph $H$ on $k$ vertices, there are positive constants $\kappa=\kappa(\epsilon,H)$ and $C=C(\epsilon,H)$ such that every graph $G=(V,E)$ on $n$ vertices with less than $\kappa\, n^{k}$ induced copies of $H$, there is an equitable partition $V=\bigcup_{i=1}^{\ell} V_i$ such that $\ell \leq C$ and the edge density in each $V_i$ is at most $\epsilon$ or at least $1-\epsilon$. \end{theorem} \noindent This extension of R\"odl's theorem was proved by Nikiforov \cite{Ni} using the regularity lemma and therefore it had quite poor (tower like) dependence of $\kappa$ and $C$ on $\epsilon$ and $k$. Obtaining a proof without using the regularity lemma was the main open problem raised in \cite{Ni} . \vspace{0.2cm} \noindent {\bf Proof of Theorem \ref{weakversion}.}\, Let $\kappa(\epsilon,H)=(\frac{\epsilon}{4})^k\kappa^*(\frac{\epsilon}{4},H)$ and $C(\epsilon,H)=4/(\epsilon \delta^*(\frac{\epsilon}{4},H))$, where $\kappa^*$ and $\delta^*$ were defined above. Take a subset $W_1 \subset V$ of size $ \delta^*(\frac{\epsilon}{4},H)\frac{\epsilon}{4}n$ whose edge density is at most $\frac{\epsilon}{4}$ or at least $1-\frac{\epsilon}{4}$, and set $U_1 =V\setminus W_1$. For $j \geq 1$, if $|U_j| \geq \frac{\epsilon}{4} n$, then by definition of $\kappa$ we have that the number of induced copies of $H$ in $U_j$ is at most (the number of such copies in $G$) $\kappa\,n^k= (\frac{\epsilon}{4})^k\kappa^*\,n^k \leq \kappa^*|U_j|^k$. Therefore by definition of $\kappa^*$ and $\delta^*$ we can find a subset $W_{j+1} \subset U_j$ of size $\delta^*\frac{\epsilon}{4}n \leq \delta^*|U_j|$ whose edge density is at most $\frac{\epsilon}{4}$ or at least $1-\frac{\epsilon}{4}$, and set $U_{j+1}=U_j \setminus W_{j+1}$. Once this process stops, we have disjoint sets $W_1,\ldots,W_{\ell}$, each with the same cardinality, and a subset $U_{\ell}$ of cardinality at most $\frac{\epsilon}{4}n$. The number $\ell$ is at most $$n/|W_1| \leq 4/(\epsilon \delta^*(\frac{\epsilon}{4},H)).$$ Partition set $U_\ell$ into $\ell$ equal parts $T_1, \ldots, T_\ell$ and let $V_j=W_j \cup T_j$ for $1 \leq j \leq \ell$. Notice that $V=V_1 \cup \ldots \cup V_{\ell}$ is an equitable partition of $V$. By definition, $|T_j|=|U_\ell|/\ell \leq \frac{\epsilon}{4}n/\ell$. On the other hand $|W_j|=(n-|U_\ell|)/\ell \geq (1-\epsilon/4)n/\ell$. Since $1-\epsilon/4 >7/8$, this implies that $$|T_j|\leq \frac{\epsilon}{4}n/\ell \leq \frac{\epsilon}{4}\big(1-\epsilon/4\big)^{-1}|W_j| \leq \frac{2\epsilon}{7}|W_j|.$$ We next look at the edge density in $V_j$. If the edge density in $W_j$ is at most $\epsilon/4$, then using the above bound on $|T_j|$, it is easy to check that the number of edges in $V_j$ is at most $${|T_j| \choose 2}+|T_j||W_j|+\frac{\epsilon}{4}{|W_j| \choose 2} \leq \epsilon{|W_j| \choose 2} \leq \epsilon{|V_j| \choose 2}.$$ Hence, the edge density in each such $V_j$ is at most $\epsilon$. Similarly, if the edge density in $W_j$ is at least $1-\frac{\epsilon}{4}$, then the edge density in $V_j$ is at least $1-\epsilon$. This completes the proof. \hfill $\Box$ \vspace{0.2cm} We next use Lemma \ref{lemmaerdoshajnal4} to prove that there is a constant $c>0$ such that every graph $G$ on $n$ vertices which contains at most $(1-\epsilon)2^{-{k \choose 2}}n^k$ labeled induced copies of some fixed $k$-vertex graph $H$ has a subset $S$ of size $|S|=\lfloor n/2 \rfloor$ with $|e(S)-\frac{n^2}{16}| \geq \epsilon c^{-k} n^2$. \vspace{0.25cm} \noindent {\bf Proof of Theorem \ref{dev}.}\, For $1 \leq i \leq k-1$, let $\epsilon_i=\frac{1}{2}(1-2^{i-k-2}\epsilon)$ and $\delta_i=2^{i-k-2}\epsilon$. Notice that for all $i \leq k-1$ \begin{eqnarray} \label{eq3} \prod_{j<i}\epsilon_j=2^{-i+1}\prod_{j<i}\big(1-2^{j-k-2}\epsilon\big)\geq 2^{-i+1}\bigg(1-\epsilon \sum_{j<k-1}2^{j-k-2}\bigg) \geq 2^{-i+1}(1-\epsilon/8)>2^{-i} \end{eqnarray} and also that \begin{eqnarray*} \prod_{i=1}^{k-1}(1-\delta_i)\epsilon_i^{k-i}&=& 2^{-{k \choose 2}}\prod_{i=1}^{k-1}\big(1-2^{i-k-2}\epsilon \big)^{k-i+1}= 2^{-{k \choose 2}}\prod_{j=2}^{k} \big(1-\epsilon2^{-j-1}\big)^{j}\\ &\geq& 2^{-{k \choose 2}}\bigg(1-\epsilon\sum_{j=2}^k \frac{j}{2^{j+1}}\bigg) > \left(1-\frac{\epsilon}{2}\right)2^{-{k \choose 2}}. \end{eqnarray*} We may assume that $\epsilon \geq k^2/n$ since otherwise by choosing constant $c$ large enough we get that $\epsilon c^{-k}n^2<1$ and the conclusion of the theorem follows easily. Therefore $$ \left(1-\frac{k^2}{2n}\right)\prod_{i=1}^{k-1} (1-\delta_i)\epsilon_i^{k-i} \geq \left(1-\frac{\epsilon}{2}\right)^22^{-{k \choose 2}} > (1-\epsilon)2^{-{k \choose 2}},$$ and we can apply Lemma \ref{lemmaerdoshajnal4} with $\epsilon_i$ and $\delta_i$ as above to our graph $G$ since it contains at most $(1-\epsilon)2^{-{k \choose 2}}n^k$ labeled induced copies of $H$. This lemma, together with (\ref{eq3}), implies that there is an index $i \leq k-1$ and disjoint subsets $A$ and $B$ with $$|A|\geq \frac{\delta_in}{k(k-i)}\prod_{j<i} \epsilon_j \geq k^{-2}2^{-k-2}n,$$ $$|B| \geq \frac{n}{k}\prod_{j < i} \epsilon_j \geq 2^{-i}k^{-1}n,$$ and every element of $A$ is adjacent to at most $\epsilon_i|B|$ elements of $B$ or every element of $A$ is adjacent to at least $(1-\epsilon_i)|B|$ elements of $B$. In either case, we have $$\left|e(A,B)-\frac{1}{2}|A||B|\right| \geq \left(\frac{1}{2}-\epsilon_i\right)|A||B|=2^{i-k-3}\epsilon|A||B| \geq k^{-3}2^{-2k-5}\epsilon n^2.$$ Note that $$e(A,B)-\frac{1}{2}|A||B|=\left(e(A \cup B)-\frac{1}{2}{|A \cup B| \choose 2}\right)-\left(e(A)-\frac{1}{2}{|A| \choose 2}\right)-\left(e(B)-\frac{1}{2}{|B| \choose 2}\right).$$ It follows from the triangle inequality that there is some subset of vertices $R \in \{A,B, A \cup B\}$ such that \begin{equation} \label{subset} \left|e(R)-\frac{1}{2}{|R| \choose 2}\right| \geq \frac{1}{3} k^{-3}2^{-2k-5}\epsilon n^2, \end{equation} i.e., it deviates by at least $\epsilon k^{-3}2^{-2k-5}n^2/3$ edges from having edge density $1/2$. To finish the proof we will use the lemma of Erd\H{o}s et al. \cite{ErGoPaSp}, mentioned in the introduction. This lemma says that if graph $G$ on $n$ vertices with edge density $\eta$ has a subset that deviates by $D$ edges from having edge density $\eta$, then it also has a subset of size $n/2$ that deviates by at least $D/5$ edges from having edge density $\eta$. Note that if the edge density of our graph $G$ is either larger than $1/2+\epsilon k^{-3}2^{-2k-5}n^2/30$ or smaller than $1/2-\epsilon k^{-3}2^{-2k-5}n^2/30$ than by averaging over all subsets of size $n/2$ we will find subset $S$ satisfying our assertion. Otherwise, if the edge density $\eta$ of $G$ satisfies $|\eta-1/2| \leq \epsilon k^{-3}2^{-2k-5}n^2/30$, then the subset $R$ from (\ref{subset}) deviates by at least $\epsilon k^{-3}2^{-2k-5}n^2/3- \epsilon k^{-3}2^{-2k-5}n^2/30 \geq \epsilon k^{-3}2^{-2k}n^2/4$ edges from having edge density $\eta$. Then, by the lemma of Erd\H{o}s et al., $G$ has a subset $S$ of cardinality $n/2$ that deviates by at least $\epsilon k^{-3}2^{-2k-5}n^2/20$ edges from having edge density $\eta$. This $S$ satisfies $$\left|e(S)-\frac{1}{4}|S|^2\right| \geq \epsilon k^{-3}2^{-2k-5}n^2/20-\epsilon k^{-3}2^{-2k-5}n^2/30= \Omega\left(\epsilon k^{-3}2^{-2k}n^2\right),$$ completing the proof. \ifvmode\mbox{ }\else\unskip\fi\hskip 1em plus 10fill$\Box$ \vspace{0.1cm} For positive integers $k$ and $n$, recall that $D(k,n)$ denotes the largest integer such that every graph $G$ on $n$ vertices that is $H$-free for some $k$-vertex graph $H$ contains a subset $S$ of size $n/2$ with $|e(S)-\frac{1}{16}n^2|>D(k,n)$. We end this section by proving the upper bound on $D(k,n)$. \begin{proposition} There is a constant $c>0$ such that for all positive integers $k$ and $n \geq 2^{k/2}$, there is a $K_k$-free graph $G$ on $n$ vertices such that for every subset $S$ of $n/2$ vertices of $G$, $$\Big|e(S)-\frac{1}{16}n^2\Big|<c2^{-k/4}n^2.$$ \end{proposition} \begin{proof} Consider the random graph $G(\ell,1/2)$ with $\ell=2^{ k /2}$. For every subset of vertices $X$ in this graph the number of edges in $X$ is a binomially distributed random variable with expectation $\frac{|X|(|X|-1)}{4}$. Therefore by Chernoff's bound (see, e.g., Appendix A in \cite{AlSp}), the probability that it deviates from this value by $t$ is at most $2e^{-t^2/|X|^2}$. Thus choosing $t=1.5\ell^{3/2}$ we obtain that the probability that there is a subset of vertices $X$ such that $\big|e(X)-\frac{|X|(|X|-1)}{4}\big|>t$ is at most $2^\ell \cdot 2e^{-t^2/\ell^2} \ll 1$. This implies that there is graph $\Gamma$ on $\ell$ vertices such that every subset $X$ of $\Gamma$ satisfies \begin{equation} \label{random} \Big|e(X)-\frac{1}{4}|X|^2\Big| \leq 2\ell^{3/2}. \end{equation} Let $G$ be the graph obtained by replacing every vertex $u$ of $\Gamma$ with an independent set $I_u$, of size $n/\ell$, and by replacing every edge $(u,v)$ of $\Gamma$ with a complete bipartite graph, whose partition classes are independent sets $I_u$ and $I_v$. Clearly, since $\Gamma$ does not contain $K_k$, then neither does $G$. We claim that graph $G$ satisfies the assertion of the proposition. Suppose for contradiction that there is a subset $S$ of $n/2$ vertices of $G$ satisfying $$e(S)-\frac{1}{16}n^2 >4\ell^{3/2}(n/\ell)^2=4\ell^{-1/2}n^2=2^{-k/4+2}n^2,$$ (the other case when $e(S)-n^2/16<-4\ell^{-1/2}n^2$ can be treated similarly). For every vertex $u \in \Gamma$ let the size of $S \cap I_u$ be $a_un/\ell$. By definition, $0 \leq a_u \leq 1$ and since $S$ has size $n/2$ we have that $\sum_u a_u=\ell/2$. We also have that $$e(S)=\sum_{(u,v)\in E(\Gamma)} a_u a_v \cdot (n/\ell)^2>\frac{1}{16}n^2+4\ell^{3/2}(n/\ell)^2,$$ and therefore $$\sum_{(u,v)\in E(\Gamma)} a_u a_v >\ell^2/16+4\ell^{3/2}=\frac{1}{4}\Big(\sum_u a_u\Big)^2 +4\ell^{3/2}.$$ Consider a random subset $Y$ of $\Gamma$ obtained by choosing every vertex $u$ randomly and independently with probability $a_u$. Since all choices were independent we have that $$\mathbb{E}\big[|Y|^2\big]= \sum_u a_u +\sum_{u\not =v}a_ua_v \leq \big(\sum_u a_u\big)^2+\ell/2.$$ We also have that the expected number of edges spanned by $Y$ is $\mathbb{E}\big[e(Y)\big]=\sum_{(u,v)\in E(\Gamma)} a_u a_v$. Then, by the above discussion, $\mathbb{E}\big[e(Y)-|Y|^2/4\big] >3\ell^{3/2}$. In particular, there is subset $Y$ of $\Gamma$ with this property, which contradicts (\ref{random}). This shows that every subset $S$ of $n/2$ vertices of $G$ satisfies $$\Big|e(S)-\frac{1}{16}n^2\Big| \leq 2^{-k/4+2}n^2$$ and completes the proof. \end{proof} \section{Induced Ramsey Numbers and Pseudorandom Graphs} \label{moreoninduced} The main result in this section is Theorem \ref{quasirandominduced}, which shows that any sufficiently pseudo-random graph of appropriate density has strong induced Ramsey properties. It generalizes Theorem \ref{quasicor1} and Corollary \ref{payley} from the introduction. Combined with known examples of pseudo-random graphs, this theorem gives various explicit constructions which match and improve the best known estimates for induced Ramsey numbers. The idea of the proof of Theorem \ref{quasirandominduced} is rather simple. We have a sufficiently large, pseudo-random graph $G$ that is not too sparse or dense. We also have $d$-degenerate graphs $H_1$ and $H_2$ each with vertex set $[k]$ and chromatic number at most $q$. We suppose for contradiction that there is a red-blue edge-coloring of $G$ without an induced red copy of $H_1$ and without an induced blue copy of $H_2$. We may view the red-blue coloring of $G$ as a red-blue-green edge-coloring of the complete graph $K_{|G|}$, in which the edges of $G$ have their original color, and the edges of the complement $\bar G$ are colored green. The fact that in $G$ there is no induced red copy of $H_1$ means that the red-blue-green coloring of $K_{|G|}$ does not contain a particular red-green coloring of the the complete graph $K_k$. Then we prove, similar to Lemma \ref{lemmaerdoshajnal} of Erd\H{o}s and Hajnal, that any large subset of vertices of $G$ contains two large disjoint subsets for which the edge density in color red between them is small. By using the key lemma from Section 3, we find $k$ large disjoint vertex subsets $V_1,\ldots,V_k$ of $G$ for which the edge density in color red is small between any pair $(V_i,V_j)$ for which $(i,j)$ an edge of $H_2$. Next we try to find an induced blue copy of $H_2$ with vertex $i$ in $V_i$ for all $i \in [k]$. Since the edge density between $V_i$ and $V_j$ in color red is sufficiently small for every edge $(i,j)$ of $H_2$, we can build an induced blue copy of $H_2$ one vertex at a time. At each step of this process we use pseudo-randomness of $G$ to make sure that the existing possible subsets for not yet embedded vertices of $H_2$ are sufficiently large and that the density of red edges does not increase a lot between any pair of subsets corresponding to adjacent vertices of $H_2$. This last part of the proof, embedding an induced blue copy of $H_2$, is the most technically involved and handled by Lemma \ref{densitylemma}. Recall that $[i]=\{1, \ldots, i\}$ and that a graph is $d$-degenerate if every subgraph has a vertex of degree at most $d$. For an edge-coloring $\Psi:E(K_{k})\rightarrow [r]$, we say that another edge-coloring $\Phi:E(K_{n})\rightarrow [s]$ is {\it $\Psi$-free} if, for every subset $W$ of size $k$ of the complete graph $K_n$, the restriction of $\Phi$ to $W$ is not isomorphic to $\Psi$. In the following lemma, we have a coloring $\Psi$ of the edges of the complete graph $K_k$ with colors $1$ and $2$ such that the graph of color $2$ is $d$-degenerate. We also have a $\Psi$-free coloring $\Phi$ of the edges of the complete graph $K_n$ such that between any two large subsets of vertices there are sufficiently many edges of color 1. With these assumptions, we show that there are two large subsets of $K_n$ which in coloring $\Phi$ have few edges of color $2$ between them. A graph $G$ is {\it bi-$(\epsilon,\delta)$-dense} if $d(A,B)>\epsilon$ holds for all disjoint subsets $A,B \subset V(G)$ with $|A|,|B| \geq \delta |V(G)|$. \begin{lemma}\label{lemmaerdoshajnal2} Let $d$ and $k$ be positive integers and $\Psi:E(K_k) \rightarrow [2]$ be a $2$-coloring of the edges of $K_k$ such that the graph of color $2$ is $d$-degenerate. Suppose that $q,\epsilon \in (0,1)$ and $\Phi:E(K_n) \rightarrow [s]$ is a $\Psi$-free edge-coloring such that the graph of color $1$ is bi-$(q,\epsilon^dq^{k}k^{-2})$-dense. Then there are disjoint subsets $A$ and $B$ of $K_n$ with $|A|,|B| \geq \epsilon^{d}q^{k} k^{-2}n$ such that every vertex of $A$ is connected to at most $\epsilon|B|$ vertices in $B$ by edges of color 2. \end{lemma} \begin{proof} Note that from definition, the vertices of every $d$-degenerate graph can be labeled $1,2, \ldots$ such that for every vertex $\ell$ the number of vertices $j<\ell$ adjacent to it is at most $d$. (Indeed, remove from the graph a vertex of minimum degree, place it in the end of the list and repeat this process in the remaining subgraph.) Therefore we may assume that the labeling $1, \ldots, k$ of vertices of $K_k$ has the property that for every $\ell \in [k]$ there are at most $d$ vertices $j <\ell$ such that the color $\Psi(j,\ell)=2$. Partition the vertices of $K_n$ into sets $V_1 \cup \ldots \cup V_{k}$ each of size $\frac{n}{k}$. For $w \in V_i$ and a subset $S \subset V_j$ with $j \not = i$, let $N(w,S)=\{s \in S ~|~ \Phi(w,s)=\Psi(i,j)\}.$ For $i<\ell $, let $D(\ell,i)$ denote the number of vertices $j \leq i$ such that the color $\Psi(j,\ell)=2$. By the above assumption, $D(\ell,i) \leq d$ for $1 \leq i < \ell \leq k$. We will try iteratively to build a copy of $K_{k}$ with coloring $\Psi$. After $i$ steps, we either find two disjoint subsets of vertices $A, B$ which satisfy the assertion of the lemma or we will have vertices $v_1,\ldots,v_{i}$ and subsets $V_{i+1,i},V_{i+2,i},\ldots,V_{k,i}$ such that \begin{enumerate} \item $V_{\ell,i}$ is a subset of $V_\ell$ of size $|V_{\ell,i}| \geq \epsilon^{D(\ell,i)}q^{i-D(\ell,i)}|V_{\ell}|$ for all $i+1 \leq \ell \leq k$, \item $\Phi(v_j,v_{\ell})=\Psi(j,\ell)$ for $1 \leq j<\ell \leq i$, \item and if $j \leq i< \ell$ and $w \in V_{\ell,i}$, then $\Phi(v_j,w)=\Psi(j,\ell)$. \end{enumerate} In the first step, we call a vertex $w\in V_1$ {\it good} if $|N(w,V_j)| \geq \epsilon|V_j|$ for all $j>1$ with $\Psi(1,j) =2$ and $|N(w,V_j)| \geq q|V_i|$ for all $j>1$ with $\Psi(1,j)=1$. If there is no good vertex in $V_1$, then there is a subset $A \subset V_1$ with $|A|\geq \frac{1}{k-1}|V_1|$ and index $j >1$ such that either $\Psi(1,j)=1$ and every vertex $w \in A$ has fewer than $q|V_j|$ edges of color $1$ to $V_j$ or $\Psi(1,j)=2$ and every vertex $w \in A$ is connected to less than $\epsilon|V_j|$ vertices in $V_j$ by edges of color 2. Letting $B=V_j$, we conclude that the first case is impossible since the graph of color $1$ is bi-$(q,\epsilon^dq^{k}k^{-2})$-dense, while in the second case we would be done, since $A$ and $B$ would satisfy the assertion of the lemma. Therefore, we may assume that there is a good vertex $v_1 \in V_1$, and we define $V_{i,1}=N(v_1,V_i)$ for $i>1$. Suppose that after step $i$ the properties 1-3 are still satisfied. Then, in step $i+1$, a vertex $w \in V_{i+1,i}$ is called {\it good} if $|N(w,V_{j,i})| \geq \epsilon|V_{j,i}|$ for each $j> i+1$ with $\Psi(i+1,j) =2$ and $|N(w,V_{j,i})| \geq q|V_{j,i}|$ for each $j >i+1$ with $\Psi(i+1,j) =1$. If there is no good vertex in $V_{i+1,i}$, then there is a subset $A \subset V_{i+1,i}$ with $|A|\geq \frac{1}{k-i-1}|V_{i+1,i}|$ and $j >i+1$ such that either $\Psi(i+1,j)=1$ and every vertex $w \in A$ has fewer than $q|V_{j,i}|$ edges of color $1$ to $V_{j,i}$ or $\Psi(1,j)=2$ and every vertex $w \in A$ is connected to less than $\epsilon|V_{j,i}|$ vertices in $V_{j,i}$ by edges of color 2. Note that even in the last step when $i+1=k$ the size of $A$ is still at least $|V_{k,k-1}|/k\geq \epsilon^dq^k|V_k|/k\geq \epsilon^dq^kk^{-2}n$. Therefore, letting $B=V_{j,i}$, we conclude that as before the first case is impossible since the graph of color $1$ is bi-$(q,\epsilon^dq^{k}k^{-2})$-dense, while the second case would complete the proof, since $A$ and $B$ would satisfy the assertion of the lemma. Hence, we may assume that there is a good vertex $v_{i+1} \in V_{i+1,i}$, and we define $V_{j,i+1}=N(v_{i+1},V_{j,i})$ for $j >i+1$. Note that $|V_{j,i+1}| \geq q|V_{j,i}|$ if $\Psi(i+1,j) =1$ and $|V_{j,i+1}| \geq \epsilon |V_{j,i}|$ if $\Psi(i+1,j) =2$. This implies that after step $i+1$ we have that $|V_{\ell,i+1}| \geq \epsilon^{D(\ell,i+1)}q^{i+1-D(\ell,i+1)}|V_{\ell}|$ for all $i+2 \leq \ell \leq k$. The iterative process must stop at one of the steps $j \leq k-1$, since otherwise the coloring $\Phi$ would not be $\Psi$-free. As we already explained above, when this happens we have two disjoint subsets $A$ and $B$ that satisfy the assertion of the lemma. \end{proof} Notice that if coloring $\Phi:K_n \rightarrow [s]$ is $\Psi$-free, then so is $\Phi$ restricted to any subset of $K_n$ of size $\alpha n$. Therefore, Lemma \ref{lemmaerdoshajnal2} has the following corollary. \begin{corollary}\label{last} Let $d$ and $k$ be positive integers and $\Psi:E(K_k) \rightarrow [2]$ be a $2$-coloring of the edges of $K_k$ such that the graph of color $2$ is $d$-degenerate. If $q,\alpha,\epsilon \in (0,1)$ and $\Phi:E(K_n) \rightarrow [s]$ is a $\Psi$-free edge-coloring such that the graph of color $1$ is bi-$(q,\alpha\rho)$-dense with $\rho=\epsilon^{d}q^{k} k^{-2}$, then the graph of color $2$ is $(\alpha,\rho,\epsilon,2)$-sparse. \end{corollary} \noindent The next statement follows immediately from Corollary \ref{last} (with $\epsilon/4$ instead of $\epsilon $) and Corollary \ref{secondcor}. \begin{corollary}\label{last2} Let $d$, $k$, and $h$ be positive integers and $\Psi:E(K_k) \rightarrow [2]$ be a $2$-coloring of the edges of $K_k$ such that the graph of color $2$ is $d$-degenerate. Suppose that $q,\alpha,\epsilon \in (0,1)$ and $\Phi:E(K_n) \rightarrow [s]$ is a $\Psi$-free edge-coloring such that the graph of color $1$ is bi-$(q,\alpha\rho)$-dense with $\rho=(\epsilon/4)^dq^{k}k^{-2}$. Then the graph of color $2$ is $((\frac{2}{\rho})^{h-1}\alpha,2^{1-h}\rho^h,\epsilon,2^h)$-sparse. \end{corollary} \noindent Pending one additional lemma, we are now ready to prove the main result of this section, showing that pseudo-random graphs have strong induced Ramsey properties. \begin{theorem}\label{quasirandominduced} Let $\chi \geq 2$ and $G$ be a $(p,\lambda)$-pseudo-random graph with $0 <p \leq 3/4$ and $\lambda \leq ((\frac{p}{10k})^d2^{-pk})^{20\log \chi}n$. Then every $d$-degenerate graph on $k$ vertices with chromatic number at most $\chi$ occurs as an induced monochromatic copy in every $2$-coloring of the edges of $G$. Moreover, all of these induced monochromatic copies can be found in the same color. \end{theorem} Taking $p=1/k$, $n=k^{cd\log \chi}$ and constant $c$ sufficiently large so that $((\frac{p}{10k})^d2^{-pk})^{20\log \chi}>n^{-0.1}$ one can easily see that this result implies Theorem \ref{quasicor1}. To obtain Corollary \ref{payley}, recall that for a prime power $n$, the Paley graph $P_n$ has vertex set $\mathbb{F}_n$ and distinct vertices $x,y \in \mathbb{F}_n$ are adjacent if $x-y$ is a square. This graph is $(1/2,\lambda)$-pseudo-random with $\lambda=\sqrt{n}$ (see e.g., \cite{KrSu}). Therefore, for sufficiently large constant $c$, the above theorem with $n=2^{ck\log^2 k}$, $p=1/2$ and $d=\chi=k$ implies that every graph on $k$ vertices occurs as an induced monochromatic copy in all $2$-edge-colorings of the Paley graph. Similarly, one can prove that there is a constant $c$ such that, with high probability, the random graph $G(n,1/2)$ with $n \geq 2^{ck\log^2 k}$ satisfies that every graph on $k$ vertices occurs as an induced monochromatic copy in all $2$-edge-colorings of $G$. \vspace{0.1cm} \noindent {\bf Proof of Theorem \ref{quasirandominduced}.}\, Suppose for contradiction that there is an edge-coloring $\Phi_0$ of $G$ with colors red and blue, and $d$-degenerate graphs $H_1$ and $H_2$ each having $k$ vertices and chromatic number at most $\chi$ such that there is no induced red copy of $H_1$ and no induced blue copy of $H_2$. Since $H_1, H_2$ are $d$-degenerate graphs on $k$ vertices we may suppose that their vertex set is $[k]$ and every vertex $i$ has at most $d$ neighbors less than $i$ in both $H_1$ and $H_2$. Consider the red-blue-green edge-coloring $\Phi$ of the complete graph $K_n$, in which the edges of $G$ have their original coloring $\Phi_0$, and the edges of the complement $\bar G$ are colored green. Let $\Psi$ be the edge-coloring of the complete graph $K_k$ where the red edges form a copy of $H_1$ and the remaining edges are green. By assumption, the coloring $\Phi$ is $\Psi$-free. Since $G$ is $(p,\lambda)$-pseudo-random, we have that the density of edges in $\bar G$ between any two disjoint sets $A, B$ of size at least $6p^{-1} \lambda$ is at least $$d_{\bar G}(A,B)=1-d_G(A,B)\geq 1-\Big(p+\frac{\lambda}{\sqrt{|A||B|}}\Big) \geq 1-\frac{7}{6}p.$$ Therefore the green graph in coloring $\Phi$ is bi-$(q,6p^{-1}\frac{\lambda}{n})$-dense for $q=1-7p/6$. Let $\epsilon= \frac{p}{1000k^6}$, $\rho=(\epsilon/4)^dq^{k}k^{-2}$, $h=\log \chi $, and $\alpha=(\rho/2)^{h-1} $. Using that $q=1-7p/6$ and $\lambda/n \leq ((\frac{p}{10k})^d2^{-pk})^{20\log \chi}$ it is straightforward to check that $6p^{-1}\frac{\lambda}{n} \leq 2^{1-h}\rho^{h}=\alpha \rho$. By Corollary \ref{last2} and Definition \ref{d31}, there are $2^h=\chi$ subsets $W_1,\ldots,W_{\chi}$ of $K_n$ with $|W_1|=\ldots=|W_\chi| \geq 2^{1-h}\rho^{h}n$, such that the sum of densities of red edges between all pairs $W_i$ and $W_j$ is at most ${\chi \choose 2}\epsilon$. Hence, the density between $W_i$ and $W_j$ is also at most $\chi^2\epsilon$ for all $1 \leq i < j \leq \chi$. Partition every set $W_i$ into $k$ subsets each of size $|W_i|/k \geq \frac{1}{k}2^{1-h}\rho^{h}n$. Since the chromatic number of $H_2$ is at most $\chi$ and it has $k$ vertices, we can choose for every vertex $i$ of $H_2$ one of these subsets, which we call $V_i$, such that all subsets corresponding to vertices of $H_2$ in the same color class (of a proper $\chi$-coloring) come from the same set $W_\ell$. In particular, for every edge $(i,j)$ of $H_2$, the corresponding sets $V_i$ and $V_j$ lie in two different sets $\{W_\ell\}$. Since the size of $V_i$'s is by a factor $k$ smaller than the size of $W_\ell$'s the density of red edges between $V_i$ and $V_j$ corresponding to an edge in $H_2$ is at most $k^2\chi^2\epsilon \leq \frac{p}{1000k^2}$ (note that it can increase by a factor at most $k^2$ compare to density between sets $\{W_\ell\}$). Notice that the subgraph $G' \subset G$ induced by $V_1 \cup \ldots \cup V_{k}$ has $n' \geq 2^{1-h}\rho^{h}n$ vertices and is also $(p,\lambda)$-pseudo-random. By the definitions of $\rho$ and $h$, and our assumption on $\lambda$, we have that $$\lambda/n'\leq 2^{h-1}\rho^{-h}\lambda/n \leq 2^{h-1}\rho^{-h} \left(\Big(\frac{p}{10k}\Big)^d2^{-pk}\right)^{20\log \chi} \leq \left(\Big(\frac{p}{10k}\Big)^d2^{-pk}\right)^{10\log \chi}.$$ Applying Lemma \ref{densitylemma} below with $H=H_2$ to the coloring $\Phi_0$ of graph $G'$ with partition $V_1 \cup \ldots \cup V_{k}$, we find an induced blue copy of $H_2$, completing the proof. \hfill $\Box$ \begin{lemma}\label{densitylemma} Let $H$ be a $d$-degenerate graph with vertex set $[k]$ such that each vertex $i$ has at most $d$ neighbors less than $i$. Let $G=(V,E)$ be a $(p,\lambda)$-pseudo-random graph on $n$ vertices with $0<p \leq 3/4$, $\lambda \leq ((\frac{p}{10k})^d2^{-pk})^{10}n$ and let $V=V_1 \cup \ldots \cup V_k$ be a partition of its vertices such that each $V_i$ has size $n/k$. Suppose that the edges of $G$ are $2$-colored, red and blue, such that for every edge $(j,\ell)$ of $H$, the density of red edges between the pair $(V_j,V_{\ell})$ is at most $\beta = \frac{p}{1000k^2}$. Then there is an induced blue copy of $H$ in $G$ for which the image of every vertex $i \in [k]$ lies in $V_i$. \end{lemma} \begin{proof} For $i<j$, let $D(i,j)$ denote the number of neighbors of $j$ that are at most $i$. Let $\epsilon_1=\frac{1}{k}$, $\epsilon_2=\frac{p}{10k}$, and $\delta=(1-p)^kp^d$. Since $p \leq 3/4$, notice that $\delta \geq 2^{-3pk}p^d$ and \begin{eqnarray} \label{eq2} \lambda \leq \left(\Big(\frac{p}{10k}\Big)^d2^{-pk}\right)^{10}n \leq \frac{p^8}{(10k)^{10}}\delta^2n. \end{eqnarray} We construct an induced blue copy of $H$ one vertex at a time. At the end of step $i$, we will have vertices $v_1,\ldots,v_i$ and subsets $V_{j,i} \subset V_j$ for $j >i $ such that the following four conditions hold \begin{enumerate} \item for $j, \ell \leq i$, if $(j,{\ell})$ is an edge of $H$, then $(v_j,v_{\ell})$ is a blue edge of $G$, otherwise $v_j$ and $v_{\ell}$ are not adjacent in $G$, \item for $ j \leq i < \ell$, if $(j,{\ell})$ is an edge of $H$, then $v_j$ is adjacent to all vertices in $V_{\ell,i}$ by blue edges, otherwise there are no edges of $G$ from $v_j$ to $V_{\ell,i}$, \item for $i < j$, we have $|V_{j,i}| \geq (1-p-\epsilon_2)^{i-D(i,j)}(p-\epsilon_2)^{D(i,j)}|V_j|$, \item and for $j, \ell>i$ if $(j,\ell)$ is an edge of $H$, then the density of red edges between $V_{j,i}$ and $V_{\ell,i}$ is at most $(1+\epsilon_1)^i \beta$. \end{enumerate} Clearly, in the end of the first $k$ steps of this process we obtain a required copy of $H$. For $i=0$ and $j \in [k]$, define $V_{j,0}=V_j$. Notice that the above four properties are satisfied for $i=0$ (the first two properties being vacuously satisfied). We now assume that the above four properties are satisfied at the end of step $i$, and show how to complete step $i+1$ by finding a vertex $v_{i+1} \in V_{i+1,i}$ and subsets $V_{j,i+1} \subset V_{j,i}$ for $j>i+1$ such that the conditions 1-4 still hold. We need to introduce some notation. For a vertex $w \in V_j$ and a subset $S \subset V_{\ell}$ with $j \not =\ell$, let \begin{itemize} \item $N(w,S)$ denote the set of vertices $s \in S$ such that $(s,w)$ is an edge of $G$, \item $B(w,S)$ denote the set of vertices $s \in S$ such that $(s,w)$ is a blue edge of $G$, \item $R(w,S)$ denote the set of vertices $s \in S$ such that $(s,w)$ is a red edge of $G$, \item $\tilde{N}(w,S)=N(w,S)$ if $(j, \ell)$ is an edge of $H$ and $\tilde{N}(w,S)=S \setminus N(w,S)$ otherwise, \item $\tilde{B}(w,S)=B(w,S)$ if $(j, \ell)$ is an edge of $H$ and $\tilde{B}(w,S):=S \setminus N(w,S)$ otherwise, and \item $p_{j,\ell}=p$ if $(j,\ell)$ is an edge of $H$ and $p_{j,\ell}=1-p$ if $(j,\ell)$ is not an edge of $H$. \end{itemize} Note that since graph $G$ is pseudo-random with edges density $p$, by the above definitions, for every large subset $S \subset V_{\ell}$ and for most vertices $w \in V_j$ we expect the size of $\tilde{N}(w,S)$ to be roughly $p_{j,\ell}|S|$. We also have for all $S \subset V_{\ell}$ and $w \in V_j$ that $\tilde{B}(w,S)=\tilde{N}(w,S) \setminus R(w,S)$. Call a vertex $w \in V_{i+1,i}$ {\it good} if for all $j >i+1$, $\tilde{B}(w,V_{j,i})\geq (p_{i+1,j}-\epsilon_2)|V_{j,i}|$ and for every edge $(j,\ell)$ of $H$ with $j, \ell>i+1$, the density of red edges between $\tilde{B}(w,V_{j,i})$ and $\tilde{B}(w,V_{\ell,i})$ is at most $(1+\epsilon_1)^{i+1}\beta$. If we find a good vertex $w \in V_{i+1,i}$, then we simply let $v_{i+1}=w$ and $V_{j,i+1}=\tilde{B}(w,V_{j,i})$ for $j > i+1$, completing step $i+1$. It therefore suffices to show that there is a good vertex in $V_{i+1,i}$. We first throw out some vertices of $V_{i+1,i}$ ensuring that the remaining vertices satisfy the first of the two properties of good vertices. For $j >i+1$ and an edge $(i+1,j)$ of $H$, let $R_{j}$ consist of those $w \in V_{i+1,i}$ for which the number of red edges $(w,w_j)$ with $w_j \in V_{j,i}$ is at least $\frac{\epsilon_2}{2}|V_{j,i}|$. Since the density of red between $V_{i+1,i}$ and $V_{j,i}$ is at most $(1+\epsilon_1)^i\beta$, then $R_{j}$ contains at most $$|R_{j}| \leq \frac{(1+\epsilon_1)^i\beta|V_{i+1,i}||V_{j,i}|}{\frac{\epsilon_2}{2}|V_{j,i}|}=2(1+\epsilon_1)^i\epsilon_2^{-1}\beta|V_{i+1,i}|$$ vertices. Let $V'$ be the set of vertices in $V_{i+1,i}$ that are not in any of the $R_j$. Using that $\epsilon_1=1/k, \epsilon_2=\frac{p}{10k}$ and $ \beta=\frac{p}{1000k^2}$ we obtain \begin{eqnarray*} |V'| &\geq& |V_{i+1,i}|-\sum_{j>i+1}|R_j| \geq |V_{i+1,i}|-k\Big(2(1+\epsilon_1)^i\epsilon_2^{-1}\beta|V_{i+1,i}|\Big)\\ &\geq& \Big(1-2k(1+\epsilon_1)^k\epsilon_2^{-1}\beta\Big)|V_{i+1,i}| \geq \frac{1}{2}|V_{i+1,i}|. \end{eqnarray*} For $j >i+1$, let $S_j$ consist of those $w \in V'$ for which $\tilde{N}(w,V_{j,i}) < (p_{i+1,j}-\frac{\epsilon_2}{2})|V_{j,i}|$. Then the density of edges of $G$ between $S_j$ and $V_{j,i}$ deviates from $p$ by at least $\frac{\epsilon_2}{2}$. Since graph $G$ is $(p,\lambda)$-pseudo-random, we obtain that $\frac{\epsilon_2}{2} \leq \frac{\lambda}{\sqrt{|V_{j,i}||S_j|}}$ and hence $|S_j| \leq \frac{4\lambda^2}{\epsilon_2^2|V_{j,i}|}$. Also using that $p \leq 3/4$ we have $1-p-\epsilon_2=1-p-\frac{p}{10k} \geq (1-\frac{1}{3k})(1-p)$. Therefore, our third condition, combined with $\delta=(1-p)^kp^d$ and $(1-x)^t \geq 1-xt$ for all $0 \leq x \leq 1$, imply that for $j \geq i+1$ \begin{eqnarray} \label{eq1} |V_{j,i}| &\geq& (1-p-\epsilon_2)^{i-D(i,j)}(p-\epsilon_2)^{D(i,j)}|V_j| \geq (1-p-\epsilon_2)^k(p-\epsilon_2)^d|V_j| \nonumber\\ &\geq& \left(\Big(1-\frac{1}{3k}\Big)(1-p)\right)^k \left(p-\frac{p}{10k}\right)^d|V_j| \nonumber \\ &\geq& \left(1-\frac{1}{3k}\right)^k\left(1-\frac{1}{10k}\right)^k(1-p)^kp^d|V_j| \nonumber\\ &\geq& \frac{1}{2}(1-p)^kp^d|V_i|= \frac{\delta n}{2k}. \end{eqnarray} Since $\lambda \leq \frac{p\delta}{100k^3}n$ (see (\ref{eq2})) and $\epsilon_2=\frac{p}{10k}$, we therefore have $|S_j| \leq \frac{4\lambda^2}{\epsilon_2^2|V_{j,i}|} \leq \frac{1}{4k}|V_{i+1,i}|$. Let $V''$ be the set of vertices in $V'$ that are not in any of the sets $S_j$. The cardinality of $V''$ is at least $$|V''|\geq |V'|-\sum_{j>i+1}|S_j| \geq |V'|-k\cdot\Big(\frac{1}{4k}|V_{i+1,i}|\Big) \geq |V'|-\frac{1}{4}|V_{i+1,i}| \geq \frac{1}{4}|V_{i+1,i}|.$$ Moreover, by definition, for every $j>i+1$ and every vertex $w \in V''$ there are $|R(w,V_{j,i})| \leq \frac{\epsilon_2}{2}|V_{j,i}|$ red edges from $w$ to $V_{j,i}$ if $(i+1,j)$ is an edge of $H$ and also $\tilde{N}(w,V_{j,i})$ has size at least $(p_{i+1,j}-\frac{\epsilon_2}{2})|V_{j,i}|$. This implies that $$|\tilde{B}(w,V_{j,i})|= |\tilde{N}(w,V_{j,i}) \setminus R(w,V_{j,i})| \geq |\tilde{N}(w,V_{j,i})|-\frac{\epsilon_2}{2}|V_{j,i}| \geq (p_{i+1,j}-\epsilon_2)|V_{j,i}|$$ and therefore the vertices of $V''$ satisfy the first of the two properties of good vertices. We have reduced our goal to showing that there is an element of $V''$ that has the second property of good vertices. For $i+1<j<\ell \leq k$ and $(j,\ell)$ an edge of $H$, let $T_{j,\ell}$ denote the set of $w \in V''$ such that the density of red edges between $\tilde{B}(w,V_{j,i})$ and $\tilde{B}(w,V_{\ell,i})$ is more than $(1+\epsilon_1)^{i+1}\beta$. Notice that any vertex of $V''$ not in any of the sets $T_{j,\ell}$ is good. Therefore, if we show that $T_{j,\ell} <\frac{|V''|}{k^2}$ for each $T_{j,\ell}$, then there is a good vertex in $V''$ and the proof would be complete. To do so we will assume without loss of generality that $p_{i+1,j}$ and $p_{i+1,\ell}$ are both $p$ (the other 3 cases can be treated similarly using the fact that $\bar G$ is $(1-p,\lambda)$-pseudo-random). Since by (\ref{eq1}) we have that $|V_{\ell,i}|, |V_{j,i}| \geq \frac{\delta n}{2k}$ and $\frac{|V''|}{k^2} \geq \frac{1}{4k^2}|V_{i+1,i}| \geq \frac{\delta n}{8k^3}$, the result follows from the following claim. \begin{claim} Let $X,Y$ and $Z$ be three disjoint subsets of our $(p,\lambda)$-pseudo-random graph $G$ such that $|X| \geq \frac{\delta n}{8k^3}$ and $|Y|,|Z| \geq \frac{\delta n}{2k}$. For every $w \in X$ let $B_1(w), B_2(w)$ be the set of vertices in $Y$ and $Z$ respectively connected to $w$ by a blue edge and suppose that $|B_1(w)|\geq (p-\frac{p}{10k})|Y|$ and $|B_2(w)| \geq (p-\frac{p}{10k})|Z|$. Also suppose that the density of red edges between $Y$ and $Z$ is at most $\eta$ for some $\eta \geq \frac{p}{1000k^2}$. Then there is a vertex $w\in X$ such that the density of red edges between $B_1(w)$ and $B_2(w)$ is at most $\frac{k+1}{k}\eta$. \end{claim} \noindent {\bf Proof.} Let $m$ denote the number of triangles $(x,y,z)$ with $x \in X, y \in Y, z \in Z$, such that the edge $(y,z)$ is red. We need an upper bound on $m$. Let $U$ be the set of vertices in $Y$ that have fewer than $p^3\delta^3(10k)^{-10}n$ red edges to $Z$. So the number $m_1$ of triangles $(x,y,z)$ which have $y\in U$ and edge $(y,z)$ red is clearly at most $m_1\leq p^3\delta^3(10k)^{-10}n^3$. Let $W_1, W_2$ denote the subsets of vertices in $Y$ whose number of neighbors in $X$ is at least $(p+\frac{p}{20k})|X|$ or respectively at most $(p-\frac{p}{20k})|X|$. Since the density of edges between $W_i$ and $X$ deviates from $p$ by more than $\frac{p}{20k}$, using $(p,\lambda)$-pseudo-randomness of $G$, we have $\frac{p}{20k} \leq \frac{\lambda}{\sqrt{|X||W_i|}}$, or equivalently, $|X||W_i| \leq 400k^2p^{-2}\lambda^2.$ Therefore, using the upper bound $\lambda \leq \frac{p^8}{(10k)^{10}}\delta^2n$ from (\ref{eq2}), the number $m_2$ of triangles $(x,y,z)$ with $y \in W=W_1 \cup W_2$ and edge $(y,z)$ red is at most $$m_2 \leq |X||W|n \leq 800k^2p^{-2}\lambda^2n \leq (10k)^{-10}p^4\delta^4n^3.$$ For $y \in Y \setminus (U \cup W)$, we have the number of neighbors of $y$ in $X$ satisfy $\big|\frac{|N(y, X)|}{|X|}-p\big|\leq \frac{p}{20k}$ and the number of red edges from $y$ to $Z$ is at least $p^3\delta^3(10k)^{-10}n$. Recall that $R(y,Z)$ denotes the set of vertices in $Z$ connected to $y$ by red edges, hence we have that $|R(y,Z)| \geq p^3\delta^3(10k)^{-10}n$ for every $y \in Y \setminus (U \cup W)$. We also have that $|N(y, X)| \geq p|X|/2 \geq \frac{p\delta n}{16k^3}$. Since $G$ is $(p,\lambda)$-pseudo-random, we can bound the number of edges between $N(y, X)$ and $R(y,Z)$ by $p|N(y, X)||R(y,Z)|+\lambda\sqrt{|N(y, X)||R(y,Z)|}$. Using the above lower bounds on $|N(y, X)|$ and $|R(y,Z)|$, and the upper bound (\ref{eq2}) for $\lambda$, one can easily check that $$\frac{\lambda}{\sqrt{|N(y, X)||R(y,Z)|}} \leq\frac{\lambda}{\sqrt{\big(p\delta n/(16k^3)\big) \big(p^3\delta^3(10k)^{-10}n\big)}} \leq \frac{p}{20k}.$$ Hence the number of edges between $N(y, X)$ and $R(y,Z)$ is at most $(p+\frac{p}{20k})|N(y, X)||R(y,Z)|$. Recall that for all $y \in Y \setminus (U \cup W)$ we have that $|N(y, X)| \leq \big(p+\frac{p}{20k}\big)|X|$. Also, since the density of red edges between $Y$ and $Z$ is at most $\eta$, we have that $\sum_y|R(y,Z)| \leq \eta|Y||Z|$. Therefore, the number $m_3$ of triangles $(x,y,z)$ with $y \in Y \setminus (U \cup W), x\in X, z\in Z$ such that the edge $(y,z)$ is red is at most $$m_3 \leq \Big(p+\frac{p}{20k}\Big)\sum_{y \in Y \setminus (U \cup W)} |N(y, X)||R(y,Z)|\leq \Big(p+\frac{p}{20k}\Big)^2|X|\sum_y|R(y,Z)| \leq \Big(p+\frac{p}{20k}\Big)^2\eta|X||Y||Z|.$$ Using the lower bounds on $|X|, |Y|, |Z|, \eta$ from the assertion of the claim we have that $$p^2\eta|X||Y||Z| \geq \frac{p^3\delta^3}{(10k)^7}n^3 \geq (10k)^3 \max\big(m_1, m_2\big).$$ This implies that the total number of triangles $(x,y,z)$ with $x \in X, y \in Y, z \in Z$, such that the edge $(y,z)$ is red is at most \begin{eqnarray*} m &=& m_1+m_2+m_3 \leq 2\frac{p^2\eta|X||Y||Z|}{(10k)^3}+\Big(p+\frac{p}{20k}\Big)^2\eta|X||Y||Z|\\ &\leq& \big(1+1/(8k)\big)p^2\eta|X||Y||Z|.\end{eqnarray*} Therefore, there is vertex $w \in X$ such that the number of these triangles through $w$ is at most $(1+1/(8k))p^2\eta|Y||Z|$. Since $B_1(w) \subset N(w,Y)$ and $B_2(w) \subset N(w,Z)$, then the number of red edges between $B_1(w)$ and $B_2(w)$ is at most $(1+1/(8k))p^2\eta|Y||Z|$. Since we have that $|B_1(w)|\geq (p-\frac{p}{10k})|Y|$ and $|B_2(w)| \geq (p-\frac{p}{10k})|Z|$, the density of red edges between $B_1(w)$ and $B_2(w)$ can be at most $$\frac{(1+1/(8k))p^2\eta|Y||Z|}{|B_1(w)||B_2(w)|} \leq \frac{(1+1/(8k))p^2\eta} {(p-\frac{p}{10k})^2} \leq \frac{k+1}{k}\eta,$$ completing the proof. \end{proof} \section{Trees with superlinear induced Ramsey numbers} \label{superlinear} In this section we prove Theorem \ref{tree}, that there are trees whose induced Ramsey number is superlinear in the number of vertices. The proof uses Szemer\'edi's regularity lemma, which we mentioned in the introduction. A red-blue edge-coloring of the edges of a graph partitions the graph into two monochromatic subgraphs, the {\it red graph}, which contains all vertices and all red edges, and the {\it blue graph}, which contains all vertices and all blue edges. The weak induced Ramsey number $r_{\textrm{weak ind}}(H_1,H_2)$, introduced by Gorgol and \L uczak \cite{GoLu}, is the least positive integer $n$ such that there is a graph $G$ on $n$ vertices such that for every red-blue coloring of the edges of $G$, either the red graph contains $H_1$ as an induced subgraph or the blue graph contains $H_2$ as an induced subgraph. Note that this definition is a relaxation of the induced Ramsey numbers since we allow blue edges between the vertices of red copy of $H_1$ or red edges between the vertices of blue copy of $H_2$. Therefore a weak induced Ramsey number lies between the usual Ramsey number and the induced Ramsey number. Using this new notion we can strengthen Theorem \ref{tree} as follows. Recall that $K_{1,k}$ denotes a star with $k$ edges. \begin{theorem}\label{weak} For each $\alpha \in (0,1)$, there is a constant $k(\alpha)$ such that if $H$ is a graph on $k \geq k(\alpha)$ vertices with maximum independent set of size less than $(1-\alpha)k$, then $r_{\textrm{weak ind}}(H,K_{1,k}) \geq \frac{k}{\alpha}$. \end{theorem} Let $T$ be a tree which is a union of path of length $k/2$ with the star of size $k/2$ such that the end point of the path is the center of the star. Since $T$ contains the path $P_{k/2}$ and the star $K_{1,k/2}$ as induced subgraphs, then $r_{\textrm{ind}}(T) \geq r_{\textrm{weak ind}}(P_{k/2},K_{1,k/2})$. By using the above theorem with $k/2$ instead of $k$, $H=P_{k/2}$, and sufficiently small $\alpha$, we obtain that $r_{\textrm{ind}}(T)/k \rightarrow \infty$. Moreover the same holds for every sufficiently large tree which contains a star and a matching of linear size as subgraphs. We deduce Theorem \ref{weak} from the following lemma. \begin{lemma}\label{twocoloring} For each $\delta>0$ there is a constant $c_{\delta}>0$ such that if $G=(V,E)$ is a graph on $n$ vertices, then there is a $2$-coloring of the edges of $G$ with colors red and blue such that the red graph has maximum degree less than $\delta n$ and for every subset $W \subset V$, either there are at least $c_{\delta}n^2$ blue edges in the subgraph induced by $W$ or there is an independent set in $W$ in the blue graph of cardinality at least $|W|-\delta n$. \end{lemma} \begin{proof} Let $\epsilon=\frac{\delta^2}{100}$. By Szemer\'edi's regularity lemma, there is a positive integer $M(\epsilon)$ together with an equitable partition $V=\bigcup_{i=1}^k V_i$ of vertices of the graph $G=(V,E)$ into $k$ parts with $\frac{1}{\epsilon}<k<M(\epsilon)$ such that all but at most $\epsilon k^2$ of the pairs $(V_i,V_j)$ are $\epsilon$-regular. Recall that a partition is equitable if $\big||V_i|-|V_j|\big| \leq 1$ and a pair $(V_i,V_j)$ is called $\epsilon$-regular if for every $X \subset V_i$ and $Y \subset V_j$ with $|X| > \epsilon |V_i|$ and $|Y| > \epsilon |V_j|$, we have $|d(X,Y)-d(V_i,V_j)|<\epsilon$. Let $c_{\delta}=\epsilon M(\epsilon)^{-2}$. Notice that to prove Lemma \ref{twocoloring}, it suffices to prove it under the assumption that $n $ is sufficiently large. So we may assume that $n \geq \epsilon^{-1}M(\epsilon)$. If a pair $(V_i,V_j)$ is $\epsilon$-regular with density $d(V_i,V_j)$ at least $2\epsilon$, then color the edges between $V_i$ and $V_j$ blue. Let $G'$ be the subgraph of $G$ formed by deleting the edges of $G$ that are already colored blue. Let $V'$ be the vertices of $G'$ of degree at least $\delta n$. Color blue any edge of $G'$ with a vertex in $V'$. The remaining edges are colored red. First notice that every vertex has red degree less than $\delta n$. We next show that $|V'|$ is small by showing that $G'$ has few edges. There are at most $$\sum_{i=1}^k {|V_i| \choose 2} \leq \frac{n^2}{k} \leq \epsilon n^2$$ edges $(v,w)$ of $G$ with $v$ and $w$ both in the same set $V_i$. Since at most $\epsilon k^2$ of the pairs $(V_i,V_j)$ are not $\epsilon$-regular, then there are at most $\epsilon n^2$ edges in such pairs. The $\epsilon$-regular pairs $(V_i,V_j)$ with density less than $2\epsilon$ contain at most a fraction $2\epsilon$ of all possible edges on $n$ vertices. So there are less than $\epsilon n^2$ edges of this type. Therefore the number of edges of $G'$ is at most $3\epsilon n^2$, and therefore there are at most $|V'| \leq 2\frac{e(G')}{\delta n} \leq 6\epsilon \delta^{-1}n<\frac{\delta n}{10}$ vertices of degree at least $\delta n$ in it. Let $W \subset V$. Let $W'=W \setminus V'$, so $W'$ has cardinality at least $|W|-\frac{\delta n}{10}$. Let $W_i =V_i \cap W'$. Let $W''=\bigcup_{|W_i| \geq \epsilon |V_i|} W_i$. Notice that for any $i \in [k]$ there are at most $\epsilon \frac{n}{k}$ vertices in $(W' \setminus W'') \cap V_i$, so there are at most $k(\epsilon \frac{n}{k}) =\epsilon n = \frac{\delta^2 n}{100}$ vertices in $W'\setminus W''$. Therefore, $W''$ has at least $|W|-\delta n$ vertices. If there are $i \not =j$ such that $|W_i|,|W_j| \geq \epsilon\frac{n}{k}$ and the pair $(V_i,V_j)$ is $\epsilon$-regular with density at least $2\epsilon$, then there are at least $$\epsilon|W_i||W_j| \geq \frac{\epsilon}{k^2}n^2 \geq \epsilon M(\epsilon)^{-2}n^2=c_{\delta}n^2$$ blue edges between $W_i$ and $W_j$. In this case the blue subgraph induced by $W$ has at least $c_{\delta}n^2$ edges. Otherwise, all the edges in $W''$ are red, and $W''$ is an independent set in the blue graph of cardinality at least $|W|-\delta n$. \end{proof} \vspace{0.15cm} \noindent {\bf Proof of Theorem \ref{weak}.}\, Let $H$ be a graph on $k$ vertices with maximum independent set of size less than $(1-\alpha)k$. Take $\delta=\alpha^2$ and $c_{\delta}$ to be as in Lemma \ref{twocoloring}. Let $G=(V,E)$ be any graph on $n$ vertices, where $n \leq \frac{k}{\alpha}$. If $H$ has at least $c_{\delta}k^2$ edges, consider a random red-blue coloring of the edges of $G$ such that the probability of an edge being red is $\frac{\alpha}{2}$. The expected degree of a vertex in the red graph is at most $\alpha n/2$. Therefore by the standard Chernoff bound for the Binomial distribution it is easy to see that with probability $1-o(1)$ the degree of every vertex in the red graph is less than $\alpha n\leq k$, i.e., it contains no $K_{1,k}$. On the other hand, for $k$ sufficiently large, the probability that the blue graph contains a copy of $H$ is at most $$ n^k (1-\alpha/2)^{e(H)} \leq n^k e^{-\alpha c_{\delta}k^2/2} \leq e^{-\alpha c_{\delta}k^2/2+k\log (k/\alpha)}=o(1).$$ Thus with high probability this coloring has no blue copy of $H$ as well. This implies that we can assume that the number of edges in $H$ is less than $c_{\delta}k^2$. By Lemma \ref{twocoloring}, there is a red-blue edge-coloring of the edges of $G$ such that the red graph has maximum degree at most $\delta n$ and every subset $W \subset V$ contains either an independent set in the blue graph of size at least $|W|-\delta n$ or contains at least $c_{\delta}n^2$ blue edges. Since $\delta n=\alpha^2 n <k$, then the red graph does not contain $K_{1,k}$ as a subgraph. Suppose for contradiction that there is an induced copy of $H$ in the blue graph, and let $W$ be the vertex set of this copy. The blue graph induced by $W$ has $e(H)<c_{\delta}k^2\leq c_{\delta}n^2$ edges. Therefore it contains an independent set of size at least $|W|-\delta n \geq |W|-\alpha k=(1-\alpha)k$, contradicting the fact that $H$ has no independent set of size $(1-\alpha)k$. Therefore, there are no induced copies of $H$ in the blue graph.\ifvmode\mbox{ }\else\unskip\fi\hskip 1em plus 10fill$\Box$ \section{Concluding Remarks} \begin{itemize} \item All of the results in this paper concerning induced subgraphs can be extended to many colors. One such multicolor result was already proved in Section \ref{moreoninduced} (see Lemma \ref{lemmaerdoshajnal2}), and we use here the notation from that section. For example, one can obtain the following generalization of Theorem \ref{main}. For $k \geq 2$, let $\Psi:E(K_{k}) \rightarrow [r]$ be an edge-coloring of the complete graph $K_{k}$ and $\Phi:E(K_n) \rightarrow [s]$ be a $\Psi$-free edge-coloring of the complete graph $K_n$. Then there is a constant $c$ so that for every $\epsilon \in (0,1/2)$, there is a subset $W \subset K_n$ of size at least $2^{-crk (\log \frac{1}{\epsilon})^2}n$ and a color $i \in [r]$ such that the edge density of color $i$ in $W$ is at most $\epsilon$. Since the proofs of this statement and other generalizations can be obtained using our key lemma in essentially the same way as the proofs of the results that we already presented (which correspond to the two color case), we do not include them here. \item It would be very interesting to get a better estimate in Theorem \ref{main}. This will immediately give an improvement of the best known result for Erd\H{o}s-Hajnal conjecture on the size of the maximum homogeneous set in $H$-free graphs. We believe that our bound can be strengthened as follows. \begin{conjecture}\label{mainconjecture} For each graph $H$, there is a constant $c(H)$ such that if $\epsilon \in (0,1/2)$ and $G$ is a $H$-free graph on $n$ vertices, then there is an induced subgraph of $G$ on at least $\epsilon^{c(H)}n$ vertices that has edge density either at most $\epsilon$ or at least $1-\epsilon$. \end{conjecture} \noindent This conjecture if true would imply the Erd\H{o}s-Hajnal conjecture. Indeed, take $\epsilon=n^{-\frac{1}{c(H)+1}}$. Then every $H$-free graph $G$ on $n$ vertices contains an induced subgraph on at least $\epsilon^{c(H)}n=n^{\frac{1}{c(H)+1}}$ vertices that has edge density at most $\epsilon$ or at least $1-\epsilon$. Note that this induced subgraph or its complement has average degree at most $1$, which implies that it contains a clique or independent set of size at least $\frac{1}{2}n^{\frac{1}{c(H)+1}}$. \item One of the main remaining open problems on induced Ramsey numbers is a beautiful conjecture of Erd\H{o}s which states that there exists a positive constant $c$ such that $r_{\textrm{ind}}(H) \leq 2^{ck}$ for every graph $H$ on $k$ vertices. This, if true, will show that induced Ramsey numbers in the worst case have the same order of magnitude as ordinary Ramsey numbers. Our results here suggest that one can attack this problem by studying 2-edge-colorings of a random graph with edge probability $1/2$. It looks very plausible that for sufficiently large constant $c$, with high probability random graph $G(n,1/2)$ with $n\geq 2^{ck}$ has the property that any of its 2-edge-colorings contains every graph on $k$ vertices as an induced monochromatic subgraph. Moreover, maybe this is even true for every sufficiently pseudo-random graph with edge density $1/2$. \item The results on induced Ramsey numbers of sparse graphs naturally lead to the following questions. What is the asymptotic behavior of the maximum of induced Ramsey numbers over all trees on $k$ vertices? We have proved $r_{\textrm{ind}}(T)$ is superlinear in $k$ for some trees $T$. On the other hand, Beck \cite{Be} proved that $r_{\textrm{ind}}(T)=O\left( k^{2}\log^2 k\right)$ for all trees $T$ on $k$ vertices. For induced Ramsey numbers of bounded degree graphs, we proved a polynomial upper bound with exponent which is nearly linear in the maximum degree. Can this be improved further, e.g., is it true that the induced Ramsey number of every $n$-vertex graph with maximum degree $d$ is at most a polynomial in $n$ with exponent independent of $d$? It is known that the usual Ramsey numbers of bounded degree graphs are linear in the number of vertices. \end{itemize} \vspace{0.2cm} \noindent {\bf Acknowledgment.}\, We'd like to thank Janos Pach and Csaba T\'oth for helpful comments on an early stage of this project and Steve Butler and Philipp Zumstein for carefully reading this manuscript.
{ "timestamp": "2007-12-27T22:18:10", "yymm": "0706", "arxiv_id": "0706.4112", "language": "en", "url": "https://arxiv.org/abs/0706.4112", "abstract": "We present a unified approach to proving Ramsey-type theorems for graphs with a forbidden induced subgraph which can be used to extend and improve the earlier results of Rodl, Erdos-Hajnal, Promel-Rodl, Nikiforov, Chung-Graham, and Luczak-Rodl. The proofs are based on a simple lemma (generalizing one by Graham, Rodl, and Rucinski) that can be used as a replacement for Szemeredi's regularity lemma, thereby giving much better bounds. The same approach can be also used to show that pseudo-random graphs have strong induced Ramsey properties. This leads to explicit constructions for upper bounds on various induced Ramsey numbers.", "subjects": "Combinatorics (math.CO)", "title": "Induced Ramsey-type theorems", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9852713887451681, "lm_q2_score": 0.8311430436757312, "lm_q1q2_score": 0.8189014608882735 }
https://arxiv.org/abs/2109.03749
Lifting methods in mass partition problems
Many results in mass partitions are proved by lifting $\mathbb{R}^d$ to a higher-dimensional space and dividing the higher-dimensional space into pieces. We extend such methods to use lifting arguments to polyhedral surfaces. Among other results, we prove the existence of equipartitions of $d+1$ measures in $\mathbb{R}^d$ by parallel hyperplanes and of $d+2$ measures in $\mathbb{R}^d$ by concentric spheres. For measures whose supports are sufficiently well separated, we prove results where one can cut a fixed (possibly different) fraction of each measure either by parallel hyperplanes, concentric spheres, convex polyhedral surfaces of few facets, or convex polytopes with few vertices.
\section{Introduction} In a standard mass partition problem, we are given measures or finite families of points in a Euclidean space and we seek to partition the ambient space into regions that meet certain conditions. Some conditions determine how we split the measures and the sets of points. For instance, in an equipartition, we ask that each part has the same size in each measure or contains the same number of points of each set. Some conditions restrict the types of partitions which are allowed, such as partition by a single hyperplane. Determining whether such partitions always exist leads to a rich family of problems. Solutions to these problems often require topological methods and can have computational applications \cites{matousek2003using, Zivaljevic2017, RoldanPensado2021}. The quintessential mass partition result is the ham sandwich theorem, conjectured by Steinhaus and proved by Banach \cite{Steinhaus1938}. \begin{theorem}[Ham sandwich theorem] Let $d$ be a positive integer and $\mu_1, \ldots, \mu_d$ be finite measures of $\mathds{R}^d$. Then, there exists a hyperplane $H$ of $\mathds{R}^d$ so that its two closed half-spaces $H^+$ and $H^-$ satisfy \begin{align*} \mu_i (H^+) & \ge \frac{1}{2}\mu_i (\mathds{R}^d), \\ \mu_i (H^-) & \ge \frac{1}{2}\mu_i (\mathds{R}^d) \qquad \mbox{for $i=1,\ldots,d$.} \end{align*} \end{theorem} If we further ask that $\mu_i(H') = 0$ for each hyperplane $H'$ and every $i=1,\ldots, d$, the inequalities above are equalities. Stone and Tukey proved the ham sandwich theorem for general measures \cite{Stone:1942hu}. They also proved the polynomial ham sandwich theorem which states that \textit{any $\binom{d+k}{k}-1$ measures in $\mathds{R}^d$ can be halved with a polynomial on $d$ variables of degree at most $k$.} Even though this is a far-reaching generalization of the ham sandwich theorem, its proof relies on a simple trick. We lift $\mathds{R}^d$ to $\mathds{R}^{\binom{d+k}{k}-1}$ by the Veronese map and apply the ham sandwich theorem in the higher-dimensional space. In this paper, we prove several mass partition results by lifting $\mathds{R}^d$ to higher-dimensional spaces, particularly $\mathds{R}^{d+1}$, in new ways. In \cref{sec:sines-and-spheres}, we revisit a known result about equipartitions of measures with spheres and prove a new result about equipartitions of three measures in $\mathds{R}^2$ using a sinusoidal curve of fixed period. Then, instead of lifting to higher-dimensional spaces via smooth maps, such as the Veronese maps, we lift to polyhedral surfaces in $\mathds{R}^{d+1}$. This forces us to use the ham sandwich theorem for general measures, which is interesting on its own. One of our main results is the following ham sandwich theorem for parallel hyperplanes. \begin{theorem}\label{thm:parallel-hyperplanes} Let $d$ be a positive integer and $\mu_1, \ldots, \mu_{d+1}$ be $d+1$ finite measures in $\mathds{R}^d$, each absolutely continuous with respect to the Lebesgue measure. Then, there exist two parallel hyperplanes $H_1, H_2$ so that the region between them contains exactly half of each measure. \end{theorem} If there is a hyperplane $H_1$ that halves all measures, we can consider $H_2$ to be at infinity. If $H_1$ and $H_2$ are not required to be parallel, we present with \cref{thm:simple-wedge} a simple proof that \textit{for any $d+1$ finite measures in $\mathds{R}^d$ there exist two closed half-spaces whose intersection contains exactly half of each measure}. The intersection of two half-spaces is called a wedge. The fact that $d+1$ measures in $\mathds{R}^d$ can be halved by a wedge was first proved by B\'ar\'any and Matou\v{s}ek in dimension two \cite{Barany:2001fs} and later generalized to $\mathds{R}^d$ by Schnider \cite{Schnider:2019ua}. For $\mathds{R}^2$, Bereg presented algorithmic approaches for the discrete version which show that more conditions can be imposed to the wedge \cite{Bereg:2005voa}. The proof requires a new Borsuk--Ulam type theorem about direct products of spheres and Stiefel manifolds, \cref{thm:new-topological-result}, which we describe in \cref{sec:polyhedral-lift}. As a corollary, we combine \cref{thm:parallel-hyperplanes} with known lifting techniques. We nickname the following result as the ``bagel ham sandwich theorem'', due to how it looks in $\mathds{R}^2$. \begin{corollary}[Bagel ham sandwich theorem]\label{cor:Bagel} Let $d$ be a positive integer and $\mu_1, \ldots, \mu_{d+2}$ be $d+2$ finite measures in $\mathds{R}^d$, each absolutely continuous with respect to the Lebesgue measure. Then, we can find two concentric spheres in $\mathds{R}^d$ so that the closed region between them has exactly half of each measure. \end{corollary} \cref{thm:parallel-hyperplanes} is optimal, as the region between the two hyperplanes is convex. One can simply take $d+1$ measures concentrated each around a vertex of a simplex and a final measure concentrated around the barycenter of the simplex to show that the result is impossible with $d+2$ measures. The problem of cutting the same fraction for a family of measures with a single convex set has been studied before \cites{Akopyan:2013jt, Blagojevic:2007ij}, which we revisit in \cref{sec:remarks}. \cref{thm:parallel-hyperplanes} is also related to the problem of halving measures in $\mathds{R}^d$ using hyperplane arrangements. Langerman conjectured that \textit{any $dn$ measures in $\mathds{R}^d$ can be simultaneously halved by a chessboard coloring induced by $n$ hyperplanes} \cites{Barba:2019to, Hubard2020}. For $n=2$, this has been confirmed for $2d-O(\log d)$ measures \cite{Blagojevic2018}. If the hyperplanes are required to be parallel, this reduces the dimension of the space of possible partitions from $2d$ to $d+1$, matching the number of measures in \cref{thm:parallel-hyperplanes}. General mass partition results like the ham sandwich theorem can halve many measures simultaneously. If we want to cut a fixed (but possibly different) fraction of each measure, conditions need to be imposed. For example, if two measures coincide, it is impossible to find a half-space that contains exactly half of one and one third of the other. The first result with arbitrary sizes for each measure was proved by Hugo Steinhaus in dimensions two and three \cite{Steinhaus1945}. He required the support of the measures to be well separated, meaning that the supports of any set of measures could be separated from the supports of the rest by a hyperplane. This condition was sufficient to guarantee the existence of a half-space cutting a fixed fraction of several measures. This result was rediscovered and extended to high dimensions independently by B\'ar\'any, Hubard, and Jer\'onimo and by Breuer \cites{Barany:2008vv, Breuer2010}. \begin{theorem}[B\'ar\'any, Hubard, Jer\'onimo 2008; Breuer 2010]\label{thm:BHJ} Let $d$ be a positive integer and $\mu_1, \ldots, \mu_d$ be $d$ finite measures in $\mathds{R}^d$, each absolutely continuous with respect to the Lebesgue measure so that their supports $K_1, \ldots, K_d$ are well separated. Let $\alpha_1, \ldots, \alpha_d$ be real numbers in $(0,1)$. Then, there exists a half-space $H$ so that \[ \mu_i(H) = \alpha_i \cdot \mu_i (\mathds{R}^d) \qquad \mbox{for }i=1,\ldots, d. \] \end{theorem} The proof of B\'ar\'any, Hubard, and Jer\'onimo uses Brouwer's fixed point theorem. Breuer's proof uses the Poincar\'e--Miranda theorem, which is equivalent to Brouwer's fixed point theorem but has a significantly different formulation. Steinhaus' proof is quite different and uses the Jordan curve theorem. In \cref{sec:well separated} we present a new proof of \cref{thm:BHJ} that uses a degree argument. We extend \cref{thm:parallel-hyperplanes} in a similar way for well separated measures. \begin{theorem}\label{thm:parallel-hyperplanes-separated} Let $d$ be a positive integer and $\mu_1, \ldots, \mu_{d+1}$ be $d+1$ finite measures in $\mathds{R}^d$, each absolutely continuous with respect to the Lebesgue measure. Suppose that the supports $K_1, \ldots, K_{d+1}$ of $\mu_1, \ldots, \mu_{d+1}$ are well separated. Let $\alpha_1, \ldots, \alpha_{d+1}$ be real numbers in $(0,1)$. Then, there exist two parallel hyperplanes $H_1, H_2$ in $\mathds{R}^d$ so that the region $A$ between them satisfies \[ \mu_i(A) = \alpha_i \cdot \mu_i(\mathds{R}^d) \qquad \mbox{for all }i=1,\ldots, d+1. \] \end{theorem} We also combine the results with partitions with few hyperplanes and those of partitions using a single convex. We exhibit conditions for measures in $\mathds{R}^d$ that guarantee the existence of a (possibly unbounded) convex polyhedron of few facets which contains a fixed fraction of each measure or the existence of a convex polytope with few vertices that contains a fixed fraction of each measure. This is done in \cref{sec:polyhedral}. These results work with an arbitrary number of measures in $\mathds{R}^d$. Finally, we revisit a mass partition result by Akopyan and Karasev that uses a lifting argument in its proof. Akopyan and Karasev proved that for any positive integer $n$ and any $d+1$ measures in $\mathds{R}^d$, there exists a convex set $K$ whose measure is exactly $1/n$ of each measure. We extend the methods from \cref{sec:polyhedral-lift} to bound the complexity of $K$ by writing it as the intersection of few half-spaces. \begin{theorem}\label{thm:same-fraction} Let $n,d $ be positive integers and $\mu_1, \ldots, \mu_{d+1}$ be $d+1$ finite measures in $\mathds{R}^d$, each absolutely continuous with respect to the Lebesgue measure. There exists a convex set $K$, such that $K$ is the intersection of $\sum_{j=1}^r k_j(p_j-1)p_j$ half-spaces and \[ \mu_i(K) = \frac{1}{n}\mu_i(\mathds{R}^d) \qquad \mbox{for all }i=1,\ldots, d+1, \] where $n=p_1^{k_1}\ldots p_r^{k_r}$ is the prime factorization of $n$. \end{theorem} We conclude in \cref{sec:remarks} with remarks and open problems. \section{Equipartition with spheres and sine curves}\label{sec:sines-and-spheres} While the traditional ham sandwich theorem simultaneously halves $d$ measures in $\mathds{R}^d$ by a hyperplane, we can simultaneously halve $d+1$ or more measures in $\mathds{R}^d$ if we increase the complexity of the cut. The following theorem is a consequence of Stone and Tukey's polynomial ham sandwich theorem \cite{Stone:1942hu} and was one of Stone and Tukey's first examples of their main results. It was also proved in dimension two by Hugo Steinhaus in 1945 \cite{Steinhaus1945} using a particular parametrizations of the space of circles in $\mathds{R}^2$. We present a new proof with a stereographic projection. \begin{theorem}\label{thm:circular-sandwich} Let $d$ be a positive integer and let $\mu_1, \ldots, \mu_{d+1}$ be $d+1$ finite measures in $\mathds{R}^d$, each absolutely continuous with respect to the Lebesgue measure. Then, there exists either a sphere or a hyperplane that simultaneously splits each measure by half. \end{theorem} \begin{proof} We first embed $\mathds{R}^d$ to $\mathds{R}^{d+1}$ by appending a coorinate $1$ to each point, so $x \mapsto (x,1) \in \mathds{R}^{d+1}$. Then, we apply $r: \mathds{R}^{d+1}\setminus\{0\} \to \mathds{R}^{d+1}\setminus\{0\}$ the inversion centered at $0$ with radius $1$. This is a transformation that sends spheres containing the origin to hyperplanes and hyperplanes to spheres containing the origin. Hyperplanes containing the origin are fixed set-wise by the inversion and we consider them as degenerate spheres as well. Restricted to the embedding of $\mathds{R}^d$, the inversion is a stereographic projection to the sphere $S$ of radius $1/2$ centered at $(0,\ldots,0,1/2)$ by rays through the origin. We also know that $r \circ r$ is the identity. When we lift the measures $\mu_1, \ldots, \mu_{d+1}$ to $\mathds{R}^{d+1}$ and apply $r$, we get measures $\sigma_1, \ldots, \sigma_{d+1}$ on $S$. By the ham sandwich theorem in $\mathds{R}^{d+1}$, there exists a hyperplane $H$ halving each of $\sigma_1, \ldots, \sigma_{d+1}$. Since $r(H)$ is a sphere in $\mathds{R}^{d+1}$, it intersects the embedding of $\mathds{R}^d$ in a $(d-1)$-dimensional sphere halving each of $\mu_1, \ldots, \mu_{d+1}$, as we wanted. The only exceptional case is if $H$ contains the origin, in which case $r(H) = H$, which gives us an equipartition by a hyperplane in $\mathds{R}^d$. \end{proof} Similarly, the idea of lifting allows for an visual and intuitive proof of the existence of equipartitions of three sets in $\mathds{R}^2$ by a sine wave. By a sine wave of period $\alpha$ we mean the graph of a function of the form $y = r + A \sin (2\pi x / \alpha + s)$, for real numbers $A,r,s$. \begin{theorem}\label{thm:cylinder} Let $\alpha > 0$ be a real number. Given three finite measures $\mu_1, \mu_2, \mu_3$ in $\mathds{R}^2$, each absolutely continuous with respect to the Lebesgue measure, there exists a sine wave of period $\alpha$ halving each measure. \end{theorem} We allow ``degenerate'' sine waves of period $\alpha$. A degenerate sine wave of period $\alpha$ is formed by taking two vertical lines intersecting the $x$-axis in the interval $[0,\alpha)$ and making a translated copy in each interval of the form $n\alpha + [0,\alpha)$. This set induces a chessboard coloring of the plane into two regions. We can think of this as the limit of a sequence of sine waves of period $\alpha$ of increasing ampitude. \begin{proof} We prove the result for $\alpha = 2\pi$, as the two cases are equivalent after a scaling argument. We wrap $\mathds{R}^d$ around the cylinder $C$ in $\mathds{R}^3$ with equation $x^2 + z^2 = 1$ with the function. \begin{align*} f: \mathds{R}^2 & \to C \\ (x,y) & \mapsto \left(\cos\left(x\right), y, \sin\left(x\right)\right) \end{align*} Let $\sigma_1, \sigma_2, \sigma_3$ be the measures that $\mu_1, \mu_2, \mu_3$ induce on $C$ by this lifting, respectively. We apply the ham sandwich theorem to these three measures in $\mathds{R}^3$. Therefore, we can find a plane $H=\{(x,y,z): ax + by + cz = d\}$ that halves each of $\sigma_1, \sigma_2, \sigma_3$. When we pull $H \cap C$ back to $\mathds{R}^2$, we get the set of points $(x,y)$ that satisfy $a \cos(x) + b y + c\sin(x) = d$. Since a linear combination of the sine and cosine functions is a sinusoid with the same period but possibly different amplitude and phase shift, we have $a \sin(x) + c \cos (x) = A\sin(x + s)$ for some $A$ and $s$. \begin{figure}[ht!] \centering \includegraphics[width=.7\textwidth]{cylinder.png} \caption{An example of a cylinder $C$ with period $\alpha=2\pi$. The lift $\mathds{R}^2 \to C$ is not an injective function but this does not cause a problem.} \label{fig:cylinder} \end{figure} \end{proof} The degenerate cases appear when $b$, the coefficient of $y$, is zero. One can prove a high-dimensional version of \cref{thm:cylinder} by wrapping $\mathds{R}^d$ around $S^{d-1}\times \mathds{R}$ to find ``sinusoidal surfaces of fixed period'' that halve $d+1$ measures in $\mathds{R}^d$. \section{Equipartitions with wedges and parallel hyperplanes}\label{sec:polyhedral-lift} In this section, we prove results regarding equipartitions of $d+1$ mass distributions in $\mathds{R}^d$ by a wedge. A \textit{wedge} in $\mathds{R}^d$ is the intersection of two closed half-spaces. Note that a single closed half-space is also considered a wedge. We say that a measure $\mu$ with support $K$ in $\mathds{R}^d$ is \textit{absolutely continuous} if it is absolutely continuous with respect to the Lebesgue measure, the interior of $K$ is connected and not empty, and for every open set $U \subset K$ we have $\mu(U) > 0$. This guarantees that there is a unique halving hyperplane in each direction for $\mu$ and that the halving hyperplane varies continuously as we change the direction of the cut. We first establish a lemma about halving hyperplanes. We only use the lemma below with $n=d+1$, but it works in general. \begin{lemma}\label{lem:separating-fixed-direction} Let $n$ be a positive integer, $\mu_1, \ldots, \mu_{n}$ be finite absolutely continuous measures in $\mathds{R}^d$, and $v$ be a unit vector in $\mathds{R}^d$. There either exists a hyperplane $H$ orthogonal to $v$ that halves each of the $n$ measures or there exists a hyperplane $H$ such that its two closed half-spaces satisfy \begin{align*} \mu_i (H^+) & < \frac{1}{2}\mu_i(\mathds{R}^d) \qquad \mbox{for some $i \in [n]$ and}\\ \mu_{i'} (H^-) & < \frac{1}{2}\mu_{i'}(\mathds{R}^d) \qquad \mbox{for some $i' \in [n]$.} \end{align*} \end{lemma} \begin{proof} For each $i$, let $H_i$ be the halving hyperplane for $\mu_i$ orthogonal to $v$. If all these hyperplanes coincide we are done. Otherwise, we can order this set of hyperplanes by the direction $v$ and any hyperplane $H$ strictly between the first $H_i$ and the last $H_{i'}$ satisfies the conditions we want. \end{proof} In the situation above, we always take the hyperplane $H'$ exactly half-way between the first $H_i$ and the last $H_{i'}$. This makes the choice of hyperplanee continuous as $v$ varies, and invariant if we replace $v$ by $-v$. Through the rest of the manuscript we denote the canonical basis of $\mathds{R}^d$ by $e_1, \ldots, e_d$. \begin{theorem}\label{thm:simple-wedge} Let $d$ be a positive integer and $\mu_1, \ldots, \mu_{d+1}$ be $d+1$ finite absolutely continuous measures in $\mathds{R}^d$. Then, there exists a wedge that contains exactly half of each measure. \end{theorem} \begin{proof} Let $v$ be a unit vector in $\mathds{R}^d$. If there is a hyperplane $H$ orthogonal to $v$ halving each measure we are done. Otherwise, by \cref{lem:separating-fixed-direction}, we can find a hyperplane $H$ such that each side contains less than half of some measure. Consider the lifting of $\mathds{R}^d$ to $\mathds{R}^{d+1}$ where we append an additional coordinate to every point $x \in \mathds{R}^d$. Formally, we lift via the map $x \mapsto (x, \operatorname{dist}(x,H))$. We denote by $S(H)$ the image of $\mathds{R}^{d}$ in this embedding. Note that the function $x \mapsto \operatorname{dist}(x,H)$ is affine on each side of $H$, so $S(H)$ is contained in the union of two hyperplanes that contain $\{(x,0): x \in H\}$. See \cref{fig:S(H)} for an illustration of the case $d=2$. We lift each measure $\mu_i$ in $\mathds{R}^d$ to a measure $\sigma_i$ in $\mathds{R}^{d+1}$. The measures $\sigma_1, \ldots, \sigma_{d+1}$ are no longer absolutely continuous. We now apply the ham sandwich theorem for general measures in $\mathds{R}^{d+1}$. Therefore, we can find a hyperplane $H'$ in $\mathds{R}^{d+1}$ so that its two closed half-spaces $(H')^+, (H')^-$ satisfy $\sigma_i((H')^+) \ge \frac{1}{2}\sigma_i(\mathds{R}^{d+1})$ and $\sigma_i((H')^-) \ge \frac{1}{2}\sigma_i(\mathds{R}^{d+1})$ for all $i =1,\ldots, d+1$. By construction, each side of $H$ has strictly less than half of one of the measures $\mu_i$. If the hyperplane $H'$ coincides with one of the two hyperplanes whose union contains $S(H)$, the half-space bounded by $H'$ that contains infinite rays in the direction $-e_{d+1}$ would have less than half the corresponding $\sigma_i$. Therefore $H'$ is not one of the two hyperplanes forming $S(H)$. As the two components of $S(H)$ were the only hyperplanes with non-zero measure for each $\sigma_i$, we conclude that $H'$ halves each of the measures in $\mathds{R}^{d+1}$. As a final observation, $H'\cap S(H)$ projects back to $\mathds{R}^d$ as the boundary of a wedge that halves all measures. \end{proof} \begin{figure} \centering \includegraphics[width=.8\textwidth]{Vshape.pdf} \caption{An example of lifting when given three sets of points in $\mathds{R}^2$. Points on the $xy$-plane are sent to points on this surface. The hyperplane $H$ in this case is the $x$-axis.} \label{fig:S(H)} \end{figure} In the proof of \cref{thm:simple-wedge} we could choose the direction $v$ arbitrarily. We now use this degree of freedom to strengthen the result. Even though \cref{thm:parallel-hyperplanes} implies \cref{thm:simple-wedge}, we state it separately as the proof requires more technical tools. In particular, a simple application of the ham sandwich theorem is insufficient. We require some additional topological tools in lieu of the Borsuk--Ulam theorem. Let $V_k(\mathds{R}^d)$ be the Stiefel manifold of orthonormal $k$-frames in $\mathds{R}^d$. Formally, \[ V_k(\mathds{R}^d) = \{(v_1,\ldots, v_k) : v_1, \ldots, v_k \in \mathds{R}^d \mbox{ are orthonormal}\}. \] The space $V_k(\mathds{R}^d)$ has a free action of the group $(\mathds{Z}_2)^{k}$, where we consider $\mathds{Z}_2 = \{+1,-1\}$ with multiplication. Given $(v_1,\ldots, v_k) \in V_k(\mathds{R}^d)$ and $(\lambda_1, \ldots, \lambda_k) \in (\mathds{Z}_2)^k$, we define \[ (\lambda_1, \ldots, \lambda_k)\cdot (v_1,\ldots, v_k) = (\lambda_1 v_1, \ldots, \lambda_k v_k) \in V_k(\mathds{R}^d). \] A similar action of $(\mathds{Z}_2)^k$ can be defined in $\mathds{R}^{d-1}\times \ldots \times \mathds{R}^{d-k}$, as the direct product of the actions of $\mathds{Z}_2$ on each $\mathds{R}^{d-i}$. A recent result of Chan, Chen, Frick, and Hull describes properties of $(\mathds{Z}_2)^k$-equivariant maps between these two spaces. \begin{theorem}[Chan, Chen, Frick, Hull 2020 \cite{Chan2020}]\label{thm:topo1} Let $k, d$ be positive integers. Every continuous $(\mathds{Z}_2)^k$-equivariant map $f:V_k(\mathds{R}^d) \to \mathds{R}^{d-1}\times \ldots \times \mathds{R}^{d-k}$ has a zero. \end{theorem} Manta and Sober\'on recently found an elementary proof of \cref{thm:topo1} \cite{Manta2021}. We use the result above in \cref{sec:well separated}. For this section, we need a slight modification. We use the product of the actions of $\mathds{Z}_2$ on the $d$-dimensional sphere $S^d$ and of $(\mathds{Z}_2)^k$ on $V_k(\mathds{R}^d)$ to define a free action of $(\mathds{Z}_2)^{k+1}$ on $S^d \times V_k(\mathds{R}^d)$. \begin{theorem}\label{thm:new-topological-result} Let $k, d$ be positive integers. Every continuous $(\mathds{Z}_2)^{k+1}$-equivariant map $f:S^d \times V_k(\mathds{R}^d) \to \mathds{R}^d \times \mathds{R}^{d-1}\times \ldots \times \mathds{R}^{d-k}$ has a zero. \end{theorem} There are several ways to prove the result above. The dimension of the image and the domain are the same and the action of $(\mathds{Z}_2)^{k+1}$ is free on $S^d \times V_k(\mathds{R}^d)$. Therefore, \cref{thm:new-topological-result} is a consequence of the general Borsuk--Ulam type results of Musin \cite{Mus12}. We simply need to find an equivariant function between these two spacse that has an odd number of orbits of zeroes. Such functions are known for $\mathds{Z}_2$-equivariant $f_1:S^d \to \mathds{R}^d$ and for $(\mathds{Z}_2)^{k}$-equivariant $f_2: V_k(\mathds{R}^d) \to \mathds{R}^{d-1}\times \ldots \times \mathds{R}^{d-k}$, so we can simply take $f_0= f_1 \times f_2$. Alternatively, one can use the methods of Chan et al. \cite{Chan2020} to prove \cref{thm:new-topological-result}. It suffices to note that $S^d \times V_k(\mathds{R}^d)$ is a space in which their topological invariants can be applied and the particular function $f_0$ is all that's needed to replace \cite{Chan2020}*{second proof of Lemma 3.2}. We only use \cref{thm:new-topological-result} for $k=d-1$. We present a short proof using the existing computations of the Fadell--Husseini index of these spaces on $\mathds{Z}_2$ cohomology \cite{Fadell:1988tm}. Given spaces $X$ and $Y$ with actions of $(\mathds{Z}_2)^{k+1}$, their indices $\ind^{(\mathds{Z}_2)^{k+1}}(X)$, $\ind^{(\mathds{Z}_2)^{k+1}}(Y)$ are ideals in the polynomial ring $\mathds{Z}_2[t_0,t_1,\ldots,t_k]$. Moreover, if there exists a continuous $(\mathds{Z}_2)^{k+1}$-equivariant map $f:X \to Y$, we must have $\ind^{(\mathds{Z}_2)^{k+1}}(Y) \subset \ind^{(\mathds{Z}_2)^{k+1}}(X)$. More details on this index and its computation for spaces and group actions common in discrete geometry can be found on recent work of Blagojevi\'c, L\"uck, and Ziegler \cite{Blagojevic2015}. \begin{proof}[Proof of \cref{thm:new-topological-result}] The result is equivalent to showing that there exists no continuous $(\mathds{Z}_2)^{k+1}$-equivariant map $f:S^d \times V_k(\mathds{R}^d) \to (\mathds{R}^d \times \mathds{R}^{d-1}\times \ldots \times \mathds{R}^{d-k}) \setminus \{0\}$. The space $(\mathds{R}^{d}\times \ldots \times \mathds{R}^{d-k}) \setminus \{0\}$ is homotopy equivalent to the join of spheres $S^{d-1}*\ldots * S^{d-k-1}$ and we know $\ind^{(\mathds{Z}_2)^{k+1}} ( S^{d-1}*\ldots * S^{d-k-1} ) \subset \mathds{Z}_2 [t_0, t_1,\ldots, t_k]$ is the ideal generated by the single monomial $t_0^dt_1^{d-1}\ldots t_{k}^{d-k}$. On the other hand, $\ind^{(\mathds{Z}_2)^{k+1}} (S^d \times V_k(\mathds{R}^d))\subset \mathds{Z}_2[t_0,\ldots, t_k]$ is the ideal generated by the polynomials $t_0^{d+1}, f_1,\ldots, f_k$ where $f_1,\ldots, f_k \subset \mathds{Z}_2[t_1,\ldots, t_k]$ generate $\ind^{(\mathds{Z}_2)^k}(V_k(\mathds{R}^d))$. These polynomials are were described completely by Fadell and Husseini \cite{Fadell:1988tm}*{Thm. 3.16}. Notably \[ f_i = t_i^{n-i+1}+ w_{i,n-i}t_i^{n-i}+ \ldots + w_{i,0}, \] where $w_{i,j} \in \mathds{Z}_2[t_1,\ldots, t_{i-1}]$ and the degree of $w_{i,j}t_i^j$ is $n-i+1$. In particular, \[ t_0^dt_1^{d-1}\ldots t_{k}^{d-k} \not\in \ind^{(\mathds{Z}_2)^{k+1}}(S^d \times V_k(\mathds{R}^d)), \] which shows that no continuous $(\mathds{Z}_2)^{k+1}$-equivariant map $f:S^d \times V_k(\mathds{R}^d) \to (\mathds{R}^d \times \mathds{R}^{d-1}\times \ldots \times \mathds{R}^{d-k})\setminus\{0\}$ exists. \end{proof} The second tool we require is a minor modification of the lift from $\mathds{R}^d$ to $\mathds{R}^{d+1}$. In the previous proof, given a hyperplane $H \subset \mathds{R}^d$ we lifted $\mathds{R}^d$ directly to $S(H) \subset \mathds{R}^{d+1}$. This has the inconvenience that the lift of an absolutely continuous measure is no longer absolutely continuous in $\mathds{R}^{d+1}$. We do not require such a strong condition, but we do require the lifted measures to assign mass zero to any hyperplane. To avoid this problem, we lift each measure $\mu_i$ to $S(H)^{\varepsilon}$, the region between $S(H)-\varepsilon \cdot e_{d+1}$ and $S(H)+\varepsilon \cdot e_{d+1}$, which we formalize below. We say that a measure $\mu$ in $\mathds{R}^d$ is smooth if it is the integral of a continuous positive function $f:\mathds{R}^d \to \mathds{R}$ (i.e., $f(A) = \int_A f$ for any measurable set $A$). We ``lift'' $f$ to a function \begin{align*} \tilde{f} : \mathds{R}^d \times \mathds{R} & \to \mathds{R} \\ (x,t) & \mapsto \begin{cases} \left(\frac{1}{2\varepsilon}\right)f(x) & \mbox{ if } |\operatorname{dist}(x,H) - t| \le \varepsilon \\ 0 & \mbox{ otherwise.} \end{cases} \end{align*} We say that the measure $\sigma^{\varepsilon}$ defined as the integral of $\tilde{f}$ in $\mathds{R}^d$ is the lift of $\mu$ to $S(H)^{\varepsilon}$. Notice that as $\varepsilon \to 0$, the measure $\sigma^{\varepsilon}$ converges weakly to the lift of $\mu$ to $S(H)$. For $\varepsilon>0$, the measure $\sigma^{\varepsilon}$ is not absolutely continuous, but it has value $0$ on each hyperplane. \begin{proof}[Proof of \cref{thm:parallel-hyperplanes}] We first assume that no hyperplane simultaneously halves all measures, or we are done. Since the set of smooth measures is dense in the set of absolutely continuous measures, we may assume without loss of generality that the measures $\mu_1, \ldots, \mu_{d+1}$ are smooth. Let $\varepsilon>0$. For $v \in S^d$ and $(v_1, \ldots, v_{d-1}) \in V_{d-1}(\mathds{R}^d)$, consider the element $(v,v_1, \ldots, v_{d-1}) \in S^d \times V_{d-1} (\mathds{R}^d)$. Let $H$ be the translate of the hyperplane $T=\operatorname{span}\{v_1,\ldots, v_{d-1}\}$ chosen from \cref{lem:separating-fixed-direction}. Let $\sigma^{\varepsilon}_1, \ldots, \sigma^{\varepsilon}_{d+1}$ be the lifts of $\mu_1, \ldots, \mu_{d+1}$ to $S(H)^{\varepsilon}$, respectively. Let $\lambda$ be the value so that the half-spaces \begin{align*} A & = \{x \in \mathds{R}^{d+1}: \langle x, v \rangle \ge \lambda \} \quad \mbox{ and } \\ B & = \{x \in \mathds{R}^{d+1}: \langle x, v \rangle \le \lambda \} \end{align*} have the same $\sigma_{d+1}^{\varepsilon}$-measure. Now we are ready to define a map \begin{align*} f : S^d \times V_{d-1}(\mathds{R}^d) & \to \mathds{R}^d \times \mathds{R}^{d-1} \times \ldots \times \mathds{R}^1 \\ (v,v_1,\ldots, v_{d-1}) & \mapsto (x_d,\ldots, x_1) \end{align*} where $x_i \in \mathds{R}^{i}$ for $i=1,\ldots, d$. First, we consider \[ x_d = \begin{bmatrix} \sigma_1^{\varepsilon}(A)-\sigma_1^{\varepsilon}(B) \\ \vdots \\ \sigma_d^{\varepsilon}(A)-\sigma_d^{\varepsilon}(B) \end{bmatrix}. \] For $i=1,\ldots, d-1$, the first coordinate of $x_i$ is $\langle v, e_{d+1}\rangle \langle v, (v_{d-i},0) \rangle$ and the rest are zero. This function is continuous by the construction of the lift of the measures. It is also $(\mathds{Z}_2)^{k+1}$-equivariant: if we flip the sign of $v$, only $x_d$ changes sign and if we flip the sign of $v_i$ only $x_{d-i}$ changes sign for $i=1,\ldots, d-1$. By \cref{thm:new-topological-result} the function $f$ has a zero. The condition $x_d = 0$ tells us that $A$ and $B$ each have exactly half of each $\sigma^{\varepsilon}_i$ for $i=1,\ldots, d+1$. The conditions of the rest of the $x_i$ being zero vectors means that either $v$ is orthogonal to $e_{d+1}$ or $v$ is orthogonal to each $(v_i,0)$ for $i=1,\ldots, d-1$. If the first condition happens, then when we project $\mathds{R}^{d+1}$ to the hyperplane $e_{d+1} = 0$, each $\sigma^{\varepsilon}_i$ projects to $\mu_i$ and $A$, $B$ project onto two half-spaces of $\mathds{R}^d$. This would mean we have a hyperplane halving each of the original measures, contradicting our initial assumption. Therefore $v$ is orthogonal to $(v_i,0)$ for all $i$. We take a sequence of positive real numbers $\varepsilon_{k} \to 0$. For each of them, we find a zero of the function induced above. As $S^d \times V_{d-1}(\mathds{R}^d)$ is compact, the zeros must have a converging subsequence. In the limit, we obtain two complementary half-spaces $A, B$ so that each contains at least half of $\sigma_i$ for $i=1,\ldots, d+1$ on $S(H)$ (where the direction of $H$ is determined by the limit of $(v_1, \ldots, v_{d-1})$). Let $H'$ be the hyperplane at the boundary of $A$ and $B$. In the limit, the vector $v$ normal to $H'$ is orthogonal to the subspace $T_0=\operatorname{span}\{(v_1,0), \ldots, (v_{d-1},0)\}$. By the construction of $H$, the hyperplane $H'$ cannot be one of the two hyperplane components of $S(H)$. The orthogonality mentioned before implies that $H' \cap S(H)$ must be two $(d-1)$-dimensional affine spaces parallel to $T_0$. When we project back to $\mathds{R}^d$, these parallel intersections form $H_1$ and $H_2$ and the region between them has exactly half of each $\mu_i$. \end{proof} \begin{proof}[Proof of \cref{cor:Bagel}] We lift $\mathds{R}^d$ to the paraboloid \[ \mathcal{P} = \left\{\left(x_1,\ldots, x_d, \sum_{i=1}^d x_i^2\right)\in \mathds{R}^{d+1}\right\}. \] We now have $d+2$ measures in $\mathds{R}^{d+1}$, and every hyperplane has measure zero in each of them. We can apply \cref{thm:parallel-hyperplanes} and find two parallel hyperplanes $H_1, H_2$ so that the region between them contains exactly half of each measure. Notice that $H_1 \cap \mathcal{P}$ and $H_2 \cap \mathcal{P}$ project onto concentric spheres in $\mathds{R}^d$. \end{proof} Of course, similar results can be obtained by applying general Veronese maps as Stone and Tukey did to prove the polynomial ham sandwich theorem \cite{Stone:1942hu}. \begin{corollary} Let $d,k$ be positive integers. For any set of $\binom{d+k}{k}$ finite absolutely continuous measures in $\mathds{R}^d$, there exists a polynomial $P$ on $d$ variables and constants $\lambda_1, \lambda_2$ so that $\{x \in \mathds{R}^d : \lambda_1 \le P(x) \le \lambda_2\}$ has exactly half of each measure. \end{corollary} The polynomial ham sandwich usually splits $\binom{d+k}{k}-1$ measures using a polynomial of degree at most $k$. If we want to restrict which monomials are used in the splitting polynomial, we just have to reduce the number of measures accordingly. \section{Fixed size partitions for well separated measures}\label{sec:well separated} As noted by B\'ar\'any et al. \cite{Barany:2008vv}, it is known that for well separated convex subsets $K_1, \ldots, K_d$ in $\mathds{R}^d$, there are $2^d$ hyperplanes tangent to all of them. If none of the hyperplanes are vertical (i.e., perpendicular to $e_d$), the tangent hyperplanes are in one-to-one correspondence with the sets of $K_i$ below the hyperplane. This is was also proved by Klee, Lewis, and Hohenbalken \cite{Klee1997}. We use this fact in our proof of \cref{thm:BHJ}. \begin{proof}[Proof of \cref{thm:BHJ}] Consider the hypercube $Q = [0,1]^d$. Each vertex of $Q$ can be assigned to a subset of $I \subset [d]$ uniquely. We denote \[ v_I = (p_1,\ldots, p_d) \qquad \mbox{where } \quad p_i = \begin{cases} 1 & \mbox{ if } i\in I \\ 0 & \mbox{ if } i\not\in I. \end{cases} \] For a point $q = (q_1, \ldots, q_d) \in Q$ and $I \subset [d]$ we consider the coefficients \begin{align*} \lambda_I(q) = \prod_{i \in I}q_i \prod_{i \not\in I} (1-q_i). \end{align*} The coefficients $\lambda_I(q)$ are the coefficients of a convex combination, as they are non-negative and their sum is $1$. Suppose we have a function $f: \{0,1\}^d \to \mathds{R}^d$. We can extend it to a function $\tilde{f}:Q \to \mathds{R}^d$ by mapping \[ q \mapsto \sum_{I \subset [d]} \lambda_I(q) f(v_I). \] Notice that if $\sigma$ is a face of $Q$, then $\tilde{f}(\sigma) \subset \conv(\{f(v_I): v_I \in \sigma\})$. In particular, $\tilde{f}(v_I) = f(v_I)$. Now, suppose we are given $d$ well separated convex sets $K_1, \ldots, K_d$ in $\mathds{R}^d$ and measures $\mu_1, \ldots, \mu_d$ so that the support of $\mu_i$ is $K_i$. We may assume without loss of generality that there is no vertical hyperplane tangent to each of $K_1, \ldots, K_d$. Each non-vertical hyperplane $H$ can be written as \[ \{(x_1, \ldots, x_d) : x_d = \alpha_1 x_1 + \ldots + \alpha_{d-1} x_{d-1} + \alpha_d \} \] for some constants $\alpha_1, \ldots, \alpha_d$. We assign the vector $r(H)=(\alpha_1, \ldots, \alpha_d)$ to the hyperplane $H$. We say that a point is above $H$ if $x_d \ge \alpha_1 x_1 + \ldots + \alpha_{d-1} x_{d-1} + \alpha_d$ and below $H$ if $x_d \le \alpha_1 x_1 + \ldots + \alpha_{d-1} x_{d-1} + \alpha_d$. Notice that if a point $x$ is below a set of hyperplanes $H_1, \ldots, H_k$, then it is also below the hyperplane $r^{-1}(y)$ for any $y \in \conv (r(H_1), \ldots, r(H_k))$. We know that for each subset $I \subset [d]$, there is a unique hyperplane $H_I$ that is tangent to each $K_i$ and so that $K_i$ is below $H$ if and only if $i \in I$. This defines a function $f:\{0,1\}^d \to \mathds{R}^d$ by simply taking $f(v_I) = r(H_I)$. We extend $f$ to a function $\tilde{f}:Q \to \mathds{R}^d$ as described above. For $q \in Q$, let $H(q)$ be the set of points below $r^{-1}(\tilde{f}(q))$. We define a final function \begin{align*} g: Q & \to Q \\ q & \mapsto (\mu_1 (H(q)), \ldots, \mu_d (H(q))) \end{align*} The function $g$ is continuous. Notice, for example, that $g(v_I) = v_I$ for each $I \subset [d]$. Moreover, using the properties of $\tilde{f}$, we have that for every face $\sigma \subset Q$, we have $g(\sigma) \subset \sigma$. This means that $g$ is of degree one on the boundary, so it is surjective. In particular, there is a point $q_0 \in Q$ such that $g(q_0) = (\alpha_1, \ldots, \alpha_d)$. Therefore, the hyperplane $H(q_0)$ is the hyperplane we were looking for. \end{proof} B\'ar\'any, Hubard, and Karasev also showed that under simple conditions the half-space $H$ from \cref{thm:BHJ} is unique. It suffices that \begin{itemize} \item each measure $\mu_i$ assigns a positive value to each open set in its support $K_i$, \item the interior of each $K_i$ is connected and not empty, \item no vertical hyperplane is tangent to $K_1, \ldots, K_d$, and \item the half-space $H$ contains infinite rays in direction $-e_d$. \end{itemize} \begin{proof}[Proof of \cref{thm:parallel-hyperplanes-separated}] We follow a process similar to the proof of \cref{thm:parallel-hyperplanes}. First we need an additional observation about our construction of $S(H)$. In $\mathds{R}^d$, there is no hyperplane $H$ intersecting each of $K_1, \ldots, K_{d+1}$. Otherwise, we can take a point $p_i \in K_i \cap H$. This gives us $d+1$ points in a $(d-1)$-dimensional space, so by Radon's lemma we can find a partition of them into two subsets $A, B$ whose convex hulls intersect. This implies that $\{K_i : p_i \in A\}$ cannot be separated from $\{K_i : p_i \in B\}$, contradicting the hypothesis. Therefore, when we are given an vector $v \in \mathds{R}^d$ and we construct $S(H)$ for $H \perp v$, each side of $H$ must have measure zero for some $\mu_i$. This is much stronger than simply having less than half of some measure. The main idea will be to lift each measure to a surface $S(H)$ for an appropriate $H$ and use \cref{thm:BHJ}. We show that by choosing $H$ carefully, we can deduce the existence of the two parallel hyperplanes we seek. Consider the Stiefel manifold $V_{d-1}(\mathds{R}^d)$. Given $(v_1, \ldots, v_{d-1}) \in V_{d-1}(\mathds{R}^d)$, we lift $\mathds{R}^d$ to $S(H) \subset \mathds{R}^{d+1}$ as in \cref{lem:separating-fixed-direction} where $H$ is parallel to $\operatorname{span}\{v_1,\ldots, v_{d-1}\}$. This lifts each measure $\mu_i$ in $\mathds{R}^d$ to a measure $\sigma_i$ in $S(H)$. Every hyperplane in $\mathds{R}^d$ separating the support vectors of two sets of the measures $\mu_{i}$ can be extended vertically in $\mathds{R}^{d+1}$ to separate the corresponding measures $\sigma_i$. A small tilting can ensure that the separating hyperplane is not vertical. Therefore, the measures $\sigma_i$ are well separated. The measures $\sigma_i$ do not satisfy the requirements of \cref{thm:BHJ}, so an additional step is necessary. For an $\varepsilon>0$, we lift the measures to $S(H)^{\varepsilon}$ as in the proof of \cref{thm:parallel-hyperplanes}, apply \cref{thm:BHJ}, and then take $\varepsilon \to 0$. This ensures that we get a half-space $H^+$ that has infinite rays in the direction $-e_{d+1}$ and such that $\sigma_i(H^+) \ge \alpha_i \cdot \sigma_i(\mathds{R}^d)$ and $\sigma_i(H^-) \ge (1-\alpha_i) \cdot \sigma_i(\mathds{R}^d)$, where $H^-$ is the complementary closed half-space of $H^+$. Since $\sigma_i(H^+)>0$ for all $i$, we know that the boundary of $H^+$ cannot be one of the two hyperplane components of $S(H)$. Therefore the measure $\sigma_i$ of the boundary of $H^+$ is zero for all $i$ and $\sigma_i(H^+) = \alpha_i \cdot \mu_i(H^+)$. The same arguments that B\'ar\'any, Hubard, and Jer\'onimo used to show the uniqueness in their theorem can be applied to show that $H^+$ is uniquely defined. The uniqueness also implies that $H^+$ changes continuously as we modify $(v_1, \ldots, v_{d-1})$. Let $n \in S^d \subset \mathds{R}^{d+1}$ be the normal vector to the boundary of $H^+$ that points in the direction of $H^+$. We can use this to construct a function \begin{align*} g: V_{d-1} (\mathds{R}^{d}) & \to \mathds{R}^{d-1} \times \ldots \times \mathds{R}^{1} \\ (v_1, \ldots, v_{d-1}) & \mapsto (x_1, \ldots, x_d). \end{align*} For each $i=1,\ldots, d-1$, the first coordinate of $x_i \in \mathds{R}^{d-i}$ is $\langle n, (v_i,0) \rangle$ and the rest are zero. This function is well defined and continuous. If we flip the sign of $v_i$, the surface $S(H)$ does not change. The vector $n \in S^d$ is not affected by this change, so only the sign of $x_i$ changes. Therefore, the function $g$ is $(\mathds{Z}_2)^{d-1}$-equivariant. By \cref{thm:topo1}, the function $g$ must have a zero. This implies that the projection of $H^+ \cap S(H)$ onto $\mathds{R}^d$ is the region between two hyperplanes parallel to $H$. \end{proof} The construction of the function $g$ only uses $d-1$ out of the $d(d-1)/2$ coordinates that \cref{thm:topo1} makes available. It would be interesting to know if much stronger conditions can be imposed on $H$. We also have consequences similar to \cref{cor:Bagel}. We say that a family of sets $K_1, \ldots, K_{d+2}$ in $\mathds{R}^d$ is \textit{well separated by spheres} if for any subset way to split them into two families $I, J$, there is a sphere that separates $I$ and $J$, i.e., it contains the union of one of the sets and leaves out the union of the other set. \begin{corollary} Let $d$ be a positive integer and $\mu_1, \ldots, \mu_{d+2}$ be measures in $\mathds{R}^d$ absolutely continuous with respect to the Lebesgue measure. Suppose that the supports $K_1, \ldots, K_{d+2}$ of $\mu_1, \ldots, \mu_{d+2}$ are well separated by spheres. Let $\alpha_1, \ldots, \alpha_{d+2}$ be real numbers in $(0,1)$. Then, there exist two concentric $S_1, S_2$ in $\mathds{R}^d$ so that the region $A$ between them satisfies \[ \mu_i(A) = \alpha_i \cdot \mu_i(\mathds{R}^d) \qquad \mbox{for all }i=1,\ldots, d+2. \] \end{corollary} \begin{proof} We lift $\mathds{R}^d$ to the paraboloid \[ \mathcal{P} = \left\{\left(x_1,\ldots, x_d, \sum_{i=1}^d x_i^2\right)\in \mathds{R}^{d+1}\right\}. \] A sphere in $\mathds{R}^d$ separating two families $I, J$ of measure supports translates to a hyperplane in $\mathds{R}^{d+1}$ separating the lift of those supports. We apply \cref{thm:parallel-hyperplanes-separated} to the family of measures induced on $\mathcal{P}$ and we are done. Even though the set of lifted measures do not satisfy the conditions of \cref{thm:parallel-hyperplanes-separated}, a standard approximation argument fixes this problem. \end{proof} \section{Equipartition with Polytopes and polyhedral surfaces of bounded complexity}\label{sec:polyhedral} In previous sections, the number of measures to be partitioned was constrained by the dimension of the ambient space, while the boundaries of the partition were relatively simple. In this section we consider mass partitions of a family of $n$ measures in $\mathds{R}^d$, where $n$ can be much larger than $d$. We do so by increasing the complexity of the boundary of the partition. We focus on partitions by polyhedral surfaces. \begin{definition}\label{def:nicely-separated} Let $\mathcal{F}=\{\mu_1,\ldots, \mu_n\}$ be a family of finite absolutely continuous measures in $\mathds{R}^d$ with support $K_i$ for each $1\leq i\leq n$. The supports are called \textit{nicely separated} if for each $1\leq i\leq n$, there exists a hyperplane $H_i$ such that $K_i\cap H_i^+=\emptyset$ and $K_j\cap H_i^-=\emptyset$ for all $j\neq i$. \end{definition} The maximum number of well separated measures is $d+1$, due to Radon's theorem. For nicely separated measures we only want to be able to separate any measure for the union of the other $n-1$, and not any two subsets. An example of nicely separated measures are $n$ measures such that each is concentrated near a vertex of a convex polytope. \begin{figure}[h!] \centering \includegraphics[width=1.2\textwidth]{nicely-separated.pdf} \caption{(a) An example of four nicely separated measures in $\mathds{R}^2$. (b) An example of five nicely separated and concentrated measures in $\mathds{R}^2$. Notice that if we take $q_1$ instead of $p_1$ to form the convex hull, the resulting polygon contains all of $K_1$.} \label{fig:nicely-separated} \end{figure} We define a polyhedron in $\mathds{R}^d$ to be a finite intersection of closed half-spaces. A facet of a polyhedron is a $(d-1)$-dimensional face, and a vertex of a polyhedron is a zero dimensional face. \begin{theorem}\label{thm:n-faces} Let $\mathcal{F}=\{\mu_1,\ldots, \mu_n\}$ be a family of finite absolutely continuous measures in $\mathds{R}^d$ with nicely separated supports $K_i$ for all $1\leq i\leq n$, and let $\alpha_1,\ldots,\alpha_{n}$ be real numbers in $(0,1)$. Then, there exists a polyhedron $P$ with at most $n$ facets such that $\mu_i(P)=\alpha_i\cdot \mu_i(\mathds{R}^d)$ for every $1\leq i\leq n$. \end{theorem} \begin{proof} Because the supports are nicely separated, for each $1\leq i \leq n$, we can fix a hyperplane $H_i$ with $K_i\cap H_i^+ = \emptyset$ and $K_j\cap H_i^-=\emptyset$ for all $j\neq i$. Notice that a polyhedron $P = \bigcap_{ i=1}^n H_i^+$ has the property $\mu_i(P) = 0$ for every $1\leq i \leq n$. Now, consider $\mu_1$. Let $v$ be the normal vector to the hyperplane $H_1$ pointing in the direction of $H^-$. We can move $H_1$ in the direction of $v$ until we have the desired portion of the measure $\mu_1$, so we can fix $H_1'\parallel H_1$ with $\mu_1(H_1'^+) = \alpha \cdot \mu_1(\mathds{R}^d)$. By letting $P' = \left(\bigcap_{i=2}^n H_i^+\right) \cap H_1'^+ $, we have $\mu_1(P') = \alpha_1 \cdot \mu_1(\mathds{R}^d)$ because $\mu_1\big(\bigcap_{i=2}^{n}H_i^+\big) = \mu_1(\mathds{R}^d)$. Moreover, because $K_j\cap H_1^-=\emptyset$ for each $j\neq 1$, moving $H_1$ to the direction of $H_1^-$ does not interfere with the rest of the measures $\mu_2,\ldots,\mu_n$. We can repeat the same process for $\mu_2,\ldots,\mu_n$ to find a convex polyhedron of at most $n$ facets with the desired property. \end{proof} While \cref{thm:n-faces} allows for a mass partition with a polyhedron of $n$ facets, we can quantify the complexity of a compact polyhedron by the number of vertices as well. \cref{thm:n-vertices} proves a mass partition with a polyhedron of $n$ vertices, but this time for $n$ measures with a slightly stronger separation condition. We will use a similar idea to the proof of \cref{thm:BHJ}. Let $\mu_1, \ldots, \mu_n$ be a family of nicely separated measures in $\mathds{R}^d$. Let $H_i$ be the hyperplane separating $K_i$ from the rest of the supports, as in \cref{def:nicely-separated}. For $n \ge d+1$, by Helly's theorem we know that $P=\bigcap_{i=1}^n H^+_i \neq \emptyset$. We say that the measures are \textit{concentrated} if the following happens. There exists a point $p \in P$ and points $p_i, q_i$ for $i = 1,\ldots, n$ so that the following holds. \begin{itemize} \item For each $i=1,\ldots,n$, $p_i \in H_i \cap P$. We denote $K_0=\conv\{p_1,\ldots, p_n\}$. \item We have $p \in K_0$. \item For each $i=1,\ldots, n$, $q_i$ is in the ray $pp_i$ and in $\bigcap_{i' \neq i}H^+_{i'}$. \item For each $i=1,\ldots, n$, we have $K_i \subset \conv (\{q_i\}\cup K_0)$. \end{itemize} An example is illustrated in \cref{fig:nicely-separated}(b). \begin{theorem}\label{thm:n-vertices} Let $n,d$ be positive integers. Let $\mathcal{F}=\{\mu_1,\ldots, \mu_n\}$ be a family of nicely separated and concentrated measures in $\mathds{R}^d$, each absolutely continuous. Let $ \alpha_1,\ldots,\alpha_{n}$ be real numbers in $(0,1)$. Then, there exists a polytope $K$ with $n$ vertices such that $\mu_i(K)=\alpha_i\cdot \mu_i(\mathds{R}^d)$ for every $1\leq i\leq n$. \end{theorem} Note that the intuitive idea we used to prove \cref{thm:n-faces} would indicate that we should slide each $p_i$ towards $q_i$ until we have the desired measure. The issue with this is that the values of other measures are no longer fixed. \begin{proof} Consider the hypercube $Q=[0,1]^n$. For $x=(x_1,\ldots, x_n) \in Q$, and $i=1,\ldots, n$, let $y_i = (1-x_i)q_i+x_i p_i$. We define \[ K(x) = \conv \{y_1,\ldots, y_n\}. \] This convex set allows us to construct a function \begin{align*} f: Q & \to Q \\ x & \mapsto \left( \frac{\mu_1(K(x))}{\mu_1(\mathds{R}^d)}, \ldots, \frac{\mu_n(K(x))}{\mu_n(\mathds{R}^d)}\right). \end{align*} The function is continuous. From the conditions of the measures, we can see that for every vertex $v$ of $Q$, we have $f(v) = v$. However, we have a stronger condition. For every face $\sigma \subset Q$, we have $f(\sigma) \subset \sigma$. This is because if a coordinate $x_i$ of $x$ equals zero, the $K(x) \subset H_i^+$, so $\mu_i(K(x)) = 0$. If $x_i=1$, then $K(x) \supset \conv\{\{q_i\}\cup K_0\}$, so $\mu_i(K(x)) = \mu_i (\mathds{R}^d)$. Therefore $f$ is of degree one on the boundary and must be surjective. There is a point $x \in Q$ such that $f(x) = (\alpha_1,\ldots,\alpha_n)$, which implies that $K(x)$ is the polytope we were looking for. \end{proof} \section{remarks and open problems}\label{sec:remarks} To prove \cref{thm:same-fraction}, we need to strengthen \cref{lem:separating-fixed-direction}. \begin{lemma}\label{lem:separating-fixed-direction-strong} Let $m, n$ be positive integers, $\mu_1, \ldots, \mu_{n}$ be $n$ finite absolutely continuous measures in $\mathds{R}^d$, and $v$ be a unit vector in $\mathds{R}^d$. There either exists $m-1$ hyperplanes orthogonal to $v$ that divide $\mathds{R}^d$ into $m$ regions $R_1,\ldots, R_m$ of equal measure for each $\mu_i$ simultaneously or there exist $m-1$ hyperplanes orthogonal to $v$ such that they divide $\mathds{R}^d$ into $m$ regions $R_1, \ldots, R_m$ such that for every $j=1,\ldots, m$ there exists an $i$ such that \begin{align*} \mu_i (R_j) & < \frac{1}{m}\mu_i(\mathds{R}^d). \end{align*} \end{lemma} \begin{proof} Given parallel hyperplanes $H_1, \ldots, H_{m-1}$ in this order, we denote by $R_1, \ldots, R_m$ the regions they divide $\mathds{R}^d$ into such that $R_j$ is bounded by $H_{j-1}$ and $H_j$. The unbounded regions $R_1$, $R_m$ are bounded only by $H_1$ and $H_m$ respectively. We can find $m-1$ hyperplanes such that $\mu_1(R_j) = (1/m)\mu_1(\mathds{R}^d)$ for every $j$. If these regions also form an equipartition for every other $\mu_i$, we are done. Otherwise, there is an $i$ and a $j$ such that $\mu_i(R_j) < (1/m)\mu_i(\mathds{R}^d)$. We can widen $R_j$ by moving $R_{j-1}$ and $R_j$ slightly apart so that we still have $\mu_i(R_j) < (1/m)\mu_i(\mathds{R}^d)$. Then, $\mu_1(R_{j-1}) < (1/m) \mu_1(\mathds{R}^d)$ and $\mu_1(R_{j+1}) < (1/m) \mu_1(\mathds{R}^d)$. We can translate $H_{j-2}$ and $H_{j+1}$ away from $H_{j-1}$ and $H_j$ respectively so that these inequalities are preserved. This makes $\mu_1(R_{j-2})$ and $\mu_1(R_{j+2})$ to be strictly reduced. We continue this way until we are done. \end{proof} Now, given $\mu_1, \ldots, \mu_{d+1}$ finite absolutely continuous measures in $\mathds{R}^d$, we construct a surface in $\mathds{R}^{d+1}$. We take $v=e_{d}$ and find the $m-1$ hyperplanes $H_1, \ldots, H_m$ such that \[ H_j = \{(x_1,\ldots, x_d) \in \mathds{R}^d: x_d = \lambda_j\}. \] For some $\lambda_1 < \ldots < \lambda_{m-1}$. We define $\lambda_0 =-\infty$ and $\lambda_m = \infty$. Let $h:\mathds{R} \to \mathds{R}$ be a convex function that is linear between $\lambda_j$ and $\lambda_{j+1}$ for each $j=0,\ldots, {m-1}$, but not between $\lambda_j$ and $\lambda_{j+2}$ for each $j=0,\ldots, m-2$. Let $V$ be the surface in $\mathds{R}^{d+1}$ defined by the equation $x_{d+1}=h(x_d)$. The set of points on or above $V$ is the intersection of $m$ closed half-spaces. To prove \cref{thm:same-fraction} we repeat the proof of Akopyan and Karasev but we lift $\mathds{R}^d$ to $V$ instead of a paraboloid. \begin{proof}[Proof of \cref{thm:same-fraction}] By a subdivision argument, it suffices to prove the result when $n=p$ a prime number. We apply \cref{lem:separating-fixed-direction-strong} with $m=p$. If there are $p-1$ parallel hyperplanes that form an equipartition of the measures, we are done. Otherwise, we lift $\mathds{R}^d$ to $\mathds{R}^{d+1}$ by lifting it to the surface $V$ defined above. Let $\sigma_1, \ldots, \sigma_{d+1}$ be the measures induced by $\mu_1,\ldots, \mu_{d+1}$ on $V$. It's known that we can split $\mathds{R}^{d+1}$ into $p$ convex sets $C_1, \ldots, C_p$ that form an equipartition of $\mu_1, \ldots, \mu_{d+1}$. Since each of the regions $R_1, \ldots, R_p$ we constructed in $\mathds{R}^{d}$ have less than a $(1/p)$-fraction of some $\mu_i$ and $V$ is the boundary of a convex set, none of the boundaries between the sets $C_{j}$ can coincide with the hyperplanes defining $V$. Moreover, the sets are $C_1, \ldots, C_p$ are induced by a generalized Voronoi diagram \cites{Karasev:2014gi, Blagojevic:2014ey}. In other words, there are points (called sites) $s_1, \ldots, s_p$ in $\mathds{R}^{d+1}$ and real number $\beta_1, \ldots, \beta_p$ such that the $p$ convex regions \[ C_j = \{x \in \mathds{R}^{d+1}: ||x-s_j||^2-\beta_j \le ||x - s_{j'}||^2-\beta_{j'} \mbox{ for } j' =1,\ldots, p\} \] form an equipartition of $\mu_1, \ldots, \mu_{d+1}$. Since the set of point above $V$ is convex, if we take the region $C_j$ whose site $s_j$ has minimal $(d+1)$-th coordinate, when we project $C_j \cap V$ back to $\mathds{R}^d$ we get a convex set. This is the set $K$ we are looking for. The boundary of the corresponding $C_j$ is the union of at most $p-1$ hyperplanes (the ones dividing it from each other $C_{j'}$). Each of those $p-1$ hyperplanes can intersect each of the $p$ hyperplanes defining $V$, forming at most $p(p-1)$ linear components of the boundary of $C_j \cap V$. This gives us the bound on the number of half-spaces whose intersection is $K$. \end{proof} Earlier proofs of the equipartition result we used do not guarantee that the partition $C_1, \ldots, C_p$ comes directly from a generalized Voronoi diagram \cite{Soberon:2012kp}, which is important in this proof. When $n$ is a prime power, the number of half-spaces we used grows logarithmically with $n$. We wonder if this holds in general. \begin{question} Let $d$ be a fixed integer. Determine if for every positive integer $n$ and any $d+1$ finite absolutely continuous measures $\mu_1, \ldots, \mu_{d+1}$ in $\mathds{R}^d$ there exists a convex set $K \subset \mathds{R}^d$ formed by the intersection of $O(\log n)$ half-spaces that contains exactly a $(1/n)$-fraction of each $\mu_i$. \end{question} We nicknamed \cref{cor:Bagel} the bagel ham sandwich theorem due to its drawing in $\mathds{R}^2$. However, since the set used is the region between two concentric spheres, it certainly does not look like a Bagel in $\mathds{R}^3$. We define a \textit{regular torus} in $\mathds{R}^3$ to be the any set of the form $\{x \in \mathds{R}^3: \operatorname{dist}(x,S)\le \alpha\}$ where $S$ is a flat circle in $\mathds{R}^3$ and $\alpha$ is a positive real number. \begin{question}[Three-dimensional bagels] Is is true that for any five absolutely continuous finite measures in $\mathds{R}^3$ there exists a regular torus containing exactly half of each measure? \end{question} With four measures the result holds, since when $S$ degenerates to a point the regular torus is a sphere. One of the questions that motivated the work on this manuscript was inspired by a conjecture by Mikio Kano. Kano conjectured that for any $n$ smooth measures in $\mathds{R}^2$ there exists a path formed only by horizontal and vertical segments, that takes at most $n-1$ turns, that simultaneously halves each measure. The conjecture is only known for $k=1,2$ or if the path is allowed to go through infinity \cites{Uno:2009wk, Karasev:2016cn}. We wonder if the following way to mix Kano's conjecture with \cref{thm:circular-sandwich} holds. \begin{question}[Existence of square sandwiches]\label{question:square} Is is true that for any three finite absolutely continuous measures in $\mathds{R}^2$ there exists a square that contains exactly half of each measure? \end{question} \cref{thm:parallel-hyperplanes} shows that we have a positive answer for rectangles (if the support of the measures are compact, we can cut the two lines given by \cref{thm:parallel-hyperplanes} by perpendicular segments sufficiently far away, otherwise we have degenerate rectangles). However, it is still possible that for squares the answer to \cref{question:square} is affirmative. \begin{bibdiv} \begin{biblist} \bib{Akopyan:2013jt}{article}{ author={Akopyan, Arseniy}, author={Karasev, Roman~N.}, title={{Cutting the Same Fraction of Several Measures}}, date={2013}, journal={Discrete Comput. Geom.}, volume={49}, number={2}, pages={402\ndash 410}, } \bib{Blagojevic:2007ij}{article}{ author={Blagojević, Pavle V.~M.}, author={Blagojević, Aleksandra~Dimitrijević}, title={{Using equivariant obstruction theory in combinatorial geometry}}, date={2007}, journal={Topology Appl.}, volume={154}, number={14}, pages={2635\ndash 2655}, } \bib{Blagojevic2018}{article}{ author={Blagojevi{\'c}, Pavle V.~M.}, author={Blagojevi{\'c}, Aleksandra~Dimitrijevi{\'c}}, author={Karasev, Roman}, author={Kliem, Jonathan}, title={{More bisections by hyperplane arrangements}}, date={2018}, journal={arXiv preprint arXiv:1809.05364}, volume={math.MG}, } \bib{Bereg:2005voa}{article}{ author={Bereg, Sergey}, title={{Equipartitions of Measures by 2-Fans}}, date={2005}, journal={Discrete Comput. Geom.}, volume={34}, number={1}, pages={87\ndash 96}, } \bib{Barany:2008vv}{article}{ author={B{\'a}r{\'a}ny, Imre}, author={Hubard, Alfredo}, author={Jerónimo, Jesús}, title={{Slicing Convex Sets and Measures by a Hyperplane}}, date={2008}, ISSN={0179-5376}, journal={Discrete Comput. Geom.}, volume={39}, number={1-3}, pages={67\ndash 75}, } \bib{Blagojevic2015}{article}{ author={Blagojevi\'{c}, Pavle V.~M.}, author={L\"{u}ck, Wolfgang}, author={Ziegler, G\"{u}nter~M.}, title={Equivariant topology of configuration spaces}, date={2015}, ISSN={1753-8416}, journal={J. Topol.}, volume={8}, number={2}, pages={414\ndash 456}, url={https://doi.org/10.1112/jtopol/jtv002}, review={\MR{3356767}}, } \bib{Barany:2001fs}{article}{ author={B{\'a}r{\'a}ny, Imre}, author={Matou\v{s}ek, Ji\v{r}\'i}, title={{Simultaneous partitions of measures by K-fans}}, date={2001}, journal={Discrete Comput. Geom.}, volume={25}, number={3}, pages={317\ndash 334}, } \bib{Barba:2019to}{article}{ author={Barba, Luis}, author={Pilz, Alexander}, author={Schnider, Patrick}, title={{Sharing a pizza: bisecting masses with two cuts}}, date={2019}, journal={arXiv preprint arXiv:1904.02502}, volume={cs.CG}, } \bib{Breuer2010}{article}{ author={Breuer, Felix}, title={{Uneven Splitting of Ham Sandwiches}}, date={2010}, ISSN={0179-5376}, journal={Discrete Comput. Geom.}, volume={43}, number={4}, pages={876\ndash 892}, } \bib{Blagojevic:2014ey}{article}{ author={Blagojević, Pavle V.~M.}, author={Ziegler, G\"unter~M.}, title={{Convex equipartitions via Equivariant Obstruction Theory}}, date={2014}, journal={Israel J. Math.}, volume={200}, number={1}, pages={49\ndash 77}, } \bib{Chan2020}{article}{ author={Chan, Yu~Hin}, author={Chen, Shujian}, author={Frick, Florian}, author={Hull, J.~Tristan}, title={{Borsuk-Ulam theorems for products of spheres and Stiefel manifolds revisited}}, date={2020}, journal={Topol. Methods Nonlinear Anal.}, volume={55}, number={2}, pages={553\ndash 564}, } \bib{Fadell:1988tm}{article}{ author={Fadell, Edward}, author={Husseini, Sufian}, title={{An ideal-valued cohomological index theory with applications to Borsuk—Ulam and Bourgin—Yang theorems}}, date={1988}, journal={Ergodic Theory Dynam. Systems}, volume={8}, pages={73\ndash 85}, } \bib{Hubard2020}{article}{ author={Hubard, Alfredo}, author={Karasev, Roman}, title={Bisecting measures with hyperplane arrangements}, date={2020}, journal={Math. Proc. Cambridge Philos. Soc.}, volume={169}, number={3}, pages={639\ndash 647}, } \bib{Karasev:2014gi}{article}{ author={Karasev, Roman~N.}, author={Hubard, Alfredo}, author={Aronov, Boris}, title={{Convex equipartitions: the spicy chicken theorem}}, date={2014}, journal={Geom. Dedicata}, volume={170}, number={1}, pages={263\ndash 279}, } \bib{Klee1997}{article}{ author={Klee, Victor}, author={Lewis, Ted}, author={Von~Hohenbalken, Balder}, title={Appollonius revisited: supporting spheres for sundered systems}, date={1997}, ISSN={0179-5376}, journal={Discrete Comput. Geom.}, volume={18}, number={4}, pages={385\ndash 395}, url={https://doi.org/10.1007/PL00009324}, } \bib{Karasev:2016cn}{article}{ author={Karasev, Roman~N.}, author={Roldán-Pensado, Edgardo}, author={Soberón, Pablo}, title={{Measure partitions using hyperplanes with fixed directions}}, date={2016}, journal={Israel J. Math.}, volume={212}, number={2}, pages={705\ndash 728}, } \bib{matousek2003using}{book}{ author={Matou\v{s}ek, Ji\v{r}\'{\i}}, title={Using the {B}orsuk-{U}lam theorem: Lectures on topological methods in combinatorics and geometry}, series={Universitext}, publisher={Springer-Verlag, Berlin}, date={2003}, ISBN={3-540-00362-2}, } \bib{Manta2021}{article}{ author={Manta, Michael~N.}, author={Sober{\'o}n, Pablo}, title={{Generalizations of the Yao--Yao partition theorem and the central transversal theorem}}, date={2021}, journal={arXiv preprint arXiv:2107.06233}, volume={math.CO}, } \bib{Mus12}{article}{ author={Musin, Oleg}, title={{Borsuk--Ulam type theorems for manifolds}}, date={2012}, journal={Proc. Amer. Math. Soc.}, volume={140}, number={7}, pages={2551\ndash 2560}, } \bib{RoldanPensado2021}{article}{ author={R{old{\'a}n-Pensado}, Edgardo}, author={Sober{\'o}n, Pablo}, title={A survey of mass partitions}, date={2021}, journal={Bull. Amer. Math. Soc.}, note={Electronically published on February 24, 2021, DOI: https://doi.org/10.1090/bull/1725 (to appear in print).}, } \bib{Schnider:2019ua}{article}{ author={Schnider, Patrick}, title={{Equipartitions with Wedges and Cones}}, date={2019}, journal={arXiv preprint arXiv:1910.13352}, volume={cs.CG}, } \bib{Soberon:2012kp}{article}{ author={Sober{\'o}n, Pablo}, title={{Balanced Convex Partitions of Measures in $R^d$}}, date={2012}, journal={Mathematika}, volume={58}, number={01}, pages={71\ndash 76}, } \bib{Stone:1942hu}{article}{ author={Stone, Arthur~H.}, author={Tukey, John~W.}, title={{Generalized “sandwich” theorems}}, date={1942}, ISSN={0012-7094}, journal={Duke Math. J.}, volume={9}, number={2}, pages={356\ndash 359}, } \bib{Steinhaus1938}{article}{ author={Steinhaus, Hugo}, title={A note on the ham sandwich theorem}, date={1938}, journal={Mathesis Polska}, volume={9}, pages={26\ndash 28}, } \bib{Steinhaus1945}{article}{ author={Steinhaus, Hugo}, title={Sur la division des ensembles de l'espace par les plans et des ensembles plans par les cercles}, date={1945}, journal={Fund. Math.}, volume={33}, number={1}, pages={245\ndash 263}, } \bib{Uno:2009wk}{article}{ author={Uno, Miyuki}, author={Kawano, Tomoharu}, author={Kano, Mikio}, title={{Bisections of two sets of points in the plane lattice}}, date={2009}, journal={IEICE Transactions on Fundamentals of Electronics, Communications and Computer Sciences}, volume={92}, number={2}, pages={502\ndash 507}, } \bib{Zivaljevic2017}{incollection}{ author={{\v{Z}}ivaljevi{\'c}, Rade~T.}, title={Topological methods in discrete geometry}, date={2017}, booktitle={{Handbook of Discrete and Computational Geometry}}, edition={Third}, publisher={CRC Press}, pages={551\ndash 580}, } \end{biblist} \end{bibdiv} \end{document}
{ "timestamp": "2021-09-09T02:24:48", "yymm": "2109", "arxiv_id": "2109.03749", "language": "en", "url": "https://arxiv.org/abs/2109.03749", "abstract": "Many results in mass partitions are proved by lifting $\\mathbb{R}^d$ to a higher-dimensional space and dividing the higher-dimensional space into pieces. We extend such methods to use lifting arguments to polyhedral surfaces. Among other results, we prove the existence of equipartitions of $d+1$ measures in $\\mathbb{R}^d$ by parallel hyperplanes and of $d+2$ measures in $\\mathbb{R}^d$ by concentric spheres. For measures whose supports are sufficiently well separated, we prove results where one can cut a fixed (possibly different) fraction of each measure either by parallel hyperplanes, concentric spheres, convex polyhedral surfaces of few facets, or convex polytopes with few vertices.", "subjects": "Combinatorics (math.CO)", "title": "Lifting methods in mass partition problems", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9904405998064016, "lm_q2_score": 0.8267117983401363, "lm_q1q2_score": 0.8188089294150336 }
https://arxiv.org/abs/1812.10850
Decomposition of Gaussian processes, and factorization of positive definite kernels
We establish a duality for two factorization questions, one for general positive definite (p.d) kernels $K$, and the other for Gaussian processes, say $V$. The latter notion, for Gaussian processes is stated via Ito-integration. Our approach to factorization for p.d. kernels is intuitively motivated by matrix factorizations, but in infinite dimensions, subtle measure theoretic issues must be addressed. Consider a given p.d. kernel $K$, presented as a covariance kernel for a Gaussian process $V$. We then give an explicit duality for these two seemingly different notions of factorization, for p.d. kernel $K$, vs for Gaussian process $V$. Our result is in the form of an explicit correspondence. It states that the analytic data which determine the variety of factorizations for $K$ is the exact same as that which yield factorizations for $V$. Examples and applications are included: point-processes, sampling schemes, constructive discretization, graph-Laplacians, and boundary-value problems.
\section{Introduction} We give an integrated approach to positive definite (p.d.) kernels and Gaussian processes, with an emphasis on factorizations, and their applications. Positive definite kernels serve as powerful tools in such diverse areas as Fourier analysis, probability theory, stochastic processes, boundary theory, potential theory, approximation theory, interpolation, signal/image analysis, operator theory, spectral theory, mathematical physics, representation theory, complex function-theory, moment problems, integral equations, numerical analysis, boundary-value problems for partial differential equations, machine learning, geometric embedding problems, and information theory. While there is no single book which covers all these applications, the reference \cite{MR3526117} goes some of the way. As for the use of RKHS analysis in machine learning, we refer to \cite{MR2327597} and \cite{MR3236858}. Here, we give a new and explicit duality for positive definite functions (kernels) on the one hand, and Gaussian processes on the other. A covariance kernel for a general stochastic process is positive definite. In general, the stochastic process in question is not determined by its covariance kernel. But in the special case when the process is Gaussian, it is. In fact (\thmref{C1}), every p.d. kernel $K$ is indeed the covariance kernel of a Gaussian process. The construction is natural; starting with the p.d. kernel $K$, there is a canonical inductive limit construction leading to the Gaussian process for this problem, following a realization of Gaussian processes dating back to Kolmogorov. The interplay between analytic properties of p.d. kernels and their associated Gaussian processes is the focus of our present study. We formulate two different factorization questions, one for general p.d. kernels $K$, and the other for Gaussian processes, say $V$. The latter notion, for Gaussian processes, is a subordination approach. Our approach to factorization for p.d. kernels is directly motivated by matrix factorizations, but in infinite dimensions, there are subtle measure theoretic issues involved. If the given p.d. kernel $K$ is already presented as a covariance kernel for a Gaussian process $V$, we then give an explicit duality for these two seemingly different notions of factorization. Our main result, \thmref{E1}, states that the analytic data which determine the variety of factorizations for $K$ is the exact same as that which yield factorizations for $V$. \section{\label{sec:pdk}Positive definite kernels} The notion of a positive definite (p.d.) kernel has come to serve as a versatile tool in a host of problems in pure and applied mathematics. The abstract notion of a p.d. kernel is in fact a generalization of that of a positive definite function, or a positive-definite matrix. Indeed, the matrix-point of view lends itself naturally to the particular factorization question which we shall address in \secref{fac} below. The general idea of p.d. kernels arose first in various special cases in the first half of 20th century: It occurs in work by J. Mercer in the context of solving integral operator equations; in the work of G. Szeg\H{o} and S. Bergmann in the study of harmonic analysis and the theory of complex domains; and in the work by N. Aronszajn in boundary value problems for PDEs. It was Aronszajn who introduced the natural notion of reproducing kernel Hilbert space (RKHS) which will play a central role here; see especially (\ref{eq:A4}) below. References covering the areas mentioned above include: \cite{MR3687240,MR0051437,MR562914,IM65,jorgensen2018harmonic,MR0277027,MR3882025}, and \cite{MR3721329}. Right up to the present, p.d. kernels have arisen as powerful tools in many and diverse areas of mathematics. A partial list includes the areas listed above in the Introduction. An important new area of application of RKHS theory includes the following \cite{MR1120274,MR1200633,MR1473250,MR1821907,MR1873434,MR1986785,MR2223568,MR2373103}. \subsection*{Positive definite kernels and their reproducing kernel Hilbert spaces} Let $X$ be a set and let $K$ be a complex valued function on $X\times X$. We say that $K$ is \emph{positive definite} (p.d.) iff (Def.) for all finite subset $F$ ($\subset X$) and complex numbers $\left(\xi_{x}\right)_{x\in F}$, we have: \begin{equation} \sum_{x\in F}\sum_{y\in F}\overline{\xi}_{x}\xi_{y}K\left(x,y\right)\geq0.\label{eq:A1} \end{equation} In other words, the $\left|F\right|\times\left|F\right|$ matrix $\left(K\left(x,y\right)\right)_{F\times F}$ is positive definite in the usual sense of linear algebra. We refer to the rich literature regarding theory and applications of p.d. functions \cite{MR2966130,MR3507188,HKL14,RAKK05,MR3290453,MR3046303,MR2982692}. We shall also need the Aronszajn \cite{MR0051437} reproducing kernel Hilbert spaces (R.K.H.S.), denoted $\mathscr{H}\left(K\right)$: It is the Hilbert completion of all functions \begin{equation} \sum_{x\in F}\xi_{x}K\left(\cdot,x\right)\label{eq:A2} \end{equation} where $F$, and $\left(\xi\right)_{x\in F}$, are as above. If $F$ (finite) is fixed, and $\left(\xi_{x}\right)_{x\in F}$, $\left(\eta_{x}\right)_{x\in F}$ are vectors in $\mathbb{C}^{\left|F\right|}$, we set \begin{equation} \left\langle \sum\nolimits _{x\in F}\xi_{x}K\left(\cdot,x\right),\sum\nolimits _{y\in F}\eta_{y}K\left(\cdot,y\right)\right\rangle _{\mathscr{H}\left(K\right)}:=\sum\sum\nolimits _{F\times F}\overline{\xi}_{x}\eta_{y}K\left(x,y\right).\label{eq:A3} \end{equation} With the definition of the R.K.H.S. $\mathscr{H}\left(K\right)$, we get directly that the functions $\left\{ K\left(\cdot,x\right)\right\} _{x\in X}$ are automatically in $\mathscr{H}\left(K\right)$; and that, for all $h\in\mathscr{H}\left(K\right)$, we have \begin{equation} \left\langle K\left(\cdot,x\right),h\right\rangle _{\mathscr{H}\left(K\right)}=h\left(x\right);\label{eq:A4} \end{equation} i.e., the \emph{reproducing} property holds. Further recall (see e.g. \cite{MR3526117}) that, given $K$, then the R.K.H.S. $\mathscr{H}\left(K\right)$ is determined uniquely, up to isometric isomorphism in Hilbert space. \begin{lem} \label{lem:B1}Let $X\times X\xrightarrow{\;K\;}\mathbb{C}$ be a p.d. kernel, and let $\mathscr{H}\left(K\right)$ be the corresponding RKHS (see (\ref{eq:A3})-(\ref{eq:A4})). Let $h$ be a function defined on $X$; then TFAE: \begin{enumerate} \item \label{enu:LemA2-1}$h\in\mathscr{H}\left(K\right)$; \item \label{enu:LemA2-2}there is a constant $C=C_{h}<\infty$ such that, for all finite subset $F\subset X$, and all $\left(\xi_{x}\right)_{x\in F}$, $\xi_{x}\in\mathbb{C}$, the following \uline{a priori} estimate holds: \begin{equation} \left|\sum\nolimits _{x\in F}\xi_{x}h\left(x\right)\right|^{2}\leq C_{h}\sum\nolimits _{x\in F}\sum\nolimits _{y\in F}\overline{\xi}_{x}\xi_{y}K\left(x,y\right).\label{eq:B5} \end{equation} \end{enumerate} \end{lem} \begin{proof} The implication (\ref{enu:LemA2-1})$\Rightarrow$(\ref{enu:LemA2-2}) is immediate, and in this case, we may take $C_{h}=\left\Vert h\right\Vert _{\mathscr{H}\left(K\right)}^{2}$. Now for the converse, assume (\ref{enu:LemA2-2}) holds for some finite constant. On the $\mathscr{H}\left(K\right)$-dense span in (\ref{eq:A2}), define a linear functional \begin{equation} L_{h}\left(\sum\nolimits _{x\in F}\xi_{x}K\left(\cdot,x\right)\right):=\sum\nolimits _{x\in F}\xi_{x}h\left(x\right).\label{eq:B6} \end{equation} From the assumption (\ref{eq:B5}) in (\ref{enu:LemA2-2}), we conclude that $L_{h}$ (in (\ref{eq:B6})) is a well defined bounded linear functional on $\mathscr{H}\left(K\right)$. Initially, $L_{h}$ is only defined on the span (\ref{eq:A2}), but by (\ref{eq:B5}), it is bounded, and so extends uniquely by $\mathscr{H}\left(K\right)$-norm limits. We may therefore apply Riesz' lemma to the Hilbert space $\mathscr{H}\left(K\right)$, and conclude that there is a unique $H\in\mathscr{H}\left(K\right)$ such that \begin{equation} L_{h}\left(\psi\right)=\left\langle \psi,H\right\rangle _{\mathscr{H}\left(K\right)}\label{eq:B7} \end{equation} for all $\psi\in\mathscr{H}\left(K\right)$. Now, setting $\psi\left(\cdot\right):=K\left(\cdot,x\right)$, for $x\in X$, we conclude from (\ref{eq:B7}) that $h\left(x\right)=H\left(x\right)$; and so $h\in\mathscr{H}\left(K\right)$, proving (\ref{enu:LemA2-1}). \end{proof} \section{\label{sec:gp}Gaussian processes} The interest in positive definite (p.d.) functions has at least three roots: (i) Fourier analysis, and harmonic analysis more generally; (ii) Optimization and approximation problems, involving for example spline approximations as envisioned by I. Sch\"oenberg; and (iii) Stochastic processes. See \cite{MR0004644,MR709376}. Below, we sketch a few details regarding (iii). A \emph{stochastic process} is an indexed family of random variables based on a fixed probability space. In some cases, the processes will be indexed by some group $G$, or by a subset of $G$. For example, $G=\mathbb{R}$, or $G=\mathbb{Z}$, correspond to processes indexed by real time, respectively discrete time. A main tool in the analysis of stochastic processes is an associated \emph{covariance function}. A process $\{X_{g}\mid g\in G\}$ is called \emph{Gaussian} if each random variable $X_{g}$ is Gaussian, i.e., its distribution is Gaussian. For Gaussian processes, we only need two moments. So if we normalize, setting the mean equal to $0$, then the process is determined by its covariance function. In general, the covariance function is a function on $G\times G$, or on a subset, but if the process is \emph{stationary}, the covariance function will in fact be a p.d. function defined on $G$, or a subset of $G$. For a systematic study of positive definite functions on groups $G$, on subsets of groups, and the variety of the extensions to p.d. functions on $G$, see e.g. \cite{MR3559001}. By a theorem of Kolmogorov \cite{MR735967}, every Hilbert space may be realized as a (Gaussian) reproducing kernel Hilbert space (RKHS), see \thmref{C1} below, and also \cite{PaSc75,IM65,NF10}. Now every positive definite kernel is also the covariance kernel of a Gaussian process; a fact which is a point of departure in our present analysis: Given a positive definite kernel, we shall explore its use in the analysis of the associated Gaussian process; and vice versa. This point of view is especially fruitful when one is dealing with problems from stochastic analysis. Even restricting to stochastic analysis, we have the exciting area of applications to statistical learning theory \cite{MR2327597,MR3236858}.\\ Let $\left(\Omega,\mathscr{F},\mathbb{P}\right)$ be a \emph{probability space}, i.e., $\Omega$ is a fixed set (sample space), $\mathscr{F}$ is a specified sigma-algebra (events) of subsets in $\Omega$, and $\mathbb{P}$ is a probability measure on $\mathscr{F}$. A Gaussian random variable is a function $V:\Omega\rightarrow\mathbb{R}$ (in the real case), or $V:\Omega\rightarrow\mathbb{C}$, such that $V$ is measurable with respect to the sigma-algebra $\mathscr{F}$ on $\Omega$, and the corresponding sigma-algebra of Borel subsets in $\mathbb{R}$ (or in $\mathbb{C}$). Let $\mathbb{E}$ denote the expectation defined from $\mathbb{P}$, i.e., \begin{equation} \mathbb{E}\left(\cdots\right)=\int_{\Omega}\left(\cdots\right)d\mathbb{P}.\label{eq:A5} \end{equation} The requirement on $V$ is that its distribution is Gaussian. If $g$ denotes a Gaussian on $\mathbb{R}$ (or on $\mathbb{C}$), the requirement is that \begin{equation} \mathbb{E}\left(f\circ V\right)=\int_{\mathbb{R}\left(\text{or \ensuremath{\mathbb{C}}}\right)}f\,dg;\label{eq:A6} \end{equation} or equivalently \begin{equation} \mathbb{P}\left(V\in B\right)=\int_{B}dg=g\left(B\right)\label{eq:A7} \end{equation} for all Borel sets $B$; see \figref{G1}. \begin{figure} \includegraphics[width=0.35\textwidth]{gr} \caption{\label{fig:G1}A Gaussian random variable and its distribution, see (\ref{eq:A7}).} \end{figure} If $N\in\mathbb{N}$, and $V_{1},\cdots,V_{N}$ are random variables, the Gaussian requirement is (see \figref{G2}) that the joint distribution of $\left(V_{1},\cdots,V_{N}\right)$ is an $N$-dimensional Gaussian, say $g_{N}$, so if $B\subset\mathbb{R}^{N}$ then \begin{equation} \mathbb{P}\left(\left(V_{1},\cdots,V_{N}\right)\in B\right)=g_{N}\left(B\right).\label{eq:A8} \end{equation} \begin{figure} \includegraphics[width=0.35\textwidth]{grn} \caption{\label{fig:G2} A Gaussian system and its joint distribution, see (\ref{eq:A8}).} \end{figure} For our present purpose we may restrict to the case where the mean (of the respective Gaussians) is assumed zero. In that case, a finite joint distribution is determined by its covariance matrix. In the $\mathbb{R}^{N}$ case, it is specified as follows (the extension to $\mathbb{C}^{N}$ is immediate) $\left(G_{N}\left(j_{1},j_{2}\right)\right)_{j_{1},j_{2}=1}^{N}$, \begin{equation} G_{N}\left(j_{1},j_{2}\right)=\int_{\mathbb{R}^{N}}x_{j_{1}}x_{j_{2}}g_{N}\left(x_{1},\cdots,x_{N}\right)dx_{1}\cdots dx_{N}\label{eq:A9} \end{equation} where $dx_{1}\cdots dx_{N}=\lambda_{N}$ denotes the standard Lebesgue measure on $\mathbb{R}^{N}$. The following is known: \begin{thm}[Kolmogorov \cite{MR0133175}, see also \cite{MR562914,MR1176778}] \label{thm:C1}A kernel $K:X\times X\rightarrow\mathbb{C}$ is positive definite if and only if there is a (mean zero) Gaussian process $\left(V_{x}\right)_{x\in X}$ indexed by $X$ such that \begin{equation} \mathbb{E}\left(\overline{V}_{x}V_{y}\right)=K\left(x,y\right)\label{eq:A10} \end{equation} where $\overline{V}_{x}$ denotes complex conjugation. Moreover (see Hida \cite{MR0301806,MR1176778}), the process in (\ref{eq:A10}) is uniquely determined by the kernel $K$ in question. If $F\subset X$ is finite, then the covariance kernel for $\left(V_{x}\right)_{x\in F}$ is $K_{F}$ given by \begin{equation} K_{F}\left(x,y\right)=G_{F}\left(x,y\right),\label{eq:A11} \end{equation} for all $x,y\in F$, see (\ref{eq:A9}) above. \end{thm} In the subsequent sections, we shall address a number of properties of Gaussian processes important for their stochastic calculus. Our analysis deals with both the general case, and particular examples from applications. We begin in \secref{sms} with certain Wiener processes which are indexed by sigma-finite measures. For this class, the corresponding p.d. kernel has a special form; see (\ref{eq:C1}) in \defref{D1}. (The case of fractal measures is part of \secref{exa} below.) In \secref{fac}, we address the general case: We prove our duality result for factorization, \thmref{E1}. The remaining sections are devoted to examples and applications. \section{\label{sec:sms}Sigma-finite measure spaces and Gaussian processes} We shall consider functions of $\sigma$-finite measure space $\left(M,\mathscr{F}_{M},\mu\right)$ where $M$ is a set, $\mathscr{F}_{M}$ a $\sigma$-algebra of subsets in $M$, and $\mu$ is a positive measure defined on $\mathscr{F}_{M}$. It is further assumed that there is a countably indexed $\left(A_{i}\right)_{i\in\mathbb{N}}$ s.t. $0<\mu\left(A_{i}\right)<\infty$, $M=\cup_{i}A_{i}$; and further that the measure space $\left(M,\mathscr{F}_{M},\mu\right)$ is complete; so the Radon-Nikodym theorem holds. We shall also restrict to the case when $\mu$ is assumed non-atomic. The case when $\mu$ is atomic is different, and is addressed in \secref{atomic} below. \begin{defn} \label{def:D1}Set \[ \mathscr{F}_{fin}=\left\{ A\in\mathscr{F}_{M}\mid0<\mu\left(A\right)<\infty\right\} . \] \end{defn} Note then \begin{equation} K^{\left(\mu\right)}\left(A,B\right)=\mu\left(A\cap B\right),\;A,B\in\mathscr{F}_{fin}\label{eq:C1} \end{equation} is positive definite. The corresponding Gaussian process $(W_{A}^{\left(\mu\right)})_{A\in\mathscr{F}_{fin}}$ is called the Wiener process \cite{MR0301806,MR1176778}. In particular, we have \begin{equation} \mathbb{E}\left(W_{A}^{\left(\mu\right)}W_{B}^{\left(\mu\right)}\right)=\mu\left(A\cap B\right),\label{eq:C2} \end{equation} and \begin{equation} \lim_{\left(A_{i}\right)}\sum_{i}\left(W_{A_{i}}^{\left(\mu\right)}\right)^{2}=\mu\left(A\right).\label{eq:C3} \end{equation} The precise limit in (\ref{eq:C3}), quadratic variation, is as follows: Given $\mu$ as above, and $A\in\mathscr{F}_{fin}$, we then take limit over the filter of all partitions of $A$ (see (\ref{eq:C4})) relative to the standard notation of refinement: \begin{equation} A=\cup_{i}A_{i},\;A_{i}\cap A_{j}=\emptyset\;\text{if \ensuremath{i\neq j}},\;\text{and}\;\lim\mu\left(A_{i}\right)=0.\label{eq:C4} \end{equation} Details: Let $\left(\Omega,Cyl,\mathbb{P}\right)$, $\mathbb{P}=\mathbb{P}^{\left(\mu\right)}$ be the probability space which realizes $W^{\left(\mu\right)}$ as a Gaussian process (or generalized Wiener process), i.e., s.t. (\ref{eq:C2}) holds for all pairs in $\mathscr{F}_{fin}$. In particular, we have that $W_{A}^{\left(\mu\right)}\underset{\left(\text{dist}\right)}{\sim}N\left(0,\mu\left(A\right)\right)$, i.e., mean zero, Gaussian, and variance = $\mu\left(A\right)$. Then: \begin{lem}[see e.g., \cite{MR3687240}] With the assumptions as above, we have \begin{equation} \lim_{\left(A_{i}\right)}\mathbb{E}\left(\big|\mu\left(A\right)\mathbbm{1}-\sum\nolimits _{i}(W_{A_{i}}^{\left(\mu\right)})^{2}\Big|^{2}\right)=0\label{eq:C4-1} \end{equation} where (in (\ref{eq:C4-1})) the limit is taken over the filter of all partitions $\left(A_{i}\right)$ of $A$, and $\mathbbm{1}$ denotes the constant function ``one'' on $\Omega$. \end{lem} As a result, we get the following Ito-integral \begin{equation} W^{\left(\mu\right)}\left(f\right):=\int_{M}f\left(s\right)\,dW_{s}^{\left(\mu\right)},\label{eq:C6} \end{equation} defined for all $f\in L^{2}\left(M,\mathscr{F},\mu\right)$, and \begin{equation} \mathbb{E}\left(\left|\int\nolimits _{M}f\left(s\right)dW_{s}^{\left(\mu\right)}\right|^{2}\right)=\int_{M}\left|f\left(s\right)\right|^{2}d\mu\left(s\right).\label{eq:C5} \end{equation} We note that the following operator, \begin{equation} L^{2}\left(M,\mu\right)\ni f\longmapsto W^{\left(\mu\right)}\left(f\right)\in L^{2}\left(\Omega,\mathbb{P}\right)\label{eq:C7} \end{equation} is isometric. In our subsequent considerations, we shall need the following precise formula (see \lemref{D3}) for the RKHS associated with the p.d. kernel \begin{equation} K^{\left(\mu\right)}\left(A,B\right):=\mu\left(A\cap B\right),\label{eq:C8} \end{equation} defined on $\mathscr{F}_{fin}\times\mathscr{F}_{fin}$. We denote the RKHS by $\mathscr{H}(K^{\left(\mu\right)})$. \begin{lem} \label{lem:D3}Let $\mu$ be as above, and let $K^{\left(\mu\right)}$ be the p.d. kernel on $\mathscr{F}_{fin}$ defined in (\ref{eq:C8}). Then the corresponding RKHS $\mathscr{H}(K^{\left(\mu\right)})$ is as follows: A function $\Phi$ on $\mathscr{F}_{fin}$ is in $\mathscr{H}(K^{\left(\mu\right)})$ if and only if there is a $\varphi\in L^{2}\left(M,\mathscr{F}_{M},\mu\right)\left(=:L^{2}\left(\mu\right)\right)$ such that \begin{equation} \Phi\left(A\right)=\int_{A}\varphi\,d\mu,\label{eq:C10} \end{equation} for all $A\in\mathscr{F}_{fin}$. Then \begin{equation} \left\Vert \Phi\right\Vert _{\mathscr{H}(K^{\left(\mu\right)})}=\left\Vert \varphi\right\Vert _{L^{2}\left(\mu\right)}.\label{eq:D11-1} \end{equation} \end{lem} \begin{proof} To show that $\Phi$ in (\ref{eq:C10}) is in $\mathscr{H}(K^{\left(\mu\right)})$, we must choose a finite constant $C_{\Phi}$ such that, for all finite subset $\left(A_{i}\right)_{i=1}^{N}$, $A_{i}\in\mathscr{F}_{fin}$, $\left\{ \xi_{i}\right\} _{i=1}^{N}$, $\xi_{i}\in\mathbb{R}$, we get the following \emph{a priori} estimate: \begin{equation} \left|\sum\nolimits _{i=1}^{N}\xi_{i}\Phi\left(A_{i}\right)\right|^{2}\leq C_{\Phi}\sum\nolimits _{i}\sum\nolimits _{j}\xi_{i}\xi_{j}K^{\left(\mu\right)}\left(A_{i},A_{j}\right).\label{eq:C11} \end{equation} But a direct application of Schwarz to $L^{2}\left(\mu\right)$ shows that (\ref{eq:C11}) holds, and for a finite $C_{\Phi}$, we may take $C_{\Phi}=\left\Vert \varphi\right\Vert _{L^{2}\left(\mu\right)}^{2}$, where $\varphi$ is the $L^{2}\left(\mu\right)$-function in (\ref{eq:C10}). The desired conclusion now follows from an application of \lemref{B1}. We have proved one implication from the statement of the lemma: Functions $\Phi$ on $\mathscr{F}_{fin}$ of the formula (\ref{eq:C10}) are in the RKHS $\mathscr{H}\left(K^{\left(\mu\right)}\right)$, and the norm $\left\Vert \cdot\right\Vert _{\mathscr{H}\left(K^{\left(\mu\right)}\right)}$ is as stated in (\ref{eq:D11-1}). In the below, we shall denote these elements in $\mathscr{H}\left(K^{\left(\mu\right)}\right)$ as pairs $\left(\Phi,\varphi\right)$. We shall also restrict attention to the case of real valued functions. For the converse implication, let $H$ be a function on $\mathscr{F}_{fin}$, and assume $H\in\mathscr{H}\left(K^{\left(\mu\right)}\right)$. Then by Schwarz applied to $\left\langle \cdot,\cdot\right\rangle _{\mathscr{H}\left(K^{\left(\mu\right)}\right)}$ we get \begin{equation} \left|\left\langle H,\Phi\right\rangle _{\mathscr{H}\left(K^{\left(\mu\right)}\right)}\right|\leq\left\Vert H\right\Vert _{\mathscr{H}\left(K^{\left(\mu\right)}\right)}\left\Vert \varphi\right\Vert _{L^{2}\left(\mu\right)},\label{eq:D11-2} \end{equation} where we used (\ref{eq:D11-1}). Hence when Schwarz is applied to $L^{2}\left(\mu\right)$, we get a unique $h\in L^{2}\left(\mu\right)$ such that \begin{equation} \left\langle H,\Phi\right\rangle _{\mathscr{H}\left(K^{\left(\mu\right)}\right)}=\int_{M}h\,\varphi\,d\mu\label{eq:D11-3} \end{equation} for all $\left(\Phi,\varphi\right)$ as in (\ref{eq:C10}). Now specialize to $\varphi=\chi_{A}$, $A\in\mathscr{F}_{fin}$, in (\ref{eq:D11-3}) and we conclude that \begin{equation} H\left(A\right)=\int_{A}h\,d\mu; \end{equation} which translates into the assertion that the pair $\left(H,h\right)$ has the desired form (\ref{eq:C10}). And hence by (\ref{eq:D11-1}) we have $\left\Vert H\right\Vert _{\mathscr{H}\left(K^{\left(\mu\right)}\right)}=\left\Vert h\right\Vert _{L^{2}\left(\mu\right)}$ as stated. This concludes the proof of the converse inclusion. \end{proof} \section{\label{sec:fac}Factorizations and stochastic integrals} In Sections \ref{sec:pdk} and \ref{sec:gp}, we introduced the related notions of positive definite (p.d.) functions (kernels) on the one hand, and Gaussian processes on the other. One notes the immediate fact that a covariance kernel for a general stochastic process is positive definite. In general, the stochastic process in question is not determined by its covariance kernel. But in the special case when the process is Gaussian, it is. In \thmref{C1}, we stated that every p.d. kernel $K$ is indeed the covariance kernel of a Gaussian process. The construction is natural; starting with the p.d. kernel $K$, there is a canonical inductive limit construction leading to the Gaussian process for this problem. The basic idea for this particular construction of Gaussian processes dates back to pioneering work by Kolmogorov \cite{MR735967,MR562914}. In the present section, we formulate two different factorization questions, one for general p.d. kernels $K$, and the other for Gaussian processes, say $V$. For details, see the respective definitions in (\ref{eq:D2}) and (\ref{eq:D3}) below. If $K$ is indeed the covariance kernel for a Gaussian process $V$, it is natural to try to relate these two seemingly different notions of factorization. (In the case of Gaussian processes, a better name is perhaps \textquotedblleft subordination\textquotedblright{} (see (\ref{eq:D6}) below), but our theorem justifies the use of factorization in both of these contexts.) Our main result, \thmref{E1}, states that the data determining factorization for $K$ is the exact same as that which yields factorization for $V$.\\ Let $K$ be a positive definite kernel $X\times X\xrightarrow{\;K\;}\mathbb{C}$; and let $V=V_{K}$ be the corresponding Gaussian (mean zero) process, indexed by $X$, i.e., $V_{x}\in L^{2}\left(\Omega,\mathbb{P}\right)$, $\forall x\in X$, and \begin{equation} \mathbb{E}\left(\overline{V}_{x}V_{y}\right)=K\left(x,y\right),\;\forall\left(x,y\right)\in X\times X.\label{eq:D1} \end{equation} We set \begin{align} \mathscr{F}\left(K\right) & :=\Big\{\left(M,\mathscr{F}_{M},\mu\right)\mid\text{s.t. }K\left(\cdot,x\right)\longmapsto k_{x}\in L^{2}\left(M,\mu\right)\label{eq:D2}\\ & \qquad\text{extends to an isometry, i.e., }\nonumber \\ & \qquad K\left(x,y\right)=\int_{M}\overline{k_{x}\left(s\right)}k_{y}\left(s\right)d\mu\left(s\right)=\left\langle k_{x},k_{y}\right\rangle _{L^{2}\left(\mu\right)},\;\forall x,y\in X\Big\}.\nonumber \end{align} Further, if $V$ is the Gaussian process (from (\ref{eq:D1})), we set \begin{align} \mathscr{M}\left(V\right) & :=\Big\{\left(M,\mathscr{F}_{M},\mu\right)\mid\text{s.t. \ensuremath{V} admits an Ito-integral representation}\nonumber \\ & \qquad V_{x}=\int_{M}k_{x}\left(s\right)dW_{s}^{\left(\mu\right)},\;\forall x\in X,\text{ where}\:\left\{ k_{x}\right\} _{x\in X}\text{ is an}\label{eq:D3}\\ & \qquad\text{indexed system in \ensuremath{L^{2}\left(M,\mu\right)}}\Big\}.\nonumber \end{align} Following parallel terminology from measure theory, we say that a Gaussian process $V$ admits a disintegration, via suitable Ito-integrals, when there is a measure space with measure $\mu$ such that the corresponding Wiener process $W^{\left(\mu\right)}$ satisfies (\ref{eq:D3}). Our theorem below (\thmref{E1}) shows that this disintegration question may be decided instead by the answer to an equivalent spectral decomposition question; the latter of course formulated for the covariance kernel for $V$. As is shown in the examples/applications below, given a Gaussian process, it is not at all clear what disintegrations hold; see for example \corref{F7}. \begin{thm} \label{thm:E1}Let $K:X\times X\rightarrow\mathbb{C}$ be given positive definite, and let $\left\{ V_{x}\right\} _{x\in X}$ be the corresponding Gaussian (mean zero) process, then \begin{equation} \mathscr{F}\left(K\right)=\mathscr{M}\left(V\right).\label{eq:D4} \end{equation} \end{thm} \begin{proof} We shall need the following: \end{proof} \begin{lem} \label{lem:F2}From the definition of $\mathscr{F}\left(K\right)$, with $K$ fixed and assumed p.d., we get to every $\left(\left(k_{x}\right)_{x\in X},\mu\right)\in\mathscr{F}\left(K\right)$ a natural isometry $T_{\mu}:\mathscr{H}\left(K\right)\longrightarrow L^{2}\left(M,\mu\right)$. It is denoted by \begin{equation} T_{\mu}(\underset{\in\mathscr{H}\left(K\right)}{\underbrace{K\left(\cdot,x\right)}}):=k_{x}\in L^{2}\left(\mu\right);\label{eq:D4-1} \end{equation} and the adjoint operator $T_{\mu}^{*}:L^{2}\left(M,\mu\right)\longrightarrow\mathscr{H}\left(K\right)$ is as follows: For all $f\in L^{2}\left(M,\mu\right)$ we have \begin{equation} \left(T_{\mu}^{*}f\right)\left(x\right)=\int_{M}f\left(s\right)\overline{k_{x}\left(s\right)}d\mu\left(s\right).\label{eq:D4-2} \end{equation} Moreover, we also have \begin{equation} T_{\mu}^{*}\left(k_{x}\right)=K\left(\cdot,x\right),\;\text{for all \ensuremath{x\in X}.}\label{eq:E7-1} \end{equation} \end{lem} \begin{proof} Since $\left(k_{x},\mu\right)\in\mathscr{F}\left(K\right)$, we have the factorization property (\ref{eq:D2}), and so it follows from (\ref{eq:D4-1}) that this extends by linearity and norm-completion to an isometry $\mathscr{H}\left(K\right)\xrightarrow{\;T_{\mu}\;}L^{2}\left(\mu\right)$ as stated. By the definition of the adjoint operator $L^{2}\left(\mu\right)\xrightarrow{\;T_{\mu}^{*}\;}\mathscr{H}\left(K\right)$, we have for $f\in L^{2}\left(\mu\right)$: \[ \left(T_{\mu}^{*}f\right)\left(x\right)=\left\langle K\left(\cdot,x\right),T_{\mu}^{*}f\right\rangle _{\mathscr{H}\left(K\right)}=\left\langle k_{x},f\right\rangle _{L^{2}\left(\mu\right)}=\int_{M}f\left(s\right)\overline{k_{x}\left(s\right)}d\mu\left(s\right), \] which is the assertion in the lemma. From the properties of $\mathscr{H}\left(K\right)$ (see \secref{pdk}), it follows that (\ref{eq:E7-1}) holds iff \begin{equation} \left\langle K\left(\cdot,y\right),T_{\mu}^{*}\left(k_{x}\right)\right\rangle _{\mathscr{H}\left(K\right)}=\left\langle K\left(\cdot,y\right),K\left(\cdot,x\right)\right\rangle _{\mathscr{H}\left(K\right)}\label{eq:E7-2} \end{equation} for all $y\in X$. But we may compute both sides in eq. (\ref{eq:E7-2}) as follows: \begin{eqnarray*} \text{LHS}_{\left(\ref{eq:E7-2}\right)} & = & \left\langle T_{\mu}K\left(\cdot,y\right),k_{x}\right\rangle _{L^{2}\left(\mu\right)}\\ & \underset{\text{by \ensuremath{\left(\ref{eq:D4-1}\right)}}}{=} & \left\langle k_{y},k_{x}\right\rangle _{L^{2}\left(\mu\right)}\\ & \underset{\text{by \ensuremath{\left(\ref{eq:D2}\right)}}}{=} & K\left(y,x\right)\\ & \underset{\text{by \ensuremath{\left(\ref{eq:A3}\right)}}}{=} & \text{RHS}_{\left(\ref{eq:E7-2}\right)}. \end{eqnarray*} \end{proof} \begin{proof}[Proof of \thmref{E1} continued] The proof is divided into two parts, one for each of the inclusions $\subseteq$ and $\supseteq$ in (\ref{eq:D4}). \textbf{Part 1 ``$\subseteq$''.} Assume a pair $\left(\left(k_{x}\right)_{x\in X},\mu\right)$ is in $\mathscr{F}\left(K\right)$; see (\ref{eq:D2}). Then by definition, the factorization (\ref{eq:D3}) holds on $X\times X$. Now let $W^{\left(\mu\right)}$ denote the Wiener process associated with $\mu$, i.e., $W^{\left(\mu\right)}$ is a Gaussian process indexed by $\mathscr{F}_{fin}$, and \begin{equation} \mathbb{E}\left(W_{A}^{\left(\mu\right)}W_{B}^{\left(\mu\right)}\right)=\mu\left(A\cap B\right),\label{eq:D5} \end{equation} for all $A,B\in\mathscr{F}_{fin}$; see (\ref{eq:C1}) above. Now form the Ito-integral \begin{equation} V_{x}:=\int_{M}k_{x}\left(s\right)dW_{s}^{\left(\mu\right)},\;x\in X.\label{eq:D6} \end{equation} We stress that then $V_{x}$, as defined by (\ref{eq:D6}), is a Gaussian process indexed by $X$. To see this, use the general theory of Ito-integration, see also \cite{MR3882025,MR3701541,MR3574374,MR3721329,MR3670916,MR0301806,MR562914}. The approximation in (\ref{eq:D6}) is over the filter of all \emph{partitions } \begin{equation} \left\{ A_{i}\right\} _{i\in\mathbb{N}}\text{ s.t. }A_{i}\cap A_{j}=\emptyset,\:i\neq j,\;\cup_{i\in\mathbb{N}}A_{i}=M,\;\text{and}\;0<\mu\left(A_{i}\right)<\infty;\label{eq:D7} \end{equation} see (\ref{eq:C4}). From the property of $W_{A_{i}}^{\left(\mu\right)}$, $i\in\mathbb{N}$, we conclude that, for all $s_{i}\in A_{i}$, we have that \begin{equation} \sum_{i\in\mathbb{N}}k_{x}\left(s_{i}\right)W_{A_{i}}^{\left(\mu\right)}\label{eq:D8} \end{equation} is Gaussian (mean zero) with \begin{align} \mathbb{E}\left|\sum\nolimits _{i}k_{x}\left(s_{i}\right)W_{A_{i}}^{\left(\mu\right)}\right|^{2} & =\sum\nolimits _{i}\sum\nolimits _{j}\overline{k_{x}\left(s_{i}\right)}k_{x}\left(s_{j}\right)\mu\left(A_{i}\cap A_{j}\right)\nonumber \\ & =\sum\nolimits _{i}\left|k_{x}\left(s_{i}\right)\right|^{2}\mu\left(A_{i}\right);\label{eq:D9} \end{align} where we used (\ref{eq:D7}). Passing to the limit over the filter of all partitions of $M$ (as in (\ref{eq:D7})), we then get \[ \mathbb{E}\left(\int_{M}\overline{k_{x}\left(s\right)}dW_{s}^{\left(\mu\right)}\int_{M}k_{y}\left(t\right)dW_{t}^{\left(\mu\right)}\right)=\int_{M}\overline{k_{x}\left(s\right)}k_{y}\left(s\right)d\mu\left(s\right); \] and with definition (\ref{eq:D6}), therefore: \begin{equation} \mathbb{E}\left(\overline{V}_{x}V_{y}\right)=\left\langle k_{x},k_{y}\right\rangle _{L^{2}\left(\mu\right)}=K\left(x,y\right),\;\forall\left(x,y\right)\in X\times X,\label{eq:D10} \end{equation} where the last step in the derivation (\ref{eq:D10}) uses the assumption that $\left(\left(k_{x}\right)_{x\in X},\mu\right)\in\mathscr{F}\left(K\right)$; see (\ref{eq:D2}). \textbf{Part 2 ``$\supseteq$''.} Assume now that some pair $\left(\left(k_{x}\right)_{x\in X},\mu\right)$ is in $\mathscr{M}\left(V\right)$ where $K$ is given assumed p.d.; and where $\left(V_{x}\right)_{x\in X}$ is ``the'' associated (mean zero) Gaussian process; i.e., with $K$ as its covariance kernel; see (\ref{eq:D1}). We claim that $\left(\left(k_{x}\right)_{x\in X},\mu\right)$ must then be in $\mathscr{F}\left(K\right)$, i.e., that the factorization (\ref{eq:D3}) holds. This in turn follows from the following chain of identities: \begin{alignat}{2} \left\langle k_{x},k_{y}\right\rangle _{L^{2}\left(\mu\right)} & =\mathbb{E}\left(\overline{V}_{x}V_{y}\right) & \quad & \big(\text{since \ensuremath{V_{x}=\int_{M}k_{x}\left(s\right)dW_{s}^{\left(\mu\right)}}}\big)\nonumber \\ & =K\left(x,y\right) & & \big(\text{since \ensuremath{K} is the covariance kernel}\label{eq:D11}\\ & & & \text{ of the Gaussian process \ensuremath{\left(V_{x}\right)_{x\in X}}}\big)\nonumber \end{alignat} valid for $\forall\left(x,y\right)\in X\times X$, and the conclusion follows. Note that the first step in the derivation of (\ref{eq:D11}) uses the Ito-isometry. Hence, initially $K$ may possibly be the covariance kernel for a mean zero Gaussian process, say $\left(V'_{x}\right)$, different from $V_{x}:=\int_{M}k_{x}\left(s\right)dW_{s}^{\left(\mu\right)}$. But we proved that the two Gaussian processes $V_{x}$, and $V_{x}'$, have the same covariance kernel. It follows then the two processes must be equivalent. This is by general theory; see e.g. \cite{MR0277027,MR2053326,MR3687240}. The last uniqueness is only valid since we can consider Gaussian processes. Other stochastic processes are typically not determined uniquely from the respective covariance kernels. \end{proof} \begin{rem} In the statement of \thmref{E1} there are two isometries: Starting with $\left(\left(k_{x}\right)_{x\in X},\mu\right)\in\mathscr{F}\left(K\right)$ we get the canonical isometry $T_{\mu}:\mathscr{H}\left(K\right)\rightarrow L^{2}\left(\mu\right)$ given by \begin{equation} T_{\mu}\left(K\left(\cdot,x\right)\right)=k_{x}; \end{equation} see (\ref{eq:D4-1}) of \lemref{F2}. But with $\mu$, we then also get the Wiener process $W^{\left(\mu\right)}$ and the Ito-integral \begin{equation} L^{2}\left(M,\mu\right)\ni f\longmapsto\int_{M}f\,dW^{\left(\mu\right)}\in L^{2}\left(\Omega,Cyl,\mathbb{P}\right) \end{equation} as an isometry. Here $\left(\Omega,Cyl,\mathbb{P}\right)$ denotes the standard probability space, with $Cyl$ abbreviation for the cylinder sigma-algebra of subsets of $\Omega:=\mathbb{R}^{M}$. For finite subsets $\left(s_{1},s_{2},\cdots,s_{k}\right)$ in $M$, and Borel subsets $B_{k}$ in $\mathbb{R}^{k}$, the corresponding cylinder set \[ Cyl\left(\left(s_{i}\right)_{i=1}^{k}\right):=\left\{ \omega\in\mathbb{R}^{M}\mathrel{;}\left(\omega\left(s_{1}\right),\cdots,\omega\left(s_{k}\right)\right)\in B_{k}\right\} . \] In summary, we get the the following diagram of isometries, corresponding to a fixed $\left(\left(k_{x}\right)_{x\in X},\mu\right)\in\mathscr{F}\left(K\right)$, where $K$ is a fixed p.d. function on $X\times X$: \begin{figure}[H] \[ \xymatrix{\mathscr{H}\left(K\right)\ar@/^{1.5pc}/[rr]^{T_{\mu}}\ar@/_{1.2pc}/[dr]_{\text{composition}} & & L^{2}\left(M,\mu\right)\ar@/^{1.2pc}/[dl]^{\text{Ito-isometry for \ensuremath{W^{\left(\mu\right)}}}}\\ & L^{2}\left(\Omega,\mathbb{P}\right) } \] \caption{The two isometries. Factorizations by isometries.} \end{figure} \end{rem} \section{\label{sec:exa}Examples and applications} Below we present four examples in order to illustrate the technical points in \thmref{E1}. In the first example $X=\left[0,1\right]$, the unit interval, and in the next two examples $X=\mathbb{D}=\left\{ z\in\mathbb{C}\mathrel{;}\left|z\right|<1\right\} $ the open complex disk. In the fourth example, the Drury-Arveson kernel, we have $X=\mathbb{C}^{k}$. We begin with a note on identifications: For $t\in\left[0,1\right]$, we set \[ e\left(t\right):=e^{i2\pi t}. \] We write $\lambda_{1}$ for the Lebesgue measure restricted to $\left[0,1\right]$; and we make the identification: \begin{equation} \left[0,1\right]\cong\mathbb{R}/\mathbb{Z}\cong\mathbb{T}^{1}=\left\{ z\in\mathbb{C}\mathrel{;}\left|z\right|=1\right\} .\label{F1} \end{equation} Hence, for $L^{2}\left(\left[0,1\right],\lambda_{1}\right)$ we have the familiar Fourier expansion: With \begin{equation} f\in L^{2}\left(\lambda_{1}\right),\;\text{and}\;c_{n}:=\int_{\left[0,1\right]}\overline{e\left(nt\right)}f\left(t\right)d\lambda_{1}\left(t\right),\;n\in\mathbb{Z}, \end{equation} \begin{equation} f\left(t\right)=\sum_{n\in\mathbb{Z}}c_{n}e\left(nt\right),\;\text{and}\;\int_{0}^{1}\left|f\left(t\right)\right|^{2}d\lambda_{1}\left(t\right)=\sum_{n\in\mathbb{Z}}\left|c_{n}\right|^{2}. \end{equation} On $\left[0,1\right]$, we shall also consider the Cantor measure $\mu_{4}$ with support equal to the Cantor set \[ C_{4}=\left\{ x\mathrel{;}x=\sum\nolimits _{k=1}^{\infty}b^{k}/4^{k},\;b_{k}\in\left\{ 0,2\right\} \right\} \subseteq\left[0,1\right]; \] see \figref{F1} and \cite{MR1655831,jorgensen2018harmonic}. \begin{figure} \includegraphics[width=0.35\textwidth]{c4} \caption{\label{fig:F1} The $4$-Cantor set with double gaps as an iterated function system. This is an iterated-function system construction: Cantor-set and measure; see (\ref{eq:F4}) below.} \end{figure} It is known that $\mu_{4}$ is the unique probability measure s.t. \begin{equation} \frac{1}{2}\int_{0}^{1}\left(f\left(\frac{x}{4}\right)+f\left(\frac{x+2}{4}\right)\right)d\mu_{4}\left(x\right)=\int_{0}^{1}f\,d\mu_{4}.\label{eq:F4} \end{equation} For the Fourier transform $\widehat{\mu}_{4}$ we have \begin{equation} \widehat{\mu}_{4}\left(t\right)=\prod_{k=1}^{\infty}\frac{1}{2}\left(1+e^{i\pi t/4^{k}}\right),\;t\in\mathbb{R}. \end{equation} In \tabref{F1}, we summarize the three examples with the data from \thmref{E1}. We now turn to the details of the respective examples: \renewcommand{\arraystretch}{1.5} \begin{table} \begin{tabular}{|c|c|c|>{\centering}p{0.35\columnwidth}|c|} \hline & $X$ & $K$ & $k_{x},M=\left[0,1\right]\cong\mathbb{T}^{1}$, $\mathscr{F}\left(K\right)=\left\{ \left(k_{x},\mu\right)\right\} $ & \multicolumn{1}{c|}{$\mu$}\tabularnewline \hline Ex 1 & $\left[0,1\right]$ & $x\wedge y$ & $k_{x}\left(s\right)=\chi_{\left[0,x\right]}\left(s\right)$ & $\lambda_{1}$\tabularnewline \hline Ex 2 & $\mathbb{D}$ & ${\displaystyle \frac{1}{1-z\overline{w}}}$ & ${\displaystyle k_{z}\left(t\right)=\frac{1}{1-z\overline{e\left(t\right)}}}$ & $\lambda_{1}$ on $\mathbb{T}^{1}$\tabularnewline \hline Ex 3 & $\mathbb{D}$ & ${\displaystyle \prod_{n=0}^{\infty}\left(1+z^{4^{n}}\overline{w}^{4^{n}}\right)}$ & ${\displaystyle k_{z}\left(t\right)=\prod_{n=0}^{\infty}\left(1+z^{4^{n}}\overline{e\left(4^{n}t\right)}\right)}$ & $\mu_{4}$\tabularnewline \hline \multicolumn{1}{c}{} & \multicolumn{1}{c}{} & \multicolumn{1}{c}{} & \multicolumn{1}{>{\centering}p{0.35\columnwidth}}{} & \multicolumn{1}{c}{}\tabularnewline \end{tabular} \caption{\label{tab:F1} Three p.d. kernels and their respective Gaussian realizations.} \end{table} \renewcommand{\arraystretch}{1} \begin{example} \label{exa:F1}If $K\left(x,y\right):=x\wedge y$ is considered a kernel on $\left[0,1\right]\times\left[0,1\right]$, then the corresponding RKHS $\mathscr{H}\left(K\right)$ is the Hilbert space of functions $f$ on $\left[0,1\right]$ such that the distribution derivative $f'=df/dx$ is in $L^{2}\left(\left[0,1\right],\lambda_{1}\right)$, $\lambda_{1}=dx$, $f\left(0\right)=0$, and \begin{equation} \left\Vert f\right\Vert _{\mathscr{H}\left(K\right)}^{2}:=\int_{0}^{1}\left|f'\left(x\right)\right|^{2}dx;\label{eq:F6} \end{equation} and it is immediate that $\left(k_{x},\lambda_{1}\right)\in\mathscr{F}\left(K\right)$ where $k_{x}\left(s\right):=\chi_{\left[0,x\right]}\left(s\right)$, the indicator function; see \figref{F2}. \begin{figure}[H] \begin{tabular}{c} \includegraphics[width=0.35\textwidth]{ind1}\tabularnewline \includegraphics[width=0.35\textwidth]{ind2}\tabularnewline \tabularnewline \end{tabular} \caption{\label{fig:F2} The generators of the Cameron-Martin RKHS. See \exaref{F1}.} \end{figure} \begin{figure}[H] \includegraphics[width=0.65\textwidth]{bm} \caption{\label{fig:F3}Brownian motion on $\left[0,1\right]$. Sample-paths by Monte Carlo. See \exaref{F1}.} \end{figure} \begin{figure}[H] \includegraphics[width=0.65\textwidth]{bm2} \caption{\label{fig:F4}A Wiener process with holding patterns in the gaps of the Cantor set $C_{4}$ in \figref{F1}: The $W^{\left(\mu_{4}\right)}$ process on $\left[0,1\right]$. Sample-paths by Monte Carlo.} \end{figure} The process $W^{\left(\lambda_{1}\right)}$ is of course the standard Brownian motion on $\left[0,1\right]$, pinned at $x=0$; see \figref{F3}, and compare with the $W^{\left(\mu_{4}\right)}$-process in \figref{F4}. For Monte Carlo simulation, see e.g. \cite{KDB,MR3884672}. The Hilbert space characterized by (\ref{eq:F6}) is called the Cameron-Martin space, see e.g., \cite{MR562914}. Moreover, to see that (\ref{eq:F6}) is indeed the precise characterization of the RKHS for this kernel, one again applies \lemref{B1}. \end{example} It immediately follows from \thmref{E1} then the Gaussian processes corresponding to the data in \tabref{F1} are as follows: \begin{example} $z\in\mathbb{D}$: \begin{equation} V_{z}=\int_{0}^{1}\frac{1}{1-z\overline{e\left(t\right)}}dW_{t}^{\left(\lambda_{1}\right)} \end{equation} realized as an Ito-integral. As an application of \thmref{E1}, we get: \[ \mathbb{E}\left(\overline{V}_{z}V_{w}\right)=\frac{1}{1-z\overline{w}},\;\forall\left(z,w\right)\in\mathbb{D}\times\mathbb{D}. \] \end{example} \begin{example} \label{exa:F3}$z\in\mathbb{D}$: \begin{equation} V_{z}=\int_{0}^{1}\prod_{n=0}^{\infty}\left(1+z^{4^{n}}\overline{e\left(4^{n}t\right)}\right)dW_{t}^{\left(\mu_{4}\right)} \end{equation} were the $W^{\left(\mu_{4}\right)}$-Ito integral is supported on the Cantor set $C_{4}\subset\left[0,1\right]$, see \figref{F1}. As an application of \thmref{E1}, we get: \[ \mathbb{E}\left(\overline{V}_{z}V_{w}\right)=\prod_{n=0}^{\infty}\left(1+\left(z\overline{w}\right)^{4^{n}}\right). \] \end{example} The reasoning of \exaref{F3} is based on a theorem of the paper \cite{MR1655831} (see also \cite{jorgensen2018harmonic}). Set \begin{align} \Lambda_{4} & =\left\{ 0,1,4,5,16,17,20,21,64,65,\cdots\right\} \nonumber \\ & =\left\{ \sum\nolimits _{k=0}^{\text{finite}}\alpha_{k}4^{k}\mid\alpha_{k}\in\left\{ 0,1\right\} ,\;\text{fintie summations}\right\} \label{eq:F9} \end{align} then the Fourier functions $\left\{ e\left(\lambda t\right)\mathrel{;}\lambda\in\Lambda_{4}\right\} $ forms an orthonormal basis in $L^{2}\left(C_{4},\mu_{4}\right)$, i.e., every $f\in L^{2}\left(C_{4},\mu_{4}\right)$ has its Fourier expansion \begin{align*} \widehat{f}\left(\lambda\right) & =\int_{C_{4}}\overline{e\left(\lambda t\right)}f\left(t\right)d\mu_{4}\left(t\right);\\ f\left(t\right) & =\sum_{\lambda\in\Lambda_{4}}\widehat{f}\left(\lambda\right)e\left(\lambda t\right); \end{align*} and \[ \int_{C_{4}}\left|f\right|^{2}d\mu_{4}=\sum_{\lambda\in\Lambda_{4}}\left|\widehat{f}\left(\lambda\right)\right|^{2}. \] \begin{lem} \label{lem:F4}Consider the set $\Lambda_{4}$ in (\ref{eq:F9}), and, for $s\in\mathbb{D}$, let \begin{equation} F\left(s\right):=\sum_{\lambda\in\Lambda_{4}}s^{\lambda} \end{equation} be the corresponding generating function. Then we have the following infinite-product representation \begin{equation} F\left(s\right)=\prod_{n=0}^{\infty}\left(1+s^{4^{n}}\right).\label{eq:F11} \end{equation} \end{lem} \begin{proof} From (\ref{eq:F9}) we have the following self-similarity for $\Lambda_{4}$: It is the following identity of sets \begin{equation} \Lambda_{4}=\left\{ 0,1\right\} +4\Lambda_{4}.\label{eq:F12} \end{equation} Note that (\ref{eq:F12}) is an algorithm for generating points in $\Lambda_{4}$. Hence, \begin{align*} F\left(s\right) & =\sum_{\lambda\in\Lambda_{4}}s^{\lambda}=\sum_{\left\{ 0,1\right\} +4\Lambda_{4}}s^{\lambda}\\ & =\sum_{4\Lambda_{4}}s^{\lambda}+s\sum_{4\Lambda_{4}}s^{\lambda}\\ & =\left(1+s\right)F\left(s^{4}\right)\\ & =\big(1+s\big)\big(1+s^{4}\big)\cdots\big(1+s^{4^{n-1}}\big)F\big(s^{4^{n}}\big) \end{align*} and by induction. Hence, if $s\in\mathbb{D}$, the infinite-product is absolutely convergent, and the desired product formula (\ref{eq:F11}) follows. \end{proof} \begin{rem} Note that, in combination with the theorem from \cite{MR1655831} (see also \cite{jorgensen2018harmonic}), this property of the generating function $F=F_{\Lambda_{4}}$ from \lemref{F4} is used in the derivation of the assertions made about the factorization properties in \exaref{F3}; this includes the two formulas (Ex 3) as stated in \tabref{F1}; as well as of the verification that $\left(k_{z},\mu_{4}\right)\in\mathscr{F}\left(K\right)$, where $k_{z}$, $\mu_{4}$, and $K$ are as stated. A direct computation of the two cases, \exaref{F1} and \exaref{F3}, is of interest. Our result, \lemref{D3}, is useful in the construction: When computing the two Wiener processes $W^{\left(\lambda_{1}\right)}$ and $W^{\left(\mu\right)}$ one notes that the covariance computed on intervals $\left[0,x\right]$ as $0<x<1$ are as follows: \begin{align} \mathbb{E}\left(\left(W_{\left[0,x\right]}^{\left(\lambda_{1}\right)}\right)^{2}\right) & =\lambda_{1}\left(\left[0,x\right]\right)=x,\;\text{and}\label{eq:F13}\\ \mathbb{E}\left(\left(W_{\left[0,x\right]}^{\left(\mu_{4}\right)}\right)^{2}\right) & =\mu_{4}\left(\left[0,x\right]\right).\label{eq:F14} \end{align} So the two functions have the representations as in \figref{F5}. \begin{figure}[H] \begin{tabular}{>{\centering}p{0.45\columnwidth}>{\centering}p{0.45\columnwidth}} \includegraphics[width=0.3\textwidth]{var1} & \includegraphics[width=0.3\textwidth]{var2}\tabularnewline The variance formula in (\ref{eq:F13}). & The Devil's staircase. The variance formula in (\ref{eq:F14}).\tabularnewline \end{tabular} \caption{\label{fig:F5}The two cumulative distributions.} \end{figure} \end{rem} \begin{example} The following example illustrates the need for a distinction between $X$, and families of choices $M$ in \thmref{E1}. \emph{A priori}, one might expect that if $X\times X\xrightarrow{\;K\;}\mathbb{C}$ is given and p.d., it would be natural to try to equip $X$ with a $\sigma$-algebra $\mathscr{F}_{X}$ of subsets, and a measure $\mu$ such that the condition in (\ref{eq:D2}) holds for $\left(X,\mathscr{F}_{X},\mu\right)$, i.e., \begin{equation} K\left(x,y\right)=\int_{X}\overline{k}_{x}k_{y}d\mu,\;\left(x,y\right)\in X\times X \end{equation} with $\left\{ k_{x}\right\} _{x\in X}$ a system in $L^{2}\left(X,\mathscr{F}_{X},\mu\right)$. It turns out that there are interesting examples where this is known to \emph{not }be feasible. The best known such example is perhaps the Drury-Arveson kernel; see \cite{MR1668582} and \cite{MR2419381,MR2648865}. Specifics. Consider $\mathbb{C}^{k}$ for $k\geq2$, and $B_{k}\subset\mathbb{C}^{k}$ the complex ball defined for $z=\left(z_{1},\cdots,z_{k}\right)\in\mathbb{C}^{k}$, \begin{equation} B_{k}:=\Big\{ z\in\mathbb{C}^{k}\mathrel{;}\underset{\left\Vert z\right\Vert _{2}^{2}}{\underbrace{\sum\nolimits _{j=1}^{k}\left|z_{j}\right|^{2}}}<1\Big\}. \end{equation} For $z,w\in\mathbb{C}^{k}$, set \begin{align} \left\langle z,w\right\rangle & :=\sum_{j=1}^{k}z_{j}\overline{w}_{j},\;\text{and}\nonumber \\ K_{DA}\left(z,w\right) & :=\frac{1}{1-\left\langle z,w\right\rangle },\;\left(z,w\right)\in B_{k}\times B_{k}.\label{eq:F17} \end{align} \end{example} \begin{cor}[{Arveson \cite[Coroll 2]{MR1668582}}] \label{cor:F7}Let $k\geq2$, and let $\mathscr{H}\left(K_{DA}\right)$ be the RKHS of the D-A kernel in (\ref{eq:F17}). Then there is \uline{no} Borel measure on $\mathbb{C}^{k}$ such that $\left(\mathbb{C}^{k},\mathscr{B}_{k},\mu\right)\in\mathscr{F}\left(K_{DA}\right)$; i.e., there is \uline{no} solution to the formula \[ \left\Vert f\right\Vert _{\mathscr{H}\left(K_{DA}\right)}^{2}=\int_{\mathbb{C}^{k}}\left|f\left(z\right)\right|^{2}d\mu\left(z\right), \] for all $f\left(z\right)$ $k$-polynomials. \end{cor} \begin{rem} It is natural to ask about disintegration properties for the Gaussian process $V_{DA}$ corresponding to the Drury-Arveson kernel (\ref{eq:F17}). Combining our \thmref{E1} above with the corollary (Coroll \ref{cor:F7}), we conclude that, in two or more complex dimensions $k$, the question of finding the admissible disintegrations this Gaussian process $V_{DA}$ is subtle. It must necessarily involve measure spaces going beyond $\mathbb{C}^{k}$. \end{rem} \section{\label{sec:atomic}The case of $\left(k_{x},\mu\right)\in\mathscr{F}\left(K\right)$ when $\mu$ is atomic} Below we present a case where $\mu$ from pairs in $\mathscr{F}\left(K\right)$ may be chosen to be atomic. The construction is general, but for the sake of simplicity we shall assume that a given p.d. $K$ is such that the RKHS $\mathscr{H}\left(K\right)$ is separable, i.e., when it has an (all) orthonormal basis (ONB) indexed by $\mathbb{N}$. \begin{defn} \label{def:G1}Let $\mathscr{H}$ be a Hilbert space (separable), and let $\left\{ g_{n}\right\} _{n\in\mathbb{N}}$ be a system of vectors in $\mathscr{H}$ such that \begin{equation} \sum_{n\in\mathbb{N}}\left|\left\langle \psi,g_{n}\right\rangle _{\mathscr{H}}\right|^{2}=\left\Vert \psi\right\Vert _{\mathscr{H}}^{2}\label{eq:F1} \end{equation} holds for all $\psi\in\mathscr{H}$. We then say that $\left\{ g_{n}\right\} _{n\in\mathbb{N}}$ is a Parseval frame for $\mathscr{H}$. (Also see \defref{J1}.) An equivalent assumption is that the mapping \begin{equation} \mathscr{H}\ni\psi\xmapsto{\quad T\quad}\left(\left\langle \psi,g_{n}\right\rangle _{\mathscr{H}}\right)\in l^{2}\left(\mathbb{N}\right) \end{equation} is isometric. One checks that then the adjoint $T^{*}:l^{2}\rightarrow\mathscr{H}$ is: \[ T^{*}\left(\left(\xi_{n}\right)\right)=\sum_{n\in\mathbb{N}}\xi_{n}g_{n}\in\mathscr{H}. \] \end{defn} For general background references on frames in Hilbert space, we refer to \cite{MR2367342,MR2538596,MR3118429,MR3005286,MR3009685,MR3204026,MR3121682,MR3275625,MR3574374}, and also see \cite{MR3005286,MR3526434,MR3688637,MR3700114,MR3800275}. \begin{lem} \label{lem:G2}Let $K$ be given p.d. on $X\times X$, and assume that $\left\{ g_{n}\right\} _{n\in\mathbb{N}}$ is a Parseval frame in $\mathscr{H}\left(K\right)$; then \begin{equation} K\left(x,y\right)=\sum_{n\in\mathbb{N}}g_{n}\left(x\right)\overline{g_{n}\left(y\right)}\label{eq:G3} \end{equation} with the sum on the RHS in (\ref{eq:G3}) absolutely convergent. \end{lem} \begin{proof} By the reproducing property of $\mathscr{H}\left(K\right)$, see \secref{pdk}, we get, for all $\left(x,y\right)\in X\times X$: \begin{eqnarray*} K\left(x,y\right) & = & \left\langle K\left(\cdot,x\right),K\left(\cdot,y\right)\right\rangle _{\mathscr{H}\left(K\right)}\\ & \underset{\text{by \ensuremath{\left(\ref{eq:F1}\right)}}}{=} & \sum_{n\in\mathbb{N}}\left\langle K\left(\cdot,x\right),g_{n}\right\rangle _{\mathscr{H}\left(K\right)}\left\langle g_{n},K\left(\cdot,y\right)\right\rangle _{\mathscr{H}\left(K\right)}\\ & \underset{\text{by \ensuremath{\left(\ref{eq:A4}\right)}}}{=} & \sum_{n\in\mathbb{N}}g_{n}\left(x\right)\overline{g_{n}\left(y\right)}. \end{eqnarray*} \end{proof} Now a direct application of the argument in the proof of \thmref{E1} yields the following: \begin{cor} Let $K$ be given p.d. on $X\times X$ such that $\mathscr{H}\left(K\right)$ is separable, and let $\left\{ g_{n}\right\} _{n\in\mathbb{N}}$ be a Parseval frame, for example an ONB in $\mathscr{H}\left(K\right)$. Let $\left\{ \zeta_{n}\right\} _{n\in\mathbb{N}}$ be a chosen system of i.i.d. (independent identically distributed) system of standard Gaussians, i.e., with $N\left(0,1\right)$-distribution $\nicefrac{1}{\sqrt{2\pi}}e^{\nicefrac{-s^{2}}{2}}$, $s\in\mathbb{R}$. Then the following sum defines a Gaussian process, \begin{equation} V_{x}\left(\cdot\right):=\sum_{n\in\mathbb{N}}g_{n}\left(x\right)\zeta_{n}\left(\cdot\right), \end{equation} i.e., $\left\{ V_{x}\right\} _{x\in X}$ is well-defined in $L^{2}\left(\Omega,Cyl,\mathbb{P}\right)$, as stated, where $\Omega=\mathbb{R}^{\mathbb{N}}$ as a realization in an infinite Cartesian product with the usual cylinder $\sigma$-algebra, and $\left\{ V_{x}\right\} _{x\in X}$ has $K$ as covariance kernel, i.e., \[ \mathbb{E}\left(\overline{V}_{x}V_{y}\right)=K\left(x,y\right),\;\forall\left(x,y\right)\in X\times X; \] see (\ref{eq:D11}). \end{cor} \begin{proof} This is a direct application of \lemref{G2}, and we leave the remaining verifications to the reader. \end{proof} \section{Point processes: The case when $\left\{ \delta_{x}\right\} \subset\mathscr{H}\left(K\right)$} Let $X\times X\xrightarrow{\;K\;}\mathbb{R}$ be a fixed positive definite kernel. We know that the RKHS $\mathscr{H}\left(K\right)$ consists of functions $h$ on $X$ subject to the \emph{a priori} estimate in \lemref{B1}. For recent work on point-processes over infinite networks \cite{MR3860446,MR3246982,MR3843552,MR3670916,MR3390972,MR3861681,MR3848251,MR3881668,MR3854505}, the case when the Dirac measures $\delta_{x}$ are in $\mathscr{H}\left(K\right)$ is of special significance. In this case there is an abstract Laplace operator $\Delta$, defined as follows: \begin{equation} \left(\Delta h\right)\left(x\right)=\left\langle \delta_{x},h\right\rangle _{\mathscr{H}\left(K\right)},\;\forall h\in\mathscr{H}\left(K\right).\label{eq:H1} \end{equation} For the $\left\Vert \cdot\right\Vert _{\mathscr{H}\left(K\right)}$-norm of $\delta_{x}$, we have \begin{equation} \left(\Delta\delta_{x}\right)\left(x\right)=\left\Vert \delta_{x}\right\Vert _{\mathscr{H}\left(K\right)}^{2};\label{eq:H2} \end{equation} immediate from (\ref{eq:H1}). For every finite subset $F\subset X$, we consider the induced $\left|F\right|\times\left|F\right|$ matrix \begin{equation} K_{F}\left(x,y\right)=\left(K\left(x,y\right)\right)_{x,y\in F}. \end{equation} Note that $K_{F}$ is a positive definite square matrix. Its spectrum consists of eigenvalues $\lambda_{s}\left(F\right)$. If $\left(K,X\right)$ is as described, i.e., $X\times X\xrightarrow{\;K\;}\mathbb{R}\left(\text{or }\mathbb{C}\right)$ p.d., and if \begin{equation} \left\{ \delta_{x}\right\} _{x\in X}\subset\mathscr{H}\left(K\right),\label{eq:H4-1} \end{equation} we shall see that $X$ must then be discrete. (In interesting cases, also countable.) If (\ref{eq:H4-1}) holds, we shall say that $\left(K,X\right)$ is a \emph{point process}. We shall further show that point processes arise by restriction as follows: Let $\left(K,X\right)$ be given with $K$ a p.d. kernel. If a countable subset $S\subset X$ is such that $K^{\left(S\right)}:=K\big|_{S\times S}$ has \begin{equation} \left\{ \delta_{x}\right\} _{x\in S}\in\mathscr{H}(K^{\left(S\right)}), \end{equation} then we shall say that $\left(K^{\left(S\right)},S\right)$ is an \emph{induced point process}. \subsection{Nets of finite submatrices, and their limits} Given $(K,X)$ as above with $K$ p.d. and defined on $X\times X$. Then the finite submatrices in the subsection header are indexed by the net of all finite subsets $F$ of $X$ as follows: Given $F$, then the corresponding $\left|F\right|\times\left|F\right|$ square matrix $K_{F}$ is simply the restriction of $K$ to $F\times F$. Of course, each matrix $K_{F}$ is positive definite, and so it has a finite list of eigenvalues. These eigenvalue lists figure in the discussion below. \begin{lem} \label{lem:H1}Let $K$, $F$, and $K_{F}$ be as above, with $\lambda_{s}\left(F\right)$ denoting the numbers in the list of eigenvalues for the matrix $K_{F}$. Then \begin{equation} 1\leq\lambda_{s}\left(F\right)\sum_{x\in F}\left\Vert \delta_{x}\right\Vert _{\mathscr{H}\left(K\right)}^{2}.\label{eq:H4} \end{equation} \end{lem} \begin{proof} Consider the eigenvalue equation \begin{equation} \left(\xi_{x}\right)_{x\in F},\quad\sum_{x\in F}\left|\xi_{x}\right|^{2}=\left\Vert \xi\right\Vert _{2}^{2}=1,\quad K_{F}\xi=\lambda_{s}\left(F\right)\xi. \end{equation} From \lemref{B1} and for $x\in F$, we then get \begin{align} \left|\xi_{x}\right|^{2} & \leq\left\Vert \delta_{x}\right\Vert _{\mathscr{H}\left(K\right)}^{2}\left\langle \xi,K_{F}\xi\right\rangle _{l^{2}\left(F\right)}\nonumber \\ & =\left\Vert \delta_{x}\right\Vert _{\mathscr{H}\left(K\right)}^{2}\lambda_{s}\left(F\right).\label{eq:H6} \end{align} Now apply $\sum_{x\in F}$ to both sides in (\ref{eq:H6}), and the desired conclusion (\ref{eq:H4}) follows. \end{proof} \begin{rem} A consequence of the lemma is that the matrices $K_{F}^{-1}$ and $K_{F}^{-1/2}$ automatically are well defined (by the spectral theorem) with associated spectral bounds. \end{rem} \begin{defn} Let $K$, $F$, and $K_{F}$ be as above; and with the condition $\delta_{x}\in\mathscr{H}\left(K\right)$ in force. Set \begin{equation} \mathscr{H}_{K}\left(F\right):=span_{x\in F}\left\{ K\left(\cdot,x\right)\right\} . \end{equation} \end{defn} It is a finite-dimensional (and therefore closed) subspace in $\mathscr{H}\left(K\right)$. The orthogonal projection onto $\mathscr{H}_{K}\left(F\right)$ will be denoted $P_{F}:\mathscr{H}\left(K\right)\rightarrow\mathscr{H}_{K}\left(F\right)$. \begin{lem} \label{lem:H4}Let $K$, $F$, $K_{F}$, and $\mathscr{H}_{K}\left(F\right)$ be as above. Then the orthogonal projection $P_{F}$ is as follows: For $h\in\mathscr{H}\left(K\right)$, set $h_{F}=h\big|_{F}$, restriction: \begin{equation} \left(P_{F}h\right)\left(\cdot\right)=\sum_{y\in F}\left(K_{F}^{-1}h_{F}\right)\left(y\right)K\left(\cdot,y\right).\label{eq:H8} \end{equation} \end{lem} \begin{proof} It is immediate from the definition that $P_{F}h$ has the form \begin{equation} P_{F}h=\sum_{y\in F}\xi_{y}K\left(\cdot,y\right) \end{equation} with $\left(\xi_{y}\right)_{y\in F}\in\mathbb{C}^{\left|F\right|}$. Since $P_{F}$ is the orthogonal projection, \begin{equation} \left(h-P_{F}h\right)\perp_{\mathscr{H}\left(K\right)}\left\{ K\left(\cdot,y\right)\right\} _{y\in F} \end{equation} (orthogonality in the $\mathscr{H}\left(K\right)$-inner product) which yields: \[ h\left(x\right)=\left(K_{F}\xi\right)\left(x\right)\left(=\sum\nolimits _{y\in F}K\left(x,y\right)\xi_{y}\right),\;\forall x\in F; \] and therefore, $\xi=K_{F}^{-1}h_{F}$, which is the desired formula (\ref{eq:H8}). \end{proof} \begin{cor} \label{cor:H5}Let $X$, $K$, $\mathscr{H}\left(K\right)$ be as above, and assume $\delta_{x}\in\mathscr{H}\left(K\right)$ for some $x\in X$. Then a function $h$ on $X$ is in $\mathscr{H}\left(K\right)$ if and only if \begin{equation} \sup_{F}\left\Vert K_{F}^{-1/2}h_{F}\right\Vert _{l^{2}\left(F\right)}<\infty,\label{eq:H11} \end{equation} where the supremum is over all finite subsets $F$ of $X$. If $h$ is finite energy, then \begin{equation} \left\Vert h\right\Vert _{\mathscr{H}\left(K\right)}^{2}=\sup_{F}\left\Vert K_{F}^{-1/2}h_{F}\right\Vert _{l^{2}\left(F\right)}^{2}.\label{eq:H12} \end{equation} \end{cor} \begin{proof} The proof follows from an application of Hilbert space geometry to the RKHS $\mathscr{H}\left(K\right)$, on the family of orthogonal projections $P_{F}$ indexed by the finite subsets $F$ in $X$. With the standard lattice operations, applied to projections, we have $\sup_{F}P_{F}=I_{\mathscr{H}\left(K\right)}$. The conclusions (\ref{eq:H11})-(\ref{eq:H12}) follow from this since, by the lemma, \begin{eqnarray} \left\Vert P_{F}h\right\Vert _{\mathscr{H}\left(K\right)}^{2} & \underset{\text{by \ensuremath{\left(\ref{eq:A3}\right)}}}{=} & \left\langle K_{F}^{-1}h_{F},K_{F}K_{F}^{-1}h_{F}\right\rangle _{l^{2}\left(F\right)}\nonumber \\ & = & \left\langle h_{F},K_{F}^{-1}h_{F}\right\rangle _{l^{2}\left(F\right)}=\left\Vert K_{F}^{-1/2}h_{F}\right\Vert _{l^{2}\left(F\right)}^{2}.\label{eq:H13} \end{eqnarray} \end{proof} \begin{rem} The advantage with the use of this system of orthogonal projections $P_{F}$, indexed by the finite subsets $F$ of $X$, is that we may then take advantage of the known lattice operations for orthogonal projections in Hilbert space. But it is important that we get approximation with respect to the canonical norm in the RKHS $\mathscr{H}\left(K\right)$. This works because by our construction, the orthogonality properties for the projections $P_{F}$ refers precisely to the inner product in $\mathscr{H}\left(K\right)$. Naturally we get the best $\mathscr{H}\left(K\right)$-approximation properties when $X$ is further assumed countable. But the formula for the $\mathscr{H}\left(K\right)$-norm holds in general. \end{rem} \begin{cor} \label{cor:H7}Let $X\times X\xrightarrow{\;K\;}\mathbb{C}$ be fixed, assumed p.d., and let $\mathscr{H}\left(K\right)$ be the corresponding RKHS. Let $x\in X$ be given. Then $\delta_{x}\in\mathscr{H}\left(K\right)$ if and only if \begin{equation} \sup_{F\subset X,\:F\text{ finite, \ensuremath{x\in F}}}\left(K_{F}^{-1}\right)_{x,x}<\infty.\label{eq:H13-1} \end{equation} In this case, we have: \[ \left\Vert \delta_{x}\right\Vert _{\mathscr{H}\left(K\right)}^{2}=\text{the supremum in }\left(\ref{eq:H13-1}\right). \] \end{cor} \begin{proof} The result is immediate from \corref{H5} applied to $h:=\delta_{x}$, where $x$ is fixed. Here the terms in (\ref{eq:H12}) are, for $F$ finite, $x\in F$: \begin{equation} \left\langle \delta_{x}\big|_{F},K_{F}^{-1}\left(\delta_{x}\big|_{F}\right)\right\rangle _{l^{2}\left(F\right)}=\left(K_{F}^{-1}\right)_{x,x}, \end{equation} and the stated conclusion is now immediate. \end{proof} \begin{cor} \label{cor:H8}Let $X$, $K$, and $\mathscr{H}\left(K\right)$ be as above, but assume now that $X$ is countable, with a monotone net of finite sets: \begin{equation} F_{1}\subset F_{2}\subset F_{3}\cdots,\;\text{and}\quad X=\cup_{i\in\mathbb{N}}F_{i}; \end{equation} then a function $h$ on $X$ is in $\mathscr{H}\left(K\right)$ iff $\sup_{i}\left\Vert K_{F_{i}}^{-1/2}h\big|_{F_{i}}\right\Vert _{l^{2}\left(F_{i}\right)}<\infty$. Moreover, \begin{equation} \left\Vert h\right\Vert _{\mathscr{H}_{E}}^{2}=\lim_{i\rightarrow\infty}\left\Vert K_{F_{i}}^{-1/2}h\big|_{F_{i}}\right\Vert _{l^{2}\left(F_{i}\right)}^{2},\label{eq:H15} \end{equation} where, the convergence in (\ref{eq:H15}) is monotone. \end{cor} \begin{proof} From the definition of the order of orthogonal projections, we have \begin{equation} P_{F_{1}}\leq P_{F_{2}}\leq P_{F_{3}}\leq\cdots, \end{equation} and therefore, \begin{equation} \left\Vert P_{F_{1}}h\right\Vert _{\mathscr{H}\left(K\right)}^{2}\leq\left\Vert P_{F_{2}}h\right\Vert _{\mathscr{H}\left(K\right)}^{2}\leq\left\Vert P_{F_{3}}h\right\Vert _{\mathscr{H}\left(K\right)}^{2}\leq\cdots,\label{eq:H17} \end{equation} with $\lim_{i\rightarrow\infty}\left\Vert P_{F_{i}}h\right\Vert _{\mathscr{H}\left(K\right)}^{2}=\left\Vert h\right\Vert _{\mathscr{H}\left(K\right)}^{2}$. But by (\ref{eq:H13}) and the proof of \corref{H5}, we have \[ \left\Vert K_{F_{i}}^{-1/2}h\big|_{F_{i}}\right\Vert _{l^{2}\left(F_{i}\right)}^{2}=\left\Vert P_{F_{i}}h\right\Vert _{\mathscr{H}\left(K\right)}^{2} \] and, so, by (\ref{eq:H17}), we get: \[ \left\Vert K_{F_{1}}^{-1/2}h\big|_{F_{1}}\right\Vert _{l^{2}\left(F_{1}\right)}^{2}\leq\left\Vert K_{F_{2}}^{-1/2}h\big|_{F_{2}}\right\Vert _{l^{2}\left(F_{2}\right)}^{2}\leq\left\Vert K_{F_{3}}^{-1/2}h\big|_{F_{3}}\right\Vert _{l^{2}\left(F_{3}\right)}^{2}\leq\cdots. \] The conclusion now follows. \end{proof} \subsection{Restrictions of p.d. kernels} Below we shall be considering pairs $(K,X)$ with $K$ a fixed p.d. kernel defined on $X\times X$, and, as before, we denote by $\mathscr{H}\left(K\right)$ the corresponding RKHS with its canonical inner product. In general, $X$ is an arbitrary set, typically of large cardinality, in particular uncountable: It may be a complex domain, a generalized boundary, or it may be a manifold arising from problems in physics, in signal processing, or in machine learning models. Moreover, for such general pairs $(K,X)$, with $K$ a fixed p.d. kernel, the Dirac functions $\delta_{x}$ are typically not in $\mathscr{H}\left(K\right)$. Here we shall turn to induced systems, indexed by suitable countable discrete subsets $S$ of $X$. Indeed, for a number of sampling or interpolation problems, it is possible to identify countable discrete subsets $S$ of $X$, such that when $K$ is restricted to $S\times S$, i.e., $K^{\left(S\right)}:=K\big|_{S\times S}$, then for $x\in S$, the Dirac functions $\delta_{x}$ will be in $\mathscr{H}\left(K^{\left(S\right)}\right)$; i.e., we get induced point processes indexed by $S$. In fact, with \corref{H8}, we will be able to identify a variety of such subsets $S$. Moreover, each such choice of subset $S$ yields point-process, and an induced graph, and graph Laplacian; see (\ref{eq:H1})-(\ref{eq:H2}). These issues will be taken up in detail in the two subsequent sections. In the following \exaref{H8}, for illustration, we identify a particular instance of this, when $X=\mathbb{R}$ (the reals), and $S=\mathbb{Z}$ (the integers), and where $K$ is the covariance kernel of standard Brownian motion on $\mathbb{R}$. \begin{example}[\textbf{Discretizing the covariance function for Brownian motion on $\mathbb{R}$}] \label{exa:H8}The present example is a variant of \exaref{F1}, but with $X=\mathbb{R}$ (instead of the interval $\left[0,1\right]$). We now set \begin{equation} K\left(x,y\right):=\begin{cases} \left|x\right|\wedge\left|y\right| & \left(x,y\right)\in\mathbb{R}\times\mathbb{R},\;xy\geq0;\\ 0 & xy<0. \end{cases}\label{eq:H22} \end{equation} It is immediate that (\ref{eq:F6}) in \exaref{F1} carries over, but now with $\mathbb{R}$ in place of $\left[0,1\right]$. The normalization $f\left(0\right)=0$ is carried over. We get that: A function $f\left(x\right)$ on $\mathbb{R}$ is in $\mathscr{H}\left(K\right)$ iff it has distribution-derivative $f'=df/dx$ in $L^{2}\left(\mathbb{R}\right)$, see (\ref{eq:H19}). As before, we conclude that the $\mathscr{H}\left(K\right)$-norm is: \begin{equation} \left\Vert f\right\Vert _{\mathscr{H}\left(K\right)}^{2}=\int_{\mathbb{R}}\left|f'\right|^{2}dx;\label{eq:H19} \end{equation} see also \lemref{D3}. Set \begin{equation} K^{\left(\mathbb{Z}\right)}=K\big|_{\mathbb{Z}\times\mathbb{Z}}, \end{equation} and consider the corresponding RKHS $\mathscr{H}\left(K^{\left(\mathbb{Z}\right)}\right)$. Using \cite{MR3450534,MR3507188}, we conclude that functions $\Phi$ on $\mathbb{Z}$ are in $\mathscr{H}\left(K^{\left(\mathbb{Z}\right)}\right)$ iff $\Phi\left(0\right)=0$, and \[ \sum_{n\in\mathbb{Z}}\left|\Phi\left(n\right)-\Phi\left(n+1\right)\right|^{2}<\infty. \] In that case, \begin{equation} \left\Vert \Phi\right\Vert _{\mathscr{H}\left(K^{\left(\mathbb{Z}\right)}\right)}^{2}=\sum_{n\in\mathbb{Z}}\left|\Phi\left(n\right)-\Phi\left(n+1\right)\right|^{2}.\label{eq:H24-1} \end{equation} For the $\mathbb{Z}$-kernel, we have: $\left\{ \delta_{n}\right\} _{n\in\mathbb{Z}}\subset\mathscr{H}\left(K^{\left(\mathbb{Z}\right)}\right)$, and \begin{equation} \delta_{n}\left(\cdot\right)=2K\left(\cdot,n\right)-K\left(\cdot,n+1\right)-K\left(\cdot,n-1\right),\;\forall n\in\mathbb{Z}.\label{eq:H26-1} \end{equation} Moreover, the corresponding Laplacian $\Delta$ from (\ref{eq:H1}) is \begin{equation} \left(\Delta\Phi\right)\left(n\right)=2\Phi\left(n\right)-\Phi\left(n+1\right)-\Phi\left(n-1\right), \end{equation} i.e., the standard discretized Laplacian. From the matrices $K_{F}^{\left(\mathbb{Z}\right)}$, $F\subset\mathbb{Z}$, we have the following; illustrated with $F=F_{N}=\left\{ 1,2,\cdots,N\right\} $. \begin{align} K_{F_{N}}^{\left(\mathbb{Z}\right)} & =\begin{bmatrix}1 & 1 & 1 & 1 & \cdots & \cdots & 1\\ 1 & 2 & 2 & 2 & \cdots & \cdots & 2\\ 1 & 2 & 3 & 3 & \cdots & \cdots & 3\\ 1 & 2 & 3 & 4 & \cdots & \cdots & 4\\ \vdots & \vdots & \vdots & \vdots & \ddots & \ddots & \vdots\\ \vdots & \vdots & \vdots & \vdots & \ddots & N-1 & N-1\\ 1 & 2 & 3 & 4 & \cdots & N-1 & N \end{bmatrix},\\ \intertext{and}\left(K_{F_{N}}^{\left(\mathbb{Z}\right)}\right)^{-1} & =\begin{bmatrix}2 & -1 & 0 & 0 & 0 & \cdots & 0\\ -1 & 2 & -1 & 0 & 0 & \cdots & 0\\ 0 & -1 & 2 & -1 & 0 & \cdots & 0\\ \vdots & \ddots & \ddots & \ddots & \ddots & \ddots & \vdots\\ \vdots & & \ddots & -1 & 2 & -1 & 0\\ 0 & 0 & \cdots & 0 & -1 & 2 & -1\\ 0 & 0 & \cdots & 0 & 0 & -1 & 1 \end{bmatrix}. \end{align} In particular, we have for $n,m\in\mathbb{Z}$: \[ \left\langle \delta_{n},\delta_{m}\right\rangle _{\mathscr{H}\left(K^{\left(\mathbb{Z}\right)}\right)}=\begin{cases} 2 & \text{if \ensuremath{n=m}}\\ -1 & \text{if \ensuremath{\left|n-m\right|}=1}\\ 0 & \text{otherwise}. \end{cases} \] \end{example} \begin{rem} The determinant of $K_{F_{N}}^{\left(\mathbb{Z}\right)}$ is 1 for all $N$. \emph{Proof}. By eliminating the first column, and then the first row, $\det(K_{F_{N}}^{\left(\mathbb{Z}\right)})$ is reduced to $\det(K_{F_{N-1}}^{\left(\mathbb{Z}\right)})$ . So by induction, the determinant is 1. Note that \[ \sum_{k\in\mathbb{Z}}\chi_{\left[1,n\right]}\left(k\right)\chi_{\left[1,m\right]}\left(k\right)=n\wedge m \] which yields the factorization \begin{equation} K_{F_{N}}^{\left(\mathbb{Z}\right)}=A_{N}A_{N}^{*},\label{eq:H29-1} \end{equation} i.e., \[ K_{F_{N}}^{\left(\mathbb{Z}\right)}\left(n,m\right)=\left(A_{N}A_{N}^{*}\right)_{n,m}=\sum A_{N}\left(n,k\right)A_{N}^{*}\left(k,m\right), \] where $A_{N}$ is the $N\times N$ lower triangular matrix given by \[ A_{N}=\begin{bmatrix}1 & 0 & \cdots & \cdots & 0\\ 1 & 1 & 0 & \cdots & 0\\ \vdots & \vdots & \ddots & \ddots & \vdots\\ \vdots & \vdots & \vdots & \ddots & 0\\ 1 & 1 & \cdots & \cdots & 1 \end{bmatrix}. \] In particular, we get that $\det(K_{F_{N}}^{\left(\mathbb{Z}\right)})=1$ immediately. This is a special case of \thmref{E1}. For the general case, let $F_{N}=\left\{ x_{j}\right\} _{j=1}^{N}$ be a finite subset of $\mathbb{R}$, assuming $x_{1}<x_{2}<\cdots<x_{N}$. Then the factorization (\ref{eq:H29-1}) holds with \begin{equation} A_{N}=\begin{bmatrix}\sqrt{x_{1}} & 0 & 0 & \cdots & 0\\ \sqrt{x_{1}} & \sqrt{x_{2}-x_{1}} & 0 & \cdots & \vdots\\ \sqrt{x_{1}} & \sqrt{x_{2}-x_{1}} & \sqrt{x_{3}-x_{2}} & \ddots & \vdots\\ \vdots & \vdots & \vdots & \ddots & 0\\ \sqrt{x_{1}} & \sqrt{x_{2}-x_{1}} & \sqrt{x_{3}-x_{2}} & \cdots & \sqrt{x_{N}-x_{N-1}} \end{bmatrix}.\label{eq:H31} \end{equation} Thus, \begin{equation} \det(K_{F_{N}}^{\left(\mathbb{Z}\right)})=x_{1}\left(x_{2}-x_{1}\right)\cdots\left(x_{N}-x_{N-1}\right). \end{equation} In the setting of \secref{fac} (finite sums of standard Gaussians), we have the following: Let $\left\{ x_{i}\right\} _{i=1}^{N}$ be as in (\ref{eq:H31}), and let $1\leq n,m\leq N$. Let $\left\{ Z_{i}\right\} _{i=1}^{N}$ be a system i.i.d. standard Gaussians $N\left(0,1\right)$, i.e., independent identically distributed. Set \begin{equation} V_{n}=Z_{1}\sqrt{x_{1}}+Z_{2}\sqrt{x_{2}-x_{1}}+\cdots+Z_{n}\sqrt{x_{n}-x_{n-1}}. \end{equation} Then one checks that \begin{equation} \mathbb{E}\left(V_{n}V_{m}\right)=x_{n}\wedge x_{m}=K\left(x_{n},x_{m}\right) \end{equation} which is the desired Gaussian realization of $K$. Alternatively, $K_{F_{N}}^{\left(\mathbb{Z}\right)}$ assumes the following factorization via non-square matrices: Assume $F_{N}\subset\mathbb{Z}_{+}$, then \begin{equation} K_{F_{N}}^{\left(\mathbb{Z}\right)}=AA^{*}, \end{equation} where $A$ is the $N\times x_{N}$ matrix such that \[ A_{n,k}=\begin{cases} 1 & \text{if \ensuremath{1\leq k\leq x_{n}}}\\ 0 & \text{otherwise} \end{cases}. \] That is, $A$ takes the form: \begin{equation} A=\begin{bmatrix}\tikzmark{1L}1 & \cdots & 1\tikzmark{1R} & 0 & \cdots & \cdots & \cdots & \cdots & \cdots & 0\\ \\ \\ \tikzmark{2L}1 & \cdots & \cdots & \cdots & 1\tikzmark{2R} & 0 & \cdots & \cdots & \cdots & 0\\ \\ \\ \tikzmark{3L}1 & \cdots & \cdots & \cdots & \cdots & \cdots & 1\tikzmark{3R} & 0 & \cdots & 0\\ \\ \\ \vdots & \vdots & & & \vdots & \vdots & & \vdots & \vdots & \vdots\\ \vdots & \vdots & & & \vdots & \vdots & & \vdots & \vdots & 0\\ \tikzmark{NL}1 & 1 & \cdots & \cdots & \cdots & \cdots & \cdots & \cdots & 1 & 1\tikzmark{NR} \end{bmatrix}. \end{equation} \tikz[overlay, remember picture, decoration={brace, amplitude=3pt}] { \draw[decorate,thick] (1R.south) -- (1L.south) node [midway,below=5pt] {$x_{1}$}; \draw[decorate,thick] (2R.south) -- (2L.south) node [midway,below=5pt] {$x_{2}$}; \draw[decorate,thick] (3R.south) -- (3L.south) node [midway,below=5pt] {$x_{3}$}; \draw[decorate,thick] (NR.south) -- (NL.south) node [midway,below=5pt] {$x_{N}$}; } \end{rem} \vspace{2em} \begin{rem}[Spectrum of the matrices $K_{F}$; see also \cite{MR3034493}] It is known that the factorization as in (\ref{eq:H29-1}) can be used to obtain the spectrum of positive definite matrices. The algorithm is as follows: Let $K$ be a given p.d. matrix. Initialization: $B:=K$; Iterations: $k=1,2,\cdots,n-1$, \begin{enumerate} \item $B=AA^{*}$; \item $B=A^{*}A$; \end{enumerate} Here $A$ in step (i) denotes the lower triangular matrix in the Cholesky decomposition of $B$ (see (\ref{eq:H29-1})). Then $\lim_{n\rightarrow\infty}B$ converges to a diagonal matrix consisting of the eigenvalues of $K$. \end{rem} We now resume consideration of the general case of p.d. kernels $K$ on $X\times X$ and their restrictions: A setting for harmonic functions. \begin{rem} \label{rem:H10}In the general case of (\ref{eq:H2}) and \lemref{H1}, we still have a Laplace operator $\Delta$. It is a densely defined symmetric operator on $\mathscr{H}\left(K\right)$. Moreover (general case), \begin{equation} \Delta_{\cdot}K\left(\cdot,x\right)=\delta_{x}\left(\cdot\right),\;\forall x\in X\label{eq:H23} \end{equation} (assuming that $\delta_{x}\in\mathscr{H}\left(K\right)$). The dot ``$\cdot$'' in (\ref{eq:H23}) refers to the action variable for the operator $\Delta$. In other words, $K\left(\cdot,\cdot\right)$ is a generalized Greens kernel. \end{rem} \begin{defn} Let $X\times X\xrightarrow{\;K\:}\mathbb{C}$ be given p.d., and assume \begin{equation} \left\{ \delta_{x}\right\} _{x\in X}\subset\mathscr{H}\left(K\right).\label{eq:H24} \end{equation} Let $\Delta$ denote the induced Laplace operator. A function $h$ (in $\mathscr{H}\left(K\right)$) is said to be \emph{harmonic} iff (Def.) $\Delta h=0$. \end{defn} \begin{cor} Let $\left(X,K,\mathscr{H}\left(K\right)\right)$ be as above. Assume (\ref{eq:H24}), and let $\Delta$ be the induced Laplace operator. Then we have the following orthogonal decomposition for $\mathscr{H}\left(K\right)$: \begin{equation} \mathscr{H}\left(K\right)=\left\{ h\mathrel{;}\Delta h=0\right\} \oplus clospan^{\mathscr{H}\left(K\right)}\left(\left\{ \delta_{x}\right\} _{x\in X}\right)\label{eq:H25} \end{equation} where ``clospan'' in (\ref{eq:H25}) refers to the norm in $\mathscr{H}\left(K\right)$. \end{cor} \begin{proof} It is immediate from (\ref{eq:H1}) that \begin{equation} \left\{ h\in\mathscr{H}\left(K\right)\mathrel{;}\Delta h=0\right\} =\left(\left\{ \delta_{x}\right\} _{x\in X}\right)^{\perp}\label{eq:H26} \end{equation} where the orthogonality ``$\perp$'' in (\ref{eq:H26}) refers to the inner product $\left\langle \cdot,\cdot\right\rangle _{\mathscr{H}\left(K\right)}$. Since, by Hilbert space geometry, $\left(\left\{ \delta_{x}\right\} _{x\in X}\right)^{\perp\perp}=clospan^{\mathscr{H}\left(K\right)}\left(\left\{ \delta_{x}\right\} _{x\in X}\right)$, we only need to observe that $\left\{ h\in\mathscr{H}\left(K\right)\mathrel{;}\Delta h=0\right\} $ is closed in $\mathscr{H}\left(K\right)$. But this is immediate from (\ref{eq:H1}). \end{proof} \begin{cor}[Duality] \label{cor:H14} Let $X\times X\xrightarrow{\;K\;}\mathbb{R}$ be given, assumed p.d., and let $S\subset X$ be a countable subset such that \begin{equation} \mathscr{D}\left(S\right):=\left\{ \delta_{x}\right\} _{x\in S}\subset\mathscr{H}(K^{\left(S\right)}). \end{equation} \begin{enumerate} \item Then the following duality holds for the two induced kernels: \begin{align} K^{\left(S\right)} & :=K\big|_{S\times S},\;\text{and}\\ D^{\left(S\right)}\left(x,y\right) & :=\left\langle \delta_{x},\delta_{y}\right\rangle _{\mathscr{H}\left(K^{\left(S\right)}\right)},\;\forall\left(x,y\right)\in S\times S; \end{align} both p.d. kernels on $S\times S$. For every pair $x,y\in S$, we have the following matrix-inversion formula: \begin{equation} \sum_{z\in S}D^{\left(S\right)}\left(x,z\right)K^{\left(S\right)}\left(z,y\right)=\delta_{x,y},\label{eq:H32} \end{equation} where the summation on the LHS in (\ref{eq:H32}) is a limit over a net of finite subsets $\left\{ F_{i}\right\} _{i\in\mathbb{N}}$, $F_{1}\subset F_{2}\subset\cdots$, s.t. $\cup_{i}F_{i}=S$; and the result is independent of choice of net. \item \label{enu:corH14-2}We get an \uline{induced graph} with $S$ as the set of vertices, and edge set $E$ as follows: $E\subset\left(S\times S\right)\backslash\left(\text{diagonal}\right)$. An edge is a pair $\left(x,y\right)\in\left(S\times S\right)\backslash\left(\text{diagonal}\right)$ such that \[ \left\langle \delta_{x},\delta_{y}\right\rangle _{\mathscr{H}\left(K^{\left(S\right)}\right)}\neq0. \] \end{enumerate} \end{cor} \begin{proof} The result follows from an application of Corollaries \ref{cor:H7} and \ref{cor:H8}, and \remref{H10}. \end{proof} Let $X$, $K$, and $S$ be as stated, $S$ countable infinite, with assumptions as in the previous two results. We showed that then the subset $S$ acquires the structure of a vertex set in an induced infinite graph (\corref{H14} (\ref{enu:corH14-2})). If $\Delta$ denotes the corresponding graph Laplacian, then the following boundary value problem is of great interest: Make precise the boundary conditions at \textquotedblleft infinity\textquotedblright{} for this graph Laplacian $\Delta$. An answer to this will require identification of Hilbert space, and limit at \textquotedblleft infinity.\textquotedblright{} The result below is such an answer, and the limit notion will be, limit over the filter of all finite subsets in $S$; see \corref{H7}. Another key tool in the arguments below will again be the net of orthogonal projections $\left\{ P_{F}\right\} $ from \lemref{H4}, and the convergence results from Corollaries \ref{cor:H5} and \ref{cor:H7}. \begin{cor} Let $X\times X\xrightarrow{\;K\;}\mathbb{R}$, and $S\subset X$ be as in the statement of \corref{H14}. Let $\mathscr{F}_{fin}\left(S\right)$ denote the filter of finite subsets $F\subset S$. Let $\Delta=\Delta_{S}$ be the graph Laplacian defined in (\ref{eq:H2}), i.e., \[ \left(\Delta h\right)\left(x\right):=\left\langle \delta_{x},h\right\rangle _{\mathscr{H}(K^{\left(S\right)})}, \] for all $x\in S$, $h\in\mathscr{H}(K^{\left(S\right)})$. Then the following equivalent conditions hold: \begin{enumerate} \item For all $h\in\mathscr{H}(K^{\left(S\right)})$, \begin{align} \left\Vert h\right\Vert _{\mathscr{H}(K^{\left(S\right)})}^{2} & =\sup_{F\in\mathscr{F}_{fin}\left(S\right)}\left\langle h\big|_{F},\Delta P_{F}h\right\rangle _{l^{2}\left(F\right)}\\ & =\sup_{F\in\mathscr{F}_{fin}\left(S\right)}\left\langle h\big|_{F},K_{F}^{-1}\left(h\big|_{F}\right)\right\rangle _{l^{2}\left(F\right)}.\nonumber \end{align} \item For $\forall F\in\mathscr{F}_{fin}\left(S\right)$, $x\in F$, $h\in\mathscr{H}(K^{\left(S\right)})$, \begin{equation} \left(\Delta\left(P_{F}h\right)\right)\left(x\right)=\left(K_{F}^{-1}\left(h\big|_{F}\right)\right)\left(x\right).\label{eq:H46} \end{equation} \item $K_{F}\Delta P_{F}h=h\big|_{F}$. \end{enumerate} \end{cor} \begin{proof} On account of \corref{H8}, we only need to verify (\ref{eq:H46}). Let $F\in\mathscr{F}_{fin}\left(S\right)$, $h\in\mathscr{H}(K^{\left(S\right)})$, then we proved that \begin{align} \left(P_{F}h\right)\left(\cdot\right) & =\sum_{y\in F}\xi_{y}K\left(\cdot,y\right)\;\text{with}\label{eq:H47}\\ \xi_{y} & =\left(K_{F}^{-1}\left(h\big|_{F}\right)\right)\left(y\right). \end{align} Now apply $\left\langle \delta_{x},\cdot\right\rangle _{\mathscr{H}(K^{\left(S\right)})}$ to both sides in (\ref{eq:H47}); and we get \begin{equation} \left(\Delta\left(P_{F}h\right)\right)\left(x\right)=\xi_{x}\label{eq:H49} \end{equation} where we used $\left\langle \delta_{x},K\left(\cdot,y\right)\right\rangle _{\mathscr{H}(K^{\left(S\right)})}=\delta_{x,y}$. The desired conclusion (\ref{eq:H46}) now follows from (\ref{eq:H49}). Also note that $\left(\Delta\left(P_{F}h\right)\right)\left(x\right)=0$ if $x\in X\backslash F$. \end{proof} \subsection{Canonical isometries computed from point processes} Below we consider p.d. kernels $K$ defined initially on $X\times X$. Our present aim is to consider restrictions to $S\times S$ when $S$ is a suitable subset of $X$. Our first observation is the identification of a canonical isometry $T_{S}$ between the respective reproducing kernel Hilbert spaces; $T_{S}$ identifying $\mathscr{H}(K^{\left(S\right)})$ as an isometric subspace inside $\mathscr{H}(K)$. This isometry $T_{S}$ exists in general. However, we shall show that, when the subset $S$ is further restricted, the respective RKHSs, and isometry $T_{S}$ will admit explicit characterizations. For example, if $S$ is countable, and is the Dirac functions $\delta_{s}$, $s\in S$, are in $\mathscr{H}(K^{\left(S\right)})$ we shall show that this setting leads to a point process. In this case, we further identify an induced (infinite) graph with the set $S$ as vertices, and with associated edges defined by an induced $\delta_{s}$ kernel. \begin{thm} \label{thm:H14}Let $X\times X\xrightarrow{\;K\;}\mathbb{C}$ be a p.d. kernel, and let $S\subset X$ be a subset. Set $K^{\left(S\right)}:=K\big|_{S\times S}$. Let $\mathscr{H}\left(K\right)$, and $\mathscr{H}(K^{\left(S\right)})$, be the respective RKHSs. \begin{enumerate} \item Then there is a canonical isometric embedding \[ \mathscr{H}(K^{\left(S\right)})\xrightarrow{\;T\;}\mathscr{H}\left(K\right), \] given by the following formula: For $s\in S$, set \begin{equation} T(K^{\left(S\right)}\left(\cdot,s\right))=K\left(\cdot,s\right).\label{eq:H35} \end{equation} (Note that $K^{\left(S\right)}\left(\cdot,s\right)$ on the LHS in (\ref{eq:H35}) is a function on $S$, while $K\left(\cdot,s\right)$ on the RHS is a function on $X$.) \item \label{enu:H14-2}The adjoint operator $T^{*}$, \begin{equation} \mathscr{H}\left(K\right)\xrightarrow{\;T^{*}\;}\mathscr{H}(K^{\left(S\right)}) \end{equation} is given by restriction, i.e., if $f\in\mathscr{H}(K)$, and $s\in S$, then $\left(T^{*}f\right)\left(s\right)=f\left(s\right)$; or equivalently, for all $f\in\mathscr{H}(K)$, \begin{equation} T^{*}f=f\big|_{S}.\label{eq:H37} \end{equation} \end{enumerate} \end{thm} \begin{proof} To show that $T$ in (\ref{eq:H35}) is isometric, proceed as follows: Let $\left\{ s_{i}\right\} _{i=1}^{N}$ be a finite subset of $S$, and $\left\{ \xi_{i}\right\} _{i=1}^{N}\in\mathbb{C}^{N}$, then \begin{eqnarray*} \left\Vert T(\sum\nolimits _{i}\xi_{i}K^{\left(S\right)}\left(\cdot,s_{i}\right))\right\Vert _{\mathscr{H}\left(K\right)}^{2} & = & \left\Vert \sum\nolimits _{i}\xi_{i}T(K^{\left(S\right)}\left(\cdot,s_{i}\right))\right\Vert _{\mathscr{H}\left(K\right)}^{2}\\ & \underset{\text{by \ensuremath{\left(\ref{eq:H35}\right)}}}{=} & \left\Vert \sum\nolimits _{i}\xi_{i}K\left(\cdot,s_{i}\right)\right\Vert _{\mathscr{H}\left(K\right)}^{2}\\ & \underset{\text{by \ensuremath{\left(\ref{eq:A3}\right)}}}{=} & \sum\nolimits _{i}\sum\nolimits _{j}\overline{\xi}_{i}\xi_{j}K\left(s_{i},s_{j}\right)\\ & = & \sum\nolimits _{i}\sum\nolimits _{j}\overline{\xi}_{i}\xi_{j}K^{\left(S\right)}\left(s_{i},s_{j}\right)\\ & = & \left\Vert \sum\nolimits _{i}\xi_{i}K^{\left(S\right)}\left(\cdot,s_{i}\right)\right\Vert _{\mathscr{H}\left(K^{\left(S\right)}\right)}^{2} \end{eqnarray*} which is the desired isometric property. We now turn to (\ref{eq:H37}), the restriction formula: Let $s\in S$, and $f\in\mathscr{H}\left(K\right)$, then \begin{eqnarray} \left\langle T(K^{\left(S\right)}\left(\cdot,s\right)),f\right\rangle _{\mathscr{H}\left(K\right)} & = & \left\langle K^{\left(S\right)}\left(\cdot,s\right),T^{*}f\right\rangle _{\mathscr{H}\left(K^{\left(S\right)}\right)}\label{eq:H38}\\ & \underset{\text{by \ensuremath{\left(\ref{eq:A4}\right)}}}{=} & \left(T^{*}f\right)\left(s\right).\nonumber \end{eqnarray} But, for the LHS in (\ref{eq:H38}), we have \[ \left\langle T(K^{\left(S\right)}\left(\cdot,s\right)),f\right\rangle _{\mathscr{H}\left(K\right)}\underset{\text{by \ensuremath{\left(\ref{eq:H35}\right)}}}{=}\left\langle K\left(\cdot,s\right),f\right\rangle _{\mathscr{H}\left(K\right)}\underset{\text{by \ensuremath{\left(\ref{eq:A4}\right)}}}{=}f\left(s\right); \] and so the desired formula (\ref{eq:H37}) follows. \end{proof} \begin{rem} \label{rem:H8}\textbf{The canonical isometry for \exaref{H8} ($\mathbb{Z}$-discretization of the covariance function for Brownian motion on $\mathbb{R}$).} From \thmref{H14}, we know that the canonical isometry $T$ maps $\mathscr{H}(K^{\left(Z\right)})$ into $\mathscr{H}\left(K\right)$; see (\ref{eq:H22}). But (\ref{eq:H19}) and (\ref{eq:H24-1}) in the Example offer exact characterization of these two Hilbert spaces. So, in the special case of \exaref{H8}, the canonical isometry $T$ maps from functions $\Phi$ on $\mathbb{Z}$ into functions on $\mathbb{R}$. In view of (\ref{eq:H19}), this assignment turns out to be a precise spline realization of the point grids realized by these sequences $\Phi$. Below we present an explicit formula, and graphics, for the spline realizations. By (\ref{eq:H26-1}), the embedding of $\delta_{n}$ from $\mathscr{H}(K^{\left(\mathbb{Z}\right)})$ into $\mathscr{H}\left(K\right)$ is given by \[ \left(T\delta_{n}\right)\left(x\right)=2K\left(x,n\right)-K\left(x,n+1\right)-K\left(x,n-1\right),\;\forall x\in\mathbb{R}. \] See \figref{H1}. Therefore, for all $h\in\mathscr{H}\left(K\right)$, we get \begin{align*} \left(T^{*}h\right)\left(m\right) & =\sum_{n\in\mathbb{Z}}h\left(n\right)\delta_{n}\left(m\right),\;m\in\mathbb{Z},\;\text{and}\\ \left(TT^{*}h\right)\left(x\right) & =\sum_{n\in\mathbb{Z}}h\left(n\right)\left(2K\left(x,n\right)-K\left(x,n+1\right)-K\left(x,n-1\right)\right),\;x\in\mathbb{R} \end{align*} which is the spline interpolation. \end{rem} \begin{figure} \includegraphics[width=0.6\columnwidth]{sp1} \caption{\label{fig:H1}Isometric extrapolation from functions on $\mathbb{Z}$ to functions on $\mathbb{R}$. An illustration of the isometric embedding of $\delta_{n}$ from $\mathscr{H}(K^{\left(\mathbb{Z}\right)})$ into $\mathscr{H}\left(K\right)$, with $n=3$.} \end{figure} \begin{cor} Let $X\times X\xrightarrow{\;K\;}\mathbb{C}$ be a p.d. kernel, and let $S\subset X$ be a subset. Assume further that $\left\{ \delta_{s}\right\} _{s\in S}\subset\mathscr{H}(K^{\left(S\right)})$. Then every finitely supported function $h$ on $S$ is in $\mathscr{H}(K^{\left(S\right)})$, and we have the following generalized spline interpolation; i.e., isometrically extending $h$ from $S$ to $X$: \begin{equation} \widetilde{h}\left(x\right)=\sup\nolimits _{F\supset F_{0}}\sum\nolimits _{y\in F}\left(K_{F}^{-1}h_{F}\right)\left(y\right)K\left(y,x\right),\;x\in X, \end{equation} where $F_{0}=suppt\left(h\right)$, and the sup is taken over the filter of all finite subsets of $X$ containing $F_{0}$. \end{cor} \begin{proof} Assume $h\in\mathscr{H}(K^{\left(S\right)})$, supported on a finite subset $F_{0}\subset S$. Then, \begin{align*} \widetilde{h}\left(x\right):=Th\left(x\right) & =T\left(\sum\nolimits _{s\in F_{0}}h\left(s\right)\delta_{s}\right)\left(x\right)\\ & =\sum\nolimits _{s\in F_{0}}h\left(s\right)\left(T\delta_{s}\right)\left(x\right)\\ & =\sum\nolimits _{s\in F_{0}}h\left(s\right)\sup\nolimits _{F\supset F_{0}}\left(P_{F}\delta_{s}\right)\left(x\right)\\ & =\sup\nolimits _{F\supset F_{0}}P_{F}\left(\sum\nolimits _{s\in F_{0}}h\left(s\right)\delta_{s}\right)\left(x\right)\\ & =\sup\nolimits _{F\supset F_{0}}\left(P_{F}h_{F_{0}}\right)\left(x\right)\\ & =\sup\nolimits _{F\supset F_{0}}\sum\nolimits _{y\in F}\left(K_{F}^{-1}h_{F}\right)\left(y\right)K\left(x,y\right), \end{align*} where the last step follows from (\ref{eq:H8}), and $P_{F}$ is the orthogonal projection from $\mathscr{H}\left(K\right)$ onto the subspace $\mathscr{H}_{K}\left(F\right)$. \end{proof} \begin{cor} \label{cor:H15}Let $X\times X\xrightarrow{\;K\;}\mathbb{C}$, p.d.. be given, and let $S\subset X$ be a subset. Let $T=T_{S}$, $\mathscr{H}(K^{\left(S\right)})\xrightarrow{\;T\;}\mathscr{H}\left(K\right)$, be the canonical isometry. Then a function $f$ in $\mathscr{H}\left(K\right)$ satisfies $\left\langle f,T(\mathscr{H}(K^{\left(S\right)}))\right\rangle _{\mathscr{H}\left(K\right)}=0$ if and only if \begin{equation} f\left(s\right)=0\;\text{for all \ensuremath{s\in S}.}\label{eq:H39} \end{equation} \end{cor} \begin{proof} Immediate from part (\ref{enu:H14-2}) in \thmref{H14}. \end{proof} \begin{rem} Let $\left(X,K,S\right)$ be as in \corref{H15}, and let $T_{S}$ be the canonical isometry. Let $P_{S}:=T_{S}T_{S}^{*}$ be the corresponding projection. Then $I_{\mathscr{H}\left(K\right)}-P_{S}$ is the projection onto the subspace given in (\ref{eq:H39}). \end{rem} \begin{cor} Let $X\times X\xrightarrow{\;K\;}\mathbb{C}$ be given p.d.; and let $S\subset X$ be a subset with induced kernel \begin{equation} K^{\left(S\right)}:=K\big|_{S\times S}.\label{eq:H40} \end{equation} Consider the two sets $\mathscr{F}\left(S\right)$ and $\mathscr{F}(K^{\left(S\right)})$ from (\ref{eq:D2}) and \thmref{E1}. Let $T_{S}:\mathscr{H}(K^{\left(S\right)})\rightarrow\mathscr{H}\left(K\right)$ be the canonical isometry (\ref{eq:H35}) in \thmref{H14}. Then the following implication holds: \begin{eqnarray} \left(\left\{ k_{x}\right\} _{x\in X},\mu\right) & \in & \mathscr{F}\left(K\right)\label{eq:H41}\\ & \Downarrow\nonumber \\ \left(\left\{ k_{s}\right\} _{s\in S},\mu\right) & \in & \mathscr{F}(K^{\left(S\right)})\label{eq:H42} \end{eqnarray} \end{cor} \begin{proof} Assuming (\ref{eq:H41}), we get the representation (\ref{eq:D2}): \begin{equation} K\left(x,y\right)=\int_{M}\overline{k}_{x}k_{y}d\mu,\;\forall\left(x,y\right)\in X\times X. \end{equation} But then, for all $\left(s_{1},s_{2}\right)\in S\times S$, we then have \begin{eqnarray*} K^{\left(S\right)}\left(s_{1},s_{2}\right) & = & \left\langle T_{S}(K^{\left(S\right)}\left(\cdot,s_{1}\right)),T_{S}(K^{\left(S\right)}\left(\cdot,s_{2}\right))\right\rangle _{\mathscr{H}\left(K\right)}\\ & \underset{\text{by \ensuremath{\left(\ref{eq:H40}\right)}}}{=} & K\left(s_{1},s_{2}\right)\\ & = & \int_{M}\overline{k}_{s_{1}}k_{s_{2}}d\mu, \end{eqnarray*} which is the desired conclusion. \end{proof} \section{Boundary value problems} Our setting in the present section is the discrete case, i.e., RKHSs of functions defined on a prescribed countable infinite discrete set $S$. We are concerned with a characterization of those RKHSs $\mathscr{H}$ which contain the Dirac masses $\delta_{x}$ for all points $x\in S$. Of the examples and applications where this question plays an important role, we emphasize two: (i) discrete Brownian motion-Hilbert spaces, i.e., discrete versions of the Cameron-Martin Hilbert space; (ii) energy-Hilbert spaces corresponding to graph-Laplacians. The problems addressed here are motivated in part by applications to analysis on infinite weighted graphs, to stochastic processes, and to numerical analysis (discrete approximations), and to applications of RKHSs to machine learning. Readers are referred to the following papers, and the references cited there, for details regarding this: \cite{MR3231624,MR2966130,MR2793121,MR3286496,MR3246982,MR2862151,MR3096457,MR3049934,MR2579912,MR741527,MR3024465}. The discrete case can be understood as restrictions of analogous PDE-models. In traditional numerical analysis, one builds discrete and algorithmic models (finite element methods), each aiming at finding approximate solutions to PDE-boundary value problems. They typically use multiresolution-subdivision schemes, applied to the continuous domain, subdividing into simpler discretized parts, called finite elements. And with variational methods, one then minimize various error-functions. In this paper, we turn the tables: our object of study are the discrete models, and analysis of suitable continuous PDE boundary problems serve as a tool for solutions in the discrete world. \begin{defn} \label{def:dmp}Let $X\times X\xrightarrow{\;K\;}\mathbb{C}$ be a given p.d. kernel on $X$. The RKHS $\mathscr{H}=\mathscr{H}\left(K\right)$ is said to have the \emph{discrete mass} property ($\mathscr{H}$ is called a \emph{discrete RKHS}), if $\delta_{x}\in\mathscr{H}$, for all $x\in X$. \end{defn} In fact, it is known (\cite{MR3507188}) that every fundamental solution for a Dirichlet boundary value problem on a bounded open domain $\Omega$ in $\mathbb{R}^{\nu}$, allows for discrete restrictions (i.e., vertices sampled in $\Omega$), which have the desired ``discrete mass'' property. We recall the following result to stress the distinction of the discrete models vs their continuous counterparts. Let $\Omega$ be a bounded, open, and connected domain in $\mathbb{R}^{\nu}$ with smooth boundary $\partial\Omega$. Let $K:\Omega\times\Omega\rightarrow\mathbb{R}$ continuous, p.d., given as the Green's function of $\Delta_{0}$, where \begin{equation} \begin{split} & \Delta_{0}:=-\sum_{j=1}^{\nu}\left(\frac{\partial}{\partial x_{j}}\right)^{2},\\ & dom\left(\Delta_{0}\right)=\left\{ f\in L^{2}\left(\Omega\right)\:\big|\:\Delta f\in L^{2}\left(\Omega\right),\;\mbox{and }f\big|_{\partial\Omega}\equiv0\right\} . \end{split} \label{eq:e4} \end{equation} for the Dirichlet boundary condition. Thus, $\Delta_{0}$ is positive selfadjoint, and \begin{align} & \Delta_{0}K=\delta\left(x-y\right)\text{ on \ensuremath{\Omega\times\Omega}}\label{eq:m1}\\ & K\left(x,\cdot\right)\big|_{\partial\Omega}\equiv0.\label{eq:m2} \end{align} Let $\mathscr{H}_{CM}\left(\Omega\right)$ be the corresponding Cameron-Martin RKHS. For $\nu=1$, $\Omega=\left(0,1\right)$, take \begin{equation} \begin{split}\mathscr{H}_{CM}\left(0,1\right)= & \Big\{ f\:\big|\:f'\in L^{2}\left(0,1\right),\;f\left(0\right)=f\left(1\right)=0,\\ & \left\Vert f\right\Vert _{CM}^{2}:=\int_{0}^{1}\left|f'\right|^{2}dx<\infty\Big\} \end{split} \label{eq:m3} \end{equation} For $\nu>1$, let \begin{equation} \begin{split}\mathscr{H}_{CM}\left(\Omega\right)= & \left\{ f\:\big|\:\nabla f\in L^{2}\left(\Omega\right),\:f\big|_{\partial\Omega}\equiv0,\:\left\Vert f\right\Vert _{CM}^{2}:=\int_{\Omega}\left|\nabla f\right|^{2}dx<\infty\right\} ,\\ & \text{ where }\nabla=\left(\frac{\partial}{\partial x_{1}},\frac{\partial}{\partial x_{2}},\cdots,\frac{\partial}{\partial x_{\nu}}\right). \end{split} \label{eq:m4} \end{equation} \begin{thm} \label{thm:main}Let $\Omega$, and $S\subset\Omega$, be given. Then \begin{enumerate} \item Discrete case: Fix $S\subset\Omega$, $\#S=\aleph_{0}$, where $S=\left\{ x_{j}\right\} _{j=1}^{\infty}$, $x_{j}\in\Omega$. Assume $\exists\varepsilon>0$ s.t. $\left\Vert x_{i}-x_{j}\right\Vert \geq\varepsilon$, $\forall i,j$, $i\neq j$. Let \[ \mathscr{H}\left(S\right)=\text{RKHS of \ensuremath{K^{\left(S\right)}:=K\big|_{S\times S}}}; \] then $\delta_{x_{j}}\in\mathscr{H}\left(S\right)$. \item Continuous case; by contrast: $K_{x}^{\left(S\right)}\in\mathscr{H}_{CM}\left(S\right)$, but $\delta_{x}\notin\mathscr{H}_{CM}\left(\Omega\right)$, $x\in\Omega$. \end{enumerate} \end{thm} \begin{proof} The result follows from an application of Corollaries \ref{cor:H7} and \ref{cor:H8}. It extends earlier results \cite{MR3450534,MR3507188} by the co-authors. \end{proof} \section{Sampling in $\mathscr{H}\left(K\right)$} In the present section, we study classes of reproducing kernels $K$ on general domains with the property that there are non-trivial restrictions to countable discrete sample subsets $S$ such that every function in $\mathscr{H}\left(K\right)$ has an $S$-sample representation. In this general framework, we study properties of positive definite kernels $K$ with respect to sampling from ``small\textquotedblright{} subsets, and applying to all functions in the associated Hilbert space $\mathscr{H}\left(K\right)$. We are motivated by concrete kernels which are used in a number of applications, for example, on one extreme, the Shannon kernel for band-limited functions, which admits many sampling realizations; and on the other, the covariance kernel of Brownian motion which has no non-trivial countable discrete sample subsets. \begin{defn} \label{def:J1}Let $X\times X\xrightarrow{\;K\;}\mathbb{C}$ be a p.d. kernel, and $\mathscr{H}\left(K\right)$ be the associated RKHS. We say that $K$ has non-trivial sampling property, if there exists a countable subset $S\subset X$, and $a,b\in\mathbb{R}_{+}$, such that \begin{equation} a\sum_{s\in S}\left|f\left(s\right)\right|^{2}\leq\left\Vert f\right\Vert _{\mathscr{H}\left(K\right)}^{2}\leq b\sum_{s\in S}\left|f\left(s\right)\right|^{2},\quad\forall f\in\mathscr{H}\left(K\right).\label{eq:sp1} \end{equation} If equality holds in (\ref{eq:sp1}) with $a=b=1$, then we say that $\left\{ K\left(\cdot,s\right)\right\} _{s\in S}$ is a Parseval frame. (Also see \defref{G1}.) \end{defn} It follows that sampling holds in the form \[ f\left(x\right)=\sum_{s\in S}f\left(s\right)K\left(x,s\right),\quad\forall f\in\mathscr{H}\left(K\right),\:\forall x\in X \] if and only if $\left\{ K\left(\cdot,s\right)\right\} _{s\in S}$ is a Parseval frame. \begin{lem} \label{lem:fr}Suppose $K$, $X$, $a$, $b$, and $S$ satisfy the condition in (\ref{eq:sp1}), then the linear span of $\left\{ K\left(\cdot,s\right)\right\} _{s\in S}$ is dense in $\mathscr{H}\left(K\right)$. Moreover, there is a positive operator $B$ in $\mathscr{H}\left(K\right)$ with bounded inverse such that \[ f\left(\cdot\right)=\sum_{s\in S}\left(Bf\right)\left(s\right)K\left(\cdot,s\right) \] is a convergent interpolation formula valid for all $f\in\mathscr{H}\left(K\right)$. Equivalently, \[ f\left(x\right)=\sum_{s\in S}f\left(s\right)B\left(K\left(\cdot,s\right)\right)\left(x\right),\;\text{for all \ensuremath{x\in X}.} \] \end{lem} \begin{proof} Define $A:\mathscr{H}\left(K\right)\rightarrow l^{2}\left(S\right)$ by $\left(Af\right)\left(s\right)=f\left(s\right)$, $s\in S$. Then the adjoint operator $A^{*}:l^{2}\left(S\right)\rightarrow\mathscr{H}\left(K\right)$ is given by $A^{*}\xi=\sum_{s\in S}\xi_{s}K\left(\cdot,s\right)$, $\forall\xi\in l^{2}\left(S\right)$, and \[ A^{*}Af=\sum_{s\in S}f\left(s\right)K\left(\cdot,s\right) \] holds in $\mathscr{H}\left(K\right)$, with $\mathscr{H}\left(K\right)$-norm convergence. Now set $B=\left(A^{*}A\right)^{-1}$, and note that $\left\Vert B\right\Vert _{\mathscr{H}\left(K\right)\rightarrow\mathscr{H}\left(K\right)}\leq a^{-1}$, where $a$ is in the lower bound in (\ref{eq:sp1}). \end{proof} \begin{thm} \label{thm:ps}Let $K:X\times X\rightarrow\mathbb{R}$ be a p.d. kernel, and let $S\subset X$ be a countable discrete subset. For all $s\in S$, set $K_{s}\left(\cdot\right)=K\left(\cdot,s\right)$. Then TFAE: \begin{enumerate} \item \label{enu:ps1}The family $\left\{ K_{s}\right\} _{s\in S}$ is a Parseval frame in $\mathscr{H}\left(K\right)$; \item \label{enu:ps2} \[ \left\Vert f\right\Vert _{\mathscr{H}\left(K\right)}^{2}=\sum_{s\in S}\left|f\left(s\right)\right|^{2},\;\forall f\in\mathscr{H}\left(K\right); \] \item \label{enu:ps3} \[ K\left(x,x\right)=\sum_{s\in S}\left|K\left(x,s\right)\right|^{2},\;\forall x\in X; \] \item \label{enu:ps4} \[ f\left(x\right)=\sum_{s\in S}f\left(s\right)K\left(x,s\right),\;\forall f\in\mathscr{H}\left(K\right),\:\forall x\in X, \] where the sum converges in the norm of $\mathscr{H}\left(K\right)$. \end{enumerate} \end{thm} \begin{proof} The proof is simple, and follows the steps in the proof of \lemref{G2}. Details are left to the reader. \end{proof} We now turn to dichotomy: Existence of countably discrete sampling sets vs non-existence. \begin{example} \label{exa:shan}Let $X=\mathbb{R}$, and let $K:\mathbb{R}\times\mathbb{R}\rightarrow\mathbb{R}$ be the Shannon kernel, where \begin{align} K\left(x,y\right) & :=\text{sinc}\,\pi\left(x-y\right)\nonumber \\ & =\frac{\sin\pi\left(x-y\right)}{\pi\left(x-y\right)},\quad\forall x,y\in\mathbb{R}.\label{eq:sp5} \end{align} We may choose $S=\mathbb{Z}$, and then $\left\{ K\left(\cdot,n\right)\right\} _{n\in\mathbb{Z}}$ is even an orthonormal basis (ONB) in $\mathscr{H}\left(K\right)$, but there are many other examples of countable discrete subsets $S\subset\mathbb{R}$ such that (\ref{eq:sp1}) holds for finite $a,b\in\mathbb{R}_{+}$. The RKHS $\mathscr{H}\left(K\right)$ in (\ref{eq:sp5}) is the Hilbert space $\subset L^{2}\left(\mathbb{R}\right)$ consisting of all $f\in L^{2}\left(\mathbb{R}\right)$ such that $suppt(\hat{f})\subset\left[-\pi,\pi\right]$, where ``suppt'' stands for support of the Fourier transform $\hat{f}$. Note $\mathscr{H}\left(K\right)$ consists of functions on $\mathbb{R}$ which have entire analytic extensions to $\mathbb{C}$. Using the above observations, we get \begin{align*} f\left(x\right) & =\sum_{n\in\mathbb{Z}}f\left(n\right)K\left(x,n\right)\\ & =\sum_{n\in\mathbb{Z}}f\left(n\right)\text{sinc}\,\pi\left(x-n\right),\quad\forall x\in\mathbb{R},\:\forall f\in\mathscr{H}\left(K\right). \end{align*} \end{example} \begin{example} Let $K$ be the covariant kernel of standard Brownian motion, with $X:=[0,\infty)$ or $[0,1)$, and \begin{equation} K\left(x,y\right):=x\wedge y=\min\left(x,y\right),\;\forall\left(x,y\right)\in X\times X.\label{eq:sp6} \end{equation} \end{example} \begin{thm} \label{thm:bm}Let $K$, $X$ be as in (\ref{eq:sp6}); then there is no countable discrete subset $S\subset X$ such that $\left\{ K\left(\cdot,s\right)\right\} _{s\in S}$ is dense in $\mathscr{H}\left(K\right)$. \end{thm} \begin{proof} Suppose $S=\left\{ x_{n}\right\} $, where \begin{equation} 0<x_{1}<x_{2}<\cdots<x_{n}<x_{n+1}<\cdots;\label{eq:sp7} \end{equation} then consider the following function \begin{equation} \raisebox{-6mm}{\includegraphics[width=0.7\textwidth]{wave1.pdf}}\label{eq:sp9} \end{equation} On the respective intervals $\left[x_{n},x_{n+1}\right]$, the function $f$ is as follows: \[ f\left(x\right)=\begin{cases} c_{n}\left(x-x_{n}\right) & \text{if }x_{n}\leq x\leq\frac{x_{n}+x_{n+1}}{2}\\ c_{n}\left(x_{n+1}-x\right) & \text{if }\frac{x_{n}+x_{n+1}}{2}<x\leq x_{n+1}. \end{cases} \] In particular, $f\left(x_{n}\right)=f\left(x_{n+1}\right)=0$, and on the midpoints: \[ f\left(\frac{x_{n}+x_{n+1}}{2}\right)=c_{n}\frac{x_{n+1}-x_{n}}{2}, \] see \figref{stooth}. \begin{figure}[H] \includegraphics[width=0.35\textwidth]{stooth} \caption{\label{fig:stooth}The saw-tooth function.} \end{figure} Choose $\left\{ c_{n}\right\} _{n\in\mathbb{N}}$ such that \begin{equation} \sum_{n\in\mathbb{N}}\left|c_{n}\right|^{2}\left(x_{n+1}-x_{n}\right)<\infty.\label{eq:sp11} \end{equation} Admissible choices for the slope-values $c_{n}$ include \[ c_{n}=\frac{1}{n\sqrt{x_{n+1}-x_{n}}},\;n\in\mathbb{N}. \] We will now show that $f\in\mathscr{H}\left(K\right)$. For the distribution derivative computed from (\ref{eq:sp9}), we get \begin{equation} \raisebox{-12mm}{\includegraphics[width=0.7\textwidth]{wave2.pdf}}\label{eq:sp9b} \end{equation} \[ \int_{0}^{\infty}\left|f'\left(x\right)\right|^{2}dx=\sum_{n\in\mathbb{N}}\left|c_{n}\right|^{2}\left(x_{n+1}-x_{n}\right)<\infty \] which is the desired conclusion, see (\ref{eq:sp9}). \end{proof} \begin{cor} For the kernel $K\left(x,y\right)=x\wedge y$ in (\ref{eq:sp6}), $X=[0,\infty)$, the following holds: Given $\left\{ x_{j}\right\} _{j\in\mathbb{N}}\subset\mathbb{R}_{+}$, $\left\{ y_{j}\right\} _{j\in\mathbb{N}}\subset\mathbb{R}$, then the interpolation problem \begin{equation} f\left(x_{j}\right)=y_{j},\;f\in\mathscr{H}\left(K\right)\label{eq:ip1} \end{equation} is solvable if \begin{equation} \sum_{j\in\mathbb{N}}\left(y_{j+1}-y_{j}\right)^{2}/\left(x_{j+1}-x_{j}\right)<\infty.\label{eq:sp2} \end{equation} \end{cor} \begin{proof} Let $f$ be the piecewise linear spline (see \figref{ip}) for the problem (\ref{eq:ip1}), see \figref{ip}; then the $\mathscr{H}\left(K\right)$-norm is as follows: \[ \int_{0}^{\infty}\left|f'\left(x\right)\right|^{2}dx=\sum_{j\in\mathbb{N}}\left(\frac{y_{j+1}-y_{j}}{x_{j+1}-x_{j}}\right)^{2}\left(x_{j+1}-x_{j}\right)<\infty \] when (\ref{eq:sp2}) holds. \end{proof} \begin{figure}[H] \includegraphics[width=0.35\textwidth]{spline} \caption{\label{fig:ip}Piecewise linear spline.} \end{figure} \begin{rem} Let $K$ be as in (\ref{eq:sp6}), $X=[0,\infty)$. For all $0\leq x_{j}<x_{j+1}<\infty$, let \begin{align*} f_{j}\left(x\right): & =\frac{2}{x_{j+1}-x_{j}}\left(K\left(x-x_{j},\frac{x_{j+1}-x_{j}}{2}\right)-K\left(x-\frac{x_{j}+x_{j+1}}{2},\frac{x_{j+1}-x_{j}}{2}\right)\right)\\ & =\raisebox{-5mm}{\includegraphics[width=0.4\textwidth]{tmp.pdf}} \end{align*} Assuming (\ref{eq:sp11}) holds, then \[ f\left(x\right)=\sum_{j}c_{j}f_{j}\left(x\right)\in\mathscr{H}\left(K\right). \] \end{rem} \begin{thm} Let $X$ be a set of cardinality $c$ of the continuum, and let $K:X\times X\rightarrow\mathbb{R}$ be a positive definite kernel. Let $S=\left\{ x_{j}\right\} _{j\in\mathbb{N}}$ be a discrete subset of $X$. Suppose there are weights $\left\{ w_{j}\right\} _{j\in\mathbb{N}}$, $w_{j}\in\mathbb{R}_{+}$, such that \begin{equation} \left(f\left(x_{j}\right)\right)\in l^{2}\left(\mathbb{N},w\right)\label{eq:c1} \end{equation} for all $f\in\mathscr{H}\left(K\right)$. Suppose further that there is a point $t_{0}\in X\backslash S$, a $y_{0}\in\mathbb{R}\backslash\left\{ 0\right\} $, and $\alpha\in\mathbb{R}_{+}$ such that the infimum \begin{equation} \inf_{f\in\mathscr{H}\left(K\right)}\left\{ \sum\nolimits _{j}w_{j}\left|f\left(x_{j}\right)\right|^{2}+\left|f\left(t_{0}\right)-y_{0}\right|^{2}+\alpha\left\Vert f\right\Vert _{\mathscr{H}\left(K\right)}^{2}\right\} \label{eq:c2} \end{equation} is strictly positive. Then $S$ is \uline{not} a interpolation set for $\left(K,X\right)$. \end{thm} \begin{proof} This results follows from \lemref{fr} and \thmref{ps} above. We also refer readers to \cite{MR3670916}. \end{proof} \begin{acknowledgement*} The co-authors thank the following colleagues for helpful and enlightening discussions: Professors Daniel Alpay, Sergii Bezuglyi, Ilwoo Cho, Myung-Sin Song, Wayne Polyzou, and members in the Math Physics seminar at The University of Iowa. \end{acknowledgement*} \bibliographystyle{amsalpha}
{ "timestamp": "2018-12-31T02:13:18", "yymm": "1812", "arxiv_id": "1812.10850", "language": "en", "url": "https://arxiv.org/abs/1812.10850", "abstract": "We establish a duality for two factorization questions, one for general positive definite (p.d) kernels $K$, and the other for Gaussian processes, say $V$. The latter notion, for Gaussian processes is stated via Ito-integration. Our approach to factorization for p.d. kernels is intuitively motivated by matrix factorizations, but in infinite dimensions, subtle measure theoretic issues must be addressed. Consider a given p.d. kernel $K$, presented as a covariance kernel for a Gaussian process $V$. We then give an explicit duality for these two seemingly different notions of factorization, for p.d. kernel $K$, vs for Gaussian process $V$. Our result is in the form of an explicit correspondence. It states that the analytic data which determine the variety of factorizations for $K$ is the exact same as that which yield factorizations for $V$. Examples and applications are included: point-processes, sampling schemes, constructive discretization, graph-Laplacians, and boundary-value problems.", "subjects": "Functional Analysis (math.FA); Probability (math.PR)", "title": "Decomposition of Gaussian processes, and factorization of positive definite kernels", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9877587272306607, "lm_q2_score": 0.8289388019824947, "lm_q1q2_score": 0.8187915359983376 }
https://arxiv.org/abs/2007.09958
Monodromy of general hypersurfaces
Let $X$ be a general complex projective hypersurface in $\mathbb{P}^{n+1}$ of degree $d>1$. A point $P$ not in $X$ is called uniform if the monodromy group of the projection of $X$ from $P$ is isomorphic to the symmetric group. We prove that all the points in $\mathbb{P}^{n+1}$ are uniform for $X$, generalizing a result of Cukierman on general plane curves.
\section{Introduction} The monodromy group of linear projections of irreducible complex projective varieties has been intensively studied. Fixed an irreducible and reduced projective hypersurface $X \subset \mathbb{P}^{n+1}$, consider its linear projections from a point $P \in \mathbb{P}^{n+1}$. We want to look at those maps from a topological point of view: in particular, we aim to classify the centres of projection through their monodromy group. We recall that we can give also an algebraic description: indeed, the monodromy group is isomorphic to the Galois group for finite dominant morphisms between irreducible complex varieties \cite[Section I]{H}. We will say that a point $P$ is \emph{uniform} for $X$ if the monodromy group of the projection from $P$ is the symmetric group, \emph{non uniform} otherwise. A direct consequence of the Castelnuovo's uniform position principle, in the formulation of Harris \cite{JHCurves}, is that a general projection has always symmetric monodromy group. In 2005 Pirola and Schlesinger \cite{PS} improved this result showing that an irreducible and reduced plane curve admits at most a finite number of non uniform points. Moreover, in \cite{CMS} it is proved that smooth surfaces in $\mathbb{P}^3$ admit at most a finite number of non uniform points. More recently the author, Cuzzucoli and Moschetti \cite{CCM} studied the case of hypersurfaces of higher dimension proving that, except for special configurations, the non uniform locus is contained in linear subspaces of codimension two. In particular, we proved that smooth hypersurfaces admit at most a finite number of non uniform points \cite[Theorem 1.3]{CCM}. Examples of smooth hypersurfaces admitting at least a non uniform point are known (see for instance \cite{Miura1} \cite{MY1} for plane curves or \cite{Yoshiara} for hypersurfaces). One may ask if every smooth hypersurface admit non uniform points, but the answer is negative. In 1999 Fernando Cukierman (\cite{Cuk}) proved that for general plane curves, all the outer points are uniform. In this work we generalize this result proving the following \begin{thm} Let $X \subset \mathbb{P}^{n+1}$ be a general hypersurface of degree $d>1$. Then all the points $P \in \mathbb{P}^{n+1}$ are uniform. \end{thm} The result was already known for a special class of non uniform points that are the Galois points (\cite[Theorem 1]{Yoshiara}). We remark also that the Theorem extends the result of Cukiermann to inner points of general plane curves. The proof combines inner and outer projections and it is based on an induction argument on the degree of the variety: we degenerate the hypersurface $X$ to a limit one given by a general hypersurface $Y$ of degree $d-1$ and a hyperplane. The base case of the induction ($d=3$, Theorem \ref{d3}) is consequence of a result of Matsumura and Monsky \cite{MM} saying that a general hypersurface has trivial automorphism group. More in general, the induction step is based on the study of the behaviour of the monodromy group of the projection $\pi_P$ of $X$ from $P$ under degenerations (Lemma \ref{induction}). \begin{lem} Let $P \in \mathbb{P}^{n+1}$ and let $\pi_0$ be the map $\pi_P$ restricted to $Y$. Then, the monodromy group $M(\pi_0)$ is contained in the monodromy group $M(\pi_P)$. \end{lem} This Lemma is based on some classical topological results on homotopy of fibrations (Proposition \ref{lem:Noriprelim}), reported in Section \ref{sec3}. In particular, we considered the case of a family of dominant maps $F: \mathcal{X} \to Y \times \mathbb{P}^1$ parametrised by $\mathbb{P}^1$, where $\mathcal{X}$ is a flat family of projective varieties of dimension $m$ in $\mathbb{P}^N$ and $Y$ is a smooth projective variety. For a general $s \in \mathbb{P}^1$, the fibre over $\mathbb{P}^1$ is a smooth projective variety $X_s$ of dimension $m$ in $\mathbb{P}^N$, together with a finite dominant morphism $f_s: X_s \to Y$ of degree $d$. We deduce a result on monodromy groups (Proposition \ref{lemmagenerale}): \begin{prop} If $X_s$ is reduced for every $s \in \mathbb{P}^1$, then $$M(F)\cong M(f_s).$$ \end{prop} More generally, if there is a non reduced fibre $X_0$ with a reduced component $Z$, we have (Proposition \ref{redcomponent}) \begin{prop} The monodromy group of $f_0$ restricted to $Z$ is contained in the monodromy group $M(f_s)$ for a general $s$ in a neighbourhood of $0$. \end{prop} To conclude the proof of the main Theorem, we use results on multiply transitive permutation groups (see Section \ref{permutazioni}). \medskip \textbf{Notations.} All the varieties are assumed to be complex and projective. Let $\mathscr{F}$ be a family of objects parametrised by a scheme $V$. We say the general element of $\mathscr{F}$ satisfies a certain property if this property holds for every element in a Zariski dense open subset of $V$. Moreover, we will always use the Zariski topology, unless stated otherwise. \section{Preliminaries} \label{sec:preliminaries} \subsection{Monodromy and Galois group} Let $f:X \to Y$ be a finite dominant morphism of degree $d$ between complex irreducible reduced varieties of the same dimension. Let $U \subset Y$ be a Zariski open set over which $f$ is \'etale, and let $y$ denote a point in $U$. We have a well defined map $$\mu: \pi_1(U,y) \to \operatorname{Aut}\big(f^{-1}(y)\big) \simeq S_d.$$ The image $M(f):=\mu\left(\pi_1(U,y)\right)$ is called \emph{monodromy group} of the map $f$; it is a transitive subgroup of the symmetric group. We can also describe this group by means of Galois extensions: let $K$ be the Galois closure of the extension $\mathbb{C}(X)/\mathbb{C}(Y)$, where $\mathbb{C}(X),\mathbb{C}(Y)$ define the fields of rational functions of $X$ and $Y$ respectively. Define the \emph{Galois group} $G(f)$ of the map $f$ to be the Galois group of the field extension $K/\mathbb{C}(Y)$. It turns out that $G(f)$ is isomorphic to $M(f)$, see \cite[Section I]{H}. We recall also that Galois group of a field extension $K/\mathbb{C}(Y)$ is defined as the group of automorphisms of $K$ fixing $\mathbb{C}(Y)$. \subsection{Automorphisms of general hypersurfaces} Let $V$ be a projective variety; we denote by $\operatorname{Aut}(V)$ the group of automorphisms of $V$. We will use the following result of Matsumura and Monsky \cite{MM}: \begin{thm}\label{autom} Let $X$ be a general hypersurface in $\mathbb{P}^{n+1}$, with $ n\geq 2$ and $d\geq 3$. Then $\operatorname{Aut}(X)$ is trivial. \end{thm} \subsection{Permutation groups} \label{permutazioni} We recall some definitions and results that we will use in the following. A group $G$ acting on a set $\ \Omega=\{1,\ldots,d\}$ is $k$-transitive, with $k \leq d$, if, given two ordered $k$-tuples $(m_1,\ldots,m_k)$ and $(t_1,\ldots,t_k)$ of distinct points in $\Omega$, there is an element $g \in G$ such that sends $g \cdot m_i=t_i$ for every $i=1,\ldots,$. If $k=1$ we say that $G$ is transitive. We state some results on transitive permutation groups that we will use in the following. See for instance \cite[Chapter 8]{Isaacs} for a more complete treatment. \begin{lem}\label{lemma1} Let $G$ be a group acting transitively on $\Omega$, let $i \in \Omega$ and $k \leq d-1$. The group $G$ is $k$-transitive on $\Omega$ if and only if the stabilizer of $i$ in $G$ is $(k-1)$-transitive on $\Omega \setminus \{i\}$. \end{lem} We recall that a {block} is a non-empty subset $B \subset \Omega$ such that either $g\cdot B=B$ or $(g\cdot B) \cap B = \emptyset$ for all $g \in G$. We say that $G$ is {imprimitive} if its action preserves non-trivial blocks and {primitive} otherwise. A 2-transitive permutation group is primitive, but the converse is not always true. \begin{lem}\label{block} Let $G$ be a group acting transitively on $\Omega$ and let $B$ be a block. Then $|B|$ divides $d=|\Omega|$ and in $\Omega$ there are exactly ${|\Omega|}/{|B|}$ disjoint blocks, all with the same cardinality. \end{lem} \begin{lem}\label{lemma2} Let $G$ be a primitive group on $\Omega$, let $A \subset \Omega$ such that $0 < |A| \leq d-2$ and the stabilizer of $A$ is transitive on $\Omega \setminus A$. Then $G$ is $2$-transitive on $\Omega$. \end{lem} We recall that the monodromy group of $\pi_P$ is imprimitive if and only if the projection is decomposable (\cite[Remark 2.2]{PS}). \section{Topology of finite morphisms}\label{sec3} We introduce the following definition of what will be for us a \emph{fibration}. For a more complete treatment see \cite[Chapter III sect. 8]{BHPVdV}. \begin{defn} A \emph{fibration} is a proper surjective morphism $f: X \to Y$ with connected fibres from a smooth complex variety to a smooth quasi-projective curve. Let $y \in Y$ be a point. A fibre of $f$ is $F:=f^*(y)=\sum n_iY_i$, where the $Y_i$'s are irreducible components and $n_i \geq 1$ are their multiplicities. A fibre $F$ is called \emph{multiple} if $\gcd\{n_i\}:=m>1$; we will write $F=mE$, where $E=\sum t_iY_i$ with $\gcd\{t_i\}=1$. \end{defn} We remark that $f$ is flat since $Y$ is smooth. We will use the following classical result on homotopy of fibrations. \begin{prop}\label{lem:Noriprelim} Let $f: X \to Y$ be a fibration. If $f$ does not have multiple fibres, then the following sequence is exact \begin{equation*} \pi_1(F) \to \pi_1(X)\to \pi_1(Y)\to 1 \end{equation*} where $F$ is a general fibre of $f$. \end{prop} The proof is based on a combination of the techniques in \cite[Lemma 1.5]{Nori} that proved the result in the case where every fibre has at least a reduced component, and in \cite{Serrano} that proved the exactness of the sequence for homology groups. \subsection{Monodromy group of families of maps} \label{sec:limit} We refer the reader to \cite[ChapterIII.9]{Hartshorne} and \cite[Chapter4.6.7]{Sernesi} for background material about families of algebraic space. Let $\mathcal{X} \to \mathbb{P}^1$ be a (flat) family of projective varieties of dimension $m$ in $\mathbb{P}^N$ parametrized by $\mathbb{P}^1$ and let $Y$ be a smooth projective variety. Consider the following diagram \begin{equation*} \xymatrix{ \mathcal{X} \ar[r]^F \ar[d]^p & Y \times \mathbb{P}^1 \ar[dl]^q \\ \mathbb{P}^1 & } \end{equation*} Let $p$ and $q$ be proper surjective maps with connected fibres. A general fibre $ X_s$ of $p$, with $s \in \mathbb{P}^1$, is a smooth projective variety of dimension $m$ in $\mathbb{P}^N$, together with a finite dominant morphism $f_s: X_s \to Y$ of degree $d$. Denote by $B_s \subset Y$ the branch divisor of $f_s$ and $R_s \subset X_s$ its ramification divisor. Let $M(f_s)$ be the monodromy group of $f_s$. Let $\mathcal{R}\subset \mathcal{X}$ be the ramification divisor of $F$, i.e. $\mathcal{X} \setminus \mathcal{R}=\lbrace X_s \setminus R_s\ |\ s \in \mathbb{P}^1 \rbrace$. Let moreover $\mathcal{B}$ be the branch divisor of $F$, i.e. $(Y \times \mathbb{P}^1) \setminus \mathcal{B}=:\mathcal{V}=\lbrace (Y \setminus B_s) \times \{s\}\ |\ s \in \mathbb{P}^1 \rbrace$, open inside $Y \times \mathbb{P}^1$. Let $M(F)$ be the monodromy group of $F$. If all the varieties $X_s$ are reduced we can deduce the following property of monodromy groups. \begin{prop}\label{lemmagenerale} In the above setting, assume that every fibre of $p$ is reduced; then $$M(F)\cong M(f_s)$$ for a general $s \in \mathbb{P}^1$. \end{prop} \begin{proof} By assumptions, there is no $s \in \mathbb{P}^1$ such that $(Y,s) \subset \mathcal{B}$. Therefore, the map $q': (Y \times \mathbb{P}^1)\setminus \mathcal{B} \to \mathbb{P}^1$ is a fibration with reduced and connected fibres. Thanks to Proposition \ref{lem:Noriprelim}, the following sequence is exact for a general $s \in \mathbb{P}^1$ \begin{equation*} \xymatrix{ \pi_1(Y\setminus B_s) \ar@{->>}[r] & \pi_1((Y \times \mathbb{P}^1) \setminus \mathcal{B}) \ar[r] & \pi_1(\mathbb{P}^1)=1. } \end{equation*} Combining this together with the monodromy map $\mu$, we have \begin{equation*} \xymatrix{ \pi_1(Y \setminus B_s) \ar@{->>}[r] \ar@{->>}[d]^\mu & \pi_1((Y \times \mathbb{P}^1) \setminus \mathcal{B}) \ar@{->>}[d]^\mu\\ M(f_s) \ar@{->>}[r] & M(F)} \end{equation*} Moreover, if we have a subvariety $Z \subset Y\times \mathbb{P}^1$ that is not contained in $\mathcal{B}$, then \begin{equation*}\label{iniettiva} \pi_1(Z) \hookrightarrow \pi_1((Y \times \mathbb{P}^1) \setminus \mathcal{B}). \end{equation*} Taking $Z$ as a general fibre of $q'$, then we have also that the map between the monodromy groups is injective. Hence $M(F) \cong M(f_s)$. \end{proof} \begin{cor} In the above assumptions, let $X_0$ be a fibre of $p$ and $f_0: X_0 \to Y$ its dominant morphism. Then $M(f_0) \subseteq M(f_s)$. \end{cor} More generally, assume that the fibration $p:\mathcal{X} \to \mathbb{P}^1$ has a singular fibre $F_0=\sum n_i Z_i$ with at least a reduced component $Z_i$. Let $g_0$ be the map $f_0$ restricted to $Z_i$, i.e. $g_0 = (f_0)_{|Z_i}:Z_i \to Y$, dominant morphism of degree strictly lower than $d=\deg(f_0)$. \begin{prop}\label{redcomponent} For a general $s$ in a neighbourhood of $\ 0$, $$M(g_0) \subset M(f_s).$$ \end{prop} \begin{proof} Let $Z:=Z_i$ and, by abuse of notation, we will still denote by $B_0$ the branch divisor of $g_0$ and $R_0$ its ramification divisor. Let $\sigma \in M(g_0)$. Then there exists $[\gamma] \in \pi_1(Y \setminus B_0,y)$ such that $\mu (\gamma)=\sigma$. Let $\gamma$ be a representative of $[\gamma]$ and let $\Tilde{\gamma}$ be its lifting to $\Tilde{Z}:=Z\setminus R_0$. The path $\Tilde{\gamma}$ is the image of $[0,1]$ inside $\Tilde{Z}$, such that $\Tilde{\gamma}(0)=z_0$ and $\Tilde{\gamma}(1)=z_1$, where $z_0,z_1$ are two distinct points in the fibre $g_0^{-1}(y)$. We can also assume that we avoid the points in which $Z$ meets the other components $Z_j$ of $F_0$. The path is compact and consider a tubular neighbourhood $U$ of it. Then, by assumptions, the fibration $p$ restricted to $U$ is a locally trivial fibration by Ehresmann's Theorem (\cite[Lemma 4.2]{Catanese}, \cite[Sec 4]{Massey} for manifolds with boundary). Hence, the path $\Tilde{\gamma}$ can be moved in $U$ to a path $\Tilde{\gamma}_s$ in a fibre $X_s,\ s \neq 0$. Therefore, $M(f_s)\owns\mu\left(p(\Tilde{\gamma}_s)\right)= \sigma$. \end{proof} \section{Projections of general hypersurfaces}\label{sec:main} We want now apply the previous construction to the following situation. Let $\mathcal{X} \to \Delta$ be a pencil of hypersurfaces in $\mathbb{P}^{n+1}$ parametrised by a disc $\Delta$, small neighbourhood of $0$. Its general element is a general hypersurface $X$ and the hypersurface $X_0$ is given by a general hypersurface $Y$ of degree $d-1$ and a hyperplane $H$. Let $P \in \mathbb{P}^{n+1}$ be a point and $\mathbb{P}^n$ an hyperplane not containing $P$ and consider \begin{equation*} \xymatrix{ \widetilde{\mathbb{P}^{n+1}} \ar[d]^\nu \ar[dr]^{\widetilde{\pi_P}} & \\ \mathbb{P}^{n+1} \ar@{-->}[r]^{\pi_P} & \mathbb{P}^n } \end{equation*} where $\nu$ is the blow up of the projective space at $P$ and $\pi_P$ is the projection of $\mathbb{P}^{n+1}$ form $P$. Consider the linear projection $\pi_s:= (\pi_P)_{|X_s}: X_s \dashrightarrow \mathbb{P}^n$ of a general element $X_s$ in $\mathcal{X}$ with $s \in \Delta$. Degenerating the hypersurface $X$ to $X_0$ as $t$ goes to $0$, the point $P$ degenerate onto a point $P_0 \in \mathbb{P}^{n+1}$. Note that, if the point $P$ is in $X$, the point $P_0$ is in $X_0$. After a change of coordinates, we can think the point $P$ as fixed. We have the following diagram \begin{equation*} \xymatrix{ \widetilde{\mathcal{X}} \ar[r]^{\widetilde{\pi_P}} \ar[d]^p & \Delta \times \mathbb{P}^n \ar[dl]^q \\ \Delta & } \end{equation*} where $\widetilde{\mathcal{X}}$ is the family of the strict transforms $\widetilde{X_s} \subset \widetilde{\mathbb{P}^{n+1}}$ for every $X_s \subset \mathcal{X}$ and $\widetilde{\pi_s}: \widetilde{X_s} \to \mathbb{P}^n $ is a dominant morphism of degree $d$. We recall that, if $P \notin X_s$, then $\widetilde{\pi_s}=\pi_s$ and moreover, the monodromy group does not change when we blow up a smooth point of $X_s$. \begin{lem}\label{induction} Let $P \in \mathbb{P}^{n+1}$ and let $\pi_0$ be the map $\pi_P$ restricted to $Y$. Then, the monodromy group $M(\pi_0)$ is contained in the monodromy group $M(\pi_s)$ for a general $s \in \Delta$. \end{lem} \begin{proof} If in the limit $P \notin Y \cap H$, the singular locus of $X_0$, then all the varieties in $\widetilde{\mathcal{X}}$ are reduced. We can apply Proposition \ref{lemmagenerale} and have that $M(\widetilde{\pi_s}) \cong M(\widetilde{\pi_P})$. Moreover, we have that $$M(\pi_0) \subset M(\widetilde{\pi}:\widetilde{X_0} \to \mathbb{P}^n) \subset M(\widetilde{\pi_s})=M(\pi_s).$$ If $P \in Y \cap H$, then we have that $\widetilde{Y}$ is a reduced component of $\widetilde{X_0}$ and $P$ is a smooth point of $Y$. By Proposition \ref{redcomponent}, the monodromy group of $\pi_0$ is still contained in the monodromy group of a general fibre $M(\pi_s)$. \end{proof} \begin{comment} Let $P \in \mathbb{P}^{n+1}$ be a point and $\mathbb{P}^n$ an hyperplane not containing $P$ and consider \begin{equation*} \xymatrix{ \widetilde{\mathbb{P}^{n+1}} \ar[d]^\nu \ar[dr]^{\widetilde{\pi_P}} & \\ \mathbb{P}^{n+1} \ar@{-->}[r]^{\pi_P} & \mathbb{P}^n } \end{equation*} where $\nu$ is the blow up the projective space at $P$ and $\pi_P$ is the projection of $\mathbb{P}^{n+1}$ form $P$. Let $X \subset \mathbb{P}^{n+1}$ be a general hypersurface of degree $d$ and let $\pi_s: X \dashrightarrow \mathbb{P}^n$, the restriction of $\pi_P$ to $X$. Let $\mathcal{X}$ be a smooth family of hypersurfaces in $\mathbb{P}^{n+1}$ of degree $d \geq 3$ parametrized by a disc $\Delta$, whose general member is a general hypersurface. Let $X_0$ be a general hypersurface $Y$ of degree $d-1$ and a hyperplane $H$. \begin{lem}\label{induction} Let $P \in \mathbb{P}^{n+1}$. Then, the monodromy group $M(\pi_0)$ is contained in the monodromy group $M(\pi_s)$ for a general $s \in \Delta$. \end{lem} \begin{proof} The variety $\Tilde{X_0}$, the blow up of $X_0$ at $P$, has at least one non reduced irreducible component. Hence we can apply Lemma \ref{lemmagenerale} and conclude that $M(\pi_0) \subset M(\pi_s)$. \end{proof} \end{comment} \medskip \subsection{Monodromy of general hypersurfaces} Let $X \subset \mathbb{P}^{n+1}$ be a general hypersurface of degree $d >1$ and let $P \in \mathbb{P}^{n+1}$ be a point. Let $\pi_P$ be the linear projection of $X$ from $P$ and let $M(\pi_P)$ be its corresponding monodromy group. We recall that if $P \in X$, then $M(\pi_P) \subseteq S_{d-1}$, while if $P \notin X$, then $M(\pi_P)\subseteq S_d$. \begin{rmk}\label{d2} Every point $P$ is uniform if $d=2$. Indeed, there are no proper transitive subgroup of $S_2$. \end{rmk} As a consequence of Theorem \ref{autom} we get the following. \begin{thm}\label{d3} Let $d=3$. Then every point is uniform. \end{thm} \begin{proof} If $P \in X$, the degree of $\pi_P: X \dashrightarrow \mathbb{P}^n$ is two and so $M(\pi_P)=S_2$. Let now $P \notin X$ and assume by contradiction that $P$ is non uniform. Then $M(\pi_P)=A_3$ and so $X$ has a non trivial automorphism. This is a contradiction of Theorem \ref{autom}. Hence $M(\pi_P)=S_3$ for every $P \notin X$. \end{proof} We are now ready to prove the main result of the paper. \begin{thm}\label{interni} Let $X \subset \mathbb{P}^{n+1}$ be a general hypersurface of degree $d>1$. Then all the points $P \in \mathbb{P}^{n+1}$ are uniform. \end{thm} \begin{proof} The result has been already proven for $d \leq 3$ (Theorem \ref{d3}). We work by induction on $d=\deg(X)$. Assume that every point is uniform for a general hypersurface of degree $d-1$. Let $P \in \mathbb{P}^{n+1}$ be a point and degenerate $X$ onto $X_0=Y \cup H$ as in Lemma \ref{induction}. Recall that the hypersurface $Y$ is general of degree $d-1$. \medskip Assume that $P \in X$, hence $P \in X_0$ by degeneration. If $P \in Y$, by induction $M((\pi_0)_{|Y})=S_{d-2}$. Moreover, $S_{d-2} \subseteq M(\pi_P)$ by Lemma \ref{induction}. Therefore, $M(\pi_P)$ is a transitive group acting on a general fibre of $\pi_P$ and, by construction, it contains a subgroup that is $d-2$ transitive on $d-2$ points of the fibre. Therefore $M(\pi_P)$ is $d-1$ transitive by Lemma \ref{lemma1}, and so $P$ is uniform. If $P \in H$, then $M((\pi_0)_{|Y})=S_{d-1}$. Therefore, applying Lemma \ref{induction} we conclude that $P$ is uniform for $X$. \medskip Assume now that $P \notin X$. We recall that, in the degeneration, the point $P$ may be in $X_0$. If $P \notin Y$, then $M((\pi_0)_{|Y})=S_{d-1}$. Moreover, by Lemma \ref{induction}, it is contained in $M(\pi_P)$. Hence it is a group acting transitively on a general fibre of $\pi_P$ and that contains a subgroup that is $d-1$ transitive on $d-1$ points of the fibre. Therefore, by Lemma \ref{lemma1}, $M(\pi_P)$ is $d$ transitive, i.e. the point $P$ is uniform. If $P \in Y $, then $S_{d-2}= M((\pi_0)_{|Y})$. By Lemma \ref{induction} we have that $M(\pi_P)$ contains a subgroup that is $d-2$ transitive on $d-2$ points of a general fibre. If moreover $M(\pi_P)$ is primitive, then by Lemma \ref{lemma2} we have that it is $2$-transitive. If we apply again Lemma \ref{lemma1} we get that it is $d$-transitive on a general fibre, i.e. $P$ is uniform. We are then left to prove that the action of $M(\pi_P)$ is primitive. If $d \geq 5$ the action of $M(\pi_P)$ is clearly primitive since $d-2$ does not divide $d$ (see Lemma \ref{block}). If $d=4$, assume by contradiction that the map $\pi_P$ is decomposable. The only possibility is that it factors via two maps of degree two $X \stackrel{2:1}{\to} Y \stackrel{2:1}{\to} \mathbb{P}^n.$ The first map can be seen as an involution of the general quartic, hence a non trivial automorphism of $X$. This contradicts Theorem \ref{autom}. Therefore, every point is uniform. \end{proof} \section*{Acknowledgements} The author is supported by MIUR: Dipartimenti di Eccellenza Program (2018-2022) - Dept. of Math. Univ. of Pavia and by PRIN 2017 "Moduli spaces and Lie Theory" code 2017YRA3LK\_003. I would like to thank Gian Pietro Pirola for introducing me to the problem and for all the help he gave during the preparation of this paper. I also thank Ciro Ciliberto, Riccardo Moschetti, Lidia Stoppino and Thomas Dedieu for helpful discussions and suggestions. \bibliographystyle{alpha}
{ "timestamp": "2020-07-21T02:29:47", "yymm": "2007", "arxiv_id": "2007.09958", "language": "en", "url": "https://arxiv.org/abs/2007.09958", "abstract": "Let $X$ be a general complex projective hypersurface in $\\mathbb{P}^{n+1}$ of degree $d>1$. A point $P$ not in $X$ is called uniform if the monodromy group of the projection of $X$ from $P$ is isomorphic to the symmetric group. We prove that all the points in $\\mathbb{P}^{n+1}$ are uniform for $X$, generalizing a result of Cukierman on general plane curves.", "subjects": "Algebraic Geometry (math.AG)", "title": "Monodromy of general hypersurfaces", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9825575152637946, "lm_q2_score": 0.8333245994514082, "lm_q1q2_score": 0.8187893478451727 }
https://arxiv.org/abs/1604.05507
The young centre of the Earth
We treat, as an illustrative example of gravitational time dilation in relativity, the observation that the center of the Earth is younger than the surface by an appreciable amount. Richard Feynman first made this insightful point and presented an estimate of the size of the effect in a talk; a transcription was later published in which the time difference is quoted as 'one or two days'. However, a back-of-the-envelope calculation shows that the result is in fact a few years. In this paper we present this estimate alongside a more elaborate analysis yielding a difference of two and a half years. The aim is to provide a fairly complete solution to the relativity of the 'aging' of an object due to differences in the gravitational potential. This solution - accessible at the undergraduate level - can be used for educational purposes, as an example in the classroom. Finally, we also briefly discuss why exchanging 'years' for 'days' - which in retrospect is a quite simple, but significant, mistake - has been repeated seemingly uncritically, albeit in a few cases only. The pedagogical value of this discussion is to show students that any number or observation, no matter who brought it forward, must be critically examined.
\section{Introduction} The gravitational potential influences the rate at which time passes. This means that a hypothetical measurement of the age of a massive object like the Sun or the Earth would yield different results depending on whether performed at the surface or near the center. In this connection, clearly, issues such as the initial assembly of cosmic dust to form the protoplanet eventually leading to the Earth is not what is alluded to when considering the age. Rather, the age is understood as e.g.\ the 'aging' of radioactive elements in the Earth, i.e.\ that fewer radioactive decays of a particular specimen have taken place in the Earth center than on its surface. Furthermore, arguments based on symmetry will convince most skeptics, including those from 'the general public', that there is no gravitational force at the Earth center. Consequently, such an effect cannot be due to the force itself, but may instead be due to the 'accumulated action of gravity' (a layman expression for the gravitational potential energy being the radial integral of the force). Thus, there is also a good deal of pedagogical value in this observation. In a series of lectures presented at Caltech in 1962-63, Feynman is reported to have shared this fascinating insight with the audience using the formulation "...since the center of the earth should be a day or two younger than the surface!"~\cite{Hatf03}. This thought experiment is just one among a plethora of fascinating observations about the physical world provided by Richard Feynman. Although this time difference has been quoted in a few papers, either the lecturer or the transcribers had it wrong; it should have been given as 'years' instead of 'days'. In this paper, we first present a simple back-of-the-envelope calculation which compares to what may have been given in the lecture series. We then present a more elaborate analysis which brings along a number of instructive points. We believe that this correction only makes the observation of age difference due to gravity even more intriguing. We stress that this paper is by no means an attempt at besmearing the reputation of neither Feynman nor any of the authors who trustingly replicated his statement (including one of the authors of the present paper, UIU). Instead the, admittedly small, mistake is used as a pedagogical point much like the example 'the human failings of genius' that Ohanian has used in his book about Einstein's mistakes \cite{Ohan08}. Realising that even geniuses make mistakes may make the scientist more inclined towards critically examining any postulate on his/her own. \section{The center of the Earth is younger than its surface} \subsection{Homogeneous Earth} We initially suppose that the object under consideration is a sphere with radius $R$ and mass $M$, homogeneously distributed. Its gravitational potential as a function of distance $r$ to its center is then given by \begin{eqnarray} \Phi&=&-G\cfrac{M}{r}~~~~~~~~~~~~~~~~~~~r\geq R\\ \Phi&=&-G\cfrac{M(3R^2-r^2)}{2R^3}~~~~~~r\leq R \label{earth_time1} \end{eqnarray} such that the potential on its surface is $\Phi(R)=-GM/R$ and the potential in its center is $\Phi(0)=-3GM/2R$. The difference between the gravitational potential at the center and at the surface is then \begin{equation} \Delta\Phi=\Phi(R)-\Phi(0)=\cfrac{1}{2}G\cfrac{M}{R}. \label{earth_time4} \end{equation} A difference in gravitational potential implies a time dilation at the point with the lower potential. This is given by the standard 'gravitational redshift' \begin{equation} \omega=\omega_0(1-\cfrac{\Delta\Phi}{c^2}), \label{work_height6} \end{equation} which here relates the (angular) frequencies at the center, $\omega$, and at the surface $\omega_0$. Being the inverse of the period, the frequency is indirectly a measure of how quickly time passes. It is customary to use the symbol $\omega$ in this connection, and we emphasise that this variable has nothing to do with the Earth rotation. We combine equation \eqref{work_height6} with the result for $\Delta\Phi$ in equation \eqref{earth_time4} and use that $\Delta\omega=\omega-\omega_0$, \begin{equation} \Delta\omega=-\cfrac{1}{2}G\omega_0\cfrac{M}{Rc^2} \label{earth_time5} \end{equation} We note that this treatment is based on equation \eqref{work_height6} which "...refers only to identically constructed clocks located at different distances from the center of mass of a gravitating body along the lines of force. All that is required is that the clocks obey the weak equivalence principle [...] and the special theory of relativity." \cite{Nobi13}. See Ref.~\cite{RSChristensen} for a recent, instructive example that can be easily performed in the undergraduate laboratory to display one aspect of the equivalence principle. For the case of the Earth, upon rewriting and setting the surface acceleration $GM_e/R_e^2=g=9.82$ m/s$^2$, with $R_e$ being the Earth radius, equation \eqref{earth_time5} becomes \begin{equation} \cfrac{\Delta\omega}{\omega_0}=-\cfrac{1}{2}\cfrac{R_eg}{c^2}, \label{earth_time6} \end{equation} such that the Earth mass $M_e$ and the gravitational constant $G$ are not needed explicitly. For the sake of a back-of-the-envelope calculation we may exploit that $c/g\simeq1$ year (within 3\%, although there is no direct connection between the motion of the Earth around the Sun and $c$). The Earth age is $T_e=4.54\cdot10^9$ years and its average radius is $R_e=$ 6371 km so that $R_e/2c$ is approximately 10 ms. A year is approximately $\pi\cdot10^7$ s. Clearly, the use here of $\pi$ is a mnemonic device, not an expression of precision, although it is precise to about half a percent (one could use $3$ instead of $\pi$ which, however, is imprecise to 5 percent). Thus the difference between the age of the Earth surface and its center becomes approximately $4.5\cdot10^9\cdot10^{-2}/\pi\cdot10^7\simeq1.4$ years, with the center being youngest. This is the \emph{type} of 'back-of-the-envelope' calculation that one could imagine that Feynman had in mind when he expressed his "...since the center of the earth should be a day [which thus should have read 'year'] or two younger than the surface!" \cite{Hatf03}. Where the mistake actually entered in the lecture and transcription process is unlikely to ever be ascertained, and its exact origin is not important for the following discussion. With tabulated values for $M_e,~G,~R_e,~c$ and $T_e$ a more precise number for the homogeneous Earth is obtained: \begin{equation} \Delta T_{eh}=T_e\cfrac{1}{2}G\cfrac{M_e}{R_ec^2}=1.58~\mathrm{years}, \label{earth_time7} \end{equation} with the center being youngest. \subsection{Realistic Earth} Rather than assuming a homogeneous Earth, we now turn to a more realistic density distribution. This yields a significantly different result and reveals some insights to the origin of the time difference. A rather precise description, but not the only one available, of the Earth density profile is tabulated in the so-called 'Preliminary Reference Earth Model' (PREM)~\cite{Dzie81}. Very recently, the PREM has been applied to give a detailed description of the Earth 'gravity tunnel' problem \cite{Klot15}. We shall consider a spherically symmetric Earth with a density only dependent on radius, $\rho (r)$, as given by the PREM, see Figure \ref{fig:Density}. \begin{figure}[hbt] \begin{center} \includegraphics[width=0.5\textwidth]{Density.pdf} \caption{The density of the Earth as a function of distance to the Earth center for two different models. The blue line shows the PREM of the Earth density and the red curve is the constant density in the approximation of a homogeneous Earth with mass $M_e$.} \label{fig:Density} \end{center} \end{figure} The gravitational potential caused by this sphere is then given by \begin{align} \Phi(r) = -\int_{\infty}^{r}\vec{f}_{\textrm{grav}}\cdot \vec{dr'}, \end{align} where $\vec f_{\textrm{grav}}=\vec F_{\textrm{grav}}/m$ is the mass specific force, or acceleration $\vec a_{\textrm{grav}}$, due to gravity, with $\vec F_{\textrm{grav}}$ being the gravitational force (which is why $\Phi$ is the gravitational potential and not the gravitational potential \emph{energy}). The gravitational potential energy is equal to the work done by taking a test particle of mass $m$ from infinity to a distance $r$ away from the center of the Earth. We split the expression in two parts: \begin{align} \Phi(r) =& -\int_{R}^{r}\vec{f}_{\textrm{grav}}\cdot \vec{dr'} - \int_{\infty}^{R}\vec{f}_{\textrm{grav}}\cdot \vec{dr'}, \end{align} with the first term being the work per unit mass done inside the object - in this case the Earth - and the last term the work per unit mass done moving the test particle from infinity to the Earth surface. The gravitational acceleration at a distance, $r$, outside a sphere of mass $M$ is $\vec a_{\textrm{grav}}=\vec f_{\textrm{grav}}=-\hat{r}GM/r^{2}$, the sign showing that it is directed towards the center. Inside the sphere, when $r < R$, only the mass closer than $r$ to the center matters. We denote this by \begin{align} M(r) = \int_0^r 4\pi r'^2 \rho(r')dr'. \label{M_r} \end{align} Now we can write the sum specifically for $r < R$ as \begin{align} \Phi(r) =& \int_{R}^{r} G\frac{M(r')}{r'^2} dr' -G\cfrac{M}{R} \label{Phi_r} \end{align} where the last term is the potential at the surface of the object. The integrand in the first term is the gravitational acceleration as a function of $r$. When evaluated at the surface, $r=R_e$, the result is the normal gravitational acceleration, $g$. This can be seen on Figure \ref{fig:Acceleration} where the acceleration felt at different distances to the Earth center is shown. Due to the mass distribution 'kink' seen in Figure \ref{fig:Density} at a radius of about 3500 km, the acceleration becomes almost constantly equal to its surface value from this radius, outwards. \begin{figure}[hbt] \begin{center} \includegraphics[width=0.5\textwidth]{Acceleration.pdf} \caption{The size of the gravitational acceleration as a function of distance to the Earth center. It reaches the familiar value of $9.82$ m/s$^2$ at the surface. The analytical curve is given by the simple scaling $g\cdot r/R_e$ by assuming a homogeneous mass distribution.} \label{fig:Acceleration} \end{center} \end{figure} Using the PREM density distribution $\rho(r')$ in eq.\ \eqref{M_r} as an input to eq.\ \eqref{Phi_r}, the more elaborate result for the age difference of the Earth center and the surface is \begin{equation} \Delta T_e=2.49~\mathrm{years}, \label{earth_time7b} \end{equation} with the center being youngest. As a, perhaps, intriguing side-effect, we show the time difference as a function of radius, see Figure \ref{fig:TvariationE}. As expected, the two theories predict similar time differences near the surface of the Earth. Closer to the center, the PREM yields a larger result than the homogeneous distribution. This is because $M(r)_{\mathrm{PREM}}>M(r)_{\mathrm{Homog.}}$ for small $r$. In fact, assuming for simplicity that the object of radius $R$ consists of a region of high density for $0\leq r\leq r_0$ and zero density for $r_0< r\leq R$, respecting that the total mass equals $M$, and with $r_0=k_0R$, $0<k_0\leq1$, the potential difference between center and surface becomes \begin{equation} \Delta\Phi=(\cfrac{3}{k_0}-2)\Delta\Phi_h, \label{DPhi_hom} \end{equation} where $\Delta\Phi_h$ is that of the homogeneous distribution. Thus, the factor $3/k_0-2$ yields the increase in time difference compared to the homogeneous model. So for the Earth, where this approximation is rather crude, we may set $r_0$ to be 3480 km as seen from the PREM curve in Figure \ref{fig:Density}, i.e.\ $k_0\simeq3480~\mathrm{km}/6371~\mathrm{km}\simeq0.55$ such that $\Delta\Phi\simeq3.5\Delta\Phi_h$ somewhat above the factor 1.7 obtained by the numerical method, as expected from the crudeness of this approximation. \begin{figure}[hbt] \begin{center} \includegraphics[width=0.5\textwidth]{TvariationEarth2x.pdf} \caption{The figure shows the time difference between a point inside the Earth and the surface (center) of the Earth. The blue curve is calculated with PREM and the red curve using a homogeneous mass distribution.} \label{fig:TvariationE} \end{center} \end{figure} We end this section by showing that the time dilation due to the rotational speed of the surface of the Earth makes a negligible contribution. The surface speed is given from the period of rotation as $v_s=R_e 2\pi/T_s\simeq464$ m$/$s where $T_s=86.164,099$ s is the stellar day (the Earth rotation period with respect to the 'fixed' stars). Since the time dilation in special relativity is given from the Lorentz factor $\gamma=1/\sqrt{1-v^2/c^2}\simeq1+v^2/2c^2$ as $\Delta T=Tv^2/2c^2$ we get with $T_e$ that $\Delta T_s\simeq5\cdot10^{-3}$ years, which can be neglected in the present discussion. \section{The case of the Sun} Clearly, the calculations performed in connection with the Earth can be performed for essentially any other cosmic object with known mass and radius, at least in the limit of a homogeneous mass distribution. However, we limit the additional cases to our cosmic neighbourhood, i.e.\ to that of the Sun, in order to demonstrate the applicability of eq.\ \eqref{DPhi_hom}. For the Sun, in analogy with the PREM which is based on seismic data, we choose the so-called 'Model S' for its density distribution, a model in good agreement with helioseismic data \cite{Chri96}. \begin{figure}[hbt] \begin{center} \includegraphics[width=0.5\textwidth]{SunDensity.pdf} \caption{The density of the Sun as a function of distance to the Sun center for two different models. The blue line shows the 'Model S' of the Sun density, obtained from helioseismic data \cite{Chri96}, and the red curve is the uniform density Sun.} \label{fig:DensitySun} \end{center} \end{figure} In the homogeneous case, the age difference between the Sun center and surface, which can be rewritten as $\Delta T=T_s v_{\mathrm{esc}}^2/4c^2$ with $v_{\mathrm{esc}}=\sqrt{2GM/R}$ being the surface escape velocity, is \begin{equation} \Delta T_{sh}=4.8\cdot10^3~\mathrm{years}, \label{sun_time1} \end{equation} whereas with the 'Model S' solar model it becomes \begin{equation} \Delta T_{s}=3.9\cdot10^4~\mathrm{years}, \label{sun_time2} \end{equation} see Figure \ref{fig:TvariationS}. The factor of $8.0$ difference between these two numbers is substantially larger than that between the same two numbers for the Earth. This is a result of the Earth being relatively homogeneous while for the Sun, a significantly larger part of its mass is located close to its center. Using eq.\ \eqref{DPhi_hom} and approximating $k_0\simeq2.5/7$ from the density distribution, the Model S curve in Figure \ref{fig:DensitySun}, we get a factor $8.4$, a much better approximation than for the case of the Earth. \begin{figure}[hbt] \begin{center} \includegraphics[width=0.5\textwidth]{TvariationSun2x.pdf} \caption{The figure shows the time difference between a point inside the Sun and the surface (center) of the Sun. The blue curve is calculated with the 'Model S' density model for the Sun and the red curve using a homogeneous mass distribution.} \label{fig:TvariationS} \end{center} \end{figure} \section{Discussion and conclusion} As a final discussion we address the question: why did famous, respectable and clever physicists publish Feynman's claim (although not verbatim, actually) that "[Feynman] concluded that the center of the Earth should be 'a day or two younger than its surface'" \cite{Okun99}, or "[Feynman] concludes that the center of the Earth should be by a day or two younger than its surface" \cite{Okun99b} and reversely "'atoms at the surface of the earth are a couple of days older than at its center'", in the latter case even with the comment "this was confirmed by airplane experiments in 1970s" \cite{Okun02}? And why did other, equally talented physicists not correct \emph{that} particular mistake in the foreword to the transcribed lectures, in spite of quite extensive discussions, spanning 24 pages, of among other things a few misconceptions etc. \cite{Pres95}? Not to mention the transcribers - postdocs with Feynman - who, along the way, probably have corrected a few mistakes here and there? Or the editor, who also provided introductory notes on quantum gravity \cite{Hatf03b}? Why did one of us (UIU), repeat the same mistake in a science book for the layman \cite{Ugge14}? This, of course, was not because any of these physicists were unable to check the original claim, or found it particularly laborious to do so. Instead, it seems likely that they knew that the qualitative effect had to be there, and simply trusted that Feynman and his transcribers had got the number right. This is here considered an example of 'proof by ethos' \cite{Faye14}. The term 'proof by ethos' refers to cases where a scientist's status in the community is so high that everybody else takes this person's calculations or results for granted. In other words, nobody questions the validity of that scientist's claim because of the particular ethos that is associated with that person. The result is accepted merely by trust. Indeed, the proof by ethos is not really a proof as it does not follow logically from a set of premises. But it is a proof in the sense that it is persuasive, and tells us something about how scientists work in practice when they accept a calculation or an experimental result. Scientists must to a large extent rely on the validation of other fellow's work, and it happens to be a psychological default condition among many (scientists), that if a famous peer has publicly announced a result, it is accepted at face value. This seems also to be the situation in the case of the flawed estimate of the relativistic age of the Earth's core. In science, one route to becoming famous is being right on some important topics. However, just because someone has become famous, this person is evidently not necessarily right on all matters. Feynman himself would most likely have agreed with this and he would probably not have fallen for his own miscalculation: For a long time, his own theory of beta decay was at odds with the then prevalent, but false, understanding of existing experimental results. Upon finally realizing and correcting this community-wide misunderstanding Feynman wrote: "Since then I never pay any attention to "experts". I calculate everything myself."~\cite{Surely}. And when faced with a mistake of his own, he put it even more bluntly: "What it says in the book [I have written] is absolutely wrong!" \cite{Hey99}.\\ In spite of the small numerical mistake, Feynman's observation that the center of the Earth is younger than its surface is a fascinating demonstration of time dilation in relativity, and as such a very illustrative example for use in the classroom.
{ "timestamp": "2016-04-20T02:08:32", "yymm": "1604", "arxiv_id": "1604.05507", "language": "en", "url": "https://arxiv.org/abs/1604.05507", "abstract": "We treat, as an illustrative example of gravitational time dilation in relativity, the observation that the center of the Earth is younger than the surface by an appreciable amount. Richard Feynman first made this insightful point and presented an estimate of the size of the effect in a talk; a transcription was later published in which the time difference is quoted as 'one or two days'. However, a back-of-the-envelope calculation shows that the result is in fact a few years. In this paper we present this estimate alongside a more elaborate analysis yielding a difference of two and a half years. The aim is to provide a fairly complete solution to the relativity of the 'aging' of an object due to differences in the gravitational potential. This solution - accessible at the undergraduate level - can be used for educational purposes, as an example in the classroom. Finally, we also briefly discuss why exchanging 'years' for 'days' - which in retrospect is a quite simple, but significant, mistake - has been repeated seemingly uncritically, albeit in a few cases only. The pedagogical value of this discussion is to show students that any number or observation, no matter who brought it forward, must be critically examined.", "subjects": "Physics Education (physics.ed-ph); Popular Physics (physics.pop-ph)", "title": "The young centre of the Earth", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9702399069145609, "lm_q2_score": 0.8438951005915208, "lm_q1q2_score": 0.8187807038435712 }
https://arxiv.org/abs/1812.03561
An Inverse Function Theorem Converse
We establish the following converse of the well-known inverse function theorem. Let $g:U\to V$ and $f:V\to U$ be inverse homeomorphisms between open subsets of Banach spaces. If $g$ is differentiable of class $C^p$ and $f$ if locally Lipschitz, then the Fréchet derivative of $g$ at each point of $U$ is invertible and $f$ must be differentiable of class $C^p$.
\section{Introduction} A general form of the well-known inverse function theorem asserts that if $g$ is a differentiable function of class $C^p$, $p\geq 1$, between two open subsets of Banach spaces and if the Fr\'echet derivative of $g$ at some point $x$ is invertible, then locally around $x$, there exists a differentiable inverse map $f$ of $g$ that is also of class $C^p$. But in various settings, one may have the inverse function $f$ readily at hand and want to know about the invertibility of the Fr\'echet derivative of $g$ at $x$ and whether $f$ is of class $C^p$. Our purpose in this paper is to present a relatively elementary proof of this converse result under the general hypothesis that the inverse $f$ is (locally) Lipschitz. Simple examples like $g(x)=x^3$ at $x=0$ on the real line show that the assumption of continuity alone is not enough. Thus it is a bit surprising that the mild strengthening to the assumption that the inverse is locally Lipschitz suffices. Helpful tools for the task at hand have been developed in the intense study of Lipschitz functions in the Banach space setting motivated by Rademacher's theorem concerning the existence of an abundance of points of differentiability of Lipschitz mappings in the setting of euclidean spaces. In particular we recall in the next section a useful generalization of the chain rule by O.\ Maleva and D.\ Preiss \cite{MP}. In Sections 3 and 4 we present our main results and provide some follow-up illustrative material in Section 5. For a comprehensive reference to the inverse function and implicit function theorems and related theory, we refer the reader to \cite{KP}. \section{A general version of the chain rule} In this section we recall some notions of differentiability of Lipschitz functions between open subsets of a Banach spaces and a generalized chain rule from the work of Maleva and Preiss \cite{MP}. This material will be crucial to the derivation of our main results in the next section. Suppose that $Y$ and $Z$ are Banach spaces, $U$ is a nonempty open subset of $Y$, $f : U \to Z$, and $y\in U$, $v \in Y $. Recall that $\lim_{t\to 0^+} (f(y+tv)-f(y))/t$, if it exists, is called the \emph{one-sided directional derivative of $f$ at $y$ in the direction} $v$. Similarly if $\lim_{t\to 0} (f(y+tv)-f(y))/t$ exists, it is called the (\emph{bilateral}) \emph{directional derivative of $f$ at $y$ in the direction} $v$. If the directional derivative of $f$ at $y$ in the direction $v$ exists for all $v\in Y$, then the mapping from $Y$ to $Z$ sending $v$ to its directional derivative is, by definition, the \emph{G\^{a}teaux derivative} (by some authors the mapping is also required to be a continuous linear map). Maleva and Preiss \cite{MP} have given the following generalization of the one-sided directional derivative. \begin{definition} The \emph{derived set of $f$ at the point $y$ in the direction of $v$} is defined as the set $\mathcal D f(y, v)$ consisting of all existing limits $\mathrm{lim}_{n\to\infty}(f(y + t_nv)-f(y))/t_n$, where $t_n\searrow 0$. The \emph{$\delta$-approximating derived set of $f$ at $y$ in the direction of $v$} is defined, for $\delta>0$, by \begin{equation*} \mathcal D_\delta f(y,v)=\Big\{ \frac{f(y + tv)-f(y)}{t}: 0<t <\delta\Big\}. \end{equation*} \end{definition} \begin{remark} It is easy to see that \begin{equation}\label{E:derived} \mathcal D f(y,v)=\bigcap_{\delta>0} \overline{\mathcal D_\delta f(y,v)} \end{equation} In general the derived set may be empty, a single point, or multi-valued. If $\mathcal D f(y,v)$ is a single point, then it is the one-sided directional derivative of $f$ at $y$ in the direction $v$, and we denote it by $f_+'(y,v)$. If the directional derivative of $f$ at $y$ in the direction $v$ exists, it is denoted $f'(y,v)$. If $f$ is G\^{a}teaux differentiable at $y$, then the G\^{a}teaux derivative at $y$ is equal to $f'(y,\cdot)=\mathcal D f(y,\cdot)$. \end{remark} We recall a general version of the chain rule from \cite[Corollary 2.6]{MP}. \begin{proposition}\label{P:chain} Suppose $X$, $Y$ are Banach spaces, $x\in U$, an open subset of $X$, and $V$ is an open subset of $Y$. If $g :U\to V$ is continuous and has a one-sided directional derivative at $x$ in the direction of $v$ and $f :V\to Z$ is Lipschitz, then $$\mathcal D(f \circ g)(x, v) = \mathcal D f(g(x), g_+'(x,v)).$$ \end{proposition} \begin{proof} Let $a\in \mathcal D(f \circ g)(x, v)$. Then there exists a sequence $t_n\to 0^+$ such that $$a=\lim_n \frac{f\circ g(x+t_nv)-f\circ g(x)}{t_n}.$$ Let $\kappa$ be a Lipschitz constant for $f$ on $V$. Let $\varepsilon>0$ and choose $N$ such that $\Vert(g(x+t_nv)-g(x))/t_n-g_+'(x,v)\Vert<\varepsilon/\kappa$ for $n\geq N$. We then note for $n\geq N$ \begin{eqnarray*} \bigg\Vert \frac{f\circ g(x+t_nv)-f\circ g(x)}{t_n}\!\!&-& \!\!\frac{f(g(x) +t_ng_+'(x,v))-f(g(x))}{t_n}\bigg\Vert \\ &=& \bigg\Vert \frac{f(g(x+t_nv))-f(g(x) +t_ng_+'(x,v))}{t_n}\bigg\Vert \\ &\leq& \frac{\kappa}{t_n}\Vert g(x+t_nv)-g(x) -t_ng_+'(x,v)\Vert\\ &=&\kappa\bigg\Vert \frac{g(x+t_nv)-g(x)}{t_n} -g_+'(x,v)\bigg\Vert\leq \kappa(\varepsilon/\kappa)=\varepsilon. \end{eqnarray*} It thus follows that the sequence $[f(g(x) +t_ng_+'(x,v))-f(g(x))]/t_n$ also converges to $a$, so $a\in \mathcal D f(g(x),g_+'(x,v))$. The argument is reversible so the equality claimed in the proposition holds. \end{proof} \section{A Converse of the Inverse Function Theorem} The inverse function theorem asserts the existence of a local inverse if the derivative is invertible. In this section we derive a converse result: a Lipschitz continuous local inverse implies an invertible derivative. \emph{In this section we work in the following setting. Let $X,Y$ be Banach spaces, let $U$ and $V$ be nonempty open subsets of $X$ and $Y$ resp., each equipped with the restricted metric from the containing Banach space, and let $g:U\to V$ and $f:V\to U$ be inverse homeomorphisms.} \begin{definition} For $g:U\to V$, the \emph{G\^ateaux derivative of $g$ at} $x\in U$ is defined by $d_x^Gg(v)=g_+'(x,v)$ for all $v\in X$, provided such one-sided directional derivatives exist for all $v\in X$ and the resulting map $d_x^Gg:E\to F$ is a continuous linear map. \end{definition} \begin{lemma}\label{L:inj} Let $x\in U$, $y=f(x)\in V$. Assume that $g$ has a G\^ateaux derivative $d_x^Gg: E\to F$ at $x$ and assume that $f$ is Lipschitz. Then $\mathcal D f(g(x),\cdot)\circ d_x^Gg$ is the identity map on $X$ and the G\^ateaux derivative $d_x^Gg$ is injective. \end{lemma} \begin{proof} Since $f\circ g$ on $U$ is the identity map, it follows directly from the definition of $\mathcal D$ that $\mathcal D(f\circ g)(x,\cdot)$ is the identity map on $X$. Hence by Proposition \ref{P:chain} for any $v\in X$, $$v=\mathcal D(f\circ g)(x,v)=\mathcal D f(g(x), g_+'(x,v))=\mathcal D f(g(x),d_x^G(v)).$$ This equality of the left-hand and right-hand sides of the equation yields the two concluding assertions. \end{proof} Lemma \ref{L:inj} yields the invertibility of the the G\^ateaux derivative in the finite dimensional setting. \begin{corollary} If $X,Y$ are both finite dimensional and $g$ is G\^ateaux differentiable at $x\in U$, then $d_x^G$ is invertible. \end{corollary} \begin{proof} Since $g$ and $f$ are homeomorphisms, $X$ and $Y$ must have the same dimension. Hence the linear map $d_x^G$ is a linear isomorphism if and only if it injective, which is the case by Lemma \ref{L:inj}. \end{proof} The infinite dimensional case requires more work and stronger hypotheses. \begin{lemma}\label{L:surj} Let $x\in U$, $y=g(x)\in V$. Assume that $g$ is Fr\'echet differentiable at $x$ and that $f$ is Lipschitz on $V$. Then the image of $X$ under the Fr\'echet derivative $dg_x:X\to Y$ is dense in $Y$. \end{lemma} \begin{proof} Let $M_f$ be a Lipschitz constant for $f$ on $V$. Let $w\in Y$. Pick $\tau>0$ small enough so that $y+tw\in V$ for $0<t<\tau$. We set $z_t=\frac{f(y+tw)-f(y)}{t} $ and note from the Lipschitz condition that \begin{equation}\label{E:lips} \Vert z_t\Vert=\Big\Vert \frac{f(y+tw)-f(y)}{t}\Big\Vert\leq \frac{1}{t} M_f\Vert (y+tw)-y\Vert=M_f\Vert w\Vert. \end{equation} We divide the remainder of the proof into steps.\\ \emph{Step 1}: $(g(x+tz_t)-g(x))/t=w$ for $0<t<\tau$. $$g(x+tz_t)-g(x)=g\bigg(f(y)+t\Big(\frac{f(y+tw)-f(y)}{t}\Big)\bigg)-y=gf(y+tw)-y=tw,$$ so $(g(x+tz_t)-g(x))/t=w$. \\ \emph{Step 2}: \emph{For $\varepsilon>0$ there exists $t<\tau$ such that $\Vert (g(x+tz_t)-g(x))/t-dg_x(z_t)\Vert<\varepsilon$.} From the Fr\'echet differentiability of $g$ at $x$, the Fr\'echet derivative $dg_x$ satisfies $$\lim_{u\to 0} \frac{\Vert g(x+u)-g(x)-dg_x(u)\Vert}{\Vert u\Vert}=0.$$ For $\varepsilon >0$ pick $\delta> 0$ such that $$\frac{\Vert g(x+u)-g(x)-dg_x(u)\Vert}{\Vert u\Vert}<\frac{\varepsilon}{M_f\Vert w\Vert} \mbox{ whenever } 0<\Vert u\Vert<\delta.$$ Pick $t>0$ such that $y+tw\in V$ and $t M_f\Vert w\Vert< \delta$. We conclude from inequality (\ref{E:lips}) and the preceding that \begin{eqnarray*} \Big\Vert\frac{g(x+tz_t)-g(x))-dg_x(tz_t)}{t}\Big\Vert&=&\Vert z_t\Vert\Big\Vert\frac{g(x+tz_t)-g(x))-dg_x(tz_t)}{\Vert tz_t\Vert}\Big\Vert\\ &<& M_f\Vert w\Vert \frac{\varepsilon}{M_f\Vert w\Vert} =\varepsilon \end{eqnarray*} Using the linearity of of the Fr\'echet derivative $dg_x$, we obtain $dg_x(tz_t)=tdg_x(z_t)$, which allows us to rewrite the first entry in the preceding string to obtain $$ \Big\Vert\frac{g(x+tz_t)-g(x)}{t}-dg_x(z_t)\Big\Vert <\varepsilon,$$ which establishes the Step 2. We note that combining Steps 1 and 2 yields $$ \Vert w-dg_x(tz_t)\Vert\leq \Big\Vert w-\frac{g(x+tz_t)-g(x)}{t}\Big\Vert+ \Big\Vert \frac{g(x+tz_t)-g(x)}{t}-dg_x(z_t)\Big\Vert\leq \varepsilon.$$ Since $w\in Y$ and $\varepsilon>0$ were chosen arbitrarily, this completes the proof. \end{proof} We come now to a central result of the paper, what we are calling a converse of the inverse function theorem. \begin{theorem}\label{T:main} Let $X$ and $Y$ be Banach spaces, let $U$ and $V$ be nonempty open subsets of $X$ and $Y$ resp., and let $g:U\to V$ and $f:V\to U$ be inverse homeomorphisms. Let $x\in U$ and $y=g(x)\in V$. Let $g$ be Fr\'echet differentiable at $x$ and let $f$ be Lipschitz continuous on $V$. Then $dg_x(\cdot): X\to Y$ is an isomorphism. \end{theorem} \begin{proof} By Lemma \ref{L:inj} the Fr\'echet derivative $dg_x$ is injective, hence a linear isomorphism onto its image $Z=g(X)$, a subspace of $Y$, and has inverse $\mathcal D f(y,\cdot):Z\to X$. (In particular in this setting for $w\in Z$ it must be the case that $\mathcal D f(y,w)$ is a singleton, which we could write alternatively as $f_+'(y,w)$.) Let $M_f$ be the Lipschitz constant for $f$ on $V$. By equation (\ref{E:lips}) every member of $\mathcal D_\delta f(y,w)$ is bounded in norm by $M_f\Vert w\Vert$, and hence the same is true for $\mathcal D f(y,w)=\bigcap_{\delta>0} \overline{\mathcal D_\delta f(y,w)}$. We conclude that $\mathcal D f(y,\cdot):Z\to X$ is Lipschitz for the Lipschitz constant $M_f$. By Lemma \ref{L:surj} $Z$ is dense in $Y$. Thus the linear Lipschitz map $\mathcal D f(y,\cdot):Z\to X$ extends uniquely to a linear Lipschitz map (hence a bounded linear operator) from $Y$ to $X$. Label $dg_x=\Gamma_X$, $\mathcal D f(y,\cdot)=\Gamma_Z$ and its extension to $Y$ by $\Gamma_Y$. We note for $0\ne z\in Z$, $$\Vert z\Vert =\Vert \Gamma_X(\Gamma_Z(z)\Vert\leq \Vert \Gamma_X\Vert \Vert \Gamma _Z(z)\Vert,$$ so $(1/\Vert\Gamma_X\Vert)\Vert z\Vert \leq \Vert \Gamma _Z(z)\Vert$. By continuity of the norm and density of $Z$ in $Y$ this inequality carries over from $\Gamma_Z$ to its extension $\Gamma_Y$. But this extended inequality implies $\Gamma_Y$ has a trivial kernel, which means that $\Gamma_Y$ is injective. But if $Z$ is proper in $Y$, then this is impossible, since $\Gamma_Z(Z)=X$. Hence $Z=Y$ and $\Gamma_Z=\Gamma_Y$ is an inverse for $dg_x$. \end{proof} \section{A Global Result} We can use Theorem \ref{T:main} to derive useful results for studying differentiable functions with locally Lipschitz inverses, in particular for deriving differentiability properties of the inverse. In the following $C^p$ means have continuous derivatives through order $p$ for $p$ a positive integer, have continuous derivatives of all order for $p=\infty$, and being analytic (having locally power series expansions) for $p=\omega$. \begin{theorem}\label{T:ift} Let $X$ and $Y$ be Banach spaces, let $U$ and $V$ be nonempty open subsets of $X$ and $Y$ resp., and let $g:U\to V$ and $f:V\to U$ be inverse homeomorphisms. Assume further that $g$ is of class $C^p$ on $U$ for some $p\geq 1$ and that $f$ is locally Lipschitz on $V$. Then $g$ and $f$ are inverse diffeomorphisms of class $C^p$. \end{theorem} \begin{proof} We fix $x\in U$ and $y=f(x)\in V$. We apply Theorem \ref{T:main} to small enough neighborhoods of $x$ and $y$ so that $f$ is Lipschitz to see that the hypotheses of the standard inverse function theorem (in the Banach space setting) are satisfied. The conclusions of this theorem then follow locally from the inverse function theorem. (See, for example, \cite[Theorem 1.23]{Up} for the $C^\omega$-case.) In this way we obtain that the conclusions of the theorem hold locally for each $x\in X$ and hence hold globally. \end{proof} We remark that the preceding results can be generalized to the setting of Banach manifolds. In order to obtain the preceding results in this setting one needs Lipschitz charts (between the manifold metric and the Banach space metric) at each point and then the preceding results readily extend to this more general setting. \section{An Application} We sketch in this section one setting in which our general converse of the inverse function theorem can be fruitfully applied and briefly consider a specific example. Suppose we are given some equation of the form $F(x,y)=0$, where $x,y$ belong to some given open subset of a Banach space. In some cases it might be possible to solve the equation for $y$ in terms of $x$, i.e., $y=g(x)$, which can seen to be of class $C^p$. Let's suppose further that $g$ has an inverse given by $x=f(y)$, but no corresponding explicit description of this function. If it can be shown, however, that $f$ is (locally) Lipschitz, then we use the results of the preceding section to show that $f$ is also of class $C^p$. \begin{example} Let $\mathcal{B}(H)$ be the $C^*$-algebra of bounded linear operators on the Hilbert space $H$. We consider the Banach space $\mathbb{S}$ of hermitian operators and the open cone of positive invertible hermitian operators $\mathbb{P}$. In recent years a useful notion of a multi-variable geometric mean $\Lambda$ on $\mathbb{P}$ has arisen \cite{LL13}, \cite{LL14}, generally called the Karcher mean. One useful characterization of this $n$-variable mean $\Lambda(A_1,\ldots,A_n)$ for $A_1,\ldots,A_n\in\mathbb{P}$ is that it is the unique solution $X$ of the equation $$\log (X^{-1/2}A_1X^{-1/2})+\cdots +\log (X^{-1/2}A_nX^{-1/2})=0.$$ If we fix $A_1,\ldots, A_{n-1}$, let $Y=A_n$, and the left-hand side of the equation be $F(X,Y)$, then we are in the general setting of the previous paragraph with $U=V=\mathbb{P}$. One can rather easily solve $F(X,Y)$ for $Y$ and see it is an analytic function $g$ of $X$, but not conversely. However the inverse $f$ is given by $f(Y)=\Lambda(A_1,\ldots, A_{n-1}, Y)$. The (local) Lipschitz property is a basic property of $\Lambda$, so the earlier results yield $X$, the geometric mean, as an analytic function of $Y$, or alternatively we can say $\Lambda$ is an analytic function of each of its variables. This turns out to be an important property of $\Lambda$. This type of analysis can be applied to other important operator means. \end{example}
{ "timestamp": "2018-12-11T02:17:14", "yymm": "1812", "arxiv_id": "1812.03561", "language": "en", "url": "https://arxiv.org/abs/1812.03561", "abstract": "We establish the following converse of the well-known inverse function theorem. Let $g:U\\to V$ and $f:V\\to U$ be inverse homeomorphisms between open subsets of Banach spaces. If $g$ is differentiable of class $C^p$ and $f$ if locally Lipschitz, then the Fréchet derivative of $g$ at each point of $U$ is invertible and $f$ must be differentiable of class $C^p$.", "subjects": "Functional Analysis (math.FA)", "title": "An Inverse Function Theorem Converse", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9850429125286726, "lm_q2_score": 0.8311430499496095, "lm_q1q2_score": 0.8187115706503274 }
https://arxiv.org/abs/2103.13462
Why Do Local Methods Solve Nonconvex Problems?
Non-convex optimization is ubiquitous in modern machine learning. Researchers devise non-convex objective functions and optimize them using off-the-shelf optimizers such as stochastic gradient descent and its variants, which leverage the local geometry and update iteratively. Even though solving non-convex functions is NP-hard in the worst case, the optimization quality in practice is often not an issue -- optimizers are largely believed to find approximate global minima. Researchers hypothesize a unified explanation for this intriguing phenomenon: most of the local minima of the practically-used objectives are approximately global minima. We rigorously formalize it for concrete instances of machine learning problems.
\section{Generalized Linear Models}\label{sec:glm} \newcommand{\widehat{L}}{\widehat{L}} \newcommand{\<}{\langle} \renewcommand{\>}{\rangle} We consider the problem of learning a \emph{generalized linear model} and we will show that the loss function for it will be non-convex, but all of its local minima are global. Suppose we observe $n$ data points $\{(x_i,y_i)\}_{i=1}^n$, where $x_i$'s are sampled i.i.d. from some distribution $D_x$ over $\mathbb R^d$. In the generalized linear model, we assume the label $y_i\in \mathbb R$ is generated from \begin{equation*} y_i = \sigma(w_\star^\top x_i) + \epsilon_i, \label{eqn:generative} \end{equation*} where $\sigma:\mathbb R\to\mathbb R$ is a known monotone activation function, $\epsilon_i\in\mathbb R$ are i.i.d. mean-zero noise (independent with $x_i$), and $w_\star\in\mathbb R^d$ is a fixed unknown ground truth coefficient vector. We denote the joint distribution of $(x_i,y_i)$ by $D$. Our goal is to recover approximately $w_{\star}$ from the data. We minimize the empirical squared risk: $ \widehat{L}(w) = \frac{1}{2n}\sum_{i=1}^n (y_i - \sigma(w^\top x_i))^2\,. $ Let $L(w)$ be the corresponding population risk: $L(w) = \frac{1}{2}\mathop{\mathbb E}\displaylimits_{(x,y)\sim D}\left[(y - \sigma(w^\top x))^2\right].$ We will analyze the optimization of $\widehat{L}$ via characterizing the property of its landscape. Our road map consists of two parts: a) all the local minima of the population risk are global minima; b) the empirical risk $\widehat{L}$ has the same property. When $\sigma$ is the identity function, that is, $\sigma(t) =t$, we have the linear regression problem and the loss function is convex. In practice, people have taken $\sigma$, e.g., to be the sigmoid function and then the objective $\widehat{L}$ is no longer convex. Throughout the rest of the section, we make the following regularity assumptions on the problem. These assumptions are stronger than what's necessary, for the ease of exposition. However, we note that some assumptions on the data are necessary because in the worst-case, the problem is intractable. (E.g., the generative assumption~\eqref{eqn:generative} on $y_i$'s is a key one.) \begin{assumption}\label{ass:1} We assume the distribution $D_x$ and activation $\sigma$ satisfy that \begin{itemize} \item[1.] The vectors $x_i$ are bounded and non-degenerate: $D_x$ is supported in $\{x:\|x\|_2 \le B\}$, and $\mathbb{E}_{x\sim D_x}[xx^\top] \succeq \lambda I$ for some $\lambda>0$, where $I$ is the identity. \item[2.] The ground truth coefficient vector satisfies $\|w_\star\|_2 \le R$, and $BR\ge 1$. \item[3.] The activation function $\sigma$ is strictly increasing and twice differentiable. Furthermore, it satisfies the bounds \begin{equation*} \sigma(t)\in[0,1],~~~\sup_{t\in\mathbb R}\left\{|\sigma'(t)|, |\sigma''(t)| \right\}\le 1,~~~{\rm and}~~~\inf_{t\in[-BR, BR]}\sigma'(t) \ge \gamma > 0. \end{equation*} \item[4.] The noise $\epsilon_i$'s are mean zero and bounded: with probability 1, we have $|\epsilon_i|\le 1$. \end{itemize} \end{assumption} \subsection{Analysis of the population risk} \newcommand{\newpart}[1]{{#1}} \newcommand{\brief}[1]{{#1}} \newcommand{\solution}[1]{{\color{red}#1}} In this section, we show that all the local minima of the population risk $L(w)$ are global minima. In fact, $L(w)$ has a unique local minimum which is also global. (But still, $L(w)$ may likely be not convex for many choices of $\sigma$.) \begin{theorem}The objective $L(\cdot)$ has a unique local minimum, which is equal to $w_{\star}$ and is also a global minimum. In particular, $L(\cdot)$ is weakly-quasi-convex. \end{theorem} The proof follows from directly checking the definition of the quasi-convexity. The intuition is that generalized linear models behave very similarly to linear models from the lens of quasi-convexity: many steps of the inequalities of the proof involves replacing $\sigma$ be an identity function effectively (or replacing $\sigma'$ be 1.) \begin{proof}[Proof Sketch] Using the property that $\mathbb{E}[y|x]=\sigma(w_\star^\top x)$, we have the following bias-variance decomposition (which can be derived by elementary manipulation) \begin{equation} L(w) = \frac{1}{2}\mathbb{E}[(y - \sigma(w^\top x))^2] = \frac{1}{2}\mathbb{E}[(y - \sigma(w_\star^\top x))^2] + \frac{1}{2}\mathbb{E}[(\sigma(w_\star^\top x) - \sigma(w^\top x))^2]\,.\label{eqn:6} \end{equation} The first term is independent of $w$, and the second term is non-negative and equals zero at $w=w_\star$. Therefore, we see that $w_\star$ is a global minimum of $L(w)$. Towards proving that $L(\cdot)$ is quasi-convex, we first compute $\nabla L(w)$: \begin{equation*} \nabla L(w) = \mathbb{E}[(\sigma(w^\top x) - y)\sigma'(w^\top x)x] = \mathbb{E}[(\sigma(w^\top x) - \sigma(w_\star^\top x))\sigma'(w^\top x)x], \end{equation*} where the last equality used the fact that $\mathbb{E}[y|x]=\sigma(w_\star^\top x)$. It follows that \begin{equation*} \<\nabla L(w), w - w_\star\> = \mathbb{E}[(\sigma(w^\top x) - \sigma(w_\star^\top x))\sigma'(w^\top x)\<w-w_\star, x\>]. \end{equation*} Now, by the mean value theorem, and bullet 3 of Assumption~\ref{ass:1}, we have that \begin{align*} (\sigma(w^\top x) - \sigma(w_\star^\top x))\<w-w_\star, x\> \ge \gamma(w^\top x - w_\star^\top x)^2. \end{align*} Using $|\sigma'(t)| \ge \gamma$ and $|\sigma'(t)| \le 1$ for every $|t|\le BR$, and the monotonicity of $\sigma$, \begin{align} \quad \<\nabla L(w), w - w_\star\> & = \mathbb{E}[(\sigma(w^\top x) - \sigma(w_\star^\top x))\sigma'(w^\top x)\<w-w_\star, x\>] \nonumber\\ & \ge \gamma\mathbb{E}(\sigma(w^\top x) - \sigma(w_\star^\top x))(w^\top x - w_\star^\top x)] \label{eqn:5}\\ & \ge \gamma \mathbb{E}[(\sigma(w^\top x) - \sigma(w_\star^\top x))^2] \nonumber\ge 2\gamma (L(w) - L(w_\star))\label{eqn:7} \end{align} where the last step uses the decomposition~\eqref{eqn:6} of the risk $L(w)$. \end{proof} \subsection{Concentration of the empirical risk} We next analyze the empirical risk $\widehat{L}(w)$. We will show that with sufficiently many examples, the empirical risk $\widehat{L}$ is close enough to the population risk $L$ so that $\widehat{L}$ also satisfies that all local minima are global. \begin{theorem}[The empirical risk has no bad local minimum] \label{thm} Under the problem assumptions, with probability at least $1-\delta$, for all $w$ with $\|w\|_2\le R$, the empirical risk has no local minima outside a small neighborhood of $w_\star$: for any $w$ such that $\|w\|_2\le R$, if $\nabla\widehat{L}(w)=0$, then \begin{equation*} \|w-w_\star\|_2 \le \frac{C_1B}{\gamma^2\lambda} \sqrt{\frac{d(C_2 + \log(nBR)) + \log\frac{1}{\delta}}{n}}. \end{equation*} where $C_1, C_2>0$ are universal constants that do not depend on $(B,R,d,n,\delta)$. \end{theorem} Theorem~\ref{thm} shows that all stationary points of $\widehat{L}(w)$ have to be within a small neighborhood of $w_\star$. Stronger landscape property can also be proved though: there is a unique local minimum in the neighborhood of $w_\star$. The main intuition is that to verify quasi-convexity or restricted secant inequality for $\widehat{L}$, it suffices to show that with high probability over the randomness of the data, $\forall w \textup{ with } \|w\|_2\le R$ \begin{align} \<\nabla L(w), w - w_\star\> \approx \<\nabla \widehat{L}(w), w - w_\star\>\,.\label{eqn:9} \end{align} Various tools to prove such concentration inequalities have been developed in statistical learning theory and probability theory community, and a thorough exposition of them is beyond the scope of this chapter. \section{Introduction} Optimizing non-convex functions has become the standard algorithmic technique in modern machine learning and artificial intelligence. It is increasingly important to understand the working of the existing heuristics for optimizing non-convex functions, so that we can design more efficient optimizers with guarantees. The worst-case intractability result says that finding a global minimizer of a non-convex optimization problem --- or even just a degree-4 polynomial --- is NP-hard. Therefore, theoretical analysis with global guarantees has to depend on the special properties of the target functions that we optimize. To characterize the properties of the real-world objective functions, researchers have hypothesized that many objective functions for machine learning problems have the property that \begin{equation}\label{eqn:11} \textup{all or most local minima are approximately global minima.} \end{equation} Optimizers based on local derivatives can solve this family of functions in polynomial time (under some additional technical assumptions that will discussed below). Empirical evidences also suggest practical objective functions from machine learning and deep learning may have such a property. In this chapter, we formally state the algorithmic result that local methods can solve objective with property~\eqref{eqn:11} in Section~\ref{sec:landscape}, and then rigorously prove that this property holds for a few objectives arising from several key machine learning problems: generalized linear models (Section~\ref{sec:glm}), principal component analysis (Section~\ref{sec:pca}), matrix completion (Section~\ref{sec:matrix}), and tensor decompositions (Section~\ref{sec:tensor}). We will also briefly touch on recent works on neural networks (Section~\ref{sec:neural_net}). \section{Analysis Technique: Characterization of the Landscape}\label{sec:landscape} In this section, we will show that a technical and stronger version of the property~\eqref{eqn:11} implies that many optimizers can converge to a global minimum of the objective function. \subsection{Convergence to a local minimum} We consider a objective function $f$, which is assumed to be twice-differentiable from $\mathbb R^d$ to $\mathbb R$. Recall that $x$ is a \textit{local minimum} of $f(\cdot)$ if there exists an open neighborhood $N$ of $x$ in which the function value is at least $f(x)$: $\forall z\in N, f(z)\ge f(x)$. A point $x$ is a \textit{stationary point} if it satisfies $\nabla f(x) = 0$. A \textit{saddle point} is a stationary point that is not a local minimum or maximum. We use $\nabla f(x)$ to denote the gradient of the function, and $\nabla^2 f(x)$ to denote the Hessian of the function ($\nabla^2 f(x)$ is an $d\times d$ matrix where $[\nabla^2 f(x)]_{i,j} = \frac{\partial^2}{\partial x_i \partial x_j} f(x)$). A local minimum $x$ must satisfy the first order necessary condition for optimality, that is, $\nabla f(x) = 0$, and the second order necessary condition for optimality, that is, $\nabla^2 f(x) \succeq 0$. (Here $A\succeq 0$ denotes that $A$ is a positive semi-definite matrix.) Thus, A local minimum is a stationary point, so is a global minimum. However, $\nabla f(x) = 0$ and $\nabla^2 f(x)\succeq 0$ is not a sufficient condition for being a local minimum. For example, the original is not a local minimum of the function $f(x_1,x_2) = x_1^2 + x_2^3$ even though $\nabla f(0) =0$ and $\nabla^3 f(0) \succeq 0$. Generally speaking, along those direction $v$ where the Hessian vanishes (that is, $v^\top \nabla^2 f(x)v = 0$), the higher-order derivatives start to matter to the local optimality. In fact, finding a local minimum of a function is NP-hard~\citep{HillarL13}. Fortunately, with the following strict-saddle assumption, we can efficiently find a local minimum of the function $f$. A strict-saddle function satisfies that every saddle point must have a strictly negative curvature in some direction. It assumes away the difficult situation in the example above where higher-order derivatives are needed to decide if a point is a local minimum. \begin{definition}\label{def:strictsaddle} For $\alpha, \beta , \gamma \ge 0$, we say $f$ is $(\alpha,\beta,\gamma)$-\textit{strict saddle} if every $x\in \mathbb R^d$ satisfies \textit{at least} one of the following three conditions: \\ 1. $\norm{\nabla f(x)}_2 \ge \alpha$. \\ 2. $\lambda_{\min}(\nabla^2 f) \le -\beta$. \\ 3. There exists a local minimum $x^{\star}$ that is $\gamma$-close to $x$ in Euclidean distance. \end{definition} \begin{figure} \centering \begin{minipage}{.45\textwidth} \centering \includegraphics[height=4cm]{localmin} \end{minipage} ~~~~ \begin{minipage}{.5\textwidth} \centering \captionof{figure}{A two-dimensional function with the property that all local minima are global minima. It also satisfies the strict-saddle condition because all the saddle points have a trictly negative curvature in some direction. } \label{fig:wordvectors} \end{minipage} \end{figure} This condition is conjectured to hold for many real-world functions, and will be proved to hold for various problems concretely. However, in general, verifying it mathematically or empirically may be difficult. Under this condition, many algorithms can converge to a local minimum of $f$ in polynomial time as stated below.\footnote{Note that in this chapter, we only require polynomial time algorithm to be polynomial in $1/\epsilon$ when $\epsilon$ is the error. This makes sense for the downstream machine learning applications because very high accuracy solutions are not necessary due to intrinsic statistical errors.} \begin{theorem}\label{thm:localmin} Suppose $f$ is a twice differentiable $(\alpha,\beta,\gamma)$-strict saddle function from $\mathbb R^d\rightarrow \mathbb R$. Then, various optimization algorithms (such as stochastic gradient descent) can converge to a local minimum with $\epsilon$ error in Euclidean distance in time $\mbox{poly}(d, 1/\alpha, 1/\beta, 1/\gamma, 1/\epsilon)$. \end{theorem} \subsection{Local optimality vs global optimality } If a function $f$ satisfies the property that ``all local minima are global'' and the strict saddle property, we can provably find one of its global minima. (See Figure~\ref{fig:wordvectors} for an example of functions with this property. ) \begin{theorem}\label{thm:global}Suppose $f$ satisfies ``all local minima are global'' and the strict saddle property in a sense that all points satisfying approximately the necessary first order and second order optimality condition should be close to a global minimum: there exist $\epsilon_0, \tau_0 > 0$ and a universal constant $c> 0$ such that if a point $x$ satisfies $\norm{\nabla f(x)}_2\le \epsilon\le \epsilon_0$ and $\nabla^2 f(x)\succeq -\tau_0\cdot I$, then $x$ is $\epsilon^c$-close to a global minimum of $f$. Then, many optimization algorithms (including stochastic gradient descent and cubic regularization) can find a global minimum of $f$ up to $\delta$ error in $\ell_2$ norm in domain in time $\poly(1/\delta,1/\tau_0, d)$. \end{theorem} The technical condition of the theorem is often succinctly referred to as ``all local minima are global'', but its precise form, which is a combination of ``all local minima are global'' and the strict saddle condition, is crucial. There are functions that satisfy ``all local minima are global'' but cannot be optimized efficiently. Ignoring the strict saddle condition may lead to misleadingly strong statements. The condition of Theorem~\ref{thm:global} can be replaced by stronger ones which may occasionally be easier to verify, if they are indeed true for the functions of interests. One of such conditions is that ``any stationary point is a global minimum." The gradient descent is known to converge to a global minimum linearly, as stated below. However, because this condition effectively rules out the existence of multiple disconnected local minima, it can't hold for many objective functions related to neural networks, which guarantees to have multiple local minima and stationary points due to a certain symmetry. \begin{theorem}\label{thm:quasi} Suppose a function $f$ has $L$-Lipschitz continuous gradients and satisfies the Polyak-Lojasiewicz condition: $\exists$ $\mu > 0$ and $x^*$ such that for every $x$, \begin{align} \Norm{\nabla f(x)}_2^2 \ge \mu(f(x)-f(x^*))\ge 0 \,.\label{eqn:pl} \end{align} Then, the errors of the gradient descent with step size less than $1/(2L)$ decays geometrically. \end{theorem} It can be challenging to verify the Polyak-Lojasiewicz condition because the quantity $\Norm{\nabla{f(x)}}_2^2$ is often a complex function of $x$. An easier-to-verify but stronger condition is the quasi-convexity. Intuitively speaking, quasi-convexity says that at any point~$x$ the gradient should be negatively correlated with the direction~$x^*-x$ pointing towards the optimum. \renewcommand{{weakly-quasi-convex}\xspace}{{weakly-quasi-convex}} \begin{definition}[Weak quasi-convexity]\label{def:cond} We say an objective function $f$ is \textit{$\tau$-{weakly-quasi-convex}\xspace} over a domain~$\mathcal{B}$ with respect to the global minimum $x^*$ if there is a positive constant $\tau>0$ such that for all $x\in \mathcal{B}$, \begin{equation} \nabla f(x)^{\top} (x-x^*) \ge \tau (f(x)-f(x^*))\,.\label{eqn:condition} \end{equation} The following one is another related condition, which is sometimes referred to as the restricted secant inequality (RSI): \begin{equation} \nabla f(x)^{\top} (x-x^*) \ge \tau \|x-x^*\|_2^2.\label{eqn:msi} \end{equation} \end{definition} We note that convex functions satisfy~\eqref{eqn:condition} with $\tau =1$. Condition~\eqref{eqn:msi} is stronger than~\eqref{eqn:condition} because for smooth function, we have $\|x-x^*\|_2^2 \ge L (f(x) - f(x^*))$ for some constant $L$.\footnote{Readers who are familiar with convex optimization may realize that condition~\eqref{eqn:msi} is an extension of the strong convexity.} Conditions~\eqref{eqn:pl},~\eqref{eqn:condition}, and~\eqref{eqn:msi} all imply that all stationary points are global minimum because $\nabla f(x) = 0$ implies that $f(x) = f(x^*)$ or $x = x^*$. \subsection{Landscape for manifold-constrained optimization}\label{sec:manifold} We can extend many of the results in the previous section to the setting of constrained optimization over a smooth manifold. This section is only useful for problems in Section~\ref{sec:tensor} and casual readers can feel free to skip it. Let $\mathcal{M}$ be a Riemannian manifold. Let $T_x \mathcal{M}$ be the tangent space to $\mathcal{M}$ at $x$, and let $P_x$ be the projection operator to the tangent space $T_x \mathcal{M}$. Let $\textup{grad } f(x)\in T_x \mathcal{M}$ be the gradient of $f$ at $x$ on $\mathcal{M}$ and $\textup{Hess } f(x)$ be the Riemannian Hessian. Note that $\textup{Hess } f(x)$ is a linear mapping from $T_x \mathcal{M}$ onto itself. \begin{theorem}[Informally stated]\label{thm:manifold} \sloppy Consider the constrained optimization problem $ \min_{x\sim \mathcal{M}} f(x) $. Under proper regularity conditions, Theorem~\ref{thm:localmin} and Theorem~\ref{thm:global} still hold when replacing $\nabla f$ and $\nabla^2 f$ by $\textup{grad } f$ and $\textup{Hess } f$, respectively. \end{theorem} \noindent\textit{Backgrounds on manifold gradient and Hessian.} Later in Section~\ref{sec:tensor}, the unit sphere in $d$-dimensional space will be our constraint set, that is, $\mathcal{M} =S^{d-1}$. We provide some further backgrounds on how to compute the manifold gradients and Hessian here. We view $f$ as the restriction of a smooth function $\bar{f}$ to the manifold $\mathcal{M}$. In this case, we have $T_x \mathcal{M}= \{z\in \mathbb R^d: z^{\top} x = 0\}$, and $P_x = I - xx^{\top}$. We derive the manifold gradient of $f$ on $\mathcal{M}$: $ \textup{grad } f(x) = P_x \nabla\bar{f}(x)\,,\label{eqn:mani-gradient} $ where $\nabla$ is the usual gradient in the ambient space $\mathbb R^d$. Moreover, we derive the Riemannian Hessian as $ \textup{Hess } f(x) = P_x \nabla^2 \bar{f}(x) P_x - (x^\top \nabla \bar{f}(x)) P_x\label{eqn:mani-hessian}. $ \subsection{Matrix completion} \label{sec:mc} Matrix completion is the problem of recovering a low-rank matrix from partially observed entries, which has been widely used in collaborative filtering and recommender systems, dimension reduction, and multi-class learning. Despite the existence of elegant convex relaxation solutions, stochastic gradient descent on non-convex objectives are widely adopted in practice for scalability. We will focus on the rank-1 symmetric matrix completion in this chapter, which demonstrates the essence of the analysis. \subsubsection{Rank-1 case of matrix completion} Let $M = zz^{\top}$ be a rank-1 symmetric matrix with factor $z \in \mathbb R^{d}$ that we aim to recover. We assume that we observe each entry of $M$ with probability $p$ independently.\footnote{Technically, because $M$ is symmetric, the entries at $(i,j)$ and $(j,i)$ are the same. Thus, we assume that, with probability $p$ we observe both entries and otherwise we observe neither.} Let $\Omega \subset [d]\times [d]$ be the set of entries observed. Our goal is to recover from the observed entries of $M$ the vector $z$ up to sign flip (which is equivalent to recovering $M$). A known issue with matrix completion is that if $M$ is ``aligned'' with standard basis, then it's impossible to recover it. E.g., when $M = e_je_j^\top$ where $e_j$ is the $j$-th standard basis, we will very likely observe only entries with value zero, because $M$ is sparse. Such scenarios do not happen in practice very often though. The following standard assumption will rule out these difficult and pathological cases: \begin{assumption}[Incoherence] \label{assump:incoherence} W.L.O.G, we assume that $\|z\|_2=1$. In addition, we assume that $z$ satisfies $ \|z\|_\infty \le \frac{\mu}{\sqrt{d}}. $ We will think of $\mu$ as a small constant or logarithmic in $d$, and the sample complexity will depend polynomially on it. \end{assumption} In this setting, the vector $z$ can be recovered exactly up to a sign flip provided $\widetilde{\Omega}(d)$ samples. However, for simplicity, in this subsection we only aim to recover $z$ with an $\ell_2$ norm error $\epsilon \ll 1$. We assume that $p = \mbox{poly}(\mu,\log d)/(d\epsilon^2)$ which means that the expected number of observations is on the order of $d/\epsilon\cdot \polylog d$. We analyze the following objective that minimizes the total squared errors on the observed entries: \begin{align} \textup{argmin}_x ~f(x) := \frac12 \sum_{(i,j)\in \Omega}(M_{ij}-x_ix_j)^2= \frac12\cdot \norm{P_{\Omega}( M-xx^{\top})}_F ^2 \,. \end{align} Here $P_\Omega(A)$ denotes the matrix obtained by zeroing out all the entries of $A$ that are not in $\Omega$. For simplicity, we only focus on characterizing the landscape of the objective in the following domain $\mathcal{B}$ of incoherent vectors that contain the ground-truth vector $z$ (with a buffer of factor of 2) \begin{align} \mathcal{B} & = \Set{x: \|x\|_{\infty} < \frac{2\mu}{\sqrt{d}}} \,.\label{eqn:x_incoherent_0} \end{align} We note that the analyzing the landscape inside $\mathcal{B}$ does not suffice because the iterates of the algorithms may leave the set $\mathcal{B}$. We refer the readers to the original paper~\citep{ge2016matrix} for an analysis of the landscape over the entire space, or to the recent work~\citep{ma2018implicit} for an analysis that shows that the iterates won't leave the set of incoherent vectors if the initialization is random and incoherent. The global minima of $f(\cdot)$ are $z$ and $-z$ with function value 0. In the rest of the section, we prove that all the local minima of $f(\cdot)$ are $O(\sqrt{\epsilon})$-close to $\pm z$. \begin{theorem}\label{lem:mc:partial} In the setting above, all the local minima of $f(\cdot)$ inside the set $\mathcal{B}$ are $O(\sqrt{\epsilon})$-close to either $z$ or $-z$.\footnote{It's also true that the only local minima are exactly $\pm z$, and that $f$ has strict saddle property. However, their proofs are involved and beyond the scope of this chapter.} \end{theorem} It's insightful to compare with the full observation case when $\Omega = [d]\times [d]$. The corresponding objective is exactly the PCA objective $g(x) = \frac12\cdot \norm{ M-xx^{\top}}_F ^2$ defined in equation~\eqref{eqn:g}. Observe that $f(x)$ is a sampled version of the $g(x)$, and therefore we expect that they share the same geometric properties. In particular, recall that $g(x)$ does not have spurious local minima and thus we expect neither does $f(x)$. However, it‘s non-trivial to extend the proof of Theorem~\ref{thm:pca} to the case of partial observation, because it uses the {\em properties of eigenvectors} heavily. Indeed, suppose we imitate the proof of Theorem~\ref{thm:pca}, we will first compute the gradient of $f(\cdot)$: \begin{align} & \nabla f(x) = P_{\Omega}(zz^{\top}-xx^{\top})x\,. \end{align} Then, we run into an immediate difficulty --- how shall we solve the equation for stationary points $f(x) = P_{\Omega}(M-xx^{\top})x= 0$. Moreover, even if we could have a reasonable approximation for the stationary points, it would be difficult to examine their Hessians without using the exact orthogonality of the eigenvectors. The lesson from the trial above is that we may need to have an alternative proof for the PCA objective (full observation) that relies less on solving the stationary points exactly. Then more likely the proof can be extended to the matrix completion (partial observation) case. In the sequel, we follow this plan by first providing an alternative proof for Theorem~\ref{thm:pca}, which does not require solving the equation $\nabla g(x) = 0$, and then extend it via concentration inequality to a proof of Theorem~\ref{lem:mc:partial}. The key intuition will be is the following: \vspace{.05in} \textit{Proofs that consist of inequalities that are linear in $\mathbf{1}_{\Omega}$ are often easily generalizable to partial observation case.} \vspace{.05in} \noindent \sloppy Here statements that are linear in $\mathbf{1}_{\Omega}$ mean the statements of the form $\sum_{ij} 1_{(i,j) \in \Omega} T_{ij} \le a$. We will call these kinds of proofs ``simple'' proofs in this section. Indeed, by the law of large numbers, when the sampling probability $p$ is sufficiently large, we have that \begin{align} \underbrace{\sum_{(i,j)\in \Omega} T_{ij}}_{\textrm{partial observation}} = \sum_{i,j} \mathbf{1}_{(i,j)\in \Omega}T_{ij} \approx p \underbrace{\sum_{i,j}T_{ij}}_{\textrm{full observation}}\label{eqn:partialfull} \end{align} Then, the mathematical implications of $p\sum T_{ij} \le a$ are expected to be similar to the implications of $\sum_{(i,j)\in \Omega} T_{ij} \le a/p$, up to some small error introduced by the approximation. What natural quantities about $f$ are of the form $\sum_{(i,j)\in \Omega} T_{ij}$? First, quantities of the form $\inner{P_\Omega(A), B}$ can be written as $\sum_{(i,j)\in \Omega} A_{ij} B_{ij}$. Moreover, both the projection of $\nabla f$ and $\nabla^2 f$ are of the form $\inner{P_\Omega(A), B}$: \begin{align} \langle v, \nabla f(x)\rangle & = \inner{v,P_{\Omega}(zz^{\top}-xx^{\top})x} = \inner{P_{\Omega}(zz^{\top}-xx^{\top}), vx^\top} \nonumber\\ \langle v, \nabla^2 f(x) v\rangle & = \inner{P_{\Omega}(vx^{\top}+xv^{\top}), vx^{\top}+xv^{\top}}_F^2 - 2\inner{P_{\Omega}(vv^{\top}-xx^{\top}), vv^\top} \nonumber \end{align} The concentration of these quantities can all be captured by the following theorem below: \begin{theorem}\label{thm:concentration} Let $\epsilon > 0$ and $ p = \mbox{poly}(\mu,\log d)/(d\epsilon^2)$. Then, with high probability of the randomness of $\Omega$, we have that for all $A = uu^\top , B = vv^\top \in \mathbb R^{d\times d}$, where $\|u\|_2 \le 1, \|v\|2\le 1$ and $\|u\|_\infty, \|v\|_\infty \leq 2\mu/\sqrt{d}$. \begin{align} |\inner{P_\Omega(A), B}/p - \inner{A,B}| \le \epsilon\,.\end{align} \end{theorem} We will provide two claims below, combination of which proves Theorem~\ref{thm:pca}. In the proofs of these two claims, all the inequalities are of the form of LHS of equation~\eqref{eqn:partialfull}. Following each claim, we will immediately provide its extension to the partial observation case. \begin{customthm}{1f}\label{claim:full1} Suppose $x\in \mathcal{B}$ satisfies $\nabla g(x) = 0$, then $\inner{x,z}^2 = \|x\|_2^4$. \end{customthm} \begin{proof} By elementary calculation \begin{align} & \nabla g(x) = (zz^{\top}-xx^{\top})x= 0 \nonumber\\ \Rightarrow ~~& \inner{x,\nabla g(x)} = \inner{x,(zz^{\top}-xx^{\top})x} = 0 \quad\quad \quad \label{eqn:91}\\ \Rightarrow ~~& \inner{x,z}^2 =\|x\|_2^4 \nonumber \end{align} Intuitively, a stationary point $x$'s norm is governed by its correlation with $z$. \end{proof} \ni The following claim is the counterpart of Claim~\ref{claim:full1} in the partial observation case. \begin{customthm}{1p}\label{claim:partial1} Suppose $x\in \mathcal{B}$ satisfies $\nabla f(x) = 0$, then $\inner{x,z}^2 \ge \|x\|^4 -\epsilon$. \end{customthm} \begin{proof} Imitating the proof of Claim~\ref{claim:full1}, \begin{align} & \nabla f(x) = P_{\Omega}(zz^{\top}-xx^{\top})x= 0 \nonumber\\ \Rightarrow ~~& \inner{x,\nabla f(x)} = \inner{x,P_{\Omega}(zz^{\top}-xx^{\top})x} = 0\quad\quad \label{eqn:92}\\ \Rightarrow ~~& \inner{x,\nabla g(x)} = |\inner{x,(zz^{\top}-xx^{\top})x}| \leq \epsilon \quad\quad \label{eqn:13} \\ \Rightarrow ~~& \inner{x,z}^2 \ge \|x\|_2^4 -\epsilon\nonumber\end{align} where derivation from line~\eqref{eqn:92} to ~\eqref{eqn:13} follows the fact that line~\eqref{eqn:92} is a sampled version of~\eqref{eqn:13}. Technically, we can obtain it by applying Theorem~\ref{thm} twice with $A = B = xx^\top$ and $A=xx^\top$ and $B=zz^\top$ respectively. \end{proof} \begin{customthm}{2f}\label{claim:full2} If $x\in \mathcal{B}$ has positive Hessian $\nabla^2 g(x)\succeq 0$, then $\|x\|_2^2 \ge 1/3$. \end{customthm} \begin{proof} By the assumption on $x$, we have that $\inner{z,\nabla^2 g(x)z} \ge 0$. Calculating the quadratic form of the Hessian (which can be done by elementary calculus and is skipped for simplicity), we have \begin{align} & \inner{z,\nabla^2 g(x)z} = \norm{zx^{\top}+xz^{\top}}_F^2 - 2z^{\top}(zz^{\top}-xx^{\top})z\ge 0 \label{eqn:14} \end{align} This implies that \begin{align} & \Rightarrow \norm{x}_2^2 + 2\inner{z,x}^2 \ge 1 \nonumber\\ & \Rightarrow \norm{x}_2^2\ge 1/3\tag{since $\inner{z,x}^2\le \norm{x}_2^2$} \end{align} \end{proof} \begin{customthm}{2p}\label{claim:partial2} If $x\in \mathcal{B}$ has positive Hessian $\nabla^2 f(x)\succeq 0$, then $\|x\|_2^2 \ge 1/3-\epsilon/3$. \end{customthm} \begin{proof} Imitating the proof of Claim~\ref{claim:full2}, calculating the quadratic form over the Hessian at $z$, we have \begin{align} \inner{z,\nabla^2 f(x)z} = \norm{P_{\Omega}(zx^{\top}+xz^{\top})}_F^2 - 2z^{\top}P_{\Omega}(zz^{\top}-xx^{\top})z\ge 0 \end{align} Note that equation above is just a sampled version of equation~\eqref{eqn:14}, applying Theorem~\ref{thm:concentration} for various times (and note that $\inner{P_{Omega}(A), P_{\Omega}(B)} = \inner{P_\Omega(A), B}$, we can obtain that \begin{align} & \norm{P_{\Omega}(zx^{\top}+xz^{\top})}_F^2 - 2z^{\top}P_{\Omega}(zz^{\top}-xx^{\top})z \nonumber\\ & = p \cdot \left(\norm{zx^{\top}+xz^{\top}}_F^2 - 2z^{\top}(zz^{\top}-xx^{\top})z \pm \epsilon \right)\nonumber \end{align} Then following the derivation in the proof of Claim~\ref{claim:full2}, we achieve the same conclusion of Claim~\ref{claim:full2} up to approximation: $ \norm{x}^2\ge 1/3-\epsilon/3\nonumber. $ \end{proof} \ni With these claims, we are ready to prove Theorem~\ref{thm:pca} (again) and Theorem~\ref{lem:mc:partial}. \begin{proof}[Proof of Theorem~\ref{thm:pca} (again) and Theorem~\ref{lem:mc:partial}] By Claim~\ref{claim:full1} and~\ref{claim:full2}, we have $x$ satisfies $\inner{x,z}^2\ge \|x\|_2^4\ge 1/9$. Moreover, we have that $\nabla g(x) = 0 $ implies \begin{align} & \inner{z,\nabla g(x)} = \inner{z,(zz^{\top}-xx^{\top})x} = 0\label{eqn:93}\\ \Rightarrow ~~& \inner{x,z} (1 - \|x\|_2^2 ) = 0\nonumber\\ \Rightarrow ~~& \|x\|_2^2 = 1 \tag{by $\inner{x,z}^2\ge 1/9$} \end{align} Then by Claim~\ref{claim:full1} again we obtain $\inner{x,z}^2 = 1$, and therefore $x = \pm z$. The proof of Theorem~\ref{lem:mc:partial} are analogous (and note that such analogy was by design). When $\epsilon\le 1/12$, we by Claim~\ref{claim:partial1} and~\ref{claim:partial2}, we have when $\epsilon\le 1/16$, \begin{align} \inner{x,z}^2\ge \|x\|_2^4 - \epsilon\ge \left(\frac{1-\epsilon}{3}\right)^2 -\epsilon \ge \frac{1}{32} \end{align} Because $|\inner{z,\nabla g(x)} - \inner{z,\nabla f(x)}/p|\le \epsilon$, we have that \begin{align} & |\inner{z,\nabla g(x)}| = |\inner{z,(zz^{\top}-xx^{\top})x}| = O(\epsilon)\nonumber\\ \Rightarrow ~~& |\inner{x,z} (1 - \|x\|_2^2 ) |= O(\epsilon)\nonumber\\ \Rightarrow ~~& \|x\|_2^2 = 1 \pm O(\epsilon) \tag{by $\inner{x,z}^2\ge 1/32$} \end{align} Then by Claim~\ref{claim:partial1} again, we have $\inner{x,z}^2 \ge 1-O(\epsilon)$ which implies that $|\inner{x,z}| \ge 1-O(\epsilon)$. Now suppose $\inner{x,z} \ge 1-O(\epsilon)$, then we have \begin{align} \|x-z\|_2^2 = \|x\|_2^2 + \|z\|_2^2 - 2\inner{x,z} \le 1+O(\epsilon) + 1 - (1-O(\epsilon) \le O(\epsilon)\nonumber \end{align} Therefore $x$ is $O(\sqrt{\epsilon})$ close to $z$. On the other hand, if $\inner{x,z} \le -(1-O(\epsilon))$, we can similarly conclude that $x$ is $O(\sqrt{\epsilon})$-close to $-z$. \end{proof} \section{Notes} \citet{HillarL13} show that a degree four polynomial is NP-hard to optimize and ~\citet{murty1987some} show that it's also NP-hard to check whether a point is not a local minimum. Our quantitative definition quasi-convexity (Definition~\ref{def:cond}) is from~\citet{hardt2016gradient}. Polyak-Lojasiewicz condition was introduced by~\citet{polyak1963gradient}, and see a recent work of ~\citet{karimi2016linear} for a proof of Theorem~\ref{thm:quasi} . The RSI condition was originally introduced in~\citet{zhang2013gradient}. The strict saddle condition was originally defined in~\citep{ge2015escaping}, and we use a variant of the definition formalized in the work of~\citet{lee2016gradient,agarwal2016finding}. Formal versions of Theorem~\ref{thm:global} and Theorem~\ref{thm:localmin} for various concrete algorithms can be found in e.g., ~\citet{nesterov2006cubic, ge2015escaping, agarwal2017finding, carmon2016accelerated,sun2015nonconvex} and their follow-up works. Theorem~\ref{thm:manifold} is due to \citet[Theorem 12]{2016arXiv160508101B}. We refer the readers to the book~\cite{absil2007optimization} for the definition of gradient and Hessian on the manifolds and for the derivation of equation~\eqref{eqn:mani-gradient} and~\eqref{eqn:mani-hessian}.\footnote{For example, the gradient is defined in~\citet[Section3.6, Equation (3.31)]{absil2007optimization}, and the Hessian is defined in~\citet[Section 5.5, Definition 5.5.1]{absil2007optimization}. ~\cite[Example 5.4.1]{absil2007optimization} gives the Riemannian connection of the sphere $S^{d-1}$ which can be used to compute the Hessian.} The results covered in Section~\ref{sec:glm} was due to~\cite{kakade2011efficient,hazan2015beyond}. The particular exposition was first written by Yu Bai for the statistical learning theory course at Stanford. The analysis of the landscape of the PCA objective was derived in~\citet{baldi1989neural,srebro2003weighted}. The main result covered in Section~\ref{sec:mc} is based on the work~\citep{ge2016matrix}. Please see~\citet{ge2016matrix} for more references on the matrix completion problem. Nonconvex optimization has also been used for speeding up convex problems, e.g., the Burer-Monteiro approach~\citep{burer2005local} was theoretically analyzed by the work of~\citet{boumal2016non,bandeira2016low}. Section~\ref{sec:tensor} is based on the work of \citet{ge2015escaping}. Recently, there have been work on analyzing more sophisticated cases of tensor decomposition, e.g., using Kac-Rice formula~\citep{ge2017optimization} for random over-complete tensors. Please see the reference in~\citet{ge2017optimization} for more references regarding the tensor problems. \section{Survey and Outlook: Optimization of Neural Networks}\label{sec:neural_net} Theoretical analysis of algorithms for learning neural networks is highly challenging. We still lack handy mathematical tools. We will articulate a few technical challenges and summarize the attempts and progresses. We follow the standard setup in supervised learning. Let $f_{\theta}$ be a neural network parameterized by parameters $\theta$.\footnote{E.g., a two layer neural network would be $f_{\theta}(x) = W_1\sigma(W_2x)$ where $\theta = (W_1,W_2)$ and $\sigma$ are some activation functions.} Let $\ell$ be the loss function, and $\{(x^{(i)},y^{(i)})\}_{i=1}^n$ be a set of i.i.d examples drawn from distribution $D$. The empirical risk is $ \widehat{L}(\theta) = \frac{1}{n}\sum_{i=1}^{n}\ell(f_\theta(x^{(i)}), y^{(i)}), $ and the population risk is $ L(\theta) = \mathop{\mathbb E}\displaylimits_{(x,y)\sim D}\left[\ell(f_\theta(x), y)\right]. $ The major challenge of analyzing the landscape property of $\widehat{L}$ or $L$ stems from the non-linearity of neural networks---$f_{\theta}(x)$ is neither linear in $x$, nor in $\theta$. As a consequence, $\widehat{L}$ and $L$ are not convex in $\theta$. Linear algebra is at odds with neural networks---neural networks do not have good invariance property with respect to rotations of parameters or data points. \vspace{0.1in} \noindent{\bf Linearized neural networks.} Early works for optimization in deep learning simplify the problem by considering linearized neural networks: $f_\theta$ is assumed to be a neural networks without any activations functions. E.g., $f_{\theta} = W_1W_2W_3 x$ with $\theta = (W_1,W_2,W_3)$ would be a three-layer feedforward linearized neural network. Now, the model $f_\theta$ is still not linear in $\theta$, but it is linear in $x$. This simplification maintains the property that $\widehat{L}$ or $L$ are still nonconvex functions in $\theta$, but allows the use of linear algebraic tools to analyze the optimization landscapes of $\widehat{L}$ or $L$. \citet{baldi1989neural,kawaguchi2016deep} show that all the local minima of $L(\theta)$ are global minima when $\ell$ is the squared loss and $f_\theta$ is a linearized feed-forward neural network (but $L(\theta)$ does have degenerate saddle points so that it does not satisfy the strict saddle property). ~\citet{hardt2016gradient,hardt17identity} analyze the landscape of learning linearized residual and recurrent neural networks and show that all the stationary points (in a region) are global minima. We refer the readers to~\citet{arora2018optimization} and references therein for some recent works along this line. There are various results on another simplification: two-layer neural networks with quadratic activations. In this case, the model $f_{\theta}(x)$ is linear in $x\otimes x$ and quadratic in the parameters, and linear algebraic techniques allow us to obtain relatively strong theory. See~\citet{li2017algorithmic, soltanolkotabi2018theoretical,du2018power} and references therein. We remark that the line of results above typically applies to the landscape of the population losses as well as the empirical losses when there are sufficient number of examples.\footnote{Note that the former implies the latter when there are sufficient number of data points compared to the number of parameters, because in this case, the empirical loss has a similar landscape to that of the population loss due to concentration properties~\citep{mei2017landscape}. } \vspace{0.1in} \noindent{\bf Changing the landscape, by, e.g., over-parameterization or residual connection.} Somewhat in contrast to the clean case covered in earlier sections of this chapter, people have empirically found that the landscape properties of neural networks depend on various factors including the loss function, the model parameterization, and the data distribution. In particular, changing the model parameterization and the loss functions properly could ease the optimization. An effective approach to changing the landscape is to over-parameterize the neural networks --- using a large number of parameters by enlarging the width, often not necessary for expressivity and often bigger than the total number of training samples. It has been empirically found that wider neural networks may alleviate the problem of bad local minima that may occur in training narrower nets~\citep{livni2014computational}. This motivates a lot of studies of on the optimization landscape of over-parameterized neural networks. Please see~\citet{safran2016quality,venturi2018neural,soudry2016no,haeffele2015global} and the references therein. We note that there is an important distinction between two type of overparameterizations: (a) more parameters than what's needed for sufficient expressivity but still fewer parameters than the number of training examples, and (b) more parameters than the number of training examples. Under the latter setting, analyzing the landscape of empirical loss no longer suffices because even if the optimization works, the generalization gap might be too large or in other words the model overfits (which is an issue that is manifested clearly in the NTK discussion below.) In the former setting, though the generalization is less of a concern, analyzing the landscape is more difficult because it has to involve the complexity of the ground-truth function. Two extremely empirically successful approaches in deep learning, residual neural networks~\citep{he15deepresidual} and batch normalization~\citep{ioffe2015batch} are both conjectured to be able to change the landscape of the training objectives and lead to easier optimization. This is an interesting and promising direction with the potential of circumventing certain mathematical difficulties, but existing works often suffers from the strong assumptions such as linearized assumption in ~\citet{hardt17identity} and the Gaussian data distribution assumption in ~\citet{ge2018learning}. \vspace{0.1in} \noindent{\bf Connection between over-parametrized model and Kernel method: the Neural Tangent Kernel (NTK) view. } Another recent line of work studies the optimization dynamics of learning over-parameterized neural networks at a special type of initialization with a particular learning rate scheme~\citep{li2018learning,du2018gradient,jacot2018neural,allen2018convergence}, instead of characterizing the full landscape of the objective function. The main conclusion is of the following form:when using overparameterization (with more parameters than training examples), under a special type of initialization, optimizing with gradient descent can converge to a zero training error solution. The results can also be viewed/interpreted as a combination of landscape results and a convergence result: (i) the landscape in a small neighborhood around the initialization is sufficiently close to be convex, (ii) in the neighborhood a zero-error global minimum exists, and (iii) gradient descent from the initialization will not leave the neighborhood and will converge to the zero-error solution. Consider a non-linear model $f_{\theta}(\cdot)$ and an initialization $\theta_0$. We can approximate the model by a linear model by Taylor expansion at $\theta_0$: \begin{align} f_\theta(x) \approx g_\theta(x) & \triangleq \langle \theta-\theta_0, \nabla f_{\theta_0}(x)\rangle + f_{\theta_0}(x) = \langle \theta, \nabla f_{\theta_0}(x)\rangle + c(x)\label{eqn:approx} \end{align} where $c(x)$ only depends on $x$ but not $\theta$. Ignoring the non-essential shift $c(x)$, the model $g_{\theta}$ can be viewed as a linear function over the feature vector $\nabla f_{\theta_0}(x)$. Suppose the approximation in~\eqref{eqn:approx} is accurate enough throughout the training, then we are essentially optimizing the linear model $g_{\theta}(x)$, which leads to the part (i). \textit{For certain settings of initialization}, it turns out that (ii) and (iii) can also be shown with some proper definition of the neighborhood. \vspace{0.1in} \noindent{\bf Limitation of NTK and beyond.} A common limitation of analyses based on NTK is that they analyze directly the empirical risk whereas they do not necessarily provide good enough generalization guarantees. This is partially because the approach cannot handle regularized neural networks and the particular learning rate and level of stochasticity used in practice. In practice, typically the parameter $\theta$ does not stay close to the initialization either because of a large initial learning rate or small batch size. When the number of parameters in $\theta$ is bigger than $n$, without any regularization, we cannot expect that $\widehat{L}$ uniformly concentrates around the population risk. This raises the question of whether the obtained solution simply memorizes the training data and does not generalize to the test data. A generalization bound can be obtained by the NTK approach, by bounding the norm of the difference between the final solution and the initialization. However, such a generalization bound can only be effectively as good as what a kernel method can provide. In fact, \citet{wei2019regularization} show that, for a simple distribution, NTK has fundamentally worse sample complexity than a regularized objective for neural networks. This result demonstrates that the NTK regime of neural nets is statistically not as powerful as regularized neural nets, but it does not show that the regularized neural net can be optimized efficiently. Many recent works aim to separate training neural net and its NTK regime in a computationally-efficient sense, that is, to present a polynomial time algorithm of training neural networks that enjoys a better generalization guarantee than what the NTK result can offer. E.g., ~\cite{li2020learning} show that gradient descent can learn a two-layer neural net with orthonormal weights on a Gaussian data distribution from small but random initialization with a sample complexity better than the NTK approach. \citet{allen2019can} present a family of functions that can be learn efficiently by three-layer neural networks but not by NTK. These results are still largely demonstrating the possibility of stronger results for neural networks than NTK on some special cases, and it remains a major open question to have more general analysis for neural networks optimization beyond NTK. \vspace{0.03in} \noindent{\bf Regularized neural networks.} Analyzing the landscape or optimization of a regularized objective is more challenging than analyzing the un-regularized ones. In the latter case, we know that achieving zero training loss implies that we reach a global minimum, whereas in the former case, we know little about the function value of the global minima. Some progresses had been made for infinite-width two-layer neural networks~\citep{chizat2018global, mei2018mean,wei2019regularization,sirignano2018mean,rotskoff2018neural}. For example,~\citet{wei2019regularization} show that polynomial number of iterations of perturbed gradient descent can find a global minimum of an $\ell_2$ regularized objective function for infinite-width two-layer neural networks with homogeneous activations. However, likely the same general result won't hold for polynomial-width neural networks, if we make no additional assumptions on the data. \vspace{0.03in} \noindent{\bf Algorithmic or implicit regularization.} Empirical findings suggest, somewhat surprisingly, that even unregularized neural networks with over-parameterization can generalize~\citep{zhang2016understanding}. Moreover, different algorithms apparently converge to essentially differently global minima of the objective function, and these global minima have \textit{different} generalization performance! This means that the algorithms have a regularization effect, and fundamentally there is a possibility to delicately analyze the dynamics of the iterates of the optimization algorithm to reason about exactly which global minimum it converges to. Such types of results are particular challenging because it requires fine-grained control of the optimization dynamics, and rigorous theory can often be obtained only for relatively simple models such as linear models~\citep{soudry2018implicit} or matrix sensing~\citep{gunasekar2017implicit}, quadratic neural networks~\citep{li2017algorithmic}, a quadratically-parameterized linear model~\citep{woodworth2020kernel,vaskevicius2019implicit,haochen2020shape}, and special cases of two-layer neural nets with relu activations~\citep{li2019towards}. \vspace{0.03in} \noindent{\bf Assumptions on data distributions.} The author of the chapter and many others suspect that in the worst case, obtaining the best generalization performance of neural networks may be computationally intractable. Beyond the worst case analysis, people have made stronger assumptions on the data distribution such as Gaussian inputs~\citep{brutzkus2017globally,ge2018learning}, mixture of Gaussians or linearly separable data~\citep{brutzkus2017sgd}. The limitations of making Gaussian assumptions on the inputs are two-fold: a) it's not a realistic assumption; b) it may both over-estimate and under-estimate the difficulties of learning real-world data in different aspects. It is probably not surprising that Gaussian assumption can over-simplify the problem, but there could be other non-Gaussian assumptions that may make the problem even easier than Gaussians (e.g. see the early work in deep learning theory~\citep{arora2014provable}). \section{Matrix Factorization Problems}\label{sec:matrix} In this section, we will discuss the optimization landscape of two problems based on matrix factorization: principal component analysis (PCA) and matrix completion. The fundamental difference between them and the generalized linear models is that their objective functions have saddle points that are \textit{not} local minima or global minima. It means that the quasi-convexity condition or Polyak-Lojasiewicz condition does not hold for these objectives. Thus, we need more sophisticated techniques that can distinguish saddle points from local minima. \subsection{Principal Component Analysis} \label{sec:pca} One interpretation of PCA is approximating a matrix by its best low-rank approximation. Given a matrix $M\in \mathbb R^{d_1\times d_2}$, we aim to find its best rank-$r$ approximation (in either Frobenius norm or spectral norm). For the ease of exposition, we take $r=1$ and assume $M$ to be symmetric positive semi-definite with dimension $d$ by $d$. In this case, the best rank-1 approximation has the form $xx^\top$ where $x\in \mathbb R^d$. There are many well-known algorithms for finding the low-rank factor $x$. We are particularly interested in the following non-convex program that directly minimizes the approximation error in Frobenius norm. \begin{align} \min_x ~g(x) := \frac12\cdot \norm{ M-xx^{\top}}_F ^2 \,. \label{eqn:g} \end{align} We will prove that even though $g$ is not convex, all the local minima of $g$ are global. It also satisfies the strict saddle property (which we will not prove formally here). Therefore, local search algorithms can solve~\eqref{eqn:g} in polynomial time.\footnote{In fact, local methods can solve it very fast. See, e.g., \citet[Thereom 1.2]{li2017algorithmic}} \begin{theorem}\label{thm:pca} In the setting above, all the local minima of the objective function $g(x) $ are global minima.\footnote{The function $g$ also satisfies the $(\alpha,\beta,\gamma)$-strict-saddle property (Definition~\ref{def:strictsaddle}) with some $\alpha, \beta, \gamma > 0$ (that may depend on $M$) so that it satisfies the condition of Theorem~\ref{thm:global}. We skip the proof of this result for simplicity.} \end{theorem} Our analysis consists of two main steps: a) to characterize all the stationary points of the function $g$, which turn out to be the eigenvectors of $M$; b) to examine each of the stationary points and show that the only the top eigenvector(s) of $g$ can be a local minimum. Step b) implies the theorem because the top eigenvectors are also global minima of $g$. We start with step a) with the following lemma. \begin{lemma}\label{lem:pcastaionary} In the setting of Theorem~\ref{thm:pca}, all the stationary points of the objective $g()$ are the eigenvectors of $M$. Moreover, if $x$ is a stationary point, then $\|x\|_2^2$ is the eigenvalue corresponding to $x$. \end{lemma} \begin{proof} By elementary calculus, we have that \begin{align} \nabla g(x) = -(M-xx^\top)x = \|x\|_2^2\cdot x - Mx \end{align} Therefore, if $x$ is a stationary point of $g$, then $Mx = \|x\|_2^2 \cdot x$, which implies that $x$ is an eigenvector of $M$ with eigenvalue equal to $\|x\|_2^2$. \end{proof} Now we are ready to prove b) and the theorem. The key intuition is the following. Suppose we are at a point $x$ that is an eigenvector but not the top eigenvector, moving in either the top eigenvector direction $v_1$ or the direction of $-v_1$ will result in a second-order local improvement of the objective function. Therefore, $x$ cannot be a local minimum unless $x$ is a top eigenvector. \begin{proof}[Proof of Theorem~\ref{thm:pca}] By Lemma~\ref{lem:pcastaionary}, we know that a local minimum $x$ is an eigenvector of $M$. If $x$ is a top eigenvector of $M$ with the largest eigenvalue, then $x$ is a global minimum. For the sake of contradiction, we assume that $x$ is an eigenvector with eigenvalue $\lambda$ that is strictly less than $\lambda_1$. By Lemma~\ref{lem:pcastaionary} we have $\lambda = \|x\|_2^2$. By elementary calculation, we have that \begin{align} \nabla^2 g(x) = 2xx^{\top} - M + \|x\|_2^2 \cdot I\,.\label{eqn:2} \end{align} Let $v_1$ be the top eigenvector of $M$ with eigenvalue $\lambda_1$ and with $\ell_2$ norm 1. Then, because $\nabla^2 g(x) \succeq 0$, we have that \begin{align} v_1^\top \nabla^2 g(x) v \ge 0 \label{eqn:1} \end{align} It's a basic property of eigenvectors of positive semidefinite matrix that any pairs of eigenvectors with different eigenvalues are orthogonal to each other. Thus we have $\inner{x, v_1} = 0$. It follows equation~\eqref{eqn:1} and~\eqref{eqn:2} that \begin{align} 0 & \leq v_1^\top (2xx^{\top} - M + \|x\|_2^2 \cdot I) v_1 = \|x\|_2^2 - v_1^\top M v_1 \tag{by $\inner{x, v_1} = 0$}\\ &= \lambda - \lambda_1 \tag{by that $v_1$ has eigenvalue $\lambda_1$ and that $\lambda=\|x\|_2^2$}\\ & < 0 \tag{by the assumption} \end{align} which is a contradiction. \end{proof} \section{Landscape of Tensor Decomposition} \label{sec:tensor} In this section, we analyze the optimization landscape for another machine learning problem, tensor decomposition. The fundamental difference of tensor decomposition from matrix factorization problems or generalized linear models is that the non-convex objective function here has multiple isolated local minima, and therefore the set of local minima does not have rotational invariance (whereas in matrix completion or PCA, the set of local minima are rotational invariant.). This essentially prevents us to only use linear algebraic techniques, because they are intrinsically rotational invariant. \subsection{Non-convex optimization for orthogonal tensor decomposition and global optimality} We focus on one of the simplest tensor decomposition problems, orthogonal 4-th order tensor decomposition. Suppose we are given the entries of a symmetric 4-th order tensor $T \in \mathbb R^{d \times d \times d\times d}$ which has a low rank structure in the sense that: \begin{align} T = \sum_{i=1}^n a_i\otimes a_i\otimes a_i \otimes a_i \label{eqn:lowrank} \end{align} where $a_1,\dots, a_n \in \mathbb R^d$. Our goal is to recover the underlying components $a_1,\dots, a_n$. We assume in this subsection that $a_1,\dots, a_n$ are orthogonal vectors in $\mathbb R^d$ with unit norm (and thus implicitly we assume $n\le d$.) Consider the objective function \begin{align} \textup{argmax} & ~~f(x) := \inner{T, x^{\otimes 4}}\label{eqn:obj}\\ \textup{s.t.} &~~ \|x\|_2^2 = 1 \nonumber \end{align} The optimal value function for the objective is the (symmetric) injective norm of a tensor $T$. In our case, the global maximizers of the objective above are exactly the set of components that we are looking for. \begin{theorem}\label{thm:tensorglobal} Suppose $T$ satisfies equation~\eqref{eqn:lowrank} with orthonormal components $a_1,\dots, a_n$. Then, the global maximizers of the objective function~\eqref{eqn:obj} are exactly $\pm a_1, \dots, \pm a_n$. \end{theorem} \subsection{All local optima are global} We next show that all the local maxima of the objective~\eqref{eqn:obj} are also global maxima. In other words, we will show that $\pm a_1,\dots, \pm a_n$ are the only local maxima. We note that all the geometry properties here are defined with respect to the manifold of the unit sphere $\mathcal{M} = S^{d-1}$. (Please see Section~\ref{sec:manifold} for a brief introduction of the notions of manifold gradient, manifold local maxima, etc.) \begin{theorem}\label{thm:tensorlocal} In the same setting of Theorem~\ref{thm:tensorglobal}, all the local maxima (w.r.t the manifold $S^{d-1}$) of the objective~\eqref{eqn:obj} are global maxima. \footnote{The function also satisfies the strict saddle property so that we can rigorously invoke Theorem~\ref{thm:manifold}. However, we skip the proof of that for simplicity.} \end{theorem} Towards proving the Theorem, we first note that the landscape property of a function is invariant to the coordinate system that we use to represent it. It's natural for us to use the directions of $a_1,\dots, a_n$ together with an arbitrary basis in the complement subspace of $a_1,\dots, a_n$ as the coordinate system. A more convenient viewpoint is that this choice of coordinate system is equivalent to assuming $a_1,\dots, a_n$ are the natural standard basis $e_1,\dots, e_n$. Moreover, one can verify that the remaining directions $e_{n+1},\dots, e_d$ are irrelevant for the objective because it's not economical to put any mass in those directions. Therefore, for simplicity of the proof, we make the assumption below without loss of generality: \begin{align} n= d, \textup{ and } a_i = e_i, ~\forall i\in [n]\,. \end{align} Then we have that $f(x) = \|x\|_4^4$. We compute the manifold gradient and manifold Hessian using the formulae of $\textup{grad } f(x)$ and $\textup{Hess } f(x)$ in Section~\ref{sec:manifold}, \begin{align} \textup{grad } f(x) &= 4P_x \nabla \bar{f}(x) = 4(I_{d\times d}-xx^\top) \begin{bmatrix} x_1^3 \\ \vdots \\ x_d^3 \end{bmatrix}= 4\begin{bmatrix} x_1^3 \\ \vdots \\ x_d^3 \end{bmatrix} - 4\|x\|_4^4 \cdot \begin{bmatrix} x_1 \\ \vdots \\ x_d \end{bmatrix}\,.\label{eqn:grad} \end{align} \begin{align} \textup{Hess } f(x) & = P_x \nabla^2 \bar{f}(x) P_x - (x^\top \nabla \bar{f}(x)) P_x \nonumber\\ & = P_x\left(12\mathop{\mathrm{diag}}(x_1^2,\dots, x_d^2) - 4\|x\|_4^4 \cdot I_{d\times d}\right) P_x \label{eqn:tensorhessian} \end{align} where $\mathop{\mathrm{diag}}(v)$ for a vector $v\in \mathbb R^d$ denotes the diagonal matrix with $v_1,\dots, v_d$ on the diagonal. Now we are ready to prove Theorem~\ref{thm:tensorlocal}. In the proof, we will first compute all the stationary points of the objective, and then examine each of them and show that only $\pm a_1,\dots, \pm a_n$ can be local maxima. \begin{proof}[Proof of Theorem~\ref{thm:tensorlocal}] We work under the assumptions and simplifications above. We first compute all the stationary points of the objective~\eqref{eqn:obj} by solving $\textup{grad } f = 0$. Using equation~\eqref{eqn:grad}, we have that the stationary points satisfy that \begin{align} x_i^3 = \|x\|_4^4 \cdot x_i, \forall i \end{align} It follows that $x_i = 0$ or $x_i = \pm \|x\|_4^{1/2}$. Assume that $s$ of the $x_i$'s are non-zero and thus take the second choice, we have that \begin{align} 1 = \|x\|_2^2 = s\cdot \|x\|_4^4 \end{align} This implies that $\|x\|_4^4 = 1/s$, and $x_i = 0$ or $\pm 1/s^{1/2}$. In other words, all the stationary points of $f$ are of the form $(\pm 1/s^{1/2}, \cdots, \pm 1/s^{1/2}, 0,\cdots, 0)$ (where there are $s$ non-zeros) for some $s\in [d]$ and all their permutations (over indices). Next, we examine which of these stationary points are local maxima. Let $\tau = 1/s^{1/2}$ for simplicity. This implies that $\|x\|_4^4 = \tau^2$. Consider a stationary point $x = (\sigma_1\tau, \cdots, \sigma_s\tau, 0,\dots, 0)$ where $\sigma_i \in \{-1,1\}$. Let $x$ be a local maximum. Thus $\textup{Hess } f(X) \preceq 0$. We will prove that this implies $s= 1$. For the sake of contradiction, we assume $s\ge 2$. We will show that the Hessian cannot be negative semi-definite by finding a particular direction in which the Hessian has positive quadratic form. The form of equation\eqref{eqn:tensorhessian} implies that for all $v$ such that $\inner{v,x} = 0$ (which indicates that $P_x v = v$), we have \begin{align} v^\top \left((12\mathop{\mathrm{diag}}(x_1^2,\dots, x_d^2) - 4\|x\|_4^4 I\right) v \leq 0 \label{eqn:3} \end{align} We take $v = (1/2, -1/2)$ to be our test direction. Then LHS of the formula above simplifies to \begin{align} 3x_1^2 - 3x_2^2 - 2 \|x\|_4^4 = 6\tau^2 - 2 \|x\|_4^4 = 4\tau^2 > 0 \end{align} which contradicts to equation~\eqref{eqn:3}. Therefore, $s = 1$, and we conclude that all the local maxima are $\pm e_1.\dots, \pm e_d$. \end{proof}
{ "timestamp": "2021-03-26T01:04:10", "yymm": "2103", "arxiv_id": "2103.13462", "language": "en", "url": "https://arxiv.org/abs/2103.13462", "abstract": "Non-convex optimization is ubiquitous in modern machine learning. Researchers devise non-convex objective functions and optimize them using off-the-shelf optimizers such as stochastic gradient descent and its variants, which leverage the local geometry and update iteratively. Even though solving non-convex functions is NP-hard in the worst case, the optimization quality in practice is often not an issue -- optimizers are largely believed to find approximate global minima. Researchers hypothesize a unified explanation for this intriguing phenomenon: most of the local minima of the practically-used objectives are approximately global minima. We rigorously formalize it for concrete instances of machine learning problems.", "subjects": "Machine Learning (cs.LG); Data Structures and Algorithms (cs.DS); Optimization and Control (math.OC); Machine Learning (stat.ML)", "title": "Why Do Local Methods Solve Nonconvex Problems?", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES\n\n", "lm_q1_score": 0.9850429156022933, "lm_q2_score": 0.8311430415844384, "lm_q1q2_score": 0.8187115649648933 }
https://arxiv.org/abs/1706.05410
On approximate Gauss-Lucas theorems
The Gauss--Lucas theorem states that any convex set $K\subset\mathbb{C}$ which contains all $n$ zeros of a degree $n$ polynomial $p\in\mathbb{C}[z]$ must also contain all $n-1$ critical points of $p$. In this paper we explore the following question: for which choices of positive integers $n$ and $k$, and positive real number $\epsilon$, will it follow that for every degree $n$ polynomial $p$ with at least $k$ zeros lying in $K$, $p$ will have at least $k-1$ critical points lying in the $\epsilon$-neighborhood of $K$. We supply an inequality relating $n$, $k$, and $\epsilon$ which, when satisfied, guarantees a positive answer to the above question.
\section{Overview and definitions.}% The Gauss--Lucas theorem states that if all $n$ of the zeros of a degree $n$ complex polynomial $p$ lie in a convex set $K\subset\mathbb{C}$, then all $n-1$ of the critical point of $p$ (that is, the zeros of $p'$) lie in $K$ as well. One might expect that if at least $k$ zeros of a degree $n$ polynomial $p$ lie in $K$, then at least $k-1$ of the critical points of $p$ will lie in $K$. In general this fails badly. Consider the example (drawn from~\cite{T}) where $K=[0,1]$, and $p$ is a polynomial with $n-1$ distinct zeros in $K$, and a single zero at $i$. A little arithmetic shows that in fact every critical point of $p$ will have strictly positive imaginary part, and thus will lie outside of $K$. However for any fixed neighborhood $\mathcal{O}$ of $K$, if $n$ is sufficiently large, then all $n-1$ of the critical points of $p$ will lie in $\mathcal{O}$ (as will follow from Theorem~\ref{thm: Main inequalities.} to come). This last observation suggests the following property, which we will explore in this paper. For positive integers $n$ and $k$, with $k\leq n$, and for $\epsilon>0$, we say that $K$ satisfies the approximate Gauss--Lucas property (denoted $\mathcal{AGL}(n,k,\epsilon)$) if: \medskip \begin{minipage}{.9\textwidth} \begin{note}[$\mathcal{AGL}(n,k,\epsilon)$:] Every degree $n$ polynomial with at least $k$ zeros lying in $K$ has at least $k-1$ critical points lying in the $\epsilon$-neighborhood of $K$. \end{note} \end{minipage} \medskip The stronger property, where ``polynomial'' is replaced by ``rational function'', and $n$ denotes not the degree of the rational function, but the combined number of zeros and poles of the rational function, will be denoted $\mathcal{RAGL}(n,k,\epsilon)$ (the ``rational approximate Gauss--Lucas property''). For a bounded convex set $K\subset\mathbb{C}$, our goal is to find conditions on $n$, $k$, and $\epsilon$ which will ensure that $K$ satisfies $\mathcal{RAGL}(n,k,\epsilon)$. Our main theorem is the following. \begin{theorem}\label{thm: Main inequalities.} Let $K\subset\mathbb{C}$ be bounded and convex, with diameter $s\geq0$. For positive integers $n$ and $k$, with $k\leq n$, and a positive real number $\epsilon$, if $n$, $k$, and $\epsilon$ satisfy the inequality $$\dfrac{16(s+\epsilon)^2}{\epsilon^2}<\dfrac{k}{(n-k)^2},$$ then $K$ satisfies $\mathcal{RAGL}(n,k,\epsilon)$. Moreover, if $K$ is known to be a disk, it suffices for $n$, $k$, and $\epsilon$ to satisfy $$\dfrac{8(s+\epsilon)}{\epsilon}<\dfrac{k}{(n-k)^2}$$ to ensure that $K$ satisfies $\mathcal{RAGL}(n,k,\epsilon)$. \end{theorem} Historically, there have been two different approaches to determining for which values of $n$, $k$, and $\epsilon$ the set $K$ will satisfy $\mathcal{RAGL}(n,k,\epsilon)$. The first, pioneered by S.~Kakeya~\cite{K} views $n$ and $k$ as fixed, and seeks the smallest possible $\epsilon$ for which $K$ satisfies $\mathcal{RAGL}(n,k,\epsilon)$. The second, which views $\epsilon$ as fixed, and seeks to find choices of $n$ and $k$ for which $K$ satisfies $\mathcal{RAGL}(n,k,\epsilon)$, is related to the~2016 asymptotic Gauss--Lucas theorem of V. Totik~\cite{T}. Let us first introduce some notation. For a set $G\subset\mathbb{C}$ and a rational function $f\in\mathbb{C}(z)$, let $\#_z(f,G)$, $\#_p(f,G)$, and $\#_c(f,G)$ denote the number of zeros, poles, and critical points of $f$ respectively which lie in $G$. For $\epsilon>0$, let $G_\epsilon$ denote the $\epsilon$-neighborhood of $G$. Thus the $\mathcal{RAGL}(n,k,\epsilon)$ property may be restated as: \medskip \begin{minipage}{.9\textwidth} \begin{note}[$\mathcal{RAGL}(n,k,\epsilon)$:] For every rational function $f$ with $\#_z(f,\mathbb{C})+\#_p(f,\mathbb{C})=n$, if $\#_z(f,K)\geq k$ then $\#_c(f,K_\epsilon)\geq k-1$. \end{note} \end{minipage} \medskip \subsection{Fixing \texorpdfstring{$n$}{TEXT} and \texorpdfstring{$k$}{TEXT}, and seeking \texorpdfstring{$\epsilon$}{TEXT}.} In 1917, S. Kakeya~\cite{K} showed that if $K$ is the disk $D_R$ centered at the origin with radius $R\geq0$, then there is a function $\epsilon=\psi_R(n,k)<\infty$ such that $D_R$ satisfies $\mathcal{AGL}(n,k,\psi_R(n,k))$, and he showed moreover that for any $n$ and $k$, $\psi_R(n,k)=R\cdot\psi_1(n,k)$. In this work Kakeya was also able to find $\psi_1(n,k)$ in the special case where $k=2$: $\psi_1(n,2)=\csc(\pi/n)-1$. (Note: in~\cite{K} and elsewhere in the literature, the quantity sought, rather than $\epsilon$, is $R'=R+\epsilon$, the radius of the larger circle guaranteed to contain the expected number of critical points. In this paper, we focus on the ``$\epsilon$-cushion'' needed around the set $K$, as this concept still makes sense in the context of bounded convex sets which are not disks.) Two upper bounds are known for $\psi_1(n,k)$ in general. First in~1939, M. Marden~\cite{M} established that \begin{equation}\label{eqn: Marden upper bound on psi.} \psi_1(n,k)\leq\csc\left(\dfrac{\pi}{2(n-k+1)}\right)-1. \end{equation} Second, in 1945, M. Biernacki~\cite{B} established that \begin{equation}\label{eqn: Biernacki upper bound on psi.} \psi_1(n,k)\leq\displaystyle\left(\prod_{j=1}^{n-k}\dfrac{n+j}{n-j}\right)-1. \end{equation} Neither of the bounds of Marden and Biernacki is universally better than the other, however since Marden's bound is a function of $n-k$, we see that if $n-k$ is held constant, and $n$ is allowed to approach $\infty$, Marden's bound remains constant, while Biernacki's bound approaches zero. If we solve the first inequality given in Theorem~\ref{thm: Main inequalities.} for $\epsilon$, we obtain an upper bound for the quantity corresponding to $\psi_R$ in the general case where $K$ is bounded and convex, but not necessarily a disk, and $f$ is rational, not necessarily a polynomial. That is, we obtain the following. \begin{corollary}\label{cor: n and k held constant, solve for epsilon.} Let $K\subset\mathbb{C}$ be bounded and convex with diameter $s$, and let $k$ and $n$ be positive integers with $k<n$, and such that $\sqrt{k}>4(n-k)$. If $$\epsilon>\dfrac{4s(n-k)}{\sqrt{k}-4(n-k)},$$ then $K$ satisfies $\mathcal{RAGL}(n,k,\epsilon)$. \end{corollary} In the case where $K$ is a disk, the second inequaltiy in Theorem~\ref{thm: Main inequalities.} gives us the following bound on $\psi_1$. \begin{corollary}\label{cor: Our bound on psi_1.} Let $k$ and $n$ be positive integers with $k<n$, and such that $k>8(n-k)^2$. Then $$\psi_1(n,k)\leq\dfrac{8(n-k)^2}{k-8(n-k)^2}.$$ \end{corollary} It is difficult to determine in exactly which cases any one of the given three upper bounds on $\psi_1$ (Marden's, Biernacki's, and Corollary~\ref{cor: Our bound on psi_1.} above) is the lowest. However, as mentioned before, Marden's bound is constant as a function of $n-k$. Biernacki's bound, on the other hand, may be written as \begin{equation} \displaystyle\left(\prod_{i=1}^{n-k}\dfrac{n+i}{n-i}\right)-1=\dfrac{p(1/n)}{n+q(1/n)} \end{equation} where $p$ and $q$ are polynomials, whose coefficients are functions of $n-k$, and the constant term of $p$ is $(n-k)(n-k+1)$. Thus Biernacki's bound implies that if $n-k$ is fixed, and $n\to\infty$, then $\psi_1(n,k)\in O(1/n)$. This fact follows also from our bound in Corollary~\ref{cor: Our bound on psi_1.}. Note that since (again when $n-k$ is fixed and $n\to\infty$) Biernacki's bound is on the order of $(n-k)(n-k+1)/n$, while the bound in Corollary~\ref{cor: Our bound on psi_1.} is on the order of $8(n-k)^2/n$, so that Biernacki's bound is superior to Corollary~\ref{cor: Our bound on psi_1.} in the asymptotic sense. For more information on this approach, see Sections~25 and~26 of~\cite{M2}. \subsection{Fixing \texorpdfstring{$\epsilon$}{TEXT}, and seeking \texorpdfstring{$n$}{TEXT} and \texorpdfstring{$k$}{TEXT}.} In 2016, V. Totik~\cite{T} established the following approximate and asymptotic version of the Gauss--Lucas theorem. \begin{theorem}\label{thm: Asymp. GL.} Let $K\subset\mathbb{C}$ be bounded and convex, and let $\epsilon>0$ be given. For any sequence of polynomials $\{p_n\}$, with $\deg(p_n)=n$, if $\dfrac{\#_z(p_n,K)}{n}\to1$, then $\dfrac{\#_c(p_n,K_\epsilon)}{n-1}\to1$. \end{theorem} (Note that another proof of Theorem~\ref{thm: Asymp. GL.} is posted on the arXiv~\cite{BKS} by R. Boegvad et. al. using more classical methods of complex analysis, where the proof of V. Totik uses results from logarithmic potential theory.) We believe that underlying this asymptotic result (that is, Theorem~\ref{thm: Asymp. GL.}) is a static principle, namely that for any $\epsilon>0$, if a sufficiently high fraction of the zeros of a polynomial $p$ lie in $K$, then a similarly high fraction of the critical points of $p$ lie in $K_\epsilon$ (where ``sufficiently high'' depends only on $K$ and $\epsilon$, and not on the degree of $p$). That is, we make the following conjecture, again including the possibility of poles. \begin{conjecture}\label{conj: epsilon fixed goal.} Let $K\subset\mathbb{C}$ be bounded and convex, and let $\epsilon>0$ be given. There is some constant $C>0$ such that if $k/(n-k)>C$, then $K$ satisfies $\mathcal{RAGL}(n,k,\epsilon)$. \end{conjecture} Clearly a positive result for this conjecture would immediately imply the conclusion of Theorem~\ref{thm: Asymp. GL.} as well. In support of Conjecture~\ref{conj: epsilon fixed goal.}, we observe that the first inequality in Theorem~\ref{thm: Main inequalities.} establishes the result of Conjecture~\ref{conj: epsilon fixed goal.} subject to the stronger condition that the ratio of $k$ to the \textbf{square} of $(n-k)$ is sufficiently large. \section{A BRIEF LEMMA}\label{sect: Lemma.}% \begin{lemma}\label{lem: Lemma.} Let $r\in\mathbb{C}(z)$ be a rational function, all of whose zeros and poles have multiplicity one. Then the zeros of the logarithmic derivative $\dfrac{r'}{r}$ are exactly the critical points of $r$, with the same multiplicities, and the poles of $\dfrac{r'}{r}$ are the zeros and poles of $r$, each with multiplicity one. \end{lemma} \begin{proof} Let $p,q\in\mathbb{C}[z]$ be polynomials such that $r=\dfrac{p}{q}$ is in lowest form. A bit of arithmetic gives that $$\dfrac{r'}{r}=\dfrac{qp'-pq'}{pq}.$$ The numerator $qp'-pq'$ is exactly the numerator of $r'$, and since each zero and pole of $r$ has multiplicity one, no zero of either $p$ or $q$ is a zero of $qp'-pq'$, so that no terms cancel in our expansion of $\dfrac{r'}{r}$. Moreover, the poles of $\dfrac{r'}{r}$ are just the zeros of $p$ and $q$, namely the zeros and poles of $r$. This completes the proof. \end{proof} \section{PROOFS}\label{sect: Proofs.}% We proceed to our proof of Theorem~\ref{thm: Main inequalities.}. Let $K\subset\mathbb{C}$ be bounded and convex, with diameter $s\geq0$. Let $n$ and $k$ be positive integers with $k\leq n$, and let $\epsilon>0$ be given. Suppose that $f$ is a rational function with $n$ zeros and poles combined (counting multiplicity), and having at least $k$ zeros lying in $K$ (that is, $\#_z(f,\mathbb{C})+\#_p(f,\mathbb{C})=n$, and $\#_z(f,K)\geq k$). In the case that $k=n$, $f$ is a polynomial with all of its zeros lying in $K$, so all of its critical points lie in $K$ by the classical Gauss--Lucas theorem. Thus we consider only the case $k<n$. We first factor $f$ as $f=g\cdot h$, where $g$ is the monic polynomial having for its zeros exactly the at least $k$ zeros of $f$ which lie in $K$, and $h$ is a rational function with at most $n-k$ zeros and poles combined. Since the locations of the critical points of $g$, $h$, and $gh$ are continuous as a function of the zeros and poles of $g$ and $h$, we may assume without loss of generality that the zeros and poles of $g$ and $h$ are all distinct. Define the set $$E=\left\{z\in\mathbb{C}:\dfrac{\epsilon}{2}< d(z,K)<\epsilon\right\},$$ where $d(z,K)$ denotes the distance $d(z,K)=\displaystyle\inf_{w\in K}(|z-w|)$. Since $K$ is convex, $E$ is a bounded conformal annulus (that is, a bounded open subset of $\mathbb{C}$ such that $E^c$ has a single bounded component) that contains $K$ in its bounded face. Let $\mathcal{A}$ denote the union of the closed balls centered at the zeros and poles of $h$, each with radius $\dfrac{\epsilon}{8(n-k)}$. $\mathcal{A}$ consists of at most $n-k$ closed balls, each with diameter $\dfrac{\epsilon}{4(n-k)}$, so the diameter of any component of $\mathcal{A}$ is at most $\dfrac{\epsilon}{4}$. However for any points $z$ in the bounded face of $E$ and $z'$ in the unbounded face of $E$, the triangle inequality implies that $|z-z'|\geq\dfrac{\epsilon}{2}$, so it follows that $E\setminus\mathcal{A}$ is still an open set with $K$ contained in a bounded face of $E\setminus\mathcal{A}$. Since $K$ is contained in a bounded face of $E\setminus\mathcal{A}$, and $E\setminus\mathcal{A}$ is open, we may find a smooth path $\gamma$ which lies in $E\setminus\mathcal{A}$, and winds once around the face of $E\setminus\mathcal{A}$ which contains $K$. Our plan is to apply Rouch\'e's theorem to the functions $\dfrac{g'}{g}$ and $\dfrac{h'}{h}$ on the path $\gamma$. Thus we wish to have $\left|\dfrac{g'}{g}\right|>\left|\dfrac{h'}{h}\right|$ on $\gamma$. Fix some point $z_0$ in $\gamma$. We will first find an upper bound for $\left|\dfrac{h'(z_0)}{h(z_0)}\right|$. Let $Z_h$ and $P_h$ denote the set of zeros of $h$ and poles of $h$ respectively (recall that $h$ was assumed to have all distinct zeros and poles). Then we have $|Z_h|+|P_h|\leq n-k$. After a little arithmetic, we have \begin{equation}\label{eqn: Breakdown of h'/h.} \dfrac{h'(z_0)}{h(z_0)}=\displaystyle\sum_{x\in Z_h}\dfrac{1}{z_0-x}-\sum_{y\in P_h}\dfrac{1}{z_0-y}. \end{equation} Since $z_0$ lies outside of $\mathcal{A}$, for any $w$ in either $Z_h$ or $P_h$, $|z_0-w|>\epsilon/8(n-k)$. Thus, using the triangle inequality on Equation~\ref{eqn: Breakdown of h'/h.}, we have \begin{equation}\label{eqn: Bound on h'/h.}\left|\dfrac{h'(z_0)}{h(z_0)}\right|<\sum_{x\in Z_h}\dfrac{1}{\epsilon/8(n-k)}+\sum_{y\in P_h}\dfrac{1}{\epsilon/8(n-k)}=\left(|Z_h|+|P_h|\right)\dfrac{8(n-k)}{\epsilon}\leq\dfrac{8(n-k)^2}{\epsilon}.\end{equation} Let us now turn our attention to a lower bound for $\left|\dfrac{g'(z_0)}{g(z_0)}\right|$. Let $w_0$ denote the point in $K$ which is closest to $z_0$. Since $\gamma\subset E$, we have $\epsilon/2<|z_0-w_0|<\epsilon$. After applying the appropriate distance preserving affine transformation to the plane, we may assume without loss of generality that $z_0=0$ and that $w_0$ lies in the real line, in the interval $(-\epsilon,-\epsilon/2)$. The choice of $w_0$ and the convexity of $K$ implies that $K$ lies in the left half of the circle centered at $w_0$ with radius $s$. Let $F$ denote this left half-circle. Let $Z_g\subset F$ denote the set of zeros of $g$. Then we have $|Z_g|\geq k$ (recall that $g$ has no poles), and similarly as in Equation~\ref{eqn: Breakdown of h'/h.}, we have \begin{equation}\label{eqn: Breakdown of g'/g.} \dfrac{g'(z_0)}{g(z_0)}=\dfrac{g'(0)}{g(0)}=\displaystyle\sum_{x\in Z_g}\dfrac{1}{-x}. \end{equation} Now, for any point $z\in \mathbb{C}$, $|z|\geq\Re(z)$, and if $z\in F$, we have $|z|\leq s+\epsilon$ and $\Re(-z)\geq\epsilon/2$, so taking absolute values in Equation~\ref{eqn: Breakdown of g'/g.}, we obtain \begin{equation}\label{eqn: First estimate on g'/g.} \left|\dfrac{g'(0)}{g(0)}\right|\geq\Re\left(\sum_{x\in Z_g}\dfrac{1}{-x}\right)=\sum_{x\in Z_g}\Re\left(\dfrac{1}{-x}\right)=\sum_{x\in Z_g}\dfrac{\Re(-x)}{|-x|^2}\geq\sum_{x\in Z_g}\dfrac{\epsilon/2}{(s+\epsilon)^2}\geq\dfrac{k\epsilon}{2(s+\epsilon)^2}. \end{equation} Since the choice of $z_0$ was arbitrary, Equations~\ref{eqn: Bound on h'/h.} and~\ref{eqn: First estimate on g'/g.} give us that on $\gamma$, \begin{equation}\label{eqn: Bounds on gamma.} \left|\dfrac{h'}{h}\right|<\dfrac{8(n-k)^2}{\epsilon}\text{ and }\dfrac{k\epsilon}{2(s+\epsilon)^2}<\left|\dfrac{g'}{g}\right|. \end{equation} Thus if \begin{equation}\label{eqn: Constants needed.}\dfrac{8(n-k)^2}{\epsilon}<\dfrac{k\epsilon}{2(s+\epsilon)^2},\end{equation} then on $\gamma$, $\left|h'/h\right|<\left|g'/g\right|$. Let $\Omega$ denote the bounded face of $\gamma$. According to Rouch\'e's theorem, we could then conclude that \begin{equation}\label{eqn: Rouche set-up.} \#_z\left(\dfrac{g'}{g},\Omega\right)-\#_p\left(\dfrac{g'}{g},\Omega\right)=\#_z\left(\dfrac{g'}{g}+\dfrac{h'}{h},\Omega\right)-\#_p\left(\dfrac{g'}{g}+\dfrac{h'}{h},\Omega\right).\end{equation} Since $g$ is a polynomial with all distinct zeros, and all of the zeros and critical points of $g$ lie in $K$, Lemma~\ref{lem: Lemma.} then gives us \begin{equation}\label{eqn: Equation g.} \#_z\left(\dfrac{g'}{g},\Omega\right)-\#_p\left(\dfrac{g'}{g},\Omega\right)=\#_c(g,\Omega)-(\#_z(g,\Omega)+\#_p(g,\Omega)). \end{equation} On the other hand, through the magic of the product rule, $$\dfrac{g'}{g}+\dfrac{h'}{h}=\dfrac{(gh)'}{gh},$$ so again by Lemma~\ref{lem: Lemma.}, we would have \begin{equation}\label{eqn: Equation g+h.} \#_z\left(\dfrac{g'}{g}+\dfrac{h'}{h},\Omega\right)-\#_p\left(\dfrac{g'}{g}+\dfrac{h'}{h},\Omega\right)=\#_c(gh,\Omega)-(\#_z(gh,\Omega)+\#_p(gh,\Omega)). \end{equation} Performing the substitutions indicated by Equations~\ref{eqn: Equation g.} and~\ref{eqn: Equation g+h.}, Equation~\ref{eqn: Rouche set-up.} becomes \begin{equation}\label{eqn: After substitutions.} \#_c(g,\Omega)-(\#_z(g,\Omega)+\#_p(g,\Omega))=\#_c(gh,\Omega)-(\#_z(gh,\Omega)+\#_p(gh,\Omega)). \end{equation} Solving Equation~\ref{eqn: After substitutions.} for $\#_c\left(gh,\Omega\right)$, and using the fact that $\Omega\subset K_\epsilon$, we obtain \begin{equation}\label{eqn: Critical points for K_epsilon.} \#_c(gh,K_\epsilon)\geq\#_c(gh,\Omega)=\#_c(g,K)+(\#_z(gh,\Omega)+\#_p(gh,\Omega))-(\#_z(g,K)+\#_p(g,K)). \end{equation} Since $\#_z(gh,\Omega)-\#_z(g,\Omega)=\#_z(h,\Omega)$, and $\#_p(gh,\Omega)-\#_p(g,\Omega)=\#_p(h,\Omega)$, Equation~\ref{eqn: Critical points for K_epsilon.} provides us with our desired inequality: \begin{equation}\label{eqn: Final inequality for critical points of gh in K_epsilon.} \#_c(gh,K_\epsilon)\geq\#_c(g,K)+(\#_z(h,\Omega)+\#_p(h,\Omega))\geq\#_c(g,K)\geq k-1. \end{equation} Recall that Equation~\ref{eqn: Final inequality for critical points of gh in K_epsilon.} was obtained subject to the assumption that $$\dfrac{8(n-k)^2}{\epsilon}<\dfrac{k\epsilon}{2(s+\epsilon)^2}.$$ This inequality may easily be rearranged into the first inequality found in the statement of Theorem~\ref{thm: Main inequalities.}. If we view $n$ and $k$ as fixed, and solve this inequality for $\epsilon$, we obtain as a sufficient condition for $\mathcal{RAGL}(n,k,\epsilon)$ the inequality \begin{equation}\label{eqn: n and k fixed inequality.} \epsilon>\dfrac{4s(n-k)}{\sqrt{k}-4(n-k)}-1. \end{equation} This is the conclusion of Corollary~\ref{cor: n and k held constant, solve for epsilon.}. Note that in solving for $\epsilon$ to obtain the above inequality, the assumption was made that $\sqrt{k}-4(n-k)>0$. We now turn our attention to Corollary~\ref{cor: Our bound on psi_1.}, and the special case where $K$ is a disk with diameter $s$. In that case, our earlier work still holds, but we are able to sharpen our estimate on $\left|\dfrac{g'}{g}\right|$. We again arrive at the situation where we assume that $z_0\in\gamma$ equals $0$, and that the closest point to $z_0$ in $K$ is $w_0\in(-\epsilon,-\epsilon/2)$ (so that $K$ is the disk with diameter $s$, centered at the real number $w_0-s/2$). For any $w\in K$, with $x=\Re(w)$ and $y=\Im(w)$, $$\Re\left(\dfrac{1}{-w}\right)=\dfrac{-x}{x^2+y^2}.$$ Using the techniques of undergraduate multi-variable calculus, we find that this function of $x$ and $y$ is minimized on $K$ at the furthest left point of the disk, namely $w_0-s$. Thus we have that for $w\in K$, $$\Re\left(\dfrac{1}{-w}\right)\geq\dfrac{1}{-w_0+s}\geq\dfrac{1}{\epsilon+s}.$$ Adopting $1/(\epsilon+s)$ as our lower bound for $\Re(1/-w)$ (for each zero $w$ of $g$), Equation~\ref{eqn: First estimate on g'/g.} becomes \begin{equation}\label{eqn: Inequality for g'(0)/g(0) in disk case.} \left|\dfrac{g'(0)}{g(0)}\right|\geq\dfrac{k}{\epsilon+s}. \end{equation} Recall that this then implies that $$\left|\dfrac{g'}{g}\right|\geq\dfrac{k}{\epsilon+s}$$ on $\gamma$. Thus the sufficient condition for $\mathcal{RAGL}(n,k,\epsilon)$ which we obtain in this case is $$\dfrac{8(n-k)^2}{\epsilon}<\dfrac{k}{\epsilon+s},$$ which, after some simple arithmetic, provides the second conclusion of Theorem~\ref{thm: Main inequalities.}. If we view $n$ and $k$ as fixed, and solve this inequality for $\epsilon$ (this arithmetic requires the assumption that $k>8(n-k)^2$), we obtain the inequality $$\epsilon>\dfrac{8s(n-k)^2}{k-8(n-k)^2}.$$ From this inequality, we derive the result of Corollary~\ref{cor: Our bound on psi_1.} that if $k>8(n-k)^2$, then $$\psi_1(n,k)\leq\dfrac{8(n-k)^2}{k-8(n-k)^2}.$$ \section{AN ADDITIONAL CONJECTURE}\label{sect: An additional conjecture.} The hueristic which underlies our proof of Theorem~\ref{thm: Main inequalities.} is that when our rational function $f$ is factored as $f=g\cdot h$ (where $g$ gets all the zeros of $f$ lying in $K$, and $h$ gets all remaining zeros and poles of $f$), then the relatively few zeros and poles of $h$ cannot drag the critical points of $g$ very far outside of $K$. This hueristic seems to accommodate perfectly well the possibility of $K$ being unbounded (this possibility is present in the classical Gauss--Lucas theorem as well), and of $g$ being a rational function (now possessing all of the zeros \textbf{and} poles of $f$ which lie in $K$). Of course, if $g$ is a rational function, its critical points need not in general lie in $K$, so that must be added as an assumption. We therefore extend Conjecture~\ref{conj: epsilon fixed goal.} (with the function $f$ already factored as $f=g\cdot h$) as follows. \begin{conjecture}\label{conj: Allow unbounded and poles.} Let $K\subset\mathbb{C}$ be convex, and let $\epsilon>0$ be given. There exists some constant $C>0$ for which the following holds. Let $g,h\in\mathbb{C}(z)$ be rational functions, and assume that all zeros, poles, and critical points of $g$ lie in $K$. If $\dfrac{\#_z(g,\mathbb{C})+\#_p(g,\mathbb{C})}{\#_z(h,\mathbb{C})+\#_p(h,\mathbb{C})}>C$, then $\#_c(gh,K_\epsilon)\geq\#_c(g,K)$. \end{conjecture}
{ "timestamp": "2017-06-20T02:01:03", "yymm": "1706", "arxiv_id": "1706.05410", "language": "en", "url": "https://arxiv.org/abs/1706.05410", "abstract": "The Gauss--Lucas theorem states that any convex set $K\\subset\\mathbb{C}$ which contains all $n$ zeros of a degree $n$ polynomial $p\\in\\mathbb{C}[z]$ must also contain all $n-1$ critical points of $p$. In this paper we explore the following question: for which choices of positive integers $n$ and $k$, and positive real number $\\epsilon$, will it follow that for every degree $n$ polynomial $p$ with at least $k$ zeros lying in $K$, $p$ will have at least $k-1$ critical points lying in the $\\epsilon$-neighborhood of $K$. We supply an inequality relating $n$, $k$, and $\\epsilon$ which, when satisfied, guarantees a positive answer to the above question.", "subjects": "Complex Variables (math.CV)", "title": "On approximate Gauss-Lucas theorems", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.990291523518526, "lm_q2_score": 0.8267117855317474, "lm_q1q2_score": 0.818685673604955 }
https://arxiv.org/abs/1209.5185
Bounds on Characteristic Polynomials
Suppose $G$ is a simple graph with $n$ vertices, $m$ edges, and rank $r$. Let $\chi_G(t)=a_0t^n-a_1t^{n-1}+\cdots +(-1)^ra_rt^{n-r}$ be the chromatic polynomial of $G$. For $q,k\in \Bbb{Z}$ and $0\le k\le q+r+1$, we obtain a sharp two-side bound for the partial binomial sum of the coefficient sequence, that is, \[ {r+q\choose k}\le \sum_{i=0}^{k}{q\choose k-i}a_{i}\le {m+q\choose k}. \] Indeed, this bound holds for the characteristic polynomial of hyperplane arrangements and matroids, and its weak version can be generalized to the characteristic polynomial of toric arrangements and arithmetic matroids. We also propose a problem on the geometric interpretation of the above bound.
\section{Introduction} We start with some notations in graph theory. Let $G=(VG, EG)$ be a simple graph (no loops and multi-edges) with the vertex set $VG$ and the edge set $EG$. Let $n=|VG|$, $m=|EG|$, and $c$ the number of connected components of $G$. Then the rank of $G$ is $r=n-c$. First appeared in \cite{Birkhoff}, the chromatic polynomial $\chi_G(t)$ counts the number of proper colorings of the graph $G$ with $t$ colors, which can be written as follows, \[\chi_G(t)=a_0t^n-a_1t^{n-1}+\cdots +(-1)^ra_rt^{n-r}.\] The chromatic polynomial is one of the most central topics in graph theory, whose coefficients are mysterious and have caught many mathematicians' interests. In 1932, Whitney \cite{Whitney2} showed that the coefficient sequence is sign-alternating, i.e., $a_i>0$. Moreover, he \cite{Whitney1} gave a combinatorial interpretation to each coefficient $a_i$, which is equal to the number of those $i$-subsets of $EG$ that contain no broken circuits, known as the \emph{no broken circuit} theorem. In 1968, Read \cite{Read} asked which polynomial is the chromatic polynomial of some graph and conjectured that the sequence $a_0,a_1,\ldots,a_r$ is unimodal. In general, it looks impossible to give all properties of the coefficient sequence. Fortunately, Read's conjecture has been positively answered by Huh \cite{Huh} presently. It turns out that the coefficient sequence $a_0,a_1,\ldots,a_r$ is logconcave. All above results are relatively big steps in the way investigating the properties of the coefficient sequence. Few results have been obtained for the coefficient sequence. After searching from the web, the next is the only general result founded on the bounds of the coefficient sequence which, in fact, can be regarded as a consequence of Whitney's combinatorial interpretation. In 1970, G.H.J. Meredith \cite{Meredith} gave an upper bound for each coefficient, which is $|a_i|\le {m\choose i}$. In this paper, we shall introduce a new bound which will generalize Whitney's sign-alternating result and Meredith's upper bound result. Next is the statement of our main result. If $q,k\in \Bbb{Z}$ with $0\le k\le q+r+1$, we have \[ {r+q\choose k}\le \sum_{i=0}^{k}{q\choose k-i}a_{i}\le {m+q\choose k}. \] We can see that Whitney's sign-alternating theorem and Meredith's upper bound theorem are direct consequences of the above inequality. When $q=0$, we have \[ {r\choose k}\le a_{k}\le {m\choose k},\sp \If 0\le k\le r. \] If $q=-1$, we obtain that \[ {r-1\choose k}\le (-1)^k\sum_{i=0}^{k}(-1)^ia_{i}\le {m-1\choose k}, \sp\If 0\le k\le r. \] So the first $r-1$ partial sums of the coefficient sequence of the chromatic polynomial are still sign-alternating. Indeed, all above results hold for a more generalized object, the characteristic polynomial of hyperplane arrangements. Hence, we shall practise our proof on hyperplane arrangements in the next section. \section{Main Results} An $n$-dimensional arrangement $\mathcal{A}$ of hyperplanes is a finite collection of codimension one subspaces in an $n$-dimensional vector space $V$. Equipped with the partial order defined by the inverse of set inclusion, the set of all nonempty intersections of hyperplanes in $\mathcal{A}$ including the ambient space $V:=\cap_{H\in \emptyset} H$ forms a semi-lattice $L(\mathcal{A})$, called the intersection semi-lattice, i.e., \[ L(\mathcal{A})=\{\cap_{H\in \mathcal{B}}H\mid \mathcal{B}\subseteq \mathcal{A}\}, \] Note that the minimal element of $L(\mathcal{A})$ is $V$. The maximal rank of the semi-lattice $L(\mathcal{A})$ is called the \emph{rank} of hyperplane arrangement $\mathcal{A}$, denoted by $r(\mathcal{A})$. In another viewpoint, the rank $r(\mathcal{A})$ is the dimension of the vector space spanned by those normal vectors of hyperplanes in $\mathcal{A}$. The \emph{characteristic polynomial} $\chi(\mathcal{A},t)\in \Bbb{C}[t]$ of $\mathcal{A}$ is defined to be \[ \chi(\mathcal{A};t):=\sum_{X\in L(\mathcal{A})}\mu(\hat{0},X)\,t^{{\small\dim(X)}}. \] where $\mu$ is the {\emph M\"{o}bius function} of $L(\mathcal{A})$. Let $G=(VG,EG)$ be a simple graph with the vertex set $VG=[n]$ and the edge set $EG\subseteq [n]\times [n]$. The \emph{graphic arrangement} $\mathcal{A}_G$ of $G$ is an $n$-dimensional arrangement of $|EG|$ hyperplanes whose members are given by \[ H_{ij}: x_i=x_j,\sp \For (i,j)\in EG. \] With these definitions, we have, see Theorem 2.7 in \cite{Stanley2}, \[ \chi(\mathcal{A}_G;t)=\chi_G(t). \] It follows that the rank of the graph $G$ is indeed the same as the rank of the graphic arrangement $\mathcal{A}_G$. It is well known that the characteristic polynomial satisfies the \emph{deletion-contraction} recurrence \[\chi(\mathcal{A};t)=\chi(\mathcal{A}\setminus H_0;t)-\chi(\mathcal{A}/H_0;t),\] where $H_0\in \mathcal{A}$ is a fixed hyperplane, $\mathcal{A}\setminus H_0$ is an $n$-dimensional subarrangement of hyperplanes in $V$ obtained by removing $H_0$ from $\mathcal{A}$, and $\mathcal{A}/H_0$ is an $(n-1)$-dimensional hyperplane arrangement in $H_0$ whose members are those restrictions of all hyperplanes of $\mathcal{A}\setminus H_0$ on $H_0$, i.e., \[ \mathcal{A}\setminus H_0=\mathcal{A}\setminus\{H_0\}, \sp \mathcal{A}/H_0=\{H\cap H_0\mid H\in \mathcal{A}\setminus H_0\}. \] A hyperplane arrangement is called \emph{central} if $\cap_{H\in \mathcal{A}}H\neq \emptyset$. We have $r(\mathcal{A})\le |\mathcal{A}|$ in general and call $\mathcal{A}$ \emph{boolean} when $r(\mathcal{A})=|\mathcal{A}|$. It is easy to show that the boolean hyperplane arrangement is central and its intersection semi-lattice is isomorphic to the boolean lattice $(2^\mathcal{A},\subseteq)$. Hence the characteristic polynomial of an $n$-dimensional boolean arrangement $\mathcal{A}$ of $m$ hyperplanes is \begin{equation}\label{boolean-formula} \chi(\mathcal{A};t)=t^{n-m}(t-1)^m=\sum_{i=0}^m(-1)^i{m\choose i}t^{n-i}. \end{equation} The graphic arrangement $\mathcal{A}_G$ is boolean if and only if the graph $G$ is a forest. Note that a subset $\mathcal{B}$ of $\mathcal{A}$ naturally defines a subarrangement of hyperplanes in the same ambient space as $\mathcal{A}$, still denoted $\mathcal{B}$ by abuse of notations. A hyperplane arrangement $\mathcal{A}$ of rank $r$ is called \emph{in general position} if the subarrangement $\mathcal{B}$ of $\mathcal{A}$ is boolean whenever $|\mathcal{B}|\le r$, or not central otherwise. If a central hyperplane arrangement is in general position if and only if it is boolean. From \cite[Proposition 2.4]{Stanley2}, the characteristic polynomial of an $n$-dimensional arrangement $\mathcal{A}$ of $m$ hyperplanes in general position is \begin{equation}\label{formula-general-position} \chi(\mathcal{A};t)=t^n-mt^{n-1}+{m\choose 2}t^{n-2}+\cdots+(-1)^r{m\choose r}t^{n-r}. \end{equation} Later, we shall prove in Proposition \ref{proposition-general-position} that the converse of the above statement is still true by using no broken circuit theorem. First we need some preparations to state no broken circuit theorem. A subset $\mathcal{B}$ of the hyperplane arrangement $\mathcal{A}$ is called \emph{dependent} if $\cap_{H\in\mathcal{B}}H\neq \emptyset$ and $r(\cap_{H\in\mathcal{B}}H)<|\mathcal{B}|$, i.e., the subarrangement $\mathcal{B}$ is central but not boolean. Let $\mathcal{A}$ be totally ordered under a given order $\prec$. A subset of $\mathcal{A}$ is called a \emph{circuit} if it is a minimal dependent subset of $\mathcal{A}$. It is obvious that each dependent subset of $\mathcal{A}$ contains at least a circuit. A \emph{broken circuit} is a subset of $\mathcal{A}$ obtained by removing the maximal element from a circuit of $\mathcal{A}$. A subset $\mathcal{B}$ of $\mathcal{A}$ is called \emph{$\chi$-independent} if $\cap_{H\in\mathcal{B}}H\neq \emptyset$ and $\mathcal{B}$ contains no broken circuits. \begin{theorem}[{\bf No Broken Circuit Theorem} \cite{Orlik}]\label{theorem-BC} Let $\mathcal{A}$ be an $n$-dimensional hyperplane arrangement of rank $r$ and its characteristic polynomial \[ \chi(\mathcal{A};t)=a_0t^n-a_1t^{n-1}+\cdots +(-1)^ra_rt^{n-r}. \] Then for $0\le k\le r$, $a_k$ is equal to the number of $\chi$-indenpendent $k$-subsets of $\mathcal{A}$. \end{theorem} \begin{proposition}\label{proposition-general-position}Let $\mathcal{A}$ be an $n$-dimensional arrangement of $m$ hyperplanes and $r(\mathcal{A})=r$. Then $\mathcal{A}$ is in general position if and only if its characteristic polynomial is exactly the same as $(\ref{formula-general-position})$. \end{proposition} \begin{proof}Note that no subset of $\mathcal{A}$ is a circuit if $\mathcal{A}$ is in general position. Then every $k$-subset of $\mathcal{A}$ is $\chi$-independent for all $k\le r$. By Theorem \ref{theorem-BC}, we have $a_i={m\choose i}$. Conversely, for $0\le k \le r$, $a_k={m\choose k}$ implies that all $k$-subsets of $\mathcal{A}$ are $\chi$-independent. It follows from the definition that if $\mathcal{B}\subseteq \mathcal{A}$ and $|\mathcal{B}|\le r$, $\mathcal{B}$ contains no broken circuits and $\cap_{H\in\mathcal{B}}H\neq \emptyset$. So the subarrangement $\mathcal{B}$ is boolean if $|\mathcal{B}|\le r$. When $k> r$, we have $a_k=0$, which implies $\cap_{H\in\mathcal{B}}H=\emptyset$ if $|\mathcal{B}|\ge r$. \end{proof} To prove our main result, we introduce a combinatorial identity first, that is, \begin{equation}\label{formula-combinatorial-identity} \sum_{i=0}^{k}{x\choose i}{y\choose k-i}={x+y\choose k},\sp \If~~ x,y\in \Bbb{C}, ~~ k\in \Bbb{Z}_{\ge 0}, \end{equation} where ${x\choose i}=\frac{1}{i!}x(x-1)\cdots(x-i+1)$. Indeed, in the case that $x\in \Bbb{C}$ is fixed and $y$ is an arbitrary nonnegative integer, we present a brief proof for $(\ref{formula-combinatorial-identity})$ by induction on $y\in\Bbb{Z}_{\ge 0}$. First, the induction basis $y=0$ is trivial. With the induction hypothesis, we have, \begin{eqnarray*} \sum_{i=0}^{k}{x\choose i}{y+1\choose k-i}&=&\sum_{i=0}^{k}{x\choose i}{y\choose k-i}+\sum_{i=0}^{k-1}{x\choose i}{y\choose k-1-i}\\ &=&{x+y\choose k}+{x+y\choose k-1}={x+y+1\choose k}, \end{eqnarray*} Notice that $(\ref{formula-combinatorial-identity})$ can be viewed as a polynomial equation in $y$ whenever $x$ is fixed. Then each $y\in \Bbb{Z}_{\ge 0}$ is a root of this polynomial equation. By the fundamental theorem of algebra, $(\ref{formula-combinatorial-identity})$ holds for all $y\in \Bbb{C}$ whenever $x\in \Bbb{C}$ is fixed. \begin{theorem}\label{theorem-main} Let $\mathcal{A}$ be an $n$-dimensional arrangement of $m$ hyperplanes and its characteristic polynomial \[ \chi(\mathcal{A};t)=a_0t^n-a_1t^{n-1}+a_2t^{n-2}+\cdots+(-1)^ra_rt^{n-r}, \] where $r=r(\mathcal{A})$. If $q,k\in \Bbb{Z}$ satisfies $0\le k\le q+r+1$, then \begin{equation}\label{formula-main} {r+q\choose k}\le \sum_{i=0}^{k}{q\choose k-i}a_{i}\le {m+q\choose k}. \end{equation} \end{theorem} \begin{proof}First if $\mathcal{A}$ is boolean, then $r=m$. From (\ref{boolean-formula}) and (\ref{formula-combinatorial-identity}), we have \begin{equation}\label{boolean-case} \sum_{i=0}^{k}{q\choose k-i}a_{i}=\sum_{i=0}^{k}{q\choose k-i}{m\choose i}={m+q\choose k}={r+q\choose k}. \end{equation} So if $\mathcal{A}$ is boolean, (\ref{formula-main}) holds for any $q,k\in \Bbb{Z}$. In general, we shall use induction on $|\mathcal{A}|$ to prove (\ref{formula-main}). Note that if $|\mathcal{A}|=0$ or $1$, $\mathcal{A}$ is a boolean arrangement. Suppose the result holds for $|\mathcal{A}|\le m$. Since (\ref{formula-main}) holds for any boolean hyperplane arrangement, it is enough to prove the result for the case that $|\mathcal{A}|=m+1$ and $\mathcal{A}$ is not boolean. In this case, we have $r=r(\mathcal{A})< |\mathcal{A}|=m+1$, that is to say, the space spanned by the $m+1$ normal vectors of hyperplanes in $\mathcal{A}$ has dimension $r<m+1$. So at least one of these $m+1$ normals can be removed without changing the spanning space. In another word, there is a hyperplane $H_0\in \mathcal{A}$ such that $r(\mathcal{A}\setminus H_0)=r(\mathcal{A})=r$. Then we can write \[ \chi(\mathcal{A}\setminus H_0;t)=b_0t^n-b_1t^{n-1}+b_2t^{n-2}+\cdots+(-1)^rb_rt^{n-r}. \] Notice that each maximal element in the intersection semi-lattice $L(\mathcal{A}/H_0)$ is a maximal element of the intersection semi-lattice $L(\mathcal{A})$. Since all maximal elements of $L(\mathcal{A})$ have the same rank $r$, it follows that the rank of $\mathcal{A}/H_0$ is $r-1$, i.e., $r(\mathcal{A}/H_0)=r-1$. Then we can write \[ \chi(\mathcal{A}/H_0;t)=c_0t^{n-1}-c_1t^{n-2}+c_2t^{n-3}+ \cdots+(-1)^{r-1}c_{r-1}t^{n-r}. \] Since $|\mathcal{A}\setminus H_0|= m$ and $r(\mathcal{A}\setminus H_0)= r$, the induction hypothesis implies that, if $0\le k\le q+r+1$, \begin{equation}\label{deletion-case} {r+q\choose k}~\le ~\sum_{i=0}^{k}{q\choose k-i}b_{i}~\le~{m+q\choose k}. \end{equation} Since $|\mathcal{A}/H_0|\le m$ and $r(\mathcal{A}/ H_0)= r-1$, the induction hypothesis implies that, if $0\le k-1\le q+r$, i.e., $1\le k\le q+r+1$, \begin{equation}\label{restriction-case} {r-1+q\choose k-1}~\le ~\sum_{i=0}^{k-1}{q\choose k-1-i}c_{i}=\sum_{i=1}^{k}{q\choose k-i}c_{i-1} ~\le~ {|\mathcal{A}/H_0|+q\choose k-1}. \end{equation} Using the deletion-contraction recurrence $\chi(\mathcal{A};t)=\chi(\mathcal{A}\setminus H_0;t)-\chi(\mathcal{A}/H_0;t)$, we have \[ a_0=b_0=1,\sp a_i=b_i+c_{i-1} ~\If~ 1\le i\le r. \] It then follows by combining with (\ref{deletion-case}) and (\ref{restriction-case}) that if $1\le k\le r+q+1$ \begin{equation}\label{combining} {r+q\choose k}+{r-1+q\choose k-1}~\le ~\sum_{i=0}^{k}{q\choose k-i}a_{i}~\le~{m+q\choose k}+{|\mathcal{A}/H_0|+q\choose k-1}. \end{equation} Since (\ref{formula-main}) is obviously true when $k=0$, it remains to show that, if $1\le k\le r+q+1$, \begin{eqnarray} {r+q\choose k}+{r-1+q\choose k-1}&\ge& {r+q\choose k},\label{inequality-r}\\ {m+q\choose k}+{|\mathcal{A}/H_0|+q\choose k-1}&\le& {m+q+1\choose k}\label{inequality-m}. \end{eqnarray} Note that ${r-1+q\choose k-1}\ge 0$ if $r-1+q\ge 0$. However if $r-1+q< 0$, then $1\le k \le r+q+1<2$ implies $k=1$ and ${r-1+q\choose k-1}=1\ge 0$. It completes (\ref{inequality-r}). Since ${m+q\choose k}+{m+q\choose k-1}= {m+q+1\choose k}$ and $|\mathcal{A}/H_0|\le m$, (\ref{inequality-m}) is obvious when $|\mathcal{A}/H_0|+q\ge 0$. Now consider the case $|\mathcal{A}/H_0|+q<0$. Since $r(\mathcal{A}/H_0)=r-1$, then we have $|\mathcal{A}/H_0|\ge r-1$. Combining with $1\le k\le r+q+1$, we obtain that $k=1$. So ${|\mathcal{A}/H_0|+q\choose k-1}={m+q\choose k-1}$ if $|\mathcal{A}/H_0|+q<0$, which completes (\ref{inequality-m}). \end{proof} By taking $q=0$ in (\ref{formula-main}), Whitney's sign-alternating theorem and Meredith's upper bound theorem become direct consequences of Theorem \ref{theorem-main}. \begin{corollary}\label{corollary-3} Under assumptions of {\rm Theorem \ref{theorem-main}}, we have \[{r\choose k}\le a_{k}\le {m\choose k}\sp ~\If~ 0\le k \le r.\] \end{corollary} Since ${-1 \choose k-i}=(-1)^{k-i}$, after taking $q=-1$ in (\ref{formula-main}), we shall obtain two-side bounds for the partial sums of the coefficient sequence. \begin{corollary}\label{corollary-4} Under assumptions of {\rm Theorem \ref{theorem-main}}, we have \[{r-1\choose k}\le (-1)^k\sum_{i=0}^{k}(-1)^ia_{i}\le {m-1\choose k}\sp ~\If~ 0\le k \le r.\] \end{corollary} When $k\le r-1$, we have $(-1)^k\sum_{i=0}^{k}(-1)^ia_{i}\ge {r-1\choose k}\ge 1$, that is to say, the first $r-1$ partial sums of the coefficient sequence form a sign-alternating sequence. \begin{corollary}\label{corollary-5} Under assumptions of {\rm Theorem \ref{theorem-main}}, we have \begin{equation}\label{formula-new} (-1)^k\sum_{i=0}^{k}(-1)^i{r-i\choose k-i}a_{i}\ge 0\sp . \end{equation} In particular, \begin{equation}\label{bound-new} a_2\ge {r\choose 2}+(m-r)(r-1)\sp \And \sp a_3\ge {r\choose 3}+(m-r){r-1\choose 2}. \end{equation} \end{corollary} \begin{proof}Taking $q=k-r-1$ in (\ref{formula-main}), we have \[\sum_{i=0}^{k}{k-r-1\choose k-i}a_{i}\ge {k-1\choose k}=0.\] Notice that ${k-r-1\choose k-i}=(-1)^{k-i}{r-i\choose k-i}$ Then (\ref{formula-new}) is obvious. It is well known from the definition of the characteristic polynomial that $a_1=|\mathcal{A}|=m$. Applying $k=2$ to (\ref{formula-new}), we have \[ {r\choose 2}-(r-1)m+a_2\ge 0, \] which gives the lower bound ${r\choose 2}+(m-r)(r-1)$ of $a_2$ as above. Applying $k=3$ to (\ref{formula-new}), we have \[ -{r\choose 3}+{r-1\choose 2}m-(r-2)a_2+a_3\ge 0. \] Since $a_2\ge {r\choose 2}+(m-r)(r-1)$, it follows that \begin{eqnarray*} a_3&\ge& {r\choose 3}-{r-1\choose 2}m-(r-2)(m-\frac{r}{2})(r-1)\\ &=&{r\choose 3}+(m-r){r-1\choose 2}. \end{eqnarray*} \end{proof} Since the chromatic polynomial $\chi_G(t)$ of a graph $G$ is the characteristic polynomial of the graphic arrangement $\mathcal{A}_G$, the two-side bound $(\ref{formula-main})$ holds for the coefficient sequence of $\chi_G(t)$. \begin{theorem}\label{theorem-chromatic} Let $\chi_G(t)=t^n-a_1t^{n-1}+\cdots+(-1)^ra_rt^{n-r}$ be the chromatic polynomial of a graph $G$ with $n$ vertices, $m$ edges, and rank $r$. Then the following three statements are equivalent, \begin{enumerate}[label={\rm (\roman*)}] \item\label{theorem-i} $a_k={m\choose k}$ for all $k$ with $1\le k\le r;$ \item\label{theorem-ii} $a_k={r\choose k}$ for all $k$ with $1\le k\le r;$ \item\label{theorem-iii} $G$ is a forest, i.e., $m=r$. \end{enumerate} \end{theorem} \begin{proof}$\ref{theorem-iii}\Rightarrow\ref{theorem-i}$ and $\ref{theorem-iii}\Rightarrow\ref{theorem-ii}$ are easy consequences of Corollary \ref{corollary-3}. Recall $a_1=|EG|=m$, then $\ref{theorem-ii}\Rightarrow\ref{theorem-iii}$ becomes obvious. From Proposition \ref{proposition-general-position}, $a_k={m\choose k}$ if and only if the graphic arrangement $\mathcal{A}_G$ is in general position. Note that the graphic arrangement $\mathcal{A}_G$ is a central hyperplane arrangement. Then $\mathcal{A}_G$ contains no subset of size larger than $r$. Hence we have $|\mathcal{A}_G|=|EG|=m=r$ which proves $\ref{theorem-i}\Rightarrow\ref{theorem-iii}$. \end{proof} \section{Discussions and Problems} Let $\mathcal{A}$ be an $n$-dimensional arrangement of $m$ hyperplanes and rank $r$. If all hyperplanes of $\mathcal{A}$ are restricted onto the $r$-dimensional subspace spanned by their normal vectors, we shall obtain an $r$-dimensional arrangement of $m$ hyperplanes and rank $r$, whose characteristic polynomial has the same coefficient sequence as $\chi(\mathcal{A};t)$. Under this restriction, the characteristic polynomial becomes $t$-free, i.e, containing no $t$ as a factor. Next we always assume the characteristic polynomial of $\mathcal{A}$ is \begin{equation}\label{formula-chi} \chi(\mathcal{A};t)=a_0t^r-a_1t^{r-1}+\cdots+(-1)^ra_r. \end{equation} Recall the no broken circuit Theorem \ref{theorem-BC} that $a_k$ counts the number of $\chi$-independent $k$-subsets of $\mathcal{A}$. It is then obvious that $a_k\le {m\choose k}$. On the other hand, $a_r\ne 0$ implies that there exists at least one $\chi$-independent $r$-subset $\mathcal{B}$ of $\mathcal{A}$. Note the fact from the definition that any subset of a $\chi$-independent set is still $\chi$-independent. Then all subsets of $\mathcal{B}$ are $\chi$-independent, which implies $a_k\ge {r\choose k}$ by the no broken circuit Theorem \ref{theorem-BC}. In this sense, the inequality ${r\choose k}\le a_k\le {m\choose k}$ of Corollary \ref{corollary-3} can be easily obtained from the no broken circuit theorem. Since theorem \ref{theorem-main} can be regarded as a generalization of Corollary \ref{corollary-3}, a natural problem is as follows, \begin{itemize} \item Can (\ref{formula-main}) be interpreted by the no broken circuit theorem? \end{itemize} Define an operation $\D$ of the characteristic polynomial $\chi(\mathcal{A};t)$ to be \[\D \chi(\mathcal{A};t)=\frac{\chi(\mathcal{A};t)-\chi(\mathcal{A};1)}{t-1}.\] Denote $\D^0$ the identity map and $\D^i \chi(\mathcal{A};t)=\underbrace{\D\circ\cdots\circ\D}_i \chi(\mathcal{A};t)$ for $i\in \Bbb{N}$. Note that \begin{eqnarray*} \chi(\mathcal{A};t)-\chi(\mathcal{A};1)&=&a_0(t^r-1)-a_1(t^{r-1}-1)+\cdots+(-1)^{r-1}a_{r-1}(t-1)\\ &=&(t-1)\left(\sum_{k=0}^{r-1}t^{r-1-k}\sum_{i=0}^{k}(-1)^ia_i\right) \end{eqnarray*} Then we have \begin{eqnarray*} \D \chi(\mathcal{A};t)=\sum_{k=0}^{r-1}\left(\sum_{i=0}^{k}(-1)^ia_i\right)t^{r-1-k} \end{eqnarray*} In general, for any non-positive integer $q$, we have \begin{eqnarray*} \D^{-q} \chi(\mathcal{A};t)&=&\sum_{k=0}^{r+q}(-1)^k\left(\sum_{i=0}^{k}{q\choose k-i}a_{i}\right)t^{r+q-k}. \end{eqnarray*} If $\D^q \chi(\mathcal{A};t)$ can be geometrically realized as the characteristic polynomial of an arrangement of $m-q$ hyperplanes and rank $r-q$, from discussions at the beginning of this section, the inequality (\ref{formula-main}) can be easily interpreted by the no broken circuit theorem, which answers the previous question. So our question is reduced to a geometric realization of $\D^{-q} \chi(\mathcal{A};t)$ as the characteristic polynomial of an arrangement of hyperplanes for all $q\in \Bbb{Z}_{\le 0}$. Moreover, it can be further reduced as follows \begin{itemize} \item Is there an arrangement of hyperplanes whose characteristic polynomial is $\D\chi(\mathcal{A};t)$? \end{itemize} Indeed, suppose we can find a geometric realization to $\D \chi(\mathcal{A};t)$ for any $\mathcal{A}$, i.e., $\D\chi(\mathcal{A};t)=\chi(\mathcal{A}_1;t)$ for some hyperplane arrangement $\mathcal{A}_1$. Similarly, we shall have a hyperplane arrangement $\mathcal{A}_2$ such that $\D\chi(\mathcal{A}_1;t)=\chi(\mathcal{A}_2;t)$. Continuing this process, the geometric realization of $\D^q\chi(\mathcal{A};t)$ can be obtained reductively for all $q\in \Bbb{Z}_{\ge 0}$. Hence, the positive answer of the above question combining with the no broken circuit theorem will provide an intuitive interpretation to the inequality (\ref{formula-main}). Fortunately, when $\mathcal{A}$ is central, we can assume all hyperplanes in $\mathcal{A}$ pass through the origin. In this case, we can construct an affine hyperplane arrangement $\d\mathcal{A}$ such that $\chi(\d\mathcal{A};t)=\D\chi(\mathcal{A};t)$. Suppose $\mathcal{A}$ is a linear arrangement of $m+1$ hyperplanes in $\Bbb{R}^n$. Given $K_0\in \mathcal{A}$ with the defining equation $K_0: \sum_{i=1}^n\alpha_ix_i=0$, the \emph{deconing} $\d\mathcal{A}$ of $\mathcal{A}$ is an arrangement of $m$ hyperplanes in the affine space $K_1:\sum_{i=1}^n\alpha_ix_i=1$, which is defined by \[\d\mathcal{A}=\{H\cap K_1\mid H\in \mathcal{A}, H\neq K_0\}.\] Since $\chi(\mathcal{A},1)=0$ when $\mathcal{A}$ is linear, we have \[\chi(\d\mathcal{A};t)=\D\chi(\mathcal{A};t),\] Namely, the deconing construction can geometrically realize $\D\chi(\mathcal{A};t)$ as the characteristic polynomial of the hyperplane arrangement $\d\mathcal{A}$ for the linear hyperplane arrangement $\mathcal{A}$. However, this construction can not be applied directly to affine cases. Hence, to answer the above question, we need to extend the deconing construction an affine hyperplane arrangement $\mathcal{A}$ such that $\chi(\d\mathcal{A};t)=\D\chi(\mathcal{A};t)$.
{ "timestamp": "2014-02-27T02:09:46", "yymm": "1209", "arxiv_id": "1209.5185", "language": "en", "url": "https://arxiv.org/abs/1209.5185", "abstract": "Suppose $G$ is a simple graph with $n$ vertices, $m$ edges, and rank $r$. Let $\\chi_G(t)=a_0t^n-a_1t^{n-1}+\\cdots +(-1)^ra_rt^{n-r}$ be the chromatic polynomial of $G$. For $q,k\\in \\Bbb{Z}$ and $0\\le k\\le q+r+1$, we obtain a sharp two-side bound for the partial binomial sum of the coefficient sequence, that is, \\[ {r+q\\choose k}\\le \\sum_{i=0}^{k}{q\\choose k-i}a_{i}\\le {m+q\\choose k}. \\] Indeed, this bound holds for the characteristic polynomial of hyperplane arrangements and matroids, and its weak version can be generalized to the characteristic polynomial of toric arrangements and arithmetic matroids. We also propose a problem on the geometric interpretation of the above bound.", "subjects": "Combinatorics (math.CO)", "title": "Bounds on Characteristic Polynomials", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9929882031659848, "lm_q2_score": 0.8244619220634456, "lm_q1q2_score": 0.818680962568555 }
https://arxiv.org/abs/2205.01032
Étale extensions of polynomial rings are faithfully flat
We apply Ohi's criterion for faithfully flatness of extensions of commutative rings to prove that any étale extension $k[Y_1, \ldots, Y_n]\subseteq k[X_1, \ldots, X_n]$ of polynomial rings (each in $n$ indeterminates) over a commutative ring $k$ is faithfully flat. In particular, if $k$ is an algebraically closed field then any étale polynomial map $k^{n} \to k^{n}$ is surjective.
\section{Introduction} Faithfully flat extensions abound among ring extensions: any polynomial extension $R\subseteq R[X]$ is faithfully flat, any completion $A\subset\hat{A}$ of a noetherian local ring $A$ is faithfully flat. However, we know little about the following problem: when an \'{e}tale extension $k[\vect Y n]\subseteq k[\vect X n]$ of polynomial rings(each in $n$ indeterminates) over a commutative ring $k$ is faithfully flat?. Let now $k$ be an algebraically closed field. The geometric counterpart of the problem: when an \'{e}tale polynomial map $k^{n} \to k^{n}$ of the affine $n$-space is surjective?, has also been sought, and even in the case of the affine plane over $\mathbb{C}$ it is unknown \cite[p. 415]{Vitushkin75}. In this paper we provide an unconditional solution to the problem. Our main theorem is the following. \begin{theorem}\label{thm_main_theorem} Let $k$ be a commutative ring. Then any \'{e}tale extension $k[\vect Y n]\subseteq k[\vect X n]$ of polynomial rings(each in $n$ indeterminates) over $k$ is faithfully flat. \end{theorem} \begin{corollary}\label{main_corollary} Let $k$ be an algebraically closed field. Then any \'{e}tale polynomial map $k^{n} \to k^{n}$ is surjective. \end{corollary} \begin{proof} See for example \cite[Theorem 46, item ii)]{Wang80}. \end{proof} The proof of our main Theorem \ref{thm_main_theorem} is given in Section \ref{sec_proof}, it is based on a characterization obtained by Ohi \cite{Ohi74} for faithfully flat extensions of commutative rings. It turns out that Ohi's criterion is well suited for \'{e}tale extensions of polynomial rings, in Section \ref{Ohi_result} we recall Ohi's result. In this paper all rings are commutative with an identity element, a ring homomorphism maps the identity to the identity. For a ring $k$, $k^{*}$ denotes the group of units of $k$, and $\operatorname{GL}_{n}\left(k\right)$ denotes the group of $n \times n$ invertible matrices over $k$. \section{Ohi's criterion for faithfully flatness of extensions}\label{Ohi_result} The setup for the next two propositions is as follows. Let $R$ and $S$ be rings such that $R \subseteq S$. Suppose that $S$ is generated as a ring by a set of elements in $S$ over $R$, then we have a presentation of $S$ over $R$, i.e., we have an $R$-isomorphism $S\cong R[\{X_{\lambda}\}_{\lambda \in \Lambda}]/I$, where $\{X_{\lambda}\}_{\lambda \in \Lambda}$ is a set of indeterminates over $R$, $R[\{X_{\lambda}\}_{\lambda \in \Lambda}]$ is the polynomial ring in the indeterminates $X_{\lambda}$ over $R$ and $I$ is the ideal of relations. If $w=r_{0}+r_{1}X^{(1)}+r_{2}X^{(2)}+\cdots+r_{n}X^{(n)}\in R[\{X_{\lambda}\}_{\lambda \in \Lambda}]$, where $r_{0}, r_{1},\,\cdots, \, r_{n} \in R$ and $X^{(i)}$ are monomials in $X_{\lambda}$'s of degree $\geq 1$ such that $X^{(i)}\neq X^{(j)}$ if $i\neq j$, then we define $c(w):=\left(r_{0}, r_{1},\,\cdots, \, r_{n}\right)$ and $c'(w):=\left(r_{1},\,\cdots, \, r_{n}\right)$, that is, the ideals of $R$, generated by $r_{0}, r_{1},\,\cdots, \, r_{n}$ and $r_{1},\,\cdots, \, r_{n}$ respectively. The following criterion for flatness in terms of the ideal $I$ and the content of polynomials $c(w)$ was given by Chase. \begin{proposition}[{\cite[Proposition 2.1]{Chase60}}]\label{Chase_Proposition} Let $R$ and $S$ be rings as in the above setup. Then $S$ is flat over $R$ if and only if $u\in c(u)I$ for each $u \in I$. \end{proposition} Ohi extended Chase's characterization for faithfully flat modules. \begin{proposition}[{\cite[Theorem 1]{Ohi74}}]\label{Ohi_Proposition} Let $R$ and $S$ be rings as in the above setup. Then $S$ is faithfully flat over $R$ if and only if $u\in c'(u)I$ for each $u \in I$. \end{proposition} The proof of Chase's criterion Proposition \ref{Chase_Proposition} relies on the following characterization for flatness of quotients of flat modules that goes back to Cartan-Eilenberg. \begin{lemma}[{\cite[p. 122, Exercise 5]{Cartan-Eilenberg56}}]\label{Cartan-Eilenberg_lemma} Let $0 \to I \to F \to M \to 0 $ be an exact sequence of $R$-modules, where $F$ is a flat $R$-module(e.g., free or projective). The following conditions are equivalent: \begin{itemize} \item[(i)] M is a flat $R$-module. \item[(ii)] For every ideal $A$ of $R$, $AM\cong AF/AI$. \item[(iii)] For every ideal $A$ of $R$, $I\cap AF\cong AI$. \end{itemize} \end{lemma} The proof of this lemma is found, for example, in \cite[Corollary 11.21]{Faith73}. There, condition $(ii)$ appears as an intermediary step in the proof of the equivalence between $(i)$ and $(iii)$. We state it explicitly because we use it in the proof of our main theorem. Chase's result is also recorded in \cite[Proposition 11.26]{Faith73}. \section{Proof of the main theorem}\label{sec_proof} Before we get into the proof, let's recall a few facts \cite[p. 478]{Wang80}. Let $k$ be a ring, $R=k[\vect Y n]\subseteq S=k[\vect X n]$ two polynomial rings(each in $n$ indeterminates) over $k$. Then $S$ is finitely presented as an $R$-algebra, in fact, write $Y_{i}=f_{i}\left(\vect X n\right) \in k[\vect X n] $, then \begin{align*} S\cong R[\vect Z n]/\left(\vect h n\right), \end{align*} where $R[\vect Z n]$ is the polynomial ring in $n$ indeterminates over $R$ and $h_{i}:=f_{i}\left(\vect Z n\right)-Y_{i}$. We will use an overline $\overline{g}$ to denote the image of $g\in R[\vect Z n]$ in $S$. For an extension $R=k[\vect Y n]\subseteq S=k[\vect X n]$ as above the following conditions are equivalent \cite[Theorem 38 and Theorem 7]{Wang80}: \begin{itemize} \item [(i)] $S$ is \'{e}tale over $R$. \item[(ii)] The Jacobian matrix $J:=\left(\frac{\partial f_{i}}{\partial X_{j}}\right)^{i=1,\cdots,n}_{j=1,\cdots,n}\in \operatorname{GL}_{n}\left(S\right)$ or equivalently $\operatorname{det}\left(J\right)\in S^{*}=k^{*}$. \item[(iii)] $S$ is separable and flat over $R$. \item[(iv)] The Jacobian matrix $\overline{J}:=\left(\overline{\frac{\partial h_{i}}{\partial Z_{j}}}\right)^{i=1,\cdots,n}_{j=1,\cdots,n}\in \operatorname{GL}_{n}\left(S\right)$ or equivalently $\operatorname{det}\left(\overline{J}\right)\in S^{*}=k^{*}$. \end{itemize} We now proceed to the proof of our main theorem. \begin{proof}[Proof of Theorem \ref{thm_main_theorem}] Assume that we have an \'{e}tale extension $R=k[\vect Y n]\subseteq S=k[\vect X n]$. By the observations above $S$ is finitely presented as an $R$-algebra, i.e., $S\cong R[\vect Z n]/I$, where $I=\left(\vect h n\right)$ is the ideal of $R[\vect Z n]$ generated by the $h_{i}$'s. By Ohi's criterion, Proposition \ref{Ohi_Proposition} $S$ is faithfully flat over $R$ if and only if $u\in c'(u)I$ for each $u \in I$. Let $u=\sum_{i=1}^{n}g_{i}h_{i}\in I$, where $g_{i}\in R[\vect Z n]$ for $i=1,\cdots,n$, in order to prove that $u\in c'(u)I$ it is enough to prove that $g_{i}\in c'(u)R[\vect Z n]$ for $i=1,\cdots,n$. We take the formal partial derivative of $u=\sum_{i=1}^{n}g_{i}h_{i}$ with respect to each $Z_{j}$ to obtain a system of $n$ equations in $R[\vect Z n]$ \begin{align*} \frac{\partial u}{\partial Z_{j}}=\sum_{i=1}^{n}\frac{\partial g_{i}}{\partial Z_{j}}h_{i} + \sum_{i=1}^{n}g_{i}\frac{\partial h_{i}}{\partial Z_{j}}, \qquad j=1, \cdots, n. \end{align*} Passing to the quotient by $I$ we obtain the following system of $n$ equations in $S$ \begin{align}\label{syst-equations} \overline{\frac{\partial u}{\partial Z_{j}}}= \sum_{i=1}^{n}\overline{g_{i}}\overline{\frac{\partial h_{i}}{\partial Z_{j}}}, \qquad j=1, \cdots, n. \end{align} As noted above $S$ is \'{e}tale over $R$ if and only if the Jacobian matrix $\left(\overline{\frac{\partial h_{i}}{\partial Z_{j}}}\right)$ is invertible. Then we use the Cramer's rule for commutative rings to solve the system (\ref{syst-equations}) for the $\overline{g_{i}}$'s, expressing them as linear combinations of the $\overline{\frac{\partial u}{\partial Z_{j}}}$'s with coefficients in $S$. This shows that $\overline{g_{i}}\in c'(u)S$ for $i=1,\cdots,n$ because by definition of $c'(u)$ it is always the case that $\frac{\partial u}{\partial Z_{j}}\in c'(u)R[\vect Z n]$ and $\overline{\frac{\partial u}{\partial Z_{j}}}\in c'(u)S$ for $j=1, \cdots, n$. At the same time we have an exact sequence of $R$-modules \begin{align*} 0 \to I \to R[\vect Z n] \to S \to 0, \end{align*} where $R[\vect Z n]$ is free over $R$. Also $S$ is flat over $R$ because $S$ is \'etale over $R$. Then Lemma \ref{Cartan-Eilenberg_lemma} item (ii) applied to the ideal $c'(u)\subset R$ gives the canonical isomorphism \begin{align*} \overline{g_{i}}\in c'(u)S\cong c'(u)R[\vect Z n]/c'(u)I, \end{align*} from which we conclude that $g_{i}\in c'(u)R[\vect Z n]$ for $i=1,\cdots,n$ as desired. \end{proof} \begin{example} Let $k$ be an algebraically closed field of characteristic $p$. Then any map of the form $f=\left(f_{1},\cdots, f_{n}\right):$ $k^{n}\to k^{n}$, $f_{i}(\vect X n)=X_{i}+\left(g_{i}(\vect X n)\right)^{p}$, where $g_{i}\in k[\vect X n]$ are arbitrary polynomials for $i=1,\cdots,n$, is surjective. In fact, this type of map is \'etale, the result follows from Corollary \ref{main_corollary}. \end{example} \bibliographystyle{abbrv}
{ "timestamp": "2022-05-03T02:48:06", "yymm": "2205", "arxiv_id": "2205.01032", "language": "en", "url": "https://arxiv.org/abs/2205.01032", "abstract": "We apply Ohi's criterion for faithfully flatness of extensions of commutative rings to prove that any étale extension $k[Y_1, \\ldots, Y_n]\\subseteq k[X_1, \\ldots, X_n]$ of polynomial rings (each in $n$ indeterminates) over a commutative ring $k$ is faithfully flat. In particular, if $k$ is an algebraically closed field then any étale polynomial map $k^{n} \\to k^{n}$ is surjective.", "subjects": "Commutative Algebra (math.AC); Algebraic Geometry (math.AG)", "title": "Étale extensions of polynomial rings are faithfully flat", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9773707946938299, "lm_q2_score": 0.8376199673867852, "lm_q1q2_score": 0.8186652931762421 }
https://arxiv.org/abs/1910.00629
Uniqueness of Optimal Point Sets Determining Two Distinct Triangles
In this paper, we show that the maximum number of points in $d\geq3$ dimensions determining exactly 2 distinct triangles is $2d$. We further show that this maximum is uniquely achieved by the vertices of the $d$-orthoplex. We build upon the work of Hirasaka and Shinohara who determined that the $d$-orthoplex is such an optimal configuration, but did not prove its uniqueness. Further, we present a more elementary argument for its optimality.
\section{Introduction} Paul Erd\H{o}s posed his distinct distance conjecture in 1946, spawning a wide array of similar problems in discrete geometry. He originally conjectured that a set of $n$ points in general position in the plane must determine $\Omega(n/\sqrt{\text{log}(n)})$ distinct distances \cite{erdos}. This conjecture was ultimately proved by Guth and Katz in 2015 \cite{guth_katz}. Instead of fixing the number of points, consider fixing the number of distinct distances $k$. A question that is closely associated to the original Erd\H{o}s distinct distance problem asks: what is the maximal number of points determining $k$ distinct distances? In 1996, Erd\H{o}s and Fishburn found all the optimal point configurations in the plane determining $k\leq 4$ distinct distances and found an optimal point configuration determining $k=5$ distinct distances (see Figure \ref{fig:distances}). \begin{figure}[!h] \includegraphics[scale=0.8]{OptimalDistanceConfigs.png} \caption{Maximal configurations determining exactly $k$ distances, for $2 \leq k \leq 6$ \cite{brass_et_al}. Note that for each $k>2$, there is an example from the triangular lattice; it is conjectured that this is always the case for $k$ large enough.} \label{fig:distances} \end{figure} They also conjectured that for each $k\geq 3$, at least one solution lies on the triangular lattice, and for $k\geq 7$, all solutions lie on the triangular lattice. This conjecture remains open \cite{erdos_fishburn}. A distance can be interpreted as a 1-simplex, so a natural generalization of this problem is to find the greatest number of points determining $k$ higher simplices. In our case, we consider 2-simplices, i.e. triangles. Some work has already been done in this direction. In \cite{epstein_et_al}, Epstein et al. showed that the optimal configuration determining one distinct triangle is the vertices of the square, and the optimal configurations determining two distinct triangles are the square with its center and the vertices of the regular pentagon. In \cite{brenner_et_al}, Brenner et al. determine that the unique optimal configuration determining one distinct triangle is the $d$-simplex in $\R^d$ for $d\geq 3$. Finally, in \cite{hirasaka_shinohara}, Hirasaka and Shinohara determined that an optimal configuration determining two triangles is the $d$-orthoplex in $\R^d$ for $d\geq 3$. As our main result, we provide a more elementary argument that the $d$-orthoplex is an optimal configuration determining two distinct triangles and offer a proof that it is in fact the unique optimal configuration determining two distinct triangles. \begin{thm}\label{thm:main_result} The vertices of the $d$-orthoplex are the unique optimal configuration determining two distinct triangles in $\R^d$ for $d\geq 3$. \end{thm} \section{Definitions, Lemmas and Setup} First, let us formalize our notion of distinct triangles. \begin{defi} Given a finite point set $P \subset \R^d$, we say two triples $(a,b,c), (a',b',c') \in P^3$ are equivalent if there is an isometry mapping one to the other, and we denote this as $(a,b,c) \sim (a',b',c')$. \end{defi} \begin{defi} Given a finite point set $P \subset \R^d$, we denote by $P_{nc}^3$ the set of noncollinear triples $(a,b,c) \in P^3$. \end{defi} \begin{defi} Given a finite point set $P \subset \R^d$, we define the set of distinct triangles determined by $P$ as \begin{equation} T(P) := P_{nc}^3 / \sim. \end{equation} \label{def:tris} \end{defi} We will need the following lemmas to proceed with our argument for our main result. Note that given a finite point set $S$ to say that a point $P$ in $S$ determines a distance $d$ means that there is another point $Q$ in $S$ such that $PQ = d$. \begin{restatable}{lem}{lemmaxdds} A finite point set determining $t$ distinct triangles determines at most $2t + 1$ distinct distances. \label{lem:max_dds} \end{restatable} \begin{restatable}{lem}{lemdistinctbound} A finite point set containing a point which determines $n$ distinct distances must determine at least $$\ceil*{\frac{n}{2}\cdot\floor*{\frac{n}{3}}}$$ distinct triangles. \label{lem:distinct_bound} \end{restatable} \begin{restatable}{lem}{lemrepeatbound} Given a finite point set $S$ and a point $P$ in $S$. Let $n$ be the number of distinct distances determined by $P$. Let $m$ denote the number of these distinct distances which are determined by $P$ and two or more distinct points. Then, $S$ determines at least $$\ceil*{\frac{n(n-1) + 4m}{6}}$$ distinct triangles. \label{lem:repeat_bound} \end{restatable} \noindent For an illustration of Lemma \ref{lem:repeat_bound}, see Figure \ref{fig:repeat_bound}. \begin{figure}[!h] \centering \input{repeat_bound.tikz} \caption{The distinct distances present are $d_1, d_2$ and $d_3$, so $n=3$. Of these, $d_2$ and $d_3$ are repeated, so $m=2$. Thus, this configuration determines at least 3 distinct triangles.} \label{fig:repeat_bound} \end{figure} The structure of our main proof will be based primarily on Lemma \ref{lem:max_dds}. By this lemma, configurations determining only two distinct triangles (our primary focus in this paper) can determine at most five distinct distances. In Section \ref{sec:elim_dds} we split into cases and show that any configuration of at least $2d$ points determining five, four or three distinct distances must determine more than two distinct triangles in $d$ dimensions. Trivially, any finite point set determining only one distinct distance can determine at most one distinct triangle. So we can conclude that any configuration of at least $2d$ points determining two distinct triangles must determine exactly two distinct distances. Then, in Section \ref{sec:geom}, we prove that any configuration of at least $2d$ points determining only two distinct triangles must, in fact, determine triangles of specific geometry. That is, one of the triangles must be equilateral and the other must be isosceles with its repeated edge length being that of the equilateral triangle. Then, in Section \ref{sec:main_result}, we show using these restrictions that any set of $2d$ points in $\R^d$ determining two distinct triangles must be the vertices of a $d$-orthoplex. As a corollary, we then show that a configuration of more points necessarily determines more distinct triangles. This proves Theorem \ref{thm:main_result}. \section{Eliminating Higher Numbers of Distinct Distances}\label{sec:elim_dds} All of the arguments below will proceed with the following setup. Given a set $S$ of $2d$ points $\R^d$, label one of the points of $S$ as $\Oc$ and the remaining points $P_1$, $P_2$, $\dots$, $P_{2d - 1}$. We will often choose $\Oc$ to be some convenient point in a manner that clearly does not lose generality by relabeling. \begin{prop}\label{prop:elim_dds} Any configuration of at least $2d$ points in $d\geq 3$ dimensions determining five, four or three distinct distances determines more than two distinct triangles. \end{prop} \begin{proof}\ \begin{description}[align=left] \item[5 Distinct Distances] Given that the two triangles must share an edge, two distinct triangles determining five distinct distances must both be scalene. This means that none of the distances from $\Oc$ to the remaining points can be repeated, otherwise the triangle formed by $\Oc$ and those two points would be isosceles (or equilateral). For $d>3$, $\Oc$ determines more than five distances, so by the pigeonhole principle, at least one of the distances must be repeated, which is a contradiction. In $d=3$ we clearly may assume all of the distances from $\Oc$ are distinct. By Lemma \ref{lem:distinct_bound}, these distances then determine at least $\ceil*{\frac{5}{3}\cdot\floor*{\frac{5}{2}}} = 4$ distinct triangles. \item[4 Distinct Distances] By similar reasoning, a configuration determining an equilateral triangle or two distinct isosceles triangles can determine at most three distinct distances. So, a configuration determining two distinct triangles and four distinct distances may determine no equilateral triangles and at most one distinct isosceles triangle. Note that when a given distance is determined by a single fixed point $P$ and at least two other distinct points $A$, $B$, etc., $\triangle PAB$ must be either equilateral or isosceles. Since it is impossible for one of the two distinct triangles to be equilateral in this case, we can assume that any repeated distances determine isosceles triangles. Since two distinct repeated distances necessarily form two non-congruent isosceles triangles and only one of the distinct triangles may be isosceles, only one distinct distance may be repeated. Finally, consider a configuration with at least two repetitions of a single distinct distance, e.g., $\Oc P_1 =\Oc P_2 = \Oc P_3= d_1$. \begin{figure}[h!] \centering \input{4distcontra.tikz} \caption{} \label{fig:icecream_cone} \end{figure} Since this configuration cannot determine an equilateral triangle and determines at most one distinct isosceles triangle, assume that $\triangle \Oc P_1 P_2$, $\triangle \Oc P_2 P_3$ and $\triangle \Oc P_1 P_3$ are all congruent to the same isosceles triangle. Then $\triangle P_1 P_2 P_3$ is equilateral (see Figure \ref{fig:icecream_cone}). So, by contradiction, the repeated distinct distance may be repeated only once. Since only one of the distances from $\Oc$ to the other $2d -1$ points may be repeated at most once, all four distinct distances must appear. So by Lemma \ref{lem:distinct_bound}, these distances then determine at least $\ceil*{\frac{4}{3}\cdot\floor*{\frac{4}{2}}} = 3$ distinct triangles. Notice that this argument depends only on having six or more points, so this holds in all cases. \item[3 Distinct Distances] In a configuration of points determining three distinct distances and two distinct triangles, at most one of the triangles may be equilateral, in which case the other must be scalene. \begin{rek*} In such a configuration of points, any distance repeating more than once from a single point must determine an equilateral triangle. \end{rek*} To see that this is true, consider a configuration in which $\Oc P_1 =\Oc P_2 = \Oc P_3= d_1$. Clearly, if only one or two of the three incomplete triangles is equilateral, we would have an equilateral and an isosceles triangle co-occurring. Otherwise, if none of the distances $P_1 P_2, P_2 P_3$ or $P_3 P_1$ are equal to $d_1$, and any of them differ, then $\triangle P_1 P_2 P_3$ must be an isosceles triangle not congruent to either of the two isosceles triangles containing $\Oc$. This is a contradiction as this configuration would then determine three distinct triangles. So, they must all be either $d_2$ or $d_3$. Then $\triangle P_1 P_2 P_3$ is equilateral, which is impossible as we already have an isosceles triangle. So, we must have $P_1 P_2 = P_2 P_3 = P_3 P_1 = d_1$. From this, it is clearly impossible for all of the distances from $\Oc$ to be the same, otherwise every triangle determined by this configuration would be equilateral, so specifically, the configuration would only determine one distinct triangle. So, there must be at least two distinct distances determined from $\Oc$. Consider one such configuration in which $\Oc P_1 =\Oc P_2 = \Oc P_3= d_1$ while $\Oc P_4 = d_2$. By the remark, $\triangle \Oc P_2 P_3$ must be equilateral, so specifically $P_2 P_3=d_1$. Since the remaining distinct triangle must be scalene, $\Oc P_3 P_4$ must be scalene, so specifically $P_3 P_4 = d_3$. By the same reasoning, $\triangle P_2 P_3 P_4$ must also be a congruent scalene triangle, so $P_2 P_4 = d_2$. Then $\triangle \Oc P_2 P_4$ is an isosceles triangle, which is a contradiction since $\triangle \Oc P_2 P_3$ is equilateral \begin{figure}[h!] \centering \input{3distcontra.tikz} \caption{} \label{fig:3distcontra} \end{figure} Thus, it is impossible for any distance to repeat more than once without determining more than two distinct triangles. For $d>3$, by a simple application of the pigeonhole principle it is clear that one of the distances must be repeated more than once, so by the above, configurations of more than six points determining three distinct distances necessarily determine at least three distinct triangles independent of dimension. In the case where $d=3$, the only possible configuration contains one repeat of two of the distinct distances (for example $d_1$ and $d_2$) and one occurrence of the third. This configuration determines $n=3$ distinct distances and two of them are repeated (i.e., $m=2$). So, by \ref{lem:repeat_bound}, this configuration must determine at least $\ceil*{\frac{3(3-1) + 4(2)}{6}} = 3$ distinct triangles. \end{description} \end{proof} \section{Geometry of Two Distinct Triangles Determined by \texorpdfstring{$2d$}{2d} Points}\label{sec:geom} By Proposition \ref{prop:elim_dds}, any configuration of at least $2d$ points in $d\geq 3$ dimensions determining two distinct triangles must also determine only two distinct distances. We can then prove the following proposition about the geometry of two distinct triangles. \begin{prop}\label{prop:tri_geom} Given a configuration of at least $2d$ points in $d\geq 3$ dimensions determining two distinct triangles and two distinct distances, one of the triangles must be equilateral, and the other triangle must be isosceles with its repeated edge length being the same as the edge length of the equilateral triangle. \end{prop} \begin{proof} We will begin by proving that one of the triangles must be equilateral. Assume to the contrary that neither of the triangles is equilateral. In this case, both of the triangles are isosceles with side lengths $(d_1, d_1, d_2)$ and $(d_2, d_2, d_1)$, which we will reference as $T_1$ and $T_2$ respectively. Clearly, since this configuration determines two distinct distances, there must be at least one point which determines both distances. Additionally, by the pigeonhole principle, one of the distances must occur at least $d$ out of $2d-1$ times from a given point. So, without loss of generality, assume $d_1$ is determined by $\Oc$ and at least three distinct points $P_1$, $P_2$, $P_3$, etc. All of $\triangle \Oc P_1 P_2$, $\triangle \Oc P_1 P_3$ and $\triangle \Oc P_2 P_3$ must be congruent to $T_1$. So, $P_1 P_2 = P_1 P_3 = P_2 P_3 = d_2$. Then $\triangle P_1 P_2 P_3$ is equilateral, which is a contradiction. Since one of the triangles must be equilateral (assume without loss of generality that it has side length $d_1$), it remains to eliminate the case where repeated distance of the second triangle is $d_2$. We will denote the equilateral triangle as $T_{equ}$ and the isosceles triangle as $T_{iso}$. For the following, we will split into two cases. For the first case we will consider $d_1$ as the distance that occurs $d$ times. Without loss of generality, assume that $\Oc P_1$,$\dots$ $\Oc P_d$ are $d_1$ and assume $\Oc P_{d+1} = d_2$. \begin{figure}[h!] \centering \input{equil_iso_setup.tikz} \caption{The setup for the 3-dimensional version of this argument} \end{figure} Notice that a single edge length of $d_2$ is sufficient to show that a given triangle is congruent to $T_{iso}$, and similarly two edge lengths of $d_1$ are sufficient to show congruence to $T_{equ}$. Then, since each triangle in $\{\triangle \Oc P_i P_j\}_{1 \leq i, j\leq d}$ for $i\neq j$ contains two edges of length $d_1$ (namely $\Oc P_i$ and $\Oc P_j$), each must be congruent to $T_{equ}$. And, specifically, we must have $P_i P_j = d_1$. This means that $\{\Oc, P_1,\dots, P_d\}$ is a set of $d+1$ mutually equidistant points in $d$ dimensions; namely, it determines a $d$-simplex. Notice now that none of the remaining points $P_{d+2},\dots,P_{2d - 1}$ may lie at distance $d_1$ from any of the points of the simplex, else, it must lie at distance $d_1$ from all of the points of the simplex. This is a contradiction as the maximum number of mutually equidistant points in $d$-dimensional Euclidean space is $d+1$, as is achieved by the regular $d$-simplex. \begin{figure}[h!] \centering \input{tetrahedron.tikz} \caption{In the 3-dimensional case, our simplex is the tetrahedron whose vertices are $\Oc$, $A$, $B$ and $C$. The point $D$ is the center of the tetrahedron. It is intuitively clear from this image that there is no consistent position for the contradictory point $E$ other than $D$.} \end{figure} So, specifically, every distance from each point that is not a vertex of the simplex to a vertex of the simplex must be $d_2$. So each point $P_{d+1},\dots, P_{2d -1}$ is equidistant from all of the vertices of the simplex. In general, $d+1$ points in general linear position uniquely determine a $(d-1)$-sphere (see Appendix \ref{app:geom}). Notice then that since the vertices of a $d$-simplex lie in general position, they uniquely determine a $(d-1)$-sphere. But, since each of the points $P_{d+1},\dots, P_{2d -1}$ by definition lie at the center of a $(d-1)$-sphere containing all of the vertices of the simplex, and there is only one such sphere, all of these points must coincide, which is a contradiction. Note that in $d=3$, there are two such points, so this argument holds for all $d\geq 3$. Consider instead the case where $d_2$ is repeated $d$ times. For clarity, relabel $P_{d+1}$ as $Q$ and without loss of generality, let $\Oc Q = d_1$. By a similar argument to the $d_1$ case, each triangle in $\{\triangle \Oc P_i P_j\}_{1 \leq i, j\leq d}$ for $i\neq j$ is congruent to $T_{iso}$. Then, specifically, each distance $P_i P_j$ is $d_1$. Clearly then, $\{P_1, \dots, P_d\}$ is a set of $d$ mutually equidistant point. So, these points are the vertices of a $(d-1)$-simplex. Now, notice that since $\Oc Q = d_1$ and each distance from $\Oc$ to a vertex of the simplex (i.e. $P_i$ for $1\leq i\leq d$) must be $d_2$, $\triangle \Oc Q P_i$ is congruent to $T_{iso}$. Thus $Q P_i = d_2$ for all $1 \leq i \leq d$. The set of all points equidistant from two points in $d$ dimensions, is a $(d-1)$-hyperplane (see Appendix \ref{app:geom}) So specifically, the vertices of the simplex lie on a $(d-1)$-hyperplane orthogonal to the line through $\Oc$ and $Q$. Suppose $\Oc P_{d+2} = d_2$. Then $\triangle \Oc P_{d+2} P_i$ must be congruent to $T_{iso}$ for $1\leq i\leq d$, so specifically the distance from $P_{d+2}$ to every vertex of the simplex must be $d_1$. So, $\{P_1, \dots, P_d, P_{d+2}\}$ is a set of $d+1$ mutually equidistant points, i.e. the vertices of a $d$-simplex. Similarly, since $QP_1 = d_2$ and $P_1 P_{d+2} = d_1$, we must have $QP_{d+2} = d_2$. Then, $P_{d+2}$ is equidistant from $\Oc$ and $Q$, so it must lie in the same $(d-1)$-hyperplane as the vertices of the simplex. This is a contradiction as a $d$-simplex clearly cannot lie within a $(d-1)$-hyperplane. So, fixing $d<j\leq 2d-1$, each $\Oc P_j$ must be $d_1$. Then each triangle made up of either $\Oc$ or $Q$, one of the vertices of the simplex and one of the remaining points $P_j$, must be congruent $T_{iso}$. So specifically, the distance between any point of the simplex and each of the remaining points must be $d_2$. Additionally, since $Q P_1 = P_1 P_j = d_2$, we must have $Q P_j = d_1$. Notice then that all such points $P_j$ are equidistant from $\Oc$ and $Q$. Namely they lie in the same $(d-1)$-hyperplane as the vertices of the simplex. Notice that each of these points is also equidistant from all of the vertices of the simplex. But as seen earlier, there is a unique point equidistant from all of the vertices of a $d$-simplex in $d$-dimensional Euclidean space. So, all of these points must coincide, which is a contradiction. So a point configuration in $d$ dimensions can only determine two distinct triangles if one of them is equilateral and the other is an isosceles triangle whose repeated edge length is the same as the side length of the equilateral triangle. \end{proof} \section{Proving the Main Result}\label{sec:main_result} In this section, we seek to prove that the $d$-orthoplex is the unique configuration of $2d$ points determining two distinct triangles in $d\geq 3$ dimensions. Notice that all previous sections assumed only that we have at least $2d$ points in $d$ dimensions, so at the end of this section, as an easy corollary to the proof of Proposition \ref{prop:unique}, we will see that any configuration of strictly more than $2d$ points in $d\geq 3$ dimensions determines more than two distinct triangles. Taking Proposition \ref{prop:unique} together with Corollary \ref{cor:optimal} will yield our main result, Theorem \ref{thm:main_result}. Note that in the following argument, we need only consider the pair of triangles determined by Proposition \ref{prop:tri_geom}. For what follows, label the equilateral triangle $T_{equ}$ and the isosceles triangle $T_{iso}$. This will be enough to directly construct the vertices of the $d$-orthoplex in $d$ dimensions. The proof proceeds inductively, so we begin by proving the following proposition which will serve as our base case. \begin{prop} The set of vertices of an octahedron is the unique configuration of six points determining two distinct triangles in three dimensions. \end{prop} \begin{proof} We begin with a similar setup as before, noting that at least one of the six points must determine both distances so we may assume by relabeling that $\Oc$ is one of these points. For convenience, we will label the remaining points $A$, $B$, $C$, $D$ and $E$. Assume $\Oc E = d_2$. Then, notice that $\triangle \Oc AE$, $\triangle \Oc BE$, $\triangle \Oc CE$ and $\triangle \Oc DE$ must all be congruent to $T_{iso}$, meaning all distances from points other than $E$ to $\Oc$ and all distances from points other than $\Oc$ to $E$ must be $d_1$. \begin{figure}[h!] \centering \input{octa_setup.tikz} \caption{} \end{figure} Notice that $A$, $B$, $C$ and $D$ are then all equidistant from $\Oc$ and $E$, so specifically they must be coplanar and lie on a circle centered on the midpoint of $\Oc$ and $E$ (see Appendix \ref{app:geom}). So, to determine the remaining distances, it remains only to determine the cyclic quadrilateral $\square ABCD$. Notice that any two adjacent edges equal to $d_2$ would necessarily determine an isosceles triangle with $d_2$ as the repeated edge, which is not one of our two distinct triangles. So, the only two cases we must eliminate are an isosceles trapezoid with three edges equal to $d_1$ and one edge equal to $d_2$, and a rectangle with opposite pairs of edges equal to $d_1$ and $d_2$. The diagonal of a rectangle is always longer than either of the two sides, so $AC$ and $BD$ could not be one of the two distances, which is a contradiction. In the case of the trapezoid, suppose $AB = d_2$. Then the diagonal cannot be $d_2$ since both of the diagonals are adjacent to the edge of length $d_2$. If the diagonal were $d_1$, then $\triangle CDA$ and $\triangle DCB$ would be congruent equilateral triangles on the same side of the same base. Specifically, they would be coincident which would imply $A = B$, a clear contradiction. The only possible remaining cyclic quadrilateral is a square of side length $d_1$. Then, $AC$ and $BD$ are necessarily $d_2$. An octahedron centered at the origin positioned along the coordinate axes is given by vertices $\{(\pm r, 0, 0), (0, \pm r, 0), (0, 0, \pm r)\}$ for some positive constant $r$. Assume without loss of generality that the shared midpoint of $\Oc E$, $AC$ and $BD$ lies at the origin and $E$ lies at $(d_2/2, 0, 0)$. This clearly presents no problem as it requires at worst a similarity transformation of any configuration satisfying the distances determined above. Then we must conclude that $\Oc = (-d_2/2, 0, 0)$. It is then easy to see that all of the remaining points must lie at the coordinates of the remaining vertices of the above octahedron. Then, we may conclude that any configuration of six points in three dimensions determining two distinct triangles must be given by the vertices of an octahedron. \end{proof} With the base case established, we can prove the following partial result by induction on number of dimensions. \begin{prop}\label{prop:unique} The set of vertices of the $d$-orthoplex is the unique configuration of $2d$ points in $d\geq 3$ dimensions determining two distinct triangles. \end{prop} \begin{proof} Assume the unique point configuration determining two distinct triangles in $k$ dimensions is determined by the vertices of the $k$-orthoplex, which specifically contains $2k$ points. Now, consider a configuration of $2k+2$ points determining two distinct triangles in $k+1$ dimensions. By Proposition \ref{prop:elim_dds}, it must determine exactly two distinct distances, and by Proposition \ref{prop:tri_geom} one of the triangles must be equilateral, and the other is isosceles with its repeated edge length equal to the side length of the equilateral triangle. For clarity, relabel $P_{2k+1}$ as $Q$, then we may assume without loss of generality that $\Oc Q = d_2$. Notice that a single edge length of $d_2$ is sufficient to determine that a given triangle is congruent to $T_{iso}$, so specifically we must have $\Oc P_i = Q P_i = d_1$ for all $1 \leq i \leq 2k$. So, all of these points are equidistant from $\Oc$ and $Q$, so as mentioned before, they lie in a $k$-hyperplane. Notice further that since the distances determined by each of the points from $\Oc$ and $Q$ are the same, specifically, all of the points $P_i$ lie on a $(k-2)$-sphere centered on the midpoint of $\Oc Q$. For the entire point set to determine two distinct triangles, these $2k$ points must determine either one or two distinct triangles. The greatest number of points in $k$ dimensions determining one distinct triangles is $k+1$ \cite{brenner_et_al}, so for $k>1$, $2k$ points in $k$ dimensions must determine more than one distinct triangle. So, specifically, the points $P_1, \dots, P_{2k}$ must determine two distinct triangles. Since they all lie in the same $k$-hyperplane, we may conclude from the inductive hypothesis that these points then must be the vertices of a $k$-orthoplex. If we then assume again that the center of the $k$-orthoplex lies at the origin and the remaining vertices lie along the $x_1,\dots, x_k$ coordinate axes, then clearly $\Oc$ and $Q$ lie on the $x_{k+1}$ coordinate axis. Specifically, they lie at coordinates $(0, \dots, 0, d_2/2)$ and $(0, \dots, 0, -d_2/2)$. So, the whole point configuration must determine the vertices of a $(k+1)$-orthoplex. \end{proof} \begin{cor}\label{cor:optimal} Any configuration of strictly greater than $2d$ points in $d\geq 3$ dimensions determines more than two distinct triangles. \end{cor} \begin{proof} Consider a configuration of $2d+1$ points in $d$ dimensions. As mentioned before, since at most $d+1$ points can determine a single triangle, any $2d$ points must determine two distinct triangles. Then, by Proposition \ref{prop:unique}, any $(2d)$-subset of the point configuration must be the vertices of a $d$-orthoplex. So, specifically, $P_1,\dots, P_{2d}$ must be given by the vertices of a $d$-orthoplex. For convenience, relabel $P_{2d+1}$ as $P'$ Consider the distance between $P'$ and an arbitrary vertex of the orthoplex. Each vertex of the orthoplex lies at distance $d_2$ from exactly one other vertex of the orthoplex, so $P' $ cannot lie at distance $d_2$ from any of the vertices of the orthoplex; otherwise, an isosceles triangle with repeated edge $d_2$ would be created. So, the distance from $P'$ to each of the vertices of the orthoplex must be $d_1$. Clearly this is the center of the unique $(d-1)$-sphere on which the vertices of the orthoplex lie. This must coincide with the common midpoint of the segments of length $d_2$. Assume without loss of generality that $P_1$ is not opposite $P_2$. Then, $P'P_1 \perp P'P_2$. That is to say, $\triangle P'P_1 P_2$ is right. But, notice that $P_1 P_2 = d_1$, so it is also equilateral, which is a contradiction. So, a configuration of at least $2d+1$ points in $d$ dimensions necessarily determines more than two distinct triangles. So the point configuration consisting of the vertices of the $d$-orthoplex in $d$ dimensions is optimal. \end{proof} Then, Theorem \ref{thm:main_result} follows directly from Proposition \ref{prop:unique} together with Corollary \ref{cor:optimal}. \section{Proofs of Lemmas} For the main proof, we used the three essentially combinatorial lemmas, Lemma \ref{lem:max_dds}, Lemma \ref{lem:distinct_bound} and Lemma \ref{lem:repeat_bound}. In this section, we restate and prove these three lemmas. \lemmaxdds* \begin{proof} Let $S$ be a finite point configuration determining $t$ distinct triangles. Consider an arbitrary finite ascending chain $A_3 \subset A_4 \subset \dots \subset A_{\abs{S} - 1} \subset S$ such that $A_3 \subseteq S$ is a three point subset of $S$, and $A_{i+1} \setminus A_i$ contains a single point. Clearly, $A_3$ determines at most 3 distinct distances and 1 distinct triangle. Consider $A_{k+1}$ for some $k$, and let $Q\in A_{k+1}$ be the unique point in $A_{k+1}$ such that $Q\notin A_{k}$. Notice that if $T(A_{k}) = T(A_{k+1})$, then the number of distinct distances determined by $A_k$ must be the same as that determined by $A_{k+1}$. So, assume that $T(A_k) \neq T(A_{k+1})$. Notice that every possible representative of any element of $T(A_{k+1}) \setminus T(A_k)$ is of the form $\triangle Q P_i P_j$ where $P_i, P_j \in A_k$. Then, the only possible new distinct distances are $QP_i$ and $QP_j$, so the difference between the sizes of the distance sets of $A_{k+1}$ and $A_k$ is at most $2\cdot (\abs{T(A_{k+1})} - \abs{T(A_k)})$. Then, proceeding inductively, it is clear that the number of distinct distances determined by $S$ is at most $2(\abs{T(S)} - 1) + 3 = 2t + 1$. \end{proof} \lemdistinctbound* \begin{proof} This result is essentially a direct consequence of a result by Fort and Hedlund (Theorem 1 in \cite{fort_hedlund}). For convenience, a paraphrasing of Fort and Hedlund's result is as follows. Given the set of all distinct unordered pairs of positive integers between 1 and $n$ inclusive, a covering of this set by triples is a set of distinct unordered triples such that each pair is a subset of at least one of the triples. The minimal number of triples needed to cover the set of distinct unordered pairs of distinct positive integers is $\ceil{n/3 \cdot \floor{n/2}}$. Consider a point $Q$ and points $P_1, \dots, P_k$ such that $Q P_i = d_i$ are all distinct distances. Then, consider each pair of these distances. The corresponding points must determine a triangle $\triangle Q P_i P_j$. For the purposes of this proof, a triangle may be thought of as the unordered triple of its edge lengths. Then, the minimal number of triangles needed to realize this configuration is clearly the minimal number of unordered triples of distinct $d_i$'s required to cover all distinct unordered pairs of distinct $d_i$'s. This problem is clearly isomorphic to the problem considered by Fort and Hedlund (i.e., by replacing $d_i$ with $i$), so we can conclude that the minimal number of triangles determined by a point set containing such a point is $\ceil{n/3 \cdot \floor{n/2}}$. \end{proof} \lemrepeatbound* \begin{proof} Although the setup differs somewhat from the setup above, this proof will proceed in a similar fashion. While Fort and Hedlund considered only pairs of distinct numbers (e.g., $(1,2)$, $(4, 1)$, etc.), for the following setup, we will also allow pairs of the same number (e.g., $(2,2)$, $(5,5)$). Then, consider a set of distinct pairs of (not necessarily distinct) integers between 1 and $n$ inclusive. Let $n$ be the number of pairs of distinct integers and $m$ be the number of pairs not distinct integers. We hope to produce a lower bound on the size of a minimal covering by triples of such an arbitrary set. Given a triple $(a, b, c)$, clearly this triple can cover at most 3 distinct pairs, $(a, b)$, $(b, c)$ and $(a, c)$. So, if we have $m=0$, we clearly require at least $n/3$ triples to cover the $n$ pairs of distinct integers. Consider now $m\neq 0$. Each pair of non-distinct integers $(a, a)$ needs to be covered by its own distinct triple $(a, a, b)$. That is to say, $(a, a)$ and $(b, b)$ cannot be covered by the same triple. Notice that the triple $(a, a, b)$ also includes the pair $(a, b)$. So, each triple required to cover a pair of non-distinct integers can also contain a pair of distinct integers. So one fewer of the distinct pairs needs to be covered separately. Thus, a lower bound on the total number of triples needed to cover such pairs is clearly $\ceil{(n-m)/3 + m} = \ceil{(n+2m)/3}$. We can apply the same analogy as used in the proof of Lemma \ref{lem:distinct_bound} to apply this bound to our problem. Notice that if a point determines $k$ distinct distances, it necessarily determines $k$ choose $2$ distinct pairs of distinct distances. Substituting this into our above bound, we finally come to the conclusion that given a point which determines $n$ distinct distances, $m$ of which are repeated, this point configuration determines at least $$\ceil*{\frac{n(n-1) + 4m}{6}}$$ distinct triangles. \end{proof}
{ "timestamp": "2019-10-03T02:02:38", "yymm": "1910", "arxiv_id": "1910.00629", "language": "en", "url": "https://arxiv.org/abs/1910.00629", "abstract": "In this paper, we show that the maximum number of points in $d\\geq3$ dimensions determining exactly 2 distinct triangles is $2d$. We further show that this maximum is uniquely achieved by the vertices of the $d$-orthoplex. We build upon the work of Hirasaka and Shinohara who determined that the $d$-orthoplex is such an optimal configuration, but did not prove its uniqueness. Further, we present a more elementary argument for its optimality.", "subjects": "Combinatorics (math.CO)", "title": "Uniqueness of Optimal Point Sets Determining Two Distinct Triangles", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9901401467331061, "lm_q2_score": 0.8267118004748678, "lm_q1q2_score": 0.818560543428176 }
https://arxiv.org/abs/1812.05791
$n$-absorbing monomial ideals in polynomial rings
In a commutative ring $R$ with unity, given an ideal $I$ of $R$, Anderson and Badawi in 2011 introduced the invariant $\omega(I)$, which is the minimal integer $n$ for which $I$ is an $n$-absorbing ideal of $R$. In the specific case that $R = k[x_{1}, \ldots, x_{n}]$ is a polynomial ring over a field $k$ in $n$ variables $x_{1},\ldots, x_{n}$, we calculate $\omega(I)$ for certain monomial ideals $I$ of $R$.
\section{Introduction} Throughout this paper, we set $\mathbb{N} := \{1,2,\ldots, \}$ and $\mathbb{N}_{0} := \{0,1,2, \ldots \}$ and $R$ will denote a commutative ring with unity. Given a nonzero ideal $I$ of $R$, Ass$(R/I)$ will denote the set of associated primes of $I$ in $R$. The primary notion we are interested in this paper is the following: \begin{defi} Let $n \in \mathbb{N}$, $R$ a commutative ring with unity, and $I$ an ideal of $R$. $I$ is said to be an \textbf{n-absorbing ideal} of a ring $R$ if for any $x_{1},\ldots, x_{n+1} \in R$ such that $x_{1} \cdots x_{n+1} \in I$, there is some $j$ with $1 \leq j \leq n+1$ such that $x_{1} \cdots x_{j-1}x_{j+1} \cdots x_{n+1} \in I$. $I$ is said to be a \textbf{strongly n-absorbing ideal} of a ring $R$ if for any ideals $I_{1},\ldots, I_{n+1}$ of $R$ such that $I_{1} \cdots I_{n+1} \subseteq I$, there is some $j$ with $1 \leq j \leq n+1$ such that $I_{1} \cdots I_{j-1}I_{j+1} \cdots I_{n+1} \subseteq I$. \end{defi} (Strongly) $2$-absorbing ideals were initially defined and investigated by Badawi in \cite{Badawi} as a generalization of prime ideals, which are precisely the proper $1$-absorbing ideals. In 2011, Anderson and Badawi together generalized this further to the notion of a (strongly) $n$-absorbing ideal for any $n \in \mathbb{N}$ defined above in \cite{Anderson}. For an ideal $I$ in a ring $R$, we let $\omega(I)$ denote the minimal integer $n \in \mathbb{N}$ such that $I$ is $n$-absorbing. In a general ring, $I$ may not be $n$-absorbing for any $n \in \mathbb{N}$, in which case we set $\omega(I) = \infty$. Similarly, we can define the invariant $\omega^{\bullet}(I)$ to be the smallest integer $n \in \mathbb{N}$ for which an ideal $I$ is strongly $n$-absorbing, and set $\omega^{\bullet}(I) = \infty$ if no such integer exists. We set $\omega(R) = \omega^{\bullet}(R) = 0$. It is easy to see that $\omega(I)\le \omega^{\bullet}(I)$ holds for each ideal $I$ of $R$. In fact, Anderson and Badawi in Conjecture $1$ of \cite[page 1669]{Anderson} postulate that $\omega(I) = \omega^{\bullet}(I)$ holds for any ideal $I$ in arbitrary ring $R$; that is, they conjecture that the notion of an $n$-absorbing ideal and strongly $n$-absorbing ideal coincide. As of this writing, this problem remains open. However, it is known that the conjecture holds true for any $n\in\mathbb{N}$ if $R$ is a Pr{\"u}fer domain (\cite[Corollary 6.9]{Anderson}) or a commutative algebra over an infinite field(\cite{Donadze}), and for any ring $R$ if $n=2$ (\cite[Theorem 2.13]{Badawi}). The interested reader may refer to the survey article \cite[Section 5]{Badawi2} for further information on strongly $n$-absorbing ideals. Recall that for an ideal $I$ in a ring $R$, the \textit{Noether exponent} of $I$, denoted by $e(I)$, is the minimal integer $\mu \in \mathbb{N}$ such that $(\sqrt{I})^{\mu} \subseteq I$. If such an integer does not exist, we set $e(I) = \infty$. We also set $ e(R) = 0$. In a Noetherian ring, since $\sqrt{I}$ is finitely generated for any ideal $I$, $e(I) < \infty$. Anderson and Badawi in \cite{Anderson} establish a connection between $\omega^{\bullet}(I)$ and Noether exponents: \begin{theorem}[Remark 2.2, Theorem 5.3, Section 6, Paragraph 2 on page 1669, \cite{Anderson}] \label{Wtheo} Let $I_{1},\cdot\cdot\cdot, I_{r}$ be ideals of a ring $R$. Then $\omega(I_{1}\cap\cdot\cdot\cdot\cap I_{r})\le\omega(I_{1})+\cdot\cdot\cdot +\omega(I_{r})$ and $\omega^{\bullet}(I_{1}\cap\cdot\cdot\cdot\cap I_{r})\le\omega^{\bullet}(I_{1})+\cdot\cdot\cdot +\omega^{\bullet}(I_{r})$. In particular, let $I$ be an ideal in a Noetherian ring $R$. If $I = Q_{1} \cap \cdots \cap Q_{n}$, where the $Q_{i}$ are primary ideals, then $\omega(I)\le\omega^{\bullet}(I) \leq \sum\limits\limits_{i=1}^{n} e(Q_{i})$. Thus every ideal in a Noetherian ring is $n$-absorbing for some $n \in \mathbb{N}$. \end{theorem} On the other hand, $e(I)$ is actually a lower bound of $\omega(I)$. \begin{lemm} \label{fund} Given an ideal $I$ of a ring $R$, $e(I) \le \omega(I)$. If $Q$ is a primary ideal of $R$, then $\omega(Q) = \omega^{\bullet}(Q) = e(Q)$. \end{lemm} \begin{proof} The first statement follows either \cite[Corollary 3]{Walker} or \cite{Donadze2}. The second statements follows from \cite[Theorem 6.3(c), Theorem 6.6]{Anderson}. \end{proof} This raises the question then if for an arbitrary ideal $I$ whether $\omega(I)$ can be described purely in terms of Noether exponents or possibly other well-known ring-theoretic invariants. This has been investigated to some extent by others in at least one case. Namely, Moghimi and Naghani \cite[Theorem 2.21(1)]{Moghimi} show that in a discrete valuation ring $R$, $\omega(I)$ is precisely the length of the $R$-module $R/I$. In this spirit, we attempt to give in this paper a description of $\omega(I)$ in terms of other ring-theoretic invariants in the special case that $I$ is a monomial ideal of a polynomial ring over a field. In some cases, our arguments are general enough to also give the same results for $\omega^{\bullet}(I)$, and thus as a side-effect we can show that in some cases the notion of $n$-absorbing ideal and strongly $n$-absorbing ideal coincide as Anderson and Badawi conjecture. The present paper is divided into two parts. In section 2, we review the definitions and facts concerning $n$-absorbing ideals and monomial ideals. Using these we calculate $\omega(I)$ for primary monomial ideals by computing Noether exponents and the standard primary decomposition of monomial ideals. These results lead to the study of how $\omega(I)$ can be explicitly computed from the generating set of $I$ when $I$ is a monomial ideal of $R=k[x_{1},\cdot\cdot\cdot, x_{n}]$ with $n\le 3$ in the following section. The second part is section 4, where we define and investigate $\omega$-linear monomial ideals, i.e., monomial ideals $I$ such that $\omega(I^{m})=m\omega(I)$ for each $m\in\mathbb{N}$. We give a characterization theorem for primary $\omega$-linear monomial ideals, and in particular show that every integrally closed monomial ideal in $R=k[x,y]$ and the edge ideal of a cycle is $\omega$-linear. \section{Some Background} As a prerequisite of the main section of this paper, we briefly review some of the basic material excerpted from \cite{Herzog} regarding monomial ideals, and show that $\omega(I)$ can be directly calculated from the generators of $I$ when $I$ is a primary monomial ideal. Let $k$ be a field and $R=k[x_{1},\cdot\cdot\cdot, x_{n}]$ be the polynomial ring with $n$ variables over $k$. An element of $R$ of the form $x_{1}^{a_{1}}\cdot\cdot\cdot x_{n}^{a_{n}}$ with $a_{i}\in \mathbb{N}_{0}$ is called a \textit{monomial}, and an ideal of $R$ generated by monomials is called a \textit{monomial ideal}. The degree of $f=x_{1}^{a_{1}}\cdot\cdot\cdot x_{n}^{a_{n}}$, denoted by $\deg(f)$, is defined to be $a_{1}+\cdot\cdot\cdot+a_{n}$. $G(I)$ will denote the set of monomials in $I$ which are minimal with respect to divisibility. Any element of $R$ can be written uniquely as a $k$-linear combination of monomials; that is, given $f\in R$, we may write $f=\sum a_{u}u$ where the sum is taken over the monomial ideals of $R$ and $a_{u}\in k$ for each monomial $u$. Then the \textit{support} of $f$, denoted by $\text{supp}(f)$, is the set of monomials $u$ such that $a_{u}\neq 0$. An ideal $I$ of a ring $R$ is \textit{irreducible} if there are no ideals $I_{1}, I_{2}$ of $R$ such that $I=I_{1}\cap I_{2}$ and $I\subsetneq I_{1}$, $I\subsetneq I_{2}$. We denote by $\mathfrak{m}$ the unique maximal homogeneous ideal of $R$. \begin{lemm} \label{support} (\cite[Chapter 1]{Herzog}) Let $R = k[x_{1},\ldots, x_{n}]$ and $I$ a monomial ideal of $R$ generated by monomials $u_{1},\cdot\cdot\cdot, u_{r}$ of $R$. Then the following hold: \begin{enumerate}[label=(\roman*)] \item Given a monomial $f\in I$, there exists $i\in\{1,\cdot\cdot\cdot, r\}$ so $u_{i}|f$. \item $G(I)$ is the unique minimal set of monomial generators of $I$. \item $I$ can be written as a finite intersection of ideals of the form $(x_{i_{1}}^{d_{1}},\ldots, x_{i_{m}}^{d_{m}})$. An irredundant presentation of this form is unique ($I=Q_{1}\cap\cdots\cap Q_{r}$ is \textrm{irredundant} if none of the ideals $Q_{i}$ can be omitted). \item $I$ is irreducible if and only if $I$ is of the form $(x_{i_{1}}^{d_{1}},\ldots, x_{i_{m}}^{d_{m}})$. Moreover, every irreducible monomial ideal of the form $(x_{i_{1}}^{d_{1}},\ldots, x_{i_{m}}^{d_{m}})$ is $(x_{i_{1}},\ldots, x_{i_{m}})$-primary. \item If $J$ is another monomial ideal of $R$, then \[ I \cap J=(\{lcm(u,v)\mid u \in G(I), v \in G(J)\}). \] In particular, if $a$ and $b$ are coprime monomials of $R$ and $I$ is a monomial ideal of $R$, then $(ab,I)=(a,I)\cap(b,I)$. \item An ideal $I'$ of $R$ is monomial if and only if for each $f\in I'$, $\supp(f)\subseteq I'$. \end{enumerate} \end{lemm} By Lemma \ref{support}(iv), the irredundant unique decomposition of Lemma \ref{support}(iii) is also a primary decomposition of $I$, which is known as the \textit{standard decomposition} of $I$ (see \cite[P. 12]{Herzog}). We will also need the following characterization of primary monomial ideals: \begin{lemm} \label{pch} \cite[Exercise 3.6]{Eisenbud} Let $R=k[x_{1},\cdot\cdot\cdot, x_{n}]$ and $P=(x_{i_{1}},\cdot\cdot\cdot, x_{i_{r}})$ a monomial prime ideal of $R$. Then given a $P$-primary monomial ideal $Q$, there exists $a_{1},\cdot\cdot\cdot, a_{r}\in \mathbb{N}$ and monomials $f_{1},\cdot\cdot\cdot, f_{s} \in k[x_{i_{1}},\cdot\cdot\cdot, x_{i_{r}}]$ so $G(Q)=\{x_{1}^{a_{1}},\cdot\cdot\cdot, x_{i_{r}}^{a_{r}}, f_{1},\cdot\cdot\cdot, f_{s}\}$. Conversely, every monomial ideal of this form is a $P$-primary ideal. \end{lemm} \begin{coro} \label{monprimary} Let $P$ be a prime monomial ideal and $I,J$ be $P$-primary monomial ideals of $R$. Then both $I\cap J$ and $IJ$ are $P$-primary monomial ideals. Moreover, $I:J$ is a $P$-primary monomial ideal provided $J \not\subset I$. \end{coro} \begin{proof} This is an immediate consequence of Lemma \ref{support}.(v) and Lemma \ref{pch}. \end{proof} We can now calculate $\omega(I)$, where $I$ is an irreducible monomial ideal. \\ \begin{lemm} \label{powermonomial} Let $R = k[x_{1},\ldots,x_{n}]$ denote a polynomial ring over a field $k$. Let $I = (x^{d_{1}}_{i_{1}},\ldots, x_{i_{m}}^{d_{m}})$, where $d_{1},\ldots, d_{n} \in \mathbb{N}$ and $1 \leq i_{1} < i_{2} < \cdots < i_{m} \leq n$. Then $\omega(I)=\omega^{\bullet}(I)=e(I)=d_{1} + \cdots + d_{m} - m + 1$. \end{lemm} \begin{proof} Since $I$ is a $(x_{i_{1}}, x_{i_{2}}, \cdot\cdot\cdot, x_{i_{m}})$-primary ideal by Lemma \ref{pch}, the first two equalities follow from Lemma \ref{fund}. Thus it suffices to show that $e(I)=r$, where $r=d_{1} + \cdots + d_{m} - m + 1$. We have $\sqrt{I} = (x_{i_{1}},\ldots, x_{i_{m}})$. For $N \in \mathbb{N}$, $(\sqrt{I})^{N} \subseteq I$ if and only if for every $c_{1},\ldots,c_{m} \in \mathbb{N}_{0}$ with $c_{1} + \cdots + c_{m} = N$, we have $x^{c_{1}}_{i_{1}} \cdots x^{c_{m}}_{i_{m}} \in I$. By Lemma \ref{support}(i), the latter happens precisely when $c_{i} \geq d_{i}$ for some $1 \leq i \leq m$. Thus $e(I)=r$. \end{proof} Next, we produce a way to calculate $\omega(I)$ when $I$ is a monomial primary ideal not necessarily generated by pure powers. \begin{lemm} \label{small} Let $I$ be an ideal of a ring $R$. Suppose there is $P \in \Spec(R)$ such that $I=J_{1}\cap\cdot\cdot\cdot \cap J_{r}$, where $J_{i}$ are ideals of $R$ with $\sqrt{J_{i}}=P$ for each $i \in \{ 1, \ldots, r\}$. Then $e(I)=\max_{1\le i \le r} \{e(J_{i})\}$. \end{lemm} \begin{proof} Note that $\sqrt{I}=\sqrt{J_{1}}\cap\cdot\cdot\cdot\cap \sqrt{J_{r}}=P$. Thus given $\mu \in \mathbb{N}$, $(\sqrt{I})^{\mu} \subseteq I$ if and only if $(\sqrt{J_{i}})^{\mu} \subseteq J_{i}$ for each $i \in \{1, \ldots, r \}$, from which the conclusion of the lemma follows. \end{proof} \begin{coro} \label{primary} Let $R = k[x_{1},\ldots,x_{n}]$ denote a polynomial ring over a field $k$. If $Q$ is a monomial primary ideal of $R$ and $Q=\cap_{i=1}^{r}Q_{i}$ is its standard decomposition, then \begin{align*} \omega(Q)=\omega^{\bullet}(Q)=\max_{1\le i\le r}\{e(Q_{i})\}. \end{align*} \end{coro} \begin{ex} Let $R=k[x,y,z]$ with a field $k$ and $I=(x^{4}, y^{3}, z^{2}, xy, y^{2}z)$. Then repeatedly applying Lemma \ref{support}(v), we obtain the standard decomposition $I=(x, y^{2}, z^{2}) \cap (x^{4}, y, z^{2}) \cap (x, y^{3}, z)$. Thus by Lemma \ref{powermonomial} and Corollary \ref{primary}, \begin{align*} \omega(I)=\omega^{\bullet}(I)=\max\{1+2+2-3+1, 4+1+2-3+1, 1+3+1-3+1\}=5. \end{align*} \end{ex} \section{When $I$ is a monomial ideal of $R=k[x_{1},\cdot\cdot\cdot, x_{n}]$ with $n\le 3$} In this section we show that when $I$ is a monomial ideal of $R=k[x_{1},\cdot\cdot\cdot, x_{n}]$ with $n\le 3$, then $\omega(I)$ can be explicitly calculated from $G(I)$. We first prove a theorem analogous to \cite[Theorem 2.5]{Anderson2}. \begin{lemm} \label{irreducible} Let $R$ be a UFD and $p$ an irreducible element of $R$. Then given $n\in\mathbb{N}$, $I$ is an $n$-absorbing ideal of $R$ if and only if $pI$ is an $(n+1)$-absorbing ideal of $R$. In particular, $\omega(pI) = \omega(I) + 1$. \end{lemm} \begin{proof} Suppose that $I$ is $n$-absorbing. Let $f_{1},\ldots, f_{n+2} \in R$ and $f_{1} \cdots f_{n+2} \in pI$. Then since $p$ is irreducible, $p \mid f_{i}$ for some $i$. Without loss of generality, suppose that $p \mid f_{1}$. Then $f_{1}/p \in R$, and so $(f_{1}/p)f_{2} \cdots f_{n+2} \in I$. Since $I$ is $n$-absorbing, and hence $(n+1)$-absorbing as well, we have that either $(f_{1}/p) f_{2} \cdots \widehat{f_{i}} \cdots f_{n+2} \in I$ for some $i \in \{2,\ldots, n+2\}$, in which case $f_{1}f_{2} \cdots \widehat{f_{i}} \cdots f_{n+2} \in pI$ and we're done, or $f_{2} \cdots f_{n+2} \in I$. This is a product of length $n+1$, so that since $I$ is $n$-absorbing, for some $j$ with $2 \leq j \neq n+1$, we have $f_{2} \cdots \widehat{f_{j}} \cdots f_{n+2} \in I$. Thus $p f_{2} \cdots \widehat{f_{j}} \cdots f_{n+2} \in pI$, and so $f_{1} f_{2} \cdots \widehat{f_{j}} \cdots f_{n+2} \in pI$. This shows that $pI$ is then $(n+1)$-absorbing, and $\omega(pI) \leq \omega(I)+1$. To show the converse, suppose that $pI$ is an $(n+1)$-absorbing ideal. If $I$ is not an $n$-absorbing ideal, then there exists $f_{1},\cdot\cdot\cdot, f_{n+1} \in R$ such that $f=f_{1}\cdot\cdot\cdot f_{n+1} \in I$ but $f_{1}\cdot\cdot\cdot \widehat{f_{i}} \cdot\cdot\cdot f_{n+1} \not\in I$ for each $i$. Since $pI$ is $(n+1)$-absorbing and $pf \in pI$, it follows that either $pf_{1}\cdot\cdot\cdot \widehat{f_{i}} \cdot\cdot\cdot f_{n+1} \in pI$ for some $i$ or $f \in pI$. But the former is impossible by our choice of $f_{i}$'s, and without loss of generality we may assume that $p|f_{1}$. Now $(f_{1}/p)f_{2}\cdot\cdot\cdot f_{n} \in I$, and neither $\widehat{(f_{1}/p)}f_{2}\cdot\cdot\cdot f_{n+1}$ nor $(f_{1}/p)f_{2}\cdot\cdot\cdot \widehat{f_{i}} \cdot\cdot\cdot f_{n+1}$ is an element of $I$ for each $i\geq 2$. Therefore, since $R$ is a UFD, we may assume that none of $f_{i}$ are divisible by $p$. Now $pf_{1}\cdot\cdot\cdot f_{n+1} \in pI$, but $pf_{1}\cdot\cdot\cdot \widehat{f_{i}} \cdot\cdot\cdot f_{n+1} \not\in pI$ and $f_{1}\cdot\cdot\cdot f_{n+1} \not\in pI$, which contradicts the assumption that $pI$ is an $(n+1)$-absorbing ideal. Hence $I$ is an $n$-absorbing ideal and $\omega(pI)\ge \omega(I)+1$. \end{proof} The following corollary is now immediate. \begin{coro} \label{prinmo} Given a monomial $f$ and an ideal $I$ of $R=k[x_{1},\cdot\cdot\cdot, x_{n}]$, $\omega(fI)=$deg$(f)+\omega(I)$. In particular, $\omega(fR)=$deg$(f)$. \end{coro} Given a monomial ideal $I$ with the standard decomposition $I=\cap_{\ell=1}^{t}T_{\ell}$, we can define an equivalence relation on $\{1,\cdot\cdot\cdot, t\}$ by defining $i\sim j$ iff $\sqrt{T_{i}}=\sqrt{T_{j}}$, and set $\{S_{i}\}_{i=1}^{r}$ to be the corresponding equivalence classes. Then $Q_{i}=\cap_{\ell\in S_{i}}T_{\ell}$ is a monomial primary ideal for each $i \in \{1, \ldots, r\}$, and $I=\cap_{i=1}^{r}Q_{i}$ is an irredundant primary decomposition of $I$. We will call this decomposition the \textit{canonical primary decomposition} of $I$. \begin{lemm} \label{l} Let $R=k[x_{1},\cdot\cdot\cdot, x_{n}]$. Let $I$ be a monomial ideal with canonical primary decomposition $I=\cap_{i=1}^{r}{Q_{i}}$. If there exists $k \in\{1,\cdot\cdot\cdot, r\}$ such that $\sqrt{Q_{i}}\subseteq \sqrt{Q_{k}}$ for all $i\in\{1,\cdot\cdot\cdot, r\}$, then $\omega(I)=\max\{e(Q_{k}), \omega(\cap_{1\le i\le r, i\neq k}Q_{i})\}$ and $\omega^{\bullet}(I)=\max\{e(Q_{k}), \omega^{\bullet}(\cap_{1\le i\le r, i\neq k}Q_{i})\}$. \end{lemm} \begin{proof} Let $t=\max\{e(Q_{k}), \omega(\cap_{1\le i\le r, i\neq k}Q_{i})\}$. We will first show that $I$ is $t$-absorbing. If not, then there are $f_{1},\ldots, f_{t+1} \in R$ such that $f=\prod_{j=1}^{t+1}f_{j} \in I$ but $g_{j} :=f/f_{j}\not\in I$ for each $j \in\{1,\ldots, t+1\}$. Hence given any $i \in\{1,\ldots, t+1\}$, there exists $\ell \in \{1,\ldots, r\}$ such that $g_{i} \not\in Q_{\ell}$, and since $f_{i}g_{i}=f \in I\subseteq Q_{\ell}$, we must have $f_{i}\in \sqrt{Q_{\ell}}\subseteq \sqrt{Q_{k}}$. Therefore, $g_{j}\in (\sqrt{Q_{k}})^{t}\subseteq (\sqrt{Q_{k}})^{e(Q_{k})}\subseteq Q_{k}$ for all $j \in\{1,\ldots, t+1\}$. On the other hand, $\cap_{1\le i\le r, i\neq k}Q_{i}$ is $t$-absorbing and $f\in \cap_{1\le i\le r, i\neq k}Q_{i}$, so that we conclude $g_{j} \in \cap_{1\le i\le r, i\neq k}Q_{i}$ for some $j \in\{1,\ldots, t+1\}$ and thereby $g_{j}\in I$, a contradiction. Thus $\omega(I) \leq t$. Next, we show that $\omega(I)\ge t$; that is, $I$ is not $(t-1)$-absorbing. We now consider two cases.\\ \\ Case $1$: $t = \omega( \cap_{1 \leq i \leq r, i \neq k} Q_{i})$. Since $\cap_{1\le i\le r, i\neq k}Q_{i}$ is not $(t-1)$-absorbing, there are $h_{1},\ldots, h_{t} \in R$ such that $h=\prod_{i=1}^{t}h_{i} \in \cap_{1\le i\le r, i\neq k}Q_{i}$ and $\ell_{j} := h/h_{j} \not\in \cap_{1\le i\le r, i\neq k}Q_{i}$ for each $j\in\{1,\ldots, t\}$. By an argument similar to the first paragraph of this proof, $h_{i}\in \sqrt{Q_{k}}$ for each $i\in \{1,\ldots, t\}$, and so $h \in Q_{k}$. Hence $h \in I$ and $\ell_{j}\not\in I$ for each $j\in \{1,\ldots, t\}$, so that $I$ is not $(t-1)$-absorbing.\\ \\ Case $2$: $t = e(Q_{k})$. Consider the standard decomposition of $I$, and choose an irreducible component $T$ of $I$ such that $e(T)=e(Q_{k})$ and $\sqrt{T}=\sqrt{Q_{k}}$. Since we obtained the canonical primary decomposition $I=\cap_{i=1}^{r} Q_{i}$ from the standard decomposition, we can choose a monomial $g\in (\cap_{1\le i\le r, i\neq k}Q_{i})\setminus T$ by Lemma \ref{support}(vi). Now $T=(x_{i_{1}}^{a_{1}},\cdot\cdot\cdot, x_{i_{l}}^{a_{l}})$ for some $a_{j} \in \mathbb{N}$ and $1\le i_{1}<\cdot\cdot\cdot<i_{l}\le n$. Note that we may assume that $g=\prod_{j=1}^{l}x_{i_{j}}^{c_{j}}$ for some $c_{j}\in\mathbb{N}_{0}$ such that $c_{j}<a_{j}$ for each $j\in\{1,\cdot\cdot\cdot, l\}$. Set \[f :=x_{i_{1}}^{a_{1}-1}\cdot\cdot\cdot x_{i_{l}}^{a_{l}-1}(x_{i_{1}}+\cdot\cdot\cdot+x_{i_{l}}).\] Then $f$ is a product of $e(T)$ elements of $\sqrt{T}$ by Lemma \ref{powermonomial}, and so $f \in (\sqrt{T})^{e(T)}= (\sqrt{Q_{k}})^{e(Q_{k})}\subseteq Q_{k}$. Since $g\mid f$ it also follows that $f \in \cap_{1\le i\le r, i\neq k}Q_{i}$. Hence $f \in I$. However, given $j \in\{1,\cdot\cdot\cdot, l\}$, $\cfrac{f}{x_{i_{j}}}\not\in T$. Indeed, $x_{i_{1}}^{a_{1}-1}\cdot\cdot\cdot x_{i_{l}}^{a_{l}-1}\in \supp(\cfrac{f}{x_{i_{j}}})\setminus T$ by Lemma \ref{support}.(i), and $\cfrac{f}{x_{i_{j}}}\not\in T$ by Lemma \ref{support}.(vi). Similarly $x_{i_{1}}^{a_{1}-1}\cdot\cdot\cdot x_{i_{l}}^{a_{l}-1}=\cfrac{f}{x_{i_{1}}+\cdot\cdot\cdot+x_{i_{l}}}\not\in T$. Therefore $I$ is not $(e(Q_{k})-1)$-absorbing, and $\omega(I)\ge e(Q_{k})=t$. Hence we have shown that $\omega(I)=\max\{e(Q_{k}), \omega(\cap_{1\le i\le r, i\neq k}Q_{i})\}$. The proof of $\omega^{\bullet}(I)=\max\{e(Q_{k}), \omega^{\bullet}(\cap_{1\le i\le r, i\neq k}Q_{i})\}$ can be obtained in a similar manner, and is omitted. \end{proof} The following corollary is immediate. \begin{coro} \label{refined} Let $R=k[x_{1},\cdot\cdot\cdot, x_{n}]$ and $I$ be a monomial ideal of $R$ with standard decomposition $I=\bigcap_{i=1}^{r}T_{i}$. Then $\omega(I)=\omega^{\bullet}(I)=\max_{1\le i\le r}\{e(T_{i})\}$ if Ass$(R/I)$ is totally ordered under set inclusion. \end{coro} In the next lemma, we give a characterization of when the upper bound of $\omega(I)$ from Theorem \ref{Wtheo} is sharp. \begin{lemm} \label{upper} Let $I$ be a monomial ideal of $R=k[x_{1},\cdot\cdot\cdot, x_{n}]$ with an irredundant primary decomposition $I = Q_{1} \cap \cdots \cap Q_{r}$. Then $\omega(I)=\omega^{\bullet}(I)=\sum\limits_{i=1}^{r}e(Q_{i})$ if and only if $I$ has no embedded associated primes \end{lemm} \begin{proof} Set $P_{i}=\sqrt{Q_{i}}$ for each $i=1,\cdot\cdot\cdot, r$.\\ \\ $\Leftarrow$: Assume that $P_{1},\cdot\cdot\cdot, P_{r}$ are incomparable prime ideals. The case when $r=1$ follows from Lemma \ref{fund}, so we may assume that $r\ge 2$. Since $\omega(I)\le\omega^{\bullet}(I)\le\sum\limits_{i=1}^{r}e(Q_{i})$ by Theorem \ref{Wtheo}, it suffices to show that $I$ is not $( \sum\limits_{i=1}^{r} e(Q_{i})-1)$-absorbing. Now given $i \in \{1,\cdot\cdot\cdot, r\}$, choose $T_{i}$ to be an irreducible component of $I$ with $\sqrt{T_{i}} = P_{i}$ and $e(T_{i}) = e(Q_{i})$. Write $T_{i} = (x_{i_{1}}^{a_{1}},\cdot\cdot\cdot, x_{i_{s_{i}}}^{a_{s_{i}}})$ with $1\le i_{1}<\cdot\cdot\cdot< i_{s_{i}}\le n$ and $a_{1},\cdot\cdot\cdot, a_{s_{i}}\in \mathbb{N}$. For $i \in \{1, \ldots, r\}$ and $j \in \{1, \ldots, s_{i}\}$, set \[f_{i,j}=x_{i_{j}}+\sum\limits_{t\neq j}x_{i_{t}}^{2} \text{ and } f_{i}= \displaystyle \Big(\sum\limits_{l=1}^{s_{i}}x_{i_{l}}\Big)\Big(\prod_{j=1}^{s_{i}}f_{i,j}^{a_{j}-1}\Big).\] It follows that $f_{i}\in P_{i}^{e(T_{i})}=(\sqrt{Q_{i}})^{e(Q_{i})}\subseteq Q_{i}$. Thus $f :=\prod_{i=1}^{r}f_{i} \in I$, and $f$ is a product of $\sum\limits_{i=1}^{r}e(Q_{i})$ elements of $R$. We wish to show that $\cfrac{f}{\sum\limits_{l=1}^{s_{i}}x_{i_{l}}}\not\in I$ and $\cfrac{f}{f_{i,j}}\not\in I$ for each $i\in\{1,\cdot\cdot\cdot, r\} \text{ and } j\in\{1,\cdot\cdot\cdot, s_{i}\}$. Without loss of generality, we let $i = 1$. Note that $\cfrac{f_{1}}{f_{1,j}}\not\in T_{1}$, since $\prod_{t=1}^{s_{1}}x_{1_{t}}^{a_{t}-1}\in \supp\Big(\cfrac{f_{1}}{f_{1,j}}\Big)\setminus T_{1}$. On the other hand, $\sum\limits_{l=1}^{s_{i}}x_{i_{l}}\not\in P_{1}$ and $f_{i,l}\not\in P_{1}$ for each $i\neq 1$ and $l\in\{1,\cdot\cdot\cdot, s_{i}\}$. Therefore $f_{i}\not\in P_{1}$ for each $i\neq 1$, and $\cfrac{f}{f_{1}}=\prod_{i=2}^{r}f_{i}\not\in P_{1}$. Hence $\cfrac{f}{f_{1,j}}=(\cfrac{f}{f_{1}})(\cfrac{f_{1}}{f_{1,j}})\not\in Q_{1}$. The proof that $\cfrac{f}{\sum\limits_{l=1}^{s_{1}}x_{1_{l}}}\not\in Q_{1}$ follows similarly. Hence we have $\omega(I)=\sum\limits_{i=1}^{r}e(Q_{i})$. \\ \\ $\Rightarrow$: We prove the contrapositive; assume that $P_{1},\cdot\cdot\cdot, P_{r}$ are not incomparable prime ideals. Then without loss of generality we may assume that $P_{1}\subsetneq P_{2}$, and we have $\omega(Q_{1}\cap Q_{2})=\max\{e(Q_{1}), e(Q_{2})\}$ by Corollary \ref{refined}. Therefore by Theorem \ref{Wtheo} we have \begin{align*} \omega(I)&=\omega \Big(Q_{1}\cap Q_{2}\cap(\bigcap_{i\neq 1,2} Q_{i})\Big)\\ &\le \omega \Big(Q_{1}\cap Q_{2} \Big)+\omega \Big(\bigcap_{i\neq 1,2} Q_{i} \Big)\\ &=\max\{e(Q_{1}), e(Q_{2})\}+\omega \Big(\bigcap_{i\neq 1,2} Q_{i}\Big)\\ &\le\max\{e(Q_{1}), e(Q_{2})\}+\sum_{i\neq 1,2} e(Q_{i})\\ &<\sum_{i=1}^{r}e(Q_{i}). \end{align*} \end{proof} Lemma \ref{l} and Lemma \ref{upper} yield the following corollary. \begin{coro} \label{n-1} Let $I$ is a monomial ideal of $R=k[x_{1},\cdot\cdot\cdot, x_{n}]$ with dim$(R/I)=1$. Let $I=\cap_{i=1}^{r} Q_{i}$ be the canonical primary decomposition of $I$. Then \[ \omega(I) = \omega^{\bullet}(I)=\left\{ \begin{array}{ll} \max\{e(Q_{k}),\sum\limits_{i\neq k}e(Q_{i})\} & \text{ if } \sqrt{Q_{k}}= \mathfrak{m} \text{ for some } k\in\{1,\cdot\cdot\cdot, r\}. \\ \sum\limits_{i=1}^{r}e(Q_{i}) & \text{ otherwise }.\\ \end{array} \right. \] \end{coro} \begin{coro} \label{realprinmo} Let $f$ be a monomial of $R$. Then $\omega^{\bullet}(fR)=deg(f)$. In particular, $\omega(fR)=\omega^{\bullet}(fR)$. \end{coro} \begin{proof} Let $f=\prod_{k=1}^{r}x_{i_{k}}^{a_{k}}$ for some $a_{1},\ldots, a_{r} \in \mathbb{N}$ and $1 \leq i_{1} < i_{2} < \cdots < i_{r} \leq n$. Then $fR=x_{i_{1}}^{a_{1}}R\cap\cdot\cdot\cdot \cap x_{i_{r}}^{a_{r}}R$, and by Lemma \ref{powermonomial} and Lemma \ref{upper} we have $\omega(fR)=\omega^{\bullet}(fR)=\sum_{i=1}^{r}e(x_{i_{k}}^{a_{k}}R)=\sum_{i=1}^{r} a_{k}=$deg$(f)$. \end{proof} Given a monomial ideal $I$ of $R=k[x,y,z]$, we can produce an algorithm that can compute $\omega(I)$. If $I$ is principal, then Corollary \ref{realprinmo} says that $\omega(I )$ is equal to the degree of a generator for $I$. Otherwise, $I=hJ$ for some monomial $h$ and a monomial ideal $J$ with dim$(R/J)\le1$. Now, $\omega(J)$ can be calculated explicitly using Corollary \ref{primary} or Corollary \ref{n-1} after obtaining a canonical primary decomposition of $I$, and we have $\omega(I)=deg(h)+\omega(J)$ by Corollary \ref{prinmo}. \begin{ex} Let $R=k[x,y,z]$ and $I=(x^{3}y^{4}, x^{2}y^{5}, x^{4}y^{3}z^{2}, x^{5}y^{3}z, x^{2}y^{4}z^{2})$. Then $I=x^{2}y^{3}J$ with canonical primary decomposition $J=(x^{2},y)\cap (y,z) \cap (x^3, y^2, z^2, xy)$. By Lemma \ref{powermonomial} and Corollary \ref{primary}, the standard decomposition $(x^3, y^2, z^2, xy)=(x, y^{2}, z^{2})\cap (x^{3}, y, z^{2})$ yields that $e((x^3, y^2, z^2, xy))=4$. Thus by Corollary \ref{n-1}, \begin{align*} \omega(I)&=deg(x^{2}y^{3})+\omega(J)\\ &=5+\max\{e((x^3, y^2, z^2, xy)), e((x^{2},y))+e((y,z))\}\\ &=5+\max\{4, 2+1\}\\ &=9. \end{align*} \end{ex} Another interesting corollary of Lemma \ref{irreducible} and Lemma \ref{l} is a formula of $\omega(I)$ and $\omega^{\bullet}(I)$ for monomial ideals of $R=k[x,y]$ where $k$ is a field and $x,y$ are indeterminates over $k$. \begin{theorem} \label{hoo} Let $R = k[x,y]$ and $J$ a monomial ideal of $R$. Write $J=(x^{a_{1}}y^{b_{1}},\cdot\cdot\cdot, x^{a_{r}}y^{b_{r}})$, where $\{a_{i} \}$ is strictly decreasing and $\{ b_{i} \}$ is strictly increasing. Then \[ \omega(J) =\omega^{\bullet}(J)= \left\{ \begin{array}{ll} a_{1}+b_{1} & \text{ if } r=1. \\ \max_{1\le i \le r-1}\{a_{i}+b_{i+1}\}-1 & \text{ if } r>1.\\ \end{array} \right. \] \end{theorem} \begin{proof} The case when $r=1$ follows from Corollary \ref{realprinmo}. For $r > 1$, first observe the standard decomposition of $J$ is $J= x^{a_{r}}R\cap y^{b_{1}}R \cap (x^{a_{1}}, y^{b_{2}})\cap (x^{a_{2}}, y^{b_{3}}) \cap \cdot\cdot\cdot \cap (x^{a_{r-1}},y^{b_{r}})$ (\cite[Proposition 3.2]{Miller}). The case $b_{1}=a_{r}=0$ follows from Corollary \ref{primary}. Suppose that at least one of $a_{r}$ and $b_{1}$ is nonzero. Thus by Lemma \ref{powermonomial} and Corollary \ref{n-1}, \begin{align*} \omega(J)=\omega^{\bullet}(J)&=\max\{e((x^{a_{1}}, y^{b_{2}})\cap (x^{a_{2}}, y^{b_{3}}) \cap \cdot\cdot\cdot \cap (x^{a_{r-1}},y^{b_{r}})), e(x^{a_{r}}R)+e(y^{b_{1}}R)\}\\ &=\max\{\max_{1\le i\le r-1}\{e((x^{a_{i}}, y^{b_{i+1}}))\}, a_{r}+b_{1}\}\\ &=\max\{\max_{1\le i\le r-1}\{a_{i}+b_{i+1}-1\}, a_{r}+b_{1}\}\\ &=\max_{1\le i\le r-1}\{a_{i}+b_{i+1}\}-1. \end{align*} \end{proof} \begin{ex} If $R=k[x,y]$ and $J=(x^{11}y^{4}, x^{8}y^{5}, x^{7}y^{9}, x^{4}y^{10}, x^{2}y^{16})$, then by Theorem \ref{hoo},\\ $\omega(J)=\omega^{\bullet}(J)=\max\{11+5, 8+9, 7+10, 4+16\}-1=19$. \end{ex} \section{$\omega$-linear ideals} Given an ideal $I$ of a ring $R$, we will say that $I$ is an \textit{$\omega$-linear ideal} if $\omega(I^{m})=m\omega(I)$ for each $m\in\mathbb{N}$. Perhaps the most common example of $\omega$-linear ideals can be found amongst those $P \in \Spec(R)$ which $P^{n}$ is $P$-primary for each $n \in \mathbb{N}$ (\cite[Theorem 3.1, Theorem 5.7]{Anderson}). For instance, \\ \\ 1. $R$ is a Pr{\"u}fer domain and $P^{2}\neq P$.\\ 2. $R$ is a Noetherian ring and $P$ is a maximal ideal that contains a nonzerodivisor.\\ 3. $R=k[x_{1},\cdot\cdot\cdot, x_{n}]$ and $P$ is a monomial ideal.\\ \\ In this section, we investigate the properties of $\omega$-linear ideals. Again, we will restrict our concern to monomial ideals of a polynomial ring $R=k[x_{1},\cdot\cdot\cdot, x_{n}]$ where $k$ is a field. We first consider a few useful inequalities regarding monomial ideals. \begin{lemm} \label{topi} Let $I$ be a monomial ideal of $R = k[x_{1},\ldots,x_{n}]$. Then $\omega(I)\ge \max\{\deg(f)\mid f\in G(I)\}$. \end{lemm} \begin{proof} Let $f\in G(I)$. Then $f=\prod_{k=1}^{r}x_{i_{k}}^{a_{k}}$ for some $a_{1},\ldots, a_{r} \in \mathbb{N}$ and $1 \leq i_{1} < i_{2} < \cdots < i_{r} \leq n$. Since $f\in I$ but $\cfrac{f}{x_{i_{k}}}\not\in I$ for each $k\in\{1,\cdot\cdot\cdot, r\}$ by minimality of $G(I)$, we have that $I$ is not $($deg$(f)-1)$-absorbing. Hence $\omega(I)\ge$ deg$(f)$, and since $f$ was chosen arbitrarily, we have the desired conclusion. \end{proof} \begin{lemm} \label{orderreversing} Let $I\subseteq J$ be ideals of a ring $R$. If $\sqrt{I}=\sqrt{J}$, then $e(J)\le \omega(I)$. In particular, if $I$ and $J$ are both $P$-primary ideals of a prime ideal $P$ of $R$, then $\omega(J)\le \omega(I)$. \end{lemm} \begin{proof} Since $\sqrt{I}=\sqrt{J}$, $(\sqrt{J})^{\omega(I)}\subseteq(\sqrt{I})^{e(I)}\subseteq I\subseteq J$ by Lemma \ref{fund} and $e(J)\le \omega(I)$. The second statement follows from that $e = \omega$ for primary ideals. \end{proof} \begin{lemm} \label{idealop} Let $P$ be a prime monomial ideal of $R = k[x_{1},\ldots,x_{n}]$. If $I,J$ are $P$-primary monomial ideals of $R$, then $\omega(I+J)\le\min\{\omega(I),\omega(J)\} \le \max\{\omega(I),\omega(J)\}=\omega(I\cap J)\le\omega(IJ)\le \omega(I)+\omega(J)$. Moreover, $\omega(I:J)\ge \omega(I) - \omega(J)$. \end{lemm} \begin{proof} Note that by Corollary \ref{monprimary}, $IJ\subseteq I\cap J\subseteq I+J$ are all $P$-primary monomial ideals. Therefore $\omega(I+J)\le \min\{\omega(I),\omega(J)\}\le\max\{\omega(I),\omega(J)\}\le \omega(I\cap J)\le\omega(IJ)$ by Lemma \ref{orderreversing}. On the other hand, let $I=\bigcap_{i=1}^{r}Q_{i}$ and $J=\bigcap_{j=1}^{s}T_{j}$ be standard decomposition of $I$ and $J$, respectively. Then $I\cap J=(\bigcap_{i=1}^{r}Q_{i})\cap(\bigcap_{j=1}^{s}T_{j})$ is an irreducible decomposition of $I\cap J$, and by throwing away any redundant component, there exists $A\subseteq \{1,\cdot\cdot\cdot, r\}$ and $B\subseteq \{1,\cdot\cdot\cdot, s\}$ so $I\cap J=(\bigcap_{i\in A}Q_{i})\cap(\bigcap_{j\in B}T_{j})$ is the standard decomposition of $I\cap J$. Thus by Corollary \ref{primary}, \begin{align*} \omega(I\cap J)&=\max\{\max_{i\in A}\{e(Q_{i})\},\max_{j\in B}\{e(T_{j})\}\}\\ &\le \max\{\max_{1\le i\le r}\{e(Q_{i})\},\max_{1\le j\le s}\{e(T_{j})\}\}\\ &=\max\{\omega(I),\omega(J)\}. \end{align*} Moreover, $(\sqrt{IJ})^{e(I)+e(J)}=P^{e(I)+e(J)}=P^{e(I)}P^{e(J)}=(\sqrt{I})^{e(I)}(\sqrt{J})^{e(J)}\subseteq IJ$, and so $e(IJ)\le e(I)+e(J)$. Combined with Lemma \ref{fund}, we have $\omega(IJ)\le \omega(I)+\omega(J)$. It remains to show that $\omega(I:J)\ge \omega(I)-\omega(J)$. When $J\subseteq I$, then we have $I:J=R$ and $\omega(I:J)=0\ge \omega(I)-\omega(J)$ by Lemma \ref{orderreversing}. If $J\not\subseteq I$, then $I:J$ is $P$-primary by Corollary \ref{monprimary}, and since $J(I:J)\subseteq I$, we have $\omega(I:J)+\omega(J)\ge \omega(I)$ by the first part of this lemma, hence the claim. \end{proof} As Anderson and Badawi pointed out (\cite[Example 2.7]{Anderson}), the conclusion of Lemma \ref{idealop} does not hold in every ring $R$. We add, that even in a polynomial ring over a field, the conclusion of the above lemma may fail if we drop any part of the hypothesis. \begin{ex} \label{tw} Let $R=k[x,y,z]$ and $I=(x^{2}, xy, y^{2}, xz^{2})$ and $J=(x^{2}, xy, y^{2}, yz^{3})$, so that neither $I$ nor $J$ are primary ideals. The standard decompositions of $I,J$ and $I\cap J$ are \begin{align*} I&=(x^{2}, y, z^{2})\cap (x, y^{2})\\ J&=(x, y^{2}, z^{3})\cap (x^{2}, y)\\ I\cap J&=(x, y^{2})\cap (x^{2}, y)\\ I+J&=(x,y)\cap (x, y^{2}, z^{3})\cap (x^{2}, y, z^{2}). \end{align*} Thus we have $\omega(I)=3$, $\omega(J)=4$, $\omega(I\cap J)=2$ and $\omega(I+J)=4$, so that $\omega(I\cap J)<\omega(I+J)=\max\{\omega(I),\omega(J)\}$. \end{ex} \begin{ex} Let $R = k[x,y,z]$ and $I=(x,y)$ and $J=(y,z^{2})$, so that $I$ and $J$ are both primary, but $\sqrt{I} \neq \sqrt{J}$. Then we have $\omega(I)=1$, $\omega(J)=2$ and $\omega(I\cap J)=3$, so that $\omega(I\cap J)>\max\{\omega(I),\omega(J)\}$. \end{ex} \begin{coro} \label{idealp} Let $I$ be a primary monomial ideal of $R = k[x_{1},\ldots,x_{n}]$. Then for each $m \in \mathbb{N}$ we have $\omega(I^{m})\le m\omega(I)$. \end{coro} \begin{proof} Follows immediately by induction on $m$ and Lemma \ref{idealop}. \end{proof} Next, we derive a characterization of primary monomial $\omega$-linear ideals. \begin{lemm} \label{limitratio} Let $R=k[x_{1},\cdot\cdot\cdot, x_{n}]$ and $Q$ a primary monomial ideal of $R$, so that $G(Q)=\{x_{i_{1}}^{a_{1}},\cdot\cdot\cdot, x_{i_{r}}^{a_{r}}, f_{1},\cdot\cdot\cdot, f_{t}\}$ where $a_{1},\ldots, a_{r} \in \mathbb{N}$, $1 \leq i_{1} < i_{2} < \cdots < i_{r} \leq n$ and $f_{1},\cdot\cdot\cdot, f_{t}$ are monomials in $k[x_{i_{1}},\cdot\cdot\cdot, x_{i_{r}}]$. Choose $s \in \{1,\cdot\cdot\cdot, r\}$ so $a_{s}=\max_{1\le j\le r}\{a_{j}\}$. \begin{enumerate} \item If $f_{1}=f_{2}=\cdot\cdot\cdot=f_{t}=0 , then $\omega(Q^{m})=(m-1)a_{s}+\omega(Q)$ for each $m \in\mathbb{N}$. \item $Q$ is $\omega$-linear if and only if $\omega(Q)=a_{s}$ \end{enumerate} \end{lemm} \begin{proof} \begin{enumerate} \item Let $Q=(x_{i_{1}}^{a_{1}},\cdot\cdot\cdot, x_{i_{r}}^{a_{r}})$. Then given $m \in \mathbb{N}$, set $S_{m}=\{(k_{1},\cdot\cdot\cdot,k_{r})\in \mathbb{N}^{r}\mid \sum\limits_{j=1}^{r}k_{j}=m+r-1\}$ and $Q_{k}=(x_{i_{1}}^{k_{1}a_{1}},\cdot\cdot\cdot, x_{i_{r}}^{k_{r}a_{r}})$ for each $k=(k_{1},\cdot\cdot\cdot,k_{r}) \in S_{m}$. Then $Q^{m}=\cap_{k \in S_{m}}Q_{k}$ (\cite[Theorem 6.2.4]{Moore}). Now by Corollary \ref{primary} and Lemma \ref{powermonomial}, $\omega(Q^{m})=\max_{k \in S_{m}}\{e(Q_{k})\}=\max_{k\in S_{m}}\{\sum\limits_{j=1}^{r} k_{j}a_{j}\}-r+1=(m-1)a_{s}+\omega(Q)$. \item Fix $m \in \mathbb{N}$ and set $I_{1}=(x_{i_{1}}^{a_{1}},\cdot\cdot\cdot, x_{i_{r}}^{a_{r}})^{m}, I_{2}=(x_{i_{1}},\cdot\cdot\cdot, x_{i_{s-1}},x_{i_{s}}^{ma_{s}}, x_{i_{s+1}}, \cdot\cdot\cdot, x_{i_{r}})$. It follows that $I_{1}\subseteq Q^{m}\subseteq I_{2}$ are $(x_{i_{1}},\cdot\cdot\cdot, x_{i_{r}})$-primary ideals, so we have\\ $ma_{s}=\omega(I_{2})\le\omega(Q^{m})\le\omega(I_{1})=(m-1)a_{s}+\sum_{j=1}^{r}a_{j}-r+1$ by Corollary \ref{orderreversing}, Lemma \ref{powermonomial} and part 1 of this lemma. Therefore if $Q$ is $\omega$-linear, then $\omega(Q)=\displaystyle \lim_{m\to\infty}\cfrac{m\omega(Q)}{m}= \displaystyle \lim_{m\to\infty}\cfrac{\omega(Q^{m})}{m}=a_{s}$. Conversely, suppose that $\omega(Q)=a_{s}$ and fix $m\in\mathbb{N}$. Then since $x_{i_{s}}^{ma_{s}}\in G(Q^{m})$ we have $\omega(Q^{m})\ge ma_{s}=m\omega(Q)$ by Lemma \ref{topi}. Hence $\omega(Q^{m})=m\omega(Q)$ by Corollary \ref{idealp} and so $Q$ is $\omega$-linear. \end{enumerate} \end{proof} \begin{coro} \label{return} Let $I$ be an irreducible monomial ideal of $R=k[x_{1},\cdot\cdot\cdot, x_{n}]$ so that $I=(x_{i_{1}}^{a_{1}},\cdot\cdot\cdot, x_{i_{r}}^{a_{r}})$ for some $a_{1},\ldots, a_{n} \in \mathbb{N}$. Set $a_{s}=\max_{1\le j\le r}\{a_{j}\}$. Then the following are equivalent. \begin{enumerate} \item $I$ is $\omega$-linear. \item $\omega(I^{m})=m\omega(I)$ for some $m>1$. \item $\omega(I)=a_{s}$. \item $a_{i}=1$ for each $i\neq s$. \end{enumerate} \end{coro} \begin{proof} $1\Rightarrow 2$: Obvious.\\ $2\Rightarrow 3$: Suppose that $\omega(I^{m})=m\omega(I)$ for some $m >1$. By Lemma \ref{limitratio}.1 we have $\omega(I^{m})=(m-1)a_{s}+\omega(I)$. Hence $\omega(I)=a_{s}$.\\ $3\Leftrightarrow 4$: Immediate consequence of Lemma \ref{powermonomial}.\\ $3\Leftrightarrow 1$: Follows from Lemma \ref{limitratio}.2 \end{proof} \begin{lemm} Let $P$ be a monomial prime ideal of $R$. If $I,J$ are $P$-primary $\omega$-linear monomial ideals of $R$, then so is $I\cap J$. \end{lemm} \begin{proof} Without loss of generality we may assume that $\omega(I)\ge \omega(J)$. By Lemma \ref{limitratio}.2, there is $j\in\{1,\cdot\cdot\cdot, r\}$ so that $x_{i_{j}}^{\omega(I)}\in G(I)$. There exists $a\in\mathbb{N}$ so $x_{i_{j}}^{a}\in G(J)$. Then again, by Lemma \ref{limitratio}.2, $a\le \omega(J)$. Now, $x_{i_{j}}^{\omega(I)}=lcm(x_{i_{j}}^{\omega(I)}, x_{i_{j}}^{a}) \in G(I\cap J)$. On the other hand, $\omega(I\cap J)=\omega(I)$ by Lemma \ref{idealop}. Hence $I\cap J$ is $\omega$-linear by Lemma \ref{limitratio}.2. \end{proof} Given a monomial ideal $I$ of $R=k[x,y]$ we will write $I=(x^{a_{1}}y^{b_{1}}, \ldots, x^{a_{r}}y^{b_{r}})$ where $\{a_{i}\}$ and $\{b_{i}\}$ are strictly decreasing and strictly increasing sequences of non-negative integers, respectively. Similarly, if $J$ is a monomial ideal of $R$ we write $J=(x^{c_{1}}y^{d_{1}}, \ldots, x^{c_{s}}y^{d_{s}})$ where $\{c_{i}\}$ and $\{d_{i}\}$ are strictly decreasing and strictly increasing sequence of non-negative integers, respectively. Hence $b_{1}=a_{r}=0$ iff $I$ is $(x,y)$-primary, and $d_{1}=c_{s}=0$ iff $J$ is $(x,y)$-primary. \begin{lemm} \label{2dimex} Let $R=k[x,y]$ and $I,J$ be $(x,y)$-primary monomial ideals with $\omega(I)\ge \omega(J)$. Then $\omega(IJ)\le \omega(I)+\max\{c_{1}, d_{s}\}$. \end{lemm} \begin{proof} We may assume that $c_{1}\ge d_{s}$. Then $e(I)=\omega(I)\ge\omega(J)\ge c_{1}$ by Lemma \ref{topi}, so $(x,y)^{e(I)+c_{1}}=(x,y)^{e(I)}(x^{c_{1}}, y^{c_{1}})=(\sqrt{I})^{e(I)}(x^{c_{1}}, y^{c_{1}})\subseteq IJ$ are $(x,y)$-primary ideals. Therefore $\omega(IJ)\le \omega((x,y)^{e(I)+c_{1}})=e(I)+c_{1}=\omega(I)+c_{1}$ by Lemma \ref{orderreversing}. \end{proof} We now classify $\omega$-linear monomial ideals $I$ in $R=k[x,y]$. \begin{lemm} \label{twodim} Let $R=k[x,y]$ and $I=(x^{a_{1}}y^{b_{1}},\cdot\cdot\cdot, x^{a_{r}}y^{b_{r}})$ be a monomial ideal of $R$. Then the following are equivalent. \begin{enumerate} \item $I$ is $\omega$-linear. \item $\omega(I^{n})=n\omega(I)$ for some $n>1$. \item $\omega(I)=\max\{a_{1}+b_{1}, a_{r}+b_{r}\}$. \end{enumerate} \end{lemm} \begin{proof} Note that given $n\in \mathbb{N}$ and a monomial $f$ of $R$, by Lemma \ref{irreducible} we have \begin{align*} \omega(I^{n})&=n\omega(I)\\ \Leftrightarrow n(deg(f))+\omega(I^{n})&=n(deg(f))+n\omega(I)\\ \Leftrightarrow deg(f^{n})+\omega(I^{n})&=n(deg(f)+\omega(I))\\ \Leftrightarrow \omega((fI)^{n})&=n\omega(fI). \end{align*} Moreover, if $I$ is a principal ideal, then $I$ satisfies all of 1,2 and 3 by Corollary \ref{prinmo}. Hence we may assume that $I$ is a $(x,y)$-primary monomial ideal of $R$. That is, $a_{r}=b_{1}=0$.\\ \\ $1\Rightarrow 2$ is trivial.\\ $2\Rightarrow 3$: Suppose that $\omega(I^{n})=n\omega(I)$ for some $n>1$. Note that $\omega(I^{n-1})+\omega(I)\ge \omega(I^{n})=n\omega(I)$ by Lemma \ref{idealop} and $\omega(I^{n-1})\le(n-1)\omega(I)$ by Corollary \ref{idealp}, and thereby $\omega(I^{n-1})=(n-1)\omega(I)$. Hence we must have $\omega(I^{2})=2\omega(I)$. Since $\omega(I^{2})\le \omega(I)+\max\{a_{1}, b_{r}\}$ by Lemma \ref{2dimex}, $\omega(I)=\omega(I^{2})-\omega(I)\le \max\{a_{1}, b_{r}\}$. On the other hand, $\omega(I)\ge\max\{a_{1}, b_{r}\}$ by Lemma \ref{topi}. Therefore $\omega(I)=\max\{a_{1}, b_{r}\}$.\\ $3 \Rightarrow 1$: Follows from Lemma \ref{limitratio}.2. \end{proof} \begin{lemm} \label{intomega} The set of monomial $\omega$-linear ideals of $R=k[x,y]$ is multiplicatively closed. \end{lemm} \begin{proof} Let $I$ and $J$ be monomial $\omega$-linear ideals of $R$. By Lemma \ref{irreducible} we may assume that $I$ and $J$ are $(x,y)$-primary ideals of $R$. Then $\omega(I)= \max\{a_{1}, b_{r}\}$, $\omega(J)=\max\{c_{1}, d_{s}\}$ by Lemma \ref{twodim}. Now, $x^{a_{1}+c_{1}}$ and $y^{b_{r}+d_{s}}$ are elements of $G(IJ)$. Hence by Lemma \ref{limitratio}.2 and Lemma \ref{topi}, to show that $IJ$ is $\omega$-linear it suffices to show that $\omega(IJ)\le\max\{a_{1}+c_{1}, b_{r}+d_{s}\}$. Suppose that $\omega(I)=a_{1}$ and $\omega(J)=c_{1}$. Then all we have to show is $\omega(IJ)\le a_{1}+c_{1}$, which follows from Lemma \ref{idealop}. The case when $\omega(I)=b_{r}$ and $\omega(J)=d_{s}$ can be derived in the exact same manner. Therefore, without loss of generality we may assume that $\omega(I)=a_{1}>b_{r}$ and $\omega(J)=d_{s}>c_{1}$. Observe now that $Ix^{c_{1}} + Jy^{b_{r}}$ is an $(x,y)$-primary ideal contained in $IJ$. Thus by Lemma \ref{orderreversing} and Theorem \ref{hoo} we have \begin{align*} \omega(IJ)&\le \omega(Ix^{c_{1}}+Jy^{b_{r}})\\ &=\max\{\max_{1\le i\le r-1}\{a_{i}+b_{i+1}+c_{1}\}-1, \max_{1\le j\le s-1}\{c_{j}+d_{j+1}+b_{r}\}-1\}\\ &=\max\{ \omega(I)+c_{1}, \omega(J)+b_{r}\}\\ &=\max\{a_{1}+c_{1}, b_{r}+d_{s}\}.\end{align*}\end{proof} Recall that given an ideal $I$ of a commutative ring $R$, an element $f \in R$ is said to be \textit{integral} over $I$ if there is some $k \in \mathbb{N}$ and $c_{i} \in I^{i}$ for each $i \in \{1, \ldots, k\}$ so that \begin{align*} f^{k}+c_{1}f^{k-1}+\cdots + c_{k-1}f+c_{k}&=0. \end{align*} The set of elements of $R$ integral over $I$ is called the \textit{integral closure} of $I$ and denoted by $\overline{I}$. $I$ is said to be \textit{integrally closed} if $I=\overline{I}$. \begin{coro} \label{int} Every integrally closed monomial ideal of $R=k[x,y]$ is $\omega$-linear. \end{coro} \begin{proof} Let $I$ be an integrally closed monomial ideal of $R$. It is well known that $R$ is an \textit{integrally closed domain} (i.e., $R$ is an integral domain that contains every nonzero element of the quotient field of $R$ that is integral over $R$), and that each principal ideal of $R$ is integrally closed, and the product of an integrally closed ideal of $R$ and a nonzero element of $R$ yields another integrally closed ideal of $R$. Hence by Lemma \ref{irreducible} we may assume that $I$ is $(x,y)$-primary. Now by \cite[Proposition 2.6]{Quinonez} there are monomial ideals $I_{1}=(\{x^{r-i}y^{b_{i}}\}_{i=0}^{r})$ and $I_{2}=(\{x^{a_{i}}y^{i}\}_{i=0}^{r})$ of $R$ with $0=b_{0}<b_{1}<\cdot\cdot\cdot< b_{r}$ and $a_{0}>a_{1}>\cdot\cdot\cdot> a_{r}=0$ so $I=I_{1}I_{2}$. Thus by Lemma \ref{intomega}, it suffices to show that $I_{1}$ and $I_{2}$ are $\omega$-linear. By Theorem \ref{hoo}, $\omega(I_{1})=\max_{0\le i\le r-1}\{c_{i}\}$, where $c_{i}=r-i+b_{i+1}-1$ for each $i\in\{0,1,\cdot\cdot\cdot, r-1\}$. Since $c_{i+1}-c_{i}=b_{i+1}-(b_{i}+1)\ge0$ for each $i\in\{0,1,\cdot\cdot\cdot, r-1\}$, we have $\omega(I_{1})=c_{r-1}=b_{r}=\max\{r,b_{r}\}$ and $I_{1}$ is $\omega$-linear by Lemma \ref{twodim}. The proof that $I_{2}$ is $\omega$-linear follows similarly. \end{proof} \begin{rem} \label{rem3} \begin{enumerate} \item Even if $I$ and $J$ are $\omega$-linear monomial primary ideals such that $\sqrt{I}=\sqrt{J}$, we may have $\omega(I\cap J)<\omega(IJ)<\omega(I)+\omega(J)$. Indeed, set $R=k[x,y]$, $I=(x^{3}, xy, y^{2})$ and $J=(x^{2}, xy, y^{3})$. Then both $I$ and $J$ are $\omega$-linear $(x,y)$-primary ideals of $R$. However, $IJ=(x^{5}, x^{3}y, x^{2}y^{2}, xy^{3}, y^{5})$, so $\omega(IJ)=5<6=\omega(I)+\omega(J)$. On the other hand, $\omega(I\cap J)=\max\{\omega(I),\omega(J)\}=3$ by Corollary \ref{primary}. \item Not every $\omega$-linear monomial ideal of $R=k[x,y]$ is integrally closed. For example, set $I=(x^{3}, xy^{2}, y^{4})$. Then $\omega(I)=4$ by Theorem \ref{hoo}, and $I$ is $\omega$-linear by Lemma \ref{twodim}. However, $(x^{2}y)^{2}=x^{3}(xy^{2})\in I^{2}$ and $x^{2}y\not\in I$. Thus $I$ is not integrally closed (\cite[Theorem 1.4.2]{Herzog}). \end{enumerate} \end{rem} So far, we have considered only primary $\omega$-linear monomial ideals, and most of the proof is solely based on the fact that $e(I)=\omega(I)$ when $I$ is a primary ideal. We now show that there exists a class of (integrally closed) nonprimary $\omega$-linear monomial ideals. In fact, some of the squarefree monomial ideals are $\omega$-linear, as we will see in the next two lemmas. Recall that a graph $G$ consists of a set of vertices $V = \{v_{1}, . . . , v_{n}\}$ and a set of edges $E \subseteq \{v_{i}v_{j} | v_{i}, v_{j} \in V \}$, and is called \textit{bipartite} if there exists two disjoint subsets $U_{1}, U_{2}$ of $V$ such that $E\subseteq\{v_{i}v_{j}\mid v_{i}\in U_{1}, v_{j}\in U_{2}\}$. The \textit{edge ideal} of $G$ is defined to be the ideal $I = (\{x_{i}x_{j} | v_{i}v_{j} \in E\})$ of $R = k[x_{1}, . . . , x_{d}]$, where $k$ be a field and $d$ is the number of vertices of $G$. Given a graph $G=(V,E)$, a subset $W$ of $V$ is said to be a \textit{vertex cover} if given $v_{i}v_{j}\in E$, either $v_{i}\in W$ or $v_{j}\in W$. A vertex cover $W$ of $G$ is said to be a \textit{minimal vertex cover} if each proper subset of $W$ is not a vertex cover of $G$. If $I$ is an edge ideal of a graph, then it is a squarefree monomial ideal and a monomial prime ideal $P$ is a minimal ideal of $I$ if and only if the set of vertices that corresponds to $P$ is a minimal vertex cover. Also, a graph is bipartite if and only if it has no cycle of odd length as its subgraph. Our first example of a nonprimary $\omega$-linear ideal is the edge ideal of a bipartite graph. \begin{lemm} \label{bipartite} Let $R=k[x_{1},\cdot\cdot\cdot, x_{n}]$. If $I$ is an ideal of $R$ that is also the edge ideal of a bipartite graph $G$, then $I$ is $\omega$-linear. \end{lemm} \begin{proof} Let $I$ be an edge ideal of a graph $G$ and let $P_{1}, \cdot\cdot\cdot, P_{r}$ be the set of (incomparable) minimal prime ideals of $I$. Recall that a graph $G$ is bipartite if and only if \[I^{m}=\bigcap_{P \text{ is a minimal prime of $I$}}P^{m}\] for each $m\in\mathbb{N}$ (\cite[Theorem 5.9]{Simis}). Hence if $G$ is bipartite, then by Lemma \ref{upper}, $\omega(I^{m})=\sum\limits_{i=1}^{r} e(P_{i}^{m})=\sum\limits_{i=1}^{r}m=mr$ for each $m \in\mathbb{N}$. Therefore the conclusion follows. \end{proof} There are nonbipartite graphs whose edge ideals are $\omega$-linear. \begin{lemm} \label{hum} Let $R=k[x_{1},\cdot\cdot\cdot, x_{n}]$. Let $I=(x_{1}x_{2}, x_{2}x_{3},\cdot\cdot\cdot, x_{n-1}x_{n}, x_{n}x_{1})$ (that is, $I$ is the edge ideal of a cycle graph of length $n$). Then $I$ is $\omega$-linear. \end{lemm} \begin{proof} Since a cycle of even length is bipartite, by Lemma \ref{bipartite} we may assume that $n=2l+1$ for some $l\in\mathbb{N}$. Fix $m\in \mathbb{N}$. $I$ is a squarefree monomial ideal, so $I=P_{1}\cap\cdot\cdot\cdot \cap P_{r}$ where $P_{1},\cdot\cdot\cdot P_{r}$ are the minimal prime ideals of $I$ (\cite[Lemma 1.3.5]{Herzog}). Thus by Lemma \ref{upper} we have $\omega(I)=\sum_{i=1}^{r}e(P_{i})=r$, and we only need to show that $\omega(I^{m})=mr$. Note that since $I$ is an edge ideal of a cycle of length $2l+1$, Ass$(R/I^{m})=\{P_{1},\cdot\cdot\cdot, P_{r}\}$ if $m\le l$ and Ass$(R/I^{m})=\{P_{1},\cdot\cdot\cdot, P_{r}, \mathfrak{m}\}$ if $m>l$ (\cite[Lemma 3.1]{Chen}). Hence if $m\le l$, then $I^{m}=\cap_{i=1}^{r}P_{i}^{m}$ and $\omega(I^{m})=\sum\limits_{i=1}^{r}e(P_{i}^{m})=mr$ by Lemma \ref{upper}, so we are done. Assume that $m>l$. Then $I^{m}=(\cap_{i=1}^{r}P_{i}^{m})\cap Q$ is the canonical primary decomposition of $I^{m}$, where $Q$ is an $\mathfrak{m}$-primary monomial ideal of $R$ (\cite[Proposition 1.4.4]{Herzog}). Now, $Q=(x_{1}^{a_{1}},\cdot\cdot\cdot, x_{n}^{a_{n}}, f_{1},\cdot\cdot\cdot, f_{t})$ for some $a_{i} \in \mathbb{N}$ and monomials $f_{i}$. Since $I$ is a squarefree monomial ideal and $Q$ is a primary component of $I^{m}$, we must have $a_{i}\le m$ for each $i\in\{1,\cdot\cdot\cdot, n\}$ and then $e(Q)\le e((x_{1}^{a_{1}},\cdot\cdot\cdot, x_{n}^{a_{n}}))\le mn-n+1\le mr$ by Lemma \ref{powermonomial} and since $n \leq r$. It follows that $\omega(I^{m})=\max\{\sum\limits_{i=1}^{r}e(P_{i}^{m}), e(Q)\}=\max\{mr, e(Q)\}=mr$ by Lemma \ref{l} and Lemma \ref{upper}. \end{proof} We were unable to show whether every edge ideal is $\omega$-linear, but we do have the following. \begin{lemm} Let $I$ be a square-free monomial ideal. Then $\omega(I^{m})\ge m\omega(I)$ for each $m\in\mathbb{N}$. \end{lemm} \begin{proof} Let $P_{1},\cdot\cdot\cdot, P_{r}$ be minimal prime ideals of $I$. Then $I=\cap_{i=1}^{r}P_{i}$ and $\omega(I)=r$ by Lemma \ref{upper}. Set $f_{i}=\sum\limits_{x_{j}\in G(P_{i})}x_{j}$ for each $i\in\{1,\cdot\cdot\cdot, r\}$. Then $f:=\prod_{i=1}^{r}f_{i}\in \prod_{i=1}^{r}P_{i}\subseteq I$, so $f^{m}\in I^{m}$. However, $\cfrac{f^{m}}{f_{i}}\not\in P_{i}^{m}$, so $\cfrac{f^{m}}{f_{i}}\not\in I^{m}$(\cite[Proposition 1.4.4]{Herzog}). Thus $I^{m}$ is not $(mr-1)$-absorbing and $\omega(I^{m})\ge m\omega(I)$. \end{proof} We close the section with the following question: Is every integrally closed monomial ideal $\omega$-linear? Integrally closed monomial ideals considered in this note (monomial ideals in $R=k[x,y]$, irreducible monomial ideals, or edge ideal of bipartite graphs) were all $\omega$-linear. Note also that if this question has an affirmative answer, then it follows that every edge ideal is $\omega$-linear.
{ "timestamp": "2018-12-17T02:07:52", "yymm": "1812", "arxiv_id": "1812.05791", "language": "en", "url": "https://arxiv.org/abs/1812.05791", "abstract": "In a commutative ring $R$ with unity, given an ideal $I$ of $R$, Anderson and Badawi in 2011 introduced the invariant $\\omega(I)$, which is the minimal integer $n$ for which $I$ is an $n$-absorbing ideal of $R$. In the specific case that $R = k[x_{1}, \\ldots, x_{n}]$ is a polynomial ring over a field $k$ in $n$ variables $x_{1},\\ldots, x_{n}$, we calculate $\\omega(I)$ for certain monomial ideals $I$ of $R$.", "subjects": "Commutative Algebra (math.AC)", "title": "$n$-absorbing monomial ideals in polynomial rings", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9848109489630493, "lm_q2_score": 0.8311430478583168, "lm_q1q2_score": 0.8185187736853902 }
https://arxiv.org/abs/1312.5369
Sign patterns of rational matrices with large rank
Let $A$ be a real matrix. The term rank of $A$ is the smallest number $t$ of lines (that is, rows or columns) needed to cover all the nonzero entries of $A$. We prove a conjecture of Li et al. stating that, if the rank of $A$ exceeds $t-3$, there is a rational matrix with the same sign pattern and rank as those of $A$. We point out a connection of the problem discussed with the Kapranov rank function of tropical matrices, and we show that the statement fails to hold in general if the rank of $A$ does not exceed $t-3$.
\section{Introduction} The problem of constructing a matrix over a given ordered field with specified sign pattern and rank deserved a significant amount of attention in recent publications, see~\cite{LiEtAl} and references therein. The present paper establishes a connection of this problem with that of computing certain rank functions arisen from tropical geometry. We prove the conjecture on sign patterns of rational matrices formulated in~\cite{LiEtAl}, and we present the examples showing the optimality of our result. \section{Preliminaries} The following notation is used throughout our paper. By $U^{m\times n}$ we denote the set of all $m$-by-$n$ matrices with entries from a set $U$, by $A_{ij}\in U$ we denote an entry of a matrix $A\in U^{m\times n}$. By $U_{(i)}$ we denote the $i$th row of $U$, and we call a \textit{line} of a matrix any of its columns or rows. A field $R$ is called \textit{ordered} if, for some subset $P\subset R$ closed under addition and multiplication, the sets $P$, $-P$, and $\{0\}$ form a partition of $R$. The elements of $P$ are then called \textit{positive}, and those from $-P$ \textit{negative}. The \textit{sign pattern} of a matrix $A\in R^{m\times n}$ is the matrix $S={\mathcal S}(A)\in\{+,-,0\}^{m\times n}$ defined as $S_{ij}=+$ if $A_{ij}$ is positive, $S_{ij}=-$ if $A_{ij}$ is negative, and $S_{ij}=0$ if $A_{ij}=0$. The \textit{minimum rank} of a sign pattern $S$ with respect to $R$ is the minimum of the ranks of matrices $B$ over $R$ satisfying ${\mathcal S}(B)=S$. There are a significant number of recent publications devoted to the study of the minimal ranks of sign patterns (see~\cite{LiEtAl} and references therein), and our paper aims to prove a conjecture formulated in~\cite{LiEtAl}. This conjecture relates the minimal rank of a pattern with a concept of the \textit{term rank} of a matrix, which is defined as the smallest number of lines needed to include all the nonzero elements of that matrix. The classical \textit{K\"{o}nig's theorem} states the the term rank of a matrix $A$ equals the maximum number of nonzero entries of $A$ no two of which belong to the same line, so the term rank of a sign pattern $S$ can be thought of as the maximum of the ranks of matrices $C$ over $R$ satisfying ${\mathcal S}(C)=S$. Now we can formulate the conjecture by Li et al.\,\,relating the concepts of minimum and term ranks for sign pattern matrices. \begin{con}\cite[Conjecture 4.2]{LiEtAl}\label{conLiEtAl} Assume $S$ is a sign pattern matrix with term rank equal to $t$, and let $r$ be the minimum rank of $S$ over the reals. If $r\geqslant t-2$, then the minimum rank of $S$ over the rationals is $r$ as well. \end{con} In Section~3 we develop a combinatorial technique which allows to prove Conjecture~\ref{conLiEtAl}. In Section~4 we establish the connection of the problem discussed with the Kapranov rank function of Boolean matrices introduced in~\cite{DSS}. We also make the use of matroid theory to prove the optimality of the bound in Conjecture~\ref{conLiEtAl} by showing that its statement fails to hold in general if $r$ is less than $t-2$. \section{Proof of the result} We start with two easy observations helpful for further considerations. \begin{observ}\label{observsign2} Multiplying a row of a real matrix $A$ by a nonzero number will not change the minimal ranks of its sign pattern. \end{observ} \begin{proof} Trivial. \end{proof} \begin{observ}\label{observsign1} Let $r$ and $t$ be, respectively, the minimum and maximum ranks of a sign pattern $S$ with respect to an ordered field $R$. Then, for any integer $h\in[r,t]$, there is a matrix over $R$ which has rank $h$ and sign pattern $S$. \end{observ} \begin{proof} Changing a single entry produces a matrix whose rank differs by at most $1$ from that of the initial matrix. \end{proof} The following lemma gives a useful description of the rank of a block matrix. We say that a linear subspace $S\subset\mathbb{R}^d$ is \textit{rational} if $S$ has a basis consisting of vectors that have rational coordinates only. \begin{lem}\label{lemsignblock} Let $V_1\in\mathbb{Q}^{p\times(p-1)}$ and $V_2\in\mathbb{Q}^{(q-1)\times q}$ be rational matrices that have ranks $p-1$ and $q-1$, respectively. Then the set $\mathcal{W}$ of all $W\in\mathbb{R}^{p\times q}$ for which the matrix $U=\left(\begin{smallmatrix} W & V_1 \\ V_2 & O \end{smallmatrix}\right)$ has rank $p+q-2$ is a rational subspace. \end{lem} \begin{proof} Note that rational elementary transformations on the first $p$ rows or first $q$ columns of $U$ can not break the property of $\mathcal{W}$ to be a rational subspace. So we can assume that $V_1$ and $V_2$ differ from the identity matrices by adding the zero column and row, which case is easy. \end{proof} Now we are ready to prove Conjecture~\ref{conLiEtAl} in a special case. \begin{lem}\label{lemsign2} For any real $m$-by-$n$ matrix $A$ of rank $n-2$, there is a rational $m$-by-$n$ matrix which has rank $n-2$ and sign pattern equal to that of $A$. \end{lem} \begin{proof} By the assumptions, there is a rank-two matrix $B\in\mathbb{R}^{n\times2}$ for which the matrix $AB$ is zero. Observation~\ref{observsign2} allows one to assume that the first column of $B$ consists of zeros and ones. Let $X$ be a matrix whose $(i,j)$th entry is a variable if $A_{ij}\neq0$ and $X_{ij}=0$ otherwise. For a sufficiently large integer $N>0$, we set $C_{jk}=[NB_{jk}]/N$. Note that, for every row index $i$, the matrix formed by the rows of $B$ with indexes $j$ satisfying $A_{ij}\neq 0$ has the same rank as the matrix formed by the rows of $C$ with the same indexes. For every $i$, we assign to every free variable $X_{ig}$ of the linear system $X_{(i)}C=(0\,0)$ the value $[NA_{ig}]/N$. Solving those systems, we get as a solution a rational matrix $X=X(N)$ which satisfies $XC=0$. Since $X(N)\rightarrow A$ as $N\rightarrow\infty$, the matrices $X(N)$ and $A$ have the same sign pattern for sufficiently large $N$. \end{proof} Now let us prove the key result of the section. \begin{thr}\label{thrmaintermrealis0} Let $A$ be a real matrix with term rank equal to $t$. If the rank of $A$ equals $t-2$, then there is a rational matrix which has rank $t-2$ and the same sign pattern as $A$. \end{thr} \begin{proof} 1. Up to row and column permutations, $A$ is an $n$-by-$m$ matrix of the form $\left(\begin{smallmatrix} B & C \\ D & O \end{smallmatrix}\right)$, where the matrix $B\in\mathbb{R}^{p\times q}$ satisfies $p+q=t$, and $O$ is the zero matrix. If the rank of $D$ is less than $q-1$, then by Lemma~\ref{lemsign2} we can construct a rational matrix $D'$ of rank $q-2$ with the same sign pattern as $D$. Choosing $B'$ and $C'$ as arbitrary matrices with sign patterns equal to those of $B$ and $C$, we get that the rank of $\left(\begin{smallmatrix} B' & C' \\ D' & O \end{smallmatrix}\right)$ is at most $t-2$, and we are done. We can assume in what follows that $D$ has rank at least $q-1$ and, similarly, that $C$ has rank at least $p-1$. Since the rank of $A$ is $t-2=p+q-2$, we conclude that the rank of $D$ exactly equals $q-1$ and the rank of $C$ is $p-1$. 2. By Step~1, the rows of $C$ are linearly dependent, and we can assume by Observation~\ref{observsign2} that the coefficients of this linear dependence are rational. In other words, the columns of $C$ generate a rational subspace in $\mathbb{R}^p$. Since rational points are dense in rational subspaces, we can assume that the matrix $C$ (and the matrix $D$, similarly) consists of rational numbers, in which case the result follows from Lemma~\ref{lemsignblock}. \end{proof} Now we are ready to prove Conjecture~\ref{conLiEtAl}. \begin{thr}\label{thrmaintermrealis} Let $A$ be a real matrix with term rank equal to $t$. If the rank of $A$ is at least $t-2$, then there is a rational matrix which has the same sign pattern and rank as those of $A$. \end{thr} \begin{proof} Note that adding a repeating row does not affect the rank of a matrix, and the term rank of the matrix obtained is either equal to or greater by one than that of the initial matrix. Therefore, adding a sufficient number of repeating rows to $A$, we get a matrix $A'$ satisfying the assumptions of Theorem~\ref{thrmaintermrealis0}. So we can find a rational matrix $B$ which has the same sign pattern as that of $A$ and rank not exceeding the rank of $A$. Now the result follows from Observation~\ref{observsign1}. \end{proof} \section{Optimality of the result} To construct sign patterns of term rank $t$ realizable by real matrices of rank $t-3$ but not by rational matrices of that rank, we need to recall the definition of another rank concept. For ${\mathbb F}$ a field, define the \textit{Kapranov rank} of a matrix $B\in\{0,1\}^{m\times n}$ with respect to ${\mathbb F}$ as the smallest possible rank of a matrix $C\in{\mathbb F}^{m\times n}$ satisfying $C_{ij}=0$ if and only if $B_{ij}=0$. The following lemma points out a connection between the quantity introduced (which we denote by ${\mathrm K}_{\mathbb F}(B)$ in what follows) and the problem of pattern realisability. \begin{lem}\label{lemKapreal} Assume $R_1$ is an ordered field, and a matrix $B\in\{0,1\}^{m\times n}$ satisfies $r={\mathrm K}_{R_1}(B)<{\mathrm K}_{R_2}(B)$ for any field $R_2$ strictly contained in $R_1$. Then there is a sign pattern $S\in\{0,+,-\}^{m\times n}$ realizable by a matrix over $R_1$ of rank $r$ but not by a matrix over $R_2$ of that rank. \end{lem} \begin{proof} By definition of Kapranov rank, there is a matrix $A\in{R_1}^{m\times n}$ which has rank $r$ and satisfies $A_{ij}=0$ if and only if $B_{ij}=0$. Denoting the sign pattern of $A$ by $S$, one can see that $S$ is not realizable by a matrix over $R_2$ of rank $r$. \end{proof} Now we see that a sign pattern realizable over $R_1$ but not over $R_2$ always exists if we have a zero-one matrix whose Kapranov rank over $R_2$ is greater than that over $R_1$. It turns out that producing zero-one matrices with this property can be performed by the use of matroid theory, and let us recall the basic definitions of this theory~\cite{Ox}. A \textit{matroid} $M$ on a finite set $E$ is defined by the set $\mathcal{B}\subset 2^E$ of its \textit{bases}, which are supposed to satisfy the following conditions: (1) $\mathcal{B}\neq\varnothing$; (2) if $A,B\in \mathcal{B}$ and $a\in A\setminus B$, then there is a $b\in B\setminus A$ such that $A\setminus\{a\}\cup\{b\}\in I$. All the bases can easily be shown to have the same cardinality, and this cardinality is called the \textit{rank} of a matroid $M$. A \textit{circuit} in $M$ is a minimal set which is a subset of no $B\in\mathcal{B}$. A \textit{dual matroid} $M^*$ has as its bases the complements of the bases of $M$, and a circuit in $M^*$ is called a \textit{cocircuit} for $M$. The matroid $M$ is \textit{representable} over a field ${\mathbb F}$ if we can assign vectors from ${\mathbb F}^d$ to the elements of $E$ in such a way that a set $B$ is a basis of the linear span of $E$ if and only if $B\in\mathcal{B}$. Finally, define a cocircuit matrix $\mathcal{C}=\mathcal{C}_M$ of $M$ as a matrix with rows indexed by elements of $E$ and columns indexed by cocircuits such that $\mathcal{C}_{ij}=1$ if $i$ belongs to the $j$th cocircuit and $\mathcal{C}_{ij}=0$ otherwise. The following theorem allows one to construct matrices whose Kapranov rank depends on a ground field. \begin{thr}\cite[Proposition~7.2 and Theorem~7.3]{DSS}\label{thrKapMatr} If $M$ is a matroid of rank $r$ and $\mathcal{C}$ its cocircuit matrix, then ${\mathrm K}_{\mathbb F}(\mathcal{C})\geqslant r$ for any field ${\mathbb F}$. For ${\mathbb F}$ infinite, the condition ${\mathrm K}_{\mathbb F}(\mathcal{C})=r$ holds if and only if $M$ is representable over ${\mathbb F}$. \end{thr} The following well-known fact connects the notions of matroid duality and representability. \begin{thr}\cite{Ox}\label{thrmatrdual} If a matroid $M$ is representable over a field ${\mathbb F}$, then so is its dual $M^*$. \end{thr} Note that the matroid duality is an involution, that is, the condition $(M^*)^*=M$ holds. This shows that Theorem~\ref{thrmatrdual} holds as well in the opposite direction. We also need the classical example of a non-representable matroid, which appeared in a foundational paper by Saunders MacLane~\cite{MacLane}. \begin{thr}\cite[Theorem~3]{MacLane}\label{rank3repres} Let $K$ be a finite algebraic field over the field of rational numbers. Then there exists a matroid $M$ of rank $3$ which is representable over $K$ but over no field strictly contained in $K$. \end{thr} Now we are ready to prove the theorem stating that the bound of $t-2$ is optimal in Theorem~\ref{thrmaintermrealis}. \begin{thr} Let $K$ be an ordered finite algebraic field over the field of rational numbers. Then there exists a matrix $A\in K^{n\times m}$ of rank $n-3$ with the following property: the entries of any matrix $A'\in K^{n\times m}$ which has the same rank and sign pattern as those of $A$ generate the whole field $K$. \end{thr} \begin{proof} Using Theorem~\ref{rank3repres}, we get a rank-three matroid $M$ representable over $K$ but not over any field strictly contained in $K$. Denoting the number of its vertices by $n$, we see that by definitions the dual $M^*$ has rank $n-3$. Then we use Theorem~\ref{thrmatrdual} to conclude that $M^*$ is representable over $K$ but not over any field strictly contained in $K$. From Theorem~\ref{thrKapMatr} it follows that the cocircuit matrix of $M^*$, which has $n$ rows, has also Kapranov rank $n-3$ with respect to $K$ and greater Kapranov rank with respect to any field strictly contained in $K$. Application of Lemma~\ref{lemKapreal} now completes the proof. \end{proof}
{ "timestamp": "2013-12-20T02:01:57", "yymm": "1312", "arxiv_id": "1312.5369", "language": "en", "url": "https://arxiv.org/abs/1312.5369", "abstract": "Let $A$ be a real matrix. The term rank of $A$ is the smallest number $t$ of lines (that is, rows or columns) needed to cover all the nonzero entries of $A$. We prove a conjecture of Li et al. stating that, if the rank of $A$ exceeds $t-3$, there is a rational matrix with the same sign pattern and rank as those of $A$. We point out a connection of the problem discussed with the Kapranov rank function of tropical matrices, and we show that the statement fails to hold in general if the rank of $A$ does not exceed $t-3$.", "subjects": "Combinatorics (math.CO)", "title": "Sign patterns of rational matrices with large rank", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9873750492324249, "lm_q2_score": 0.8289388083214155, "lm_q1q2_score": 0.8184734966770253 }
https://arxiv.org/abs/1206.4111
On the numerical stability of Fourier extensions
An effective means to approximate an analytic, nonperiodic function on a bounded interval is by using a Fourier series on a larger domain. When constructed appropriately, this so-called Fourier extension is known to converge geometrically fast in the truncation parameter. Unfortunately, computing a Fourier extension requires solving an ill-conditioned linear system, and hence one might expect such rapid convergence to be destroyed when carrying out computations in finite precision. The purpose of this paper is to show that this is not the case. Specifically, we show that Fourier extensions are actually numerically stable when implemented in finite arithmetic, and achieve a convergence rate that is at least superalgebraic. Thus, in this instance, ill-conditioning of the linear system does not prohibit a good approximation.In the second part of this paper we consider the issue of computing Fourier extensions from equispaced data. A result of Platte, Trefethen & Kuijlaars states that no method for this problem can be both numerically stable and exponentially convergent. We explain how Fourier extensions relate to this theoretical barrier, and demonstrate that they are particularly well suited for this problem: namely, they obtain at least superalgebraic convergence in a numerically stable manner.
\section{Introduction}\label{s:introduction} Let $f : [-1,1] \rightarrow \bbR$ be an analytic function. When periodic, an extremely effective means to approximate $f$ is via its truncated Fourier series. This approximation converges geometrically fast in the truncation parameter $N$, and can be computed efficiently via the Fast Fourier Transform (FFT). Moreover, Fourier series possess high resolution power. One requires an optimal $2$ modes per wavelength to resolve oscillations, making Fourier methods well suited for (most notably) PDEs with oscillatory solutions \cite{naspec}. For these reasons, Fourier series are extremely widely used in practice. However, the situation changes completely when $f$ is nonperiodic. In this case, rather than geometric convergence, one witnesses the familiar Gibbs phenomenon near $x = \pm 1$ and only linear pointwise convergence in $(-1,1)$. \subsection{Fourier extensions}\label{ss:FE} For analytic and nonperiodic functions, one way to restore the good properties of a Fourier series expansion (in particular, geometric convergence and high resolution power) is to approximate $f$ with a Fourier series on an \textit{extended} domain $[-T,T]$. Here $T>1$ is a user-determined parameter. Thus we seek an approximation $F_N(f)$ to $f$ from the set \bes{ \cG_{N} := \spn \left \{ \phi_n : |n| \leq N \right \},\qquad \phi_n(x) : = \frac{1}{\sqrt{2T}} \E^{\I \frac{n \pi}{T} x }. } Although there are many potential ways to define $F_N(f)$, in \cite{BoydFourCont,brunoFEP,huybrechs2010fourier} it was proposed to compute $F_N(f)$ as the best approximation to $f$ on $[-1,1]$ in a least squares sense: \be{ \label{L2} F_N(f) : = \underset{\phi \in \cG_N}{\operatorname{argmin}} \| f - \phi \| . } Here $\nm{\cdot}$ is the standard norm on $\rL^2(-1,1)$---the space of square-integrable functions on $[-1,1]$. Henceforth, we shall refer to $F_N(f)$ as the \textit{continuous} Fourier extension (FE) of $f$. In \cite{BADHFEResolution,huybrechs2010fourier} it was shown that the continuous FE $F_N(f)$ converges geometrically fast in $N$ and has a resolution constant (number of degrees of freedom per wavelength required to resolve an oscillatory wave) that ranges between $2$ and $\pi$ depending on the choice of the parameter $T$, with $T \approx 1$ giving close to the optimal value $2$ (see \S\ref{ss:Tchoice} for a discussion). Thus the continuous FE successfully retains the key properties of rapid convergence and high resolution power of a standard Fourier series in the case of nonperiodic functions. We note that one does not usually compute the continuous FE \R{L2} in practice. A more convenient approach \cite{BADHFEResolution,huybrechs2010fourier} is to replace \R{L2} by the discrete least squares \be{ \label{discL2} \tilde{F}_N(f) : = \underset{\phi \in \cG_N}{\operatorname{argmin}} \sum_{|n| \leq N} | f(x_n) - \phi(x_n) |^2, } for nodes $\{ x_n \}_{|n| \leq N} \subseteq [-1,1]$. We refer to $\tilde{F}_N(f)$ as the \textit{discrete} Fourier extension of $f$. When chosen suitably---in particular, as in \R{mapCheb}---such nodes ensure that the difference in approximation properties between the extensions \R{L2} and \R{discL2} is minimal (for details, see \S \ref{s:FEtypes}). \subsection{Numerical convergence and stability of Fourier extensions}\label{ss:intro_summary} The approximation properties of the continuous and discrete FEs were analyzed in \cite{BADHFEResolution,huybrechs2010fourier}. Therein it was also observed numerically that the condition numbers of the matrices $A$ and $\tilde{A}$ of the least squares \R{L2} and \R{discL2} are exponentially large in $N$. We shall confirm this observation later in the paper. Thus, if $a = (a_{-N},\ldots,a_{N})^{\top}$ is the vector of \textit{coefficients} of the continuous or discrete FE, i.e.\ $F_N(f)$ or $\tilde{F}_N(f)$ is given by $\sum_{|n| \leq N} a_n \phi_n$, one expects small perturbations in $f$ to lead to large errors in $a$. In other words, the computation of the coefficients of the continuous or discrete FE is ill-conditioned. Because of this ill-conditioning, it is tempting to think that FEs will be useless in applications. At first sight it is reasonable to expect that the good approximation properties of \textit{exact} FEs (i.e.\ those obtained in exact arithmetic) will be destroyed when computing \textit{numerical} FEs in finite precision. However, previous numerical studies \cite{BADHFEResolution,BoydFourCont,brunoFEP,huybrechs2010fourier,LyonFESVD,LyonFast} indicate otherwise. Despite very large condition numbers, one typically obtains an extremely good approximation with a numerical FE, even for poorly behaved functions and in the presence of noise. The aim of this paper is to give a full explanation of this phenomenon. This explanation can be summarized as follows. In computations, one's interest does not lie with the accuracy in computing the coefficient vector $a$, but rather the accuracy of the numerical FE approximation $\sum_{|n| \leq N} a_n \phi_n$. As we show, although the mapping from a function to its coefficients is ill-conditioned, the mapping from $f$ to its numerical FE is, in fact, well-conditioned. In other words, whilst the small singular values of $A$ (or $\tilde{A}$) have a substantial effect on $a$, they have a much less significant, and completely quantifiable, effect on the FE itself. Although this observation explains the apparent stability of numerical FEs, it does not address their approximation properties. In \cite{BADHFEResolution,huybrechs2010fourier} it was shown that the exact continuous and discrete FEs $F_N(f)$ and $\tilde{F}_N(f)$ converge geometrically fast in $N$. However, the fact that there may be substantial differences between the coefficients of $F_N(f)$, $\tilde{F}_N(f)$ and those of the numerical FEs, which henceforth we denote by $G_N(f)$ and $\tilde{G}_N(f)$, suggests that geometric convergence may not be witnessed in finite arithmetic for large $N$. As we show later, for a large class of functions, geometric convergence of $F_N(f)$ (or $\tilde{F}_N(f)$) is typically accompanied by geometric growth of the norm $\nm{a}$ of the exact (infinite-precision) coefficient vector. Hence, whenever $N$ is sufficiently large, one expects there to be a discrepancy between the exact coefficient vector and its numerically computed counterpart, meaning that the numerical extensions $G_N(f)$ and $\tilde{G}_N(f)$ may not exhibit the same convergence behaviour. In the first half of this paper, besides showing stability, we also give a complete analysis and description of the convergence of $G_N(f)$ and $\tilde{G}_N(f)$, and discuss how this differs from that of $F_N(f)$ and $\tilde{F}_N(f)$. We now summarize the main conclusions of the first half of the paper. Concerning stability, we have: \begin{enumerate} \item The condition numbers of the matrices $A$ and $\tilde{A}$ of the continuous and discrete FEs are exponentially large in $N$ (see \S \ref{ss:condnumb}). \item The condition number $\kappa(F_N)$ of the exact continuous FE mapping is exponentially large in $N$. The condition number of the exact discrete FE mapping satisfies $\kappa(\tilde{F}_N) = 1$ for all $N$ (see \S \ref{ss:cond_numb_exact}). \item The condition number of the numerical continuous and discrete FE mappings $G_N$ and $\tilde{G}_N$ satisfy \bes{ \kappa(G_N) \lesssim 1/\sqrt{\epsilon},\quad \kappa(\tilde{G}_N) \lesssim 1,\qquad \forall N \in \bbN, } where $\epsilon = \epsilon_{\mathrm{mach}}$ is the machine precision used (see \S \ref{ss:stabfn}). \end{enumerate} To state our main conclusions regarding convergence, we first require some notation. Let $\cD(\rho)$, $\rho \geq 1$, be a particular one-parameter family of regions in the complex plane related to Bernstein ellipses (see \R{Mapped_Bernstein} and Definition \ref{d:Bernstein}), and define the Fourier extension \textit{constant} \cite{BADHFEResolution,huybrechs2010fourier} by \be{ \label{cteET} E(T) = \cot^2 \left ( \frac{\pi}{4 T} \right ). } We now have the following: \begin{enumerate} \item[4.] Suppose that $f$ is analytic in $\cD(\rho^*)$ and continuous on its boundary. Then the exact continuous and discrete FEs satisfy \bes{ \| f - F_N(f) \|, \ \| f - \tilde{F}_{N} f \| \leq c_f \rho^{-N}, } where $\rho = \min \left \{ \rho^*,E(T) \right \}$ and $c_f$ is proportional to $\max_{x \in \cD(\rho)} | f(x) |$ (see \S \ref{s:FEconv}). \item[5.] For $f$ as in 4.\ the errors of the numerical continuous and discrete FEs satisfy (see \S \ref{ss:numsolnanalysis}): \enum{ \item[(i)] For $N \leq N_0$ (continuous) or $N \leq N_1 : = 2 N_0$ (discrete), where $N_0$ is a function-independent breakpoint depending on $\epsilon$ and $T$ only, both $\| f - G_N(f) \|$ and $\| f - \tilde{G}_N f \|$ decay like $\rho^{-N}$, where $\rho$ is as in 4. \item[(ii)] When $N = N_0$ or $N = N_1$, the errors \bes{ \| f - G_{N_0}(f) \| \approx c_f (\sqrt{\epsilon})^{d_f},\quad \| f - \tilde{G}_{N_1}(f) \| \approx c_f \epsilon^{d_f}, } where $c_f$ is as in 4.\ and $d_f = \frac{\log \rho}{\log E(T)} \in (0,1]$. \item[(iii)] When $N>N_0$ or $N > N_1$, the errors decay at least superalgebraically fast down to \textit{maximal achievable accuracies} of order $\sqrt{\epsilon}$ and $\epsilon$ respectively. In other words, \bes{ \limsup_{N \rightarrow \infty} \| f - G_{N}(f) \| \lesssim \sqrt{\epsilon},\qquad \limsup_{N \rightarrow \infty} \| f - \tilde{G}_{N}(f) \| \lesssim \epsilon. } } \end{enumerate} \rem{ In this paper we refer to several different types of convergence of an approximation $f_N \approx f$. We say that $f_N$ converges \textit{algebraically} fast to $f$ at rate $k$ if $\| f - f_N \| = \ord{N^{-k}}$ as $N \rightarrow \infty$. If $\nm{f-f_N}$ decays faster than any algebraic power of $N^{-1}$ then $f_N$ is said to converge \textit{superalgebraically} fast. We say that $f_N$ converges \textit{geometrically} fast to $f$ if there exists a $\rho > 1$ such that $\nm{f-f_N} = \ord{\rho^{-N}}$. We shall also occasionally use the term \textit{root-exponential} to describe convergence of the form $\nm{f-f_N} = \ordu{\rho^{-\sqrt{N}}}$. } As we explain in \S \ref{s:stability}, the reason for the disparity between the exact and numerical FEs can be traced to the fact that the system of functions $\{ \E^{\I \frac{n \pi}{T} \cdot} \}_{n \in \bbZ}$ forms a \textit{frame} for $\rL^2(-1,1)$. The inherent redundancy of this frame, i.e.\ the fact that any function $f$ has infinitely many expansions in this system, leads to both the ill-conditioning in the coefficients, as well as the differing convergence between the exact and numerical approximations $F_N$, $\tilde{F}_N$ and $G_N$, $\tilde{G}_N$ respectively. This aside, observe that conclusion 5.\ asserts that the numerical continuous FE $G_N(f)$ converges geometrically fast in the regime $N < N_0$ down to an error of order $(\sqrt{\epsilon})^{d_f}$, and then at least superalgebraically fast for $N > N_0$ down to a best achievable accuracy of order $\sqrt{\epsilon}$. Note that $d_f = 1$ whenever $f$ is analytic in $\cD(\rho)$ with $\rho \geq E(T)$. Thus $G_N$ approximates all sufficiently analytic functions possessing moderately small constants $c_f$ with geometric convergence down to order $\sqrt{\epsilon}$, and this is achieved at $N = N_0$. For functions only analytic in regions $\cD(\rho)$ with $\rho < E(T)$, or possessing large constants $c_f$, this accuracy is obtained after a further regime of at least superalgebraic convergence. Note that $c_f$ is large typically when $f$ is oscillatory or possessing boundary layers. Hence for such functions, even though they may well be entire, one usually still sees the second phase of superalgebraic convergence. The limitation of $\sqrt{\epsilon}$ accuracy for the numerical continuous FE is undesirable. Since $\epsilon = \epsilon_{\mathrm{mach}} \approx 10^{-16}$ in practice, this means that one cannot expect to obtain more than $7$ or $8$ digits of accuracy in general. The condition number is also large---specifically, $\kappa(G_N) \approx 10^{8}$ (see 3.)---and hence the continuous FE has limited practical value. This is in addition to $G_N(f)$ being difficult to compute in practice, since it requires calculation of $2N+1$ Fourier integrals of $f$ (see \S \ref{sss:continuous}). On the other hand, conclusion 3.\ shows that the discrete FE is completely stable when implemented numerically. Moreover, it possesses the same qualitative convergence behaviour as the continuous FE, but with two key differences. First, the region of guaranteed geometric convergence is precisely twice as large, $N_1 = 2 N_0$. Second, the maximal achievable accuracy is on the order of machine precision, as opposed to its square root (see 5.). Thus, an important conclusion of the first half of this paper is the following: it is possible to compute a numerically stable FE of any analytic function which converges at least superalgebraically fast in $N$ (in particular, geometrically fast for all small $N$), and which attains close to machine accuracy for $N$ sufficiently large. \rem{ \label{r:numsolver} This paper is about the discrepancy between theoretical properties of solutions to \R{L2} and \R{discL2} and their numerical solutions when computed with standard solvers. Throughout we shall consistently use \textit{Mathematica}'s \texttt{LeastSquares} routine in our computations, though we would like to stress that \textit{Matlab}'s command $\backslash$ gives similar results. Occasionally, to compare theoretical and numerical properties, we shall carry out computations in additional precision to eliminate the effect of round-off error. When done, this will be stated explicitly. Otherwise, it is to be assumed that all computations are carried out as described in standard precision. } \subsection{Fourier extensions from equispaced data}\label{ss:intro_summary_equispaced} In many applications, one is faced with the problem of recovering an analytic function $f$ to high accuracy from its values on an equispaced grid $\left \{ f \left ( \tfrac{n}{M} \right ) : n=-M,\ldots,M \right \}$. This problem turns out to be quite challenging. For example, the famous Runge phenomenon states that the polynomial interpolant of this data will diverge geometrically fast as $M \rightarrow \infty$ unless $f$ is analytic in a sufficiently large region. Numerous approaches have been proposed to address this problem, and thereby `overcome' the Runge phenomenon (see \cite{BoydRunge,TrefPlatteIllCond} for a comprehensive list). Whilst many are quite effective in practice, ill-conditioning is often an issue. This was recently explained by Platte, Trefethen \& Kuijlaars in \cite{TrefPlatteIllCond} (see also \S \ref{ss:PTKrelation}), wherein it was shown that any exponentially convergent method for recovering analytic functions $f$ from equispaced data must also be exponentially ill-conditioned. As was also proved, the best possible that can be achieved by a stable method is root-exponential convergence. This profound result, most likely the first of its kind for this type of problem, places an important theoretical benchmark against which all such methods must be measured. As we show in the first half of this paper, the numerical discrete FE is well-conditioned and has good convergence properties. Yet it relies on particular interpolation points \R{mapCheb} which are not equispaced. In the second half of this paper we consider Fourier extensions based on equispaced data. In particular, if $x_n = \frac{n}{M}$ we study the so-called \textit{equispaced} Fourier extension \be{ \label{equiExt} F_{N,M}(f) : = \underset{\phi \in \cG_N}{\operatorname{argmin}} \sum_{|n| \leq M} | f(x_n) - \phi(x_n) |^2, } and its finite-precision counterpart $G_{N,M}(f)$. Our primary interest shall lie with the case where $M = \gamma N$ for some $\gamma \geq 1$, i.e.\ where the number of points $M$ scales linearly with $N$. In this case we refer to $\gamma$ as the \textit{oversampling} parameter. Observe that \R{equiExt} results in an $(2M+1) \times (2N+1)$ least squares problem for the coefficients of $F_{N,M}(f)$. We shall denote the corresponding matrix by $\bar{A}$. Our main conclusions concerning the exact equispaced FE $F_{N,M}(f)$ are as follows (see \S \ref{ss:equiFEthy}): \begin{enumerate} \item[6.] The condition number of $\bar{A}$ is exponentially large as $N,M \rightarrow \infty$ with $M \geq N$. \item[7.] The condition number of exact equispaced FE mapping $\kappa(F_{N,\gamma N})$ is exponentially large in $N$ whenever $M = \gamma N$ for $\gamma \geq 1$ fixed. Moreover, the approximation $F_{N,\gamma N}(f)$ suffers from a Runge phenomenon for any fixed $\gamma \geq 1$. In particular, the error $\| f - F_{N,\gamma N}(f)\|$ may diverge geometrically fast in $N$ for certain analytic functions $f$. \item[8.] The scaling $M = \ord{N^2}$ is required to overcome the ill-conditioning and the Runge phenomenon in $F_{N,M}$. In this case, $F_{N,M}(f)$ converges at the same rate as the exact continuous FE $F_N(f)$, i.e.\ geometrically fast in $N$. Although the condition number of $\bar{A}$ remains exponentially large, the condition number of the mapping $\kappa(F_{N,M})$ is $\ord{1}$ for this scaling. \end{enumerate} These results lead to the following conclusion. The exact (infinite-precision) equispaced FE $F_{N,M}$ with $M = \ord{N^2}$ attains the stability barrier of Platte, Trefethen \& Kuijlaars: namely, it is well-conditioned and converges root-exponentially fast in the parameter $M$. However, since the matrix $\bar{A}$ is always ill-conditioned, one expects there to be differences between the exact equispaced extension $F_{N,M}(f)$ and its numerical counterpart $G_{N,M}(f)$. In practice, one sees both differing stability and convergence behaviour of $G_{N,M}(f)$, much like in the case of continuous and discrete FEs. Specifically, in \S \ref{ss:equianalysis} we show the following: \begin{enumerate} \item[9.] The condition number $\kappa(G_{N,\gamma N})$ satisfies \bes{ \kappa(G_{N,\gamma N}) \lesssim \epsilon^{-a(\gamma;T)},\quad \forall N \in \bbN, } where $\epsilon = \epsilon_{\mathrm{mach}}$ is the machine precision used, and $0<a(\gamma;T) \leq 1$ is independent of $N$ and satisfies $a(\gamma;T) \rightarrow 0$ as $\gamma \rightarrow \infty$ for fixed $T$ (see \R{agammaT} for the definition of $a(\gamma;T)$). \item[10.] The error $\| f - G_{N,\gamma N} ( f )\|$ behaves as follows: \enum{ \item[(i)] If $N<N_2$, where $N_2$ is a function-independent breakpoint, $\| f - G_{N,\gamma N} ( f ) \|$ converges or diverges exponentially fast at the same rate as $\nm{f-F_{N,\gamma N} ( f ) }$. \item[(ii)] If $N_2 \leq N < N_1$, where $N_1$ is as introduced previously in \S\ref{ss:intro_summary}, then $\nm{f-G_{N,\gamma N} ( f )}$ converges geometrically fast at the same rate as $\nm{f-F_N ( f )}$, where $F_N ( f )$ is the exact continuous FE. \item[(iii)] When $N = N_1$ the error \bes{ \| f - G_{N_1,\gamma N_1}(f) \| \approx c_f \epsilon^{d_f-a(\gamma;T)}, } where $c_f$ and $d_f$ are as in 5.\ of \S \ref{ss:intro_summary}. \item[(iii)] If $N > N_1$ then $\nm{f-G_{N,\gamma N} ( f )}$ decays at least superalgebraically fast in $N$ down to a maximal achievable accuracy of order $\epsilon^{1-a(\gamma;T)}$. } \end{enumerate} These results show that the condition number of the numerical equispaced FE is bounded whenever $M = \gamma N$, unlike for its exact analogue. Moreover, after a (function-independent) regime of possible divergence, we witness geometric convergence of $G_{N,\gamma N}(f)$ down to a certain accuracy. As in the case of the continuous or discrete FEs, if the function $f$ is sufficiently analytic with small constant $c_f$ then the convergence effectively stops at this point. If not, we witness a further regime of guaranteed superalgebraic convergence. But in both cases, the maximal achievable accuracy is of order $\epsilon^{1-a(\gamma;T)}$, which, since $a(\gamma;T) \rightarrow 0$ as $\gamma \rightarrow \infty$, can be made arbitrarily close to $\epsilon$ by increasing $\gamma$. Note that doing this both improves the condition number of the numerical equispaced FE and yields a less severe rate of exponential divergence in the region $N<N_2$. As we show via numerical computation of the relevant constants, double oversampling $\gamma = 2$ with $T=2$ gives perfectly adequate results in most cases. The main conclusion of this analysis is that numerical equispaced FEs, unlike their exact counterparts, are able to circumvent the stability barrier of Platte, Trefethen \& Kuijlaars to an extent (see \S \ref{ss:PTKrelation} for a more detailed discussion). Specifically, the numerical FE $F_{N,\gamma N}$ has a bounded condition number, and for all sufficiently analytic functions---namely, those analytic in the region $\cD(E(T))$---the convergence is geometric down to a finite accuracy of order $c_f \epsilon^{1-a(\gamma;T)}$. This latter observation, namely the fact that the maximal accuracy is nonzero, is precisely the reason why the stability theorem, which requires geometric convergence for all $N$, does not apply. On the other hand, for all other analytic functions (or those possessing large constants $c_f$) the convergence is at least superalgebraic for $N > N_1$ down to roughly $\epsilon^{1-a(\gamma;T)}$; again not in contradiction with the theorem. Importantly, one never sees divergence of the numerical FE after the finite breakpoint $N_2$. For this reason, we conclude that equispaced FEs are an attractive method for approximations from equispaced data. To further support this conclusion we also remark that although the primary concern of this paper is analytic functions, equispaced FEs are also applicable to functions of finite regularity. In this case, one witnesses algebraic convergence, with the precise order depending solely on the degree of smoothness (see Theorem \ref{t:specconv}). \subsection{Relation to previous work} One-dimensional FEs for overcoming the Gibbs and Runge phenomena were studied in \cite{BoydFourCont} and \cite{BoydRunge}, and applications to surface parametrizations considered in \cite{brunoFEP}. Analysis of the convergence of the exact continuous and discrete FEs was presented by Huybrechs in \cite{huybrechs2010fourier} and Adcock \& Huybrechs in \cite{BADHFEResolution}. The issue of resolution power was also addressed in the latter. The content of the first half of this paper, namely analysis of exact/numerical FEs, follows on directly from this work. A different approach to FEs, known as the FC--Gram method, was introduced in \cite{lyon2010high}. This approach forms a central part of an extremely effective method for solving PDEs in complex geometries \cite{albin2011,bruno2010high}. For previous work on using FEs for PDE problems (so-called Fourier \textit{embeddings}) see \cite{boyd2005fourier,pasquettiFourEmbed}. Equispaced FEs of the form studied in this paper were first independently considered by Boyd \cite{BoydFourCont} and Bruno \cite{bruno2003review}, and later by Bruno et al.\ \cite{brunoFEP}. In particular, Boyd \cite{BoydFourCont} describes the use of truncated singular value decompositions (SVDs) to compute equispaced FEs, and gives extensive numerical experiments (see also \cite{BoydRunge}). Bruno focuses on the use of Fourier extensions (also called \emph{Fourier continuations} in the above references) for the description of complicated smooth surfaces. He suggested in \cite{bruno2003review} a weighted least squares to obtain a smooth extension for this purpose, with numerical evidence supporting convergence results in \cite{brunoFEP}. Most recently Lyon has presented an analysis of equispaced FEs computed using truncated SVDs \cite{LyonFESVD}. In particular, numerical stability and convergence (down to close to machine precision) were shown. In \S \ref{ss:equianalysis} we discuss this work in more detail (see, in particular, Remark \ref{r:Lyon}), and give further insight into some of the questions raised in \cite{LyonFESVD}. \subsection{Outline of the paper} The outline of the remainder of this paper is as follows. In \S \ref{s:FourExt} we recap properties of the continuous and discrete FEs from \cite{BADHFEResolution,huybrechs2010fourier}, including convergence and how to choose the extension parameter $T$. Ill-conditioning of the coefficient map is proved in \S \ref{s:instability}, and in \S \ref{s:stability} we consider the stability of the numerical extensions and their convergence. Finally, in \S \ref{s:equispacedI} we consider the case of equispaced FEs. A comprehensive list of symbols is given at the end of the paper. \section{Fourier extensions}\label{s:FourExt} In this section we introduce FEs, and recap salient important aspects of \cite{BADHFEResolution,huybrechs2010fourier}. \subsection{Two interpretations of Fourier extensions}\label{ss:FEinterp} There are two important interpretations of FEs which inform their approximation properties and their stability, respectively. These are described in the next two sections. \subsubsection{Fourier extensions as polynomial approximations}\label{sss:FEpoly} The space $\cG_{N}$ can be decomposed as $\cG_{N} = \cC_{N} \oplus \cS_{N}$, where \bes{ \cC_{N} = \spn \left \{ \cos \tfrac{n \pi}{T} x : n=0,\ldots,N \right \},\quad \cS_{N} = \spn \left \{ \sin \tfrac{n \pi}{T} x : n=1,\ldots,N \right \}, } consist of even and odd functions respectively. Likewise, for $f$ we have \bes{ f(x) = f_{e}(x) + f_{o}(x),\qquad f_e(x) = \tfrac{1}{2} \left [ f(x) + f(-x) \right ],\quad f_{o}(x)= \tfrac{1}{2} \left [ f(x) - f(-x) \right ], } and for any FE $f_N$ of $f$: \be{ \label{FEdecomp} f_{N} = f_{e,N} + f_{o,N},\quad f_{e,N} \in \cC_N,\ f_{o,N} \in \cS_{N}. } Throughout this paper we shall use the notation $f_N$ to denote an arbitrary FE of $f$ when not wishing to specify its particular construction. From \R{FEdecomp}, it follows that the problem of approximating $f$ via a FE $f_N$ decouples into two problems $f_{e,N}\approx f_e $ and $f_{o,N} \approx f_o$ in the subspaces $\cC_N$ and $\cS_N$ respectively on the half-interval $[0,1]$. Let us define the mapping $y = y(x) : [0,1] \rightarrow [c(T),1]$ by $ y = \cos \tfrac{\pi}{T} x$, where $c(T) = \cos \frac{\pi}{T}$. The functions $\cos \frac{n \pi}{T} x$ and $\sin \frac{(n+1) \pi}{T} x / \sin \frac{\pi}{T} x$ are algebraic polynomials of degree $n$ in $y$. Therefore $\cC_{N}$ and $\cS_{N}$ are (up to multiplication by $\sin \frac{\pi}{T} x$ for the latter) the subspaces $\bbP_N$ and $\bbP_{N-1}$ of polynomials of degree $N$ and $N-1$ respectively in the transformed variable $y$. Letting \bes{ g_1(y) = f_e(x),\quad g_2(y) = \frac{f_o(x)}{\sin \frac{\pi}{T} x},\qquad g_{1,N}(y) = f_{e,N}(x),\quad g_{2,N}(y) = \frac{f_{o,N}(x)}{\sin \tfrac{\pi}{T} x} , } with $g_{1,N}(y) \in \bbP_{N}$ and $g_{2,N}(y) \in \bbP_{N-1}$, we conclude that the FE approximation $f_N$ in the variable $x$ is completely equivalent to two polynomial approximations in the transformed variable $y \in [c(T),1]$. This fact is central to the analysis of FEs. It allows one to use the rich literature on polynomial approximations to determine the theoretical behaviour of the continuous and discrete FEs (see \S \ref{s:FEconv}). \rem{ The interpretation of $f_N$ in terms of polynomials is solely for the purposes of analysis. We always perform computations in the $x$-domain using the standard trigonometric basis for $\cG_N$ (see \S \ref{s:FEtypes}). } The interval $[c(T),1] \subseteq (-1,1]$ is not standard. It is thus convenient to map it affinely to $[-1,1]$. Let \bes{ z: = z(y)= 2 \frac{ y-c(T) } { 1-c(T) } - 1 \in [-1,1]. } Observe that $y = y(z) = c(T) + \frac{1-c(T)}{2}(z+1)$. Let $m : [0,1] \rightarrow [-1,1]$ be the mapping $x \mapsto z$, i.e.\ \be{ \label{mdef} z = m(x) = 2 \frac{\cos \frac{\pi}{T} x - c(T)}{1-c(T)} - 1. } Note that $x = m^{-1}(z) = \frac{T}{\pi} \arccos \left [ c(T) + \frac{1-c(T)}{2}(z+1) \right ]$. If we now define \be{ \label{hidef} h_{i}(z) = g_{i}(y(z)), \quad i=1,2, } then the FE $f_N$ is equivalent to the two polynomial approximations \be{ \label{E:hNdef} h_{1,N}(z) = g_{1,N}(y(z)) = f_{e,N}(m^{-1}(z)),\quad h_{2,N}(z) = g_{2,N}(y(z)) = \frac{f_{o,N}(m^{-1}(z))}{\sin \left ( \frac{\pi}{T} m^{-1}(z) \right)}, } of degree $N$ and $N-1$ respectively in the new variable $z \in [-1,1]$. \subsubsection{Fourier extensions as frame approximations}\label{sss:FEframe} \defn{ Let $\rH$ be a Hilbert space with inner product $\ip{\cdot}{\cdot}$ and norm $\nm{\cdot}$. A set $\{ \phi_n \}^{\infty}_{n =1} \subseteq \rH$ is a frame for $\rH$ if (i) $\spn \{ \phi_n \}^{\infty}_{n=1}$ is dense in $\rH$ and (ii) there exist $c_1,c_2 > 0$ such that \be{ \label{frameprop} c_1 \nm{f}^2 \leq \sum^{\infty}_{n=1} | \ip{f}{\phi_n} |^2 \leq c_2 \nm{f}^2,\quad \forall f \in \rH. } If $c_1 = c_2$ then $\{ \phi_n \}^{\infty}_{n=1}$ is referred to as a tight frame. } Introduced by Duffin \& Schaeffer \cite{duffinshaeffer}, frames are vitally important in signal processing \cite{christensen2003introduction}. Note that all orthonormal, indeed Riesz, bases are frames, but a frame need not be a basis. In fact, frames are typically \textit{redundant}: any element $f \in \rH$ may well have infinitely many representations of the form $f = \sum^{\infty}_{n=1} \alpha_n \phi_n$ with coefficients $\{ \alpha_n \}^{\infty}_{n=1} \in l^2(\bbN)$. The relevance of frames to Fourier extensions is due to the following observation: \lem{ [\cite{BADHFEResolution}] The set $\{ \frac{1}{\sqrt{2T}} \E^{\I \frac{n \pi}{T} x} \}_{n \in \bbZ}$ is a tight frame for $\rL^2(-1,1)$ with $c_1=c_2=1$. } Note that $\{ \frac{1}{\sqrt{2T}} \E^{\I \frac{n \pi}{T} x} \}_{n \in \bbZ}$ is an orthonormal basis for $\rL^2(-T,T)$: it is precisely the standard Fourier basis on $[-T,T]$. However, it forms only a frame when considered as a subset of $\rL^2(-1,1)$. This fact means that ill-conditioning may well be an issue in numerical algorithms for computing FEs, due to the possibility of redundancies. As it happens, it is trivial to see that the set $\{ \frac{1}{\sqrt{2T}} \E^{\I \frac{n \pi}{T} x} \}_{n \in \bbZ}$ is redundant: \lem{ Let $f \in \rL^2(-1,1)$ be arbitrary, and suppose that $\tilde{f} \in \rL^2(-T,T)$ is such that $f = \tilde{f}$ a.e.\ on $[-1,1]$. If $\phi_n(x) = \frac{1}{\sqrt{2 T}} \E^{\I \frac{n \pi}{T} x}$ and $\alpha_n = \ip{\tilde{f}}{\phi_n}_{[-T,T]}$, then \be{ \label{f_rep} f = \sum_{n \in \bbZ} \alpha_n \phi_n\ \mbox{a.e.} } In particular, there are infinitely many sequences $\{ \alpha_n \}_{n \in \bbZ} \in l^2(\bbZ)$ for which $f = \sum_{n \in \bbZ} \alpha_n \phi_n$. } \prf{ The sum $\sum_{n \in \bbZ} \alpha_n \phi_n$ is the Fourier series of $\tilde{f}$ on $[-T,T]$. Thus it coincides with $\tilde{f}$ a.e.\ on $[-T,T]$, and hence $f$ when restricted to $[-1,1]$. Since there are infinitely many possible $\tilde{f}$, each giving rise to a different sequence $ \{ \alpha_n \}_{n \in \bbZ}$, the result now follows. } This lemma is valid for arbitrary $f \in \rL^2(-1,1)$. When $f$ has higher regularity---say $f \in \rH^k(-1,1)$, where $\rH^k(-1,1)$ is the $k^{\rth}$ standard Sobolev space on $(-1,1)$---it is useful to note that there exist extensions $\tilde f$ with the same regularity on the torus $\bbT = [-T,T)$. This is the content of the next result. For convenience, given a domain $I$, we now write $\nm{\cdot}_{\rH^k(I)}$ for the standard norm on $\rH^k(I)$: \lem{ \label{l:smoothextension} Let $f \in \rH^k(-1,1)$ for some $k \in \bbN$. Then there exists an extension $\tilde f \in \rH^k(\bbT)$ of $f$ satisfying $\| \tilde f\|_{\rH^k(\bbT)} \leq c_k(T) \| f \|_{\rH^k(-1,1)}$, where $c_k(T) > 0$ is independent of $f$. Moreover, $f = \sum_{n \in \bbZ} \alpha_n \phi_n$, where $\alpha_n = \ip{\tilde{f}}{\phi_n}_{[-T,T]}$ satisfies $\alpha_n = \ord{n^{-k}}$ as $|n| \rightarrow \infty$. } \prf{ The first part of the lemma follows directly from the proof of Theorem 2.1 in \cite{BADHFEResolution}. The second follows from integrating by parts $k$ times and using the fact that $\tilde f$ is periodic. } This lemma, which shall be important later when studying numerical FEs, states that there exist representations of $f$ in the frame $\{ \frac{1}{\sqrt{2T}} \E^{\I \frac{n \pi}{T} x} \}_{n \in \bbZ}$ that have nice (i.e.\ rapidly decaying) coefficients and which cannot grow large on the extended region $[-T,T] $. \subsection{The continuous and discrete Fourier extensions}\label{s:FEtypes} We now describe the two types of FEs we consider in the first part of this paper. \subsubsection{The continuous Fourier extension}\label{sss:continuous} The continuous FE of $f \in \rL^2(-1,1)$, defined by \R{L2}, is the orthogonal projection onto $\cG_N$. Computation of this extension involves solving a linear system. Let us write $F_{N}(f) = \sum^{N}_{n=-N} a_n \phi_{n}$ with unknowns $\{ a_n \}^{N}_{n=-N}$. If $a = (a_{-N},\ldots,a_{N})^{\top}$ and $b=(b_{-N},\ldots,b_{N})^{\top}$, where \be{ \label{E:fcoef} b_{n} =\ip{f}{\phi_n} = \int^{1}_{-1} f(x) \overline{\phi_{n}(x)} \D x,\quad n=-N,\ldots,N, } and $A \in \bbC^{(2N+1) \times (2N+1)}$ is the matrix with $(n,m)^{\rth}$ entry \be{ \label{E:A} A_{n,m} = \ip{\phi_m}{\phi_n} = \int^{1}_{-1} \phi_{m}(x) \overline{\phi_n(x)} \D x,\quad n,m = -N,\ldots,N, } then $a$ is the solution of the linear system $A a = b$. We refer to the values $\{ a_n \}^{N}_{n=-N}$ as the \textit{coefficients} of the FE $F_N(f)$. Note that the matrix $A$ is a Hermitian positive-definite, Toeplitz matrix with $A_{n,m} = A_{n-m}$, where $A_0 = \frac{1}{T}$ and $A_n = \frac{\sin \frac{n \pi }{T}}{n \pi}$ otherwise. In fact, $A$ coincides with the so-called \textit{prolate} matrix \cite{SlepianV,varah}. We shall discuss this connection further in \S \ref{ss:ASVD}. For later use, we also note the following characterization of $F_N(f)$: \prop{ [\cite{BADHFEResolution,huybrechs2010fourier}] \label{p:exactinterp} Let $F_N(f)$ be the continuous FE \R{L2} of a function $f$, and let $h_{i}(z)$ and $h_{i,N}(z)$ be given by \R{hidef} and \R{E:hNdef} respectively (i.e.\ the symmetric and anti-symmetric parts of $f$ and $f_N$ with the coordinate transformed from the trigonometric argument $x$ to the polynomial argument $z$). Then $h_{1,N}(z)$ and $h_{2,N}(z)$ are the truncated expansions of $h_1(z)$ and $h_2(z)$ respectively in polynomials orthogonal with respect to the weight functions \be{ \label{weightfns} w_{1}(z) = \left [ (1-z)(z-m(T)) \right ]^{-\frac12},\quad w_{2}(z) = \left [ (1-z)(z-m(T)) \right ]^{\frac12},\quad z \in [-1,1], } where $m(T) = 1 - 2 \mathrm{cosec}^2 \left ( \tfrac{\pi}{2T} \right ) < -1$. In other words, $h_{i,N}(z)$, $i=1,2$, is the orthogonal projection of $h_i(z)$ onto $\bbP_{N+1-i}$ with respect to the weighted inner product $\ip{\cdot}{\cdot}_{w_i}$ with weight function $w_i$. } \subsubsection{The discrete Fourier extension}\label{sss:discrete} The discrete FE $\tilde{F}_N(f)$ is defined by \R{discL2}. To use this extension it is first necessary to choose nodes $\{ x_n \}^{N}_{n=-N}$. This question was considered in \cite{BADHFEResolution}, and a solution was obtained by exploiting the characterization of FEs as polynomial approximations in the transformed variable $z$. A good system of nodes for polynomial interpolation is given by the Chebyshev nodes \be{ \label{Cheb} z_n = \cos \left ( \frac{(2n+1) \pi}{2N+2} \right ),\quad n=0,\ldots,N. } Mapping these back to the $x$-variable and symmetrizing about $x=0$ leads to the so-called \textit{mapped symmetric Chebyshev} nodes \be{ \label{mapCheb} x_n = - x_{-n-1} = \frac{T}{\pi} \arccos \left [ \frac{1}{2} (1-c(T)) \cos \left ( \frac{(2n+1) \pi}{2N+2} \right )+ \frac{1}{2}(1+c(T)) \right ],\quad n=0,\ldots,N. } This gives a set of $2N+2$ nodes. Therefore, rather than \R{discL2}, we define the discrete FE by \be{ \label{discL2alt} \tilde{F}_N(f) : = \underset{\phi \in \cG'_N}{\operatorname{argmin}} \sum^{N}_{n=-N-1} | f(x_n) - \phi(x_n) |^2, } from now on, where $ \cG'_N = \cC_N \oplus \cS_{N+1}$. Exploiting the relation between FEs and polynomial approximations once more, we now obtain the following: \prop{ \label{p:discinterp} Let $f_N = \tilde{F}_N(f) \in \cG'_N$ be the discrete FE \R{discL2alt} based on the nodes \R{mapCheb}, and let $h_i(z)$ and $h_{i,N}(z) \in \bbP_N$ be given by \R{hidef} and \R{E:hNdef} respectively. Then $h_{i,N}(z)$, $i=1,2$ is the $N^{\rth}$ degree polynomial interpolant of $h_i(z)$ at the Chebyshev nodes \R{Cheb}. } Write $\phi_{n}(x) = \cos \frac{n \pi}{T} x$, $\phi_{-(n+1)}(x) = \sin \frac{n+1}{T} \pi x$, $n \in \bbN$, and let $\tilde{F}_{N}(f)(x) = \sum^{N}_{n=-N-1} a_n \phi_n(x)$. If $a = (a_{-N-1},\ldots,a_{N})^{-T}$ and $\tilde A \in \bbR^{(2N+2) \times (2N+2)}$ has $(n,m)^{\rth}$ entry \be{ \label{E:A_tilde} \tilde{A}_{n,m} = \sqrt{\frac{\pi}{N+1}}\phi_{m}(x_n), \qquad n,m=-N-1,\ldots,N, } then we have $\tilde A a =\tilde b$, where $\tilde b = (\tilde b_{-N-1},\ldots,\tilde b_N)^{\top}$ and $\tilde b_n = \sqrt{\frac{\pi}{N+1}} f(x_n)$. The following lemma concerning the matrix $\tilde A$ will prove useful in what follows: \lem{[\cite{BADHFEResolution}] \label{l:weightedOpt} The matrix $A_W = (\tilde A)^* \tilde A$ has entries \bes{ \ip{\phi_n}{\phi_m}_W := \int^{1}_{-1} \phi_n(x) { \phi_m(x) } W(x) \D x,\quad n,m=-N-1,\ldots,N, } where $W$ is the positive, integrable weight function given by $W(x) = \frac{\sqrt{2} \pi}{T} \frac{\cos \frac{\pi}{2T} x}{\sqrt{ \cos \frac{\pi}{T} x-\cos \frac{\pi}{T}}}$.} This lemma implies that the left-hand side of the normal equations of the discrete FE are the equations of a continuous FE based on the weighted least-squares minimization with weight function $W$. \subsection{Convergence of exact Fourier extensions}\label{s:FEconv} A detailed analysis of the convergence of the exact continuous FE, which we now recap, was carried out in \cite{BADHFEResolution,huybrechs2010fourier}. We commence with the following theorem: \thm{ [\cite{BADHFEResolution}] \label{t:specconv} Suppose that $f \in \rH^{k}(-1,1)$ for some $k \in \bbN$ and that $T>1$. If $F_N(f)$ is the continuous FE of $f$ defined by \R{L2}, then \begin{equation} \label{specineq} \| f - F_N(f) \| \leq c_{k}(T) N^{-k} \| f \|_{\rH^{k}(-1,1)},\quad \forall N \in \bbN, \end{equation} where $c_{k}(T) > 0 $ is independent of $f$ and $N$. } This theorem confirms \textit{algebraic} convergence of $F_N(f)$ whenever the approximated function $f$ has finite degrees of smoothness, and \textit{superalgebraic} convergence, i.e.\ faster than any fixed algebraic power of $N^{-1}$, whenever $f \in \rC^\infty[-1,1]$. Suppose now that $f$ is analytic. Although superalgebraic convergence is guaranteed by Theorem \ref{t:specconv}, it transpires that the convergence is actually geometric. This is a direct consequence of the interpretation of the $F_N(f)$ as the sum of two polynomial expansions in the transformed variable $z$ (Proposition \ref{p:exactinterp}). To state the corresponding theorem, we first require the following definition: \defn{ \label{d:Bernstein} The Bernstein ellipse $\cB(\rho) \subseteq \bbC$ of index $\rho \geq 1$ is given by \bes{ \cB(\rho) = \left \{ \tfrac{1}{2} \left ( \rho^{-1} \E^{\I \theta} + \rho \E^{-\I \theta} \right ) : \theta \in [-\pi,\pi ] \right \}. } } Given a compact region bounded by the Bernstein ellipse $\cB(\rho)$, we shall write \be{ \label{Mapped_Bernstein} \cD(\rho) \subseteq \bbC } for its image in the complex $x$-plane under the mapping $x = m^{-1}(z)$, where $m$ is as in \R{mdef}. \thm{ [\cite{BADHFEResolution}, \cite{huybrechs2010fourier}] \label{t:expconv} Suppose that $f$ is analytic in $\cD(\rho^*)$ and continuous on its boundary. Then $\| f - F_N(f) \|_{\infty} \leq c_f \rho^{-N}$, where $\rho = \min \left \{ \rho^*, E(T) \right \}$, $c_f > 0$ is proportional to $\max_{x \in \cD(\rho)}| f(x) |$, and $E(T)$ is as in \R{cteET}. } \prf{ A full proof was given in \cite[Thm 2.3]{BADHFEResolution}. The expansion $g_N$ of an analytic function $g$ in a system of orthogonal polynomials with respect to some integrable weight function satisfies $\| g - g_N \|_{\infty} \leq c_g \rho^{-N}$, where $c_g$ is proportional to $\max_{z \in \cB(\rho)} | g(z)|$ \cite{rivlin1990chebyshev}. In view of Proposition \ref{p:exactinterp}, it remains only to determine the maximal parameter $\rho$ of Bernstein ellipse $\cB(\rho)$ within which $h_1(z)$ and $h_2(z)$ are analytic. The mapping $z = m(x)$ introduces a square-root type singularity into the functions $h_{i}(z)$ at the point $z = m(T) < -1$. Hence the maximal possible value of the parameter $\rho$ satisfies \be{ \label{Bernequiv} \tfrac{1}{2} (\rho + \rho^{-1}) = - m(T). } Observe that if $\psi(t) = t + \sqrt{t^2-1} $ then \be{ \label{Emequiv} \psi(m(T)) = E(T). } Thus, since $\rho > 1$, the solution to \R{Bernequiv} is precisely $\rho = E(T)$. Conversely, any singularity of $f$ introduces a singularity of $h_i(z)$, which also limits this value. Hence we obtain the stated minimum. } Theorem \ref{t:expconv} shows that if $f$ is analytic in a sufficiently large region (for example, if $f$ is entire) then the rate of geometric convergence is precisely $E(T)$. Recall that the parameter $T$ can be chosen by the user. In the next section we consider the effect of different choices of $T$. \rem{ Although Theorems~\ref{t:specconv} and \ref{t:expconv} are stated for $F_N(f)$, they also hold for the discrete FE $\tilde{F}_N(f)$, since the latter is equivalent to a sum of Chebyshev interpolants (Proposition \ref{p:discinterp}). } \subsection{The choice of $T$} \label{ss:Tchoice} Note that $E(T) \sim 1 + \pi(T-1)$ as $T \rightarrow 1^+$ and $E(T) \sim \frac{16}{\pi^2} T^2$ when $T \rightarrow \infty$. Thus, small $T$ leads to a slower rate of geometric convergence, whereas large $T$ gives a faster rate. As discussed in \cite{BADHFEResolution}, however, a larger value of $T$ leads to a worse resolution power, meaning that more degrees of freedom are required to resolve oscillatory behaviour. On the other hand, setting $T$ sufficiently close to $1$ yields a resolution power that is arbitrarily close to optimal. In \cite{BADHFEResolution} a number of fixed values of $T$ were used in numerical experiments. These typically give good results, with small values of $T$ being particularly well suited to oscillatory functions. Another approach for choosing $T$ was also discussed. This involves letting \be{ \label{Tkte} T =T(N;\epsilon_{\mathrm{tol}})= \frac{\pi}{4} \left ( \arctan \left ( (\epsilon_{\mathrm{tol}})^{\frac{1}{2N}} \right ) \right )^{-1}, } where $\epsilon_{\mathrm{tol}} \ll 1$ is some fixed tolerance (note that this is very much related to the Kosloff Tal--Ezer map in spectral methods for PDEs \cite{boyd,KTEmapped}---see \cite{BADHFEResolution} for a discussion). This choice of $T$, which now depends on $N$, is such that $E(T)^{-N} = \epsilon_{\mathrm{tol}}$. Although this limits the best achievable accuracy of the FE with this approach to $\ord{\epsilon_{\mathrm{tol}}}$, setting $\epsilon_{\mathrm{tol}} = 10^{-14}$ is normally sufficient in practice. Numerical experiments in \cite{BADHFEResolution} indicate that this works well, especially for oscillatory functions. In fact, since \be{ \label{Tkteasymp} T(N;\epsilon_{\mathrm{tol}}) \sim 1 - \frac{\log (\epsilon_{\mathrm{tol}})}{\pi N} + \ord{N^{-2}},\quad N \rightarrow \infty,} this approach has formally optimal resolution power. \rem{ The strategy \R{Tkte} is particularly good for oscillatory problems. However, if this is not a concern, a practical choice appears to be $T=2$. In this case, the FE has a particular symmetry that can be exploited to allow for its efficient computation in only $\ord{N (\log N)^2}$ operations \cite{LyonFast}. } \section{Condition numbers of exact Fourier extensions}\label{s:instability} The redundancy of the frame $\{ \frac{1}{\sqrt{2 T}} \E^{\I \frac{n \pi}{T} \cdot} \}_{n \in \bbZ}$ means that the matrices associated with the continuous and discrete FEs are ill-conditioned. We next derive bounds for the condition number of these matrices. The spectrum of $A$ is considered further in \S \ref{ss:ASVD}, and the condition numbers of the FE mappings $f \mapsto F_{N}(f)$ and $f \mapsto \tilde{F}_N(f)$ are discussed in \S \ref{ss:cond_numb_exact}. \subsection{The condition numbers of the continuous and discrete FE matrices} \label{ss:condnumb} \thm{ \label{t:exactCondNumb} Let $A$ be the matrix \R{E:A} of the continuous FE. Then the condition number of $A$ is $\ord{E(T)^{2N}}$ for large $N$. Specifically, the maximal and minimal eigenvalues satisfy \ea{ \label{exactCondNumb} T^{-1} \leq \lambda_{\max}(A) \leq 1,\qquad c_1(T) N^{-3} E(T)^{-2N} \leq \lambda_{\min}(A) \leq c_2(T) N^{2} E(T)^{-2N}, } where $c_1(T)$ and $c_2(T)$ are positive constants with $c_1(T),c_2(T) = \ord{1}$ as $T \rightarrow 1^+$. } \prf{ It is a straightforward exercise to verify that \be{ \label{exactmateval} \lambda_{\min}(A) = \min_{\phi \in \cG_N} \left \{\| \phi \|^2 : \| \phi \|_{[-T,T]} = 1 \right \},\quad \lambda_{\max}(A) = \max_{\phi \in \cG_N} \left \{\| \phi \|^2 : \| \phi \|_{[-T,T]} = 1 \right \}. } Using the fact that $\| \phi \| \leq \| \phi \|_{[-T,T]}$, we first notice that $\lambda_{\max}(A) \leq 1 $. On the other hand, setting $\phi =\frac{1}{\sqrt{2 T}}$, we find that $\lambda_{\max}(A) \geq T^{-1}$, which completes the result for $\lambda_{\max}(A)$. We now consider $\lambda_{\min}(A)$. Recall that any $\phi \in \cG_{N}$ can be decomposed into even and odd parts $\phi_{e}$ and $\phi_{o}$, with each function corresponding to a polynomial in the transformed variable $z$. Hence, \be{ \label{char} \lambda_{\min}(A) = \min_{\substack{\phi \in \cG_N \\ \phi \neq 0}} \left \{\frac{\| \phi \|^2}{\| \phi \|^2_{[-T,T]} } \right \} = \min_{\substack{p_1 \in \bbP_N, p_2 \in \bbP_{N-1} \\ \|p_1\| + \|p_2\| \neq 0}} \left \{ \frac{\| p_1 \|^2_{w_1} + \| p_2 \|^2_{w_2}}{\| p_1 \|^2_{w_1,[m(T),1]} + \| p_2 \|^2_{w_2,[m(T),1]}} \right \}, } where $w_i$, $i=1,2$, is given by \R{weightfns}. Since the weight function $w_{i}$ is integrable, we have \be{ \label{intweight} \| p_{i} \|_{w_i,[m(T),1]} \leq \sqrt{C_{i}(T)} \| p_i \|_{\infty,[m(T),1]},\quad i=1,2, } where $C_{i}(T) = \int^{1}_{m(T)} \D w_{i}$, $i=1,2$. Moreover, by Remez's inequality, \bes{ \| p \|_{\infty,[m(T),1]} \leq \| T_{N} \|_{\infty,[m(T),1]} \| p \|_{\infty},\quad \forall p \in \bbP_N, } where $T_{N} \in \bbP_N$ is the $N^{\rth}$ Chebyshev polynomial. Since $T_{N}$ is monotonic outside $[-1,1]$, we have $\| T_{N} \|_{\infty,[m(T),1]} = | T_{N}(m(T)) |$. Moreover, due to the formula \bes{ T_{N}(x) = \frac{1}{2} \left [ \left ( x - \sqrt{x^2-1} \right )^n + \left ( x + \sqrt{x^2-1} \right )^n \right ], } an application of \R{Emequiv} gives \be{ \label{remez} \| T_{N} \|_{\infty,[m(T),1]} = \frac{1}{2} \left [ E(T)^N + E(T)^{-N} \right ] < E(T)^N,\quad \forall N \in \bbN, T>1. } Next we note that $w_{1}(z) \geq D_1(T) $ and $w_{2}(z) \geq D_2(T) \sqrt{1-z^2} $, $ \forall z \in [-1,1]$, for positive constants $D_1(T)$ and $D_2(T)$. Moreover, there exist constants $d_1,d_2 > 0$ independent of $T$ such that \bes{ \| p \|_{\infty} \leq d_1 N \| p \|,\quad \| p \|_{\infty} \leq d_2 N^{\frac{3}{2}} \| p \|_v,\quad p \in \bbP_N, } where $v(z) = \sqrt{1-z^2}$ (this follows from expanding $p$ in orthonormal polynomials $\{ p_n \}_{n \in \bbN}$ on $[-1,1]$ corresponding to the weight function $w(z) = 1$, i.e.\ Legendre polynomials, or $w(z) = v(z)$, i.e.\ Chebyshev polynomials of the second kind, and using the known estimate $\| p_n \|_{\infty} = \ordu{n^{\frac12}}$ for the former and $\| p_n \|_{\infty} = \ordu{n^{\frac32}}$ for the latter \cite[chpt.\ X]{bateman}). Therefore \be{ \label{claim} \| p \|_{\infty} \leq \frac{d_i}{\sqrt{D_i(T)}} N^{\frac{1+i}{2}} \| p \|_{w_i},\quad \forall p \in \bbP_N,\quad i=1,2. } Substituting \R{intweight}, \R{remez} and \R{claim} into \R{char} now gives \bes{ \lambda_{\min}(A)\geq \frac{1}{\max \{ C_1(T)/D_1(T),C_2(T)/D_2(T) \} } N^{-3} E(T)^{-2N}, } which gives the lower bound in \R{exactCondNumb}. For the upper bound, we set $p_2 = 0$ and $p_1 = T_N$ in \R{char} to give \be{ \label{lminupper} \lambda_{\min}(A) \leq \frac{ \| T_{N} \|^2_{w_1}}{\| T_{N} \|^2_{w_1,[m(T),1]}} \leq \frac{C_1(T)}{\| T_{N} \|^2_{w_1,[m(T),1]}}. } Using \R{remez} we note that $\| T_N \|_{\infty,[m(T),1]} \geq \frac{1}{2} E(T)^N$. Recall also that $\| p \|_{\infty} \leq d_1 N \| p \|$,$\forall p \in \bbP_N$. Scaling this inequality to the interval $[m(T),1]$ now gives \bes{ \| p \|_{\infty,[m(T),1]} \leq d_1 \sqrt{\frac{2}{1-m(T)}} N \| p \|_{[m(T),1]} = \sqrt{C_3(T)} N\| p \|_{[m(T),1]}. } Note also that $w_1(z) \geq D_3(T)$, $\forall z \in [m(T),1]$. Therefore, \eas{ \| T_N \|_{w_1,[m(T),1]} \geq \sqrt{D_3(T)} \| T_N \|_{[m(T),1]} \geq \frac{\sqrt{D_3(T)}}{\sqrt{C_3(T)} N} \| T_N \|_{\infty,[m(T),1]} \geq \frac{\sqrt{D_3(T)} }{2 \sqrt{C_3(T)} N} E(T)^N. } Substituting this into \R{lminupper} now gives the result. } We now consider the case of the discrete FE: \thm{ \label{t:discCondNumb} Let $\tilde A$ be the matrix \R{E:A_tilde} of the discrete FE. Then the condition number of $\tilde A$ is $\ord{E(T)^{N}}$ for large $N$. Specifically, the maximal and minimal singular values of $\tilde{A}$ satisfy \ea{ \label{discCondNumb} c_1(T) \leq \sigma_{\max}(\tilde A) \leq c_2(T) N^{\frac{3}{2}},\qquad d_1(T) N^{-\frac32} E(T)^{-N} \leq \sigma_{\min}(\tilde A) \leq d_2(T) N^{\frac52} E(T)^{-N}, } where $c_1(T),c_2(T),d_1(T),d_2(T)$ are positive constants that are $\ord{1}$ as $T \rightarrow 1^+$. } \prf{ Using Lemma \ref{l:weightedOpt}, the values $\sigma^2_{\min}(\tilde A)$ and $\sigma^2_{\max}(\tilde A)$ may be expressed as in \R{exactmateval} (with $\nm{\cdot}$ replaced by $\nm{\cdot}_W$). Note that $W(0) \| \phi \|^2 \leq \| \phi \|^2_W \leq \| \phi \|^2_{\infty} \int^{1}_{-1} \D W.$ It is a straightforward exercise (using the bound \R{claim} and the fact that $\phi$ can be expressed as the sum of two polynomials) to show that $\| \phi \|_{\infty} \leq C_1(T) N^{\frac32} \| \phi \|$, where $C_1(T) = \ord{1}$ as $T \rightarrow 1^+$. Thus we obtain \bes{ W(0) \frac{\| \phi \|^2}{\| \phi \|^2_{[-T,T]}} \leq \frac{\| \phi \|^2_W}{\| \phi \|^2_{[-T,T]}} \leq \left ( C_1(T)^2 \int^{1}_{-1} \D W \right ) N^3 \frac{\| \phi \|^2}{\| \phi \|^2_{[-T,T]}} . } The result now follows immediately from the bounds \R{exactCondNumb}. } Theorems \ref{t:exactCondNumb} and \ref{t:discCondNumb} demonstrate that the condition numbers of the continuous and discrete FE matrices grow exponentially in $N$. This establishes conclusion 1.\ of \S \ref{s:introduction}. \rem{ \label{r:factorization} Although exponentially large, the matrix of the discrete FE is substantially less poorly conditioned than that of the continuous FE. In particular, the condition number is of order $E(T)^{N}$ as opposed to $E(T)^{2N}$. This can be understood using Lemma \ref{l:weightedOpt}. The normal form $A_W = (\tilde A)^* \tilde A$ of the discrete FE matrix is a continuous FE matrix with respect to the weight function $A_W$. Hence $\kappa(\tilde A) =\sqrt{\kappa(A_W)} \approx \sqrt{\kappa(A)} \approx E(T)^N$. As we shall see later, this property also translates into superior performance of the numerical discrete FE over its continuous counterpart (see \S \ref{ss:numsolnanalysis}). } Since the constants in Theorems \ref{t:exactCondNumb} and \ref{t:discCondNumb} are bounded as $T \rightarrow 1^+$, this allows one also to determine the condition number in the case that $T \rightarrow 1^+$ as $N \rightarrow \infty$ (see \S \ref{ss:Tchoice}). In particular, if $T$ is given by \R{Tkte}, then $\kappa(A)$ and $\kappa(\tilde A)$ are (up to possible small algebraic factors in $N$) of order $(\epsilon_{\mathrm{tol}})^{-2}$ and $(\epsilon_{\mathrm{tol}})^{-1}$. \subsection{The singular value decomposition of $A$}\label{ss:ASVD} Although we have now determined the condition number of $A$, it is possible to give a rather detailed analysis of its spectrum. This follows from the identification of $A$ with the well-known prolate matrix, which was analyzed in detail by Slepian \cite{SlepianV,varah}. We now review some of this work. Following Slepian \cite{SlepianV}, let $P(N,W) \in \bbC^{N \times N}$ be the prolate matrix with entries \bes{ P(N,W)_{m,n} = \left \{ \begin{array}{cl} \frac{\sin 2 \pi W (m-n)}{\pi (m-n)} & m \neq n \\ 2 W & m = n, \end{array} \right . \quad m,n=0,\ldots,N-1, } where $0<W <\frac{1}{2}$ is fixed, and write $1> \lambda_0(N,W)> \ldots> \lambda_{N-1}(N,W)> 0$ for its eigenvalues. Note that \be{ \label{lambdasymmetry} \lambda_{k}(N,\tfrac{1}{2}-W) = 1 - \lambda_{N-1-k}(N,W). } The following asymptotic results are found in \cite{SlepianV}: \begin{itemize} \item[(i)] For fixed and small $k$, \be{ \label{smalleval} 1- \lambda_{k}(N,W) \sim \sqrt{\pi} (k!)^{-1} 2^{(14k+9)/4} \alpha^{(2k+1)/4} (2-\alpha)^{-(k+1/2)} N^{k+1/2} \beta ^{-N}, } where $\alpha = 1 - \cos 2 \pi W$ and $\beta = \frac{\sqrt{2}+\sqrt{\alpha}}{\sqrt{2}-\sqrt{\alpha}} $. \item[(ii)] For large $N$ and $k$ with $k = \lfloor 2 W N (1-\epsilon) \rfloor$ and $0 < \epsilon < 1$, $1 - \lambda_{k}(N,W) \sim \E^{-c_1 - c_2 N}$ for explicitly known constants $c_1,c_2$ depending only on $W$ and $\epsilon$. \item[(iii)] For large $N$ and $k$ with $ k = \lfloor 2 W N + (b/\pi) \log N \rfloor$, $\lambda_{k}(N,W) \sim \frac{1}{1+\E^{\pi b}}$. \end{itemize} (Slepian also derives similar asymptotic results for the eigenvectors of $P(N,W)$ \cite{SlepianV}). From these results we conclude that the eigenvalues of the prolate matrix cluster exponentially near $0$ and $1$ and have a transition region of width $\ord{\log N}$ around $k = 2 W N$. This is shown in Figure \ref{f:ExactEvals}. \begin{figure} \begin{center} $\begin{array}{ccc} \includegraphics[width=6.25cm]{Fig1Cts} & \hspace{1pc} & \includegraphics[width=6.25cm]{Fig1Disc} \end{array}$ \caption{\small Eigenvalues of the matrices \R{E:A} (left) and \R{E:A_tilde} (right) for $N=200$ and $T=2$.} \label{f:ExactEvals} \end{center} \end{figure} The matrix $A$ of the continuous FE is precisely the prolate matrix $P(2N+1,\frac{1}{2T})$. In this case, the parameter $\beta$ in \R{smalleval} is given by \bes{ \beta = \frac{\sqrt{2}+\sqrt{\alpha}}{\sqrt{2}-\sqrt{\alpha}} = \cot^2 \left (\frac{\pi}{4 T} \right ) =E(T).} Applying Slepian's analysis, we now see that the eigenvalues of $A$ cluster exponentially at rate $E(T)^2$ near zero and one (note that $A$ corresponds to a prolate matrix of size $2N$), and in particular, that the condition number is $\ord{E(T)^{2N}}$. The latter estimate agrees with that given in Theorem \ref{t:exactCondNumb}. We remark, however, that Theorem \ref{t:exactCondNumb} gives bounds for the minimal eigenvalue of $A$ that hold for all $N$ and $T$, unlike \R{smalleval}, which holds only for fixed $T$ and sufficiently large $N$. Hence Theorem \ref{t:exactCondNumb} remains valid when $T$ is varied with $N$, an option which, as discussed in \S \ref{ss:Tchoice}, can be advantageous in practice. Since the matrix $\tilde A$ of the discrete FE is related to $A$ (see Lemma \ref{l:weightedOpt}), we expect a similar structure for its singular values. This is illustrated in Figure \ref{f:ExactEvals}. As is evident, the only qualitative difference between $\tilde A$ and $A$ is found in the large singular values. The other features---the narrow transition region and the exponential clustering of singular values near $0$---are much the same. \rem{ \label{sss:T2} The choice $T=2$ ($W = \frac{1}{4}$) is special. As shown by \R{lambdasymmetry}, the eigenvalues $\lambda_{k}(N,W)$ are symmetric in this case, and the transition region occurs at $k = \frac{1}{2} N$. This is unsurprising. When $T=2$, the frame $\{ \E^{\I \frac{n \pi}{2} x} \}_{n \in \bbZ}$ decomposes into two orthogonal bases, related to the sine and cosine transforms. Using this decomposition and the associated discrete transforms for each basis, M. Lyon has introduced an $\ord{N (\log N)^2 }$ complexity algorithm for computing FEs \cite{LyonFast}. } \subsection{Numerical examples}\label{ss:S5numexamp} We now consider several numerical examples of the continuous and discrete FEs. In Figure \ref{f:FnApp} we plot the error $\| f - f_N \|_{\infty}$ against $N$ for various choices of $f$. Here the extension $f_N$ is the numerically computed continuous or discrete FE---i.e.\ the result of solving the corresponding linear system in standard precision (recall Remark \ref{r:numsolver}). Henceforth, we use the notation $G_N(f)$ and $\tilde{G}_N(f)$ for these \textit{numerical} extensions, so as to distinguish them from their \textit{exact} counterparts $F_N(f)$ and $\tilde{F}_N(f)$. Note that the word `exact' in this context refers to exact arithmetic. We do not mean exact in the sense that $F_N(f) = f$ for $f \in \cG_N$. At first sight, Figure \ref{f:FnApp} appears somewhat surprising: for all three functions we obtain good accuracy, and there is no drift or growth in the error, even in the case where $f$ is nonsmooth or has a complex singularity near $x=0$. Evidently the ill-conditioning of the FE matrices established in Theorems \ref{t:exactCondNumb} and \ref{t:discCondNumb} appears to have little effect on the numerical extensions $G_N(f)$ and $\tilde{G}_{N}(f)$. The purpose of \S \ref{s:stability} is to offer an explanation of this phenomenon. In Figure \ref{f:FnApp} we also compare two choices of $T$: fixed $T=2$ and the $N$-dependent value \R{Tkte} with $\epsilon_{\mathrm{tol}} = 10^{-14}$. Note that the latter typically outperforms the fixed value $T=2$, especially for oscillatory functions. This is unsurprising in view of the discussion in \S \ref{ss:Tchoice}. Figure \ref{f:FnApp} also illustrates an important disadvantage of the continuous FE: namely, the approximation error levels off at around $\sqrt{\epsilon_{\mathrm{mach}}}$, where $\epsilon_{\mathrm{mach}} \approx 10^{-16}$ is the machine precision used, as opposed to around $\epsilon_{\mathrm{mach}} $ for the discrete extension. Our analysis in \S \ref{s:stability} will confirm this phenomenon. Note that the differing behaviour between the continuous and discrete extensions in this respect can be traced back to the observation made in Remark \ref{r:factorization}. \begin{figure} \begin{center} $\begin{array}{ccc} \includegraphics[width=6.25cm]{Fig2Fn1} & \hspace{1pc} &\includegraphics[width=6.25cm]{Fig2Fn2} \\ f(x) = \E^{25 \sqrt{5} \pi \I x} && f(x) = \frac{1}{1+25 x^2} \\ \\ \includegraphics[width=6.25cm]{Fig2Fn3} & &\includegraphics[width=6.25cm]{Fig2Fn4} \\ f(x) = \frac{1}{8-7 x} & & f(x) = | x|^7 \end{array}$ \caption{\small The error $\| f - f_N \|_{\infty}$, where $f_N = G_N(f)$ (squares and circles) or $f_N = \tilde{G}_N(f)$ (crosses and diamonds) and $T=2$ (squares/crosses) or $T = T(N;\epsilon_{\mathrm{tol}})$ (circles/diamonds) with $\epsilon_{\mathrm{tol}} = 10^{-14}$.} \label{f:FnApp} \end{center} \end{figure} \subsection{Condition numbers of the exact continuous and discrete FE mappings}\label{ss:cond_numb_exact} The exponential growth in the condition numbers of the continuous and discrete FE matrices imply extreme sensitivity in the FE coefficients to perturbations. However, the numerical results of Figure \ref{f:FnApp} indicate that the FE approximations themselves are far more robust. Although we shall defer a full explanation of this difference to \S \ref{s:stability}, it is possible to give a first insight by determining the condition numbers of the mappings $F_N$ and $\tilde{F}_N$. For vectors $b \in \bbC^{2N+1}$ and $\tilde{b} \in \bbC^{2N+2}$ let us write, with slight abuse of notation, $F_N(b)$ and $\tilde{F}_N(\tilde b)$ for the corresponding continuous and discrete Fourier extensions whose coefficient vectors are the solutions of the linear systems $A a = b$ and $\tilde{A} a = \tilde{b}$ respectively. We now define the condition numbers \be{ \label{conditionFE} \kappa(F_N) = \sup \left \{ \| F_{N}(b) \| : b \in \bbC^{2N+1},\ \nm{b} = 1 \right \},\ \kappa(\tilde{F}_N) = \sup \left \{ \| F_{N}(b) \|_W : b \in \bbC^{2N+2},\ \nm{b} = 1 \right \}. } Here $\nm{\cdot}$ denotes the usual $l^2$ vector norm, and $W$ is the weight function of Lemma \ref{l:weightedOpt}. Note that \R{conditionFE} gives the \textit{absolute} condition numbers of $F_N$ and $\tilde{F}_N$, as opposed to the more standard \textit{relative} condition number \cite{TrefethenBau}. The key results of this paper can easily be reformulated for the latter. However, we shall use \R{conditionFE} throughout, since it coincides with the definition given in \cite{TrefPlatteIllCond} for linear mappings such as FEs. The work of \cite{TrefPlatteIllCond} will be particularly relevant when considering equispaced FEs in \S \ref{s:equispacedI}. We now have the following result: \lem{ \label{l:ExactCondNumb} The condition numbers of the exact continuous and discrete FEs satisfy \bes{ \kappa(F_N)= 1/\sqrt{\lambda_{\min}(A)},\qquad \kappa(\tilde{F}_N) = 1. } } \prf{ Write $F_N(b) = \sum^{N}_{n=-N} a_n \phi_n$, where $A a = b$. We have $\| F_N(b) \|^2 = a^* A a = b^* A^{-1} b$, and therefore $\kappa(F_N) = 1/\sqrt{\lambda_{\min}(A)}$, as required. For the second result, we note that $\| \tilde{F}_N(\tilde{b}) \|^2 = a^* A_W a$, where $A_W = (\tilde{A})^* \tilde{A}$ is the matrix of Lemma \ref{l:weightedOpt}. Since $\tilde A a = \tilde b$ the second result now follows. } As with the FE matrices, this lemma shows that condition number of the discrete mapping $\tilde F_N$, which is identically one, is much better than that of the continuous mapping $F_N$. Similarly, the reason can be traced back to Remark \ref{r:factorization}. Note that this lemma establishes 2.\ of \S \ref{s:introduction}. At first, it may seem that the fact that $\kappa(\tilde{F}_N) =1 $ explains the observed numerical stability in Figure \ref{f:FnApp}. However, since $\lambda_{\min}(A)$ is exponentially small (Theorem \ref{t:exactCondNumb}), the above lemma clearly does not explain the lack of drift in the numerical error in the case of the continuous FE. This is symptomatic of a larger issue: in general, the exact FEs $F_N(f)$ and $\tilde{F}_N(f)$ differ substantially from their numerical counterparts $G_N(f)$ and $\tilde{G}_N(f)$. As we show in the next section, there are important differences in both their stability \textit{and} their convergence. In particular, any analysis based solely on $F_N$ and $\tilde{F}_N$ is insufficient to describe the behaviour of the numerical extensions $G_N$ and $\tilde{G}_N$. \section{The numerical continuous and discrete Fourier extensions}\label{s:stability} We now analyze the numerical FEs $G_N$ and $\tilde{G}_N$, and describe both when and how they differ from the exact extensions $F_N$ and $\tilde{F}_N$. \subsection{The norm of the exact FE coefficients}\label{ss:magnitude} In short, the reason for this difference is as follows. Since the FE matrices $A$ and $\tilde A$ are so ill-conditioned, the coefficients of the exact FEs $F_N$ and $\tilde{F}_N$ will not usually be obtained in finite precision computations. To explain exactly how this affects stability and convergence, we first need to determine when this will occur. We require the following theorem: \thm{ \label{t:coefficients} Suppose that $f$ is analytic in $\cD(\rho^*)$ and continuous on its boundary. If $a \in \bbC^{2N+1}$ is the vector of coefficients of the continuous FE $F_N(f)$ then \be{ \label{coefbd1} \| a \| \leq c_f \left \{ \begin{array}{ll} \left ( \frac{E(T)}{\rho^*} \right )^N & \rho^* < E(T ), \\ N & \rho^* \geq E(T), \end{array} \right. } where $c_f$ is proportional to $\max_{x \in \cD(\rho)} | f(x) |$. If $f \in \rL^2(-1,1)$, then \be{ \label{coefbd2} \| a \| \leq c \| f \| E(T)^N, } for some $c>0$ independent of $f$ and $N$. } \prf{ Write $F_N(f) = f_N = f_{e,N} + f_{o,N}$, where $f_{e,N}$ and $f_{o,N}$ are the even and odd parts of $f_N$ respectively. Since the set $\{ \phi_n \}_{n \in \bbZ}$ is orthonormal over $[-T,T]$ we find that \eas{ \| a \| = \| f_N \|_{[-T,T]} \leq 2 \left ( \| f_{e,N} \|_{[0,T]} + \| f_{o,N} \|_{[0,T]} \right ) \leq 2 \sqrt{T} \left ( \| f_{e,N} \|_{\infty,[0,T]} + \| f_{o,N} \|_{\infty,[0,T]} \right ). } Recall from \S \ref{sss:FEpoly} that $f_{e,N}(x) = h_{1,N}(z)$ and $f_{o,N}(x) =\sin \left ( \tfrac{\pi}{T} m^{-1}(z) \right ) h_{2,N}(z)$, where $h_{i,N} \in \bbP_{N+1-i}$, $i=1,2$, is defined by \R{E:hNdef}. Thus, $\| a \| \leq c \left ( \| h_{1,N} \|_{\infty,[m(T),1]} + \| h_{2,N} \|_{\infty,[m(T),1]} \right )$ for some $c >0$. Consider $h_{1,N}(z)$. This is precisely the expansion of the function $h_1(z) = f_1(m^{-1}(z))$ in polynomials $\{ p_n \}^{\infty}_{n=0}$ orthogonal with respect to the weight function $w_{1}$: i.e.\ $h_{1,N} = \sum^{N}_{n=0} \ip{h_1}{p_n}_{w_1} p_n$. Therefore \bes{ \| h_{1,N} \|_{\infty,[m(T),1]} \leq \sum^{N}_{n=0} | \ip{h_1}{p_n}_{w_1} | \| p_n \|_{\infty,[m(T),1]}. } It is known that $\| p_n \|_{\infty,[m(T),1]} \leq c E(T)^n$ \cite{huybrechs2010fourier}. Also, since $h_1$ is analytic in $\cB(\rho^*)$ we have $| \ip{h_1}{p_n}_{w_1} | \leq c_{f} (\rho^*)^{-n}$. Hence \bes{ \| h_{1,N} \|_{\infty,[m(T),1]} \leq c_f \sum^{N}_{n=0} \left ( E(T) / \rho^* \right )^{n}, } which gives \R{coefbd1}. For \R{coefbd2} we use the bound $| \ip{h_1}{p_n}_{w_1} | \leq \| h_1 \|_{w_1} \leq c \| f \|$ instead. } \cor{ Let $f$ be as in Theorem \ref{t:coefficients}. Then the vector of coefficients $a \in \bbC^{2N+2}$ of the discrete Fourier extension $\tilde{F}_N(f)$ of $f$ satisfies the same bounds as those given in Theorem \ref{t:coefficients}. } \prf{The functions $h_{i,N}$, $i=1,2$ are the polynomial interpolants of $h_i$ at the nodes \R{Cheb} (Proposition \ref{p:discinterp}). Write $h_{i,N}(z) = \sum^{N}_{n=0} \tilde d_n T_n(z)$, where $T_n(z)$ is the $n^{\rth}$ Chebyshev polynomial, and let $\hat{d}_n = \ip{h_i}{T_n}_{w}$ be the Chebyshev polynomial coefficient of $h_i$. Note that $| \hat{d}_n | \leq c_f (\rho^*)^{-n}$. Due to aliasing formula $\tilde d_n = \hat{d}_n + \sum_{k \neq 0} (\hat{d}_{2kN+n} + \hat{d}_{2kN-n})$ (see \cite[Eqn. (2.4.20)]{SMSD}) we obtain \bes{ | \tilde d_n | \leq c_f \left ( (\rho^*)^{-n} + \sum^{\infty}_{k=1} (\rho^*)^{-2 k N - n} +\sum^{\infty}_{k=1}(\rho^*)^{-2kN + n} \right ) \leq c_f \left ( (\rho^*)^{-n} + (\rho^*)^{n-2N} \right ) \leq c_f (\rho^*)^{-n}.} The result now follows along the same lines as the proof of Theorem \ref{t:coefficients}. } To compute the continuous or discrete FE we need to solve the linear system $A a = b$ (respectively $\tilde A a =\tilde b$). When $N$ is large, the columns of $A$ ($\tilde A$) become near-linearly dependent, and, as shown in \S \ref{ss:ASVD}, the numerical rank of $A$ is roughly $1/T$ times its dimension. Now suppose we solve $A a = b$ with a standard numerical solver. Loosely speaking, the solver will use the extra degrees of freedom to construct approximate solutions $a'$ with small norms. The previous theorem and corollary therefore suggest the following. In general, only in those cases where $f$ is analytic with $\rho^* \geq E(T)$ can we expect the theoretical coefficient vector $a$ to be produced by the numerical solver for all $N$. Outside of this case, we may well have that $a' \neq a$ for sufficiently large $N$, due to the potential for exponential growth of the latter. Hence, in this case, the numerical extension $G_N(f)$ will not coincide with the exact extension $F_N(f)$. This raises the following question: if the numerical solver does not give the exact coefficients vector, then what does it yield? The following proposition confirms the existence of infinitely many approximate solutions of the equations $A a = b$ with small norm coefficient vectors: \prop{ \label{p:appcoeff} Suppose that $f \in \rH^k(-1,1)$. Then there exist $a^{[N]} \in \bbC^{2N+1}$, $N \in \bbN$, satisfying \be{ \label{coeffbd} \| a^{[N]} \| \leq c_k(T) \| f \|_{\rH^k(-1,1)}, } and \be{ \label{coefferr} \| A a^{[N]} - b \| \leq c_k(T) N^{-k} \| f \|_{\rH^k(-1,1)}, } where $c_k(T)$ is the constant of Lemma \ref{l:smoothextension}. Moreover, if $g_N = \sum_{|n| \leq N} a^{[N]}_{n} \phi_n$ then \be{ \label{apperr} \| f - g_N \| \leq c_k(T) N^{-k} \| f \|_{\rH^k(-1,1)}. } } \prf{ Let $\tilde f \in \rH^k(\bbT)$ be the extension guaranteed by Lemma \ref{l:smoothextension}, and write $a^{[N]}$ for the vector of its first $2N+1$ Fourier coefficients on $\bbT = [-T,T)$. By Bessel's inequality, $\| a^{[N]} \| \leq \| \tilde f \|_{[-T,T]} \leq c_k(T) \| f \|_{\rH^k(-1,1)}$ which gives \R{coeffbd}. For \R{coefferr}, we merely note that $(A a^{[N]} - b)_n = \ip{f - g_N}{\phi_n}$. Using the frame property \R{frameprop} we obtain $\| A a^{[N]} - b \| \leq \| f - g_N \|$. Thus, \R{coefferr} follows directly from \R{apperr}, and the latter is a standard result of Fourier analysis (see \cite[eqn. (5.1.10)]{SMSD}, for example). } This proposition states that there exist vectors with norms bounded independently of $N$ that solve the equations $A a =b$ up to an error of order $N^{-k}$. Moreover, these vectors yield extensions which converge algebraically fast to $f$ at rate $k$. Whilst it does not imply that these are the vectors produced by the numerical solver, it does indicate that, in the case where the exact extension $F_N(f)$ or $\tilde F_N(f)$ has a large coefficient norm, geometric convergence of the numerical extension $G_N(f)$ or $\tilde G_N(f)$ may be sacrificed for superalgebraic convergence so as to retain boundedness of the computed coefficients. \begin{figure}[t] \begin{center} $\begin{array}{ccc} \includegraphics[width=6.25cm]{Fig3Fn2Err} & \hspace{1pc} & \includegraphics[width=6.25cm]{Fig3Fn2Coef} \\ \includegraphics[width=6.25cm]{Fig3Fn1Err} & \hspace{1pc} & \includegraphics[width=6.25cm]{Fig3Fn1Coef} \\ \includegraphics[width=6.25cm]{Fig3Fn3Err} & \hspace{1pc} & \includegraphics[width=6.25cm]{Fig3Fn3Coef} \end{array}$ \caption{\small Comparison of the numerical continuous and discrete FEs $G_N(f)$ and $\tilde{G}_{N}(f)$ (squares and circles) and their exact counterparts $F_N(f)$ and $\tilde{F}_N(f)$ (crosses and diamonds) for $T=2$. Left: the uniform error $\| f - f_N \|_{\infty}$ against $N$. Right: the norm $\| a \|$ of the coefficient vector. Top row: $f(x) = \frac{1}{1+16 x^2}$. Middle row: $f(x) = \frac{1}{8-7 x}$. Bottom row: $f(x) = 1+\frac{\cosh 40 x}{\cosh 40}$. } \label{f:FnAppCoeff} \end{center} \end{figure} This hypothesis is verified numerically in Figure \ref{f:FnAppCoeff} (all computations were carried out in \textit{Mathematica}, with additional precision used to compute the exact FEs and standard precision used otherwise). Geometric convergence of the exact extension is replaced by slower, but still high-order convergence for sufficiently large $N$. Note that the `breakpoint' occurs at roughly the same value of $N$ regardless of the function being approximated. Moreover, the breakpoint occurs at a larger value of $N$ for the discrete extension than for the continuous extension. These observations will be established rigorously in the next section. However, we now make several further comments on Figure \ref{f:FnAppCoeff}. First, note that the breakdown of geometric convergence is far less severe for the classical Runge function $f(x) = \frac{1}{1+16 x^2}$ than for the functions $f(x) = \frac{1}{8-7 x}$ and $f(x) = 1 + \frac{\cosh 40 x}{\cosh 40}$. This can be explained by the behaviour of these functions near $x=\pm 1$. The Runge function $f(x) = \frac{1}{1+16 x^2}$ is reasonably flat near $x = \pm 1$. Hence it possesses extensions with high degrees of smoothness which do not grow large on the extended domain $[-T,T]$. Conversely, the other two functions have boundary layers near $x=1$ (also $x=-1$ for the latter). Therefore any smooth extension will be large on $[-T,T]$, and by Parseval's relation, the coefficient vectors corresponding to the Fourier series of this extension will also have large norm. Second, although it is not apparent from Figure \ref{f:FnAppCoeff} that the convergence rate beyond the breakpoint is truly superalgebraic, this is in fact the case. This is confirmed by Figure \ref{f:FnAppSpectral}. In the right-hand diagram we plot the error against $N$ in log-log scale. The slight downward curve in the error indicates superalgebraic convergence. Had the convergence rate been algebraic of fixed order then the error would have followed a straight line. \begin{figure}[t] \begin{center} $\begin{array}{ccc} \includegraphics[width=6.25cm]{Fig4Log} & \hspace{1pc} &\includegraphics[width=6.25cm]{Fig4LogLog} \\ \end{array}$ \caption{\small Comparison of the numerical continuous and discrete FEs $G_N(f)$ and $\tilde{G}_{N}(f)$ (squares and circles) and their exact counterparts $F_N(f)$ and $\tilde{F}_N(f)$ (crosses and diamonds) for $T=2$ and $f(x) = \frac{1}{101-100 x}$. Left: the uniform error in log scale. Right: the uniform error in log-log scale. } \label{f:FnAppSpectral} \end{center} \end{figure} \subsection{Analysis of the numerical continuous and discrete FEs}\label{ss:numsolnanalysis} We now wish to analyze the numerical extensions $G_N(f)$ and $\tilde{G}_N(f)$. Since the numerical solvers used in environments such as \textit{Matlab} or \textit{Mathematica} are difficult to analyze directly, we shall look at the result of solving $A a = b$ (or $\tilde A a =\tilde b$) with a truncated singular value decomposition (SVD). This represents an idealization of the numerical solver. Indeed, neither \textit{Matlab}'s $\backslash$ or \textit{Mathematica}'s \texttt{LeastSquares} actually performs a truncated SVD. However, in practice, this simplification appears reasonable: numerical experiments indicate that these standard solvers give roughly the same approximation errors as the truncated SVD with suitably small truncation parameter (typically $\epsilon = 10^{-14}$). We shall also assume throughout that the truncated SVD is computed without error. However, this also seems fair: in experiments, we observe that the finite-precision SVD gives similar results to the numerical solver whenever the tolerance is sufficiently small. Suppose that $A$ (respectively $\tilde A$) has SVD $U S V^*$ with $S$ being the diagonal matrix of singular values. Given a truncation parameter $\epsilon > 0$, we now consider the solution \be{ \label{aeps} a_{\epsilon} = V S^{\dag} U^* b, } where $S^\dag$ is the diagonal matrix with $n^{\rth}$ entry $1/\sigma_n$ if $\sigma_n > \epsilon$ and $0$ otherwise. We write \bes{ H_{N,\epsilon}(f) = \sum_{|n| \leq N} (a_{\epsilon})_n \phi_n, } for the corresponding FE. Suppose that $ v_n \in \bbC^{2N+1}$ is the right singular vector of $A$ with singular value $\sigma_n$, and let \bes{ \Phi_n = \sum_{|m| \leq N} (v_n)_m \phi_m \in \cG_N, } be the Fourier series corresponding to $v_n$. Note that the functions $\Phi_n$ are orthonormal with respect to $\ip{\cdot}{\cdot}_{[-T,T]}$ and span $\cG_N$. Also, if we define $\cG_{N,\epsilon} = \spn \{ \Phi_n : \sigma_n > \epsilon \} \subseteq \cG_N$, then we have $H_{N,\epsilon}(f) \in \cG_{N,\epsilon}$. We now consider the cases of the continuous and discrete FEs separately. \subsubsection{The continuous Fourier extension}\label{sss:SVDexact} In this case, since $A$ is Hermitian and positive definite, the singular vectors $v_n$ are actually eigenvectors of $A$ with $A v_n = \sigma_n v_n$. By definition, we have $\ip{\Phi_n}{\Phi_m} = (v_n)^* A v_m = \sigma_n \delta_{n,m}$, and therefore \be{ \label{exactTSVD} H_{N,\epsilon}(f) = \sum_{n : \sigma_n > \epsilon} \frac{1}{\sigma_n} \ip{f}{\Phi_n} \Phi_n. } Our main result is as follows: \thm{ \label{t:exactSVD} Let $f \in \rL^2(-1,1)$ and suppose that $H_{N,\epsilon}(f)$ is given by \R{exactTSVD}. Then \be{ \label{exactTSVDerr} \| f - H_{N,\epsilon}(f) \| \leq \| f - \phi \| + \sqrt{\epsilon} \| \phi \|_{[-T,T]},\quad \forall \phi \in \cG_N, } and \be{ \label{exactTSVDcoeff} \| a_{\epsilon} \| = \| H_{N,\epsilon}(f) \|_{[-T,T]} \leq \frac{1}{\sqrt{\epsilon}} \| f - \phi \| + \| \phi \|_{[-T,T]},\quad \forall \phi \in \cG_N. } } \prf{ The function $H_{N,\epsilon}(f)$ is the orthogonal projection of $f$ onto $\cG_{N,\epsilon}$ with respect to $\ip{\cdot}{\cdot}$. Hence for any $\phi \in \cG_{N}$ we have $\| f - H_{N,\epsilon}(f) \| \leq \| f - H_{N,\epsilon}(\phi)\| \leq \| f - \phi \| + \| \phi - H_{N,\epsilon}(\phi) \|$. Consider the latter term. Since $\phi \in \cG_N$, the observation that the functions $ \Phi_n$ are also orthonormal on $[-T,T]$ gives \bes{ \| \phi - H_{N,\epsilon}(\phi) \|^2 = \left \| \sum_{n : \sigma_n < \epsilon} \ip{\phi}{\Phi_n}_{[-T,T]} \Phi_n \right \|^2= \sum_{n: \sigma_n < \epsilon} \sigma_n |\ip{\phi}{\Phi_n}_{[-T,T]} |^2 \leq \epsilon \| \phi \|^2_{[-T,T]}. } This yields \R{exactTSVDerr}. For \R{exactTSVDcoeff} we first write $\| H_{N,\epsilon}(f) \|_{[-T,T]} \leq \| H_{N,\epsilon}(f-\phi) \|_{[-T,T]} + \| H_{N,\epsilon}(\phi) \|_{[-T,T]}$. By orthogonality, \bes{ \| H_{N,\epsilon}(f-\phi) \|^2_{[-T,T]} = \sum_{n: \sigma_n > \epsilon} \frac{1}{\sigma^2_n} | \ip{f-\phi}{\Phi_n} |^2 \leq \frac{1}{\epsilon} \sum_{n: \sigma_n > \epsilon} \frac{1}{\sigma_n} | \ip{f-\phi}{\Phi_n} |^2 = \frac{1}{\epsilon} \| H_{N,\epsilon}(f-\phi) \|^2. } Since $H_{N,\epsilon}$ is an orthogonal projection, we conclude that $\| H_{N,\epsilon}(f-\phi) \|^2_{[-T,T]} \leq \frac{1}{\epsilon} \| f - \phi \|^2$, which gives the first term in \R{exactTSVDcoeff}. For the second, we notice that \bes{ \| H_{N,\epsilon}(\phi) \|^2_{[-T,T]}=\sum_{n : \sigma_n > \epsilon} | \ip{\phi}{\Phi_n}_{[-T,T]} |^2 \leq \| \phi \|^2_{[-T,T]}, } since $\phi \in \cG_N$. } This theorem allows us to explain the behaviour of the numerical FE $G_N(f)$. Suppose that $f$ is analytic in $\cD(\rho)$ and continuous on its boundary, where $\rho < E(T)$ and $\cD(\rho)$ is as in Theorem \ref{t:expconv}. Set $\phi = F_N(f)$ in \R{exactTSVDerr}, where $F_N(f)$ is the exact continuous FE. Then Theorems \ref{t:expconv} and \ref{t:coefficients} give \be{\label{one} \| f - H_{N,\epsilon}(f) \| \leq c_f \left ( 1 + \sqrt{\epsilon} E(T)^N \right ) \rho^{-N}. } For small $N$, the first term in the brackets dominates, and we see geometric convergence of $H_{N,\epsilon}(f)$, and therefore also $G_N(f)$, at rate $\rho$. Convergence continues as such until the breakpoint \be{ \label{exactbreakpt} N_0 = N_0(\epsilon,T) : = - \frac{\log \epsilon}{2 \log E(T)}, } at which point the second term dominates and the bound begins to increase. On the other hand, Proposition \ref{p:appcoeff} establishes the existence of functions $\phi \in \cG_N$ with bounded coefficients which approximate $f$ to any given algebraic order. Substituting such a function $\phi$ into \R{exactTSVDerr} gives \be{ \| f - H_{N,\epsilon}(f) \| \leq c_k(T) \left ( N^{-k} + \sqrt{\epsilon} \right ) \| f \|_{\rH^k(-1,1)},\quad \forall N, k \in \bbN. } Therefore, once $N > N_0(\epsilon,T)$ we expect at least superalgebraic convergence of $H_{N,\epsilon}(f)$ down to a maximal achievable accuracy of order $\sqrt{\epsilon} \| f \|$. Note that at the breakpoint $N = N_0$, the error satisfies \be{ \label{four} \| f - H_{N_0,\epsilon} (f) \| \leq 2 c_f (\sqrt{\epsilon})^{d_f},\qquad d_f = \frac{\log \rho}{\log E(T)} \in (0,1]. } If $f$ is analytic in $\cD(E(T))$, and if $c_f = \max_{x \in \cD(\rho)} | f(x) |$ is not too large, then $f$ is already approximated to order $\sqrt{\epsilon}$ accuracy at this point. It is only in those cases where either $\rho < E(T)$ or where $c_f$ is large (or both) that one sees the second phase of superalgebraic convergence. Theorem \ref{t:coefficients} also explains the behaviour of the coefficient norm $\| a_{\epsilon} \|$. Observe that breakpoint $N_0(\epsilon,T)$ is (up to a small constant) the largest $N$ for which all singular values of $A$ are included in its truncated SVD (see Theorem \ref{t:exactCondNumb}). Thus, when $N < N_0(\epsilon,T)$, we have $H_{N,\epsilon}(f) = F_{N}(f)$, and Theorem \ref{t:coefficients} indicates exponential growth of $\| a_{\epsilon} \|$. On the other hand, once $N > N_0(\epsilon,T)$, we use \R{exactTSVDcoeff} to obtain \bes{ \| a_{\epsilon} \| \leq c_k(T) \left ( N^{-k} / \sqrt{\epsilon} + 1 \right ) \| f \|_{\rH^k(-1,1)},\quad \forall N, k \in \bbN. } In particular, for $N > N_{0}(\epsilon,T)$, we expect decay of $\| a_{\epsilon} \|$ down from its maximal value at $N = N_{0}(\epsilon,T)$. \begin{figure}[t] \begin{center} $\begin{array}{ccc} \includegraphics[width=6.25cm]{Fig5Fn2Err} & \hspace{1pc} & \includegraphics[width=6.25cm]{Fig5Fn2Coef} \\ \includegraphics[width=6.25cm]{Fig5Fn1Err} & \hspace{1pc} & \includegraphics[width=6.25cm]{Fig5Fn1Coef} \\ \includegraphics[width=6.25cm]{Fig5Fn3Err} & \hspace{1pc} & \includegraphics[width=6.25cm]{Fig5Fn3Coef} \end{array}$ \caption{\small Error (left) and coefficient norm (right) against $N$ for the continuous FE with $T=2$, where $f(x) = \frac{1}{1+16 x^2}$ (top row), $f(x)= \frac{1}{8-7 x} $ (middle row) and $f(x) = x$ (bottom row). Squares, circles, crosses and diamonds correspond to the truncated SVD extension $H_{N,\epsilon}(f)$ with $\epsilon = 10^{-6},10^{-12},10^{-18},10^{-24}$ respectively, and pluses correspond to the exact extension $F_N(f)$.} \label{f:ManSVDexact} \end{center} \end{figure} This analysis is corroborated in Figure \ref{f:ManSVDexact}, where we plot the error and coefficient norm for the truncated SVD extension for various test functions. Note that the maximal achievable accuracy in all cases is order $\sqrt{\epsilon}$, consistently with our analysis. Moreover, for the meromorphic functions $f(x) = \frac{1}{1+16 x^2}$ and $f(x) = \frac{1}{8-7 x}$ we see initial geometric convergence followed by slower convergence after $N_0$, again as our analysis predicts. The qualitative difference in convergence for these functions in the regime $N > N_0$ is due to the contrasting behaviour of their derivatives (recall the discussion in \S \ref{ss:magnitude}). On the other hand, the convergence effectively stops at $N_0$ for $f(x) = x$, since this function has small constant $c_f$ and is therefore already resolved down to order $\sqrt{\epsilon}$ when $N = N_0$. Since $N_0(10^{-6},2) \approx 4$, $N_0(10^{-12},2) \approx 8$, $N_0(10^{-18},2) \approx 12,$ and $N_0(10^{-24},2) \approx 16$, Figure \ref{f:ManSVDexact} also confirms the expression \R{exactbreakpt} for the breakpoint in convergence. In particular, the breakpoint is independent of the function being approximated. This latter observation is unsurprising. As noted, $N_0(\epsilon,T)$ is the largest value of $N$ for which $H_{N,\epsilon}(f)$ coincides with $F_N(f)$. Beyond this point, $H_{N,\epsilon}(f)$ will not typically agree with $F_N(f)$, and thus we cannot expect further geometric convergence in general. Note that our analysis does not rule out geometric convergence for $N > N_0$. There may well be certain functions for which this occurs. However, extensive numerical tests suggest that in most cases, one sees only superalgebraic convergence in this regime, and indeed, this is all that we have proved. \rem{ \label{r:singularfunctions} At first sight, it may appear counterintuitive that one can still obtain good accuracy when excluding all singular values below a certain tolerance. However, recall that we are not interested in the accuracy of computing $a$, but rather the accuracy of $F_N(f)$ on the domain $[-1,1]$. Since the $n^{\rth}$ singular value $\sigma_n$ is equal to $\| \Phi_n \|^2 / \| \Phi_n \|^2_{[-T,T]}$, the functions $\Phi_n$ excluded from $H_{N,\epsilon}(f)$ are precisely those for which $\| \Phi_n \|^2 < \epsilon \| \Phi_n \|^2_{[-T,T]}$. In other words, they have little effect on $F_{N}(f)$ in $[-1,1]$. In Figure \ref{f:SingFnPlot} we plot the functions $\Phi_n$ for several $n$. Note that these functions are precisely the discrete prolate spheroidal wavefunctions of Slepian \cite{SlepianV}. As predicted, when $n$ is small, the function $\Phi_n$ is large in $[-1,1]$ and small in $[-T,T] \backslash [-1,1]$. When $n$ is in the transition region ($n \approx 2N/T$, see \S \ref{ss:ASVD}), the function $\Phi_n$ is roughly of equal magnitude in both regions, and for $n \approx 2N$, $\Phi_n$ is much smaller in $[-1,1]$ than on $[-T,T]$. Note also that $\Phi_n$ is increasingly oscillatory in $[-1,1]$ as $n$ increases, and decreasingly oscillatory in $[-T,T] \backslash [-1,1]$. This follows from the fact that $\Phi_n$ has precisely $n$ zeroes in $[-1,1]$ and $2N-n$ zeroes in $[-T,T] \backslash [-1,1]$ \cite{SlepianV}. Such behaviour also implies that any `nice' function will eventually be well approximated by functions $\Phi_n$ corresponding to `nice' eigenvalues, as expected. \begin{figure} \begin{center} $\begin{array}{ccc} \includegraphics[width=4.75cm]{Sing1} & \includegraphics[width=4.75cm]{Sing2} & \includegraphics[width=4.75cm]{Sing3} \end{array}$ \caption{\small The SVD functions $|\Phi_n(x)|$ for $n=0$, $n=20$ and $n=40$, where $N = 20$ and $T=2$.} \label{f:SingFnPlot} \end{center} \end{figure} } \subsubsection{The discrete Fourier extension}\label{sss:discSVDanalysis} In this case, we have $( \Phi_n , \Phi_m)_N = \sigma^2_n \delta_{n,m}$, where \bes{ (f,g)_N = \frac{\pi}{N+1} \sum^{N}_{n=-N-1} f(x_n) \overline{g(x_n)}, } is the discrete inner product corresponding to the quadrature nodes $\{ x_n \}^{N}_{n=-N-1}$. Therefore \be{ \label{discTSVD} \tilde{H}_{N,\epsilon}(f) = \sum_{n : \sigma_n > \epsilon} \frac{1}{\sigma^2_n} (f,\Phi_n)_N \Phi_n \in \cG'_{N,\epsilon} : = \spn \left \{ \Phi_n : \sigma_n > \epsilon \right \}, } is the orthogonal projection of $f$ onto $\cG'_{N,\epsilon}$ with respect to the discrete inner product $(\cdot,\cdot)_{N}$. \thm{ Let $f \in \rL^\infty(-1,1)$ and $\tilde{H}_{N,\epsilon}(f)$ be given by \R{discTSVD}. Then\be{ \label{discTSVDerr} \| f - \tilde{H}_{N,\epsilon}(f) \|_{W} \leq \| f - \phi \|_W + \sqrt{2 \pi Q(N;\epsilon)} \| f - \phi \|_{\infty} + \epsilon \| \phi \|_{[-T,T]},\quad \forall \phi \in \cG_N, } and \be{ \label{discTSVDcoeff} \| a_{\epsilon} \| = \| \tilde{H}_{N,\epsilon}(f) \|_{[-T,T]} \leq \frac{1}{\epsilon} \sqrt{2 \pi Q(N;\epsilon)} \| f - \phi \|_{\infty} + \| \phi \|_{[-T,T]},\quad \forall \phi \in \cG_N, } where $Q(N;\epsilon) = | \{ n : \sigma_n > \epsilon \} | \leq 2(N+1)$ and $W$ is the weight function of Lemma \ref{l:weightedOpt}. } \prf{ By the triangle inequality, \bes{ \| f - \tilde{H}_{N,\epsilon}(f) \|_W \leq \| f - \phi \|_W + \| \phi - \tilde{H}_{N,\epsilon}(\phi) \|_W + \| \tilde{H}_{N,\epsilon}(f-\phi) \|_W,\quad \forall \phi \in \cG'_N. } Consider the second term. Since $\phi \in \cG'_{N}$, and the quadrature is exact on $\cG'_N$, we have \bes{ \| \phi - \tilde{H}_{N,\epsilon}(\phi) \|^2_W = (\phi - \tilde{H}_{N,\epsilon}(\phi),\phi - \tilde{H}_{N,\epsilon}(\phi))_N = \sum_{n : \sigma_n < \epsilon} \sigma^2_n |\ip{\phi}{\Phi_n}_{[-T,T]}|^2 \leq \epsilon^2 \| \phi \|^2_{[-T,T]}. } For the third term, let $g$ be arbitrary. Then $ (\tilde{H}_{N,\epsilon}(g),\tilde{H}_{N,\epsilon}(g))_N = \sum_{n : \sigma_n > \epsilon} \frac{1}{\sigma^2_n} | (g,\Phi_n)_N |^2$. Hence \be{ \label{step1} \| \tilde{H}_{N,\epsilon}(g) \|^2_{W} =(\tilde{H}_{N,\epsilon}(g),\tilde{H}_{N,\epsilon}(g))_N \leq (g,g)_N \sum_{n : \sigma_n > \epsilon} \frac{1}{\sigma^2_n} (\Phi_n,\Phi_n)_N =(g,g)_N Q(N;\epsilon), } since $(\Phi_n,\Phi_n)_N = \sigma^2_n$. It is straightforward to show that $(g,g)_N \leq 2 \pi \| g \|^2_{\infty}$. Setting $g = f - \phi$ now gives the corresponding term in \R{discTSVDerr}, and completes its proof. For \R{discTSVDcoeff}, we proceed as in the proof of Theorem \ref{t:exactSVD}. Note that \be{ \label{step2} \| \tilde{H}_{N,\epsilon}(g) \|^2_{[-T,T]} = \sum_{n : \sigma_n > \epsilon} \frac{1}{\sigma^4_n} | (g,\Phi_n)_N |^2 \leq \frac{1}{\epsilon^2} \| \tilde{H}_{N,\epsilon}(g) \|^2_W, } for any $g \in \rL^\infty(-1,1)$. Also, \be{ \label{step3} \| \tilde{H}_{N,\epsilon}(\phi) \|_{[-T,T]} \leq \| \phi \|_{[-T,T]}, \quad \phi \in \cG_N. } The result now follows by writing $\| \tilde{H}_{N,\epsilon}(f) \|_{[-T,T]} \leq \| \tilde{H}_{N,\epsilon}(f-\phi) \|_{[-T,T]} +\| \tilde{H}_{N,\epsilon}(\phi) \|_{[-T,T]}$ and using \R{step1}--\R{step3}. } \begin{figure}[t] \begin{center} $\begin{array}{ccc} \includegraphics[width=6.25cm]{Fig7Fn2Err} & \hspace{1pc} & \includegraphics[width=6.25cm]{Fig7Fn2Coef} \\ \includegraphics[width=6.25cm]{Fig7Fn1Err} & \hspace{1pc} & \includegraphics[width=6.25cm]{Fig7Fn1Coef} \\ \includegraphics[width=6.25cm]{Fig7Fn3Err} & \hspace{1pc} & \includegraphics[width=6.25cm]{Fig7Fn3Coef} \end{array}$ \caption{\small Error (left) and coefficient norm (right) against $N$ for the discrete FE with $T=2$, where $f(x) = \frac{1}{1+16 x^2}$ (top row), $f(x) = \frac{1}{8-7 x}$ (middle row) and $f(x) = x$ (bottom row). Squares, circles, crosses and diamonds correspond to the truncated SVD extension $H_{N,\epsilon}(f)$ with $\epsilon = 10^{-6},10^{-12},10^{-18},10^{-24}$ respectively, and pluses correspond to the exact extension $F_N(f)$.} \label{f:ManSVDdisc} \end{center} \end{figure} As with the continuous FE, this theorem allows us to analyze the numerical discrete extension $\tilde{G}_N(f)$. Once more we deduce geometric convergence in $N$ up to the function-independent breakpoint \be{ \label{N1def} N_1(T;\epsilon) : = - \frac{\log \epsilon}{\log E(T)} \equiv 2 N_0(T;\epsilon), } with superalgebraic convergence beyond this point. These conclusions are confirmed in Figure \ref{f:ManSVDdisc}. Note, however, two key differences between the continuous and discrete FE. First, the bound \R{discTSVDerr} involves $\epsilon$, as opposed to $\sqrt{\epsilon}$, meaning that we expect convergence of $\tilde{G}_N(f)$ down to close to machine precision. Second, the breakpoint $N_{1}(T;\epsilon)$ is precisely twice $N_0(T;\epsilon)$. Hence, the regime of geometric convergence of $\tilde{G}_N(f)$ is exactly twice as large as that of the continuous FE. These observations are in close agreement with the behaviour seen in the numerical examples in \S \ref{ss:magnitude}. \subsection{The condition numbers of the numerical continuous and discrete FEs}\label{ss:stabfn} Having analyzed the convergence of the numerical FE---and in particular, established 5.\ of \S \ref{s:introduction}---we next address its condition number. Once more, we do this by considering the extensions $H_{N,\epsilon}$ and $\tilde{H}_{N,\epsilon}$: \thm{ \label{t:stabfn} Let $H_{N,\epsilon}$ be the continuous truncated SVD FE given by \R{exactTSVD}. Then \bes{ \kappa(H_{N,\epsilon})=1/ \min \{ \sqrt{\sigma_n} : \sigma_n > \epsilon \} \leq \min \left \{ 1/\sqrt{\epsilon} , c(T) N^{\frac32} E(T)^N \right \},\quad N \in \bbN,\ \epsilon > 0, } where $c(T)$ is a positive constant independent of $N$. Conversely, if $\tilde{H}_{N,\epsilon}$ is the discrete extension \R{discTSVD}, then $\kappa( \tilde{H}_{N,\epsilon})= 1$ for all $N \in \bbN$ and $\epsilon > 0$. } \prf{ The proof of the equalities is similar to that of Lemma \ref{l:ExactCondNumb} with $A$ and $\tilde{A}$ replaced by their truncated SVD versions. The upper bound for $\kappa(H_{N,\epsilon})$ is a consequence of Theorem \ref{t:exactCondNumb}. } This theorem, which establishes 3.\ of \S \ref{s:introduction}, has some interesting consequences. First, the discrete FE is perfectly stable. On the other hand, the numerical continuous FE is far from stable. The condition number grows exponentially fast at rate $E(T)$ until it reaches $1/\sqrt{\epsilon}$, where $\epsilon$ is the truncation parameter in the SVD. Thus, with the continuous FE, we may see perturbations being magnified by a factor of $1/\sqrt{\epsilon_{\mathrm{mach}}} \approx 10^8$ in practice. Note that $G_N$ and $\tilde{G}_N$ are both substantially better conditioned than the corresponding coefficient mappings. The explanation for this difference comes from Remark \ref{r:singularfunctions}. A perturbation $\eta$ in the input vector $b$ gives large errors in the FE coefficients if $\eta$ has a significant component in the direction of a singular vector $v_n$ associated with a small singular value $\sigma_n$. However, since the corresponding function $\Phi_n$ is small on $[-1,1]$, this error is substantially reduced (in the case of the continuous FE) or cancelled out altogether (for the discrete FE) in the resulting extension. Another implication of Theorem \ref{t:stabfn} is the following: varying $T$ has no substantial effect on stability. Although the condition number of the FE matrices depends on $T$ (recall Theorems \ref{t:exactCondNumb} and \ref{t:discCondNumb}), as does the condition number of the exact continuous FE (see Lemma \ref{l:ExactCondNumb}), the condition numbers of the numerical mappings $\tilde{G}_N$ and, for all large $N$, $G_N$ are actually independent of this parameter. It is important to confirm that the results of this theorem on the condition number of the truncated SVD extensions predict the behaviour of the numerical extensions $G_N$ and $\tilde{G}_N$. It is easiest to do this by computing upper bounds for $\kappa(G_N)$ and $\kappa(\tilde{G}_N)$. Let $\{ e_n \}^{2N+1}_{n=1}$ be the standard basis for $\bbC^{2N+1}$. Then a simple argument gives that \be{ \| G_N(b) \| \leq \| b \| \sqrt{\sum^{2N+1}_{n=1} \| G_N(e_n) \|^2 },\quad \forall b \in \bbC^{2N+1}, } and therefore \be{ \label{KGN_def} \kappa(G_N) \leq K(G_N) : = \sqrt{\sum^{2N+1}_{n=1} \| G_N(e_n) \|^2 }. } We define the upper bound $K(\tilde{G}_N)$ in a similar manner: \bes{ \kappa(\tilde{G}_N) \leq K(\tilde{G}_N) : = \sqrt{\sum^{2N+2}_{n=1} \| \tilde{G}_N(e_n) \|^2_W}. } In Table \ref{t:FEStabFn} we show $K(G_N)$ and $K(\tilde{G}_N)$ for various choices of $N$. As we see, the discrete FE is extremely stable: not only is there no blowup in $N$, but the value of $K(\tilde{G}_N)$ is also close to one in magnitude. For the continuous extension, we see that $K(G_N) \approx 5 \times 10^{6} =1/\sqrt{\epsilon}$, where $\epsilon = 2.5 \times 10^{-13}$. This behaviour is in good agreement with Theorem \ref{t:stabfn}. \begin{table} \begin{center} \begin{tabular}{|c||c|c|c|c|c|} \hline $N$ & 40 & 80 & 120 & 160 & 200\\ \hline $K(G_N)$ & $4.93 \times10^6$ & $4.22 \times 10^6$ & $3.30 \times 10^6$ & $3.82 \times 10^6$ & $5.28\times 10^6$ \\ \hline $K(\tilde{G}_N)$ & $8.00\times10^0$ & $1.04\times10^1$ & $1.23\times10^1$ & $1.39\times10^1$ & $1.53\times10^1$ \\ \hline \end{tabular} \caption{The functions $K(G_N)$ and $K(\tilde{G}_N)$ for $T=2$.}\label{t:FEStabFn} \end{center} \end{table} The difference in stability between the continuous and discrete FEs is highlighted in Figure \ref{f:Noisy}. Here we perturbed the right-hand side $b$ of the function $f(x) = \E^x$ by noise of magnitude $\delta$, and then computed its FE. As is evident, the discrete extension approximates $f$ to an error of magnitude roughly $\delta$, whereas for the continuous extension the error is of magnitude $\approx 10^{6} \delta$, as predicted by Table \ref{t:FEStabFn}. \begin{figure} \begin{center} $\begin{array}{ccc} \includegraphics[width=6.25cm]{NoisyFnAppExact_New} & \hspace{1pc} & \includegraphics[width=6.25cm]{NoisyFnAppDisc_New} \end{array}$ \caption{\small The error $| f(x) - f_N(x) |$ against $x$, where $f_N = G_N(f)$ (left) or $f_N = \tilde{G}_N(f)$ (right), for $N = 30$, $T=2$ and $f(x) = \E^x$, with noise at amplitudes $\delta = 10^{-4},10^{-8},10^{-12},0$.} \label{f:Noisy} \end{center} \end{figure} \section{Fourier extensions from equispaced data}\label{s:equispacedI} We now turn our attention to the problem of computing FEs when only equispaced data is prescribed. As discussed in \S \ref{s:introduction}, a theorem of Platte, Trefethen \& Kuijlaars \cite{TrefPlatteIllCond} states that any exponentially-convergent method for this problem must also be exponentially ill-conditioned (see \S \ref{ss:PTKrelation} for the precise result). However, as we show in this section, FEs give rise to a method, the so-called \textit{equispaced Fourier extension}, that allows this barrier to be circumvented to a substantial extent. Namely, it achieves rapid convergence in a numerically stable manner. \subsection{The equispaced Fourier extension}\label{ss:equiFEdef} Let \be{ \label{equinodes} x_{n} = \frac{n}{M},\quad n=-M,\ldots,M, } be a set of $2M+1$ equispaced points in $[-1,1]$, where $M \geq N$. We define the \textit{equispaced Fourier extension} of a function $f \in \rL^\infty[-1,1]$ by \be{ \label{equiL2} F_{N,M}(f) : = \underset{\phi \in \cG_N}{\operatorname{argmin}} \sum_{|n| \leq M} | f(x_n) - \phi(x_n) |^2. } If $F_{N,M}(f) = \sum_{|n| \leq N} a_n \phi_n$, then the vector $a = (a_{-N},\ldots,a_N)^{\top}$ is the least squares solution to $\bar{A} a \approx \bar{b}$, where $\bar{A} \in \bbC^{(2M+1) \times (2N+1)}$ has $(n,m)^{\rth}$ entry $\frac{1}{\sqrt{M+1/2}} \phi_m(x_n)$ and $\bar{b}$ has $n^{\rth}$ entry $\frac{1}{\sqrt{M+1/2}} f(x_n)$. Note that $F_{N,M}(f)$, as defined by \R{equiL2}, is (up to minor changes of parameters/notation) identical to the extensions considered in the previous papers \cite{BoydFourCont,BoydRunge,brunoFEP,LyonFESVD,LyonFast} on equispaced FEs. \subsection{The exact equispaced Fourier extension}\label{ss:equiFEthy} Consider first the case $M=N$. Then $F_{N,N}(f)$ is equivalent to polynomial interpolation in $z$: \prop{ Let $F_{N,N}(f)= f_N = f_{e,N}+f_{o,N} \in \cG_N$ be defined by \R{equiL2} with $N=M$ and let $h_{i,N}(z)$ be given by \R{E:hNdef}. Then $h_{i,N}(z)$, $i=1,2$ is the $(N+1-i)^{\rth}$ degree polynomial interpolant of $h_i(z)$ at the nodes $\{ z_n \}^{N}_{n=i-1} \subseteq [-1,1]$, where \be{ \label{unodes} z_n = m(x_{n}) = 2 \frac{\cos \left ( \tfrac{n \pi}{NT} \right ) - c(T)}{1-c(T)} -1,\quad n=0,\ldots,N. }} This proposition allows us to analyze the theoretical convergence/divergence of $F_{N,N}(f)$ using standard results on polynomial interpolation. Recall that associated with a set of nodes $\{ z_n \}^{N}_{n=0}$ is a \textit{node density function} $\mu(z)$, i.e.\ a function such that (i) $\int^{1}_{-1} \mu(z) \D z = 1$ and (ii) each small interval $[z,z+h]$ contains a total of $N \mu(z) h$ nodes for large $N$ \cite{Fps}. In the case of \R{unodes} we have \lem{ The nodes \R{unodes} have node density function $\mu(z) = T /(\pi \sqrt{(1-z)(z-m(T))})$. } \prf{ Note first that $\int^{1}_{-1} \mu(z) \D z = 1$. Now let $I = [z,z+h] \subseteq [-1,1]$ be an interval. Then the node $z_n \in I$ if and only if $m^{-1}(z+h) \leq x_n \leq m^{-1}(z)$. Therefore, as $N \rightarrow \infty$, the proportion of nodes lying in $I$ tends to $m^{-1}(z) - m^{-1}(z+h)$. Now suppose that $h \rightarrow 0$. Then \eas{ m^{-1}(z+h)&=\frac{T}{\pi} \arccos \left [ c(T) + \frac{1-c(T)}{2} (z+h+1) \right ] = m^{-1}(z) - \mu(z) h + \ord{h^2}. } Thus $m^{-1}(z) - m^{-1}(z+h)= \mu(z) h + \ord{h^2}$, as required. } It is useful to consider the behaviour of $\mu(z)$. When $z \rightarrow 1^{-}$, $\mu(z) \sim T /(\pi \sqrt{1-z})$. On the other hand, $\mu$ is continuous at $z = -1$ with $\mu(-1) = \frac{T}{2 \pi} \tan \left ( \frac{\pi}{2T} \right )$. Hence the nodes $\{ z_n \}^{N}_{n=0}$ cluster quadratically near $z = 1$ and are linearly distributed near $z=-1$. It is well known that to avoid the Runge phenomenon in a polynomial interpolation scheme, it is essentially necessary for the nodes to cluster quadratically near both endpoints (as is the case with Chebyshev nodes) \cite{Fps}. If this is not the case, one expects the Runge phenomenon: that is, divergence (at a geometric rate) of the interpolant for any function having a singularity in a certain complex region containing $[-1,1]$ (the \textit{Runge region} for the interpolation scheme). Since the nodes \R{unodes} do not exhibit the correct clustering at the endpoint $z=-1$, we consequently expect this behaviour in the equispaced FE $F_{N,N}(f)$. As it transpires, the corresponding Runge region $\cR = \cR(T)$ for $F_{N,N}$ can be defined in terms of the potential function $\phi(t) = - \int^{1}_{-1} \mu(z) \log | t - z | \D z + c$. Here $c$ is an arbitrary constant. Standard polynomial interpolation theory \cite{Fps} then gives that \bes{ \cR(T)= \left \{ x \in \bbC : \phi(m(x)) = \phi(-1) \right \}, } (observe that this is a subset of the complex $x$-plane). We note also that the convergence/divergence of $F_{N,N}(f)$ at a point $x$ will be exponential at a rate $\E^{\phi(m(x_0)) - \phi(m(x))}$, where $x_0$ is the limiting singularity of $f$. This follows from a general result on polynomial interpolation \cite{Fps}. In particular, if $f$ has a singularity in $\cR(T)$, then there will be some points $x \in [-1,1]$ for which $F_{N,N}(f)$ diverges. We next discuss two approaches for overcoming the Runge phenomenon in $F_{N,N}(f)$. \subsubsection{Overcoming the Runge phenomenon I: varying $T$}\label{sss:oversamp} One way to attempt to overcome (or, at least, mitigate) the Runge phenomenon observed above is to vary the parameter $T$. Note that: \lem{ The Runge region $\cR(T)$ satisfies $\cR(T) \rightarrow [-1,1]$ as $T \rightarrow 1^+$, and $\cR(T) \rightarrow \cR$ as $T \rightarrow \infty$, where $\cR$ is the Runge region for equispaced polynomial interpolation. } \prf{ Suppose first that $T \rightarrow 1^+$. Since $m(T) \sim -1$, we have $\mu(z) \sim \frac{1}{\pi \sqrt{1-z^2}}$. The right-hand side is the potential function for Chebyshev interpolation, and thus the first result follows. For the second result, we first recall that $\phi(m(x)) = - \int^{1}_{-1} \mu(z) \log | m(x) - z | \D z$. Define the change of variable $z = m(s)$. Since $m'(s) = - 1/\mu(m(s))$ we have \bes{ \phi(m(x)) =- \int^{1}_{0} \log | m(x) - m(s) | \D s . } Note that \bes{ m(x) - m(s) = \frac{\cos \frac{\pi x}{T} - \cos \frac{\pi s}{T}}{\sin^2 \frac{\pi}{2T}} =- \frac{2 \sin \tfrac{\pi(x-s)}{2T} \sin \tfrac{\pi(x+s)}{2T} }{\sin^2 \frac{\pi}{2T}} \sim - 2 (x-s)(x+s),\quad T \rightarrow \infty. } Therefore \eas{ \phi(m(x)) \sim - \int^{1}_{-1} \log \left | x - s\right | \D s + C,\quad T \rightarrow \infty, } which is the potential function of equispaced polynomial interpolation, as required. } This lemma comes as no surprise. As $T \rightarrow 1^+$ for fixed $N$, the system $\{ \E^{\I \frac{n \pi}{T} \cdot } \}_{|n| \leq N}$ tends to the standard Fourier basis on $[-1,1]$. The problem of equispaced interpolation with trigonometric polynomials is well-conditioned and convergent. On the other hand, when $T \rightarrow \infty$, the subspaces $\cC_{N}$ and $\cS_N$ both resemble spaces of algebraic polynomials in $x$. Thus, in the large $T$ limit, $F_{N,N}(f)$ is an algebraic polynomial interpolant of $f$ at equispaced nodes. Since the Runge region $\cR(T)$ can be made arbitrarily small by letting $T \rightarrow 1^+$, one way to overcome the Runge phenomenon is to vary $T$ in the way described in \S \ref{ss:Tchoice} and set $T = T(N;\epsilon)$. One could also take $T \approx 1$ fixed. However, this will always lead to a nontrivial Runge region, and consequently divergence of $F_{N,N}$ for some nonempty class of analytic functions. \subsubsection{Overcoming the Runge phenomenon II: oversampling} An alternative means to overcome the Runge phenomenon in $F_{N,M}(f)$ is to allow $M \geq N$. Oversampling is known to defeat the Runge phenomenon in equispaced polynomial interpolation \cite{boyd2009divergence,BoydRunge,TrefPlatteIllCond}, and the same is true in this context (see \cite{BoydFourCont,brunoFEP} for previous discussions on oversampling for equispaced FEs). It is now useful to introduce some notation. For nodes $\{ x_n \}_{|n| \leq M}$ given by \R{equinodes}, let $(\cdot,\cdot)_M$ be the discrete bilinear form $(g,h)_M = \frac{1}{M+\frac12} \sum_{|n| \leq M} g(x_n) \overline{h(x_n)}$, and denote the corresponding discrete semi-norm by $\nm{\cdot}_{M}$. Much as before, we define the condition number of $F_{N,M}$ by \be{ \label{FNM_cond} \kappa(F_{N,M} ) = \sup \left \{ \| F_{N,M}(b) \| : b \in \bbC^{2M+1},\ \nm{b} = 1 \right \}. } We now have: \thm{ \label{t:oversamp} Let $F_{N,M}(f)$ be given by \R{equiL2}, and suppose that \be{ \label{DNM} D(N,M) = \sup \left \{ \|\phi \| : \phi \in \cG_N,\ \| \phi \|_M = 1 \right \}, } then \bes{ \| f - F_{N,M}(f) \| \leq \sqrt{2}\left ( 1+D(N,M) \right ) \inf_{\phi \in \cG_N} \| f - \phi \|_{\infty}. } Moreover, the condition number $\kappa(F_{N,M}) = D(N,M)$. } \prf{ For the sake of brevity, we omit the first part of the proof (a very similar argument is given in \cite{boyd2009divergence} for the case of polynomial interpolation). For the second part, we first notice that \bes{ \kappa(F_{N,M}) = \sup \left \{ \| F_{N,M}(f) \| : f \in \rL^\infty(-1,1),\ \nm{f}_M = 1 \right \}. } Since $F_{N,M}(\phi) = \phi$ for $\phi \in \cG_N$ we have $\kappa(F_{N,M}) \geq D(N,M)$. Conversely, since $F_{N,M}(f) \in \cG_N$, and since $F_{N,M}$ is an orthogonal projection with respect to the bilinear form $(\cdot,\cdot)_M$, we have $\| F_{N,M}(f) \| \leq D(N,M) \| F_{N,M}(f) \|_M \leq D(N,M) \| f \|_M$. Hence $\kappa(F_{N,M}) \leq D(N,M)$, and we get the result. } This theorem implies that $F_{N,M}(f)$ will converge, regardless of the analyticity of $f$, provided $M$ is chosen such that $D(N,M)$ is bounded. Note that this is always possible: $D(N,M) \rightarrow 1$ as $M \rightarrow \infty$ for fixed $N$ since $\nm{\cdot}_M$ is a Riemann sum approximation to $\nm{\cdot}$ and $\cG_N$ is finite-dimensional. Up to small algebraic factors in $M$ and $N$, the quantity $D(N,M)$ is equivalent to \be{ \label{tilDNM} \tilde D(N,M) = \sup \left \{ \| p \|_{\infty} : p \in \bbP_N,\ | p(z_n) | \leq 1,\ n=0,\ldots,M \right \}. } Note the meaning of $\tilde D(N,M)$: it informs us how large a polynomial of degree $N$ can be on $[-1,1]$ if that polynomial is bounded at the $M$ points $z_n$. Unfortunately, numerical evidence suggests that \be{ \label{coppriv} \alpha^{\frac{N^2}{M}} \leq \tilde D(N,M) \leq \beta^{\frac{N^2}{M}}, } for constants $\beta \geq \alpha > 1$. Thus one requires $M = \ord{N^2}$ nodes for boundedness of $D(N,M)$. This is clearly less than ideal: it means that we require many more samples of $f$ to compute its $N$-term equispaced FE. In particular, the exact equispaced FE $F_{N,M}(f)$ of an analytic function $f$ will converge only root-exponentially fast in the number $M$ of equispaced grid values. Had the nodes $\{ z_n \}^{M}_{n=0}$ clustered quadratically near $z= \pm 1$, then $M = \ord{N}$ would be sufficient to ensure boundedness of $\tilde D(N,M)$. Note that when $N=M$, $\tilde D(N,M)$ is precisely the Lebesgue constant of polynomial interpolation. On the other hand, if $\{ z_n \}^{M}_{n=0}$ were equispaced nodes on $[-1,1]$ then \R{coppriv} would coincide with a well-known result of Coppersmith \& Rivlin \cite{copprivlinpolygrowth}. The intuition for a bound of the form \R{coppriv} for the nodes \R{unodes} comes from the fact that these nodes are linearly distributed near $z=-1$. Thus, at least near $z=-1$ they behave like equispaced nodes. We remark that it is straightforward to show that the scaling $M = \ord{N^2}$ is sufficient for boundedness of $\tilde D(N,M)$. This is based on Markov's inequality for polynomials. Necessity of this condition would follow directly from the lower bound in \R{coppriv}, provided \R{coppriv} were shown to hold. It may be possible to adapt the proof of \cite{copprivlinpolygrowth} to establish this result. Since the scaling $M = \ord{N^2}$ is undesirable, one can ask what happens when $M = \gamma N$ for some fixed oversampling parameter $\gamma \geq 1$. Using potential theory arguments, one can show that $\tilde D(N, \gamma N)$ grows exponentially in $N$ (with the constant of this growth becoming smaller as $\gamma$ increases), as predicted by the conjectured bound \R{coppriv}. In other words, \be{ \label{cPotentialDef} N^{-1} \log D(N , \gamma N ) \sim \log c(\gamma;T),\quad N \rightarrow \infty, } for some $c(\gamma ; T) >1$\footnote{The constant of growth was obtained in private communication with A.\ Kuijlaars. A closed expression (up to several integrals involving the potential function $\phi$ for the nodes $z_n$) can be found for $c(\gamma;T)$. We omit the full argument as it is rather lengthy, but note that it is based on standard results in potential theory. A general reference is \cite{ransford1995potentialtheory}.}. In view of this behaviour, Theorem \ref{t:oversamp} guarantees convergence of the FE \R{equiL2}, provided $\rho \geq c(\gamma ; T)$, where $\rho$ is as in Theorem \ref{t:expconv}. In other words, $f$ needs to be analytic in the region $\cD(c(\gamma;T))$ (recall $\cD$ from Theorem \ref{t:expconv}) to ensure convergence. Therefore, one expects a Runge phenomenon whenever $f$ has a complex singularity lying in the corresponding Runge region $\cR(\gamma ; T) = \cD(c(\gamma;T))$. Naturally, a larger value of $\gamma$ leads to a smaller (but still nontrivial) Runge region. However, regardless of the choice of $\gamma$, there will always be analytic functions for which one expects divergence of $F_{N,\gamma N}(f)$ (see \cite{boyd2009divergence} for a related discussion in the case of equispaced polynomial interpolation). Moreover, the mapping $f \mapsto F_{N,\gamma N}$ will always be exponentially ill-conditioned for any fixed $\gamma$, since the condition number is precisely $D(N,\gamma N)$ (Theorem \ref{t:oversamp}). Primarily for later use, we now note that it is also possible to study the condition number of the equispaced FE matrix $\bar{A}$ in a similar way. Straightforward arguments show that \be{ \label{minsingBNM} 1/\sigma_{\min}(\bar A)= B(N,M),\qquad B(N,M) = \sup \left \{ \| \phi \|_{[-T,T]} : \phi \in \cG_N, \| \phi \|_M=1 \right \}. } Using the fact that $1/\sigma_{\min}(A) = \sup \left \{ \| \phi \|_{[-T,T]} : \phi \in \cG_N, \| \phi \|=1 \right \}$, where $A$ is the matrix of the continuous FE, one can show that \bes{ 1/\sigma_{\min}(A) \lesssim B(N,M) \leq D(N,M) / \sigma_{\min}(A), } (here we use $\lesssim$ to mean up to possible algebraic factors in $N$). Theorem \ref{t:exactCondNumb} now shows that $\bar{A}$ is always exponentially ill-conditioned in $N$, regardless of $M \geq N$. Much like the case of $D(N,M)$ and $\tilde D(N,M)$, one can show that the quantity $B(N,M)$ is, up to algebraic factors, equivalent to \be{ \label{tilBNM} \tilde B(N,M) = \sup \left \{ \| p \|_{\infty,[m(T),1]} : p \in \bbP_N, | p(z_n) | \leq 1, n=0,\ldots,M \right \}. } Potential theory can be used once more to determine the exact behaviour of $\tilde B(N,\gamma N)$. In particular, \be{ \label{dPotentialDef} N^{-1} \log B(N,\gamma N) \sim d(\gamma;T),\quad N \rightarrow \infty, } for some constant $d(\gamma;T) \geq c(\gamma;T) > 1$. \subsubsection{Numerical examples}\label{sss:numexampEqui} In the previous section we established (up to the conjecture \R{coppriv}) 6., 7.\ and 8.\ of \S \ref{s:introduction}. The main conclusion is as follows. In order to obtain a convergent FE in exact arithmetic using equispaced data one either needs to oversample quadratically (and thereby reduce the convergence rate to only root-exponential), or scale the extension parameter $T$ suitably with $N$ or both. However, recall from \S \ref{s:stability} that a FE obtained from a finite precision computation may differ quite dramatically from the corresponding infinite-precision extension. Is it therefore possible that the unpleasant effects described in the previous section may not be witnessed in finite precision? The answer transpires to be yes, and consequently FEs can safely be used for equispaced data, even in situations where divergence is expected in exact arithmetic. To illustrate, consider the function $f(x) = \frac{1}{1+100 x^2}$. When $T=2$, this function has a singularity lying in the Runge region $\cR(1;2)$. The predicted divergence of its exact (i.e.\ infinite-precision) equispaced FE is shown in Figure \ref{f:EquiFnApp}. Note that double oversampling also gives divergence, whilst with quadruple oversampling the singularity of $f$ no longer lies in $\cR(\gamma ; T)$. We therefore witness geometric convergence, albeit at a very slow rate. This behaviour is typical. Given a function $f$ it is always possible to select the oversampling parameter $\gamma$ in such a way that $F_{N,\gamma N}(f)$ converges geometrically. However, such a $\gamma$ depends on $f$ in a nontrivial manner (i.e.\ the location of the nearest complex singularity of $f$) and therefore cannot in practice be determined from the given data. Note that this phenomenon---namely, the fact that careful tuning of a particular parameter in a function-dependent way allows geometric convergence to be restored---is also seen in other methods for approximating functions to high accuracy, such as the Gegenbauer reconstruction technique \cite{GottGibbsRev,GottGibbs1} (see Boyd \cite{BoydGegen} for a description of the phenomenon) and polynomial least squares \cite{boyd2009divergence}. Fortunately, and unlike for these other methods, the situation changes completely for Fourier extensions when we carry out computations in finite precision. This is shown in Figure \ref{f:EquiFnApp}. For all choices of $\gamma$ used, the finite precision FE, which we denote $G_{N,\gamma N}(f)$, converges geometrically fast, and there is no drift in the error once the best achievable accuracy is attained. Note that oversampling by a constant factor improves the approximation, but in all cases we still witness convergence. In particular, no careful selection of $\gamma$, such as that discussed above, appears to be necessary in finite precision. \begin{figure} \begin{center} $\begin{array}{ccc} \includegraphics[width=6.25cm]{Figure10Exact} & \hspace{1pc} & \includegraphics[width=6.25cm]{Figure10Numer} \end{array}$ \caption{\small The error $\| f - f_N \|_{\infty}$ against $N$ for the equispaced FEs $f_N = F_{N,\gamma N}(f)$ (left) and $f_{N} = G_{N,\gamma N}(f)$ (right) of $f(x) = \frac{1}{1+100 x^2}$ with oversampling factor $\gamma = 1,2,4$ (squares, circles and crosses) and $T=2$.} \label{f:EquiFnApp} \end{center} \end{figure} \subsection{The numerical equispaced Fourier extension}\label{ss:equianalysis} We now explain these results by analysing the numerical equispaced FE. Proceeding as in \S \ref{ss:numsolnanalysis} we shall consider the truncated SVD approximation, which we denote $H_{N,M,\epsilon}(f)$. Note that a similar analysis has also recently been presented in \cite{LyonFESVD}; see Remark \ref{r:Lyon} for further details. Let $\Phi_n \in \cG_N$ be the function corresponding to the right singular vector $v_n$ of the matrix $\bar{A}$. Write $\cG_{N,M,\epsilon} = \spn \left \{ \Phi_n : \sigma_n > \epsilon \right \}$ and $\cG^{\perp}_{N,M,\epsilon} = \spn \{ \Phi_n : \sigma_n\leq\epsilon \}$, and note that $H_{N,M,\epsilon}$ is the orthogonal projection onto $\cG_{N,M,\epsilon}$ with respect to $(\cdot,\cdot)_{M}$. Since $(\Phi_n,\Phi_m)_M = \sigma^2_n \delta_{n,m}$, we have \be{ \label{equiTSVD} H_{N,M,\epsilon}(f) = \sum_{n: \sigma_n > \epsilon} \frac{1}{\sigma^2_n} (f,\Phi_n)_M \Phi_n. } Our main result is as follows: \thm{ \label{t:equiTSVD} Let $f \in \rL^\infty(-1,1)$ and $H_{N,M,\epsilon}(f)$ be given by \R{equiTSVD}. Then \be{ \label{equiTSVDerr} \| f - H_{N,M,\epsilon}(f) \| \leq \sqrt{2} \left ( 1+ C_1(N,M;T,\epsilon) \right ) \| f - \phi \|_{\infty} + C_2(N,M;T,\epsilon) \| \phi \|_{[-T,T]},\quad \forall \phi \in \cG_N, } and \be{ \label{equiTSVDcoeff} \nm{a_{\epsilon}} = \| H_{N,M,\epsilon}(f) \|_{[-T,T]} \leq \frac{\sqrt{2}}{\epsilon} \| f - \phi \|_{\infty} + \| \phi \|_{[-T,T]},\quad \forall \phi \in \cG_N, } where \ea{ \label{C1} C_{1}(N,M;T,\epsilon) = \sup_{\substack{\phi \in \cG_{N,M,\epsilon} \\ \phi \neq 0}} \left \{ \frac{\| \phi \|}{\| \phi \|_M} \right \}, \quad C_{2}(N,M;T,\epsilon) = \sup_{\substack{\phi \in \cG^{\perp}_{N,M,\epsilon} \\ \phi \neq 0}} \left \{ \frac{\| \phi \|}{\| \phi \|_{[-T,T]}} \right \}. } } \prf{ Let $\phi \in \cG_{N}$. Then \be{ \label{eqstep1} \| f - H_{N,M,\epsilon}(f) \| \leq \| f - \phi \| + \| H_{N,M,\epsilon}(f-\phi) \| + \| \phi - H_{N,M,\epsilon}(\phi) \|. } Consider the second term. By definition of $C_1(N,M;T,\epsilon)$, \bes{ \| H_{N,M,\epsilon}(f - \phi) \| \leq C_{1}(N,M,\epsilon) \| H_{N,M,\epsilon }(f- \phi) \|_M \leq C_{1}(N,M,\epsilon) \| f - \phi \|_M, } where the second inequality follows from the fact that $H_{N,M,\epsilon}$ is an orthogonal projection with respect to $(\cdot,\cdot)_M$. Noting that $\| g \|,\| g \|_M \leq \sqrt{2} \| g \|_{\infty}$ for any function $g \in \rL^\infty(-1,1)$ now gives the corresponding term in \R{equiTSVDerr}. The bound for the third term of \R{eqstep1} follows immediately from the definition of $C_2(N,M;T,\epsilon)$ and the inequality $\| \phi - H_{N,M,\epsilon}(\phi) \|_{[-T,T]} \leq \| \phi \|_{[-T,T]}$. For \R{equiTSVDcoeff}, we first write $\| H_{N,M,\epsilon}(f) \|_{[-T,T]} \leq \| H_{N,M,\epsilon}(f - \phi) \|_{[-T,T]} + \| H_{N,M,\epsilon}(\phi) \|_{[-T,T]}$. Observe that for any $g \in \rL^\infty(-1,1)$ we have \bes{ \| H_{N,M,\epsilon}(g) \|^2_{[-T,T]} = \sum_{n:\sigma_n > \epsilon} \frac{1}{\sigma^4_n} | (g,\Phi_n)_M |^2 \leq \frac{1}{\epsilon^2} \| H_{N,M,\epsilon}(g) \|^2_M \leq \frac{1}{\epsilon^2} \| g \|^2_M \leq \frac{2}{\epsilon^2} \| g \|^2_{\infty}. } Also, $\| H_{N,M,\epsilon}(\phi) \|_{[-T,T]} \leq \| \phi \|_{[-T,T]}$ for $\phi \in \cG_N$. Setting $g=f-\phi$ and combining these two bounds now gives \R{equiTSVDcoeff}. } \cor{ If $f \in \rL^\infty(-1,1)$ then $\| H_{N,M,\epsilon}(f) \| \leq \sqrt{2} / \epsilon \| f \|_{\infty}$, $\forall N \in \bbN$, $M \geq N$. Moreover, if $f \in \rH^1(-1,1)$, $\bbT=[-T,T)$ is the $T$-torus and $c_1(T) > 0$ is as in Lemma \ref{l:smoothextension}, then \bes{ \limsup_{\substack{N,M \rightarrow \infty \\ M \geq N}} \| H_{N,M,\epsilon}(f) \| \leq \inf \left \{ \| \tilde f \|_{[-T,T]} : \tilde{f} \in \rH^1(\bbT),\ \tilde{f} |_{[-1,1]} = f \right \} \leq c_1(T) \| f \|_{\rH^1(-1,1)}. } } \prf{ By \R{equiTSVDcoeff}, we have \be{ \label{normbd} \| H_{N,M,\epsilon}(f) \| \leq \| H_{N,M,\epsilon}(f) \|_{[-T,T]} \leq \frac{\sqrt{2}}{\epsilon} \| f - \phi \|_{\infty} + \| \phi \|_{[-T,T]},\quad \forall \phi \in \cG_N. } Setting $\phi = 0$ gives the first result. For the second, we let $\phi$ be the $N$-term Fourier series of $\tilde{f}$ on $\bbT$, so that $\| f - \phi \|_{\infty} \rightarrow 0$ as $N \rightarrow \infty$. The final inequality follows from Lemma \ref{l:smoothextension}. } This corollary shows that the equispaced FE cannot suffer from a Runge phenomenon in finite precision, since it is bounded in $N$ and $M$. This should come as no surprise. Divergence of $H_{N,M,\epsilon}(f)$ would imply unboundedness of the coefficients $a_{\epsilon}$, a behaviour which is prohibited by truncating the singular values of $\bar{A}$ at level $\epsilon$. Note that this corollary actually shows a much stronger result, namely that $H_{N,M,\epsilon}(f)$ is bounded on the extended domain $[-T,T]$, not just on $[-1,1]$. Although this corollary demonstrates lack of divergence of $H_{N,M,\epsilon}(f)$, it says littles about its convergence besides the observation that $\|H_{N,M,\epsilon}(f)\|$ is asymptotically bounded by $\| f \|_{\rH^1(-1,1)}$. To study convergence we shall use \R{equiTSVDerr}. For this we first need to understand the constants $C_{i}(N,M;T,\epsilon)$. \subsubsection{Behaviour of $C_{i}(N,M;T,\epsilon)$}\label{Cibehaviour} \begin{figure} \begin{center} $\begin{array}{ccc} \includegraphics[width=4.75cm]{Fig11_l1_gm1_alt} & \includegraphics[width=4.75cm]{Fig11_l2_gm1_alt} & \includegraphics[width=4.7cm]{Fig11_l3_gm1_alt} \\ \includegraphics[width=4.75cm]{Fig11_l1_gm2_alt} & \includegraphics[width=4.75cm]{Fig11_l2_gm2_alt} & \includegraphics[width=4.7cm]{Fig11_l3_gm2_alt} \\ T=\frac43 & T=2 & T=4 \end{array}$ \caption{\small The quantity $C_1(N,\gamma N;T,\epsilon)$ against $N$ for $\gamma=1$ (top row) or $\gamma = 2$ (bottom row) and $\epsilon = 10^{-6},10^{-12},10^{-18},10^{-24},10^{-30}$ (squares, circles, crosses, diamonds and dashes respectively).} \label{f:C1} \end{center} \end{figure} Although Theorem \ref{t:equiTSVD} holds for arbitrary $M \geq N$, we now focus on the case of linear oversampling, i.e.\ $M = \gamma N$ for some $\gamma \geq 1$. Let $N_2(\gamma,T,\epsilon)$ be the largest $N$ such that all the singular values of $\bar{A}$ are at least $\epsilon$ in magnitude: \bes{ N_2(\gamma,T,\epsilon) = \max \left \{ N : \sigma_{\min}(\bar{A}) > \epsilon \right \}. } For $N \leq N_2(\gamma,T,\epsilon)$ we have $\cG_{N, \gamma N,\epsilon} = \cG_N$ and therefore $C_1(N,\gamma N;T,\epsilon) = D(N,\gamma N)$, where $D(N,M)$ is given by \R{DNM}. Thus we witness exponential divergence of $C_1(N,\gamma N;T,\epsilon)$ at rate $c(\gamma;T)$, where $c(\gamma;T)$ is as in \R{cPotentialDef}. This is shown numerically in Figure \ref{f:C1}. However, once $N > N_2(\gamma,T,\epsilon)$ the numerical results in Figure \ref{f:C1} indicate a completely different behaviour: namely, $C_1(N,\gamma N;T,\epsilon)$ appears to be bounded. Although we have no proof of this fact, these results strongly suggest the following conjecture: \be{ \label{conjecture1} C_1(N,\gamma N;T,\epsilon) \lesssim C_1(N_2,\gamma N_2;T,\epsilon) \sim c(\gamma;T)^{N_2},\quad \forall N \in \bbN. } In other words, $C_1(N,\gamma N;T,\epsilon)$ achieves its maximal value in $N$ at $N \approx N_2$. Recalling \R{minsingBNM} and \R{dPotentialDef}, we note that \be{ \label{N2} N_2(\gamma ,T,\epsilon) \approx - \frac{\log \epsilon}{\log d(\gamma ; T)}. } Thus, substituting this into bound \R{conjecture1} results in the following conjecture: \be{ \label{C1bd} C_1(N,\gamma N;T,\epsilon) \lesssim \min \left \{ c(\gamma;T)^N, \epsilon^{-\frac{\log c(\gamma;T)}{\log d(\gamma;T)}} \right \},\quad \forall N \in \bbN. } In particular, $C_1(N,\gamma N;T,\epsilon)$ is bounded for all $N$ by some power of $\epsilon^{-1}$. Importantly, this power cannot be too large. Note that $c(\gamma;T) \leq d(\gamma;T)$, $\forall T > 1$, since the maximum of a polynomial on $[m(T),1]$ is at least as large as its maximum on the smaller interval $[-1,1]$---compare \R{tilBNM} to \R{tilDNM}. Therefore the ratio $\frac{\log c(\gamma;T)}{\log d(\gamma;T)}$ is at most one. Moreover, by varying either $\gamma$ or $T$ we may decrease this ratio to arbitrarily close to $1$. We discuss this further in the next section. The quantity $C_2(N,M;T,\epsilon)$ is harder to analyze, although clearly we have $C_2(N,M;T,\epsilon) = 0$ when $N < N_2$. Figure \ref{f:C2} demonstrates that $C_2(N,\gamma N,\epsilon)$ is also bounded in $N$. Moreover, closer comparison with Figure \ref{f:C1} suggests the existence of a bound of the form \be{ \label{conjecture2} C_2(N,\gamma N;T,\epsilon) \lesssim \epsilon C_1(N,\gamma N;T,\epsilon). } Once more, we have no proof of this observation. \begin{figure} \begin{center} $\begin{array}{ccc} \includegraphics[width=4.75cm]{Fig12_l1_gm1_alt} & \includegraphics[width=4.75cm]{Fig12_l2_gm1_alt} & \includegraphics[width=4.7cm]{Fig12_l3_gm1_alt} \\ \includegraphics[width=4.75cm]{Fig12_l1_gm2_alt} & \includegraphics[width=4.75cm]{Fig12_l2_gm2_alt} & \includegraphics[width=4.7cm]{Fig12_l3_gm2_alt} \\ T=\frac43 & T=2 & T=4 \end{array}$ \caption{\small The quantity $C_2(N,\gamma N;T,\epsilon)$ against $N$ for $\gamma=1$ (top row) or $\gamma = 2$ (bottom row) and $\epsilon = 10^{-6},10^{-12},10^{-18},10^{-24},10^{-30}$ (squares, circles, crosses, diamonds and dashes respectively).} \label{f:C2} \end{center} \end{figure} \rem{ The quantities $C_1(N,M;T,\epsilon)$ and $C_2(N,M;T,\epsilon)$ have the explicit expressions \bes{ C_1(N,M;T,\epsilon) = \sqrt{\| (S^{\epsilon})^{\dag} V^* A V (S^{\epsilon})^{\dag} \|},\quad C_2(N,M;T,\epsilon) = \sqrt{\| (V^{\epsilon})^* A V^{\epsilon} \|}, } where $A$ is the continuous FE matrix, $U S V^*$ is the singular value decomposition of the equispaced FE matrix $\bar{A}$, $S^{\epsilon}$ is formed by replacing the $n^{\rth}$ column of $S$ by the zero vector whenever $\sigma_n \leq \epsilon$, and $V^{\epsilon}$ is formed by doing the same for columns of $V$ corresponding to indices $n$ with $\sigma_n > \epsilon$. These expressions were used to obtain the numerical results in Figures \ref{f:C1} and \ref{f:C2}. Computations were carried out with additional precision to avoid effects due to round-off. } \subsubsection{Behaviour of the truncated SVD Fourier extension} Combining the analysis of the previous section with Theorem \ref{t:equiTSVD}, we now conjecture the bound \be{ \label{exactnumerr} \| f -H_{N,\gamma N,\epsilon}(f) \| \leq C(\gamma,T,\epsilon) \left ( \| f - \phi \|_{\infty} + \epsilon \| \phi \|_{[-T,T]} \right ),\quad \forall \phi \in \cG_N, } where $C(\gamma,T,\epsilon)$ is proportional to $\epsilon^{-a(\gamma;T)}$ and $a(\gamma;T)$ is given by \be{ \label{agammaT} a(\gamma;T) = \frac{\log c(\gamma;T)}{\log d(\gamma;T)}. } This estimate allows us to understand the behaviour of the numerical equispaced FE $G_{N,\gamma N}(f)$. When $N < N_2$ we have $G_{N,\gamma N}(f) = F_{N,\gamma N}(f)$ and therefore $G_{N,\gamma N}(f)$ will diverge geometrically fast in $N$ whenever $f$ has a singularity in the Runge region $\cR(\gamma;T)$ (see \S \ref{sss:oversamp}). However, once $N$ exceeds $N_2$, one obtains convergence. Indeed, setting $\phi = F_N(f)$ in \R{exactnumerr}, we find that the convergence is geometric up to the breakpoint $N_1$ (see \R{N1def}), and then, much as before, at least superalgebraic beyond that point. Note that the maximal achievable accuracy of order $C(\gamma,T,\epsilon) \epsilon \approx \epsilon^{ 1-a(\gamma;T) }$. In summary, we have now identified the following convergence behaviour for $H_{N,\gamma N,\epsilon}(f)$: \begin{enumerate} \item[(i)] $N < N_2(\gamma,T,\epsilon) \approx - \frac{\log \epsilon}{\log d(\gamma;T)}$. Geometric divergence/convergence of $H_{N,\gamma N,\epsilon}(f)$ at a rate of, at worst, $c(\gamma;T) / \rho$, where $\rho$ is as in Theorem \ref{t:expconv} and $c(\gamma;T)$ is given by \R{cPotentialDef}. \item[(ii)] $N_2(\gamma,T,\epsilon) \leq N < N_1(T,\epsilon) \approx - \frac{\log \epsilon}{\log E(T)}$. Geometric convergence at a rate of at least $\rho$. \item[(iii)] $N= N_1(T,\epsilon)$. The error \bes{ \| f - H_{N,\gamma N,\epsilon}(f) \| \approx c_f \epsilon^{d_f-a(\gamma;T)}, } where $a(\gamma;T)$ is as in \R{agammaT} and $d_f = \frac{\log \rho}{\log E(T)} \in (0,1]$. \item[(iv)] $N \geq N_1(\gamma,T)$. Superalgebraic convergence of $H_{N,\gamma N,\epsilon}(f)$ down to a maximal achievable accuracy proportional to $\epsilon^{1-a(\gamma;T)}$. \end{enumerate} (This establishes 10.\ of \S \ref{s:introduction}). Much as in the case of the discrete FE, we see that if $f$ is analytic in $\cD(E(T))$, and if $c_f$ is not too large, then convergence stops at $N = N_1$ with maximal accuracy of order $c_f \epsilon^{1-a(\gamma;T)}$. Otherwise, we have a further regime of at least superalgebraic convergence before this accuracy is reached. An important question is the role of the oversampling parameter $\gamma$ in this convergence. We note: \lem{ \label{l:a_behaviour} Let $a(\gamma;T)$ be given by \R{agammaT}. Then $a(\gamma;T)$ satisfies $0 \leq a(\gamma;T) \leq 1$ for all $\gamma$ and $T$. Moreover, $a(\gamma;T) \rightarrow 0$ as $\gamma \rightarrow \infty$ for fixed $T$, and $a(\gamma;T) \rightarrow 0$ as $T \rightarrow \infty$ for fixed $\gamma$. } \prf{ Note that $c(\gamma;T) \leq d(\gamma;T)$. Also $c(\gamma;T) \rightarrow 1$ and $d(\gamma;T) \rightarrow E(T)$ as $\gamma \rightarrow \infty$ for fixed $T$, and $d(\gamma;T) \rightarrow \infty$ as $T \rightarrow \infty$ for fixed $\gamma$, whereas $c(\gamma;T)$ is bounded. } This lemma suggests that increasing $\gamma$ will lead to a smaller constant $C(\gamma,T,\epsilon)$ in \R{exactnumerr}. In fact, numerical results (Figures \ref{f:C1} and \ref{f:C2}) indicate that using $T=2$ and $\gamma =2$ gives a bound of a little over $1$ in magnitude for $\epsilon = 10^{-14}$. Note that the effect of even just double oversampling is quite dramatic. Without oversampling (i.e.\ $\gamma = 1$), the constant $C(\gamma,T,\epsilon)$ is approximately $10^4$ in magnitude when $\epsilon = 10^{-14}$ (see Figures \ref{f:C1} and \ref{f:C2}). Let us make several further remarks. First, in practice the regime $N < N_1$ is typically very small---recall that $N_1$ is around $20$ for $T=2$ (see \S \ref{sss:discSVDanalysis})---and therefore one usually does not witness all three types of behaviour in numerical examples. Second, as $\gamma \rightarrow \infty$, we have $N_2 \rightarrow N_1$ (recall that $d(\gamma;T) \rightarrow E(T)$ as $\gamma \rightarrow \infty$). Thus, with a sufficient amount of oversampling, the regime (ii) will be arbitrarily small. On the other hand, oversampling decreases $c(\gamma;T)$, and therefore the rate of divergence in the regime (i) is also lessened by taking $\gamma > 1$. Indeed, the numerical results in Figure \ref{f:ManSVDequi}, as well as in \S \ref{ss:NumExp} later, indicate that oversampling by a factor of $2$ is typically sufficient in practice to mitigate the effects of divergence for most reasonable functions. \begin{figure} \begin{center} $\begin{array}{ccc} \includegraphics[width=6.25cm]{Figure13_gamma1_alt} & \hspace{1pc} & \includegraphics[width=6.25cm]{Figure13_gamma2_alt} \end{array}$ \caption{\small Error against $N$ for $H_{N,\gamma N , \epsilon}(f)$, where $f(x) = \frac{1}{1+16 x^2}$, $T=2$, $\gamma = 1$ (left) or $\gamma = 2$ (right) and $\epsilon = 10^{-6} , 10^{-12}, 10^{-18}$ (squares, circles, crosses). Diamonds correspond to the exact equispaced FE $F_{N,\gamma N}(f)$.} \label{f:ManSVDequi} \end{center} \end{figure} Figure \ref{f:ManSVDequi} confirms these observations for the function $f(x) = \frac{1}{1+16 x^2}$. For $\gamma = 1$ the initial exponential divergence is quite noticeable. However, this effect largely vanishes when $\gamma =2$. Notice that a larger cutoff $\epsilon$ actually gives a smaller error initially, since there is a smaller regime of divergence. However, the maximal achievable accuracy is correspondingly lessened. We note also that maximal achievable accuracies for $\epsilon = 10^{-6},10^{-12},10^{-18}$ are roughly $10^{-4}$, $10^{-8}$ and $10^{-12}$ respectively when $\gamma = 1$ and $10^{-7}$, $10^{-12}$ and $10^{-16}$ when $\gamma = 2$. These are in close agreement with the corresponding numerical values of $C_2(N,\gamma N; T , \epsilon)$ (see Figure \ref{f:C2}), as predicted by Theorem \ref{t:equiTSVD}. \rem{\label{r:scaling} A central conclusion of this section is that one requires a lower asymptotic scaling of $M$ with $N$ for the numerical equispaced FE than for its exact counterpart. Since $\cG_{N,M,\epsilon}$ is a subset of $\cG_N$, we clearly have $C_{1}(N,M;T,\epsilon) \leq D(N,M)$, where $D(N,M)$ is given by \R{DNM}. Hence quadratic scaling $M = \ord{N^2}$ is sufficient (see \S \ref{sss:oversamp}) to ensure boundedness of $C_{1}(N,M;T,\epsilon)$, and one can make a similar argument for $C_{2}(N,M;T,\epsilon)$. However, Figures \ref{f:C1} and \ref{f:C2} indicate that this condition is not necessary, and that one can get away with the much reduced scaling $M=\ord{N}$ in practice. This difference can be understood in terms of the singular values of $\bar{A}$. Recall that small singular values correspond to functions $\phi \in \cG_N$ with $\| \phi \|_{[-T,T]} \gg \| \phi \|_M$. Now consider an arbitrary $\phi \in \cG_N$. If the ratio $\| \phi \| / \| \phi \|_M$ is large, it suggests that $\phi$ lies approximately in the space $\cG^{\perp}_{N,M,\epsilon}$ corresponding to small singular values. Hence, $\| \phi \| / \| \phi \|_M$ cannot be too large over $\phi \in \cG_{N,M,\epsilon}$, and thus we see boundedness of $C_{1}(N,M,\epsilon)$, even when $D(N,M)$---the supremum of this ratio over the whole of $\cG_N$---is unbounded. } \rem{ \label{r:Lyon} A similar analysis of the equispaced FE, also based on truncated SVDs, was recently presented by M. Lyon in \cite{LyonFESVD}. In particular, our expressions \R{equiTSVDerr} and \R{exactnumerr} are similar to equations (30) and (31) of \cite{LyonFESVD}. Lyon also provides extensive numerical results for his analogues of the quantities $C_1(N,M;T,\epsilon)$ and $C_2(N,M;T,\epsilon)$, and describes a bound which is somewhat easier to use in computations. The main contributions of our analysis are the conjectured scaling of the constant $C(\gamma,T,\epsilon)$ in terms of $\epsilon$, $\gamma$ and $T$, the description and analysis of the breakpoints $N_2$ and $N_1$, and the differing convergence/divergence in the corresponding regions. } \subsubsection{The condition number of the numerical equispaced FE}\label{ss:equistabfn} \begin{table} \begin{center} \begin{tabular}{|c||c|c|c|c|c|} \hline N & 40 & 80 & 120 & 160 & 200 \\ \hline $\gamma = 1$ & $2.37 \times 10^4$ & $3.50 \times 10^4$ & $2.24 \times 10^4$ & $2.47 \times 10^4$ & $1.93 \times 10^4$ \\ \hline $\gamma = 2$ & $2.18 \times 10^1$ & $2.66 \times 10^1$ & $2.40 \times 10^1$ & $2.56 \times 10^1$ & $2.47 \times 10^1$ \\ \hline $\gamma = 4$ & $8.03 \times 10^0$ & $1.05 \times 10^1$ & $1.23 \times 10^1$ & $1.39 \times 10^1$ & $1.54 \times 10^1$ \\ \hline \end{tabular} \caption{ The function $K(G_{N,\gamma N})$ against $N$ with $T=2$ and $\gamma = 1, 2, 4$.}\label{t:FEStabFnEquispaced} \end{center} \end{table} We now consider the condition number $\kappa(G_{N,M})$ (defined as in \R{FNM_cond}) of the numerical equispaced extension. In Table \ref{t:FEStabFnEquispaced} we plot $K(G_{N,\gamma N})$ against $N$, where $K(G_{N,M})$ is an upper bound for $\kappa(G_{N,M})$ defined analogously to \R{KGN_def}. The results indicate numerical stability, and, as we expect, improved stability with more oversampling. Besides oversampling it is also possible to improve stability by varying the extension parameter $T$. In Figure \ref{f:contours} we give a contour plot of $K(G_{N,\gamma N})$ in the $(\gamma,T)$-plane. Evidently, increasing $T$ improves stability. Recall, however, that a larger $T$ corresponds to worse resolution power (see \S \ref{ss:Tchoice}). Conversely, increasing $\gamma$ also leads to worse resolution when measured in terms of the total number $M = \gamma N$ of equispaced function values required. Hence a balance must be struck between the two quantities. Figure \ref{f:contours} suggests that $\gamma = T = 2$ is a reasonable choice in practice. Recall also that the choice $T=2$ allows for fast computation of the equispaced FE (Remark \ref{sss:T2}), and hence is desirable to use in computations. \begin{figure} \begin{center} $\begin{array}{c} \includegraphics[width=6.25cm]{ContourPlotNew2} \end{array}$ \caption{\small Contour plot of the quantity $K(G_{N,\gamma N})$ against $1\leq \gamma \leq 4$ and $1 < T \leq 4$ for $N=200$.} \label{f:contours} \end{center} \end{figure} The behaviour of the condition number can be investigated with the following theorem (the proof is similar to that of Theorem \ref{t:stabfn} and hence omitted): \thm{ \label{t:equistabfn} The condition number $\kappa(H_{N,M,\epsilon})$ of the truncated SVD equispaced FE $H_{N,M,\epsilon}$ satisfies $\kappa(H_{N,M,\epsilon}) =C_1(N,M;T,\epsilon)$, where $C_{1}(N,M;T,\epsilon)$ is given by \R{C1}. } From the analysis of \S \ref{Cibehaviour} we conclude that $\kappa(H_{N,\gamma N,\epsilon}) \lesssim \epsilon^{-a(\gamma;T)}$, where $a(\gamma;T)$ is as in \R{agammaT}. Lemma \ref{l:a_behaviour} therefore shows that $\kappa(H_{N,\gamma N,\epsilon}) \lesssim 1$ as $\gamma \rightarrow \infty$ for fixed $T$, and $\kappa(H_{N,\gamma N,\epsilon}) \lesssim 1$ as $T \rightarrow \infty$ for fixed $\gamma$. This confirms the behaviour described above. \subsubsection{Numerical examples}\label{ss:NumExp} In Figure \ref{f:EquiFnApp2} we consider the equispaced FE for four test functions. In all cases we use $\gamma =2$ and $T=2$. As is evident, all choices of $T$ give good, stable numerical results, with the best achievable accuracy being at least $10^{-12}$. Robustness in the presence of noise is shown in Figure \ref{f:NoisyEqui}. Observe that when $\gamma = 1$, noise of amplitude $\delta$ is magnified by around $10^{5}$, in a manner consistent with Theorem \ref{t:equistabfn}. Conversely, with double oversampling, this factor drops to less than $10^{2}$, again in agreement with Theorem \ref{t:equistabfn}. \begin{figure} \begin{center} $\begin{array}{ccc} \includegraphics[width=6.25cm]{Fig16Fns14} & \hspace{1pc} & \includegraphics[width=6.25cm]{Fig16Fns23} \end{array}$ \caption{\small The error $\| f - G_{N,\gamma N}(f) \|_{\infty}$, where $\gamma = 2$ and $T=2$. Left: $f(x) = \E^{25 \sqrt{5} \pi \I x}$ (squares), $f(x) = | x |^7$ (circles). Right: $f(x) = \frac{1}{1+25 x^2}$ (squares), $f(x) = \frac{1}{8-7 x}$ (circles). } \label{f:EquiFnApp2} \end{center} \end{figure} \begin{figure} \begin{center} $\begin{array}{ccc} \includegraphics[width=6.25cm]{NoisyFnAppEquig1_New} & \hspace{1pc} & \includegraphics[width=6.25cm]{NoisyFnAppEquig2_New} \end{array}$ \caption{\small The error $| f(x) - G_{N,\gamma N}(f)(x) |$ against $x$, where $\gamma =1$ (left) or $\gamma = 2$ (right), for $N = 30$, $T=2$ and $f(x) = \E^x$, with noise at amplitudes $\delta = 10^{-4},10^{-6},10^{-8},10^{-10},0$.} \label{f:NoisyEqui} \end{center} \end{figure} \subsection{Relation to the theorem of Platte, Trefethen \& Kuijlaars}\label{ss:PTKrelation} We are now in a position to explain how FEs relate to the impossibility theorem of Platte, Trefethen \& Kuijlaars \cite{TrefPlatteIllCond}. A restatement of this theorem (with minor modifications to notation) is as follows: \thm{[\cite{TrefPlatteIllCond}] \label{t:impossibility} Let $F_M$, $M \in \bbN$, be a sequence of approximations such that $F_M(f)$ depends only on the values of $f$ on an equispaced grid of $M$ points. Let $\cE \subseteq \bbC$ be compact and suppose that there exists $C < \infty$, $\alpha>1$ and $\tau \in (\frac12,1]$ such that \be{ \label{FMerror} \| f - F_M(f) \|_{\infty} \leq C c_f \alpha^{-M^{\tau}},\qquad c_f = \max_{x \in \cE} | f(x) |, } for all $M \in \bbN$ and all $f$ that are continuous on $\cE$ and analytic in its interior. Then there exists a $\beta > 1$ such that the condition numbers $\kappa(F_M) \geq \beta^{M^{2 \tau -1}}$ for all sufficiently large $M$. } This theorem has two important consequences. First, any exponentially convergent method is also exponentially ill-conditioned. Second, the best possible convergence for a stable method is root-exponential in $M$. Note that the theorem is valid for all methods, both linear and nonlinear, that satisfy \R{FMerror}. Consider now the exact equispaced Fourier extension $F_{N,M}$. As shown in \S \ref{ss:equiFEthy}, when $N = \ordu{\sqrt{M}}$ this method is stable and root-exponentially convergent. Hence equispaced FEs in infinite precision \textit{attain the maximal possible convergence rate for stable methods} satisfying the conditions of the theorem. Now consider the numerical equispaced FE $G_{\eta M,M}$, where $0 < \eta \leq 1$ is the reciprocal of the oversampling parameter $\gamma$ used in the previous sections. We have shown that this approximation is stable, so at least one condition in Theorem \ref{t:impossibility} must be violated. Suppose that we take $\cE = \cD(E(T))$, for example. Then \R{exactnumerr} shows that \be{ \label{conv_nice} \| f - G_{\eta M,M}(f) \|_{\infty} \lesssim c_f \epsilon^{-a(\eta^{-1} ; T)}\left ( \left ( E(T)^{\eta} \right )^{-M} + \epsilon \right ). } The finite term $\epsilon$ in the brackets means that this approximation does not satisfy \R{FMerror}, and hence Theorem \ref{t:impossibility} does not apply. Recall that if $c_f$ is small then \R{conv_nice} describes the full convergence behaviour for all $M$. On the other hand, if $c_f$ is large, or if $f \in \cD(\rho)$ with $\rho < E(T)$, then the convergence is, after initial geometric convergence, at least superalgebraic down to the maximal achievable accuracy $\epsilon^{1-a(\eta^{-1};T)}$. This is also not in contradiction with the conditions of Theorem \ref{t:impossibility}. To summarize, equispaced FEs, when implemented in finite precision, possess both numerical stability and rapid convergence, and hence allow one to circumvent the impossibility theorem to an extent. In particular, for all functions $f \in \cD(E(T))$ possessing small constants $c_f$, the approximations converge geometrically fast down to a maximal accuracy of order $\epsilon^{1-a(\eta^{-1};T)}$. In all other cases, the convergence is at least superalgebraic down to the same accuracy. \section{Conclusions and challenges}\label{s:conclusions} We conclude by making the following remark. Extensive numerical experiments \cite{BoydFourCont,BoydRunge,brunoFEP,LyonFESVD,LyonFast} have shown the effectiveness of FEs in approximating even badly behaved functions to high accuracy in a stable fashion. The purpose of this paper has been to provide analysis to explain these results. In particular, we have shown numerical stability for all three types of extensions considered, and analyzed their convergence. The reason for this robustness, despite the presence of exponentially ill-conditioned matrices, is due to the fact that the FE is a frame approximation and that for all functions $f$, even those with oscillations or large derivatives, there eventually exist coefficient vectors with small norms which approximate $f$ to high accuracy. The main outstanding theoretical challenge is to understand the constants $C_i(N,M;T,\epsilon)$ of the equispaced FE. In particular, we wish to show that linear scaling $M = \gamma N$ is sufficient to ensure boundedness of these constants in $N$, with a larger $\gamma$ corresponding to a smaller bound. Note that the analysis of \S \ref{sss:oversamp} implies the suboptimal result that $M = \ord{N^2}$ is sufficient (Remark \ref{r:scaling}). It is also a relatively straightforward exercise to show that if $M = c N / \epsilon$ for suitable $c > 0$, then $C_i(N,M;T,\epsilon)$ is bounded. This is based on making rigorous the arguments given in Remark \ref{r:scaling}---we do not report it here for brevity's sake. Unfortunately, although this estimate gives the correct scaling $M = \ord{N}$, it is wildly pessimistic. It implies that $M$ should scale like $\approx 10^{16} N$, whereas the numerics in \S \ref{Cibehaviour} indicate that $M= \gamma N$ is sufficient for \textit{any} $\gamma \geq 1$. One approach towards establishing a more satisfactory result is to perform a closer analysis of the singular values of the matrix $\bar{A}$. Some preliminary insight into this problem was given in \cite{EdelmanFuture}. Therein it was proved that (whenever $M=N$ and $2T \in \bbN$) the singular values cluster near zero and one, and the transition region is $\ord{\log N}$ in width, much like for the prolate matrix $A$. Unfortunately, little is known outside of this result. There is no existing analysis for $\bar A$ akin to that of Slepian's for the prolate matrix---see \cite{EdelmanFuture} for a discussion. Note, however, that the normal form $B = \bar{A}^*\bar{A}$ has entries $B_{n,m} = \frac{\sin \frac{(n-m) \pi}{T}}{M T \sin \frac{(n-m) \pi}{M T}}$, and can therefore be viewed as a discretized version of the prolate matrix $A$. Indeed, $B \rightarrow A$ as $M \rightarrow \infty$ for fixed $N$. Given the similarities between the two matrices, there is potential for Slepian's analysis to be extended to this case. However, this remains an open problem. Another issue is that of understanding how to choose the parameters $T$ and $\gamma$ in the case of the equispaced extension. As discussed in \S \ref{ss:Tchoice}, the choice of $T$ is reasonably clear for the continuous and discrete FEs (where there is no $\gamma$). If resolution of oscillatory functions is a concern, one should choose a small value of $T$ (in particular, \R{Tkte}). Otherwise, a good choice appears to be $T=2$. However, for the equispaced FE, small $T$ adversely affects stability (see \S \ref{ss:equistabfn}). Hence it must be balanced by taking a larger value of the oversampling parameter $\gamma$, which has the effect of reducing the effective resolution power. In practice, however, a reasonable choice appears to be $T = \gamma = 2$. Investigating whether or not this is optimal is a topic for further investigation. \section*{Acknowledgements} The authors would like to thank John Boyd, Doug Cochran, Laurent Demanet, Anne Gelb, Anders Hansen, Arieh Iserles, Arno Kuijlaars, Mark Lyon, Nilima Nigam, Sheehan Olver, Rodrigo Platte, Jie Shen and Nick Trefethen for useful discussions and comments. They would also like to thank the anonymous referees for their constructive and helpful remarks. \bibliographystyle{abbrv} \small
{ "timestamp": "2013-05-14T02:02:14", "yymm": "1206", "arxiv_id": "1206.4111", "language": "en", "url": "https://arxiv.org/abs/1206.4111", "abstract": "An effective means to approximate an analytic, nonperiodic function on a bounded interval is by using a Fourier series on a larger domain. When constructed appropriately, this so-called Fourier extension is known to converge geometrically fast in the truncation parameter. Unfortunately, computing a Fourier extension requires solving an ill-conditioned linear system, and hence one might expect such rapid convergence to be destroyed when carrying out computations in finite precision. The purpose of this paper is to show that this is not the case. Specifically, we show that Fourier extensions are actually numerically stable when implemented in finite arithmetic, and achieve a convergence rate that is at least superalgebraic. Thus, in this instance, ill-conditioning of the linear system does not prohibit a good approximation.In the second part of this paper we consider the issue of computing Fourier extensions from equispaced data. A result of Platte, Trefethen & Kuijlaars states that no method for this problem can be both numerically stable and exponentially convergent. We explain how Fourier extensions relate to this theoretical barrier, and demonstrate that they are particularly well suited for this problem: namely, they obtain at least superalgebraic convergence in a numerically stable manner.", "subjects": "Numerical Analysis (math.NA)", "title": "On the numerical stability of Fourier extensions", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9899864281950694, "lm_q2_score": 0.8267117983401363, "lm_q1q2_score": 0.818433460385474 }
https://arxiv.org/abs/1604.02672
On non-maximal prime ideals of $C[0,1]$
We first show a counter intuitive result that in the ring of real valued continuous functions on $[0,1]$ non maximal prime ideals exist. This is a standard proof and a well known result. Interestingly, a non maximal prime ideal in this ring is actually contained inside a unique maximal ideal. We arrive at this result merely by looking at the zero set of ideals in this ring and by making simple geometrical observations. We end by leaving the reader with an interesting open problem that logically follows from this article.
\section*{Introduction} Consider the ring $R = C[0,1] = \{f:[0,1] \rightarrow \mathbb{R} : \textit{f is continuous}\}$ under pointwise addition and multiplication. Consider any two nonzero functions $f$ and $g$ in $R$ whose zero sets are complements of each other in $[0,1]$, then the product $fg$ is the zero function. So, $C[0,1]$ is not an integral domain. Further, we know all the maximal ideals of $R$. All maximal ideals of $R$ are of the form $ M_{\gamma}$ for some $\gamma \in [0,1]$ where $M_{\gamma} = \{f \in R : f(\gamma) = 0\}$. Now, given the points $\gamma_1, \gamma_2 \in [0,1]$, consider the ideal $I = \{f \in R : f(\gamma_1) = f(\gamma_2) = 0 \}$. Is it a prime ideal? No. This is because the polynomial $(x-\gamma_1)(x-\gamma_2) \in I$ but both $x-\gamma_1$ and $x-\gamma_2$ do not individually belong to $I$. \section*{Towards the question} We ask if non maximal prime ideals exist in $R$ and if so where can we locate them. From the above discussion it appears that ideals vanishing at $2$ or more points are not prime ideals. This gives us a nice tool of associating with a given ideal $I$ of $R$, a subset of $[0,1]$ namely $V(I) = \bigcap_{f \in I} V(f)$ where $V(f) = f^{-1}(0)$. For example $V(M_{\gamma}) = \{\gamma\}$. Note that $V(I)$ is compact in $[0,1]$. Also note that if $I_1 \subset I_2$ then $V(I_2) \subset V(I_1)$. We first investigate the existence of non maximal prime ideals in $R$. \section*{An existential proof} Since ideals $I$ with $|V(I)| = 2$ are not prime ideals, one would guess that ideals with $|V(I)| \geq 2$ are not prime ideals as well. Based on this, our intuition might tell us that $R$ has no non maximal prime ideals which would be rather interesting since $R$ is not a principal ideal domain (in fact not even an integral domain). However, we show the existence of non maximal prime ideals. But first a definition. \begin{definition*} A nonempty set $S$ is said to be multiplicative if $1 \in S$ and given any two elements in $S$, the product of these two elements also lies in $S$. \end{definition*} \textbf{Proof} - Let $S$ be the set of all polynomials in $C[0,1]$. Note that $S$ is a multiplicative set. Now consider all the ideals in $R$ with the property that they are disjoint from $S$. Call this set $A$ with the partial ordering of set inclusion. Note that $A$ is nonempty since the zero ideal belongs to it. Consider a chain $I_1 \subset I_2 \subset \ldots I_n \ldots$ of ideals in $A$. Then the ideal $\bigcup I_j$ is clearly an upper bound. By Zorn's Lemma, $A$ has a maximal element. Call it $P$.\\ We claim that $P$ is a prime ideal. If not, then there exist $a,b \in R$ (outside $P$) such that $ab \in P$. Consider the ideals $\langle P,a \rangle$ and $\langle P,b \rangle$. Both these ideals strictly contain $P$ and therefore must intersect with $S$. Hence, there exist $f,g \in R$ and $p,p' \in P$ such that $p+fa, p'+gb \in S$. As $S$ is multiplicative $(p+fa)(p'+gb) \in S$. Now $(p+fa)(p'+gb) = pp' + pgb + p'fa + fgab$. As $P$ is an ideal $pp', pgb, p'fa \in P$. By assumption $ab \in P$ so $fgab \in P$. Therefore, we get that $(p+fa)(p'+gb) \in P$. But this is a contradiction to the fact that $S \cap P = \phi$. Hence, $P$ is a prime ideal and the claim is proved.\\ We further claim that $P$ is not a maximal ideal. We prove this by contradiction. If $P$ is a maximal ideal then $P = M_{\gamma}$ for some $\gamma \in [0,1]$. Then the nonzero polynomial $p(x) = x - \gamma \in P$. But $P$ does not contain any nonzero polynomials. This contradiction proves our claim that $P$ is not a maximal ideal. \section*{Locating non maximal prime ideals} We now know that non maximal prime ideals do exist in $R$. We ask how does the zero set $V(P)$ of a non maximal prime ideal $P$ look like. \begin{theorem*} If $P$ is a prime ideal of $R$ then $V(P)$ cannot be the empty set or a set with more than $1$ element. \end{theorem*} \begin{proof} Since $R$ is a ring with $1$, any ideal $I$ is contained in a maximal ideal. So, $I \subset M_{\gamma}$ for some $\gamma \in [0,1]$. Then, $|V(I)| \geq |V(M_{\gamma})| = 1$. So, $|V(I)| = 0$ is not possible for any ideal $I$ of $R$, and hence in particular not possible for prime ideals of $R$.\\ Now consider the case when $|V(P)| \geq 2$. Let $\gamma_1, \gamma_2 \in V(P)$ with $\gamma_1 < \gamma_2$. Find points $x_1$ and $x_2$ in $(\gamma_1, \gamma_2)$ such that $x_1 < x_2$. Define \[ f(x) = \left\{ \begin{array}{ll} 0 & x\leq x_2 \\ x-x_2 & x\geq x_2 \\ \end{array} \right. \] and \[ g(x) = \left\{ \begin{array}{ll} x-x_1 & x\leq x_1 \\ 0 & x\geq x_1 \\ \end{array} \right. \] Then $f(\gamma_2) \neq 0$ and $g(\gamma_1) \neq 0$, so $f$ and $g$ do not belong to $I$ whereas $fg = 0$ belongs to $I$. Thus $I$ is not a prime ideal. \end{proof} Note that the above argument works equally well regardless of whether $V(P)$ is a finite set or an infinite set. \section*{Conclusion} Since non maximal prime ideals exist in $R$, from the above theorem it directly follows that for such an ideal $P$, $|V(P)| =1$ as all other possibilities are eliminated. \begin{corollary*} A non maximal prime ideal of $C[0,1]$ is (strictly)contained in a unique maximal ideal ($M_\gamma$). \end{corollary*} In conclusion, we remark that we just showed the existence of non maximal prime ideals and located them in some sense. It would be a good follow up if someone can come up with a constructive proof of a non maximal prime ideal in this ring. One can prove that the maximal ideals of $C[0,1]$ are actually unaccountably generated [$\textbf{3}$, p. 404] and it is probably difficult to come up with an explicit set of generators for these ideals. Therefore, it doesn't seem a very easy job to find a generating set for non maximal prime ideals in this ring.
{ "timestamp": "2016-04-12T02:08:55", "yymm": "1604", "arxiv_id": "1604.02672", "language": "en", "url": "https://arxiv.org/abs/1604.02672", "abstract": "We first show a counter intuitive result that in the ring of real valued continuous functions on $[0,1]$ non maximal prime ideals exist. This is a standard proof and a well known result. Interestingly, a non maximal prime ideal in this ring is actually contained inside a unique maximal ideal. We arrive at this result merely by looking at the zero set of ideals in this ring and by making simple geometrical observations. We end by leaving the reader with an interesting open problem that logically follows from this article.", "subjects": "Rings and Algebras (math.RA)", "title": "On non-maximal prime ideals of $C[0,1]$", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9899864281950694, "lm_q2_score": 0.8267117919359419, "lm_q1q2_score": 0.8184334540454086 }
https://arxiv.org/abs/math/0507395
A solution to de Groot's absolute cone conjecture
A compactum X is an `absolute cone' if, for each of its points x, the space X is homeomorphic to a cone with x corresponding to the cone point. In 1971, J. de Groot conjectured that each n-dimensional absolute cone is an n-cell. In this paper, we give a complete solution to that conjecture. In particular, we show that the conjecture is true for n<4 and false for n>4. For n=4, the absolute cone conjecture is true if and only if the 3-dimensional Poincare Conjecture is true.
\section{Introduction} A compactum $X$ is an \emph{absolute suspension} if for any pair of points $x,y\in X$, the space $X$ is homeomorphic to a suspension with $x$ and $y$ corresponding to the suspension points. Similarly, $X$ is an \emph{absolute cone} if, for each point $x\in X$, the space $X$ is homeomorphic to a cone with $x$ corresponding to the cone point. At the 1971 Prague Symposium, J. de Groot \cite{Gr} made the following two conjectures: \begin{conjecture} \label{susp-conj}Every $n$-dimensional absolute suspension is homeomorphic to the $n$-sphere. \end{conjecture} \begin{conjecture} \label{cone-conj}Every $n$-dimensional absolute cone is homeomorphic to an $n$-cell. \end{conjecture} In 1974, Szyma\'{n}ski \cite{Sz} proved Conjecture \ref{susp-conj} in the affirmative for $n=1,2$ or $3$. Later, Mitchell \cite{Mi1} reproved Szyma\'{n}ski's results, and at the same time shed some light on higher dimensions, by showing that every $n$-dimensional absolute suspension is an ENR homology $n$-manifold homotopy equivalent to the $n$-sphere. Still, the `absolute suspension conjecture' remains open for $n\geq4$. In 2005, Nadler \cite{Na} announced a proof of Conjecture \ref{cone-conj} in dimensions $1$ and $2$. In this paper we provide a complete solution to the `absolute cone conjecture'. In particular, we verify Conjecture \ref{cone-conj} for $n\leq3$ and provide counterexamples for all $n\geq5$. For $n=4$ we show that the conjecture is equivalent to the $3$-dimensional Poincar\'{e} Conjecture, that has recently been claimed by Perelman. \bigskip \section{Definitions, notation, and terminology} \subsection{Cones} For any topological space $L$, the \emph{cone} on $L$ is the quotient space \[ cone\left( L\right) =L\times\left[ 0,1\right] /L\times\left\{ 0\right\} \text{.}% \] Let $q:L\times\left[ 0,1\right] \rightarrow cone(L)$ be the corresponding quotient map. We refer to $q\left( L\times\left\{ 0\right\} \right) $ as the \emph{cone point} and we view $L$ as a subspace of $cone\left( L\right) $ via the embedding $L\leftrightarrow L\times\left\{ 1\right\} \hookrightarrow cone\left( L\right) $. We refer to this copy of $L$ as the \emph{base} of the cone. For any $\left( z,t\right) \in L\times\left[ 0,1\right] $, denote $q\left( z,t\right) $ by $t\cdot z$. Thus, $0\cdot z$ represents the cone point and $1\cdot z=z$ for all $z\in L$. For each $z\in L$, the \emph{cone line} corresponding to $z$ is the arc \[ I_{z}=q\left( \left\{ z\right\} \times\left[ 0,1\right] \right) =\left\{ \left. t\cdot z\ \right\vert 0\leq t\leq1\right\} , \] while the \emph{open cone line} corresponding to $z$, denoted by $\overset{\circ}{I}_{z}$, is the set% \[ \overset{\circ}{I}_{z}=q\left( \left\{ z\right\} \times(0,1\right) )=\left\{ \left. t\cdot z\ \right\vert 0<t<1\right\} \text{.}% \] For $\varepsilon\in\left( 0,1\right) $, the \emph{subcone of radius }$\varepsilon$ is the set \[ cone\left( L,\varepsilon\right) =q\left( L\times\left[ 0,\varepsilon \right] \right) =\left\{ \left. t\cdot z\ \right\vert z\in L\text{ and }0\leq t\leq\varepsilon\right\} \] Clearly, each subcone is homeomorphic to $cone(L)$. More generally, if $\lambda:L\rightarrow(0,1)$ is continuous, the $\lambda$\emph{-warped subcone }is defined by \[ cone\left( L,\lambda\right) =\left\{ \left. t\cdot z\ \right\vert z\in L\text{ and }0\leq t\leq\lambda\left( z\right) \right\} \text{.}% \] It also is homeomorphic to $cone(L)$. In fact, the following is easy to prove. \begin{lemma} \label{epsilon-neighborhoods}Let $L$ be a space, $\varepsilon\in\left( 0,1\right) $, and $\lambda:L\rightarrow(0,1)$. Then there is a homeomorphism (in fact, an ambient isotopy) $f:cone(L)\rightarrow cone(L)$ fixed on $L\cup\left\{ \text{cone point}\right\} $ such that $f\left( cone\left( L,\varepsilon\right) \right) =cone\left( L,\lambda\right) $. \end{lemma} By applying the above lemma, or by a similar direct proof, we also have: \begin{lemma} \label{partial-homogeneity}Let $L$ be a space and suppose $t\cdot z$ and $t^{\prime}\cdot z$ are points on the same open cone line of $cone\left( L\right) $. Then there is a homeomorphism (in fact, an ambient isotopy) $f:cone(L)\rightarrow cone(L)$ fixed on $L\cup\left\{ \text{cone point}\right\} $ such that $f\left( t\cdot z\right) =t^{\prime}\cdot z$. \end{lemma} On occasion, we will have use for the \emph{open cone} on $L$, which we view as a subspace of $cone\left( L\right) $. It is defined by% \[ opencone\left( L\right) =L\times\lbrack0,1)/L\times\left\{ 0\right\} . \] \subsection{Suspensions and mapping cylinders} For a topological space $L$, the \emph{suspension} of $L$ is the quotient space \[ susp\left( L\right) =L\times\left[ 0,1\right] /\left\{ L\times\left\{ 0\right\} ,L\times\left\{ 1\right\} \right\} \text{.}% \] In other words, the suspension of $L$ is obtained by separately crushing out the top and bottom levels of the product $L\times\left[ 0,1\right] $. The images of these two sets under the quotient map are called the \emph{suspension points }of $susp\left( L\right) $. Given a map $f:L\rightarrow K$ between disjoint topological spaces, the \emph{mapping cylinder} of $f$ is the quotient space \[ Map\left( f\right) =((L\times\left[ 0,1\right] )\sqcup K)/\sim \] where $\sim$ is the equivalence relation on the disjoint union $(L\times \left[ 0,1\right] )\sqcup K$ induced by the rule: $(x,0)\sim f\left( x\right) $ for all $x\in L$. We view $L$ and $K$ as subsets of $Map\left( f\right) $ via the embeddings induced by \begin{align*} L & \leftrightarrow L\times\left\{ 1\right\} \hookrightarrow (L\times\left[ 0,1\right] )\sqcup K\text{, and}\\ K & \hookrightarrow(L\times\left[ 0,1\right] )\sqcup K. \end{align*} In addition, for each $z\in L$, the inclusion \[ \left\{ z\right\} \times\left[ 0,1\right] \hookrightarrow(L\times\left[ 0,1\right] )\sqcup K \] induces an embedding of an arc into $Map\left( f\right) $. We call the image arc a \emph{cylinder line} and denote it by $E_{z}$. Clearly, if the above range space $K$ consists of a single point, then $Map\left( f\right) =cone\left( L\right) $. Similarly, $cone\left( L\right) $ can always be obtained as a quotient space of $Map\left( f\right) $ by crushing $K$ to a point. The following lemma will be useful later. It allows us to view certain cones as mapping cylinders having one of the cone lines as the range space. \begin{lemma} \label{cone-to-cylinder}Let $Y$ be a space and suppose $y\in Y$ has a $k$-dimensional euclidean neighborhood $U$ in $Y$. Let $B_{0}^{k}$ and $B_{1}^{k}$ be (tame) $k$-cell neighborhoods of $y$ lying in $U$ such that $B_{1}^{k}\subseteq int\left( B_{0}^{k}\right) $. Then the pair $\left( cone\left( Y\right) ,I_{y}\right) $ is homeomorphic to $\left( Map\left( f\right) ,I_{y}\right) $ for some map $f:Y-int\left( B_{1}^{k}\right) \rightarrow I_{y}$. The homeomorphism may be chosen to be the identity on $(Y-int\left( B_{1}^{k}\right) )\cup I_{y}$. \begin{proof} Choose a homeomorphism $h:B_{0}^{k}-int\left( B_{1}^{k}\right) \rightarrow S^{k-1}\times\left[ 0,1\right] $ taking $\partial B_{0}^{k}$ to $S^{k-1}\times\left\{ 0\right\} $ and $\partial B_{1}^{k}$ to $S^{k-1}% \times\left\{ 0\right\} $. Then define $f:Y-int\left( B_{1}^{k}\right) \rightarrow I_{y}$ by\medskip% \[ f\left( x\right) =\left\{ \begin{tabular} [c]{ll}% $t\cdot y$ & if $x\in B_{0}^{k}-int\left( B_{1}^{k}\right) $ and $h\left( x\right) \in S^{k-1}\times\left\{ t\right\} $\\ $0\cdot y\quad$ & if $x\in Y-int\left( B_{0}^{k}\right) $% \end{tabular} \ \ \medskip\ \right. . \] Since $f$ sends all points of $Y-int\left( B_{0}^{k}\right) $ to the cone point $0\cdot y$, we may identify the `sub-mapping cylinder' $Map\left( \left. f\right\vert _{Y-int\left( B_{0}^{k}\right) }\right) $ with the `subcone' $cone\left( Y-int\left( B_{1}^{k}\right) \right) $. In addition, it is easy to build a homeomorphism between the $\left( k+1\right) $-cell $Map\left( \left. f\right\vert _{B_{0}^{k}-int\left( B_{1}^{k}\right) }\right) $ and the $\left( k+1\right) $-cell $cone\left( B_{0}^{k}\right) $ taking $I_{y}$ identically onto $I_{y}$, and each cylinder line emanating from an $x\in\partial B_{0}^{k}$ identically onto the corresponding cone line. Fitting these pieces together yields the desired homeomorphism between $Map\left( f\right) $ and $cone\left( Y\right) $. See Figure 1.% \begin{figure} [ptb] \begin{center} \includegraphics[ height=2.4898in, width=5.7683in ]% {cones-fig1.eps}% \caption{ }% \end{center} \end{figure} \end{proof} \end{lemma} \subsection{ENR homology manifolds} A space is a \emph{euclidean neighborhood retract (ENR)} if it is a retract of some open subset of euclidean space. This is equivalent to being a finite-dimensional separable metric ANR. A space that is a retract of $\mathbb{R}^{n}$ (for some $n$) is called a \emph{euclidean retract (ER)}. This is equivalent to being a contractible ENR. A locally compact ENR $X$ is an \emph{ENR homology }$n$\emph{-manifold} if, for every $x\in X$,% \begin{equation} H_{\ast}\left( X,X-x\right) \cong\left\{ \begin{tabular} [c]{ll}% $\mathbb{Z}$ & if $\ast=n$\\ $0$ & otherwise \end{tabular} \ \ \ \right. \text{.} \tag{\dag$_n$}% \end{equation} We call $X$ an \emph{ENR homology }$n$\emph{-manifold with boundary} if, for every $x\in X$,% \[ H_{\ast}\left( X,X-x\right) \cong\left\{ \begin{tabular} [c]{ll}% $0$ or $\mathbb{Z}$ & if $\ast=n$\\ $0$ & otherwise \end{tabular} \ \ \ \right. \text{.}% \] In this case, the \emph{boundary of }$X$ is the set% \[ \partial X=\left\{ \left. x\in X\ \right\vert H_{\ast}\left( X,X-x\right) \equiv0\right\} \text{,}% \] and the \emph{interior of }$X$ is the set% \[ int\left( X\right) =X-\partial X\text{.}% \] In all of the above and throughout this paper, except where stated otherwise, homology is singular with integer coefficients. By \cite{Mi2}, $\partial X$ is a closed subset of $X$; hence, $int\left( X\right) $ is an ENR homology $n$-manifold. In addition, if Borel-Moore homology is used, $\partial X$ satisfies the algebraic condition for being a homology $\left( n-1\right) $-manifold, i.e., $\partial X$ satisfies (\dag$_{n-1}$). For ENRs, Borel-Moore homology agrees with singular homology, so if $\partial X$ is an ENR then it is an ENR homology $\left( n-1\right) $-manifold. \begin{remark} There are interesting situations where, although $X$ is an ENR homology manifold with boundary, $\partial X$ is not an ENR. See, for example, \cite{AG} or \cite{Fi}. For the spaces of interest in this paper, existing conditions will prevent this from happening. \end{remark} \section{Absolute cones} Suppose a compactum $X$ is an absolute cone. For each $x\in X$, choose $L_{x}\subseteq X$ and a homeomorphism $h_{x}:cone\left( L_{x}\right) \rightarrow X$ which is the identity on $L_{x}$ and sends the cone point to $x$. We will refer to $L_{x}$ as the \emph{link} of $x$ in $X$. (\textbf{Note.} The choice of $L_{x}$ and $h_{x}$ may not be unique; however, for each $x$ we make a choice and stick with it.) For $\varepsilon\in\left( 0,1\right) $ and $\lambda:L\rightarrow(0,1)$ let $N\left( x,\varepsilon \right) =h_{x}\left( cone\left( L,\varepsilon\right) \right) $ and $N\left( x,\lambda\right) =h_{x}\left( cone\left( L,\lambda\right) \right) $. We refer to these as the $\varepsilon$\emph{-cone neighborhood} and the \emph{warped }$\lambda$\emph{-cone neighborhood} of $x$, respectively. Clearly, each point of $x$ has arbitrarily small $\varepsilon$-cone neighborhoods. In a similar vein, for any $x\in X$ and $z\in L_{x}$, let $J_{x}\left( z\right) $ and $\overset{\circ}{J}_{x}\left( z\right) $ denote $h_{x}\left( I_{z}\right) $ and $h_{x}(\overset{\circ}{I}_{z})$, respectively. We refer to these as \emph{[open] cone lines of} $X$ with respect to $x$. The following proposition lists several easy properties of absolute cones. \begin{proposition} \label{ANR}Let $X$ be a finite dimensional absolute cone, $x\in X$ and $z\in L_{x}$. Then \begin{enumerate} \item $X$ is a compact ER, \item $L_{x}$ is a compact ENR, \item $H_{\ast}\left( X,X-x\right) \cong\widetilde{H}_{\ast-1}\left( L_{x}\right) $, \item $L_{z}$ is contractible, and \item $H_{\ast}\left( X,X-z\right) \equiv0$. \end{enumerate} \begin{proof} Since each point of $X$ has arbitrarily small $\varepsilon$-cone neighborhoods, $X$ is locally contractible; so by \cite[V.7.1]{Hu}, $X$ is an ENR. Since $X$ is also contractible, it is an ER. Since $L_{x}$ is a retract of its neighborhood $X-x\approx L_{x}\times \lbrack0,1)$ in $X$, it too is an ENR \cite[III.7.7]{Hu}. Being a closed subset of $X$, $L_{x}$ is also compact. To prove 3), we again use $X-x\approx L_{x}\times(0,1]$. Since $X$ is contractible, the desired isomorphisms may be obtained from the long exact sequence for the pair $\left( X,X-x\right) $. The canonical contraction of $cone\left( L_{x}\right) $ along cone lines restricts to a contraction of $cone\left( L_{x}\right) -z$, since $z$ lies in the base. Thus, $X-z$ is contractible. Since $X-z\approx L_{z}\times(0,1]$, it follows that $L_{z}$ is contractible. Assertion 5) follows from 3) and 4). \end{proof} \end{proposition} The next proposition is a key ingredient in our understanding of absolute cones. \begin{proposition} \label{links}Let $X$ be a finite dimensional absolute cone and \[ B_{X}=\left\{ \left. z\in X\ \right\vert H_{\ast}\left( X,X-z\right) =0\right\} . \] Then \begin{enumerate} \item $L_{x}\subseteq B_{X}$ for all $x\in X$, \item $X-B_{X}\neq\varnothing$, and \item For all $x\in X-B_{X}$, $L_{x}=B_{X}$. \end{enumerate} \begin{proof} Assertion 1) just restates part of Proposition \ref{ANR}, while Assertion 2) is a basic fact in dimension theory. In particular, if $\dim X=n$, then there exists $x\in X$ such that $H_{n}\left( X,X-x\right) \neq0$; see, for example, \cite[Lemma 2]{Mi2}. To prove 3), fix $x\in X-B_{X}$ and suppose $y\in X-L_{x}$. We must show that $y\notin B_{X}$, i.e., that $H_{\ast}\left( X,X-y\right) $ is non-trivial. Choose $\varepsilon<1$ sufficiently small that $N\left( y,\varepsilon\right) \cap L_{x}=\varnothing$. Choose $z\in L_{y}$ such that $x$ lies on the open cone line $\overset{\circ}{J}_{y}\left( z\right) $. By Lemma \ref{partial-homogeneity}, $H_{\ast}\left( X,X-x\right) \cong H_{\ast }\left( X,X-x^{\prime}\right) $ for all $x^{\prime}\in\overset{\circ}{J}% _{y}\left( z\right) $. Therefore, $\overset{\circ}{J}_{y}\left( z\right) \cap L_{x}=\varnothing$. By `pushing out along' $\overset{\circ}{J}_{y}\left( z\right) $ we may expand the $\varepsilon$-cone neighborhood $N\left( y,\varepsilon\right) $ about $y$ to a warped $\lambda$-cone neighborhood $N\left( y,\lambda\right) $ which contains $x$ in its interior and is disjoint from $L_{x}$. See Figure 2.% \begin{figure} [ptb] \begin{center} \includegraphics[ height=2.7069in, width=3.0165in ]% {cones-fig2.eps}% \caption{ }% \end{center} \end{figure} Since the inclusions $\left( X,L_{x}\right) \hookrightarrow\left( X,X-x\right) $ and $\left( X,X-N\left( y,\lambda\right) \right) \hookrightarrow\left( X,X-y\right) $ are both homotopy equivalences of pairs, we have inclusion induced isomorphisms% \begin{align*} & H_{\ast}\left( X,L_{x}\right) \overset{\cong}{\longrightarrow}H_{\ast }\left( X,X-x\right) \text{, and}\\ & H_{\ast}\left( X,X-N\left( y,\lambda\right) \right) \overset{\cong }{\longrightarrow}H_{\ast}\left( X,X-y\right) \end{align*} The first of these can be factored via inclusions as such:% \[ H_{\ast}\left( X,L_{x}\right) \overset{\phi}{\longrightarrow}H_{\ast}\left( X,X-N\left( y,\lambda\right) \right) \overset{\psi}{\longrightarrow}% H_{\ast}\left( X,X-x\right) \] Then $\phi$ is necessarily injective and $\psi$ surjective, so $H_{\ast }\left( X,X-N\left( y,\lambda\right) \right) $ is non-trivial. Thus, $H_{\ast}\left( X,X-y\right) \neq0$. \end{proof} \end{proposition} \begin{corollary} \label{1}For all $x\in X-B_{X}$, $H_{\ast}\left( X,X-x\right) \cong \widetilde{H}_{\ast-1}\left( B_{X}\right) $. This homology is finitely generated. \begin{proof} Since $B_{X}=L_{x}$, the isomorphism follows from Proposition \ref{ANR}. Since $B_{X}=L_{x}$ is a compact ENR, it has the homotopy type of a finite CW complex \cite{We}; hence, finitely generated homology. \end{proof} \end{corollary} \begin{corollary} \label{2}For any $x,y\in X-B_{X}$, there exists a homeomorphism $f:X\rightarrow X$ which is the identity on $B_{X}$ and sends $x$ to $y$. \begin{proof} Since $L_{x}=B_{X}=L_{y}$, we may let $f=h_{y}\circ h_{x}^{-1}$. \end{proof} \end{corollary} \begin{theorem} \label{gen-manifold-theorem}If $X$ is an $n$-dimensional absolute cone, then \begin{enumerate} \item $X$ is an ENR homology $n$-manifold with boundary, \item $\partial X$ is precisely the link $L_{x}$ of any point $x\in int(X)$, \item $\partial X$ is an ENR homology $\left( n-1\right) $-manifold homotopy equivalent to $S^{n-1}$, and \item for each $z\in\partial X$, $L_{z}$ is a contractible ENR homology $\left( n-1\right) $-manifold with boundary. \end{enumerate} \begin{proof} As above, let $B_{X}=\left\{ \left. z\in X\ \right\vert H_{\ast}\left( X,X-z\right) \equiv0\right\} $. By Proposition \ref{ANR} and Corollaries \ref{1} and \ref{2}, $X-B_{X}$ is a homogeneous $n$-dimensional ENR with finitely generated local homology. By an application of \cite{Bre} or \cite{Bry}, $X-B_{X}$ is an ENR homology $n$-manifold. Therefore, $X$ is an ENR homology $n$-manifold with boundary, and $\partial X=B_{X}$. Assertion 2) is a restatement of Proposition \ref{links}. Proposition \ref{ANR} and an application of \cite{Mi2} (as discussed earlier) tells us that $\partial X$ is an ENR homology $\left( n-1\right) $-manifold. Moreover, by Assertion 1) and Corollary \ref{1}, for any $x\in int\left( X\right) $, \[ \widetilde{H}_{k-1}\left( \partial X\right) \cong H_{k}\left( X,X-x\right) \cong\left\{ \begin{tabular} [c]{ll}% $\mathbb{Z}$ & if $k=n$\\ $0$ & if $k\neq n$% \end{tabular} \ \ \ \ \ \ \ \ \ \right. \] Thus, $\partial X$ has the homology of $S^{n-1}$. If $n=1,2$ or $3$ it is known that every homology $\left( n-1\right) $-manifold is an actual $\left( n-1\right) $-manifold \cite[Ch.IX]{Wi}; and in those dimensions a manifold is determined by its homology. Hence $\partial X$ is homeomorphic to $S^{n-1}$. In higher dimensions we only claim a homotopy equivalence between $S^{n-1}$ and $\partial X$. This may be obtained in the usual way if we can show that $\partial X$ is simply connected. In particular, the Hurewicz Theorem would then assure us that $\pi_{n-1}\left( \partial X\right) \cong H_{n-1}\left( \partial X\right) \cong\mathbb{Z}$. A generator of $\pi _{n-1}\left( \partial X\right) $ provides a degree $1$ map from $S^{n-1}$ to $\partial X$. Since $\partial X$ is an ANR---and thus has the homotopy type of a CW complex---a theorem of Whitehead shows that this map is a homotopy equivalence. We hold off proving simple connectivity of $\partial X$ until after we verify Assertion 4). To prove 4), note that the homeomorphism $h_{z}:cone\left( L_{z}\right) \rightarrow X$ induces a homeomorphism of $L_{z}\times(0,1]$ onto $X-z$, taking $L_{z}\times\left\{ 1\right\} $ onto $L_{z}$. Since $L_{z}% \times(0,1]$ is an ENR homology $n$-manifold with boundary, $L_{z}$ is an ENR homology $\left( n-1\right) $-manifold with boundary. This is an application of \cite[Th.6]{Ra}. We have already observed (Proposition \ref{ANR}) that $L_{z}$ is contractible. Lastly we complete Assertion 3) by showing that $\partial X$ is simply connected when $n\geq2$. Since the above mentioned homeomorphism $L_{z}% \times(0,1]\rightarrow X-z$ must take (homology) boundary to boundary, it follows that $\partial X-z$ has the structure of $L_{z}$ with an open collar attached to $\partial L_{z}$. Thus, $\partial X-z$ is contractible; and $z$ has a neighborhood in $\partial X$ homeomorphic to a cone over $\partial L_{z}$. Therefore, $\partial X$ may be viewed as the union of open sets $\partial X-z$ and $U$, where $\partial X-z$ is contractible and $U$ is homeomorphic to the open cone on $\partial L_{z}$. Since the intersection of these sets is connected, simple connectivity follows from Van Kampen's theorem. \end{proof} \end{theorem} The above proof provides some additional structure information about absolute cones which we record as: \begin{theorem} If $X$ is an $n$-dimensional absolute cone, then $X$ is a contractible ENR homology $n$-manifold with boundary and $\partial X$ is a locally conical ENR homology $\left( n-1\right) $-manifold; more specifically, each point of $\partial X$ has a neighborhood in $\partial X$ which is a cone over an ENR homology $\left( n-2\right) $-manifold with the homology of $S^{n-2}$. \begin{proof} By the above proof, each $z\in\partial X$ has a neighborhood in $\partial X$ homeomorphic to $cone\left( \partial L_{z}\right) $. Since this cone lies in $\partial X$, $\partial L_{z}$ must be an ENR. Moreover, since $L_{z}$ is an ENR homology $\left( n-1\right) $-manifold with boundary, then $\partial L_{z}$ is an ENR homology $\left( n-2\right) $-manifold. Lastly, since $\partial X$ has the local homology of an $\left( n-1\right) $-manifold at $z$, the homology type of $\partial L_{z}$ must be that of an $\left( n-2\right) $-sphere. \end{proof} \end{theorem} \begin{corollary} If $X$ is an $n$-dimensional absolute cone and $n\leq3$, then $X$ is an $n$-cell. The same is true for $n=4$, provided the $3$-dimensional Poincar\'{e} Conjecture is true. \begin{proof} If $n\leq3$, we have already observed in the proof of Theorem \ref{gen-manifold-theorem} that $B_{X}=\partial X$ is homeomorphic to $S^{n-1}$. Thus, $X\approx cone\left( S^{n-1}\right) \approx B^{n}$. If $n=4$, we have shown that $\partial X$ is an ENR homology $3$-manifold homotopy equivalent to $S^{3}$. In addition, we know that each point of $\partial X$ has a neighborhood in $\partial X$ homeomorphic to the cone over an ENR homology $2$-manifold having the homology of $S^{2}$. As above, such ENR homology $2$-manifolds are $2$-spheres. Thus, $\partial X$ is an actual $3$-manifold. Assuming the $3$-dimensional Poincar\'{e} Conjecture, $\partial X\approx S^{3}$; so $X$ is a $4$-cell. \end{proof} \end{corollary} \begin{remark} At the conclusion of the next section, we will show that if there exists a homotopy $3$-sphere $H^{3}$, not homeomorphic to $S^{3}$, then $cone\left( H^{3}\right) $ is a $4$-dimensional absolute cone that is not a $4$-cell. \end{remark} \section{Counterexamples in higher dimensions} The main goal of this section is to construct, for all $n\geq5$, $n$-dimensional absolute cones which are not $n$-cells. In all cases, we begin with a non-simply connected $k$-manifold $\Sigma^{k}$ having the same $\mathbb{Z}$-homology as $S^{k}$. Existence of such manifolds for all $k\geq3$ is well-known. Our counterexamples are obtained by first coning over $\Sigma^{k}$, then suspending that cone. This section is primarily devoted to proving that the resulting spaces are absolute cones, but not cells. For completeness, we will conclude this section by showing that---if there is a counterexample to the $3$-dimensional Poincar\'{e}---then there is also a $4$-dimensional absolute cone that is not a $4$-cell. We begin our construction of counterexamples in dimensions $\geq5$ with a very general lemma. \begin{lemma} \label{cones-suspensions}For any space $Y$, the following are homeomorphic. \begin{enumerate} \item $susp\left( cone\left( Y\right) \right) $ \item $cone\left( susp\left( Y\right) \right) $ \item $cone\left( cone\left( Y\right) \right) $ \item $cone\left( Y\right) \times\left[ 0,1\right] $. \end{enumerate} \end{lemma} Before proving this lemma, consider the map $f:cone\left( Y\right) \times\left[ 0,1\right] \rightarrow cone\left( Y\right) \times\left[ 0,1\right] $ defined by \[ f\left( t\cdot z,s\right) =\left( (st)\cdot z,s\right) . \] This map takes $cone\left( Y\right) \times\left\{ 1\right\} $ identically onto $cone\left( Y\right) \times\left\{ 1\right\} $, and each level set $cone\left( Y\right) \times\left\{ s\right\} $ to the subcone of radius $s$ contained in $cone\left( Y\right) \times\left\{ s\right\} $; finally $cone\left( Y\right) \times\left\{ 0\right\} $ is taken to the cone point of $cone\left( Y\right) \times\left\{ 1\right\} $. This map induces a level-preserving embedding of $cone\left( cone\left( Y\right) \right) $ into $cone\left( Y\right) \times\left[ 0,1\right] $. The image of the embedding is a particularly nice realization of $cone\left( cone\left( Y\right) \right) $ which we will denote by $\mathcal{C}^{2}\left( Y\right) $. A similar map $g:cone\left( Y\right) \times\left[ 0,1\right] \rightarrow cone\left( Y\right) \times\left[ 0,1\right] $ can be used to induce an embedding of $susp\left( cone\left( Y\right) \right) $ into $cone\left( Y\right) \times\left[ 0,1\right] $. We will denote the image of that map by $\mathcal{SC}\left( Y\right) $. Each of the spaces $cone\left( Y\right) \times\left[ 0,1\right] $, $\mathcal{C}^{2}\left( Y\right) $ and $\mathcal{SC}\left( Y\right) $ contain as a subspace $\left\{ \text{cone point}\right\} \times\left[ 0,1\right] $, which we call the \emph{axis }and denote by\emph{ }$A$. In addition, the points $($cone point$,0)$, $($cone point,$1)$, and $($cone point$,\frac{1}{2})$ will be denoted $p_{0}$, $p_{1}$ and $p_{\ast}$, respectively. See Figure 3.% \begin{figure} [ptb] \begin{center} \includegraphics[ height=4.2704in, width=4.2134in ]% {cones-fig3.eps}% \caption{ }% \end{center} \end{figure} \begin{proof} [Proof of Lemma \ref{cones-suspensions}]We first show that $cone\left( cone\left( Y\right) \right) \approx cone\left( Y\right) \times\left[ 0,1\right] $ by illustrating a homeomorphism from $cone\left( Y\right) \times\left[ 0,1\right] $ to $\mathcal{C}^{2}\left( Y\right) $. Each cone line $I_{y}$ of $cone\left( Y\right) $ determines a `square' $S_{y}% =I_{y}\times\left[ 0,1\right] $ in $cone\left( Y\right) \times\left[ 0,1\right] $. Similarly, $I_{y}$ determines a `right triangle' $T_{y}$ in $\mathcal{C}^{2}\left( Y\right) $ such that $T_{y}$ and $S_{y}$ have two common sides: $I_{y}\times\left\{ 1\right\} $ and $A$. See Figure 4.% \begin{figure} [ptb] \begin{center} \includegraphics[ height=3.122in, width=5.2641in ]% {cones-fig4.eps}% \caption{ }% \end{center} \end{figure} For a given $y$ choose a homeomorphism \[ k_{y}:S_{y}\rightarrow T_{y}% \] which is the identity on their common sides. By using the `same' homeomorphism for each $y\in Y$, we may combine these into a single homeomorphism.% \[ k:cone\left( Y\right) \times\left[ 0,1\right] \rightarrow\mathcal{C}% ^{2}\left( Y\right) . \] A similar strategy produces a homeomorphism of $cone\left( Y\right) \times\left[ 0,1\right] $ onto $\mathcal{SC}\left( Y\right) $. Thus we have $cone\left( cone\left( Y\right) \right) \approx cone\left( Y\right) \times\left[ 0,1\right] \approx susp\left( cone\left( Y\right) \right) $. Lastly, we show that $susp\left( cone\left( Y\right) \right) \approx cone\left( susp\left( Y\right) \right) $ by observing that $\mathcal{SC}% \left( Y\right) $ may be given a cone structure with base $\mathcal{S}% \left( Y\right) $ and cone point $p_{\ast}$. To this end, we view $\mathcal{SC}\left( Y\right) $ as the union of its `suspension triangles' intersecting in $A$; each triangle being the suspension of a cone line. To place a cone structure on $\mathcal{SC}\left( Y\right) $ we place the obvious common cone structure on each of these triangles, with $p_{\ast}$ serving as cone point. See Figure 5.% \begin{figure} [ptb] \begin{center} \includegraphics[ height=2.674in, width=4.9848in ]% {cones-fig5.eps}% \caption{ }% \end{center} \end{figure} \end{proof} \begin{notation} For convenience, we denote the each of the homeomorphic topological spaces described by 1)-4) of Lemma \ref{cones-suspensions} by $\Theta\left( Y\right) $. \end{notation} The following lemma will be useful later. We include it now because its proof is similar to the last part of the above argument. \begin{lemma} \label{cylinder-product}For any map $f:Y\rightarrow K$, the pair $\left( Map\left( f\right) \times\left[ 0,1\right] ,K\times\left\{ \frac{1}% {2}\right\} \right) $ is homeomorphic to $\left( Map\left( F\right) ,K\right) $ for some map $F:\left( Y\times\left[ 0,1\right] \right) \cup(Map\left( f\right) \times\{0,1\})\rightarrow K$. The homeomorphism may be chosen to take each point $\left( k,\frac{1}{2}\right) \in K\times \left\{ \frac{1}{2}\right\} $ to $k$. \end{lemma} \begin{proof} To simplify matters, we assume that $f$ is surjective. If it is not, the argument requires only minor adjustments. By surjectivity, $Map\left( f\right) $ is the union of its cylinder lines $\left\{ E_{y}\mid y\in Y\right\} $. Hence, $Map\left( f\right) \times\left[ 0,1\right] $ is a union of `squares' $\left\{ S_{y}\mid y\in Y\right\} $, where $S_{y}$ denotes $E_{y}\times\left[ 0,1\right] $. Furthermore, each $S_{y}$ intersects the proposed domain of our map $F$ in three of its four boundary edges. More precisely, \[ S_{y}\cap\left( \left( Y\times\left[ 0,1\right] \right) \cup(Map\left( f\right) \times\{0,1\})\right) =(\left\{ y\right\} \times\left[ 0,1\right] )\cup\left( E_{y}\times\left\{ 0,1\right\} \right) . \] Due to its shape, we denote the right-hand side of the above equation by $C_{y}$. Then $\left( Y\times\left[ 0,1\right] \right) \cup(Map\left( f\right) \times\{0,1\})$ is the union of the collection $\left\{ C_{y}\mid y\in Y\right\} $ and, for all $y,y^{\prime}\in Y$, $C_{y}\cap C_{y^{\prime}% }=\varnothing$ unless $f(y)=f\left( y^{\prime}\right) $. Define $F$ by sending each point of $C_{y}$ to $f\left( y\right) $. Then% \[ F:\left( Y\times\left[ 0,1\right] \right) \cup(Map\left( f\right) \times\{0,1\})\rightarrow K \] is continuous, and for each $y\in Y$, the sub-mapping cylinder $Map\left( \left. F\right\vert _{C_{y}}\right) $ is simply $cone(C_{y})$, with cone point $f\left( y\right) \in K$. Thus, $Map\left( F\right) $ is a union of the collection $\left\{ cone\left( C_{y}\right) \mid y\in Y\right\} $. To produce the desired homeomorphism from $Map\left( f\right) \times\left[ 0,1\right] $ to $Map\left( F\right) $, choose a coherent collection of homeomorphisms from the squares $\{S_{y}\}$ making up $Map\left( f\right) \times\left[ 0,1\right] $ to the cones $\{cone(C_{y})\}$ making up $Map\left( F\right) $. In particular, for each $y$, define a homeomorphism $h_{y}:S_{y}\rightarrow cone(C_{y})$ which: takes $\left( f\left( y\right) ,\frac{1}{2}\right) $ to the cone point $f\left( y\right) $, is the identity on $C_{y}$, and is linear on segments in between these subspaces. The union of these homeomorphisms is the desired homeomorphism from $Map\left( f\right) \times\left[ 0,1\right] $ to $Map\left( F\right) $. \end{proof} We are now ready to prove the main theorem of this section. \begin{theorem} If $\Sigma^{k}$ is a closed $k$-manifold with the same $\mathbb{Z}$-homology as the $k$-sphere, then $\Theta\left( \Sigma^{k}\right) $ is a $\left( k+2\right) $-dimensional absolute cone. If $\Sigma^{k}$ is not simply connected, then $\Theta\left( \Sigma^{k}\right) $ is not homeomorphic to an $\left( k+2\right) $-cell. \begin{proof} We begin by observing that $\Theta\left( \Sigma^{k}\right) $ is ENR homology $\left( k+2\right) $-manifold with boundary. To experts on homology manifolds, this may be obvious; otherwise, proceed as follows. First note that $cone\left( \Sigma^{k}\right) $ is an actual $\left( k+1\right) $-manifold with boundary at all points except the cone point $p$. At that point \begin{align*} H_{\ast}\left( cone\left( \Sigma^{k}\right) ,cone\left( \Sigma^{k}\right) -p\right) & \cong H_{\ast}\left( cone\left( \Sigma^{k}\right) ,\Sigma^{k}\right) \\ & \cong\widetilde{H}_{\ast-1}\left( \Sigma^{k}\right) \\ & \cong\widetilde{H}_{\ast-1}\left( S^{k}\right) \cong\left\{ \begin{tabular} [c]{ll}% $\mathbb{Z}$ & if $\ast=k+1$\\ $0$ & otherwise \end{tabular} \ \ \ \ \ \right. \end{align*} Thus, $cone\left( \Sigma^{k}\right) $ is an ENR homology $\left( k+1\right) $-manifold with boundary; moreover, $int\left( cone\left( \Sigma^{k}\right) \right) =opencone\left( \Sigma^{k}\right) $. By \cite{Ra} or straightforward calculation, $cone\left( \Sigma^{k}\right) \times\left[ 0,1\right] $ is an ENR homology $\left( k+2\right) $-manifold with boundary, and \[ int\left( cone\left( \Sigma^{k}\right) \times\left[ 0,1\right] \right) =opencone\left( \Sigma^{k}\right) \times\left( 0,1\right) . \] \smallskip \noindent\textsc{Claim 1. }$\partial\left( \Theta\left( \Sigma^{k}\right) \right) $ \emph{is not a} $\left( k+1\right) $\emph{-manifold. Therefore} $\Theta\left( \Sigma^{k}\right) $ \emph{is not an} $\left( n+2\right) $\emph{-cell}. Under the realization of $\Theta\left( \Sigma^{k}\right) $ as $susp(cone\left( \Sigma^{k}\right) )$, the boundary is $susp\left( \Sigma^{k}\right) $. If $susp\left( \Sigma^{k}\right) $ were an actual $\left( k+1\right) $-manifold, then removing a finite collection of points would not change its fundamental group. But $susp\left( \Sigma^{k}\right) $ is simply connected, and $susp\left( \Sigma^{k}\right) -\left\{ p_{0}% ,p_{1}\right\} \approx\Sigma^{k}\times\left( 0,1\right) $ is not simply connected. The claim follows.\bigskip We now work toward showing that $\Theta\left( \Sigma^{k}\right) $ is an absolute cone. For each $x\in\Theta\left( \Sigma^{k}\right) $, we must identify a subspace $L_{x}$ of $\Theta\left( \Sigma^{k}\right) $ and a homeomorphism of $cone\left( L_{x}\right) $ onto $\Theta\left( \Sigma ^{k}\right) $ taking the cone point to $x$. The proof splits into the following three cases: \begin{itemize} \item $x\in int\left( \Theta\left( \Sigma^{k}\right) \right) ,$ \item $x=p_{0}$ or $p_{1},$ \item $x\in\partial\left( \Theta\left( \Sigma^{k}\right) \right) -\left\{ p_{0},p_{1}\right\} $ \end{itemize} \begin{case} $x\in int\left( \Theta\left( \Sigma^{k}\right) \right) .$ \end{case} In this case we know from Theorem \ref{gen-manifold-theorem} that, if $L_{x}$ exists, it must equal $\partial\left( \Theta\left( \Sigma^{k}\right) \right) $. By applying Lemma \ref{cones-suspensions} to realize $\Theta\left( \Sigma^{k}\right) $ as $cone\left( susp\left( \Sigma ^{k}\right) \right) $, we see that $\Theta\left( \Sigma^{k}\right) $ is indeed homeomorphic to the cone over $\partial\left( \Theta\left( \Sigma ^{k}\right) \right) $. Let $p_{\ast}\in int\left( \Theta\left( \Sigma ^{k}\right) \right) $ be the corresponding cone point and $h_{p_{\ast}% }:cone(\partial\left( \Theta\left( \Sigma^{k}\right) )\right) \rightarrow\Theta\left( \Sigma^{k}\right) $ the homeomorphism. We must show that all other elements of $int\left( \Theta\left( \Sigma^{k}\right) \right) $ can be viewed similarly. The following is the essential ingredient.\bigskip \noindent\textsc{Claim 2. }$int\left( \Theta\left( \Sigma^{k}\right) \right) $ \emph{is an actual} $\left( k+2\right) $\emph{-manifold. In fact, }$int\left( \Theta\left( \Sigma^{k}\right) \right) \approx\mathbb{R}% ^{k+2}$. As noted earlier, we may view $int\left( \Theta\left( \Sigma^{k}\right) \right) $ as $opencone\left( \Sigma^{k}\right) \times\left( 0,1\right) $. That this space is a $\left( k+2\right) $-manifold is a direct application of work by J.W. Cannon and R.D. Edwards on the `double suspension problem' \cite{Ca}, \cite{Ed}. A nice exposition of the relevant result may be found in \cite[Cor. 24.3D]{Da}. Once we know that $int\left( \Theta\left( \Sigma ^{k}\right) \right) $ is a manifold, an application of \cite{St} gives us the homeomorphism to $\mathbb{R}^{k+2}$.\bigskip Now let $x$ be an arbitrary element of $int\left( \Theta\left( \Sigma ^{k}\right) \right) $. By a standard homogeneity argument for manifolds, there is a homeomorphism $u:int\left( \Theta\left( \Sigma^{k}\right) \right) \rightarrow int\left( \Theta\left( \Sigma^{k}\right) \right) $ taking $p_{\ast}$ to $x$. Moreover, we may choose $u$ to be the identity outside some compact neighborhood of $\left\{ p_{\ast},x\right\} $. This allows us to extend $u$ to a homeomorphism $\overline{u}$ of $\Theta\left( \Sigma^{k}\right) $ to itself. The homeomorphism $\overline{u}\circ h_{p_{\ast}}:cone(\partial\left( \Theta\left( \Sigma^{k}\right) )\right) \rightarrow\Theta\left( \Sigma^{k}\right) $ now realizes $x$ as the cone point. \begin{case} $x=p_{0}$ or $p_{1}$. \end{case} Use Lemma \ref{cones-suspensions} to view $\Theta\left( \Sigma^{k}\right) $ as $cone\left( cone\left( \Sigma^{k}\right) \right) $ (or more precisely $\mathcal{C}^{2}\left( \Sigma^{k}\right) $). Then $p_{0}$ corresponds to the cone point; and the base, $cone\left( \Sigma^{k}\right) $, serves as $L_{p_{0}}$. Furthermore, the realizations of $\Theta\left( \Sigma ^{k}\right) $ provided by 1) or 3) of Lemma \ref{cones-suspensions} reveal an involution of $\Theta\left( \Sigma^{k}\right) $ interchanging $p_{0}$ and $p_{1}$. Thus, $p_{1}$ also may be viewed as a cone point. \begin{case} $x\in\partial\left( \Theta\left( \Sigma^{k}\right) \right) -\left\{ p_{0},p_{1}\right\} $. \end{case} By realization 1) of Lemma \ref{cones-suspensions}, $\Theta\left( \Sigma ^{k}\right) -\left\{ p_{1},p_{0}\right\} \approx cone\left( \Sigma ^{k}\right) \times\left( 0,1\right) $, which by another application of Cannon-Edwards, is a $\left( k+2\right) $-manifold with boundary. That boundary corresponds to $\Sigma^{k}\times\left( 0,1\right) $. By another standard homogeneity argument for manifolds, any two points of $\Sigma ^{k}\times\left( 0,1\right) $ can be interchanged by a homeomorphism of $cone\left( \Sigma^{k}\right) \times\left( 0,1\right) $. This homeomorphism can be arranged to be the identity off a compact set; and, thus, extends to a self-homeomorphism of $susp\left( cone\left( \Sigma^{k}\right) \right) $. This means that it suffices to find a single point $x_{0}% \in\partial\left( \Theta\left( \Sigma^{k}\right) \right) -\left\{ p_{1},p_{0}\right\} $ at which $\Theta\left( \Sigma^{k}\right) $ is conical. To this end, we return to the realization of $\Theta\left( \Sigma^{k}\right) $ as $cone\left( \Sigma^{k}\right) \times\left[ 0,1\right] $ and choose $x_{0}\in\Sigma^{k}\times\left\{ \frac{1}{2}\right\} $. Choose nice $k$-cell neighborhoods $B_{1}^{k}\subseteq B_{0}^{k}$ of $x_{0}$ in $\Sigma^{k}% \times\left\{ \frac{1}{2}\right\} $ and, by a mild abuse of notation, let $I_{x_{0}}$ denote the corresponding cone line as a subset of $cone(\Sigma ^{k})\times\left\{ \frac{1}{2}\right\} $. By a combination of Lemmas \ref{cone-to-cylinder} and \ref{cylinder-product}, \[ \left( cone\left( \Sigma^{k}\right) \times\left[ 0,1\right] ,I_{x_{0}% }\right) \approx\left( Map\left( F\right) ,I_{x_{0}}\right) \] for some map \[ F:((\Sigma^{k}-int(B_{1}^{k}))\times\lbrack0,1])\cup\left( cone\left( \Sigma^{k}\right) \times\left\{ 0,1\right\} \right) \rightarrow I_{x_{0}% }\text{.}% \] Crushing out the range end of the mapping cylinder yields the cone over its domain. Hence, \begin{equation} cone\left( \Sigma^{k}\right) \times\left[ 0,1\right] /I_{x}\approx cone\left( ((\Sigma^{k}-int(B^{k}))\times\left[ 0,1\right] )\cup\left( cone\left( \Sigma^{k}\right) \times\left\{ 0,1\right\} \right) \right) \text{.} \tag{\ddag}% \end{equation} Note that, as a subspace of the manifold with boundary $cone\left( \Sigma ^{k}\right) \times\left( 0,1\right) $, $I_{x_{0}}$ is a tame arc intersecting the boundary in a single end point. (Local tameness for this arc is obvious at all except the interior end point of $I_{x}$. This can be seen from the nice tubular neighborhood provided by the obvious structure of $cone\left( \Sigma^{k}\right) \times\left[ 0,1\right] $. But, since our manifold has dimension $\geq4$, $I_{x_{0}}$ cannot have a wild set consisting of a single point (see \cite{CE} or \cite[Th.3.2.1]{Ru}). Thus $I_{x_{0}}$ is tame in $cone\left( \Sigma^{k}\right) \times\left( 0,1\right) $.) The tameness of $I_{x_{0}}$ in $cone\left( \Sigma^{k}\right) \times\left[ 0,1\right] $ ensures that \[ cone\left( \Sigma^{k}\right) \times\left[ 0,1\right] /I_{x_{0}}\approx cone\left( \Sigma^{k}\right) \times\left[ 0,1\right] \text{.}% \] This homeomorphism can to chosen to take the equivalence class of $I_{x_{0}}$ to $x_{0}$. Combining this homeomorphism with (\ddag) induces the necessary cone structure on $cone\left( \Sigma^{k}\right) \times\left[ 0,1\right] $ with $x_{0}$ corresponding to the cone point and \[ L_{x_{0}}=cone\left( ((\Sigma^{k}-int(B^{k})\times\left[ 0,1\right] )\cup\left( cone\left( \Sigma^{k}\right) \times\left\{ 0,1\right\} \right) \right) \text{.}% \] \textbf{Note. }By the equivalence of tame $\left( k+1\right) $-cells in a $\left( k+1\right) $-manifold, $L_{x_{0}}$ may be more easily visualized as the complement of any tame open $\left( k+1\right) $-cell neighborhood of $x_{0}$ in the manifold portion of $\partial\left( \Theta\left( \Sigma ^{k}\right) \right) $. \end{proof} \end{theorem} Lastly, we show that the $4$-dimensional version of the absolute cone conjecture is equivalent to the $3$-dimensional Poincar\'{e} Conjecture. \begin{theorem} Every $4$-dimensional absolute cone is the cone over a homotopy $3$-sphere; moreover, if $H^{3}$ is a $3$-manifold homotopy equivalent to $S^{3}$, then $cone\left( H^{3}\right) $ is an absolute cone. If $H^{3}$ is not homeomorphic to $S^{3}$, then $cone\left( H^{3}\right) $ is not a $4$-cell. \begin{proof} We have already shown that, if $X$ is a $4$-dimensional absolute cone, then $\partial X$ is a $3$-manifold homotopy equivalent to $S^{3}$; thus $X$ is homeomorphic to the cone over a homotopy $3$-sphere. The last statement of the theorem is obvious. Therefore, it remains only to show that $cone\left( H^{3}\right) $ is, indeed, an absolute cone. Our proof utilizes Freedman's breakthrough work on $4$-dimensional manifolds. Aside from that application, the proof is just a simpler version of work we have already done.\bigskip \textsc{Claim. }$cone\left( H^{3}\right) $ \emph{is an absolute cone.} By \cite[Cor.1.3]{Fr}, $H^{3}\times\mathbb{R\cong S}^{3}\times\mathbb{R}$. It follows that $cone\left( H^{3}\right) $ is a $4$-manifold with boundary---that boundary being the base $H^{3}$. By a homogeneity argument similar to one used earlier, any point of the interior, $opencone\left( H^{3}\right) $, may be realized as a cone point with $H^{3}$ as its link. If $x_{0}\in H^{3}$, use Lemma \ref{cone-to-cylinder} to realize $cone\left( H^{3}\right) $ as a mapping cylinder with $I_{x_{0}}$ corresponding to the `range end'. As before, this arc is tame in the manifold $cone\left( H^{3}\right) $. Therefore, $cone\left( H^{3}\right) /I_{x_{0}}\approx cone\left( H^{3}\right) $; moreover, $cone\left( H^{3}\right) /I_{x_{0}}$ is also homeomorphic to a cone with the equivalence class of $I_{x_{0}}$ as its cone point (and the complement of an open $3$-cell as its base). These homeomorphisms yield a cone structure on $cone\left( H^{3}\right) $ with $x_{0}$ as the cone point. \end{proof} \end{theorem}
{ "timestamp": "2005-07-19T18:18:44", "yymm": "0507", "arxiv_id": "math/0507395", "language": "en", "url": "https://arxiv.org/abs/math/0507395", "abstract": "A compactum X is an `absolute cone' if, for each of its points x, the space X is homeomorphic to a cone with x corresponding to the cone point. In 1971, J. de Groot conjectured that each n-dimensional absolute cone is an n-cell. In this paper, we give a complete solution to that conjecture. In particular, we show that the conjecture is true for n<4 and false for n>4. For n=4, the absolute cone conjecture is true if and only if the 3-dimensional Poincare Conjecture is true.", "subjects": "Geometric Topology (math.GT)", "title": "A solution to de Groot's absolute cone conjecture", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9820137931962462, "lm_q2_score": 0.8333246015211008, "lm_q1q2_score": 0.8183362529034864 }
https://arxiv.org/abs/0811.2449
Points in a triangle forcing small triangles
An old theorem of Alexander Soifer's is the following: Given five points in a triangle of unit area, there must exist some three of them which form a triangle of area 1/4 or less. It is easy to check that this is not true if "five" is replaced by "four", but can the theorem be improved in any other way? We discuss in this article two different extensions of the original result.First, we allow the value of "small", 1/4, to vary. In particular, our main result is to show that given five points in a triangle of unit area, then there must exist some three of them determining a triangle of area 6/25 or less.Second, we put bounds on the minimum number of small triangles determined by n points in a triangle, and make a conjecture about the asymptotic right answer as n tends to infinity.
\section{Introduction} This article is in honor of Alexander ``Sasha'' Soifer's sixtieth birthday. I have known Sasha for about twenty years now; I first met him when I competed in his Colorado Math Olympiad in 1988. I remember that he teased me about looking even younger than I was (and also about my poor grades in school), but when we talked about mathematics he always spoke to me as his peer. We could speak a common language and share our curiosity and passion for solving problems. Professor Soifer has solved many outstanding math problems, but posed many more. He has composed most of the problems for twenty-five (and counting) Colorado Math Olympiads, and written numerous research articles and five books about open problems in combinatorics and geometry. Perhaps the most important thing that I learned from him is that we are free to ask our own mathematical questions and pursue them as far as we are able. This was a very empowering idea to me when I was young. Sasha has nice taste in problems too, and tells a good mathematical story \cite{Coloring}; I have sometimes felt a little haunted by problems which I first heard from him. One problem I have always remembered for its simple and elegant statement (and somewhat tricky proofs) is the following, which appeared on the Colorado Math Olympiad in 1988 \cite{Olympiad}. It stayed with me, and I have occasionally sat down trying to see a new solution to it. \begin{theorem} \label{soif} Given five points in a triangle of unit area, there must exist some three of them forming a triangle of area $1/4$ or less. \end{theorem} If the reader has never seen this problem, they are warmly invited to try it. The problem becomes easier if ``five'' is replaced by ``six'' or ``seven,'' but it seems to take some work to get ``five''. Solutions by Alexander Soifer, Royce Peng, and Cecil Rosseau all appear in the book, ``How does one cut a triangle?'' which will soon be reprinted by Springer \cite{HowDoes}. One wonders if this is the best possible result, and in a sense it is: it is easy to check that ``five'' can not be replaced by ``four''. But what about the $1/4$? I must admit that I did not make any real progress toward proving Theorem \ref{soif} during the Olympiad, but now I will have a (very) small revenge on the problem by improving $1/4=0.25$ to $6/25=0.24$. In the spirit of Sasha Soifer and his wonderful journal Geombinatorics, we leave the reader with some open problems as well. \section{Varying the definition of small} Our main result is the following. \begin{theorem} Given five points in a triangle of unit area, then some three of them form a triangle of area less than or equal to $6/25$. \end{theorem} Since ratios of areas are preserved under affine transformation, we assume without loss of generality that our unit triangle $T$ is equilateral. Let $L$ denote the length of one side of $T$ (i.e. $L=2/\sqrt[4]{3}$). We partition $T$ into $100$ congruent equilateral triangles of side length $L/10$, as in Figure \ref{fig:hund}. Label the five points by $P=\{p_1, p_2, \ldots, p_5\}$. \begin{figure} \includegraphics[width=2.25in]{hundred.eps} \caption{Partitioning $T$ into $100$ congruent triangles.} \label{fig:hund} \end{figure} { \bf We use the word ``small'' to refer to triangles of area $\le 6/25$ for the rest of the section.} \begin{lemma}\label{dist4} If for any pair of points in $p_i, p_j \in P$, $d(p_i, p_j) \le 4L/10$ then three of the points in $S$ form a small triangle. (In particular, if $d(p_1,p_2) < 4L / 10$ then either $\{ p_1, p_2, p_k \}$ forms a small triangle for some $k \in \{3, 4,5\}$, or else $\{p_3, p_4, p_5\}$ forms a small triangle.) \end{lemma} \begin{proof}[Proof of Lemma \ref{dist4}] For any three points $x,y,z$ let $A(x,y,z)$ denote the area of the triangle they form. The locus of points $x$ such that $A(p_i,p_j,x) \le 24/100$ is a closed infinite strip $S$ centered on line $p_ip_j$, and of width at least $$2 ( \sqrt{3} /2)(6L/10)=(3\sqrt{3}/5)L > L.$$ Since the diameter of $T$ is $L$, the intersection $I$ of the complement of the strip with $T$, has at most one component. The main point is that if a third point is inside $S$ we are done, and otherwise all three points lie in $I$, and we need only check that this forces the area of the triangle they form to be less than $24/100$. By moving $p_i$ and $p_j$ along line $p_ip_j$ until they intersect the boundary of $T$ if necessary, we only increase the width of the strip $S$, so we may assume without loss of generality that $p_i$ and $p_j$ are both on the boundary of $T$. If $p_i$ and $p_j$ are on the same side of $T$ then $I$ is a triangle with maximal area when $d(p_i,p_j)=4L/10$, and this area is only $(4/10)^2=16/100$, in which case we are done, so we may assume that $p_i$ and $p_j$ are on two different sides of $T$. Now suppose that $d(p_i,p_j) < 4L / 10$. Then we can replace $p_i$ and $p_j$ by $p_i'$ and $p_j'$, respectively, so that the line $p_i'p_j'$ is parallel to $p_ip_j$, and $d(p_i', p_j') = 4L / 10$. This results in a new strip $S'$ and denote the intersection of $S'$ with $T$ by $I'$. We claim that $I'$ strictly contains $I$, as follows. Suppose that the distance between line $p_i p_j$ and $p_i' p_j'$ is $\Delta$. The strip $S$ is parallel to the strip $S'$, and the width $w'$ of strip $S'$ is related to the width $w$ of $S$ by $$(4L/10)w'=d(p_i,p_j) w.$$ We also have by basic trigonometry that $$4L/10 - d(p_i,p_j) = \Delta \left( \frac{1}{\tan{A}} + \frac{1}{\tan{2 \pi /3 - A}} \right),$$ where $A$ is the angle that line $p_ip_j$ makes with the sides of $T$, at $p_j$, as in Figure \ref{fig:seglemma}. \begin{figure} \includegraphics[width=4in]{seglemma.eps} \caption{Proof of Lemma \ref{dist4}.} \label{fig:seglemma} \end{figure} The minimum of the function $f(x) = 1 / \tan(x) + 1 / \tan(2\pi/3 - x)$ on the interval $x \in [0, 2 \pi /3 ]$ is attained when $x= \pi/3$ and $f(x)= 2/ \sqrt{3} > 1$. Now \begin{eqnarray*} w-w' & = & w \left( 1- \frac{d(p_i,p_j) }{4L/10} \right) \\ &= & w \left( \frac{4L/10 - d(p_i,p_j)}{4L/10} \right) \\ & \ge & \frac{w}{4L/10}(2/\sqrt{3})\Delta\\ & \ge & \Delta,\\ \end{eqnarray*} since $w \ge (3\sqrt{3} /10) L > 4L/10$. This implies that $I'$ contains $I$, and we may assume that $d(p_i,p_j) = 4L /10$. Now label the corners of $T$ by $p$, $q$, and $r$, so that $p_i$ lies on side $pr$ and $p_j$ on side $pq$, as in Figure \ref{fig:seglemma}. If $d(p,p_i)$ and $d(p,p_j)$ are both greater than $4L/10$ then $d(p_i,p_j) > 4L/10$, and if they are both less than $4L/10$ then $d(p_i,p_j) < 4L/10$, but we are assuming that $d(p_i,p_j)=4L/10$. So by symmetry, we assume without loss of generality that $d(p,p_i) \le 4L / 10$ and $d(p,p_j) \ge 4L / 10$. In particular that $0 \le A \le \pi /3$. Let $q_i$ and $q_j$ denote the other two intersections of the boundary of $S$ with the boundary of $T$. Note that line $q_iq_j$ intersects side $qr$, and in particular $I$ is a triangle. Let $a=d(p,p_i)$. Then by the Law of Sines, $$a= \sin{A(4L/10)}{\sqrt{3}/2} =\frac{4 \sin{A}}{5\sqrt{3}} L.$$ Denote $b=d(p_i,q_i)$ and $c=d(q_i,r)$, so in particular $a+b+c=L$. We compute $b$ in terms of $A$ by $$b=\frac{3\sqrt{3}/10}{\cos{(A-\pi/6)}},$$ so $$c= L \left( 1-\frac{4}{5\sqrt{3}} \sin{A} - \frac{3\sqrt{3}}{10\cos{(A-\pi/6)}} \right).$$ The angles of triangle $q_iq_jq$ are $\pi/3$, $\pi /3 -A$, and $\pi/3 +A$, and applying the Law of Sines and a familiar formula for area of a triangle, its area is given by \begin{eqnarray*} |I| &= &\frac{1}{2} c^2 \frac{\sin{(\pi/3)}\sin{(\pi/3+A)}}{\sin{(\pi/3-A)}}.\\ \end{eqnarray*} It seems that this is probably a monotone decreasing function of $A$ on the interval $A \in [0,2\pi/3]$ , but we did not care to take derivatives of this function to prove it. So we instead take a slightly more indirect approach, and one that does not involve calculus. The idea is to start with the fact that $c$ cannot be too large, which forces $A$ large, which in turn forces $c$ even smaller, etc., and we quickly reach a contradiction to the supposition that $|I| \ge 24/100$. Since $0 \le A \le 2\pi /3$, we have $c \le 4L/10$, so $c^2 \le 16L^2/100$. Then if $|I| \ge 24/100$, $$\frac{\sin{(\pi/3)}\sin{(\pi/3+A)}}{2\sin{(\pi/3-A)}} \ge \frac{3}{2L^2},$$ so $$\frac{\sin{(\pi/3+A)}}{\sin{(\pi/3-A)}} \ge \frac{3}{2}.$$ Now $\sin{(\pi/3 + A )} \le 1$, so this implies in particular that $\sin{(\pi/3 -A)} < 2/3$, and then since $0 \le A \le \pi /3$ this gives that $0.317 < A$. Plugging back in to the formula for $c$ gives \begin{eqnarray*} c & = & L \left( 1-\frac{4}{5\sqrt{3}} \sin{A} - \frac{3\sqrt{3}}{10\cos{(A-\pi/6)}} \right) \\ & \le & L \left( 1-\frac{4}{5\sqrt{3}} \sin{A} - \frac{3\sqrt{3}}{10} \right) \\ & \le & L \left( 1-\frac{4}{5\sqrt{3}} \sin{0.317} - \frac{3\sqrt{3}}{10} \right) \\ & < & 0.337 L \\ \end{eqnarray*} We repeat the argument from before. If $c < 0.337L$ and $|I| \ge 24/100$, then $\sin{(\pi/3 -A)} \le c^2/0.24 < 0.474$, so $A > 0.553$. This gives in turn that \begin{eqnarray*} c & \le & L \left( 1-\frac{4}{5\sqrt{3}} \sin{0.553} - \frac{3\sqrt{3}}{10} \right) \\ & < & (24/100)L.\\ \end{eqnarray*} a contradiction to the assumption that $|I| > 24/ 100$, since $$|I| = c d(r,q_j) / L^2 < c / L.$$ \end{proof} Then we immediately have the following. \begin{lemma}\label{trihex} If any two points of $P$ lie in a triangle of side length $4L/10$, or a hexagon of diameter $4L/10$, then some three points form a small triangle. \end{lemma} These are many such triangles and hexagons suggested by our partition of $T$, as in Figure \ref{fig:trihex}. \begin{figure} \includegraphics[width=2in]{4tri.eps} \includegraphics[width=2in]{hex.eps} \caption{A triangle and hexagon of diameter $4L/10$.} \label{fig:trihex} \end{figure} We further restrict the possible arrangements of points by using the fact that a triangle in a parallelogram of area $A$ has area at most $A/2$. (This observation is used in Soifer's original proof of Theorem \ref{soif}.) \begin{lemma} \label{parallel} If any three points of $P$ lie in a parallelogram of area $A \le 48/100$, then they form a triangle of area no more than $A/2 \le 24/100$. \end{lemma} We make use of three such kinds of parallelograms, illustrated in Figure \ref{fig:parallel}, which we will call $(2,10)-$, $(3,8)-$, and $(4,6)-$ parallelograms. \begin{figure} \includegraphics[width=1.5in]{parallel210.eps} \includegraphics[width=1.5in]{parallel38.eps} \includegraphics[width=1.5in]{parallel46.eps} \caption{Parallelograms of area $A\le48/100$. From left to right, $(2,10)-$, $(3,8)-$, and $(4,6)-$ parallelograms.} \label{fig:parallel} \end{figure} We consider three cases. Case I is that there is a point of $P$ inside the central shaded triangle in Figure \ref{fig:case1}. \begin{figure} \includegraphics[width=2in]{case1.eps} \caption{Case I: At least one point is in the shaded triangle.} \label{fig:case1} \end{figure} Assuming Case I, by the hexagon case of Lemma \ref{trihex}, there can be no other points in the wider shaded region in Figure \ref{fig:wide1}, or else we are already done. But that leaves at least four points in the three corner white triangles. Two of them must lie in the same triangle, but then by the triangle case of Lemma \ref{trihex} we are done. \begin{figure} \includegraphics[width=2in]{wide1.eps} \caption{Case I: There can be no more than the given point in the greater shaded region. Hence four points lie in the three white triangles.} \label{fig:wide1} \end{figure} Case II is that there is a point $p_1$ in the shaded region in Figure \ref{fig:case2}. By symmetry, assume that $p_1$ is in the upper triangle. There are no other points in the wider shaded region of Figure \ref{fig:wide2} or we are done, by the hexagon case of Lemma \ref{trihex}. There is at most one point in the top white triangle, and so there are at least three points in the lower white region. The case of no points in the top triangle and four in the bottom is easy to dispense with by using Lemma \ref{trihex}. So assume that there is one point, $p_2$, in the top triangle. \begin{figure} \includegraphics[width=2in]{case2.eps} \caption{Case II: At least one point $p_1$ lies in one of the three shaded triangles. Assume that $p_1$ in the top triangle.} \label{fig:case2} \end{figure} Now $p_3, p_4,$ and $p_5$ are all in the lower white region of Figure \ref{fig:case2}. They can not all be in the lower two strips, or we are done by the parallelogram lemma with a $(2,10)$-parallelogram, so there is at least one point $p_3$ on the third or fourth row, by symmetry assume it is in the triangle on the left, as in Figure \ref{fig:case22}. \begin{figure} \includegraphics[width=2in]{wide2.eps} \caption{Case II: There can be no more than the given point in the greater shaded region, and there is at most one point in the top white triangle. So at least three points lie in the white region at the bottom.} \label{fig:wide2} \end{figure} \begin{figure} \includegraphics[width=2in]{wide22.eps} \caption{Case II: $p_1$, $p_2$, and $p_3$ are in the three shaded triangles: $p_1$ in the center, $p_2$ at the top, and $p_3$ on the bottom-left.} \label{fig:case22} \end{figure} By Lemma \ref{parallel} with a $(3,9)$-parallelogram along the left edge of $T$ and containing points $p_2$ and $p_3$, $p_1$ must be in the bottom right corner triangle of side length $L/10$, as in Figure \ref{fig:case24}. Then applying Lemma \ref{parallel} with $(3,8)-$ and $(4,6)$-parallelograms gives that $p_4$ and $p_5$ are in the white region at the bottom of Figure \ref{fig:case24}. \begin{figure} \includegraphics[width=2in]{wide23.eps} \caption{Case II: Then by the parallelogram lemma, with a $(3,9)$-parallelogram, point $p_1$ is in the bottom right corner of the center triangle, or we are done.} \label{fig:case23} \end{figure} \begin{figure} \includegraphics[width=2in]{wide24.eps} \caption{Case II: Applying Lemma \ref{parallel} with $(3,8)-$ and $(4,6)-$ parallelograms gives than $p_4$ and $p_5$ are in the white regions at the bottom of $T$. Then $p_3$, $p_4$, and $p_5$ form a small triangle.} \label{fig:case24} \end{figure} We claim that then $p_3$, $p_4$, and $p_5$ form a triangle of area $24/100$ or less. Given any particular location of $p_4$ and $p_5$, triangle $p_3p_4p_5$ attains its maximum value (given $p_4$ and $p_5$) with $p_3$ at a corner of its boundary triangle. If $p_3$ is one of the bottom two corners, then the three points are in a $(2,10)$-parallelogram, and if it is the top corner, they are in a $(3,8)$-parallelogram. In either case, the area of the triangle is less than $24/100$ by Lemma \ref{parallel}. Then Case III is the only possibility left, that all five points contained in the outer shaded region in Figure \ref{fig:case3}. \begin{figure} \includegraphics[width=2in]{case3.eps} \caption{Case III: All five points lie in the shaded area.} \label{fig:case3} \end{figure} We break the outer strip into three diamond and trapezoid shaped regions, Figure \ref{fig:case32}. \begin{figure} \includegraphics[width=2in]{case32.eps} \caption{Case III: We partition the outer strip into six smaller regions, diamonds and trapezoids. Lemma \ref{trihex} places severe restrictions on where the points may lie, if they are to form no small triangles. In fact without loss of generality we assume that $p_1$ is in the diamond on top, $p_2$ and $p_3$ in the trapezoids on the left and right, and $p_4$ and $p_5$ in the bottom trapezoid.} \label{fig:case32} \end{figure} Any pair of the trapezoids can be covered with a hexagon and two triangles of diameter $4L/10$, so by Lemma \ref{trihex}, they contain a total of at most three points from $S$. It is quick to deduce from this, that the points are distributed $2-1-1$ in the trapezoids, with one point in the diamond opposite the trapezoid with two points, or we are done. By symmetry we assume $p_1$ is in the diamond on top, $p_2$ and $p_3$ are in the trapezoids on the left and right, respectively, and $p_4$ and $p_5$ are in the bottom trapezoid. By $(3,8)$-parallelograms at the top vertex of $T$ and going down the left and right sides, $p_4$ and $p_5$ are in an even smaller trapezoid, as seen in Figure \ref{fig:case33}, but then we are done by Lemma \ref{trihex}.\\ \begin{figure} \includegraphics[width=2in]{case33.eps} \caption{Case III: $(3,8)$-parallelograms force $p_4$ and $p_5$ into in an even smaller trapezoid, shaded at bottom, but then we are done by Lemma \ref{trihex}.} \label{fig:case33} \end{figure} This completes the proof of the main result, but begs the question: \begin{question} What is the minimum $s$ of all $\sigma$ such that among every five points in a triangle of unit area, some three of them form a triangle of area less than or equal to $\sigma$? (Note that $s$ is actually a minimum and not just an infinium, by compactness. \cite{Etudes}) \end{question} \label{inf} By Figure \ref{fig:s16}, we have $s \ge 1/6$. This might be best possible. \begin{figure} \includegraphics[width=2.5in]{s16.eps} \caption{Five points in a unit area triangle, with none of the triangles having area less than $1/6$. Point $a$ is at the corner, $b$ and $e$ are on midpoints, and $c$ and $d$ partition the bottom side into three equal segments. Note that triangles $abc$, $ade$, $bcd$, and $cde$ all have area equal to $1/6$.} \label{fig:s16} \end{figure} \begin{conjecture} $s=1/6$. \end{conjecture} \section{Many small triangles} {\bf In this section we return to using ``small'' to refer to triangles of area $\le 1/4$.} There is another way in which Theorem \ref{soif} might be improved, besides varying the $1/4$. Is it possible that given five points in a triangle of unit area, they must form not only one small triangle, but more than one? We conjecture that this is the case; in particular we believe the following holds. \begin{conjecture} Given five points in a triangle of unit area, then they form at least {\it three} small triangles. \end{conjecture} This is not true with ``three'' replaced by ``four'', by the arrangement in Figure \ref{fig:3small}. If the conjecture holds then the following corollary would follow, by averaging: Given $n \ge 5$ points in a triangle of unit area, there they form at least $$ \frac{3}{{5 \choose 2}} {n \choose 3} = \frac{3}{10} {n \choose 3}$$ small triangles. This leads us to the following question. \begin{figure} \includegraphics[width=2in]{3small.eps} \caption{There are only three small triangles.} \label{fig:3small} \end{figure} \begin{question} What is the supremum $A$ of all $\alpha \in [0,1]$ such that given (sufficiently large) $n$ points in a triangle of unit area, they form at least $$ \alpha {n \choose 3}$$ small triangles? \end{question} We observe that $A \le 5/8$, by the construction in Figure \ref{fig:58small}, and conjecture that asymptotically this arrangement gives the smallest number of small triangles. \begin{figure} \includegraphics[width=2in]{58small.eps} \caption{If there approximately $n/4$ points in each of the small circles, then there are $4 (n/4)^3 = n^3 / 16 \approx (3/8) {n \choose 3}$ big triangles, hence approximately $(5/8){n \choose 3}$ small triangles. } \label{fig:58small} \end{figure} \begin{conjecture} $A=5/8$. \end{conjecture}
{ "timestamp": "2008-11-15T01:28:08", "yymm": "0811", "arxiv_id": "0811.2449", "language": "en", "url": "https://arxiv.org/abs/0811.2449", "abstract": "An old theorem of Alexander Soifer's is the following: Given five points in a triangle of unit area, there must exist some three of them which form a triangle of area 1/4 or less. It is easy to check that this is not true if \"five\" is replaced by \"four\", but can the theorem be improved in any other way? We discuss in this article two different extensions of the original result.First, we allow the value of \"small\", 1/4, to vary. In particular, our main result is to show that given five points in a triangle of unit area, then there must exist some three of them determining a triangle of area 6/25 or less.Second, we put bounds on the minimum number of small triangles determined by n points in a triangle, and make a conjecture about the asymptotic right answer as n tends to infinity.", "subjects": "Combinatorics (math.CO); Metric Geometry (math.MG)", "title": "Points in a triangle forcing small triangles", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9871787879966232, "lm_q2_score": 0.828938806208442, "lm_q1q2_score": 0.8183108060362174 }
https://arxiv.org/abs/math/9802046
Economical numbers
A number $n$ is said to be economical if the prime power factorisation of $n$ can be written with no more digits than $n$ itself. We show that under a plausible hypothesis, related to the twin prime conjecture, there are arbitrarily long sequences of consecutive economial numbers, and exhibit such a sequence of length 9.
\section{Introduction} In \cite{San:equidigit}, Bernardo Recam\'an Santos defined a number $n$ to be {\it equidigital}\/\footnote{We prefer this spelling to ``equadigital''.} if the prime power factorisation of $n$ requires the same number of decimal digits as $n$, and {\it economical} if the prime power factorisation requires no more digits. He asked whether there were arbitrarily long sequences of consecutive economical numbers. In \cite{Hes:equidigit}, Richard Hess observed that the five consecutive integers $13$, $14 = 2 \cdot 7$, $15 = 3 \cdot 5$, $16 = 2^4$ and $17$ are economical. We note that the sequence $157$, $158 = 2 \cdot 79$, $159 = 3 \cdot 53$, $160 = 2^5 \cdot 5$, $161 = 7 \cdot 23$, $162 = 2 \cdot 3^4$, $163$ is of length seven. Hess also suggested that there was a longest such string. In this article, we shall show that there are strings of consecutive economical numbers of arbitrary length if Dickson's conjecture, on the simultaneous primality of linear forms, is true. This is a generous assumption, since the conjecture includes the twin prime conjecture as a special case. We use the idea of the proof to find a sequence of nine consecutive economical numbers from 1034429177995381247 to 10344291779953812455. \section{Economical and frugal numbers} There is no need to restrict to decimal expansions, so we begin by working with a general base $B$. Let $\delta(n)$ denote the number of digits required to write $n$ in base $B$, so that $\delta(n) = k$ if and only if $B^{k-1} \le n < B^k$. We can express this conveniently as $\delta(n) = \xceil[\log_B n]$, where\footnote{Note that this differs from the usual $\ceil[x]$ when $x$ is an integer: indeed, $\xceil[x] = \floor[x]+1$.} $\xceil[x] = \min\brace{ k \in {\mathbb Z} : k > x }$. We let $\phi(n)$ denote the number of digits required to write down the prime power factorisation of $n$, so that if $n = \prod_{i=1}^r p_i^{a_i}$ then $\phi(n) = \sum_{i=1}^r \delta(p_i) + \delta'(a_i)$, where $\delta'(a) = \delta(a)$ for $a > 1$ and $\delta'(1) = 0$. We let $h(n) = \delta(n) - \phi(n)$. We define $n$ to be {\it equidigital (base $B$)} if $h(n) = 0$, {\it economical (base $B$)} if $h(n) \ge 0$ and {\it frugal (base $B$)} if $h(n) > 0$. For $k > 0$, if $h(n) \ge k$ we say that $n$ is {\it $k$--frugal}, so that 1-frugal is the same as frugal. We include $1$ as a frugal number. We begin with some obvious properties. \begin{proposition}\label{prop:1} \begin{enumerate} \item A prime number is equidigital in every base; \item $\max\brace{\delta(m), \delta(n)} \le \delta(m+n) \le 1 + \max\brace{\delta(m), \delta(n)}$; \item $\delta(m) + \delta(n) - 1 \le \delta(mn) \le \delta(m) + \delta(n)$; \item $b \delta(n) - b \le \delta\paren(n^b) \le b \delta(n)$; \item If $m$ and $n$ are coprime, then $\phi(mn) = \phi(m) + \phi(n)$; \item $\phi(mn) \le \phi(m) + \phi(n)$; \item $\phi(n^b) \le \phi(n) + \delta(b) \log_2 n$; \item $h(mn) \le h(m) + h(n) - 1$; \item If $m$ is frugal and $n$ is economical, then $mn$ is economical; \item If $m$ is frugal then $h(mn) \le h(n)$. \end{enumerate} \end{proposition} \begin{proof} Part \therosteritem1 follows immediately from the definition. For part \therosteritem2, suppose that $m \le n$ and $\delta(n) = k$. Then $B^{k-1} < m+n \le 2n \le 2 B^k \le B^{k+1}$. For part \therosteritem3, we note that if $\delta(m) = l$ and $\delta(n) = k$ then $B^{l-1} \le m < B^l$ and $B^{k-1} \le n < B^k$, so that $B^{k+l-2} \le mn < B^{k+l}$. Part \therosteritem4 follows by induction. For part \therosteritem5, we note that if $m$ and $n$ are coprime, then their prime power factorisation are disjoint. For part \therosteritem6, we need to consider only the prime power factors that $m$ and $n$ have in common. If $p^a$ occurs in $m$ and $p^b$ in $n$, then $p^{a+b}$ occurs in $mn$. Assume that $a \ge b$. We claim that $\delta'(a+b) + \delta(p) \le \delta'(a) + \delta'(b) + 2\delta(p)$, that is, $\delta'(a+b) \le \delta'(a) + \delta'(b) + \delta(p)$. If $a = b = 1$, then $\delta'(a+b) = \delta'(2) = \delta(2) \le \delta(p)$. If $a > b = 1$, then $\delta'(a+b) = \delta'(a+1) = \delta(a+1) \le \delta(a) + 1$ by \therosteritem2. If $a \ge b > 1$ then $\delta'(a+b) = \delta(a+b) \le \delta(a) + \delta(b)$, again by \therosteritem2. For part \therosteritem7, let $n = \prod_{i=1}^r p_i^{a_i}$. Assume $b > 1$. Then $\phi(n^b) = \sum_{i=1}^r \delta\paren(p_i) + \delta\paren(a_i b)$ and using \therosteritem3 this is at most $\sum_{i=1}^r \delta\paren(p_i) + \delta(b) + \delta'\paren(a_i) = \phi(n) + r \delta(b)$. Now $r \le \log_2 n$, so $\phi(n^b) \le \phi(n) + \delta(b) \log_2 n$. We obtain \therosteritem8 by combining \therosteritem 3 and \therosteritem6: parts \therosteritem9 and \therosteritem{10} are special cases of \therosteritem8. \end{proof} We use the following proposition to prove the result we shall use in the next section. \begin{proposition}\label{prop:2} Fix a base $B$. \begin{enumerate} \item Suppose that $r$ and $s$ are coprime. For each $k \ge 0$ there are infinitely many $k$-frugal numbers coprime to $s$ and divisible exactly by the prime powers in $r$. \item Given $r$, for each $k \ge 0$ there are infinitely many $k$-frugal numbers divisible by $r$ and containing only the same prime factors as $r$. \end{enumerate} \end{proposition} \begin{proof} For \therosteritem1, take $p$ to be a prime greater than $rs$ and let $n = r p^a$. Assume $a \ge 2$. We have $\phi(n) = \phi(r) + \delta(a) + \delta(p)$ and $\delta(n) \ge \delta(r) + \delta\paren(p^a) \ge \delta(r) + a(\delta(p) - 1)$. So \begin{eqnarray*} h(n) &\ge& \delta(r) + a(\delta(p) - 1) - \phi(r) - \delta(a) - \delta(p) \\ & &\quad\quad= a(\delta(p) - 1) - \delta(a) +h(r) - \delta(p) \end{eqnarray*} and this tends to infinity with $a$. For \therosteritem2, consider $n = r^a$ for some $a > 1$. By Proposition \ref{prop:1} \therosteritem4 and \therosteritem7 we have \begin{eqnarray*} h(n) &=& \delta(r^a) - \phi(r^a) \ge a \delta(r) - a - \paren({\phi(r) + \delta(a) \log_2 r}) \\ &\ge& a(\delta(r)-1) - \delta(a) \log_2 r - \phi(r), \end{eqnarray*} and again this tends to infinity with $a$. \end{proof} Although we shall not need it, we can prove the converse of Proposition \ref{prop:1} \therosteritem{10}. We use the following result of Baker and Harman \cite{BH:primegaps}. \begin{proposition}\label{prop:3} For $x$ sufficiently large, there is always a prime between $x$ and $x + x^\theta$, where we can take $\theta = 0.535$. \qed \end{proposition} For our purposes any $\theta < 1$ would suffice. See Ribenboim \cite{Rib:rec},4.II.C for other results in this direction. \begin{proposition}\label{prop:4} Suppose that $m$ has the property that $mn$ is economical whenever $n$ is. Then $m$ is frugal. \end{proposition} \begin{proof} Suppose that $m$ has the property stated, and let $\delta(m) = l$. Then $B^{l-1} \le m < B^l$, so $\frac{B^l}{m} > 1$. Let $k$ be large enough that $(B^k + B^{\theta k}) / B^k = 1 + B^{(\theta-1)k} < B^l / m$: we may also assume that $k > l$ and that $B^k$ is large enough for Proposition \ref{prop:3} to hold. Take a prime $p$ in the interval $B^k < p \le B^k + B^{\theta k}$: by Proposition \ref{prop:1} \therosteritem1, $p$ is economical, so $mp$ is economical by the assumption on $m$. But $mp < m\paren({1 + B^{(\theta-1)k}}) < B^{l+k}$, so $\delta(mp) = \delta(m) + \delta(p) - 1 = k+l$. We have $p > m$, so $p$ is coprime to $m$ and hence $\phi(mp) = \phi(m) + \phi(p) = \phi(m) + \delta(p) = \phi(m) + k+1$. So $h(mp) = \delta(mp) - \phi(mp) = k + l - \phi(m) - k - 1 = l - \phi(m) - 1 = h(m) - 1$. But by our assumptions, $mp$ is economical, so that $h(mp) \ge 0$ and so we must have $h(m) \ge 1$: that is, $m$ is frugal. \end{proof} \section{Extravagant numbers} We can use Proposition \ref{prop:3} to show that there are numbers $n$ with $h(n) \ll 0$: presumably we should call these {\it extravagant}. \begin{proposition}\label{prop:5} Fix a base $B$. For $k > 0$ there are infinitely many numbers with $h(n) \le -k$. \end{proposition} \begin{proof} Choose $t$ large enough that $B^t$ exceeds the bound of Proposition \ref{prop:3}, and also $B^{t(1-\theta)} > 2^\theta (k+2)^2$. Put $L = 1 + \frac{2^\theta}{B^{t(1-\theta)}}$. Then $$ L < 1 + \frac{1}{(k+2)^2} < 1 + \frac{1}{2(k+1)} $$ so $$ L^{k+1} < \paren(1 + \frac{1}{2(k+1)})^{k+1} < \exp\paren({1/2}) < 2 , $$ since $\paren(1 + x/n)^n$ tends to $e^x$ from below for $x > 0$. Consider the intervals $[B^t L^i, B^t L^{i+1}]$, for $i = 0,\ldots,k$. The length of the $i$-th interval is $B^t L^i(L-1)$ and $$ B^t L^i (L-1) \ge B^t (L-1) = 2^\theta B^{t \theta} > L^{k+1} B^{t \theta} \ge L^i B^{t \theta} > (L^i B^t)^\theta , $$ so that Proposition \ref{prop:3} applies and each interval contains a prime, say $p_i$. Further, $B^k L^{k+1} < 2 B^k \le B^{k+1}$. Hence $\delta(p_i) = t+1$ for each $i$. If we put $n = \prod_{i=0}^k p_i$, then $\phi(n) = \sum_{i=0}^k \delta(p_i) = (k+1)(t+1)$. But \begin{eqnarray*} n &<& B^{t(k+1)} L^{(k+1)(k+2)/2} < B^{t(k+1)} \paren(1 + \frac{1}{(k+1)(k+2)})^{(k+1)(k+2)/2} \\ &<& B^{t(k+1)} \exp\paren({1/2}) < B^{t(k+1)+1}, \end{eqnarray*} so that $\delta(n) = t(k+1)+1$. So $h(n) = t(k+1) + 1 - (k+1)(t+1) = -k$. \end{proof} We note that the idea of the proof can be applied to intervals of the form $[B^t L^{-i-1}, B^t L^{-i}]$ to show that there are infinitely many squarefree economical numbers with distinct prime factors, providing an alternative solution to \cite{San:equidigit} in any base. \section{Dickson's conjecture} Dickson \cite{Dic:conjecture} conjectured that a family of linear functions $f_i(n) = a_i n + b_i$, $i = 1,...,t$, with integer coefficients $a_i$, $b_i$, would be simultaneously prime unless they ``obviously cannot'': that is, unless there is an integer $m > 1$ such that $m$ divides the product $f_1(n)f_2(n)\cdots f_t(n)$ for every value of $n$. This rather powerful conjecture would imply the {\it twin primes conjecture}, that there are infinitely many twin prime pairs $(p,p+2)$, and the {\it Sophie Germain primes conjecture}, that there are infinitely primes $p$ for which $2 p+1$ is also prime\footnote{If $p$ is a Sophie Germain prime, then $2p+1$ is a {\it safe} prime.}. Schinzel \cite{SS:hypH} extended Dickson's conjecture to the analogous ``Hypothesis H'' for arbitrary integer polynomials $f_i(x)$. \section{Consecutive economical numbers} \begin{theorem} Fix a base $B$, and $k \ge 0$. If Dickson's conjecture holds, there are arbitrarily long sequences of consecutive $k$-frugal numbers. \end{theorem} \begin{proof} Suppose we wish to find a sequence of $t$ consecutive $k$-frugal numbers, say $N, N+1, \ldots, N+t-1$. Let $p_h$ be the largest prime less than $t$ and for the primes $p_i \le p_h$ let $p_i^{a_i}$ be the largest power of $p_i$ less than $t$. We shall insist that $N$ be divisible by $f_0 = \prod_{i=1}^h p_i^{a_i + 1}$. This implies that the $N+j$ must be divisible by certain powers of the $p_i$: in fact, for $j=1,\ldots,t-1$, if $p_i^{b_{i,j}}$ exactly divides $j$ then the same power of $p_i$ must exactly divide $N+j$. Let $f_j = \prod_i p_i^{b_{i,j}}$. (In fact, $f_j = j$ but we shall not use that.) We have $f_1 = 1$. For each $j = 2,\ldots,t-1$ we use Proposition \ref{prop:2} \therosteritem1 to find $m_j$ such that $h(m_j) \ge k+2 - h(f_j) \ge 1$ and each $m_j$ is coprime to $f_0$ and the other $m_{j'}$. We thus ensure that $m_j f_j$ is $(k+1)$-frugal. Since $f_0$ and the $m_j$ are coprime, we can use the Chinese Remainder Theorem to solve the simultaneous congruences $N+j \equiv 0 \bmod f_j m_j$ to find infinitely many solutions of the form $N \equiv N_0 \bmod M$ where $M = \lcm\brace{f_0, m_j}$. The cofactors $C_j = \frac{N+j}{f_j m_j}$ will satisfy congruences of the form $C_j \equiv S_j \bmod M_j$ where $S_j = \frac{N_0+j}{f_j m_j}$ and $M_j = \frac{M}{f_j m_j}$. By Dickson's conjecture, the forms $C_j = S_j + M_j x$ will be simultaneously prime for infinitely many values for $x$. For such values, we have $h\paren(f_j m_j) \ge k+1$ and $C_j$ economical, so by Proposition \ref{prop:1} \therosteritem8 the values $N+j = f_j m_j C_j$ are all $k$-frugal, as required. \end{proof} \begin{corollary} Fix a base $B$. If Dickson's conjecture holds, there are arbitrarily long sequences of consecutive economical numbers.\qed \end{corollary} In practice there are modifications one could make to the construction given in the proof of the Theorem. For example, it may be more efficient to take the sequence of numbers $N-1,\ldots,N+t-2$, since none of the $p_i$ will divide $N-1$ and we can take $f_{-1} = 1$. A further variant would be to take $m_0$ to be composed of further powers of the primes $p_i$, using Proposition \ref{prop:2} \therosteritem2 rather than \therosteritem1 to ensure that $f_0 m_0$ is frugal: for example, we might take $m_0$ to be a power of $f_0$. \section{Examples} We illustrate with some example. We take base $B = 10$. Suppose we wish to find seven consecutive economical numbers. We have $p_3 = 5$ and $2^2, 3^1, 5^1$ the highest powers less than $7$, so that $f_0 = 2^3 \cdot 3^2 \cdot 5^2 = 1800$. We find that $f_j = j$, $j=1,\ldots,6$. We take $m_0 = 7^6$, $m_1 = 1$, $m_2 = 11^4$, $m_3 = 13^4$, $m_4 = 17^4$, $m_5 = 19^3$, $m_6 = 23^4$. We have $M = 14196220211350791776356766371800$ and $N \equiv N_0 = 5599355285926686611723646146400 \bmod M$. For $N = 133365337188083812598934543492600 = N_0 + 9M$ we have {\small \begin{eqnarray*} 133365337188083812598934543492599 &=& 29 \cdot 598426817561 \cdot 7684823934473500571\\ 133365337188083812598934543492600 &=& 2^{3} \cdot 3^{2} \cdot 5^{2} \cdot 7^{6} \cdot 649567 \cdot 969523340521703729\\ 133365337188083812598934543492601 &=& 133365337188083812598934543492601\\ 133365337188083812598934543492602 &=& 2 \cdot 11^{4} \cdot 89 \cdot 113 \cdot 6701 \cdot 67582485977340653273\\ 133365337188083812598934543492603 &=& 3 \cdot 13^{4} \cdot 170173187 \cdot 832591957 \cdot 10985629799\\ 133365337188083812598934543492604 &=& 2^{2} \cdot 17^{4} \cdot 73 \cdot 16141 \cdot 338792660124002968867\\ 133365337188083812598934543492605 &=& 5 \cdot 19^{3} \cdot 46399 \cdot 188563981 \cdot 444472402866601\\ 133365337188083812598934543492606 &=& 2 \cdot 3 \cdot 23^{5} \cdot 3453444945058703174533907 \end{eqnarray*} } We find that $N-1,\ldots,N+6$ form a sequence of eight consecutive economical numbers: even though we only wanted seven! We note that not all the cofactors are prime, although they are all economical. If we adopt the variant of choosing $m_0$ to be a power of $f_0$, then in this example we would have $m_0 = f_0 = 1800$, $m_2 = 7^4$, $m_3 = 11^4$, $m_4 = 13^4$, $m_5 = 17^3$ and $m_6 = 19^4$. We have $M = 2082775632877914851396520000$ and $N \equiv N_0 = 1625787524296851742054440000 \bmod M$. For $N = 1625787524296851742054440000$ we have the sequence of length seven \begin{eqnarray*} 1625787524296851742054440000&=&2^{6} \cdot 3^{4} \cdot 5^{4} \cdot 501786272931127080881 \\ 1625787524296851742054440001&=&2237 \cdot 726771356413433948169173 \\ 1625787524296851742054440002&=&2 \cdot 7^{4} \cdot 678077 \cdot 499301208591113813 \\ 1625787524296851742054440003&=&3 \cdot 11^{4} \cdot 23 \cdot 37030463 \cdot 43459508766889 \\ 1625787524296851742054440004&=&2^{2} \cdot 13^{4} \cdot 183907 \cdot 77380605917154763 \\ 1625787524296851742054440005&=&5 \cdot 17^{3} \cdot 227 \cdot 65239 \cdot 4469036127622909 \\ 1625787524296851742054440006&=&2 \cdot 3 \cdot 19^{4} \cdot 53 \cdot 6297281 \cdot 1751557 \cdot 3556681 \end{eqnarray*} Again we note that not all the cofactors are prime, although they are all economical. The precise assignment of the powers of the small primes $p_i$ to the number $N+j$ in the proof of the Theorem is quite arbitrary. If we have different elements of the sequence divisible by the highest powers of each $p_i$, it seems that we can do better: \begin{eqnarray*} 1034429177995381247 &=& 51551 \cdot 20066132140897 \\ 1034429177995381248 &=& 2^{9} \cdot 3 \cdot 88651 \cdot 7596716293 \\ 1034429177995381249 &=& 17 \cdot 60848775176198897 \\ 1034429177995381250 &=& 2 \cdot 5^{5} \cdot 76757 \cdot 2156268073 \\ 1034429177995381251 &=& 3^{5} \cdot 19 \cdot 383 \cdot 584981475541 \\ 1034429177995381252 &=& 2^{2} \cdot 7^{5} \cdot 154267 \cdot 99741877 \\ 1034429177995381253 &=& 394007 \cdot 2625408122179 \\ 1034429177995381254 &=& 2 \cdot 3 \cdot 23 \cdot 59^{3} \cdot 1367 \cdot 26699131 \\ 1034429177995381255 &=& 5 \cdot 26264543 \cdot 7877001157 . \end{eqnarray*} \section{Distribution of frugal numbers} Again we take base $B = 10$. We computed $h(n)$ for $2 \le n \le 1000000$. The number of integers up to $5 \cdot 10^8$ which have $h(n) = k$ is {\small $$\begin{array}{rrrrrrrrrrrrr} k =-6 & -5 & -4 & -3 & -2 & -1 \\ 1313 & 195341 & 5101112 & 44435592 & 153988692 & 208380123 \\ \end{array} $$ $$\begin{array}{rrrrrrrrrrrrr} k = 0 & 1 & 2 & 3 & 4 & 5 & 6 \\ 86441875 & 1297001 & 140575 & 16670 & 1483 & 207 & 16 \\ \end{array} $$ } The smallest 6-frugal number is $40353607 = 7^9$: the smallest 6-extravagant number is $8314020 = 2^2 3^2 5 \cdot 11 \cdot 13 \cdot 17 \cdot 19$. We examined consecutive economical numbers up to $10^6$. The longest strings of consecutive economical numbers we found were of length 7, starting at 157, 108749, 109997, 121981, 143421. The longest strings of consecutive frugal numbers were only of length two, the first starting with $4374$. { } \ifx\undefined\leavevmode\hbox to3em{\hrulefill}\, \newcommand{\leavevmode\hbox to3em{\hrulefill}\,}{\leavevmode\hbox to3em{\hrulefill}\,} \fi
{ "timestamp": "1998-02-09T15:56:09", "yymm": "9802", "arxiv_id": "math/9802046", "language": "en", "url": "https://arxiv.org/abs/math/9802046", "abstract": "A number $n$ is said to be economical if the prime power factorisation of $n$ can be written with no more digits than $n$ itself. We show that under a plausible hypothesis, related to the twin prime conjecture, there are arbitrarily long sequences of consecutive economial numbers, and exhibit such a sequence of length 9.", "subjects": "Number Theory (math.NT)", "title": "Economical numbers", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9898303443461076, "lm_q2_score": 0.8267118026095991, "lm_q1q2_score": 0.8183044282520509 }
https://arxiv.org/abs/1902.04055
On the number of pancake stacks requiring four flips to be sorted
Using existing classification results for the 7- and 8-cycles in the pancake graph, we determine the number of permutations that require 4 pancake flips (prefix reversals) to be sorted. A similar characterization of the 8-cycles in the burnt pancake graph, due to the authors, is used to derive a formula for the number of signed permutations requiring 4 (burnt) pancake flips to be sorted. We furthermore provide an analogous characterization of the 9-cycles in the burnt pancake graph. Finally we present numerical evidence that polynomial formulae exist giving the number of signed permutations that require $k$ flips to be sorted, with $5\leq k\leq9$.
\section{Introduction} The idea of sorting permutations is a classical one in combinatorics and computer science. Of particular interest is when the sorting is performed utilizing only prefix reversals (see, for example,~\cite{Chitt,Cibulka, GatesPapa}). Problems like the \emph{pancake problem} that ask to determine the minimum number of prefix-reversal flips that are needed to sort any permutation in $S_n$ are computationally hard to solve~\cite{BulFerRusu}. In this context a natural question arises: How many permutations in $S_n$ require $k$ prefix-reversal flips to be sorted? One of our main contributions is the answer to the case $k=4$ (the cases $k\leq 3$ are trivial) for both permutations and signed permutations. Our methods rely on an existing classification of the 6-,7-, and 8-cycles in the pancake graph and a classification of the 8-cycles in the burnt pancake graph, which is due to the authors~\cite{BBP2018}. In particular, our main results are the following. \begin{enumerate} \item[(I)]\label{main:one} We provide an explicit formula that gives the number of permutations in $S_n$ that require exactly four prefix reversal flips to be sorted. This formula is given by a simple integer-valued polynomial \begin{displaymath}\frac{1}{2} \left(2 n^4 - 15 n^3 + 29 n^2 + 6 n - 34\right),\end{displaymath} where $n\geq4$. The details are given in Section~\ref{sec:sn}. \item[(II)] We give an explicit formula for the number of signed permutations in $B_n$ that require exactly four prefix-reversal flips to be sorted. This formula is also given by a simple integer-valued polynomial, namely, \[ \frac{1}{2} n(n-1)^2(2n-3), \] with $n\geq1$. The details are given in Section~\ref{sec:bn}. \item[(III)] We also provide a classification of all the 9-cycles in the burnt pancake graph. Concretely, we prove that all of these 9-cycles can be described by two canonical forms. The details are given in Section~\ref{sec:9cycles}. \end{enumerate} We point out that the polynomial in (I) can also be derived from the algorithm described in~\cite{HV16}, where the authors utilize structural properties of certain permutations to obtain generating functions. Our methods are entirely elementary and rely on the classification of cycles inside the pancake and burnt pancake graph and on the principle of inclusion-exclusion. Presently, the algorithm in~\cite{HV16} cannot be applied to signed permutations, and in particular, no other proof for our second main result (II) is known. The reader is referred to Section~\ref{sec:preliminaries} for the basic definitions and notation used throughout the paper. We end the paper with conjectures for closed formulas giving the number of signed permutations requiring exactly $k$ prefix-reversal flips to be sorted with $5\leq k\leq 9$, and other conjectures. \section{Preliminaries and notation}\label{sec:preliminaries} Throughout this note, $n$ will denote a positive integer greater than 1, and for a positive integer $k\leq n$, $[k]$ will denote the set $\{1,2,\ldots,k\}$. We will use $S_n$ to denote the set of permutations of the set $[n]$. Furthermore, if $2\leq i\leq n$, we denote by $r_i$ the following permutation, written in one-line notation: \[ r_i= i\,(i-1)\,\cdots\,2\,1\,(i+1)\,\cdots\,n. \] The elements of the set $R:=\{r_i\}_{i=2}^n$ are referred to as \textit{prefix reversals} or \textit{pancake flips}. Similarly one can define prefix reversals on the group of signed permutations. A \textit{signed permutation} is a permutation $w$ of the set $[\pm n]:=\{-n,-(n-1),\ldots,-2,-1,1,2,\ldots n-1,n\}$ satisfying $w(-i)=-w(i)$ for all $i\in[n]$. For convenience, we will use $\underline{i}$ instead of $-i$. We will use \textit{window notation} (see~\cite[Section 8.1]{BjornerBrenti}) to denote signed permutations. More specifically, we will use $[w(1)\,w(2)\,\ldots\,w(n)]$ to denote $w$. In Coxeter groups literature, the group of signed permutations is denoted by $B_n$ (see~\cite[Chapter 8]{BjornerBrenti}). In this context, if $i\in [n]$, a \text{signed prefix reversal} is given by \[ r_i^{B}=[\underline{i}\,\underline{i-1}\,\cdots\,\underline{1}\,(i+1)\,(i+2)\,\cdots\,n]. \] We refer to the elements of the set $R^B:=\{r_i^B\}_{i=1}^n$ as \textit{signed prefix reversals} or \textit{burnt pancake flips}. If clear from the context, we may drop the words ``signed," ``burnt," and the $B$ superscript. It is worth pointing out that for any $i\in[n]$, $r_i$ and $r_i^B$ are both involutions (elements of order two) of $S_n$ and $B_n$, respectively. That is, \[ \left(r_i\right)^2 = 1\, 2\, \cdots\, (i-1)\, i\, (i+1) \cdots n\text{, and}\] \[ \left(r_i^B\right)^2 = [1\, 2\, \cdots\, (i-1)\, i\, (i+1) \cdots n]. \] The graph $P_n:=(S_n,E_n)$ where \[ E_n:= \{\{\pi,\pi r_i\}: \pi\in S_n,2\leq i\leq n\} \] is called the \emph{pancake graph} of order $n$. Similarly, the graph $BP_n:=(B_n,E_n^B)$, where \[ E^B_n:= \{\{\pi,\pi r^B_i\}: \pi\in B_n,i\in[n]\}, \] is called the \textit{burnt pancake graph}. Both $P_n$ and $BP_n$ are Cayley graphs of $S_n$ and $B_n$, respectively, since the groups $S_n$ and $B_n$ are generated by $R$ and $R^B$, respectively. Therefore, both $P_n$ and $BP_n$ are \textit{vertex transitive} graphs; that is, given any two vertices $u,v$ in a vertex transitive graph, there exists a graph isomorphism $f$ such that $v=f(u)$. We will use $P_{n-1}(q)$ ($BP_{n-1}(q)$, respectively) to denote the subgraph of $P_n$ ($BP_n$, respectively) induced by the subset of $S_n$ ($B_n$, respectively) of all permutations that end with $q$, with $q\in[n]$ ($[\pm n]$, respectively) in one-line notation. Furthermore, for $1 < k < n$, we use $P_{k-1}(p)$ ($BP_{k-1}(p)$, respectively) to denote the subgraph of $P_{n-1}(n)$ ($BP_{n-1}(n)$, respectively) whose vertices are the set of all $\pi \in S_{n}$ with \begin{displaymath}\pi = \pi_1\, \pi_2\, \cdots \,\pi_{k-1}\, p \,(k+1)\, (k+2)\, \cdots\, n,\end{displaymath} where $p\in[k]$ and $\pi_i \in [k] \setminus \{p\}$ or, respectively, $\pi \in B_{n}$ with \begin{displaymath}\pi = [\pi_1\, \pi_2\, \cdots \,\pi_{k-1}\, p\, (k+1)\, (k+2)\, \cdots \,n],\end{displaymath} where $p \in [\pm k]$ and $\pi_i \in [\pm k] \setminus \{p\}$. The edges of $P_{k}(p)$ being $\{\{\pi,\pi r_i\}: \pi \in P_{k}(p), 2\leq i \leq k-1\}$ ($\{\{\pi,\pi r_i\}: \pi \in BP_{k}(p), i \in [k-1]\}$, respectively). One readily notices that each $P_k(p)$ ($BP_k(p)$, respectively) is isomorphic to $P_{k-1}$ ($BP_{k-1}$, respectively). A key result that we will use is the following classification of the 6-, 7-, and 8-cycles in $P_n,n\geq4$. We will refer to each cycle $C$ in $P_n$ ($BP_n$, respectively) by listing the edges that form $C$ consecutively, and we label each edge $\{\pi,\pi r_i\}$ ($\{\pi,\pi r^B_i\}$, respectively), with $\pi\in S_n,r_i\in R$ (with $\pi\in B_n,r_i^B\in R^B$, respectively) by $r_i$ ($r_i^B$, respectively). Since there are multiple ways to refer to a cycle, we choose a canonical form for every cycle. We say that a cycle $C$ is in \emph{canonical form} if $C=r_{i_1}\cdots r_{i_{\ell}}$ ($C=r^B_{i_1}\cdots r^B_{i_{\ell}}$, if in $BP_n$) and the sequence $(i_1,\ldots,i_{\ell})$ is lexicographically maximal among all sequences corresponding to indices of prefix reversals that would also traverse $C$. For example, $r_3 r_2 r_3 r_2 r_3 r_2$ is in canonical form whereas $r_2 r_3 r_2 r_3 r_2 r_3$ is not since $(232323)\underset{\text{lex}}{<}(323232)$. We are now ready to spell out the cycle classifications that are used in the proofs of our main results. The following is an amalgam of results spanning three articles written by Konstantinova and Medvedev, which classifying the canonical forms of small cycles (6-,7-,8-, and 9-cycles) in $P_n$. This single theorem is actually a restatement of four separate results found in \cite[Lemma 3]{KM10}, \cite[Theorem 1]{KM10}, \cite[Theorem 1.3]{KonMed}, and \cite[Theorem 4]{KM11}. \begin{thm}\label{t:cycleclassification} If $n\geq3$, then there is only one canonical 6-cycle in $P_n$: \begin{align} \label{P6-1}&r_3 r_2 r_3 r_2 r_3 r_2. \end{align} Furthermore, the 7-cycles in $P_n$ have the following canonical forms: \begin{align} \label{P7-1}&r_k r_{k-1} r_k r_{k-1} r_{k-2} r_k r_2, &\quad 4 \leq k \leq n. \end{align} Moreover, the 8-cycles in $P_n$ have the following canonical forms: \begin{align} \label{P8-1}&r_{k}r_{j}r_{i}r_{j}r_{k}r_{k-j+i}r_{i}r_{k-j+i}, &\quad 2 \leq i < j \leq k-1, 4 \leq k \leq n,\\ \label{P8-2}&r_{k}r_{k-1}r_{2}r_{k-1}r_{k}r_{2}r_{3}r_{2}, &\quad 4 \leq k \leq n,\\ \label{P8-3}&r_{k}r_{k-i}r_{k-1}r_{i}r_{k}r_{k-i}r_{k-1}r_{i}, &\quad 2 \leq i \leq k-2, 4 \leq k \leq n,\\ \label{P8-4}&r_{k}r_{k-i+1}r_{k}r_{i}r_{k}r_{k-i}r_{k-1}r_{i-1}, &\quad 3 \leq i \leq k-2, 5 \leq k \leq n,\\ \label{P8-5}&r_{k}r_{k-1}r_{i-1}r_{k}r_{k-i+1}r_{k-i}r_{k}r_{i}, &\quad 3 \leq i \leq k-2, 5 \leq k \leq n,\\ \label{P8-6}&r_{k}r_{k-1}r_{k}r_{k-i}r_{k-i-1}r_{k}r_{i}r_{i+1}, &\quad 2 \leq i \leq k-3, 5 \leq k \leq n,\\ \label{P8-7}&r_{k}r_{k-j+1}r_{k}r_{i}r_{k}r_{k-j+1}r_{k}r_{i}, &\quad 2 \leq i < j \leq k-1, 4 \leq k \leq n,\text{ and}\\ \label{P8-8}&r_4r_3r_4r_3r_4r_3r_4r_3. \end{align} Additionally, the 9-cycles in $P_n$ have the following canonical forms: \begin{align} \label{P9-1}&r_{k}r_{k-1}r_{i}r_{k-1}r_{k}r_{i}r_{i-1}r_{i+1}r_{2}, &\quad 3 \leq i \leq k-2, 5 \leq k \leq n,\\ \label{P9-2}&r_{2}r_{k-i+2}r_{k}r_{i-2}r_{i-1}r_{i}r_{i-1}r_{k}r_{k-i+2}, &\quad 4 \leq i \leq k-1, 5 \leq k \leq n,\\ \label{P9-3}&r_{k}r_{k-i}r_{k-1}r_{k-j+i-1}r_{k-j}r_{k}r_{j-i+1}r_{j}r_{i}, &\quad 2 \leq i < j \leq k-2, 5 \leq k \leq n,\\ \label{P9-4}&r_{k}r_{k-1}r_{i}r_{i-1}r_{k-1}r_{k}r_{i}r_{i+1}r_{2}, &\quad 3 \leq i \leq k-2, 5 \leq k \leq n,\\ \label{P9-5}&r_{k}r_{k-1}r_{k-2}r_{k-1}r_{k-2}r_{k}r_{3}r_{k}r_{k-2}, &\quad 4 \leq k \leq n,\\ \label{P9-6}&r_{k}r_{k-1}r_{k-2}r_{i}r_{k}r_{2}r_{k}r_{i}r_{k-1}, &\quad 2 \leq i \leq k-3, 5 \leq k \leq n,\\ \label{P9-7}&r_{k}r_{k-j+i}r_{k}r_{j}r_{i}r_{k}r_{k-j}r_{k-i}r_{j-i}, &\quad 2 \leq i \leq j-2, i+2 \leq j \leq k-2, 6 \leq k \leq n,\\ \label{P9-8}&r_{k}r_{k-j+i}r_{k-j}r_{k}r_{j}r_{i}r_{k}r_{k-i}r_{j-i}, &\quad 2 \leq i \leq j-2, i+2 \leq j \leq k-2, 6 \leq k \leq n,\\ \label{P9-9}&r_{k}r_{k-j+i}r_{k-j+1}r_{k}r_{j}r_{i}r_{k}r_{k-i+1}r_{j-i+1}, &\quad 2 \leq i < j \leq k-1, 4 \leq k \leq n,\\ \label{P9-10}&r_{k}r_{k-1}r_{k}r_{k-1}r_{k}r_{k-1}r_{k-3}r_{k}r_{3}, &\quad 5 \leq k \leq n. \end{align} \end{thm} In the same spirit as the classification of the cycles in $P_n$, the authors proved the following theorem classifying the 8-cycles in $BP_n$, for $n\geq2$. \begin{thm}[Theorem 4.1 in \cite{BBP2018}]\label{t:burntpancakeclassification} If $n\geq2$ then each 8-cycle in $BP_n$ has one of the following canonical forms: \begin{align} \label{BP-8-1} &r_{k}r_{j}r_{i}r_{j}r_{k}r_{k-j+i}r_{i}r_{k-j+i}, & 1 \leq i < j \leq k-1, 3 \leq k \leq n,\\ \label{BP-8-2} &r_{k}r_{j}r_{k}r_{i}r_{k}r_{j}r_{k}r_{i}, & 2 \leq i,j \leq k-2, i+j \leq k, 4 \leq k \leq n,\\ \label{BP-8-3} &r_{k}r_{i}r_{k}r_{1}r_{k}r_{i}r_{k}r_{1}, &2 \leq i \leq k-1, 3 \leq k \leq n,\text{ and}\\ \label{BP-8-4} &r_{k}r_{1}r_{k}r_{1}r_{k}r_{1}r_{k}r_{1} & 2 \leq k \leq n. \end{align} \end{thm} Our question is a very natural one: How many pancake stacks require $k$ flips to be sorted? Equivalently, how many permutations in $S_n$ require composition with $k$ prefix reversals to be sorted? Naturally, we will think of permutations instead of pancake stacks for convenience in notation. In this light, we define the \emph{pancake distance} between two permutations as follows. \begin{defn} Given $\pi_1,\pi_2\in S_n$, we write $d(\pi_1,\pi_2)=k$ if $\pi_2=\pi_1 r_{i_1}\cdots r_{i_k}$ for some $r_{i_1},\ldots,r_{i_k}\in R$, and $k$ is minimal. Namely, if $\pi_2=\pi_1 r_{j_1}r_{j_2}\cdots r_{j_{k'}}$ then $k\leq k'$. We call $d(\cdot,\cdot)$, the pancake distance. By the same token, if given $\sigma_1,\sigma_2\in B_n$ such that $\sigma_1=\sigma_2r^B_{i_1}r^B_{i_2}\cdots r^B_{i_{\ell}}$ and $\ell$ is minimal, we say that the \textit{burnt pancake distance} between $\sigma_1$ and $\sigma_2$ is $\ell$ and write $d^B(\sigma_1,\sigma_2)=\ell$. \end{defn} If $k\geq1$, we denote by $R_k(n)$ the number of permutations in $S_n$ that require $k$ flips to be sorted, that is, $R_k(n)=|\{\pi\in S_n: d(e,\pi)=k\}|$. Similarly, we use $R_k^B(n)$ to denote the cardinality of the set $|\{\pi\in B_n: d^B(e,\pi)=k\}|$ One easily sees that $R_0(n)=1$ (corresponding to the identity permutation that is already sorted) and that $R_1(n)=n-1$, as the only stacks that can be sorted with one flip are the prefix-reversal permutations. For $k=2,3$ the cycle structure of $P_n$ allows us to conclude that \[ R_2(n)=(n-1)(n-2)\text{ and }R_3(n)=(n-1)(n-2)^2-1\text { if }n\geq3. \] Indeed, the smallest cycle in $P_n$ is a 6-cycle and \textit{there is only one} such a cycle, namely, $r_3 r_2 r_3 r_2 r_3 r_2$ (see~\cite[Theorem 1.1-1.2]{KonMed}). Hence, since there are $n-1$ prefix reversals, $R_2(n)=(n-1)(n-2)$ and $R_3(n)=(n-1)(n-2)^2-1$. The first non-trivial computation is $R_4(n)$. We have computed values of $R_k(n)$ and $R^B_k(n)$ for several instances of $n,k$ utilizing a system with Dual Xeon CPUs and 256GB of RAM for the largest computations. We have summarized the values found in Table~\ref{tab:sn} and Table~\ref{tab:bn} and remark that due to computing limitations, several entries are still unknown. However, we were able to compute enough values to be able to offer some conjectures, which we present in Section~\ref{s:conclusion}. We are happy to share our code upon request. We are now ready to prove the first main result of this paper, an explicit description of $R_4(n)$. \section{Permutations requiring four flips to be sorted}\label{sec:sn} The approach we use to obtain the number of pancake stacks that require four flips is the following: we will use the principle of inclusion-exclusion (PIE) using a family of sets $A_i$ and we will use the classification of 7- and 8-cycles in $P_n$ when obtaining the cardinality of the intersections of said sets. More formally, our aim is to obtain the cardinality of the set $A^4:=\{\pi\in S_n: d(e,\pi)=4\}.$ We furthermore let $A_i\subseteq A^4$, for $0\leq i\leq 4$, be the following sets. \begin{enumerate} \item If $0\leq i\leq 3$, \begin{displaymath}A_i := \{\pi=r_{j_1}r_{j_2}r_{j_3}r_{j_4} \in A^4: j_{i+1}=n,j_k\neq n\text{ for }k<i+1\}.\end{displaymath} \item If $i=4$, \begin{displaymath}A_4 := \{\pi=r_{j_1}r_{j_2}r_{j_3}r_{j_4} \in A^4: j_k\neq n \text{ for all }k\in[4]\}.\end{displaymath} \end{enumerate} In other words, if one were to think of a path between $e$ and $\pi\in A^4$ in $P_n$, $A_i$ would contain all the paths that have $i+1$ vertices inside $P_{n-1}(n)$. It follows that \begin{equation}\label{eq:PIE} R_4(n)=|A^4|=\left|\bigcup_{i=0}^n A_i\right|. \end{equation} We will utilize PIE to compute the cardinality of the union $\bigcup_{i=0}^n A_i$. In the proof of Theorem~\ref{t:snk=4}, we will need to determine substrings of larger strings. It will be more clear in the exposition if we establish the following convention. \begin{defn} Let $r_{i_1}r_{i_2}\cdots r_{i_m}$ represent a cycle in $P_n$ or $BP_n$. We say that a string $s$ is a \emph{continuous substring} of $r_{i_1}r_{i_2}\cdots r_{i_m}$ if $s$ or its reversal can be written in the form $r'_{j_1}r'_{j_2}\cdots r'_{j_{\ell}}$ where $j_{k+1}=j_k+1\pmod m$ for $1\leq k\leq \ell-1$. \end{defn} \begin{figure} \begin{center} \begin{tikzpicture}[scale=1,every node/.style={scale=1}] \node (e) at (-1.5,3.5) {$1234$}; \node (2) at (-2.5,2.5) {$3214$}; \node (1) at (-0.5,2.5) {$2134$}; \node (3) at (1.5,3.5) {$4321$}; \node (32) at (2.5,-2.5) {$4123$}; \node (12) at (-2.5,1.5) {$2314$}; \node (21) at (-0.5,1.5) {$3124$}; \node (31) at (-0.5,-2.5) {$4312$}; \node (13) at (0.5,2.5) {$3421$}; \node (23) at (2.5,2.5) {$2341$}; \node (232) at (1.5,-3.5) {$2143$}; \node (132) at (2.5,-1.5) {$1423$}; \node (312) at (-2.5,-1.5) {$4132$}; \node (212) at (-1.5,0.5) {$1324$}; \node (321) at (0.5,-1.5) {$4213$}; \node (231) at (-0.5,-1.5) {$1342$}; \node (131) at (-1.5,-3.5) {$3412$}; \node (213) at (0.5,1.5) {$2431$}; \node (313) at (0.5,-2.5) {$1243$}; \node (123) at (2.5,1.5) {$3241$}; \node (323) at (-2.5,-2.5) {$1432$}; \node (1321) at (1.5,-0.5) {$2413$}; \node (1231) at (-1.5,-0.5) {$3142$}; \node (1213) at (1.5,0.5) {$4231$}; \draw[line width=2.pt,red](e)--(1) (2)--(12) (3)--(13) (31)--(131) (21)--(212) (32)--(132) (23)--(123) (231)--(1231) (321)--(1321) (213)--(1213); \draw[line width=2.pt,blue] (e)--(2) (1)--(21) (3)--(23) (31)--(231) (12)--(212) (32)--(232) (13)--(213) (312)--(1231) (132)--(1321) (123)--(1213); \draw[line width=2.pt,purple!50!black] (e)--(3) (2)--(32) (21)--(321) (12)--(312) (23)--(323) (212)--(1213) \draw[line width=2.pt,purple!50!black,domain=-72:72] plot ({0.25+2*cos(\x)},{2.5*sin(\x)}) plot ({-0.25+2*cos(\x+180)},{2.5*sin(\x+180)}); \draw[line width=2.pt] (232) edge[red] (313) (232) edge[purple!50!black] (131) (132) edge[purple!50!black] (123) (321) edge[blue] (313) (231) edge[purple!50!black] (213) (312) edge[red] (323) (131) edge[blue] (323) (1321) edge[purple!50!black] (1231); \draw[line width=2.pt] (e) edge[preaction={draw,yellow,-,double=yellow,double distance=2\pgflinewidth,},purple!50!black] (3) (3) edge[preaction={draw,yellow,-,double=yellow,double distance=2\pgflinewidth,},blue] (23) (23) edge[preaction={draw,yellow,-,double=yellow,double distance=2\pgflinewidth,},red] (123) (123) edge[preaction={draw,yellow,-,double=yellow,double distance=2\pgflinewidth,},purple!50!black] (132) (132) edge[preaction={draw,yellow,-,double=yellow,double distance=2\pgflinewidth,},red] (32) (32) edge[preaction={draw,yellow,-,double=yellow,double distance=2\pgflinewidth,},blue] (232) (232) edge[preaction={draw,yellow,-,double=yellow,double distance=2\pgflinewidth,},purple!50!black] (131) (131) edge[preaction={draw,yellow,-,double=yellow,double distance=2\pgflinewidth,},blue] (323) (323) edge[preaction={draw,yellow,-,double=yellow,double distance=2\pgflinewidth,},red] (312) (312) edge[preaction={draw,yellow,-,double=yellow,double distance=2\pgflinewidth,},purple!50!black] (12) (12) edge[preaction={draw,yellow,-,double=yellow,double distance=2\pgflinewidth,},red] (2) (2) edge[preaction={draw,yellow,-,double=yellow,double distance=2\pgflinewidth,},blue] (e); \end{tikzpicture} \caption{Pancake graph $P_4$. The different colors indicate the different pancake generators. The 12-cycle $r_4r_3r_2r_4r_2r_3r_4r_3r_2r_4r_2r_3$ is highlighted.} \label{f:network} \end{center} \end{figure} For example, consider the cycle $r_4r_3r_2r_4r_2r_3r_4r_3r_2r_4r_2r_3$ (highlighted in Figure~\ref{f:network}.) Then $r_2r_4r_2r_3r_4$ is a continuous substring as is $r_2r_3r_4r_3$. Our intention with this definition is to emphasize that if $r_{i_1}r_{i_2}\cdots r_{i_m}$ is a cycle, then $r_{i_m}$ and $r_{i_1}$ should be considered consecutive edges. \begin{observation}\label{ob:S1} \begin{enumerate} \item[(a)] There are six continuous substrings of the form $r_n r_a r_b r_c$ in the only canonical form for a 7-cycle containing $r_n$, $r_n r_{n-1} r_n r_{n-1} r_{n-2} r_n r_2$. These are indicated by an arrow on top of the longer string (the directions indicate how to read the substring): \begin{align*} \overrightarrow{r_n r_{n-1} r_n r_{n-1}} r_{n-2} r_n r_2,\qquad& r_n r_{n-1} \overrightarrow{r_n r_{n-1} r_{n-2} r_n}r_2,\\ \overrightarrow{r_n r_{n-1}} r_n r_{n-1} r_{n-2} \overline{r_n r_2},\qquad& \overline{r_n r_{n-1} r_n }r_{n-1} r_{n-2} r_n \overleftarrow{r_2},\\ r_n r_{n-1} \overleftarrow{ r_n r_{n-1} r_{n-2} r_n} r_2,\text{ and }& \overline{r_n} r_{n-1} r_n r_{n-1} \overleftarrow{r_{n-2} r_n r_2}. \end{align*} We let $Q_0$ denote the set of these continuous substrings. That is, \begin{align*} Q_0 := \{ &r_n r_{n-1} r_n r_{n-1}, r_n r_{n-1} r_{n-2} r_n,\\ &r_{n} r_2 r_n r_{n-1}, r_n r_{n-1} r_n r_2, \\ &r_n r_{n-2}r_{n-1} r_n, r_{n} r_2 r_n r_{n-2} \}. \end{align*} \item[(b)] There are six continuous substrings of the form $r_a r_n r_b r_c$ in $r_n r_{n-1} r_n r_{n-1} r_{n-2} r_n r_2$, the only canonical form for a 7-cycle containing $r_n$. These are indicated by an arrow on top of the longer string (the directions indicate how to read the substring): \begin{align*} r_n \overrightarrow{r_{n-1} r_n r_{n-1} r_{n-2}} r_n r_2,\qquad& \overrightarrow{r_n} r_{n-1} r_n r_{n-1} \overline{r_{n-2} r_n r_2},\\ \overrightarrow{r_n r_{n-1} r_n} r_{n-1} r_{n-2} r_n \overline{r_2},\qquad& \overleftarrow{r_n r_{n-1} r_n r_{n-1}} r_{n-2} r_n r_2,\\ r_n r_{n-1} r_n \overleftarrow{r_{n-1} r_{n-2} r_n r_2},\text{ and }& \overline{r_n r_{n-1}} r_n r_{n-1} r_{n-2} \overleftarrow{r_n r_2}. \end{align*} We let $Q_1$ denote the set of these continuous substrings. That is, \begin{align*} Q_1:= \{ &r_{n-1} r_n r_{n-1} r_{n-2}, r_{n-2} r_n r_2 r_n, \\ &r_2 r_n r_{n-1} r_n, r_{n-1} r_n r_{n-1} r_n, \\ &r_2 r_n r_{n-2} r_{n-1}, r_{n-1} r_n r_2 r_n \}. \end{align*} \item[(c)] There are two continuous substrings of the form $r_a r_b r_n r_c$ with $a \neq n$ in $r_n r_{n-1} r_n r_{n-1} r_{n-2} r_n r_2$. These are shown by an arrow over the top of the canonical form of the 7-cycles (the directions indicate how to read the substring): $r_n r_{n-1} r_n \overrightarrow{r_{n-1} r_{n-2} r_n r_2}$ and $r_n \overleftarrow{r_{n-1} r_n r_{n-1} r_{n-2}} r_n r_2$. We let $Q_2$ denote the set of these continuous substrings. That is \[Q_2 := \{r_{n-1} r_{n-2} r_n r_2, r_{n-2} r_{n-1} r_n r_{n-1}\}.\] \item[(d)] Since there is only once 6-cycle, $r_2r_3r_2r_3r_2r_3$, then $r_2r_3r_2r_n=r_3r_2r_3r_n$. We define \begin{displaymath}Q_3 :=\{r_2 r_3 r_2 r_n\}.\end{displaymath} \end{enumerate} \end{observation} The sets $Q_0, Q_1, Q_2,$ and $Q_3$ are important to identify since they are those length 4 paths in $P_n$ for which there is a ``short cut,'' of length 3, in $Q_0, Q_1,$ and $Q_2$. While $Q_3$ represents a length 4 path that potentially may be double counted since it has another representation. Thus these paths will need to be accounted for in the cardinalities of the sets $A_i$. \begin{thm}\label{t:snk=4} If $n\geq4$, then \[ R_4(n)=\frac{1}{2} (2 n^4 - 15 n^3 + 29 n^2 + 6 n - 34). \] \end{thm} \begin{proof} We will use PIE with the sets $A_i$, for $0\leq i\leq 4$ and (\ref{eq:PIE}). We verify each of the cardinalities, one by one. To obtain the relevant cardinality of intersections of the sets, we will use Theorem~\ref{t:cycleclassification} since some permutations that can be sorted in four flips can also be sorted in three, and even if said permutations require four flips to be sorted, there might be more than one way of doing so. We explain each of the cases in detail in the proof. To obtain a recurrence for $R_k(n)$, we will assume that $n>4$. \begin{description} \item[$|A_4|$.] Due to the recursive structure of $P_n$, it follows that $|A_4| = R_4(n-1)$. \item[$|A_3|$.] Any permutation in $A_3$ will be of the form $r_a r_b r_c r_n$ where $2 \leq a,b,c \leq n-1, a \neq b,$ and $b\neq c$. We can choose $a$ in $n-2$ ways and $b,c$ in $n-3$ ways. By Observation~\ref{ob:S1}(d), $r_2r_3r_2r_n=r_3r_2r_3r_n$ so to avoid double counting we remove the string in $Q_3$ from the set of strings of the form $r_a r_b r_c r_n$ to get $A_3$. Hence, $|A_3|=(n-2)(n-3)^2-1$. \item[$|A_2|$.] Any permutation in $A_2$ will be of the form $r_a r_b r_n r_c$ where $2 \leq a,b,c \leq n-1$ and $a\neq b$. We can choose $a,c$ in $n-2$ ways each and $b$ in $n-3$ ways, so there are $(n-2)^2(n-3)$ strings of this form. All of these strings represent permutations that can be sorted in four flips, but some of these can be also be sorted in three flips. So we need to exclude those permutations that can be sorted in three flips to get $A_2$. Notice that any permutation that can be sorted in four flips and also in three flips will be part of a 7-cycle in $P_n$. By Observation~\ref{ob:S1}(c), there are two strings of this form, those in $Q_2$, that lead to the same permutation. Therefore, $|A_2| = (n-2)^2(n-3) - 2$. \item[$|A_1|$.] Any permutation in $A_1$ will be of the form $r_a r_n r_b r_c$ where $2 \leq a,b \leq n-1$, $2 \leq c \leq n$, and $b\neq c$. We can choose $a, b, c$ in $n-2$ ways each, and so there are $(n-2)^3$ strings of this form. All of these strings represent permutations that can be sorted in four flips but some of these can be also be sorted in three flips. Notice that any permutation that can be sorted in four flips and also in three flips will be part of a 7-cycle in $P_n$. So if we can find $r_a r_n r_b r_c$ as a continuous substring of $r_n r_{n-1} r_n r_{n-1} r_{n-2} r_n r_2$ (the only canonical form for a 7-cycle containing $r_n$), then the permutation obtained from $r_a r_n r_b r_c$ can be sorted in three flips, and should not be counted in $A_1$. By Observation~\ref{ob:S1}(b), there are six strings that should not be counted in $A_1$, those in $Q_1$. Furthermore, unlike in the previous case, there are some 8-cycles to consider: The two strings of the form $r_a r_n r_b r_c$ and $r_z r_n r_y r_x$ (with $2\leq x\leq n, 2\leq y,z\leq n-1$) would represent the same permutation, if they are part of an 8-cycle of the form $r_a r_n r_b r_c r_x r_y r_n r_z$. Comparing with the canonical forms of the 8-cycles in Theorem~\ref{t:cycleclassification}. First, there are $2(n-4)$ cycles of the form~(\ref{P8-5}), \begin{align*} &r_{i-1} r_n r_{n-i+1} r_{n-i} \ r_n r_i r_n r_{n-1}, \quad \text{for } 3 \leq i \leq n-2, \text{ and}\\ &r_{n-i} r_n r_{i} r_{n} \ r_{n-1} r_{i-1} r_n r_{n-i+1}, \quad \text{for } 3 \leq i \leq n-2. \end{align*} Second, there are $n-4$ cycles of the form~(\ref{P8-6}): $r_{n-i-1} r_n r_{i} r_{i+1} \ r_{n} r_{n-1} r_n r_{n-i}$ with $2 \leq i \leq n-3$. For a total of $3(n-4)$ such strings that are double counted of the form $r_ar_nr_br_c$. Therefore, $|A_1| = (n-2)^3-3(n-4)-6$. \item[$|A_0|$.] Any permutation in $A_0$ will be of the form $r_n r_a r_b r_c$ where $2 \leq a \leq n-1$, $2 \leq b,c \leq n$ and $b\neq c$. We can choose $a, b, c$ in $n-2$ ways each, so there are $(n-2)^3$ strings that follow this form. The two strings $r_n r_2 r_3 r_2 $ and $r_n r_3 r_2 r_3 $ lead to the same permutation because of the only 6-cycle $r_3r_2r_3r_2r_3r_2$ in $P_n$ (so $r_3r_2r_3=r_2r_3r_2$) so we will exclude the string $r_n r_2 r_3 r_2$ to avoid double counting. As before, we need to account for strings that yield permutations that are part of 7-cycles, since these can be sorted in three flips. Let $\pi=r_n r_a r_b r_c$ denote a permutation that can be sorted in three flips, then $\pi$ must form part of a 7-cycle. By Observation~\ref{ob:S1}(a), there are six strings, those in $Q_0$, that should not be counted as part of $A_0$. In this case, no 8-cycle leads to double counting as if there exists $r_n r_x r_y r_z $ with $2\leq x\leq n-1$ and $2\leq y,z\leq n$ with $r_nr_ar_br_c=r_nr_xr_yr_z$, then $r_n r_a r_b r_c r_z r_y r_x r_n$ must be an 8-cycle, which is not possible by observing the canonical forms for 8-cycles in $P_n$ given in Theorem~\ref{t:cycleclassification}. Therefore, $|A_0| = (n-2)^3 -7$. \item[$|A_0 \cap A_1|$.] If $\pi\in A_0\cap A_1$, then there must be an 8-cycle of the form $r_x r_n r_y r_z r_a r_b r_c r_n$ with $2 \leq a,b,z \leq n$, $2 \leq c,x,y \leq n-1$, $y \neq z$, $a \neq z$, $a \neq b$, and $c \neq b$ where $r_x r_n r_y r_z \not\in Q_1$ and $r_n r_c r_b r_a \notin Q_0$. Since in this cycle the two $r_n$s are separated by one reversal, no such 8-cycles exist of the form (\ref{P8-1}), (\ref{P8-2}), nor (\ref{P8-3}). By comparing with (\ref{P8-4}), we see that the form \begin{align*} &r_{n-i+1} r_n r_i r_k \ r_{n-i} r_{n-1} r_{i-1} r_n, \end{align*} with $ 3 \leq i \leq n-2$, contributes $n-4$ to $|A_0 \cap A_1|$. By comparing with (\ref{P8-5}) we see that the forms \begin{align*} &r_i r_n r_{n-1} r_{i-1} \ r_n r_{n-i+1} r_{n-i} r_n, \quad \text{for } 3 \leq i \leq n-2,\text{ and }\\ &r_{i} r_n r_{n-i} r_{n-i+1} \ r_n r_{i-1} r_{n-1} r_n, \quad \text{for } 3 \leq i \leq n-2, \end{align*} contribute $2(n-4)$ to $|A_0 \cap A_1|$. By comparing with (\ref{P8-6}) we see that the form \begin{align*} &r_{n-1} r_n r_{n-i} r_{n-i-1} \ r_n r_i r_{i+1} r_n, \end{align*} with $2 \leq i \leq n-3$, contributes $n-4$ to $|A_0 \cap A_1|$. By comparing with (\ref{P8-7}) we get that the following cycles \begin{align*} &r_{n-j+1} r_n r_i r_n \ r_{n-j+1} r_n r_i r_n, \end{align*} with $2 \leq i < j\leq n-1$, contribute $\frac{1}{2}(n-2)(n-3)$ to $|A_0 \cap A_1|$. However, when $i=2$ and $j=3$ the cycle is $r_{n-2}r_nr_2r_nr_{n-2}r_nr_2r_n$. So $r_{n-2}r_nr_2r_n=r_nr_2r_nr_{n-2}$, but the left-hand-side string is in $Q_1$ and the right-hand-side string is in $Q_0$. Thus this cycle should be excluded, since both strings actually represent the same permutation $r_{n-1}r_nr_{n-1}$. Therefore, $|A_0 \cap A_1| = \frac{1}{2}(n-2)(n-3) + 4(n-4) -1 = \frac{1}{2}(n^2 + 3n - 28)$. \item [$|A_0 \cap A_2|$.] If $\pi \in A_0 \cap A_2$, then there must be an 8-cycle of the form $r_x r_y r_n r_z r_a r_b r_c r_n$ with $2 \leq a,b,x \leq n$, $2 \leq c,y,z \leq n-1$, $x \neq y$, $z \neq a$, $a \neq b$, and $b \neq c$ where $r_x r_y r_n r_z \notin Q_2$ and $r_n r_c r_b r_a \notin Q_0$. Since in this cycle a pair of $r_n$s are separated by two reversals, no such cycle of the form (\ref{P8-1}), (\ref{P8-2}), (\ref{P8-3}), (\ref{P8-4}), nor (\ref{P8-7}) exists. Upon comparison with (\ref{P8-5}) we see that the forms \begin{align*} &r_{n-1}r_{i-1}r_{n}r_{n-i+1}\ r_{n-i}r_{n}r_{i}r_{n}, \quad \text{for } 3 \leq i \leq n-2,\\ &r_{n-i+1}r_{n-i}r_{n}r_{i}\ r_{n}r_{n-1}r_{i-1}r_{n}, \quad \text{for } 3 \leq i \leq n-2,\\ &r_{n-i}r_{n-i+1}r_{n}r_{i-1}\ r_{n-1}r_{n}r_{i}r_{n}, \quad \text{for } 3 \leq i \leq n-2, \text{ and }\\ &r_{i-1}r_{n-1}r_{n}r_{i}\ r_{n}r_{n-i}r_{n-i+1}r_{n}, \quad \text{for } 3 \leq i \leq n-2,\\ \end{align*} contribute $4(n-4)$ to $|A_0 \cap A_2|$. Furthermore, upon comparing with (\ref{P8-6}), we see that \begin{align*} &r_{n-i}r_{n-i-1}r_{n}r_{i}\ r_{i+1}r_{n}r_{n-1}r_{n}, \quad \text{for } 2 \leq i \leq n-3, \text{ and }\\ &r_{i}r_{i+1}r_{n}r_{n-1}\ r_{n}r_{n-i}r_{n-i-1}r_{n}, \quad \text{for } 2 \leq i \leq n-3, \end{align*} contribute $2(n-4)$ to $|A_0 \cap A_2|$. Hence $|A_0 \cap A_2| = 6(n-4)$. \item[$|A_0 \cap A_3|$.] If $\pi\in A_0\cap A_3$, then there must be an 8-cycle of the form $r_x r_y r_z r_n r_a r_b r_c r_n$ with $2 \leq a,c,x,y,z \leq n-1$, $2 \leq b \leq n$, $x \neq y$, $y \neq z$, $a \neq b$, and $b \neq c$ where $r_x r_y r_z r_n \notin Q_1$ and $r_n r_c r_b r_a \notin Q_0$. Since in this cycle a pair of $r_n$s are separated by three reversals, no such cycle of the form (\ref{P8-5}), (\ref{P8-6}), nor (\ref{P8-7}) exists. Let us compare with form (\ref{P8-1}), we see that the form \begin{align*} &r_{j}r_{i}r_{j}r_{n}\ r_{n-j+i}r_{i}r_{n-j+i}r_{n}, \end{align*} with $2 \leq i <j \leq n-1$, contributes $\frac{1}{2}(n-2)(n-3)$ to $|A_0\cap A_3|$. When we compare with form (\ref{P8-2}), $r_{2}r_{3}r_{2}r_{n}\ r_{n-1}r_{n}r_{i}r_{n}$ actually does not contribute to $|A_0\cap A_3|$ since $r_{2}r_{3}r_{2}r_{n}\in Q_3$. We now compare with (\ref{P8-3}), \begin{align*} r_{n-i}r_{n-1}r_{i}r_{n}r_{n-i}r_{n-1}r_{i}r_{n}, \end{align*} with $ 2 \leq i \leq n-2$. Each value of $i$ gives a valid cycle (neither of the paths are in $Q_0 \ nor \ Q_3$). We now compare with (\ref{P8-4}), \begin{align*} &r_{n-i}r_{n-1}r_{i-1}r_{n}r_{n-i+1}r_{n}r_{i}r_{n}, \quad \text{for } 3 \leq i \leq n-2. \end{align*} This form contributes $n-4$ to $|A_0 \cap A_3|$. Hence, \begin{displaymath}|A_0 \cap A_3| = \frac{1}{2}(n-2)(n-3) + (n-4) +(n-3) = \frac{1}{2}(n^2 - n - 8). \end{displaymath} \item[$|A_0 \cap A_4|$.] If a permutation $\pi \in A_0 \cap A_4$, then there must be an 8-cycle of the form $r_w r_x r_y r_z r_a r_b r_c r_n$, where $2 \leq w,x,y,z \leq n-1$ and $r_n r_c r_b r_a \notin Q_0$. No such 8-cycle exist since there are no cycles with four reversals between a pair of $r_{n}$ in Theorem \ref{t:cycleclassification}. So $|A_0 \cap A_4|=0$. \item[$|A_1 \cap A_2|$.] If a permutation $\pi \in A_1 \cap A_2$, then there must be an 8-cycle of the form $r_x r_y r_n r_z r_a r_b r_n r_c$ with $2 \leq a \leq n$, $2 \leq b,c,x,y,z \leq n-1$, $x \neq y$, $z \neq a$, and $a \neq b$ where $r_x r_y r_n r_z \notin Q_2$ and $r_c r_n r_b r_a \notin Q_1$. Since in this cycle a pair of $r_{n}$ are separated by three reversals, no such cycle of the form (\ref{P8-5}), (\ref{P8-6}), nor (\ref{P8-7}) exists. From form (\ref{P8-1}), we get: $r_{i}r_{j}r_{n}r_{n-j+i} \ r_{i}r_{n-j+i}r_{n}r_{j}$ with $2 \leq i < j \leq n-1$. Each value of $i$ and $j$ give a valid cycle (the paths are not in $Q_1 \ or \ Q_2$) except when $i=n-2$ and $j=n-1$. Hence the contribution of (\ref{P8-1}) is $\frac{1}{2}(n-2)(n-3)-1$. From form (\ref{P8-2}), we get: $r_{2}r_{n-1}r_{n}r_{2}r_{3}r_{2}r_{n}r_{n-1}$ and $r_{3}r_{2}r_{n}r_{n-1}r_{2}r_{n-1}r_{n}r_{2}$. Thus this form contributes $2$ to $|A_1 \cap A_2|$. From form (\ref{P8-3}), we get: $r_{n-1}r_{i}r_{n}r_{n-i} \ r_{n-1}r_{i}r_{n}r_{n-i}$ with $2 \leq i \leq n-2$. Each value of $i$ except for $i=n-2$ gives a valid cycle (the paths are not in $Q_1 \ or \ Q_2$). Thus this form contributes $n-4$ to $|A_1 \cap A_2|$. From form (\ref{P8-4}) we get: $r_{n-1}r_{i-1}r_{n}r_{n-i+1} \ r_{n}r_{i}r_{n}r_{n-i}$ with $3 \leq i \leq n-2$. If we replace $i-1$ with $i$ in this form, we obtain the cycles whose first four reversals are the same as those obtained from the form (\ref{P8-3}) in the previous paragraph. So the permutations that were contributed from form (\ref{P8-4}) were already considered and would be double counted. Thus this particular form contributes nothing to $|A_1 \cap A_2|$. Therefore, $|A_1 \cap A_2| = \frac{1}{2}(n-2)(n-3) + (n-4)+1 = \frac{1}{2}(n^2-3n)$. \item[Other intersections of two sets.] All other intersections $A_1 \cap A_3,A_1 \cap A_4,A_2 \cap A_3,A_2 \cap A_4$, and $A_3 \cap A_4$ are empty. Otherwise, there would be an 8-cycle that could not be matched to any of the canonical forms of Theorem~\ref{t:cycleclassification}. Specifically, it can seen that the number of reversals between any two $r_n$ would be greater than or equal to four, which does not occur in any of the 8-cycles in Theorem~\ref{t:cycleclassification}. \item[$|A_0 \cap A_1 \cap A_2|$.] If a permutation belongs to these three sets, $\pi\in A_0\cap A_1\cap A_2$, then it must be that the permutation can be written in the forms $\pi=r_n r_w r_v r_u$, $\pi=r_c r_n r_b r_a$ and $\pi=r_x r_y r_n r_z$ with $2 \leq b,c,x,y,w,z \leq n-1$ and $2 \leq a,u,v \leq n$ and none in $Q_0, Q_1,$ or $Q_2$. Thus, $P_n$ must contain 8-cycles of the following forms: \begin{align} \label{3int1} &r_x r_y r_n r_z r_a r_b r_n r_c, \\ \label{3int2} &r_x r_y r_n r_z r_u r_v r_w r_n, \text{ or}\\ \label{3int3} &r_c r_n r_a r_b r_u r_v r_w r_n. \end{align} Notice that if there are cycles of the form (\ref{3int1}) and of the form (\ref{3int2}), there will be cycles that can be written in the form (\ref{3int3}) as well. Once again we will use the classification of the 8-cycles given in Theorem~\ref{t:cycleclassification} to see if there are cycles that can be written in the form (\ref{3int1}), that would share the first four reversals from (\ref{3int2}) at the same time. Comparing (\ref{P8-1}) with (\ref{3int1}) we obtain the cycles of the form \begin{equation}\label{eq:3int-1} r_i r_j r_n r_{n-j+i} \ r_i r_{n-j+i}r_n r_j, \end{equation} with $2 \leq i < j \leq n-1$. With the substitutions of $n-1$ for $j$ and $i-1$ for $i$, these cycles become \begin{equation*} r_{i-1} r_{n-1} r_{n} r_{i} \ r_{i-1} r_{i} r_{n} r_{n-1}, \end{equation*} with $3 \leq i \leq n-1$. Notice when comparing with the form (\ref{P8-5}), \begin{equation*} r_{i-1}r_{n-1}r_n r_i r_n r_{n-i} r_{n-i+1} r_n, \end{equation*} with $2 \leq i \leq n-3$ the first four reversals match. So each value of $i$ in the overlap of indices, $3 \leq i \leq n-2$, corresponds to a permutation in this intersection and contributes a total of $n-4$ to $|A_0 \cap A_1 \cap A_2|$. Now substituting $i+1$ for $j$ in (\ref{eq:3int-1}), we obtain the following form \begin{equation*} r_{i} r_{i+1} r_{n} r_{n-1} r_{i} r_{n-1} r_{n} r_{i+1}, \end{equation*} which matches (\ref{P8-6}) \begin{equation*} r_i r_{i+1} r_n r_{n-1} r_n r_{n-i} r_{n-i-1} r_n, \end{equation*} in the first four reversals, with $2 \leq i \leq n-3$. So each value $i$ in the above form contributes a permutation to the intersection for a total of $n-4$ to $|A_0 \cap A_1 \cap A_2|$. No other canonical 8-cycles would have the first four reversals from the form (\ref{3int1}), and so the total contribution to $|A_0 \cap A_1 \cap A_2|$ starting from (\ref{P8-1}) is $2(n-4)$. Moreover, comparing (\ref{P8-2}) with (\ref{3int1}), we do not find any additional cycles that are also of the form (\ref{3int2}). Furthermore, comparing (\ref{P8-3}) with (\ref{3int1}), we find the form \begin{equation}\label{eq:3int-4} r_{n-1}r_{i}r_{n}r_{n-i} \ r_{n-1}r_{i}r_{n}r_{n-i}, \end{equation} with $2 \leq i \leq n-2$. By substituting $i-1$ for $i$ in (\ref{eq:3int-4}), we obtain \begin{equation*} r_{n-1}r_{i-1}r_{n}r_{n-i+1} \ r_{n-1}r_{i-1}r_{n}r_{n-i+1}, \end{equation*} with $3 \leq i \leq n-1$. The first four reversals of which match with form (\ref{P8-5}) \begin{equation*} r_{n-1}r_{i-1}r_{n}r_{n-i+1}\ r_{n-i}r_{n}r_{i}r_{n}, \end{equation*} with $3 \leq i \leq n-2$. So each value of $i$ in the overlap of these intervals, $3 \leq i \leq n-2$, contribute a permutation to this intersection for a total of $n-4$ added to $|A_0 \cap A_1 \cap A_2|$. No other matches with form (\ref{P8-3}) are obtained, and thus the contribution to $|A_0 \cap A_1 \cap A_2|$ from (\ref{P8-3}) is $n-4$. Finally, we compare (\ref{P8-4}) with (\ref{3int1}) and obtain the form \begin{equation}\label{eq:3int-6} r_{n-1}r_{i-1}r_{n}r_{n-i+1} \ r_{n}r_{i}r_{n}r_{n-i}, \end{equation} with $3 \leq i \leq n-2$. If we compare form (\ref{eq:3int-6}) with any of the other canonical forms of 8-cycles in $P_n$, we would find a match with form (\ref{P8-3}) after substituting $i-1$ for $i$ in (\ref{eq:3int-6}). However, these cycles have already been counted. This completes the count of permutations found in $A_0 \cap A_1 \cap A_2$, and thus $|A_0 \cap A_1 \cap A_2|=3(n-4)$. \end{description} An exhaustive argument gives that any other intersections of three of the sets $A_0,\ldots,A_4$ will be empty, since no cycle matching three forms at the same time exists. Therefore if $n>4$, \begin{align*} R_4(n) &= \left|\bigcup_{i=0}^4 A_i\right| \\ &= \sum_{S \subseteq \{0,1,2,3,4\}, S \neq \emptyset} (-1)^{|S|+1} \bigcap_{i \in S} A_i\\ &= |A_0| + |A_1| + |A_2| + |A_3| + |A_4| - |A_0 \cap A_1| - |A_0 \cap A_2| - |A_0 \cap A_3| - |A_1 \cap A_2|\\ & \quad + |A_0 \cap A_1 \cap A_2|\\ &=R_4(n-1) + \frac{1}{2}(8n^3 - 57n^2 + 111n - 40). \end{align*} Using the initial condition $R_4(4)=3$ (see Table~\ref{tab:sn}) and solving the recurrence, it follows that if $n\geq4$, \[ R_4(n)=\frac{1}{2} (2 n^4 - 15 n^3 + 29 n^2 + 6 n - 34), \] as desired. \end{proof} \begin{table \centering \tiny \scalebox{1.1}{ \begin{tabular}{|*{13}{c|}} \hline \diagbox{$n$}{$k$} & 0 & 1 & 2 & 3 & 4 & 5 & 6 & 7 & 8 & 9 & 10 & 11 \\ \hline 1 & 1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0\\ \hline 2 & 1 & 1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0\\ \hline 3 & 1 & 2 & 2 & 1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0\\ \hline 4 & 1 & 3 & 6 & 11 & 3 & 0 & 0 & 0 & 0 & 0 & 0 & 0\\ \hline 5 & 1 & 4 & 12 & 35 & 48 & 20 & 0 & 0 & 0 & 0 & 0 & 0\\ \hline 6 & 1 & 5 & 20 & 79 & 199 & 281 & 133 & 2 & 0 & 0 & 0 & 0\\ \hline 7 & 1 & 6 & 30 & 149 & 543 & 1357 & 1903 & 1016 & 35 & 0 & 0 & 0\\ \hline 8 & 1 & 7 & 42 & 251 & 1191 & 4281 & 10561 & 15011 & 8520 & 455 & 0 & 0\\ \hline 9 & 1 & 8 & 56 & 391 & 2278 & 10666 & 38015 & 93585 & 132697 & 79379 & 5804 & 0\\ \hline 10 & 1 & 9 & 72 & 575 & 3963 & 22825 & 106461 & 377863 & 919365 & 1309756 & 814678 & 73232\\ \hline 11 & 1 & 10 & 90 & 809 & 6429 & 43891 & 252737 & 1174766 & 4126515 & 9981073 & 14250471 & 9123648\\ \hline 12 & 1 & 11 & 110 & 1099 & 9883 & 77937 & 533397 & 3064788 & 14141929 & 49337252 & 118420043 & 169332213\\ \hline 13 & 1 & 12 & 132 & 1451 & 14556 & 130096 & 1030505 & 7046318 & 40309555 & 184992275 & 639783475 & 1525125357\\ \hline 14 & 1 & 13 & 156 & 1871 & 20703 & 206681 & 1858149 & 14721545 & 100464346 & 572626637 &&\\ \hline 15 & 1 & 14 & 182 & 2365 & 28603 & 315305 & 3169675 & 28528986 & 226016576 &&&\\ \hline 16 & 1 & 15 & 210 & 2939 & 38559 & 465001 & 5165641 & 52027677 & 468966948 &&&\\ \hline 17 & 1 & 16 & 240 & 3599 & 50898 & 666342 & 8102491 & 90238067 & 911274131 &&&\\ \hline 18 & 1 & 17 & 272 & 4351 & 65971 & 931561 & 12301949 & 150044655 & 1677036683 &&&\\ \hline 19 & 1 & 18 & 306 & 5201 & 84153 & 1274671 & 18161133 & 240665410 & 2947991637 &&&\\ \hline 20 & 1 & 19 & 342 & 6155 & 105843 & 1711585 & 26163389 &374193014 & 4982872347 &&&\\ \hline 21 & 1 & 20 & 380 & 7219 & 131464 & 2260236 & 36889845 &566212968 & 8141208511 &&&\\ \hline \end{tabular} } \caption{Numbers of the form $R_k(n)$ for several values of $n$ and $k$. In particular, notice that if $n\geq4$, $R_4(n)=\frac{1}{2} (2 n^4 - 15 n^3 + 29 n^2 + 6 n - 34)$. The empty entries are unknown to us.} \label{tab:sn} \end{table} \section{Signed permutations requiring four flips to be sorted}\label{sec:bn} We write $R_k^B(n)$ to denote the number of sign permutations that require $k$ burnt pancake flips to be sorted. Since the smallest cycle that can be found in $BP_n$ has length 8 (See~\cite{Compeau2011}), it follows that if $n\geq1$, $R_1^B(n)=n,R_2^B(n)=n(n-1)$, and $R_3^B(n)=n(n-1)^2.$ So the first non-trivial case is the computation of $R_4^B(n)$. We will follow the same method used in Section~\ref{sec:sn}: We will define certain sets whose union will the set of all signed permutations requiring four flips to sort. The cardinality of this union will then equal the number of burnt pancake stacks that require four flips to be sorted. The computation of the cardinality of the union of the sets that we will define is carried out utilizing the principle of inclusion-exclusion. We will use the classification of canonical forms of the 8-cycles, due to the authors~\cite{BBP2018}, from Theorem~\ref{t:burntpancakeclassification}. Our aim is to obtain the cardinality of the set $BA^4:=\{\pi\in B_n: d^B(e,\pi)=4\}$. We furthermore let $BA_i\subseteq BA^4$, $0\leq i\leq 4$, be the following sets. \begin{enumerate} \item If $0\leq i\leq 3$, \begin{displaymath}BA_i := \{\pi=r^B_{j_1}r^B_{j_2}r^B_{j_3}r^B_{j_4} \in BA^4: j_{i+1}=n,j_k\neq n\text{ for }k<i+1\}.\end{displaymath} \item If $i=4$, \begin{displaymath}BA_4 := \{\pi=r^B_{j_1}r^B_{j_2}r^B_{j_3}r^B_{j_4} \in BA^4: j_k\neq n \text{ for all }k\in[4]\}.\end{displaymath} \end{enumerate} In other words, if one were to think of a path between the identity and $\pi\in BA^4$ in $BP_n$, $BA_i$ would contain all the paths that have $i+1$ vertices inside $BP_{n-1}(n)$. It follows that \begin{equation}\label{eq:bPIE} R^B_4(n)=|BA^4|=\left|\bigcup_{i=0}^n BA_i\right|. \end{equation} Just like in the previous section, we will use PIE to compute the cardinality of the union $\bigcup_{i=0}^n BA_i$. In the proof of the main theorem of this section, the following two results will be used. \begin{lem}[Lemma 4.5 in~\cite{BBP2018}]\label{l:4.5} If $\pi_1, \pi_2 \in V (BP_{n-1}(p))$, for any $p \in [\pm n]$, with $d^B(\pi_1, \pi_2) \leq 2$, then $\pi_1r^B_n$ and $\pi_2r^B_n$ must belong to distinct copies of $BP_{n-1}$ in $BP_n$. \end{lem} Moreover, the following corollary also follows. \begin{cor}\label{c:4vertices} Let $C$ be an 8-cycle in $BP_n$, with $n\geq2$. If $C$ has vertices in exactly two copies $BP_{n-1}(i)$ and $BP_{n-1}(j)$ with $i,j\in[\pm n]$, then $C$ has four vertices in $BP_{n-1}(i)$ and four vertices in $BP_{n-1}(j)$. \end{cor} \begin{proof} By Lemma \ref{l:4.5}, if the endpoints in the $BP_{k-1}(p)$ copy (say $\pi_1$ and $\pi_2$) are at a distance of at most two, then $\pi_1r_k$ and $\pi_2r_k$ will belong to distinct copies of $BP_{k-1}$. Hence an 8-cycle cannot occur in such a way that it has six vertices in one copy of $BP_{n-1}$ and two in the other, or with five vertices in one copy and two in the other. Therefore an 8-cycle with vertices in exactly two copies of $BP_{n-1}$ can only have four vertices in each of the copies. \end{proof} We now state and prove the main theorem of this section, that is, an explicit formula for $R^B_4(n)$. \begin{thm}\label{t:bsnk=4} If $n\geq1$, then \[ R_4^B(n) = \frac{1}{2} n(n-1)^2(2n-3). \] \end{thm} \begin{proof} We will use PIE with the sets $BA_i$, $0\leq i\leq 4$, and (\ref{eq:bPIE}), and the canonical forms for the 8-cycles from Theorem~\ref{t:burntpancakeclassification}. We analyze each of the cardinalities individually. To derive a recurrence for $R^B_k(n)$, we will first assume that $n>3$. \begin{description} \item[$|BA_4|$.] Due to the recursive structure of $BP_n$, it follows that $|BA_4| = R_4(n-1)$. \item[$|BA_3|$.] Any permutation in $BA_3$ will be of the form $r^B_ar^B_br^B_cr^B_n$ where $1 \leq a,b,c \leq n-1$, $a\neq b$, and $b\neq c$. We can choose $a$ in $n-1$ ways and $b, \ c$ in $n-2$ ways each. Since there are no 6-, nor 7-cycles in $BP_n$, each choice of $a,b,c$ will give a different signed permutation that requires four flips to be sorted. Indeed, if there were two strings $r^B_dr^B_er^B_fr^B_n$ and $r^B_xr^B_by^B_cz^B_n$ corresponding to the same permutation, $r^B_dr^B_er^B_fr^B_zr^B_yr^B_x$ would be a 6-cycle in $BP_n$, and these do not exist. Therefore, $|BA_3|=(n-1)(n-2)^2$. \item[$|BA_2|$.] Any permutation in $BA_2$ will be of the form $r^B_ar^B_br^B_nr^B_c$ where $1 \leq a,b,c \leq n-1$, $a\neq b$. All of these signed permutations require four flips to be sorted, as there are no 6-, nor 7-cycles. Furthermore, if two strings of this form produced the same signed permutation, then there would be an 8-cycle of the form $r^B_d r^B_e r^B_n r^B_f r^B_z r^B_n r^B_y r^B_x$, which cannot be placed in any of the canonical forms given in Theorem~\ref{t:burntpancakeclassification}. Therefore, $|BA_2|=(n-1)^2(n-2)$. \item[$|BA_1|$.] Any permutation in $BA_1$ will be of the form $r^B_a r^B_n r^B_b r^B_c$ where $1 \leq a,b \leq n-1$, $1 \leq c \leq n$, and $b\neq c$. We can choose $a, b, c$ in $n-1$. The same arguments presented in the previous cases yield that each of these strings give a different signed permutation, and so $|BA_1|=(n-1)^3$. \item[$|BA_0|$.] Any permutation in $BA_0$ will be of the form $r^B_n r^B_a r^B_b r^B_c$ where $1\leq a\leq n-1$, $1 \leq b,c \leq n$, and $b\neq c$. Since there are no 6-, nor 7- cycles, each of these strings will lead to a different signed permutation, and so $|BA_0|=(n-1)^3$. \item[$|BA_0 \cap BA_1|$.] If $\pi \in BA_0 \cap BA_1$, then there must be an 8-cycle of the form $r^B_n r^B_a r^B_b r^B_c r^B_z r^B_y r^B_n r^B_x$ where $1 \leq a,x,y \leq n-1$, $1 \leq b,c,y,z \leq n$, $a \neq b$, $b \neq c$, and $y \neq z$. We find contributions to this form from the canonical forms (\ref{BP-8-2}), (\ref{BP-8-3}), and (\ref{BP-8-4}) only. For form (\ref{BP-8-2}), we obtain the following cycles \begin{equation}\label{eq:b1} r^B_{n}r^B_{j}r^B_{n}r^B_{i}r^B_{n}r^B_{j}r^B_{n}r^B_{i}, \end{equation} with $2 \leq i,j \leq n-2, i+j \leq n$. Considering the possible values of $i,j$ can take in (\ref{eq:b1}), we get $\frac{1}{2}(n-3)(n-2)$ 8-cycles. Similarly, comparing with (\ref{BP-8-3}) we get $2(n-2)$ 8-cycles arising from all possible values of $i$ in the forms below. \begin{align*} r^B_n r^B_i r^B_n r^B_1 r^B_n r^B_i r^B_n r^B_1, &\text{ with }2 \leq i \leq n-1, \text{ and}\\ r^B_n r^B_1 r^B_n r^B_i r^B_n r^B_1 r^B_n r^B_i, &\text{ with }2 \leq i \leq n-1. \end{align*} Furthermore, by comparing with (\ref{BP-8-4}) we get only one 8-cycle: $r^B_n r^B_1 r^B_n r^B_1 r^B_n r^B_1 r^B_n r^B_1 $. Putting the pieces together, we have \begin{displaymath}|BA_0 \cap BA_1| = \frac{1}{2}(n-2)(n-3)+2(n-2)+1.\end{displaymath} \item[$|BA_0 \cap BA_2|$.] If $\pi \in BA_0 \cap BA_2$, then there must be an 8-cycle of the form $r^B_n r^B_a r^B_b r^B_c r^B_z r^B_n r^B_y r^B_x$. If such a cycle existed, it would have five vertices in one copy of $BP_{n-1}$, which contradicts Corollary~\ref{c:4vertices}. Hence, $|BA_0 \cap BA_2|=0$. \item[$|BA_0 \cap BA_3|$.] If $\pi \in BA_0 \cap BA_3$, then there must be an 8-cycle of the form $r^B_n r^B_a r^B_b r^B_c r^B_n r^B_z r^B_y r_x$ where $1 \leq a,c,x,y,z \leq n-1$, $1 \leq b \leq n$, $a \neq b$, $b \neq c$, $x \neq y$, and $y \neq z$. The only canonical form that can match this is (\ref{BP-8-1}), obtaining \begin{equation}\label{eq:b2} r^B_{n}r^B_{j}r^B_{i}r^B_{j}r^B_{n}r^B_{n-j+i}r^B_{i}r^B_{n-j+i}, \end{equation} with $1\leq i<j \leq n-1$. There are $\frac{1}{2}(n-1)(n-2)$ possible values for $i,j$ in (\ref{eq:b2}), and so \begin{displaymath}|BA_0 \cap BA_3| = \frac{1}{2}(n-1)(n-2).\end{displaymath} \item[$|BA_0 \cap BA_4|$.] Essentially the same argument as the case $BA_0 \cap BA_2$ gives that $|BA_0 \cap BA_4|=0$. \item[$|BA_1 \cap BA_2|$.] If $\pi \in BA_1 \cap BA_2$, then there must be an 8-cycle of the form $r^B_a r^B_n r^B_b r^B_c r^B_z r^B_n r^B_y r^B_x$ where $1 \leq a,b,x,y,z \leq n-1$, $1 \leq c \leq n$, $a \neq x$, $b \neq c$, $c \neq z$, and $x \neq y$. The only canonical form from Theorem~\ref{t:burntpancakeclassification} that matches this form is (\ref{BP-8-1}), we get \begin{equation}\label{eq:b3} r^B_{n-j+i}r^B_{n}r^B_{j}r^B_{i}r^B_{j}r^B_{n}r^B_{n-j+i}r^B_{i}, \end{equation} with $1\leq i<j \leq n-1$. Considering all possible values for $i,j$ in (\ref{eq:b3}), we have \begin{displaymath}|BA_1 \cap BA_2| = \frac{1}{2}(n-1)(n-2).\end{displaymath} \item[Other intersections] All other intersections $BA_1 \cap BA_3, BA_1 \cap BA_4,BA_2 \cap BA_3, BA_2 \cap BA_4$, and $BA_3 \cap BA_4$ are empty. By the same token, all the intersections of three distinct sets from $\{BA_i\}_{i=0}^4$ are empty as well. Indeed, if one of these intersections were not empty, then there would be an 8-cycle that could not be matched to any of the canonical forms of Theorem~\ref{t:burntpancakeclassification}. \end{description} Now, using PIE, if $n>3$ it follows that \begin{align*} R^B_4(n) &= \left|\bigcup_{i=0}^4BA_i\right| \\ &= \sum_{S \subseteq \{0,1,2,3,4\}, S \neq \emptyset} (-1)^{|S|+1} \bigcap_{i \in S} BA_i\\ &= |BA_0| + |BA_1| + |BA_2| + |BA_3| + |BA_4| - |BA_0 \cap BA_1| -|BA_0 \cap BA_3|\\ &\quad - |BA_1 \cap BA_2|\\ &=R^B_4(n-1) + \frac{1}{2}(8n^3-33n^2+45n-20). \end{align*} After solving the recurrence relation, using the initial condition $R^B_4(3)=18$ (see Table~\ref{tab:bn}), we obtain that for $n\geq3$, \[ R^B_4(n)=\frac{1}{2} n(n-1)^2(2n-3). \] Upon further inspection, it turns out that the if we plug in $n=1,2$ into $\frac{1}{2} n(n-1)^2(2n-3)$ we obtain $0,1$ respectively. Since these are indeed the true values of $R^B_4(1)$ and $R^B_4(2)$, we have that for $n\geq1$, $R^B_4(n)=\frac{1}{2} n(n-1)^2(2n-3)$. This completes the proof of the theorem. \end{proof} \begin{table \centering \tiny \scalebox{1.1}{ \begin{tabular}{|*{13}{c|}} \hline \diagbox{$n$}{$k$} & 0 & 1 & 2 & 3 & 4 & 5 & 6 & 7 & 8 & 9 & 10 & 11\\ \hline 1 & 1 & 1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0\\ \hline 2 & 1 & 2 & 2 & 2 & 1 & 0 & 0 & 0 & 0 & 0 & 0 & 0\\ \hline 3 & 1 & 3 & 6 & 12 & 18 & 6 & 2 & 0 & 0 & 0 & 0 & 0\\ \hline 4 & 1 & 4 & 12 & 36 & 90 & 124 & 96 & 18 & 3 & 0 & 0 & 0\\ \hline 5 & 1 & 5 & 20 & 80 & 280 & 680 & 1214 & 1127 & 389 & 40 & 4 & 0\\ \hline 6 & 1 & 6 & 30 & 150 & 675 & 2340 & 6604 & 12795 & 15519 & 6957 & 959 & 43\\ \hline 7 & 1 & 7 & 42 & 252 & 1386 & 6230 & 24024 & 71568 & 159326 & 222995 & 136301 & 21951\\ \hline 8 & 1 & 8 & 56 & 392 & 2548 & 14056 & 68656 & 276136 & 901970 & 2195663 & 3531887 & 2743477\\ \hline 9 & 1 & 9 & 72 & 576 & 4320 & 28224 & 166740 & 843822 & 3636954 & 12675375 & 33773653 & 60758618\\ \hline 10 & 1 & 10 & 90 & 810 & 6885 & 51960 & 359928 & 2193534 & 11738418 & 53257425 & 198586153 & \\ \hline 11 & 1 & 11 & 110 & 1100 & 10450 & 89430 & 710358 & 5060220 & 32328648 & 180577749 & & \\ \hline 12 & 1 & 12 & 132 & 1452 & 15246 & 145860 & 1306448 & 10645866 & 79016157 & & & \\ \hline 13 & 1 & 13 & 156 & 1872 & 21528 & 227656 & 2269410 & 20812077 & 175905015 & & & \\ \hline 14 & 1 & 14 & 182 & 2366 & 29575 & 342524 & 3760484 & 38319281 & 363216425 & & & \\ \hline 15 & 1 & 15 & 210 & 2940 & 39690 & 499590 & 5988892 & 67117596 & & & & \\ \hline 16 & 1 & 16 & 240 & 3600 & 52200 & 709520 & 9220512 & 112694400 & & & & \\ \hline 17 & 1 & 17 & 272 & 4352 & 67456 & 984640 & 13787272 & 182483644 & & & & \\ \hline 18 & 1 & 18 & 306 & 5202 & 85833 & 1339056 & 20097264 & 286341948 & & & & \\ \hline 19 & 1 & 19 & 342 & 6156 & 107730 & 1788774 & 28645578 & & & & & \\ \hline 20 & 1 & 20 & 380 & 7220 & 133570 & 2351820 & 40025856 & & & & & \\ \hline 21 & 1 & 21 & 420 & 8400 & 163800 & 3048360 & 54942566 & & & & & \\ \hline 22 & 1 & 22 & 462 & 9702 & 198891 & 3900820 & 74223996 & & & & & \\ \hline 23 & 1 & 23 & 506 & 11132 & 239338 & 4934006 & 98835968 & & & & & \\ \hline 24 & 1 & 24 & 552 & 12696 & 285660 & 6175224 & 129896272 & & & & & \\ \hline 25 & 1 & 25 & 600 & 14400 & 338400 & 7654400 & 168689820 & & & & & \\ \hline \end{tabular} } \caption{Numbers of the form $R_k^B(n)$ for several values of $n$ and $k$. In particular, notice that if $n\geq1$, then $R_4^B(n)=\frac{1}{2} n(n-1)^2(2n-3)$. The empty entries are unknown to us.} \label{tab:bn} \end{table} \section{Classification of the 9-cycles in the Burnt Pancake Graph}\label{sec:9cycles} In this section, we present classification of any 9-cycle in $BP_n$, with $n\geq2$. This presentation is in the same spirit as \cite{BBP2018,KM10,KM11,KonMed,KM16} where similar forms for 6-,7-,8- and 9-cycles in the pancake graphs $P_n$ and 8-cycles in the burnt pancake graph $BP_n$ are given. We start the description of $9$-cycles in $BP_n$, with $n\geq2$, by giving some preliminary definitions, notation, and lemmas. In classifying the 9-cycles we will look at decomposing the window notation of the a signed permutation $\sigma \in B_n$ into substrings, $\sigma = [X Y Z]$. A convenient notation that will be employed is for a signed reversal of a substring, that is if $X=[x_1\;x_2\;\cdots\;x_{i-1}\;x_i]$, then $\overline{X}=[\uline{x_i}\;\uline{x_{i-1}}\;\cdots\;\uline{x_2}\;\uline{x_1}]$. As is customary, for a graph $G$ we shall use $V(G)$ for its set of vertices and $E(G)$ for its set of edges. In this section, we generally follow the convention of using names of signed permutations based on the last character, e.g., $\pi \in V(BP_{n-1}(p))$, $\uline{\pi} \in V(BP_{n-1}(\uline{p}))$, $\rho \in V(BP_{n-1}(q))$, etc. In addition to Lemma~\ref{l:4.5}, a few other lemmas from \cite{BBP2018} will be necessary in the classification of all the 9-cycles in $BP_n$, with $n\geq3$. We recall that $BP_2$ is itself an 8-cycle~\cite[Theorem 10]{Compeau2011}, so if a 9-cycle exists in $BP_n$, then $n\geq3$. \begin{lem}[Lemma 4.2 in \cite{BBP2018}]\label{lem:31} If $\pi \in V(BP_{n-1}(p))$ and $\pi r^B_{n} \in V(BP_{n-1}(q))$, then $|p| \neq |q|$. \end{lem} Moreover, the following lemma is also used. \begin{lem}[Lemma 4.3 in \cite{BBP2018}]\label{lem:32} Let $\pi,\tau\in B_n$ have the same first element $q\in[\pm n]$ in window notation. Then $d(\pi,\tau) = 3$ if and only if $\tau = \pi r^B_{j} r^B_{i} r^B_{j}$, $1 \leq i < j \leq n$ where $\pi = [A B C]$, $\tau=[A \overline{B} C]$, $|A|=j-i$, $|B|=i$, and $|C| \geq 0$. \end{lem} We are now ready to state and prove the main result of this section, the classification of all the 9-cycles in $BP_n$, with $n\geq3$. \begin{thm}\label{t:canonical9cyclesB} If $n\geq3$, then the canonical forms of 9-cycles in $BP_n$ are as follows: \begin{align} \label{9-1} &r^B_{k}r^B_{k-i}r^B_{k}r^B_{k-j}r^B_{k-i-j}r^B_{k}r^B_{j}r^B_{i+j}r^B_{i} &\quad 1 \leq i,j \leq k-2, \quad i+j \leq k-1, \quad 3 \leq k \leq n;\\ \label{9-2} &r^B_{k}r^B_{i+j}r^B_{i}r^B_{k}r^B_{k-i}r^B_{j}r^B_{k}r^B_{k-j}r^B_{k-i-j} &\quad 1 \leq i,j \leq k-2, \quad i+j \leq k-1, \quad 3 \leq k \leq n. \end{align} \end{thm} \begin{proof} As mentioned before, $BP_k$ has a recursive structure and we can find $2k$ copies of $BP_{k-1}$ embedded into $BP_k$, with $3\leq k\leq n$. Recall that we use $BP_{k-1}(x)$, with $x\in[\pm k]$, to denote the subgraph isomorphic to $BP_{k-1}$ induced by looking at the vertices of $BP_k$ that end with the string $x\,(k+1)\,(k+2)\,\cdots\,n$. We will make use of this recursive structure to classify all 9-cycles in $BP_{n-1}(n)$, which is the copy including the identity, by considering the vertices of the 9-cycles in different copies of $BP_{k-1}$ embedded in $BP_k$. Due to the vertex transitive nature of $BP_{n}$, if there is a cycle $C$ in $BP_n$, there would be a cycle with the same labels as $C$ that includes the identity, and therefore it is enough to consider the cycles that contain the identity. Since each vertex in $BP_{k-1}(x)$ is connected to exactly one other vertex in some $BP_{k-1}(y)$, with $y\in[\pm k]\setminus\{x\}$, any 9-cycle will share at least two vertices with any copy of $BP_{k-1}$. We will identify a 9-cycle with a partition $(a_1+a_2+\cdots+a_m)$ of 9. That is, $a_1+a_2+\cdots+a_m=9$, where $a_i$ indicates the number of vertices in the $i${th} copy of $BP_{k-1}$ the cycle is incident upon. As noted above, $a_i \geq 2$ for all $i$. Thus a 9-cycle can be formed by using two, three, or four copies of $BP_{k-1}$. Enumerating through each possible partition will exhaust all possible 9-cycles. \begin{enumerate} \item[CASE I :-] A cycle incident upon two copies of $BP_{k-1}$. We know from Lemma \ref{l:4.5} that if two permutations $\pi_1$ and $\pi_2$ belong to the same copy of $BP_{k-1}$ and are at a distance of less than 3, then $\pi_1r^B_k$ and $\pi_2r^B_k$ belong to different copies of $BP_{k-1}$. Hence it is necessary that five vertices are in one copy and four vertices are in the other. \begin{figure} \centering \begin{tikzpicture}[scale=0.8] \draw (2,2) ellipse (1.2cm and 2cm); \draw (6,2) ellipse (1.2cm and 2cm); \draw node at (-0.2,2) {$BP_{k-1}(q)$}; \draw node at (8.2,2) {$BP_{k-1}(p)$}; \filldraw (2.15,3.50) circle (2pt) node[align=left,xshift = -0.5cm] {$\rho_{1}$}; \filldraw (1.85,2.75) circle (2pt) node[align=left,xshift = -0.5cm] {$\rho_{2}$}; \draw [thick] (2.15,3.50) -- (1.85,2.75); \filldraw (2.15,2.0) circle (2pt) node[align=left,xshift = -0.5cm] {$\rho_{3}$}; \draw [thick] (1.85,2.75) -- (2.15,2); \filldraw (1.85,1.25) circle (2pt) node[align=left,xshift = -0.5cm] {$\rho_{4}$}; \draw [thick] (2.15,2) -- (1.85,1.25); \filldraw (2.15,0.5) circle (2pt) node[align=left,xshift = -0.5cm] {$\rho_{5}$}; \draw [thick] (1.85,1.25) -- (2.15,0.5); \filldraw (5.75,0.50) circle (2pt) node[align=left,xshift = 0.5cm] {$\pi_{4}$}; \draw [thick] (2.15,0.50) -- (5.75,0.50); \filldraw (6.25,1.50) circle (2pt) node[align=left,xshift = 0.5cm] {$\pi_{3}$}; \draw [thick] (5.75,0.50) --(6.25,1.50); \filldraw (6.25,2.50) circle (2pt) node[align=left,xshift = 0.5cm] {$\pi_{2}$}; \draw [thick] (6.25,1.50) -- (6.25,2.50); \filldraw (5.75,3.50) circle (2pt) node[align=left,xshift = 0.5cm] {$\pi_{1}$}; \draw [thick] (6.25,2.50) -- (5.75,3.50); \draw [thick] (2.25,3.50) -- (5.75,3.50); \end{tikzpicture} \caption{A 9-cycle incident on two copies of $BP_{k-1}$ would need to be a (5+4) cycle.} \label{f:(5+4)} \end{figure} Let the two copies used be $BP_{k-1}(p)$ and $BP_{k-1}(q)$. By Lemma~\ref{lem:31}, $|p| \neq |q|$. So we may track the position and sign of both $p$ and $q$ in a every permutation of the cycle. Suppose that four vertices of such a 9-cycle belong to $BP_{k-1}(p)$, and the other five vertices belong to $BP_{k-1}(q)$ (see Figure~\ref{f:(5+4)}). The four vertices of $BP_{k-1}(p)$ form a path of length three whose endpoints are adjacent to vertices from $BP_{k-1}(q)$, which means both vertices should have $\uline{q}$ in their first positions. Starting with one of these vertices in $BP_{k-1}(p)$ we have the form $\pi_{1}=[\uline{q} X p]$. By Lemma~\ref{lem:32}, we can describe the forms of the remaining vertices of $BP_{k-1}(p)$. With $1 \leq i < j \leq k-1$ we have $\pi_{2}=\pi_{1}r^B_j=[\overline{X_1} q X_2 p]$, $\pi_{3}=\pi_{2}r^B_i=[X_{12} \overline{X_{11}} q X_2 p]$, $\pi_{4}=\pi_{3}r^B_j=[\uline{q} X_{11} \overline{X_{12}} X_2 p]$ where $X=X_1X_2$, $X_1 = X_{11}X_{12}$, $|X_1|=j-1$, and $|X_{12}|=i$. Continuing in $BP_{k-1}(q)$ we see $\rho_{1} = [\uline{p} \overline{X_2} \ \overline{X_{12}} \ \overline{X_{11}} q]$ and $\rho_{5} = [\uline{p} \overline{X_2} X_{12} \overline{X_{11}} q]$. Taking $A = \uline{p}X_2$ and $B=\overline{X_{11}}q$ it is clear that $|A|,|B|,|X_{12}| \geq 1$. We need a path of length four from $\rho_5=[A X_{12} B]$ to $\rho_1=[A \overline{X_{12}} B]$. $r^B_{|A|+|X_{12}|}r^B_{|X_{12}|}r^B_{|A|+|X_{12}|}$ is a path of length three. Thus if a path of length four existed, there would be a 7-cycle in $BP_{k-1}(q)$, which is not possible since the length of the smallest cycle in $BP_n$ is eight. Hence cycles of form (5+4) are not possible. \item[CASE II :-] A cycle incident upon three copies of $BP_{k-1}$. There can be three possibilities for the partition of vertices (5+2+2) or (4+3+2) or (3+3+3). Let the three copies incident upon be $BP_{k-1}(p)$, $BP_{k-1}(q)$, and $BP_{k-1}(s)$. By Lemma \ref{lem:31}, it follows that $|p|\neq |q|$, $|p| \neq |s|$, and $|q| \neq |s|$. \begin{figure} \centering \begin{tikzpicture}[scale=0.8] \draw (2,2) ellipse (1.2cm and 2cm); \draw node at (-0.2,2) {$BP_{k-1}(p)$}; \draw [rotate around={55:(5,3.2)}](5,3.2) ellipse (0.75cm and 1.45cm); \draw node at (7,3.5) {$BP_{k-1}(q)$}; \draw [rotate around={-55:(5,0.8)}](5,0.8) ellipse (0.75cm and 1.45cm); \draw node at (7,0.5) {$BP_{k-1}(s)$}; \filldraw (2.15,3.50) circle (2pt) node[align=left,xshift = -0.5cm] {$\pi_{1}$}; \filldraw (1.85,2.75) circle (2pt) node[align=left,xshift = -0.5cm] {$\pi_{2}$}; \draw [thick] (2.15,3.50) -- (1.85,2.75); \filldraw (2.15,2.0) circle (2pt) node[align=left,xshift = -0.5cm] {$\pi_{3}$}; \draw [thick] (1.85,2.75) -- (2.15,2); \filldraw (1.85,1.25) circle (2pt) node[align=left,xshift = -0.5cm] {$\pi_{4}$}; \draw [thick] (2.15,2) -- (1.85,1.25); \filldraw (2.15,0.5) circle (2pt) node[align=left,xshift = -0.5cm] {$\pi_{5}$}; \draw [thick] (1.85,1.25) -- (2.15,0.5); \filldraw (4.6,0.35) circle (2pt) node[align=left,xshift = 0.45cm] {$\sigma_{2}$}; \draw [thick] (2.15,0.50) -- (4.6,0.35); \filldraw (5.4,1.15) circle (2pt) node[align=left,xshift = 0.4cm] {$\sigma_{1}$}; \draw [thick] (4.6,0.35) -- (5.4,1.15); \filldraw (5.4,2.85) circle (2pt) node[align=left,xshift = 0.4cm] {$\rho_{2}$}; \draw [thick] (5.4,1.15) -- (5.4,2.85); \filldraw (4.6,3.65) circle (2pt) node[align=left,xshift = 0.45cm] {$\rho_{1}$}; \draw [thick] (5.4,2.85) -- (4.6,3.65); \draw [thick] (4.6,3.65) -- (2.15,3.50); \end{tikzpicture} \caption{A 9-cycle incident upon three copies of $BP_{k-1}$ with vertex partition (5+2+2).} \label{f:(5+2+2)} \end{figure} Suppose five vertices of such a 9-cycle belong to the copy $BP_{k-1}(p)$, two vertices belong to a copy $BP_{k-1}(q)$ and the other two vertices belong to a copy $BP_{k-1}(s)$ (see Figure~\ref{f:(5+2+2)}). As $\rho_{2}$ will have $\uline{s}$ in its first position and that $\pi_1$ is exactly two edges away, we see that $\pi_{1}=[\uline{q} X \uline{s} Y p]$. This gives $\rho_{1}=\pi_{1}r^B_k=[\uline{p} \overline{Y} s \overline{X} q]$, $\rho_{2}=\rho_{1}r^B_{|Y|+2}=[\uline{s} Y p \overline{X} q]$, $\sigma_{1}=\rho_{2}r^B_k=[\uline{q} X \uline{p} \overline{Y} s]$. $\sigma_{2}$ must have $\uline{p}$ in its first position, which is not possible in one edge from $\sigma_{1}$. Hence cycles of form (5+2+2) do not exist in the burnt pancake graph. \begin{figure} \centering \begin{tikzpicture}[scale=0.8] \draw (2,2) ellipse (1.2cm and 2cm); \draw node at (-0.2,2) {$BP_{k-1}(p)$}; \draw [rotate around={65:(5,3.2)}](5,3.2) ellipse (0.75cm and 1.05cm); \draw node at (7,3.5) {$BP_{k-1}(q)$}; \draw [rotate around={-55:(5,0.8)}](5,0.8) ellipse (0.95cm and 1.45cm); \draw node at (7.1,0.35) {$BP_{k-1}(s)$}; \filldraw (2.25,3.50) circle (2pt) node[align=left,xshift = -0.5cm] {$\pi_{1}$}; \filldraw (1.75,2.50) circle (2pt) node[align=left,xshift = -0.5cm] {$\pi_{2}$}; \draw [thick] (2.25,3.50) -- (1.75,2.50); \filldraw (1.75,1.50) circle (2pt) node[align=left,xshift = -0.5cm] {$\pi_{3}$}; \draw [thick] (1.75,2.50) -- (1.75,1.50); \filldraw (2.25,0.50) circle (2pt) node[align=left,xshift = -0.5cm] {$\pi_{4}$}; \draw [thick] (1.75,1.50) -- (2.25,0.50); \filldraw (4.3,0.2) circle (2pt) node[align=left,xshift = 0.45cm,yshift=-0.15cm] {$\sigma_{3}$}; \draw [thick] (2.25,0.50) -- (4.3,0.2); \filldraw (5.3,0.5) circle (2pt) node[align=left,xshift = 0.35cm,yshift = 0.1cm] {$\sigma_{2}$}; \draw [thick] (4.3,0.2) -- (5.3,0.5); \filldraw (5.5,1.5) circle (2pt) node[align=left,xshift = 0.35cm] {$\sigma_{1}$}; \draw [thick] (5.3,0.5) -- (5.5,1.5); \filldraw (5.4,2.85) circle (2pt) node[align=left,xshift = 0.3cm,yshift = 0.15cm] {$\rho_{2}$}; \draw [thick] (5.5,1.5) -- (5.4,2.85); \filldraw (4.6,3.65) circle (2pt) node[align=left,xshift = 0.35cm] {$\rho_{1}$}; \draw [thick] (5.4,2.85) -- (4.6,3.65); \draw [thick] (4.6,3.65) -- (2.25,3.50); \end{tikzpicture} \caption{A 9-cycle incident upon three copies of $BP_{k-1}$ with vertex partition (4+3+2).} \label{f:(4+3+2)} \end{figure} Suppose four vertices of such a 9-cycle belong to the copy $BP_{k-1}(p)$, two vertices belong to a copy $BP_{k-1}(q)$, and the other three vertices belong to a copy $BP_{k-1}(s)$ (see Figure~\ref{f:(4+3+2)}). As $\rho_{2}$ will have $\uline{s}$ in its first position and that $\pi_1$ is exactly two edges away, we see that $\pi_{1}=[\uline{q} X \uline{s} Y p]$. This gives $\rho_{1}=\pi_{1}r^B_k=[\uline{p} \overline{Y} s \overline{X} q]$, $\rho_{2}=\rho_{1}r^B_{|Y|+2}=[\uline{s} Y p \overline{X} q]$, $\sigma_{1}=\rho_{2}r^B_k=[\uline{q} X \uline{p} \overline{Y} s]$. Now, $\sigma_{3}$ must have $\uline{p}$ in its first position, so in the path from $\sigma_{1}$ to $\sigma_{3}$, $p$ should be involved in both the flips. $\sigma_{2}=\sigma_{1}r^B_{|X|+|Y_2|+2}=[Y_2 p \overline{X} q \overline{Y_1} s]$ where $Y=Y_1Y_2$. This gives $\sigma_{3}=\sigma_{2}r^B_{|Y_2|+1}=[\uline{p} \overline{Y_2}\ \overline{X} q \overline{Y_1} s]$, $\pi_{4}=\sigma_{3}r^B_k=[\uline{s} Y_1 \uline{q} X Y_2 p]$. Taking $A=\uline{s}Y_1$, $B=\uline{q}X$, $C=Y_2 p$ where $|A|,|B|,|C| \geq 1$ we need to find a path of length three between $\pi_4=[ABC]$ and $\pi_1=[BAC]$. One may verify that $\pi_4r^B_{|A|}r^B_{|A|+|B|}r^B_{|B|}=\pi_1$. Taking $|X|=i-1,|Y_1|=j-1$ we get $|Y_2|=k-i-j-1 \geq 0$ and a cycle corresponding to (\ref{9-1}). \begin{figure} \centering \begin{tikzpicture}[scale=0.8] \draw (2,2) ellipse (1.15cm and 1.5cm); \draw node at (-0.2,2) {$BP_{k-1}(p)$}; \draw [rotate around={65:(4.5,3.5)}](4.5,3.5) ellipse (0.95cm and 1.75cm); \draw node at (7,3.7) {$BP_{k-1}(q)$}; \draw [rotate around={-65:(4.5,0.5)}](4.5,0.5) ellipse (0.95cm and 1.75cm); \draw node at (7,0.35) {$BP_{k-1}(s)$}; \filldraw (2.15,3.0) circle (2pt) node[align=left,xshift = -0.35cm] {$\pi_{1}$}; \filldraw (1.65,2.0) circle (2pt) node[align=left,xshift = -0.35cm] {$\pi_{2}$}; \draw [thick] (2.15,3) -- (1.65,2); \filldraw (2.15,1) circle (2pt) node[align=left,xshift = -0.35cm] {$\pi_{3}$}; \draw [thick] (1.65,2) -- (2.15,1); \filldraw (3.8,-0.1) circle (2pt) node[align=left,xshift = 0.35cm,yshift=-0.15cm] {$\sigma_{1}$}; \draw [thick] (2.15,1) -- (3.8,-0.1); \filldraw (4.9,0.35) circle (2pt) node[align=left,xshift = 0.35cm] {$\sigma_{2}$}; \draw [thick] (3.8,-0.1) -- (4.9,0.35); \filldraw (5.4,1.2) circle (2pt) node[align=left,xshift = 0.3cm,yshift=-0.1] {$\sigma_{3}$}; \draw [thick] (4.9,0.35) -- (5.4,1.2); \filldraw (5.4,2.85) circle (2pt) node[align=left,xshift = 0.3cm] {$\rho_{3}$}; \draw [thick] (5.4,1.2) -- (5.4,2.85); \filldraw (4.8,3.85) circle (2pt) node[align=left,xshift = 0.35cm] {$\rho_{2}$}; \draw [thick] (5.4,2.85) -- (4.8,3.85); \filldraw (3.8,4) circle (2pt) node[align=left,yshift = 0.3cm] {$\rho_{1}$}; \draw [thick] (4.8,3.85) -- (3.8,4); \draw [thick] (3.8,4) -- (2.15,3.0); \end{tikzpicture} \caption{A 9-cycle incident upon three copies of $BP_{k-1}$ with vertex partition (3+3+3).} \label{f:(3+3+3)} \end{figure} Suppose three vertices of such a 9-cycle belong to a copy $BP_{k-1}(p)$, three vertices belong to a copy $BP_{k-1}(q)$ and the other three vertices belong to a copy $BP_{k-1}(s)$ (see Figure~\ref{f:(3+3+3)}). The vertex in $BP_{k-1}(p)$ that is adjacent to a vertex in $BP_{k-1}(q)$, $\pi_1$, can be of the form (a) $\pi_1=[\uline{q} X s Y p]$ or (b) $\pi_1=[\uline{q} X \uline{s} Y p]$. \begin{enumerate} \item Since $\pi_{1} = [\uline{q} X s Y p]$, then $\rho_{1}=\pi_{1}r^B_k=[\uline{p} \overline{Y} \uline{s} \overline{X} q]$. As $\pi_{3}$ must have $\uline{s}$ in its first position, $s$ should be involved in only one reversal in the path from $\pi_{1}$ to $\pi_{3}$. So the first reversal must not involve $s$. This gives $\pi_{2}=\pi_{1}r^B_{|X_1|+1}=[\overline{X_1} q X_2 s Y p]$ where $X=X_1X_2$, $\pi_{3}=\pi_{2}r^B_{|X|+2}=[\uline{s} \overline{X_2} \uline{q} X_1 Y p]$, and $\sigma_{1}=\pi_{3}r^B_k=[\uline{p} \overline{Y} \ \overline{X_1} q X_2 s]$. As $\sigma_{3}$ must have $\uline{q}$ in its first position, $q$ should be involved in one reversal in the path from $\sigma_{1}$ to $\sigma_{3}$. So the first reversal must not involve $q$. This gives two possibilities: \begin{enumerate} \item $\sigma_{2}=\sigma_{1}r^B_{|Y_2|+1}[Y_2 p \overline{Y_1} \ \overline{X_1} q X_2 s ]$ where $Y=Y_1Y_2$ and $\sigma_{3}=\sigma_{2}r^B_{|Y|+|X_1|+2} = [\uline{q} X_1 Y_1 \uline{p}$ $\overline{Y_2}X_2 s]$. Following through with this possibility we get $\rho_{3}=[\uline{s} \overline{X_2} Y_2 p \overline{Y_1}\ \overline{X_1} q]$. We need a path of length two from $\rho_{3}$ to $\rho_{1}=[\uline{p} \overline{Y_2}\ \overline{Y_1} \uline{s} \overline{X_2}\ \overline{X_1} q]$. As $\rho_{1}$ has $\uline{p}$ in its first position, $p$ should be involved in only the second reversal. In order for this to be so, without having to exchange the positions of $Y_1$, it must be that $|Y_1|=0$. Then we get $\rho_{3}r^B_{|X_2|+1}r^B_{|X_2|+|Y_2|+2}=\rho_{1}$. Taking $|X|=i-1$ and $|X_2|=j-1$ we get $|Y_2|=k-i-j-1 \geq 0$ and a cycle corresponding to (\ref{9-2}). \item $\sigma_{2}=\sigma_{1}r^B_{|Y|+|X_{12}|+1}=[X_{12} Y p \overline{X_{11}} q X_2 s]$ where $X_1=X_{11}X_{12}$ with $|X_{12}|\geq 1$ and $\sigma_{3}=\sigma_{2}r^B_{|Y|+|X_1|+2} = [\uline{q} X_{11} \uline{p} \overline{Y} p \overline{X_{12}} X_2 s]$. Following through with this possibility we get $\rho_{3}=[\uline{s} \overline{X_2} X_{12} Y p \overline{X_{11}} q]$. We need a path of length two from $\rho_{3}$ to $\rho_{1}=[\uline{p} \overline{Y}\ \uline{s} \overline{X_2} \ \overline{X_{12}}\ \overline{X_{11}}q]$. As $\rho_{1}$ has $\uline{p}$ in its first position, $p$ should be involved in only the second reversal. We have $\rho_{3}r^B_{|X_2|+|X_{12}|+1}r^B_{|X_{12}}|r^B_{|X_2|+|X_{12}|+|Y|+2}=\rho_{1}$. This is a path of length three which can be reduced to length two if and only if $|X_{12}|=0$. Since $|X_{12}| \geq 1$ this possibility does not give any 9-cycle. \end{enumerate} \item If $\pi_{1} = [\uline{q} X \uline{s} Y p]$, then $\rho_{1}=\pi_{1}r^B_k=[\uline{p} \overline{Y} s \overline{X} q]$. As $\pi_{3}$ must have $\uline{s}$ in its first position, $\uline{s}$ should be involved in both the reversals. This gives $\pi_{2}=\pi_{1}r^B_{|X|+|Y_1|+2}=[\overline{Y_1} s \overline{X} q Y_2 p]$ where $Y=Y_1 Y_2$, $\pi_{3}=\pi_{2}r^B_{|Y_1|+1}=[\uline{s} Y_1 \overline{X} q Y_2 p]$, and $\sigma_{1}=\pi_{3}r^B_k=[\uline{p} \overline{Y_2} \uline{q} X \overline{Y_1} s]$. As $\sigma_{3}$ must have $\uline{q}$ in its first position, $q$ should be involved in both the reversals in the path from $\sigma_{1}$ to $\sigma_{3}$. This gives two possibilities: \begin{enumerate} \item $\sigma_{2}=\sigma_{1}r^B_{|X_1|+|Y_2|+2}=[\overline{X_1} q Y_2 p X_2 \overline{Y_1}s]$ where $X=X_1X_2$ and $\sigma_{3}=\sigma_{2}r^B_{|X_1|+1}= [\uline{q} X_1 Y_2 p X_2 \overline{Y_1} s]$. Following through with this possibility we get $\rho_{3}=\sigma_{3}r^B_k=[\uline{s} Y_1 \overline{X_2} \uline{p}$ $\overline{Y_2}$ $\overline{X_1} q]$. We need a path of length two from $\rho_{3}$ to $\rho_{1}=[\uline{p}\overline{Y_2}\ \overline{Y_1} s \overline{X_2}\ \overline{X_1}q]$. As $\rho_{1}$ has $\uline{p}\overline{Y_2}$ in its prefix, $\overline{X_2}$ with the same sign and ordering of characters, but $s$ will have the opposite sign in $\rho_{3}$, then $\uline{p}\overline{Y_2}$ and $\overline{X_2}$ must be part of both reversals but $s$ must only be part of one. This is only possible if $|X_2|=0$. Then $\rho_{3}r^B_{|Y|+2}r^B_{|Y_2|+1}=\rho_{1}$. Taking $|X_1|=j-1$, $|Y_1|=i-1$ we get $|Y_2|=k-i-j-1$ and a cycle corresponding to (\ref{9-2}). \item $\sigma_{2}=\sigma_{1}r^B_{|X|+|Y_2|+|Y_{12}|+2}=[Y_{12}\overline{X}qY_2p\overline{Y_{11}}s]$ where $Y_1=Y_{11}Y_{12}$ with $|Y_{12}| \geq 1$ and $\sigma_{3}=\sigma_{2}r^B_{|X|+|Y_{12}|+1}= [\uline{q}X \overline{Y_{12}} Y_2 p\overline{Y_{11}}s]$. Following through with this possibility we get $\rho_{3}=\sigma_{3}r^B_k=[\uline{s}Y_{11}\uline{p}\overline{Y_2}Y_{12}\overline{X}q]$. We need a path of length two from $\rho_{3}$ to $\rho_{1}=[\uline{p} \overline{Y_2}\ \overline{Y_{12}}\ \overline{Y_{11}}s\overline{X}q]$. As $\rho_{1}$ has $\uline{p}$ in its first position, $p$ should be involved in both the reversals in this path. However, it can be seen that $\rho_{3}r^B_{|Y_{11}|+|Y_{12}|+|Y_2|+2}r^B_{|Y_{12}|}r^B_{|Y_{12}|+|Y_2|+1}=\rho_{1}$. This is a path of length three within $BP_{k-1}(q)$ which can be reduced to length two if and only if $|Y_{12}|=0$ but by assumption $|Y_{12}|\geq 1$. Therefore this possibility does not yield a 9-cycle. \end{enumerate} \end{enumerate} \begin{figure} \ContinuedFloat* \centering \begin{subfigure}{0.45\textwidth} \begin{tikzpicture}[scale=0.8] \draw (2,2) ellipse (0.9cm and 1cm); \draw node at (0.7,0.8) {$BP_{k-1}(p)$}; \draw (5,2) ellipse (0.9cm and 1cm); \draw node at (3.5,4.8) {$BP_{k-1}(q)$}; \draw (3.5,3.8) ellipse (1cm and 0.65cm); \draw node at (6.45,3.15) {$BP_{k-1}(s)$}; \draw (3.5,0.2) ellipse (1cm and 0.65cm); \draw node at (3.5,-0.8) {$BP_{k-1}(t)$}; \filldraw (2.10,2.6) circle (2pt) node[align=left,xshift = -0.35cm] {$\pi_{1}$}; \filldraw (1.8,2) circle (2pt) node[align=left,xshift = -0.35cm] {$\pi_{2}$}; \draw [thick] (2.1,2.6) -- (1.8,2); \filldraw (2.1,1.4) circle (2pt) node[align=left,xshift = -0.35cm] {$\pi_{3}$}; \draw [thick] (1.8,2.0) -- (2.1,1.4); \filldraw (3,0.2) circle (2pt) node[align=left,yshift = -0.25cm] {$\tau_{2}$}; \draw [thick] (2.1,1.4) -- (3,0.20); \filldraw (4,0.2) circle (2pt) node[align=left,yshift = -0.25cm] {$\tau_{1}$}; \draw [thick] (3,0.20) -- (4,0.2); \filldraw (5,1.5) circle (2pt) node[align=left,xshift = 0.35cm] {$\sigma_{2}$}; \draw [thick] (4,0.2) -- (5,1.50); \filldraw (5,2.5) circle (2pt) node[align=left,xshift = 0.35cm] {$\sigma_{1}$}; \draw [thick] (5,1.5) -- (5,2.5); \filldraw (4,3.8) circle (2pt) node[align=left,yshift = 0.25cm] {$\rho_{2}$}; \draw [thick] (5,2.5) -- (4,3.8); \filldraw (3,3.8) circle (2pt) node[align=left,yshift = 0.25cm] {$\rho_{1}$}; \draw [thick] (4,3.8) -- (3,3.8); \draw [thick] (3,3.8) -- (2.1,2.6); \end{tikzpicture} \caption{pairwise different absolute values.} \label{f:(3+2+2+2)d} \end{subfigure} \begin{subfigure}{0.45\textwidth} \begin{tikzpicture}[scale=0.8] \draw (2,2) ellipse (0.9cm and 1cm); \draw node at (0.7,0.8) {$BP_{k-1}(p)$}; \draw (5,2) ellipse (0.9cm and 1cm); \draw node at (3.5,4.8) {$BP_{k-1}(q)$}; \draw (3.5,3.8) ellipse (1cm and 0.65cm); \draw node at (6.45,3.15) {$BP_{k-1}(\uline{p})$}; \draw (3.5,0.2) ellipse (1cm and 0.65cm); \draw node at (3.5,-0.8) {$BP_{k-1}(t)$}; \filldraw (2.10,2.6) circle (2pt) node[align=left,xshift = -0.35cm] {$\pi_{1}$}; \filldraw (1.8,2) circle (2pt) node[align=left,xshift = -0.35cm] {$\pi_{2}$}; \draw [thick] (2.1,2.6) -- (1.8,2); \filldraw (2.1,1.4) circle (2pt) node[align=left,xshift = -0.35cm] {$\pi_{3}$}; \draw [thick] (1.8,2.0) -- (2.1,1.4); \filldraw (3,0.2) circle (2pt) node[align=left,yshift = -0.25cm] {$\tau_{2}$}; \draw [thick] (2.1,1.4) -- (3,0.20); \filldraw (4,0.2) circle (2pt) node[align=left,yshift = -0.25cm] {$\tau_{1}$}; \draw [thick] (3,0.20) -- (4,0.2); \filldraw (5,1.5) circle (2pt) node[align=left,xshift = 0.3cm] {$\uline{\pi_{2}}$}; \draw [thick] (4,0.2) -- (5,1.50); \filldraw (5,2.5) circle (2pt) node[align=left,xshift = 0.3cm] {$\uline{\pi_{1}}$}; \draw [thick] (5,1.5) -- (5,2.5); \filldraw (4,3.8) circle (2pt) node[align=left,yshift = 0.25cm] {$\rho_{2}$}; \draw [thick] (5,2.5) -- (4,3.8); \filldraw (3,3.8) circle (2pt) node[align=left,yshift = 0.25cm] {$\rho_{1}$}; \draw [thick] (4,3.8) -- (3,3.8); \draw [thick] (3,3.8) -- (2.1,2.6); \end{tikzpicture} \caption{equal absolute values non-adjacent.} \label{f:(3+2+2+2)3-2} \end{subfigure}\\ \caption{9-cycles incident upon four copies of $BP_{k-1}$ with vertex partition (3+2+2+2).} \end{figure} \item[CASE III :-] A cycle incident upon four copies of $BP_{k-1}$. Due to the constraints on the part sizes in our partition, there can be only one possibility (3+2+2+2). Let the four copies used be $BP_{k-1}(p)$, $BP_{k-1}(q)$, $BP_{k-1}(s)$, and $BP_{k-1}(t)$ with three vertices in $BP_{k-1}(p)$ and two vertices in each of the other copies. Let us assume, without loss of generality, that one vertex of $BP_{k-1}(p)$ is adjacent to a vertex of $BP_{k-1}(q)$. Here the absolute values of $p,q,s,$ and $t$ may not be distinct. By Lemma \ref{lem:31} only non-adjacent copies can have the same absolute value. This gives rise to three subcases, which we describe below. \begin{enumerate} \item The absolute values of $p,q,s$ and $t$ are pairwise different (see Figure~\ref{f:(3+2+2+2)d}). \begin{figure} \ContinuedFloat \begin{subfigure}{0.45\textwidth} \begin{tikzpicture}[scale=0.8] \draw (2,2) ellipse (0.9cm and 1cm); \draw node at (0.7,0.8) {$BP_{k-1}(p)$}; \draw (5,2) ellipse (0.9cm and 1cm); \draw node at (3.5,4.8) {$BP_{k-1}(q)$}; \draw (3.5,3.8) ellipse (1cm and 0.65cm); \draw node at (6.45,3.15) {$BP_{k-1}(s)$}; \draw (3.5,0.2) ellipse (1cm and 0.75cm); \draw node at (3.5,-0.8) {$BP_{k-1}(\uline{q})$}; \filldraw (2.10,2.6) circle (2pt) node[align=left,xshift = -0.35cm] {$\pi_{1}$}; \filldraw (1.8,2) circle (2pt) node[align=left,xshift = -0.35cm] {$\pi_{2}$}; \draw [thick] (2.1,2.6) -- (1.8,2); \filldraw (2.1,1.4) circle (2pt) node[align=left,xshift = -0.35cm] {$\pi_{3}$}; \draw [thick] (1.8,2.0) -- (2.1,1.4); \filldraw (3,0.2) circle (2pt) node[align=left,yshift = -0.25cm] {$\uline{\rho_{2}}$}; \draw [thick] (2.1,1.4) -- (3,0.20); \filldraw (4,0.2) circle (2pt) node[align=left,yshift = -0.25cm] {$\uline{\rho_{1}}$}; \draw [thick] (3,0.20) -- (4,0.2); \filldraw (5,1.5) circle (2pt) node[align=left,xshift = 0.35cm] {$\sigma_{2}$}; \draw [thick] (4,0.2) -- (5,1.50); \filldraw (5,2.5) circle (2pt) node[align=left,xshift = 0.35cm] {$\sigma_{1}$}; \draw [thick] (5,1.5) -- (5,2.5); \filldraw (4,3.8) circle (2pt) node[align=left,yshift = 0.25cm] {$\rho_{2}$}; \draw [thick] (5,2.5) -- (4,3.8); \filldraw (3,3.8) circle (2pt) node[align=left,yshift = 0.25cm] {$\rho_{1}$}; \draw [thick] (4,3.8) -- (3,3.8); \draw [thick] (3,3.8) -- (2.1,2.6); \end{tikzpicture} \caption{equal absolute values non-adjacent} \label{f:(3+2+2+2)2-2} \end{subfigure} \begin{subfigure}{0.45\textwidth} \begin{tikzpicture}[scale=0.8] \draw (2,2) ellipse (0.9cm and 1cm); \draw node at (0.7,0.8) {$BP_{k-1}(p)$}; \draw (5,2) ellipse (0.9cm and 1cm); \draw node at (3.5,4.8) {$BP_{k-1}(q)$}; \draw (3.5,3.8) ellipse (1cm and 0.65cm); \draw node at (6.45,3.15) {$BP_{k-1}(\uline{p})$}; \draw (3.5,0.2) ellipse (1cm and 0.75cm); \draw node at (3.5,-0.8) {$BP_{k-1}(\uline{q})$}; \filldraw (2.10,2.6) circle (2pt) node[align=left,xshift = -0.35cm] {$\pi_{1}$}; \filldraw (1.8,2) circle (2pt) node[align=left,xshift = -0.35cm] {$\pi_{2}$}; \draw [thick] (2.1,2.6) -- (1.8,2); \filldraw (2.1,1.4) circle (2pt) node[align=left,xshift = -0.35cm] {$\pi_{3}$}; \draw [thick] (1.8,2.0) -- (2.1,1.4); \filldraw (3,0.2) circle (2pt) node[align=left,yshift = -0.25cm] {$\uline{\rho_{2}}$}; \draw [thick] (2.1,1.4) -- (3,0.20); \filldraw (4,0.2) circle (2pt) node[align=left,yshift = -0.25cm] {$\uline{\rho_{1}}$}; \draw [thick] (3,0.20) -- (4,0.2); \filldraw (5,1.5) circle (2pt) node[align=left,xshift = 0.35cm] {$\uline{\pi_{2}}$}; \draw [thick] (4,0.2) -- (5,1.50); \filldraw (5,2.5) circle (2pt) node[align=left,xshift = 0.35cm] {$\uline{\pi_{1}}$}; \draw [thick] (5,1.5) -- (5,2.5); \filldraw (4,3.8) circle (2pt) node[align=left,yshift = 0.25cm] {$\rho_{2}$}; \draw [thick] (5,2.5) -- (4,3.8); \filldraw (3,3.8) circle (2pt) node[align=left,yshift = 0.25cm] {$\rho_{1}$}; \draw [thick] (4,3.8) -- (3,3.8); \draw [thick] (3,3.8) -- (2.1,2.6); \end{tikzpicture} \caption{two pairs of equal absolute values.} \label{f:(3+2+2+2)p} \end{subfigure} \caption{9-cycles incident upon four copies of $BP_{k-1}$ with vertex partition (3+2+2+2).} \end{figure} Since none are opposites, $p,q,s,$ and $t$, or their opposites, are present in all the signed permutations in the cycle. Then depending upon the relative position and signs of $s$ and $t$ in our first signed permutation four cases arise. Since $\rho_2$ is two reversals away from $\pi_1$ with one reversing all elements and $\rho_2$ must begin with $\uline{s}$ it is necessary that $\uline{s}$ be in $\pi_1$. The four possible cases of $\pi_1$ are $[\uline{q} X \uline{s} Y t Z p]$, $[\uline{q} X \uline{s} Y \uline{t} Z p]$, $[\uline{q} X t Y \uline{s} Z p]$, and $[\uline{q} X \uline{t} Y \uline{s} Z p]$. \begin{enumerate} \item If $\pi_{1}=[\uline{q} X \uline{s} Y t Z p]$, then $\rho_{1}=\pi_{1}r^B_k=[\uline{p} \overline{Z} \uline{t} \overline{Y} s \overline{X} q]$, $\rho_{2}=\rho_{1}r^B_{|Y|+|Z|+3}=[\uline{s} Y t Z p$ $\overline{X} q]$, and $\sigma_{1}=\rho_{2}r^B_k=[\uline{q} X \uline{p} \overline{Z} \uline{t} \overline{Y} s]$. Now $\sigma_{2}$ must have $\uline{t}$ in its first position, which is not possible in one reversal from $\sigma_{1}$. So this case does not yield any 9-cycles. \item If $\pi_{1}=[\uline{q} X \uline{s} Y \uline{t} Z p]$, then $\rho_{1}=\pi_{1}r^B_k=[\uline{p} \overline{Z} t \overline{Y} s \overline{X} q]$, $\rho_{2}=\rho_{1}r^B_{|Y|+|Z|+3}=[\uline{s} Y \uline{t} Z p$ $\overline{X} q]$, $\sigma_{1}=\rho_{2}r^B_k=[\uline{q} X \uline{p} \overline{Z} t \overline{Y} s]$, $\sigma_{2}=\sigma_{1}r^B_{|X|+|Z|+3}=[\uline{t} Z p \overline{X} q \overline{Y} s]$, and $\tau_{1}=\sigma_{2}r^B_k=[\uline{s} Y \uline{q} X \uline{p} \overline{Z} t]$. Now $\tau_{2}$ must have $\uline{p}$ in its first position, which is not possible in one reversal from $\tau_{1}$. So this case does not yield any 9-cycles. \item If $\pi_{1}=[\uline{q} X t Y \uline{s} Z p]$, then $\rho_{1}=\pi_{1}r^B_k=[\uline{p} \overline{Z} s \overline{Y} \uline{t} \overline{X} q]$, $\rho_{2}=\rho_{1}r^B_{|Z|+2}=[\uline{s} Z p \overline{Y} \uline{t} \overline{X} q]$, $\sigma_{1}=\rho_{2}r^B_k=[\uline{q} X t Y \uline{p} \overline{Z} s]$, $\sigma_{2}=\sigma_{1}r^B_{|X|+2}=[\uline{t} \overline{X} q Y \uline{p} \overline{Z} s]$, $\tau_{1}=\sigma_{2}r^B_k=[\uline{s} Z p \overline{Y} \uline{q} X t]$, $\tau_{2}=\tau_{1}r^B_{|Z|+2}=[\uline{p} \overline{Z} s \overline{Y} \uline{q} X t]$, and $\pi_{3}=\tau_{2}r^B_k=[\uline{t} \overline{X} q Y \uline{s} Z p]$. We need a path of length two from $\pi_{3}$ to $\pi_{1}$, however, $\pi_{3}r^B_{|X|+2}=\pi_{1}$ which is a path of length one. As there are no 3-cycles in the burnt pancake graph, a path of length two between $\pi_{3}$ and $\pi_{1}$ does not exist. Hence this case does not yield any 9-cycles. \item If $\pi_{1}=[\uline{q} X \uline{t} Y \uline{s} Z p]$, then $\rho_{1}=\pi_{1}r^B_k=[\uline{p} \overline{Z} s \overline{Y} t \overline{X} q]$, $\rho_{2}=\rho_{1}r^B_{|Z|+2}=[\uline{s} Z p \overline{Y} t \overline{X} q]$, and $\sigma_{1}=\rho_{2}r^B_k=[\uline{q} X \uline{t} Y \uline{p} \overline{Z} s]$. Now $\sigma_{2}$ must have $\uline{t}$ in its first position, which is not possible in one reversal from $\sigma_{1}$. So this case does not yield any 9-cycles. \end{enumerate} \item The absolute values of only one pair among $p,q,s,t$ are the same. This gives rise to two cases. One where the pair of copies with opposite signed last elements have only two vertices each. The other where the pair of copies with opposite signed last elements includes the one copy with three vertices. \begin{enumerate} \item Say that $s=\uline{p}$, $|q| \neq |t|$ (see Figure~\ref{f:(3+2+2+2)3-2}). In this case $\pi_{1}$ can be either $[\uline{q} X t Y p]$ or $[\uline{q} X \uline{t} Y p]$. If $\pi_{1}=[\uline{q} X t Y p]$, then $\rho_{1}=\pi_{1}r^B_k=[\uline{p} \overline{Y} \uline{t} \overline{X} q]$, $\rho_{2}=\rho_{1}r^B_1=[p \overline{Y} \uline{t} \overline{X} q]$, $\uline{\pi_{1}}=\rho_{2}r^B_k=[\uline{q} X t Y \uline{p}]$, $\uline{\pi_{2}}=\uline{\pi_{1}}r^B_{|X|+2}=[\uline{t} \overline{X} q Y \uline{p}]$, $\tau_{1}=\uline{\pi_{2}}r^B_k=[p \overline{Y} \uline{q} X t]$, $\tau_{2}=\tau_{1}r^B_1=[\uline{p} \overline{Y} \uline{q} X t]$, and $\pi_{3}=\tau_{2}r^B_k=[\uline{t} \overline{X} q Y p]$. We need a path of length two from $\pi_{3}$ to $\pi_{1}$ but $\pi_{3}r^B_{|X|+2}=\pi_{1}$ which is a path of length one. As there are no 3-cycles in the burnt pancake graph a path of length two between $\pi_{3}$ and $\pi_{1}$ does not exist. Hence this case does not yield any 9-cycle. If $\pi_{1}=[\uline{q} X \uline{t} Y p]$, then $\rho_{1}=\pi_{1}r^B_k=[\uline{p} \overline{Y} t \overline{X} q]$, $\rho_{2}=\rho_{1}r^B_1=[p \overline{Y} t \overline{X} q]$, $\uline{\pi_{1}}=\rho_{2}r^B_k=[\uline{q} X \uline{t} Y \uline{p}]$. As $\uline{\pi_{2}}$ must have $\uline{t}$ in its first position, which is not possible in one reversal from $\uline{\pi_{2}}$. So this case does not yield a 9-cycle. \item Say that $t=\uline{q}$, $|p| \neq |s|$ (see Figure~\ref{f:(3+2+2+2)2-2}). As $\rho_{2}$ will have $\uline{s}$ at first position $\pi_{1}=[\uline{q}X \uline{s}Y p]$. This gives $\rho_{1}=\pi_{1}r^B_k=[\uline{p}\overline{Y}s\overline{X}q]$, $\rho_{2}=\rho_{1}r^B_{|Y|+2}=[\uline{s}Y p\overline{X}q]$, $\sigma_{1}=\rho_{2}r^B_k=[\uline{q}X \uline{p} \overline{Y}s]$, $\sigma_{2}=\sigma_{1}r^B_1=[qX \uline{p} \overline{Y}s]$, $\uline{\rho_{1}}=\sigma_{2}r^B_k=[\uline{s}Y p\overline{X}\uline{q}]$, $\uline{\rho_{2}}=\uline{\rho_{1}}r^B_{|Y|+2}=[\uline{p}\overline{Y}s\overline{X}\uline{q}]$, $\pi_{3}=\uline{\rho_{2}}r^B_k=[qX \uline{s}Y p]$. We need a path of length two from $\pi_{3}$ to $\pi_{1}$ but $\pi_{3}r^B_{1}=\pi_{1}$ which is a path of length one. As there are no 3-cycles in burnt pancake graph a path of length two between $\pi_{3}$ and $\pi_{1}$ does not exist. Hence this case does not give any 9-cycle. \end{enumerate} \item The absolute values of the two pairs among $p,q,s,t$ are the same, i.e $s=\uline{p}$ and $t=\uline{q}$ (see Figure~\ref{f:(3+2+2+2)p}). Let $\pi_{1}=[\uline{q} X p]$. This gives $\rho_{1}=\pi_{1}r^B_k=[\uline{p} \overline{X} q]$, $\rho_{2}=\rho_{1}r^B_1=[p \overline{X} q]$, $\uline{\pi_{1}}=\rho_{2}r^B_k=[\uline{q} X \uline{p}]$, $\uline{\pi_{2}}=\uline{\pi_{1}}r^B_1=[q X \uline{p}]$, $\uline{\rho_{1}}=\uline{\pi_{2}}r^B_k=[p \overline{X} \uline{q}]$, $\uline{\rho_{2}}=\uline{\rho_{1}}r^B_1=[\uline{p} \overline{X} \uline{q}]$, and $\pi_{3}=\uline{\rho_{2}}r^B_k=[q X p]$. We need a path of length two from $\pi_{3}$ to $\pi_{1}$ but $\pi_{3}r^B_{1}=\pi_{1}$, which is a path of length one. As there are no 3-cycles in the burnt pancake graph a path of length two between $\pi_{3}$ and $\pi_{1}$ does not exist. Hence this case does not yield any 9-cycle. \end{enumerate} So there are no cycles of the form (3+2+2+2) in the burnt pancake graph. \end{enumerate} This finalizes all possible partitions of the vertices, and thus gives all possible 9-cycles in $BP_{k}$. \end{proof} \section{Concluding Remarks}\label{s:conclusion} In the preceding sections, we have provided explicit formulas for the number of pancake and burnt pancake stacks with $n$ pancakes that require four flips to be sorted, utilizing entirely elementary methods using cycle classification of the pancake and burnt pancake graph and the principle of inclusion-exclusion, as well as providing a classification of all 9-cycles in the burnt pancake graph. Having a classification of longer cycles in $P_n$ and $BP_n$ might allow us to take a similar approach to what we used in the proof of Theorem~\ref{t:snk=4} and Theorem~\ref{t:bsnk=4}. Currently, there is no published classification of the 10-cycles for either $P_n$ and $BP_n$. The authors have worked out said classification, though the number of canonical forms alone makes it hard to do a systematic, manual approach using PIE. For example, we found 59 canonical forms for 10-cycles in $P_n$ and 8 canonical forms for 10-cycles in $BP_n$. However, the process would be extremely tedious. Utilizing structural characteristics of permutations satisfying certain properties, Homberger and Vatter~\cite{HV16} gave an algorithm that can be used to enumerate certain permutations. In particular, they prove that for ``sufficiently large" $n$, the number of permutations of $[n]$ that can be sorted with $k$ prefix reversals is given by a polynomial. More specifically, if one defines \begin{equation}\label{eq:tilde} \widetilde{R}_k(n):=\sum_{i=0}^kR_i(n), \end{equation} then for sufficiently large $n$, $\widetilde{R}_k(n)$ is given by a polynomial. From (\ref{eq:tilde}) it is possible to compute $R_k(n)$ from $\widetilde{R}_k(n)$ if $R_i(n)$ is known for $0\leq i<k$. In this light, and using the output of the Homberger-Vatter algorithm from~\cite{HV16} for $\widetilde{R}_k(n)$ with $k=5,6$, one obtains the following. \begin{enumerate} \item If $n\geq 5$, then \begin{displaymath}R_5(n)= \frac{1}{6}\left(6n^5-65n^4+173n^3+296n^2-1724n+1590\right).\end{displaymath} \item If $n\geq6$, then \begin{displaymath}R_6(n)= \frac{1}{60}\left(60n^6 - 883n^5 + 3140n^4 + 10775 n^3 - 91400 n^2 + 171068n -58020\right).\end{displaymath} \end{enumerate} These polynomials explain all the nonzero values for the columns $k=5,6$ in Table~\ref{tab:sn}. The situation for $k>6$ is a bit more interesting, as the polynomial obtained by the Homberger-Vatter algorithm does not explain all nonzero values of $R_k(n)$. Indeed, their algorithm produces polynomials that are valid for ``sufficiently large" $n$. For example, the algorithm correctly computes $\widetilde{R}_7(n)$ if $n>7$. Indeed, if $n>7$, \begin{align}\label{eq:r7} R_7(n)&=\widetilde{R}_7(n)-\sum_{i=1}^6R_i(n)\notag\\ &=\frac{1}{240} (240 n^7- 4619 n^6+21881 n^5+ 109275 n^4 -1372445 n^3+\notag\\ &\hspace{5cm} 4476344 n^2-4550196 n-850320). \end{align} However, (\ref{eq:r7}) does not account for $R_7(6)=2$ and $R_7(7)=1016$. Therefore, there is no polynomial that would explain all the nonzero values of $R_7(n)$. After computing $\widetilde{R}_8(n)$ using the Homberger-Vatter algorithm, which took several days using a system with Dual Xeon CPUs and 256GB of RAM, we found a polynomial that explains most of the nozero entries of the column $k=8$ in Table~\ref{tab:sn}. Namely, if $n>7$, \begin{align}\label{eq:r8} R_8(n)&=\widetilde{R}_8(n)-\sum_{i=1}^7R_i(n)\notag\\ &=\frac{1}{5040}(5040 n^8 - 122683 n^7 + 759857 n^6 + 4519067 n^5 -79101715 n^4 +\notag\\ &\hspace{3cm} 364661948 n^3 - 561161062 n^2 - 267373812 n + 844945920). \end{align} Once again, (\ref{eq:r8}) does not account for $R_8(7)=35$, so once again, there is no polynomial that explains all the nonzero values of $R_8(k)$. The situation for $R^B_k(n)$ does not seem to be as mysterious. While the Homberger-Vatter algorithm does not apply to signed permutations, the numbers $R^B_k(n)$ seem to be given by polynomials. Using a standard polynomial fitting procedure with the data from Table~\ref{tab:bn}, we obtain the following conjecture. \begin{conjecture}\label{con:bn} If $n\geq1$, then \begin{enumerate} \item[(i)] $R^B_5(n)=\frac{1}{6}n(n-1)(n-2)(6n^2-17n+3)$, \item[(ii)] $R^B_6(n)=\frac{1}{60}n(n-1)(n-2)(60n^3-343n^2+401n+284)$, \item[(iii)] $R^B_7(n)=\frac{1}{240}n(n-1)(n-2)(n-3)(240n^3-1499n^2+925n+5104)$, \item[(iv)] $R^B_8(n)=\frac{1}{5040}n(n-1)(n-2)(n-3)(5040n^4-52123n^3+113415n^2+314716n-1027242)$, and \item[(v)] $R^B_9(n)=\frac{1}{40320}(n-1) (n-2) (n-3) (n-4) (40320 n^5-444061 n^4+644746 n^3+6638777 n^2-18991470 n).$ \end{enumerate} \end{conjecture} These polynomials also conveniently explain the zero entries of Table~\ref{tab:bn}, which seems to indicate that $R^B_k(n)$ are better behaved than those of $R_k(n)$. Another question that might give interesting results is proving any summation identities that $R_k(n)$ or $R^B_k(n)$ satisfy. From their definitions, it is clear that \[ \sum_{k\geq0}R_k(n)=|S_n|=n! \] and that \[ \sum_{k\geq0}R^B_k(n)=|B_n|=2^nn!. \] Notice that determining the maximum value of $k$, for a given $n$, such that $R_k(n)$ (or $R^B_k(n)$) is not zero is equivalent to the classic pancake problem (or the burnt pancake problem). A less trivial identity can be found by looking at the explicit formulas that describe the nonzero values of $R_k(n)$ for $0\leq k\leq 6$, namely, we have the following Corollary. \begin{cor}\label{cor:sum} If $k\leq6$ and $R_k(n-i)>0$ for $1\leq i\leq k+1$, then \[ R_k(n)=\sum_{i=1}^{k+1}(-1)^{i+1}\binom{k+1}{i}R_{k}(n-i). \] \end{cor} Since the polynomials (\ref{eq:r7}) and (\ref{eq:r8}) do not cover the values $R_7(6)$, $R_7(7)$, and $R_8(7)$, Corollary~\ref{cor:sum} would once again require $n$ to be ``sufficiently large.'' For signed permutations, we have observed the following identity, which would follow from $R^B_k(n)$ being integer-valued polynomials and by using the Gregory-Newton interpolation formula for integer-valued polynomials (for background see~\cite{CC97}). \begin{conjecture}\label{con:sum} If $k,n\geq1$, then \begin{displaymath}R_k^B(n) =\sum_{j=1}^k \left(\sum_{i=0}^{k-j} (-1)^i \binom{i+j-1}{i}\binom{n}{i+j}\right)R_k^B(j), \text{ for } k \geq 1.\end{displaymath} \end{conjecture} We have verified Conjecture~\ref{con:sum} for all the values in Table~\ref{tab:bn}. This conjecture would follow from a result similar to Homberger-Vatter giving that the $R^B_k(n)$ are all polynomials and continue to explain the zero and nonzero values. \section{Acknowledgments} A. Patidar was supported by the Global Talent Attraction Program (GTAP) administered by the School of Informatics, Computing, and Engineering at Indiana University. Furthermore, the authors thank V. Vatter for pointing out his and C. Homberger's paper~\cite{HV16} and for making their code publicly available. The authors also wish to thank the anonymous referees for their comments that helped improved the presentation of this paper. \bibliographystyle{alpha}
{ "timestamp": "2019-10-29T01:10:04", "yymm": "1902", "arxiv_id": "1902.04055", "language": "en", "url": "https://arxiv.org/abs/1902.04055", "abstract": "Using existing classification results for the 7- and 8-cycles in the pancake graph, we determine the number of permutations that require 4 pancake flips (prefix reversals) to be sorted. A similar characterization of the 8-cycles in the burnt pancake graph, due to the authors, is used to derive a formula for the number of signed permutations requiring 4 (burnt) pancake flips to be sorted. We furthermore provide an analogous characterization of the 9-cycles in the burnt pancake graph. Finally we present numerical evidence that polynomial formulae exist giving the number of signed permutations that require $k$ flips to be sorted, with $5\\leq k\\leq9$.", "subjects": "Discrete Mathematics (cs.DM); Combinatorics (math.CO)", "title": "On the number of pancake stacks requiring four flips to be sorted", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9898303398461494, "lm_q2_score": 0.8267118026095991, "lm_q1q2_score": 0.8183044245318822 }
https://arxiv.org/abs/1203.4809
The Effect of Coherence on Sampling from Matrices with Orthonormal Columns, and Preconditioned Least Squares Problems
Motivated by the least squares solver Blendenpik, we investigate three strategies for uniform sampling of rows from m x n matrices Q with orthonormal columns. The goal is to determine, with high probability, how many rows are required so that the sampled matrices have full rank and are well-conditioned with respect to inversion.Extensive numerical experiments illustrate that the three sampling strategies (without replacement, with replacement, and Bernoulli sampling) behave almost identically, for small to moderate amounts of sampling. In particular, sampled matrices of full rank tend to have two-norm condition numbers of at most 10.We derive a bound on the condition number of the sampled matrices in terms of the coherence {\mu} of Q. This bound applies to all three different sampling strategies; it implies a, not necessarily tight, lower bound of O(m {\mu} ln(n)) for the number of sampled rows; and it is realistic and informative even for matrices of small dimension and the stringent requirement of a 99 percent success probability.For uniform sampling with replacement we derive a potentially tighter condition number bound in terms of the leverage scores of Q. To obtain a more easily computable version of this bound, in terms of just the largest leverage scores, we first derive a general bound on the two-norm of diagonally scaled matrices.To facilitate the numerical experiments and test the tightness of the bounds, we present algorithms to generate matrices with user-specified coherence and leverage scores. These algorithms, the three sampling strategies, and a large variety of condition number bounds are implemented in the Matlab toolbox kappa_SQ_v3.
\section{Introduction} Our paper was inspired by Avron, Maymounkov and Toledo's \textsl{Blendenpik} algorithm and analysis \cite{AMTol10}. \textsl{Blendenpik} is an iterative method for solving overdetermined least squares/regres\-sion problems $\min_x{\|Ax-b\|_2}$ with the Krylov space method \textsl{LSQR} \cite{PS82}. In order to accelerate convergence, \textsl{Blendenpik} constructs a preconditioner $R_s$ and solves instead the preconditioned least squares problem $\min_z{\|AR_s^{-1}z-b\|_2}$. The solution to the original problem is recovered by solving a linear system with coefficient matrix~$R_s$. The innovative feature is the construction of the preconditioner $R_s$ by a random sampling method. \subsection{Motivation} The purpose of our paper is a thorough experimental and analytical investigation of random sampling strategies for producing efficient preconditioners. The challenge is to ensure not only that $R_s$ is nonsingular, but also that $AR_s^{-1}$ is well-conditioned with respect to inversion, which is required for fast convergence and numerical stability. Here is a conceptual point of view of how \textsl{Blendenpik} constructs the preconditioner: First it ``smoothes out'' the rows of $A$ by applying a randomized unitary transform $F$, and then it uniformly samples (i.e. selects) a small number of rows $M_s$ from $FA$. At last it computes a QR factorization of the smaller sampled matrix, $M_s=Q_sR_s$, where the triangular factor $R_s$ serves as the preconditioner. The neat and crucial observation in \cite{AMTol10} is to realize that sampling rows from $FA$ amounts, conceptually, to sampling rows from an orthonormal basis of $FA$. That is, if the columns of $Q$ represent an orthonormal basis for the column space of $FA$, and if $S$ is a sampling matrix then $SQ$ has the same two-norm condition number as $AR_s^{-1}$. This means, it suffices to consider sampling from matrices $Q$ with orthonormal columns. The analysis in \cite{AMTol10} suggests that $SQ$ is well conditioned, if $Q$ has low ``coherence''. Intuitively, coherence gives information about the localization or ``uniformity'' of the elements of $Q$. Mathematically, coherence is the largest (squared) norm of any row of $Q$. For instance, if $Q$ consists of canonical vectors, then the non-zero elements are concentrated in only a few rows, so that $Q$ has high coherence. However, if $Q$ is a submatrix of a Hadamard matrix, then all elements have the same magnitude, so that $Q$ has low coherence. If $Q$ has low coherence, then, in the context of sampling, all rows are equally important. Hence any sampled matrix $SQ$ with sufficiently many rows is likely to have full rank. The purpose of the randomized transform $F$ is to produce a matrix $FA$ whose orthonormal basis $Q$ has low coherence. We were intrigued by the analysis of \textsl{Blendenpik} because it appears to be the first to exploit the concept of coherence for numerical purposes. We also wanted to get a better understanding of the condition number bound for $SQ$ in \cite[Theorem 3.2]{AMTol10}, which contains an unspecified constant, and of the effect of uniform sampling strategies. \subsection{Overview and main results} We survey the contents of the paper, with a focus on the main results. \subsubsection*{From preconditioned matrices to sampled matrices with orthonormal columns (Section~\ref{s_bpik})} We start with a brief sketch of the \textsl{Blendenpik} least squares solver (Section~\ref{s_bpikalg}), and make the important transition from preconditioned matrices $AR_s^{-1}$ to sampled matrices $SQ$ with orthonormal columns, made possible by the observation (\cite{AMTol10,RokT08} and Lemma~\ref{l_1}) that both have the same two-norm condition number\footnote{Here $\kappa(X)\equiv \|X\|_2\>\|X^{\dagger}\|_2$ denotes the Euclidean two-norm condition number with respect to inversion of a full rank matrix $X$. The matrix $X^{\dagger}$ is the Moore-Penrose inverse of $X$.}, $$\kappa(AR_s^{-1})=\kappa(SQ).$$ Then we discuss the notion of coherence and its properties (Section~\ref{s_coh}). For a $m\times n$ matrix $Q$ with orthonormal columns, $Q^TQ=I_n$, the \textit{coherence} $$\mu\equiv\max_{1\leq j\leq m}{\|e_j^TQ\|_2^2}$$ is the largest squared row norm\footnote{The superscript $T$ denotes transpose, and $I_n$ is the $n\times n$ identity matrix with columns $e_j$.}. \subsubsection*{Sampling methods (Section~\ref{s_samp})} We discuss three randomized methods for producing sampling matrices~$S$: Sampling without replacement (Section~\ref{s_without}), sampling with replacement (Section~\ref{s_with}), and Bernoulli sampling (Section~\ref{s_bernoulli}). We show that Bernoulli sampling can be viewed as a form of sampling without replacement (Section~\ref{s_relate}). The sampling matrices $S$ from all three methods are constructed so that $S^TS$ is an unbiased estimator of the identity matrix. The action of applying $S$ to a matrix $Q$ with orthonormal columns, $SQ$, amounts to randomly sampling rows from~$Q$. The numerical experiments (Section~\ref{s_sampcomp}) illustrate two points: First, the three sampling methods behave almost identically, in terms of the percentage of sampled matrices $SQ$ that have full rank and their condition numbers, in particular for small to moderate sampling amounts. Second, those sampled matrices $SQ$ that have full rank tend to be very well-conditioned, with condition numbers $\kappa(SQ)\leq 10$. As a consequence (Section~\ref{s_sampconc}), we recommend sampling with replacement for \textsl{Blendenpik}, because it is fast, and it is easy to implement. \subsubsection*{Numerical experiments} Since random sampling methods can be expected to work well in the asymptotic regime of very large matrix dimensions, we restrict all numerical experiments to matrices of small dimension. Furthermore, we consider only matrices that have many more rows than columns, $m\gg n$. This is the situation where random sampling methods can be most efficient. In contrast, random sampling methods are not efficient for matrices that are almost square, because the number of rows in $SQ$ has to be at least equal to $n$, otherwise $\rank(SQ)=n$ is not possible. \subsubsection*{Condition number bounds based on coherence (Section~\ref{s_cohbound})} We derive a probabilistic bound, in terms of coherence, for the condition numbers of the sampled matrices (Theorem~\ref{t_cohbound} in Section~\ref{s_subcohbound}). The bound applies to all three sampling methods. From this we derive the following lower bound, not necessarily tight, on the required number of sampled rows. \begin{preview}[Corollary~\ref{c_cohbound}] Given a failure probability $0<\delta<1$, and a tolerance $0\leq \epsilon<1$. To achieve the condition number bound $\kappa(SQ)\leq \sqrt{\tfrac{1+\epsilon}{1-\epsilon}}$, the number of rows from $Q$, sampled by any of three methods, should be at least \begin{eqnarray}\label{e_cohbound} c\geq 3m\mu\>\frac{\ln(2n/\delta)}{\epsilon^2}. \end{eqnarray} \end{preview} This suggests that one has to sample more rows for $SQ$ if $Q$ has high coherence ($\mu$ close to 1), if one wants a low condition number bound (small $\epsilon$), or if one wants a high success probability (small $\delta$). Numerical experiments (Section~\ref{s_cohex}) illustrate that the bounds are informative for matrices with sufficiently low coherence $\mu$ and sufficiently high aspect ratio $m/n$. Our bounds have the following advantages (Section~\ref{s_cohconc}): \begin{enumerate} \item They are tighter than those in \cite[Theorem 3.2]{AMTol10} because they are non-asymp\-totic, with all constants explicitly specified. \item They apply to three different sampling methods. \item They imply a lower bound, of $\Omega\left(m\mu\ln{n}\right)$, on the required number of sampled rows. \item They are realistic and informative -- even for matrices of small dimension and the stringent requirement of a 99 percent success probability. \end{enumerate} \subsubsection*{Condition number bounds based on leverage scores, for uniform sampling with replacement (Section~\ref{s_leverage})} The goal is to tighten the coherence-based bounds from Section~\ref{s_cohbound} by making use of all the row norms of $Q$, instead of just the largest one. To this end we introduce \textit{leverage scores} (Section~\ref{s_levscore}), which are the squared row norms of $Q$, $$\ell_j=\|e_j^TQ\|_2^2, \qquad 1\leq j\leq m.$$ We use them to derive a bound for uniform sampling with replacement (Theorem~\ref{t_levbound} in Section~\ref{s_levbound}). Then we present a more easily computable bound, in terms of just a few of the largest leverage scores (Section~\ref{s_clevbound}). It implies the following lower bound, not necessarily tight, on the number of samples. \begin{preview}[Corollary~\ref{c_levbound}] Given a failure probability $0<\delta<1$, a tolerance $0\leq \epsilon<1$, and a labeling of leverage scores in non-increasing order, $$\mu=\ell_{[1]}\geq \cdots\geq \ell_{[m]}.$$ To achieve the condition number bound $\kappa(SQ)\leq \sqrt{\tfrac{1+\epsilon}{1-\epsilon}}$, the number of rows from $Q$, sampled uniformly with replacement, should be at least \begin{eqnarray}\label{e_lsbound} c\geq \tfrac{2}{3}m\>(3\tau+\epsilon\mu)\> \frac{\ln(2n/\delta)}{\epsilon^2}, \end{eqnarray} where $t\equiv\left\lfloor 1/\mu\right\rfloor$ and $\tau\equiv \mu\>\sum_{j=1}^t{\ell_{[j]}}+(1-t\,\mu)\,\ell_{[t+1]}$. \end{preview} \medskip We show (Section~\ref{s_ancomp}) that (\ref{e_lsbound}) is indeed tighter than (\ref{e_cohbound}). This is confirmed by numerical experiments (Section~\ref{s_numcomp}). The difference becomes more drastic for matrices $Q$ with widely varying non-zero leverage scores, and can be as high as ten percent. Hence (Section~\ref{s_lqconc}), when it comes to lower bounds for the number of rows sampled uniformly with replacement, we recommend (\ref{e_lsbound}) over~(\ref{e_cohbound}). \subsubsection*{Algorithms for generating matrices with prescribed coherence and leve\-rage scores (Section~\ref{s_genalg})} The purpose is to make it easy to investigate the efficiency of the sampling methods in Section~\ref{s_samp}, and test the tightness of the bounds in Sections \ref{s_cohbound} and~\ref{s_leverage}. To this end we present algorithms for generating matrices with prescribed leverage scores and coherence (Section~\ref{s_alg}), and for generating particular leverage score distributions with prescribed coherence (Section~\ref{s_precoh}). Furthermore we present two classes of structured matrices with prescribed coherence that are easy and fast to generate (Section~\ref{s_mat}). The basis for the algorithms is the following majorization result. \begin{preview}[Theorem~\ref{t_maj}] Given integers $m\geq n$ and a vector $\ell$ with $m$ elements that satisfy $0\leq \ell_j \leq 1$ and $\sum_{j=1}^m{\ell_j}=n$, there exists a $m\times n$ matrix $Q$ with orthonormal columns that has leverage scores $\|e_j^TQ\|_2^2=\ell_j$, $1\leq j\leq m$, and coherence $\mu=\max_{1\leq j\leq m}{\ell_j}$. \end{preview} \subsubsection*{Bound for two-norms of diagonally scaled matrices (Section~\ref{s_app2})} The bound (\ref{e_lsbound}) is based on a special case of the following general bound for the two-norm of diagonally scaled matrices. \begin{preview}[Theorem~\ref{t_2}] Let $Z$ be a $m\times n$ matrix with $\rank(Z)=n$ and largest squared row norm $\mu_z\equiv\max_{1\leq j\leq m}{\|e_j^TZ\|_2^2}$. Let $D$ be a $m\times m$ non-negative diagonal matrix, and a labeling of diagonal elements in non-increasing order, $$\|D\|_2=d_{[1]}\geq \cdots\geq d_{[m]}\geq 0.$$ If $t\equiv\left\lfloor (\|Z^{\dagger}\|_2^2\>\mu_z)^{-1}\right\rfloor$, then either $$\|DZ\|_2^2\leq \mu_z\sum_{j=1}^t{d_{[j]}^2}+ \left(\|Z\|_2^2-t\,\mu_z\right)\,d_{[t+1]}^2\qquad \text{ if } \|Z\|_2^2-t\,\mu_z\leq \mu_z$$ or $$\|DZ\|_2^2\leq \mu_z\sum_{j=2}^{t+1}{d_{[j]}^2}+ \left(\|Z\|_2^2-t\,\mu_z\right)\,d_{[1]}^2.\qquad \text{ if } \|Z\|_2^2-t\,\mu_z > \mu_z.$$ \end{preview} \subsubsection*{Matlab toolbox} In order to perform the experiments in this paper, we developed a Matlab toolbox \texttt{kappaSQ\_v3} with a user-friendly interface \cite{impl}. The toolbox contains implementations of the three random sampling methods in Section~\ref{s_samp}, the matrix generation algorithms in Section~\ref{s_genalg}, the bounds in Sections \ref{s_cohbound} and~\ref{s_leverage}, and a variety of other condition number bounds. It also allows the user to input her/his own matrices. \subsubsection*{Proofs (Sections \ref{s_app}, \ref{s_app2} and~\ref{s_pre})} All proofs, except those for Sections \ref{s_bpik} and~\ref{s_samp}, have been relegated to these three sections, which form the appendix. Section~\ref{s_app} contains the proofs for Sections \ref{s_cohbound} and~\ref{s_leverage}, which are based on two matrix concentration inequalities: A Chernoff bound (Section~\ref{s_chernoff}), and a Bernstein bound (Section~\ref{s_bernstein}). Section~\ref{s_app2} contains the proofs for the easily computable bounds in Sections \ref{s_clevbound} and~\ref{s_ancomp}, together with the majorization results (Section~\ref{s_tns}) required for the proofs. The majorization results in Section~\ref{s_pre} represent the foundation for the algorithms in Section~\ref{s_genalg}. \subsubsection*{Future work (Section~\ref{s_future})} We list a few issues that suggest themselves immediately as a follow up to this paper. \subsection{Literature} Existing randomized least squares methods are based on randomized projections. This means, conceptually they multiply $A$ by a random matrix~$F$, and then sample a few rows from $FA$. The algorithms in \cite{BDr09,DMM06a,DMMS10} solve a smaller sampled problem by a direct method. Like \textsl{Blendenpik} \cite{AMTol10}, the algorithm in \cite{RokT08} computes a preconditioner from the QR factorization of a sampled submatrix, but then solves the preconditioned problem by applying the conjugate gradient method to the normal equations. The parallel solver \textsl{LSRN} \cite{MSM11} computes a preconditioner from the SVD of a sampled submatrix, and then solves the preconditioned problem with an iterative method. This solver applies to general matrices rather than just those of full column rank. As for randomized algorithms in general, the excellent surveys \cite{HMT09,Mah11} provide clear analyses and good intuition. \subsection{Notation} The norm $\|\cdot\|_2$ denotes the Euclidean two-norm, and the two-norm \textit{condition number} with respect to inversion of a real $m\times n$ matrix $Z$ with $\rank(Z)=n$ is denoted by $\kappa(Z)\equiv\|Z\|_2\|Z^{\dagger}\|_2$, where $Z^{\dagger}$ is the Moore-Penrose inverse. The $k\times k$ \textit{identity matrix} is $I_k=\begin{pmatrix}e_1 &\ldots &e_k\end{pmatrix}$, and its columns are the \textit{canonical vectors} $e_j$, $1\leq j\leq k$. The \textit{probability} of an event $\mathcal{X}$ is denoted by $\mathbf{Pr}[\mathcal{X}]$, and the \textit{expected value} of a random variable $X$ is denoted by $\mathbf{E}[X]$. \section{Existence of matrices with prescribed coherence and leverage scores}\label{s_pre} This section is the basis for Algorithm~\ref{alg_gen}. We review a well-known majorization result (Theorem~\ref{t_horn}). We use it to show (Theorem~\ref{t_maj}) that, given prescribed matrix dimensions and leverage scores, there always exists a matrix~$Q$ with orthonormal columns that has the required dimensions and (squared) row norms equal to the leverage scores. Our approach is again based on majorization, see Definition~\ref{d_maj}, and in particular on the fact that the eigenvalues of a real symmetric matrix majorize its diagonal elements. \begin{theorem}[Theorem 4.3.48 in \cite{HoJ12}]\label{t_horn} Let $a$ and $\lambda$ be vectors with real elements $a_j$ and $\lambda_j$, respectively, $1\leq j\leq m$. If $\lambda$ majorizes $a$, then there exists a $m\times m$ real symmetric matrix with eigenvalues $\lambda_j$ and diagonal elements $a_j$, $1\leq j\leq m$. \end{theorem} With the help of Theorem~\ref{t_horn} we show that there exists a matrix with orthonormal columns that has prescribed leverage scores and coherence. \begin{theorem}\label{t_maj} Given integers $m$ and $n$ with $m\geq n\geq 1$; and a vector $\ell$ with $m$ elements $\ell_j$ that satisfy $0\leq \ell_j \leq 1$ and $\sum_{j=1}^m{\ell_j}=n$. Then there exists a $m\times n$ matrix $Q$ with orthonormal columns that has leverage scores $\|e_j^TQ\|_2^2=\ell_j$, $1\leq j\leq m$, and coherence $\mu=\max_{1\leq j\leq m}{\ell_j}$. \end{theorem} \begin{proof} Let $\lambda$ be a vector with $m$ elements that satisfy $\lambda_j=1$ for $1\leq j\leq n$, and $\lambda_j=0$ for $n+1\leq j\leq m$. We are going to construct a matrix $Q$ by applying Theorem~\ref{t_horn} to $\lambda$ and $\ell$. To this end, we first need to show that $\lambda$ majorizes $\ell$. \paragraph{Majorization} We distinguish the cases $1\leq k\leq n$ and $n+1\leq k\leq m$. \begin{description} \item[Case $1\leq k\leq n$:\ ] From $\ell_j\leq 1$ follows $$\sum_{j=1}^{k}{\lambda_j}=k\geq \sum_{j=1}^k{\ell_{[j]}}.$$ \item[Case $n+1\leq k\leq m$:\ ] From $\ell_j\geq 0$ and $\sum_{j=1}^m{\ell_j}=n$ follows $$\sum_{j=1}^{k}{\lambda_j}=n=\sum_{j=1}^{k}{\ell_{[j]}} + \sum_{j=k+1}^{m}{\ell_{[j]}} \geq \sum_{j=1}^{k}{\ell_{[j]}}.$$ \end{description} Hence $$\sum_{j=1}^{k}{\lambda_j}\geq \sum_{j=1}^k{\ell_{[j]}}, \qquad 1\leq k\leq m,$$ which means that $\lambda$ weakly majorizes $\ell$. Since also $\sum_{j=1}^m{\lambda_j}=n=\sum_{j=1}^{m}{\ell_{[j]}}$, we can conclude that $\lambda$ majorizes $\ell$. \paragraph{Construction of $Q$} Theorem \ref{t_horn} implies that there exists a real symmetric matrix $W$ with eigenvalues $\lambda_j$ and diagonal elements $W_{jj}=\ell_j$, $1\leq j\leq m$. Since $W$ has $n$ eigenvalues equal to one, and all other eigenvalues equal to zero, it has an eigenvalue decomposition $$W=\hat{Q}\begin{pmatrix}I_n &0\\ 0&0\end{pmatrix}\hat{Q}^T=QQ^T,$$ where $\hat{Q}$ is a $m\times m$ real orthogonal matrix, and $Q\equiv\hat{Q}\begin{pmatrix}I_n&0\end{pmatrix}^T$ has $n$ orthonormal columns. Therefore $Q$ has leverage scores $\|e_j^TQ\|_2^2=e_j^TQQ^Te_j=W_{jj}=\ell_j$ and coherence $\mu=\max_{1\leq j\leq m}{\ell_j}$. \end{proof} \section{The Blendenpik algorithm, and coherence}\label{s_bpik} We describe the \textsl{Blendenpik} algorithm for solving least squares problems (Section~\ref{s_bpikalg}), and present the notion of coherence (Section~\ref{s_coh}). \subsection{Algorithm}\label{s_bpikalg} The \textsl{Blendenpik} algorithm \cite[Algorithm 1]{AMTol10} solves full column rank least squares problems with the Krylov space method \textsl{LSQR} \cite{PS82} and a randomized preconditioner. Algorithm~\ref{alg_bpik} presents a conceptual sketch of \textsl{Blendenpik}. The subscript ``$s$'' denotes quantities associated with the sampled matrix. \begin{algorithm} \caption{Sketch of \textsl{Blendenpik} \cite{AMTol10}}\label{alg_bpik} \begin{algorithmic} \REQUIRE $m\times n$ matrix $A$ with $m\geq n$ and $\rank(A)=n$, $m\times 1$ vector $b$\\ $\qquad$ $m\times m$ random unitary matrix $F$\\ $\qquad$ $k\times n$ sampling matrix $S$ with $k\geq n$\\ \ENSURE Solution of $\min_x{\|Ax-b\|_2}$\\ $\qquad$\\ \STATE $M=FA$ $\qquad$ \COMMENT{Improve coherence}\\ \STATE $M_s=SM$ $\qquad$ \COMMENT{Sample for preconditioner}\\ \STATE Thin QR factorization $M_s=Q_sR_s$ $\qquad$ \COMMENT{Generate preconditioner}\\ \STATE Determine solution $y$ to $\min_z{\|AR_s^{-1}z-b\|_2}$ $\qquad$ \COMMENT{Solve preconditioned problem}\\ \STATE Solve $R_s\hat{x}=y$ $\qquad$ \COMMENT{Recover solution to original problem} \end{algorithmic} \end{algorithm} The matrix $F$ is the product of a random diagonal matrix with $\pm 1$ entries, and a unitary transform, such as a Walsh Hadamard transform, or a discrete Fourier, Hartley or cosine transform \cite[Section 3.2]{AMTol10}. The transformed matrix $M=FA$ is $m\times n$ with $m\geq n$ and $\rank(M)=n$. The sampling matrix $S$ selects $k\geq n$ rows from the transformed matrix $M$. We discuss different types of sampling matrices in Section~\ref{s_samp}. The $k\times n$ sampled matrix $M_s$ has a thin QR decomposition $M_s=Q_sR_s$ where $Q_s$ is $k\times n$ with orthonormal columns and $R_s$ is $n\times n$ upper triangular. The basis for the analysis is the thin QR decomposition $M=QR$, where $Q$ is $m\times n$ with orthonormal columns and $R$ is $n\times n$ upper triangular. This QR decomposition is \textit{not} computed. The next result links the condition number of the preconditioned matrix to that of the matrix $SQ$, see also \cite[Section~3.1]{AMTol10} and \cite[Theorem 1]{RokT08}. \begin{lemma}\label{l_1} With the notation in Algorithm~\ref{alg_bpik}, if $\rank(M_s)=n$, then $$\kappa(AR_s^{-1})=\kappa(SQ).$$ \end{lemma} \begin{proof} From $FA=M=QR$ and the fact that the 2-norm is invariant under premultiplication by matrices with orthonormal columns, it follows that \begin{eqnarray*} \kappa(AR_s^{-1})&=&\kappa(MR_s^{-1})=\kappa(RR_s^{-1}) =\kappa(R_sR^{-1})=\kappa(M_sR^{-1})=\kappa(SMR^{-1})\\ &=&\kappa(SQ). \end{eqnarray*} \end{proof} In Sections \ref{s_cohbound} and~\ref{s_leverage} we derive bounds for the condition number of the preconditioned matrix, $\kappa(AR_s^{-1})$. Our bounds are tighter than those in \cite[Theorem 3.2]{AMTol10}, because they have all constants explicitly specified, and apply to three different sampling strategies. Since Lemma~\ref{l_1} implies $\kappa(AR_s^{-1})=\kappa(SQ)$, we state the bounds for $\kappa(SQ)$ only. An important ingredient in these bounds is the coherence of~$Q$. \subsection{Coherence}\label{s_coh} Coherence gives information about the localization or ``uniformity'' of the elements in an orthonormal basis. The more general concept of \textit{mutual coherence} between two orthonormal bases was introduced in \cite[\S VII]{DoH01}, in the context of signal processing and computational harmonic analysis, to describe a condition for the existence of sparse representations of signals. What we use here is a special case, and can be viewed as a measure for how close an orthonormal basis is to sharing a vector with a canonical basis. \begin{definition}[Definition 3.1 in \cite{AMTol10}, Definition 1.2 in \cite{CanR09}]\label{d_co} Let $Q$ be a real $m\times n$ matrix with orthonormal columns, $Q^TQ=I_n$, then the {\rm coherence} of $Q$ is \begin{eqnarray*}\label{e_coherence} \mu\equiv \max_{1\leq j\leq m}\|e_j^TQ\|_2^2. \end{eqnarray*} If the columns of $Q$ are an orthonormal basis for the column space of a matrix~$M$, then the coherence of $M$ is $\mu$. \end{definition} The second part of Definition~\ref{d_co} emphasizes that coherence is really a property of the column space, hence basis-independent. In other words, if $\hat{Q}=QV$, where $V$ is a real $n\times n$ orthogonal matrix, then $\hat{Q}$ and $Q$ have the same coherence. The range for coherence is $\tfrac{n}{m}\leq \mu\leq 1$. If $Q$ is a $m\times n$ submatrix of the $m\times m$ Hadamard matrix, then $\mu=n/m$. If a column of $Q$ is a canonical vector, then $\mu=1$. Hence an orthonormal basis has high coherence if it shares a vector with a canonical basis. There are other definitions of coherence that differ from the above by factors depending on the matrix dimensions \cite[Definition 1]{Recht11}, \cite[Definition 1]{TalR10}. However, the notion of \textit{statistical coherence} in Bayesian analysis \cite{Lind78} appears to be unrelated. \section{Sampling Methods}\label{s_samp} We present three different types of sampling methods: Sampling without replacement (Section~\ref{s_without}), sampling with replacement (Section~\ref{s_with}), and Bernoulli sampling (Section~\ref{s_bernoulli}). We show that Bernoulli sampling can be viewed as a form of sampling without replacement (Section~\ref{s_relate}). The numerical experiments illustrate that there is little difference among the three methods for small to moderate amounts of sampling (Section~\ref{s_sampcomp}). Hence we recommend sampling with replacement for Algorithm~\ref{alg_bpik} (Section~\ref{s_sampconc}). The sampling matrices $S$ in all three methods are scaled so that $S^TS$ is an unbiased estimator of the identity matrix. \subsection{Sampling without replacement}\label{s_without} The obvious sampling strategy, in Algorithm~\ref{alg_without}, picks the requested number of rows, so that the sampling matrix~$S$ is just a scaled submatrix of a permutation matrix. Uniform sampling without replacement can be implemented via \textit{random permutations}\footnote{We thank an anonymous reviewer for this advice.}. A permutation $\pi_1,\ldots, \pi_m$ of the integers $1,\ldots, m$ is a \textit{random permutation}, if it is equally likely to be one of $m!$ possible permutations \cite[pages 41 and 48]{MitzUpf}. \begin{algorithm} \caption{Uniform sampling without replacement \cite{GT11,GrN10}}\label{alg_without} \begin{algorithmic} \REQUIRE Integers $m\geq 1$ and $1\leq c\leq m$ \ENSURE $c\times m$ sampling matrix $S$ with $\mathbf{E}[S^TS]=I_m$ \STATE \STATE Let $k_1,\ldots,k_m$ be a random permutation of $1,\ldots, m$ \smallskip \STATE $S=\sqrt{\tfrac{m}{c}}\> \begin{pmatrix}e_{k_1} & \ldots & e_{k_c}\end{pmatrix}^T$ \end{algorithmic} \end{algorithm} The following lemma presents the probability that sampling without replacement picks a particular row. \begin{lemma}\label{l_without} If Algorithm~\ref{alg_without} samples $c$ out of $m$ indices, then the probability that a particular index is picked equals $c/m$. \end{lemma} \begin{proof} The probability that some index, say~$r$, is not sampled in the first trial is $1-\tfrac{1}{m}=\tfrac{m-1}{m}$. Now there are only $m-1$ indices left. So the probability that index~$r$ is not sampled in the second trial is $1-\tfrac{1}{m-1}=\tfrac{m-2}{m-1}$. Repeating this argument shows that with probability $\prod_{t=1}^c\frac{m-t}{m-t+1} = \frac{m-c}{m}$ index~$r$ is not sampled in $c$ trials. The complementary event, the probability that index~$r$ is sampled, equals $1-\tfrac{m-c}{m} = \tfrac{c}{m}$. \end{proof} \subsection{Sampling with replacement}\label{s_with} This is the sampling strategy that appears to be analyzed in~\cite{AMTol10}. It samples exactly the requested number of rows, but with replacement, which means a row may be sampled more than once. Algorithm~\ref{alg_with} is the same as the \textsl{EXACTLY(c)} algorithm \cite[Algorithm 3]{DMMS10} with uniform probabilities, which is also used in the \textsl{BasicMatrixMultiplication Algorithm} \cite[Fig. 2]{DKM06}. \begin{algorithm} \caption{Uniform sampling with replacement \cite{DKM06,DMMS10}}\label{alg_with} \begin{algorithmic} \REQUIRE Integers $m\geq 1$ and $1\leq c\leq m$ \ENSURE $c\times m$ sampling matrix $S$ with $\mathbf{E}[S^TS]=I_m$ \STATE \FOR{$t = 1 : c$} \STATE Sample $k_t$ from $\{1,\ldots,m\}$ with probability $1/m$,\\ \STATE independently and with replacement\\ \ENDFOR \smallskip \STATE $S=\sqrt{\tfrac{m}{c}}\> \begin{pmatrix}e_{k_1} & \ldots & e_{k_c}\end{pmatrix}^T$ \end{algorithmic} \end{algorithm} Sampling with replacement (Algorithm~\ref{alg_with}) is often easier to analyze and implement than sampling without replacement (Algorithm~\ref{alg_without}), and it can also be more robust to errors \cite[\S 1.2]{MitzUpf}. \subsection{Bernoulli sampling}\label{s_bernoulli} The sampling strategy in Algorithm~\ref{alg_bernoulli} is implemented in \textsl{Blendenpik} \cite[Algorithm 1]{AMTol10}. Following \cite[Section A]{GrN10}, we use the term \textsl{Bernoulli sampling}, because the strategy treats each row as an independent, identically distributed Bernoulli random variable. Each row is either sampled or not, with the same probability for each row. Algorithm~\ref{alg_bernoulli} produces a $m\times m$ square matrix $S$ -- in contrast to Algorithms~\ref{alg_without} and~\ref{alg_with}, which produce $c\times m$ matrices. \begin{algorithm} \caption{Bernoulli sampling \cite{AMTol10,GT11,GrN10}}\label{alg_bernoulli} \begin{algorithmic} \REQUIRE Integers $m\geq 1$ and $1\leq c\leq m$ \ENSURE $m\times m$ sampling matrix $S$ with $\mathbf{E}[S^TS]= I_m$ \STATE \STATE $S=0_{m\times m}$ \FOR{$t = 1 : m$} \STATE \medskip $S_{tt} = \sqrt{\tfrac{m}{c}}\>\begin{cases} 1 & \text{with probability $\tfrac{c}{m}$} \\ 0 & \text{with probability $1-\tfrac{c}{m}$} \end{cases}$ \ENDFOR \end{algorithmic} \end{algorithm} The number of sampled rows, which is equal to the number of non-zero diagonal elements in $S$, is not known a priori, but the expected number of sampled rows is~$c$. The lemma below shows that the actual number of rows picked by Bernoulli sampling is characterized by a binomial distribution \cite[Section 2.2.2]{Ross}. \begin{lemma}\label{l_bernoulli} If Algorithm~\ref{alg_bernoulli} samples from $m$ indices with probability $\gamma\equiv c/m$, then the probability that it picks exactly $k$ indices equals $\binom{m}{k}\>\gamma^{k}\>(1-\gamma)^{m-k}$. \end{lemma} \begin{proof} Determining the diagonal elements of the $m\times m$ sampling matrix $S$ in Algorithm~\ref{alg_bernoulli} can be viewed as performing $m$ independent trials, where trial $t$ is a success ($S_{tt}\neq 0$) with probability $\gamma$, and a failure ($S_{tt}=0$) with probability $1-\gamma$. The probability of $k$ successes is given by the binomial distribution $\binom{m}{k}\>\gamma^{k}\>(1-\gamma)^{m-k}$. \end{proof} \subsection{Relating Bernoulli sampling and sampling without replacement}\label{s_relate} We show that Bernoulli sampling (Algorithm~\ref{alg_bernoulli}) is the same as first determining the number of samples with a binomial distribution (motivated by Lemma~\ref{l_bernoulli}), and then sampling without replacement (Algorithm~\ref{alg_without}). This is described in Algorithm~\ref{alg_equiv} below. \begin{algorithm} \caption{Simulating Algorithm~\ref{alg_bernoulli} with Algorithm~\ref{alg_without}}\label{alg_equiv} \begin{algorithmic} \REQUIRE Integers $m\geq 1$ and $1\leq c\leq m$ \ENSURE $\tilde{c}\times m$ sampling matrix $S$ with $\mathbf{E}[S^TS]= I_m$\\ $\qquad\quad$ that ``behaves like'' a sampling matrix generated by Algorithm~\ref{alg_bernoulli} \STATE \STATE $\gamma\equiv c/m$ \smallskip \STATE Sample $\tilde{c}$ from $\{1,\ldots,m\}$ where $\mathbf{Pr}[\tilde{c}=k] = \binom{m}{k}\>\gamma^k\>(1-\gamma)^{m-k}$ \smallskip \STATE Use Algorithm~\ref{alg_without} to sample $\tilde{c}$ indices $k_1,\ldots,k_{\tilde{c}}$ uniformly and without replacement \smallskip \STATE $S=\sqrt{\tfrac{m}{\tilde{c}}}\> \begin{pmatrix}e_{k_1} & \ldots & e_{k_{\tilde{c}}}\end{pmatrix}^T$ \end{algorithmic} \end{algorithm} Below we describe the sense in which Algorithm~\ref{alg_equiv} ``behaves like'' Bernoulli sampling in Algorithm~\ref{alg_bernoulli}. \begin{lemma} The probability that Algorithm~\ref{alg_equiv} picks a particular index equals $\gamma=c/m$. \end{lemma} \begin{proof} Motivated by Lemma~\ref{l_bernoulli}, the actual number of samples $k$ in Algorithm~\ref{alg_equiv} is given by a binomial distribution. Once a specific $k$ has emerged, one applies Lemma~\ref{l_without} to conclude that the probability that Algorithm~\ref{alg_without} picks some index $r$ is $k/m$. Now the probability that Algorithm~\ref{alg_equiv} picks some index $r$ is obtained by conditioning \cite[Section 3.5]{Ross} on the number of samples, $k$, and equals \begin{eqnarray*} \lefteqn{\sum_{k=0}^m{\mathbf{Pr}\left[\text{$k$ indices sampled}\right]\> \mathbf{Pr}\left[\text{index $r$ sampled} {\large{|}}\> \text{$k$ indices sampled}\right]}}\hspace{15pt}\\ & = &\sum_{k=1}^m{\binom{m}{k}\>\gamma^k\>\left(1-\gamma\right)^{m-k} \>\frac{k}{m}}\\ & = &\gamma\> \sum_{k=0}^{m-1}{\binom{m-1}{k}\gamma^{k}\left(1-\gamma\right)^{m-1-k}} =\gamma\left(\gamma + (1-\gamma)\right)^{m-1}=\gamma, \end{eqnarray*} where the first equality follows from the zero summand for $k=0$. \end{proof} Finally, we can conclude that sampling with Algorithm~\ref{alg_equiv} is the same as sampling with Algorithm~\ref{alg_bernoulli}. \begin{theorem} Both, Algorithms \ref{alg_equiv} and~\ref{alg_bernoulli} pick a particular set of indices $i_1, \ldots, i_c$ with probability $\gamma^c(1-\gamma)^{m-c}$. \end{theorem} \begin{proof} The probability that Algorithm~\ref{alg_bernoulli} samples indices $i_1, \ldots, i_c$ is equal to $\gamma^c(1-\gamma)^{m-c}$. We show that the same is true for Algorithm~\ref{alg_equiv}. The choice of the sampling distribution in Algorithm~\ref{alg_equiv} implies that it samples $\tilde{c}=c$ indices with probability $\binom{m}{c}\>\gamma^c\>(1-\gamma)^{m-c}$. Since there are $\binom{m}{c}$ ways to sample $c$ out of $m$ indices, the probbility that the particular index set $i_1, \ldots, i_c$ is picked, given that $c$ indices are being sampled is $1/\binom{m}{c}$. Thus, the probability that Algorithm~\ref{alg_equiv} picks indices $i_1, \ldots, i_c$ equals $$\frac{1}{\binom{m}{c}}\> \binom{m}{c}\>\gamma^c(1-\gamma)^{m-c} = \gamma^c(1-\gamma)^{m-c}.$$ \end{proof} \subsection{Numerical experiments}\label{s_sampcomp} We present two representative comparisons of the three sampling strategies, with two plots for each strategy: The condition numbers of full-rank sampled matrices $SQ$, and the failure percentage, that is the percentage of sampled matrices $SQ$ that are numerically rank deficient (as determined by the Matlab command \texttt{rank}). The experiments are limited to very tall and skinny matrices (with many more rows than columns, $m\gg n$), because that's when the sampling strategies are most efficient. In particular, since $c\geq n$ is required for $SQ$ to have full column rank, sampling methods are inefficient when $n$ is not much smaller than $m$, in which case a deterministic algorithm would be preferable. \subsubsection*{Experimental setup} The $m\times n$ matrices $Q$ with orthonormal columns have $m=10^4$ rows and $n=5$ columns. The condition numbers and failure percentages are plotted against various sampling amounts $c$, with 30 runs for each $c$. For the failure percentages we display only those sampling amounts $c$ that give rise to rank-deficient matrices, in these particular 30 runs. For Algorithm~\ref{alg_bernoulli} the horizontal axis represents the numerator $c$ in the probability, that is, the expected number of sampled rows. All three strategies sample from the same matrix. We consider two different types of matrices: Matrices with low coherence $\mu=1.5n/m$ in Figure~\ref{f_fig1}; and matrices with higher coherence $\mu=150n/m$ and many zero rows in Figure~\ref{f_fig2}. Our numerical experiments indicate that these coherence values are representative, in the sense that different values of coherence would not produce any other interesting effects. \begin{figure} \begin{center} \resizebox{4.3in}{!} {\includegraphics{pics/fig1_wo.eps}} \\ {\klein{(a)\ Algorithm~\ref{alg_without}: Sampling without replacement}} \end{center} \begin{center} \resizebox{4.3in}{!} {\includegraphics{pics/fig1_wr.eps}} \\ {\klein{(b)\ Algorithm~\ref{alg_with}: Sampling with replacement}} \end{center} \begin{center} \resizebox{4.3in}{!} {\includegraphics{pics/fig1_be.eps}} \\ {\klein{(c)\ Algorithm~\ref{alg_bernoulli}: Bernoulli sampling}} \end{center} \caption{Condition numbers and percentage of rank-deficiency for matrices with low coherence and small amounts of sampling. Here $Q$ is $m\times n$ with orthonormal columns, $m=10,000$, $n=5$, coherence $\mu=1.5n/m$, and generated with Algorithm~\ref{alg_geno}. Left panels: Horizontal coordinate axes represent amounts of sampling $n\leq c\leq 1,000$. Vertical coordinate axes represent condition numbers $\kappa(SQ)$; the maximum is 10. Right panels: Horizontal coordinate axes represent amounts of sampling that give rise to numerically rank deficient matrices $SQ$. Vertical coordinate axes represent percentage of numerically rank deficient matrices. }\label{f_fig1} \end{figure} \subsubsection*{Figure~\ref{f_fig1}} Shown are condition numbers and percentage of rank deficient matrices for a matrix $Q$ with low coherence $\mu=1.5n/m$ generated by Algorithm~\ref{alg_geno}. At most 10~percent of the rows are sampled. The three strategies exhibit almost identical behavior: The sampled matrices $SQ$ of full rank are very well conditioned, with $\kappa(SQ)\leq 5$. Numerically rank-deficient matrices $SQ$ occur only for sampling amounts $c\leq 47$. \begin{figure} \begin{center} \resizebox{4.3in}{!} {\includegraphics{pics/fig2_wo.eps}} \\ {\klein{(a)\ Algorithm~\ref{alg_without}: Sampling without replacement}} \end{center} \begin{center} \resizebox{4.3in}{!} {\includegraphics{pics/fig2_wr.eps}} \\ {\klein{(b)\ Algorithm~\ref{alg_with}: Sampling with replacement}} \end{center} \begin{center} \resizebox{4.3in}{!} {\includegraphics{pics/fig2_be.eps}} \\ {\klein{(c)\ Algorithm~\ref{alg_bernoulli}: Bernoulli sampling}} \end{center} \caption{Condition numbers and percentage of rank-deficiency for matrices with higher coherence and large amounts of sampling. Here $Q$ is $m\times n$ with orthonormal columns, $m=10,000$, $n=5$, coherence $\mu=150n/m$, and generated with Algorithm~\ref{alg_zero}. Left panels: Horizontal coordinate axes represent amounts of sampling $4,000\leq c\leq m$. Vertical coordinate axes represent condition numbers $\kappa(SQ)$; the maximum is 10. Right panels: Horizontal coordinate axes represent amounts of sampling that give rise to numerically rank deficient matrices $SQ$. Vertical coordinate axes represent percentage of numerically rank deficient matrices $SQ$; the maximum is 10 percent. }\label{f_fig2} \end{figure} \subsubsection*{Figure~\ref{f_fig2}} Shown are condition numbers and percentage of rank deficient matrices for a matrix $Q$, generated by Algorithm~\ref{alg_zero}, with coherence $150n/m$ and many zero rows. The number of sampled rows ranges from $c=4000$ to $m$. The sampled matrices $SQ$ of full rank are very well conditioned, with $\kappa(SQ)\leq 10$. Even for $c=4,000$, as many 10 percent of the sampled matrices can still be rank-deficient. All three algorithms have to sample more than half of the rows of $Q$ in order to always produce matrices $SQ$ with full column rank. Specifically in these particular runs, Algorithms \ref{alg_without} and~\ref{alg_bernoulli} need to sample $c\geq 5,222$ and $c\geq 5,301$ rows, respectively, while Algorithm~\ref{alg_without} needs $c\geq 7732$. Note that the condition numbers of matrices from Algorithms \ref{alg_without} and~\ref{alg_bernoulli} approach~1 as more and more rows are sampled. This is because no row is sampled more than once; and for $c=m$ all rows are sampled. Again, the three strategies exhibit almost identical behavior: The sampled matrices $SQ$ of full rank are very well conditioned, with $\kappa(SQ)\leq 10$. However, due to the higher coherence, numerically rank-deficient matrices occur more frequently. \subsection{Conclusions for Section~\ref{s_samp}}\label{s_sampconc} The numerical experiments illustrate that the three sampling strategies behave almost identically, in particular for small to moderate sampling amounts, and that sampled matrices of full rank tend to be very well-conditioned\footnote{We have not been able to show rigorously why the condition numbers tend to be less than~10.}. Furthermore, Section~\ref{s_relate} shows that Bernoulli sampling can be viewed as a form of sampling without replacement, and the numerical experiments confirm the similarity in behavior. Among the three strategies, we recommend sampling with replacement (Algorithm~\ref{s_with}) for small to moderate amounts of sampling in Algorithm~\ref{alg_bpik}. It is fast and easy to implement in both. \section{Condition number bounds based on coherence}\label{s_cohbound} We derive bounds for the condition numbers of matrices produced by the sampling strategies in section~\ref{s_samp}, in terms of coherence. These bounds are based on a specific concentration inequality and imply a, not necessarily tight, lower bound for the number of sampled rows (Section~\ref{s_subcohbound}). Numerical experiments illustrate that the bounds are informative (Section~\ref{s_cohex}). We end this section by summarizing the main features of the bounds (Section~\ref{s_cohconc}). \subsection{Bounds}\label{s_subcohbound} We show that the three sampling strategies in Section~\ref{s_samp} all have the same condition number bound, in terms of coherence. Theorem~\ref{t_cohbound} below is based on a matrix Chernoff concentration inequality (Section~\ref{s_chernoff}). We chose this particular inequality because extensive numerical experiments with our Matlab toolbox \texttt{kappaSQ\_v3} \cite{impl} suggest that it tends to produce the tightest bound. \begin{theorem}\label{t_cohbound} Let $Q$ be a real $m\times n$ matrix with $Q^TQ=I_n$ and coherence~$\mu$. Let $S$ be a sampling matrix produced by Algorithms~\ref{alg_without}, \ref{alg_with}, or \ref{alg_bernoulli} with $n\leq c\leq m$. For $0<\epsilon<1$ and $f(x)\equiv e^x(1+x)^{-(1+x)}$ define $$\delta\equiv n\left(f(-\epsilon)^{c/(m\mu)}+ f(\epsilon)^{c/(m\mu)}\right).$$ If $\delta<1$, then with probability at least $1-\delta$ we have $\rank(SQ)=n$ and $$\kappa(SQ)\leq \sqrt{\frac{1+\epsilon}{1-\epsilon}}.$$ \end{theorem} \begin{proof} The proof is based on results from \cite{GT11,tropp11b,tropp11} and is relegated to Section~\ref{s_tcohproof}. \end{proof} Since $0<f(\pm\epsilon)<1$ for $0<\epsilon<1$, Theorem~\ref{t_cohbound} implies that the sampling strategies in Section~\ref{s_samp} are more likely to produce full-rank matrices as the number $c$ of sampled rows increases. Furthermore, for a given total number of rows $m$, matrices $Q$ with fewer columns $n$ and lower coherence $\mu$ are more likely to give rise to sampled matrices $SQ$ that have full rank. Theorem~\ref{t_cohbound} implies the following lower bound on the number of samples, but we make no claims about the tightness of this bound. \begin{corollary}\label{c_cohbound} Under the assumptions of Theorem~\ref{c_cohbound}, $$c\geq 3m\mu\>\frac{\ln(2n/\delta)}{\epsilon^2}$$ samples are sufficient to achieve $\kappa(SQ)\leq \sqrt{\tfrac{1+\epsilon}{1-\epsilon}}$ with probability at least $1-\delta$. \end{corollary} \begin{proof} See Section~\ref{s_ccohproof}. \end{proof} Corollary~\ref{c_cohbound} implies that the sampling strategies in Section~\ref{s_samp} should sample at least $c=\Omega\left(m\mu\ln{n}\right)$ rows to produce a full rank, well-conditioned matrix. In particular, if $Q$ has minimal coherence $\mu=n/m$, then Corollary~\ref{c_cohbound} implies that the number of sampled rows should be at least \begin{eqnarray}\label{e_nmcoh} c\geq 3n \> \frac{\ln(2n/\delta)}{\epsilon^2}, \end{eqnarray} that is $c=\Omega\left(n\ln{n}\right)$. To achieve $\kappa(SQ)\leq 10$ with probability at least .99 requires that the number of sampled rows be at least \begin{eqnarray}\label{c_lb} c\geq 3.2\,m\mu\>\left(\ln(2n)+4.7\right). \end{eqnarray} Here we chose $\epsilon_0=99/101$, so that the condition number bound equals $\sqrt{\frac{1+\epsilon_0}{1-\epsilon_0}}=10$. \begin{remark}\label{r_lowcoh} Theorem~\ref{t_cohbound} is informative only for sufficiently low coherence values. For instance, consider the higher coherence matrices from Figure~\ref{f_fig2} in Section~\ref{s_sampcomp} with $m=10,000$, $n=5$ and coherence $\mu =150n/m$. Choose $\epsilon=99/101$ so that $\kappa(SQ)\leq 10$, and a failure probability $\delta = .01$. Then Corollary~\ref{c_cohbound} implies the lower bound $c\geq 12,408$, which means that the number of sampled rows would have to be larger than the total number of rows. \end{remark} \subsection{Numerical experiments}\label{s_cohex} We compare the bound for the condition numbers of the sampled matrices (Theorem~\ref{t_cohbound}) with the true condition numbers of matrices produced by sampling with replacement (Algorithm~\ref{alg_with}). There are several reasons why it suffices to consider only a single sampling strategy: The three sampling methods all have the same bound (Theorem~\ref{t_cohbound}); Bernoulli sampling is a form of sampling without replacement (Section~\ref{s_relate}); and all three sampling methods exhibit very similar behavior for matrices of low coherence (Sections \ref{s_sampcomp} and~\ref{s_sampconc}). Furthermore, this allows a clean comparison with the bounds in Section~\ref{s_leverage} which apply only to Algorithm~\ref{alg_with}. \subsubsection*{Experimental setup} The $m\times n$ matrices $Q$ with orthonormal columns have $m=10^4$ rows and $n=5$ columns. The left panels in Figure~\ref{f_fig3} show the condition numbers of the full-rank sampled matrices $SQ$ produced by Algorithm~\ref{alg_with} against different sampling amounts $c$, with 30 runs for each $c$. The right panels in Figure~\ref{f_fig3} show the percentage of rank deficient matrices $SQ$ against different sampling amounts~$c$. We display only those sampling amounts $c$ that give rise to rank-deficient matrices, in these particular 30 runs. The left panels in Figure~\ref{f_fig3} also show the condition number bound $\kappa_{\epsilon}\equiv \sqrt{\tfrac{1+\epsilon}{1-\epsilon}}$ from Theorem~\ref{t_cohbound}. For each value of $c$, we obtain $\epsilon$ as the solution of the nonlinear equation $F_c(x)^2=0$ associated with Theorem~\ref{t_cohbound} and defined as $$F_c(x)\equiv \delta - n\left(f(-x)^{c/(m\mu)}+ f(x)^{c/(m\mu)}\right).$$ We impose the stringent requirement of $\delta = .01$, corresponding to a 99 percent success probability. Since an explicit expression seems out of reach, we use unconstrained nonlinear optimization (a Nelder-Mead simplex direct search) to solve $F_c(x)^2=0$. This is done in Matlab with a code equivalent to $$\epsilon=\left|\texttt{fminsearch}(F_c(x)^2,0, 10^{-30})\right|,$$ where \texttt{fminsearch} starts at the point~0, and terminates when $|F_c(\epsilon)|^2\leq 10^{-30}$. If $0<\epsilon<1$ then $\kappa_{\epsilon}$ is plotted, otherwise nothing is plotted. As explained in Remark~\ref{r_lowcoh}, Theorem~\ref{t_cohbound} is not informative for higher coherence values, so we consider matrices with the following properties: Minimal coherence $\mu=n/m$ in Figure~\ref{f_fig3}(a); low coherence $\mu=1.5n/m$ in Figure~\ref{f_fig3}(b); slightly higher coherence $\mu=15n/m$ with many zero rows in Figure~\ref{f_fig3}(c). The matrices for Figures~\ref{f_fig3}(a) and~\ref{f_fig3}(b) were generated with Algorithm~\ref{alg_geno}, while the matrix for Figure~\ref{f_fig3}(c) was generated with Algorithm~\ref{alg_zero}. \subsubsection*{Figure~\ref{f_fig3}} The left panels illustrate that Theorem~\ref{t_cohbound}, constrained to a 99 percent success probability, correctly predicts the magnitude of the condition numbers, i.e. $\kappa(SQ)\leq 10$. Hence Theorem~\ref{t_cohbound} provides informative qualitative bounds for matrices with very low coherence, as well as for matrices with slightly higher coherence and many zero rows. \begin{figure} \begin{center} \resizebox{4in}{!} {\includegraphics{pics/fig3a.eps}}\\ {\klein{(a)\ $Q$ has minimal coherence $\mu = n/m$. Sampling amounts are $n\leq c\leq 1,000$.}} \end{center} \begin{center} \resizebox{4in}{!} {\includegraphics{pics/fig3b.eps}}\\ {\klein{(b)\ $Q$ has low coherence $\mu = 1.5 n/m$. Sampling amounts are $n\leq c\leq 1,000$.}} \end{center} \begin{center} \resizebox{4in}{!} {\includegraphics{pics/fig3c.eps}}\\ {\klein{(c)\ $Q$ has slightly higher coherence $\mu = 15n/m$ and many zero rows. Sampling sampling amounts are $n\leq c\leq 3,000$.}} \end{center} \caption{Condition numbers and bound from Theorem~\ref{t_cohbound}, and percentage of rank-deficiency. Here $Q$ is $m\times n$ with orthonormal columns, $m=10,000$ and $n=5$. Left panels: The horizontal coordinate axes represent amounts of sampling~$c$. The vertical coordinate axes represent condition numbers $\kappa(SQ)$; the maximum is 10. The dots at the bottom represent the condition numbers of matrices sampled with Algorithm~\ref{alg_with}, while the upper line represents the bound from Theorem~\ref{t_cohbound}. Right panels: The horizontal coordinate axes represent amounts of sampling that produce numerically rank deficient matrices $SQ$. The vertical coordinate axes represent the percentage of numerically rank deficient matrices~$SQ$. }\label{f_fig3} \end{figure} \subsubsection*{Table~\ref{t_tab1}} This is a comparison of the numerical experiments in Figure~\ref{f_fig3} with the bounds from Theorem~\ref{t_cohbound} and Corollary~\ref{c_cohbound}, both restricted to a 99 percent success probability. The third column depicts the highest values of $c$ for which a rank-deficient matrix occurs, during these particular 30 runs. It should be kept in mind that these values are highly dependent on the particular sampling runs. This column is to be compared to the fourth column which contains the lowest values of $c$ where Theorem~\ref{t_cohbound} starts to apply. Although there is a gap between the occurrence of the last rank deficiency and the onset of Theorem~\ref{t_cohbound}, the values have qualitatively the same order of magnitude. The rightmost column in Table~\ref{t_tab1} contains the values of the lower bound (\ref{c_lb}), and is to be compared to the column with the starting values for Theorem~\ref{t_cohbound}. Although (\ref{c_lb}) is weaker than Theorem~\ref{t_cohbound}, its values are close to the starting values of Theorem~\ref{t_cohbound}, especially for lower coherence. Hence, the lower bound (\ref{c_lb}) captures the correct magnitude of the sampling amounts where Theorem starts to become informative. Table~\ref{t_tab1} illustrates that, although Theorem~\ref{t_cohbound} and Corollary~\ref{c_cohbound} tend to become more pessimistic with increasing coherence, they still provide qualitative information for matrices with low coherence -- even when restricted to a 99 percent success probability. \begin{table} \begin{center} \begin{tabular}{|l||c|c|c|c|} \hline Figure& coherence $\mu$ & last rank deficiency & Theorem~\ref{t_cohbound} & (\ref{c_lb})\\ & & occurs at $c=$ & starts at $c=$ & \\ \hline\hline \ref{f_fig3}(a) & $n/m$ & 31 & 81 & 83 \\ \ref{f_fig3}(b) & $1.5\>n/m$ & 31 & 121 & 125\\ \ref{f_fig3}(c) & $15\>n/m$ & 740 & 1207 & 1241\\ \hline \end{tabular} \end{center} \bigskip \caption{Comparison of information from Figure~\ref{f_fig3}, with Theorem~\ref{t_cohbound} and Corollary~\ref{c_cohbound}.}\label{t_tab1} \end{table} \subsection{Conclusions for Section~\ref{s_cohbound}}\label{s_cohconc} The bounds in Theorem~\ref{t_cohbound} and Corollary~\ref{c_cohbound} have the following advantages: \begin{enumerate} \item They are non-asymptotic bounds, where all constants have explicit numerical values, hence they are tighter than the bounds in \cite[Theorem 3.2]{AMTol10}. \item They apply to three different sampling methods. \item They imply a lower bound, of $\Omega\left(m\mu\ln{n}\right)$, on the required number of sampled rows. Although we did not give a formal proof of tightness, numerical experiments illustrate that sampling only the required number of rows implied by the bound is realistic. numerical experiments illustrate that the bound is realistic. \item Even under the stringent requirement of a 99 percent success probability, they are informative for matrices of small dimension because they correctly predict the magnitude of the condition numbers for the sampled matrices. \end{enumerate} Note that the bounds in Theorem~\ref{t_cohbound} and Corollary~\ref{c_cohbound} are informative only for matrices that are tall and skinny ($m\gg n$) and have low coherence. The restriction to tall and skinny matrices is not an imposition, because it is required for the effectiveness of the sampling strategies, see Section~\ref{s_sampcomp}. In the next section we try to relax the restriction to low coherence matrices, by more thoroughly exploiting the information available from the row norms of $Q$. \section{Condition number bounds based on leverage scores, for uniform sampling with replacement}\label{s_leverage} The goal is to tighten Theorem~\ref{t_cohbound} by making use of all the row norms of $Q$, instead of just the largest one. To this end we introduce leverage scores (Section~\ref{s_levscore}), which are the squared row norms of $Q$. We use them to derive a bound for uniform sampling with replacement (Section~\ref{s_levbound}), and for more easily computable versions of the bound (Section~\ref{s_clevbound}). Analytical (Section~\ref{s_ancomp}) and experimental (Section~\ref{s_numcomp}) comparisons demonstrate that the implied lower bound on the number of sampled rows is better than the coherence-based bounds in Section~\ref{s_cohbound}. A review with some reflection ends this section (Section~\ref{s_lqconc}). \subsection{Leverage scores}\label{s_levscore} So-called \textit{statistical leverage scores} were first introduced in 1978 by Hoaglin and Welsch \cite{HoagW78} to detect outliers when computing regression diagnostics, see also \cite{ChatterH86,VelleW81}. Mahoney and Drineas pioneered the use of leverage scores for importance sampling strategies in randomized matrix computations \cite{Mah11}. Specifically, if $M$ is a real $m\times n$ matrix with $\rank(M)=n$, then the $m\times m$ \textit{hat matrix} $$H\equiv M(M^TM)^{-1}M^T$$ is the orthogonal projector onto the column space of $M$, and its diagonal elements are called \textit{leverage scores} \cite[Section~2]{HoagW78}. Hence, leverage scores are basis-independent. For our purposes, though, it suffices to define them in terms of a thin QR decomposition $M=QR$, so that the hat matrix can be expressed as $H=QQ^T$. \begin{definition} If $Q$ is a $m\times n$ matrix with $Q^TQ=I_n$, then its {\rm leverage scores} are $$\ell_j\equiv \|e_j^TQ\|_2^2, \qquad 1\leq j\leq m.$$ The $m\times m$ diagonal matrix of leverage scores is $$L\equiv\diag(\ell_1,\ldots,\ell_m).$$ \end{definition} Note that the coherence is the largest leverage score, $$\mu=\max_{1\leq j\leq m}{\ell_j}=\|L\|_2.$$ \subsection{Bounds}\label{s_levbound} The bound in Theorem~\ref{t_levbound} below involves leverage scores and is based on a matrix Bernstein concentration inequality (Section~\ref{s_bernstein}), rather than on the matrix Chernoff concentration inequality (Section~\ref{s_chernoff}) for Theorem~\ref{t_cohbound}. Although the Bernstein inequality may not always be as tight, we did not see how to insert leverage scores into the Chernoff inequality. \begin{theorem}\label{t_levbound} Let $Q$ be a $m\times n$ real matrix with $Q^TQ=I_n$, leverage scores $\ell_j$, $1\leq j\leq m$, and coherence $\mu$. Let $S$ be a sampling matrix produced by Algorithm~\ref{alg_with} with $n\leq c\leq m$. For $0<\epsilon<1$ set $$\delta \equiv 2n\exp\left(-\tfrac{3}{2}\> \frac{c \epsilon^2}{m\>(3\|Q^TLQ\|_2+\epsilon\mu)}\right).$$ If $\delta<1$, then with probability at least $1-\delta$ we have $\rank(SQ)=n$ and $$\kappa(SQ) \leq \sqrt{\frac{1+\epsilon}{1-\epsilon}}.$$ \end{theorem} \begin{proof} The proof uses results from \cite{BalzRN11,Recht11} and is relegated to Section~\ref{s_tlevproof}. \end{proof} Like Theorem~\ref{t_cohbound}, Theorem~\ref{t_levbound} implies that sampling with replacement is more likely to produce full-rank matrices as the number $c$ of sampled rows increases. Furthermore, for a given total number of rows $m$, matrices $Q$ with fewer columns $n$ and lower coherence $\mu$ are more likely to yield sampled matrices $SQ$ that have full rank. The dependence of $\|Q^TLQ\|_2$ on $\mu$ is discussed below. \begin{remark} \label{r_lq} The norm $\|Q^TLQ\|_2$ has simple and tight bounds in terms of the coherence, \begin{eqnarray}\label{e_mu} \mu^2\leq \|Q^TLQ\|_2\leq \mu. \end{eqnarray} The lower bound follows from $\|Q^TLQ\|_2=\|L^{1/2}Q\|_2^2$ and $$\|L^{1/2}Q\|_2\geq \|e_j^TL^{1/2}Q\|_2=\ell_j^{1/2}\>\|e_j^TQ\|_2=\ell_j, \qquad 1\leq j\leq m,$$ which implies $\|L^{1/2}Q\|_2\geq\mu$. \smallskip The bounds (\ref{e_mu}) are attained for extreme values of the coherence: \begin{itemize} \item In case of minimal coherence $\mu=\ell_j$ for all $1\leq j\leq m$, we have $L=\mu I_m$. Thus $\|Q^TLQ\|_2=\mu\|Q^TQ\|_2=\mu$, and the upper bound is attained. \item In case of maximal coherence $\mu=1$, we have $\mu^2=\mu$. Thus $\|Q^TLQ\|_2=\mu^2=\mu$, and both, lower and upper bounds are attained. \end{itemize} \end{remark} \subsection{Computable bounds}\label{s_clevbound} We present easily computable bounds for $\|Q^TLQ\|_2$, based on coherence and several of the largest leverage scores. To this end, we use a labeling of the leverage scores in non-increasing order, $$\mu=\ell_{[1]}\geq \cdots\geq \ell_{[m]}.$$ \begin{corollary}\label{c_1} Under the assumptions of Theorem~\ref{t_levbound}, if $t\equiv\left\lfloor 1/\mu\right\rfloor$, then $$\|Q^TLQ\|_2\leq \mu\>\sum_{j=1}^t{\ell_{[j]}}+(1-t\,\mu)\,\ell_{[t+1]} \leq \mu.$$ If, in addition, $t$ is an integer, then $\|Q^TLQ\|_2\leq \mu\>\sum_{j=1}^t{\ell_{[j]}}$. \end{corollary} \begin{proof} See Section~\ref{s_lqproof}. \end{proof} The number of large leverage scores appearing in Corollary~\ref{c_1} depends on the coherence: Few leverage scores for high coherence, but more for low coherence. Henceforth we will use the approximation from Corollary~\ref{c_1} instead of the true value $\|Q^TLQ\|_2$, for two reasons: First, numerical experiments show that the approximation tends to be very accurate. Second, the approximation is convenient, because it requires only a leverage score distribution rather than a full-fledged matrix~$Q$. \begin{remark} Corollary~\ref{c_1} is tight for the extreme cases of minimal and maximal coherence. \begin{itemize} \item In case of minimal coherence $\mu=\ell_j$ for all $1\leq j\leq m$, Remark~\ref{r_lq} implies $\|Q^TLQ\|_2=\mu$. The bound in Corollary~\ref{c_1} is $\|Q^TLQ\|_2\leq \mu$, thus tight. \item In case of maximal coherence $\mu=1$, Remark~\ref{r_lq} implies $\|Q^TLQ\|_2= \mu^2=\mu$. Corollary~\ref{c_1} holds with $t=1$ and gives the bound $\|Q^TLQ\|_2\leq \mu$, which is tight as well. \end{itemize} \end{remark} Inserting this approximation for $\|Q^TLQ\|_2$ into the expression for $\delta$ in Theorem~\ref{t_levbound} yields a, not necessarily tight, lower bound on the number of samples. \begin{corollary}\label{c_levbound} Under the assumptions of Theorem~\ref{t_levbound}, $$c\geq \tfrac{2}{3}m\>(3\tau+\epsilon\mu)\> \frac{\ln(2n/\delta)}{\epsilon^2},$$ where $\tau\equiv \mu\>\sum_{j=1}^t{\ell_{[j]}}+(1-t\,\mu)\,\ell_{[t+1]}$, samples are sufficient to achieve $\kappa(SQ)\leq \sqrt{\tfrac{1+\epsilon}{1-\epsilon}}$ with probability at least $1-\delta$. \end{corollary} In particular, if $Q$ has minimal coherence $\mu=n/m$, then Corollary~\ref{c_levbound} implies that the number of sampled rows should be at least $$c\geq 3n \> \frac{\ln(2n/\delta)}{\epsilon^2}.$$ This is the same as the coherence-based lower bound (\ref{e_nmcoh}). To achieve $\kappa(SQ)\leq 10$ with probability at least .99 requires that the number of sampled rows be at least \begin{eqnarray}\label{c_lbl} c\geq m\>(2.1\tau+.7\mu)\>\left(\ln(2n)+4.7\right). \end{eqnarray} \subsection{Analytical comparison of the bounds in Sections \ref{s_subcohbound} and~\ref{s_levbound}}\label{s_ancomp} An analytical comparison between Theorems \ref{t_cohbound} and~\ref{t_levbound} is not obvious, because they are based on different concentration inequalities. Instead we compare the implied lower bounds for the number of sampled rows, and show that the leverage-score based bound in Corollary~\ref{c_levbound} is at least as tight as the coherence-based bound in Corollary~\ref{c_cohbound}. \begin{corollary}\label{c_comp} Under the assumptions of Theorem~\ref{t_levbound} and Corollary~\ref{c_1}, $$\tfrac{2}{3}m\>(3\tau+\epsilon\mu)\> \frac{\ln(2n/\delta)}{\epsilon^2}\leq 3m\mu\>\frac{\ln(2n/\delta)}{\epsilon^2}.$$ Hence Corollary~\ref{c_levbound} is at least as tight as Corollary~\ref{c_cohbound}. \end{corollary} \begin{proof} See Section~\ref{s_compproof}. \end{proof} \subsection{Experimental comparison of the bounds in Sections \ref{s_subcohbound} and~\ref{s_levbound}}\label{s_numcomp} We present numerical experiments to compare the lower bounds for the number of sampled rows in Corollaries~\ref{c_cohbound} and~\ref{c_levbound}, for different values of coherence. This gives quantitative insight into the comparison in Corollary~\ref{c_comp}, and illustrates the reduction in the number of sampled rows from Corollary~\ref{c_levbound}, as compared to Corollary~\ref{c_cohbound}. \subsubsection*{Experimental setup} As in previous sections, we use $m\times n$ matrices with $m=10^4$ rows and $n=5$ columns. The success probability is .99; and $\epsilon=99/101$, so that the bound for $\kappa(SQ)$ is equal to 10. Hence the bounds in Corollaries \ref{c_cohbound} and~\ref{c_levbound} amount to (\ref{c_lb}) and (\ref{c_lbl}), respectively. We consider two different leverage scores distributions: A distribution generated by Algorithm~\ref{alg_geno} with one large leverage score in Table~\ref{t_tab2}; and a distribution generated by Algorithm~\ref{alg_zero} with as many zeros as possible in Table~\ref{t_tab3}. \subsubsection*{Table~\ref{t_tab2}} This table shows the lower bounds on the number of sampled rows, for a leverage score distribution generated with Algorithm~\ref{alg_geno} that consists of one large leverage score, equal to the coherence, and all remaining leverage scores being non-zero and identical. The bounds, as well as the approximation $\tau$ to $\|Q^TLQ\|_2$, are displayed for eight different values of coherence, ranging from minimal coherence $\mu=n/m$ to $\mu =100n/m$. Table~\ref{t_tab2} illustrates that with increasing coherence, the number of sampled rows implied by Corollary~\ref{c_levbound} is only about 20 percent of that from Corollary~\ref{c_cohbound}. This is because $\tau$ increases much more slowly than $\mu$. For instance, $\tau\approx \mu/10$ when $\mu=100n/m$. \begin{table} \begin{center} \begin{tabular}{|c||c|c|c|c|c|c|c|c|} \hline $\mu/(n/m)$ & 1 & 5 & 10 & 15 & 20 &25 &50& 100\\ \hline\hline Cor. \ref{c_cohbound} & 108 & 540 & 1,079 & 1,618 & 2,157 &2,697 &5,393 &10,786\\ Cor. \ref{c_levbound} & 96 & 191 & 310 & 432 & 556 &682 &1,3343 & 2,777\\ \hline $\tau/(n/m)$ & 1.00 & 1.01 & 1.04 & 1.10 &1.19 & 1.30 & 2.22 &9.95\\ \hline \end{tabular} \end{center} \bigskip \caption{Lower bounds for number of sampled rows in Corollaries \ref{c_cohbound} and~\ref{c_levbound} and approximation $\tau$, for different values of coherence~$\mu$. The first value represents minimal coherence $\mu=n/m$. Here $m=10,000$, $n=5$, $\delta=.01$, $\epsilon=99/101$, with leverage scores generated by Algorithm~\ref{alg_geno}.}\label{t_tab2} \end{table} \subsubsection*{Table~\ref{t_tab3}} This table shows the lower bounds on the number of sampled rows. The corresponding leverage score distribution is generated with Algorithm~\ref{alg_zero} and consists of as many zeros as possible. All non-zero leverage scores, expect possibly one, are equal to the coherence $\mu$, so that $\tau\approx\mu$. The bounds are displayed for eight different values of coherence, ranging from minimal coherence $\mu=n/m$ to $\mu =100n/m$. The bounds for Corollary~\ref{c_cohbound} are the same as in Table~\ref{t_tab2}, because the coherence values are the same. Since $\tau=\mu$, the difference between Corollaries \ref{c_cohbound} and~\ref{c_levbound} is not as drastic as in Table~\ref{t_tab2}, yet it increases with increasing coherence. For $\mu=100n/m$, Corollary~\ref{c_levbound} remain informative, while Corollary~\ref{c_cohbound} does not. \begin{table} \begin{center} \begin{tabular}{|c||c|c|c|c|c|c|c|c|} \hline $\mu/(n/m)$ & 1 & 5 & 10 & 15 & 20 &25 &50& 100\\ \hline\hline Cor. \ref{c_cohbound} & 108 & 540 & 1,079 & 1,618 & 2,157&2,697 &5,393&10,787\\ Cor. \ref{c_levbound}& 96 & 477 & 954 & 1,431 & 1,908 &2,385 &4,770 & 9,539\\ \hline \end{tabular} \end{center} \bigskip \caption{Lower bounds for number of sampled rows in Corollaries \ref{c_cohbound} and~\ref{c_levbound}, for different values of coherence~$\mu$. The first value represents minimal coherence $\mu=n/m$. Here $m=10,000$, $n=5$, $\delta=.01$, $\epsilon=99/101$, with leverage scores generated by Algorithm~\ref{alg_zero}.}\label{t_tab3} \end{table} \subsection{Conclusions for Section~\ref{s_leverage}}\label{s_lqconc} The goal of this section was to derive condition number bounds that are based on leverage scores rather than just coherence, when rows are sampled uniformly with replacement (Algorithm~\ref{alg_with}). Corollary~\ref{c_comp} and the numerical experiments illustrate that the lower bound on the number of sampled rows implied by Corollary~\ref{c_levbound} is smaller than that from Corollary~\ref{c_cohbound}. Although the coherence based bound in Theorem~\ref{t_cohbound} is derived from a stronger concentration inequality than the one for Theorem~\ref{t_levbound}, this difference disappears in the weakening necessary to obtain lower bounds for the amount of sampling. Even in cases when the leverage score measure $\tau$ is the same as the coherence, Corollary~\ref{c_levbound} still retains a small advantage, which can increase with increasing coherence. Hence Corollary~\ref{c_levbound} tends to remain informative for larger values of coherence, even when Corollary~\ref{c_cohbound} fails. The difference in implied sampling amounts becomes more drastic in the presence widely varying non-zero leverage scores, and can be as high as ten percent. This is because the coherence-based bound in Corollary~\ref{c_cohbound} cannot take advantage of the distribution of the leverage scores. Hence, when it comes to lower bounds for the number of rows sampled uniformly with replacement, we recommend Corollary~\ref{c_levbound}. We have yet to derive leverage score based bounds for the other two sampling strategies, uniform sampling without replacement (Algorithm~\ref{alg_without}) and Bernoulli sampling (Algorithm~\ref{alg_bernoulli}). \section{Algorithms for generating matrices with prescribed coherence and leverage scores}\label{s_genalg} In order to investigate the efficiency of the sampling methods in Section~\ref{s_samp}, and test the tightness of the bounds in Sections \ref{s_cohbound} and~\ref{s_leverage}, we need to generate matrices with orthonormal columns that have prescribed leverage scores and coherence. The algorithms are implemented in the Matlab package \textsl{kappa\_SQ\_v3} \cite{impl}. We present algorithms for generating matrices with prescribed leverage scores and coherence (Section~\ref{s_alg}), and for generating particular leverage score distributions with prescribed coherence (Section~\ref{s_precoh}). Such distributions can then, in turn, serve as inputs for the algorithm in Section~\ref{s_alg}. Furthermore we present two classes of structured matrices with prescribed coherence that are easy and fast to generate (Section~\ref{s_mat}). \subsection{Matrices with prescribed leverage scores}\label{s_alg} We present an algorithm that generates matrices with orthonormal columns that have prescribed leverage scores. In Section~\ref{s_pre} we prove an existence result to show that this is always possible. Algorithm~\ref{alg_gen} is a transposed version of \cite[Algorithm 3]{DHST05}. It repeatedly applies $m\times m$ Givens rotations $G_{ij}$ that rotate two rows $i$ and $j$, and are computed from numerically stable expressions \cite[section 3.1]{DHST05}. At most $m-1$ such rotations are necessary. Since each rotation affects only two rows, Algorithm~\ref{alg_gen} requires $\mathcal{O}(mn)$ arithmetic operations. \begin{algorithm} \caption{Generating a matrix with prescribed leverage scores \cite{DHST05}}\label{alg_gen} \begin{algorithmic} \REQUIRE Integers $m$ and $n$ with $m\geq n\geq 1$\\ $\qquad\ $ Vector $\ell$ with elements $0\leq \ell_1\leq\cdots \leq \ell_m\leq 1$ and $\sum_{j=1}^m{\ell_j}=n$ \ENSURE $m\times n$ matrix $Q$ with $Q^TQ=I_n$ and leverage scores $\|e_j^TQ\|_2^2 =\ell_j$, $1\leq j\leq m$\\ $\qquad$\\ \STATE $Q=\begin{pmatrix}I_n & 0_{n\times (m-n)}\end{pmatrix}^T$ $\qquad$ \COMMENT{Initialization} \REPEAT \STATE Determine indices $i<k<j$ with\\ $\|e_i^TQ\|_2^2<\ell_i$, $\|e_k^TQ\|_2^2=\ell_k$, $\|e_j^TQ\|_2^2>\ell_j$ \IF{$\ell_i-\|e_i^TQ\|_2^2\leq \|e_j^TQ\|_2^2-\ell_j$} \STATE Apply rotation $G_{ij}$ to rows $i$ and $j$ so that $\|e_i^TG_{ij}\,Q\|_2^2=\ell_i$ \ELSE \STATE Apply rotation $G_{ij}$ to rows $i$ and $j$ so that $\|e_j^TG_{ij}\,Q\|_2^2=\ell_j$ \ENDIF \STATE $Q=G_{ij}\,Q$ $\qquad$ \COMMENT{Update} \UNTIL{no more such indices exist} \end{algorithmic} \end{algorithm} \subsection{Leverage score distributions with prescribed coherence}\label{s_precoh} We present algorithms that generate leverage score distributions for prescribed coherence. The resulting distributions then serve as inputs for Algorithm~\ref{alg_gen}. These particular leverage score distribution help to distinguish the effect of coherence, which is the largest leverage score, from that of the remaining leverage scores. \subsubsection*{One large leverage score} Given a prescribed coherence $\mu$, Algorithm~\ref{alg_geno} generates a distribution consisting of one large leverage score equal to $\mu$ and the remaining leverage scores being identical and non-zero. \begin{algorithm} \caption{Generating a leverage score distribution with prescribed coherence: One large leverage score}\label{alg_geno} \begin{algorithmic} \REQUIRE Integers $m$ and $n$ with $m\geq n\geq 1$\\ \STATE $\qquad\ $ Real number $\mu$ with $n/m\leq \mu\leq 1$\\ \ENSURE Vector $\ell$ with elements $\ell_1=\mu$, $0< \ell_j \leq 1$ and $\sum_{j=1}^m{\ell_j}=n$\\ $\qquad$\\ \STATE $\ell_1=\mu$ \FOR{$j=2:m$} \STATE $\ell_j=\tfrac{n-\mu}{m-1}$ \ENDFOR \end{algorithmic} \end{algorithm} In the special case of minimal coherence $\mu=n/m$, Algorithm~\ref{alg_geno} generates $m$ identical leverages equal to $\mu$, which is the only possible leverage score distribution in this case. \subsubsection*{Many zero leverage scores} Given a prescribed coherence, Algorithm~\ref{alg_zero} generates a distribution with as many zero leverage scores as possible. This serves as an ``adversarial'' distribution for the sampling algorithms in Section~\ref{s_samp}. Given a prescribed coherence $\mu$, Algorithm~\ref{alg_zero} first determines the smallest number of rows $m_s$ that can realize this coherence, sets $m_s-1$ leverage scores equal to $\mu$, assigns another leverage score to to take up the possibly non-zero slack, and sets the remaining leverage scores to zero. \begin{algorithm} \caption{Generating a leverage score distribution with prescribed coherence: Many zero leverage scores}\label{alg_zero} \begin{algorithmic} \REQUIRE Integers $m$ and $n$ with $m\geq n\geq 1$\\ $\qquad\ $ Real number $\mu$ with $n/m\leq \mu\leq 1$\\ \ENSURE Vector $\ell$ with elements $\ell_1=\mu$, $0\leq \ell_j \leq 1$ and $\sum_{j=1}^m{\ell_j}=n$\\ $\qquad$\\ \STATE $m_s=\lceil n/\mu\rceil$ $\qquad$ \COMMENT{Number of nonzero rows} \FOR{$j=1:m_s-1$} \STATE $\ell_j=\mu$ \ENDFOR \STATE $\ell_{m_s}= n-(m_s-1)\,\mu$ \FOR{$j=m_s+1:m$} \STATE $\ell_j=0$ \ENDFOR \end{algorithmic} \end{algorithm} \subsection{Structured matrices with prescribed coherence}\label{s_mat} We present two clas\-ses of structured matrices with orthonormal columns that have prescribed coherence. Although the structure puts constraints on the matrix dimensions, the generation of these matrices is faster than running Algorithm~\ref{alg_gen}. Note that the matrices produced by Algorithm~\ref{alg_gen} also have structure, but it is not easily characterized. \paragraph{Stacks of diagonal matrices} Given matrix dimensions $m$ and $n$, where $s=m/n$ is an integer, and prescribed coherence $\mu$. The $m\times n$ matrix $Q$ below has orthonormal columns and coherence $\mu$, and consists of $s$ stacks of $n\times n$ diagonal matrices, $$Q= \begin{pmatrix}\sqrt{\mu}\,I_n\\ \phi\, I_n\\ \vdots \\ \phi\, I_n\end{pmatrix} \qquad where \qquad \phi\equiv \sqrt{\frac{1-\mu}{\frac{m}{n}-1}}.$$ \paragraph{Matrices with Hadamard structure} Given matrix dimensions $m$ and $n$, where $m=2^k$ and $n<m$ is also a power of two, and prescribed coherence $\mu$. The $m \times n$ matrix $$Q= D_k\begin{pmatrix}I_n\\0\end{pmatrix}$$ has orthonormal columns and coherence $\mu$, and is defined recursively as follows. For $$\alpha\equiv\sqrt{\frac{\mu-\frac{n-1}{m-1}}{1-\frac{n-1}{m-1}}},\qquad \beta\equiv \sqrt{\frac{1-\alpha^2}{m-1}}$$ define square matrices $B_j$ of dimension $2^j$ and square matrices $D_j$ of dimension $2^{j+1}$ as follows, \begin{eqnarray*} &B_0=\beta, \qquad &B_{j+1}=\begin{pmatrix} -B_j & B_j \\ B_j & B_j\end{pmatrix} \quad\qquad 0\leq j\leq k-1\\ &D_1=\begin{pmatrix}\alpha & -\beta \\ \beta & \alpha\end{pmatrix},\qquad &D_{j+1}=\begin{pmatrix}D_j & -B_j \\ B_j & D_j\end{pmatrix}. \end{eqnarray*} Note that only the final matrix $Q$ has orthonormal columns and coherence $\mu$ while, in general, the intermediate matrices $B_j$ and $D_j$ do not. We omit the messy induction proof, because it does not provide much insight. \section{Future work}\label{s_future} We have investigated three strategies for uniform sampling of rows from matrices with orthonormal columns: Without replacement, with replacement, and Bernoulli sampling. We derived bounds on the condition numbers of the sampled matrices, in terms of coherence and leverage scores. Numerical experiments confirm that the bounds are realistic, even for high success probabilities and matrices with small dimensions. The following work still needs to be done. \begin{itemize} \item Conversion of the \textsl{kappa\_SQ\_v3} MATLAB toolbox from a research code to a robust, flexible, and user-friendly GUI that facilitates reproducible research in the randomized algorithms community. \item Tightening of Corollary~\ref{c_cohbound} so that it retains the strength of the Chernoff concentration inequality inherent in Theorem~\ref{t_levbound}. \item Extension of the condition number bounds in Section~\ref{s_leverage} to uniform sampling without replacement (Algorithm~\ref{alg_without}) and Bernoulli sampling (Algorithm~\ref{alg_bernoulli}). \item Determination of a statistically significant number of runs for each sampling amount $c$, for two purposes: \begin{enumerate} \item To assert, within a specific confidence interval, bounds on the condition numbers of the \textit{actually sampled} matrices. \item To assert with a specific confidence that the probabilistic expressions in Sections \ref{s_cohbound} and~\ref{s_leverage} do indeed represent bounds. \end{enumerate} \end{itemize} \subsection*{Acknowledgements} We are very grateful to John Holodnak, Petros Drineas and two anonymous reviewers for reading our paper so carefully and for providing many helpful suggestions. \section{Proofs for Sections \ref{s_cohbound} and \ref{s_levbound}}\label{s_app} For the coherence-based bounds in Section~\ref{s_cohbound} we first present a matrix concentration inequality (Section~\ref{s_chernoff}), and then the proofs of Theorem~\ref{t_cohbound} (Section~\ref{s_tcohproof}) and Corollary~\ref{c_cohbound} (Section~\ref{s_ccohproof}). For the bound based on leverage scores in Section~\ref{s_levbound}, we first present a matrix concentration inequality (Section~\ref{s_bernstein}), and then the proof of Theorem~\ref{t_levbound} (Section~\ref{s_tlevproof}). \subsection{Matrix Chernoff Concentration inequality}\label{s_chernoff} The matrix concentration inequality below is the basis for Theorem~\ref{t_cohbound} and Corollary~\ref{c_cohbound}. Denote the eigenvalues of a Hermitian matrix $Z$ by $\lambda_j(Z)$, and the smallest and largest eigenvalues by $\lambda_{min}(Z)\equiv \min_j{\lambda_j(Z)}$ and $\lambda_{max}(Z) \equiv \max_j{\lambda_j(Z)}$, respectively. \begin{theorem}[Corollary 5.2 in \cite{tropp11}]\label{t_tropp} Let $X_j$ be a finite number of independent random $n\times n$ Hermitian positive semidefinite matrices with $\max_j{\|X_j\|_2}\leq \tau$. Define $$\omega_{min}\equiv\lambda_{min}\left(\sum_j{\mathbf{E}[X_j]}\right) \qquad \omega_{max}\equiv\lambda_{max}\left(\sum_j{\mathbf{E}[X_j]}\right),$$ and $f(x)\equiv e^x(1+x)^{-(1+x)}$. Then for any $0\leq \epsilon< 1$ $$\mathbf{Pr}\left[\lambda_{min}\left(\sum_j{X_j}\right) \leq (1-\epsilon) \>\omega_{min} \right]\leq n \> f(-\epsilon)^{\omega_{min}/\tau},$$ and for any $\epsilon\geq 0$ $$\mathbf{Pr}\left[\lambda_{max}\left(\sum_j{X_j}\right) \geq (1+\epsilon) \>\omega_{max} \right]\leq n \> f(\epsilon)^{\omega_{max}/\tau}.$$ \end{theorem} \subsection{Proof of Theorem~\ref{t_cohbound}}\label{s_tcohproof} We present a separate proof for each sampling method. \paragraph{Algorithm~\ref{alg_without}: Sampling without replacement} The proof follows directly from \cite[Lemma 3.4]{tropp11b}. \paragraph{Algorithm~\ref{alg_with}: Sampling with replacement} The proof is based on Theorem~\ref{t_tropp}, and turns out to be somewhat similar to that of \cite[Lemma 3.4]{tropp11b}. Set $X_t\equiv\frac{m}{c}\>Q^Te_{k_t}e_{k_t}^TQ$, $1\leq t\leq c$. Then $X_t$ is $n\times n$ Hermitian positive semidefinite and $\|X_t\|_2\leq\tfrac{m}{c}\|e_{k_t}^TQ\|_2^2\leq \tfrac{m\mu}{c}$. Hence we set $\tau=m\mu/c$. Furthermore, $$\mathbf{E}[X_t]=\sum_{j=1}^m{\tfrac{1}{m}\> \left(\tfrac{m}{c}\>Q^Te_je_j^TQ\right)}= \tfrac{1}{c}\>\sum_{j=1}^m{Q^Te_je_j^TQ}=\tfrac{1}{c} I_n.$$ Hence the eigenvalues of the sum are $\lambda_j\left(\sum_{t=1}^c{\mathbf{E}[X_t]}\right)=\lambda_j(I_n)=1$, $1\leq j\leq n$, and we set $\omega_{min}=\omega_{max}=1$. Applying Theorem~\ref{t_tropp} to $\sum_{t=1}^c{X_t}=Q^TS^TSQ$ gives \begin{eqnarray*} \mathbf{Pr}\left[\lambda_{min}\left(Q^TS^TSQ\right) \leq 1-\epsilon \right]\leq\ & n f(-\epsilon)^{c/(m\mu)}\\ \mathbf{Pr}\left[\lambda_{max}\left(Q^TS^TSQ\right) \geq 1+\epsilon \right]\leq\ & n f(\epsilon)^{c/(m\mu)}. \end{eqnarray*} The result follows from Boole's inequality \cite[p. 16]{Ross}. \paragraph{Algorithm~\ref{alg_bernoulli}: Bernoulli sampling} The proof is similar to the one above, and a special case of \cite[Theorem 6.1]{GT11}. Set $$X_j\equiv\tfrac{m}{c}\> \begin{cases} Q^Te_je_j^TQ & \text{with probability $\frac{c}{m}$}\\ 0_{n\times n}& \text{with probability $1-\frac{c}{m}$} \end{cases}, \qquad 1\leq j\leq m.$$ Then $X_j$ is $n\times n$ Hermitian positive semidefinite, $\|X_j\|_2\leq\tfrac{m}{c}\|e_j^TQ\|_2^2\leq \tfrac{m\mu}{c}$. As above, we set $\tau=m\mu/c$. Furthermore, $$\mathbf{E}[X_j]=\tfrac{c}{m}\cdot \tfrac{m}{c}Q^Te_je_j^TQ +(1-\tfrac{c}{m})\cdot 0_{n\times n}=Q^Te_je_j^TQ,$$ which implies $\sum_{j=1}^m{\mathbf{E}[X_j]}=\sum_{j=1}^m{Q^Te_je_j^TQ}=I_n$. Now proceed as in the above proof for Algorithm~\ref{alg_with}, and apply Theorem~\ref{t_tropp} to $\sum_{j=1}^m{X_j}=Q^TS^TSQ$. \subsection{Proof of Corollary~\ref{c_cohbound}}\label{s_ccohproof} First we simplify the bound in Theorem~\ref{t_cohbound} based on the inequality $f(-x)\leq f(x)$ for $0<x<1$ . This implies for Theorem~\ref{t_cohbound} that $$\delta\equiv n\left(f(-\epsilon)^{c/(m\mu)}+ f(\epsilon)^{c/(m\mu)}\right) \delta\leq 2n\>f(\epsilon)^{c/(m\mu)}.$$ Solving for $c$ gives \begin{eqnarray*} c\geq m\mu\>\frac{\ln(2n/\delta)}{-\ln{f(\epsilon)}}. \end{eqnarray*} If we can show that $-\ln{f(\epsilon)}>\epsilon^2/3$, then the above lower bound for $c$ definitely holds if $$c\geq 3m\mu\>\frac{\ln(2n/\delta)}{\epsilon^2}.$$ To show $-\ln{f(\epsilon)}>\epsilon^2/3$ for $0<\epsilon<1$, apply the definition $f(x)= e^x(1+x)^{-(1+x)}$ so that $h(x)\equiv -\ln{f(x)} = (1+x)\ln{(1+x)}-x$. Expand into the power series $\ln{(1+x)}=\sum_{j=1}^{\infty}{(-1)^{j+1} \tfrac{x^j}{j}}$. For $0<x< 1$ this yields $h(x)=\frac{1}{2}x^2-\frac{1}{6}x^3+E(x)$, where $$E(x)\equiv\sum_{j=4}^{\infty}{(-1)^j\>\frac{x^j}{(j-1)j}}= \sum_{j=2}^{\infty}{ \left(\frac{2j+1-(2j-1)x}{(2j-1)2j(2j+1)}\right)\> x^{2j}}>0,$$ since each summand is positive for $0<x< 1$. Thus for $0<x< 1$ we obtain $$h(x)>\frac{1}{2}x^2-\frac{1}{6}x^3= \frac{3-x}{6}\>x^2\geq \frac{x^2}{3}.$$ \subsection{Matrix Bernstein concentration inequality}\label{s_bernstein} The matrix concentration inequality below is the basis for Theorem~\ref{t_levbound}. It is a version specialized to square matrices of \cite[Theorem 4]{Recht11}. In numerical experiments we found it to be tighter than \cite[Theorem 4]{DMMS10} and the Frobenius norm bound \cite[Theorem 2]{DKM06}. \begin{theorem}[Theorem 4 in \cite{Recht11}] Let $X_j$ be $m$ independent random $n\times n$ matrices with $\mathbf{E}[X_j]=0_{n\times n}$, $1\leq j\leq m$. Let $\rho_j\equiv \max\{\|\mathbf{E}[X_jX_j^T]\|_2, \, \|\mathbf{E}[X_j^TX_j]\|_2\}$ and $\max_{1\leq j\leq m}{\|X_j\|_2}\leq \tau$. Then for any $\epsilon >0$ $$\mathbf{Pr}\left[\|\sum_{j=1}^m{X_j}\|_2> \epsilon\right]\leq 2n\>\exp\left(-\tfrac{3}{2}\>\frac{\epsilon^2}{3\sum_{j=1}^m{\rho_j} +\tau\epsilon}\right).$$ \end{theorem} \subsection{Proof of Theorem~\ref{t_levbound}}\label{s_tlevproof} The proof is similar to that of \cite[Lemma 3]{BalzRN11}. Represent the outcome of uniform sampling with replacement in Algorithm~\ref{alg_with} by $Q^TS^TSQ=\sum_{t=1}^c{Y_t}$, where $Y_t\equiv \frac{m}{c}\>Q^Te_{k_t}e_{k_t}^TQ$ are $n\times n$ matrices, $1\leq t\leq c$, with expected value $$\mathbf{E}[Y_t]=\sum_{j=1}^m{\frac{1}{m}\>\frac{m}{c} Q^Te_je_j^TQ} =\frac{1}{c}\>\sum_{j=1}^m{Q^Te_je_j^TQ}=\frac{1}{c} I_n.$$ Thus, the zero mean versions are $X_t\equiv Y_t-\tfrac{1}{c}I_n$. To apply Theorem \cite[Theorem 4]{Recht11} to the $X_t$ we need to verify that they fulfill the required conditions. First, by construction, $\mathbf{E}[X_t]=0$, $1\leq t\leq c$. Second, since $Y_t$ and $I_n$ are symmetric positive semidefinite, $$\|X_t\|_2\leq \max\{\|Y_t\|_2, \|\frac{1}{c}I_n\|_2\}= \frac{1}{c}\max\{m\>\|e_{k_t}^TQ\|_2^2,1\}\leq \frac{m\mu}{c},$$ where the last inequality follows from the definition of $\mu$, and $\mu\geq n/m$. Hence we set $\tau=m\mu/c$. Third, since $X_t$ is symmetric, $$X_t^TX_t=X_tX_t^T=X_t^2=Y_t^2 -\frac{2}{c} Y_t+\frac{1}{c^2}I_n.$$ From $\mathbf{E}[Y_t]=\tfrac{1}{c} I_n$ follows \begin{eqnarray}\label{e_ys} \mathbf{E}[X_t^2]=\mathbf{E}[Y_t^2]-\frac{2}{c}\>\mathbf{E}[Y_t]+\frac{1}{c^2}I_n= \mathbf{E}[Y_t^2]-\frac{1}{c^2}I_n. \end{eqnarray} Since $Y_t^2=\tfrac{m^2}{c^2}\>\ell_{k_t}\>Q^Te_{k_t}e_{k_t}^TQ$, we obtain \begin{eqnarray*} \mathbf{E}[Y_t^2]=\sum_{j=1}^m{\frac{1}{m}\> \frac{m^2}{c^2}\>\ell_j Q^Te_je_j^TQ} =\frac{m}{c^2}\>Q^T \> \left(\sum_{j=1}^m{\ell_je_je_j^T}\right)\>Q = \frac{m}{c^2}\> Q^TLQ. \end{eqnarray*} Substituting this into (\ref{e_ys}) yields $$\mathbf{E}[X_t^2]=\tfrac{1}{c^2} \> \left(m\> Q^TLQ - I_n\right).$$ Positive semi-definiteness gives $$\|\mathbf{E}[X_t^2]\|_2\leq \frac{1}{c^2}\>\max\{m\>\|Q^TLQ\|_2,\, 1\}= \frac{m}{c^2}\>\|Q^TLQ\|_2.$$ We set $\rho_t=\tfrac{m}{c^2}\> \|Q^TLQ\|_2$. Applying \cite[Theorem 4]{Recht11} to $$\sum_{t=1}^c{X_t}=\sum_{t=1}^c{\left(Y_t-\tfrac{1}{c}I_n\right)}= (SQ)^T(SQ)- I_n$$ shows that $\|\sum_{t=1}^c{X_t}\|_2\leq \epsilon$ with probability at least $1-\delta$. \section{Two-norm bound for scaled matrices, and proofs for Sections \ref{s_clevbound} and~\ref{s_ancomp}}\label{s_app2} We derive a bound for the two-norm of diagonally scaled matrices (Section~\ref{s_tns}), which leads immediately to the proofs of Corollary~\ref{c_1} (Section~\ref{s_lqproof}), and Corollary~\ref{c_comp} (Section~\ref{s_compproof}). \subsection{Bound}\label{s_tns} We present two majorization bounds for Hadamard products of vectors (Lemmas \ref{l_s1} and~\ref{l_s2}), and use them to derive a bound for the two-norm of diagonally scaled matrices (Theorem~\ref{t_2}). \begin{definition}[Definition 4.3.41 in \cite{HoJ12}]\label{d_maj} Let $a$ and $b$ be vectors with $m$ real elements. The elements, labelled in algebraically decreasing order, are $a_{[1]}\geq \cdots\geq a_{[m]}$ and $b_{[1]}\geq \cdots\geq b_{[m]}$. The vector $a$ {\rm weakly majorizes} the vector $b$, if $$\sum_{j=1}^k{a_{[j]}}\geq \sum_{j=1}^k{b_{[j]}}, \qquad 1\leq k\leq m.$$ The vector $a$ {\rm majorizes} the vector $b$, if $a$ weakly majorizes $b$ and also $\sum_{j=1}^m{a_{[j]}}=\sum_{j=1}^m{b_{[j]}}$. \end{definition} The first lemma follows from a stronger majorization inequality for functions that are monotone and lattice superadditive. \begin{lemma}[Theorem II.4.2 in \cite{Bha97}]\label{l_s2} If $b$ and $x$ are vectors with $m$ non-negative elements, then $$\sum_{j=1}^k{b_j\, x_j}\leq \sum_{j=1}^k{b_{[j]}\, x_{[j]}}, \qquad 1\leq k\leq m.$$ \end{lemma} The second lemma is a variant of a well-known majorization result for Hadamard products of vectors \cite[Lemma 4.3.51]{HoJ12}. Since the version below is slightly different, we include a proof from first principles. \begin{lemma}\label{l_s1} Let $x$, $a$ and $b$ be vectors with $m$ non-negative elements. If $a$ weakly majorizes $b$, then $$\sum_{j=1}^k{a_{[j]}\, x_{[j]}}\geq \sum_{j=1}^k{b_{[j]}\, x_{[j]}}, \qquad 1\leq k\leq m.$$ \end{lemma} \begin{proof} The following arguments hold for $1\leq k\leq m-1$. Start out with the upper bound, and separate the last summand, \begin{eqnarray}\label{e_m3} \sum_{j=1}^{k+1}{a_{[j]}\,x_{[j]}} &=& \sum_{j=1}^{k}{a_{[j]}\,x_{[j]}} + a_{[k+1]}\,x_{[k+1]}. \end{eqnarray} Re-writing the right sum and applying $x_{[j]}\geq x_{[k+1]}\geq 0$, $1\leq j\leq k$, gives \begin{eqnarray*} \sum_{j=1}^{k}{a_{[j]}\,x_{[j]}} &=& \sum_{j=1}^{k}{b_{[j]}\,x_{[j]}} + \sum_{j=1}^{k}{(a_{[j]}-b_{[j]})\,x_{[j]}}\\ &\geq & \sum_{j=1}^{k}{b_{[j]}\,x_{[j]}} + \sum_{j=1}^{k}{(a_{[j]}-b_{[j]})}\, x_{[k+1]} \\ &= & \sum_{j=1}^{k}{b_{[j]}\,x_{[j]}} + \left(\sum_{j=1}^{k}{a_{[j]}}-\sum_{j=1}^{k}{b_{[j]}}\right)\, x_{[k+1]}.\\ \end{eqnarray*} Insert this into (\ref{e_m3}) and gather common terms, \begin{eqnarray*} \sum_{j=1}^{k+1}{a_{[j]}\,x_{[j]}} &\geq & \sum_{j=1}^{k}{b_{[j]}\,x_{[j]}} + \left(\sum_{j=1}^{k+1}{a_{[j]}} -\sum_{j=1}^{k}{b_{[j]}}\right)\, x_{[k+1]}\\ &\geq &\sum_{j=1}^{k}{b_{[j]}\,x_{[j]}} + b_{[k+1]}\,x_{[k+1]} =\sum_{j=1}^{k+1}{b_{[j]}\,x_{[j]}}, \end{eqnarray*} where the second inequality follows from the majorization $\sum_{j=1}^{k+1}{a_{[j]}}\geq \sum_{j=1}^{k+1} {b_{[j]}}$. \end{proof} Now we are ready to bound the two norm of a row scaled matrix $DZ$, where $Z$ is $m\times n$ of full column rank, and $D=\diag\begin{pmatrix}d_1 &\ldots & d_m\end{pmatrix}$ is a non-negative $m\times m$ diagonal matrix. The obvious bound is \begin{eqnarray}\label{e_sub} \|DZ\|_2\leq \|D\|_2\>\|Z\|_2=d_{[1]}\>\|Z\|_2. \end{eqnarray} However, the bound in Theorem~\ref{t_2} below, which incorporates the largest row norm of $Z$ and several of the largest (in magnitude) diagonal elements of~$D$, turns out to be tighter. \begin{theorem}\label{t_2} Let $Z$ be a real $m\times n$ matrix with $\rank(Z)=n$, smallest singular value $\sigma_z=1/\|Z^{\dagger}\|_2$, and largest squared row norm $\mu_z\equiv\max_{1\leq j\leq m}{\|e_j^TZ\|_2^2}$. If $t\equiv\left\lfloor\sigma_z^2/\mu_z\right\rfloor$, then $$\|DZ\|_2^2\leq \begin{cases} \mu_z\sum_{j=1}^t{d_{[j]}^2}+\left(\|Z\|_2^2-t\,\mu_z\right)\,d_{[t+1]}^2 & \text{if}~ \|Z\|_2^2-t\,\mu_z\leq \mu_z \\ \mu_z\sum_{j=2}^{t+1}{d_{[j]}^2}+\left(\|Z\|_2^2-t\,\mu_z\right)\,d_{[1]}^2 & \text{otherwise}.\end{cases}$$ \end{theorem} \begin{proof} Let $z$ be a $n\times 1$ vector with $\|z\|_2=1$ and $\|DZ\|_2=\|DZz\|_2$. Furthermore let $z_j\equiv e_j^TZz$, $1\leq j\leq m$, be the elements of $Zz$, so that $\|Zz\|_2^2=\sum_{j=1}^m{z_j^2}$. \smallskip \paragraph{Apply Lemma~\ref{l_s2}} Since $d_j^2\geq 0$ and $z_{j}^2\geq 0$, $1\leq j\leq m$, we can apply Lemma~\ref{l_s2} with $x_j=d_{j}^2$ and $b_j=z_{j}^2$, to obtain \begin{eqnarray*} \|DZ\|_2^2=\|DZz\|_2^2= \sum_{j=1}^m{d_{j}^2\,z_{j}^2}= \sum_{j=1}^m{b_j\, x_j}\leq \sum_{j=1}^m{b_{[j]}\,x_{[j]}}. \end{eqnarray*} \paragraph{Verify assumptions of Lemma~\ref{l_s1}} In order to apply Lemma~\ref{l_s1} with $$a_j=\mu_z, \quad 1\leq j\leq t,\qquad a_{t+1}=\|Zz\|_2^2-t\,\mu_z, \qquad a_j=0, \quad t+2\leq j\leq m,$$ we need show that the assumptions are satisfied, meaning all vector elements are non-negative and the majorization condition holds. Clearly $a_j\geq 0$ for $1\leq j\leq t$ and $t+2\leq j\leq m$. This leaves $a_{t+1}$. From $\rank(Z)=n$ follows that $\sigma_z>0$. The definition of $t$ implies $0\leq t\leq \sigma_z^2/\mu_z$, so that $$0\leq \sigma_z^2-t\,\mu_z=\min_{\|y\|_2=1}{\|Zy\|_2^2}-t\,\mu_z\leq \|Zz\|_2^2-t\,\mu_z=a_{t+1}.$$ Thus, all vector elements are non-negative. To show the majorization condition, start with the Cauchy-Schwartz inequality, $$b_j=z_{j}^2=(e_{j}^TZ\>z)^2\leq \|e_{j}^TZ\|_2^2\>\|z\|_2^2=\|e_{j}^TZ\|_2^2\leq \mu_z, \qquad 1\leq j\leq m.$$ This yields, regardless of whether $a_{t+1}=\|Zz\|_2^2-t\,\mu_z\leq \mu_z$ or not, $$\sum_{j=1}^k{a_{[j]}}\geq\sum_{j=1}^k{\mu_z}\geq \sum_{j=1}^k{z_{[j]}^2}= \sum_{j=1}^k{b_{[j]}},\qquad 1\leq k\leq t.$$ Moreover, for $1\leq k\leq m-t$, $$\sum_{j=1}^{t+k}{a_{[j]}}=\sum_{j=1}^t{\mu_z}+(\|Zz\|_2^2-t\,\mu_z) =\|Zz\|_2^2\geq \sum_{j=1}^{t+k}{z_{[j]}^2}=\sum_{j=1}^{t+k}{b_{[j]}}.$$ This gives the weak majorization condition $\sum_{j=1}^k{a_{[j]}}\geq \sum_{j=1}^k{b_{[j]}}$, $1\leq k\leq m$. \smallskip \paragraph{Apply Lemma \ref{l_s1}} Since the assumptions of Lemma~\ref{l_s1} are satisfied, we can conclude that $\sum_{j=1}^{m}{b_{[j]}\,x_{[j]}}\leq \sum_{j=1}^{m}{a_{[j]}\,x_{[j]}}$. At last, substitute into this majorization relation the expressions for $a$ and $b$. If $\|Z\|_2^2-t\,\mu_z \leq \mu_z$, then $$\sum_{j=1}^m{a_{[j]}\,x_{[j]}}=\mu_z\sum_{j=1}^t{d_{[j]}^2}+ (\|Zz\|_2^2-t\,\mu_z)\,d_{[t+1]}^2\leq \mu_z\sum_{j=1}^t{d_{[j]}^2}+(\|Z\|_2^2-t\,\mu_z)\,d_{[t+1]}^2,$$ otherwise $$\sum_{j=1}^m{a_{[j]}\,x_{[j]}}=(\|Zz\|_2^2-t\,\mu_z)\,d_{[1]}^2+ \mu_z\sum_{j=2}^{t+1}{d_{[j]}^2}\leq (\|Z\|_2^2-t\,\mu_z)\,d_{[1]}^2+ \mu_z\sum_{j=2}^{t+1}{d_{[j]}^2}.$$ \end{proof} Theorem~\ref{t_2} is tighter than (\ref{e_sub}) because $d_{[j]}^2\leq \|D\|_2^2$ implies \begin{eqnarray*} \|DZ\|_2^2 & \leq & \begin{cases} \mu_z\sum_{j=1}^t{d_{[j]}^2}+\left(\|Z\|_2^2-t\,\mu_z\right)\,d_{[t+1]}^2 & \text{if}~ \|Z\|_2^2-t\,\mu_z\leq \mu_z \\ \mu_z\sum_{j=2}^{t+1}{d_{[j]}^2}+\left(\|Z\|_2^2-t\,\mu_z\right)\,d_{[1]}^2 & \text{otherwise}\end{cases}\\ & \leq & t\mu_z \|D\|_2^2 -\left(\|Z\|_2^2-t\,\mu_z\right)\,\|D\|_2^2 = \|D\|_2^2\|Z\|_2^2. \end{eqnarray*} \begin{corollary}\label{c_t2} Let $Z$ be a real $m\times n$ matrix with $Z^TZ=I_n$, and coherence $\mu_z\equiv\max_{1\leq j\leq m}{\|e_j^TZ\|_2^2}$. If $t\equiv\left\lfloor 1/\mu_z\right\rfloor$, then $$\|DZ\|_2^2\leq \mu_z\sum_{j=1}^t{d_{[j]}^2}+ \left(1-t\,\mu_z\right)\,d_{[t+1]}^2.$$ \end{corollary} \begin{proof} Applying Theorem \ref{t_2} and assuming $1-t\,\mu_z\leq \mu_z$ gives $$\|DZ\|_2^2\leq \mu_z\sum_{j=1}^t{d_{[j]}^2}+ \left(1-t\,\mu_z\right)\,d_{[t+1]}^2.$$ The assumption $1-t\,\mu_z\leq \mu_z$ is justified because $$1-t\,\mu_z = 1-\left\lfloor 1/\mu_z\right\rfloor\,\mu_z \leq 1-(1/\mu_z - 1)\mu_z = \mu_z.$$ \end{proof} \subsection{Proof of Corollary~\ref{c_1}}\label{s_lqproof} Apply Corollary~\ref{c_t2} with $D=L^{1/2}$, $Z=Q$, $\mu_z=\mu$, and $t=\lfloor 1/\mu\rfloor$ to prove the first inequality, $$\|Q^TLQ\|_2=\|L^{1/2}Q\|_2^2\leq \mu \> \sum_{j=1}^t{\ell_{[j]}}+(1-t\,\mu)\,\ell_{[t+1]}.$$ As for the second inequality, $\ell_{[j]}\leq \mu$ implies $$\mu\>\sum_{j=1}^t{\ell_{[j]}}+(1-t\,\mu)\,\ell_{[t+1]}\ \leq t\mu^2+ (1-t\mu)\mu=\mu.$$ If, in addition, $t$ is an integer, then $t=1/\mu$ and $1-t\,\mu=0$. \subsection{Proof of Corollary~\ref{c_comp}}\label{s_compproof} Define the common term $\phi\equiv m\>\ln(2n/\delta)/\epsilon^2$ in both bounds, and write Corollary~\ref{c_cohbound} as $c\geq 3\mu\>\phi$, and Corollary~\ref{c_levbound} as $c\geq (2\tau+\tfrac{2}{3}\epsilon\,\mu)\>\phi$. From $\epsilon<1$ and $\tau\leq \mu$ follows $$2\tau+\tfrac{2}{3}\epsilon\,\mu\leq 3\mu.$$
{ "timestamp": "2014-03-06T02:02:14", "yymm": "1203", "arxiv_id": "1203.4809", "language": "en", "url": "https://arxiv.org/abs/1203.4809", "abstract": "Motivated by the least squares solver Blendenpik, we investigate three strategies for uniform sampling of rows from m x n matrices Q with orthonormal columns. The goal is to determine, with high probability, how many rows are required so that the sampled matrices have full rank and are well-conditioned with respect to inversion.Extensive numerical experiments illustrate that the three sampling strategies (without replacement, with replacement, and Bernoulli sampling) behave almost identically, for small to moderate amounts of sampling. In particular, sampled matrices of full rank tend to have two-norm condition numbers of at most 10.We derive a bound on the condition number of the sampled matrices in terms of the coherence {\\mu} of Q. This bound applies to all three different sampling strategies; it implies a, not necessarily tight, lower bound of O(m {\\mu} ln(n)) for the number of sampled rows; and it is realistic and informative even for matrices of small dimension and the stringent requirement of a 99 percent success probability.For uniform sampling with replacement we derive a potentially tighter condition number bound in terms of the leverage scores of Q. To obtain a more easily computable version of this bound, in terms of just the largest leverage scores, we first derive a general bound on the two-norm of diagonally scaled matrices.To facilitate the numerical experiments and test the tightness of the bounds, we present algorithms to generate matrices with user-specified coherence and leverage scores. These algorithms, the three sampling strategies, and a large variety of condition number bounds are implemented in the Matlab toolbox kappa_SQ_v3.", "subjects": "Numerical Analysis (math.NA)", "title": "The Effect of Coherence on Sampling from Matrices with Orthonormal Columns, and Preconditioned Least Squares Problems", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9793540722737479, "lm_q2_score": 0.8354835289107309, "lm_q1q2_score": 0.8182341963563659 }
https://arxiv.org/abs/1006.3049
Long paths and cycles in subgraphs of the cube
Let $Q_n$ denote the graph of the $n$-dimensional cube with vertex set $\{0,1\}^n$ in which two vertices are adjacent if they differ in exactly one coordinate. Suppose $G$ is a subgraph of $Q_n$ with average degree at least $d$. How long a path can we guarantee to find in $G$?Our aim in this paper is to show that $G$ must contain an exponentially long path. In fact, we show that if $G$ has minimum degree at least $d$ then $G$ must contain a path of length $2^d-1$. Note that this bound is tight, as shown by a $d$-dimensional subcube of $Q_n$. We also obtain the slightly stronger result that $G$ must contain a cycle of length at least $2^d$ and prove analogous results for other `product-type' graphs.
\section{Introduction} Given a graph $G$ of average degree at least $d$, a classical result of Dirac \cite{dirac} guarantees a path of length $d$ in $G$. Moreover, this bound is best possible as can be seen from $K_{d+1}$. Inside the cube $Q_n$ can we improve this bound? That is, given a subgraph $G$ of $Q_n$ with average degree at least $d$, what is the length of the longest path in $G$? The edge isoperimetric inequality for the cube (\cite{bernstein}, \cite{harper}, \cite{hart}, \cite{lindsey}, see \cite{com} for background) says that any subgraph of average degree at least $d$ must have size at least $2^d$. In light of this, the above linear bound seems very weak. A natural subgraph of $Q_n$ with average degree at least $d$ is the $d$-dimensional cube $Q_d$, the analogue of the complete graph in $Q_n$, which contains a path of length $2^d-1$. Must the size of the longest path in $G$ also be exponential? The main result of this paper answers this question in the affirmative. \begin{thm} { \label{mainthm} Every subgraph $G$ of $Q_n$ with minimum degree $d$ contains a path of length $2^{d}-1$. } \end{thm} Note that this is best possible as shown by a $d$-dimensional subcube of $Q_n$. In fact, the proof of Theorem \ref{mainthm} shows that we can always find a longer path in $G$ unless it is isomorphic to $Q_d$. Using the well known fact that every graph with average degree at least $d$ contains a subgraph with minimum degree at least $\frac{d}{2}$ we obtain the following corollary to Theorem \ref{mainthm}. \begin{cor} { \label{corollary} Every subgraph $G$ of $Q_n$ with average degree at least $d$ contains a path of length at least $2^{\frac{d}{2}}-1$. } \end{cor} We do not know a tight bound for average degree $d$. We also obtain the corresponding result for the length of the longest cycle in subgraphs of $Q_n$ with large minimum degree. \begin{thm} { \label{cyclethm} Every subgraph $G$ of $Q_n$ with minimum degree $d$ contains a cycle of length at least $2^d$. } \end{thm} In Section 2 we give an overview of the proofs of Theorems \ref{mainthm} and \ref{cyclethm}. The theorems themselves are then proved in Sections 3-7. In Section 8 we show that the lower bound from Theorems \ref{mainthm} and \ref{cyclethm} also extends to subgraphs of the grid graph $\Z^n$ and the discrete torus $C_k^n$, for all $k\geq 4$. We also give a generalization of Theorem \ref{mainthm} and \ref{cyclethm} to general `product-type' graphs and make some conjectures. \section{Overview} As in the statement of Theorem \ref{mainthm}, let $G$ be a subgraph of $Q_n$ with $\delta (G)\geq d$. We will view the vertices of $Q_n$ as elements of the power set of $[n]$, $\mathcal{P}[n]$. A plausible approach to proving Theorem \ref{mainthm} is to split $G$ along some direction $i$ to obtain two induced subgraphs $G_1$ and $G_2$ consisting of those vertices of $G$ respectively containing and not containing $i$, for some $i\in [n]$. Provided such a direction is chosen to ensure that $G_1, G_2\neq \emptyset $, we have $\delta (G_i)\geq d-1$ and by induction on Theorem \ref{mainthm} we have a path of length $2^{d-1}-1$ in each subgraph. If we could join these two paths into one we would clearly be done. However, as Theorem \ref{mainthm} provides no information on where these paths start or end, we can not expect to be able to do this. This suggests that we strengthen Theorem \ref{mainthm} to guarantee an exponentially long path between \emph{any} two vertices $x$ and $y$ of $G$. In general this is not possible -- for example, consider the graph $G'$ obtained by removing all but one edge $xy$ of direction $d+1$ from the ($d+1$)-dimensional cube $Q_{d+1}$. However this graph is not 2-connected. The following theorem says that this is the only obstruction to such a strengthening. \begin{thm} { \label{conn} Let $G$ be a 2-connected subgraph of $Q_n$ and $a$ and $b$ be distinct vertices of $G$. Suppose that $d_G(z) \geq d$ for all $z \in G-\lbrace a,b\rbrace$. Then $a$ and $b$ are joined by a path of length at least $2^{d}-2$. Furthermore, unless $G$ is isomorphic to $Q_d$ with $a$ and $b$ at even Hamming distance, $G$ contains an $a-b$ path of length at least $2^{d}-1$. } \end{thm} Note that we do not assume that $a$ or $b$ have degree at least $d$ in Theorem \ref{conn}. This slight weakening of the minimum degree condition will allow us to use induction on various subgraphs of $G$ which would otherwise not be available. Before continuing with the overview we make a small diversion to introduce some definitions: these are standard (e.g. see \cite{mgt}). \\ A subgraph $B$ of a graph $G$ is a \emph{block} of $G$ if $B$ is either a bridge of $G$ or forms a maximal 2-connected subgraph of $G$. By maximality, $|B_1\cap B_2|\leq 1$ for any two blocks $B_1$ and $B_2$ of $G$ and $G-E(B)$ contains no $x-y$ path between distinct vertices $x,y$ in a block $B$. Therefore if any two blocks intersect, their common vertex must be a cutvertex and conversely every cutvertex lies in at least two blocks. Since every cycle is 2-connected and an edge is a bridge iff it does not lie in any cycle, every graph $G$ decomposes uniquely into its blocks $B_1,\ldots, B_p$ in the sense that: \begin{equation*} { E(G)=\bigcup_{i=1}^p E(B_i) \mbox{ and } E(B_i)\cap E(B_j)=\emptyset \mbox{ if } i\neq j. } \end{equation*} Suppose now that $G$ is connected. Let $\mathcal{B}(G)$, the \emph{block-cutvertex graph} of $G$, be the bipartite graph with bipartition $(\mathcal{B},\mathcal{C})$ where $\mathcal{B}$ is the set of blocks of $G$, $\mathcal{C}$ is the set of cutvertices of $G$ with $Bc$ an edge if $c\in B$. For a connected graph $G$, $\mathcal{B}(G)$ is a tree. The leaves of this tree are all elements of $\mathcal{B}$ and are called \emph{endblocks}. Given an endblock $E$ we will denote its unique cutvertex by cutv$(E)$. Note that a graph $G$ has only one endblock iff it is 2-connected. \vspace{.2cm} We now return to the overview of the proof of Theorem \ref{conn}. \begin{lem} { \label{splitting} Let $G$ be a connected subgraph of $Q_n$ with $a$ and $b$ distinct vertices of $G$. Then there exists a partition of $G$ into two connected subgraphs $G_a$ and $G_b$ such that $a\in G_a$, $b\in G_b$ and for all $v\in G_c$, $d_{G_c}(v)\geq d_G(v)-1$, where $c\in \{a,b\}$. } \end{lem} \begin{proof} { Picking a splitting direction $i$ such that $a$ and $b$ differ in coordinate $i$ and forming $G_1$ and $G_2$ as before, we have $a\in G_1$ and $b\in G_2$. Let $C_b$ be the connected component of $G_2$ containing $b$. Taking $G_a$ to be the connected component of $G-C_b$ containing $a$ and $G_b=G-G_a$ we are done. } \end{proof} A central observation in the proof of Theorem \ref{conn} is that, provided $d\geq 3$, given any endblock $E$ of $G_a$ with $a\notin E$, by induction on Theorem \ref{conn}, $E$ contains a path of length at least $2^{d-1}-2$ from $\mbox{cutv}(E)$ to any $y\in E-\mbox{cutv}(E)$ -- here $d\geq 3$ guarantees the $E$ is 2-connected and not a bridge. Since $G$ is 2-connected there must exist $y\in E-\mbox{cutv}(E)$ with a neighbour in $G_b$. Thus endblocks like $E$ guarantee `endblock paths' of length at least $2^{d-1}-1$ from a point in $G_a$ to one in $G_b$. If we can find a path from $a$ to $b$ containing at least two such endblock paths we would almost be done (we might still be short two or three vertices to give the $2^{d}-2$ or $2^d-1$ bound in $G$). For ease of exposition we will prove the following weakening of Theorem \ref{conn} first. It will allow the reader to focus on the main ideas of the proof of Theorem \ref{conn} without some distracting details needed to ensure that an $a-b$ path formed from endblock paths is not slightly too short. \begin{thm} { \label{connmod} Let $G$ be a 2-connected subgraph of $Q_n$ and $a,b\in V(G)$. Suppose that $d_G(z) \geq d$ for all $z \in V(G)-\lbrace a,b\rbrace$. Then $G$ contains an $a-b$ path of length at least $2^{d-1}$. } \end{thm} Another slight technicality that creeps into the proof of Theorem \ref{conn} and \ref{connmod} is the possibility that the only partitions of $G$ into $G_a$ and $G_b$ as in Lemma \ref{splitting} above, have $a$ with just one neighbour in $G_a$ or $b$ with just one neighbour in $G_b$. While all cases can be dealt with simultaneously, we felt for clarity's sake it is easier to first restrict attention to the case where a partition direction $i$ exists for which both $d_{G_a}(a)\geq 2$ and $d_{G_b}(b)\geq 2$. \\ Theorem \ref{connmod} is proved in Sections 3-6. Sections 3-5 will focus on the above case, that is, where we can find a partition direction $i$, such that $d_{G_a}(a)\geq 2$ and $d_{G_b}(b)\geq 2$. Section 3 will describe the block-cutvertex decomposition structure of $G_a$ and $G_b$ on the absence of an $a-b$ path of length $2^{d-1}$ formed by joining at least two endblock paths together and Section 4 describes how the endblocks of $G_a$ interact with those of $G_b$. In Section 5 we show that if $G$ does not contain a path from $a$ to $b$ containing at least two endblock paths then the conditions of Theorem \ref{connmod} hold for a smaller subgraph of $G$. This allows for an inductive step and completes the proof of Theorem \ref{connmod} in this case. Section 6 will allow us, using a small modification of the argument from Sections 3-5, to extend from the case $d_{G_a}(a)\geq 2$ and $d_{G_b}(b)\geq 2$ to the general case, therefore proving Theorem \ref{connmod}. Finally in Section 7 we show how to adjust the approach in Sections 3-6 to obtain the optimal bound of Theorem \ref{conn}. To close this section we note that Theorem \ref{conn} implies Theorem \ref{cyclethm}. \vspace{0.2cm} \noindent \textit{Proof of Theorem \ref{cyclethm}:} Take an endblock $E$ in the block cutvertex decomposition of $G$. Clearly $E$ is 2-connected and all vertices in $E-\mbox{cutv}(E)$ have at least $d$ neighbours in $E$. Pick a neighbour $v$ of $\mbox{cutv}(E)$ in $E$. Then by Theorem \ref{conn} $G$ contains a $\mbox{cutv}(E)-v$ path $P$ of length at least $2^d-1$. Combining $P$ with the edge $\mbox{cutv}(E)v$ we obtain the desired cycle. \hspace{5cm} $\square$ \section{Endblocks in $G_a$ and $G_b$} To begin we introduce some useful definitions. \begin{dfn} { \emph{ Let $E$ be an endblock in the block-cutvertex decomposition of $G_a$ ($G_b$). The \emph{interior} of $E$ is the set $\mbox{int}(E)=E-\mbox{cutv}(E)$. A vertex $x\in \mbox{int}(E)$ is said to be an \emph{exit vertex} of $E$ if $x$ has a neighbour in $G_b$ ($G_a$). If this neighbour exists, it is unique and is denoted by $p(x)$, $x$'s \emph{partner}.} } \end{dfn} \begin{dfn} { \label{why} \emph { Body$(a)$ is the intersection of all blocks of $G_a$ containing $a$. Let $\mbox {Core}(a)$ consist of those vertices in $\mbox {Body}(a)$ that are not cutvertices of $G_a$. } } \end{dfn} \begin{dfn} { \emph { A subgraph $K$ of $G_a$ is said to be a \emph{limb} of $a$ if:} \begin{itemize} \item \emph{$a$ is a cutvertex of $G_a$ and $K=G\lbrack C\cup \lbrace a\rbrace \rbrack $ where $C$ is a connected component of $G_a-a $;} \item \emph{$a$ is not a cutvertex of $G_a$ and $K=G\lbrack C \rbrack $ where $C$ is a connected component of $G_a-\mbox {Core}(a)$.} \end{itemize} \emph{ The \emph{joint} of a limb $K$, Joint$(K)$, is the unique vertex $v\in K\cap \mbox {Body}(E)$. } } \end{dfn} The reader may find it helpful to examine Figure \ref{intro}. The circles and ellipses will always denote blocks in the block-cutvertex decomposition of graph. \vspace{0.3cm} \noindent \emph{Proof of Theorem \ref{connmod}.} The proof is by induction on $d$ where the base case $d=2$ is trivial. The proof will last until the end of Section 6. Suppose for contradiction, the theorem fails for $d$ and take $G$ to be a minimal counterexample so that Theorem \ref{connmod} holds for all smaller degrees and all graphs $G'$ with $|G'|<|G|$. The following theorem restricts the possibilities for $G$. \vspace{0.3cm} \noindent \textbf{Theorem 2.3$'$.} \emph{Suppose $G$ is a counterexample to Theorem \ref{connmod} so that Theorem \ref{connmod} holds for all smaller degrees and all graphs $G'$ with $|G'|<|G|$. Then there does not exist a direction $i$ such that forming $G_a$ and $G_b$ as in Lemma \ref{splitting}, $d_{G_a}(a)\geq 2$ and $d_{G_b}(b)\geq 2$.} \vspace{0.2cm} \noindent \emph{Proof of Theorem 2.3$'$.} The proof will last until the end of Section 5. Suppose for contradiction that such a $G$ and $i$ exist and form $G_a$ and $G_b$ as above from direction $i$. Our first lemma describes the block structure of $G_a$ provided we cannot use endblock paths to form an $a-b$ path of length at least $2^{d-1}$. \begin{figure} \centering \input{intro.pstex_t} \caption{The diagram displays various parts of $G_a$ and $G_b$. The broken line separates $G_a$ and $G_b$. In $G_a$, $\mbox{Body}(a)\neq \{a\}$ and $a$ has three limbs. In $G_b$, $b$ is a cutvertex and one of its limbs $L$ contains an endblock $F$ with exit vertex $x$.\label{intro}} \end{figure} \begin{lem} { \label{list} Given $G$ the following hold: \begin{itemize} { \item[\emph{(i)}] Every endblock of $G_a$ which does not contain $a$ in its interior must contain at least two exit vertices. \item[\emph{(ii)}] $G_a$ is not 2-connected. \item[\emph{(iii)}] $a$ does not lie in the interior of an endblock in $G_a$. \item[\emph{(iv)}] $a$ must have at least two limbs. } \end{itemize} } \end{lem} \begin{proof} { (i) Suppose not and let $E$ be such an endblock. By the 2-connectivity of $G$, $E$ must contain an exit vertex $x$. If $x$ were its only exit vertex then every $v\in E-\lbrace {\mbox{cutv}(E),x}\rbrace $ has degree least $d$ in $G[E]$ -- such $v$ must exist since $d\geq 3$. Then by choice of $G$, $G[E]$ contains a path $P_2$ of length at least $2^{d-1}$ from $\mbox{cutv}(E)$ to $x$. Joining $a$ to $\mbox{cutv}(E)$ in $G_a$ by a path $P_1$ and $p(x)$ to $b$ in $G_b$ by a path $P_3$ we have created a path $P_1P_2P_3$ of length $2^{d-1}$ from $a$ to $b$, a contradiction. (ii) Suppose $G_a$ is 2-connected. First consider the case where $G_b$ is not 2-connected. Let $E$ be an endblock in $G_b$ not containing $b$ in its interior and take $x$ to be an exit vertex of $E$ with $p(x)\neq a$ -- this exists by (i). Then by induction on $d$, there are paths $P_1$ in $G_a$ from $a$ to $p(x)$ and $P_2$ in $G[E]$ from $x$ to $\mbox{cutv}(E)$ both of length at least $2^{d-2}$. Taking a path $P_3$ from $\mbox{cutv}(E)$ to $b$ in $G_b$ we have constructed a path $P=P_1p(x)xP_2P_3$ from $a$ to $b$ of length at least $2^{d-1}$, a contradiction. If $G_b$ is 2-connected then the same proof as in (i) shows that $G_b$ must contain two exit vertices one of which $x$ has $x\neq b$ and $p(x) \neq a$. By induction on $d$ we obtain endblock paths from $a$ to $p(x)$ in $G_a$ and from $x$ to $b$ in $G_b$ both of length at least $2^{d-2}$. Joining the two with edge $xp(x)$, $G$ contains the required path, a contradiction. (iii) Suppose $a$ did lie in the interior of an endblock $E$ of $G_a$. As $d_{G_a}(a)\geq 2$ $E$ is 2-connected, so by induction on $d$ we have an endblock path $P_1$ from $a$ to $\mbox{cutv}(E)$ in $G[E]$ of length at least $2^{d-2}$. Now by (ii) $G_a$ is not 2-connected and so it contains a second endblock $E'$, with an exit vertex $x$. Again by induction on $d$, $G[E']$ contains an endblock path $P_3$ from $\mbox{cutv}(E')$ to $x$ of length $2^{d-2}$. Join $\mbox{cutv}(E)$ to $\mbox{cutv}(E')$ by a path $P_2$ in $G_a$ and $p(x)$ to $b$ by a path $P_4$ in $G_b$. Combining all of these paths we have a path $P_1P_2P_3xp(x)P_4$ from $a$ to $b$ of length at least $2^{d-1}$, a contradiction. (iv) This follows from (ii) and (iii) as if $G_a$ is not 2-connected and $a$ does not lie in the interior of any endblock, $a$ must have at least two limbs. } \end{proof} \begin{figure} \centering \input{List.pstex_t} \caption{Path $P$ constructed in Lemma \ref{list}(ii). Curved paths like $P_1$ and $P_2$ will represent endblock paths of length at least $2^{d-2}$ throughout.} \end{figure} Note that by symmetry of $a$ and $b$, Lemma \ref{list} also applies on replacing $a$ with $b$. The next proposition dispenses with the simplest case we can use endblock paths to build our path of length $2^{d-1}$ from $a$ to $b$. \begin{prop} { \label{int} For an exit vertex $x$ in an endblock $E$ of $G_a$, $p(x)$ can never lie in the interior of an endblock $F$ of $G_b$. } \end{prop} \begin{proof} { From Lemma \ref{list}(iii) $a\notin \mbox{int}(E)$ and $b\notin \mbox{int}(F)$. Pick a path $P_1$ in $G_a$ from $a$ to $\mbox{cutv}(E)$ and a path $P_4$ in $G_b$ from $\mbox{cutv}(F)$ to $b$. Since $E$ is 2-connected and all $v\in E-\lbrace \mbox{cutv}(E),x\rbrace $ have degree at least $d-1$ in $G[E]$, by induction $G[E]$ contains a path $P_2$ of length at least $2^{d-2}$ from $\mbox{cutv}(E)$ to $x$. Similarly $G[F]$ contains a path $P_3$ of length at least $2^{d-2}$ from $p(x)$ to $\mbox{cutv}(F)$. Combining these gives an $a-b$ path $P=P_1P_2xp(x)P_3P_4$ of length at least $2^{d-1}$, a contradiction. } \end{proof} \section{The Interaction Digraph} Let $K_1,\ldots ,K_r$ be the limbs of $a$ and $L_1,\ldots ,L_s$ be the limbs of $b$. Note that by Lemma \ref{list}(iv) $r,s\geq 2$. We form an auxiliary bipartite multidigraph $H=(A,B,\overrightarrow {E})$ which will represent the interaction between the limbs and cores of $a$ and $b$. Let $A=\lbrace {K_1,\ldots ,K_r }\rbrace $ and $B=\lbrace {L_1,\ldots ,L_s }\rbrace $. Additionally, adjoin $\mbox {Core}(a)$ to $A$ and $\mbox {Core}(b)$ to $B$ if they are non-empty. Given an endblock $E$ of $G_a$ there exists an exit vertex $x$ with $x\neq a$ and $p(x)\neq b$ by Lemma \ref{list}(i) and (iii). Pick exactly one such exit vertex $x_E$ for each such endblock $E$ and adjoin a directed edge to $H$ from $K$ to $W\in B$ where $E$ is contained in limb $K$ and $p(x_E)\in W$. Similarly, for each endblock $F$ in $L$ we pick an exit vertex $y_F\in F$ with $p(y_F) \neq a$ and add a directed edge to $H$ from $L$ to $V$ where $p(y_F)\in V$. Note that by Proposition \ref{int} we never choose an exit vertex $x_E$ for some $E$ and $y_F$ for some $F$ such that $p(x_E)=y_F$. Also since any limb of $a$ or $b$ contains an endblock, every limb vertex in $H$ must have outdegree at least one and core vertices have no outneighbours. We shall study the component structure of $H$. The next two lemmas say that this must be very restricted. Together they will allow us to find a connected component $C$ of $H$ consisting entirely of limbs. The inductive step in Section 5 will take place on the subgraph of $G$ corresponding to this $C$. \begin{lem} { \label{paths3} $H$ cannot contain an undirected path of length three. } \end{lem} \begin{proof} { Suppose we have such an undirected path $Q=V_0V_1V_2V_3$ in $H$ and assume $V_0\in A$. Each directed edge $\overrightarrow {VW}$ of $Q$ gives an endblock in $V$ with exit vertex $x$, such that $p(x)\neq b$ and $p(x)\in W$. These endblocks are distinct by the construction of $H$ and in each we can find an endblock path of length $2^{d-2}$ from its cutvertex to this exit vertex by induction on $d$. We claim that we can form an $a-b$ path $P$ which extends all three of these paths. As such a path has length at least $3(2^{d-2})>2^{d-1}$, this contradicts our choice of $G$ and proves the lemma. We will construct our path by forming paths $P_i$ in each $V_i$ and eventually join them into one. The start point of $P_i$ will be denoted by $a_i$ and its end point by $b_i$. We first choose these vertices. If $\overrightarrow {V_iV_{i+1}}$ is an edge of $Q$ there is an endblock $E$ in $V_i$ with an exit vertex $x$ such that $p(x)\in V_{i+1}$. In this case let $b_i=x$ and $a_{i+1}=p(x)$. If $\overleftarrow {V_iV_{i+1}}$ is an edge of $Q$ this gives an endblock $E$ in $V_{i+1}$ with an exit vertex $x$ such that $p(x)\in V_{i}$. In this case let $b_i=p(x)$ and $a_{i+1}=x$. We set \[a_0 = \left\{ \begin{array}{ll} \mbox{Joint$(V_0)$} & \mbox{if $V_0$ is a limb of $a$};\\ b_0 & \mbox{if $V_0=$Core$(a)$}\end{array} \right. \] \[b_3 = \left\{ \begin{array}{ll} \mbox{Joint$(V_3)$} & \mbox{if $V_3$ is a limb of $b$};\\ a_3 & \mbox{if $V_3=$Core$(b)$}.\end{array} \right. \] \noindent Note that $b_i$ and $a_{i+1}$ are adjacent for $i\in \lbrace 0,1,2\rbrace $ and $a,b\notin \{b_0,a_1,b_1,a_2,b_2,a_3\}$. We now build the paths $P_i$ from $a_i$ to $b_i$ in each $V_i$, where $V_i$ is a limb. We claim we can choose $P_i$ so that neither $a$ nor $b$ are interior vertices of $P_i$ (that is, they can lie on $P_i$, but only as end vertices) and $P_i$ has length at least $2^{d-2}$ if $V_i$ has one outneighbour on $Q$ and $2^{d-1}$ if $V_i$ has two. Indeed, if $V_i$ has one outneighbour in $Q$ then one of $a_i$ or $b_i$ must be an exit vertex of an endblock $E$ of $V_i$. Without loss of generality this is $a_i$. Then $b_i\notin \mbox{int}(E)$ by Proposition \ref{int} and $G\lbrack E\rbrack $ contains a path of length $2^{d-2}$ from $a_i$ to the cutv$(E)$. Since $V_i-\lbrace a,b\rbrace $ is connected for all $i$ from the definition of a limb, we can extend this path from cutv$(E)$ to $b_i$ as required. The case where $V_i$ has two outneighbours on $Q$ is identical, using the same argument in two endblocks of $V_i$ and joining their cutvertices in $V_i$. Finally we combine the $P_i$ paths. We first deal with the case where neither Core$(a)$ nor Core$(b)$ occur as interior vertices of $Q$. Combining the paths above we have an $a_0-b_3$ path $P'=P_0b_0a_1P_1b_1a_2P_2b_2a_3P_3$. If $\mbox{Body}(a)=\lbrace a\rbrace $ then $P'$ starts at $a$ so we only need to extend $P'$ to start at $a$ when $\mbox{Body}(a)\neq \lbrace a\rbrace $. In $P'$ as constructed above, $\mbox {Body}(a)\cap P'$ contains $a_0$ and at most one other vertex. Since $\mbox{Body}(a)$ is 2-connected it contains a path $P_1'$ from $a$ to $a_0$ avoiding this vertex. Finding a similar path $P_2'$ from $b_3$ to $b$ in $\mbox {Body}(b)$ if $\mbox {Body}(b)\neq \lbrace b\rbrace $ we may take $P=P_1'P'P_2'$. \begin{figure} \centering \input{Paths3.pstex_t} \caption{An illustration of Lemma \ref{paths3} in the case where $V_2=\mbox{Core}(a)$ and $V_0V_1$, $V_1V_2$ and $V_3V_2$ are directed edges of $Q$. As in the proof of Lemma \ref{paths3}, 2-connectivity can be used in $\mbox{Body}(a)$ to find vertex disjoint paths from $\{a_0,a_2\}$ to $\{b,b_2\}$.} \label{picpaths3} \end{figure} If $Q$ contains one of the Core vertices, without loss of generality let it be Core$(a)$. If Core$(a)$ occurs as an interior vertex of $Q$, it must be $V_2$. Body$(a)$ then contains distinct $a_0,a_2,b_2$ and we have two paths $P_1'=P_0b_0a_1P_1b_1a_2$ from $a_0$ to $a_2$ and $P_2'=b_2a_3P_3$ from $b_2$ to $b_3$ as in Figure \ref{picpaths3}. From the choice of the $a_2$ and $b_2$ above and the fact that $a$ is not a cutvertex we have $a\notin \lbrace a_0,a_2,b_2\rbrace $. Then by 2-connectivity Body$(a)$ contains two vertex disjoint paths from $\lbrace a_0,a_2\rbrace $ to $\lbrace a,b_2\rbrace $. Piecing these paths together with $P_1'$ and $P_2'$ we obtain an $ab_3$-path $P'$. If Body$(b)=\lbrace b\rbrace $ we are done since $b=b_3$. Otherwise we extend $P'$ using 2-connectivity as above to find an $a-b$ path of length at least $2^{d-1}$, a contradiction to the choice of $G$. } \end{proof} Note that Lemma \ref{paths3} guarantees that $H$ has at least two connected components. The next lemma further limits $H$. Its proof is very similar to that of Lemma \ref{paths3}. \begin{lem} { \label{body} Suppose that we have $\mbox{Body}(a)\neq \lbrace a\rbrace $. Then no component of $H$ contains two vertices of $A$. } \end{lem} \begin{proof} { Suppose $H$ has such a component $C$. Then by Lemma \ref{paths3}, $C$ consists of vertices $V_1,\ldots ,V_t$ in $A$ and vertex $W$ in $B$. At most one of $V_1,\ldots ,V_t, W$ can be a core vertex as there is no edge between Core$(a)$ and Core$(b)$ in $H$. If $W=\mbox{Core}(b)$ then $V_1$ and $V_2$ must be limbs and these guarantee two vertex disjoint paths $P_1$, $P_2$ from vertices $a_1,a_2\in \mbox{Body}(a)$ to vertices $b_1,b_2\in \mbox{Body}(b)$ both of length at least $2^{d-2}$. By Lemma \ref{paths3} and Lemma \ref{list}(iv) $H$ must contain a second component $C'$ containing a limb of $b$ which guarantees the existence of a third path $P_3$ from a vertex $a_3\in \mbox{Body}(a)$ to $b_3\in \mbox{Body}(b)$ of length $2^{d-2}$. Using identical 2-connectivity arguments in both $\mbox{Body}(a)$ and $\mbox{Body}(b)$ as in Lemma \ref{paths3} we can combine these three paths into one from $a$ to $b$, a contradiction. If $W\neq \mbox{Core}(b)$ then $C$ guarantees a path $P_1$ of length $2^{d-1}$ between two vertices $a_1$ and $a_2$ in $\mbox{Body}(a)$ with $b\notin P_1\cap \mbox{Body}(b)$ and $|P_1\cap \mbox{Body}(b)|\leq 1$. Again from a second connected component of $H$ we obtain a disjoint path $P_2$ from an element $a_3\in \mbox{Body}(a)$ to $b_1\in \mbox{Body}(b)$. Once more, with an application of 2-connectivity in $\mbox{Body}(a)$ and a possible application in $\mbox{Body}(b)$ we find an $a-b$ path extending both $P_1$ and $P_2$, a contradiction. } \end{proof} Again the same applies switching $a$ with $b$. As mentioned before Lemma \ref{paths3} the previous two lemmas imply that $H$ contains a connected component $C$ consisting entirely of limbs. If not, each component of $H$ would contain one of Core$(a$) or Core$(b)$ and thus $H$ would contain at most two connected components. Since $H$ contains no path of length 3 by Lemma \ref{paths3}, it must have exactly two components, one containing Core$(a)$ and the other containing Core$(b)$. But as $A$ contains Core$(a)$ and at least two limbs, two of these must lie in the same connected component contradicting Lemma \ref{body}. \section {The Inductive Step} Let ${C}$ be the component of $H$ consisting entirely of limbs of $a$ and $b$ guaranteed from Section 4 and write $G_C$ for the subgraph $G[\bigcup _{W\in C}V(W)]$ of $G$. Notice that $G_{C}$ must contain exactly one vertex $a_C$ in $\mbox{Body}(a)$ and one vertex $b_C$ in $\mbox{Body}(b)$ - if Body$(a)= \lbrace a\rbrace$ then $a_C=a$, if not ${C}\cap A=\lbrace V\rbrace $ by Lemma \ref{body} and $a_C=\mbox{Joint}(V)$. Our final lemma before we complete the proof of Theorem 2.3$'$ allows us to find a subgraph of $G_C$ which will either also satisfy the conditions of Theorem \ref{connmod} or build at least half of the path we are looking for from any edge entering it. Before stating it we give one last definition. \begin{dfn} { \emph{Given a graph $G$ and $S\subset V(G)$ define the span$(S)$ in $G$ to be the subset of $V(G)$ consisting of all vertices which lie on a path between two elements of $S$.} } \end{dfn} Note that we include paths of length zero in this definition, so that $S\subset $ span($S$). \begin{lem} { \label{nice} Let $C$ and $G_C$ be as above. Then $G_C$ has a 2-connected subgraph $J$ containing two vertices $a'\in G_a$ and $b'\in G_b$ with the following properties: \begin{itemize} { \item[\emph{(i)}] every vertex $v\in J - \lbrace a',b'\rbrace$ has degree at least $d-1$ in $J$ and all the neighbours of $v$ in $G_C$ are contained in $J$. \item[\emph{(ii)}] for any vertex $v\in J-\lbrace a',b'\rbrace $, $J$ contains an $a'-v$ path not containing $b$ and a $b'-v$ path not containing $a$, both of length at least $2^{d-2}$. } \end{itemize} } \end{lem} \begin{proof} { Suppose first that $C$ consists of two vertices with a limb $K$ in $A\cap C$ and $L$ in $B\cap C$. Let $S$ be the subset of vertices of $K-a$ with a neighbour in $L-b$ and $T$ the subset of $L-b$ with a neighbour in $K-a$. Note that by definition of $C$ each endblock in $K$ has an exit vertex $x$ with $p(x)\in L-a$ and by Proposition \ref{int} $p(x)$ cannot lie in the interior of an endblock of $L$. As the same holds true for the endblocks in $L$, both $|S|,|T|\geq 2$. Consider the graph $G' = G[\mbox{span}(S)\cup \mbox{span}(T)]$. This is 2-connected and contains all vertices of the endblocks of $K$ in $G_a$ and $L$ in $G_b$. Furthermore, $G'$ restricted to $G_a$ is a union of blocks of $G_a$. Such a subgraph must be joined to the rest of the $G_a$ by a single cutvertex $a_0'$. The same holds true for $G'$ restricted to $G_b$ with cutvertex $b_0'$. While this is almost the graph we will take as $J$, $a$ and $b$ may be vertices of $K$ and $L$ respectively, not lie in $G'$ but still have a neighbour in $G'-\lbrace a_0',b_0'\rbrace $ which would not allow (i) to hold. If this occurs with $a$ say, then adjoin it to $S$ and run the same construction as above with this new $S$ to obtain a new graph $G''$. Let the $a_1'$ be the cutvertex with the rest of $G_a$ and $b_1'$ the cutvertex with the rest of $G_b$. The same problem may now occur with with $G''$ and $b$. If so, adjoining $b$ to $T$ we obtain our final graph $G'''$, which again is 2-connected with cutvertex $a_2'$ with the rest of $G_a$ and cutvertex $b_2'$ with the rest of $G_b$. It should be clear that (i) now certainly holds for this graph and we may take $J=G'''$, $a'=a_2'$ and $b'=b_2'$. To see that $J$ contains an $a'-v$ path as claimed in (ii) we first consider the case with $b'\neq b$. Now $J$ contains an endblock $E$ from $G_a$ and an endblock $F$ from $G_b$ and as $v$ is in at most one of these, assume $v\notin F$. Taking $\widetilde {J}$ to be the graph formed from $J$ by contracting int$(F)$ to a single point $f$ it is easily seen $\widetilde {J}$ is still 2-connected. As neither $a'$ nor $b'$ can lie in the interior of $F$, by 2-connectivity $\widetilde{J}$ contains two vertex disjoint paths from the set $\lbrace a',v\rbrace $ to $\lbrace {\mbox{cutv}(F), f} \rbrace $. These paths give two paths in $J$, $P_1$ from $a'$ to say $w\in \mbox{int}(F)$ and $P_3$ from $v$ to $\mbox{cutv}(F)$. By induction on Theorem \ref{connmod} $G[F]$ contains a path $P_2$ of length $2^{d-2}$ from $\mbox{cutv}(F)$ to $w$. $P=P_1P_2P_3$ now works for the $a'-v$ path claimed in (ii). If $b'=b$, using the same argument as above we might use $b$ in one of the paths $P_1$ or $P_3$. However in this case we have the following: \begin{quotation} { \noindent $J-b$ has a 2-connected subgraph $J'$ containing all of $J\cap G_a$ and an endblock of $G_b$. } \end{quotation} To see this, look at the block-cutvertex decomposition of $J-b$. This contains exactly one block $J'$ with vertices in both $G_a$ and $G_b$ and this block contains all of $J\cap G_a$ by construction. Since any endblock of $G_b$ contained in $J$ has an exit vertex whose partner lies in $J\cap G_a$, all such endblocks lie in $J'$. Given $v$ as above, choose a minimal path $P_0$ not containing $b$ from $v$ to some $v'\in J'$. We may now run the same argument as above in $J'$ to obtain a $a'-v'$ path $P'$ of length at least $2^{d-2}$. Combining the resulting path with $P_0$ we are done. By symmetry this finishes the case where $C$ consists of two vertices. We now deal with the case where $C$ consists of more than two vertices. By Lemma \ref{paths3} and Lemma \ref{body} we may assume $C$ consists of a limb $K$ of $a$, limbs $L_1,\ldots L_t$ of $b$ and that Body$(b)=\lbrace b\rbrace $. Let $L_1,\ldots ,L_{t'}$, $t'\leq t$, be the limbs containing at least two vertices other than $b$ with neighbours in $K-a$. Note that $t'\geq 1$ since some $L_i$ must be an outneighbour of $K$ in $H$. We will first work with these limbs and add in the rest if needed later. As above, for each $i\in \lbrack 1,t'\rbrack $ let $S_i\subset V(K)$ consist of those vertices in $K-a$ with a neighbour in $L_i-b$ and $T_i\subset V(L_i)$ consist of the vertices of $L_i-b$ with a neighbour in $K-a$. Taking $G_i=G\lbrack \mbox{span}(S_i)\cup \mbox{span}(T_i)\rbrack $ this graph is 2-connected for all $i\in \lbrack 1,t'\rbrack $. Note that $G_i\cap G_b$ must again have exactly one cutvertex with the rest of $G_b$, say $c_i$. Once more there are concerns that $a\in K$ or $b\in L_i$ are not contained in $G_i$ but have neighbours in it as this would violate (i) above. We will again adjoin them to $S_i$ and $T_i$ as needed. While the condition to adjoin $a$ to $S_i$ is exactly the same as above, that is adjoin $a$ to $S_i$ if $a$ has a neighbour in $G_i\cap G_b-c_i$, the condition to adjoin $b$ to $T_i$ needs a slight variation since $G_i\cap G_a$ may have more than one cutvertex separating it from the rest of $G_a$. With a little foresight, take $c_i'\in G_i\cap G_a$ to be the unique cutvertex of $G_a$ separating $G_i\cap G_a$ from $a$ (possibly $a$ itself) and adjoin $b$ to $T_i$ if $b$ has a neighbour in $G_i\cap G_a-c_i'$. For each $i\in[ 1,t'] $, having added the vertices to $S_i$ and $T_i$ as demanded by these conditions, let $S_i$ and $T_i$ now refer to the new sets and $G_i$ to the new graphs formed from them. For all $M\subset \lbrack 1,t\rbrack $ define $S_M=\bigcup _{j\in M}S_j$. Beginning with an initial list $\lbrace S_1,\ldots ,S_{t'}\rbrace $ repeatedly remove $S_M$ and $S_{N}$ from the list and adjoin $S_{M\cup {N}}$ if $|\mbox {span}(S_M)\cap \mbox {span}(S_{N})|\geq 2$. This procedure eventually terminates, and in the final list there must exist some $S_M$ for which $G[\mbox{span}(S_M)]$ has a single cutvertex separating it from the rest of $G_a$. To see this note that for each $I$, $G[\mbox{span}(S_I)]$ is a connected union of blocks of $G_a$ and so corresponds to a subtree of the block graph $\mathcal{B}(G_a)$. Since each leaf of $\mathcal{B}(G_a)$ is contained in some $G[\mbox{span}(S_I)]$ and no two $G[\mbox{span}(S_I)]$ and $G[\mbox{span}(S_{I'})]$ share a block, one of these, $G[\mbox{span}(S_M)]$ say, must have a single cutvertex $a'$ with the rest of $G_a$. Now we deal with the limbs $L_j$ where $j\in \lbrack t'+1,t\rbrack $. Form a set $N\subset \lbrack t'+1,t\rbrack $ with $i\in N$ if some $v\in L_i-b$ has a neighbour in $\mbox{span}(S_{M})-a'$ and let $I=M\cup N$. If $I=\lbrace i\rbrace$ for some $i$ then $N=\emptyset $ and we can run the same argument as in the case where $C$ consists of two vertices replacing $S$ with $S_i$ and $T$ with $T_i$. If not, take $J=G[\bigcup _{i\in I} V(L_i) \cup \mbox{span}(S_I)]$. $J$ is 2-connected and taking $b'=b$, $J$ satisfies (i). Finally we show that (ii) holds for $J$. Given $v\in J-\lbrace a',b'\rbrace $ suppose we are looking for an $a'-v$ path avoiding $b$ of length at least $2^{d-2}$. We again claim: \begin{quotation} { \noindent $J-b$ has a 2-connected subgraph containing all of $J\cap G_a$ and an endblock $F$ of $G_b$. } \end{quotation} It suffices to show that $G[\bigcup _{i\in M}V(L_i)\cup \mbox{span}(S_{M})]$ contains such a subgraph. Exactly as in the case where $C$ consists of two limbs, each $G[\mbox{span}(S_i)\cup \mbox{span}(T_i)]-b$, $i \in [1,t']$, has a 2-connected subgraph $G_i'$ containing all of $\mbox{span}(S_i)$. Moreover one of the subgraphs $G[\mbox{span}(T_i)]-b$ must contain an entire endblock $F$ (since $t'\geq 1$). Now the subgraph $G[\bigcup _{i\in M}V(G_{i}')]$ of $J-b$ is 2-connected by construction of $M$, contains all vertices of $J\cap G_a$ and endblock $F$ in $G_b$, proving the claim. Using this subgraph, exactly as in the case where $C$ consists of two limbs, we obtain an $a'-v$ path of desired length. An identical argument gives the $b'-v$ path as claimed in (ii). } \end{proof} \begin{figure} \centering \input{together.pstex_t} \caption{An illustration of the case $w\in \mbox{int}(E)$ in the proof of Theorem 2.3$'$. The broken dotted pieces represent vertices in $G_C$ that are left out of $J$.} \label{together} \end{figure} Now we are ready to finish the proof of Theorem 2.3$'$. Let $J$ be the subgraph of $G$ guaranteed from Lemma \ref{nice}. If there is no edge between $J-\lbrace a',b'\rbrace $ and $G-J$, all $v\in J-\lbrace a',b'\rbrace $ have degree at least $d$ in $J$ and by choice of $G$, $J$ contains an $a'-b'$ path $P$ of length at least $2^{d-1}$. Extending this path from $a'$ to $a$ and $b'$ to $b$, $G$ contains an $a-b$ path of length $2^{d-1}$, a contradiction. So such an edge must exist, joining say $v\in J-\lbrace {a',b'}\rbrace $ to $w\notin G_C$ by Lemma \ref{nice}(i). We may assume $w\in G_a$. First suppose $w\in \mbox{Body}(a)$ -- this is only possible if $\mbox{Body}(a)\neq \{a\}$, by Lemma \ref{nice}(i). Applying Lemma \ref{nice}(ii) we have an $a'-v$ path in $J$ not containing $b$ of length at least $2^{d-2}$. This path extends in $C$ to an $a_C-v$ path $P_1$, where again $a_C=G_C\cap \mbox{Body}(a)$. Now taking a limb $K$ of $a$ not in $C$, $K$ guarantees a path $P_2$ from Joint$(K)$ to $b$ of length at least $2^{d-2}$ disjoint from $P_1$. Thus we have a path $P_1vw$ joining two vertices $a_C$ and $w$ in Body$(a)$ and a path $P_2$ from Joint$(K)$ to $b$. By 2-connectivity, Body$(a)$ contains two vertex disjoint paths $P_3$ and $P_4$ from $\lbrace a, \mbox{Joint}(K)\rbrace $ to $\lbrace a_C,w\rbrace $. Combining all four of these paths we obtain an $a-b$ path $P$ extending both $P_1$ and $P_2$. But this path has length at least $2^{d-1}$, contradicting our choice of $G$. So we may assume $w\notin $ Body$(a)$. Then $w\in K$ for some limb $K$ of $a$, where by Lemma \ref{nice}(i) $K\notin C$. Let $E$ be an endblock of $K$ and take $x_E$ to be the exit vertex of $E$ chosen in the construction of $H$ in Section 4. Note that $p(x_E)\notin G_C$, since $K\notin C$. Depending on whether or not $w$ lies in int$(E)$ we can construct a path $P$ from $w$ to either Joint$(K)$ or to $x$ of length $2^{d-2}$ by induction on $d$ in Theorem \ref{connmod}. If $P$ is from $w$ to Joint$(K)$ adjoin it via the edge $vw$ to the $v-b'$ path in $J$ guaranteed by Lemma \ref{nice}(ii) (see Figure \ref{together}). This gives a Joint$(K)-b'$ path $P'$ of length at least $2^{d-1}$ which extends to an $a-b$ path. If $P$ is from $w$ to $x_E$ then we combine this path with the $v-a'$ path in $J$ guaranteed by Lemma \ref{nice}(ii). Again this gives a $a'-p(x_E)$ path of length at least $2^{d-1}$ which extends to an $a-b$ path. This proves Theorem 2.3$'$. \hspace{5cm} $\square$ \section{Removing the Degree Assumption} Recall in Lemma \ref{splitting}, we obtained a splitting of $G$ into two pieces, one containing $a$ and the other containing $b$ by choosing a direction $i$ on which $a$ and $b$ differ when viewed as elements of $\mathcal{P}[n]$. In general it is not possible to choose such a direction to ensure that both $d_{G_a}(a)\geq 2$ and $d_{G_b}(b)\geq 2$ -- for example, $a$ and $b$ could be adjacent with both having only one other neighbour in $G$. It is however always possible to choose such a direction $i$ if $d_G(a)\geq 3$ and $d_G(b)\geq 3$. We will therefore assume that $d_G(a)=2$. The condition $d_{G_a}(a)\geq 2$ and $d_{G_b}(a)\geq 2$ as used above allowed us to ensure that all endblocks of $G_a$ and $G_b$ contain long paths, which is clearly false if $a$ has a single neighbour $a'$ in $G_a$. This in turn guaranteed that $a$ had at least two limbs in $G_a$ which was crucially used numerous times in our analysis of $H$ e.g. Lemma \ref{body}. In this section, we extend the arguments of Theorem 2.3$'$ to prove Theorem \ref{connmod}. \begin{lem} \label{remove} The following hold: \begin{itemize} { \item[\emph{(i)}] $G-a$ is a 2-connected graph. \item[\emph{(ii)}] We can choose a splitting direction $i$ such that after forming $G_a$ and $G_b$ as above from this $i$, $d_{G_a}(a)\geq 2$ or $d_{G_b}(b)\geq 2$ } \end{itemize} \end{lem} \begin{proof} { (i) If $G-a$ is not 2-connected it has at least two endblocks in its block-cutvertex decomposition, one of which $E$ has $b\notin \mbox{int}(E)$. Now since $G$ is 2-connected, $a$ must be joined to the interior of all the endblocks of $G-a$. As $d_G(a)=2$, $G-a$ has exactly two endblocks with $a$ having exactly one neighbour in the interior of each. Let $w$ be this neighbour in $E$. Now $E$ is 2-connected (as $d\geq 3$) and all vertices in $G[E]-\lbrace \mbox{cutv}(E), w\rbrace $ have degree at least $d$ in $G[E]$. Since $G$ is the smallest counterexample and $G[E]$ is a non-spanning subgraph of $G$, it contains a path $P$ of length $2^{d-1}$ from $w$ to $\mbox{cutv}(E)$. Extending this path on either side to $a$ and $b$ respectively, we have an $a-b$ path of length at least $2^{d-1}$, a contradiction. (ii) We can always choose such a direction if $a$ and $b$ are at Hamming distance at least three in $Q_n$ or if one of $a$ or $b$ have degree greater than 2 in $G$. So $a$ and $b$ must be at Hamming distance one or two in $G$ and both have degree exactly two. First consider $a$ and $b$ at Hamming distance one. If they are not adjacent in $G$ we may choose the direction on which they differ for $i$ so we can assume they are adjacent. Then $a$ and $b$ both have one other neighbour in $G$, $a'$ and $b'$ respectively. Now if $G-\{a,b\}$ is 2-connected we can apply Theorem \ref{connmod} to $G-\{a,b\}$ with $a'$ and $b'$ replacing $a$ and $b$ to obtain an $a'-b'$ path of length at least $2^{d-1}$. Adjoining the edges $aa'$ and $bb'$ to this path we have an $a-b$ path of length $2^{d-1}+2$, more than enough. If $G-\{a,b\}$ is not 2-connected it is easily seen that $a'$ and $b'$ must lie in the interior of different endblocks of $G-\{a,b\}$. We can therefore find a path from $a'$ to $b'$ in $G-\{a,b\}$ which extends two endblock paths. Adjoining the edges $aa'$ and $bb'$ to this path again we have an $a-b$ path of length at least $2^{d}+2$. If $a$ and $b$ are at Hamming distance two in the cube, we can always find such a direction $i$ unless $a$ and $b$ are joined to the same two neighbours in $G$, $a'$ and $b'$ say. Then $\lbrace a,a',b,b'\rbrace $ form a $C_4$ with $a$ opposite $b$. Working with $G-\lbrace a,b\rbrace $, $a'$ and $b'$ as above, we again obtain an $a-b$ path of desired length in $G$. } \end{proof} From Lemma \ref{remove}(ii) we can now assume that we have chosen a partition direction $i$ such that $\mbox{deg}_{G_b}(b)\geq 2$. Lemma \ref{list}(i)-(iv) still hold for $G_b$ with the same proofs as above. In particular $b$ has at least two limbs. In order for our main argument in Theorem 2.3$'$ to be inapplicable, $a$ must have exactly one neighbour $a'$ in $G_a$ and one neighbour $v$ in $G_b$. We may also assume that $v\neq b$ as otherwise by Lemma \ref{remove}(i) we could apply Theorem \ref{connmod} to $G-a$ taking $a'$ in place of $a$. \begin{lem} { \label{pick} We have $v\in \mbox{int}(E_v)$ for some endblock $E_v$ of $G_b$. } \end{lem} \begin{proof} { Suppose not. From Lemma \ref{list}(iii) $b$ does not lie in the interior of an endblock of $G_b$ and by Lemma \ref{list}(iv) $G_b$ contains two vertex disjoint paths $P_1$ from $v$ to $\mbox{cutv}(E_1)$ and $P_5$ from $\mbox{cutv}(E_2)$ to $b$, where $E_1$ and $E_2$ are two endblocks of $G_b$. Taking exit vertices $x_1$ and $x_2$ of $E_1$ and $E_2$ respectively, by induction on $d$, $G[E_1]$ contains a path $P_2$ of length at least $2^{d-2}$ from $\mbox{cutv}(E_1)$ to $x_1$ and $G[E_2]$ contains a path $P_4$ of length at least $2^{d-2}$ from $x_2$ to $\mbox{cutv}(E_2)$. Taking a path $P_3$ from $p(x_1)$ to $p(x_2)$ in $G_a-a$ and combining the paths, $G$ contains an $a-b$ path $avP_1P_2x_1p(x_1)P_3p(x_2)x_2P_4P_5$ of length at least $2^{d-1}$, a contradiction. } \end{proof} We again construct an interaction digraph $H$ but this time it is built from the limbs of $a'$ and $b$ instead of those of $a$ and $b$. Note that $\{a,a'\}$ is a limb of $a'$ and so both $a'$ and $b$ have at least two limbs. Take $H=(A',B,\overrightarrow{E})$ to be a bipartite multidigraph on vertex sets $A'=\{K_1,\ldots ,K_r\}$ and $B=\{L_1,\ldots L_s\}$, the set of limbs of $a'$ and $b$ respectively. We also adjoin $\mbox{Core}(b)$ to $B$ if it is non-empty ($\mbox{Core}(a')=\emptyset$ since $a'$ is a cutvertex of $G_a$). Now each endblock of $G_a$ or other than $\{a,a'\}$ contains at least two exit vertices, as in Lemma \ref{list}(i). Therefore for each endblock $E$ of $G_a$ or $G_b$ other than $\{a,a'\}$ we can pick an exit vertex $x_E$ with $p(x_E)\neq a',b$. From Lemma \ref{pick} we can pick $x_{E_v}=v$. Now adjoin a directed edge from $K\in A'$ to $L\in B$ for each endblock $E$ in $K$ with $p(x_E)\in L$ and a directed edge from $L\in B$ to $K\in A'$ for each endblock $E$ in $L$ with $p(x_E)\in K$. Note that every limb other than $\{a,a'\}$ still has an outneighbour in $H$. For this $H$ Lemma \ref{paths3} and Lemma \ref{body} still hold with the same proofs as before. These two ensure that $H$ still contains a connected component $C$ consisting entirely of limbs and not containing the limb $\{a,a'\}$. Indeed, since $b$ has at least two limbs, pick a limb $L\in B$ not containing $E_v$ and let $C$ be the connected component of $H$ with $L\in C$. As $v$ is the unique neighbour of $a$ in $B$ and $v\notin L$, if $\{a,a'\}\in C$ then $H$ would contain a path of length three, contradicting Lemma \ref{paths3}. Furthermore, since $\mbox{Core}(a')\notin A'$, if $C$ did not consist entirely of limbs of $a'$ and $b$, $\mbox{Body}(b)\neq \{b\}$ and $C$ contains two vertices of $B$, contradicting Lemma \ref{body}. Following the argument as in Section 5, an application of Lemma \ref{nice} to $G_C$ finishes the proof of Theorem \ref{connmod}. \hspace{5cm} $\square$ \section {A Tight Bound} Theorem \ref{conn} is proved in the exact same manner as Theorem \ref{connmod} but requires a more care and attention to detail in various arguments. We are now ready for its proof. \vspace{0.3cm} \noindent \emph{Proof of Theorem \ref{conn}.} The proof is again by induction on $d$. The base case $d=2$ is immediate unless $a$ and $b$ are at Hamming distance 2 apart. If this is the case and $G$ is not isomorphic to $Q_2$ pick any vertex $v$ of $G$ not in the unique 2-cube containing $a$ and $b$. By 2-connectivity $G$ contains vertex disjoint $a-v$ and $v-b$ paths, which when combined give a path of length at least 3, as required. Suppose for contradiction that Theorem \ref{conn} fails for some minimal $d$ and that $G$ is the smallest counterexample. We now explain how the proofs above in Sections 3-6 can be altered to give this stronger result. To begin we again assume there is some splitting direction $i$ of $G$ such that forming $G_a$ and $G_b$ as before, $d_{G_a}(a)\geq 2$ and $d_{G_b}(b)\geq 2$. The first lemma of this section says that Lemma \ref{list} still applies to $G_a$ and $G_b$. \begin{lem} { \label{listextension} \item[\emph{(i)}] Every endblock of $G_a$ which does not contain $a$ in its interior must contain at least two exit vertices. \item[\emph{(ii)}] $G_a$ is not 2-connected. \item[\emph{(iii)}] $a$ does not lie in the interior of an endblock in $G_a$. \item[\emph{(iv)}] $a$ must have at least two limbs. } \end{lem} \begin{proof} { (i) The proof in this case needs only a slight variation, as if we apply our strengthened induction hypothesis to $E$ as Lemma \ref{list}(i), we are guaranteed to find an $a-x$ path of length at least $2^d-2$ and continuing this path as before we adjoin at least one more vertex $x$ with edge $xp(x)$ giving a bound of at least $2^d-1$, as required. (ii) The modification for the proof of this case is a little more demanding. Suppose for contradiction that $G_a$ is 2-connected. First suppose $G_b$ is not 2-connected. If there exists an endblock $E$ in $G_b$ such that $b\notin E$, we have a path in $G_b$ of length at least 1 from $b$ to cutv($E$). We claim $E$ contains a cutv($E$)$-x$ path of length at least $2^{d-1}-1$ where $x$ is some exit vertex of $E$ with $p(x)\neq a$. If $E$ is isomorphic to the ($d-1$)-dimensional cube $Q_{d-1}$ every interior vertex of $E$ must be an exit vertex of $E$. Therefore we can take any $x\in \mbox{int}(E)$ with $p(x)\neq a$ at odd distance from $\mbox{cutv}(E)$. If $E$ is not isomorphic to the ($d-1$)-dimensional cube $Q_{d-1}$, by (i) $E$ must contain an exit vertex $x$ with $p(x)\neq a$ and by induction on Theorem \ref{conn} E contains a path of length $2^{d-1}-1$ from $\mbox{cutv}(E)$ to $x$. Combining the appropriate one of these two with the $a-p(x)$ path of length $2^{d-1}-2$ in $G_a$ guaranteed by induction on Theorem \ref{conn}, $G$ contains an $a-b$ path of length at least $1+(2^{d-1}-1)+1+(2^{d-1}-2)=2^d-1$ a contradiction. So $b$ must lie in \emph{every} endblock $E_1,\ldots ,E_t$ of $G_b$. Note that since $t\geq 2$ this implies $b\notin \mbox{int}(E_i)$ for any $i$. As with $E$ above, $E_1$ must have an exit vertex $x$ such that $E_1$ contains a $x-b$ path of length at least $2^{d-1}-1$, with $p(x)\neq a$. If $G_a$ were not isomorphic to $Q_{d-1}$, it contains a path of length $2^{d-1}-1$ from $a$ to $p(x)$. Combining these two with the edge $xp(x)$ we obtain an $a-b$ path of length $2^{d}-1$. Therefore we can assume $G_a$ is isomorphic to $Q_{d-1}$. For $t\geq 2$ none of the $E_1,\ldots ,E_t$ can be isomorphic to $Q_{d-1}$ as $G_a$ would have to receive too many edges. Now since $G_a$ is isomorphic to the cube, some endblock $E_i$ must contain an exit vertex $x$ with $p(x)\neq a$ such that $x$ is at even Hamming distance from $b$. But then since $E_i$ is not isomorphic to $Q_{d-1}$, by induction $E_i$ contains an $x-b$ path of length at least $2^{d-1}$. By induction on Theorem \ref{conn} applied to $G_a$, it contains an $a-p(x)$ path of length at least $2^{d-1}-2$. Combining these paths via the edge $xp(x)$ gives a desired path from $a$ to $b$ of length $2^d-1$. The case when $G_b$ is 2-connected is very similar. We can obtain two paths of length at least $2^{d-1}-1$ in $G_a$ and $G_b=E_1$ if neither of the two are isomorphic to $Q_{d-1}$ and if one is isomorphic to $Q_{d-1}$ we use the same argument as in the case where $G_a$ is isomorphic to $Q_{d-1}$ and $t\geq 2$ above. (iii) This is similar to (ii) but a little easier. Again, taking the endblocks $E$ and $E'$ as in the proof of Lemma \ref{list}(iii), we can find a $a-\mbox{cutv}(E)$ path $P_1$ in $E$ of length at least $2^{d-1}-2$ and a $\mbox{cutv}(B)-x$ path $P_3$ in $E'$ of length at least $2^{d-1}-1$ where $x$ is an exit vertex of $E'$. Now if $p(x)=a$ then combining these we may only obtain an $a-b$ path of length $2^d-2$. Instead, using (i) we can choose the exit vertex $x$ so that $p(x)\neq a$. This gives the extra edge on the path required. (iv) Again follows from (ii) and (iii). } \end{proof} \vspace{0.1cm} The above modifications demonstrate the main problem moving from the bounds in Theorem \ref{connmod} to Theorem \ref{conn} -- on combining endblock paths together without any care as before, we are usually left short by one vertex. To get around this there are two small tricks, both of which were demonstrated above: we try to ensure that we form a path from $a$ to $b$ with at least two endblock paths, so that one of these paths has length at least $2^{d-1}-1$ and squeeze in another vertex on the way as in Lemma \ref{listextension}(iii) above or using a parity argument as in the case where $G_a$ is isomorphic to $Q_{d-1}$ we may be able to show that a given endblock $E$ which we know is not isomorphic to $Q_{d-1}$ contains an exit vertex at even distance away from its cutvertex. This second approach was used in Lemma \ref{listextension}(ii) and allows us to find an endblock path of length $2^{d-1}$ in $E$ which when combined with any other endblock path would give the required bound. As neither of these options can be guaranteed given the statement of Proposition \ref{int} we should not expect to be able to prove it. Now Proposition \ref{int} was important in defining our interaction graph $H$ in Section 4 and 5. Can we find a way to define $H$ not depending on this? We look towards an slightly altered construction for $H$. Note that unlike our first construction of $H$ in Section 4, since we cannot in general appeal to Proposition \ref{int}, we now have the possibility that an exit vertex of an endblock in $G_a$ is joined to an exit vertex of an endblock in $G_b$. It is necessary to pick these exit vertices $x_E$ such that $x_F\neq p(x_E)$ for every two endblocks $E$ and $F$ in $G_a$ and $G_b$ respectively in order to ensure that we can still find 2-connected subgraphs of $G$ in connected components of $H$ as in Lemma \ref{nice}. Can we always find such $x_E$ and $x_F$? The answer is that we can. For each endblock $E$ of $a$ or $b$, as was seen in Lemma \ref{listextension}(ii), there exists at least two exit vertices $x$ of $E$ such that $E$ contains a path of length at least $2^{d-1}-1$ from $\mbox{cutv}(E)$ to $x$. We choose one of these for $x_E$ with the condition $p(x_E)\neq a,b$, which is always possible. What prevents us from choosing such an $x_E$ and $x_F$ with $x_F=p(x_E)$? Because joining the endblock paths from $\mbox{cutv}(E)$ to $x_E$ and the path from $x_F$ to $\mbox {cutv}(F)$ in $E$ and $F$ respectively, with the edge $x_Ex_F$ we would obtain a path of length at least $(2^{d-1}-1)+1+(2^{d-1}-1)=2^d-1$ which could be extended to an $a-b$ path. Therefore we can choose our $x_E$ for all endblocks $E$ of $G_a$ and $G_b$ such that $x_E\neq p(x_F)$. We now take our interaction graph $H= \{A,B,\overrightarrow E\} $ to be a bipartite multidigraph whose bipartition consists of the limbs of $a$ and $b$ respectively. Again we additionally adjoin Core$(a)$ and Core$(b)$ to $A$ and $B$ respectively if they are non-empty. Add a directed edge to $H$ for every endblock $E$ in $G_a$ from $K\in A$ to $L\in B$ if $E$ is an endblock of limb $K$ with $x_E\in K$ and $p(x_E)\in L$. Similarly add a directed edge to $H$ for every endblock $F$ in $G_b$ from $L\in B$ to $K\in A$ if $F$ is an endblock of limb $L$ with $x_F\in L$ and $p(x_F)\in K$. Note again that every limb in $H$ has outdegree at least 1 as it contains an endblock. We can actually say a lot more about $H$ in the case where one of $a$ or $b$ is not a cutvertex in $G_a$ or $G_b$ - we can guarantee that $p(x_E)\notin \mbox{int}(F)$ for all endblocks $E$ and $F$ of $G_a$ and $G_b$. Indeed, suppose we have $p(x_E)=y\in \mbox{int}(F)$ for some endblocks $E$ of $G_a$ and $F$ of $G_b$ say in limbs $K$ and $L$ respectively. Then by choice of $x_E$ and induction on Theorem \ref{conn} $E$ contains a $\mbox{cutv}(E)-x_E$ path of length at least $2^{d-1}-1$ and $F$ contains a $y-\mbox{cutv}(F)$ path of length $2^{d-1}-2$. This gives a path of length $2^d-2$ from $\mbox{cutv}(E)$ to $\mbox{cutv}(F)$, which extends to a path from $\mbox{Joint}(K)$ to $\mbox{Joint}(L)$. We must have that either $\mbox{Joint}(K)\neq a$ or $\mbox{Joint}(L)\neq b$ since both are not cutvertices. Therefore, extending this path we obtain a path of length at least $2^d-1$ from $a$ to $b$, a contradiction. The property that $p(x_E)\notin \mbox{int}(F)$ for all endblocks $E$ and $F$ of $G_a$ and $G_b$ was used crucially throughout the proof of Theorem \ref{connmod}, and knowing it in the case where one of $a$ or $b$ is not a cutvertex actually allows the proofs of Lemma \ref{paths3} and \ref{body} to proceed identically in these cases. The focus therefore will be on obtaining Lemma \ref{paths3} in the case where both $a$ and $b$ are cutvertices of $G_a$ and $G_b$ respectively -- we need not worry about Lemma \ref{body} as it does not apply to this case. While Lemma \ref{paths3} again holds true, a little more work is required than in the original proof in Section 4. \begin{lem} { \label{path3extension} $H$ cannot contain an undirected path of length three. } \end{lem} \begin{proof} { From the above discussion we need only consider the case in which $a$ and $b$ are cutvertices of $G_a$ and $G_b$ respectively. If one of the interior vertices on this path has two out-neighbours in $Q$ the same argument as in the original proof will create a path between two exit vertices in this limb of length at least $2^d-2$. Extending this path through $Q$ as in Theorem \ref{paths3} gives us an $a-b$ path of length at least $2^d-1$. Similarly, if both endvertices on this path have out-neighbours in $Q$ (that is $\overrightarrow{V_0V_1}$ and $\overleftarrow{V_2V_3}$ are edges of $Q$) we can find a paths of length at least $2^{d-1}-1$ in both $V_0$ and $V_3$, which again can be joined through $Q$ to give an $a-b$ path of length at least $2^d-1$ as required. This just leaves the case of a directed path \begin{equation} { \label{description} \overrightarrow{V_0V_1}, \overrightarrow{V_1V_2} \mbox{ and } \overrightarrow{V_2V_3}. } \end{equation} While here we obtain a path of length at least $2^{d-1}-1$ from the edge $\overrightarrow{V_0V_1}$ as before, in $V_1,V_2,V_3$ we might not be able to guarantee a full endblock path. Indeed, we may now have the possibility that the edges $\overrightarrow{V_{i-1}V_i}$ and $\overrightarrow{V_iV_{i+1}}$ correspond to one edge entering an endblock $E$ by a vertex $x$ in its interior and the other edge leaving E by a vertex $y$ in its interior. This does not allow us to use induction on Theorem \ref{conn} as $\mbox{cutv}(E)\in E-\{x,y\}$ may have degree lower than $d-1$. However, if this does not happen at one of $V_1$ or $V_2$ the same proof applies. Also if the exit vertex $x$ of $V_2$ guaranteed by $\overrightarrow{V_{2}V_3}$ had $p(x)$ in the interior of an endblock of $V_3$ we would be able to find our two endblock paths, one in $V_0$ as mentioned already of length at least $2^{d-1}-1$ and the other in $V_3$ of length at least $2^{d-1}-2$. Since joining both of these through $V_1$ and $V_2$ joins at least four more vertices onto these paths, we can extend them to form an $a-b$ path of length at least $2^{d}+1$, more than required. Therefore we must have $p(x)\notin \mbox{int}(E)$. But now take any outneighour of $V_3$ in $H$. Combined with our path $Q$ above it is easily seen we can obtain a path $Q'$ of length three which is either (i) not of the form (\ref{description}), or (ii) contains $V_3$ as an interior vertex and allows for a full endblock path to be build through it. In both cases we are done. } \end{proof} The fact that Theorem \ref{path3extension} holds for $G$ again enables us to guarantee that $H$ has at least two connected components. As before Lemma \ref{body} again gives a connected component $C$ of $H$ consisting entirely of limbs. Now Theorem \ref{nice} still holds with an identical proof -- the slight variation in the definition of $H$ allows us to guarantee that $|S|,|T|\geq 2$ in the case where $C$ consist of exactly two vertices and that $t'\geq 1$ when it consists of at least three, which ensures 2-connectivity. This time however, it guarantees a path of length at least $2^{d-1}-2$ from any $v\in J-\{a',b'\}$ to $a'$ not containing $b$ or to $b'$ not containing $a$. Take $J$ in $G_C$ as guaranteed by Lemma \ref{nice}. \begin{lem} { \label{noedge} There does not exist an edge in $G$ from $v\in J-\{a',b'\}$ to $G-J$ } \end{lem} \begin{proof} { For contradiction suppose $v\in J-\{a',b'\}$ had a neighbour $w$ outside of $J$. Without loss of generality take $v\in G_b$. Then $w\notin G_C$ by Theorem \ref{nice}(i) and so $w\in K$ for some limb $K$ of $a$ say or $w\in \mbox{Core}(a)$. First take $w\in K$, $K\in C'$ for some connected component $C'$ of $H$. If $w\notin \mbox{int}(E)$ for some endblock $E$ of $K$ take a $w-x_E$ path $P_2$ in $K$ of length at least $2^{d-1}-1$, where $x_E$ is an exit vertex of $E$, which exists by induction on Theorem \ref{conn}. Combining this with the path $P_1$ given from Lemma \ref{nice}(ii) in $J$ from $a'$ to $v$ of length $2^{d-1}-2$ and the edge $x_Ep(x_E)$ we have a path $P_1vwP_2x_Ep(x_E)$ from $a'$ to $p(x_E)$ of length at least $(2^{d-1}-2)+1+(2^{d-1}-1)+1= 2^d-1$. But this path extends to a path from $a$ to $b$, a contradiction. If $w\in \mbox{int}(E)$ for some endblock $E$ of $K$, we would like to combine the $\mbox{Joint}(K)-w$ path guaranteed by induction on $E$ with the $v-b'$ path in $J$ as given by Lemma \ref{nice}(ii) using the edge $wv$. This path extends to an $a-b$ path but may only have length $(2^{d-1}-2)+1+(2^{d-1}-2)=2^d-3$, too little for us. Instead, look at an outneighbour of $K$ in $H$. If this is a limb then $C'$ must consist entirely of limbs (by Lemma \ref{body}) and therefore contains a path of the form $\overleftarrow{KW}$ or of the form $KVW$ where $\overleftarrow{VW}$ is an edge of $H$ as all limbs have at least one outneighbour in $H$. This allows us to build a path $P$ from $w$ to $\mbox{Joint}(W)$ of length at least $2^{d-1}+1$ in $C'$. Now combining $P$ with an appropriate path from Lemma \ref{nice}(ii) via the edge $vw$ we obtain a path that extends to an $a-b$ path of length at least $(2^{d-1}-2)+1+(2^{d-1}+1)=2^d$, as required -- take this path to be the $a'-v$ path if $W\in B$ or the $b'-v$ path if $W\in A$. If the outneighbour of $K$ in $H$ is $\mbox{Core}(b)$, again using Lemma \ref{nice}(ii) we can find a $b'-v$ path in $J$ of length at least $2^{d-1}-2$ which extends through $C'$ to give a $y-b'$ path $P_2$ of length at least $2^{d-1}$, where $y\in \mbox{Body}(b)$. Now $H$ must have a third connected component $C''$ containing a limb of $b$ since $b$ has at least two limbs and only one element of $B$ can lie in a component by Lemma \ref{body}. This component gives an $a-z$ path $P_1$ of length at least $2^{d-1}-2$ where again $z\in \mbox{Body}(b)$ and $P_1$ and $P_2$ are disjoint. As in the proof of Lemma \ref{paths3} we can join $P_1$ and $P_2$ together in $\mbox{Core}(a)$ with a small use of 2-connectivity to give an $a-b$ path of length at least $2^d-1$ as required. This completes the case when $w\in K$. Finally, the case where $w\in \mbox{Core}(a)$ follows a similar argument. } \end{proof} Using Lemma \ref{noedge} we can apply Theorem \ref{conn} to $J$ taking $a'$ in place of $a$ and $b'$ in place of $b$. This guarantees a path of length $2^{d}-2$ from $a'$ to $b'$. Moreover, unless $J$ is isomorphic to $Q_d$ with $a=a'$ and $b=b'$ where $a$ and $b$ are at even Hamming distance, $J$ contains a path of length at least $2^{d}-1$ between $a$ and $b$, so we may assume this is the case. Since $G$ is not isomorphic to $Q_d$, the graph $G'=G[V(G)-J\cup \{a,b\}]$ is non-empty and all $v$ in $G'-\{a,b\}$ have degree at least $d$ in $G'$. If $a$ and $b$ both have more than two limbs in $G'$, they guarantee that $G'$ is 2-connected. In this case since $|G'|<|G|$ we can apply Theorem \ref{conn} to $G'$. This gives an $a-b$ path in $G'$ of length at least $2^d-1$ unless $G'$ is isomorphic to $Q_d$. If $G'$ was isomorphic to $Q_d$ then $J$ and $G'$ would both contain the subcube containing $a$ and $b$, which contains at least four points since $a$ and $b$ are at even Hamming distance. But from construction $G'$ and $J$ only share $a$ and $b$, so $G'$ is not isomorphic to $Q_d$ and therefore contains an $a-b$ path of length at least $2^d-1$. If one of $a$ and $b$ has exactly one limb, $a$ say, $G'$ must be of the form $G_C$ for some component $C$ of $H$ (as all limbs of $b$ must have a out-neighbour in $H$). But then we can apply Theorem \ref{nice} to $G'$ to obtain a 2-connected subgraph $\widetilde{J}$ and vertices $\widetilde{a}\in G'_a$ and $\widetilde{b}\in G'_b$. As in Lemma \ref{nice}(i) for any $v\in \widetilde{J}-\{ \widetilde{a},\widetilde{b}\}$, $\widetilde{J}$ contains all neighbours of $v$ in $G_C=G'$. As such $v$ can have no neighbours in $G$ other than those in $G'$ we have $d_{\widetilde{J}}(v)=d_{G}(v) \geq d$. Theorem \ref{conn} now holds for $\widetilde{J}$ taking $\widetilde{a}$ and $\widetilde{b}$ in place of $a$ and $b$ so $\widetilde{J}$ contains a $\widetilde{a}-\widetilde{b}$ path of length at least $2^d-2$ which extends to an $a-b$ path in $G'$ of length at least $2^{d}-2$ in $G$. Again as above, since $J$ and $\widetilde{J}$ cannot both be isomorphic to $Q_{d}$ if $a$ and $b$ are at even Hamming distance, $\widetilde{J}$ must contain an $a-b$ path of length at least $2^d-1$, as required.\\ Lastly, we show that the degree condition can again be removed. As in Section 6 we can assume that $d_{G_b}(b)\geq 2$ and that $a$ has a neighbour $v\in G_b$, $v\neq b$ with $v\in E_v$ an endblock of $G_b$ -- the proof of Lemma \ref{remove} is identical. In order for the previous argument to be inapplicable $a$ must have exactly one neighbour $a'$ in $G_a$. Again we switch attention to the limbs of $a'$ and $b$ to construct the interaction digraph $H$. Note that both $a'$ and $b$ have at least 2 limbs. Now each endblock $F\neq E_v, \{a,a'\}$ of $a'$ or $b$ has at least two exit vertices $x$ and $y$ such that $F$ contains a path of length at least $2^{d-1}-1$ from $\mbox{cutv}(F)$ to $x$ and to $y$. We choose one of these for $x_F$ with the condition that $p(x_F)\notin \{a,a',b\}$, which is possible since $p(a)\in E_v$. Additionally, let $x_{E_v}=v$. Using these vertices construct the interaction digraph $H$ as before. We claim that again $p(x_E)\notin \mbox{int}(F)$ for all endblocks $E$ and $F$ of $a'$ and $b$, when $F\neq \{a,a'\}$. Otherwise there exists a path $P_1$ from $a'$ to $x_E$ of length at least $2^{d-1}-1$ and a path $P_2$ from $p(x_E)$ to $b$ of length at least $2^{d-1}-2$. But then we have a path $aa'P_1x_Ep(x_E)P_2$ of length at least $1+(2^{d-1}-1)+1+(2^{d-1}-2)=2^d-1$ from $a$ to $b$, a contradiction. This allows us to again establish Lemma \ref{paths3} and Lemma \ref{body} for $G$ since the problem of entering and leaving an endblock $E$ through vertices in its interior cannot happen. These give a component $C$ of $H$ consisting entirely of limbs of $a'$ and $b$, with $\{a,a'\}\notin C$. Now we can apply Lemma \ref{nice} to $G_C$ to obtain $J$, $a''$, $b''$ as before. If no $u\in J-\{a'',b''\}$ has a neighbour outside of $J$ then by induction on Theorem \ref{conn} $J$ contains an $a''-b''$ path $P$ of length at least $2^d-2$. As extending $P$ to an $a-b$ path adds at least one more edge $aa'$, $G$ contains an $a-b$ path of length at least $2^d-1$, a contradiction. Therefore we may assume some $u \in J-\{a'',b''\}$ has a neighbour $w$ outside of $J$. Now notice that the proof of Lemma \ref{nice}(ii) actually gives an $a''-u$ path not containing $b$ and a $b''-u$ path not containing $a$ in $J$ of length at least $2^{d-1}$ in this case. Indeed if we are looking for such an $a''-u$ path, as before $J$ contains an endblock $F$ on the opposite side of $J$ to $u$ (that is $F\subset J\cap G_a$ if $u\in G_b$ and $F\subset J\cap G_b$ if $u\in G_a$) for which we can find vertex disjoint paths between $\{a'',u\}$ and $\{\mbox{cutv}(F),y\}$ for some vertex $y\in \mbox{int}(F)$. The path containing $u$ here must have length at least 2 since $p(u)=w$ which lies outside of $J$. Combining these with the $\mbox{cutv}(F)-y$ path of length $2^{d-1}-2$ in $F$ gives the required path. If $w\in G_{C'}$ with $C'$ a component of $H$ not containing $\{a,a'\}$ then the remainder of the proof is exactly as in the case where $d_{G_a}(a)\geq 2$. So we can assume $\{a,a'\}\in C'$. Clearly $w\neq a$ as $u\neq a',v$. If $w\in \mbox{int}(E_v)$ then $E_v$ contains a $\mbox{cutv}(E_v)-w$ path of length at least $2^{d-1}-2$. Combining this path via the edge $uw$ with the $a''-u$ path of length at least $2^{d-1}$ in $J$ we have an $a''-\mbox{cutv}(E_v)$ path of length at least $2^d-1$ which extends to an $a-b$ path, as required. If $w\notin \mbox{int}(E_v)$ then $G_{C'}$ contains a $w-a$ path $P_1$ of length at least $2^{d-1}-1$ passing through $E_v$. Combining this with the $u-b'$ path $P_2$ in $J$ of length at least $2^{d-1}$ with the edge $uw$ we have an $a-b'$ path $P_1wuP_2$ of length at least $2^d$. As this extends to an $a-b$ path, this finishes the proof of Theorem \ref{conn}. \hspace{11cm} $\square$ \section{Generalizations} The reader might notice that we have used very little about $Q_n$ in the proof of Theorem \ref{conn}. The $n$-dimensional grid ${\Z}^n$ is the graph whose vertex set consists of $n$-tuples with entries in $\Z$ and in which two vertices $x$ and $y$ are adjacent if $|x_i-y_i|=1$ for some $i\in [n]$ and $x_j=y_j$ for all $j\neq i$. The next theorem extends Theorem \ref{conn} (and therefore Theorems \ref{mainthm} and \ref{cyclethm}) to subgraphs of $\Z^d$. \begin{thm} { Let $G$ be a 2-connected subgraph of ${\Z}^n$ and $a,b\in V(G)$. Suppose that $d(z) \geq d$ for all $z \in V(G)-\lbrace a,b\rbrace$. Then $a$ and $b$ are joined by a path of length at least $2^d-2$. Furthermore unless $G$ is isomorphic to $Q_d$ with $a$ and $b$ at even Hamming distance from each other, $G$ contains an $a-b$ path of length $2^d-1$. } \end{thm} \begin{proof} { The crucial property of $\Z ^n$ here is that we can always find a splitting of $G$ into two connected pieces, $G_a$ and $G_b$ with $a\in G_a$ and $b\in G_b$ such that $d_{G_a}(a)\geq 1$ and $d_{G_b}(b)\geq 1$ and all $v\in G$ lose at most one neighbour in their piece. Indeed, taking some coordinate $j$ on which $a$ and $b$ differ, say with $a_j>b_j$, let $G_1$ be the induced subgraph of $G$ consisting all vertices $v$ with $v_j\geq a_j$ and $G_2$ be the induced subgraph of $G$ consisting of all $w$ for which $w_j<a_j$. Again with the same modification to these graphs as in Lemma \ref{splitting} we obtain connected graphs $G_a$ and $G_b$ with the required degree conditions. From here on the proof is identical to that of Theorem \ref{conn}. } \end{proof} Moreover, the same proof also extends to subgraphs of the discrete torus $C_k^{n}$ provided $k\geq 4$. Now we cannot expect a bound of the form $C2^d$ as above for subgraphs of the discrete torus $C_3^{d}$ as this graph has minimum degree $2d$ but only $3^d$ points. This shows that given a subgraph $G$ of $C_3^{n}$ of minimal degree at least $d$ we cannot in general guarantee a path of length more than $3^{\frac{d}{2}}-1$ in $G$. Why does our approach not work in this case? The main reason is that we cannot guarantee a partition into two subgraphs such that all vertices lose at most one neighbour in their piece. Can we still guarantee an exponentially long path in this case? The following general result shows that we can. \begin{thm} { \label{gen} Let $k\in \N$ and $G$ be a 2-connected graph with $a,b\in V(G)$. Suppose $d(v)\geq d$ for all $v\in V(G)-\{a,b\}$. Furthermore, suppose that $G$ has the following property: \begin{quotation} {\noindent Given any two vertices $x,y\in G$, there is a partition of $V(G)$ into two sets $X$ and $Y$ with $x\in X$ and $y\in Y$ such that $d_{G[X]}(v)\geq d(v)-k$ for all $v\in X$ and $d_{G[Y]}(v)\geq d(v)-k$ for all $v\in Y$.} \end{quotation} Then $G$ contains an $a-b$ path of length at least $2^{\frac{d}{k+2}}$. } \end{thm} Note that if the property above holds for $G$, it also holds for all subgraphs of $G$. As an immediate corollary of Theorem \ref{gen} we have the following: \begin{cor} { Every subgraph of $C_3^n$ of minimum degree at least $d$ contains a path of length at least $2^{\frac{d}{4}}$. } \end{cor} It would be interesting to decided what the correct lower bounds for the length of the longest path in subgraphs of $C_3^n$ with minimum degree at least $d$. \begin{conjecture} { Given a subgraph $G$ of $C_3^n$ with minimum degree at least $d$, $G$ must contain a path of length at least $3^{\frac{d}{2}}-1$. } \end{conjecture} Another consequence of Theorem \ref{gen} is the following result for product graphs. \begin{thm} { Let $G_1,\ldots ,G_l$ be graphs with maximum degree at most $k$. Then given any subgraph $G$ of the Cartesian product graph $\prod_{i=1}^{l}G_i$ of minimum degree at least $d$, $G$ contains a path of length at least $2^{\frac{d}{k+2}}$. } \end{thm} The proof of Theorem \ref{gen} is similar to that of Theorem \ref{conn} but shorter. \begin{proof} { The proof is again by induction on $d$. It suffices to prove the result for $d\geq k+4$ as otherwise it follows from 2-connectivity. As in the proof of Theorem \ref{conn} we wish to split $G$ into two subgraphs $G_a$ and $G_b$ with $a\in G_a$ and $b\in G_b$, which is the motivation for the above splitting property. However, simply taking $a$ and $b$ in place of $x$ and $y$ might not be useful as both $a$ and $b$ can have degree as low as two in $G$ in which case in the partition guaranteed above $a$ may end up with all its neighbours in $Y$. Instead we pick a neighbour $a'\neq b$ of $a$ and a neighbour $b'\neq a$ of $b$. The fact that $G$ is 2-connected ensures it is possible to pick $a'\neq b'$. Now take the partition guaranteed from our splitting property above with $x=a'$ and $y=b'$. Moving $a$ to $X$ and $b$ to $Y$ as needed we have that $d_{G[X]}(v)\geq d(v)-k-1$ for all $v\in X-a$ and similarly for $v\in Y-b$. Both $a$ and $b$ now have at least 1 neighbour in $G[X]$ and $G[Y]$ respectively. Finally, denoting the connected component of $G[X]$ containing $a$ by $C_a$, let $G_b$ be the connected component of $G-C_a$ containing $b$ and $G_a=G[V(G)-V(G_b)]$. Note that $G_a$ and $G_b$ are connected with $a\in G_a$, $b\in G_b$. Moreover, $d_{G_a}(v)\geq d(v)-k-1$ for all $v\in G_a-a$ and $d_{G_b}(v)\geq d(v)-k-1$ for $v\in G_b-b$. We will again analyse the block-cutvertex decompositions of $G_a$ and $G_b$. The following lemma will be very useful below. \begin{lem} { \label{endblock} Let $E$ be an endblock of $G_a$ or $G_b$ with $a,b\notin \mbox{int}(E)$. Then given any two vertices $u,v\in \mbox{int}(E)$, $G[E]$ contains a path of length at least $2^{\frac{d-k-2}{k+2}}$ from $u$ to $v$. } \end{lem} \begin{figure} \centering \input{smallerlongpath.pstex_t} \caption{Cases where $G'$ is not 2-connected in Lemma \ref{endblock}} \label{smallerlongpath} \end{figure} \begin{proof} { Look at the block-cutvertex decomposition of $G'=G[E]-\mbox{cutv}(E)$. Since $E$ is 2-connected (as $d\geq k+4$), $G'$ is connected and $\mbox{cutv}(E)$ must have a neighbour in the interior of every endblock of $G'$. Note that every vertex $v\in G'$ has $d_{G'}(v) \geq d_G(v)-k-2$. In particular, since $d\geq k+4$ each endblock $F$ of $G'$ is 2-connected and has at least three vertices so that we can by induction apply Theorem \ref{gen} to it. If $G'$ is 2-connected then by induction on Theorem \ref{gen} $G'$ contains the desired path from $u$ to $v$. Thus we may assume that $G'$ is not 2-connected. If $u\in \mbox{int}(F_1)$ and $v\in \mbox{int}(F_2)$ where $F_1$ and $F_2$ are two distinct endblocks of $G'$ then by induction on Theorem \ref{gen}, $G[F_1]$ and $G[F_2]$ contain $u-\mbox{cutv}(F_1)$ and $\mbox{cutv}(F_2)-v$ paths respectively, each of length at least $2^{\frac{d-k-2}{k+2}}$. Joining $\mbox{cutv}(F_1)$ to $\mbox{cutv}(F_2)$ by a third path in $G'$ and combining all three of these paths, we get a $u-v$ path of length at least $2^{\frac{d}{k+2}}$, as required. Therefore since $G'$ contains at least two endblocks, we can assume that one of these, say $F$, does not contain $u$ or $v$ in its interior. Contracting $\mbox{int}(F)$ down to a single vertex in $G[E]$, the resulting graph is still 2-connected. Therefore, as in the proof of Lemma \ref{nice}, $G[E]$ contains two vertex disjoint paths $P_1$ and $P_2$ from the set $\{u,v\}$ to $\{\mbox{cutv}(F),w\}$ for some $w\in \mbox{int}(F)$, with $(P_1\cup P_2)\cap (F-\{\mbox{cutv}(F),w\})=\emptyset$. Now using induction on Theorem \ref{gen} in $F$, it contains a path $P_3$ of length $2^{\frac{d-k-2}{k+2}}$ from $\mbox{cutv}(F)$ to $w$. Piecing $P_1$, $P_2$ and $P_3$ together we obtain our desired path. } \end{proof} Again we have: \begin{prop} { \label{intnew} Let $E$ be an endblock of $G_a$ not containing $a$ and $F$ an endblock of $G_b$ not containing $b$. Then $G$ does not contain an edge from $\mbox{int}(E)$ to $\mbox{int}(F)$ } \end{prop} \begin{proof} { Exactly as in Proposition \ref{int}. } \end{proof} \begin{lem} { \label{newlist} We have the following: \begin{itemize} { \item[\emph{(i)}] Given any endblock $E$ of $G_a$ not containing $a$, there are two disjoint edges from $\mbox{int}(E)$ to $G_b$ in $G$. \item[\emph{(ii)}] $G_a$ contains an endblock not containing $a$. } \end{itemize} } \end{lem} \begin{proof} { (i) $E$ must have an exit vertex $x_1$, with neighbour $y\in G_b$, as $G$ is 2-connected. If it had only one, $G'=G[E]$ is 2-connected and every $v\in G'-\{x_1,\mbox{cutv}(E)\}$ has degree at least $d$. Therefore, by induction on Theorem \ref{gen}, $G'$ contains a path of length at least $2^{\frac{d}{k+2}}$ from $x_1$ to $\mbox{cutv}(E)$. Extending this path from $\mbox{cutv}(E)$ to $a$ in $G_a$ and from $y$ to $b$ in $G_b$ we obtain an $a-b$ path of desired length. Therefore we may assume $E$ contains a second exit vertex $x_2$. Now if the vertices in ${\mbox{int}(E)}$ were only adjacent to $y$ in $G_b$, $x_1y$ and $x_2y$ must be edges of $G$. Then $G''=G[E\cup \{y\}]$ is 2-connected and $d_{G''}(v)\geq d$ for every $v\in G''-\{y,\mbox{cutv}(E)\}$. By Theorem \ref{gen} $G''$ contains a $\mbox{cutv}(E)-y$ path of length at least $2^{\frac{d}{k+2}}$. Again, extending this to a path from $a$ to $b$, we have an $a-b$ path of length at least $2^{\frac{d}{k+2}}$. Therefore we may assume the two edges exist or we are done. (ii) The proof is almost identical to the proof of Lemma \ref{list}(ii). } \end{proof} Take an endblock $E$ of $G_a$ not containing $a$, as guaranteed by Lemma \ref{newlist}(ii). We can choose $E$ such that $a$ and all $v\in G_a-E$ not contained in the interior of an endblock of $G_a$ lie in the same connected component of $G_a-E$ (e.g. pick a block $B$ in $G_a$ containing $a$ and choose $E$ to be a block at maximum distance from $B$ in $\mathcal{B}(G_a)$). Let $x_1y_1$ and $x_2y_2$ be the disjoint edges of $G$ with $x_1,x_2\in \mbox{int}(E)$ and $y_1,y_2\in G_b$ guaranteed by Lemma \ref{newlist}(i). By Proposition \ref{intnew}, $y_1,y_2\notin \mbox{int}(F)$ for all endblocks $F$ of $G_b$ not containing $b$ in its interior. \begin{figure} \centering \input{Generalized.pstex_t} \caption{Path created in Theorem \ref{gen}} \label{Generalized} \end{figure} Now looking at the block-cutvertex decomposition of $G_b$ we can choose two vertex disjoint paths in $G_b$ from $\{y_1,y_2\}$ to $\{b, \mbox{cutv}(F)\}$ where $F$ is some endblock of $G_b$ not containing $b$. Lets say that these paths are $P_3$ from $\mbox{cutv}(F)$ to $y_1$ and $P_5$ from $y_2$ to $b$. Applying Lemma \ref{newlist}(i) to $F$ we see that there exists $u\in \mbox{int}(F)$ adjacent to some $v\in G_a$, $v\neq \mbox{cutv}(E)$. Furthermore, by Proposition \ref{intnew} $v\notin \mbox{int}(E')$ for any endblock $E'$ of $G_a$. From our choice of $E$ there exists an $a-v$ path $P_1$ in $G_a-E$. Finally by induction on Theorem \ref{gen}, $F$ contains a $u-\mbox{cutv}(F)$ path $P_2$ of length at least $2^{\frac{d-k-1}{k+2}}$ and by Lemma \ref{endblock} $E$ contains an $x_1x_2$ path $P_4$ of length at least $2^{\frac{d-k-2}{k+2}}$. Combining these five paths we obtain an $a-b$ path $P=P_1vuP_2P_3y_1x_1P_4x_2y_2P_5$ of length at least $2^{\frac{d-k-1}{k+2}}+2^{\frac{d-k-2}{k+2}}>2^{\frac{d}{k+2}}$ as required. } \end{proof} The cycle analogues of the above theorems can be obtained in a similar fashion to the proof of Theorem \ref{cyclethm} from Theorem \ref{conn}. As mentioned in the Introduction, we do not know the correct bound for the length of the longest path in a subgraph of $Q_n$ when the minimum degree condition in Theorem \ref{mainthm} is replaced by an average degree condition. Is the following possible? \vspace{0.25cm} \begin{conjecture} { Every subgraph of $Q_n$ with average degree at least $d$ contains a path of length at least $2^d-1$. } \end{conjecture} \section*{Acknowledgements} { I would like to thank my supervisor Imre Leader for many helpful conversations and David Conlon for reading through an earlier version of this paper. }
{ "timestamp": "2010-06-16T02:01:53", "yymm": "1006", "arxiv_id": "1006.3049", "language": "en", "url": "https://arxiv.org/abs/1006.3049", "abstract": "Let $Q_n$ denote the graph of the $n$-dimensional cube with vertex set $\\{0,1\\}^n$ in which two vertices are adjacent if they differ in exactly one coordinate. Suppose $G$ is a subgraph of $Q_n$ with average degree at least $d$. How long a path can we guarantee to find in $G$?Our aim in this paper is to show that $G$ must contain an exponentially long path. In fact, we show that if $G$ has minimum degree at least $d$ then $G$ must contain a path of length $2^d-1$. Note that this bound is tight, as shown by a $d$-dimensional subcube of $Q_n$. We also obtain the slightly stronger result that $G$ must contain a cycle of length at least $2^d$ and prove analogous results for other `product-type' graphs.", "subjects": "Combinatorics (math.CO)", "title": "Long paths and cycles in subgraphs of the cube", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9896718477853187, "lm_q2_score": 0.8267117919359419, "lm_q1q2_score": 0.8181733867111556 }
https://arxiv.org/abs/1608.02558
On Functions Whose Mean Value Abscissas Are Midpoints, with Connections to Harmonic Functions
We investigate functions with the property that for every interval, the slope at the midpoint of the interval is the same as the average slope. More generally, we find functions whose average slopes over intervals are given by the slope at a weighted average of the endpoints of those intervals. This is equivalent to finding functions satisfying a weighted mean value property. In the course of our exploration, we find connections to harmonic functions that prompt us to explore multivariable analogues and the existence of "weighted harmonic functions."
\section{Introduction and Statement of the Problem} Recall the mean value theorem of calculus, which states that if $f$ is a differentiable function on $[a,b]$, then there is a $c \in (a,b)$, which we call the \emph{mean value abscissa}, such that \begin{equation*} \frac{f(b) - f(a)}{b-a} = f'(c). \end{equation*} In typical applications, such as in most proofs of the fundamental theorems of calculus, we are given $f$ and know there exists such a $c$. But what if we place conditions on $c$ and ask when there exist functions $f$ satisfying the mean value theorem with that $c$? For instance, what if $c$ is exactly halfway between $a$ and $b$? \begin{question}\label{q1} For which differentiable functions $f$ is \begin{equation}\label{eq:MAIN} \frac{f(b) - f(a)}{b - a} = f'\left( \frac{b+a}{2} \right) \end{equation} for all $b > a$? \end{question} \noindent In other words, for which functions $f$ does the midpoint of each interval serve as the mean value abscissa for that interval? The midpoint of $[a,b]$ is the average of $a$ and $b$. What about more general averages, like weighted averages? \begin{question}\label{q2} For a fixed $\lambda \in (0,1)$, which differentiable functions $f$ satisfy \begin{equation}\label{eq:MAIN2} \frac{f(b) - f(a)}{b - a} = f' \left(\lambda a + (1 - \lambda)b \right) \end{equation} for all $b > a$? \end{question} \noindent In other words, for which functions $f$ do weighted averages of the endpoints of intervals serve as mean value abscissas for those intervals? We explore the answers to these two questions. Along the way, we'll find strong connections between functions satisfying these constrained mean value theorem requirements and harmonic functions which will prompt us to investigate high dimensional analogues. \section{A First Attempt at the First Question} Before trying to find all such functions, we should note that if $f(x)$ is an affine function $f(x) = ax + b$ for any $a,b$, then $f$ satisfies the conditions for both questions. So there are infinite families of solutions. But without further details, it might be hard to find all functions satisfying~\eqref{eq:MAIN}. \begin{remark} In fact, it was known to Archimedes that parabolas also satisfy~\eqref{eq:MAIN}. This appears in Proposition~1 in his work on the quadrature of the parabola~\cite{heath1897works}. In modern times, some calculus textbook authors include verifying that parabolas satisfy~\eqref{eq:MAIN} as an exercise. For instance, it appears as an ``Additional and Advanced'' exercise in~\cite{thomas2010thomas}. \end{remark} When first encountering a problem, it can be useful to consider simplifying assumptions. We might be able to find more functions by expanding the left- and right-hand sides of~\eqref{eq:MAIN}. Suppose that $f$ is a function that satisfies~\eqref{eq:MAIN}, and suppose $f$ is at least $3$ times continuously differentiable. Then we have the Taylor polynomial $f(x) = f(y) + f'(y)(x-y) + \tfrac{1}{2}f''(y)(x-y)^2 + \tfrac{1}{6}f'''(y)(x-y)^3 + h(x)(x-y)^3$, where $h(x) \to 0$ as $x \to y$. Subtracting $f(y)$ and dividing by $x-y$ yields an expression for the left-hand side of~\eqref{eq:MAIN}: \begin{equation*} \frac{f(x)- f(y)}{x-y} = f'(y) + \tfrac{1}{2}f''(y)(x-y) + \tfrac{1}{6}f'''(y)(x-y)^2 + h(x)(x-y)^2. \end{equation*} Taking the Taylor polynomial for $f'(x)$ centered at $y$ and substituting $(x+y)/2$, we see that the right-hand side of~\eqref{eq:MAIN} is \begin{equation*} f'(y) + \tfrac{1}{2} f''(y) (x-y) + \tfrac{1}{8} f'''(y)(x-y)^2 + \tfrac{1}{4}\widetilde{h}(\tfrac{x+y}{2})(x-y)^2, \end{equation*} where $\widetilde{h}(x) \to 0$ as $x \to y$. Setting these two expressions equal, we cancel the $f'(y)$ and $f''(y)$ terms and divide each side by $(x-y)^2$. Letting $x \to y$, we see that $\tfrac{1}{6}f'''(y) = \tfrac{1}{8} f'''(y)$, or rather that $f'''(y) = 0$. Since this is true for all $y$, the third derivative is identically $0$. So $f$ is at most a quadratic polynomial. We can check that if $f(x) = a + bx + cx^2$, then \begin{equation*} \frac{f(x) - f(y)}{x-y} = \frac{b(x-y) + c(x^2 - y^2)}{x-y} = b + c(x+y) = f'\left(\frac{x+y}{2}\right), \end{equation*} so every polynomial of degree at most $2$ is an answer to Question 1. Exploration through Taylor polynomials led us only to parabolas. But we made a big assumption. What about $f$ that satisfy the conditions of the first question but which are not necessarily three times differentiable? \subsection{Connection to harmonic functions.} By fixing various $a$ and allowing $b$ to vary in~\eqref{eq:MAIN}, one can show that $f'(x)$ is continuous. We now use the fundamental theorem of calculus to rewrite~\eqref{eq:MAIN} as \begin{equation} \frac{1}{b-a} \int_a^b f'(t) \, dt = f' \left(\frac{b+a}{2}\right), \end{equation} or equivalently as \begin{equation}\label{eq:fderiv_is_harmonic} \frac{1}{2h} \int_{x-h}^{x+h} f'(t)\, dt = f'(x). \end{equation} This can be interpreted to mean that the value $f'(x)$ is given by the average of $f'$ in any symmetric interval around $x$, a very restrictive property. This is precisely the mean value property satisfied by harmonic functions, \begin{equation}\label{eq:harmonic_mvt} g(x) = \frac{1}{|B_h(x)|} \int_{B_h(x)} g(t)\, dV, \end{equation} where $g$ is a harmonic function on $\mathbb{R}^n$, $B_h(x)$ denotes the ball of radius $h$ centered at $x$, $dV$ denotes the standard Euclidean volume measure, and $|B_h(x)|$ denotes the volume of $B_h(x)$. Harmonic functions appear in many areas of mathematics, but are particularly fundamental to the theory of differential equations. Recall that a twice differentiable function $g$ is called \emph{harmonic} if the sum of the second partial derivatives of $g$ is identically $0$. At least, that's the standard definition. Harmonic functions can be thought of in more intuitive ways. One particularly illuminating overview on harmonic functions is given by Needham~\cite{needham1994geometry}, who explores geometric interpretations of harmonic functions by exploring relations similar to~\eqref{eq:harmonic_mvt}. In the memorably titled~\cite{kac1966can}, Kac gives a different approach to harmonic functions while explaining why they are called ``harmonic.'' The study of harmonic functions usually begins with $\mathbb{R}^2$, as it is trivial to identify the harmonic functions on $\mathbb{R}$. The only functions on $\mathbb{R}$ with identically zero second derivative are affine functions $ax + b$. It is a remarkable theorem that any function satisfying~\eqref{eq:harmonic_mvt} is itself harmonic (for the standard proof, see Lemma~4.6 and subsequent discussion in~\cite{stein2009real}). In fact, the condition~\eqref{eq:MAIN} is the degenerate one-dimensional form of another integral mean value property shared by harmonic functions, namely \begin{equation}\label{eq:meanvalueshell} g(x) = \frac{1}{|\partial B_h(x)|} \int_{\partial B_h(x)} g(t) \, dS, \end{equation} where $\partial B_h(x)$ denotes the boundary of $B_h(x)$, $|\partial B_h(x)|$ is the surface area of $B_h(x)$, and $dS$ denotes the standard Euclidean surface area measure. So for any function $f$ satisfying~\eqref{eq:MAIN}, its derivative $f'$ satisfies~\eqref{eq:fderiv_is_harmonic}, and so is harmonic. Thus $f'$ is at most a linear polynomial, so that $f$ is at most a quadratic polynomial. We have filled in the gap left from Taylor expansions by appealing to harmonic functions. \section{Investigating Weighted Mean Values} Let us now try to understand the more general Question~\ref{q2}. Pursuing the connection to harmonic functions, we can rewrite~\eqref{eq:MAIN2} as the integral weighted mean value property \begin{equation}\label{eq:weighted_mean_integral_badcoords} f'\left(\lambda a + (1-\lambda) b\right) = \frac{1}{b-a} \int_a^b f'(t)\, dt. \end{equation} This is similar to the regular harmonic mean value property~\eqref{eq:fderiv_is_harmonic}, and so we might ask whether functions satisfying~\eqref{eq:weighted_mean_integral_badcoords} are harmonic. This turns out to be the case. To prove this, we first show that any function satisfying~\eqref{eq:weighted_mean_integral_badcoords} is infinitely differentiable. We take $x=\frac{a+b}{2}$ and $h=\frac{b-a}{2}$ and rewrite~\eqref{eq:weighted_mean_integral_badcoords} as \begin{equation}\label{eq:weighted_mean_integral_goodcoords} f'\left(x+(1-2\lambda) h\right)=\frac{1}{2h} \int_{x-h}^{x+h} f'(t)\, dt, \end{equation} which now holds for any $x$ and any $h > 0$. We can center the left hand side at $x$ through a change of variables, getting \begin{equation*} f'(x) = \frac{1}{2h} \int_{x - (2 - 2\lambda)h}^{x + 2\lambda h} f'(t)\, dt. \end{equation*} By the fundamental theorem of calculus, we can directly differentiate $f'(x)$ to see that $f''(x) = \frac{1}{2h} \left( f'(x + 2\lambda h) - f'(x - (2 - 2\lambda)h )\right)$. By the same argument, each of the two terms appearing in the derivative are differentiable with derivatives expressible as linear combinations of $f'$. Inductively, we can show that $f'$ is infinitely differentiable. \begin{remark} We can see the infinite differentiability in a different way. We can compose the integral weighted mean value property with itself to write $f'$ as the $2$-fold integral \begin{equation*} f'(x) = \frac{1}{2h} \int_{x - (2 - 2\lambda)h}^{x + 2\lambda h} \left( \frac{1}{2h} \int_{t - (2 - 2\lambda h)}^{t + 2\lambda h} f'(u) \, du \right) \, dt, \end{equation*} which is clearly twice-differentiable, again by the fundamental theorem of calculus. Composing $n$ times, we can represent $f'$ as an $n$-fold integral which is $n$ times differentiable. So $f'$, and therefore $f$, is infinitely differentiable. \end{remark} As $f$ is at least twice continuously differentiable, we can use Taylor's Theorem as before to write $f(x) = f(y) + f'(y)(x-y) + \tfrac{1}{2}f''(y)(x-y)^2 + h(x)(x-y)^2$, where $h(x) \to 0$ as $x \to y$. Similar to when we first tackled Question~1, we expand the left-hand side and right-hand side of~\eqref{eq:MAIN2}. The left-hand side remains \begin{equation*} f'(y) + \tfrac{1}{2}f''(y)(x-y) + h(x)(x-y). \end{equation*} Using the linear Taylor polynomial for $f'(x)$ and simplifying, the right-hand side becomes \begin{equation*} f'(y) + \lambda f''(y)(x-y) + \lambda \widetilde{h}(\lambda x + (1-\lambda)y)(x-y), \end{equation*} where $\widetilde{h}(x) \to 0$ as $x \to y$. Setting these equal, we may cancel $f'(y)$ and divide by $(x-y)$. Letting $x \to y$, we see that $\tfrac{1}{2}f''(y) = \lambda f''(y)$. There are two possibilities. If $\lambda = \tfrac{1}{2}$, then our ``weighted average'' is just the normal average from Question 1. Otherwise, we must have $f''(y) = 0$. This is true for all $y$, and so the second derivative is identically $0$. We conclude that the only twice differentiable functions satisfying the conditions from Question~2 are linear polynomials $f(x) = ax + b$ unless $\lambda = \tfrac{1}{2}$, when $f$ can be a quadratic polynomial. We have shown that the normally-weighted mean value property on $\mathbb{R}$ is distinguished as the only weighted mean value property that leads to a nontrivial family. It was not obvious (to the authors, at least) that there would be no analogous ``weighted harmonic functions'' corresponding to a weighted mean value property, and this adds to the list of special properties held by harmonic functions. We summarize the answers to the two original questions in the following theorem. \begin{theorem}\label{thm:firsttwoquestions} Fix a $\lambda \in (0,1)$, and suppose $f:\mathbb{R} \longrightarrow \mathbb{R}$ satisfies the weighted mean value condition \begin{equation} \frac{f(b) - f(a)}{b-a} = f'(\lambda a + (1-\lambda) b) \end{equation} for all $a < b$. Then $f$ is a quadratic polynomial. Further, if $\lambda \neq \tfrac{1}{2}$, then $f$ is merely a linear polynomial. \end{theorem} \begin{remark} We have used that~\eqref{eq:MAIN2} is supposed to hold for all $a < b$, allowing both $a$ and $b$ to vary. Another interesting question comes from fixing $a=0$ and letting $b$ vary. Then we might look for $f$ with the mean value abscissa on the intervals $[0, b]$ being given by weighted averages of $0$ and $b$. In this case, there \emph{are} additional families of functions, but only for $\lambda = 1/(k+1)^{1/k}$. The reader might try to explore this approach using Taylor expansions, or perhaps another method entirely. \end{remark} \section{Weighted Harmonic Functions in Higher Dimensions} Our initial questions led us to study functions satisfying the integral weighted mean value property for $\mathbb{R}$, and we found that the non-weighted average has a distinguished connection to harmonic functions. Since there are many more harmonic functions on $\mathbb{R}^n$ for $n \geq 2$ than there are for $\mathbb{R}$, there might be enough wiggle room for a ``weighted harmonic function'' to exist corresponding to a weighted mean value property on $\mathbb{R}^n$. In analogy with~\eqref{eq:weighted_mean_integral_goodcoords}, we can ask about the following integral weighted mean value property \begin{equation}\label{eq:weighted_mean_Rn} g(\mathbf{x} + (1 - 2\lambda)h\mathbf{v}) = \frac{1}{|B_h(\mathbf{x})|} \int_{B_h(\mathbf{x})} g(\mathbf{t}) \, dV. \end{equation} Here, $\lambda \in (0,1)$ is the weight, and $\mathbf{v}$ is a unit vector indicating the direction of the weighting. The mean value abscissa $(\mathbf{x} + (1 - 2\lambda)h\mathbf{v})$ is a point in the ball $B_h(\mathbf{x})$ that differs from the center of the ball by a distance proportional to the radius. Then~\eqref{eq:weighted_mean_Rn} can be interpreted to mean that the average value of $g$ on the ball $B_h(\mathbf{x})$ is given by the value of $g$ at the point $(\mathbf{x} + (1 - 2\lambda)h\mathbf{v})$ inside that ball. So we ask the following question. \begin{question}\label{q3} For a fixed $\lambda \in (0,1)$ and a fixed unit vector $\mathbf{v} \in \mathbb{R}^n$, which functions $g \in C^1(\mathbb{R}^n)$ satisfy \begin{equation*} g(\mathbf{x} + (1 - 2\lambda)h\mathbf{v}) = \frac{1}{|B_h(\mathbf{x})|} \int_{B_h(\mathbf{x})} g(\mathbf{t}) \, dV \end{equation*} for all $\mathbf{x}$ and all $h > 0$? \end{question} Notice that when $n = 1$ and $v = 1$, this is exactly Question~\ref{q2}. As this generalizes Question~\ref{q2}, our method also generalizes our approach to answering Question~\ref{q2}. When $\lambda = \frac{1}{2}$, the mean value abscissa is exactly the center of the ball, and so the weighted mean value property becomes the ordinary mean value property~\eqref{eq:harmonic_mvt}. Then for $\lambda = \frac{1}{2}$, it is exactly harmonic functions $g$ which answer Question~\ref{q3}. Let us now consider $\lambda \neq \frac{1}{2}$. We first show that any function satisfying~\eqref{eq:weighted_mean_Rn} is infinitely differentiable. Rewrite~\eqref{eq:weighted_mean_Rn} as \begin{equation}\label{eq:weighted_mean_Rn_shifted} g(\mathbf{x}) = \frac{1}{|B_h|} \int_{B_h(\mathbf{x}-(1 - 2\lambda)h \mathbf{v})} g(\mathbf{t}) \, dV, \end{equation} where $\lvert B_h \rvert$ denotes the volume of the ball of integration. As $g$ is continuously differentiable, we can now compute any partial derivative of $g$ by differentiating~\eqref{eq:weighted_mean_Rn_shifted} using the Leibniz rule for higher dimensions (sometimes called the Reynolds transport theorem, see~\cite{flanders1973} for additional exposition) to obtain \begin{align*} \partial_ig(\mathbf{x}) &= \frac{1}{|B_h|} \int_{\partial B_h(\mathbf{x}-(1 - 2\lambda)h \mathbf{v})} g(\mathbf{t}) \mathbf{e}_i\cdot \mathbf{n} \, dS \\ &=\frac{1}{|B_h|} \int_{\partial B_h(0)} g(\mathbf{t}+\mathbf{x}-(1 - 2\lambda)h \mathbf{v}) \mathbf{e}_i\cdot \mathbf{n} \, dS, \end{align*} where $\mathbf{e}_i$ denotes the $i$-th standard basis vector of $\mathbb{R}^n$, $\partial_i$ denotes the partial derivative with respect to the $i$-th coordinate, and $\mathbf{n}$ is the outward-pointing unit normal vector field of the boundary $\partial B_h$. To compute the second partial derivative $\partial_{ij}$, we can now differentiate under the integral sign \begin{align*} \partial_{ij}g(\mathbf{x}) &= \frac{1}{|B_h|} \int_{\partial B_h(0)} \partial_jg(\mathbf{t}+\mathbf{x}-(1 - 2\lambda)h \mathbf{v}) \mathbf{e}_i\cdot \mathbf{n}\, dS \end{align*} since the first partial derivatives of $g$ exist and are continuous. Continuing inductively, we conclude that $g$ is smooth. \begin{remark} In fact, it should be possible to relax the assumption that $g\in C^1$ and deduce this from the argument above under the assumption that $g$ is at least continuous. However, for simplicity in applying the Reynolds transport theorem, we assume $g$ has continuous partial derivatives. \end{remark} As $g$ is at least twice continuously differentiable, we can again use Taylor's Theorem, which states \begin{equation} g(\mathbf{x}) = g(\mathbf{y}) + (\mathbf{x}-\mathbf{y})\cdot \nabla g(\mathbf{y}) +\mathcal{O}(|\mathbf{x}-\mathbf{y}|^2), \end{equation} where $\nabla g$ denotes the gradient of $g$ and $\mathcal{O}$ is big-oh notation, indicating roughly that the remainder vanishes at least as quickly as $\lvert \mathbf{x}-\mathbf{y} \rvert^2$ as $\mathbf{x} \to \mathbf{y}$. We now compute \begin{align*} g(\mathbf{x}+(1-2\lambda)h \mathbf{v})&= \frac{1}{|B_h|} \int_{B_h(\mathbf{x})} g(\mathbf{t}) \, dV \\ &= \frac{1}{|B_h|} \int_{B_h(0)} g(\mathbf{x}+\mathbf{t}) \, dV_\mathbf{t} \\ &= \frac{1}{|B_h|} \int_{B_h(0)} \bigg( g(\mathbf{x})+\mathbf{t}\cdot \nabla f(\mathbf{x}) +\mathcal{O}(|\mathbf{t}|^2) \bigg) \, dV_\mathbf{t} \\ &= g(\mathbf{x})+\nabla g(\mathbf{x})\cdot\left(\frac{1}{|B_h|} \int_{B_h(0)} \mathbf{t}\, dV_\mathbf{t}\right)+\mathcal{O}(h^2). \end{align*} We use the notation $dV_\mathbf{t}$ to remind ourselves that the variable of integration is $\mathbf{t}$, not $\mathbf{x}$. The integral of the vector field $\mathbf{t}$ over the ball $B_h(0)$ is zero, so the integral term above vanishes. We have thus shown that \begin{equation*} g(\mathbf{x} + (1 - 2\lambda)h \mathbf{v}) = g(\mathbf{x}) + \mathcal{O}(h^2). \end{equation*} Using this expression, we can show that $g$ is constant in the $\mathbf{v}$-direction, that is, $\frac{\partial g}{\partial \mathbf{v}}=0$. We compute \begin{equation*} \frac{\partial g}{\partial \mathbf{v}}=\lim_{h\to 0}\frac{g(\mathbf{x}+(1-2\lambda)h\mathbf{v})-g(\mathbf{x})}{(1-2\lambda)h} = \lim_{h\to 0}\frac{g(\mathbf{x})+\mathcal{O}(h^2)-g(\mathbf{x})}{(1-2\lambda)h} =0. \end{equation*} As $g$ is constant in the $\mathbf{v}$-direction, and the mean value abscissa $(\mathbf{x} + (1-2\lambda)h\mathbf{v})$ deviates from the center of the ball $B_h(\mathbf{x})$ in exactly the $\mathbf{v}$-direction, we can show that $g$ is harmonic. At each $\mathbf{x}$, the function $g$ satisfies \begin{equation*} g(\mathbf{x})=g(\mathbf{x}+(1-2\lambda)h\mathbf{v}) = \frac{1}{|B_h(\mathbf{x})|} \int_{B_h(\mathbf{x})} g(\mathbf{t}) \, dV, \end{equation*} which is precisely the mean value property for harmonic functions. By choosing an orthonormal basis for $\mathbb{R}^n$ with $\mathbf{v}$ as a basis vector, we can think of $g$ as coming from a function $\widetilde{g}$ on $\mathbb{R}^{n-1}$, which is extended to $\mathbb{R}^n$ trivially by having no dependence on the $\mathbf{v}$-coordinate. Noting that the sums of the second partial derivatives of $g$ and $\widetilde{g}$ are the same, we have that $\widetilde{g}$ is also harmonic. So a function $g$ satisfying the integral weighted mean value theorem~\eqref{eq:weighted_mean_Rn} for $\lambda \neq \frac{1}{2}$ is really a harmonic function on $\mathbb{R}^{n-1}$, extended to $\mathbb{R}^n$ by being constant in the $\mathbf{v}$-direction. For example, in the case $n=2$, without loss of generality we may take $\mathbf{v}=(0,1)$, i.e., the unit vector in the $y$-direction, so that $g$ is constant in $y$. From this we see that the weighted harmonic functions are simply those which are harmonic functions of the single variable $x$, that is, linear functions $g(x,y)=ax+b$. In conclusion, we have shown the following \begin{theorem}\label{thm:thirdquestion} Fix $\lambda \in (0,1)$ and a unit vector $\mathbf{v} \in \mathbb{R}^n$. Suppose $g: \mathbb{R}^n \longrightarrow \mathbb{R}$ satisfies \begin{equation*} g(\mathbf{x}+(1-2\lambda)h\mathbf{v}) = \frac{1}{|B_h(\mathbf{x})|} \int_{B_h(\mathbf{x})} g(\mathbf{t})\, dV \end{equation*} for all $\mathbf{x} \in \mathbb{R}^n$ and $h > 0$. Then $g$ is a harmonic function. Further, if $\lambda \neq \tfrac{1}{2}$, then $g$ is constant in the $\mathbf{v}$-direction, i.e., $\frac{\partial}{\partial \mathbf{v}} g = 0$. \end{theorem} This is consistent with our answers to Questions~\ref{q1} and~\ref{q2}, and includes Theorem~\ref{thm:firsttwoquestions} as the $\mathbb{R}^1$ case. This is also consistent with our earlier observation that the normally-weighted mean value property is distinguished among weighted mean value properties by including a strictly larger family of functions. \begin{remark} As a final note, it is also very natural to think of those $g$ satisfying the alternate weighted mean value property \begin{align}\label{eq:meanvalueshellweighted} g(\mathbf{x} + (1 - 2\lambda)h\mathbf{v}) = \frac{1}{|\partial B_h(\mathbf{x})|} \int_{\partial B_h(\mathbf{x})} g(\mathbf{t}) \, dS, \end{align} analogous to the harmonic mean value property~\eqref{eq:meanvalueshell}. We claim that, similar to the case of harmonic functions, this condition is equivalent to~\eqref{eq:weighted_mean_Rn} and hence leads to a result equivalent to Theorem~\ref{thm:thirdquestion}. Again assuming for simplicity that $g\in C^1$, using a very similar argument as above, it is not hard to show that $g$ is in fact infinitely differentiable and satisfies $\frac{\partial g}{\partial \mathbf{v}}=0$. We do not repeat the details here, but just note that to show differentiability, we rewrite the property~\eqref{eq:meanvalueshellweighted} as \begin{equation} g(\mathbf{x}) = \frac{1}{|\partial B_h|} \int_{\partial B_h(0)} g(\mathbf{t}+\mathbf{x}-(1 - 2\lambda)h\mathbf{v}) \, dV. \end{equation} We can now differentiate this expression to obtain smoothness of $g$. From this, we use a similar argument involving Taylor's theorem as above to find that $\frac{\partial g}{\partial \mathbf{v}}=0$. From this we deduce the harmonic mean value property~\eqref{eq:meanvalueshell}, and hence the conclusion of Theorem~\ref{thm:thirdquestion} remains valid. \end{remark} \section*{Acknowledgements} D. L.-D.\ was supported by the National Science Foundation Graduate Research Fellowship Program under Grant No. DGE 0228243. P. C.\ gratefully acknowledges support by the National Science Foundation through grant DMS-1148284.
{ "timestamp": "2016-11-14T02:00:53", "yymm": "1608", "arxiv_id": "1608.02558", "language": "en", "url": "https://arxiv.org/abs/1608.02558", "abstract": "We investigate functions with the property that for every interval, the slope at the midpoint of the interval is the same as the average slope. More generally, we find functions whose average slopes over intervals are given by the slope at a weighted average of the endpoints of those intervals. This is equivalent to finding functions satisfying a weighted mean value property. In the course of our exploration, we find connections to harmonic functions that prompt us to explore multivariable analogues and the existence of \"weighted harmonic functions.\"", "subjects": "Classical Analysis and ODEs (math.CA)", "title": "On Functions Whose Mean Value Abscissas Are Midpoints, with Connections to Harmonic Functions", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9896718462621992, "lm_q2_score": 0.826711791935942, "lm_q1q2_score": 0.8181733854519748 }
https://arxiv.org/abs/1109.4067
Finite lattices and Gröbner bases
Gröbner bases of binomial ideals arising from finite lattices will be studied. In terms of Gröbner bases and initial ideals, a characterization of finite distributive lattices as well as planar distributive lattices will be given.
\section*{Introduction} Let $L$ be a finite lattice and $K[L]$ the polynomial ring in $|L|$ variables over a field $K$ whose variables are the elements of $L$. A binomial of $K[L]$ of the form $ab - (a\wedge b)(a\vee b)$ is called a basic binomial. Let $I_L$ denote the ideal of $K[L]$ generated by basic binomials of $K[L]$. The ideal $I_L$ was first introduced by \cite{H}. It is shown in \cite{H} that $I_L$ is a prime ideal if and only if $L$ is a distributive lattice, see also \cite[Theorem 10.1.3]{HH}. When $L$ is distributive, the set of basic binomials of $K[L]$ is a Gr\"obner basis with respect to any rank reverse lexicographic order. In the present paper, by studying Gr\"obner bases of $I_L$, a Gr\"obner basis characterization of distributive lattices as well as planer distributive lattices will be obtained. Moreover, we discuss the problem when $I_L$ has a quadratic Gr\"obner basis with respect to any monomial order. \section{Characterization of distributive lattices} We refer the reader to \cite{St} for fundamental materials on finite lattices. Let $\hat{0}$ (resp.\ $\hat{1}$) denote a unique minimal (resp.\ maximal) element of a finite lattice. Recall that a finite lattice $L$ is called {\em modular} if $x \leq b$ implies $x\vee(a\wedge b) = (x \vee a)\wedge b$ for all $x,a,b \in L$. A finite lattice is modular if and only if no sublattice of $L$ is isomorphic the pentagon lattice of Figure $1$. A finite lattice is called {\em distributive} if, for all $x,y,z \in L$, the distributive laws $x\wedge(y\vee z) = (x\wedge y)\vee(y\wedge z)$ and $x\vee(y\wedge z) = (x\vee y)\wedge(y\vee z)$ hold. Every modular lattice is distributive. A modular lattice is distributive if and only if no sublattice of $L$ is isomorphic to the diamond lattice of Figure $1$. \begin{figure}[h] \begin{center} \psset{unit=0.5cm} \begin{pspicture}(-10.3,-2.5)(4,3.5) \psline(2,3)(0.5,1.5) \psline(0.5,1.5)(0.5,0) \psline(0.5,0)(2,-1) \psline(2,-1)(3.5,0.75) \psline(3.5,0.75)(2,3) \rput(2,3){$\bullet$} \put(1.9,3.3){$a$} \rput(0.5,1.5){$\bullet$} \put(-0.1,1.5){$b$} \rput(0.5,0){$\bullet$} \put(-0.1,0){$c$} \rput(2,-1){$\bullet$} \put(1.9,-1.6){$e$} \rput(3.5,0.75){$\bullet$} \put(3.8,0.75){$d$} \psline(-9,3)(-11,1) \psline(-9,3)(-9,1) \psline(-9,3)(-7,1) \psline(-11,1)(-9,-1) \psline(-9,1)(-9,-1) \psline(-7,1)(-9,-1) \rput(-9,3){$\bullet$} \put(-9.1,3.3){$a$} \rput(-11,1){$\bullet$} \put(-11.6,0.9){$b$} \rput(-9,1){$\bullet$} \put(-8.6,0.9){$c$} \rput(-7,1){$\bullet$} \put(-6.7,1){$d$} \rput(-9,-1){$\bullet$} \put(-9.1,-1.6){$e$} \rput(-8.7,-2.5){Diamond Lattice} \rput(1.7,-2.5){Pentagon Lattice} \end{pspicture} \end{center} \caption{}\label{Fig1} \end{figure} A finite lattice is called {\em pure} if all maximal chains between $\hat{0}$ and $\hat{1}$ have the same length. When a finite lattice is pure, then the {\em rank function} of $L$ can be defined. More precisely, if $L$ is a finite pure lattice and $a \in L$, then the {\em rank} of $a$ in L, denoted by $\rank(a)$, is the largest integer $r$ for which there exists a chain of $L$ of the form \[ \hat{0} = a_0 < a_1 < \cdots < a_r = a. \] If a finite lattice $L$ is modular, then one has the equality \begin{eqnarray} \label{modular} \rank(p) + \rank(q) = \rank(p \wedge q) + \rank(p \vee q) \end{eqnarray} for all $p, q \in L$. \medskip Let $K$ be field and $L$ a finite lattice. We consider the polynomial ring $K[L]$ whose variables correspond to the elements of $L$, and define the ideal $I_L\subset K[L]$ as the binomial ideal whose generators are all basic binomials attached to $L$. As explained in the introduction, a binomial of the form $ab-cd$ with $c=a \vee b$ and $c=a\wedge b$ is called a {\em basic binomial}. The residue class ring $K[L]/I_L$ will be denoted by $R(L)$ We refer the reader to \cite{HH} for basic terminologies and notation on Gr\"obner bases. A {\em rank reverse lexicographic order} on $K[L]$ is the reverse lexicographic order with the property that $a>b$ if $\rank(a)>\rank(b)$. \begin{Theorem} \label{easy} Let $L$ be a finite modular lattice. Then the following conditions are equivalent: \begin{enumerate} \item[{\em (i)}] $L$ is a distributive lattice; \item[{\em (ii)}] $I_L$ has a squarefree Gr\"obner basis with respect to any rank reverse lexicographic order. \end{enumerate} \end{Theorem} \begin{proof} The implication (i) $\Rightarrow$ (ii) is well known, see \cite{H} and \cite[Theorem 10.1.3]{HH}. (ii) $\Rightarrow$ (i): Suppose that $L$ is a finite modular lattice which is not distributive. Then $L$ contains the diamond lattice of Figure \ref{Fig1}. Since $L$ is modular, one has the equality (\ref{modular}) for all $p, q \in L$. Hence if $g = pq - p'q'$ is a basic monomial of $I_L$, then \[ \rank(p) + \rank(q) = \rank(p') + \rank(q'). \] In particular the ranks of $b,c$ and $d$ coincide. We fix a rank reverse lexicographic order $<$ with the property that $d < q$ for all $q \in L$ with $\rank(q) = \rank(d)$. Our work is to show that $\ini_<(I_L)$ cannot be squarefree. Suppose, on the contrary, that $\ini_<(I_L)$ is squarefree. First we claim $ad^2e\in \ini_<(I_L)$. In fact, $ad^2e-a^2e^2\in I_L$, because \[ ad^2e-a^2e^2=d\bigl(d(ae-bc)+c(bd-ae)\bigr)+ae(cd-ae). \] Since $\ini_<(I_L)$ is squarefree and since $ad^2e\in \ini_<(I_L)$, it follows that $ade \in \ini_<(I_L)$. Hence there exists a binomial $f = ade - u \in I$, where $u$ is a monomial of degree $3$, with $\ini_<(f) = ade$. Let $f = \sum_{i=1}^{N} x_if_i$, where each $x_i$ is a variable and where each $f_i = v_i - w_i$ is a basic binomial of $I_L$, such that $x_1v_1 = ade$ and $x_iw_i = x_{i+1}v_{i+1}$ for all $1 \leq i < N$. A crucial fact is that each variable appearing in $x_if_i$ belongs to the interval $[e,a]$ of $L$. To see why this is true, we observe that if $f_i = v_i - w_i$ is a basic binomial of $I_L$ and if each variable appearing in $v_i$ belongs to $[e,a]$, then each variable appearing in $w_i$ must belong to $[e,a]$. Now, since $x_1v_1 = ade$ and $x_iw_i = x_{i+1}v_{i+1}$ for all $1 \leq i < N$, this observation guarantees that each variable appearing in $x_if_i$ belongs to the interval $[e,a]$ of $L$. In particular $u = x_Nw_N$ consists of variables belonging to $[e,a]$, say $u = \ell m n$. Now, one has $f = ade - \ell m n \in I$, where $\ell, m$ and $n$ belong to $[e,a]$. Let $\ell \geq m \geq n$. Since we are working with a rank reverse lexicographic order, it follows that $e$ is the smallest variable among all variables belonging to $[e,a]$. Since $\ini_<(f) = ade$, one has $n = e$. On the other hand, since $I_L$ is generated by basic binomials of $L$, it follows easily that if $g = p_1p_2\cdots p_r - q_1q_2 \cdots q_r$ is a binomial belonging to $I_L$, then \[ \sum_{i=1}^{r} \rank(p_i) = \sum_{i=1}^{r} \rank(q_i). \] Thus in particular one has \[ \rank(a) + \rank(d) = \rank(\ell) + \rank(m). \] Since $a$ is a unique maximal element of $[a,e]$, it follows that $\rank(a) \geq \rank(\ell) (\geq \rank(m))$. Hence $\rank(d) \leq \rank(m)$. If $\rank(d) =\rank(m)$, then $d<m$ by the given order of the variables. On the other hand, if $\rank(d) < \rank(m)$, then $d<m$, since we use a rank reverse lexicographic order. Thus in any case $d<m$, and this implies that $\ini_<(f) = \ell m n$, a contradiction. Consequently, the monomial $ade$ cannot belong to $\ini_<(I_L)$. Hence $ad^2e$ belongs to a unique minimal set of monomial generators of $\ini_<(I_L)$. Thus $\ini_<(I_L)$ cannot be squarefree. \end{proof} It can be easily checked that for any monomial order, $\ini_<(I_{N_5})$ is squarefree where $N_5$ is the pentagon lattice, while $\ini_<(I_{N_3})$ is not squarefree where $N_3$ is the diamond lattice. \begin{Conjecture}[Squarefree conjecture] \label{believeitornot} {\em Let $L$ be a modular lattice. Then for any monomial order $\ini_<(I_L)$ is not squarefree, unless $L$ is distributive.} \end{Conjecture} \section{Characterization of planar distributive lattices} Let ${\NZQ N}^2$ denote the (infinite) distributive lattice consisting of all pairs $(i, j)$ of nonnegative integers with the partial order $(i, j)\leq (k, l)$ if and only if $i\leq k$ and $j\leq l$. A {\em planar distributive lattice} is a finite sublattice $L$ of ${\NZQ N}^2$ with $(0, 0)\in D$ \begin{Theorem} \label{alsoeasy} Let $L$ be a finite modular lattice. Then the following conditions are equivalent: \begin{enumerate} \item[{\em (i)}] $L$ is a planar distributive lattice; \item[{\em (ii)}] $I_L$ has a squarefree initial ideal with respect to the lexicographic order for any order of the variables; \item[{\em (iii)}] $I_L$ has a squarefree initial ideal with respect to any monomial order. \end{enumerate} \end{Theorem} \begin{proof} (i) \implies (iii): Let $m$ and $n$ be the smallest integers such that $L\subset [m]\times [n]$, and consider the polynomial ring $T=K[t_1,\ldots,t_m,s_1,\ldots,s_n]$. We define a $K$-algebra homomorphism $\varphi\: R(L)\to T$ which assigns to $a=(i,j)$ the monomial $s_it_j\in T$. The image $A$ of $\varphi$ is the edge ring of a bipartite graph $G$. The basic relations $ab=(a\vee b)(a\wedge b)$ of $R(L)$ are mapped under $\varphi$ to relations of $A$ corresponding to $4$-cycles of the bipartite graph $G$. It is shown in \cite[Theorem 1.2]{OH} that the defining ideal of the edge ring $A$ is generated by the binomials corresponding to 4-cycles, if each even cycle of length $\geq 6$ has a chord. That $G$ satisfies this property is shown by Querishi \cite{Q}. It follows that $R(L)\iso A$. Hence we may identify $I_L$ with the toric edge ideal $J$ defining $A$. Next we use a result of Sturmfels (see \cite[Chapter 9]{Stu}) which says the universal Gr\"obner basis of the toric edge ideal of bipartite graph consists of the binomials corresponding to the even cycles with no chords. From this it follows that $\ini_<(J)$ is squarefree for any monomial order. (iii) \implies (ii) is trivial. (ii) \implies (i): Suppose $L$ is not planar, then $L$ contains a sublattice which is isomorphic to the Boolean lattice $B_3$ of rank 3 as shown in Figure~\ref{Fig2}. \begin{figure}[h] \begin{center} \psset{unit=0.7cm} \begin{pspicture}(1,-1.5)(5,3) \psline(3,3)(1.5,1.5) \psline(3,3)(3,1.5) \psline(3,3)(4.5,1.5) \psline(1.5,1.5)(1.5,0) \psline(1.5,1.5)(3,0) \psline(3,1.5)(1.5,0) \psline(3,1.5)(4.5,0) \psline(4.5,1.5)(4.5,0) \psline(4.5,1.5)(3,0) \psline(1.5,0)(3,-1.5) \psline(3,0)(3,-1.5) \psline(4.5,0)(3,-1.5) \rput(3,3){$\bullet$} \put(3,3.2){$a$} \rput(1.5,1.5){$\bullet$} \put(1.05,1.4){$b$} \rput(3,1.5){$\bullet$} \put(2.6,1.5){$c$} \rput(1.5,0){$\bullet$} \put(1.1,-0.1){$e$} \rput(4.5,1.5){$\bullet$} \put(4.7,1.4){$d$} \rput(3,0){$\bullet$} \put(2.55,-0.25){$f$} \rput(4.5,0){$\bullet$} \put(4.7,-0.1){$g$} \rput(3,-1.5){$\bullet$} \put(2.8,-2){$h$} \end{pspicture} \end{center} \caption{}\label{Fig3} \end{figure} Let $<$ be the lexicographic order induced by an ordering such that $g<f<e<h<a<d<c<b$ and $b<q$ for any other $q\in L$. The initial ideal of $I_{B_3}$ contains the monomial $ah^2$ in the minimal set of monomial generators. Since $<$ is an elimination order (see \cite[Exercise 2.9]{HH}) it follows that $ah^2$ belongs to the minimal set of monomial generators of $I_L$. \end{proof} In contrast to the order given in the proof of the preceding theorem, there exist lexicographic orders such that $I_{B_3}$ is quadratic, or not quadratic but squarefree. The question arises whether for any finite distributive lattice there exists a lexicographic monomial order such that $\ini_<(I_L)$ is squarefree. Recall that the {\em divisor lattice} of a positive integer $n$ is the lattice $D_n$ consisting of all divisors of $n$ ordered by divisibility. Every divisor lattice is a distributive lattice. Let, in general, $L$ be a finite pure lattice. A {\em cut edge} of $L$ is a pair $(a,b)$ of elements of $L$ with $\rank(b) = \rank(a)+1$ such that $$|\{c\in L\:\; \rank(c)=\rank(a)\}|= |\{c\in L\:\; \rank(c)=\rank(b)\}|=1.$$ \begin{Theorem} \label{againeasy} Let $L$ be a finite lattice with no cut edges. Then the following conditions are equivalent: \begin{enumerate} \item[{\em (i)}] $L$ is the divisor lattice of $2\cdot 3^r$ for some $r\geq 1$; \item[{\em (ii)}] $I_L$ has a quadratic Gr\"obner basis with respect to any monomial order. \end{enumerate} \end{Theorem} \begin{proof} (i)\implies (ii): The ideal $I_L$ can be identified with the toric edge ideal of the complete bipartite graph of type $(2,r)$, since $L$ has no cut edges. Each cycle in a bipartite graph of type $(2,r)$ is of length $4$. Hence by using again \cite[Chapter 9]{Stu}, the basic binomials of $I_L$ form a universal Gr\"obner basis. This yields the desired conclusion. (ii)\implies (i): As we have seen in the proofs of Theorem~\ref{easy} and in Theorem~\ref{alsoeasy} that the lattice $L$ cannot contain as a sublattice the diamond lattice and the Boolean lattice $B_3$. It also cannot contain the pentagon lattice $N_5$ of Figure~\ref{Fig1}. Indeed, if we choose the lexicographic order induced by $c<e<a<d<b$, then $bae$ is a minimal generator of $\ini_<(I_{N_5})$. This implies that $L$ is a planar distributive lattice. Finally, $L$ cannot contain as a sublattice the planar distributive lattice $C_2$ of Figure~\ref{Fig3}. \begin{figure}[h] \begin{center} \psset{unit=0.6cm} \begin{pspicture}(1,-1.5)(5,3) \pspolygon(3,3)(4,2)(2,0)(3,-1)(4,0)(2,2)(3,3) \rput(3,3){$\bullet$} \put(2.9,3.3){$a$} \rput(2,2){$\bullet$} \put(1.5,1.9){$b$} \rput(3,1){$\bullet$} \put(3.3,0.9){$d$} \rput(4,0){$\bullet$} \put(4.2,-0.1){$f$} \rput(3,-1){$\bullet$} \put(2.9,-1.5){$g$} \rput(2,0){$\bullet$} \put(1.5,-0.1){$e$} \rput(4,2){$\bullet$} \put(4.2,1.9){$c$} \end{pspicture} \end{center} \caption{}\label{Fig2} \end{figure} In fact, if we choose the lexicographic order induced by $g<f<e<c<b<a<d$, then $aef$ is a minimal generator of $\ini_<(I_{N_5})$. Hence $L$ must be the divisor lattice of $2\cdot 3^r$ for some $r \geq 2$. \end{proof} As a strengthening of Theorem~\ref{againeasy} we expect \begin{Conjecture}[Quadratic conjecture] {\em Let $I$ be an ideal generated by binomials such that $I$ has a quadratic Gr\"obner basis with respect to any monomial order. Then either the generators of $I$ are binomials in pairwise different sets of variables or $I=I_L$ where $L$ is the divisor lattice of $2\cdot 3^r$ for some $r\geq 1$.} \end{Conjecture}
{ "timestamp": "2011-09-20T02:03:55", "yymm": "1109", "arxiv_id": "1109.4067", "language": "en", "url": "https://arxiv.org/abs/1109.4067", "abstract": "Gröbner bases of binomial ideals arising from finite lattices will be studied. In terms of Gröbner bases and initial ideals, a characterization of finite distributive lattices as well as planar distributive lattices will be given.", "subjects": "Commutative Algebra (math.AC)", "title": "Finite lattices and Gröbner bases", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9869795087371921, "lm_q2_score": 0.8289388125473628, "lm_q1q2_score": 0.8181456219811875 }
https://arxiv.org/abs/1803.07497
Discrete Cubical and Path Homologies of Graphs
In this paper we study and compare two homology theories for (simple and undirected) graphs. The first, which was developed by Barcelo, Caprano, and White, is based on graph maps from hypercubes to the graph. The second theory was developed by Grigor'yan, Lin, Muranov, and Yau, and is based on paths in the graph. Results in both settings imply that the respective homology groups are isomorphic in homological dimension one. We show that, for several infinite classes of graphs, the two theories lead to isomorphic homology groups in all dimensions. However, we provide an example for which the homology groups of the two theories are not isomorphic at least in dimensions two and three. We establish a natural map from the cubical to the path homology groups which is an isomorphism in dimension one and surjective in dimension two. Again our example shows that in general the map is not surjective in dimension three and not injective in dimension two. In the process we develop tools to compute the homology groups for both theories in all dimensions.
\section{Introduction} For a simple finite undirected graph $G$, we study a discrete cubical singular homology theory $\mathcal{H}_\bullet^{\normalfont {\textrm{Cube}}}(G)$. This theory is a special case of the discrete cubical homology theory $DH_{\bullet,r}(X)$ that was defined by Barcelo, Caprano and White \cite{BCW} for any metric space $X$ and any real number $r > 0$. Their work builds on a discrete homotopy theory for undirected graphs introduced earlier by Barcelo, Kramer, Laubenbacher, and Weaver in \cite{BKLW}. Later work by Babson, Barcelo, de Longueville, and Laubenbacher \cite{BBLL} connects this theory to classical homotopy theory of cubical sets and asks for a corresponding homology theory. The homology theory developed in \cite{BCW} is an answer to that question. A more general but closely related homotopy theory for directed graphs was developed by Grigor'yan, Lin, Muranov, and Yau in \cite{GLMY2}, which also introduces a corresponding homology theory based on directed paths. The homotopy theories in \cite{BBLL} and \cite{GLMY2} are identical when $G$ is undirected and from \cite{BCW} and \cite{GLMY2} it follows that the homology theories yield isomorphic homology groups in dimension $1$. In this paper we explore both the similarities and differences between the two homology theories, showing that they agree in all dimensions for many infinite classes of undirected graphs but disagree in general. Both theories differ markedly from classical singular/simplicial homology of graphs seen as $1$-dimensional complexes or their clique complexes. For example, when $G$ is a $4$-cycle, both the cubical and path homologies are trivial in all dimensions greater than zero. In \ref{sec-definitions}, following \cite{BCW} and \cite{GLMY2}, we give precise definitions of both cubical and path homology for undirected graphs, and discuss the differences between these theories and classical simplicial homology of a graph (as a $1$-dimensional simplicial complex as well as of the clique complex of the graph). In \ref{sec-homotopyhomology} we give proofs that both cubical and path homology are preserved under homotopy equivalence, along lines that essentially appear in \cite{BCW} and \cite{GLMY2}. These results are used in \ref{sec-computations} to compute homology for a large number of examples, showing in the process that cubical and path homology agree in all of these cases. \ref{sec-map} constructs a natural homomorphism from $\mathcal{H}_\bullet^{\normalfont {\textrm{Cube}}}(G)$ to $\mathcal{H}_\bullet^{\normalfont {\textrm{Path}}}(G)$. We show that the homomorphism is an isomorphism in dimension $0$ and $1$ and surjective in dimension $2$, hence fueling speculation that this might explain the isomorphisms observed in \ref{sec-computations}. However, \ref{sec-counter} gives a counterexample: a graph $G$ for which $\mathcal{H}_\bullet^{\normalfont {\textrm{Cube}}}(G) \not\cong \mathcal{H}_\bullet^{\normalfont {\textrm{Path}}}(G)$, and for this example the map defined in \ref{sec-map} is neither injective nor surjective. \ref{sec-questions} suggests several natural questions for further study. \section{Background: Discrete Homology of Graphs} \label{sec-definitions} Throughout the paper let $R$ denote a commutative ring with unit which shall be the ring of coefficients. For any positive integer $n$, let $[n] := \{1,\dots,n\}$. For graph theory definitions and terminology we refer the reader to \cite{Diestel00}. \subsection*{Discrete Cubical Homology} \begin{Definition} For $n \geq 1$, the discrete $n$-cube $Q_n$ is the graph whose vertex set $V(Q_n)$ is $\{0,1\}^n := \{(a_1,\dots, a_n)\,|\,a_i \in \{0,1\} \mbox{ for all } i \in [n]\}$, with an edge between two vertices $a$ and $b$ if and only if their Hamming distance is exactly one, that is, there exists $i \in [n]$ such that $a_{i} \neq b_{i}$ and $a_{j} = b_{j}$ for all $j \neq i$. For $n=0$, we define $Q_0$ to be the $1$-vertex graph with no edges. \end{Definition} \begin{Definition} Let $G$ and $H$ be simple graphs, i.e. undirected graphs without loops or multiple edges. A graph homomorphism $\sigma: G \longrightarrow H$ is a map from $V(G)$ to $V(H)$ such that, if $\{a,b\}\in E(G)$ then either $\sigma(a)=\sigma(b)$ or $\{\sigma(a),\sigma(b)\}\in E(H)$. \end{Definition} \begin{Definition} Let $G$ be a simple graph, a graph homomorphism $\sigma: Q_n \longrightarrow G$ is called a \textit{singular} $n$-cube on $G$. \end{Definition} For each $n \geq 0$, let ${\mathcal{L}}^{\normalfont {\textrm{Cube}}}_n(G)$ be the free $R$-module generated by all singular $n$-cubes on $G$. For $n \geq 1$ and each $i \in [n]$, we define two face maps $f_i^+$ and $f_i^-$ from ${\mathcal{L}}^{\normalfont {\textrm{Cube}}}_n(G)$ to ${\mathcal{L}}^{\normalfont {\textrm{Cube}}}_{n-1}(G)$ such that, for $\sigma \in {\mathcal{L}}^{\normalfont {\textrm{Cube}}}_n(G)$ and $(a_1,\dots,a_{n-1}) \in Q_{n-1}$: \begin{eqnarray*} f_{i}^{+}\sigma(a_1,\dots,a_{n-1})&:=&\sigma(a_1,\dots,a_{i-1},1,a_i,\dots,a_{n-1}), \\ f_{i}^{-}\sigma(a_1,\dots,a_{n-1})&:=& \sigma(a_1,\dots,a_{i-1},0,a_i,\dots,a_{n-1}). \end{eqnarray*} For $n \geq 1$, a singular $n$-cube $\sigma$ is called \textit{degenerate} if $f_{i}^{+}\sigma = f_{i}^{-}\sigma$, for some $i \in [n]$. Otherwise, $\sigma$ is called \textit{non-degenerate}. By definition every $0$-cube is non-degenerate. For each $n\geq 0$, let ${\mathcal{D}}^{\normalfont {\textrm{Cube}}}_n(G)$ be the $R$-submodule of ${\mathcal{L}}^{\normalfont {\textrm{Cube}}}_n(G)$ that is generated by all degenerate singular $n$-cubes, and let ${\mathcal{C}}^{\normalfont {\textrm{Cube}}}_n(G)$ be the free $R$-module ${\mathcal{L}}^{\normalfont {\textrm{Cube}}}_n(G)/{\mathcal{D}}^{\normalfont {\textrm{Cube}}}_n(G)$, whose elements are called $n$-chains. Clearly, the cosets of non-degenerate $n$-cubes freely generate ${\mathcal{C}}^{\normalfont {\textrm{Cube}}}_n(G)$. Furthermore, for each $n\geq 1$, define the boundary operator \[ \partial_n^{\normalfont {\textrm{Cube}}} :{\mathcal{L}}^{\normalfont {\textrm{Cube}}}_n(G) \longrightarrow {\mathcal{L}}^{\normalfont {\textrm{Cube}}}_{n-1}(G) \] such that, for each singular $n$-cube $\sigma$, \[ \partial_n^{\normalfont {\textrm{Cube}}}(\sigma) = \sum_{i=1}^n (-1)^i \big( f_i^-\sigma - f_i^+\sigma \big) \] and extend linearly to all chains in ${\mathcal{L}}^{\normalfont {\textrm{Cube}}}_n(G)$. When there is no danger of confusion, we will abbreviate $\partial^{\normalfont {\textrm{Cube}}}_n$ as $\partial_n$. If one sets ${\mathcal{L}}^{\normalfont {\textrm{Cube}}}_{-1} (G) ={\mathcal{D}}^{\normalfont {\textrm{Cube}}}_{-1} (G)= (0)$ then one can define $\partial_0^{\normalfont {\textrm{Cube}}}$ as the trivial map from ${\mathcal{L}}_0^{\normalfont {\textrm{Cube}}}(G)$ to ${\mathcal{L}}_{-1}^{\normalfont {\textrm{Cube}}}(G)$. It is easy to check that, for $n\geq 0$, $\partial_n [{\mathcal{D}}^{\normalfont {\textrm{Cube}}}_n(G)] \subseteq {\mathcal{D}}^{\normalfont {\textrm{Cube}}}_{n-1}(G)$ and $\partial_{n} \partial_{n+1} \sigma= 0$ (see \cite{BCW}). Hence, $\partial_n: {\mathcal{C}}^{\normalfont {\textrm{Cube}}}_n(G) \longrightarrow {\mathcal{C}}^{\normalfont {\textrm{Cube}}}_{n-1}(G)$ is a boundary operator, and ${\mathcal{C}}^{\normalfont {\textrm{Cube}}}(G) = ({\mathcal{C}}_\bullet^{\normalfont {\textrm{Cube}}},\partial_\bullet)$ is a chain complex of free $R$-modules. \begin{Definition} For $n \geq 0$, denote by $\mathcal{H}^{\normalfont {\textrm{Cube}}}_n(G)$ the $n$\textsuperscript{th} homology group of the chain complex ${\mathcal{C}}^{\normalfont {\textrm{Cube}}}(G)$. In other words, $\mathcal{H}^{\normalfont {\textrm{Cube}}}_n(G) := \mathrm{Ker}\,\partial_n/\mathrm{Im}\,\partial_{n+1}$. \end{Definition} We represent singular $n$-cubes $\sigma: Q_n \to G$ by sequences of length $2^n$, where the $i$\textsuperscript{th} term is the value of $\sigma$ on the $i$\textsuperscript{th} vertex, and the vertices of $Q^n$ are indexed in colexicographic order. For example, if $G$ is defined as in \ref{fig2}, then the sequence $(1,2,2,1,2,3,3,2)$ represents the singular $3$-cube with labels as illustrated in \ref{fig1}. \begin{figure} \begin{center} \begin{tikzpicture}[scale=1.5] \Vertex[x=0 ,y=0,L=$000$]{1} \Vertex[x=2 ,y=0,L=$100$]{2} \Vertex[x=1 ,y=1,L=$010$]{3} \Vertex[x=3 ,y=1,L=$110$]{4} \Vertex[x=0 ,y=2,L=$001$]{5} \Vertex[x=2 ,y=2,L=$101$]{6} \Vertex[x=1 ,y=3,L=$011$]{7} \Vertex[x=3 ,y=3,L=$111$]{8} \Edge(1)(2) \Edge(1)(3) \Edge(2)(4) \Edge(3)(4) \Edge(5)(6)\Edge(5)(7)\Edge(6)(8)\Edge(7)(8) \Edge(1)(5)\Edge(2)(6)\Edge(3)(7)\Edge(4)(8) \node [color=blue] at ([shift={(-.3,.3)}]1) {1}; \node [color=blue] at ([shift={(-.3,.3)}]2) {2}; \node [color=blue] at ([shift={(-.3,.3)}]3) {2}; \node [color=blue] at ([shift={(-.3,.3)}]4) {1}; \node [color=blue] at ([shift={(-.3,.3)}]5) {2}; \node [color=blue] at ([shift={(-.3,.3)}]6) {3}; \node [color=blue] at ([shift={(-.3,.3)}]7) {3}; \node [color=blue] at ([shift={(-.3,.3)}]8) {2}; \end{tikzpicture} \end{center} \caption{Singular $3$-cube represented by $(1,2,2,1,2,3,3,2)$.} \label{fig1} \end{figure} We represent each coset in ${\mathcal{C}}_n^{\normalfont {\textrm{Cube}}}(G)$ by the unique coset representative in which all terms are non-degenerate. \begin{Example} Let $G$ be a $4$-cycle, with vertices labeled cyclically, as illustrated in \ref{fig2}. \begin{figure} \begin{center} \begin{tikzpicture}[scale=1.5] \Vertex[x=0 ,y=0]{1} \Vertex[x=1 ,y=0]{2} \Vertex[x=1 ,y=1]{3} \Vertex[x=0 ,y=1]{4} \Edge(1)(2) \Edge(2)(3) \Edge(3)(4) \Edge(4)(1) \end{tikzpicture} \caption{Graph $G =$ $4$-cycle.} \label{fig2} \end{center} \end{figure} Then \begin{gather*} {\mathcal{C}}_0^{\normalfont {\textrm{Cube}}} = \big\langle (1),(2),(3),(4) \big\rangle \\ {\mathcal{C}}_1^{\normalfont {\textrm{Cube}}} = \big\langle (1,2), (2,1), (2,3), (3,2), (3,4), (4,3), (4,1), (1,4) \big\rangle \\ {\mathcal{C}}_2^{\normalfont {\textrm{Cube}}} = \big\langle (1,1,1,2), (1,1,1,4), \dots<60 \textrm{ more}>\dots, (4,4,4,1), (4,4,4,3) \big\rangle. \end{gather*} The matrix of $\partial_1$ with respect to the above bases is the standard vertex-directed edge incidence matrix of the corresponding directed graph in which each edge is replaced by directed edges in both directions. An easy computation shows that $\partial_1$ has rank $|V|-1 = 4-1 = 3$ and nullity $8-3 = 5$. Cycles in ${\mathcal{C}}_1^{\normalfont {\textrm{Cube}}}(G)$ correspond to circulations in $G$, that is, weighted sums of edges in which the net flow out of each vertex equals zero. A basis of ${\mathcal{C}}_1^{\normalfont {\textrm{Cube}}}(G)$ may be obtained from any directed cycle basis of $G$, e.g., for the graph $G$ defined in \Ref{fig2} we may take $(1,2)+(2,1), (2,3)+(3,2), (3,4)+(4,3), (4,1)+(1,4)$, and $ (1,2)+(2,3)+(3,4)+(4,1)$. Each of these $1$-cycles is the boundary of a $2$-chain: \begin{gather*} (1,2)+(2,1) = \partial_2((2,1,2,2))\\ (2,3)+(3,2) = \partial_2((3,2,3,3))\\ (3,4)+(4,3) = \partial_2((4,3,4,4))\\ (4,1)+(1,4) = \partial_2((1,4,1,1))\\ (1,2)+(2,3)+(3,4)+(4,1) = \partial_2((1,2,4,3) + (3,4,3,3) + (1,4,1,1)). \end{gather*} Hence $\mathcal{H}_0^{\normalfont {\textrm{Cube}}}(G)= R$ and $\mathcal{H}_1^{\normalfont {\textrm{Cube}}}(G) = (0)$. By somewhat tedious computations one can also show that $\mathcal{H}_2^{\normalfont {\textrm{Cube}}}(G)=(0)$. Here we have $\mathrm{rank}\,({\mathcal{C}}_2^{\normalfont {\textrm{Cube}}}(G))=64$ and $\mathrm{rank}\,({\mathcal{C}}_3^{\normalfont {\textrm{Cube}}}(G))=2432$, and for higher dimensions the problem of computing $\mathcal{H}_n^{\normalfont {\textrm{Cube}}}(G)$ becomes increasingly more difficult. Fortunately, we are able to prove more general results (in \ref{sec-computations}) implying that $\mathcal{H}_n^{\normalfont {\textrm{Cube}}}(G)=(0)$ for all $n>0$, for the graph $G$ defined above in \ref{fig2}. \end{Example} \subsection*{Discrete Path Homology} In a series of papers \cite{GLMY1,GLMY2,GLMY3} a (co)homology and a homotopy theory for directed graphs are developed. In these theories, an undirected graph is interpreted as the directed graph, with each undirected edge replaced by two oppositely directed edges between its endpoints. It is shown in \cite[Thm. 4.22]{GLMY2} that the first homology group of a directed graph is the abelianization of its fundamental group, where both homology and homotopy groups are taken in the sense of \cite{GLMY1,GLMY2}. We now recall the homology theory from \cite{GLMY1}, confining ourselves to the setting of simple (undirected) graphs. Let $V$ be a finite set. For $n\ge 0$ we denote by ${\mathcal{L}}_n^{\normalfont {\textrm{Path}}}(V) $ the $R$-module freely generated by the set of all $(n+1)-$tuples $(v_0, \dots, v_n)$ of elements in $V$. For each $n$, let ${\mathcal{D}}_n^{\normalfont {\textrm{Path}}}(V)$ denote the submodule generated by the {\em degenerate} $n$-tuples $(v_0,\dots,v_n)$ where $v_i=v_{i+1}$ for some $i$. For $n\ge 1$, let $\partial_n^{\normalfont {\textrm{Path}}}:{\mathcal{L}}_n^{\normalfont {\textrm{Path}}}(V) \to {\mathcal{L}}_{n-1}^{\normalfont {\textrm{Path}}}(V) $ be defined by \[ (v_0,\dots,v_n) \longmapsto \sum_{i=0}^n (-1)^i (v_0,\dots,\widehat{v_i},\dots,v_n). \] If we set ${\mathcal{L}}_{-1}^{\normalfont {\textrm{Path}}}(V) = {\mathcal{D}}_{-1}^{\normalfont {\textrm{Path}}}(V) = (0)$ we can also define $\partial_{0}^{\normalfont {\textrm{Path}}}$ as the trivial map from ${\mathcal{L}}_0^{\normalfont {\textrm{Path}}}(V)$ to ${\mathcal{L}}_{-1}^{\normalfont {\textrm{Path}}}(V)$. Again, we will write $\partial_n^{\normalfont {\textrm{Path}}} = \partial_n$ when there is no ambiguity. For $n\ge 0$, it is easy to verify that $\partial_{n} \partial_{n+1} = 0$ and $\partial_n[{\mathcal{D}}_n^{\normalfont {\textrm{Path}}}(V)] \subseteq {\mathcal{D}}_{n-1}^{\normalfont {\textrm{Path}}}(V)$. Hence if we define a sequence of quotients \[ {\mathcal{C}}_n^{\normalfont {\textrm{Path}}}(V) = \frac{{\mathcal{L}}_n^{\normalfont {\textrm{Path}}}(V) }{{\mathcal{D}}_n^{\normalfont {\textrm{Path}}}(V) }, \; n = -1,0,1,\dots, \] then ${\mathcal{C}}^{\normalfont {\textrm{Path}}} (V) = ({\mathcal{C}}^{\normalfont {\textrm{Path}}}_\bullet(V), \partial_\bullet^{\normalfont {\textrm{Path}}})$ forms a chain complex. Now let $G=(V,E)$ be a simple graph. Define ${\mathcal{L}}_n^{\normalfont {\textrm{Path}}}(G) \subseteq {\mathcal{L}}_n^{\normalfont {\textrm{Path}}}(V) $ to be the submodule of ${\mathcal{L}}_n^{\normalfont {\textrm{Path}}}(V) $ spanned by all $(v_0,\dots,v_n)$ such that $\{v_i, v_{i+1}\} \in E(G)$ for all $i<n$. Thus, ${\mathcal{L}}_n^{\normalfont {\textrm{Path}}}(G) = {\mathcal{L}}_n^{\normalfont {\textrm{Path}}}(V) $ when $G$ is the complete graph on vertex set $V$. For all $n \geq 0$, define $\widetilde{{\mathcal{C}}}_n^{\normalfont {\textrm{Path}}}(G) \subseteq {\mathcal{C}}_n^{\normalfont {\textrm{Path}}}(V)$ to be the submodule of ${\mathcal{C}}_n^{\normalfont {\textrm{Path}}}(V)$ generated by cosets of the form \[ (v_0, \dots, v_n) + {\mathcal{D}}_n^{\normalfont {\textrm{Path}}} (V), \] where $(v_0,\dots,v_n) \in{\mathcal{L}}_n^{\normalfont {\textrm{Path}}}(G)$, and set $\widetilde{{\mathcal{C}}}_{-1}^{\normalfont {\textrm{Path}}}(G) = (0)$. The sequence $\{\widetilde{\mathcal{C}}_n^{\normalfont {\textrm{Path}}}(G) \}_{n\ge -1}$ is not always a chain complex, since boundaries of paths $(v_0,\dots,v_n) \in {\mathcal{L}}_n^{\normalfont {\textrm{Path}}}(G)$ may contain terms that are not paths in $G$. However, if we define, for $n\ge 0$, \[ {\mathcal{C}}_n^{\normalfont {\textrm{Path}}}(G) = \partial_n^{-1}[\widetilde{{\mathcal{C}}}_{n-1}^{\normalfont {\textrm{Path}}}(G)] \] and ${\mathcal{C}}_{-1}^{\normalfont {\textrm{Path}}}(G) = (0)$, then $\partial_{n}\partial_{n+1} = 0$ immediately implies that $\partial_n[{\mathcal{C}}_n^{\normalfont {\textrm{Path}}}(G)] \subseteq {\mathcal{C}}_{n-1}^{\normalfont {\textrm{Path}}}(G)$, and ${\mathcal{C}}^{\normalfont {\textrm{Path}}}(G)=({\mathcal{C}}_\bullet^{\normalfont {\textrm{Path}}},\partial_\bullet)$ is a chain complex. \begin{Definition} For $n \geq 0$, denote by $\mathcal{H}^{\normalfont {\textrm{Path}}}_n(G)$ the $n$\textsuperscript{th} homology group of the chain complex ${\mathcal{C}}^{\normalfont {\textrm{Path}}}(G)$. In other words, $\mathcal{H}^{\normalfont {\textrm{Path}}}_n(G) := \mathrm{Ker}\,\partial_n/\mathrm{Im}\,\partial_{n+1}$. \end{Definition} We again identify cosets in ${\mathcal{C}}_n^{\normalfont {\textrm{Path}}}(G)$ with their unique representatives whose terms are all non-degenerate. Using this notation, if $G$ is the $4$-cycle graph in \ref{fig2}, then \begin{gather*} {\mathcal{C}}_0^{\normalfont {\textrm{Path}}} = \big\langle (1),(2),(3),(4) \big\rangle \\ {\mathcal{C}}_1^{\normalfont {\textrm{Path}}} = \big\langle (1,2), (2,1), (2,3), (3,2), (3,4), (4,3), (4,1), (1,4) \big\rangle \\ {\mathcal{C}}_2^{\normalfont {\textrm{Path}}} = \big\langle (1,2,1), (2,1,2), (2,3,2), (3,2,3), (3,4,3), (4,3,4), (4,1,4), (1,4,1), \qquad\qquad \qquad\\\ \qquad\qquad\qquad(1,2,3)-(1,4,3), (2,3,4)-(2,1,4), (3,4,1)-(3,2,1), (4,1,2) - (4,3,2) \big\rangle. \end{gather*} Represented in this notation, the chain groups ${\mathcal{C}}_0^{\normalfont {\textrm{Path}}}(G)$ and ${\mathcal{C}}_1^{\normalfont {\textrm{Path}}}(G)$ are identical to ${\mathcal{C}}_0^{\normalfont {\textrm{Cube}}}(G)$ and ${\mathcal{C}}_1^{\normalfont {\textrm{Cube}}}(G)$. The boundary map $\partial_1$ again has rank $5$ and its kernel is spanned by $(1,2)+(2,1), (2,3)+(3,2), (3,4)+(4,3), (4,1)+(1,4)$, and $(1,2)+(2,3)+(3,4)+(4,1)$. As before, each of these $1$-cycles is the boundary of a $2$-chain: \begin{gather*} (1,2)+(2,1) = \partial_2((1,2,1))\\ (2,3)+(3,2) = \partial_2((2,3,2))\\(3,4)+(4,3) = \partial_2((3,4,3))\\ (4,1)+(1,4) = \partial_2((1,4,1))\\ (1,2)+(2,3)+(3,4)+(4,1) = \partial_2(((1,2,3)-(1,4,3))+(3,4,3)+(1,4,1)). \end{gather*} It follows that $\mathcal{H}_0^{\normalfont {\textrm{Path}}}(G)= R$ and $\mathcal{H}_1^{\normalfont {\textrm{Path}}}(G)=(0)$. Again it is possible to prove directly that $\mathcal{H}_2^{\normalfont {\textrm{Path}}}(G) = (0)$, but more general results in \ref{sec-computations} will show that, for this example, $\mathcal{H}_n^{\normalfont {\textrm{Path}}}(G)=(0)$ for all $n > 0$. \subsection*{Classical Homology of a Graph and its Clique Complex} We mention two other homology theories of graphs that have a substantial presence in the literature. Given any undirected graph $G$, we may regard $G$ as a $1$-dimensional simplicial complex and compute its singular (or equivalently, simplicial) homology $\mathcal{H}^\textrm{Sing}_\bullet (G)$. It is elementary and classical (e.g., \cite{Massey91}, Chapter 8) that if $G$ is connected, then $\mathcal{H}^\textrm{Sing}_0(G) \cong R$, $\mathcal{H}^\textrm{Sing}_1(G) \cong R^{|E(G)|-|V(G)|+1}$, and $\mathcal{H}^\textrm{Sing}_n(G)\cong(0)$ for $n>1$. Given $G$, we may also construct the clique complex $K_G$ of $G$ (also called the flag complex of $G$; see \cite{St}), whose faces are the subsets of $V(G)$ forming cliques, and compute the simplicial (or equivalently, singular) homology $\mathcal{H}_\bullet^{\textrm{Clique}}(G)$ of $K_G$. If $N=\omega(G)$ is the size of the largest clique in $G$, then $\mathcal{H}_n^\textrm{Clique}(G) \cong (0)$ for $n>N$, but $\mathcal{H}_n^\textrm{Clique}(G)$ can be nonzero for any $n\le N$. If $G$ is a $4$-cycle as in \ref{fig2}, then $\mathcal{H}^\textrm{Sing}_n(G)$ and $\mathcal{H}^\textrm{Clique}_n(G)$ are isomorphic for all $n$, but this is not true in general (for example, when $G$ is a $3$-cycle). We note that a theory analogous to $\mathcal{H}^{\normalfont {\textrm{Cube}}}_\bullet(G)$ can be defined by considering chain groups spanned in dimension $n$ by graph maps from the complete graph on $n+1$ vertices to $G$ and differential given by the alternating sum over the restrictions to complete subgraphs on $n-1$ vertices. This theory can be seen to be equivalent to $\mathcal{H}_\bullet^\textrm{Clique}(G)$ (see \cite[p. 76]{Mu}). \subsection*{Relationships } This paper will explore connections between the two homology theories $\mathcal{H}_\bullet^{\normalfont {\textrm{Cube}}}(G)$ and $\mathcal{H}_\bullet^{\normalfont {\textrm{Path}}}(G)$ defined above. For many classes of graphs $G$ we have, $\mathcal{H}^{\normalfont {\textrm{Cube}}}_\bullet(G) \cong \mathcal{H}^{\normalfont {\textrm{Path}}}_\bullet(G)$, and we will give several more examples of this phenomenon (see especially \ref{sec-computations}). In \ref{sec-map} we define a homomorphism from ${\mathcal{C}}^{\normalfont {\textrm{Cube}}}(G)$ to ${\mathcal{C}}^{\normalfont {\textrm{Path}}}(G)$ that may explain some of these connections. However, $\mathcal{H}^{\normalfont {\textrm{Cube}}}_\bullet $ and $\mathcal{H}^{\normalfont {\textrm{Path}}}_\bullet$ are not isomorphic in general, and we give an example illustrating this in \ref{sec-counter}. Connections with $\mathcal{H}^{\normalfont {\textrm{Sing}}}_\bullet$ and $ \mathcal{H}^{\normalfont {\textrm{Clique}}}_\bullet$ seem to be less close; for example, when $G$ is a $4$-cycle, the discrete cubical and path homologies are trivial in dimension $1$, but the singular and clique homologies are nontrivial. A combination of results in \cite{BBLL} and \cite{BCW} proves that for any graph $G$, $\mathcal{H}_1^{\normalfont {\textrm{Cube}}}(G) \cong \mathcal{H}_1^{\normalfont {\textrm{Sing}}}(K)$, where $K$ is the CW-complex obtained from $G$ by ``filling in'' all of its triangles and quadrilaterals with $2$-cells. A similar construction in higher dimensions is conjectured in \cite{BBLL} to give the correct higher homotopy groups, and the authors have proposed (private communication) that this might also produce a CW-complex $K^*$ such that $\mathcal{H}_n^{\normalfont {\textrm{Cube}}}(G) \cong \mathcal{H}_n^{\normalfont {\textrm{Sing}}}(K^*)$ when $n>1$. Note that $K^*$ has cells of arbitrary high dimension for most graphs. \section{Homotopy Equivalence Preserves Homology \label{sec-homotopyhomology}} This section describes the connection between the graph homotopy theory introduced in \cite{BBLL} and \cite{GLMY2} and the cubical and path homologies introduced in \cite{BCW} and \cite{GLMY2}. First we recall several basic definitions. \begin{Definition}\label{boxdef} (See \cite{Hammack11}) If $G$ and $H$ are graphs, the {\em Cartesian (or box) product} $G\,\Box\, H$ is the graph whose vertex set is the Cartesian product set $V(G)\times V(H)$, and whose edges are pairs $\{(g_1,h_1),(g_2,h_2)\}$ such that either $g_1=g_2$ and $\{h_1,h_2\}\in E(H)$ or $h_1=h_2$ and $\{g_1,g_2\}\in E(G)$. \end{Definition} \begin{Definition}\label{homotopy-def} Suppose that $G$ and $H$ are graphs, and $f$ and $g$ are graph homomorphisms from $G$ to $H$. Then $f$ and $g$ are {\em homotopic} if there exists a graph homomorphism $\Phi$ from $G\,\Box\, I_m$ to $H$, where $I_m$ denotes the $m$-path with vertex set $\{0,1,\dots,m\}$, such that $\Phi(\bullet,0)=f$ and $\Phi(\bullet,m)=g$. \end{Definition} \begin{Definition} Two simple undirected graphs $G$ and $H$ are {\em homotopy equivalent} if there exist graph homomorphisms $\phi: G\to H$ and $\theta: H \to G$ such that $\theta \phi$ is homotopic to $id_G$ and $\phi \theta$ is homotopic to $id_H$. Here $id_G$ and $id_H$ denote the identity maps on $G$ and $H$, respectively. \end{Definition} The connection between the discrete homotopy theory in \cite{BBLL} and \cite{GLMY2} and the homology theories introduced in \cite{BCW} and \cite{GLMY2} is expressed by the following theorem, which also provides a key computational tool. \begin{Theorem}\label{equiv-thm} Let $G$ and $H$ be simple, undirected graphs. If $G$ and $H$ are homotopically equivalent, then, for all $n\ge 0$, \begin{enumerate} \item[(i)] $\mathcal{H}^{\normalfont {\textrm{Cube}}}_n(G)\cong\mathcal{H}^{\normalfont {\textrm{Cube}}}_n(H)$, and \item[(ii)] $\mathcal{H}^{\normalfont {\textrm{Path}}}_n(G)\cong\mathcal{H}^{\normalfont {\textrm{Path}}}_n(H)$. \end{enumerate} \end{Theorem} For both parts of \Ref{equiv-thm} it suffices to prove that if $\alpha$ and $\beta$ are homotopically equivalent maps from $G$ to $H$, then $\alpha$ and $\beta$ induce identical maps on homology. For cubical homology, this result is contained in Theorem 3.8(1) of \cite{BCW}, where it is proved for any discrete metric space. Since details of that argument do not appear in \cite{BCW} we provide them here (in the case of graphs) for completeness. For path homology, \Ref{equiv-thm}(ii) is stated and proved explicitly in \cite{GLMY2} (Theorem 3.3(ii)), for directed graphs. We include a sketch of that argument in the undirected case. Both proofs employ a chain homotopy argument, and are structurally similar. \begin{proof}[Proof of \Ref{equiv-thm}(i)] Suppose that $\alpha$ and $\beta$ are homomorphisms from $G$ to $H$, and $\Phi$ is a homotopy from $\alpha$ to $\beta$ with $\Phi(x,0)=\alpha(x)$ and $\Phi(x,m) = \beta(x)$ for all $x\in V(G)$, as in \ref{homotopy-def}. If $\sigma\in {\mathcal{C}}^{\normalfont {\textrm{Cube}}}_n(G)$, let $\Phi(\sigma,j)$ denote the map defined by $ \Phi(\sigma,j)(q) = \Phi(\sigma(q),j) $ for all $q\in Q_n$, and define $\tilde{\alpha}_n, \tilde{\beta}_n: {\mathcal{C}}_n^{\normalfont {\textrm{Cube}}}(G)\to {\mathcal{C}}_n^{\normalfont {\textrm{Cube}}}(H)$ by \[ \tilde{\alpha}_n(\sigma) = \Phi(\sigma,0), \quad \tilde{\beta}_n(\sigma) = \Phi(\sigma,m). \] It is straightforward to show that $\tilde{\alpha}$ and $\tilde{\beta}$ are chain maps, i.e., $\tilde{\alpha}_{n-1} \partial_n = \partial_n \tilde{\alpha}_n$ and similarly for $\tilde{\beta}$. We will construct a sequence of maps $h_n: {\mathcal{C}}_n^{\normalfont {\textrm{Cube}}}(G)\to {\mathcal{C}}_{n+1}^{\normalfont {\textrm{Cube}}}(H)$ such that \begin{equation}\label{chomotopy1} \tilde{\beta}_n - \tilde{\alpha}_n = \partial_{n+1} h_n + h_{n-1} \partial_n, \end{equation} for all $n$. In other words, the sequence $\{h_n\}$ defines a chain homotopy between $\{\tilde{\alpha}_n\}$ and $\{\tilde{\beta}_n\}$. It follows that if $z\in {\mathcal{C}}_n^{\normalfont {\textrm{Cube}}}(G)$ is a cycle, then \[ \tilde{\beta}_n(z) - \tilde{\alpha}_n(z) = \partial_{n+1} h_n (z). \] In particular, $\tilde{\beta}_n(z) - \tilde{\alpha}_n(z) \in \mathrm{Im}\,\partial_{n+1}$ and hence $\tilde{\alpha}_n(z)$ and $\tilde{\beta}_n(z)$ lie in the same homology class for all $z$, implying that $\alpha$ and $\beta$ induce the same maps on homology. Given a singular $n$-cube $\sigma \in {\mathcal{C}}_n^{\normalfont {\textrm{Cube}}}(G)$, the map $h_n(\sigma) \in {\mathcal{C}}_{n+1}^{\normalfont {\textrm{Cube}}}(H)$ is constructed as follows. For $j=1,\dots,m$, let $h_n^{(j)}(\sigma)\in{\mathcal{C}}_{n+1}^{\normalfont {\textrm{Cube}}}(H)$ be the unique labeled $(n+1)$-cube such that \begin{gather} f^+_1 h_n^{(j)}(\sigma) (q) = \Phi(\sigma(q),j)\nonumber \\ f^-_1 h_n^{(j)}(\sigma) (q) = \Phi(\sigma(q),j-1) \nonumber, \end{gather} for all $q\in Q_n$. Finally, define \begin{equation} h_n(\sigma) = h_n^{(1)}(\sigma) + \cdots + h_n^{(m)}(\sigma).\nonumber \end{equation} It is immediate from the definition of $h_n^{(j)}$ that \begin{gather} f^+_1 h_n^{(m)}(\sigma) (q) =\beta(\sigma(q)) \nonumber\\ f^-_1 h_n^{(1)}(\sigma) (q) = \alpha(\sigma(q)), \nonumber \end{gather} for all $q\in Q_n$. A few moments of reflection show that for $i=2,\dots, n$, we have \begin{equation}\label{equiv-identity2} f^\epsilon_{i} (h_n^{(j)}(\sigma))=h_{n-1}^{(j)}(f_{i-1}^\epsilon \sigma) \end{equation} for $j \in [m]$ and $\epsilon \in \{-,+\}$. Computing the right hand side of \Ref{chomotopy1}, we get \begin{eqnarray}\nonumber h_{n-1}^{(j)}(\partial_n(\sigma) )&=& h_{n-1}^{(j)}\bigg(\sum_{i=1}^n (-1)^{i} (f^-_i \sigma - f^+_i \sigma)\bigg)\\ &=& \sum_{i=1}^n (-1)^{i} \big(h_{n-1}^{(j)}(f^-_i \sigma) - h_{n-1}^{(j)}(f^+_i \sigma)\big)\label{equiv-RHSsecond2} \end{eqnarray} and \begin{eqnarray}\label{equiv-RHSfirst} \partial_{n+1} (h_n^{(j)}(\sigma)) &=&\sum_{i=1}^{n+1} (-1)^{i}\big( f^-_i (h_n^{(j)}(\sigma)) - f^+_i (h_n^{(j)}(\sigma)) \big). \end{eqnarray} It follows from \Ref{equiv-identity2} that terms $i=1,\dots,n$ in \Ref{equiv-RHSsecond2} are identical to terms $i=2,\dots, n+1$ in \Ref{equiv-RHSfirst}, but have opposite signs. Hence they cancel, leaving only the first term in \Ref{equiv-RHSfirst}, and we obtain \begin{eqnarray} \partial_{n+1}h_n^{(j)}(\sigma) + h_{n-1}^{(j)}\partial_n(\sigma) & =& f_1^+(h_n^{(j)}(\sigma)) - f_1^-(h_n^{(j)}(\sigma)) \nonumber\\ \label{finalidentity} &=& \Phi(\sigma,j) - \Phi(\sigma,j-1). \end{eqnarray} Summing \ref{finalidentity} over $j$ gives \begin{eqnarray} \partial_{n+1}h_n(\sigma) + h_{n-1}\partial_n(\sigma) & =& \sum_{j=1}^m \Phi(\sigma,j) - \Phi(\sigma,j-1) \nonumber \\ &=& \Phi(\sigma,m) - \Phi(\sigma,0) \nonumber\\ &=& \tilde{\beta}(\sigma) - \tilde{\alpha}(\sigma) \label{finalstep} \end{eqnarray} as desired, and \Ref{chomotopy1} is proved. \end{proof} \begin{proof}[Proof of \ref{equiv-thm}(ii)] The proof in \cite{GLMY2} has essentially the same structure as the proof of part (i) given above. We will sketch the argument, using similar notation but focusing on the important differences. Again assume that $\alpha, \beta: G\to H$ are graph homomorphisms, with a homotopy $\Phi$ such that $\Phi(x,0) = \alpha(x)$ and $\Phi(x,m) = \beta(x)$ for all $x\in V(G)$. It is shown in \cite[Theorem 2.10]{GLMY2} that $\alpha$ and $\beta$ induce chain maps $\tilde{\alpha}_n$ and $\tilde{\beta}_n$ from ${\mathcal{C}}_n^{\normalfont {\textrm{Path}}}(G)$ to ${\mathcal{C}}_n^{\normalfont {\textrm{Path}}}(H)$. As before, the key step in the present proof is to construct a chain homotopy between the sequences $\{\tilde{\alpha}_n\}$ and $\{\tilde{\beta}_n\}$. For $\sigma = (v_0,v_1,\dots,v_n) \in {\mathcal{C}}_n^{\normalfont {\textrm{Path}}}(G)$ and $j \in [m]$, define $h_n^{(j)}(\sigma) \in {\mathcal{C}}^{\normalfont {\textrm{Path}}}_{n+1}(H)$ as follows: \[ h_n^{(j)}(\sigma) = \sum_{k=0}^n (-1)^k (\Phi(v_0,j-1),\dots, \Phi(v_k,j-1),\Phi(v_k,j),\dots,\Phi(v_n,j)), \] and define \[ h_n(\sigma) = h_n^{(1)}(\sigma) + \cdots + h_n^{(m)}(\sigma). \] \smallskip \noindent At this point it is essential to check that $h_n^{(j)}(\sigma) \in {\mathcal{C}}^{\normalfont {\textrm{Path}}}_{n+1}(H)$ for all $j$, since not every linear combination of elements of $\tilde{{\mathcal{C}}}^{\normalfont {\textrm{Path}}}_{n+1}(H)$ is an element of ${\mathcal{C}}^{\normalfont {\textrm{Path}}}_{n+1}(H)$. An argument proving this fact can be found in \cite[Proposition 2.12]{GLMY2}, and is omitted here. The proof is completed by showing that identity \Ref{finalidentity} holds for the maps $h_n$ just defined, exactly as it did in part (i). This argument is technical but straightforward, and is omitted here. With \Ref{finalidentity} in hand, \Ref{finalstep} follows, and we are done. \end{proof} \section{Computations of Homology Groups}\label{sec-computations} With \ref{equiv-thm} in hand, we have tools that will allow us to compute $\mathcal{H}_\bullet^{\normalfont {\textrm{Cube}}}(G)$ and $\mathcal{H}_\bullet^{\normalfont {\textrm{Path}}}(G) $ for large classes of graphs. We give many examples in this section. Most involve {\em deformation retraction}, a special kind of homotopy equivalence that is frequently easy to recognize. \begin{Definition}\label{retractx-def} Let $G$ be a graph, and let $H$ be an induced subgraph of $G$. That is, $V(H) \subseteq V(G)$ and $E(H)$ consists of all edges in $E(G)$ for which both endpoints belong to $V(H).$ \begin{enumerate} \item A {\em retraction of $G$ onto $H$} is a graph homomorphism $r:G\to H$ such that $r(y) = y$ for all $y\in V(H)$. \item A {\em deformation retraction of $G$ onto $H$} is a retraction $r:G \to H$ such that $ir$ is homotopic to $id_G$, where $i$ denotes the inclusion map from $H$ to $G$. \item A {\em one-step deformation retraction from $G$ to $H$} is a deformation retraction $r$ for which $m=1$ in the homotopy (\Ref{homotopy-def}) between $ri$ and $id_G.$ Equivalently, $r$ is a retraction such that $\{x,r(x)\}$ is an edge for all $x\in V(G)$. \end{enumerate} \end{Definition} If $r$ is a deformation retraction from $G$ to $H$, then, since $ri= id_H$, the following lemma is an immediate consequence of $\Ref{equiv-thm}$. \begin{Lemma} If $r$ is a deformation retraction from $G$ onto a subgraph $H$, and $i$ denotes the inclusion map from $H$ to $G$, then $r$ and $i$ define a homotopy equivalence between $G$ and $H$. Consequently, $\mathcal{H}^{\normalfont {\textrm{Cube}}}_n(G) \cong \mathcal{H}^{\normalfont {\textrm{Cube}}}_n(H)$ and $\mathcal{H}^{\normalfont {\textrm{Path}}}_n(G) \cong \mathcal{H}^{\normalfont {\textrm{Path}}}_n(H)$, for all $n \geq 0$. \end{Lemma} \smallskip This result immediately gives several infinite classes of graphs for which the cubical and path (reduced) homology is trivial in all dimensions. \begin{Corollary}\label{cor-examples} If $G$ is a tree, or a complete graph, or a hypercube, then $\mathcal{H}_n^{\normalfont {\textrm{Cube}}}(G)\cong\mathcal{H}_n^{\normalfont {\textrm{Path}}}(G)\cong(0) $ for all $n>0$. \end{Corollary} \begin{proof}If $G$ is a tree and $x\in V(G)$ is a leaf connected to a unique vertex $y$, then the map $r: V(G)\to V(G)\backslash\{x\}$ defined by \begin{equation}\label{retract-def} r(v) = \begin{cases} v & v \ne x, \\ y & v = x, \end{cases} \end{equation} is a one-step deformation retraction from $G$ onto the subgraph $G\backslash x$. If $G$ is a complete graph, $x\in V(G)$ and $y\ne x$ is any other vertex, then \ref{retract-def} again defines a one-step deformation retraction from $G$ to $G\backslash x$. If $G$ is a hypercube of dimension $n$, then the map $r$ defined by collapsing any facet onto its opposite facet is a one-step deformation retraction onto a hypercube of dimension $n-1$. In all three cases, the process can be repeated, eventually showing that the homology (both cubical and path) is the same as that of a graph with a single vertex. \end{proof} These arguments can be extended to a larger class of examples: \begin{Theorem} Let $G$ be a graph, and $K_1$ and $K_2$ are induced nonempty subgraphs of $G$ such that $V(G) = V(K_1)\cup V(K_2)$ and $V(K_1)\cap V(K_2) = \emptyset$. Suppose there exist vertices $a\in V(K_1)$ and $b\in V(K_2)$ such that $\{a,b\}\in E(G)$, every vertex in $K_1$ is connected to $b$, and every vertex in $K_2$ is connected to $a$. Then $\mathcal{H}_n^{\normalfont {\textrm{Path}}}(G) \cong \mathcal{H}_n^{\normalfont {\textrm{Cube}}}(G) \cong (0)$ for $n> 0$. \end{Theorem} \begin{proof} Let $H$ be the subgraph of $G$ with vertices $a$ and $b$ and the single edge $\{a,b\}$. Define $r: V(G)\to V(H)$ by \[ r(x) = \begin{cases} a & \textrm{if $x \in K_2-\{b\}$},\\ b & \textrm{if $x \in K_1-\{a\}$},\\ x & \textrm{if $x \in H$.} \end{cases} \] An easy argument shows that $r$ is a one-step deformation of $G$ onto $H$, and since $H$ has trivial reduced homology in both the cubical and path case, the result follows from \ref{equiv-thm}. \end{proof} \begin{Corollary} For all $s,t >0$, let $K_{s,t}$ denote a complete bipartite graph with $s+t$ vertices. Then $\mathcal{H}_n^{\normalfont {\textrm{Path}}}(K_{s,t}) \cong \mathcal{H}_n^{\normalfont {\textrm{Cube}}}(K_{s,t}) \cong (0)$ for $n>0$. \hfill \qed \end{Corollary} \begin{Corollary} \label{cor-join}Let $K_1$ and $K_2$ be graphs with disjoint vertex sets. Consider the join graph $G = K_1 \ast K_2,$ where $V(G) = V(K_1) \cup V(K_2)$ and $E(G)$ consists of $E(K_1)$ and $E(K_2)$ together with all edges $\{p,q\}$ connecting a vertex $p\in V(K_1)$ with a vertex $q \in V(K_2)$. Then $\mathcal{H}_n^{\normalfont {\textrm{Path}}}(K_1\ast K_2) \cong \mathcal{H}_n^{\normalfont {\textrm{Cube}}}(K_1\ast K_2) \cong (0)$ for $n>0$. \hfill \qed \end{Corollary} The last corollary includes two elementary but important examples, the {\em cone} $G*\{p\}$ of $G$ over $p$, and the {\em suspension} $G*\{p,q\}$ of $G$ over a pair of non-adjacent vertices $p$ and $q.$ \Ref{cor-join} shows that the reduced cubical and path homologies in both cases are trivial. \begin{Definition} The {\em disjoint sum} $K_1\oplus K_2$ of graphs $K_1$ and $K_2$ is the graph with vertex set $V(K_1\oplus K_2) = V(K_1)\cup V(K_2)$ and edge set $E(K_1\oplus K_2) = E(K_1)\cup E(K_2)$. \end{Definition} \begin{Theorem}\label{disjointsum} For any graphs $K_1$ and $K_2$, $\mathcal{H}_n^{\normalfont {\textrm{Cube}}}(K_1\oplus K_2) \cong \mathcal{H}_n^{\normalfont {\textrm{Cube}}}(K_1)\oplus \mathcal{H}_n^{\normalfont {\textrm{Cube}}}(K_2)$ and $\mathcal{H}_n^{\normalfont {\textrm{Path}}}(K_1\oplus K_2) \cong \mathcal{H}_n^{\normalfont {\textrm{Path}}}(K_1)\oplus \mathcal{H}_n^{\normalfont {\textrm{Path}}}(K_2)$ for all $n\ge 0$. \end{Theorem} \begin{proof} The proof is elementary in both cases. \end{proof} \begin{Definition}\label{def-chordal} A graph $G$ is {\em chordal} if every cycle of length greater than three contains a chord. Equivalently, $G$ is chordal if and only if there exists an ordering of its vertices $v_1,\dots,v_m$ such that for each $j>1$, the set of vertices $v_k$ adjacent to $v_j$ with $k<j$ form a clique (possibly empty). \end{Definition} \begin{Theorem} If $G$ is a chordal graph, then $\mathcal{H}_n^{\normalfont {\textrm{Path}}}(G) \cong \mathcal{H}_n^{\normalfont {\textrm{Cube}}}(G) \cong (0)$ for $n>0$. \end{Theorem} \begin{proof} Suppose that $n>0$ and $v_1,\dots,v_m$ is an ordering of $V(G)$ satisfying the condition of \Ref{def-chordal}. For $j \in [m]$, let $G^{(j)}$ denote the induced subgraph of $G$ whose vertex set is $\{v_1,\dots,v_j\}$. Proceeding by induction, suppose that $\mathcal{H}_n^{\normalfont {\textrm{Path}}}(G^{(j)}) \cong \mathcal{H}_n^{\normalfont {\textrm{Cube}}}(G^{(j)}) \cong (0)$. If $v_{j+1}$ has no neighbors in $G^{(j)}$, it follows from \Ref{disjointsum} that $\mathcal{H}_n^{\normalfont {\textrm{Path}}}(G^{(j+1)}) \cong \mathcal{H}_n^{\normalfont {\textrm{Cube}}}(G^{(j+1)}) \cong (0)$. Otherwise, suppose that $v_{j+1}$ has neighbors in $G^{(j)}$ and let $v_k$ with $k<j+1$ be one of them. It is easy to check that the map from $G^{(j+1)}$ to $G^{(j)}$ defined by sending $ v_{j+1}$ to $v_k$ and fixing the remaining elements of $G^{(j)}$ is a $1$-step deformation retraction. Hence $G^{(j+1)}$ has trivial homology, by \Ref{equiv-thm}. \end{proof} The next theorem shows how the homology theories $\mathcal{H}^{\normalfont {\textrm{Cube}}}_\bullet$ and $\mathcal{H}_\bullet^{\normalfont {\textrm{Path}}}$ behave with respect to three well-known types of graph products. One of these, the {\em box product} $G\,\Box\, H$ has already been defined in \Ref{boxdef}. The next definition introduces two more. For a more complete treatment of these constructions, see \cite{Hammack11}. \begin{Definition} Suppose that $G$ and $K$ are graphs. Define the {\em strong product} $G\boxtimes K$ and the {\em lexicographic product} $G[K]$ as graphs whose vertex set is the Cartesian product set $V(G)\times V(K)$, and whose edges are pairs $\{(g_1,k_1), (g_2,k_2)\}$ defined by the following rules: \begin{enumerate} \item $(g_1,k_1) \sim_\boxtimes (g_2,k_2) \textrm{ iff } ((g_1=g_2)\wedge (k_1 \sim k_2))\vee ((g_1 \sim g_2) \wedge (k_1=k_2))\vee ((g_1 \sim g_2) \wedge (k_1\sim k_2))$ \item $(g_1,k_1) \sim_\textrm{lex} (g_2,k_2)\textrm{ iff } ((g_1\sim g_2))\vee ( (g_1 = g_2) \wedge (k_1\sim k_2))$. \end{enumerate} \end{Definition} \begin{Theorem}\label{retract-products} Suppose that $G$ and $K$ are graphs, and $H$ is an induced subgraph of $K$. Suppose that $r: V(K)\to V(H)$ is a retraction of $K$ onto $H$. Then for all $n\ge 0$, \begin{enumerate} \item $\mathcal{H}_n^{\normalfont {\textrm{Cube}}}(G\,\Box\, K) \cong \mathcal{H}_n^{\normalfont {\textrm{Cube}}}(G\,\Box\, H)$ \textrm{ and } $\mathcal{H}_n^{\normalfont {\textrm{Path}}}(G\,\Box \, K) \cong \mathcal{H}_n^{\normalfont {\textrm{Path}}}(G\, \Box\, H)$, \item $\mathcal{H}_n^{\normalfont {\textrm{Cube}}}(G\boxtimes K) \cong \mathcal{H}_n^{\normalfont {\textrm{Cube}}}(G \boxtimes H)$ \textrm{ and } $\mathcal{H}_n^{\normalfont {\textrm{Path}}}(G\boxtimes K) \cong \mathcal{H}_n^{\normalfont {\textrm{Path}}}(G \boxtimes H)$, \item $\mathcal{H}_n^{\normalfont {\textrm{Cube}}}(G[K]) \cong \mathcal{H}_n^{\normalfont {\textrm{Cube}}}(G[H])$ \textrm{ and } $\mathcal{H}_n^{\normalfont {\textrm{Path}}}(G[K]) \cong \mathcal{H}_n^{\normalfont {\textrm{Path}}}(G[H])$. \end{enumerate} \end{Theorem} \begin{proof} It suffices to prove that the map \[ (g,k) \longmapsto (g,r(k)) \] defines a retraction from $G\,\Box\, H$ to $G\, \Box\, K$, from $G\boxtimes H$ to $G \boxtimes K$, and from $G[K]$ to $G[H]$. The arguments in each case are straightforward. \end{proof} \section{A map between the chain complexes ${\mathcal{C}}^{\normalfont {\textrm{Cube}}}(G)$ and ${\mathcal{C}}^{\normalfont {\textrm{Path}}}(G)$} \label{sec-map} In this section we establish a map between the chain complexes ${\mathcal{C}}^{\normalfont {\textrm{Cube}}}(G)$ and ${\mathcal{C}}^{\normalfont {\textrm{Path}}}(G)$. Consider a singular $n$-cube $\sigma : Q_n \rightarrow G$. In order to define a map from ${\mathcal{C}}^{\normalfont {\textrm{Cube}}}_n(G)$ to ${\mathcal{C}}_n^{\normalfont {\textrm{Path}}}(G)$, we first associate to any permutation $\tau \in S_n$ a path $p_\tau$ from $(0,\ldots,0) \in V(Q_n)$ to $(1,\ldots,1) \in V(Q_n)$. The path $p_\tau$ is defined as the path of length $n-1$ which in its $i$\textsuperscript{th} step flips the $\tau(i)$\textsuperscript{th} coordinate from $0$ to $1$. We write $p_\tau(i)$ for the $i$\textsuperscript{th} vertex in the path $p_\tau$, $0 \leq i \leq n$. To $\sigma \in {\mathcal{C}}_n^{\normalfont {\textrm{Cube}}}(G)$, we assign the element $$\psi(\sigma) := \sum_{\tau \in S_n} \mathrm{sign}(\tau) \, \sigma \circ p_\tau$$ \noindent of ${\mathcal{C}}_n^{\normalfont {\textrm{Path}}}(G)$, where $\sigma \circ p_\tau$ denotes the path in $G$ whose $i$\textsuperscript{th} vertex is $\sigma(p_\tau(i))$. Note that we can have $\sigma \circ p_\tau = 0$ as there may be adjacent vertices in $Q_n$ that are sent by $\sigma$ to the same vertex of $G$. \begin{Lemma} \label{map} Let $\sigma \in {\mathcal{C}}_n^{\normalfont {\textrm{Cube}}}(G)$. Then { \setstretch{1.5} \begin{itemize} \item[(i)] $\partial_{n}^{\normalfont {\textrm{Path}}} \psi(\sigma) \in {\mathcal{L}}_{n-1}^{\normalfont {\textrm{Path}}}(G).$ In particular, $\psi(\sigma) \in {\mathcal{C}}_n^{\normalfont {\textrm{Path}}}(G)$. \item[(ii)] $\partial_{n}^{\normalfont {\textrm{Path}}} \psi(\sigma) = \psi(\partial_{n}^{\normalfont {\textrm{Cube}}} (\sigma)).$ \end{itemize}} \end{Lemma} \begin{proof} For part (i) we have \begin{eqnarray} \label{eq:yaudiff} \partial_n^{\normalfont {\textrm{Path}}} (\sigma \circ p_\tau) & = & \sum_{i=0}^n (-1)^i \cdot \sigma (p_\tau(0)) \cdots \widehat{\sigma( p_\tau(i))} \cdots \sigma(p_\tau(n)). \end{eqnarray} Let $1 \leq \ell \leq n-1$. If $\tau'$ is constructed from $\tau$ by interchanging $\tau(\ell)$ and $\tau(\ell+1)$ then $p_\tau(i) = p_{\tau'}(i)$ for $i \neq \ell$. In particular, the $\ell$\textsuperscript{th} summands of \ref{eq:yaudiff} for $p_\tau$ and $p_{\tau'}$ coincide. In addition, we have $\mathrm{sign}(\tau) = -\mathrm{sign}(\tau')$. This shows that \begin{eqnarray} \label{eq:yaudiff2} \partial_n^{\normalfont {\textrm{Path}}} \psi(\sigma) & = & \sum_{\tau \in S_n} \mathrm{sign}(\tau) \, \partial_n^{\normalfont {\textrm{Path}}} (\sigma \circ p_\tau) \\ \nonumber & = & \sum_{\tau \in S_n} \mathrm{sign}(\tau) \, \Big( (\sigma(p_\tau (1)) ,\ldots, \sigma( p_\tau (n))) + \\ \nonumber & & \qquad\qquad\qquad\qquad(-1)^n (\sigma(p_\tau(0)) ,\ldots, \sigma(p_\tau(n-1))) \Big). \end{eqnarray} Since both $(\sigma(p_\tau (1)), \ldots ,\sigma( p_\tau (n)))$ and $(\sigma(p_\tau (0)) \cdots \sigma( p_\tau (n-1)))$ are paths it follows that $\partial_{n}^{\normalfont {\textrm{Path}}} \psi(\sigma) \in {\mathcal{L}}_{n-1}^{\normalfont {\textrm{Path}}}(G)$. As a consequence $\psi(\sigma) \in {\mathcal{C}}_n^{\normalfont {\textrm{Path}}}(G)$, and we have proved (i). For part (ii), suppose that $\tau \in S_{n}$. Define \begin{itemize} \item $\tau^- \in S_{n-1}$ to be the permutation where $\tau^- (j) = \tau(j)$ if $\tau(j) < \tau(n)$ and $\tau(j)-1$ if $\tau(j) > \tau(n)$. \item $\tau^+ \in S_{n-1}$ to be the permutation where $\tau^+ (j) = \tau(j+1)$ if $\tau(j+1) < \tau(1)$ and $\tau(j+1)-1$ if $\tau(j+1) > \tau(1).$ \end{itemize} Now for $j = \tau(n)$ and $j' = \tau(1)$ we have \begin{eqnarray*} (\sigma(p_\tau(0)),\ldots, \sigma(p_\tau(n-1))) & = & (\sigma (p_{\tau^-} (0)) \ldots \sigma (p_{\tau^-} (n-1))), \\ (\sigma(p_\tau(1)),\ldots, \sigma(p_\tau(n))) & = & (\sigma (p_{\tau^+} (0), \ldots , \sigma (p_{\tau^+}(n-1))). \end{eqnarray*} Now \begin{eqnarray*} \psi(f_i^- \sigma) & = & \sum_{\tau' \in S_{n-1}} \mathrm{sign}(\tau') \, (f_i^- \sigma) \circ p_{\tau'} \\ \psi(f_i^+ \sigma) & = & \sum_{\tau' \in S_{n-1}} \mathrm{sign}(\tau') \, (f_i^+ \sigma) \circ p_{\tau'} \\ \end{eqnarray*} Since for $\tau \in S_n$ we have that $\tau$ is determined by $\tau^-$ and $\tau(n)$, as well as by $\tau^+$ and $\tau(n)$ it follows that: \begin{eqnarray*} \psi(f_i^- \sigma) & = & \sum_{{\tau \in S_{n}},{\tau(n) = i}} \mathrm{sign}(\tau^-) \, (f_i^- \sigma) \circ p_{\tau^-} \\ \psi(f_i^+ \sigma) & = & \sum_{{\tau \in S_{n}}, {\tau(1) = i}} \mathrm{sign}(\tau^+) \, (f_i^+ \sigma) \circ p_{\tau^+} \end{eqnarray*} Since $\tau(1)=i$ contributes $i-1$ inversions and $\tau(n) = i$ in the last position $n-i$ inversions to $\tau$ we obtain \begin{eqnarray} \label{eq:faceim1} \psi(f_i^- \sigma) & = & (-1)^{n-i} \sum_{{\tau \in S_{n}},{\tau(n) = i}} \mathrm{sign}(\tau) \, (f_i^- \sigma) \circ p_{\tau^-} \\ \label{eq:faceim2} \psi(f_i^+ \sigma) & = & (-1)^{i-1} \sum_{{\tau \in S_{n}} ,{\tau(1) = i}} \mathrm{sign}(\tau) \, (f_i^+ \sigma) \circ p_{\tau^+} \end{eqnarray} We have \begin{eqnarray*} \partial_n^{\normalfont {\textrm{Cube}}} \sigma & = & \sum_{i=1}^n (-1)^i (f_i^- \sigma - f_i^+ \sigma). \end{eqnarray*} Thus by \ref{eq:faceim1}, \ref{eq:faceim2} and \ref{eq:yaudiff2} we get \begin{eqnarray*} \psi(\partial_n^{\normalfont {\textrm{Cube}}} \sigma) & = & \sum_{i=1}^n (-1)^i \, (\psi(f_i^- \sigma) - \psi(f_i^+ \sigma)) \\ & = & \sum_{i=1}^n (-1)^i \Big( (-1)^{n-i} \sum_{{\tau \in S_{n}}, {\tau(\ell) = i}} \mathrm{sign}(\tau) \, (f_i^- \sigma) \circ p_{\tau^-} \, - \\ & & (-1)^{i-1} \sum_{{\tau \in S_{n}} , {\tau(1) = i}} \mathrm{sign}(\tau) \, (f_i^+ \sigma) \circ p_{\tau^+} \Big) \\ & = & \sum_{\tau \in S_n} \mathrm{sign}(\tau) \Big( (-1)^n (\sigma(p_\tau(0)), \ldots,\sigma(p_\tau(n-1))) + \\ & & (\sigma (p_\tau (1),\ldots, \sigma(p_\tau (n)) \Big) \\ & = & \partial_n^{\normalfont {\textrm{Path}}} \psi(\sigma) . \end{eqnarray*} \end{proof} Next we show that degenerate cubes are mapped to zero under $\psi$. \begin{Lemma} \label{degen} For any degenerate singular $n$-cube $\sigma\in {\mathcal{L}}_n^{\normalfont {\textrm{Cube}}}(G)$ we have $\psi(\sigma) = 0$. In particular, $\psi$ induces a map $\psi : {\mathcal{C}}_n^{\normalfont {\textrm{Cube}}}(G) \rightarrow {\mathcal{C}}_n^{\normalfont {\textrm{Path}}}(G)$. \end{Lemma} \begin{proof} If $\sigma \in {\mathcal{L}}_n^{\normalfont {\textrm{Cube}}}(G)$ is degenerate, then there is an $i\in[n]$ such that $f_i^- \sigma = f_i^+ \sigma$. Now each path $\sigma \circ p_\tau$ in $\psi(\sigma)$ at some point passes from the facet $f_i^{-} \sigma$ to the facet $f_i^{+} \sigma$ of the singular $n$-cube $\sigma$. Since $f_i^- \sigma = f_i^+ \sigma$, it follows that there are two consecutive vertices in the path $\sigma \circ p_\tau$ that are identical. This shows $\sigma \circ p_\tau = 0$ in ${\mathcal{L}}_n^{\normalfont {\textrm{Path}}}(G)$. \end{proof} For small $n$ we have good control over $\psi$. \begin{Lemma} \label{lem:smalln} The map $\psi : {\mathcal{C}}_n^{\normalfont {\textrm{Cube}}}(G) \rightarrow {\mathcal{C}}_n^{\normalfont {\textrm{Path}}}(G)$ is an isomorphism for $n \leq 1$ and surjective for $n =2$. \end{Lemma} \begin{proof} For $n = 0$ both ${\mathcal{C}}_0^{\normalfont {\textrm{Cube}}}(G)$ and ${\mathcal{C}}_0^{\normalfont {\textrm{Path}}}(G)$ are freely generated by the vertices of $G$. Since $\psi$ maps a vertex considered a the image of a $0$-cube to the path consisting of that vertex, it is clearly an isomorphism. Consider $n = 1$. A non-degenerate $1$-cube is represented by an edge with a direction chosen. Hence ${\mathcal{C}}_1^{\normalfont {\textrm{Cube}}}(G)$ has a basis given by pairs of vertices $(v_1,v_2)$ where $\{v_1,v_2\}$ is an edge in the graph. But this is also a basis for ${\mathcal{C}}_1^{\normalfont {\textrm{Path}}}(G)$. Thus ${\mathcal{C}}_1^{\normalfont {\textrm{Cube}}}(G)$ and ${\mathcal{C}}_1^{\normalfont {\textrm{Path}}}(G)$ are isomorphic and $\psi$ is an isomorphism. Finally we turn to $n = 2$. Let $\sigma \in {\mathcal{C}}_2^{\normalfont {\textrm{Path}}}(G)$. Then, by the proof of \cite[Proposition 4.2]{GLMY1}, $\sigma$ is a linear combination of paths $(v,v',v)$ and $(v,v',v'')$ where $\{v,v''\}$ is an edge in $G$, and of chains of the form $(v,w,v') -(v,w',v')$ where $w \neq w'$ and $\{v,v'\}$ is not an edge in $G$. We show that all those chains are in the image of $\psi$. Indeed: \begin{center} \begin{tabular}{lcr} \begin{minipage}[t]{0.4\textwidth} $$\psi \Big( \begin{array}{ccc} v & \rule[3pt]{1cm}{1pt} & v \\ \vrule & & \vrule \\ v & \rule[3pt]{1cm}{1pt}& v' \end{array} \Big) = (v,v',v), $$ \end{minipage} & & \begin{minipage}[t]{0.3\textwidth} $$\psi \Big( \begin{array}{ccc} v & \rule[3pt]{1cm}{1pt} & v'' \\ \vrule & & \vrule \\ v & \rule[3pt]{1cm}{1pt}& v' \end{array} \Big) = (v,v',v''),$$ \end{minipage} \\ \begin{minipage}[t]{0.4\textwidth} $$\psi \Big( \begin{array}{ccc} w' & \rule[3pt]{1cm}{1pt} & v' \\ \vrule & & \vrule \\ v & \rule[3pt]{1cm}{1pt}& w \end{array} \Big) = (v,w,v')-(v,w',v').$$ \end{minipage} & ~~~~ & \end{tabular} \end{center} \end{proof} Now we are in position to state and prove the consequence of the preceding lemmas on the relation of the two homology theories. \begin{Theorem} \label{pr:yaubarcelo} For any $n \geq 0$, the map $\psi$ induces a homomorphism $\psi_* : \mathcal{H}_n^{\normalfont {\textrm{Cube}}} (G) \rightarrow \mathcal{H}_n^{\normalfont {\textrm{Path}}}(G)$. For $n \leq 1$ the map $\psi_*$ is an isomorphism and for $n =2$ it is surjective. \end{Theorem} \begin{proof} The fact that $\psi$ induces a homomorphism $\psi_* : \mathcal{H}_n^{\normalfont {\textrm{Cube}}} (G) \rightarrow \mathcal{H}_n^{\normalfont {\textrm{Path}}}(G)$ is immediate from \ref{map} and \ref{degen}. Next we consider $n=0,1,2$. By \ref{lem:smalln} we can identify ${\mathcal{C}}_1^{\normalfont {\textrm{Cube}}}(G)$ and ${\mathcal{C}}_1^{\normalfont {\textrm{Path}}}(G)$ by considering a non-degenerate $1$-cube $\sigma$ as the edge $\psi(\sigma) = (\sigma(0),\sigma(1))$. After this identification, the differentials $\partial_1^{\normalfont {\textrm{Cube}}}$ and $\partial_1^{\normalfont {\textrm{Path}}}$ are easily seen to be identical. In particular, we have $\mathrm{Ker}\,\partial_1^{\normalfont {\textrm{Path}}} = \mathrm{Ker}\, \partial_1^{\normalfont {\textrm{Cube}}}$. Thus we need to show that $\mathrm{Im}\, \partial_2^{\normalfont {\textrm{Cube}}} = \mathrm{Im}\, \partial_2^{\normalfont {\textrm{Path}}}$. By \ref{lem:smalln}, $\psi$ is an isomorphism in dimension $1$, and since $\psi \circ \partial_2^{\normalfont {\textrm{Cube}}} = \partial_2^{\normalfont {\textrm{Path}}} \circ \psi$, it follows that $\mathrm{Im}\,\partial_2^{\normalfont {\textrm{Cube}}} \subseteq \mathrm{Im}\, \partial_2^{\normalfont {\textrm{Path}}}$. Conversely, let $\sigma \in \mathrm{Im}\, \partial_2^{\normalfont {\textrm{Path}}}$. Then there is $\sigma' \in {\mathcal{C}}_2^{\normalfont {\textrm{Path}}}(G)$ such that $\partial_2^{\normalfont {\textrm{Path}}}(\sigma') = \sigma$. By \ref{lem:smalln} there is $\sigma'' \in {\mathcal{C}}_2^{\normalfont {\textrm{Cube}}}(G)$ such that $\psi(\sigma'') = \sigma'$. Then, again by $\psi \circ \partial_2^{\normalfont {\textrm{Cube}}} = \partial_2^{\normalfont {\textrm{Path}}} \circ \psi$, it follows that $\sigma = \partial_2^{\normalfont {\textrm{Cube}}}$ and $\mathrm{Im}\, \partial_2^{\normalfont {\textrm{Path}}} \subseteq \mathrm{Im}\, \partial_2^{\normalfont {\textrm{Cube}}}$. This implies that $\psi_* : \mathcal{H}_1^{\normalfont {\textrm{Cube}}}(G) \rightarrow \mathcal{H}_1^{\normalfont {\textrm{Path}}}(G)$ is an isomorphism. Now consider homological dimension $2$. Let $\sigma + \mathrm{Im}\, \partial_2^{\normalfont {\textrm{Path}}}$ be an element of $\mathcal{H}_2^{\normalfont {\textrm{Path}}}(G)$. By \ref{lem:smalln} we know that $\phi$ is surjective in dimension $2$. Hence $\phi^{-1}(\sigma)$ is non-empty. For $\sigma' \in \phi^{-1}(\sigma)$ we have \begin{eqnarray*} \psi (\partial_2^{\normalfont {\textrm{Cube}}}(\sigma')) & = & \partial_2^{\normalfont {\textrm{Path}}} (\psi(\sigma')) \\ & = & \partial_2^{\normalfont {\textrm{Path}}} (\sigma) \\ & = & 0. \end{eqnarray*} From \ref{lem:smalln} we know that $\psi$ is bijective in dimension $1$. From that we deduce $\sigma' \in \mathrm{Ker}\, \partial_2^{\normalfont {\textrm{Cube}}}$. Thus $\sigma' + \mathrm{Im}\,\partial_3^{\normalfont {\textrm{Cube}}}(G) \in \mathcal{H}_2^{\normalfont {\textrm{Cube}}}(G)$ and $\psi_*(\sigma' + \mathrm{Im}\, \partial_2^{\normalfont {\textrm{Cube}}}) = \sigma+ \mathrm{Im}\, \partial_2^{\normalfont {\textrm{Path}}}$. Thus $\psi_*$ is surjective. \end{proof} In \cite[Theorem 1.2]{BCW} it is shown that that the abelianization of the discrete fundamental group $A_1(G)$ (see for example \cite{BBLL} for definitions) of a graph is isomorphic to $\mathcal{H}_1^{\normalfont {\textrm{Cube}}}(G)$. In \cite[Theorem 4.23]{GLMY2} it is shown that the abelianization of the discrete fundamental group of a graph is isomorphic to $\mathcal{H}_1^{\normalfont {\textrm{Path}}}(G)$. Indeed their result is more general and captures all directed graphs. Now \ref{pr:yaubarcelo} can be used to deduce either result from the other. \begin{Corollary} Let $G = (V,E)$ be a graph then there are isomorphisms: $$\mathcal{H}_1^{\normalfont {\textrm{Cube}}}(G) \cong A_1(G)/[A_1(G),A_1(G)] \cong \mathcal{H}_1^{\normalfont {\textrm{Path}}}(G).$$ \end{Corollary} \ref{pr:yaubarcelo} also raises the question if $\mathcal{H}_n^{\normalfont {\textrm{Cube}}}(G) \cong \mathcal{H}_n^{\normalfont {\textrm{Path}}}(G)$ for all $n$. While we have shown in Section 4 that this is true for many graphs $G$, it is false in general. \section{Example: $\mathcal{H}^{\normalfont {\textrm{Cube}}}_\bullet$ and $\mathcal{H}^{\normalfont {\textrm{Path}}}_\bullet$ Are Not Always Isomorphic} \label{sec-counter} Results in the previous two sections might suggest conjecturing that $\mathcal{H}_\bullet^{\normalfont {\textrm{Cube}}}(G)$ and $\mathcal{H}_\bullet^{\normalfont {\textrm{Path}}}(G)$ are the same for all graphs $G$. In this section we construct an example showing that this is not always the case. From \ref{pr:yaubarcelo} we know that $\psi_* : \mathcal{H}_2^{\normalfont {\textrm{Cube}}}(G) \rightarrow \mathcal{H}_2^{\normalfont {\textrm{Path}}}(G)$ is surjective. The next theorem shows that it is not always injective. \medskip \noindent \begin{Theorem} \label{thm:counterexample} Let $G$ be the following graph: \begin{center} \begin{tikzpicture}[scale=1.5] \Vertex[x=1.5,y=0]{1} \Vertex[x=0 ,y=1]{2} \Vertex[x=1,y=1]{3} \Vertex[x=2,y=1]{4} \Vertex[x=3,y=1]{5} \Vertex[x=0,y=2]{6} \Vertex[x=1,y=2]{7} \Vertex[x=2,y=2]{8} \Vertex[x=3,y=2]{9} \Vertex[x=1.5,y=3]{10} \Edge(1)(2) \Edge(1)(3) \Edge(1)(4) \Edge(1)(5) \Edge(2)(6) \Edge(2)(7) \Edge(3)(8)\Edge(3)(6) \Edge(4)(7) \Edge(4)(9)\Edge(5)(8)\Edge(5)(9) \Edge(6)(10)\Edge(7)(10)\Edge(8)(10)\Edge(9)(10) \end{tikzpicture} \end{center} Then $\mathcal{H}^{\normalfont {\textrm{Path}}}_2(G) \cong (0)$ and $\mathcal{H}^{\normalfont {\textrm{Cube}}}_2(G) \not\cong (0)$. \end{Theorem} \begin{proof} We first prove that $\mathcal{H}^{\normalfont {\textrm{Path}}}_2(G) \cong (0)$, by showing explicitly that every $2$-cycle is a boundary. Suppose that $\theta\in {\mathcal{C}}^{\normalfont {\textrm{Path}}}_2(G)$ is a $2$-cycle, i.e. $\partial_2 \theta = 0$. We claim that $\theta$ is a linear combination of cycles of one of the following three types: \begin{itemize} \item[(1)] $(b,a,b) - (a,b,a)$, where $\{a,b\}$ is an edge of $G$, \item[(2)] $((a,b,c)-(a,d,c)) + ((d,c,b)-(d,a,b)) + (d,a,d) - (c,b,c)$, where $a,b,c,d$, are consecutive vertices of a quadrilateral, \item[(3)] the cycle \begin{gather*} ((6,2,1)-(6,3,1))+ ((8,3,1)-(8,5,1)) + ((9,5,1)-(9,4,1))\qquad\qquad \\\ \qquad\qquad+ ((7,4,1)-(7,2,1)) + ((10,6 ,3)- (10,8, 3)) + ((10,8, 5) - (10,9, 5)) \\ \qquad \qquad\qquad\qquad +((10,9,4) - (10,7 ,4)) +((10, 7, 2)-(10, 6, 2)). \end{gather*} \end{itemize} One can visualize $G$ as the $1$-skeleton of a polytope with eight quadrilateral facets. The cycle (3) is obtained by giving each quadrilateral an outward orientation, then assigning signs to paths around each quadrilateral using a right-hand rule. Each of (1), (2), and (3) is easily seen to be a boundary (and hence a cycle): \begin{gather*} (b,a,b) - (a,b,a) = \partial_3 \big( (a,b,a,b) \big) \\[+1ex] ((a,b,c)-(a,d,c)) + ((d,c,b)-(d,a,b)) + (d,a,d) - (c,b,c) \qquad\qquad\qquad\qquad \\ \qquad\qquad\qquad\qquad\qquad= \partial_3 \big( (d,a,b,c) - (d,a,d,c) - (d,c,b,c) \big), \end{gather*} and (3) is equal to \begin{gather*} \partial_3 \big((10,6,2,1) - (10 ,6 ,3 ,1) + (10 , 7 ,4 ,1)- (10 , 7 ,2, 1) \qquad\qquad \\ \qquad\qquad+ (10 ,8, 3, 1)- (10,8, 5, 1)+ (10,9, 5, 1) - (10,9 ,4 ,1)\big). \end{gather*} To complete the argument, we must show that every $2$-cycle can be expressed as a linear combination of cycles of type (1), (2), and (3). If $\theta$ is a $2$-cycle, let $(a,b,c)$ be the lexicographically first term in $\theta$ with the following properties: (i) it is injective, i.e. not of the form $(a,b,a)$, and (ii) it is not monotone decreasing, i.e. not satisfying $a>b>c$. Let us call an injective term $(a,b,c)$ ``bad" if it satisfies property (ii), and ``good" otherwise (i.e. if it is decreasing). Necessarily, such an $(a,b,c)$ must be paired in $\theta$ with another opposite-signed term $(a,d,c)$ where $\{d,c\}$ and $\{a,d\}$ form the edges opposite to $\{a,b\}$ and $\{b,c\}$ in a quadrilateral of $G$ (otherwise the term $(a,c)$ in $\partial_2((a,b,c))$ does not cancel). Since $(a,b,c)$ is lexicographically first in $\theta$, we must have $b<d$. We claim further that $a<d$. If $a<b$ this is immediate; if $a>b<c$, it is easy to check that $a>d$ does not hold in any of the eight quadrilaterals in $G$. If $\tau$ is the canonical cycle of type (2) above, then since $b<d$ and $a<d$, every injective term in $\tau$ follows $(a,b,c)$ in lexicographic order. Hence we can use $\tau$ to eliminate $(a,b,c)$ from $\theta$, and by repeating the process eventually arrive at a cycle $\theta^*$ in which every injective term is ``good" , i.e., of the form $(a,b,c)$ with $a>b>c$. In the boundary $\partial_2 \theta^*$, all $1$-chains arising from good injective terms must be of the form $(x,y)$ with $x>y$. Hence if $\theta^*$ contains a non-injective term $(a,b,a)$, it must also contain a corresponding term of the form $(b,a,b)$, and we can cancel them both out by subtracting a cycle of type (2). Eventually we arrive at a cycle $\theta^{**}$ in which every term is injective and ``good''. As noted above, all injective terms must appear in pairs $(a,b,c), (a,d,c)$ arising from one of the eight quadrilaterals in $G$, with opposite-signed coeffients of equal magnitude. We may thus regard the equation $\partial_2 \theta^{**} = 0$ as a homogeneous linear system with eight unknowns (corresponding to the ``good'' quadrilateral boundary pairs appearing in (3)) and 16 equations corresponding to the coefficients of the $1$-chains \begin{gather*}(2,1), (3,1), (4,1), (5,1), (6,2), (6,3), (7,2), (7,4), (8,3), (8,5), \\ (9,4), (9,5), (10,6),(10,7),(10,8),(10,9). \end{gather*} This system has a matrix \[ \left( \begin{array}{cccccccc} 1 & 0 & 0 & -1 & 0 & 0 & 0 & 0 \\ -1 & 1 & 0 & 0 & 0 & 0 & 0 & 0 \\ 0 & 0 & -1 & 1 & 0 & 0 & 0 & 0 \\ 0 & -1 & 1 & 0 & 0 & 0 & 0 & 0 \\ 1 & 0 & 0 & 0 & 0 & 0 & 0 & -1 \\ -1 & 0 & 0 & 0 & 1 & 0 & 0 & 0 \\ 0 & 0 & 0 & -1 & 0 & 0 & 0 & 1 \\ 0 & 0 & 0 & 1 & 0 & 0 & -1 & 0 \\ 0 & 1 & 0 & 0 & -1 & 0 & 0 & 0 \\ 0 & -1 & 0 & 0 & 0 & 1 & 0 & 0 \\ 0 & 0 & -1 & 0 & 0 & 0 & 1 & 0 \\ 0 & 0 & 1 & 0 & 0 & -1 & 0 & 0 \\ 0 & 0 & 0 & 0 & 1 & 0 & 0 & -1 \\ 0 & 0 & 0 & 0 & 0 & 0 & -1 & 1 \\ 0 & 0 & 0 & 0 & -1 & 1 & 0 & 0 \\ 0 & 0 & 0 & 0 & 0 & -1 & 1 & 0 \\ \end{array} \right), \] which is easily seen to have rank $7$, and hence all solutions to $\partial \theta^{**} = 0$ are constant multiples of (3). Since cycles of type (1), (2), and (3) are all boundaries, this completes the proof that every $2$-cycle is a boundary, and hence $\mathcal{H}^{\normalfont {\textrm{Path}}}_2(G) \cong (0)$. \end{proof} Next we turn to proving that $\mathcal{H}^{\normalfont {\textrm{Cube}}}_2(G) \not\cong (0)$, which will be done by finding an explicit $2$-cycle $\theta\in {\mathcal{C}}_2^{\normalfont {\textrm{Cube}}}(G)$ that is not a boundary. Define \begin{equation}\label{theta} \begin{gathered} \theta =(1,2,3,6) - (1,2,4,7) + (1,3,5,8) - (1,4,5,9) \qquad\qquad \\ \qquad\qquad- (2,6,7,10) + (3,6,8,10) - (4,7,9,10) + (5,8,9,10). \end{gathered} \end{equation} Recall that each sequence denotes a labeling of the canonical $2$-cube by vertices in $G$, proceeding recursively by dimension. For example, $(1,2,3,6)$ represents the labeling \begin{center} \begin{tikzpicture}[scale=2] \Vertex[x=0 ,y=0, L=$$]{1} \Vertex[x=1 ,y=0, L=$$]{2} \Vertex[x=0,y=1, L=$$]{3} \Vertex[x=1,y=1, L=$$]{6} \Edge(1)(2) \Edge(1)(3) \Edge(2)(6) \Edge(3)(6) \node [color=blue] at ([shift={(-.2,.2)}]1) {1}; \node [color=blue] at ([shift={(-.2,.2)}]2) {2}; \node [color=blue] at ([shift={(-.2,.2)}]3) {3}; \node [color=blue] at ([shift={(-.2,.2)}]6) {6}; \end{tikzpicture} \end{center} One can interpret $\theta$ as the result of wrapping quadrilaterals around a $3$-polytope in an orientation-preserving way. When labeled properly by their vertex names, each face can be viewed as graph homomorphisms from a $2$-cube into $G$. It is easy to check that $\theta$ is a cycle, i.e., $\partial_2 \theta= 0$. We will show that $\theta$ is not a boundary, by constructing a linear invariant $\Psi$ on ${\mathcal{C}}^{\normalfont {\textrm{Cube}}}_2(G)$ that is zero on every $2$-boundary but is nonzero on $\theta$. Fix a quadrilateral $Q_0$ in $G$, say $Q_0=(1,2,3,6)$ with vertices listed in increasing order. We say that a $2$-cell $F = (a,b,c,d)$ (that is, a $G$-labeled $2$-cube, with labels in the standard reading order) is {\em supported by $Q_0$} if its labels agree with those of $Q_0$ in some order. Since $F$ is a graph map, the permutation $\sigma$ mapping $(1,2,3,6)$ onto $(a,b,c,d)$ is an element of the dihedral group $D_4$. Define the {\em weight} $w(F)$ of $F$ to be $\chi(\sigma)$, where $\chi$ is the reflection character of $D_4$. In other words, $\chi(\sigma) = \pm 1$ according to whether $\sigma$ is a reflection. If $X$ is any $2$-chain, define $\Psi(X)$ to be the sum over $X$ of the coefficient of each $2$-cell times the weight $w(F)$ of that cell. In this computation, $w(F)=0$ if $F$ is not supported by $Q_0$. The following rules define $\Psi$ explicitly on the eight $2$-cells supported by $Q_0$: \[ \Psi: \begin{cases} (1,2,3,6) & \longmapsto +1 \\ (2,6,1,3) & \longmapsto +1 \\ (3,1,6,2) & \longmapsto +1 \\ (6,3,2,1) & \longmapsto +1\\ (1,3,2,6) & \longmapsto -1 \\ (2,1,6,3) & \longmapsto -1 \\ (3,6,1,2) & \longmapsto -1 \\ (6,2,3,1) & \longmapsto -1 \end{cases} \] As an illustration, note that $\Psi(\theta) = 1$, since $(1,2,3,6)$ is the only $2$-cell appearing in $\theta$ supported by $Q_0$. As another illustration, consider the $3$-cell \[ Y = (3,6,1,3,1,2,3,6), \] which has $Q_0 = (1,2,3,6)$ as its top face. Its boundary is \[ \partial_3 Y = (1,2,3,6) - (1,3,3,6) - (3,1,1,3)+(3,6,1,2) - (3,6,1,3) + (6,3,2,6). \] In this case, two terms are supported by $Q_0$ but they appear with opposite signs when $\Psi$ is applied, and we get $\Psi (\partial_3 Y) = 0$. This turns out to be a general phenomenon: \begin{Lemma} For any non-degenerate $3$-cell $Y$, we have $\Psi(\partial_3 Y) = 0$. Consequently, $\Psi(X) = 0 $ for any $2$-boundary $X$. \end{Lemma} \begin{proof} Suppose that $Y$ is a non-degenerate $3$-cell. We claim first that the number of $2$-faces supported by $Q_0= (1,2,3,6)$ is equal to $2$ or $4$. To prove this, we systematically eliminate the other cases. It is easy to see that the number of $Q_0$-supported $2$-faces cannot be $3, 5$, or $6$, since if two such faces are adjacent dihedrally, the two faces adjacent to both of of those faces have a repeated label, and hence cannot be supported by $Q_0$. It remains to show that the number of $Q_0$-supported faces cannot be $1$. Suppose that $Y$ contains only one $Q_0$-supported face, e.g. as indicated in the following picture where the bottom four vertices are labeled by $1,2,3$ and $6$, and the others are labeled generically by $A,B,C$ and $D$. \begin{center} \begin{tikzpicture}[scale=.8] \Vertex[x=0 ,y=0]{1} \Vertex[x=3 ,y=0]{2} \Vertex[x=2,y=2]{3} \Vertex[x=5,y=2]{6} \Vertex[x=0,y=3]{A} \Vertex[x=3,y=3]{B} \Vertex[x=2,y=5]{C} \Vertex[x=5,y=5]{D} \Edge(1)(2) \Edge(1)(3) \Edge(2)(6) \Edge(3)(6) \Edge(A)(B) \Edge(C)(D) \Edge(A)(C) \Edge(B)(D) \Edge(1)(A) \Edge(2)(B) \Edge(3)(C) \Edge(6)(D) \end{tikzpicture} \end{center} The label $B$ can be either $1,2, 6$, or $7$, since these are the vertices adjacent to $2$ in $G$. We will argue that $B$ must be equal to $2$. First suppose $B=1$. Then $D$ must be either $2$ or $3$, since $D$ must be adjacent to $1$ and $6$. If $D=3$, this contradicts the assumption that only one face is $Q_0$-supported. Hence $D=2$. If $D=2$, then $C$ must be either $1$ or $6$. If $C=1$, we again have two $Q_0$-supported faces and hence a contradiction which implies $C=6$. Now $A$ must be either $2$ or $3$, since it is a vertex adjacent to $1$ and $6$. Either choice produces a second $Q_0$-supported face, and hence all cases lead to a contradiction, and thus $B\ne 1$. A similar (symmetrical) argument using the front face shows that $B\ne 6$. Continuing, suppose that $B=7$. The possible values for $A,C,D$, determined by adjacencies in $G$, are $D\in\{2, 10\}, A \in \{2,4\}, C \in \{1,6\}$. Note that the possibilities $C=3$ and $C=8$ are not included because neither is adjacent to either $2$ or $4$ in $G$. We proceed systematically: If $D=10$, then $C=6$ since $10$ is not adjacent to $1$. This implies $A = 2$, since $4$ is not adjacent to $6$, yielding a second $Q_0$-supported face, which is a contradiction and hence $D = 2$. If $C=1$, we obtain a second $Q_0$-supported face, hence $C=6$, which forces $A=2$ since $4$ is not adjacent to $6$. This choice produces a second $Q_0$-supported face, and hence a contradiction. This implies $B\ne 7$, which leaves $B=2$ as the only possible choice. Since $2$ and $3$ are related by an automorphism of $G$, a symmetrical argument shows that $C=3$. From that point it is easy to show that $A=1$ and $D = 6$, thus forcing a second $Q_0$-supported face on top -- as well as a degenerate cube. This double contradiction completes the proof our first claim. Next we claim that if the number of $Q_0$-supported faces, is $4$, then those faces must ``wrap around'' the $3$-cell, i.e. they avoid one of the three coordinate axes. If the number is $2$, the faces may either be adjacent (and ``hinged"), or opposite, in which case their labels differ by a 90 degree rotation. We can conclude that if $Y$ is a $3$-cell, then the $Q_0$-supported faces can occur in arrangements of three types. If the number of such faces, is $4$, then those faces must ``wrap around'' the $3$-cell, i.e. they avoid one of the three coordinate axes. This follows since if two $Q_0$-supported faces are dihedrally adjacent, then the two faces adjacent to both of them cannot be $Q_0$-supported. If the number of $Q_0$-supported faces is $2$, the faces may either be adjacent (and ``hinged"), or opposite, in which case an easy argument shows that their labels differ by a 90 degree rotation. With this information in hand, we can proceed to the proof of the main result. Suppose that $Y$ is a non-degenerate $3$-cell with two $Q_0$-supported faces $F_1$ and $F_2$ that are dihedrally adjacent. Let $\textrm{sgn}_\partial(F_1)$ and $\textrm{sgn}_\partial(F_2)$ denote the signs associated to $F_1$ and $F_2$ by the boundary operator $\partial_3$. Then we claim that \begin{equation}\label{psi} \chi(F_1)\, \textrm{sgn}_\partial(F_1) = - \chi(F_2)\, \textrm{sgn}_\partial(F_2). \end{equation} If we prove \ref{psi}, it will follow that $\Psi(\partial_3 Y) = 0$ in two of the three cases, i.e. either four $Q_0$-supported faces or two such faces that are dihedrally adjacent. We will prove \ref{psi} in the form \begin{equation}\label{psi2} \chi(F_1) \chi(F_2) = - \textrm{sgn}_\partial(F_1) \textrm{sgn}_\partial(F_2). \end{equation} Since $\chi$ is a multiplicative character, we can regard $\chi(F_1) \chi(F_2)$ as $\chi(\sigma)$ where $\sigma$ is the permutation mapping a generic set of (distinct) labels on the vertices of $F_1$ to the corresponding set of labels on $F_2$ obtained by flipping 90 degrees through the dihedral edge. On both faces, the labels are read in standard order (recursively by dimension). Denote the six faces of $Y$ by $F_{0**}, F_{1**}, F_{*0*}, F_{*1*},F_{**0}, F_{**1}$, with the obvious notation, e.g., $F_{*1*}$ denotes the face $\{(x,1,z)\}$, where $0 \le x,z \le 1$. Let us say that $F$ is a {\em positive face} if $F$ is one of $F_{1**}, F_{*0*}, F_{**1}$, and a {\em negative face} if it is one of $F_{0**}, F_{*1*}, F_{**0}$. These designations correspond exactly to the signs of $\textrm{sgn}_\partial(F)$. Hence \ref{psi2} can be interpreted as saying that if $F_1$ and $F_2$ are dihedrally adjacent faces of a $3$-cell, both supported by $Q_0$, then \begin{equation}\label{psi3} \chi(F_1) \chi(F_2) = \begin{cases} -1 & \textrm{if $F_1$ and $F_2$ are both positive or both negative}\\ +1 & \textrm{otherwise}\\ \end{cases} \end{equation} For each dihedrally adjacent pair $F_1,F_2$, we can compute the left hand side of \ref{psi3} as the reflection character $\chi(\sigma)$ of the permutation $\sigma$ that maps the labeling of $F_1$ onto the labeling of $F_2$, where each labeling is read in the standard order. There are 12 cases (one for each edge), which can be grouped into four classes of permutations: \begin{itemize} \item the identity $id$, with $\chi(id) = +1$, \item a rotation $R$ through 90 degrees, with $\chi(R)=+1$, \item a reflection $\phi$ around a diagonal axis, with $\chi(\phi) = -1$, \item a reflection $\psi$ around an horizontal or vertical axis, with $\chi(\psi)=-1$. \end{itemize} These four types are indicated on the edges of the following diagram. Edges in red correspond to permutations with $\chi(\sigma) = +1$ and edges in black correspond to permutations with $\chi(\sigma) = -1$. \begin{center} \begin{tikzpicture}[scale=1] \Vertex[x=0 ,y=0,L=$1$]{1} \Vertex[x=3 ,y=0]{2} \Vertex[x=2,y=2,L=$3$]{3} \Vertex[x=5,y=2]{4} \Vertex[x=0,y=3,L=$5$]{5} \Vertex[x=3,y=3,L=$6$]{6} \Vertex[x=2,y=5,L=$7$]{7} \Vertex[x=5,y=5,L=$8$]{8} \Edge[style={ultra thick},color=red,label=$id$](1)(2) \Edge[style={ultra thick},color=black,label=$\phi$](1)(3) \Edge[style={ultra thick},color=red,label=$R$](2)(4) \Edge[style={ultra thick},color=black,label=$\psi$](3)(4) \Edge[style={ultra thick},color=black,label=$\psi$](5)(6) \Edge[style={ultra thick},color=red,label=$R$](5)(7) \Edge[style={ultra thick},color=black,label=$\phi$](6)(8) \Edge[style={ultra thick},color=red,label=$id$](7)(8) \Edge[style={ultra thick},color=red,label=$id$](1)(5) \Edge[style={ultra thick},color=black,label=$\psi$](2)(6) \Edge[style={ultra thick},color=black,label=$\psi$](3)(7) \Edge[style={ultra thick},color=red,label=$id$](4)(8) \end{tikzpicture} \end{center} \noindent The values of $\chi(\sigma)$ in each of the four cases can be verified in a straightforward manner. Furthermore, it is easy to verify that for dihedral edges labeled in the diagram by $id$ or $R$, the corresponding pairs of faces $F_1, F_2$ have opposite boundary parity, i.e. $\textrm{sgn}_\partial(F_1) = - \textrm{sgn}_\partial(F_2)$, and for dihedral edges labeled by $\phi$ or $\psi$, the faces have the same boundary parity. This is exactly the content of \ref{psi3}. In order to complete the proof, it is only necessary to consider the case where $Y$ has exactly two opposite $Q_0$-supported faces $F_1$ and $F_2$, with labels differing by a 90 degree rotation. In this case $\chi(F_1) = \chi(F_2)$ and $\textrm{sgn}_\partial(F_1) = - \textrm{sgn}_\partial(F_2)$, implying \ref{psi} immediately. We have shown that $\Psi(\partial_3 Y) = 0$ for every $3$-cell $Y$. Since $\Psi(\theta) = 1$, where $\theta$ is defined in \ref{theta}, it follows that $\theta$ is not a boundary, and hence $\mathcal{H}^{\normalfont {\textrm{Cube}}}_2(G) \not\cong (0)$. Further computation (not included here) shows that $\mathcal{H}^{\normalfont {\textrm{Cube}}}_2(G) \cong R$, i.e. the homology in dimension $2$ is generated by $\theta$. \end{proof} In fact, the same graph can be used to show that when $n\ge 3$, the map $\psi_* : \mathcal{H}_n^{\normalfont {\textrm{Cube}}}(G) \rightarrow \mathcal{H}_n^{\normalfont {\textrm{Path}}}(G)$ defined \ref{sec-map} is not is not always surjective. Computations using \cite{Maerte17} for $\mathcal{H}^{\normalfont {\textrm{Cube}}}_\bullet$ and Mathematica for $\mathcal{H}^{\normalfont {\textrm{Path}}}_\bullet$ (not displayed here) have shown that if $G$ is the graph from \ref{thm:counterexample} then \[ \mathcal{H}_3^{\normalfont {\textrm{Cube}}}(G) \cong (0) \textrm{ and }\mathcal{H}_3^{\normalfont {\textrm{Path}}}(G) \cong R. \] It would be interesting to find self-contained, accessible proofs of these results. \section{Comments, Examples, and Questions for Further Study\label{sec-questions}} An important family of examples for which we have less than perfect information consists of the $k$-cycles $Z_k$, with $V(Z_k) = \{1,2,\dots, k\}$ and $E(Z_k) = \{ \{i,i+1\} \;|\; 1 \le i <k\} \cup \{\{k,1\}\}$. The following proposition states what we know about these graphs. \begin{Proposition}\label{cycles} Let $Z_k$ be a $k$-cycle. Then \begin{enumerate} \item If $k=3$ or $k=4$, then $\mathcal{H}_n^{\normalfont {\textrm{Cube}}}(Z_k) \cong \mathcal{H}_n^{\normalfont {\textrm{Path}}}(Z_k) \cong (0)$ for all $n>0$. \item $\mathcal{H}_1^{\normalfont {\textrm{Cube}}}(Z_5) \cong R$, $\mathcal{H}_2^{\normalfont {\textrm{Cube}}}(Z_5) \cong (0)$, $\mathcal{H}_3^{\normalfont {\textrm{Cube}}}(Z_5) \cong (0)$. \item $\mathcal{H}_1^{\normalfont {\textrm{Path}}}(Z_k) \cong R$, and $\mathcal{H}_n^{\normalfont {\textrm{Path}}}(Z_k) \cong (0)$ for $k>5,$ $n\ge 2$. \end{enumerate} \end{Proposition} \begin{proof} Since $Z_3$ is a complete graph and $Z_4$ is a $2$-cube, statement (1) follows from \Ref{cor-examples}, in both cases. In statement (2), the dimension 1 case follows by an easy direct calculation, or by applying Theorem 2.7 of \cite{BKLW} (see also Theorem 5.2 in \cite{BBLL}) and Theorem 4.1 of \cite{BCW}. In dimension 2 and 3, the results were obtained by computer computations which we do not include here. In statement (3), the dimension 1 case again follows by an easy direct calculation, or by applying statement (2) together with \Ref{pr:yaubarcelo}. The remaining cases in statement (3) follow from a more general result stated in the next proposition. \end{proof} \begin{Proposition}\label{no4cycles} Suppose that $G$ is an undirected graph containing no 3-cycles and no 4-cycles. Then $\mathcal{H}_n^{\normalfont {\textrm{Path}}}(G) \cong (0)$ for $n\ge 2$. \end{Proposition} \begin{proof} If $n\ge 2$, then every generator of ${\mathcal{C}}^{\normalfont {\textrm{Path}}}_n(G)$ must be an alternating path, that is, it has the form $(a,b,a,b,\cdots ,b)$ or $(a,b,a,b,\cdots, a)$ for some pair $\{a,b\}\in E(G)$. Denote this path by $w_{ab}^n$, where $a$ and $b$ are the first two elements and the final element is determined by the parity of $n$. An easy computation shows that \begin{equation}\label{wbdy} \partial_n w^n_{ab}= w^{n-1}_{ba} + (-1)^n w^{n-1}_{ab} . \end{equation} Suppose that \[ \gamma = \sum_ {\{a,b\}\in E(G)} c_{ab}\, w^n_{ab} + c_{ba} \,w^n_{ba} \] is a cycle in ${\mathcal{C}}_n^{\normalfont {\textrm{Path}}}(G)$. It follows from \Ref{wbdy} that \[ c_{ba} + (-1)^{n} c_{ab} = 0 \] for all $\{a,b\}\in E(G)$, implying that $\gamma$ may be expressed as a linear combination of terms of the form \[ w_{ba}^n + (-1)^{n+1} w_{ab}. \]Since $w_{ba}^{n+1} + (-1)^{n+1} w_{ab} = \partial_{n+1} w^{n+1}_{ba}$, every cycle $\gamma\in {\mathcal{C}}_n^{\normalfont {\textrm{Path}}}(G)$ is a boundary, and hence $\mathcal{H}_n^{\normalfont {\textrm{Path}}}(G) = (0).$ \end{proof} We conjecture that \Ref{no4cycles} also holds for $\mathcal{H}_n^{\normalfont {\textrm{Cube}}}(G)$: \begin{Conjecture} Suppose that $G$ is an undirected graph containing no 3-cycles and no 4-cycles. Then $\mathcal{H}_n^{\normalfont {\textrm{Cube}}}(G) \cong (0)$ for $n\ge 2$. \end{Conjecture} Although our computational evidence is somewhat limited, it seems natural to conjecture that a weaker property holds for all graphs $G$ and for both cubical and path homology. \begin{Conjecture} For any undirected graph $G$, there exists an integer $N$ such that we have $\mathcal{H}_n^{\normalfont {\textrm{Cube}}}(G)\cong \mathcal{H}_n^{\normalfont {\textrm{Path}}}(G) \cong (0)$ for $n\ge N$. \end{Conjecture} In cases where the methods of Sections 3 and 4 do not apply, computation of $\mathcal{H}_n^{\normalfont {\textrm{Cube}}}(G)$ remains a significant challenge, since the problem size increases rapidly with dimension. For example, when $G = Z_5$, $\mathrm{rank}\, {\mathcal{C}}_3^{\normalfont {\textrm{Cube}}}(G) = 2230$ and $\mathrm{rank}\, {\mathcal{C}}_4^{\normalfont {\textrm{Cube}}}(G) = 978350$. For the graph $G$ in \ref{thm:counterexample}, $\mathrm{rank}\, {\mathcal{C}}_3^{\normalfont {\textrm{Cube}}}(G) = 21552$ and $\mathrm{rank}\, {\mathcal{C}}_4^{\normalfont {\textrm{Cube}}}(G) = 21745744$. It would be useful to develop more effective tools to compute $\mathcal{H}^{\normalfont {\textrm{Cube}}}_n(G)$ for all $n$ in these cases. \bibliographystyle{siam}
{ "timestamp": "2018-03-21T01:10:52", "yymm": "1803", "arxiv_id": "1803.07497", "language": "en", "url": "https://arxiv.org/abs/1803.07497", "abstract": "In this paper we study and compare two homology theories for (simple and undirected) graphs. The first, which was developed by Barcelo, Caprano, and White, is based on graph maps from hypercubes to the graph. The second theory was developed by Grigor'yan, Lin, Muranov, and Yau, and is based on paths in the graph. Results in both settings imply that the respective homology groups are isomorphic in homological dimension one. We show that, for several infinite classes of graphs, the two theories lead to isomorphic homology groups in all dimensions. However, we provide an example for which the homology groups of the two theories are not isomorphic at least in dimensions two and three. We establish a natural map from the cubical to the path homology groups which is an isomorphism in dimension one and surjective in dimension two. Again our example shows that in general the map is not surjective in dimension three and not injective in dimension two. In the process we develop tools to compute the homology groups for both theories in all dimensions.", "subjects": "Combinatorics (math.CO); Algebraic Topology (math.AT)", "title": "Discrete Cubical and Path Homologies of Graphs", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9869795098861571, "lm_q2_score": 0.8289388083214156, "lm_q1q2_score": 0.818145618762686 }
https://arxiv.org/abs/1206.0320
Expected Patterns in Permutation Classes
In the set of all patterns in $S_n$, it is clear that each k-pattern occurs equally often. If we instead restrict to the class of permutations avoiding a specific pattern, the situation quickly becomes more interesting. Miklós Bóna recently proved that, surprisingly, if we consider the class of permutations avoiding the pattern 132, all other non-monotone patterns of length 3 are equally common. In this paper we examine the class $\Av (123)$, and give exact formula for the occurrences of each length 3 pattern. While this class does not break down as nicely as $\Av (132)$, we find some interesting similarities between the two and prove that the number of 231 patterns is the same in each.
\section{Background} Let $p = p_1 p_2 \ldots p_n$ be a permutation in the symmetric group $S_n$ written in one-line notation. Given a permutation $q \in S_k$, say that $p$ contains $q$ as a pattern (denoted $q \prec p$) if there exist $k$ indices $ 1 \leq i_1 \leq i_2 \leq \ldots \leq i_k \leq n$ such that the entries $p_{i_1} p_{i_2} \ldots p_{i_k}$ are in the same relative order as the entries of $q$ ($q_j < q_k$ if and only if $p_{i_j} < p_{i_k}$). If $p$ does not contain $q$ as a pattern, we say that $p$ \emph{avoids} $q$. The set of all permutations equipped with this ordering can be viewed as a partially ordered set which is graded with respect to permutations length. With this in mind, we define a \emph{permutation class} to be a downset (or ideal) of this poset. That is, a class is a collection of permutations $\mathcal{C}$ for which, if $p \in \mathcal{C}$ and $q \prec p$, then $q \in \mathcal{C}$. Given a pattern $q$, the set $\Av (q)$ of all permutations avoiding $q$ forms a natural permutation class, and much study has been devoted to understanding and enumerating classes of this form. An early result in the area from Knuth \cite{knuth3}, is that the number of $n$-permutations avoiding the pattern $231$ is equal to the Catalan number $c_n = \frac{1}{n+1} \binom{2n}{n}$, and these are exactly the stack sortable permutations. A more comprehensive introduction to permutation patterns can be found in \cite{bonabook}. A question of Joshua Cooper and a recent result of Mik\'os B\'ona have opened up a new line of research: in a given permutation class, what can be said about the average number of occurences of each pattern? or equivalently: what is the total number of occurences of each pattern in the class? It is simple to show that in the class of all permutations, all patterns of a given length are equally common. The situation becomes much more interesting as we restrict our attention to smaller classes. \section{Preliminaries} \begin{definition} Let $p,q$ be permutations. Denote by $\num_{q}(p)$ the number of occurences of $q$ in $p$ as a pattern. \end{definition} For example, $\num_{213}(462513) = 2$ since the first third and fourth entries as well as the third fifth and sixth entries form $213$ patterns. Also, for any permutation $p$, $\num_{21}(p)$ counts the number of inversions of $p$. Note that every permutation statistic can be expressed through combinations of counts of permutation patterns, as described in \cite{claesson11}. We'll be concerned primarily with the total number of patterns in a class of permutations. For simplicity, we use similar notation. \begin{definition} Let $\mathcal{C}$ denote a permutation class, and $q$ a pattern. Define $\num_{q}(\mathcal{C})= \sum_{p \in \mathcal{C}} \num_q(p)$. We will omit the $\mathcal{C}$ when the class in question is unambiguous. \end{definition} \begin{example} \label{linearity} Let $q \in S_k$. Then it follows by linearity of expectation that $$\num_q(S_n) = \frac{n!}{k!} \binom{n}{k}.$$ \end{example} The Catalan numbers will appear frequently in our enumeration, and so it will be useful to establish some standard notation and a few simple identities. \begin{definition} Let $c_n = \frac{1}{n+1} \binom{2n}{n}$ denote the $n$th Catalan number. Also, let $$ C(x) = \sum_{n\geq 0} c_n x^n = \frac{1 - \sqrt{1-4x}}{2x}.$$ \end{definition} \begin{fact} The following identities follow directly from the recurrence $C(x) = xC(x)^2 +1$. $$ C(x)^2 = \frac{C(x)}{1-xC(x)} = \frac{1}{(1-xC(x))^2} \text{ \ and \ } \frac{C(x)-1}{C(x)} = xC(x).$$ \end{fact} In \cite{bona10} and \cite{bona12}, Mikl\'os B\'ona studied the class $\Av_n (132)$ and found some surprising symmetries. He also gave exact formula and generating functions for the expectation of all length three patterns. In this paper we give a similar classification of the class $\Av_n(123)$, with some equally surprising connections to $\Av_n (132)$. Because both $132$ and $123$ are involutions and $231^{-1} = 312$, we have the following identity. Further identites, however, require considerably more effort. \begin{fact} In both $\Av_n (132)$ and $\Av_n(123)$, $\num_{231} = \num_{312}$. \end{fact} For a given integer $k$, inversion provides a simple bijection from the set of permutations containing exactly $k$ $231$ patterns to the set containing exactly $k$ $312$ patterns, proving not only that the total number of each pattern is the same, but that these numbers are equidistributed across the sets $\Av_n (132)$ and $\Av_n(123)$. B\'ona showed that equidistribution is not required for the total number of patterns to be equal. \begin{theorem}[B\'ona] \label{bonasthm} In $\Av_n (132) $, the total numbers of $231$, $213$, and $312$ patterns are equal, and their numbers, with respect to $n$, are given by the generating function $$ \frac{x^2C(x)^3}{(1-2xC(x))(1-4x)^\frac32}.$$ Furthermore, $321$ is the most common pattern and $123$ is the least common. \end{theorem} The first few values of $\num_q(\Av_n(123))$ and $\num_q(\Av_n (132))$ for $q$ of length $3$ are shown in the table below $$ \begin{array}{c|c|c|c|c|c|c} \multicolumn{7}{c}{\Av_n(123) } \\ \text{length} & \num_{123} & \num_{132} & \num_{213} & \num_{231} & \num_{312} & \num_{321} \\ \hline 3 & 0 & 1 & 1 & 1 & 1 & 1 \\ 4 & 0 & 9 & 9 & 11 & 11 & 16 \\ 5 & 0 & 57 & 57 & 81 & 81 & 144 \\ 6 & 0 & 312 & 312 & 500 & 500 & 1016 \\ 7 & 0 & 1578 & 1578 & 2794 & 2794 & 6271 \end{array} $$ \vspace{1pc} $$ \begin{array}{c|c|c|c|c|c|c} \multicolumn{7}{c}{\Av_n (132) } \\ \text{length} & \num_{123} & \num_{132} & \num_{213} & \num_{231} & \num_{312} & \num_{321} \\ \hline 3 & 1 & 0 & 1 & 1 & 1 & 1 \\ 4 & 10 & 0 & 11 & 11 & 11 & 13 \\ 5 & 68 & 0 & 81 & 81 & 81 & 109 \\ 6 & 392 & 0 & 500 & 500 & 500 & 748 \\ 7 & 2063 & 0 & 2794 & 2794 & 2794 & 4570 \end{array} $$ \vspace{1pc} Note that $\num_{231}$ and $\num_{312}$ are \emph{not} equidistributed as statistics in $\Av_n (132)$. Theorem \ref{bonasthm} was proved with a bijection from patterns to patterns, not necessarily respecting the underlying permutation. Turning our attention to the class $\Av(123)$, we will similarly classify the expectation of all length $3$ patterns and provide both generating functions and exact formula. In addition, we show some interesting and surprising connections to patterns in $\Av_n (132)$. In particular, we will show that the number of $231$ patterns in each set is equal, as suggested by the numerical evidence. \section{The class Av$(123)$} \subsection{Patterns of length $2$} The simplest place to start is with patterns of length $2$. The number of $21$ patterns corresponds to the total number of inversions, and these numbers have already been studied, most notably in \cite{cheng7}. Clearly, the total number of $21$ patterns plus the number of $12$ patterns gives the total number of pairs of entries in all permutations in the class, which is given by $\binom{n}{2} c_n$. \begin{theorem}[Cheng, Eu, Fu] \label{inv} Let $\mathcal{C}_n = \Av_n(123)$. Then $$ \sum_{n \geq 0} f_{12}(\mathcal{C}_n) x^n = \frac{x^2 C(x)^2}{1-4x}.$$ Furthermore, we have that $$ \num_{12}(\mathcal{C}_n) = 4^{n-1} - \binom{2n-1}{n}.$$ \end{theorem} \begin{corollary} The number of $21$ patterns in the class of all $n$-permutations avoiding $123$ is given by $$ \num_{21}(\Av_n(123)) = \binom{n}{2} c_n + \binom{2n-1}{n} - 4^{n-1}.$$ \end{corollary} \subsection{Patterns of length $3$} We turn our attention now to patterns of length $3$, and provide a similar classification. To start, using the fact that $123$ is an involution provides some immediate identities, since inversion provides a natural map from the set $\Av_n(123)$ to itself. \begin{fact} In $\Av_n(123)$, $\num_{132} = \num_{213}$ and $\num_{231} = \num_{312}$. \end{fact} Numerical data and intuition suggest that $\num_{132} < \num_{231} < \num_{321}$. We begin by establishing some basic relationships between these numbers which will eventually combine to give us exact formulae. First, note that the total number of all length $k$ patterns is exactly $\binom{n}{k} c_n$. For $k = 3$ this gives the following fact. \begin{fact} \label{relation1} In the class $\Av_n(123)$ we have that $$ \num_{132} + \num_{213} + \num_{231} + \num_{312} + \num_{321} = \binom{n}{3} c_n.$$ Note that since $\num_{132} = \num_{213}$ and $\num_{312} = \num_{231}$, we can rewrite this as $$ 2 \num_{132} + 2 \num_{231} + \num_{321} = \binom{n}{3}c_n.$$ \end{fact} Our next uses Theorem \ref{inv} to provide another linear relationship between these three numbers. \begin{proposition} \label{relation2} In the class $\Av_n(123)$, we have $$ 4 \num_{132} + 2 \num_{231} = (n-2)\num_{12}.$$ \end{proposition} \begin{proof} Rewrite the equation as $$(n-2)\num_{12} - (\num_{132} + \num_{213}) = \num_{132}+\num_{213}+\num_{231}+\num_{312} .$$ Claim that both sides count the total number of length $3$ patterns which contain at least one $12$ pattern. Indeed, the right hand side counts all length $3$ patterns other than $321$. The left hand side first takes a $12$ pattern and adds another entry to it. However, this double counts each triple which has two $12$ patterns, and these are exactly the patterns $132$ and $213$. Subtracting these off yields the desired identity. \end{proof} Note that we now have two linear relationships between the three unknown quantities, and so some new information would completely solve the system. We summarize this in the following lemma. \begin{lemma}\label{linsys} Let $\mathcal{C} = \Av_n(123)$, and let $a_n = \num_{132}(\mathcal{C}) = \num_{213}(\mathcal{C})$, $b_n = \num_{231}(\mathcal{C}) = \num_{312}(\mathcal{C})$, and $d_n = \num_{321}(\mathcal{C})$. Then we have $$ \begin{array}{ll} 2a_n + 2b_n + d_n & = \binom{n}{3} c_n \\ 4a_n + 2b_n & = 4^{n-1} - \binom{2n-1}{n} \end{array}. $$ \end{lemma} We note that Proposition \ref{relation2} has a complementary analogue, obtained by counting inversions instead of noninversions. However, this leads to a relation which is linearly dependent on the first two. It takes a new approach to yield new information, which requires a few new definitions. \begin{definition} A permutation $p = p_1p_2 \ldots p_n$ is \emph{decomposable} (sometimes referred to as skew-decomposable) if there exists $k \in [n]$ such that for all $i \leq k$ and all $j < k$, we have that $p_i > p_j$. An \emph{indecomposable} permutation is one for which no such $k$ exists. \end{definition} \begin{definition} Denote the class of all indecomposable $123$ avoiding permutations by $\Av^* (123)$, and $\Av^* (123) \cap S_n$ by $\Av_n^*(123)$. \end{definition} In general, for simplicity of notation, indecomposability will be denoted with a star. Our first step, naturally, is to find the size of the set $\Av_n^*(123)$. \begin{proposition} For all $n \geq 1$, $$|\Av_n^*(123)| = \frac{1}{n} \binom{2n-2}{n-1} = c_{n-1}.$$ \end{proposition} \begin{proof} We know that $|\Av_n(123)| = c_n$, so $$ \sum_{n \geq 0} |\Av_n(123)| x^n = \frac{1 - \sqrt{1 - 4x}}{2x} = C(x).$$ Let $C^*(x) = \sum_{n\geq 1} |\Av_n^*(123)| x^n$. Every permutation in $\Av_n(123)$ can be expressed as a skew sum of indecomposable $123$ avoiding permutations, so it follows that $$ C(x) = 1 + C^*(x) + (C^*(x))^2 + (C^*(x))^3 + \ldots = \frac{1}{1-C^*(x)}.$$ Solving this algebraically gives that $C^*(x) = \frac{C(x)-1}{C(x)} = xC(x)$, which finishes the proof. \end{proof} Lemma \ref{linsys} now has an immediate indecomposable analogue, and the numbers $\num_q \left(\Av_n(123)\right)$ and $\num_q \left( \Av_n^*(123) \right)$ can be related relatively easily for each pattern $q$. However, this alone does not allow us to solve for an exact formula. Our new information will come from exactly counting the number of $213$ patterns in the set $\Av_n^*(123)$ by building a bijection to Dyck paths. We start by defining these paths, which are counted by the Catalan numbers. \begin{definition} A \emph{Dyck path} of length $2n$ (or of semilength $n$) is defined as a sequence of steps from the set $\{(1,1),(-1,1)\}$ which begins at $(0,0)$, ends at $(2n,0)$, and never steps below the line $x=0$. \end{definition} \begin{lemma} \label{biglemma} The generating function $A^*(x)$ for the number of $213$ patterns in $\Av_n^*(123)$ is given by $$ A^*(x) = \sum_{n\geq 0} \num_{213}(\Av_n^*(123)) = \frac{x^3C(x)}{(1-4x)^{3/2}} = \frac{x^2}{2(1-4x)^{3/2}} - \frac{x^2}{2(1-4x)}.$$ \end{lemma} \begin{proof} The proof consists of three parts: First, we examine the structure of permutations in $\Av_n^*(123)$, and find a simple way of counting the number of $213$ patterns. Second, we build a bijection onto Dyck paths which maps $213$ patterns to a path statistic. Finally, we find the weighted sum of all Dyck paths with respect to this statistic. Fix $n$, and let a permutation $p \in \Av_n^*(123)$. Note that since $p$ avoids $123$, $p$ can be viewed as a union of two descending sequences, so every entry in $p$ is a left-to-right minima or a right-to-left maxima, and by indecomposability no entry is both. Graph $p$ on an $n \times n$ lattice by plotting $(i,p(i))$ for each $i \in [n]$, and color each left-to-right minima red and each right-to-left maxima blue. Denote the sequence of red entries (ordered from left to right) by $\mathcal{R}=(r_1,r_2,\ldots r_j)$, and the sequence of blue entries by $\mathcal{B}=(b_1,b_2,\ldots b_k)$. Denote by $\Sp b_i$ the number of red entries below and to the left of $b_i$. Note that $\Sp b_i \geq 1$ for all $b_i$ by indecomposability. Now, we count the number of $213$ patterns in $p$. It follows that for any such pattern $q$, the $2$ entry and $1$ entry must be red, and the $3$ entry blue. It is also clear that each blue entry is contained in $\binom{\Sp b_i}{2}$ $213$ patterns. Therefore we have that $$ \num_{213} (p) = \sum_{i=1}^{k} \binom{\Sp b_i}{2}.$$ Now we are ready to build our bijection $\phi: \Av_n^*(123) \rightarrow \mathcal{D}_{n-1}$, where $\mathcal{D}_{n-1}$ denotes the set of Dyck paths of semilength $n-1$. From each blue vertex, extend a vertical line to the $x$-axis and a horizontal line to the $y$-axis, and color each point of intersection of these lines green. Define a path $P'$ from $(1,n)$ to $(n,1)$ by the following rules: \begin{enumerate}[1)] \item Begin by walking east from $(1,n)$ \item At a blue vertex, turn south and continue walking \item At a green vertex, turn eash and continue walking \item End at $(n,1)$ \end{enumerate} Rotate the path $P'$ by $\pi/4$ radians counter-clockwise to obtain a Dyck path $P$. This path is a slight modification of the path given by Krattenthaler's bijection \cite{krat01}, taking advantage of the indecomposability of the permutation to yield a more geometric description. This geometric interpretation of the bijection gives some additional insight into the number of $213$ patterns. \begin{figure}[ht] \centering \label{permtopath} \subfloat{ \fbox{ \begin{tikzpicture} [scale=.6, auto=center, every node/.style={circle, fill=black, inner sep=2pt, minimum size=2pt}] \foreach \x/\y in {1/4,2/7,3/6,4/2,5/5,6/1,7/3}{ \node (p-\x) at (\x,\y) {}; } \end{tikzpicture} } } \subfloat{ \fbox{ \begin{tikzpicture} [scale=.6, auto=center] \foreach \x/\y in {1/4,4/2,6/1}{ \node[circle,fill=black,inner sep=2pt, minimum size=2pt] (r\x) at (\x,\y) {}; } \foreach \num/\x/\y in {1/2/7,2/3/6,3/5/5,4/7/3}{ \node[circle,fill=black,inner sep=2pt, minimum size=2pt] (b\num) at (\x,\y) {};} \foreach \num/\x/\y in {1/1/7,2/2/6,3/3/5,4/5/3,5/7/1}{ \node[circle,fill=black, inner sep=0pt, minimum size=0pt] (g\num) at (\x,\y) {};} \foreach \to/\from in {g1/b1,b1/g2,g2/b2,b2/g3, g3/b3,b3/g4,g4/b4,b4/g5}{ \draw (\to) -- (\from);} \end{tikzpicture} } } \\ \subfloat{ \fbox{ \begin{tikzpicture} [scale=.5, auto=center] \draw (0,0) -- (1,1); \draw (1,1) -- (2,0); \draw (2,0) -- (3,1); \draw (3,1) -- (4,0); \draw (4,0) -- (6,2); \draw (6,2) -- (8,0); \draw (8,0) -- (10,2); \draw (10,2) -- (12,0); \end{tikzpicture} } } \caption{$\phi(4762513) = UDUDUUDDUUDD$} \end{figure} Note that each blue entry in $p$ produces a peak in $P$. Furthermore, $b_i$ corresponds to a peak of height $\Sp b_i$ above the $x$-axis in $P$. Therefore, if we let $h_{n,k}$ denote the total number of peaks of height $k$ in all Dyck paths of semilength $n$, we have that $$ \num_{213}(\Av_n^*(123)) =\sum_{k = 1}^{n-1} \binom{k}{2} h_{n-1,k}.$$ Finally, we can compute $H(x,u) = \sum_{n,k \geq 0} h_{n,k} x^n u^k$ as follows. First, note that since each Dyck path begins with an upstep it has a unique first point at which the path returns to the $x$-axis, so we can decompose each path $P$ of length $n$ into the concatenation of two shorter paths $Q$ and $R$. This gives that $P = uQdR$, where $u$ denotes an upstep and $d$ a downstep, and each peak of height $k-1$ in $Q$ and height $k$ in $R$ leads to a peak of height $k$ in $P$. With this in mind, we have the following generating function relation: $$ H(x,u) = ux(H(x,u)+1)C(x) + xH(x,u)C(x) .$$ Here the first term counts the peaks from the $uQd$ part, including the case when $Q$ is empty. The second term counts the contribution from the $R$ part. Rearranging leads to $$ H(x,u) = \frac{uxC(x)}{1-uxC(x)-xC(x)}.$$ Now, to count $213$ patterns, we need to count each peak with weight $\binom{k}{2}$. By taking derivatives twice with respect to $u$, setting $u=1$, dividing by two and scaling by $x$, we find that $$ \begin{aligned} \sum_{n,k \geq 0} \binom{k}{2} h_{n-1,k}x^n & = x \frac{\left. \partial_u ^2 H(x,u)\right|_{u=1}}{2} = \frac{x^3C(x)}{(1-4x)^{\frac32}} \\ & = x^3 + 7x^4 + 38x^5 + 187x^6 + 874x^7 + \ldots \ . \end{aligned} $$ The sequence $0,0,1,7,38,187\ldots$ is entry A000531 in the OEIS. Finally, the correspondence between peaks and $213$ patterns completes the proof. \end{proof} Now, it is relatively simple to move from the class of indecomposable $123$ avoiding permutations to the larger class of all $123$ avoiding permutations. \begin{theorem} \label{213pats} Let $a_n$ be the number of $213$ patterns in $\Av_n 123$. Then $$ \sum_{n\geq 0} a_n x^n = \frac{x^3C(x)^3}{(1-4x)^{3/2}} = \frac{x-1}{2(1-4x)} - \frac{3x-1}{2(1-4x)^{3/2}} .$$ \end{theorem} \begin{proof} Let $A(x)$ be the generating function for the numbers $a_n$, and let $A^*(x)$ denote the generating function for the number of $213$ patterns in \emph{indecomposable} $123$ avoiding permutations. Now, any permutation $p$ in $\Av(123)$ can be written uniquely as a skew sum of a nonempty indecomposable $123$ avoiding permutation $q$ and another, possibly empty, $123$ avoiding permutation $r$. Now, it is clear that any $213$ pattern in $p$ must be contained entirely in either $q$ or $r$. This leads to the following relation: $$ A(x) = A^*(x)C(x) + xC(x)A(x).$$ Solving for $A$ gives $$ A(x) = \frac{A^*(x) C(x)}{1-xC(x)} = C^2(x) A^*(x).$$ Lemma \ref{biglemma} now implies $$ A(x) = \frac{x^3 C(x)^3}{(1-4x)^{3/2}}.$$ \end{proof} Theorem \ref{213pats} combined with Lemma \ref{linsys} allows us to obtain both generating functions and exact formula for the occurrence of all length $3$ patterns in $\Av_n(123)$. We start with $231$ patterns, which reveal a striking connection to the class $\Av(132)$. \begin{corollary} \label{231pats} Let $b_n$ denote the number of $231$ (or $312$) patterns in all $123$ avoiding $n$-permutations. Then $$\sum_{n \geq 0} b_n x^n= \frac{3x-1}{(1-4x)^{2}} - \frac{4x^2 - 5x + 1}{(1-4x)^{5/2}}.$$ \end{corollary} \begin{proof} Let $B(x)$ be the generating function for the numbers $b_n$, let $A(x)$ be the generating function for the number of $213$ patterns, and let $j_n$ be the number of $12$ patterns with corresponding generating function $J(x)$. We know from Lemma \ref{linsys} that $$ 4A(x) + 2B(x) = \sum_{n \geq 0} (n-2) j_n x^n = (J(x)/x^2)' x^3 .$$ Solving this for $B(x)$ using elementary algebra and a bit of calculus yields $$ \begin{aligned} B(x) &= \frac{x^2C(x)^3}{(1 - 2xC(x))(1-4x)^{3/2}} \\ & = \frac{3x-1}{(1-4x)^{3/2}} - \frac{4x^2 - 5x + 1}{(1-4x)^{5/2}} \end{aligned}.$$ \end{proof} The connection between the classes $\Av(123)$ and $\Av(132)$ is now immediate. \begin{corollary} \label{bridge} Theorem \ref{231pats} together with Theorem \ref{bonasthm} imply immediately that the total number of $231$ patterns in $\Av_n (123)$ is equal to the total number of $231$ patterns in $\Av_n (132)$. \end{corollary} We can similarly apply Lemma \ref{linsys} to $321$ patterns. \begin{corollary} Let $d_n = f_{321}(\Av_n(123))$. Then we have that $$ \sum_{n\geq 0} d_n x^n = \frac{ 8x^3 - 20x^2 + 8x - 1}{(1-4x)^{2}} - \frac{36x^3 - 34x^2 + 10x - 1}{(1-4x)^{5/2}}. $$ \end{corollary} Before analyzing these generating functions, we note also that Lemma \ref{linsys} and its indecomposable analogue produce several other interesting identities. We summarize some of them here for completeness. \begin{corollary} The following identities hold. $$ \num_{21}(\Av_n(123)) = 2\num_{213}(\Av_n^*(123)) $$ $$ \num_{213}(\Av_n(123)) + \num_{231}(\Av_n(123)) = \num_{231} (\Av _{n-1}^*(123))$$ $$ C(x) \left(\sum_{n\geq 0} \num_{213}(\Av_n(123)) x^n \right) = x C'(x) \left(\sum_{n \geq 0} \num_{12}(\Av_n(123))x^n \right) $$ $$ \sum_{n\geq 0} \num_{213} (\Av_n^* (132) x^n) = \sum_{n \geq 0} \big(\num_{132}(\Av_n^*(123)) + \num_{231}(\Av_n^*(123))\big)x^n. $$ \end{corollary} Note that each of these identities are equivalent. That is, combined with Lemma \ref{linsys}, a combinatorial proof of any of them would imply all of the others (including Lemma \ref{biglemma}). Now we can do some analysis of the main sequences. Using some standard generating function analysis \cite{flajolet}, we find that the asymptotic growth of the number of length $3$ patterns are as follows: \vspace{1pc} $$ \num_{213} (\Av_n(123)) \sim \sqrt{\frac{n}{\pi}} 4^n$$ $$ \num_{231} (\Av_n(123)) \sim \frac{n}{2} 4^n $$ $$ \num_{321} (\Av_n(123)) \sim \frac{8}{3} \sqrt{\frac{n^3}{\pi}} 4^n. $$ We see that the three sequences each differ by a factor of approximately $\sqrt{n}$. Surprisingly, this is the same factor that the sequences $\num_{123}, \num_{231}, \num_{321}$ differ by in the class $\Av (132)$, as seen in \cite{bona10}. Each of these generating functions are simple enough that exact formulas can be obtained with relatively little hassle. One could argue that the asymptotic values are more interesting and provide more insight than the complicated formulas, but we present them here for completeness. \begin{corollary} Let $a_n = \num_{132}(\Av_n(123))$, $b_n = \num_{213}(\Av_n(123))$, and $d_n = \num_{321}(\Av_n(123))$. Then we have that $$ a_n = \frac{n+2}{4} \binom{2n}{n} - 3 \cdot 2^{2n-3} $$ $$ b_n = (2n-1) \binom{2n-3}{n-2} - (2n+1)\binom{2n-1}{n-1} + (n+4) \cdot 2^{2n-3}$$ $$ \begin{aligned} d_n = \frac{1}{6} \binom{2n+5}{n+1} \binom{n+4}{2} &- \frac{5}{3} \binom{2n+3}{n} \binom{n+3}{2} + \frac{17}{3} \binom{2n+1}{n-1} \binom{n+2}{2} \\ &- 6\binom{2n-1}{n-2} \binom{n+1}{2} - (n+1) \cdot 4^{n-1}. \end{aligned} .$$ \end{corollary} \subsection{Larger Patterns} Some of these same techniques are applicable to larger patterns. For example, we can easily modify Lemma \ref{linsys} to for patterns of all sizes. This leads to increasingy complicated expressions, but this simple idea can be used to prove the following proposition. \begin{proposition} Let $k, l \in \mathbb{Z}^+$, and $q$ be any permutation in $S_k$ other than the decreasing permutation. Then for $n$ large enough, we have that $$\num_{k \ldots 3 2 1}(\Av_n(123)) > \num_{q}(\Av_n(123)).$$ \end{proposition} \begin{proof} Let $T$ be the set of permutation in $S_k$ which are not the descending permutation. As in Fact \ref{relation1} and Proposition \ref{relation2}, we can express the number $(n-k+1)\num_{12}(\Av_n(123))$ as a positive linear combination of all of $\num_{q} (\Av_n(123))$ where $q \in T$, and we can express $\binom{n}{k} c_n$ as the sum of all $\num_r (\Av_n(123))$ where $r \in S_n$. It follows that there is a positive integer $m$ and positive integers $e_i$ such that $$ \binom{n}{k} c_n - (n-m+1) \num_{12} (\Av_n(123)) = \num_{k \ldots 321} - \sum_{q \in T} e_i \num_q (\Av_n(123)).$$ Asymptotic analysis shows that the left hand side is eventually positive, and so the first term on the right side eventually outgrows the second term, which completes the proof. \end{proof} \section{Further Directions} The numbers $\num_q (\Av (p))$ for permutations $p,q$ exhibit numerous symmetries and produce many new questions. All of the generating presented here and in \cite{bona10} are almost rational, in the sense that they lie in the ring $\mathbb{Q}(x,\sqrt{1-4x})$. This allows for easy asymptotic analysis, and leaves open the possibility of bijections to other Catalan-related objects. Building on what was mentioned in \cite{bona12}, we have instances of the same sequence of numbers which correspond sums of statistics with different distributions in objects counted by the Catalan numbers. Do these sequences and statistics have anologues in other such objects? Thus far, to the author's knowledge, the expectation of patterns has only been studied for the classes $\Av (123)$ and $\Av (132)$ (and their symmetries). Applying these ideas to more general classes could yield similarly interesting identities. Note that the increasing and decreasing permutation do not always provide the opposite extreme cases: it is simple to show that $\num_{123}(\Av (2413)) = \num_{321}(\Av(2413))$. This leads to the natural question: in the set of permutations avoiding a specific pattern (or a set of patterns), can we easily determine what pattern is most common? And how large can the difference be between the most and least common? Finally, are there other occurences of the same sequence of patterns arising in different classes? Or within the same class, as in $\Av_n (132)$? Taking this to the extreme, is there a proper, non-trivial permutation class for which each pattern of a given length is equally as common, as in the class of all permutations? \nocite{*} \bibliographystyle{plain}
{ "timestamp": "2012-07-13T02:04:52", "yymm": "1206", "arxiv_id": "1206.0320", "language": "en", "url": "https://arxiv.org/abs/1206.0320", "abstract": "In the set of all patterns in $S_n$, it is clear that each k-pattern occurs equally often. If we instead restrict to the class of permutations avoiding a specific pattern, the situation quickly becomes more interesting. Miklós Bóna recently proved that, surprisingly, if we consider the class of permutations avoiding the pattern 132, all other non-monotone patterns of length 3 are equally common. In this paper we examine the class $\\Av (123)$, and give exact formula for the occurrences of each length 3 pattern. While this class does not break down as nicely as $\\Av (132)$, we find some interesting similarities between the two and prove that the number of 231 patterns is the same in each.", "subjects": "Combinatorics (math.CO)", "title": "Expected Patterns in Permutation Classes", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9869795102691454, "lm_q2_score": 0.8289388040954683, "lm_q1q2_score": 0.8181456149092363 }
https://arxiv.org/abs/1609.02210
Enumerating cycles in the graph of overlapping permutations
The graph of overlapping permutations is a directed graph that is an analogue to the De Bruijn graph. It consists of vertices that are permutations of length $n$ and edges that are permutations of length $n+1$ in which an edge $a_1\cdots a_{n+1}$ would connect the standardization of $a_1\cdots a_n$ to the standardization of $a_2\cdots a_{n+1}$. We examine properties of this graph to determine where directed cycles can exist, to count the number of directed $2$-cycles within the graph, and to enumerate the vertices that are contained within closed walks and directed cycles of more general lengths.
\section{Introduction} In this paper we will discuss an analogue to a classical object in combinatorics, De Bruijn graphs. The graphs in this paper are all directed graphs with arcs (edges with an orientation) between vertices. For a set $\{0,1,\ldots,q-1\}$, let $\{0,1,\ldots,q-1\}^n$ be the set of all strings of length $n$ with elements from $\{0,1,\ldots,q-1\}$. A De Bruijn graph has the vertex set $\{0,1,\ldots,q-1\}^n$ and a directed edge from each vertex $x_1x_2\cdots x_{n}$ to the vertex $x_2x_3\cdots x_{n+1}$. In other words, there is an edge from ${\bf a}$ to ${\bf b}$ if and only if the last $n-1$ elements of~${\bf a}$ and the first $n-1$ elements of ${\bf b}$ are the same. One of the properties of De Bruijn graphs with vertex set $\{0,1,\ldots,q-1\}^n$ that has been studied is the number of directed cycles of length $k$, for $k\leq n$, which is shown in~\cite{golomb}. For ease of notation, we will call a directed cycle of length~$k$ a $k$-cycle. We will examine the number of $k$-cycles in the graph of overlapping permutations, $G(n)$, which was introduced in~\cite{CDG} and was studied in~\cite{Kitaev1} and Chapter 5 of~\cite{Kitaev2}. The existence of overlapping cycles on a similar graph can also be seen in~\cite{HoranHurlbert}. For a permutation ${\bf a}=a_1a_2\cdots a_n$, we denote the \textit{standardization} of the substring $a_{s+1}a_{s+2}\cdots a_{s+t}$ by ${\text{st}}(a_{s+1}a_{s+2}\cdots a_{s+t})=b_1b_2\cdots b_{t}$ where $b_i\in \{1,\ldots,t\}$ with $b_i<b_j$ if and only if $a_{s+i}< a_{s+j}$ for all $1\leq i,j\leq t$. Let $G(n)$ be the graph with vertex set as the set of all permutations of length $n$ and consisting of a directed edge from each vertex ${\bf a}=a_1a_2\cdots a_n$ to ${\bf b}=b_1b_2\cdots b_n$ if ${\text{st}}(a_2a_3\cdots a_n)={\text{st}}(b_1b_2\cdots b_{n-1})$. This adjacency condition is why this is called a graph of overlapping permutations and is analogous to the overlapping condition of strings in the De Bruijn graph. Be aware that $G(n)$ can have multiple directed edges between the same vertices and in the same direction. Enumerating the cycles in the graph of overlapping permutations was first attempted by Ehrenborg, Kitaev, and Steingr\'\i msson in~\cite{EKS}. The authors focused on the graph $G(n,312)$, which is the graph of overlapping permutations where the vertices and edges are permutations which avoid the pattern $312$, i.e., a subsequence of three entries in the permutations of the vertices and edges whose standardization is $312$. They determined the number of closed walks of length $k$ for $k\leq n$, as well as the number of $k$-cycles. See Theorems 5.1 and 5.2 in~\cite{EKS} for these results. An important observation for these enumerations is that they do not depend upon $n$. This is not the case for $G(n)$, as increasing the length of the permutations will increase the number of $k$-cycles when looking at the entire graph $G(n)$ instead of a subgraph that avoids certain patterns. In the following section, we will discuss some pertinent graph theory terminology and present an example of $G(3)$. In Section~\ref{properties}, we will introduce conditions on which a closed walk exists at a particular vertex, as well as discuss the existence of two edges between the same two vertices and multiple closed walks stemming from a single vertex. In Section~\ref{2cycles}, we count the $2$-cycles in $G(n)$. Section 5 includes enumerations for the number of vertices contained within closed walks and a correspondence between the number of closed $k$-walks in $G(n)$ and the number of $k$-cycles in $G(n)$ as long as $k$ is prime. Finally, Section 6 consists of further research questions involving cycles within the graph of overlapping permutations. \section{Preliminaries and an Example} Before we introduce our results, we must first establish some graph theory terminology. The following definitions are for directed graphs. We consider a sequence $(v_1, e_1, v_2, e_2, v_3,\ldots, v_{k},e_k, v_{k+1})$ of vertices $v_i$ and edges $e_i=(v_i,v_{i+1})$ to be a \textit{walk} of length $k$, or a $k$-walk. We will only list the vertices in the walk for simplification within our proofs, although we do consider two walks (and two cycles) to be distinct if the sequence of vertices are the same, but edges differ due to pairs of vertices being connected by multiple edges. If $v_1=v_{k+1}$, the walk is called a \textit{closed walk} and will usually be written as the sequence $(v_1, v_2,\ldots, v_k)$ to represent the closed $k$-walk. If the vertices are all distinct, the closed $k$-walk is a \textit{$k$-cycle}. For more information about graph terminology not mentioned in this article, see~\cite{west}. When counting closed walks, we consider the closed walk $(v_1, e_1, v_2, e_2, v_3,\ldots, v_{k},e_k, v_{1})$ to be the same as the closed walk $(v_j, e_j, v_{j+1}, e_{j+1}, v_{j+2},\ldots, v_{k},e_k, v_{1},e_1,v_2,\ldots, v_{j-1},e_{j-1},v_j)$ that has a different starting vertex. Likewise, they are counted as one cycle if the vertices are distinct. More formally, we are attempting to count equivalence classes of closed walks (and cycles) by considering shifted sequences of vertices and edges within a closed walk to be equivalent. Given a vertex ${\bf a}=a_1 a_2\cdots a_n$, we consider two important vertices related to it. First, the \textit{complement} of ${\bf a}$ is the vertex $\overline{{\bf a}}=(n+1-a_1) (n+1-a_2)\cdots (n+1-a_n)$. Second, we define the \textit{cyclic shift} of the vertex ${\bf a}$ by $\sigma({\bf a})=a_2 a_3\cdots a_n a_1$. Figure 1 displays the graph $G(3)$ with permutations of length $3$ as the six vertices. The twenty-four edges, although unlabeled in the figure, correspond to the permutations of length $4$. Recall the definition of the edges in $G(n)$ as $c_1\cdots c_{n+1}$ connecting the standardization ${\bf a}=a_1\cdots a_n={\text{st}}(c_1\cdots c_n)$ to the standardization ${\bf b}=b_1\cdots b_n={\text{st}}(c_2\cdots c_{n+1})$. This directly implies there is an edge from ${\bf a}$ to~${\bf b}$ if and only if ${\text{st}}(a_2\cdots a_n)={\text{st}}(b_1\cdots b_{n-1})$. Hence we see in this example that there exists an edge from $123$ to $132$ since ${\text{st}}(23)={\text{st}}(13)=12$, whereas there is no returning edge from $132$ to $123$ since ${\text{st}}(32)=21\neq {\text{st}}(12)=12$. The number of cycles in the graph are as follows: two $1$-cycles at the trivial vertices $123$ and $321$, six $2$-cycles, and twenty-six $3$-cycles. Observe that with $2$-cycles some pairs of vertices create two $2$-cycles because there are multiple edges between the same two vertices. An example is the cycle $(132, 213)$ made from either the edge pairs $1324$ and $2143$ or the pair $1324$ and $3142$. As many as eight $3$-cycles can be made from a triple of vertices, which occurs with the $3$-cycle $(132, 321, 213)$, which can make the enumeration of long cycles rather difficult. Additionally, note that vertices can be included in multiple cycles with differing vertices. For instance, the vertex $231$ is contained in the $2$-cycles $(231, 312)$ and $(231,213)$. A final observation is that the trivial vertices are within $k$-cycles for each $1\leq k\leq 6$ except for $k=2$. The fact that they are not in $2$-cycles will be generalized in the next section. \begin{figure}\label{G3} \begin{center} \includegraphics[scale=1]{G3.png} \caption{The graph of overlapping permutations of length $3$, $G(3)$} \end{center} \end{figure} \section{Properties of Closed Walks and Cycles in $G(n)$}\label{properties} We will now introduce properties regarding the inclusion of vertices within closed walks and cycles. Unless otherwise noted, assume $k$ is an integer with $2\leq k\leq n-1$. We first establish a necessary condition for the existence of a closed $k$-walk through a vertex. \begin{theorem}\label{cycleNecessary} If a vertex ${\bf a}=a_1a_2\cdots a_n$ is in some closed $k$-walk in $G(n)$, then \[ {\text{st}}(a_1a_2\cdots a_{n-k})={\text{st}}(a_{k+1}a_{k+2}\cdots a_n). \] \end{theorem} \begin{proof} Assume ${\bf a}=a_1a_2\cdots a_n$ is a vertex in the closed $k$-walk $\left({\bf a}, {\bf a}^{(2)}, \ldots, {\bf a}^{(k)}\right)$ where ${\bf a}^{(i)}=a^{(i)}_1a^{(i)}_2\cdots a^{(i)}_n$. The directed edge from ${\bf a}$ to ${\bf a}^{(2)}$ in the walk above gives us the equality ${\text{st}}(a_2\cdots a_n)={\text{st}}\left(a^{(2)}_1\cdots a^{(2)}_{n-1}\right)$, which when narrowing down to the last $n-k$ elements of ${\bf a}$ becomes $${\text{st}}(a_{k+1}\cdots a_n)={\text{st}}\left(a^{(2)}_k\cdots a^{(2)}_{n-1}\right).$$ Then we use the second directed edge in this walk, which provides the equality ${\text{st}}\left(a^{(2)}_2\cdots a^{(2)}_{n}\right)={\text{st}}\left(a^{(3)}_1\cdots a^{(3)}_{n-1}\right)$. By combining this with the previous equality we attain $${\text{st}}(a_{k+1}\cdots a_n)={\text{st}}\left(a^{(3)}_{k-1}\cdots a^{(3)}_{n-2}\right).$$ Continuing this reasoning for the subsequent edges leading up to the vertex ${\bf a}^{(k)}$, we have $${\text{st}}(a_{k+1}\cdots a_n)={\text{st}}\left(a^{(k)}_2\cdots a^{(k)}_{n-k+1}\right).$$ The final edge from ${\bf a}^{(k)}$ to ${\bf a}$ in the cycle implies ${\text{st}}\left(a^{(k)}_2\cdots a^{(k)}_{n}\right)={\text{st}}(a_1\cdots a_{n-1})$. Focusing on the standardizations of length $n-k$, we combine the previous two equalities to obtain the desired result: $${\text{st}}(a_{k+1}\cdots a_n)={\text{st}}\left(a^{(k)}_2\cdots a^{(k)}_{n-k+1}\right)={\text{st}}(a_1\cdots a_{n-k}).$$ \end{proof} \begin{corollary}\label{secondVertex} For a closed $k$-walk $\left({\bf a}^{(1)}, {\bf a}^{(2)}, \ldots, {\bf a}^{(k)}\right)$ with ${\bf a}^{(i)}=a^{(i)}_1a^{(i)}_2\cdots a^{(i)}_n$, the vertex $a^{(2)}$ satisfies the following equality: ${\text{st}}\left(a^{(2)}_k\cdots a^{(2)}_n\right)={\text{st}}\left(a^{(1)}_1\cdots a^{(1)}_{n-k+1}\right)$. \end{corollary} \begin{proof} Following the same logic as the proof of Theorem~\ref{cycleNecessary}, we begin with the left-hand side and use subsequent edges to obtain the chain of equalities $${\text{st}}\left(a^{(2)}_k\cdots a^{(2)}_n\right)={\text{st}}\left(a^{(3)}_{k-1}\cdots a^{(3)}_{n-1}\right)=\cdots ={\text{st}}\left(a^{(k)}_{2}\cdots a^{(k)}_{n-k+2}\right)={\text{st}}\left(a^{(1)}_1\cdots a^{(1)}_{n-k+1}\right).$$ \end{proof} Before we prove a sufficient condition for when a vertex is contained within a closed $k$-walk, we first provide notation for two standardizations of substrings of a vertex ${\bf a}=a_1\cdots a_n$. We define $$y_{{\bf a}}=y_1\cdots y_{n-1}={\text{st}}(a_2\cdots a_n)$$ and $$z_{{\bf a}}=z_1\cdots z_{n-k+1}={\text{st}}(a_1\cdots a_{n-k+1}).$$ If ${\bf b}=b_1\cdots b_n$ is the second vertex in a closed $k$-walk, it clearly must satisfy ${\text{st}}(b_1\cdots b_{n-1})=y_{{\bf a}}$ to have an edge from ${\bf a}$ to ${\bf b}$, and it also must have ${\text{st}}(b_k\cdots b_n)=z_{{\bf a}}$ according to Corollary~\ref{secondVertex}. To clarify this point, Figure~\ref{yaza} provides an example of vertices ${\bf a}$ and ${\bf b}$ when the permutations have length $n=7$ and $k=4$ along with $y_{\bf a}$ and~$z_{\bf a}$ which are placed over or under their corresponding $a_i$ and $b_i$ values. These standardizations will be used frequently in the upcoming results of this section and as such we will use the next result without citing it in many of our results. \begin{figure}[htb] \begin{center} \includegraphics[width=2in]{yaza.pdf} \end{center} \caption{Permutations ${\bf a}$ and ${\bf b}$ with their corresponding $y_{\bf a}$ and $z_{\bf a}$}\label{yaza} \end{figure} \begin{theorem}\label{walkExistence} Let ${\bf a}=a_1a_2\cdots a_n$ be a vertex in $G(n)$. If ${\text{st}}(a_1a_2\cdots a_{n-k})={\text{st}}(a_{k+1}a_{k+2}\cdots a_n)$ then there exists a closed $k$-walk starting at ${\bf a}$. \end{theorem} \begin{proof} We first find a vertex ${\bf a}^{(2)}=b_1\cdots b_n$ that is the second vertex on the closed $k$-walk starting at ${\bf a}$. We must have ${\text{st}}(b_1\cdots b_{n-1})=y_{{\bf a}}={\text{st}}(a_2\cdots a_n)$ and ${\text{st}}(b_k\cdots b_n)=z_{{\bf a}}={\text{st}}(a_1\cdots a_{n-k+1})$. To determine ${\bf a}^{(2)}$, we pursue one of the following procedures depending on the last element of $z_{{\bf a}}$: \begin{itemize} \item[$(A)$] If $z_{n-k+1}=1$, then set $b_n=1$ and $b_i=y_{i}+1$ for $i=1,\ldots, n-1$. \item[$(B)$] Otherwise, let $\ell$ be the index with $z_{\ell}=z_{n-k+1}-1$. Set $b_i=y_{i}$ for all indices $i$ with $y_{i}\leq y_{\ell+k-1}$, $b_n=y_{\ell+k-1}+1$, and if $z_{n-k+1}<n-k+1$, set $b_j=y_{j}+1$ for all remaining elements. \end{itemize} In Case $(A)$, ${\text{st}}(b_1\cdots b_{n-1})=y_{{\bf a}}$ since each $b_i$ is simply attained by increasing $y_i$ by $1$. We also have ${\text{st}}(b_1\cdots b_{n-1})=y_{{\bf a}}$ in Case $(B)$ because the values within ${\bf a}^{(2)}$ and $y_{{\bf a}}$ are identical from $1$ to $y_{\ell+k-1}$ and are attained by increasing $y_i$ by $1$ for the remaining elements in ${\bf a}^{(2)}$. Because of the assumption of ${\text{st}}(a_1a_2\cdots a_{n-k})={\text{st}}(a_{k+1}a_{k+2}\cdots a_n)$, we have ${\text{st}}(y_{k}\cdots y_{n-1})={\text{st}}(z_1\cdots z_{n-k})$. Therefore, ${\text{st}}(b_k\cdots b_{n-1})={\text{st}}(z_{1}\cdots z_{n-k})$ since we already demonstrated that ${\text{st}}(b_1\cdots b_{n-1})=y_{{\bf a}}$. If Case $(A)$ was used, we can extend this standardization equality to be ${\text{st}}(b_k\cdots b_n)=z_{{\bf a}}$ since both $b_n$ and $z_{n-k+1}$ are the smallest elements in these strings, respectively. In Case $(B)$ with $z_{n-k+1}=n-k+1$, we obtain the same equality since $b_n$ and $z_{n-k+1}$ are the largest elements in those strings. Finally, in Case $(B)$ with $z_{n-k+1}<n-k+1$, if $i_1$ and $i_2$ are the indices with $z_{i_1}+1=z_{n-k+1}=z_{i_2}-1$, then $b_n$ is assigned so that $b_{i_1}<b_n<b_{i_2}$. These inequalities imply that the standardization equality can also be extended in this case to ${\text{st}}(b_k\cdots b_n)=z_{{\bf a}}$. Thus, each case maintains both of the required standardization equalities involving $y_{{\bf a}}$ and $z_{{\bf a}}$. We now have that there is an edge from ${\bf a}$ to ${\bf a}^{(2)}$, and it remains to show that we can continue with this walk and return to ${\bf a}$ after a total of $k$ steps. The procedure outlined above, using $y_{{\bf a}^{(2)}}$ and $z'_{{\bf a}}=z_1z_2\cdots z_{n-k+1}z_{n-k+2}$ where $z_{n-k+1}$ is replaced with $z_{n-k+2}$ in the breakdown of Cases $(A)$ or $(B)$, can be applied again to find a vertex ${\bf a}^{(3)}$. Let ${\bf a}^{(t)}=a_1^{(t)}a_2^{(t)}\cdots a_n^{(t)}$ for each $t\in\{3,\ldots,k\}$. There is an edge connecting ${\bf a}^{(2)}$ to ${\bf a}^{(3)}$ in which ${\text{st}}(a^{(3)}_1\cdots a^{(3)}_{n-1})=y_{{\bf a}^{(2)}}$ and ${\text{st}}(a^{(3)}_{k-1}\cdots a^{(3)}_n)={\text{st}}(a_1\cdots a_{n-k+2})$. We continue this process of determining a vertex ${\bf a}^{(t)}$ such that ${\text{st}}(a^{(t)}_1\cdots a^{(t)}_{n-1})=y_{{\bf a}^{(t-1)}}$ and ${\text{st}}(a^{(t)}_{k-t+2}\cdots a^{(t)}_n)={\text{st}}(a_1\cdots a_{n-k+t-1})$. When we reach $t=k$, we have that the vertex ${\bf a}^{(k)}$ satisfies ${\text{st}}(a^{(k)}_{2}\cdots a^{(k)}_n)={\text{st}}(a_1\cdots a_{n-1})$. This implies there is a directed edge from ${\bf a}^{(k)}$ to ${\bf a}$, completing the closed $k$-walk $({\bf a}, {\bf a}^{(2)}, {\bf a}^{(3)}, \ldots, {\bf a}^{(k)})$. \end{proof} \begin{corollary}\label{finishWalk} For vertices ${\bf a}=a_1a_2\cdots a_n$ and ${\bf b}=b_1b_2\cdots b_n$, if ${\text{st}}(a_{t+1}\cdots a_n)={\text{st}}(b_1\cdots b_{n-t})$, then there is a $t$-walk from ${\bf a}$ to ${\bf b}$. \end{corollary} \begin{proof} The $t$-walk $({\bf a}, {\bf a}^{(2)},\ldots,{\bf a}^{(t)},{\bf b})$ can be created by following a similar process as described in Theorem~\ref{walkExistence}. The cases for finding the second vertex are determined by the last element of ${\text{st}}(b_{1}\cdots b_{n-t})$ similar to when using Cases $(A)$ and $(B)$ in Theorem~\ref{walkExistence}. \end{proof} Note that by the symmetric property of permutations the following result holds. \begin{proposition}\label{cycleComplement} $({\bf a}^{(1)},{\bf a}^{(2)},\ldots,{\bf a}^{(k)})$ is a closed $k$-walk in $G(n)$ if and only if $(\overline{{\bf a}^{(1)}},\overline{{\bf a}^{(2)}},\ldots,\overline{{\bf a}^{(k)}})$ is a closed $k$-walk. \end{proposition} Notice that Theorems~\ref{cycleNecessary} and~\ref{walkExistence} are converses of each other. However, Theorem~\ref{cycleNecessary} (and Proposition~\ref{cycleComplement}) can be stated for $k$-cycles since each $k$-cycle is by definition a closed $k$-walk, but Theorem~\ref{walkExistence} cannot be extended to guarantee the existence of a $k$-cycle. Part of the next result is an example of why such a cycle might not exist. \begin{theorem}\label{trivialVertices} Let ${\bf a}=1\,2\cdots n$ and ${\bf b}=n\,(n-1)\cdots 1$ be vertices in $G(n)$. Then \begin{enumerate}[(a)] \item the vertices ${\bf a}$ and ${\bf b}$ are the only vertices contained in a $1$-cycle, and \item the vertices ${\bf a}$ and ${\bf b}$ are not contained within a $k$-cycle for $2\leq k<n$. \end{enumerate} \end{theorem} \begin{proof} For part $(a)$, Theorem~\ref{cycleNecessary} implies when $k=1$ that the first $n-1$ and last $n-1$ elements of a vertex in a $1$-cycle must have the same standardization. This property is only true for the two trivial vertices ${\bf a}$ and ${\bf b}$. For part $(b)$, by Proposition~\ref{cycleComplement}, we only need to consider ${\bf a}=1\,2\cdots n$ since ${\bf b}=\overline{{\bf a}}$. The vertex ${\bf a}$ clearly satisfies the assumption of Theorem~\ref{walkExistence}; hence, there is a closed $k$-walk starting at ${\bf a}$. If we assume ${\bf c}=c_1\cdots c_n$ is the second vertex of this walk, then since $y_{{\bf a}}=1\cdots (n-1)$ and $z_{{\bf a}}=1\cdots (n-k+1)$, we get ${\text{st}}(c_1\cdots c_{n-1})=1\cdots (n-1)$ and ${\text{st}}(c_k\cdots c_n)=1\cdots (n-k+1)$. Since $k<n$, these two standardizations overlap and imply that ${\bf c}=1\,2\cdots n$. Thus, the closed $k$-walk is simply a $1$-cycle repeated $k$ times instead of a $k$-cycle. \end{proof} The trivial permutations are not the only vertices which satisfy the sufficiency condition for a closed $k$-walk and are not included in a $k$-cycle, as shown in the following example. \begin{example}{\rm Consider the vertex ${\bf a}=162534$. For $k=4$, ${\bf a}$ satisfies the condition in Theorem~{\normalfont\ref{walkExistence}} of ${\text{st}}(16)={\text{st}}(34)$. This vertex is not included within a $4$-cycle though, as the only closed $4$-walk that contains it is the repetition of the $2$-cycle $(162534, 615243)$.} \end{example} To attempt to count how many $k$-cycles are in $G(n)$, we first need to examine how many closed walks or cycles can be created through a given vertex. The following results will provide conditions for when a vertex is adjacent to multiple vertices whose directed edges begin closed $k$-walks. \begin{theorem}\label{multipleVertices} Let $k<n<2k$ and ${\bf a}=a_1\cdots a_n$ be a vertex in $G(n)$ contained within at least one closed $k$-walk. Let $2\leq m\leq k$ be an integer. If $z_{n-k+1}=1$ and the elements in $\{1,\ldots, m-1\}$ are within positions $1,\ldots ,k-1$ of $y_{{\bf a}}$, or likewise if $z_{n-k+1}=n-k+1$ and $\{n-m+2,\ldots, n\}$ are within positions $1,\ldots ,k-1$ of $y_{{\bf a}}$, then the vertex ${\bf a}$ is contained in $m$ closed $k$-walks with distinct vertices adjacent to ${\bf a}$. \end{theorem} \begin{proof} We will prove the result for the case of $y_{1},\ldots, y_{k-1}$ including $1,\ldots, m-1$, and the other case follows by symmetry. Assume ${\bf a}$ is a vertex in a closed $k$-walk $\left({\bf a}, {\bf a}^{(2)}, {\bf a}^{(3)}, \ldots, {\bf a}^{(k)}\right)$, and let ${\bf a}^{(2)}=b_1b_2\cdots b_n$. By Corollary~\ref{secondVertex}, we have ${\text{st}}(b_1\cdots b_{n-1})=y_{{\bf a}}=y_1\cdots y_{n-1}$ and ${\text{st}}(b_k\cdots b_n)=z_{{\bf a}}=z_1\cdots z_{n-k+1}$. Let $i_1,\ldots, i_{m-1}$ be the indices containing the smallest $m-1$ elements of $y_1\cdots y_{n-1}$. Then the positions $b_{i_1},\ldots, b_{i_{m-1}}$ are the smallest elements in $b_1\cdots b_{k-1}$, and $b_n$ is the smallest element in $b_{k}\cdots b_n$ since $z_1\cdots z_{n-k+1}={\text{st}}(b_k\cdots b_n)$. Since these portions of ${\bf a}^{(2)}$ do not overlap, we can choose $b_n$ to be any element in $\{1,\ldots, m\}$ and place the remaining elements from $\{1,\ldots,m\}$ in the positions $b_{i_1},\ldots, b_{i_{m-1}}$ according to the standardization $y_1\cdots y_{n-1}$. The other elements larger than~$m$ will be fixed based on the standardization $y_1\cdots y_{n-1}$. With $m$ different options for the vertex ${\bf a}^{(2)}$, we obtain $m$ distinct closed $k$-walks that include the vertex ${\bf a}$ by following the process in the proof of Theorem~\ref{walkExistence} to complete the closed walk. \end{proof} The assumption of $k<n<2k$ in the previous theorem is necessary for the existence of the conditions on the standardizations of ${\bf a}=a_1a_2\cdots a_n$. In the case of $n\geq 2k$, $z_{n-k+1}=1$ implies that $a_{n-k+1}<a_i$ for all $i<n-k+1$. If $y_j=1$ for some $j\in \{1,\ldots k-1\}$, then $a_j<a_{n-k+1}$ since $n-k+1\geq 2k-k+1=k+1$, which is a contradiction. Theorem~\ref{multipleVertices} only guarantees the existence of $m$ closed $k$-walks rather than $k$-cycles, because it is possible that some of these $k$-walks are combinations of shorter cycles, as shown in the following example. \begin{example}\label{multipleWalks}{\rm For $n=5$ and $k=4$, the vertex $21435$ satisfies the assumptions of Theorem~\ref{multipleVertices} since $z_2=1$ and $y_{\bf a}=1324$, resulting in $m=4$ distinct closed $4$-walks that include this vertex. However, one of these is a repetition of the $2$-cycle $(21435,13254)$ and only $3$ are $4$-cycles: $(21435, 14253, 31425, 13254)$, $(21435, 14352, 32415, 23154)$, and $(21435, 24351, 32415, 23154)$.} \end{example} The vertices satisfying the conditions in Theorem~\ref{multipleVertices} are not the only vertices that are contained in multiple closed walks. The following result provides further conditions for this situation. An example of a vertex that fulfills the criteria for this next theorem is included in Example~\ref{manyWalks}. \begin{theorem}\label{otherMultipleVertices} Let ${\bf a}=a_1 a_2\cdots a_n$ be a vertex in $G(n)$ that is contained within at least one closed $k$-walk, and let $2\leq m\leq k$. Assume there exists some indices $i, j, \ell_1,\ell_2,\ldots, \ell_{m-1}$ with $k\leq i,j\leq n-1$ and $1\leq \ell_1,\ell_2,\ldots, \ell_{m-1}\leq k-1$ for which $y_i= y_{\ell_1}-1=y_{\ell_2}-2=\ldots= y_{\ell_{m-1}}-(m-1)= y_j-m$ and $z_{i-k+1}+1= z_{n-k+1}= z_{j-k+1}-1$. Then ${\bf a}$ is contained in $m$ closed $k$-walks with distinct vertices adjacent to ${\bf a}$. \end{theorem} \begin{proof} Let $({\bf a},{\bf a}^{(2)},{\bf a}^{(3)},\ldots,{\bf a}^{(k)})$ be a closed $k$-walk containing ${\bf a}$, and let \[ {\bf a}^{(2)}=b_1b_2\cdots b_n. \] Then by Corollary~\ref{secondVertex} we have $y_{\bf a}={\text{st}}(b_1b_2\cdots b_{n-1})$ and $z_{{\bf a}}={\text{st}}(b_{k}b_{k+1}\cdots b_n)$. Following the proof of Theorem~\ref{walkExistence}, since $z_{n-k+1}\neq 1$, we would use Case $(B)$ with $\ell=i-k+1$. This results in a vertex ${\bf a}^{(2)}$ with $b_s=y_{s}$ for all $s\in\{1,\ldots, n-1\}\setminus \{\ell_1,\ldots,\ell_{m-1},j\}$, $b_n=y_i+1$, and $b_s=y_{s}+1$ for $s\in \{\ell_1,\ldots,\ell_{m-1},j\}$. Based on our assumptions for $y_{{\bf a}}$ and $z_{{\bf a}}$, we have $b_{i}\leq b_{\ell_1}\leq\cdots \leq b_{\ell_{m-1}}\leq b_{j}$ and $b_{i}\leq b_n\leq b_{j}$. However, there is no necessary inequality relating $b_n$ with $b_{\ell_1},\ldots, b_{\ell_{m-1}}$. Thus, we can permute the sequence $(b_{\ell_1},\ldots, b_{\ell_{m-1}},b_n)$ to result in $m-1$ different sequences of the form $(c_{\ell_1},\ldots, c_{\ell_{m-1}},c_n)$ that maintains the inequalities $c_{\ell_1}\leq\cdots \leq c_{\ell_{m-1}}$ with the element $c_n=b_{\ell_r}$ for some $1\leq r\leq m-1$. As such, there is a directed edge from ${\bf a}$ to each of the distinct $m-1$ vertices of the form ${\bf a}^{(2)}=d_1d_2\cdots d_n$ where $$d_s=\begin{cases} b_s \hspace{.5cm} \text{if }s\notin \{\ell_1,\ldots, \ell_{m-1}, n\}\\ c_s \hspace{.5cm} \text{if }s\in \{\ell_1,\ldots, \ell_{m-1}, n\}.\\ \end{cases} $$ Since the equality $c_n=b_{\ell_r}$ satisfies $b_{i}\leq c_n\leq b_{j}$, swapping $b_n$ with the elements $b_{\ell_r}$ won't change the fact that ${\text{st}}(d_{k}\cdots d_n)={\text{st}}(a_1\cdots a_{n-k+1})$. Thus, by Corollary~\ref{finishWalk} we can find the remaining $k-2$ vertices to complete a closed $k$-walk from each of those $m-1$ vertices, giving us a total of $m$ closed $k$-walks with distinct second vertices. \end{proof} As shown in the next theorem, the two conditions in Theorems~\ref{multipleVertices} and~\ref{otherMultipleVertices} are the only way for ${\bf a}$ to be contained in multiple closed $k$-walks with differing vertices adjacent to ${\bf a}$. \begin{theorem}\label{reverse} Let $2\leq m\leq k$. Let ${\bf a}=a_1 a_2\cdots a_n$ be a vertex in $G(n)$ that is contained within $m$ closed $k$-walks with distinct elements adjacent to ${\bf a}$. Then one of the following is true regarding the standardizations $y_{{\bf a}}$ and $z_{{\bf a}}$: \begin{enumerate}[(a)] \item $z_{n-k+1}=1$ and $\{1,\ldots, m-1\}$ are within positions $1,\ldots ,k-1$ of $y_{{\bf a}}$, or $z_{n-k+1}=n-k+1$ and $\{n-m+2,\ldots, n\}$ are within positions $1,\ldots ,k-1$ of $y_{{\bf a}}$, \item There exists some $i, j, \ell_1,\ell_2,\ldots, \ell_{m-1}$ with $k-1<i,j\leq n-1$ and $1\leq \ell_1,\ell_2,\ldots, \ell_{m-1}\leq k-1$ for which $y_i= y_{\ell_1}-1=y_{\ell_2}-2=\cdots= y_{\ell_{m-2}}-(m-1)= y_j-m$ and $z_{i-k+1}+1= z_{n-k+1}= z_{j-k+1}-1$. \end{enumerate} \end{theorem} \begin{proof} Suppose that ${\bf a}$ is contained in $m$ closed $k$-walks, say $\left({\bf a},{\bf a}^{(2,r)},{\bf a}^{(3,r)},\ldots,{\bf a}^{(k,r)}\right)$ for $r\in\{1,\ldots, m\}$. Let ${\bf a}^{(2,r)}=a_1^{(2,r)}\;a_2^{(2,r)}\;\cdots\; a_n^{(2,r)}$. Since each ${\bf a}^{(2,r)}$ is in a closed $k$-walk with ${\bf a}$, we have the following for each $r$: \begin{align} \label{*} {\text{st}}\left(a_1^{(2,r)}\cdots a_{n-1}^{(2,r)}\right)&=y_{{\bf a}}\\ \label{**} {\text{st}}\left(a_k^{(2,r)}\cdots a_n^{(2,r)}\right)&=z_{{\bf a}}. \end{align} Equation~\eqref{*} implies that the vertex ${\bf a}^{(2,r)}$ is determined once a choice is made for the final element of $a_{n}^{(2,r)}$, and it is assumed that there are $m$ such choices. Together, the two equalities cause the elements $a_{k}^{(2,r)},\ldots, a_{n-1}^{(2,r)}$ to be identical for all $m$ vertices; hence, the only differences between the $m$ vertices are possibly in the final position and in the first $k-1$ positions. Consider one of the vertices ${\bf a}^{(2,1)}$, which we will simplify notationally as ${\bf a}^{(2)}$. If $z_{n-k+1}=1$, then by Equation~\eqref{**}, $a_{n}^{(2)}<a_{s}^{(2)}$ for all $s\in\{k,\ldots, n-1\}$. Therefore $a_{n}^{(2)}$ can only be interchanged with any of the first~$k-1$ elements if those elements are also smaller than each $a_{s}^{(2)}$ for all $s\in\{k,\ldots, n-1\}$. The only way for this to produce $m-1$ other vertices that maintains the Equation~\eqref{*} is for there to be the elements $1,\ldots, m-1$ in the first $k-1$ positions of $y_{{\bf a}}$. Then $a_{n}^{(2)}$ can be swapped with any of those $m-1$ positions, followed by a shuffling of the $m-1$ elements to keep their proper order. This is exactly the scenario described in Condition~$(a)$ of this theorem. If $z_{n-k+1}=n-k+1$, it follows analogously by symmetry. If $z_{n-k+1}\neq 1,n-k+1$, then consider the positions in $z_{{\bf a}}$ that are 1 greater and 1 less than $z_{n-k+1}$, whose indices we will denote as $i-k+1$ and $j-k+1$ respectively where $k-1<i,j\leq n-1$. These indices lie within the overlap of the two Equations~\eqref{*} and \eqref{**} and correspond to the positions $i$ and $j$ in $y_{{\bf a}}$, meaning $z_{i-k+1}$ and~$y_i$ match up with $a_{i}^{(2)}$ and likewise for $j$ within the standardizations. Since $a_{n}^{(2)}$ must remain between $a_{i}^{(2)}$ and~$a_{j}^{(2)}$, then it can only be interchanged with elements in the first $k-1$ positions. In order to keep Equation~\eqref{*} true and have $m-1$ other vertices that are part of a closed $k$-walk, then if these elements are in positions $\ell_1,\ldots, \ell_{m-1}$, we must have $1\leq \ell_s\leq k-1$ and $y_i<y_{\ell_s}<y_j$ for each $s\in\{1,\ldots,m-1\}$. As before, we can create $m-1$ different vertices by having $a_{n}^{(2)}$ swapped with any of these $m-1$ elements and then permuting those $m-1$ elements to maintain their required order. This set up is described in Condition $(b)$ of this theorem. With every possibility for $z_{n-k+1}$ covered, the result is proven. \end{proof} We have now fully characterized how a vertex ${\bf a}$ can branch off to $m$ different vertices to begin closed $k$-walks, and in many cases, $k$-cycles. This does not mean there are only $m$ sets of vertices containing ${\bf a}$ that can be connected to form a closed $k$-walk, since further branching can occur from vertices later in the walk, as illustrated by the following example. \begin{example}\label{manyWalks}{\rm Consider the vertex ${\bf a}=14263758$ with the cycle length of $k=6$. This vertex satisfies the condition in Theorem~\ref{otherMultipleVertices} with $m=3$. Since $z_{{\bf a}}=132$ and $y_{{\bf a}}=3152647$, we have $i=6$, $j=7$, $\ell_1=3$, and $\ell_2=5$, leading to three closed $6$-walks containing distinct vertices adjacent to ${\bf a}$. This vertex, however, is actually contained in five closed $6$-walks using those three options for the second vertex: \begin{align*} \text{the repetition of the $2$-cycle } &(14263758, 31527486),\\ \text{this $2$-cycle followed by the $4$-cycle } &(14263758, 31627485, 15263748, 41527386), \text{ and}\\ \text{the three distinct $6$-cycles: } &(14263758, 31526487, 14253768, 31427586, 13264758, 21537486),\\ &(14263758, 31627485, 15263748, 41627385, 15263748, 41527386),\\ &(14263758, 31627485, 15263748, 51627384, 15263748, 41527386). \end{align*} } \end{example} Thus far, our results are centered on when vertices are within closed walks and cycles, and on how many vertices that a vertex can branch off to form other closed walks. However, we also need to consider when edges exist with the same head and tail between consecutive vertices in a cycle. The following result presents a condition for when adjacent vertices are joined by more than one edge. \begin{lemma}\label{doubleEdges} In $G(n)$, there are two directed edges from vertex ${\bf a}$ to vertex ${\bf b}$ if and only if ${\bf a}=a_1a_2\cdots a_n$ and ${\bf b}=\sigma({\bf a})=a_2a_3\cdots a_na_1$. \end{lemma} \begin{proof} If we assume ${\bf a}=a_1a_2\cdots a_n$ and ${\bf b}=a_2a_3\cdots a_na_1$, then it can easily be verified that the following are edges from ${\bf a}$ to ${\bf b}$: $a_1c_2c_3\cdots c_n(a_1+1)$ and $(a_1+1)c_2c_3\cdots c_na_1$ where $$ c_i=\begin{cases} a_i+1 &\mbox{if $a_i>a_1$}\\ a_i &\mbox{if $a_i<a_1$.}\\ \end{cases}$$ Now we assume that there are two directed edges from the vertex ${\bf a}$ to ${\bf b}$, labeled as ${\bf w}=w_1w_2\cdots w_nw_{n+1}$ and ${\bf x}=x_1x_2\cdots x_nx_{n+1}$. Then ${\text{st}}(w_1\cdots w_n)={\text{st}}(x_1\cdots x_n)={\bf a}$ and ${\text{st}}(w_2\cdots w_{n+1})={\text{st}}(x_2\cdots x_{n+1})={\bf b}$. We claim that $w_2\cdots w_n=x_2\cdots x_n$. If there is an $i\in$~$\{2,\ldots,n\}$ such that $w_i\neq x_i$, then without loss of generality, we can assume $w_i<x_i$. Then there must also be some $j\in \{1,\ldots, n+1\}$ with $w_j>x_j$ for which $w_i<w_j$ and $x_j\leq w_i<x_i$. If $j\in\{1,\ldots, n\}$, then the opposite direction of the inequalities relating the $i$th and $j$th elements of ${\bf w}$ and ${\bf x}$ cause a change in the ordering of the first $n$ elements, that is, ${\text{st}}(w_1w_2\cdots w_n)\neq {\text{st}}(x_1x_2\cdots x_n)$. Otherwise, if $j=n+1$, then this change in order results in ${\text{st}}(x_2\cdots x_nx_{n+1})\neq {\text{st}}(w_2\cdots w_nw_{n+1})$. In either case, this contradicts the fact that both edges are from ${\bf a}$ to ${\bf b}$, and thus we have $w_2\cdots w_n=x_2\cdots x_n$. With this equality, the only way for the two edges to be distinct is for $x_1=w_{n+1}$ and $x_{n+1}=w_1$. This implies that the vertices the two edges connect are equal to the standardizations $a_1\cdots a_n={\text{st}}(w_1\cdots w_n)$ and $b_1\cdots b_n={\text{st}}(w_2\cdots w_nw_1)$, where the edge ${\bf w}$ is used for the first equality and ${\bf x}$ is used for the second one. Therefore, the two vertices are of the expected form ${\bf a}=a_1a_2\cdots a_n$ and ${\bf b}=a_2a_3\cdots a_na_1$. \end{proof} The final steps in the proof of Lemma~\ref{doubleEdges} describe how multiple edges connecting a pair of vertices in the same direction are identical with the exception of the first and last element being swapped. This reasoning directly proves the following corollary. \begin{corollary}\label{afterDoubleEdges} There cannot be more than two edges in $G(n)$ from a vertex ${\bf a}$ to a vertex ${\bf b}$. \end{corollary} We will see the conditions in which there is a directed edge from ${\bf a}$ to ${\bf b}$ and a directed edge from ${\bf b}$ to ${\bf a}$ in the next section (since these are $2$-cycles). We end this section with conditions in which permutations cannot possibly be in some closed $k$-walk, or likewise $k$-cycle, in $G(n)$. \begin{lemma}\label{necessaryPerm} Let ${\bf a}=a_1a_2\cdots a_n$ be a vertex in $G(n)$. Let one of the following conditions be satisfied: \begin{packedItem} \item[$(a)$] For some positive integer $t\geq 2$, $n\geq tk+1$, $n-(t-2)\in\{a_{k+1},a_{k+2},\ldots,a_{n-(t-1)k}\}$, and for each $j\in \{n-t+3,n-t+4,\ldots,n\}$, $j\not\in \{a_1,\ldots,a_k\}$. \item[$(b)$] For some positive integer $t\geq 2$, $n\geq tk+1$, $t-1\in\{a_{k+1},a_{k+2},\ldots,a_{n-(t-1)k}\}$, and for each $j\in \{1,2,\ldots,t-2\}$, $j\not\in \{a_1,\ldots,a_k\}$. \item[$(c)$] For some positive integer $t\geq 2$, $n\geq tk+1$, $n-(t-1)\in \{a_1,a_2,\ldots,a_s\}$ where $s=\min(\{k,n-tk\})$, and for each $j\in \{n-t+2,n-t+3,\ldots,n\}$, $j\not\in \{a_1,\ldots,a_k\}$. \item[$(d)$] For some positive integer $t\geq 2$, $n\geq tk+1$, $t\in \{a_1,a_2,\ldots,a_s\}$ where $s=\min(\{k,n-tk\})$, and for each $j\in \{1,2,\ldots,t-1\}$, $j\not\in \{a_1,\ldots,a_k\}$. \end{packedItem} Then ${\bf a}$ cannot be a vertex in any closed $k$-walks in $G(n)$. \end{lemma} \begin{proof} We will proceed by contradiction, so suppose ${\bf a}$ is in some closed $k$-walk in $G(n)$. First suppose that Condition $(a)$ is satisfied. If $t=2$, that is, if $n\in \{a_{k+1},\ldots,a_{n-k}\}$, by Theorem~\ref{cycleNecessary} it is clear that an element larger than $n$ must appear as an element later in the permutation than where $n$ appeared in the permutation ${\bf a}$, but this is impossible, hence $n\in \{a_{n-k+1},a_{n-k+2},\ldots,a_n\}$. We will proceed by induction on $t$, so assume that for each positive integer $\ell\in\{2,\ldots,t\}$, ${\bf a}$ cannot be a vertex in any closed $k$-walks in $G(n)$ when $n\geq \ell k+1$, $n-(\ell-2)\in\{a_{k+1},a_{k+2}\ldots,a_{n-(\ell-1))k}\}$ and for each $j\in\{n-\ell+3,\ldots,n\}$, $j\not\in\{a_1,\ldots,a_k\}$. Let $a_m\geq n-(t-2)$ where $k+1< m <n-(t-1)k$. If $a_{m+k}> a_m$ then $a_m$ is less than $a_{m+2k},a_{m+3k},\ldots,a_{m+(t-1)k}$, but there are only $t-2$ elements in $\{1,2,\ldots,n\}$ larger than $a_m$, so $a_{m+k}<a_m$. Since all elements in $\{a_1,\ldots,a_{m-1}\}$ are less than $a_m$, we have $a_{m-k}<a_m$, so by Theorem~\ref{cycleNecessary}, $a_m<a_{m+k}$, which is a contradiction. A similar argument can be made for Condition $(b)$ by Proposition~\ref{cycleComplement}. Second suppose that Condition $(c)$ is satisfied, so let $a_m=n-(t-1)$. Then $a_{n-(t-1)+1}=a_{n-(t-2)}$ which is exactly the same as Condition $(a)$, so we reach a contradiction. A similar argument can be made for Condition~$(d)$ by Proposition~\ref{cycleComplement}. \end{proof} \section{Number of $2$-cycles}\label{2cycles} We now focus our attention on $2$-cycles, which can be completely enumerated using the results from the previous section. One important fact that we rely on in this section is Theorem~\ref{trivialVertices}$(a)$, which states that only the trivial vertices are contained in a $1$-cycle. Therefore, any non-trivial vertex satisfying Theorem~\ref{walkExistence}, which guarantees the vertex is in closed $2$-walk, is in fact in a $2$-cycle. Let $v_{n,k}$ denote the number of vertices contained in a $k$-cycle in $G(n)$, and let $C_{n,k}$ be the number of $k$-cycles. We now present our results for the number of vertices in $2$-cycles, followed by the number of these $2$-cycles in the graph of overlapping permutations. \begin{theorem}\label{numberOf2cycleVertices} For $n\geq 4$, the number of vertices contained in $2$-cycles in $G(n)$ is $v_{n,2}=2n+2$. \end{theorem} \begin{proof} Let ${\bf a}=a_1a_2\cdots a_n$ be a vertex in $G(n)$ contained within a $2$-cycle. By Theorem~\ref{cycleNecessary}, we know that ${\text{st}}(a_1\cdots a_{n-2})={\text{st}}(a_3\cdots a_n)$. Looking at the first two elements of these strings, we see ${\text{st}}(a_1 a_2)={\text{st}}(a_3 a_4)$, which results in only $12$ permutations of length $4$ that could possibly be ${\text{st}}(a_1\cdots a_4)$: $1234$, $1324$ $1423$, $2143$, $2314$, $2413$, $3142$, $3241$, $3412$, $4132$, $4231$, $4321$. We will prove the result separately depending on whether $n$ is even or odd. First suppose $n$ is even. Additionally, we first assume ${\text{st}}(a_1\cdots a_4)=1423$. Since ${\text{st}}(a_1\cdots a_{n-2})={\text{st}}(a_{3}\cdots a_n)$, it follows that \begin{align}\label{1423} {\text{st}}(a_1\cdots a_4)={\text{st}}(a_3\cdots a_6)={\text{st}}(a_5\cdots a_8)=\cdots={\text{st}}(a_{n-3}\cdots a_n) \end{align} are all equal to $1423$. As a consequence, all even elements in ${\bf a}$ are larger than the odd elements, $a_2>a_4>a_6>\cdots> a_n$, and $a_1<a_3<a_5<\cdots< a_{n-1}$. Therefore, the only possible permutation that fits this criteria is ${\bf a}=1\; n\; 2\; (n-1)\; 3\cdots \frac{n}{2}\; (\frac{n}{2}+1)$. A similar argument can be made for vertices where ${\text{st}}(a_1\cdots a_4)\in\{2143,2314,3241,3412,4132\}$ so that each of these standardizations correspond to exactly one distinct vertex in $G(n)$ that is contained in a $2$-cycle. Now suppose ${\text{st}}(a_1\cdots a_4)=1234$. The equalities ${\text{st}}(a_3\cdots a_6)=\cdots={\text{st}}(a_{n-3}\cdots a_n)=1234$ imply $a_1<a_2<a_3<\cdots<a_n$. Thus, $1\,2\,3\cdots n$ is the only permutation which matches this standardization, but this is not a $2$-cycle by Theorem~\ref{trivialVertices}$(b)$. Similarly, the only vertex with ${\text{st}}(a_1\cdots a_4)=4321$ is ${\bf a}=n\,(n-1)\,(n-2)\cdots 1$, which similarly does not generate a $2$-cycle. Suppose ${\text{st}}(a_1\cdots a_4)=1324$. As reasoned above, ${\text{st}}(a_3\cdots a_6)=\cdots={\text{st}}(a_{n-3}\cdots a_n)=1324$, so $a_1<a_3<\ldots< a_{n-1}$ and $a_2<a_4<\cdots< a_n$, and consequently, $a_1=1$, $a_3=2$, $a_{n-2}=n-1$, and $a_n=n$. If $a_2=\frac{n}{2}+2$ then $\{a_4,a_6,\ldots,a_n\}\subseteq\{\frac{n}{2}+3,\frac{n}{2}+4,\ldots,n\}$, but there are $\frac{n}{2}-1$ even positioned elements between $a_4$ and $a_n$ inclusively while there are $\frac{n}{2}-2$ values for the even positioned elements to choose from, thus $a_2\leq \frac{n}{2}+1$. In fact, for each $a_2\in\{3,\ldots,\frac{n}{2}+1\}$ there is exactly one permutation of length $n$ satisfying the inequalities $a_1<a_3<\cdots <a_{n-1}$ and $a_2<a_4<\cdots<a_n$, and Equation~\eqref{1423}. Thus, there are $\frac{n}{2}-1$ permutations with ${\text{st}}(a_1\cdots a_4)=1324$. A similar argument can be made for permutations with ${\text{st}}(a_1\cdots a_4)\in \{2413,3142,4231\}$. Therefore, when $n$ is even there are $4\left(\frac{n}{2}-1\right)+6=2n+2$ permutations that satisfy the conditions in Theorem~{\normalfont\ref{cycleNecessary}}. Next suppose that $n$ is odd. Theorem~\ref{cycleNecessary} implies the standardization equalities $${\text{st}}(a_1\cdots a_4)={\text{st}}(a_3\cdots a_6)=\cdots={\text{st}}(a_{n-4}\cdots a_{n-1})$$ $$\text{and }{\text{st}}(a_2\cdots a_5)={\text{st}}(a_4\cdots a_7)=\cdots={\text{st}}(a_{n-3}\cdots a_{n}).$$ If ${\text{st}}(a_1\cdots a_4)\in\{1423,2143,2314,3241,3412,4132\}$, then a similar argument as that made in the even $n$ case will show that there is exactly one permutation of length $n$ corresponding to each permutation of length~$4$. The same can be done for when ${\text{st}}(a_1\cdots a_4)\in\{1234,4321\}$, although the trivial vertices associated with these standardizations are not within a $2$-cycle. Suppose ${\text{st}}(a_1\cdots a_4)=1324$. As reasoned above, ${\text{st}}(a_3\cdots a_6)=\cdots={\text{st}}(a_{n-4}\cdots a_{n-1})=1324$, so $a_1<a_3<\ldots< a_{n-2}$ and $a_2<a_4<\cdots< a_{n-1}$, and consequently, $a_1=1$ and $a_3=2$. If $a_2=\frac{n-1}{2}+3$ then $a_4,a_6,\ldots,a_n$ are values from the set $\{\frac{n-1}{2}+4,\frac{n-1}{2}+5,\ldots,n\}$, but there are more even positions in ${\bf a}$ not including $a_2$ than values to choose from, which is a contradiction. Thus $a_2\leq \frac{n-1}{2}+2$. For each $a_2\in\{3,4,\ldots,\frac{n-1}{2}+2\}$ there is exactly one permutation of length $n$ with the necessary characteristics. Therefore, there are $\frac{n-1}{2}$ permutations with ${\text{st}}(a_1\cdots a_4)=1324$. We obtain the same number of permutations with ${\text{st}}(a_1\cdots a_4)=4231$ by symmetry. Suppose ${\text{st}}(a_1\cdots a_4)=2413$. It is clear that $a_2=n$ and $a_4=n-1$. By using a similar argument to the one above, $a_1\geq \frac{n+1}{2}$ and $a_1\leq n-2$. There is exactly one permutation for each $a_1\in \{\frac{n+1}{2},\frac{n+1}{2}+1,\ldots,n-2\}$ by the inequalities $a_1<a_3<\cdots <a_{n}$ and $a_2<a_4<\cdots<a_{n-1}$, and Equation~\eqref{1423}, so there are $n-2-\frac{n+1}{2}+1=\frac{n-1}{2}-1$ permutations with ${\text{st}}(a_1\cdots a_4)=2413$. By symmetry, there are also $\frac{n-1}{2}-1$ permutations with ${\text{st}}(a_1\cdots a_4)=3142$. In total, when $n$ is odd there are $2\left(\frac{n-1}{2}\right)+2\left(\frac{n-1}{2}-1\right)+6=2n+2$ vertices contained in a $2$-cycle, completing our second and final case. \end{proof} An interesting property to note regarding vertices ${\bf a}=a_1\cdots a_n$ within $2$-cycles is that they are alternating permutations; that is, if $a_i<a_{i+1}$ when $i$ is even and $a_i>a_{i+1}$ when $i$ is odd, or if $a_i>a_{i+1}$ when $i$ is even and $a_i<a_{i+1}$ when $i$ is odd. It can also be stated as ${\bf a}$ having ascents only in even positions and descents only in odd positions, or vice versa. Now that we know how many vertices are contained in $2$-cycles, we will prove how many $2$-cycles exist in~$G(n)$. This time the cases of even and odd length permutations result in different enumerations. \begin{theorem}\label{numberOf2cycles} For $n\geq 4$, the number of $2$-cycles in $G(n)$ is the following $$C_{n,2}=\begin{cases} n+2 &\mbox{if $n$ is even} \\ n+3 &\mbox{if $n$ is odd.} \\ \end{cases}$$ Furthermore, when $n$ is even, there are no $2$-cycles featuring a multiedge with the same head and tail, and each vertex is contained in one $2$-cycle except for two vertices that are contained in exactly two $2$-cycles. These vertices are ${\bf a}=1\;(\frac{n}{2}+1)\;2\;(\frac{n}{2}+2)\cdots (\frac{n}{2}-1)\;(n-1)\;\frac{n}{2}\;n$ and $\overline{{\bf a}}$. When $n$ is odd, each vertex is paired with only one other vertex for form a $2$-cycle, and exactly two cycles contain edges with the same head and tail, namely from ${\bf a}=\frac{n+1}{2}\; 1 \;\frac{n+3}{2}\, 2\;\frac{n+5}{2} \cdots (n-1)\;\frac{n-1}{2}\;n $ to $\sigma({\bf a})$ and from $\overline{{\bf a}}$ to $\sigma(\overline{{\bf a}})$. \end{theorem} \begin{proof} From the previous theorem, we know in each case that $C_{n,2}\geq n+1$ once these $2n+2$ vertices are paired up into $2$-cycles. We first assume $n$ is even and begin by showing there does not exist any distinct cycles of the form $({\bf a},{\bf b})$ and $({\bf a},{\bf b})$ when $n$ is even; that is, there are no pairs of vertices connected by two edges in the same direction. By Lemma~{\normalfont\ref{doubleEdges}}, if such a pair exists, then ${\bf a}=a_1\cdots a_n$ and ${\bf b}=a_2a_3\cdots a_n a_1$. Since ${\text{st}}(a_1\cdots a_{n-2})={\text{st}}(a_3\cdots a_n)$, $a_1>a_3>a_5>\cdots>a_{n-1}$ or $a_1<a_3<a_5<\cdots<a_{n-1}$. Likewise, ${\text{st}}(a_3\cdots a_n a_1)={\text{st}}(a_1a_2\cdots a_{n-1})$ implies that $a_3>a_5>\cdots > a_{n-1} > a_1$ or $a_3<a_5<\cdots<a_{n-1}<a_1$, which in either case is a contradiction. Thus, there are no $2$-cycles with edges that have the same head and tail in $G(n)$ when $n$ is even. We now consider vertices ${\bf a}$ that are contained in multiple $2$-cycles with distinct second vertices. One can observe that $1\; (\frac{n}{2}+1)\; 2\; (\frac{n}{2}+2)\cdots(\frac{n}{2}-1)\;(n-1)\;\frac{n}{2}\; n$ and $n\;\frac{n}{2}\;(n-1)\;(\frac{n}{2}-1)\cdots(\frac{n}{2}+2)\;2\;(\frac{n}{2}+1)\;1$ are each in two such $2$-cycles according to Theorem~\ref{otherMultipleVertices} with $m=2$. Since these were included in the $2n+2$ vertices, the number of $2$-cycles in $G(n)$ is now at least $n+2$. It remains to show that there are no other vertices that are in two or more edge-disjoint $2$-cycles. By Theorem~\ref{reverse}, if such a vertex exists, then it must satisfy the conditions of either Theorem~\ref{multipleVertices} or~\ref{otherMultipleVertices}. The former of these theorems clearly does not apply when $n\geq 4$ and $k=2$ since it requires $n<2k$. Then we assume there is a vertex ${\bf a}=a_1\cdots a_n$ satisfying the assumptions of Theorem~\ref{otherMultipleVertices} for $k=2$, which based on the cycle length would require $m=2$ and the index $\ell=1$. To satisfy the condition, the values $y_1$ and $z_{n-1}$ cannot be $1$ or~$n-1$. Recall from the proof of Theorem~\ref{numberOf2cycleVertices} that ${\text{st}}(a_1\cdots a_4)={\text{st}}(a_3\cdots a_6)=\cdots {\text{st}}(a_{n-3}\cdots a_n)$ and these are equal to one of ten possible length $4$ permutations. If we first assume ${\text{st}}(a_1\cdots a_4)=3142$, we obtain inequalities from these standardizations of the form $a_1<a_3<\cdots <a_{n-1}$ and $a_2<a_4<\cdots <a_n$. Since $a_{n-1}>a_n$ because ${\text{st}}(a_{n-3}\cdots a_n)=3142$, the inequalities imply $a_{n-1}=n$, hence $z_{n-1}=n-1$, which we stated could not occur when satisfying Theorem~\ref{otherMultipleVertices}. Using the analogous inequalities for the following seven other options $\{1423, 2143, 2314, 2413, 3142, 3241, 4132\}$ for the standardization of the first four elements, we arrive at similar conclusions that $a_2$ or $a_{n-1}=1$ or $n$, causing $y_1$ or $z_{n-1}$ to be $1$ or $n-1$. After eliminating these possibilities, we now have that a vertex ${\bf a}$ contained within two distinct $2$-cycles must satisfy ${\text{st}}(a_1\cdots a_4)=1324$ or $4231$, which we note are the cases for our two previously mentioned vertices that do satisfy Theorem~\ref{otherMultipleVertices}. We have from the assumptions of Theorem~\ref{otherMultipleVertices} that there is some $i,j$ between $2$ and $n-1$ with $y_i+1=y_1=y_j-1$ and $z_{i-1}+1=z_{n-1}=z_{j-1}-1$. Then since ${\text{st}}(y_2\cdots y_{n-1})={\text{st}}(z_1\cdots z_{n-2})$ as a consequence of Theorem~\ref{cycleNecessary}, we claim that $y_2\cdots y_{n-1}=z_1\cdots z_{n-2}$, forcing $y_1=z_{n-1}$ as well. If there was some differing element between the overlapping portion of $y_{{\bf a}}$ and $z_{{\bf a}}$, say $z_r>y_r$, the only way to maintain the standardization equality between these portions is for $z_{n-1}$ to correspondingly decrease, or increase if $z_r<y_r$. Either situation would prevent $z_{i-1}+1=z_{n-1}=z_{j-1}-1$, confirming our claim. As shown in the proof of Theorem~\ref{numberOf2cycleVertices}, if ${\text{st}}(a_1\cdots a_4)=1324$, then $a_1=1$ and $a_n=n$, resulting in $z_r=a_r$ and $y_r=a_{r+1}-1$ for each index $r=1,\ldots, n-1$. Thus, we need $a_{2}-1=a_{n-1}$ to have $y_1=z_{n-1}$. Additionally from that proof, we know $3\leq a_2\leq \frac{n}{2}+1$ along with the inequalities $a_1<a_3<\cdots <a_{n-1}$ and $a_2<a_4<\cdots <a_n$. Since there are $\frac{n}{2}-1$ elements $a_3, a_5,\ldots, a_{n-1}$ in increasing order, all of which are larger than $a_1=1$, it follows that if $a_2<\frac{n}{2}-1$, it is not possible for $a_{n-1}<a_2$. Thus, $a_2=\frac{n}{2}-1$ is the only possible vertex ${\text{st}}(a_1\cdots a_4)=1324$ that satisfies Theorem~\ref{otherMultipleVertices}. The case of $4231$ follows by symmetry. This completes the proof for the even case, as the $2n+2$ vertices contained in $2$-cycles, added with the two vertices that are contained in two distinct $2$-cycles, are paired together make the number of $2$-cycles be $\frac{2n+4}{2}=n+2$. Now we assume that $n$ is odd. We will show that there does not exist any cycles of the form $({\bf a},{\bf b})$ and $({\bf a},{\bf c})$; that is, there is no vertex contained in two $2$-cycles with distinct second vertices. As with the even case, Theorem~\ref{reverse} implies such a vertex would satisfy Theorem~\ref{multipleVertices}, which does not apply when $k=2$, or Theorem~\ref{otherMultipleVertices}. Once again, there are only ten potential permutations of length $4$ for ${\text{st}}(a_1\cdots a_4)$. If ${\text{st}}(a_1\cdots a_4)=1324$, then unlike the even case, we will show that there are no vertices that satisfy Theorem~\ref{otherMultipleVertices}. We have $a_1<a_3<\cdots <a_n$ and $a_2<a_4<\cdots <a_{n-1}$, in addition to the fact that Equation~\eqref{1423} implies ${\text{st}}(a_{n-2}a_{n-1}a_n)=132$ in the $n$ odd case. Hence $a_{n-1}>a_n$, combining with the inequalities to result in $a_{n-1}=n$ and $z_{n-1}=n-1$. This vertex cannot satisfy Theorem~\ref{otherMultipleVertices}, and similar arguments can be made for each of the nine remaining length $4$ standardizations. Thus, there are no vertices contained in multiple $2$-cycles with distinct second vertices. It is clear that there are two edges with the same head and tail from ${\bf a}=1 \;\frac{n+3}{2}\, 2\;\frac{n+5}{2} \cdots (n-1)\;\frac{n-1}{2}\;n \; \frac{n+1}{2}$ to ${\bf b}=\frac{n+3}{2}\, 2\;\frac{n+5}{2} \cdots (n-1)\;\frac{n-1}{2}\;n \; \frac{n+1}{2}\; 1$ that creates two $2$-cycles between these vertices. Likewise, there are two $2$-cycles between their complements $\overline{{\bf a}}$ and $\overline{{\bf b}}$. It remains to show that there are no other vertices that have this property. If there did exist another such pair, then by Lemma~\ref{doubleEdges}, the pair is ${\bf a}=a_1\cdots a_n$ and ${\bf b}=a_2\cdots a_n a_1$, and the edge from ${\bf b}$ to ${\bf a}$ would imply that ${\text{st}}(a_1\cdots a_{n-1})={\text{st}}(a_3\cdots a_{n}a_1)$. Additionally, since ${\text{st}}(a_1\cdots a_{n-2})={\text{st}}(a_3\cdots a_{n})$, $a_1>a_3>\cdots>a_n$ or $a_1<a_3<\cdots <a_n$. As a result, $a_{n-1}>a_1$ or $a_{n-1}<a_1$ respectively. It follows that $a_2>a_4>\cdots>a_{n-1}$ or $a_2<a_4<\cdots<a_{n-1}$. The only permutations that fit all of these criteria above are the ones indicated. With all $2n+2$ vertices being paired with exactly one vertex to form a $2$-cycle, and with only two of these containing edges with the same head and tail, the number of $2$-cycles in $G(n)$ for odd $n$ is $\frac{2n+2}{2}+2=n+3$. \end{proof} Note that each of the previous two theorems assumed that $n\geq 4$. When $n=3$, the results are not true, as seen in the example in Section 2, since there are only $4<2n+2$ vertices contained in $2$-cycles. Even though the six $2$-cycles in $G(3)$ does match with the expected $n+3$ from Theorem~\ref{numberOf2cycles}, there are cycles of the form $({\bf a},{\bf b})$ and $({\bf a},{\bf c})$ with different pairs of vertices, which is a case that doesn't occur for larger odd values of $n$. If you try to extend the proof technique from $k=2$ in Theorems~\ref{numberOf2cycleVertices} and \ref{numberOf2cycles} to $k=3$, the problem becomes quite difficult. Instead of examining cases based on the standardization of the first $4$ elements, it would depend on the first $6$ elements. There are $120$ permutations of length $6$ that could possibly start a permutation where the standardization of the first $n-3$ elements is the same as the standardization of the last $n-3$ elements. \section{Number of Vertices Within Cycles}\label{countVertices} In this section, we will show several ways of attaining or bounding the number of closed $k$-walks and the number of $k$-cycles when $k$ is prime. We first define the number $w_{n,k}$ to be the number of vertices of $G(n)$ contained in a closed $k$-walk. \begin{theorem}\label{countingVertices2k} If $n\leq 2k$ then the number of vertices contained in a closed $k$-walk is $\displaystyle w_{n,k}=\frac{n!}{(n-k)!}$. \end{theorem} \begin{proof} Let ${\bf a}=a_1a_2\cdots a_n$ be a permutation contained in a closed $k$-walk, so ${\text{st}}(a_1\cdots a_{n-k})={\text{st}}(a_{k+1}\cdots a_n)$. Since the first $n-k$ and last $n-k$ elements do not overlap when $n\leq 2k$, choosing the first $n-k$ elements is independent of choosing the middle $n-2(n-k)=2k-n$ elements. There are $n!/k!$ ways to choose the first $n-k$ elements and $\frac{k!}{(n-k)!}$ ways to choose the middle $2k-n$ elements. Once the first $n-k$ elements and middle $2k-n$ elements of a permutation are chosen, the remaining elements are all determined by Theorem~\ref{cycleNecessary}. Thus, there are $\left(\frac{n!}{k!}\right)\left( \frac{k!}{(n-k)!}\right)=\frac{n!}{(n-k)!}$ ways to build a permutation ${\bf a}$. \end{proof} \begin{theorem}\label{countingVertices2k3k} Let $n>2k$ be odd, $k\geq 3$, and ${\bf a}=a_1a_2\cdots a_n$ be a vertex in $G(n)$. Then the number of vertices contained in a closed $k$-walk is bounded above as follows: \begin{align*} w_{n,k}&\leq \frac{(n-2)!}{(n-k)!}\left(\left(n+\frac{1}{2}\right)(n+1)(n-1)+k+\frac{n-5}{2}-2-(n-1)\left\lceil \frac{n-1}{4}\right\rceil\right). \end{align*} \end{theorem} \begin{proof} Let ${\bf a}=a_1a_2\cdots a_n$ be a permutation in a closed $k$-walk, so ${\text{st}}(a_1\cdots a_{n-k})={\text{st}}(a_{k+1}\cdots a_n)$ by Theorem~\ref{cycleNecessary}. Note that at times our analysis is done by assuming $a_1\leq n/2$. This is because by Proposition~\ref{cycleComplement}, there are the same number of permutations such that $a_1<n/2$ as there are there are with $a_1>n/2$. We will further break this case down into smaller cases based on the relationship between $a_1$ and $a_2$. Suppose that $a_1$ is odd. There are $(n+1)/2$ choices for $a_1$ and $(n-1)!/((n-1)-(k-1))!=(n-1)!/(n-k)!$ choices for $a_2,a_3,\ldots,a_k$. However, some of these will not fit the standardization condition in Theorem~\ref{cycleNecessary}. For this reason, $w_{n,k}$ is an inequality rather than an equality. Suppose $a_1$ is even. If $a_2<a_1$ then there are at most $(n-2)!/((n-2)-(k-2))!=(n-2)!/(n-k)!$ ways to choose $a_3,\ldots,a_k$, $a_1-1$ ways to choose $a_2$, and $(n-1)/2$ ways to choose $a_1$. Next suppose that $a_2>a_1$ and $a_1=2$. If $1\not\in \{a_3,\ldots,a_k\}$ then $a_{k+1}=1$, since otherwise $a_i>a_1$ for all $i\in\{k+1,k+2,\ldots,n\}$, which would not contain the element $1$. However, $1$ cannot be in $\{a_{k+1},\ldots,a_{n-k}\}$ by Lemma~\ref{necessaryPerm}, so $1\in\{a_3,\ldots,a_k\}$. There are $k-2$ choices for the position of this element. So there are $(k-2)((n-2)!)/((n-2)-(k-2))!=(k-2)((n-2)!)/(n-k)!$ permutations with $a_2>a_1$ and $a_1=2$. Finally suppose that $a_2>a_1$ and $a_1\neq 2$. So for each odd $a_1\neq 2$, the number of permutations of this specified form is $(n-2)!/((n-2)-(k-2))!=(n-2)!/(k-2)!$. Notice that we have $2(n-5)/4=(n-5)/2$ or $2(n-3)/4-1=(n-3)/2-1=(n-5)/2$ values for $a_1$ in this case if $n\equiv 1\pmod{4}$ or $n\equiv 3\pmod{4}$, respectively. The sum of these enumerations for the vertices possibly satisfying Theorem~\ref{cycleNecessary} provides an upper bound for the actual number of vertices in closed $k$-walks. Thus, the number of vertices contained in a closed $k$-walk is \begin{align*} w_{n,k}&\leq \frac{(n+1)((n-1)!)}{2((n-k)!)} + 2\left(\frac{n-1}{2}\right)\sum_{i=1}^{\lceil(n-1)/4\rceil}(2i-1)\left(\frac{(n-2)!}{(n-k)!}\right)\\ &\phantom{\leq} \; + (k-2)\left(\frac{(n-2)!}{(n-k)!}\right) + \left(\frac{n-5}{2}\right) \frac{(n-2)!}{(n-k)!} \\ &= \frac{(n-2)!}{(n-k)!}\left(\left(n+\frac{1}{2}\right)(n+1)(n-1)+k+\frac{n-5}{2}-2-(n-1)\left\lceil \frac{n-1}{4}\right\rceil\right). \end{align*} \end{proof} When $n$ is even, it is more difficult to calculate a bound for the number of closed $k$-walks. For example, besides using a case-by-case argument for each $n$, it is not clear to the authors why there are $14$ vertices that are in closed $4$-walks when $n=10$, $a_1=3$, and $a_2=2$ while there are $56$ vertices that are in closed $4$-walks when $n=10$, $a_1=3$, and $a_2=1$. There are many other instances similar to this that make it difficult to count the number of closed $k$-walks when $n$ is even. \begin{example} {\rm There are instances when the counting method in Theorem~\ref{countingVertices2k3k} is not enough to get an exact bound. One such example is when $n=11$ and $k=3$. The permutation $3615827a49b$ satisfies Theorem~\ref{cycleNecessary} but there is no permutation that satisfies Theorem~\ref{cycleNecessary} that has $a_1=3$, $a_2=6$, and $a_3=4$. In fact, this permutation is the only permutation that starts with $36$ and satisfies Theorem~\ref{cycleNecessary}. It is not clear at this time how to ensure that we account for each possible contingency in the way provided in the proof of Theorem~\ref{countingVertices2k3k}}. \end{example} We now examine the relationship of the cycle lengths if a vertex is in cycles of two different lengths. This will help in converting the results in the previous two theorems about closed $k$-walks into counting vertices in $k$-cycles, which was denoted as $v_{n,k}$. \begin{theorem}\label{combiningCycles} Assume ${\bf a}=a_1a_2\cdots a_n$ is a vertex that is in some $k$-cycle in $G(n)$. If $\gcd(k,j)=1$ and $k+j<n$, then ${\bf a}$ is not in a $j$-cycle. \end{theorem} \begin{proof} Assume that ${\bf a}$ is in both an $k$-cycle and a $j$-cycle, hence by Theorem~\ref{cycleNecessary}, ${\bf a}$ satisfies the following: \begin{align} {\text{st}}(a_1\cdots a_{n-k})&={\text{st}}(a_{k+1}\cdots a_{n}) \label{eq_i}\\ \text{and } {\text{st}}(a_1\cdots a_{n-j})&={\text{st}}(a_{j+1}\cdots a_{n}). \label{eq_j} \end{align} Assume without loss of generality that $a_1<a_2$. Equation~\eqref{eq_i} implies that $a_{k+1}<a_{k+2}$ and can be applied repeatedly to attain $a_{mk +1}<a_{mk +2}$ for any positive integer $m$ with $mk+2\leq n$. Equation~\eqref{eq_j} can likewise be used to shift the indices in the inequality by multiples of $j$. The equations are also capable of shifting the indices in the negative direction so that $a_\ell<a_{\ell+1}$ would imply $a_{\ell-j}<a_{\ell+1-j}$ if $\ell>j$. Since it is assumed that $\gcd(k,j)=1$, for each $c\in \{2,\ldots, n-1\}$ there exists $d,e\in\mathbb{Z}$ such that $d\cdot k+e\cdot j=c-1$. Therefore, we can achieve the inequality $a_{c}<a_{c+1}$ from $a_1<a_2$ by applying Equation~\eqref{eq_i} $d$ times and Equation~\eqref{eq_j} $e$ times. If either of $d$ or $e$ are negative, the corresponding equation is applied each time to shift the indices in the negative direction. Since $k+j<n$ we can order the applications of the equations to maintain that the first index of the inequality is always within $\{1,\ldots, n-1\}$. Since we now have $a_c<a_{c+1}$ for any $c\in \{2,\ldots, n-1\}$, the vertex ${\bf a}$ must be the trivial vertex ${\bf a}=1 2\cdots n$, which is not within any $j$-cycle by Theorem~\ref{trivialVertices}$(b)$, contradicting that assumption. \end{proof} The following result shows that counting the number of closed $k$-walks is enough to count the number of $k$-cycles in $G(n)$ for any prime cycle length $k$. \begin{corollary}\label{primek} For a prime number $k$, the number of vertices that are within a $k$-cycle is given by $v_{n,k}=w_{n,k}-2$. \end{corollary} \begin{proof} First, assume ${\bf a}$ is a nontrivial vertex, so ${\bf a}\neq 1\cdots n$ and ${\bf a}\neq n\cdots 1$. If there exists a closed $k$-walk at vertex ${\bf a}$, it is either a $k$-cycle or it is a sequence of cycles of lengths $i_1,\cdots, i_\ell$ where $i_1+\cdots +i_\ell=k$. By Theorem~\ref{combiningCycles}, ${\bf a}$ is only in cycles of two different lengths if those lengths are not relatively prime. Thus each pair of lengths shares a common factor, resulting in some integer $d>1$ that divides each of $i_1,\ldots,i_\ell$. Thus, $d$ must also divide their sum $k$, which is not possible since $k$ is prime. Therefore, every closed $k$-walk that includes the nontrivial vertex ${\bf a}$ is in fact a $k$-cycle. In the case of ${\bf a}=1\cdots n$ or $n\cdots 1$, a closed $k$-walk must only be a repetition of the $1$-cycle since ${\bf a}$ is not included in any $j$-cycle for $1<j<n$ by Theorem~\ref{trivialVertices}$(b)$. Since the trivial vertices are still included in $w_{n,k}$ in the count within Theorem~\ref{countingVertices2k} and the bound in Theorem~\ref{countingVertices2k3k}, we must subtract those two vertices from $w_{n,k}$. \end{proof} \section{Concluding Remarks}\label{remarks} Our methods for counting the number of cycles in $G(n)$ fall short when it comes to finding a generalized method, though we were able to find the number of $2$-cycles in $G(n)$ and this same method could possibly work for finding the number of $3$-cycles in $G(n)$. We were able to establish a list of conditions for when a closed $k$-walk can and cannot exist. An intriguing result we found was Corollary~\ref{primek}. Combined with Theorem~\ref{countingVertices2k}, it provides an exact count for the number of vertices in $k$-cycles for an infinite set of pairs of $k$ and $n$ with $k\geq 3$, $k$ being prime, and $n\leq 2k$. Furthermore, if we can pin down the number of closed $k$-walks in $G(n)$ when $n>2k$, then we can use this corollary to immediately find the number of cycles in a graph when $k$ is prime for any size of permutations. There are several unanswered questions, some of which are shown below. \begin{question} How many vertices are in $k$-cycles when $k$ is not prime? \end{question} \begin{question} How many $k$-cycles (or closed $k$-walks) are there in $G(n)$ for $3\leq k< n$? \end{question} An interesting extension to this question which has received no attention are the following two questions when $k\geq n$. Note that the enumerations of cycles for $G(n,312)$ in~\cite{EKS} also only counted cycles with length at most $n$. \begin{question} How many $k$-cycles (or closed $k$-walks) are there in $G(n)$ for $k\geq n\geq 3$? \end{question} It can be seen in Figure~\ref{G3} that there does exist a $4$-cycle $(132,213,231,312)$, so would it be possible to count the number of $k$-cycles in $G(n)$ when $k$ is larger than $n$? To this same point, it is quite a common problem to find a Hamilton cycle in graphs which leads us to our next question. Note that a similar question was answered by Horan and Hurlbert~\cite{HoranHurlbert} for $s$-overlap cycles, but their work centered on $k$-permutations where $k<n$. \begin{question} Does there exist a Hamilton cycle in $G(n)$ for all $n$? If so, how many? \end{question} As described in~\cite{EKS}, when avoiding length $3$ patterns in $G(n)$, several patterns result in identical numbers of cycles. The other distinct case that was left unanswered in their work involves avoiding the pattern $321$. \begin{question} Can our results for closed $k$-walks and cycles on the entire graph $G(n)$ be used to assist in determining the number of $k$-cycles (or closed $k$-walks) in $G(n,321)$ or in other subgraphs $G(n,\pi)$ where $\pi$ is a pattern of length at least $4$? \end{question} \bibliographystyle{amsplain}
{ "timestamp": "2016-09-09T02:01:12", "yymm": "1609", "arxiv_id": "1609.02210", "language": "en", "url": "https://arxiv.org/abs/1609.02210", "abstract": "The graph of overlapping permutations is a directed graph that is an analogue to the De Bruijn graph. It consists of vertices that are permutations of length $n$ and edges that are permutations of length $n+1$ in which an edge $a_1\\cdots a_{n+1}$ would connect the standardization of $a_1\\cdots a_n$ to the standardization of $a_2\\cdots a_{n+1}$. We examine properties of this graph to determine where directed cycles can exist, to count the number of directed $2$-cycles within the graph, and to enumerate the vertices that are contained within closed walks and directed cycles of more general lengths.", "subjects": "Combinatorics (math.CO)", "title": "Enumerating cycles in the graph of overlapping permutations", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9843363545048393, "lm_q2_score": 0.831143054132195, "lm_q1q2_score": 0.8181243239765031 }
https://arxiv.org/abs/1904.06998
An optimal polynomial approximation of Brownian motion
In this paper, we will present a strong (or pathwise) approximation of standard Brownian motion by a class of orthogonal polynomials. The coefficients that are obtained from the expansion of Brownian motion in this polynomial basis are independent Gaussian random variables. Therefore it is practical (requires $N$ independent Gaussian coefficients) to generate an approximate sample path of Brownian motion that respects integration of polynomials with degree less than $N$. Moreover, since these orthogonal polynomials appear naturally as eigenfunctions of an integral operator defined by the Brownian bridge covariance function, the proposed approximation is optimal in a certain weighted $L^{2}(\mathbb{P})$ sense. In addition, discretizing Brownian paths as piecewise parabolas gives a locally higher order numerical method for stochastic differential equations (SDEs) when compared to the standard piecewise linear approach. We shall demonstrate these ideas by simulating Inhomogeneous Geometric Brownian Motion (IGBM). This numerical example will also illustrate the deficiencies of the piecewise parabola approximation when compared to a new version of the asymptotically efficient log-ODE (or Castell-Gaines) method.
\section{Introduction} Brownian motion is a central object for modelling real-world systems that evolve under the influence of random perturbations \cite{Brownianuseful}. In applications where methods discretize Brownian motion, usually only increments of the path are generated \cite{Higman}. In this setting, the best $L^{2}(\mathbb{P})$ approximation of Brownian motion that is measurable with respect to these increments is given by the piecewise linear path that agrees on discretization points \cite{CameronClark}. This motivates the following natural question:\vspace{1.5mm} \textit{Are there better discrete approximations of Brownian motion than piecewise linear?}\vspace{1.5mm} The next simplest approximant would be a piecewise polynomial, though it is not clear whether this would be advantageous for tackling problems such as SDE simulation. This paper can be viewed as a logical continuation of \cite{Multiscalebm}, where a polynomial wavelet representation of Brownian motion was proposed. These wavelets were constructed to capture certain ``geometrical features'' of the path, namely the integrals of the Brownian motion against monomials. We shall investigate the practical applications of these polynomials and their geometrical features in the numerical analysis of SDEs.\\ \vspace{-4mm} \begin{figure}[h]\label{polynomialdiagams} \centering \includegraphics[width=0.975\textwidth]{"Polynomial_diagrams".pdf}\vspace{-2mm} \caption{Sample paths of Brownian motion with corresponding polynomial approximations.}\vspace{-3.mm} \end{figure}\newpage\noindent The paper is organised as follows. In Section 2, we shall state and prove the main result of the paper (Theorem \ref{waveletthm}). This will be a Karhunen-Lo\`{e}ve theorem for the Brownian bridge, where the orthogonal functions used in the approximation are polynomials. Furthermore, we shall explicitly show that each basis function is proportional to a shifted $(\alpha, \beta)$-Jacobi polynomial but with the nonstandard exponents $\alpha = \beta = -1$. This enables us to construct these orthogonal polynomials using recurrence relations, or as the difference of two shifted Legendre polynomials whose degrees differ by two. The resulting polynomial expansion of Brownian motion was independently discovered by Habermann in \cite{Habermann}, where a sharp $L^{2}\hspace{-0.075mm}(\mathbb{P})$ convergence rate of $O\big(\frac{1}{\sqrt{n}}\big)$ is established\footnote{\hspace*{-0.25mm}A Matlab demonstration can be found at \href{https://www.chebfun.org/examples/stats/RandomPolynomials.html}{chebfun.org/examples/stats/RandomPolynomials.html}}.\vspace{0.5mm} \noindent In\hspace{0.35mm} Section\hspace{0.35mm} 3,\hspace{0.35mm} we\hspace{0.35mm} shall\hspace{0.35mm} investigate\hspace{0.35mm} some\hspace{0.35mm} significant\hspace{0.35mm} consequences\hspace{0.35mm} of\hspace{0.35mm} the\hspace{0.35mm} main\hspace{0.35mm} theorem.\vspace{0.75mm} \begin{theorem} Let $W$ denote a standard real-valued Brownian motion on $[\hspace{0.1mm}0,1]$. Let $W^{n}$ be the unique $n$-th degree random polynomial with a root at $0$ and satisfying \begin{align} \int_{0}^{1}u^{k}\,dW_{u}^{n} & = \int_{0}^{1}u^{k}\,dW_{u}\hspace{0.125mm},\hspace{3mm}\text{for}\hspace{2mm}k = 0\hspace{0.25mm}, 1\hspace{0.25mm}, \cdots\hspace{0.25mm}, n - 1\hspace{0.25mm}. \end{align} Then\hspace{0.25mm} $W =\hspace{0.25mm} W^{n} + Z^{n}$\hspace{0.25mm}, where\hspace{0.25mm} $Z^{n}$ is a centered Gaussian process independent of\hspace{0.5mm} $W^{n}$. \end{theorem}\smallbreak \noindent The above theorem has a simple yet striking conclusion, namely that polynomials can be unbiased approximants of Brownian motion. In addition, the first non-trivial case ($n=2$) already has interesting applications within the numerical analysis of SDEs. One reason is that parabolas can capture the ``\hspace{0.25mm}space-time area'' of Brownian motion.\vspace{-7.5mm} \begin{figure}[h]\label{paraboladiagram} \centering \includegraphics[width=0.975\textwidth]{"Brownian_decomposition_diagram".pdf}\vspace{-5mm} \caption{Brownian motion can be expressed as a (random) parabola plus independent noise. Moreover, the approximating parabola has the same increment and time integral as the original path. } \end{figure}\vspace{-4mm}\smallbreak\noindent Therefore discretizing Brownian motion using a piecewise parabola gives a locally high order methodology for numerically solving one-dimensional SDEs. However, since certain triple iterated integrals of Brownian motion and time are partially matched by these parabolas, we expect this method to have only an $O(h)$ rate of convergence (where $h$ denotes the step size used). This gives motivation for the following theorem:\smallbreak \begin{theorem}\label{introintegral} Let\hspace{0.25mm} $\wideparen{W}$ be the (unique) quadratic polynomial with a root at $0$\hspace{0.25mm} and\vspace{-0.75mm} \begin{align*} \wideparen{W}_{1} = W_{1}\hspace{0.25mm}, \hspace{5mm}\int_{0}^{1}\wideparen{W}_{u}\, du = \int_{0}^{1}W_{u}\,du\hspace{0.25mm}. \end{align*} Then the following third order iterated integral of Brownian motion can be estimated:\vspace{-0.75mm} \begin{align}\label{introintegralestimate} \mathbb{E}\left[\hspace{0.25mm}\int_{0}^{1}W_{u}^{2}\,du\,\Big|\,W_{1}\hspace{0.25mm}, \int_{0}^{1}W_{u}\,du\hspace{0.25mm}\right] = \int_{0}^{1}\wideparen{W}_{u}^{2}\,du + \frac{1}{15}\hspace{0.25mm}. \end{align} \end{theorem}\vspace{-3mm}\smallbreak \noindent The above theorem can be directly incorporated into the stochastic Taylor method as well as the log-ODE or Castell-Gaines method (see \cite{CastellGaines}, \cite{StocLie}). We will show that by estimating this non-trivial iterated integral with its conditional expectation, we can design numerical methods that enjoy high orders of both strong and weak convergence. Specifically, for a general SDE that is driven by a one-dimensional Brownian motion and governed by sufficiently regular vector fields (smooth with bounded derivatives), the numerical methods that correctly utilize the above conditional expectation will have a strong convergence rate of $O(h^{\frac{3}{2}})$ as well as a weak convergence rate of $O(h^{2})$. \newpage\noindent These high orders of convergence can also be achieved in the multidimensional setting provided the vector fields governing the SDE satisfy certain commutativity conditions. For example, this estimator has applications for simulating SDEs with additive noise: \begin{align*} dy_{t} & = f(y_{t})\,dt + \sigma\hspace{0.25mm}dW_{t}\hspace{0.25mm}, \end{align*} where $f$ is a smooth vector field on $\mathbb{R}^{d}$, $\sigma > 0$ is constant and $W$ is now $d$-dimensional. By considering Theorem \ref{introintegral}, we expect that $y_{t}$ is well approximated (for small $t$) by \begin{align*} \wideparen{y}_{t} + \frac{1}{30}\hspace{0.25mm}t^{2}\sigma^{2}\Delta f(y_0)\hspace{0.25mm}, \end{align*} where $\wideparen{y}$ denotes the solution of the below ODE driven by a ``\hspace{0.25mm}Brownian parabola'' $\wideparen{W}$, \begin{align*} d\hspace{0.25mm}\wideparen{y}_{t} & = f(\hspace{0.25mm}\wideparen{y}_{t})\,dt + \sigma\hspace{0.25mm}d\wideparen{W}_{t}\hspace{0.25mm},\\[3pt] \wideparen{y}_{0} & = y_{0}\hspace{0.25mm}. \end{align*} This parabola-driven ODE can then be discretized using a three-stage Runge-Kutta method and the resulting SDE approximation shall be investigated in a future work.\medbreak \noindent Since these methods are based on the conditional expectation given by Theorem \ref{introintegral}, they are designed to minimize the leading error term within local Taylor expansions. This sense of optimality is conceptually similar to that of the asymptotically efficient SDE approximations developed by Clark \cite{Clark}, Newton \cite{Newton1, Newton2} and Castell \& Gaines \cite{CastellGaines}. The key difference is that we are employing additional integral information about $W$. Hence, this line of research could provide a further insight into the approximation of It\^{o} integrals using linear path information, where there already are a number of results concerning the computational complexity of methods (see \cite{Xiao}, \cite{LinearApprox} and \cite{Dickinson}).\\ Most notably, Tang and Xiao \cite{Xiao} consider the same triple iterated integral as in (\ref{introintegralestimate}) and present an asymptotically optimal approximation that performs well when a limited number of random variables are used (see Table 2 for these numerical results). Whilst there are other senses of optimality (such as those discussed in \cite{Lee} and \cite{Spline}) that could be used when analysing the proposed approximations of Brownian motion and SDE solutions, we shall estimate errors in an $L^{2}(\mathbb{P})$ sense throughout the paper. In particular, we can apply the main result to quantify the error of the new estimator.\medbreak \begin{theorem}\label{introintegralvar} Using the same notation as before, we have the following variance: \begin{align*} \operatorname{Var}\left(\int_{0}^{1}W_{u}^{2}\,du\,\Big|\,W_{1}\hspace{0.25mm}, \int_{0}^{1}W_{u}\,du\right) = \frac{11}{6300} + \frac{1}{180}\hspace{0.25mm}W_{1}^{2} + \frac{1}{175}\left(\hspace{0.25mm}\int_{0}^{1}W_{u}\,du - \frac{1}{2}\hspace{0.25mm}W_{1}\right)^{2}. \end{align*} \end{theorem}\bigbreak \noindent In Section 4, we demonstrate the applicability of these ideas to SDE simulation through various discretizations of Inhomogeneous Geometric Brownian Motion (IGBM) \begin{align*} dy_{t} = a(b-y_{t})\,dt+\sigma\hspace{0.125mm}y_{t}\,dW_{t}\hspace{0.25mm}, \end{align*} where $a \geq 0$ and $b\in \mathbb{R}$ are the mean reversion parameters and $\sigma \geq 0$ is the volatility.\medbreak\noindent In mathematical finance, IGBM is an example of a short rate model that can be both mean-reverting and non-negative. It is therefore suitable for modelling interest rates, stochastic volatilities and default intensities \cite{IGBMapproximation}. From a mathematical viewpoint, IGBM is one of the simplest SDEs that has no known method of exact simulation \cite{IGBMapplication}. By incorporating the ideas provided by the main theorem into the log-ODE method, we will produce a state-of-the-art numerical approximation of IGBM. Although the vector fields for IGBM are not bounded, our numerical evidence indicates that the method has a strong convergence rate of $O(h^{\frac{3}{2}})$ and a weak convergence rate of $O(h^{2})$. \subsection{Notation} Below is some of the notation that is used throughout the paper.\vspace{-0.75mm} \begin{center} \renewcommand{\arraystretch}{1.1} \begin{longtable}{cp{9.6cm}l} \toprule Symbol & Meaning & Page \vspace{0.5mm}\\ \toprule $W$ & a standard real-valued Brownian motion. & \hspace{2mm} 2\vspace{0.5mm}\\ $B$ & a standard real-valued Brownian bridge on $[\hspace{0.1mm}0\hspace{0.05mm},1\hspace{0.1mm}]$. & \hspace{2mm} 5\vspace{1.5mm}\\ $\mu$ & a Borel measure on $[\hspace{0.1mm}0,1]$ defined by a singular weight function. & \hspace{2mm} 5\vspace{1.5mm}\\ &\begin{math}\begin{aligned} \hspace{30mm}\mu(a,b) = \int_{a}^{b}\frac{1}{x(1-x)}\,dx, \end{aligned}\end{math} & \vspace{1.5mm} \\ & for all open intervals $(a,b)\subset [\hspace{0.1mm}0,1]$. &\vspace{1.5mm}\\ $\left\{e_{k}\right\}_{k\geq 1}$ & a family of Jacobi-like polynomials with $\deg\left(e_{k}\right) = k + 1$ that are orthogonal with respect to weight function $w(x) :=\hspace{0.25mm} \frac{1}{x(1-x)}\hspace{0.25mm}.$& \hspace{2mm} 5 \vspace{1mm} \\ $I_{k}$ & a time integral of $B$ times the polynomial $e_{k}(t)\,w(t)$ over $[\hspace{0.1mm}0\hspace{0.05mm},1\hspace{0.1mm}]$, & \hspace{2mm} 5 \vspace{1.5mm}\\ &\begin{math}\begin{aligned} \hspace{30mm}I_{k} = \int_{0}^{1}B_{t}\cdot\frac{e_{k}\left(t\right)}{t(1-t)}\,dt. \end{aligned}\end{math} & \vspace{1.5mm} \\ $K_B$ & the covariance function of $B$, that is $K_B(s,t) = \min(s,t) - st$. & \hspace{2mm} 5\vspace{1.5mm}\\ \hspace{2mm}$P_{k}^{(\alpha, \beta)}$ & the $k$-th order $(\alpha, \beta)$-Jacobi polynomial on $[-1,1\hspace{0.1mm}]$ $\left(\alpha, \beta > -1\right)$. & \hspace{1.25mm} 11\vspace{0.5mm}\\ \hspace{2mm}$Q_{k}$ & the $k$-th order Legendre polynomial on $[-1,1\hspace{0.1mm}]$, i.e. $Q_{k} = P_{k}^{(0, 0)}$. & \hspace{1.25mm} 13\vspace{0.5mm}\\ $y$ & a solution of the Stratonovich SDE on the finite interval $[\hspace{0.1mm}0\hspace{0.05mm},T\hspace{0.1mm}]$, & \hspace{1.25mm} 14 \vspace{1.5mm}\\ &\begin{math}\begin{aligned} \hspace{26mm} dy_{t} & = f_{0}(y_{t})\,dt + f_{1}(y_{t})\circ dW_{t}\hspace{0.25mm},\\[3pt] y_{0} & = \xi, \end{aligned}\end{math} & \vspace{1.5mm} \\ & where $y, \xi\in\mathbb{R}^{e}$, and $f_{i}:\mathbb{R}^{e}\rightarrow\mathbb{R}^{e}$ denote smooth vector fields. & \\ & (It\^{o} SDEs will be defined on fixed intervals with the same form) & \vspace{0.5mm}\\ $[s,t]$ & a general closed subinterval of $[\hspace{0.15mm}0\hspace{0.05mm},T\hspace{0.1mm}]$, usually considered small. & \hspace{1.25mm} 14\vspace{0.5mm} \\ $h$ & the step size that a numerical method uses, typically $h = t- s$. & \hspace{1.25mm} 14\vspace{0.25mm}\\ $W_{s,t}$ & the increment of Brownian motion over $[\hspace{0.1mm}s\hspace{0.05mm},t\hspace{0.1mm}]$, $W_{s,t} := W_{t} - W_{s}\hspace{0.05mm}$. & \hspace{1.25mm} 14 \vspace{0.5mm} \\ $\wideparen{W}$ & the Brownian parabola corresponding to $W$\hspace{-0.5mm} over some interval. & \hspace{1.25mm} 15 \vspace{1.5mm} \\ &\begin{math}\begin{aligned} \hspace{3.5mm}\wideparen{W}_{u} = W_{s} + \frac{u-s}{h}\,W_{s,t} + \frac{6\hspace{0.25mm}(u-s)(t-u)}{h^{2}}\,H_{s,t}\hspace{0.25mm},\hspace{2.9mm}\forall u\in [\hspace{0.1mm}s\hspace{0.05mm},t\hspace{0.1mm}]. \end{aligned}\end{math}& \vspace{1.5mm}\\ $Z$ & the Brownian arch corresponding to $W$ defined as $Z := W - \wideparen{W}$. & \hspace{1.25mm} 15 \vspace{0.5mm}\\ $H_{s,t}$ & the rescaled space-time L\'{e}vy area of Brownian motion on $[\hspace{0.1mm}s\hspace{0.05mm},t\hspace{0.1mm}]$, & \hspace{1.25mm} 15 \vspace{1.5mm} \\ &\begin{math}\begin{aligned} \hspace{22.5mm}H_{s,t} & = \frac{1}{h}\int_{s}^{t}W_{s,u} -\frac{u-s}{h}\,W_{s,t}\,du. \end{aligned}\end{math} & \vspace{1.5mm} \\ $L_{s,t}$ & the space-space-time L\'{e}vy area of Brownian motion over $[\hspace{0.1mm}s\hspace{0.05mm},t\hspace{0.1mm}]$, & \hspace{1.45mm} 17 \vspace{1.5mm}\\ &\begin{math}\begin{aligned} L_{s,t} & = \frac{1}{6}\hspace{-0.3mm}\left(\int_{s}^{t}\hspace{-1mm}\int_{s}^{u}\hspace{-1.5mm}\int_{s}^{v}\hspace{-1mm}\circ\,dW_{r}\hspace{-0.25mm}\circ dW_{v}\, du - 2\hspace{-0.3mm}\int_{s}^{t}\hspace{-1mm}\int_{s}^{u}\hspace{-1.5mm}\int_{s}^{v}\hspace{-1mm}\circ\,dW_{r}\, dv\circ dW_{u}\right. \\ &\hspace{10mm}\left. + \int_{s}^{t}\hspace{-1mm}\int_{s}^{u}\hspace{-1.5mm}\int_{s}^{v}\hspace{-1mm}\,dr \circ dW_{v}\hspace{-0.25mm}\circ dW_{u}\right)\hspace{-0.5mm}, \end{aligned}\end{math}\vspace{1.5mm}\\ $Y$ & an approximation for the true solution $y$ of a Stratonovich SDE. & \hspace{1.25mm} 19 \vspace{0.5mm}\\ $[\,\cdot\hspace{0.5mm}, \cdot\,]$ & the standard Lie bracket of vector fields, $[f_{0}, f_{1}] = f_{1}^{\prime}\hspace{0.125mm}f_{0} - f_{0}^{\prime}\hspace{0.125mm}f_{1}$. & \hspace{1.25mm} 19 \end{longtable} \end{center}\vspace{-5.75mm} \section{Main result} It was shown in \cite{Multiscalebm} that Brownian motion can be generated using Alpert-Rokhlin multiwavelets (see \cite{WaveletTheory}). The mother functions that generate this wavelet basis are supported on $[\hspace{0.1mm}0,1]$ and are defined using polynomials as follows:\medbreak \begin{definition}[Alpert-Rokhlin wavelets]\label{ARWavelet} For $q\geq 1$, define the $q$ functions $\phi^{q,1}, \cdots, \phi^{q,q} : [\hspace{0.15mm}0,1\hspace{0.15mm}]\hspace{-0.3mm}\rightarrow \mathbb{R}$ as piecewise polynomials of degree $q-1$ with pieces on $[\hspace{0.15mm}0,\frac{1}{2}\hspace{0.15mm}]$, $[\hspace{0.15mm}\frac{1}{2}, 1\hspace{0.15mm}]$ that satisfy the following conditions for all $p\in\left\{1,\,\cdots\,, q\hspace{0.15mm}\right\}$ and $t\in [\hspace{0.15mm}0, \frac{1}{2}\hspace{0.15mm})$\hspace{0.25mm}$:$ \begin{align} \phi^{q,p}(t) & = (-1)^{q+p-1}\phi^{q,p}(1 - t),\label{momentcondition1}\\[2pt] \int_{0}^{1}\phi^{q,p}(t)\phi^{q,r}(t)\,dt & = \delta_{qr}\hspace{0.25mm}, \hspace{3mm}\text{for}\hspace{2mm}1 \leq r \leq q, \label{momentcondition2}\\ \int_{0}^{1}t^{k}\phi^{q,p}(t)\,dt & = 0, \hspace{5.65mm}\text{for}\hspace{2.2mm}0 \leq k \leq q - 1. \label{momentcondition3} \end{align} The Alpert-Rokhlin multiwavelets of order $q$ can now be generated by translating and scaling the mother functions $\phi^{q,p}$. \begin{align*} \phi_{nk}^{q,p}(t) := \frac{1}{\sqrt{2^{n}}}\,\phi^{q,p}(2^{n}t - k), \end{align*} for $n\geq 0$ and $k\in\left\{0,\cdots, 2^{n}-1\right\}$. \end{definition}\medbreak\noindent Whilst our results will not be presented in terms of the above wavelets, we shall see that the polynomials of interest are directly related to conditions (\ref{momentcondition1}), (\ref{momentcondition2}) and (\ref{momentcondition3}). The main result of this paper gives an effective method for approximating sample paths of Brownian motion by a class of Jacobi-like polynomials. The proof is based on the interpretation of these polynomials as eigenfunctions of an integral operator defined by the Brownian bridge covariance function\footnote{The Brownian bridge is the centered Gaussian process with covariance $K_{B}(s,t) = \min(s,t)-st$.}. These orthogonal polynomials, which lie at the heart of this paper, will also help us interpret the geometrical features that certain normally distributed iterated integrals encode about the Brownian path.\medbreak \begin{theorem}[A polynomial Karhunen-Lo\`{e}ve theorem for the Brownian bridge]\label{waveletthm} Let $B$ denote a Brownian bridge on $[\hspace{0.1mm}0,1]$ and consider the Borel measure $\mu$ given by \begin{align*} \mu(a,b) := \int_{a}^{b}\frac{1}{x(1-x)}\,dx, \hspace{3.5mm}\text{for all open intervals}\hspace{1mm} (a,b)\subset [\hspace{0.1mm}0,1]. \end{align*}\noindent Then there exists a family of orthogonal polynomials $\{e_{k}\}_{k\geq 1}\hspace{-0.25mm}$ with $\deg\left(e_{k}\right)\hspace{-0.25mm} =\hspace{-0.25mm} k\hspace{-0.25mm} +\hspace{-0.25mm} 1$ and \begin{align*} \int_{0}^{1}e_{i}\,e_{j}\,d\mu = \delta_{ij}\hspace{0.125mm}, \end{align*}\noindent with $\delta_{ij}$ denoting the Kronecker delta, such that $B$ admits the following representation \begin{align}\label{polyapprox} B = \sum_{k=1}^{\infty}I_{k}e_{k}\hspace{0.125mm}, \end{align}\noindent where $\{I_{k}\}$ is the collection of independent centered Gaussian random variables with \begin{align}\label{polyintegrals} I_{k} := \int_{0}^{1}B_{t}\cdot\frac{e_{k}\left(t\right)}{t(1-t)}\,dt, \end{align} and \begin{align*} \operatorname{Var}(I_{k}) = \frac{1}{k(k+1)}\,. \end{align*} Furthermore, $\left\{e_{k}\right\}$ is an optimal orthonormal basis of $L^{2}([\hspace{0.1mm}0,1], \mu)$ for approximating $B$ by truncated series expansions with respect to the following weighted $L^{2}(\mathbb{P})$ norm \begin{align*} \left\|X\right\|_{L_{\mu}^{2}(\mathbb{P})} := \sqrt{\hspace{0.25mm}\mathbb{E}\hspace{-0.25mm}\left[\hspace{0.25mm}\int_{0}^{1}\left(X_{s}\right)^{2}\,d\mu(s)\right]}, \end{align*} where $X$ is a square $\mu$-integrable process. \end{theorem}\medbreak \begin{proof} Our argument is that of the Karhunen-Lo\`{e}ve theorem in general $L^{2}$ spaces. Note that $B$ is a square $\mu$-integrable process as \begin{align*} \mathbb{E}\left[\hspace{0.25mm}\int_{0}^{1}(B_{s})^{2}\,d\mu(s)\right] = \int_{0}^{1}\mathbb{E}\big[(B_{s})^{2}\big]\,d\mu(s) = \int_{0}^{1}s(1-s)\cdot\frac{1}{s(1-s)}\,ds = 1 < \infty. \end{align*} Let $K_{B}$ denote the covariance function for the standard Brownian bridge on $[\hspace{0.1mm}0, 1]$. Since $K_{B}(s,t) = \min(s,t) - st$, it can be shown by direct calculation that $K_{B}$ satisfies \begin{align*} \left\|K_{B}\right\|_{L^{2}([\hspace{0.1mm}0,1]^{2},\,\mu^{2})}^{2} = \int_{0}^{1}\int_{0}^{1}(\min(s,t) - st)^{2}\,d\mu(s)d\mu(t) = \frac{1}{3}\pi^{2} - 3 < \infty. \end{align*} Hence, it follows that the integral operator $T_{K} : L^{2}([\hspace{0.1mm}0,1],\,\mu)\rightarrow L^{2}([\hspace{0.1mm}0,1],\,\mu)$ given by \begin{align*} (T_{K}f)(t) := \int_{0}^{1}K_{B}(s,t)f(s)\,d\mu(s), \end{align*} is well-defined and continuous. In addition, the variance function $k_{B}(x) := K_{B}(x,x)$ for $x\in[\hspace{0.1mm}0,1]$ is $\mu$-integrable as \begin{align*} \int_{0}^{1}|k_{B}(x)|\,d\mu(x) = \int_{0}^{1}x(1-x)\cdot\frac{1}{x(1-x)}\,dx = 1 < \infty. \end{align*} Therefore, we can apply Mercer's theorem for kernels on general $L^{2}$ spaces (see \cite{Mercer}). It then follows from Mercer's theorem that there exists an orthonormal set $\{e_{k}\}_{k\geq 1}$ of $L^{2}([\hspace{0.1mm}0,1],\,\mu)$ consisting of eigenfunctions of $T_{K}$ such that the corresponding sequence of eigenvalues $\{\lambda_{k}\}_{k\geq 1}$ is non-negative. Moreover, the eigenfunctions corresponding to non-zero eigenvalues are continuous on $[\hspace{0.1mm}0,1]$ and the kernel $K_{B}$ has the representation \begin{align}\label{mercer} K_{B}(s,t) = \sum_{k=1}^{\infty}\lambda_{k} e_{k}(s)e_{k}(t), \end{align} where the series (\ref{mercer}) converges absolutely and uniformly on compact subsets of $[\hspace{0.1mm}0,1]$.\medbreak\noindent In the next part of the proof, we will see that each $e_{k}$ is a polynomial of degree $k+1$. As each $e_{k}$ is an eigenfunction of $T_{K}$, we have\vspace{0.5mm} \begin{align}\label{mercerv2} \int_{0}^{1}\frac{\min(s,t)-st}{s(1-s)}\,e_{k}(s)\,ds = \lambda_{k}e_{k}(t). \end{align} Since $e_{k}\in L^{2}([\hspace{0.1mm}0,1],\,\mu)$, it follows that $e_{k}(0) = 0$ and $e_{k}(1) = 0$ for each $k\geq 1$. Therefore by using the Leibniz integral rule to twice differentiate both sides of (\ref{mercerv2}) and then multiplying by $t(1-t)$, we observe that $e_{k}$ satisfies the differential equation \begin{align}\label{mercerv3} t(1-t)\lambda_{k}e_{k}^{\prime\prime}(t) + e_{k}(t) = 0. \end{align} Since $e_{k} \neq 0$, we have that $\lambda_{k}\neq 0$. Differentiating the LHS of the ODE (\ref{mercerv3}) produces \begin{align*} t(1-t)\frac{d^{2}}{dt^{2}}(e_{k}^{\prime}) + (1-2t)\frac{d}{dt}(e_{k}^{\prime})+\frac{1}{\lambda_{k}}e_{k}^{\prime} = 0. \end{align*} For $x\in[-1,1]$, we define the function \begin{align*} y_{k}(x) := e_{k}^{\prime}\left(\frac{1}{2}(1+x)\right)\hspace{-0.5mm}. \end{align*} Thus $y_{k}$ satisfies the following differential equation\vspace{1mm} \begin{align}\label{legendrede} (1-x^{2})y_{k}^{\prime\prime}(x) - 2xy_{k}^{\prime}(x) + \frac{1}{\lambda_{k}}y_{k}(x) = 0. \end{align} Remarkably, this is the Legendre differential equation \cite{ClassicalTheory}. It then follows using classical Sturm-Liouville theory that $\frac{1}{\lambda_{k}} = k(k+1)$ and $y_{k}$ is proportional to the $k$-th Legendre polynomial. Therefore, the derivative $e_{k}^{\prime}$ will be a constant multiple of the $k$-th shifted Legendre polynomial and hence each $e_{k}$ is a polynomial of degree $k+1$.\medbreak\noindent We can now define the following integrals for $k\geq 1$, \begin{align*} I_{k} := \int_{0}^{1}B_{t}\cdot\frac{e_{k}\left(t\right)}{t(1-t)}\,dt. \end{align*} It follows from Fubini's theorem that \begin{align*} \mathbb{E}[I_{k}] & = 0,\\[3pt] \mathbb{E}[I_{i}I_{j}] & = \mathbb{E}\left[\hspace{0.25mm}\int_{0}^{1}\int_{0}^{1}B_{s}B_{t}\,e_{i}(s)\,e_{j}(t)\,d\mu(s)\,d\mu(t)\right]\\ & = \int_{0}^{1}\int_{0}^{1}\mathbb{E}[B_{s}B_{t}]\,e_{i}(s)\,e_{j}(t)\,d\mu(s)\,d\mu(t)\\ & = \int_{0}^{1}\,e_{j}(t)\left(\int_{0}^{1}K_{B}(s,t)\,e_{i}(s)\,d\mu(s)\right)\,d\mu(t)\\[3pt] & = \lambda_{i}\delta_{ij}. \end{align*} Since each $I_{k}$ is defined by a linear functional on the same Gaussian process $B$, we see from the above that $\{I_{k}\}$ is a collection of uncorrelated (and therefore independent) Gaussian random variables with \begin{align*} \mathbb{E}[I_{k}] & = 0,\\[3pt] \operatorname{Var}(I_{k}) & = \frac{1}{k(k+1)}. \end{align*} Finally, the $L^{2}(\mathbb{P})$ convergence we require follows as \begin{align*} \mathbb{E}\left[\left(B_{t}-\sum_{k=1}^{N}I_{k}e_{k}(t)\right)^{2}\right] & = k_{B}(t) + \mathbb{E}\left[\hspace{0.25mm}\sum_{i,j=1}^{N}I_{i}I_{j}\,e_{i}(t)\hspace{0.25mm}e_{j}(t)\right] - 2\hspace{0.25mm}\mathbb{E}\left[\hspace{0.25mm}B_{t}\sum_{k=1}^{N}I_{k}e_{k}(t)\right]\\ & = k_{B}(t) + \sum_{i=1}^{N}\lambda_{i}e_{i}^{2}(t) - 2\hspace{0.25mm}\mathbb{E}\left[\hspace{0.25mm}\sum_{i=1}^{N}\int_{0}^{1}B_{s}B_{t}\,e_{i}(s)\hspace{0.25mm}e_{i}(t)\,d\mu(s)\right]\\[2pt] & = k_{B}(t) - \sum_{i=1}^{N}\lambda_{i}e_{i}^{2}(t), \end{align*} which converges to $0$ by Mercer's theorem (\ref{mercer}).\medbreak\noindent All that remains is to prove optimality for the truncated series expansions of (\ref{polyapprox}). Let $\left\{f_{k}\right\}_{k\geq 1}$ denote an orthonormal basis of $L^{2}([\hspace{0.1mm}0,1], \mu)$ such that \begin{align*} B = \sum_{k=1}^{\infty}J_{k}f_{k},\hspace{3mm}\text{where}\hspace{3mm} J_{k}:= \int_{0}^{1}B_{t}\,f_{k}(t)\,d\mu(t),\hspace{3mm}\forall k\geq 1. \end{align*} For $n\geq 1$, we consider an error process associated with the above: $r_{n} := \sum\limits_{k=n+1}^{\infty}J_{k}f_{k}\,$.\medbreak\noindent Then the square $L^{2}(\mathbb{P})$ norm of the $n$-th error process admits the following expansion, \begin{align*} \left\|r_{n}(t)\right\|_{L^{2}(\mathbb{P})}^{2} & = \mathbb{E}\hspace{-0.25mm}\left[\hspace{0.25mm}\sum_{i=n+1}^{\infty}\sum_{j=n+1}^{\infty}J_{i}J_{j}f_{i}(t)f_{j}(t)\right]\\[2pt] & = \sum_{i=n+1}^{\infty}\sum_{j=n+1}^{\infty}\mathbb{E}\hspace{-0.25mm}\left[\hspace{0.25mm}\int_{0}^{1}\int_{0}^{1}B_{s}B_{t}\,f_{i}(s)f_{j}(t)\,d\mu(s)\,d\mu(t)\right]\hspace{-0.25mm}f_{i}(t)f_{j}(t)\\[2pt] & = \sum_{i=n+1}^{\infty}\sum_{j=n+1}^{\infty}\left(\int_{0}^{1}\int_{0}^{1}K_{B}(s,t)\,f_{i}(s)f_{j}(t)\,d\mu(s)\,d\mu(t)\right)f_{i}(t)f_{j}(t). \end{align*} Integrating the above with respect to $\mu$ and using the orthogonality of $\{f_{k}\}_{k\geq 1}$ gives \begin{align*} \left\|r_{n}\right\|_{L_{\mu}^{2}(\mathbb{P})}^{2} = \int_{0}^{1}\left\|r(t)\right\|_{L^{2}(\mathbb{P})}^{2}\,d\mu(t) = \sum_{k=n+1}^{\infty}\int_{0}^{1}\int_{0}^{1}K_{B}(s,t)f_{k}(s)f_{k}(t)\,d\mu(s)\,d\mu(t). \end{align*} Note that any optimal orthonormal basis of $L^{2}([\hspace{0.1mm}0,1], \mu)$ solves the following problem: \begin{align*} \min_{f_{k}} \left\|r_{n}\right\|_{L_{\mu}^{2}(\mathbb{P})}^{2} \hspace{3mm} \text{subject to}\hspace{3mm} \left\|f_{k}\right\|_{L^{2}([\hspace{0.1mm}0,1],\mu)} = 1. \end{align*} By introducing Lagrange multipliers $\nu_{k}$, we wish to find functions $\{f_{k}\}$ that minimize \begin{align*} E_{n}[\left\{f_{k}\right\}]:=\hspace{-1mm}\sum_{k=n+1}^{\infty}\int_{0}^{1}\int_{0}^{1}\hspace{-0.35mm}K_{B}(s,t)f_{k}(s)f_{k}(t)d\mu(s)d\mu(t)- \nu_{k}\left(\int_{0}^{1}\hspace{-0.5mm}\big(f_{k}(s)\big)^{2}d\mu(s) - 1\right)\hspace{-0.35mm}. \end{align*} We will now consider the following square integrable functions, defined for $s,t\in\left(0,1\right)$: \begin{align*} \tilde{f}_{k}(t) := f_{k}(t)\cdot \frac{1}{\sqrt{t(1-t)}}\,, \hspace{5mm} \tilde{K}_{B}(s,t) := K_{B}(s,t)\cdot \frac{1}{\sqrt{s(1-s)}}\cdot\frac{1}{\sqrt{t(1-t)}}\,. \end{align*} Therefore it is enough to find a family of functions $\{\tilde{f}_{k}\}$ in $L^{2}([\hspace{0.1mm}0,1])$ which minimizes \begin{align*} \tilde{E}_{n}[\{\tilde{f}_{k}\}] := \sum_{k=n+1}^{\infty}\int_{0}^{1}\int_{0}^{1}\tilde{K}_{B}(s,t)\tilde{f}_{k}(s)\tilde{f}_{k}(t)\,ds\,dt - \nu_{k}\left(\int_{0}^{1}\big(\tilde{f}_{k}(s)\big)^{2}\,ds - 1\right)\hspace{-0.25mm}. \end{align*} To find a minimizer, we set the functional derivative of $\tilde{E}_{n}$ with respect to $\tilde{f}_{k}$ to zero. \begin{align*} \frac{\partial \tilde{E}_{n}}{\partial \tilde{f}_{k}(t)} = 2\int_{0}^{1}\tilde{K}_{B}(s,t)\tilde{f}_{k}(s)\,ds - 2\nu_{k}\tilde{f}_{k}(t) = 0. \end{align*} By using the definitions of $\tilde{f}_{k}$ and $\tilde{K}_{B}$, it is trivial to show the above is equivalent to \begin{align*} \int_{0}^{1}K_{B}(s,t)f_{k}(s)\,d\mu(s) = \nu_{k}f_{k}(t), \end{align*} which is satisfied if and only if $f_{k}$ are eigenfunctions of $T_{K}$. \end{proof}\medbreak\noindent This result can naturally be extended to express Brownian motion using polynomials.\medbreak \begin{theorem}\label{bmversion} If $W$ is a standard Brownian motion and $B$ is the associated bridge process on $[\hspace{0.1mm}0,1]$, then by Theorem \ref{waveletthm}, we have the below representation of $W$: \begin{align}\label{secpolyapprox} W = W_{1}e_{0} + \sum_{k=1}^{\infty}I_{k}e_{k}\hspace{0.125mm}, \end{align} where $e_{0}(t) := t\hspace{0.25mm}$ for $\hspace{0.25mm}t\in[\hspace{0.1mm}0,1]$, and the random variables $\{I_{k}\}$ are independent of $W_{1}$.\vspace{-4mm} \begin{figure}[h]\label{secparaboladiagram} \centering \includegraphics[width=0.975\textwidth]{"Second_brownian_decomposition_diagram".pdf}\vspace{-1.5mm} \caption{Brownian motion can be expressed as a sum of polynomials with independent weights. Moreover, these polynomials are orthogonal and capture different time integrals of the original path.} \end{figure} \end{theorem} \noindent In the rest of this section, we shall study the key objects introduced in Theorem \ref{waveletthm}. Since each orthogonal polynomial lies in $L^{2}([\hspace{0.1mm}0,1],\,\mu)$, it must have roots at $0$ and $1$. Therefore $e_{k}\cdot\frac{1}{t(1-t)}$ is itself a polynomial but with degree $k-1$, and one can repeatedly apply the integration by parts formula to the stochastic integrals $\{I_{k}\}$ defined by ($\ref{polyintegrals}$). This enables us to express each $I_{k}$ in terms of iterated integrals of Brownian motion. Moreover, as $e_{k}\cdot\frac{1}{t(1-t)}$ has precisely $k-2$ non-zero derivatives, the highest order iterated integral that is required to fully describe $I_{k}$ is $\int_{0<s_{1}<\cdots<s_{k}<1}B_{s_{1}}\,ds_{1}\,\cdots\,ds_{k}\hspace{0.125mm}$.\medbreak\noindent So by applying the integration by parts formula as above, we can construct a lower triangular $n\times n$ matrix $M_{n}$ with non-zero diagonal entries that characterizes the relationship between $\{I_{k}\}_{1\leq k\leq n}$ and a set of $n$ iterated integrals of Brownian motion.\medbreak\noindent Hence, for $n\geq 1$, we can express the $n$ independent Gaussian integrals $\{I_{k}\}_{1\leq k\leq n}$ as \begin{align}\label{itintrelate} \begin{pmatrix} I_{1} \\ \vdots \\ I_{n} \end{pmatrix} = M_{n} \begin{pmatrix} \int_{0<s_{1}<1}B_{s_{1}}\,ds_{1} \\ \vdots \\ \,\int_{0<s_{1}<\cdots<s_{n}<1}B_{s_{1}}\,ds_{1}\,\cdots\,ds_{n}\, \end{pmatrix}. \end{align} Since $M_{n}$ is an invertible matrix, it follows that the column vectors appearing in (\ref{itintrelate}) both encode the same information about the Brownian bridge. This enables us to establish a connection between Brownian motion, iterated integrals and polynomials.\medbreak \begin{theorem}\label{polyapproxthm} Consider\hspace{0.25mm} the\hspace{0.25mm} below\hspace{0.25mm} conditional\hspace{0.25mm} expectation\hspace{0.25mm} of\hspace{0.25mm} Brownian\hspace{0.25mm} motion, \begin{align}\label{unbiasedestimate} \hspace{-0.85mm}W_{t}^{n} := \mathbb{E}\hspace{-0.25mm}\left[W_{t}\,\Big|\,W_{1}\hspace{0.25mm}, \int_{0<s_{1}<1}W_{s_{1}}\,ds_{1}\hspace{0.25mm}, \cdots, \int_{0<s_{1}<\cdots<s_{n-1}<1}W_{s_{1}}\,ds_{1}\cdots \,ds_{n-1}\right]\hspace{-0.5mm}. \end{align}\medbreak\noindent where $t\in[\hspace{0.1mm}0,1]$. Then $W^{n}$ is the unique polynomial of degree $n$ with a root at $0$ that matches the increment $W_{1}$ and $n-1$ iterated time integrals of the path $\,W$ given by: \begin{align}\label{nminusoneintegrals} \int_{0<s_{1}<1}W_{s_{1}}\,ds_{1}\hspace{0.25mm}, \cdots, \int_{0<s_{1}<\cdots<s_{n-1}<1}W_{s_{1}}\,ds_{1}\cdots \,ds_{n-1}\hspace{0.125mm}. \end{align} \end{theorem}\medbreak \begin{proof} It is a direct consequence of (\ref{itintrelate}) that $W_{t}^{n} = \mathbb{E}[W_{t}\,|\,W_{1}, I_{1}, \cdots, I_{n-1}]$. Hence by (\ref{secpolyapprox}) and independence of the random variables $\{W_{1}, I_{1}, \cdots\}$, we have that \begin{align}\label{brownianpoly} W^{n} = W_{1}e_{0} + \sum_{k=1}^{n-1}I_{k}e_{k}\hspace{0.125mm}. \end{align} Thus $W^{n}$ is indeed a polynomial of degree $n$ with a root at $0$ and that matches the increment of the Brownian path. Without loss of generality we can now assume $n\geq 2$. All that remains is to argue $W^{n}$ matches the $n-1$ iterated integrals given in (\ref{nminusoneintegrals}). Using the orthogonality of $\{e_{k}\}$, it follows directly from (\ref{brownianpoly}) that for $1 \leq k \leq n -1$: \begin{align*} I_{k} & = \int_{0}^{1}\Big(W_{t} - W_{1}e_{0}(t)\Big)\cdot\frac{e_{k}\left(t\right)}{t(1-t)}\,dt\\[2pt] & = \int_{0}^{1}\left(W_{t}^{n} + \sum_{m=n}^{\infty}I_{m}e_{m}(t) - W_{1}e_{0}(t)\right)\cdot\frac{e_{k}\left(t\right)}{t(1-t)}\,dt\\[2pt] & = \int_{0}^{1}\Big(W_{t}^{n} - W_{1}e_{0}(t)\Big)\cdot\frac{e_{k}\left(t\right)}{t(1-t)}\,dt + \underbrace{\sum_{m=n}^{\infty}I_{m}\int_{0}^{1}\frac{e_{m}\left(t\right)e_{k}\left(t\right)}{t(1-t)}\,dt}_{=\,0.}\hspace{-0.3mm}. \end{align*} Hence $W^{n}$ matches the integrals of Brownian motion against polynomials with degree at most $n-1$. By the same argument used in the derivation of (\ref{itintrelate}), it follows that $W^n$ matches the various iterated time integrals given in the statement of the theorem. The uniqueness of $W^n$ is now a consequence of having $n+1$ different constraints. \end{proof}\medbreak \subsection{Properties of orthogonal polynomials} Although Theorem \ref{waveletthm} and Theorem \ref{bmversion} are interesting results from a theoretical point of view, both lack an explicit construction of the polynomials $\left\{e_{k}\right\}$ that could be implemented in practice. On the other hand, it was shown that the defining eigenfunction property of each $e_{k}$ implies that its derivative $e_{k}^{\prime}$ is proportional to the $k$-th shifted Legendre polynomial. Hence the family $\left\{e_{k}\right\}$ is the (normalized) shifted $(\alpha, \beta)$-Jacobi polynomials but with $\alpha = \beta = -1$. Since Jacobi polynomials are typically studied with $\alpha, \beta > -1$, it is necessary to show there exists a well-defined limit when the parameters approach $-1$.\medbreak \begin{figure}[h]\label{polynomialdiagams2} \centering \includegraphics[width=\textwidth]{"Polynomial_diagrams_2".pdf} \caption{Sample paths of Brownian motion with corresponding polynomial approximations.} \end{figure} \begin{definition}\label{jacobidef} For $k\geq 2$, the $k$-th degree $(\text{-}1,\text{-}1)$-Jacobi polynomial $P_{k}^{(\text{-}1, \text{-}1)}$ is \begin{align*} P_{k}^{(\text{-}1, \text{-}1)} := \lim_{\alpha, \beta \rightarrow -1^{+}}P_{k}^{(\alpha, \beta)}. \end{align*} \end{definition}\noindent Naturally, for this definition to be unambiguous, we will require the following lemma. \medbreak \begin{lemma}\label{jacobicheck} Let $P_{k}^{(\alpha, \beta)}$ denote the $k$-th degree $(\alpha, \beta)$-Jacobi polynomial on $[-1,1]$. Then for $k\geq 2$, there exists a real-valued polynomial $P_{k}$ such that $\big\|P_{k} - P_{k}^{(\alpha, \beta)}\big\|_{\infty} \rightarrow 0$ as $\alpha, \beta \rightarrow -1^{+}$. \end{lemma}\medbreak \begin{proof} Below is an identity for Jacobi polynomials, with $\alpha, \beta > - 1$, given in \cite{JacobiPoly}. \begin{align}\label{jacobifact} P_{k}^{(\alpha, \beta)}(x) = \frac{k+\alpha+\beta+1}{2}\int_{-1}^{x}P_{k-1}^{(\alpha + 1, \beta + 1)}(u)\,du, \hspace{5mm}\text{for all}\hspace{2mm} k \geq 2. \end{align} Therefore, we shall define the $k$-th degree polynomial $P_{k}$ over the interval $[-1,1]$ by \begin{align}\label{jacobifact2} P_{k}(x) := \frac{k-1}{2}\int_{-1}^{x}P_{k-1}^{(0, 0)}(u)\,du, \hspace{5mm}\text{for all}\hspace{2mm} k \geq 2. \end{align} It is straightforward to verify that $\lim_{\hspace{0.25mm}\alpha, \beta\, \rightarrow\, 0}\|\hspace{0.5mm}P_{n}^{(\alpha, \beta)} -\hspace{0.25mm} P_{n}^{(0, 0)}\hspace{0.25mm}\|_{\infty} = 0$ for $n\in\{0, 1\}$. So by induction and the recurrence relation for Jacobi polynomials (see \cite{JacobiPoly}), we have: \begin{align*} \big\|\hspace{0.5mm}P_{n}^{(\alpha, \beta)} -\hspace{0.25mm} P_{n}^{(0, 0)}\hspace{0.25mm}\big\|_{\infty} \rightarrow 0\hspace{2.5mm}\text{as}\,\,\,\alpha, \beta\, \rightarrow\, 0, \end{align*} for all $n\geq 0$. Hence by the dominated convergence theorem with (\ref{jacobifact}) and (\ref{jacobifact2}), it follows that $P_{k}^{(\alpha, \beta)}$ will converge pointwise to $P_{k}$ as $\alpha, \beta\, \rightarrow\, -1$ for each $k\geq 1$. Finally, the result follows as $P_{k}^{(\alpha, \beta)}$ and $P_{k}$ are always polynomials with degree $k$. \end{proof}\medbreak\noindent Using the above definition for $(\text{-}1,\text{-}1)$-Jacobi polynomials, we can give an explicit formula for the orthonormal polynomials $\{e_{k}\}_{k\geq 1}$ appearing in Theorems \ref{waveletthm} and \ref{bmversion}.\medbreak \begin{theorem}\label{explicitformulathm} Suppose each $e_{k}$ has a positive leading coefficient. Then for $k\geq 1$, \begin{align}\label{explicitformula} e_{k}(t) & = \frac{1}{k}\sqrt{k(k+1)(2k+1)}\,P_{k+1}^{(\text{-}1, \text{-}1)}(2t - 1), \hspace{4mm}\forall t\in[\hspace{0.1mm}0,1]. \end{align} \end{theorem} \begin{proof} The following identity for $(\alpha, \beta)$-Jacobi polynomials is stated in \cite{JacobiPoly}: \begin{align*} \int_{-1}^{1}(1-x)^{\alpha}(1+x)^{\beta}\big(P_{n}^{(\alpha,\beta)}(x)\big)^{2}\,dx = \frac{2^{\alpha + \beta + 1}}{2n+\alpha+\beta+1}\frac{\Gamma(n+\alpha+1)\,\Gamma(n+\beta+1)}{\Gamma(n+\alpha+\beta+1)\,n!}\,, \end{align*} for $n\geq 1$ and $\alpha, \beta > -1$. Applying the change of variables, $t := \frac{1}{2}(x+1)$, we have \begin{align*} \int_{0}^{1}t^{\beta}(1-t)^{\alpha}\big(P_{n}^{(\alpha,\beta)}(2t-1)\big)^{2}dt = \frac{1}{2n+\alpha+\beta+1}\frac{\Gamma(n+\alpha+1)\,\Gamma(n+\beta+1)}{\Gamma(n+\alpha+\beta+1)\,n!}\,, \end{align*} for $n\geq 1$ and $\alpha, \beta > -1$. By definition \ref{jacobidef}, taking the limit $\alpha, \beta \rightarrow -1^{+}$ will yield \begin{align*} \int_{0}^{1}\frac{1}{t(1-t)}\left(P_{n}^{(\text{-}1,\text{-}1)}(2t-1)\right)^{2}dt & = \frac{1}{2n-1}\frac{(n-1)!\,(n-1)!}{n!\,(n-2)!}\\ & = \frac{1}{2n-1}\frac{n-1}{n}\,,\hspace{5mm}\text{for all}\hspace{2mm} n\geq 2. \end{align*} Therefore by setting $k := n - 1$, we have \begin{align*} \int_{0}^{1}\frac{1}{t(1-t)}\left(\frac{1}{k}\sqrt{k(k+1)(2k+1)}\,P_{n}^{(\text{-}1,\text{-}1)}(2t-1)\right)^{2}\, dt = 1, \hspace{5mm}\text{for all}\hspace{1.25mm} k \geq 1. \end{align*} Recall that $e_{k}^{\prime}$ is proportional to the $k$-th shifted Legendre polynomial $P_{k}^{(0,0)}(2t-1)$. Similarly, we saw in the proof of Lemma \ref{jacobicheck} that the derivative of $P_{k+1}^{(\text{-}1,\text{-}1)}$ is $\frac{k}{2}\hspace{0.25mm}P_{k}^{(0,0)}$. As $e_{k}$ and $P_{k+1}^{(\text{-}1,\text{-}1)}$ are zero at their respective endpoints, we have that each $e_{k}$ must be proportional to $P_{k+1}^{(\text{-}1,\text{-}1)}(2t-1)$. The result now follows from the above calculations. \end{proof}\medbreak\noindent Having identified an explicit formula for the eigenfunctions $\{e_k\}$ in (\ref{explicitformula}), we shall now describe two methodologies for computing the Jacobi-like polynomials $\big\{P_{k}^{(\text{-}1,\text{-}1)}\big\}$.\smallbreak\noindent The first approach is to use the three-term recurrence relation in the theorem below.\medbreak \begin{theorem}[Recurrence relation for $(\text{-}1,\text{-}1)$-Jacobi polynomials] For $n\geq 2$\hspace{0.25mm}, \begin{align}\label{jacobirecurrence} n(n+2)P_{n+2}^{(\text{-}1,\text{-}1)}(x) & = (n+1)(2n+1)\hspace{0.25mm}xP_{n+1}^{(\text{-}1,\text{-}1)}(x) - n(n+1)P_{n}^{(\text{-}1,\text{-}1)}(x)\hspace{0.25mm}, \end{align} where the initial polynomials are given by \begin{align*} P_{2}^{(\text{-}1,\text{-}1)}(x) & = \frac{1}{4}\hspace{0.25mm}(x-1)(x+1)\hspace{0.25mm},\\[3pt] P_{3}^{(\text{-}1,\text{-}1)}(x) & = \frac{1}{2}\hspace{0.25mm}x(x-1)(x+1)\hspace{0.25mm}. \end{align*} \end{theorem}\vspace{-1.5mm} \begin{proof} The below recurrence relation for Jacobi polynomials is presented in \cite{JacobiPoly}, \begin{align*} & 2(k+1)(k+\alpha+\beta + 1)(2k+\alpha+\beta)P_{k+1}^{(\alpha,\beta)}(x)\\[3pt] &\hspace{5mm} = (2k + \alpha + \beta + 1)\hspace{-0.35mm}\left((2k + \alpha + \beta)(2k + \alpha + \beta + 2)x + \alpha^{2} - \beta^{2}\right)P_{k}^{(\alpha,\beta)}(x)\\[3pt] &\hspace{8.5mm} - 2(k+\alpha)(k+\beta)(2k+\alpha+\beta+2)\,P_{k-1}^{(\alpha,\beta)}(x)\hspace{0.25mm}, \end{align*} for $k\geq 1$ and $\alpha, \beta > -1$. By definition \ref{jacobidef}, it is possible to take the limit $\alpha,\beta\rightarrow 1^{+}$ provided that $k\geq 3$. Therefore, taking this limit and setting $n = k - 1 \geq 2$ produces \begin{align*} 4n^{2}(n+2)P_{n+2}^{(\text{-}1,\text{-}1)}(x) = 4n(n + 1)(2n + 1)xP_{n+1}^{(\text{-}1,\text{-}1)}(x) - 4n^{2}(n+1)\,P_{n}^{(\text{-}1,\text{-}1)}(x)\hspace{0.25mm}, \end{align*} for $n\geq 2$. Dividing the above by $4n$ gives the required recurrence relation (\ref{jacobirecurrence}). Finally the below formula, stated in \cite{JacobiPoly}, can be used to compute $P_{2}^{(\text{-}1,\text{-}1)}$ and $P_{3}^{(\text{-}1,\text{-}1)}$: \begin{align*} P_{n}^{(\alpha, \beta)}(x) = \frac{(-1)^{n}}{2^{n}n!}\left(1-x\right)^{-\alpha}\left(1+x\right)^{-\beta}\frac{d^{n}}{dx^{n}}\hspace{-0.25mm}\left(\left(1-x\right)^{\alpha + n}\left(1+x\right)^{\beta + n}\right),\hspace{3mm}\text{for}\,\, n\geq 2\hspace{0.25mm}. \end{align*} As before we take $\alpha,\beta\rightarrow 1^{+}$ in the above to obtain an explicit formula for $P_{n}^{(\text{-}1,\text{-}1)}$. \end{proof}\smallbreak\noindent In addition to computing these polynomials via a recurrence relation, it is also possible to represent each $P_{n}^{(\text{-}1,\text{-}1)}$ as the difference of two (rescaled) Legendre polynomials. Since the Legendre polynomials already have efficient implementations in the majority of high-level programming languages, this second approach is particularly appealing.\medbreak \begin{theorem}[Relationship between the Jacobi-like and Legendre polynomials] For $n\geq 1$, we have \begin{align*} P_{n+1}^{(\text{-}1,\text{-}1)}(x) = \frac{n}{4n+2}\hspace{0.25mm}\big(Q_{n+1}(x) - Q_{n-1}(x)\big), \end{align*} where $Q_{k}$ denotes the $k$-th degree Legendre polynomial defined on $[-1,1]$. \end{theorem}\medbreak \begin{proof} Recall that $\frac{d}{dx}\big(P^{(\text{-}1,\text{-}1)}_{n+1}\hspace{0.25mm}\big) = \frac{n}{2}P^{(0,0)}_{n}$ for $n\geq 1$, where $P^{(0,0)}_{n} (= Q_{n})$ is the $n$-th degree Legendre polynomial. Therefore differentiating both sides of (\ref{jacobirecurrence}) yields \begin{align*} \frac{1}{2}\hspace{0.25mm}n(n+1)(n+2)\hspace{0.25mm}Q_{n+1}(x) & = (n+1)(2n+1)P_{n+1}^{(\text{-}1,\text{-}1)}(x) + \frac{1}{2}\hspace{0.25mm}n(n+1)(2n+1)\hspace{0.25mm}x\hspace{0.25mm}Q_{n}(x) \\ &\hspace{5mm} - \frac{1}{2}\hspace{0.25mm}n(n-1)(n+1)\hspace{0.25mm}Q_{n-1}(x)\hspace{0.25mm}. \end{align*} Hence by simplifying and rearranging the above, we have that for $n\geq 1$, \begin{align*} (2n+1)P_{n+1}^{(\text{-}1,\text{-}1)}(x) & = \frac{1}{2}\hspace{0.25mm}n\hspace{0.25mm}\big(Q_{n+1}(x) - Q_{n-1}(x)\big)\\ &\hspace{5mm} + \frac{1}{2}\hspace{0.25mm}n\hspace{0.25mm}\big((n+1)\hspace{0.25mm}Q_{n+1}(x) - (2n+1)\hspace{0.25mm}x\hspace{0.25mm}Q_{n}(x) + n\hspace{0.25mm}Q_{n-1}(x)\big). \end{align*} We see the last term is zero by a recurrence relation for Legendre polynomials \cite{ClassicalTheory}. \end{proof}\medbreak\noindent In addition to viewing the polynomials $\{e_{k}\}$ as orthogonal with respect to the weight function $w(x) := \frac{1}{x(1-x)}$\hspace{0.25mm}, we can characterize them via their iterated time integrals. In particular, for $1\leq k\leq n -1$\hspace{0.25mm}, it follows from the integration by parts formula that \begin{align*} \hspace{10mm}\int_{0<s_{1}<\cdots<s_{k}<1}e_{n}(s_{1})\,ds_{1}\cdots \,ds_{k} & = \int_{0}^{1}e_{n}(s)\,d\hspace{-0.25mm}\left(\frac{1}{k!}s^{k}\right)\\ & = \frac{1}{(k-1)!}\int_{0}^{1}s^{k-1}e_{n}(s)\,ds\\ & = -\frac{1}{k!}\int_{0}^{1}s^{k}e_{n}^{\prime}(s)\,ds\\[4pt] & = 0\hspace{0.25mm}, \hspace{19.5mm}(\text{by the orthogonality of $e_{n}^{\prime}$}). \end{align*} Hence for $k\geq 1$, $e_{k}$ is a polynomial with degree $k+1$ that has roots at $0$ and $1$ as well as $k-1$ trivial iterated integrals against time. By additionally specifying the $k$-th iterated time integral, it is then possible to characterize the $k$-th polynomial $e_{k}$.\medbreak\noindent To conclude this section, we will address the relationship between the orthogonal Jacobi-like polynomials $\{e_{k}\}$ and the Alpert-Rokhlin wavelets given in definition \ref{ARWavelet}. Since each $e_{k}^{\prime}$ is proportional to the $k$-th shifted Legendre polynomial, the family of polynomials $\{e_{k}^{\prime}\}$ is orthogonal with respect to the standard $L^{2}([\hspace{0.1mm}0,1])$ inner product. This orthogonality is exactly what is needed to satisfy the conditions (\ref{momentcondition2}) and (\ref{momentcondition3}). Hence for any $q\geq 1$ there exists an Alpert-Rokhlin mother function of order $q$ that is a piecewise polynomial where both pieces can be rescaled and translated to give $e_{q-1}^{\prime}$. \section{Applications to SDEs} Consider the Stratonovich SDE on the interval $[\hspace{0.1mm}0,T]$ \begin{align}\label{ogsde} dy_{t} & = f_{0}(y_{t})\,dt + f_{1}(y_{t})\circ dW_{t}\hspace{0.125mm},\\ y_{0} & = \xi,\nonumber \end{align} where $\xi\in\mathbb{R}^{e}$ and $f_{i}$ denote bounded $C^{\infty}$ vector fields on $\mathbb{R}^{e}$ with bounded derivatives. It then follows from the standard Picard iteration argument that there exists a unique strong solution $y$ to (\ref{ogsde}). An important tool in the numerical analysis of this solution is the stochastic Taylor expansion (see chapter 5 of \cite{KloePlat} for a comprehensive review). For the purposes of this paper, we only require the following specific Taylor expansion.\medbreak \begin{theorem}[High order Stratonovich-Taylor expansion]\label{stochtaylorthm} Let $y$ denote the unique strong solution to (\ref{ogsde}) and let $0\leq s \leq t$. Then $y_{t}$ can be expanded as follows: \begin{align}\label{stochtaylorexp} y_{t} & = y_{s} + f_{0}(y_{s})\hspace{0.25mm}h + f_{1}(y_{s})\hspace{0.25mm}W_{s,t} + \frac{1}{2}\hspace{0.25mm}f_{1}^{\prime}(y_{s})f_{1}(y_{s})\hspace{0.25mm}W_{s,t}^{2} + \frac{1}{2}\hspace{0.25mm}f_{0}^{\prime}(y_{s})f_{0}(y_{s})\hspace{0.25mm}h^{2}\\ &\hspace{3.5mm} + f_{0}^{\prime}(y_{s})f_{1}(y_{s})\int_{s}^{t}\int_{s}^{u}\circ\, dW_{v}\,du + f_{1}^{\prime}(y_{s})f_{0}(y_{s})\int_{s}^{t}\int_{s}^{u}\,dv\circ\, dW_{u}\nonumber \\ &\hspace{3.5mm} + \frac{1}{6}\hspace{0.25mm}\big(\hspace{0.25mm}f_{1}^{\prime}(y_{s})f_{1}^{\prime}(y_{s})f_{1}(y_{s}) + f_{1}^{\prime\prime}(y_{s})(f_{1}(y_{s}), f_{1}(y_{s}))\big)W_{s,t}^{3}\nonumber \\ &\hspace{3.5mm} + \big(\hspace{0.25mm}f_{0}^{\prime}(y_{s})f_{1}^{\prime}(y_{s})f_{1}(y_{s}) + f_{0}^{\prime\prime}(y_{s})(f_{1}(y_{s}), f_{1}(y_{s}))\big)\int_{s}^{t}\int_{s}^{u}\int_{s}^{v}\circ\,dW_{r}\circ\, dW_{v}\, du\nonumber \\ &\hspace{3.5mm} + \big(\hspace{0.25mm}f_{1}^{\prime}(y_{s})f_{0}^{\prime}(y_{s})f_{1}(y_{s}) + f_{1}^{\prime\prime}(y_{s})(f_{0}(y_{s}), f_{1}(y_{s}))\big)\int_{s}^{t}\int_{s}^{u}\int_{s}^{v}\circ\,dW_{r}\, dv\circ dW_{u}\nonumber \\ &\hspace{3.5mm} + \big(\hspace{0.25mm}f_{1}^{\prime}(y_{s})f_{1}^{\prime}(y_{s})f_{0}(y_{s}) + f_{1}^{\prime\prime}(y_{s})(f_{1}(y_{s}), f_{0}(y_{s}))\big)\int_{s}^{t}\int_{s}^{u}\int_{s}^{v}\,dr \circ dW_{v}\circ dW_{u}\nonumber \\ &\hspace{3.5mm} + \frac{1}{24}\hspace{0.25mm}\big(\hspace{0.25mm}f_{1}^{\prime}(y_{s})f_{1}^{\prime}(y_{s})f_{1}^{\prime}(y_{s})f_{1}(y_{s}) + f_{1}^{\prime}(y_{s})f_{1}^{\prime\prime}(y_{s})(f_{1}(y_{s}), f_{1}(y_{s}))\nonumber \\ &\hspace{14.6mm} +\hspace{0.25mm} 3\hspace{0.25mm}f_{1}^{\prime\prime}(y_{s})(f_{1}^{\prime}(y_{s})f_{1}(y_{s}),f_{1}(y_{s})) + f_{1}^{\prime\prime\prime}(y_{s})(f_{1}(y_{s}),f_{1}(y_{s}),f_{1}(y_{s})) \big)\hspace{0.25mm}W_{s,t}^{4}\nonumber \\[3pt] &\hspace{3.5mm} + R(h, y_{s}),\nonumber \end{align} where $h:=t-s$ and the remainder term has the following uniform estimate for $h < 1$, \begin{align}\label{highorderestimate} \sup_{y_{s}\in\mathbb{R}^{e}}\left\|R(h, y_{s})\right\|_{L^{2}(\mathbb{P})} \leq C\, h^{\frac{5}{2}}, \end{align} where the constant $C > 0$ depends only on the vector fields of the differential equation. \end{theorem}\medbreak\noindent From a numerical perspective, the most challenging terms presented in (\ref{stochtaylorexp}) are those that involve non-trivial third order iterated integrals of Brownian motion and time. Moreover, the most significant source of discretization error that high order numerical methods will experience is generally due to approximating these stochastic integrals. By representing Brownian motion as a (random) polynomial plus independent noise, we shall derive a new optimal and unbiased estimator for these third order integrals.\medbreak \begin{theorem}\label{polyplusnoise} Let $W$ denote a standard real-valued Brownian motion on $[\hspace{0.1mm}0,1]$. Let $W^{n}$ be the unique $n$-th degree random polynomial with a root at $0$ and satisfying \begin{align*} \int_{0}^{1}u^{k}\,dW_{u}^{n} & = \int_{0}^{1}u^{k}\,dW_{u}\hspace{0.125mm},\hspace{3mm}\text{for}\hspace{2mm}k = 0, 1, \cdots, n - 1. \end{align*} Then\hspace{0.25mm} $W =\hspace{0.25mm} W^{n} + Z^{n}$\hspace{0.25mm}, where\hspace{0.25mm} $Z^{n}$ is a centered Gaussian process independent of\hspace{0.5mm} $W^{n}$.\medbreak\noindent Furthermore, $Z^{n}$ has the following covariance function: \begin{align*} \operatorname{cov}(Z^{n}_{s}, Z^{n}_{t}) = K_{B}(s,t) - \sum_{k=1}^{n-1}\lambda_{k}e_{k}(s)e_{k}(t), \hspace{3mm}\text{for}\hspace{2mm}s,t\in[\hspace{0.1mm}0,1], \end{align*} where $K_{B}$ denotes the standard Brownian bridge covariance function and $\{\lambda_{k}\}$, $\{e_{k}\}$ are the eigenvalues and eigenfunctions that were defined in the proof of Theorem \ref{waveletthm}. \end{theorem}\medbreak \begin{proof} It follows from the integration by parts formula that $W^{n}$ matches the increment and $n-1$ iterated time integrals of Brownian motion that appear in (\ref{nminusoneintegrals}). Hence $W^{n}$ is also the polynomial defined in Theorem \ref{polyapproxthm} and $W = W^{n} + Z^{n}$ where\vspace{-0.25mm} \begin{align*} W^{n} & = W_{1}e_{0} + \sum_{k=1}^{n-1}I_{k}e_{k}\hspace{0.125mm},\\ Z^{n} & = \sum_{k=n}^{\infty}I_{k}e_{k}\hspace{0.125mm}. \end{align*} Then by Theorem \ref{waveletthm}, $Z^{n}$ is a centered Gaussian process that is independent of $W^{n}$. In addition, the covariance function defining $Z^{n}$ can be directly computed as follows: \begin{align*} \operatorname{cov}(Z^{n}_{s}, Z^{n}_{t}) & = \operatorname{cov}\left(\sum_{i=n}^{\infty}I_{i}e_{i}(s), \sum_{j=n}^{\infty}I_{j}e_{j}(t)\right)\\ & = \sum_{k=n}^{\infty}\lambda_{k}e_{k}(s)e_{k}(t)\\ & = K_{B}(s,t) - \sum_{k=1}^{n-1}\lambda_{k}e_{k}(s)e_{k}(t),\hspace{3mm}\text{for}\hspace{2mm}s,t\in[\hspace{0.1mm}0,1]. \end{align*} Note that the final line is achieved using the representation of $K_{B}$ given by (\ref{mercer}). \end{proof}\medbreak\noindent The above theorem has an interesting conclusion, namely that there exist unbiased polynomial approximants of Brownian motion for which the error process can be independently estimated in an $L^{2}(\mathbb{P})$ sense. In particular, this theorem already has numerical applications in the case when $n=2$ and motivates the following definitions:\medbreak \begin{definition}\label{brownianparabola} The \textbf{standard Brownian parabola} $\,\wideparen{W}$ is the unique quadratic polynomial on $[\hspace{0.1mm}0,1]$ with a root at $0$ and satisfying \begin{align*} \wideparen{W}_{1} = W_{1}\hspace{0.125mm}, \hspace{5mm}\int_{0}^{1}\wideparen{W}_{u}\, du = \int_{0}^{1}W_{u}\,du. \end{align*} \end{definition}\vspace{-1mm} \begin{definition}\label{brownianarch} The \textbf{standard Brownian arch} $\,Z$ is the process $Z := W - \wideparen{W}$. By Theorem \ref{polyplusnoise}, $Z$ is the centered Gaussian process on $[\hspace{0.1mm}0,1]$ with covariance function \begin{align*} K_{Z}\left(s,t\right) = \min(s,t) - st - 3st(1-s)(1-t), \hspace{3mm}\text{for}\hspace{2mm}s,t\in[\hspace{0.1mm}0,1]. \end{align*} \end{definition}\vspace{-1mm} \begin{definition}\label{spacetimelevy} The rescaled \textbf{space-time L\'{e}vy area} of Brownian motion over an interval $[s,t]$ with length $h$ encodes the signed area of the associated bridge process, \begin{align*} H_{s,t} := \frac{1}{h}\int_{s}^{t}W_{s,u} - \frac{u-s}{h}\,W_{s,t}\,du. \end{align*} \end{definition}\vspace{-0.5mm} \begin{remark} Since $e_{1}(t) = \sqrt{6}\,t(1-t)$, we have that $H_{0,1}$ corresponds to $\frac{\sqrt{6}}{6}I_{1}$ as defined in Theorem \ref{waveletthm}. Thus, $H_{s,t}\sim \mathcal{N}\hspace{-0.25mm}\left(0,\frac{1}{12}h\right)$ and $H_{s,t}$ is independent of $\hspace{0.25mm}W_{s,t}\hspace{0.25mm}$. \end{remark}\medbreak\noindent By applying the natural scaling of Brownian motion, one can define the Brownian parabola and Brownian arch processes over any interval $[s,t]$ with finite size $h = t - s$. Whilst the Brownian arch can be viewed in a similar light to the Brownian bridge, there are clear qualitative and quantitative differences in their covariance functions. In particular, the Brownian arch has less variance at its midpoint compared to most points in $[s, t]$ $\big($by which we mean that $|\{u\in[s,t] : \operatorname{Var}(Z_{u}) \leq \operatorname{Var}(Z_{\frac{1}{2}(s+t)})\}| < \frac{1}{2}h\big)$. This is in contrast to the Brownian bridge, which has most variance at its midpoint. In fact, the Brownian parabola gives a relatively uniform estimate of the original path. \vspace{-2.5mm} \begin{figure}[h]\label{variancediagram} \centering \includegraphics[width=0.825\textwidth]{"Variance_Diagram".pdf} \caption{Variance profile of the standard Brownian arch.}\vspace{-2.5mm} \end{figure}\medbreak\noindent Using these new definitions, we can study the high order integrals appearing in (\ref{stochtaylorexp}).\medbreak \begin{theorem}[Conditional expectation of a non-trivial Brownian time integral]\vspace{-2.5mm}\label{firstcondthm} \begin{align}\label{firstcondexp} \mathbb{E}\hspace{-0.25mm}\left[\hspace{0.25mm}\int_{s}^{t}W_{s,u}^{2}\,du\, \Big|\, W_{s,t}\hspace{0.125mm}, H_{s,t}\right] = \frac{1}{3}hW_{s,t}^{2} + hW_{s,t}H_{s,t} + \frac{6}{5}hH_{s,t}^{2} + \frac{1}{15}h^{2}. \end{align} \end{theorem} \begin{proof} By the natural Brownian scaling it is enough to prove the result on $[\hspace{0.1mm}0,1]$. Recall that $W = \wideparen{W} + Z$ where the parabola $\wideparen{W}$ is completely determined by $\left(W_{1}, H_{1}\right)$ and $Z$ is independent of $\left(W_{1}, H_{1}\right)$. This leads to a decomposition for the LHS of (\ref{firstcondexp}). \begin{align*} \mathbb{E}\hspace{-0.25mm}\left[\hspace{0.25mm}\int_{0}^{1}\hspace{-0.25mm}W_{u}^{2}\,du\, \Big|\, W_{1}, H_{1}\right] & = \mathbb{E}\hspace{-0.25mm}\left[\hspace{0.25mm}\int_{0}^{1}\left(\hspace{0.25mm}\wideparen{W}_{u}+Z_{u}\right)^{2}\,du\, \Big|\, W_{1}, H_{1}\right]\\[2pt] & = \mathbb{E}\hspace{-0.25mm}\left[\hspace{0.25mm}\int_{0}^{1}\wideparen{W}_{u}^{2}\,du + 2\int_{0}^{1}\wideparen{W}_{u}\hspace{0.25mm}Z_{u}\,du + \int_{0}^{1}Z_{u}^{2}\,du\,\Big|\, W_{1}, H_{1}\right]\\[2pt] & = \int_{0}^{1}\wideparen{W}_{u}^{2}\,du + 2\int_{0}^{1}\wideparen{W}_{u}\,\mathbb{E}\left[Z_{u}\right]\,du + \int_{0}^{1}\mathbb{E}\hspace{-0.25mm}\left[Z_{u}^{2}\hspace{0.25mm}\right]\,du\\[3pt] & = \int_{0}^{1}\hspace{-1mm}\big(uW_{1} + 6u(1-u)H_{1}\big)^{2}du +\hspace{-0.1mm}\int_{0}^{1}\hspace{-0.75mm}u - u^{2} - 3u^{2}(1-u)^{2}\,du. \end{align*} The result now follows by evaluating the above integrals. \end{proof}\medbreak\noindent The above theorem has practical applications for SDE simulation as $W_{s,t}$ and $H_{s,t}$ are independent Gaussian random variables and can be easily generated or approximated. That said, we should first discuss how the iterated integrals within (\ref{stochtaylorexp}) are connected. \medbreak \begin{definition}\label{spacespacetimelevy} The \textbf{space-space-time L\'{e}vy area} of Brownian motion over an interval $[s,t]$ is defined as\vspace{-0.5mm} \begin{align*} L_{s,t} & := \frac{1}{6}\left(\int_{s}^{t}\int_{s}^{u}\int_{s}^{v}\circ\,dW_{r}\circ\, dW_{v}\, du - 2\int_{s}^{t}\int_{s}^{u}\int_{s}^{v}\circ\,dW_{r}\, dv\circ dW_{u}\right. \\ &\hspace{11mm}\left. + \int_{s}^{t}\int_{s}^{u}\int_{s}^{v}\,dr \circ dW_{v}\circ dW_{u}\right).\nonumber \end{align*} \end{definition}\noindent We can interpret $L_{s,t}$ as an area between the processes $\{W_{s,u}\}_{u\in[s,t]}$ and $\{H_{s,u}\}_{u\in[s,t]}$. Moreover, rough path theory provides an algebraic structure (called the log-signature) that relates $(W_{s,t}, H_{s,t}, L_{s,t})$ to the iterated integrals of space-time Brownian motion and ultimately to SDE solutions via the log-ODE method (see \cite{LogSig} for an overview). For our purposes, it is enough to give formulae relating these L\'{e}vy areas to integrals.\medbreak \begin{theorem}\label{logsigrelation} Let $H_{s,t}$ and $L_{s,t}$ denote the L\'{e}vy areas of Brownian motion given by definitions \ref{spacetimelevy} and \ref{spacespacetimelevy} respectively. Then the following integral relationships hold, \begin{align*} \int_{s}^{t}\int_{s}^{u}\circ\, dW_{v}\,du & = \frac{1}{2}hW_{s,t} + hH_{s,t}\hspace{0.125mm},\\ \int_{s}^{t}\int_{s}^{u}\,dv\circ\, dW_{u} & = \frac{1}{2}hW_{s,t} - hH_{s,t}\hspace{0.125mm},\\ \int_{s}^{t}\int_{s}^{u}\int_{s}^{v}\circ\,dW_{r}\circ\,dW_{v}\, du & = \frac{1}{6}hW_{s,t}^{2} + \frac{1}{2}hW_{s,t}H_{s,t} + L_{s,t}\hspace{0.125mm},\\ \int_{s}^{t}\int_{s}^{u}\int_{s}^{v}\circ\,dW_{r}\, dv\circ dW_{u} & = \frac{1}{6}hW_{s,t}^{2} - 2\hspace{0.125mm}L_{s,t}\hspace{0.125mm},\\ \int_{s}^{t}\int_{s}^{u}\int_{s}^{v}\,dr \circ dW{v}\circ dW_{u} & = \frac{1}{6}hW_{s,t}^{2} - \frac{1}{2}hW_{s,t}H_{s,t} + L_{s,t}\hspace{0.125mm}. \end{align*} \end{theorem}\vspace{-0.5mm} \begin{proof} The result follows from numerous applications of integration by parts. \end{proof}\medbreak\noindent We can now present the new unbiased estimator for third order iterated integrals of Brownian motion and time. The proposed estimator is fast to compute and the best $L^{2}(\mathbb{P})$ approximation of these integrals that is measurable with respect to $\left(W_{s,t}, H_{s,t}\right)$.\medbreak \begin{theorem}[Conditional moments of Brownian space-space-time L\'{e}vy area]\vspace{-3.5mm}\label{secondcondthm} \begin{align} \mathbb{E}\left[L_{s,t}\,\big|\,W_{s,t}\hspace{0.125mm}, H_{s,t}\right] & = \frac{1}{30}\hspace{0.25mm}h^{2} + \frac{3}{5}\hspace{0.25mm}hH_{s,t}^{2}\,,\label{secondcondexp}\\[3pt] \operatorname{Var}\left(L_{s,t}\,\big|\,W_{s,t}\hspace{0.125mm}, H_{s,t}\right) & = \frac{11}{25200}\hspace{0.25mm}h^{4} + h^{3}\left(\frac{1}{720}\hspace{0.25mm}W_{s,t}^{2} + \frac{1}{700}\hspace{0.25mm}H_{s,t}^{2}\right)\hspace{-0.5mm}.\label{firstcondvar} \end{align} \end{theorem}\vspace{-1mm} \begin{proof} The expectation (\ref{secondcondexp}) is simply a consequence of Theorems \ref{firstcondthm} and \ref{logsigrelation}. Without loss of generality, we will consider the above conditional variance on $[\hspace{0.1mm}0,1]$. Since $\wideparen{W}$ is determined using the increment $W_{1}$ and space-time L\'{e}vy area $H_{1}$, we have \begin{align*} \operatorname{Var}\left(\hspace{0.25mm}\int_{0}^{1}W_{u}^{2}\,du\, \Big|\, W_{1}, H_{1}\right) & = \operatorname{Var}\left(\hspace{0.25mm}\int_{0}^{1}\wideparen{W}_{u}^{2}\,du + 2\int_{0}^{1}\wideparen{W}_{u}\hspace{0.25mm}Z_{u}\,du +\hspace{-0.25mm}\int_{0}^{1}\hspace{-0.5mm}Z_{u}^{2}\,du\, \Big|\, W_{1}, H_{1}\right)\\[2pt] & = 4\operatorname{Var}\left(\hspace{0.25mm}\int_{0}^{1}\wideparen{W}_{u}\hspace{0.25mm}Z_{u}\,du\, \Big|\, W_{1}, H_{1}\right) + \operatorname{Var}\left(\hspace{0.25mm}\int_{0}^{1}\hspace{-0.5mm}Z_{u}^{2}\,du\right)\\ &\hspace{7.5mm} + 4\operatorname{cov}\left(\hspace{0.25mm}\int_{0}^{1}\wideparen{W}_{u}\,Z_{u}\,du\hspace{0.25mm}, \int_{0}^{1}\hspace{-0.5mm}Z_{u}^{2}\,du\,\Big|\, W_{1}\hspace{0.125mm}, H_{1}\right)\hspace{-0.25mm}. \end{align*} Recall $Z = \sum\limits_{k=2}^{\infty}I_{k}\hspace{0.25mm}e_{k}$ where $\left\{I_{k}\right\}$ are independent centered Gaussian random variables.\medbreak\noindent In particular, this means that $Z$ and $-Z$ have the same law. Therefore, we have that \begin{align*} \mathbb{E}\hspace{-0.25mm}\left[\hspace{0.25mm}\int_{0}^{1}\wideparen{W}_{u}\hspace{0.25mm}Z_{u}\,du\, \Big|\, W_{1}\hspace{0.125mm}, H_{1}\hspace{0.25mm}\right] = 0\hspace{0.25mm}, \end{align*}\vspace{-2mm} \begin{align*} \operatorname{cov}\left(\hspace{0.25mm}\int_{0}^{1}\wideparen{W}_{u}\hspace{0.25mm}Z_{u}\,du\hspace{0.25mm}, \int_{0}^{1}Z_{u}^{2}\,du\,\Big|\, W_{1}\hspace{0.125mm}, H_{1}\right) = 0\hspace{0.25mm}. \end{align*} The remaining two terms were resolved with assistance from Wolfram Mathematica. \begin{align*} &\operatorname{Var}\left(\hspace{0.25mm}\int_{0}^{1}\wideparen{W}_{u}\hspace{0.25mm}Z_{u}\,du\, \Big|\, W_{1}\hspace{0.125mm}, H_{1}\right)\\ &\hspace{10mm} = \int_{0}^{1}\int_{0}^{1}\wideparen{W}_{u}\hspace{0.25mm}\wideparen{W}_{v}\,\mathbb{E}\left[Z_{u}\hspace{0.25mm}Z_{v}\,|\, W_{1}\hspace{0.125mm}, H_{1}\right]\,du\,dv\\ &\hspace{10mm} = \int_{0}^{1}\int_{0}^{1}\wideparen{W}_{u}\hspace{0.25mm}\wideparen{W}_{v}\big(\hspace{-0.25mm}\min(u,v) - uv - 3uv(1-u)(1-v)\big)\,du\,dv\\[1pt] &\hspace{10mm} = 2\int_{0}^{1}\wideparen{W}_{v}\int_{0}^{v}u\hspace{0.25mm}\wideparen{W}_{u}\,du\,dv - \left(\int_{0}^{1}u\hspace{0.25mm}\wideparen{W}_{u}\,du\right)^{2} - 3\left(\hspace{0.25mm}\int_{0}^{1}u(1-u)\wideparen{W}_{u}\,du\right)^{2}\\[1pt] &\hspace{10mm} = 2\left(\hspace{0.125mm}\frac{1}{15}\hspace{0.25mm}W_{1}^{2}+\frac{13}{60}\hspace{0.25mm}W_{1}H_{1}+\frac{13}{70}\hspace{0.25mm}H_{1}^{2}\right) - \left(\hspace{0.125mm}\frac{1}{3}\hspace{0.25mm}W_{1} + \frac{1}{2}\hspace{0.25mm}H_{1}\right)^{2} -3\left(\hspace{0.125mm}\frac{1}{12}\hspace{0.25mm}W_{1}+\frac{1}{5}\hspace{0.25mm}H_{1}\right)^{2}\\[3pt] &\hspace{10mm} = \frac{1}{720}\hspace{0.25mm}W_{1}^{2} + \frac{1}{700}\hspace{0.25mm}H_{1}^{2}\hspace{0.25mm}. \end{align*}\vspace{-3.5mm} \begin{align*} \operatorname{Var}\left(\int_{0}^{1}Z_{u}^{2}\,du\right) & = \mathbb{E}\hspace{-0.25mm}\left[\left(\hspace{0.25mm}\int_{0}^{1}Z_{u}^{2}\,du\right)^{2}\hspace{0.25mm}\right] - \left(\mathbb{E}\hspace{-0.25mm}\left[\hspace{0.25mm}\int_{0}^{1}Z_{u}^{2}\,du\right]\right)^{2}\\[1pt] & = \int_{0}^{1}\int_{0}^{1}\mathbb{E}\hspace{-0.25mm}\left[Z_{u}^{2}\hspace{0.25mm}Z_{v}^{2}\right]\,du\,dv - \left(\mathbb{E}\left[\hspace{0.25mm}\int_{0}^{1}Z_{u}^{2}\,du\right]\right)^{2}\\[1pt] & = \int_{0}^{1}\int_{0}^{1}\mathbb{E}\hspace{-0.25mm}\left[Z_{u}^{2}\hspace{0.25mm}\right]\mathbb{E}\hspace{-0.25mm}\left[Z_{v}^{2}\hspace{0.25mm}\right] + 2\hspace{0.25mm}\big(\hspace{0.25mm}\mathbb{E}[Z_{u}\hspace{0.25mm}Z_{v}]\hspace{0.25mm}\big)^{2}\,du\,dv - \left(\mathbb{E}\hspace{-0.25mm}\left[\hspace{0.25mm}\int_{0}^{1}Z_{u}^{2}\,du\right]\right)^{2}\\[3pt] & = 4\int_{0}^{1}\int_{0}^{v}\big(u - uv - 3uv(1-u)(1-v)\big)^{2}\,du\,dv \\[3pt] & = \frac{11}{6300}\hspace{0.25mm}. \end{align*} By Theorem \ref{logsigrelation}, the above gives an explicit formula for the conditional variance (\ref{firstcondvar}). \begin{align*} \operatorname{Var}\left(L_{1}\,\big|\,W_{1}\hspace{0.125mm}, H_{1}\right) = \operatorname{Var}\left(\frac{1}{2}\int_{0}^{1}W_{u}^{2}\,du\, \Big|\, W_{1}\hspace{0.125mm}, H_{1}\right) = \frac{11}{25200} + \frac{1}{720}\hspace{0.25mm}W_{1}^{2} + \frac{1}{700}\hspace{0.25mm}H_{1}^{2}\hspace{0.25mm}. \end{align*} By the natural Brownian scaling, the result on the interval $[s,t]$ directly follows. \end{proof}\medbreak \begin{remark} The conditional variance (\ref{firstcondvar}) allows one to estimate local $L^{2}(\mathbb{P})$ errors for certain numerical methods and thus may be useful when choosing step sizes. \end{remark}\medbreak\noindent Therefore in order to propagate a numerical solution of (\ref{ogsde}) over an interval $[s,t]$, one can generate $\left(W_{s,t}, H_{s,t}\right)$ exactly and then approximate $L_{s,t}$ using Theorem \ref{secondcondthm}. However, there are many numerical methods that could be used to solve a given SDE. \subsection{Examples of ODE methods} We will consider the following two methods:\medbreak \begin{definition}[High order log-ODE method]\label{logodedef} For a fixed number of steps $N$ we can construct a numerical solution $\left\{Y_{k}\right\}_{0\leq k \leq N}$ of (\ref{ogsde}) by setting $Y_{0} := \xi$ and for each $k \in [0 \mathrel{{.}\,{.}}\nobreak N-1]$, defining $Y_{k+1}$ to be the solution at $u=1$ of the following ODE: \begin{align}\label{logode} \frac{dz}{du} & = f_{0}(z)\hspace{0.25mm}h + f_{1}(z)\hspace{0.25mm}W_{t_{k},t_{k+1}} + [f_{1}, f_{0}](z)\cdot hH_{t_{k},t_{k+1}}\\ &\hspace{14.5mm} + [f_{1}, [f_{1}, f_{0}]](z)\cdot \mathbb{E}\big[\hspace{0.25mm}L_{t_{k},t_{k+1}}\, \big|\, W_{t_{k},t_{k+1}}\hspace{0.125mm}, H_{t_{k},t_{k+1}}\hspace{0.25mm}\big],\nonumber\\ z_{0} & = Y_{k}\hspace{0.125mm},\nonumber \end{align}\noindent where $h := \frac{T}{N}$, $t_{k} := kh$ and $[\,\cdot\hspace{0.5mm}, \cdot\,]$ denotes the standard Lie bracket of vector fields. \end{definition}\medbreak \begin{definition}[The parabola-ODE method]\label{parabolaodedef} For a fixed number of steps $N$ we can construct a numerical solution $\left\{Y_{k}\right\}_{0\leq k \leq N}$ of (\ref{ogsde}) by setting $Y_{0} := \xi$ and for each $k \in [0 \mathrel{{.}\,{.}}\nobreak N-1]$, defining $Y_{k+1}$ to be the solution at $u=1$ of the following ODE: \begin{align}\label{paraode1} \frac{dz}{du} & = f_{0}(z)\hspace{0.25mm}h + f_{1}(z)\hspace{0.25mm}\big(W_{t_{k},t_{k+1}} + (6 - 12u)\hspace{0.25mm}H_{t_{k},t_{k+1}}\big),\\ z_{0} & = Y_{k}\hspace{0.125mm},\nonumber \end{align} where\hspace{0.125mm} $h := \frac{T}{N}$ and\hspace{0.125mm} $t_{k} := kh$. \end{definition}\medbreak\noindent In both numerical methods the true solution $y$ at time $t_{k}$ can be approximated by $Y_{k}$. Whilst there are different ways of interpolating between the successive approximations $Y_{k}$ and $Y_{k+1}$, for this paper we will simply interpolate between such points linearly. To analyse the above methods, we shall first note the key differences between them. The first important distinction between the two methods is a purely practical one. Although both of these methods involve computing a numerical solution of an ODE, the parabola method does not require one to explicitly resolve vector field derivatives. The second significant difference can be seen in the Taylor expansions of the methods.\medbreak \begin{theorem}\label{localerrors} Let $Y^{\text{log}}$ be the one-step approximation defined by the log-ODE method on the interval $[s,t]$ with initial value $Y_{0}^{\text{log}} = y_{s}$. Then for sufficiently small $h$ \begin{align}\label{logodestep} Y_{1}^{\text{log}} = y_{t} - [f_{1}, [f_{1}, f_{0}]] (y_{s})\hspace{0.5mm}\big(\hspace{0.25mm}L_{s,t} - \mathbb{E}\big[\hspace{0.25mm}L_{s,t}\, \big|\, W_{s,t}\hspace{0.125mm}, H_{s,t}\hspace{0.25mm}\big]\big) + O(h^{\frac{5}{2}}). \end{align} Similarly, let $Y^{\text{para}}$ denote the one-step approximation given by the parabola-ODE method on the interval $[s,t]$ with the same initial value. Then for sufficiently small $h$ \begin{align}\label{parabolaodestep} Y_{1}^{\text{para}} = y_{t} - [f_{1}, [f_{1}, f_{0}]] (y_{s})\left(L_{s,t} - \frac{3}{5}H_{s,t}^{2}\right) + O(h^{\frac{5}{2}}). \end{align} Note that $O(h^{\frac{5}{2}})$ denotes terms which can be estimated in an $L^{2}(\mathbb{P})$ sense as in (\ref{highorderestimate}). \end{theorem}\medbreak \begin{proof} In order to derive (\ref{logodestep}), we must compute the Taylor expansion of (\ref{logode}). Let $F$ denote the vector field defined in (\ref{logode}) that was constructed from $f_{0}$ and $f_{1}$. Then $F$ is smooth, and it follows from the classical Taylor's theorem for ODEs that \begin{align*} Y_{1}^{\text{log}} & = y_{s} + F(y_{s}) + \frac{1}{2}F^{\prime}(y_{s})F(y_{s}) + \frac{1}{6}F^{\prime\prime}(y_{s})(F(y_{s}), F(y_{s})) + \frac{1}{6}F^{\prime}(y_{s})F^{\prime}(y_{s})F(y_{s})\\ &\hspace{8mm} + \frac{1}{24}F^{\prime}(y_{s})F^{\prime}(y_{s})F^{\prime}(y_{s})F(y_{s}) + \frac{1}{24}F^{\prime}(y_{s})F^{\prime\prime}(y_{s})(F(y_{s}), F(y_{s})) \\[2pt] &\hspace{8mm} + \,\frac{1}{8}F^{\prime\prime}(y_{s})(F^{\prime}(y_{s})F(y_{s}),F(y_{s})) + \frac{1}{24}F^{\prime\prime\prime}(y_{s})(F(y_{s}),F(y_{s}), F(y_{s}))\\ &\hspace{8mm} + \frac{1}{24}\int_{0}^{1}(1-u)^{4}\frac{d^5}{du^{5}}\big(Y^{\text{log}}\big)(u)\,du. \end{align*} We shall first consider the remainder term, which can be directly estimated as follows: \begin{align*} \left\|\,\int_{0}^{1}(1-u)^{4}\frac{d^5}{du^{5}}\big(Y^{\text{log}}\big)(u)\,du\,\right\|_{L^{2}(\mathbb{P})} & \leq \sup_{u\in[\hspace{0.1mm}0,1]} \left\|\frac{d^5}{du^{5}}\big(Y^{\text{log}}(u)\big)\right\|_{L^{2}(\mathbb{P})}. \end{align*} One can define the degree of each term in the above Taylor expansion by counting the number of times functions from $\{F, F^{\prime}, F^{\prime\prime}, \cdots\}$ appear. Therefore, after expanding the fifth derivative of $Y^{\text{log}}$ we can see that the remainder term has a degree of five. Since the largest component of $F$ is $f_{1}(\cdot)\hspace{0.25mm}W_{s,t}$, both $F$ and its derivatives are $O(h^{\frac{1}{2}})$. Hence the remainder term in the above Taylor expansion will be $O(h^{\frac{5}{2}})$ as in (\ref{highorderestimate}). Moreover, the only terms of degree four that are not $O(h^{\frac{5}{2}})$ are those involving $W_{s,t}^{4}$.\\ It is now enough to analyse just the terms appearing in the first line of the expansion. By substituting the formula for $F$ given by (\ref{logode}) into the first line and then rearranging the resulting terms, we can obtain a Taylor expansion for $Y_{1}^{\text{log}}$ that resembles (\ref{stochtaylorexp}) as \begin{align*} F(y_{s}) & = f_{0}(y_{s})\hspace{0.25mm}h + f_{1}(y_{s})\hspace{0.25mm}W_{s,t} + \big(f_{0}^{\prime}(y_{s})f_{1}(y_{s}) - f_{1}^{\prime}(y_{s})f_{0}(y_{s})\big)\hspace{0.25mm} hH_{s,t}\\[4pt] &\hspace{5mm} + \big(f_{0}^{\prime}(y_{s})f_{1}^{\prime}(y_{s})f_{1}(y_{s}) + f_{0}^{\prime\prime}(y_{s})(f_{1}(y_{s}), f_{1}(y_{s}))\\[4pt] &\hspace{10mm} - 2\hspace{0.25mm}f_{1}^{\prime}(y_{s})f_{0}^{\prime}(y_{s})f_{1}(y_{s}) - 2\hspace{0.25mm}f_{1}^{\prime\prime}(y_{s})(f_{0}(y_{s}), f_{1}(y_{s}))\\ &\hspace{10mm} + f_{1}^{\prime}(y_{s})f_{1}^{\prime}(y_{s})f_{0}(y_{s}) + f_{1}^{\prime\prime}(y_{s})(f_{1}(y_{s}), f_{0}(y_{s}))\big)\hspace{-0.5mm}\left(\frac{1}{30}\hspace{0.25mm}h^{2} + \frac{3}{5}\hspace{0.25mm}hH_{s,t}^{2}\right)\hspace{-0.5mm},\\[3pt] F^{\prime}(y_{s})F(y_{s}) & = f_{1}^{\prime}(y_{s})f_{1}(y_{s})\hspace{0.25mm} W_{s,t}^{2} + \big(f_{0}^{\prime}(y_{s})f_{1}(y_{s}) + f_{1}^{\prime}(y_{s})f_{0}(y_{s})\big)\hspace{0.25mm} hW_{s,t}\\[3pt] &\hspace{5mm} + \big(f_{0}^{\prime}(y_{s})f_{1}^{\prime}(y_{s})f_{1}(y_{s}) - f_{1}^{\prime}(y_{s})f_{1}^{\prime}(y_{s})f_{0}(y_{s})\\[3pt] &\hspace{10mm} + f_{0}^{\prime\prime}(y_{s})(f_{1}(y_{s}), f_{1}(y_{s})) - f_{1}^{\prime\prime}(y_{s})(f_{0}(y_{s}), f_{1}(y_{s}))\big)\hspace{0.25mm} hW_{s,t}H_{s,t}\\[3pt] &\hspace{5mm} + f_{0}^{\prime}(y_{s})f_{0}(y_{s})\hspace{0.25mm}h^{2} + O\big(h^{\frac{5}{2}}\big), \end{align*}\vspace{-6mm} \begin{align*} F^{\prime}(y_{s})F^{\prime}(y_{s})F(y_{s}) & = f_{1}^{\prime}(y_{s})f_{1}^{\prime}(y_{s})f_{1}(y_{s})\hspace{0.25mm}W_{s,t}^{3} + f_{0}^{\prime}(y_{s})f_{1}^{\prime}(y_{s})f_{1}(y_{s})\hspace{0.25mm}hW_{s,t}^{2}\\[3pt] &\hspace{5mm} + \big(f_{1}^{\prime}(y_{s})f_{0}^{\prime}(y_{s})f_{1}(y_{s}) + f_{1}^{\prime}(y_{s})f_{1}^{\prime}(y_{s})f_{0}(y_{s})\big)\hspace{0.25mm}hW_{s,t}^{2} + O\big(h^{\frac{5}{2}}\big),\\[8pt] F^{\prime\prime}(y_{s})\big(F(y_{s}), F(y_{s})\big) & = f_{1}^{\prime\prime}(y_{s})(f_{1}(y_{s}), f_{1}(y_{s}))\hspace{0.25mm}W_{s,t}^{3} + 2\hspace{0.25mm}f_{1}^{\prime\prime}(y_{s})(f_{0}(y_{s}), f_{1}(y_{s}))\hspace{0.25mm}hW_{s,t}^{2}\\[2pt] &\hspace{5mm} + f_{0}^{\prime\prime}(y_{s})(f_{1}(y_{s}), f_{1}(y_{s}))\hspace{0.25mm}hW_{s,t}^{2} + O\big(h^{\frac{5}{2}}\big). \end{align*} Therefore, by summing the above formulae for $F$ (and its derivatives) we can derive an expansion of $Y_{1}^{\text{log}}$ in terms of $f_{0}, f_{1}$ and $\big(h, W_{s,t}, H_{s,t}\big)$ that has an $O\big(h^{\frac{5}{2}}\big)$ remainder. By comparing this with the stochastic Taylor expansion (\ref{stochtaylorexp}), the result (\ref{logodestep}) follows.\medbreak\noindent Arguing (\ref{parabolaodestep}) is fairly straightforward and does not require extensive computations. Using the substitution $\wideparen{Y}_{u} = z_{\frac{1}{h}(u-s)}$ for $u\in[s,t]$, the ODE (\ref{paraode1}) can be rewritten as \begin{align}\label{paraode2} d\wideparen{Y}_{u} & = f_{0}\big(\hspace{0.5mm}\wideparen{Y}_{u}\big)\hspace{0.25mm}du + f_{1}\big(\hspace{0.5mm}\wideparen{Y}_{u}\big)\hspace{0.25mm}d\wideparen{W}_{u}\hspace{0.125mm},\\ \wideparen{Y}_{s} & = y_{s}\hspace{0.125mm},\nonumber \end{align} where $\wideparen{W}$ denotes the Brownian parabola defined by $(W_{s,t}, H_{s,t})$ on the interval $[s,t]$.\medbreak\noindent By emulating the derivation of the Stratonovich-Taylor expansion (\ref{stochtaylorexp}), it is possible to Taylor expand (\ref{paraode2}) in the same fashion. The only difference is that Stratonovich integrals with respect to $W$ are replaced with Riemann-Stieltjes integrals against $\wideparen{W}$.\medbreak\noindent In particular, by the change-of-variable formula for ODEs (exercise 3.17) given in \cite{Friz}, we see that the remainder term of such a Taylor expansion will have the below form: \begin{align*} \wideparen{R}\, = \sum_{\substack{i_{1},\hspace{0.25mm}\cdots,\hspace{0.25mm} i_{n}\in\{0,1\} \\ i_{1}\hspace{0.25mm}+\hspace{0.25mm}\cdots\hspace{0.25mm}+\hspace{0.25mm}i_{n}\hspace{0.25mm} =\hspace{0.35mm} 2n-4}}\int_{s\hspace{0.25mm} <\hspace{0.125mm} r_{1}\hspace{0.125mm} <\hspace{0.125mm}\cdots\hspace{0.125mm} <\hspace{0.125mm} r_{n} <\hspace{0.25mm} t}f_{i_{1},\hspace{0.25mm}\cdots,\hspace{0.25mm} i_{n}}\big(\hspace{0.5mm}\wideparen{Y}_{r_{1}}\big) - f_{i_{1},\hspace{0.25mm}\cdots,\hspace{0.25mm} i_{n}}\big(\hspace{0.5mm}\wideparen{Y}_{s}\hspace{0.25mm}\big)\,d\wideparen{W}_{r_{1}}^{i_{1}}\,\cdots\, d\wideparen{W}_{r_{n}}^{i_{n}}\hspace{0.25mm}, \end{align*} where we have identified an additional ``\hspace{0.2mm}zero'' coordinate of $\wideparen{W}$ with time, $\wideparen{W}_{t}^{0} := t$, and for each index $(i_{1}, \cdots i_{n})$, the function $f_{i_{1},\hspace{0.25mm}\cdots,\hspace{0.25mm} i_{n}} : \mathbb{R}^{d}\rightarrow \mathbb{R}^{d}$ consists of finitely many compositions of $f_{0}, f_{1}$ along with their derivatives (and thus is Lipschitz continuous).\medbreak\noindent Therefore each term in the expansion of (\ref{paraode2}) can be estimated in $L^{2}(\mathbb{P})$ by applying the natural Brownian scaling to the corresponding iterated integral of $\wideparen{W}$ with time. As before, the largest differences are the $O(h^{2})$ terms involving third order integrals. Fortunately, iterated integrals of the Brownian parabola can be computed explicitly: \begin{align*} \int_{s}^{t}\int_{s}^{u}\int_{s}^{v}\circ\,dW_{r}\circ\, dW_{v}\, du - \int_{s}^{t}\int_{s}^{u}\int_{s}^{v}\,d\wideparen{W}_{r}\, d\wideparen{W}_{v}\, du & = L_{s,t} - \frac{3}{5}H_{s,t}^{2}\hspace{0.125mm}. \end{align*} The result (\ref{parabolaodestep}) is now a direct consequence of Theorem \ref{logsigrelation} along with the above. \end{proof} \medbreak\noindent Theorem \ref{localerrors} shows that both methods give a one-step approximation error of $O(h^{2})$. This means that the log-ODE and parabola-ODE methods are both locally high order; however there is a significant difference in how these methods propagate local errors. The reason is that the $O(h^{2})$ components of the log-ODE local errors give a martingale, whilst the $O(h^{2})$ part for each parabola-ODE local error has non-zero expectation. Thus the log-ODE method is globally high order whilst the parabola method is not. However, since the parabola-ODE method is straightforward to implement and locally high order, one could expect it to perform well compared to other low order methods. In the numerical example, we shall see that the parabola method has the same order of convergence as the piecewise linear approach but gives significantly smaller errors. To conclude this section, we will present the orders of convergence for both methods.\medbreak \begin{definition}[Strong convergence] A numerical solution $Y$ for (\ref{ogsde}) is said to converge in a strong sense with order $\alpha$ if there exists a constant $C > 0$ such that \begin{align*} \left\|Y_{N} - y_{T}\right\|_{L^{2}\left(\mathbb{P}\right)} \leq Ch^{\alpha}, \end{align*} for all sufficiently small step sizes $h = \frac{T}{N}$. \end{definition}\medbreak \begin{definition}[Weak convergence] A numerical solution $Y$ for (\ref{ogsde}) is said to converge in a weak sense with order $\beta$ if for any polynomial $p$ there exists $C_{p} > 0$ such that \begin{align*} \big|\,\mathbb{E}\hspace{-0.25mm}\left[\,p\hspace{0.25mm}(\hspace{0.25mm}Y_{N})\right] - \mathbb{E}\hspace{-0.25mm}\left[\,p\hspace{0.25mm}(\hspace{0.25mm}y_{T})\right]\big|\leq C_{p}\,h^{\beta}, \end{align*} for all sufficiently small step sizes $h = \frac{T}{N}$. \end{definition}\medbreak \begin{theorem}[Orders of convergence] For a general SDE (\ref{ogsde}), the log-ODE method converges in a strong sense with order 1.5 and a weak sense with order 2.0. The parabola-ODE method converges in both a strong and weak sense with order 1.0. \end{theorem}\medbreak \begin{proof} Note that Theorem \ref{localerrors} establishes the Taylor expansions of both methods. The strong convergence can then be shown as in the proof of Theorem 11.5.1 in \cite{KloePlat}. Moreover, the proof of Theorem 11.5.1 also provides the orders of strong convergence. Similarly weak convergence follows directly from the Taylor expansions (\ref{logodestep}) \& (\ref{parabolaodestep}), and the rate of convergence can be shown as in the proof of Theorem 14.5.2 in \cite{KloePlat}. \end{proof} \section{A numerical example} We shall demonstrate the ideas presented so far using various discretizations of Inhomogeneous Geometric Brownian Motion (IGBM) \begin{align}\label{IGBM} dy_{t} = a(b-y_{t})\,dt+\sigma\hspace{0.125mm} y_{t}\,dW_{t}\hspace{0.125mm}, \end{align} where $a \geq 0$ and $b\in \mathbb{R}$ are the mean reversion parameters and $\sigma \geq 0$ is the volatility. As the vector fields are smooth, the SDE (\ref{IGBM}) can be expressed in Stratonovich form: \begin{align}\label{stratIGBM} dy_{t} = \tilde{a}(\tilde{b} - y_{t})\,dt + \sigma\hspace{0.125mm} y_{t}\circ dW_{t}\hspace{0.125mm}, \end{align} where $\tilde{a} := a + \frac{1}{2}\sigma^{2}$ and $\tilde{b} := \frac{2ab}{2a+\sigma^{2}}$ denote the ``adjusted'' mean reversion parameters.\medbreak\noindent IGBM is an example of a one-factor short rate model and has seen recent attention in the mathematical finance literature as an alternative to popular models \cite{IGBMapproximation, IGBMapplication}. IGBM is also one of the simplest SDEs that has no known method of exact simulation. We will investigate the strong and weak convergence rates of the following methods:\medbreak \begin{enumerate} \item \textbf{Log-ODE method} (see definition \ref{logodedef})\smallskip\\ Since the vector fields of (\ref{stratIGBM}) give constant Lie brackets, this method becomes \begin{align*} Y_{k+1}^{\text{log}} & := Y_{k}^{\text{log}}e^{-\tilde{a}h + \sigma W_{t_{k},t_{k+1}}}\\ &\hspace{4mm} + abh\left(1 - \sigma H_{t_{k},t_{k+1}}\hspace{-0.5mm} + \sigma^{2}\left(\frac{3}{5}H_{t_{k},t_{k+1}}^{2}\hspace{-0.5mm} + \frac{1}{30}h\right)\right) \frac{e^{-\tilde{a}h + \sigma W_{t_{k},t_{k+1}}} - 1}{-\tilde{a}h + \sigma W_{t_{k},t_{k+1}}}\,,\nonumber\\ Y_{0}^{\text{log}} & := y_{0}\hspace{0.125mm}.\nonumber \end{align*} \item \textbf{Parabola-ODE method} (see definition \ref{parabolaodedef})\smallskip\\ As the SDE (\ref{stratIGBM}) is quite analytically tractable, this method is expressible as \begin{align*} Y_{k+1}^{\text{para}} & := e^{-\tilde{a}h + \sigma W_{t_{k},t_{k+1}}}\left(Y_{k}^{\text{para}} + ab\int_{t_{k}}^{t_{k+1}}e^{\tilde{a}\left(s-t_{k}\right) - \sigma \wideparen{W}_{t_{k}, s}}\,ds\right),\\ Y_{0}^{\text{para}} & := y_{0}\hspace{0.125mm}.\nonumber \end{align*} The integral above will be computed by $3$-point Gauss-Legendre quadrature.\medskip \item \textbf{Piecewise linear method} (see \cite{WongZakai} for definition and proof of convergence)\smallskip\\ Just as above, this method can be simplified to give a straightforward formula. \begin{align*} Y_{k+1}^{\text{lin}} & := Y_{k}^{\text{lin}}e^{-\tilde{a}h + \sigma W_{t_{k},t_{k+1}}} + abh\,\frac{e^{-\tilde{a}h + \sigma W_{t_{k},t_{k+1}}} - 1}{-\tilde{a}h + \sigma W_{t_{k},t_{k+1}}}\,,\\ Y_{0}^{\text{lin}} & := y_{0}\hspace{0.125mm}.\nonumber \end{align*} \item \textbf{Milstein method} (see section 6 of \cite{Higman} and section 10.3 of \cite{KloePlat} for overviews)\smallskip\\ For this method, we shall take the positive part to guarantee non-negativity. \begin{align*} Y_{k+1}^{\text{mil}} & := \Big(Y_{k}^{\text{mil}} + \tilde{a}(\tilde{b} - Y_{k}^{\text{mil}})h + \sigma \,Y_{k}^{\text{mil}}\,W_{t_{k}, t_{k+1}}\hspace{-0.5mm} + \frac{1}{2}\sigma^{2}\,Y_{k}^{\text{mil}}\,W_{t_{k}, t_{k+1}}^{2}\Big)^{+},\\ Y_{0}^{\text{mil}} & := y_{0}\hspace{0.125mm}.\nonumber \end{align*} \item \textbf{Euler-Maruyama method} (see sections 4, 5 of \cite{Higman} and section 10.2 of \cite{KloePlat})\smallskip\\ Just as above, we take the positive part of each step to ensure non-negativity. \begin{align*} Y_{k+1}^{\text{eul}} & := \Big(Y_{k}^{\text{eul}} + a(b - Y_{k}^{\text{eul}})h + \sigma \,Y_{k}^{\text{eul}}\,W_{t_{k}, t_{k+1}}\Big)^{+},\\ Y_{0}^{\text{eul}} & := y_{0}\hspace{0.125mm}.\nonumber \end{align*} \end{enumerate} \medbreak\noindent Note that the explicit formula for the log-ODE method comes from the Lie brackets: \begin{align*} [f_{1}, f_{0}](y) & = f_{0}^{\prime}(y)f_{1}(y) - f_{1}^{\prime}(y)f_{0}(y)\\[3pt] & = -\tilde{a}\sigma y - \tilde{a}\sigma(\tilde{b} - y)\\[3pt] & = -ab\sigma\hspace{0.25mm},\\[6pt] [f_{1}, [f_{1}, f_{0}]](y) & = f_{0}^{\prime}(y)f_{1}^{\prime}(y)f_{1}(y) - 2f_{1}^{\prime}(y)f_{0}^{\prime}(y)f_{1}(y) + f_{1}^{\prime}(y)f_{1}^{\prime}(y)f_{0}(y)\\[2pt] &\hspace{8mm} + f_{0}^{\prime\prime}(y)(f_{1}(y), f_{1}(y)) - 2\hspace{0.25mm}f_{1}^{\prime\prime}(y)(f_{0}(y), f_{1}(y)) + f_{1}^{\prime\prime}(y)(f_{1}(y), f_{0}(y))\\[3pt] & = -\tilde{a}\sigma^{2}y + 2\hspace{0.25mm}\tilde{a}\sigma^{2}y + \tilde{a}\sigma^{2}(\tilde{b} - y)\\[3pt] & = ab\sigma^{2}\hspace{0.25mm}, \end{align*} and the formula for the parabola-ODE method was derived using a change of variable.\medbreak\noindent The Euler-Maruyama and Milstein methods are included in the numerical experiment as benchmarks to test how the proposed methods compare to well-known methods. As before, we will be discretizing the SDE over a uniform partition with mesh size $h$.\vspace{-3.5mm} \begin{figure}[h]\label{IGBMsamplepaths} \centering \includegraphics[width=0.95\textwidth]{"IGBM_Sample_Paths".pdf}\vspace{-2.5mm} \caption{Log-ODE sample paths of IGBM where $a=0.1$, $b=0.04$, $\sigma = 0.6$ and $y_{0} = 0.06$.\newline\hspace*{5mm} The above sample paths experience larger fluctuations the further away from zero they are.} \end{figure}\medbreak\vspace{-2.5mm}\noindent Below is the definition of the error estimators used to analyse the numerical methods.\medbreak \begin{definition}[Strong and weak error estimators]\label{errorestimators} For each $N\geq 1$, let $Y_{N}$ denote a numerical solution of (\ref{IGBM}) computed at time $T$ using a fixed step size $h = \frac{T}{N}$. We can define the following estimators for quantifying strong and weak convergence: \begin{align} S_{N} & := \sqrt{\hspace{0.25mm}\mathbb{E}\hspace{-0.25mm}\left[\left(Y_{N} - Y_{T}^{fine}\right)^{2}\hspace{0.5mm}\right]}\hspace{0.25mm},\label{strongestimator}\\ E_{N} & := \left|\,\mathbb{E}\big[\big(Y_{N} - b\hspace{0.25mm}\big)^{+}\hspace{0.25mm}\big] - \mathbb{E}\big[\big(Y_{T}^{fine} - b\hspace{0.25mm}\big)^{+}\hspace{0.25mm}\big]\right|,\label{weakestimator} \end{align} where the above expectations are approximated by standard Monte-Carlo simulation and $Y_{T}^{fine}$ is the numerical solution of (\ref{IGBM}) obtained at time $T$ using the log-ODE method with a ``\hspace{0.5mm}fine'' step size of $\min\left(\frac{h}{10}, \frac{T}{1000}\right)$. The fine step size is chosen so that the $L^{2}(\mathbb{P})$ error between $Y_{T}^{fine}$ and the true solution $y$ is negligible compared to $S_{N}$. Note that $Y_{N}$ and $Y_{T}^{fine}$ are both computed with respect to the same Brownian paths. \end{definition}\medbreak\noindent In this numerical example, we shall use the same parameter values as in \cite{IGBMapproximation}, namely $a = 0.1$, $b=0.04$, $\sigma = 0.6$ and $y_{0} = 0.06$. We will also fix the time horizon at $T = 5$.\medbreak\noindent We will now present our results for the numerical experiment that is described above. (Code for this example can be found at \href{https://github.com/james-m-foster/igbm-simulation}{github.com/james-m-foster/igbm-simulation}) \begin{figure}[h] \centering \hspace*{-1mm}\includegraphics[width=1.025\textwidth]{"Strong_Convergence_Graph".pdf} \caption{$S_{N}$ computed with 100,000 sample paths as a function of step size $h = \frac{T}{N}$.}\label{strongconvergediagram} \end{figure}\smallbreak\noindent From the above graph we see that the log-ODE method is by far the most accurate. This is epitomized by the fact that the numerical error produced by $100$ steps of the log-ODE method is comparable to the error of the parabola method with $1000$ steps. In addition, whilst there are three methods that share the same order of convergence it is evident there are magnitudes of difference between their respective accuracies. For example, the parabola method is seven times more accurate than piecewise linear. As one might expect, the Euler-Maruyama and Milstein schemes both perform poorly.\medbreak\noindent Nevertheless, in order to truly measure the performance of these numerical methods, we should consider the computational costs required for achieving a specified accuracy.\vspace{-1mm} \begin{table}[tbhp] \caption{\hspace{7mm}Estimated simulation times for computing 100,000 sample paths that achieve a\newline \centerline{given accuracy using a single-threaded C++ program on a desktop computer.}} \begin{center} \begin{tabular}{cccccc} \cline{2-6} \multicolumn{1}{c}{} & \multicolumn{1}{|c|}{} & \multicolumn{1}{c|}{} & \multicolumn{1}{c|}{} & \multicolumn{1}{c|}{} & \multicolumn{1}{c|}{} \\ [-1em] \multicolumn{1}{c}{} & \multicolumn{1}{|c|}{Log-ODE} & \multicolumn{1}{c|}{Parabola} & \multicolumn{1}{c|}{\hspace{0.5mm}Linear\hspace*{0.5mm}} & \multicolumn{1}{c|}{Milstein} & \multicolumn{1}{c|}{Euler} \\ \hline \multicolumn{1}{|c}{} & \multicolumn{1}{|c|}{} & \multicolumn{1}{c|}{} & \multicolumn{1}{c|}{} & \multicolumn{1}{c|}{} & \multicolumn{1}{c|}{} \\ [-1em] \multicolumn{1}{|c}{Estimated time to achieve} & \multicolumn{1}{|c|}{0.179} & \multicolumn{1}{c|}{0.405} & \multicolumn{1}{c|}{1.47} & \multicolumn{1}{c|}{15.4} & \multicolumn{1}{c|}{0.437} \\ [0.1em] \multicolumn{1}{|c}{an accuracy of $S_{N} = 10^{-4}$} & \multicolumn{1}{|c|}{(s)} & \multicolumn{1}{c|}{(s)} & \multicolumn{1}{c|}{(s)} & \multicolumn{1}{c|}{(s)} & \multicolumn{1}{c|}{(days)} \\ \multicolumn{1}{|c}{} & \multicolumn{1}{|c|}{} & \multicolumn{1}{c|}{} & \multicolumn{1}{c|}{} & \multicolumn{1}{c|}{} & \multicolumn{1}{c|}{} \\ [-1.1em] \hline \multicolumn{1}{|c}{} & \multicolumn{1}{|c|}{} & \multicolumn{1}{c|}{} & \multicolumn{1}{c|}{} & \multicolumn{1}{c|}{} & \multicolumn{1}{c|}{} \\ [-1em] \multicolumn{1}{|c}{Estimated time to achieve} & \multicolumn{1}{|c|}{0.827} & \multicolumn{1}{c|}{3.90} & \multicolumn{1}{c|}{14.9} & \multicolumn{1}{c|}{157} & \multicolumn{1}{c|}{61.2} \\ [0.1em] \multicolumn{1}{|c}{an accuracy of $S_{N} = 10^{-5}$} & \multicolumn{1}{|c|}{(s)} & \multicolumn{1}{c|}{(s)} & \multicolumn{1}{c|}{(s)} & \multicolumn{1}{c|}{(s)} & \multicolumn{1}{c|}{(days)}\\ [0.1em] \hline \end{tabular} \end{center}\medbreak\medbreak The above times are estimated by the graph shown in Fig \ref{strongconvergediagram} and the following table: \end{table} \begin{table}[tbhp] \caption{\hspace{8mm}Simulation times for computing 100,000 sample paths with $100$ steps per path\newline \centerline{using a single-threaded C++ program on a desktop computer.}}\label{simtimes} \begin{center} \begin{tabular}{cccccc} \cline{2-6} \multicolumn{1}{c}{} & \multicolumn{1}{|c|}{} & \multicolumn{1}{c|}{} & \multicolumn{1}{c|}{} & \multicolumn{1}{c|}{} & \multicolumn{1}{c|}{} \\ [-1em] \multicolumn{1}{c}{} & \multicolumn{1}{|c|}{Log-ODE} & \multicolumn{1}{c|}{Parabola} & \multicolumn{1}{c|}{\hspace{0.5mm}Linear\hspace*{0.5mm}} & \multicolumn{1}{c|}{Milstein} & \multicolumn{1}{c|}{\hspace{0.5mm}Euler\hspace*{0.5mm}} \\ \hline \multicolumn{1}{|c}{} & \multicolumn{1}{|c|}{} & \multicolumn{1}{c|}{} & \multicolumn{1}{c|}{} & \multicolumn{1}{c|}{} & \multicolumn{1}{c|}{} \\ [-1em] \multicolumn{1}{|c}{Computation time (s) } & \multicolumn{1}{|c|}{2.44} & \multicolumn{1}{c|}{2.95} & \multicolumn{1}{c|}{1.48} & \multicolumn{1}{c|}{1.18} & \multicolumn{1}{c|}{1.17} \\ [-1.1em] \multicolumn{1}{|c}{} & \multicolumn{1}{|c|}{} & \multicolumn{1}{c|}{} & \multicolumn{1}{c|}{} & \multicolumn{1}{c|}{} & \multicolumn{1}{c|}{} \\ \hline \end{tabular} \end{center}\medbreak\medbreak Finally, we will investigate the rates of weak convergence for these numerical methods. \end{table} \begin{figure}[h]\label{weakconvergediagram} \centering \hspace*{-0.5mm}\includegraphics[width=1\textwidth]{"Weak_Convergence_Graph".pdf}\vspace{-1.5mm} \caption{$E_{N}$ computed with 500,000 sample paths as a function of step size $h = \frac{T}{N}$.} \end{figure}\newpage\noindent The above graph demonstrates that the log-ODE method is especially well-suited for weak approximation as it achieves a second order convergence rate in this example. Surprisingly, the middle three methods exhibit almost identical convergence profiles. As before, we can estimate the computational time needed to achieve given accuracies.\vspace{-2mm} \begin{table}[tbhp] \caption{\hspace{7mm}Estimated simulation times for computing 100,000 sample paths that achieve a\newline \centerline{given accuracy using a single-threaded C++ program on a desktop computer.}} \begin{center} \begin{tabular}{cccccc} \cline{2-6} \multicolumn{1}{c}{} & \multicolumn{1}{|c|}{} & \multicolumn{1}{c|}{} & \multicolumn{1}{c|}{} & \multicolumn{1}{c|}{} & \multicolumn{1}{c|}{} \\ [-1em] \multicolumn{1}{c}{} & \multicolumn{1}{|c|}{Log-ODE} & \multicolumn{1}{c|}{Parabola} & \multicolumn{1}{c|}{\hspace{0.5mm}Linear\hspace*{0.5mm}} & \multicolumn{1}{c|}{Milstein} & \multicolumn{1}{c|}{Euler} \\ \hline \multicolumn{1}{|c}{} & \multicolumn{1}{|c|}{} & \multicolumn{1}{c|}{} & \multicolumn{1}{c|}{} & \multicolumn{1}{c|}{} & \multicolumn{1}{c|}{} \\ [-1em] \multicolumn{1}{|c}{Estimated time to achieve} & \multicolumn{1}{|c|}{$<$ 0.240} & \multicolumn{1}{c|}{1.69} & \multicolumn{1}{c|}{2.15} & \multicolumn{1}{c|}{2.78} & \multicolumn{1}{c|}{25.5} \\ [0.1em] \multicolumn{1}{|c}{an accuracy of $E_{N} = 10^{-5}$} & \multicolumn{1}{|c|}{(s)} & \multicolumn{1}{c|}{(s)} & \multicolumn{1}{c|}{(s)} & \multicolumn{1}{c|}{(s)} & \multicolumn{1}{c|}{(s)} \\ \multicolumn{1}{|c}{} & \multicolumn{1}{|c|}{} & \multicolumn{1}{c|}{} & \multicolumn{1}{c|}{} & \multicolumn{1}{c|}{} & \multicolumn{1}{c|}{} \\ [-1.1em] \hline \multicolumn{1}{|c}{} & \multicolumn{1}{|c|}{} & \multicolumn{1}{c|}{} & \multicolumn{1}{c|}{} & \multicolumn{1}{c|}{} & \multicolumn{1}{c|}{} \\ [-1em] \multicolumn{1}{|c}{Estimated time to achieve} & \multicolumn{1}{|c|}{0.240} & \multicolumn{1}{c|}{16.9} & \multicolumn{1}{c|}{21.6} & \multicolumn{1}{c|}{24.1} & \multicolumn{1}{c|}{252} \\ [0.1em] \multicolumn{1}{|c}{an accuracy of $E_{N} = 10^{-6}$} & \multicolumn{1}{|c|}{(s)} & \multicolumn{1}{c|}{(s)} & \multicolumn{1}{c|}{(s)} & \multicolumn{1}{c|}{(s)} & \multicolumn{1}{c|}{(s)}\\ [0.1em] \hline \end{tabular} \end{center} \end{table}\smallbreak\vspace{-2.5mm}\noindent We expect the log-ODE and parabola methods to have about twice the computational cost as the other methods because each step requires generating two random variables. Table \ref{simtimes} confirms this and thus sampling may be a bottleneck for these methods. So overall, the numerical evidence supports our claim that the high order log-ODE method is currently a state-of-the-art method for the pathwise discretization of IGBM. \section{Conclusion} There are primarily three new results established in this paper:\medbreak \begin{itemize} \item \textbf{An efficient strong polynomial approximation of Brownian motion}\smallbreak The main result allows one to construct a ``smoother'' Brownian motion as a finite sum of $(\text{-}1,\text{-}1)$-Jacobi polynomials with independent Gaussian weights. Moreover, it was shown that the approximation is optimal in a weighted $L^{2}(\mathbb{P})$ sense and the surrounding noise is an independent centered Gaussian process.\medbreak \item \textbf{Unbiased approximation of third order Brownian iterated integrals}\smallbreak Iterated integrals of Brownian motion and time are important objects in the study of SDEs as they appear naturally within stochastic Taylor expansions. We have derived the $L^{2}(\mathbb{P})$-optimal estimator for a class of such integrals that is measurable with respect to the path's increment and space-time L\'{e}vy area.\medbreak \item \textbf{Simulation of Inhomogeneous Geometric Brownian Motion (IGBM)}\smallbreak IGBM is a mean-reverting short rate model used in mathematical finance and also one of the simplest SDEs that has no known method of exact simulation.\\ By incorporating the new iterated integral estimator into the log-ODE method we have developed a high order state-of-the-art numerical method for IGBM.\bigbreak \end{itemize} Furthermore, the results of this paper naturally lead to the following open questions:\medbreak \begin{itemize} \item Which weight functions give ``\hspace{0.125mm}explicit eigenfunctions'' for Brownian motion? (For example, we could try $w(x) = x$ or $w(x) = \frac{1}{x}$ with $K_{W}(s,t) = \min(s,t)$)\medbreak \item Is it possible to generalize the main theorem to fractional Brownian motion?\medbreak \item What are the most efficient Runge-Kutta methods for general one-dimensional SDEs that correctly use the new estimator for third order iterated integrals?\medbreak \item Is this polynomial expansion optimal for approximating L\'{e}vy area? (see \cite{Dickinson})\medbreak \item Which conditional moments can be computed for a given stochastic integral?\medbreak \item How might we construct a piecewise linear path $\gamma$ with the below properties? \begin{align*} &1.\,\,\,\, \gamma_{s} = W_{s}\hspace{0.25mm},\hspace{1.75mm} \gamma_{t} = W_{t}\hspace{0.25mm}.\\[5pt] &2.\,\,\, \int_{s}^{t}\gamma_{s,u}\,du = \int_{s}^{t}W_{s,u}\,du\hspace{0.25mm}.\\[2pt] &3.\,\,\, \int_{s}^{t}\gamma_{s,u}^{2}\,du = \mathbb{E}\hspace{-0.25mm}\left[\hspace{0.25mm}\int_{s}^{t}W_{s,u}^{2}\,du\,\Big|\,\cdots\right]\hspace{-0.5mm}. \end{align*} \item Would this method of construction lead to effective cubature paths? (see \cite{Cubature})\medbreak Given such a path, we can approximate (\ref{ogsde}) with a ``\hspace{0.125mm}piecewise linear'' ODE. \begin{align}\label{piecelinode} \frac{dY}{du} = f_{0}(\hspace{0.125mm}Y) + f_{1}(\hspace{0.125mm}Y)\hspace{0.5mm}\frac{d\gamma}{du}\hspace{0.5mm}. \end{align} (Along each piece of $\gamma$, we would discretize (\ref{piecelinode}) using an appropriate solver)\medbreak \item How effective is the above piecewise linear ODE method for simulating SDEs?\medbreak \item Can we extend the approximations given in this paper to the SPDE setting? \end{itemize} \newpage\bigbreak \bibliographystyle{amsplain}
{ "timestamp": "2020-05-21T02:12:18", "yymm": "1904", "arxiv_id": "1904.06998", "language": "en", "url": "https://arxiv.org/abs/1904.06998", "abstract": "In this paper, we will present a strong (or pathwise) approximation of standard Brownian motion by a class of orthogonal polynomials. The coefficients that are obtained from the expansion of Brownian motion in this polynomial basis are independent Gaussian random variables. Therefore it is practical (requires $N$ independent Gaussian coefficients) to generate an approximate sample path of Brownian motion that respects integration of polynomials with degree less than $N$. Moreover, since these orthogonal polynomials appear naturally as eigenfunctions of an integral operator defined by the Brownian bridge covariance function, the proposed approximation is optimal in a certain weighted $L^{2}(\\mathbb{P})$ sense. In addition, discretizing Brownian paths as piecewise parabolas gives a locally higher order numerical method for stochastic differential equations (SDEs) when compared to the standard piecewise linear approach. We shall demonstrate these ideas by simulating Inhomogeneous Geometric Brownian Motion (IGBM). This numerical example will also illustrate the deficiencies of the piecewise parabola approximation when compared to a new version of the asymptotically efficient log-ODE (or Castell-Gaines) method.", "subjects": "Numerical Analysis (math.NA); Probability (math.PR)", "title": "An optimal polynomial approximation of Brownian motion", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.984336353585837, "lm_q2_score": 0.8311430499496096, "lm_q1q2_score": 0.81812431909561 }
https://arxiv.org/abs/1506.07131
Lie Transformation Groups - An Introduction to Symmetry Group Analysis of Differential Equations
These are lecture notes of a course on symmetry group analysis of differential equations, based mainly on P. J. Olver's book 'Applications of Lie Groups to Differential Equations'. The course starts out with an introduction to the theory of local transformation groups, based on Sussman's theory on the integrability of distributions of non-constant rank. The exposition is self-contained, pre-supposing only basic knowledge in differential geometry and Lie groups.
\chapter*{Preface} \addcontentsline{toc}{part}{Preface} These lecture notes cover the material of a four hour course on Lie transformation groups. The goal of this course is to provide the foundations of symmetry group analysis of differential equations, a vast and very active research area with connections to Lie group theory, differential geometry, differential equations, calculus of variations, integrable systems, and mathematical physics, among others. The standard reference in this field is P.\ J.\ Olver's seminal work \cite{O}, and to a large part this set of notes will closely follow his exposition. We do, however, take a slightly different approach to the foundations of the field and try to give a reasonably complete exposition of the underlying theory of local transformation groups, which in turn is based on H.\ Sussmann's study \cite{Sussmann} of the integrability of distributions of non-constant rank. For connecting this material with Olver's approach, we have found V.\ Schmidt's doctoral thesis \cite{S} very helpful. I have tried to make these notes self-contained, pre-supposing only that the reader be familiar with the material covered in basic courses on Differential Geometry and Lie Groups. More precisely, the presentation is a direct continuation of my courses \cite{DG1} and \cite{KLie} (which in turn is mostly based on \cite{BC}), as well as the chapter on submanifolds in \cite{DG2}. In particular, these notes supply proofs for many results that are left to the reader in \cite{O}. Clearly what we can provide here is only a first step into the field, but it should allow the reader to delve into the subject, e.g., by further studying \cite{O}. I am greatly indebted to Roman Popovych for carefully reading the entire manuscript and for suggesting numerous improvements. \vskip1cm \hspace*{\fill} Michael Kunzinger, summer term 2015 \ \newpage \tableofcontents \newpage \pagenumbering{arabic} \setcounter{page}{1} \pagestyle{plain} \chapter{Lie transformation groups} \label{ltg_chapter} \section{Basic concepts} Throughout this course, by a manifold we mean a smooth ($C^\infty$) manifold. However, we do not a priori assume further properties of the natural manifold topology like Hausdorff or second countable. In this chapter we will follow \cite{BC}. We begin by recalling \cite[Def.\ 16.1]{KLie}: \begin{thr} {\bf Definition. }\label{transgroup} A {\em transformation}\index{transformation} of a manifold $M$ is a diffeomorphism from $M$ onto $M$. A group $G$ acts on $M$ as a transformation group\index{transformation group} (on the left) if there exists a map $\Phi\colon G\times M \to M$ satisfying: \begin{itemize} \item[(i)] For every $g\in G$ the map $\Phi_g\colon m\mapsto \Phi(g,m)$ is a transformation of $M$. \item[(ii)] For all $g$, $h\in G$, $\Phi_g\circ \Phi_h = \Phi_{gh}$. \end{itemize} In particular, $\Phi_e= \mathrm{id}_M$. $G$ acts {\em effectively} on $M$ if $\Phi_g(m) = m$ for all $m\in M$ implies $g=e$. $G$ acts {\em freely} on $M$ if it has no fixed points, i.e., if $\Phi_g(m) = m$ for some $m\in M$ implies that $g=e$. \et Our main object of study is introduced next. \begin{thr} {\bf Definition. }\label{ltgdef} A Lie group $G$ acts as a {\em Lie transformation group}\index{Lie transformation group} on a manifold $M$ if there exists a smooth surjection $\Phi\colon G\times M \to M$ such that for all $g$, $h\in G$, $\Phi_g\circ \Phi_h = \Phi_{gh}$. \et We first note that any Lie transformation group is a transformation group in the sense of \ref{transgroup}. To see this, note first that $\Phi$ being surjective implies that $\Phi_e=\mathrm{id}_M$. In fact, given $m\in M$ there exist $m'\in M$, $h\in G$ with $\Phi_h(m')=m$. Therefore, $\Phi_e(m) = \Phi_e\circ \Phi_h(m') = \Phi_h(m')=m$. Consequently, $\Phi_g^{-1} = \Phi_{g^{-1}}$ for all $g\in G$, so each $\Phi_g$ is a transformation of $M$, and we have verified all conditions from Def.\ \ref{transgroup}. We will often briefly write $gm$ instead of $\Phi_g(m)$. \begin{thr}{\bf Example. }\rm\label{complvf} Let $X$ be a complete vector field on a Hausdorff manifold $M$. Then the flow of $X$ defines a smooth map $\Psi:=\mathrm{Fl}^X:\mathbb R \times M \to M$ with $\Psi(s,\Psi(t,m))$ $=$ $\Psi(s+t,m)$ for all $s$, $t\in \mathbb R$ and all $m\in M$. $\Psi$ therefore induces a Lie transformation group of the Lie group $(\mathbb R,+)$ on $M$. Each $\Psi_t := m\mapsto \Psi(t,m)$ is a transformation on $M$ and $t\mapsto \Psi_t$ is a homomorphism of $\mathbb R$ into the group of transformations on $M$. \et \blem\label{subgr} Let $G$ be a Lie group acting as a Lie transformation group on $M$ and let $H$ be a Lie subgroup of $G$. Then also $H$ acts as a Lie transformation group on $M$. \et \pr Let $j: H \hookrightarrow G$ be the natural injection and let $\Phi\colon G\times M \to M$ be as in \ref{ltgdef}. Then $\Phi_H: H\times M\to M$, $\Phi_H:=\Phi\circ(j\times \mathrm{id}_M)$ does the trick. \ep If $G$ acts as a transformation group on $M$ via $\Phi$ then a subset $A$ of $M$ is called {\em invariant}\index{invariant} under $G$ if $\Phi(G\times A) \subseteq A$. \bp \label{regsub} If a regular submanifold $M'$ of $M$ is invariant under the action of a Lie transformation group $G$ on $M$ then $G$ acts naturally on $M'$ as a Lie transformation group. \et \pr Let $j:G\times M' \hookrightarrow G\times M$ be the natural injection. Then since $M'$ is regular, the map $\Phi':G\times M' \to M'$ induced by $\Phi\circ j$ is smooth (see \cite[1.1.14]{DG2}), hence it defines an action of $G$ on $M'$. \hspace*{\fill}$\Box$ \begin{thr}{\bf Example. }\rm\label{onex} $GL(n,\mathbb R)$ acts on $\mathbb R^n$ as a Lie transformation group via $(A,x)\mapsto Ax$. By \ref{subgr}, so does $O(n,\mathbb R)$. Since $S^{n-1}$ is a regular submanifold of $\mathbb R^n$ invariant under $O(n,\mathbb R)$, \ref{regsub} implies that $O(n,\mathbb R)$ acts on $S^{n-1}$ as a Lie transformation group. \et For the notion of quotient manifold, cf.\ \cite[Sec.\ 15]{KLie}. We say that a transformation group $G$ {\em preserves}\index{preserving an equivalence relation} an equivalence relation $\rho$ on $M$ if $(m_1,m_2)\in \rho$ and $g\in G$ imply $(gm_1,gm_2)\in \rho$. \bp If $M/\rho$ is a quotient manifold of $M$ and if the equivalence relation $\rho$ is preserved by a Lie transformation group $G$ on $M$ then $G$ acts naturally on $M/\rho$ as a Lie transformation group. \et \pr Denote by $\pi: M \to M/\rho$ the natural surjection. Then $\mathrm{id}\times \pi: G\times M \to G\times (M/\rho)$ is a submersion, so $G\times (M/\rho)$ is a quotient manifold of $G\times M$. Moreover, the smooth map $\pi\circ \Phi\colon G\times M \to M/\rho$ is an invariant of the corresponding equivalence relation on $G\times M$ since $(g,m_1)\sim (g,m_2)$ if and only if $(gm_1,gm_2)\in \rho$, which implies $\pi\circ \Phi(g,m_1) = \pi(gm_1) = \pi(gm_2) = \pi\circ \Phi(g,m_2)$. By \cite[15.13]{KLie} it therefore projects to a smooth map $$ \Phi_\rho: G\times (M/\rho) \to M/\rho, \qquad (g,\pi(m)) \mapsto \pi(gm) $$ which defines the required action of $G$ on $M/\rho$ since \begin{equation*} \begin{split} \Phi_\rho(g_1,\Phi_\rho(g_2,\pi(m))) &= \Phi_\rho(g_1,\pi\circ\Phi(g_2,m)) = \pi\circ \Phi(g_1,\Phi(g_2,m)) \\ &= \pi\circ \Phi(g_1g_2,m) = \Phi_\rho(g_1g_2,\pi(m)). \end{split} \end{equation*} \hspace*{\fill}$\Box$ \begin{thr}{\bf Example. }\rm \label{gmodh} Any Lie group $G$ acts on itself as a Lie transformation group via multiplication $\mu\colon G\times G\to G$. A subgroup $H$ of $G$ determines an equivalence relation $\rho$ on $G$ by $$ (g_1,g_2)\in \rho :\Leftrightarrow \exists h\in H: g_1 = g_2h $$ and this relation is preserved by the transformation group. If the set of orbits $G/H$ (i.e., the set of left cosets) is a quotient manifold of $G$ (cf.\ \cite[Sec.\ 15, 20]{KLie}) then $G$ acts naturally on $G/H$ as a Lie transformation group with action $\Phi$ given by $(g,aH) \mapsto (ga)H$. A sufficient condition for $G/H$ to be a quotient manifold of $G$ is that $H$ is closed, not open, and connected (see \cite[20.5]{KLie}). \et \bp If $H$ is a closed subgroup of a Lie group $G$ such that $G/H$ is a quotient manifold of $G$, then $G/H$ is a Hausdorff manifold. \et \pr Recall from \cite[15.9]{KLie} that the natural manifold topology of $G/H$ is the quotient topology with respect to the natural projection $\pi: G \to G/H$. Suppose first that $gH\in G/H$ is distinct from $H$. Then $gH$ is a closed subset of $G$ that does not contain $e$. Hence there exists a neighborhood $U$ of $e$ in $G$ which is disjoint from $gH$. Pick a neighborhood $V$ of $e$ in $G$ such that $V^{-1}V\subseteq U$. We show that $\pi(Vg)$ and $\pi(V)$ are disjoint neighborhoods of $gH$ and $H$, respectively. Indeed, if $\pi(Vg)\cap \pi(V)\not=\emptyset$ then there exist $a$, $b\in V$ such that $agH=bH$, so $a^{-1}b\in gH$. But $a^{-1}b\in V^{-1}V\subseteq U$, so we obtain a contradiction. In general, if $gH$ and $g'H$ are distinct elements of $G/H$ then so are $H$ and $(g^{-1}g')H$ and by what we have shown above they possess disjoint neighborhoods $W$ and $W'$. Denoting by $\phi$ the action of $G$ on $G/H$ from \ref{gmodh}, it follows that $\Phi_g(W)$ and $\Phi_g(W')$ are disjoint neighborhoods of $gH$ and $g'H$. \ep Recall from \cite[Sec.\ 4, 6]{KLie} that a vector field $X\in {\mathfrak X}(G)$ is called left- (resp.\ right)-invariant if $L_g^*X=X$ (resp.\ $R_g^*X=X$) for each $g\in G$. We denote by ${\mathfrak X}_L(G)$ (resp.\ ${\mathfrak X}_R(G)$) the space of left- (resp.\ right)-invariant vector fields on $G$. Also, we identify ${\mathfrak X}_L(G)$ with the Lie algebra ${\mathfrak g}$ via $T_eG \ni v \mapsto L^v\in {\mathfrak X}_L(G)$, where $L^v(g):=T_eL_g(v)$. The inverse of this map is $X\mapsto X(e)$. Similarly, ${\mathfrak g}$ is isomorphic to ${\mathfrak X}_R(G)$ via $T_eG \ni v \mapsto R^v\in {\mathfrak X}_R(G)$ with $R^v(g):=T_eR_g(v)$. We usually equip ${\mathfrak g}$ with the Lie algebra structure induced by $[v,w]:=[L^v,L^w](e)$. Alternatively, we may set $[v,w]_R:=[R^v,R^w](e)$. The resulting Lie algebra will be denoted by ${\mathfrak g}_R$. We have: \bp\label{leftright} Let $G$ be a Lie group. Then \begin{itemize} \item[(i)] The map $F: L^v \mapsto - R^v$ is a Lie algebra isomorphism from ${\mathfrak X}_L(G)$ onto ${\mathfrak X}_R(G)$. \item[(ii)] The map $f\colon v\mapsto -v$ is a Lie algebra isomorphism from ${\mathfrak g}$ onto ${\mathfrak g}_R$. \end{itemize} \et \pr (i) By what was said above, $F$ is a linear isomorphism, since both $L^v \mapsto v\in {\mathfrak g}$ and $v\mapsto R^v$ are. Furthermore, by \cite[6.1 (iii)]{KLie} we have: \begin{equation*} [F(L^v),F(L^w)] = [R^v,R^w] = R^{-[v,w]} = -R^{[v,w]} = F(L^{[v,w]}) = F([L^v,L^w]), \end{equation*} so $F$ is a Lie algebra isomorphism.\medskip (ii) We have $[v,w]_R=[R^v,R^w](e)=R^{-[v,w]}(e)=-[v,w]$. Hence $$ [f(v),f(w)]_R = [-v,-w]_R = [v,w]_R = - [v,w] = f([v,w]). $$ \ep \begin{thr} {\bf Remark. }\rm By \cite[6.1]{KLie} The map $F$ from \ref{leftright} can also be written as $X\mapsto \nu^*X$. In fact, if $X\in {\mathfrak X}_L(G)$ then since $\nu\circ R_g = L_{g^{-1}}\circ \nu$ we have $$ R_g^*(\nu^*X) = (\nu\circ R_g)^*X = (L_{g^{-1}}\circ \nu)^*X = \nu^*(L_{g^{-1}}^*X) = \nu^*X, $$ so $\nu^*X\in {\mathfrak X}_R(G)$, and by symmetry $\nu^*:{\mathfrak X}_L(G)\to {\mathfrak X}_R(G)$ is an isomorphism. Moreover, $\nu^*(L^v)(e) = (T_e\nu)^{-1} \circ L^v\circ \nu(e) = - v$, so $\nu^*(L^v)=R^{-v}=F(L^v)$. \et Let now $\Phi\colon G\times M \to M$ be a Lie transformation group and let $m\in M$. Then the map \begin{equation}\label{phigdef} \begin{split} \Phi_m: G &\to M\\ \Phi_m(g) &:= \Phi(g,m) \end{split} \end{equation} is smooth. Given any $v\in T_eG = {\mathfrak g}$, let \begin{equation}\label{phidef} \begin{split} \Phi(v): M &\to TM\\ \Phi(v)(m) &:= T_e\Phi_m(v) \end{split} \end{equation} Then $\Phi(v)(m) = T\Phi(v,0_m)$, so $\Phi(v)$ is smooth. Also, with $\pi:TM\to M$ and $\tilde \pi:T(G\times M)\to G\times M$ the canonical projections, $\pi\circ \Phi(v)(m) = \Phi(\tilde\pi(v,0_m)) = \Phi(e,m) = m$, so $\Phi(v)$ is a smooth section of $TM$, i.e., $\Phi(v)\in {\mathfrak X}(M)$. We set \begin{equation}\label{rightvf} {\mathscr R}(G,M) := \{\Phi(v) \mid v\in {\mathfrak g}\}. \end{equation} ${\mathscr R}(G,M)$ is called the {\em Killing algebra}\index{Killing algebra} of $\Phi$. Since $v\mapsto \Phi(v)$ is linear, it is a finite-dimensional vector space. In fact, we even have: \bp\label{psihomo} The map $\Phi\colon v\mapsto \Phi(v)$ is a Lie algebra homomorphism from ${\mathfrak g}_R$ onto ${\mathscr R}(G,M)$. \et \pr Given any $v\in {\mathfrak g}$, we first show that for any $m\in M$, the right-invariant vector field $Y$ on $G$ with $Y(e)=v$ is $\Phi_m$-related to $\Phi(v)$. In fact, noting that $\Phi_m\circ R_g(h) = \Phi_m(hg) = \Phi(h,\Phi_m(g)) = \Phi_{m'}(h)$ with $m':=\Phi_m(g)$, we obtain: $$ T_g\Phi_m(Y(g)) = T_g\Phi_m(T_eR_g(v)) = T_e(\Phi_m\circ R_g)(v) = T_e\Phi_{m'}(v) = \Phi(v)(\Phi_m(g)). $$ Thus for any $v\in {\mathfrak g}$, \begin{equation}\label{olver1.46.2} T_g\Phi_m(Y(g)) = \Phi(v)(\Phi_m(g)). \end{equation} From this, by \cite[4.4]{KLie} we conclude that for $Z=R^w$ another right-invariant vector field we have $$ [\Phi(v),\Phi(w)]\circ \Phi_m = T\Phi_m\circ[Y,Z]. $$ Inserting $e$, this gives $$ [\Phi(v),\Phi(w)]_m = T_e\Phi_m([Y,Z]_e) = T_e\Phi_m([v,w]_R) $$ (where we used that $[R^v,R^w]=R^{[v,w]_R}$, cf.\ \ref{leftright}). Thus, finally, $[\Phi(v),\Phi(w)] = \Phi([v,w]_R)$, as claimed. \ep In the following result, we will start denoting the time variable in the flow map of a vector field with $\varepsilon$, for the sake of compatibility with the case of symmetry groups of differential equations later on. \bp\label{flowaction} If $M$ is $T_2$, then every vector field $\Phi(v)\in {\mathscr R}(G,M)$ is complete, with flow $$ \mathrm{Fl}^{\Phi(v)}_\varepsilon(m) = \Phi(\exp(\varepsilon v),m) = \exp(\varepsilon v)\cdot m. $$ In particular, \begin{equation}\label{olver1} \Phi(v)(m) = \left.\frac{d}{d \eps}\right|_0 \Phi(\exp(\varepsilon v),m). \end{equation} \et \pr In the proof of \ref{psihomo} we have seen that $R^v$ is $\Phi_m$-related to $\Phi(v)$. Thus by \cite[7.2]{KLie} it follows that $\Phi_m\circ \mathrm{Fl}^{R^v}_\varepsilon = \mathrm{Fl}^{\Phi(v)}_\varepsilon\circ \Phi_m$. Moreover, by \cite[8.2]{KLie}, $\mathrm{Fl}^{R^v}_\varepsilon(g) = \exp(\varepsilon v)\cdot g$. Therefore, $$ \mathrm{Fl}^{\Phi(v)}_\varepsilon(m) = \mathrm{Fl}^{\Phi(v)}_\varepsilon\circ \Phi_m(e) = \Phi_m(\exp(\varepsilon v)\cdot e) = \Phi(\exp(\varepsilon v),m). $$ Completeness of $\Phi(v)$ is immediate from this description. \ep \begin{thr}{\bf Example. }\rm\label{rflowex} Let $\Phi\colon \mathbb R\times M \to M$ be a Lie transformation group with Lie group $(\mathbb R,+)$ and $T_2$-manifold $M$. Then ${\mathscr R}(\mathbb R,M)$ is generated by $\Phi(\left.\frac{d}{d \eps}\right|_0)$. Moreover, the flow of $\Phi(\left.\frac{d}{d \eps}\right|_0)$ is given by $\Phi$ itself: $\mathrm{Fl}^{\Phi(\left.\frac{d}{d \eps}\right|_0)}_\varepsilon(m) = \Phi(\varepsilon,m)$: this is immediate from \ref{flowaction} since $\exp(\varepsilon\left.\frac{d}{d \eps}\right|_0) = \varepsilon$. \et Recall from \cite[Sec.\ 16]{KLie} that, given a Lie transformation group $\Phi\colon G\times M \to M$, we set $$ K:=\{k\in G \mid \Phi_k = \mathrm{id}_M\}. $$ Then $K$ is a normal subgroup of $G$ and \cite[16.3]{KLie} demonstrates that the quotient group $G/K$ acts effectively on $M$. Moreover, $K$ is closed in $G$ since we can write $K=\bigcap_{m\in M} \Phi_m^{-1}(m)$. Hence by \cite[21.6]{KLie}, if $K$ is not open then the quotient manifold $G/K$ is a Lie group. The exceptional case here is not very interesting though: \begin{thr} {\bf Remark. }\rm If $K$ is open, then ${\mathscr R}(G,M)=0$. Indeed, if $K$ is open (and closed) then since $e\in K$ we must have $G_e\subseteq K$. Then because for all $v\in {\mathfrak g}$, $\exp(\varepsilon v)\in G_e$ it follows that $\Phi_m(\exp(\varepsilon v)) = m$ for all $m\in M$, and differentiating with respect to $\varepsilon$ at $\varepsilon=0$ implies that $\Phi(v)(m)=T_e\Phi_m(v) = 0$ for all $m$. Thus $\Phi(v)=0$ and thereby ${\mathscr R}(G,M)=0$. \et \bp If $K$ is not open then $G/K$ acts effectively on $M$ as a Lie transformation group and ${\mathscr R} (G/K,M)= {\mathscr R}(G,M)$. \et \pr Let \begin{equation*} \begin{split} \Phi': (G/K) \times M &\to M\\ (gK,m) &\mapsto \Phi(g,m) \end{split} \end{equation*} To see that $\Phi'$ is well-defined and smooth let $\pi: G\to G/K$ be the natural projection. Then $\pi\times \mathrm{id}_M: G\times M \to (G/K)\times M$ is a submersion (cf.\ \cite[20.5]{KLie}) and $\Phi$ is an invariant of the corresponding equivalence relation on $G\times M$: if $(g_1,m_1)\sim (g_2,m_2)$ then $g_1K=g_2K$ and $m_1=m_2$, so $\Phi(g_1,m_1) = \Phi'(g_1K,m_1) = \Phi'(g_2K,m_2) = \Phi(g_2,m_2)$. Thus by \cite[15.13]{KLie}, the map $\Phi'$, being the projection of $\Phi$, is itself smooth. That $\Phi'$ gives an effective action has been shown in \cite[16.3]{KLie}. Finally, for any $m\in M$ we have $\Phi_m = \Phi'_m\circ \pi$, so for any $v\in {\mathfrak g}$ we get $T_e\Phi_m(v) = T_e\Phi'_m \circ T_e\pi(v)$, i.e., $\Phi(v)=\Phi'(T_e\pi(v))$. Since $T_e\pi$ is a surjection, this gives ${\mathscr R} (G/K,M)={\mathscr R}(G,M)$. \ep \bp\label{phiiso} Let $\Phi$ be a transformation group that acts effectively on a $T_2$-manifold $M$. Then the map $$ \Phi\colon {\mathfrak g}_R \to {\mathscr R}(G,M) $$ defined by \eqref{phidef} is a Lie algebra isomorphism. \et \pr Using \ref{psihomo} it only remains to show that $\Phi$ is injective. Thus let $\Phi(v)=0$. Then by \ref{flowaction} it follows that $\exp(\varepsilon v)\cdot m = m$ for all $m\in M$ and all $\varepsilon\in \mathbb R$. Since $G$ acts effectively we conclude that $\exp(\varepsilon v)= e$ for all $\varepsilon$, giving $v=0$. \ep \begin{thr}{\bf Example. }\rm Let $G = GL(n,\mathbb R)$ and $M=\mathbb R^n$ with $\Phi\colon (A,x)\mapsto A\cdot x$. To determine the map \eqref{phidef} we let $A^{ij}$ be the standard coordinates on $GL(n,\mathbb R)$ and $x^k$ those on $\mathbb R^n$. Then \begin{equation*} \begin{split} T_e\Phi_x\Big(\frac{\partial}{\partial A^{ij}}\Big) &= \sum_k \Big(\frac{\partial}{\partial A^{ij}}(x^k\circ \Phi_x)\Big) \Big(\left. \frac{\partial}{\partial x^k}\right|_x\Big) \\ & = \sum_{k,r} \Big(\pd{}{A^{ij}}(A^{kr}x_r)\Big)\pdl{}{x^k}{x} = x_j \pdl{}{x^i}{x}, \end{split} \end{equation*} so $\Phi = [B^{ij}] \mapsto \sum_{i,j} B^{ij}x^j \pd{}{x^i}$. Since $G$ act effectively, $\Phi$ is an isomorphism. \et \begin{thr} {\bf Remark. }\rm There is a strong converse of \ref{phiiso}, due to R.\ Palais (see \cite[p.\ 95]{Palais}): {\em If $\mathscr A$ is a finite-dimensional Lie algebra of complete vector fields on a $T_2$-manifold $M$, then there exists a connected Lie group $G$ which acts effectively on $M$ as a Lie transformation group and such that $\mathscr{A} = {\mathscr R}(G,M)$.} \et \section{Orbits under a Lie transformation group} From \cite[Sec.\ 16]{KLie} we know that any transformation group $\Phi$ acting on a manifold $M$ induces an equivalence relation on $M$: $m\sim m'$ iff there exists some $g\in G$ with $m'=gm$. The equivalence class of any $x\in M$ is called the {\em orbit}\index{orbit} of $m$ under $G$. It is the range of the map $\Phi_m = g\mapsto \Phi(g,m)$, $G\to M$. \begin{thr} {\bf Definition. } Let $\Phi\colon G\times M \to M$ be a transformation group on a manifold $M$. For any $m\in M$, the subgroup $G_m:=\Phi_m^{-1}(m)=\{g\in G\mid g\cdot m = m\}$ is called the {\em isotropy group}\index{isotropy group} of $m$. \et \begin{thr} {\bf Remark. }\rm\label{isotropyrem} (i) Isotropy groups at equivalent points in $M$ are conjugate subgroups of $G$. In fact, suppose that $m'=gm$. Then $(gG_mg^{-1})m' = m'$, so $gG_mg^{-1}\subseteq G_{m'}$, and analogously $g^{-1}G_{m'}g\subseteq G_m$, so $gG_mg^{-1} = G_{m'}$. (ii) For any $m\in M$, the map $\Phi_m: G\to M$ projects to a map $\Psi_m:G/G_m \to M$ defined by $gG_m \mapsto gm$. The range of $\Psi_m$ is the orbit of $m$. Also, $\Psi_m$ is injective: if $\Psi_m(g_1G_m) = \Psi_m(g_2G_m)$ then $g_1m=g_2m$, so $g_1^{-1}g_2\in G_m$, i.e., $g_1G_m = g_2G_m$. \et \begin{thr} {\bf Remark. }\rm\label{genltg} Let $\Phi\colon G\times M\to M$ be a Lie transformation group on a $T_2$-manifold $M$. Then any $G_m$ is a closed subgroup of $G$. Hence by \cite[21.7]{KLie}, $G_m$ is either discrete or it admits a unique structure as a (regular) submanifold of $G$. In the latter case it is also a Lie subgroup of $G$. If $m'=gm$ then by \ref{isotropyrem} (i), $G_m$ is mapped onto $G_{m'}$ by the diffeomorphism $L_g\circ R_{g^{-1}}$. Thus the isotropy groups at points of an orbit are either all discrete or are regular submanifolds and Lie subgroups of $G$ that are pairwise diffeomorphic (since $G_m$, $G_{m'}$ are regular submanifolds, the restriction of $L_g\circ R_{g^{-1}}$ is also a diffeomorphism from $G_m$ onto $G_{m'}$). \et If $G_m$ is open then it must be a union of connected components of $G$, which, by \cite[2.4]{KLie} are precisely the cosets of the normal subgroup $G_e$. If, for example, $G=g_1G_e\cup\dots\cup g_kG_e$ and $G_m = g_1G_e\cup\dots\cup g_lG_e$, then $G/G_m = \{g_{1}G_m,\dots,g_kG_m\}$, since for $g\in g_jG_e$ we have $$ \pi(g)=g\cdot \bigcup_{i=1}^l g_iG_e = \bigcup_{i=1}^l g G_e g_i = \bigcup_{i=1}^l g_jG_e g_i = g_j G_m. $$ Also, the orbit of $m$ only consists of finitely many points (namely $\Psi_m(G/G_m)=\{g_{1}\cdot m,\dots,g_k\cdot m\}$). Otherwise, we have: \bt\label{ltgiso} Let $\Phi\colon G\times M \to M$ be a transformation group on a $T_2$-manifold $M$ and let $m\in M$. If the isotropy group $G_m$ of $m$ is not open in $G$ then the map $\Psi_m$ from \ref{isotropyrem} (ii) is an injective immersion of the quotient manifold $G/G_m$ into $M$. \et \pr Since $\Phi_m = \Psi_m\circ \pi$ (with $\pi:G\to G/G_m$), $\Psi_m$ is the projection of the smooth map $\Phi_m$, hence is itself smooth (see \cite[1.1.9]{DG2}). Also, $G/G_m$ is a quotient manifold of $G$ by \cite[21.5]{KLie}. Since $\Psi_m$ is injective by \ref{isotropyrem} (ii), it remains to show that its rank in any point equals the dimension of $G/G_m$. Since $\pi$ is a submersion, this is the case if and only if the rank of $\Phi_m$ is everywhere equal to $\dim(G/G_m)$. We begin by showing that this is true at $e$. Let $v\in T_eG$ such that $T_e\Phi_m(v)=0$. By \eqref{phidef} this means that the vector field $\Phi(v)$ has a zero at $m$. Therefore \ref{flowaction} implies that $\mathrm{Fl}_\varepsilon^{\Phi(v)}(m)=\exp(\varepsilon v)m =m$ for all $\varepsilon$, i.e., $\exp(\varepsilon v)\in G_m$ for all $\varepsilon$. Now if $G_m$ is discrete then the image of $\varepsilon\to \exp(\varepsilon v)$, being connected, must consists solely of $e\in G$, so $v = \left.\frac{d}{d \eps}\right|_0\exp(\varepsilon v)=0$. In this case, then, $T_e\Phi_m$ is injective, and so the rank of $T_e\Phi_m$ equals the dimension of $G$, and thereby the dimension of $G/G_m$. If $G_m$ is non-discrete then by \ref{genltg} it is a regular submanifold of $G$. Hence $\varepsilon\mapsto \exp(\varepsilon v)$ is smooth as a map into $G_m$ (see \cite[1.1.14]{DG2}), and so $v = \left.\frac{d}{d \eps}\right|_0\exp(\varepsilon v) \in T_eG_m$. Altogether, we obtain that $\ker(T_e\Phi_m) \subseteq T_eG_m$. Conversely, $\Phi_m$ is constant on $G_m$, so $T_e\Phi_m|_{T_eG_m} \equiv 0$, hence in fact $\ker(T_e\Phi_m) = T_eG_m$. Consequently, using \cite[21.7]{KLie} we obtain $$ \mathrm{rk}(T_e\Phi_m) = \dim G - \dim G_m = \dim G/G_m. $$ Finally, if $g$ is an arbitrary point in $G$ then $\Phi_m\circ R_g = \Phi_{m'}$, where $m'=gm$. Then since $R_g$ is a diffeomorphism we have $$ \mathrm{rk}_g(\Phi_m) = \mathrm{rk}_e(\Phi_{m'}) = \dim G/G_{m'}. $$ Now by \ref{genltg} $G_m$ and $G_{m'}$ are either both discrete or they are diffeomorphic, so we conclude that the rank of $\Phi_m$ equals $\dim G/G_{m}$ for every $m\in M$. \ep \begin{thr} {\bf Corollary. }\label{ltgisocor} Under the assumptions of \ref{ltgiso}, the orbit $G\cdot m$ of any $m\in M$ can be endowed with the structure of an immersive submanifold of $M$ diffeomorphic to $G/G_m$. \et \pr For clarity, we write $\tilde \Psi_m$ for $\Psi_m$, viewed as a (bijective) map from $G/G_m$ to $G\cdot m$. Declaring $\tilde \Psi_m$ to be a diffeomorphism provides $G\cdot m$ with a differentiable structure with respect to which the inclusion map $j:G\cdot m \hookrightarrow M$ is an immersion since $j\circ \tilde\Psi_m = \Psi_m: G/G_m \to M$ is an immersion. \ep \begin{thr} {\bf Remark. }\rm\label{conrem} Suppose that $G$ is connected and let $m\in M$. By \ref{ltgiso} the orbit $G\cdot m$ can be discrete only if $G_m$ is open. In this case, $G_m$ is open and closed in $G$, so $G_m=G$ and therefore $G\cdot m = \{m\}$. \et \begin{thr}{\bf Example. }\rm\label{rotex} Let $M=\mathbb R^2$, $G=(\mathbb R,+)$, and $X=-y\partial_x+x\partial_y$ the rotation vector field on $M$. Then we obtain a Lie transformation group $\Phi$ by $$ \Phi\colon (\varepsilon,(x,y))\mapsto \mathrm{Fl}^X_\varepsilon(x,y) = (x\cos\varepsilon - y\sin\varepsilon, x\sin\varepsilon + y\cos\varepsilon). $$ The orbits of $\Phi$ are the regular submanifolds $\{x^2+y^2=\text{const}\}$ and the singleton $\{(0,0)\}$. \et Recall that $G$ is said to act transitively\index{transitive group action} on $M$ if for any $m$, $m'\in M$ there exists some $g\in G$ with $gm=m'$. Such a group action possesses only a single orbit, namely the manifold $M$ itself. By \ref{isotropyrem} (i) this means that any $\Psi_m$ is a bijection from $G/G_m$ onto $M$. To further elaborate on this, we will need the following auxiliary result: \blem \label{immerlem} Let $M^m$ and $N^n$ be manifolds and suppose that $M$ is second countable and that $m<n$. Then an immersion $\Psi: M\to N$ cannot be onto any open subset $W$ of $N$. \et \pr It suffices to take for $W$ the domain of a chart $\chi$ in $N$. Let $p$ be any point in $\Psi^{-1}(W)$ and choose charts $(U,\varphi=(x^1,\dots,x^m))$ around $p$ in $M$ and $(V,\eta=(y^1,\dots,y^n))$ around $\Psi(p)$ such that $\Psi(U)\subseteq V\subseteq W$ and such that $\eta\circ\Psi\circ \varphi^{-1} = (x^1,\dots,x^m) \to (x^1,\dots,x^m,0,\dots,0)$. Then $\eta(\Psi(U))\subseteq \mathbb R^m\times \{0\} \subseteq \mathbb R^n$, so it has Lebesgue measure $0$. It follows that also the image $\chi(\Psi(U))$ of this set under the smooth map $\chi\circ \eta^{-1}$ has Lebesgue measure $0$. As $p$ varies in $\Psi^{-1}(W)$, the domains $U$ cover the set $\Psi^{-1}(W)$. As $M$ is second countable we may extract a countable subcover $\{U_k\mid k\in \N\}$ from this collection. Then the sets $\Psi(U_k)$ cover $W\cap \Psi(M)$, entailing $$ \chi(W\cap \Psi(M)) = \bigcup_{k\in \N} \chi(\Psi(U_k)). $$ But then $\chi(W\cap \Psi(M))$ has Lebesgue measure $0$ and so it cannot be all of $\chi(W)$. It follows that $\Psi(M)$ cannot contain all of $W$. \ep Using this, we can prove: \bp\label{diffeotrans} Let $G$ be a second countable Lie group that acts transitively as a Lie transformation group on the $T_2$-manifold $M$. Then $M$ is diffeomorphic to $G/G_m$, for any $m\in M$. \et \pr Fix any $m\in G$. If $G_m$ is open, then so is any $gG_m$, hence any point in $G/G_m$ (because $\pi^{-1}(\pi(g)) = gG_m$), which is therefore discrete (cf.\ \cite[Rem.\ 20.1]{KLie}). Otherwise, by \cite[21.5]{KLie} $G/G_m$ possesses a differentiable structure as a quotient manifold of $G$. In both cases, the quotient map $\pi: G\to G/G_m$ is open and continuous, so also the topology of $G/G_m$ is second countable. Hence if $G/G_m$ were discrete it would be countable. But $\Psi_m$ is bijective, so this would imply that $M$ was countable, which is impossible. Hence $G/G_m$ is not discrete, and so it is a quotient manifold of $G$ with a countable basis for its topology. Also, by \ref{ltgiso} $\Psi_m$ is an injective immersion of $G/G_m$ into $M$. Hence $\dim G/G_m \le \dim M$. Since $\Psi_m$ is onto $M$, \ref{immerlem} implies that the dimensions in fact are equal. As $\Psi_m:G/G_m\to M$ is an immersion, it follows that its tangent map is bijective at any point. Thus by the inverse function theorem it is a local diffeomorphism, hence a global diffeomorphism since it is bijective. \ep \begin{thr} {\bf Corollary. } Let $G$ be a second countable Lie group that acts transitively and freely as a Lie transformation group on the $T_2$-manifold $M$. Then $M$ is diffeomorphic to $G$. \et \pr If $G$ acts freely then $G_m=\{e\}$ for every $m\in M$. Hence $G/G_m=G$ and the result follows from \ref{diffeotrans}. \ep \begin{thr} {\bf Corollary. } Let $G$ be a compact Lie group that acts transitively as a Lie transformation group on the $T_2$-manifold $M$. Then $M$ is compact. \et \pr Since $G$ is compact, it is second countable. By \ref{diffeotrans}, $M$ is diffeomorphic to $G/G_m$, for any $m\in M$. Let $\pi:G\to G/G_m$ be the quotient map. Then $\pi(G)=G/G_m$ is compact, hence so is $M$. \ep \begin{thr}{\bf Example. }\rm We continue our study of the action of $O(n,\mathbb R)$ on $S^{n-1}$ as a Lie transformation group from \ref{onex}. This action is given by \begin{equation*} \begin{split} \Phi\colon O(n,\mathbb R)\times S^{n-1} &\to S^{n-1}\\ (A,x) &\mapsto Ax. \end{split} \end{equation*} For any vector $v\in S^{n-1}$ we can find an orthogonal matrix $T$ with $v$ as its first column, so $Te_1=v$. Hence $\Phi$ has only a single orbit, $S^{n-1}$, i.e., it is transitive. The isotropy group $G_{e_1}$ of $e_1$, i.e., those elements of $O(n,\mathbb R)$ that leave $e_1$ unchanged, are the matrices of the form $$ A = \begin{pmatrix} 1 & 0\\ 0 & D \end{pmatrix} $$ with $D\in O(n-1,\mathbb R)$. By \ref{diffeotrans} we conclude that $S^{n-1}$ is diffeomorphic to the quotient manifold $O(n,\mathbb R)/G_{e_1}$, i.e., $S^{n-1}\cong O(n,\mathbb R)/O(n-1,\mathbb R)$. \et \section{Groups of transformations on a $T_2$-manifold} \begin{thr} {\bf Definition. } A set $G$ of transformations on a $T_2$-manifold $M$ which is a group under composition $(g_1,g_2)\mapsto g_1\circ g_2$ $(g_1,\,g_2\in G)$ is called a group of transformations\index{group of transformations} of $M$. \et According to \ref{transgroup}, the map $\Phi\colon (g,m)\mapsto g(m)\equiv gm$ then defines a transformation group on $M$. Moreover, this action obviously is effective. In this section we want to analyze whether $G$ can be endowed with a Lie group structure such that $\Phi$ is smooth. The following example shows that such a structure may not be unique in general: \begin{thr}{\bf Example. }\rm Let $G$ be the group of translations on $\mathbb R^2$ and let $\phi_a:= z\mapsto z+a$. Then $G$ is bijectively mapped onto $\mathbb R^2$ by $\phi_a\mapsto a$, and this defines a ${\mathcal C}^\infty$-structure on $G$, for any given ${\mathcal C}^\infty$-structure on $\mathbb R^2$. But on $\mathbb R^2$ there are different such structures, e.g., the standard one and the one from \cite[Ex.\ 14.3]{KLie}. Both of these induce corresponding structures on $G$ such that $\Phi$ is smooth. \et Recall from \ref{complvf} that a complete vector field $X$ on $M$ induces an action of $(\mathbb R,+)$ on $M$ as a Lie transformation group via the flow of $X$, $\Psi_\varepsilon:=m\mapsto \mathrm{Fl}^X_\varepsilon(m)$. We call the vector field $X$ {\em tangent}\index{vector field!tangent to group of transformations} to $G$ if all the resulting transformations $\Psi_\varepsilon$ belong to $G$. \begin{thr} {\bf Definition. } \label{unicond} A group of transformations on a $T_2$-manifold $M$ is called a {\em Lie group of transformations}\index{Lie group of transformations} if it admits a Lie group structure such that \begin{itemize} \item[(i)] The map $\Phi\colon(g,m)\mapsto gm$ is smooth. \item[(ii)] If $X\in {\mathfrak X}(M)$ is complete and tangent to $G$ then the group homomorphism $\varepsilon\mapsto \Psi_\varepsilon$ is a one-parameter subgroup of $G$. \end{itemize} \et The above is the desired property entailing uniqueness: \bp A group of transformations of a $T_2$-manifold $M$ admits at most one structure as a Lie group of transformations of $M$. \et \pr Let $G$ and $G'$ be two Lie group structures on a group of transformations that both satisfy \ref{unicond}. We have to show that $\mathrm{id}: G\to G'$ is a diffeomorphism. We first note that by \ref{unicond} (i), $G$ is a Lie transformation group on $M$, denoted by $\Phi\colon G\times M \to M$. Hence by \ref{flowaction} any $v\in T_eG$ defines a complete vector field $\Phi(v)\in {\mathfrak X}(M)$ with $\mathrm{Fl}_\varepsilon^{\Phi(v)}(m)= \exp(\varepsilon v)m$. It follows that $\Phi(v)$ is tangent to $G$. Now \ref{unicond} (ii) implies that $\varepsilon\mapsto \mathrm{id}\circ \exp(\varepsilon v):\mathbb R\to G'$ is a one-parameter subgroup of $G'$. By \cite[8.3]{KLie} there is a unique $v'\in T_eG'$ such that $\mathrm{id}\circ \exp(\varepsilon v) = \exp'(\varepsilon v')$ (with $\exp'$ the exponential map of $G'$). Thereby we obtain a well-defined map $v\mapsto v'$ from $T_eG$ to $T_eG'$. Fixing a basis $(v_1,\dots,v_n)$ of $T_eG$, let $\varphi=(x^1,\dots,x^n)$ be the corresponding canonical chart of the second kind, cf.\ \cite[Rem.\ 8.5]{KLie}. Then since $\mathrm{id}$ is a group homomorphism, $$ \mathrm{id}\circ(\exp(x^1 v_1)\dots\exp(x^n v_n)) = \exp'(x^1 v_1')\dots\exp(x^n v_n'), $$ implying that $\mathrm{id}$ is smooth on the domain of $\varphi$. Smoothness at an arbitrary point of $G$ then follows by writing $\mathrm{id} = L'_g\circ \mathrm{id} \circ L_{g^{-1}}$. Interchanging $G$ and $G'$ in the above argument shows that $\mathrm{id}$ in fact is a diffeomorphism. \ep \begin{thr} {\bf Remark. }\rm By \ref{unicond} (i), any Lie group $G$ of transformations is also a Lie transformation group that acts effectively on $M$. Therefore, \ref{phiiso} implies that ${\mathscr R}(G,M)$ is Lie algebra-isomorphic to ${\mathfrak g}_R$. \et \bp\label{gtransnec} If a group $G$ of transformations of a $T_2$-manifold $M$ admits the structure of a Lie group of transformations then the set ${\mathfrak X}_t(M)$ of complete vector fields on $M$ tangent to $G$ is a finite-dimensional Lie algebra (namely ${\mathscr R}(G,M)$). \et \pr Denote the action of $G$ on $M$ by $\Phi\colon G\times M\to M$. By \ref{flowaction}, every $\Phi(v)\in{\mathscr R}(G,M)$ is complete with corresponding transformations $\Psi_\varepsilon: m\mapsto \mathrm{Fl}_\varepsilon^{\Phi(v)}(m)= \exp(\varepsilon v)m$, hence belonging to $G$. Thus $\Phi(v)$ is a complete vector field on $M$ that is tangent to $G$, i.e., ${\mathscr R}(G,M) \subseteq {\mathfrak X}_t(M)$. Conversely, let $X\in {\mathfrak X}_t(M)$ and let $\Psi_\varepsilon:= m\mapsto \mathrm{Fl}^X_\varepsilon(m)$ be the corresponding transformation of $M$. By \ref{unicond} (ii), $\varepsilon\mapsto \Psi_\varepsilon$ is a one-parameter subgroup of $G$, so by \cite[8.3]{KLie} there is a unique $v\in T_eG$ with $\Psi_\varepsilon(m) = \exp(\varepsilon v)m$ for all $m\in M$. Thus the vector fields $X$ and $\Phi(v)$ have the same maximal integral curves and hence coincide, so $X\in {\mathscr R}(G,M)$. \ep \begin{thr} {\bf Remark. }\rm There is in fact also a converse to \ref{gtransnec}, see \cite[p.\ 103]{Palais}: Let $G$ be a group of transformations of a $T_2$-manifold $M$. If ${\mathfrak X}_t(M)$ generates a finite-dimensional Lie algebra $\mathcal A$ of vector fields on $M$ then $G$ admits the structure of a Lie group of transformations of $M$ and ${\mathcal A}={\mathscr R}(G,M)$. \et \chapter{Integrability of distributions of non-constant rank}\label{nonconstchap} \section{Distributions of non-constant rank} In \cite[17.32]{KLie} we proved the classical Frobenius theorem\index{Frobenius Theorem} on the integrability of distributions of constant rank: \bt\label{classfrob} Let $M$ be an $n$-dimensional $T_2$-manifold and let $\Delta$ be a $k$-dimensional distribution on $M$. Then the following are equivalent: \begin{itemize} \item[(i)] $\Delta$ is involutive. \item[(ii)] Every point in $M$ lies in the domain of a flat chart\index{flat chart} $\varphi=(x^1,\dots,x^n)$ for $\Delta$, i.e., such that $\partial_{x^1}|_m,\dots,\partial_{x^k}|_m$ forms a basis of $\Delta(m)$ for each $m$ in the domain of $\varphi$. \item[(iii)] Every point of $M$ is contained in an integral manifold of $\Delta$. \item[(iv)] Every point $m$ of $M$ lies in a cubic chart $(\varphi=(x^1,\dots,x^n),U)$, $\varphi(U)=[-c,c]^n$ centered around $m$ such that the slices $U_a=\varphi^{-1}(\mathbb R^k\times \{a\})$ are integral manifolds of $\Delta$. If $M'$ is a connected integral-manifold of $\Delta$ with $M'\subseteq U$ then $M'$ is contained in one such slice. \end{itemize} \et Note that in \cite{KLie} we called a distribution integrable if it satisfies (ii). However, in this chapter we will use (iii) instead, cf.\ \ref{intedef} below. For the application to orbits of (local) transformation groups we have in mind, \ref{classfrob} is too restrictive in that it requires $\Delta$ to have constant rank (i.e., dimension) $k$ everywhere. In this chapter we closely follow H.\ Sussmann's article \cite{Sussmann}, as well as P.\ Michor's exposition in \cite{M} to develop a theory of integrability for distributions of non-constant rank. Throughout this chapter, all manifolds are supposed to be $T_2$ and paracompact. \begin{thr} {\bf Definition. } We call $$ {\mathfrak X}_{\mathrm{loc}}(M) := \bigcup \{{\mathfrak X}(U)\mid U\subseteq M \text{ open}\} $$ the space of local vector fields\index{local vector field} on $M$. \et If $X$, $Y$ are local vector fields on $M$, then so is $[X,Y]$, defined on the intersection of the domains of $X$ and $Y$. We agree to consider the `empty vector field' an element of ${\mathfrak X}_{\mathrm{loc}}(M)$, to avoid having to add formulations like `if the domains of $X$ and $Y$ intersect', and similar for further local notions to be introduced below. For any $X\in {\mathfrak X}_{\mathrm{loc}}(M)$ we denote the maximal domain of $\mathrm{Fl}^X$ in $\mathbb R\times M$ by $U_X$. $U_X$ is an open subset of $\mathbb R\times M$ (see \cite[2.5.17]{DG1}). For each $t\in \mathbb R$, $\mathrm{Fl}^X_t$ is a diffeomorphism of some maximal open set $U_t(X)$ (which may be empty) onto some open set $U_t'(X)$. Note that $U_t'(X) = U_{-t}(X)$. For any $n\ge 1$ in $\N$, any $\xi=(X_1,\dots,X_n)\in {\mathfrak X}_{\mathrm{loc}}(M)^n$, any $T=(t_1,\dots,t_n)\in \mathbb R^n$, and any $m\in M$ we set \begin{equation}\label{xitdef} \xi_T(m) := \mathrm{Fl}^{X_1}_{t_1}(\mathrm{Fl}^{X_2}_{t_2}(\dots \mathrm{Fl}^{X_n}_{t_n}(m)\dots)). \end{equation} The maximal domain of $(T,m)\mapsto \xi_T(m)$ then is an open subset of $\mathbb R^n\times M$ that will be denoted by $U_\xi$, and we let $U_T(\xi)$ be the set of all $m\in M$ such that $\xi_T(m)$ is defined, i.e., $U_T(\xi) = \{m\in M\mid (T,m)\in U_\xi\}$. \begin{thr} {\bf Definition. } A diffeomorphism $f\colon U\to U'$ between open subsets of $M$ is called a local diffeomorphism.\index{local diffeomorphism} \et If $f_i: U_i\to U_i'$ ($i=1,2$) are local diffeomorphisms, then so is $f_1\circ f_2$, with domain $f_2^{-1}(U_1)$ and range $f_1(U_2'\cap U_1)$. Moreover, $f_1^{-1}:U_1'\to U_1$ is a local diffeomorphism as well. A group of local diffeomorphisms\index{local diffeomorphisms!group of} is a set $G$ of local diffeomorphisms that is closed under composition and inverses. Our main examples of local diffeomorphisms are flows of local vector fields. For any $X\in {\mathfrak X}_{\mathrm{loc}}(M)$ the set of all $\mathrm{Fl}^X_t$ ($t\in \mathbb R$) is called the group of local diffeomorphisms generated by $X$, and is denoted by $G_X$. More generally, if $D$ is any subset of ${\mathfrak X}_{\mathrm{loc}}(M)$ then there exists a smallest group of local diffeomorphisms containing $\bigcup\{G_X\mid X\in D\}$. This group (more precisely, pseudogroup) will be denoted by $G_D$. It is called the group of local diffeomorphisms generated by $D$.\index{group of local diffeomorphisms} By definition, we have \begin{equation}\label{gddef} G_D = \{\xi_T\mid \exists n:\, \xi\in D^n,\, T\in \mathbb R^n\}. \end{equation} Given finite tuples $\lambda=(\lambda_1,\dots,\lambda_m)$, $\mu=(\mu_1,\dots,\mu_k)$, by $\lambda \mu$ we denote their concatenation $(\lambda_1,\dots,\lambda_m,\mu_1,\dots,\mu_k)$, and by $\hat\lambda$ we denote the reverse tuple $(\lambda_m,\dots,\lambda_1)$. With these notations we have $$ \xi_T\eta_{T'}:=\xi_T\circ\eta_{T'} = (\xi\eta)_{TT'}\quad \text{and}\quad (\xi_T)^{-1} = \xi_{-\hat T}. $$ A subset $D$ of ${\mathfrak X}_{\mathrm{loc}}(M)$ is called {\em everywhere defined}\index{everywhere defined} if the union of all domains of elements of $D$ covers $M$. An analogous definition applies to groups of local diffeomorphisms. \begin{thr} {\bf Definition. }\label{orb} Let $G$ be an everywhere defined group of local diffeomorphisms. Two elements $m_1$, $m_2$ of $M$ are called $G$-equivalent if there exists some $g\in G$ such that $g(m_1)=m_2$. The equivalence classes of the resulting equivalence relation on $M$ are called the orbits\index{orbit} of $G$. If $D\subseteq {\mathfrak X}_{\mathrm{loc}}(M)$ is everywhere defined then the $G_D$-orbits are also called $D$-orbits. The orbit of $m\in M$ is called {\em trivial}\index{orbit!trivial} if it equals $\{m\}$. \et Thus $m_1$, $m_2$ belong to the same $D$-orbit if and only if there exists some $n\ge 1$, some $\xi\in D^n$ and some $T\in \mathbb R^n$ such that $\xi_T(m_1)=m_2$. From this we immediately conclude: \blem Two points $m_1$, $m_2$ belong to the same $D$-orbit if and only if there exists a piecewise smooth curve $\gamma:[a,b]\to M$ such that $\gamma(a)=m_1$, $\gamma(b)=m_2$, with the following property: There exist $a=t_0<t_1<\dots < t_r=b$ and $X_i\in D$ ($i=1,\dots,r$) such that, for each $i$, $\gamma|_{[t_{i-1},t_i]}$ is an integral curve of $X_i$ or of $-X_i$. \et Such a curve $\gamma$ will be called a piecewise integral curve\index{piecewise integral curve} of $D$. \begin{thr} {\bf Remark. }\rm\label{orbittop} We now want to endow the orbits of $D$ with a natural topology. To this end, for any $m\in M$ and $\xi\in D^n$, let $\rho_{\xi,m}:= T \mapsto \xi_T(m)$, and let $U_{\xi,m}\subseteq\mathbb R^n$ be its domain. Then the $D$-orbit $S_m$ of $m$ is the union of all the images of the mappings $\rho_{\xi,m}$. We equip $S:=S_m$ with the finest topology such that each $\rho_{\xi,m}$ ($n\ge 1$, $\xi\in D^n$) is continuous. Since each $\rho_{\xi,m}$ is continuous for the trace topology of $M$ on $S$ it follows that the topology of $S$ is finer than the trace topology, i.e., $S\hookrightarrow M$ is continuous. In particular, $S$ is $T_2$. In general, the topology on $S$ will be strictly finer than the trace topology. Since all $U_{\xi,m}$ are connected and their images all contain $m$, $S$ is connected. \et \begin{thr} {\bf Remark. }\rm The topology on $S$ does not depend on the choice of $m\in S$. To see this, denote by $S_m$ the set $S$ equipped with the topology induced by the maps $\rho_{\xi,m}$. By symmetry, it will suffice to show that $\mathrm{id}:S_m\to S_{m'}$ is continuous for all $m$, $m'\in S$. Pick $\eta$, $T_0$ such that $\eta_{T_0}(m')=m$. Then $\rho_{\xi,m} = T\mapsto \xi_T\eta_{T_0}(m')$, which is the composition of $T\to TT_0$ and $\rho_{\xi\eta,m'}$. Now $\rho_{\xi\eta,m'}$ is continuous into $S_{m'}$, so $\rho_{\xi,m}:U_{\xi,m}\to S_{m'}$ is continuous. By the universal property of the finest topology, the claim follows. \et We next generalize the definition (\cite[Def.\ 17.1]{KLie}) of distribution\index{distribution} to the variable rank setting: \begin{thr} {\bf Definition. } A distribution on a manifold $M$ is a mapping $\Delta$ that assigns to every $m\in M$ a linear subspace $\Delta(m)$ of $T_mM$. A set $D$ of local vector fields is said to span $\Delta$ if, for every $m\in M$, $\Delta(m) = \text{span}\{X(m)\mid X\in D\}$. \et If $D\subseteq {\mathfrak X}_{\mathrm{loc}}(M)$ is everywhere defined then it spans a unique distribution, which will be denoted by $\Delta_D$. Any distribution which is of the form $\Delta_D$ for some everywhere defined family $D\subseteq {\mathfrak X}_{\mathrm{loc}}(M)$ is called smooth.\index{distribution!smooth} We say that a local vector field $X$ on $M$ belongs to a distribution $\Delta$ if $X(m)\in \Delta(m)$ for every $m$ in the domain of $X$. Let $$ D_\Delta := \{X\in {\mathfrak X}_{\mathrm{loc}}(M) \mid X \text{ belongs to } \Delta\}. $$ Then $\Delta$ is ${\mathcal C}^\infty$ if and only if $\Delta$ is spanned by $D_\Delta$. A distribution\index{distribution!invariant} $\Delta$ is called invariant under a group of local diffeomorphisms $G$ if \begin{equation}\label{ginvdef} \forall m\in M: \quad T_mg(\Delta(m))\subseteq \Delta(g(m)) \end{equation} for all $g\in G$ such that $m$ is in the domain of $g$. In this case also $T_{gm}g^{-1}$ maps $\Delta(g(m))$ into $\Delta(m)$, i.e., $T_mg(\Delta(m)) = \Delta(g(m))$. It follows that the dimension of $\Delta(m)$ is the same for all points $m$ in the same $G$-orbit. A distribution $\Delta_1$ is said to be contained in a distribution $\Delta_2$ if $\Delta_1(m) \subseteq\Delta_2(m)$ for all $m\in M$. If $\Delta$ is a distribution and $G$ is a group of local diffeomorphisms on $M$ then there is a smallest distribution $\Delta^G$ which contains $\Delta$ and is $G$-invariant (namely the intersection of all such distributions). More precisely, $\Delta^G(m)$ is the linear span of all vectors $v\in T_mM$ such that $v\in \Delta(m)$ or $v=T_{m'}g(w)$ for some $g\in G$ and $m'\in M$ with $w\in \Delta(m')$ and $m=gm'$. \begin{thr} {\bf Remark. }\rm\label{deltagcinfty} Let $\Delta$ be spanned by $D\subseteq {\mathfrak X}_{\mathrm{loc}}(M)$. Then $\Delta^G$ is spanned by the family $$ D \cup \{g_*X \mid X\in D,\, g\in G \text{ s.t. } g_*X \text{ is defined }\}. $$ It follows that if $\Delta$ is ${\mathcal C}^\infty$ then so is $\Delta^G$. \et Note that if $G=G_D$ then in fact $\Delta^{G_D}$ is spanned by \begin{equation}\label{gxd} {\mathcal D}:=\{g_*X \mid X\in D,\, g\in G \text{ s.t. } g_*X \text{ is defined }\} \end{equation} (i.e., the union with $D$ is not required). To see this, note that any $X\in D$ with domain, say, $U$ can be written as $X=g_*X$ for $g=\mathrm{Fl}^X_0=\mathrm{id}_U\in G_D$. If $D\subseteq {\mathfrak X}_{\mathrm{loc}}(M)$ then a $G_D$-invariant distribution is called $D$-invariant. The smallest $D$-invariant distribution which contains $\Delta$ is denoted by $\Delta^D$, i.e., $$ \Delta^D := \Delta^{G_D}. $$ Next, let $D$ be an everywhere defined subset of ${\mathfrak X}_{\mathrm{loc}}(M)$. Then we set \begin{equation}\label{pddef} P_D := \Delta_D^D := (\Delta_D)^D. \end{equation} Thus $P_D$ is the smallest distribution that is $D$-invariant and contains $\Delta_D$. By \ref{deltagcinfty}, $P_D$ is smooth and the dimension of $P_D(m)$ depends only on the $D$-orbit of $m$. \begin{thr} {\bf Remark. }\rm\label{imponote} It is important to note that the $D$-orbits are precisely the $P_D$-orbits: to see this, by \eqref{gddef} and the remark following \ref{orb} it suffices to note that if $g = \xi_T\in G_D$ and $X\in D$, then $X\sim_{\xi_T} g_*X$, and therefore (by \cite[17.8]{KLie}), $$ \mathrm{Fl}^{g_*X}_t = \xi_T\circ\mathrm{Fl}^X_t\circ\xi_{-\hat T}. $$ Hence the flows of the $g_*X$ do not alter the $D$-orbits. \et Thus the following definition makes sense: \begin{thr} {\bf Definition. }\label{orbitrank} Let $S$ be an orbit of an everywhere defined subset $D$ of ${\mathfrak X}_{\mathrm{loc}}(M)$. For any $m\in S$, the dimension of $P_D(m)$ is called the rank\index{orbit!rank of} of $S$. \et Trivial orbits are characterized by the following result: \blem\label{trivorb} Let $D\subseteq {\mathfrak X}_{\mathrm{loc}}(M)$ be everywhere defined and let $S$ be the orbit of $m\in M$. The following are equivalent: \begin{itemize} \item[(i)] $P_D(m)=\{0\}$. \item[(ii)] The orbit of $m$ is trivial\index{orbit!trivial}, i.e., $S=\{m\}$. \end{itemize} \et \pr (i)$\Rightarrow$(ii): By \eqref{gxd}, for any $X\in D$ and $g\in G_D$ such that $g_*X$ is defined, $g_*X(m)=0$. Therefore, the flow of any element of $P_D$ leaves $m$ unchanged. (ii)$\Rightarrow$(i): Suppose that $P_D(m)\not=\{0\}$, then by \eqref{gxd} there would exist some $g\in G_D$ and some $X\in D$ with $g_*X(m)\not=0$. But then the flow of $g_*X$ would leave $\{m\}$, contradicting (ii). \ep \begin{thr}{\bf Example. }\rm As in \ref{rotex}, let $M=\mathbb R^2$ and $X:=-y\partial_x + x\partial_y$. Now set $D:=\{X\}$. Then the $D$-orbit of any $(x_0,y_0)$ is a circle through $(x_0,y_0)$ with center $(0,0)$, while the orbit at $(0,0)$ is trivial. In this example, $P_D = \Delta_D$: In fact, this is always true when $D$ consists of only one vector field $X$ because in that case \begin{equation*} \mathrm{Fl}^{(\mathrm{Fl}^X_t)_*X}_s = \mathrm{Fl}^X_t\circ \mathrm{Fl}^X_s\circ \mathrm{Fl}^X_{-t} = \mathrm{Fl}^X_s, \end{equation*} so $X$ and $(\mathrm{Fl}^X_t)_*X$ have the same flow and therefore coincide. \et \begin{thr} {\bf Definition. }\label{dstar} A set $D\subseteq{\mathfrak X}_{\mathrm{loc}}(M)$ is called involutive\index{involutive} if for any $X,\, Y\in {\mathfrak X}_{\mathrm{loc}}(M)$ that belong to $D$ also $[X,Y]$ belongs to $D$. If $D$ is any subset of ${\mathfrak X}_{\mathrm{loc}}(M)$ then the smallest involutive subset of ${\mathfrak X}_{\mathrm{loc}}(M)$ that contains $D$ will be denoted by $D^*$. A smooth distribution $\Delta$ is called involutive if the corresponding set $D_\Delta$ is involutive. \et \blem\label{incldist} Let $D$ be an everywhere defined subset of ${\mathfrak X}_{\mathrm{loc}}(M)$. Then $$ \Delta_D \subseteq \Delta_{D^*} \subseteq P_D. $$ \et \pr The first inclusion is immediate since $D\subseteq D^*$. The second inclusion follows since $P_D$ is involutive by \ref{sussmain1} (iv) below. \ep \begin{thr}{\bf Example. }\rm\label{rankdist} Clearly the first inclusion in \ref{incldist} can be proper. This example shows that the same may happen for the second inclusion. Let $M=\mathbb R^2$, $X_1:=\frac{\partial}{\partial x}$, $X_2=\phi\frac{\partial}{\partial y}$, where $\phi(x,y)=\psi(x)$ and $\psi$ is a smooth function with $\psi(x)=0$ for $x\le 0$ and $\psi(x)>0$ for $x>0$. Let $D:=\{X_1,\,X_2\}$. Then since $D(x,y)$ has dimension $2$ for $x>0$, the same is true for $P_D\supseteq \Delta_D$. Moreover, any point in $\mathbb R^2$ can be joined to a point $(x,y)$ with $x>0$ by a piecewise integral curve of $D$, so in fact $P_D$ has dimension $2$ everywhere. However, for $x\le 0$ the distribution $\Delta_{D^*}$ has dimension $1$. \et \begin{thr} {\bf Definition. }\label{intedef} An immersive submanifold $S$ of $M$ is called an integral mani\-fold\index{integral manifold} of a distribution $\Delta$ on $M$ if, for all $s\in S$, we have $T_sS = \Delta(s)$. A $C^\infty$-distribution $\Delta$ is said to be integrable\index{integrable}, or to have the integral manifold property\index{integral manifold property} if, for every $m\in M$, either the orbit of $m$ is trivial or there exists an integral manifold $S$ of $\Delta$ such that $m\in S$. \et Note that we notationally suppress the inclusion map $j: S\hookrightarrow M$ here. If a ${\mathcal C}^\infty$-distribution $\Delta$ is integrable then a smooth vector field $X$ belongs to $\Delta$ if and only if $X$ is tangent to every integral manifold of $\Delta$. By \cite[17.16]{KLie}, any such $X$ can locally be viewed as a smooth vector field on any given integral manifold. Moreover, by \cite[17.22]{KLie} and \ref{trivorb} we have: \blem\label{intinv} Any integrable ${\mathcal C}^\infty$-distribution is involutive. \et Note, however, that the converse is not true in the present situation (contrary to constant rank distributions!). In fact, \ref{rankdist} provides an example of a distribution $\Delta_D$ that is involutive: \begin{equation*} [X_1,X_2](x,y) = \psi'(x)\partial_y = \left\{ \begin{array}{rl} \frac{\psi'(x)}{\psi(x)}X_2 & x>0\\ 0 & x\le 0 \end{array} \right. \end{equation*} but cannot have the integral manifold property (the dimension of the integral manifolds at $x=0$ would have to be $1$ and $2$). \begin{thr} {\bf Definition. } Let $\Delta$ be a smooth distribution on $M$. A maximal integral manifold\index{integral manifold!maximal} of $\Delta$ is a connected immersive submanifold $S$ of $M$ such that \begin{itemize} \item[(i)] $S$ is an integral manifold of $\Delta$, and \item[(ii)] every connected integral submanifold of $\Delta$ that intersects $S$ is an open submanifold of $\Delta$. \end{itemize} $\Delta$ is said to have the maximal integral manifold property if through each point of $M$ with $P_{D_\Delta}(m)\not=0$ (cf.\ \ref{trivorb}) there passes a maximal integral manifold of $\Delta$. \et In particular, any two maximal integral submanifolds through the same point $m$ must coincide. For the discussion below we will need the following auxiliary result: \blem\label{eduard} Let $X$, $Y$ be smooth local vector fields on $M$. Then for any $m$ in the intersection of the domains of $X$, $Y$ we have: $$ [X,Y](m)=\left.\frac{d}{d t}\right|_0 \mathrm{Fl}^Y_{-\sqrt{t}}(\mathrm{Fl}^X_{-\sqrt{t}}(\mathrm{Fl}^Y_{\sqrt{t}}(\mathrm{Fl}^X_{\sqrt{t}}(m)))) $$ \et \pr For any local smooth function $f$ we have \begin{equation}\label{floff} \frac{d}{dt} (\mathrm{Fl}^X_t)^*f = X(f)\circ \mathrm{Fl}^X_t = (\mathrm{Fl}^X_t)^*(L_Xf) = (\mathrm{Fl}^X_t)^*(Tf(X)). \end{equation} Since $$ \left.\frac{d}{d t}\right|_0 f(\mathrm{Fl}^Y_{-\sqrt{t}}(\mathrm{Fl}^X_{-\sqrt{t}}(\mathrm{Fl}^Y_{\sqrt{t}}(\mathrm{Fl}^X_{\sqrt{t}}(m))))) = Tf\left(\left.\frac{d}{d t}\right|_0 \mathrm{Fl}^Y_{-\sqrt{t}}(\mathrm{Fl}^X_{-\sqrt{t}}(\mathrm{Fl}^Y_{\sqrt{t}}(\mathrm{Fl}^X_{\sqrt{t}}(m))))\right) $$ and $Tf([X,Y])(m) = ([X,Y](f))(m)$, the claim will follow if we can show that for any such $f$ we have $$ \left.\frac{d}{d t}\right|_0\left((\mathrm{Fl}^X_{\sqrt{t}})^*(\mathrm{Fl}^Y_{\sqrt{t}})^*(\mathrm{Fl}^X_{-\sqrt{t}})^* (\mathrm{Fl}^Y_{-\sqrt{t}})^*f\right)(m) = ([X,Y](f))(m). $$ Using \eqref{floff} we obtain \begin{equation*} \begin{split} \frac{d}{dt} \Big((\mathrm{Fl}^X_{\sqrt{t}})^*(\mathrm{Fl}^Y_{\sqrt{t}})^*&(\mathrm{Fl}^X_{-\sqrt{t}})^* (\mathrm{Fl}^Y_{-\sqrt{t}})^*f\Big)\\ = & \big((\mathrm{Fl}^X_{\sqrt{t}})^*L_X((\mathrm{Fl}^Y_{\sqrt{t}})^*(\mathrm{Fl}^X_{-\sqrt{t}})^* (\mathrm{Fl}^Y_{-\sqrt{t}})^*f) \\ &+ (\mathrm{Fl}^X_{\sqrt{t}})^*(\mathrm{Fl}^Y_{\sqrt{t}})^*L_Y((\mathrm{Fl}^X_{-\sqrt{t}})^* (\mathrm{Fl}^Y_{-\sqrt{t}})^*f)\\ &- (\mathrm{Fl}^X_{\sqrt{t}})^*(\mathrm{Fl}^Y_{\sqrt{t}})^*(\mathrm{Fl}^X_{-\sqrt{t}})^*L_X((\mathrm{Fl}^Y_{-\sqrt{t}})^*f)\\ &- (\mathrm{Fl}^X_{\sqrt{t}})^*(\mathrm{Fl}^Y_{\sqrt{t}})^*(\mathrm{Fl}^X_{-\sqrt{t}})^* (\mathrm{Fl}^Y_{-\sqrt{t}})^*(L_yf) \big)\cdot \frac{1}{2\sqrt{t}} =: \frac{g(\sqrt{t})}{2\sqrt{t}} \end{split} \end{equation*} We need to calculate the limit as $t\searrow 0$ of this expression. Now since $g(0)=0$ it follows that $$ \lim_{t\searrow 0} \frac{g(\sqrt{t})}{2\sqrt{t}} = \frac{1}{2}g'(0). $$ Again using \eqref{floff} we calculate: \begin{equation*} \begin{split} g'(0) =& \big(L_XL_X + L_X(L_Y-L_X-L_Y) + (L_X+L_Y)L_Y + L_Y(-L_X-L_Y) \\ &- (L_X+L_Y-L_X)L_X + L_XL_Y -L_XL_Y-L_YL_Y+L_XL_Y+L_YL_Y\big)f \\ &= 2(L_XL_Y-L_YL_X)f = 2[X,Y]f, \end{split} \end{equation*} giving the claim. \ep \begin{thr} {\bf Remark. }\rm To clarify the geometric meaning of $\Delta_{D^*}$ and $P_D$, and also to motivate the structure of the main results in the following section, suppose that a subset $D$ of ${\mathfrak X}_{\mathrm{loc}}(M)$ is given and that we want to find a distribution $\Delta$ with the property that the orbits of $D$ are precisely the maximal integral manifolds of $\Delta$. It is then geometrically natural to define $\Delta(m)$ as the set of all tangent vectors of smooth curves that pass through $m$ and lie entirely in the $D$-orbit of $m$. Call this set of curves $\Gamma_m$. For any $X\in D$, the integral curve $t\mapsto \mathrm{Fl}^X_t(m)$ belongs to $\Gamma_m$. Consequently, $\Delta(m)$ must contain $\left.\frac{d}{d t}\right|_0\mathrm{Fl}^X_t(m) = X(m)$. Moreover, for $X$, $Y\in D$ the curve $$ t \mapsto \mathrm{Fl}^Y_{-\sqrt{t}}(\mathrm{Fl}^X_{-\sqrt{t}}(\mathrm{Fl}^Y_{\sqrt{t}}(\mathrm{Fl}^X_{\sqrt{t}}(m)))) $$ belongs to $\Gamma_m$. By \ref{eduard}, the derivative of this curve at $t=0$ is $[X,Y](m)$, which therefore must also lie in $\Delta(m)$. Iterating this procedure it follows that $\Delta_{D^*}$ must be contained in $\Delta$. There may, however, be further vectors beside those in $\Delta_{D^*}(m)$ that have to be contained in $\Delta(m)$: Let $X\in D$, fix $t'\in \mathbb R$ and set $m':=\mathrm{Fl}^X_{-{t'}}(m)$. If $\gamma$ is a smooth curve with $\gamma(0)=m'$ and $\gamma\in \Gamma_{m'}$ then the curve $\sigma:=t\mapsto \mathrm{Fl}^X_{t'}(\gamma(t))$ belongs to $\Gamma_m$. Setting $v:=\gamma'(0)$ it follows that $\sigma'(0)= T_{m'}\mathrm{Fl}^X_{t'}(v)$. We conclude that for any $v\in \Delta(m')$ we must have $T_{m'}\mathrm{Fl}^X_{t'}(v)\in \Delta(m)$, i.e., $\Delta$ must be $D$-invariant (i.e., $G_D$-invariant). These considerations suggest to define $\Delta$ as the smallest $D$-invariant distribution that contains $\Delta_{D^*}$. We shall see below that this distribution coincides with the smallest $D$-invariant distribution that contains $\Delta_{D}$, i.e., with $P_D$. The above also explains why $\Delta_{D^*}$ by itself may be too small to serve our purpose: it may not contain sufficiently many directions: one may move within the orbit of $m$ along an integral curve of some $X\in D$, catch up a new direction there and come back. Only $P_D$ is large enough to also contain these directions. \et \section{Orbit structure and integrability} Throughout this section, let $D$ be an everywhere defined subset of ${\mathfrak X}_{\mathrm{loc}}(M)$, and let $S$ be an orbit of $D$ (cf.\ \ref{trivorb}). We equip $S$ with the natural topology introduced in \ref{orbittop}, and we will use the notations introduced there. We set $$ D^\infty := \bigcup_{n\in \N_{>0}} D^n, $$ the set of all finite tuples of elements of $D$. If $\xi\in D^n$ then $U_{\xi,m}$ is open in $\mathbb R^n$ and $\rho_{\xi,m}: U_{\xi,m} \to M$, $T \mapsto \xi_T(m)$ is ${\mathcal C}^\infty$. Also, for any $m\in S$, $\rho_{\xi,m}: U_{\xi,m} \to S$ is continuous. Given $\xi\in D^\infty$, $m\in M$, and $T\in U_{\xi,m}$, we set $$ V(\xi,m,T) := T_T\rho_{\xi,m}(T_TU_{\xi,m}), $$ the image of the tangent space of $U_{\xi,m}$ at $T$ under the tangent map of $\rho_{\xi,m}$. Setting $m_0:=\xi_T(m)$, $V(\xi,m,T)$ is a linear subspace of $T_{m_0}M$. \blem\label{susslem1} Let $\xi\in D^\infty$, $m\in S$, $T\in U_{\xi,m}$, and $m_0:=\xi_T(m)$. Then $$ V(\xi,m,T)\subseteq P_D(m_0). $$ \et \pr To begin with, let $n=1$, $\xi=X$ and $T=t_0$. Then $U_{\xi,m}=\{t\in \mathbb R\mid \exists \mathrm{Fl}^X_t(m)\}$ and $T_{t_0}U_{\xi,m}=\mathbb R$. Also, $\rho_{\xi,m}(t)=\mathrm{Fl}^X_t(m)$, so setting $m':=\mathrm{Fl}^X_{t_0}(m)$ we have $T_{t_0}\rho_{\xi,m}=X(m')$ and therefore $$ V(X,m,t_0) = \text{span}(X(m')) \subseteq \Delta_D(m')\subseteq P_D(m'). $$ Suppose now that the claim is already true for $n-1$. Let $\xi\in D^n$ and $T\in U_{\xi,m}$. Then we may write $\xi=X\eta$, $T=t_0T'$ for suitable $X\in D$ and $\eta\in D^{n-1}$, $T'\in U_{\eta,m}$ and $t_0\in \mathbb R$. By definition, $V(\xi,m,T)=\text{im}(T_T\rho_{\xi,m})$. Here, we have $\rho_{\xi,m}(t_0,T')=\mathrm{Fl}^X_{t_0}(\rho_{\eta,m}(T'))$, so that \begin{equation}\label{tangsplit} \begin{split} T_{(t_0,T')}\rho_{\xi,m} &= X(\mathrm{Fl}^X_{t_0}(\rho_{\eta,m}(T'))) \oplus T_{\rho_{\eta,m}(T')}\mathrm{Fl}^X_{t_0}(T_{T'}\rho_{\eta,m})\\ &=X(\xi_T(m)) \oplus T_{\rho_{\eta,m}(T')}\mathrm{Fl}^X_{t_0}(T_{T'}\rho_{\eta,m}). \end{split} \end{equation} Therefore, $V(\xi,m,T)\subseteq \text{span}(X(\xi_T(m))) + T_{\rho_{\eta,m}(T')}\mathrm{Fl}^X_{t_0}(V(\eta,m,T'))$. By our induction assumption, $V(\eta,m,T')\subseteq P_D(\eta_{T'}(m))$, and since $P_D$ is $D$-invariant we obtain $$ T_{\rho_{\eta,m}(T')}\mathrm{Fl}^X_{t_0}(V(\eta,m,T')) \subseteq P_D(\xi_T(m)). $$ Moreover, $X(\xi_T(m))\in \Delta_D(\xi_T(m))\subseteq P_D(\xi_T(m))$, so altogether $V(\xi,m,T)\subseteq P_D(\xi_T(m))$, concluding the proof for $n$. \ep \blem\label{susslem2} Let $m_0\in S$. Then there exist $\xi\in D^\infty$, $m\in S$, and $T\in U_{\xi,m}$ such that $\xi_T(m)=m_0$ and $V(\xi,m,T)=P_D(m_0)$. \et \pr We will see that the claim follows once we establish the following two statements: \begin{itemize} \item[(i)] If $\xi,\,\eta\in D^\infty$, $m,\,m'\in S$, $T\in U_{\xi,m}$, and $T'\in U_{\eta,m'}$ are such that $\xi_T(m)=\eta_{T'}(m')=m_0$ then there exist $\sigma\in D^\infty$, $m''\in S$, and $T''\in U_{\sigma,m''}$ such that $$ V(\xi,m,T)\cup V(\eta,m',T') \subseteq V(\sigma,m'',T''). $$ \item[(ii)] There exists a finite subset $A$ of $P_D(m_0)$ that spans $P_D(m_0)$ and satisfies: for every $v\in A$ there exist $\xi\in D^\infty$, $m\in S$, and $T\in U_{\xi,m}$ such that $\xi_T(m)=m_0$ and $v\in V(\xi,m,T)$. \end{itemize} In fact, let us suppose that (i) and (ii) have already been established. By (ii), any element of $P_D(m_0)$ is a linear combination of the elements of $A=:\{v_1,\dots,v_k\}$ and there exist $\xi_i$, $m_i$, $T_i$ ($1\le i\le k$) such that $v_i\in V(\xi_i,m_i,T_i)$ and $\xi_{T_i}(m_i)=m_0$ for all $i$. By (i), then, there exists one $V(\xi,m,T)$ containing all $V(\xi_i,m_i,T_i)$, so $P_D(m_0)\subseteq V(\xi,m,T)$. Together with \ref{susslem1} this finishes the proof. It therefore only remains to show (i) and (ii). To see (i), set $m'':=m'$, $\sigma:=\xi\hat\xi\eta$, and $T'':=T(-\hat T)T'$. Then $$ \rho_{\sigma,m''}(T'') = \sigma_{T''}(m'') = \xi_T(\hat\xi_{-\hat T}(\eta_{T'}(m'))) = \eta_{T'}(m') = m_0. $$ Analogously to \eqref{tangsplit} we may therefore split the tangent map of $\rho_{\sigma,m''}$ at $T''$ into a direct sum of three maps, corresponding to differentiation with respect to $T$, $-\hat T$, and $T'$, respectively. Therefore $V(\sigma,m'',T'')=\text{im}(T_{T''}\rho_{\sigma,m''})$ contains the sum of the images of these partial maps. Since $\hat\xi_{-\hat T}\eta_{T'}(m'')=m$, the first of these maps is $T_T\rho_{\xi,m}$, so $V(\xi,m,T)=\text{im}(T_T\rho_{\xi,m})\subseteq V(\sigma,m'',T'')$. Moreover, since $\xi_T\hat\xi_{-T}$ is the identity, the third of these maps has image $\text{im}(T_{T'}\rho_{\eta,m'}) =V(\eta,m',T')$, so also $V(\eta,m',T')\subseteq V(\sigma,m'',T'')$. This proves (i). To see (ii), let $\tilde A$ be the set of all vectors $Y(m_0)$, where $Y$ is of the form $g_*X$ for some $X\in D$ and some $g\in G_D$. By \eqref{gxd}, $\tilde A$ spans $P_D(m_0)$. Since $\dim P_D(m_0)\le \dim(M)$ we may extract a finite subset $A$ from $\tilde A$ that still spans $P_D(m_0)$. Given any $v\in A$ we have $v=T_mg(w)$, where $g\in G_D$, $m\in S$, $g(m)=m_0$ and $w\in T_mM$ is of the form $w=X(m)$ for some $X\in D$. Also, since $g\in G_D$ we have $g=\xi_T$ for some $\xi\in D^\infty$ and some $T\in U_{\xi,m}$. Now set $\eta:=\xi X$ and $T':=(T,0)$. Then $\eta_{T'}(m)=\xi_T(m)=m_0$, so $\rho_{\eta,m}(T') = \eta_{T'}(m) = \xi_T\circ \mathrm{Fl}^X_0(m)$, and splitting the tangent map as above we find $$ T_{T'}\rho_{\eta,m} = T_T\rho_{\xi,m} \oplus T_m\xi_T(X(m)). $$ In particular, $v = T_m\xi_T(X(m))\in \text{im}(T_{T'}\rho_{\eta,m}) = V(\eta,m,T')$, which gives (ii). \ep \blem\label{susslem3} If a connected integral manifold $N$ of $P_D$ intersects $S$ then $N$ is an open subset of $S$ (in the topology introduced in \ref{orbittop}). \et \pr Note first that if $N$ intersects $S$ in, say, $m$ then $S$ must be non-trivial by \ref{trivorb} because $1\le \dim T_mN$ and $T_mN = P_D(m)$. Let ${\mathcal D}$ as in \eqref{gxd}, then from \ref{imponote} we know that the ${\mathcal D}$-orbits are precisely the $D$-orbits. By \eqref{gxd}, for any $m\in N$ there exist $X_1,\dots,X_p\in \mathcal D$ such that $\{X_1(m),\dots,X_p(m)\}$ form a basis for $P_D(m)=T_mN$. In particular, $p=\dim(N)$. Let \begin{equation*} \begin{split} \Phi\colon \mathbb R^p &\to N \\ (t_1,\dots,t_p) &\mapsto \mathrm{Fl}^{X_1}_{t_1}\circ\mathrm{Fl}^{X_2}_{t_2}\circ\dots\circ \mathrm{Fl}^{X_p}_{t_p}(m). \end{split} \end{equation*} Then $\Phi$ is a diffeomorphism from an open neighborhood of $0$ in $\mathbb R^p$ onto a neighborhood of $m$ in $N$: In fact, since $X_1,\dots,X_p$ are tangential to $N$ it follows from \cite[17.16]{KLie} that their restrictions to $N$ can be viewed as vector fields on $N$, and since $\partial_i\Phi(0) = X_i(m)$, the inverse function theorem gives the claim. From what was said above it follows that every point in the image of $\Phi$ lies in the same $D$-orbit as $m$. Thus any point in $N$ has a neighborhood that is contained in one orbit of $D$. Since $N$ is connected, it follows that $N$ is contained in a single $D$-orbit $S$ (given $n_1$, $n_2\in N$ we may connect them by a smooth curve $\gamma$ in $N$. Now covering $\gamma$ by neighborhoods as above it follows that the entire curve lies in the same orbit). Consequently, if $N\cap S\not=\emptyset$ for some $D$-orbit $S$, then $N\subseteq S$. It remains to show that $N$ (as a set) is open in $S$. By \ref{orbittop} we need to see that for any $m\in S$ and $\xi\in D^n$ ($n\in \N$), $\rho_{\xi,m}^{-1}(N)$ is open in $\mathbb R^n$. Thus let $T=(t_1,\dots,t_n)\in U_{\xi,m}\subseteq \mathbb R^n$ be such that $\rho_{\xi,m}(T)\in N$. Since $\xi = (X_1,\dots,X_n)\in D^n$, it follows that $$ \rho_{\xi,m}(T) = \mathrm{Fl}^{X_1}_{t_1}\circ\mathrm{Fl}^{X_2}_{t_2}\circ\dots\circ \mathrm{Fl}^{X_n}_{t_n}(m). $$ Note that $X_i\in D\subseteq \mathcal D$ for all $i$, so as in our considerations concerning $\Phi$ above we may view the restrictions to $N$ of the $X_i$ as local vector fields on $N$. Thus for $1\le i\le n$, $X_i|_N\in {\mathfrak X}_{\mathrm{loc}}(N)$. As in \eqref{xitdef} it follows from this that the maximal domain of the map $\mathbb R^n\times N\to N$, $(T,m')\mapsto \rho_{\xi,m'}(T)$ is open in $\mathbb R^n\times N$, so in particular there exists an open neighborhood of $T$ in $\mathbb R^n$ which is mapped by $\rho_{\xi,m}$ into $N$. We conclude that $\rho_{\xi,m}^{-1}(N)$ is open in $\mathbb R^n$, as claimed. \ep After these preparations we are now ready to prove the first main result of this section: \bt\label{sussmain1} Let $D\subseteq {\mathfrak X}_{\mathrm{loc}}(M)$ be everywhere defined. Then \begin{itemize} \item[(i)] If $S$ is a non-trivial orbit of $D$, then $S$, equipped with the topology from \ref{orbittop}, admits a unique ${\mathcal C}^\infty$-structure such that $S$ becomes an immersive submanifold of $M$. The dimension of $S$ equals the rank of $S$ (see \ref{orbitrank}). \item[(ii)] With the topology and differentiable structure from (i), each non-trivial orbit of $D$ is a maximal integral submanifold of $P_D$. In fact, the non-trivial orbits of $D$ are exactly the maximal integral submanifolds of $P_D$. \item[(iii)] $P_D$ has the maximal integral manifold property. \item[(iv)] $P_D$ is involutive. \end{itemize} \et \pr We note first that (iii) is immediate from (ii), and that (iv) follows from (iii) via \ref{intinv}. (i) Let $m_0\in S$. Then by \ref{susslem2} there exist $m\in S$, $\xi\in D^\infty$, and $T\in U_{\xi,m}$ such that $\xi_T(m)=m_0$ and $V(\xi,m,T)=P_D(m_0)$. It follows that $$ k:=\mathrm{rk}(S)=\dim(P_D(m_0))=\dim(V(\xi,m,T)) = \dim(\text{im}(T_T\rho_{\xi,m})), $$ so $\mathrm{rk}_T(\rho_{\xi,m})=k$. As we noted before \ref{orbitrank}, $k=\dim(P_D(p))$ for any $p\in S$. Therefore, if $T'\in U_{\xi,m}$ then since $\text{im}(T_{T'}\rho_{\xi,m})=V(\xi,m,T')\subseteq P_D(\rho_{\xi,m}(T'))$ by \ref{susslem1}, it follows that $\mathrm{rk}(T_{T'}\rho_{\xi,m})\le k$, i.e., the rank of $\rho_{\xi,m}$ cannot exceed $k$ anywhere in $U_{\xi,m}$. On the other hand, the rank of $\rho_{\xi,m}$ locally can only increase, so by the rank theorem (\cite[1.1.3]{DG2}) there exist open neighborhoods $U$ of $T$ in $\mathbb R^n$ and $V$ of $m_0$ in $M$ and diffeomorphisms $\varphi: U \to \varphi(U)\subseteq \mathbb R^n$ (where $n$ is such that $\xi\in D^n$), $\varphi(T)=0$ $\psi: V\to \psi(V)\subseteq \mathbb R^l$ (with $l=\dim(M)$), $\psi(m_0)=0$, such that the following diagram commutes: \begin{equation}\label{rhodia} \begin{CD} U @>\rho_{\xi,m}>> V \\ @V\varphi VV @VV \psi V \\ (-1,1)^n @>i_{n,l,k}>> (-1,1)^l \end{CD} \end{equation} Here, $i_{n,l,k}$ is the map $$ (x_1,\dots,x_n)\mapsto (x_1,\dots,x_k,\underbrace{0,\dots,0}_{l-k}). $$ Denote by $N$ the (regular) submanifold $\psi^{-1}(i_{n,l,k}((-1,1)^n))$ of $M$. Then, as a set, $N = \rho_{\xi,m}(U)$. If $T'\in U$ and $m':=\rho_{\xi,m}(T')$, then since \eqref{rhodia} commutes we obtain $$ T_{m'}N = T\psi^{-1}(i_{n,l,k}(T_{T'}\varphi(\mathbb R^n))) = T_{T'}\rho_{\xi,m}(\mathbb R^n) = V(\xi,m,T'). $$ By \ref{susslem1}, $V(\xi,m,T')\subseteq P_D(m')$ and since both spaces have dimension $k$ they must be equal. Thus $N$ is an integral manifold of $P_D$. By definition of $\rho_{\xi,m}$, $N\subseteq S$, so \ref{susslem3} shows that, as a set, $N$ is open in $S$. In the same way, \ref{susslem3} applies to any open connected (in the natural manifold topology of $N$) subset of $N$. As these sets form a basis of the manifold topology of $N$, the natural inclusion map $I:N\hookrightarrow S$ is an open map. Moreover, due to \eqref{rhodia} $I$ can be decomposed as \[ \begin{CD} U @>\rho_{\xi,m}>> \rho_{\xi,m}(U)=N @>I'>> V\\ @V\varphi VV & & @VV \psi V \\ (-1,1)^n & @< {\hphantom{\text{xxxxxx}}} i_{l,n,k} {\hphantom{\text{xxxxxx}}} << & (-1,1)^l \end{CD} \] Here, $I'$ is the inclusion map and (reversing the roles of $l$ and $n$ above) $$ i_{l,n,k} = (x_1,\dots,x_l) \mapsto (x_1,\dots,x_k,\underbrace{0,\dots,0}_{n-k}) $$ (Note that $k\le \min(l,n)$ by the above). Hence $$ I = \rho_{\xi,m}\circ \varphi^{-1}\circ i_{l,n,k} \circ \psi \circ I' $$ By definition of the topology on $S$, $\rho_{\xi,m}$ is continuous as a map into $S$, so also $I$ is. We conclude that $I$ is a homeomorphism onto its image and that this image is open in $S$. Denote by ${\mathcal N}$ the family of all manifolds $N$ constructed as above. Also, let $\text{set}(N)$ denote the underlying set of the manifold $N$. By the above, $\{\text{set}(N)\mid N\in {\mathcal N}\}$ is an open cover of $S$ and each $I:N\hookrightarrow S$ is a homeomorphism onto its image. This provides a family of differentiable structures on the elements of an open cover of $S$. Our aim is to define a differentiable structure on $S$ such that each $N\in {\mathcal N}$ becomes an open submanifold of $S$. To this end it suffices to show that for any $N_1,N_2\in {\mathcal N}$ the differentiable structures of $\text{set}(N_1)\cap \text{set}(N_2)$ as an open submanifold of $N_1$ resp.\ $N_2$ coincide. Call these manifolds $W_1$, $W_2$ and let $j:W_1\to W_2$ be the identity. By symmetry we only have to show that $j$ is ${\mathcal C}^\infty$. Since both $i_1: W_1\hookrightarrow M$ and $i_2: W_2\hookrightarrow M$ are immersions, by \cite[1.1.8]{DG2} to see this it suffices to show that $j$ is continuous. This, however, is immediate since both $i_1$ and $i_2$ are homeomorphisms onto the same open subset of $S$ and $j=i_2^{-1}\circ i_1$. It follows that $S$ possesses a structure of a smooth manifold whose natural manifold topology is precisely the topology from \ref{orbittop}. Also, since each $N$ as above is an open submanifold of $M$ it follows that $S$ itself is an immersive submanifold of $M$ which is an integral manifold of $P_D$. Suppose that $S'$ is another such smooth structure on $S$. Then $i:S\to S'$ is a homeomorphism and both $S\hookrightarrow M$ and $S'\hookrightarrow M$ are immersions, so again by \cite[1.1.8]{DG2} $i$ is a diffeomorphism. Hence the smooth structure on $S$ is unique. (ii) That $S$ is an integral manifold of $P_D$ was already shown in (i). In particular, $\dim(S)=k=\dim(P_D(m_0))$. Also, $S$ is connected by \ref{orbittop}. Now let $R$ be any connected integral manifold of $P_D$ with $R\cap S\not = \emptyset$. By \ref{susslem3}, $\text{set}(R)$ is an open subset of $S$. As we did in the proof of (i) for $S$, we may apply \ref{susslem3} to each open connected subset of $R$ to see that the inclusion map $R\hookrightarrow S$ is an open map. Denote by $R'$ the open submanifold of $S$ with underlying set $\text{set}(R)$ and let $I:R'\to R$ be the identity map. Then by what we have just shown, $I$ is continuous. Note that both $i_R:R\hookrightarrow M$ and $i_{R'}:R'\hookrightarrow M$ are immersions, and $i_{R'}=i_R\circ I$. Therefore, again by \cite[1.1.8]{DG2} it follows that $I$ is smooth, and is in fact an immersion. Since both $R$ and $R'$, being integral manifolds of $P_D$, have the same dimension $k$, $I$ is even a local diffeomorphism and, due to its injectivity, a diffeomorphism. We conclude that $R=R'$ as a manifold, i.e., $R$ is an open submanifold of $S$. Thus $S$ is indeed a maximal integral manifold of $P_D$, as claimed. Finally, let $m\in M$ and let $R$ be a maximal integral manifold of $P_D$ containing $m$. Then by what we have just shown, $R$ is contained in the $D$-orbit $S$ of $m$. Since, conversely, $S$ is a connected integral manifold of $P_D$ and $R$ is maximal, we in fact have $R=S$. Thus the maximal integral manifolds of $P_D$ are precisely the orbits of $D$. \ep The second main result is as follows: \bt\label{sussmain2} Let $\Delta$ be a smooth distribution on $M$ and let $D\subseteq {\mathfrak X}_{\mathrm{loc}}(M)$ span $\Delta$ (in particular, $D$ is everywhere defined). Then the following statements are equivalent: \begin{itemize} \item[(i)] $\Delta$ is integrable. \item[(ii)] $\Delta$ has the maximal integral manifold property. \item[(iii)] $\Delta$ is $D$-invariant. \item[(iv)] For every $X\in D$, $t\in \mathbb R$, and $m\in M$ such that $\mathrm{Fl}^X_t(m)$ is defined, $$ T_m \mathrm{Fl}^X_t(\Delta(m)) \subseteq \Delta(\mathrm{Fl}^X_t(m)). $$ \item[(v)] For every $m\in M$ there exists $k\ge 1$ and elements $X^1,\dots,X^k$ of $D$ such that \begin{itemize} \item[(a)] $\Delta(m) = \text{\rm span}(X^1(m),\dots,X^k(m))$, and \item[(b)] for every $X\in D$ there exists some $\varepsilon>0$ and smooth functions $f^i_j: (-\varepsilon,\varepsilon) \to \mathbb R$ ($1\le i,j\le k$) such that $$ [X,X^i](\mathrm{Fl}^X_t(m)) = \sum_{j=1}^k f^i_j(t)X^j(\mathrm{Fl}^X_t(m)) $$ for $i=1,\dots,k$ and $t\in (-\varepsilon,\varepsilon)$. \end{itemize} \item[(vi)] $\Delta = P_D$. \end{itemize} \et \pr Clearly, (iv)$\Rightarrow$(iii)$\Rightarrow$(vi), and (ii)$\Rightarrow$(i). Also, (vi)$\Rightarrow$(ii) follows from \ref{sussmain1}. (i)$\Rightarrow$(v): Let $m\in M$. If $P_D(m)=0$ then we may pick any $X^1\in D$. Then $X^1(m)=0$, and (v) (a) holds trivially. Next, let $X\in D$ and set $f^1_1\equiv 0$. By (i) also $[X,X^i]\in D$, so $[X,X^i](m)\in \Delta(m)=\{0\}$. Also, $X(m)=0$, so $\mathrm{Fl}^X_t(m)=m$ for all $t$ and also (v) (b) is satisfied trivially. So let $P_D(m)\not=0$ and let $S$ be an integral manifold of $\Delta$ through $m$. Also, let $X^1,\dots,X^k$ be elements of $D$ such that $\{X^1(m),\dots,X^k(m)\}$ is a basis of $\Delta(m)=T_mS$. Then $k=\dim(S)$ and the restriction of the $X^i$ to $S$ can be viewed as elements of ${\mathfrak X}_{\mathrm{loc}}(S)$. By continuity of the determinant it follows that there exists a neighborhood $U$ of $m$ in $S$ such that $\{X^1(m'),\dots,X^k(m')\}$ is a basis of $T_{m'}S=\Delta(m')$ for all $m'\in U$. If $X\in D$ then $X$ is tangent to $S$, hence also restricts to a local vector field on $S$. It follows that also all $[X,X^i]$ are tangent to $S$, and so they must be smooth linear combinations of the $X^i$ on $U$. This implies (v) once we pick $\varepsilon>0$ so small that the curve $t\mapsto \mathrm{Fl}^X_t(m)$ remains in $U$. (v)$\Rightarrow$(iv): By the flow property, it suffices to show (iv) for $t\in (-\varepsilon,\varepsilon)$. For any such $t$, let $W^i(t):=(\mathrm{Fl}^X_t)^*(X^i)(m)$. Then by \cite[17.7]{KLie} we get $$ \frac{dW^i}{dt} = (\mathrm{Fl}^X_t)^*(L_XX^i)(m) = (\mathrm{Fl}^X_t)^*([X,X^i])(m) = \sum_{j=1}^k f^i_j(t)(\mathrm{Fl}^X_t)^*(X^j)(m). $$ Hence $W^1,\dots,W^k$ satisfy the system of linear ODE $$ \frac{dW^i}{dt} = \sum_{j=1}^k f^i_j(t) W^j(t) $$ on $(-\varepsilon,\varepsilon)$. Since $\{W^1(0),\dots,W^k(0)\}$ forms a basis of $\Delta(m)$, the same is therefore true for $\{W^1(t),\dots,W^k(t)\}$ for any $t\in (-\varepsilon,\varepsilon)$. Now since $X^i\in D$, $$ T_m\mathrm{Fl}^X_t(W^i(t)) = X^i(\mathrm{Fl}^X_t(m)) \in \Delta(\mathrm{Fl}^X_t(m)), $$ so indeed $T_m\mathrm{Fl}^X_t(\Delta(m))\subseteq \Delta(\mathrm{Fl}^X_t(m))$. \ep Next, we want to connect the above results to the approach taken in \cite{M}. To do this, we need to introduce some notions first. \begin{thr} {\bf Definition. } $D\subseteq {\mathfrak X}_{\mathrm{loc}}(M)$ is called stable\index{stable} if for all $X,\, Y\in D$ we have $(\mathrm{Fl}^X_t)^*Y\in D$, for all $t$ such that this expression is defined. A local vector field $X$ on $M$ is called an {\em infinitesimal automorphism}\index{infinitesimal automorphism} of a distribution $\Delta$ if $T_m\mathrm{Fl}^X_t(\Delta(m))\subseteq \Delta(\mathrm{Fl}^X_t(m))$ whenever defined. The set of infinitesimal automorphisms of $\Delta$ is denoted by $\text{\rm aut}(\Delta)$. \et For any $D\subseteq {\mathfrak X}_{\mathrm{loc}}(M)$, let ${\mathcal D}\subseteq {\mathfrak X}_{\mathrm{loc}}(M)$ be as in \eqref{gxd} (for $G=G_D$). \blem \label{michcomp} Let $D\subseteq {\mathfrak X}_{\mathrm{loc}}(M)$. Then \begin{itemize} \item[(i)] ${\mathcal D}$ is the smallest stable subset of ${\mathfrak X}_{\mathrm{loc}}(M)$ containing $D$. \item[(ii)] $D$ is stable if and only if ${\mathcal D}=D$. \item[(iii)] If $D$ is stable then $P_D=\Delta_D$. \end{itemize} \et \pr (i) By \eqref{gxd}, $D\subseteq{\mathcal D}$. Moreover, ${\mathcal D}$ is stable: if $g_*X$, $h_*Y\in {\mathcal D}$ (for $g=\xi_T, h=\eta_{T'}\in G_D$ and $X, Y\in D$) then by \ref{imponote} we obtain $$ (\mathrm{Fl}^{g_*X}_t)^*(h_*Y) = (\xi_T\circ \mathrm{Fl}^X_t\circ \xi_{-\hat T})^*(h_*Y)= (\eta_{-\hat T'}\circ \xi_T\circ \mathrm{Fl}^X_t\circ \xi_{-\hat T})^*Y \in {\mathcal D}. $$ On the other hand, if $D'\supseteq D$ is stable then given $X\in D$ and $g=\xi_T \in G_D$ for $T=(t_1,\dots,t_n)$ then $(\mathrm{Fl}^{X_n}_{-t_n})^*X\in D'$, so $(\mathrm{Fl}^{X_{n-1}}_{-t_{n-1}})^*(\mathrm{Fl}^{X_n}_{-t_n})^*X\in D'$, etc., leading to $ g_*X = (\mathrm{Fl}^{X_1}_{-t_1})^*\dots (\mathrm{Fl}^{X_n}_{-t_n})^*X\in D', $ i.e., ${\mathcal D}\subseteq D'$. (ii) is immediate from (i), and (iii) follows since $\Delta_{\mathcal D}=P_D$ by \eqref{gxd}. \ep After these preparations, we have (see \cite[3.24]{M}): \bt Let $M$ be a smooth manifold and let $\Delta$ be a smooth distribution on $M$. Then the following statements are equivalent: \begin{itemize} \item[(i)] $\Delta$ is integrable. \item[(ii)] $D_\Delta$ is stable. \item[(iii)] There exists a subset $D\subseteq {\mathfrak X}_{\mathrm{loc}}(M)$ such that ${\mathcal D}$ spans $\Delta$ (i.e., s.t.\ $P_D=\Delta$). \item[(iv)] $\text{\rm aut}(\Delta)\cap D_\Delta$ spans $\Delta$. \end{itemize} \et \pr (i)$\Leftrightarrow$(iii): this is \ref{sussmain2}, (i)$\Leftrightarrow$(vi), applied to $D:=D_\Delta$. (i)$\Rightarrow$(ii): Let $X, Y\in D_\Delta$ and $m\in M$. Then $Y(\mathrm{Fl}^X_t(m))\in \Delta(\mathrm{Fl}^X_t(m))$, and so \ref{sussmain2} (i)$\Rightarrow$(iv), applied to $D=D_\Delta$ gives $$ (\mathrm{Fl}^X_t)^*Y(m) = T\mathrm{Fl}^X_{-t}(Y(\mathrm{Fl}^X_t(m)))\subseteq \Delta(m). $$ Thus $(\mathrm{Fl}^X_t)^*Y\in D_\Delta$. (ii)$\Rightarrow$(iii): Set $D:=D_\Delta$, then the claim follows from \ref{michcomp} (ii). (i)$\Rightarrow$(iv): By \ref{sussmain2}, (i)$\Rightarrow$(iv), applied to $D_\Delta$, it follows that $D_\Delta\subseteq \mathrm{aut}(\Delta)$. Thus $\mathrm{aut}(\Delta)\cap D_\Delta = D_\Delta$, which spans $\Delta$ by definition. (iv)$\Rightarrow$(iii): Set $D:=\text{\rm aut}(\Delta)\cap D_\Delta$. Then $D$ spans $\Delta$ by assumption, and for $X\in D$ and $Y\in D_\Delta$ we have $T\mathrm{Fl}^X_{-t}(Y(\mathrm{Fl}^X_t(m)))\in \Delta(m)$ since $X\in \mathrm{aut}(\Delta)$, so $(\mathrm{Fl}^X_t)^*Y\in D_\Delta$. Iterating this argument it follows that for $n\in \N$, $X_i, Y\in D$ ($1\le i\le n$), $\xi=(X_1,\dots,X_n)$ and $T\in \mathbb R^n$, $(\xi_T)_*Y\in D_\Delta$. Therefore, $D\subseteq {\mathcal D}\subseteq D_\Delta$, so ${\mathcal D}$ spans $\Delta$ and the claim follows. \ep Next, following \cite[3.25]{M}, we want to analyze the local structure of the integral manifolds of an integrable distribution. For this we first introduce an important class of immersive submanifolds: \begin{thr} {\bf Definition. }\label{inimandef} Let $A$ be any subset of $M^l$, and for $m\in A$ denote by $C_m(A)$ the set of all points in $A$ that can be joined to $m$ by a smooth curve\footnote{By a smooth curve here we simply mean a ${\mathcal C}^\infty$-map from some interval into $M$.} in $M$ lying in $A$. A subset $N$ of $M$ is called an {\em initial submanifold}\index{initial submanifold} of $M$ of dimension $n$, if for each $m\in N$ there exists a chart $(U_m,\varphi_m)$ of $M$ centered at $m$ such that \begin{equation}\label{initman} \varphi_m(C_m(U_m\cap N)) = \varphi_m(U_m)\cap (\mathbb R^n\times \{0\}) \subseteq \mathbb R^n\times \mathbb R^{l-n}. \end{equation} \et In order to see that calling such subsets submanifolds is justified, we need an auxiliary result: \blem\label{walschap} Any piecewise smooth curve $c:[a,b]\to M$ admits a reparametrization as a $C^\infty$-curve. \et \pr Let $a=t_0<t_1<\dots <t_k=b$ be such that each curve $c|_{[t_{i-1},t_i]}$ is smooth. For each $i$ pick a smooth map $\phi_i:[a,b]\to \mathbb R$ such that $\phi_i(t)=0$ for $t\le t_{i-1}$, $\phi_i(t)=1$ for $t\ge t_{i}$, and $\phi$ is strictly increasing on $[t_{i-1},t_i]$. Then the map $$ \phi := t_0 + \sum_{i=1}^k (t_i-t_{i-1})\phi_i $$ is smooth, strictly increasing on $[a,b]$, $\phi([a,b])=[a,b]$, $\phi(t_i)=t_i$, and $\phi^{(m)}(t_i)=0$ for all $i$ and all $m\ge 1$. Thus $c\circ\phi$ is the desired reparametrization. \ep \blem\label{openlem} Under the assumptions of \ref{inimandef}, let $m_1,m_2\in N$ and $U_{m_1}$, $U_{m_2}$ be as in \eqref{initman}. Then $\varphi_{m_1}(C_{m_1}(U_{m_1}\cap N)\cap C_{m_2}(U_{m_2}\cap N))$ is open in $\mathbb R^n\times \{0\}$. \et \pr We may suppose that the intersection is nonempty, so let $p\in C_{m_1}(U_{m_1}\cap N)\cap C_{m_2}(U_{m_2}\cap N)$. Then $$ W:=C_p(U_{m_1}\cap U_{m_2}\cap N)\subseteq C_{m_1}(U_{m_1}\cap N)\cap C_{m_2}(U_{m_2}\cap N): $$ In fact, there is a smooth curve $c_1$ from $m_1$ to $p$ in $U_{m_1}\cap N$ and if $q\in W$ then there is also a smooth curve $c_2$ from $p$ to $q$ in $U_{m_1}\cap U_{m_2}\cap N$. The concatenation of $c_1$ and $c_2$ can be reparametrized smoothly by \ref{walschap}, so $q\in C_{m_1}(U_{m_1}\cap N)$, and analogously for $m_2$. We claim that \begin{equation}\label{dings} \varphi_{m_1}(W) = C_{\varphi_{m_1}(p)}(\varphi_{m_1}(U_{m_1}\cap U_{m_2})\cap (\mathbb R^n\times \{0\})). \end{equation} $\subseteq$: $\varphi_{m_1}(W)\subseteq \varphi_{m_1}(C_{m_1}(U_{m_1}\cap N))= \varphi_{m_1}(U_{m_1})\cap (\mathbb R^n\times \{0\})$ and $\varphi_{m_1}(W)\subseteq \varphi_{m_1}(U_{m_1}\cap U_{m_2})$. Moreover, any $q\in W$ is connected to $p$ within $W$ by a smooth curve $c$, and so $\varphi_{m_1}\circ c$ connects the corresponding images. $\supseteq$: Let $\tilde c$ be a smooth curve from $\varphi_{m_1}(p)$ to $\varphi_{m_1}(q)$ in $\varphi_{m_1}(U_{m_1}\cap U_{m_2})\cap (\mathbb R^n\times \{0\})$. Then $c:=\varphi_{m_1}^{-1}\circ \tilde c$ is a smooth curve in $U_{m_1}\cap U_{m_2}\cap N$ from $p$ to $q$. The right hand side of \eqref{dings} is precisely the connected component of $\varphi_{m_1}(p)$ in $\varphi_{m_1}(U_{m_1}\cap U_{m_2})\cap (\mathbb R^n\times \{0\})$ in $\mathbb R^n\times \{0\}$, hence is open. Since $p$ was arbitrary, this shows that $\varphi_{m_1}(C_{m_1}(U_{m_1}\cap N)\cap C_{m_2}(U_{m_1}\cap N))$ is open, as claimed. \ep \bt\label{initmanuniv} Let $N$ be an initial submanifold of dimension $n$ of $M^l$. Then there is a unique ${\mathcal C}^\infty$-structure on $N$ such that the inclusion map $i:N\hookrightarrow M$ becomes an injective immersion and such that the following universal property holds: For any manifold $R^k$ and any map $f\colon R\to N$, $f$ is smooth if and only if $i\circ f\colon R\to M$ is smooth. Moreover, $N$ is paracompact. \et \pr For any $m\in N$, by \ref{inimandef} there exists a chart $(U_m,\varphi_m)$ of $M$ around $m$ such that $\varphi_m(C_m(U_m\cap N)) = \varphi_m(U_m)\cap (\mathbb R^n\times \{0\})$. We define a chart for $N$ at $m$ by $(C_m(U_m\cap N), \psi_m:=\varphi_m|_{C_m(U_m\cap N)})$. Then the chart transition functions $\psi_{m_1}\circ\psi_{m_2}^{-1}$ are the restrictions of the smooth transition functions $\varphi_{m_1}\circ\varphi_{m_2}^{-1}$ of $M$ to subsets of $\mathbb R^n\times \{0\}$ that are open by \ref{openlem}, hence we obtain a differentiable structure on $N$. From this choice of charts it immediately follows that $i$ is an immersion. Thus $N$ is an immersive submanifold of $M$. In particular, the natural manifold topology of $N$ is finer than the trace topology of $M$ on $N$. To show (the non-trivial part of) the universal property, let $f\colon R\to N$ be such that $i\circ f\colon R\to M$ is smooth, let $r\in R$ and choose a chart $(U,\varphi)$ at $f(r)$ in $M$ such that $\varphi(C_{f(r)}(U\cap N)) = \varphi(U)\cap (\mathbb R^n\times \{0\})$. Then since $f^{-1}(U)$ is open in $R$ we may pick a chart $(V,\psi)$ in $R$ at $r$ with $V\subseteq f^{-1}(U)$ such that $\psi(V)$ is a ball in $\mathbb R^k$. Then since $V$ is ${\mathcal C}^\infty$-contractible and $i\circ f$ is smooth it follows that any point in $f(V)$ can be connected by a smooth curve in $M$ that lies entirely in $U\cap N$ with $f(r)$. Therefore, $f(V)\subseteq C_{f(r)}(U\cap N)$, and we can write $$ (\varphi|_{C_{f(r)}(U\cap N)})\circ f\circ \psi^{-1} = \varphi\circ f\circ \psi^{-1}. $$ Here, the right hand side is smooth by assumption and the left hand side is the chart representation of $f$ around $r$, as a map from $R$ to $N$, so indeed $f\colon R\to N$ is ${\mathcal C}^\infty$. To see uniqueness, denote by $N'$ another differentiable structure on $N$ with the universal property. Then $\mathrm{id}: N\to M$ and $\mathrm{id}: N'\to M$ are smooth, hence so are $\mathrm{id}: N\to N'$ and $\mathrm{id}: N'\to N$, i.e., the smooth structures in fact coincide. Concerning the paracompactness of $N$, note that any connected component of $N$ is contained in a connected component of $M$, hence is second countable by \cite[14.9]{KLie}. Alternatively, $M$ can be equipped with a Riemannian metric $g$ (simply by gluing local Riemannian metrics on charts via a partition of unity) and this induces a Riemannian metric $i^*g$ on $N$. The claim then follows from \cite[Satz 1.1.2]{H}. \ep Next we show that also a converse of the previous theorem holds. \bt\label{imfc} Let $f\colon M^k \to N^l$ be an injective immersion between manifolds which has the universal property from \ref{initmanuniv}, i.e.: if $h\colon P\to N$ is smooth and $h(P)\subseteq f(M)$, then the induced map $\bar h\colon P\to M$ with $f\circ \bar h = h$ is smooth. Then $f(M)$ is an initial submanifold of $N$. \et \pr By \cite[1.1.3]{DG2}, given any $m\in M$, there exist charts $(\varphi,W)$ centered at $m$ in $M$ and $(\tilde\psi,V)$ centered at $f(m)$ in $N$ such that $$ (\tilde\psi\circ f\circ\varphi^{-1}) = i:=(x^1,\dots,x^k) \mapsto (x^1,\dots,x^k,0,\dots,0)\in \mathbb R^l. $$ We may assume that $W = f^{-1}(V)$, so $i:\varphi(W)\to \tilde\psi(V)$. Pick $r>0$ so small that $\{x\in \mathbb R^k\mid |x|<2r\}\subseteq\varphi(W)$ and $\{y\in \mathbb R^l\mid |y|<2r\}\subseteq\tilde\psi(V)$. Define $B^k_r(0)$ resp.\ $B^l_r(0)$ to be the open ball of radius $r$ in $\mathbb R^k$ resp.\ $\mathbb R^l$, and set \begin{equation*} U := \tilde\psi^{-1}(B^l_r(0))\subseteq N,\quad W_1 := \varphi^{-1}(B^k_r(0))\subseteq M \end{equation*} We show that $(\psi:=\tilde\psi|_U,U)$ satisfies \eqref{initman}. In fact, since \begin{equation*} \begin{split} i(i^{-1}(\{(y^1,\dots,y^k,0,\dots,0)\mid |y|<r\})) &= \{(y^1,\dots,y^k,0,\dots,0)\mid |y|<r\}\cap i(\varphi(W))\\ &= \{(y^1,\dots,y^k,0,\dots,0)\mid |y|<r\}, \end{split} \end{equation*} we have \begin{equation*} \begin{split} &\psi^{-1}(\psi(U)\cap (\mathbb R^k\times \{0\})) = \psi^{-1}(\{(y^1,\dots,y^k,0,\dots,0)\mid |y|<r\})\\ &\hphantom{xxx}=f\circ\varphi^{-1}( (\psi\circ f\circ\varphi^{-1})^{-1}(\{(y^1,\dots,y^k,0,\dots,0)\mid |y|<r\}))\\ &\hphantom{xxx}= f\circ\varphi^{-1}(B^k_r(0))=f(W_1). \end{split} \end{equation*} Now $\tilde\psi\circ f(W_1)=i(B^k_r(0))\subseteq B^l_r(0)$, so $f(W_1)\subseteq U\cap f(M)$. Also, $f(W_1)$ is ${\mathcal C}^\infty$-contractible, so altogether we get $$ \psi^{-1}(\psi(U)\cap (\mathbb R^k\times \{0\}))= f(W_1) \subseteq C_{f(m)}(U\cap f(M)). $$ Conversely, let $n\in C_{f(m)}(U\cap f(M))$. This means that there exists a smooth curve $c:[0,1]\to N$ with $c(0)=f(m)$, $c(1)=n$, and $c([0,1])\subseteq U\cap f(M)$. By the universal property of $f$ it follows that the unique curve $\bar c:[0,1]\to M$ with $c=f\circ\bar c$ is smooth. We show that $\bar c([0,1])\subseteq W_1$. Since $\bar c([0,1])\subseteq f^{-1}(U)\subseteq f^{-1}(V)=W$ and $\bar c(0) = m$ (since $f$ is injective), it follows that $\varphi(\bar c(0))=0$. Thus if $\bar c([0,1])\not\subseteq W_1$, by continuity of $\bar c$ there must exist some $t\in(0,1]$ where $\varphi\circ \bar c$ intersects $\partial B^k_r(0)$, i.e., with $\bar c(t)\in \varphi^{-1}(\partial B^k_r(0))$. But then \begin{equation*} \begin{split} \psi\circ f(\bar c (t)) \in \psi\circ f\circ\varphi^{-1}(\partial B^k_r(0))=i(\partial B^k_r(0))\subseteq \partial B^l_r(0), \end{split} \end{equation*} and so $\psi\circ c(t)=\psi\circ f\circ \bar c (t)\not\in B^l_r(0)$, i.e., $c(t)\not\in U$, a contradiction. We conclude that $\bar c([0,1])\subseteq W_1$, and therefore $n=f(\bar c(1))\in f(W_1)$. Altogether, $$ C_{f(m)}(U\cap f(M)) = f(W_1) = \psi^{-1}(\psi(U)\cap (\mathbb R^k\times \{0\})), $$ which shows that $f(M)$ is an initial submanifold of $N$. \ep The previous result in particular applies to the situation where $f=j:M\hookrightarrow N$, i.e., where $M$ is an immersive submanifold with the universal property. In fact, the setup of \ref{imfc} is only seemingly more general: if $f\colon M\to N$ is an injective immersion, then by transporting the manifold structure of $M$ to $f(M)$ via $f$, i.e., by declaring $f\colon M\to f(M)$ to be a diffeomorphism it follows that $j:f(M)\hookrightarrow N$ becomes an immersion, hence $f(M)$ turns into an immersive submanifold of $N$, and the universal property from \ref{imfc} translates into the one from \ref{initmanuniv}. We now return to the study of integrable distributions. \bt\label{locfol} Let $\Delta$ be an integrable distribution on $M^l$. Let $S$ be a non-trivial orbit of $D_\Delta$ and let $m\in S$. Then there exists a cubic chart $(U,\varphi=(x_1,\dots,x_l))$ centered at $m$, $\varphi(U)=(-\varepsilon,\varepsilon)^l$ for some $\varepsilon>0$, some $k\ge 1$ and a countable set $A\subseteq \mathbb R^{l-k}$ such that $$ \varphi(U\cap S) = \{x\in \varphi(U)\mid (x^{k+1},\dots,x^l)\in A\}. $$ If the distribution is of constant rank $k$ then the above holds for every non-trivial orbit intersecting $U$, with the same $k$. Moreover, each non-trivial orbit is an initial submanifold of $M$. \et \pr Let $k:=\dim(S)$ and pick $X_1,\dots,X_k\in D_\Delta$ such that $\{X_1(m),\dots,X_k(m)\}$ is a basis of $\Delta(m)$. Next, choose a chart $(\chi=(y^1,\dots,y^l),W)$ around $m$ in $M$ such that $X_1(m),\dots,X_k(m),\pdl{}{y^{k+1}}{m},\dots,\pdl{}{y^l}{m}$ is a basis of $T_mM$. Let $$ f(t^1,\dots,t^l):=(\mathrm{Fl}^{X_1}_{t^1}\circ\dots \circ \mathrm{Fl}^{X_k}_{t^k})(\chi^{-1}(0,\dots,0,t^{k+1},\dots,t^l)). $$ Then $f$ is a diffeomorphism from some neighborhood of $0\in \mathbb R^l$ onto a neighborhood of $m$ in $M$, and we take $\varphi:=f^{-1}$ on a suitable neighborhood $U$ of $m$, for which we may suppose that $\varphi(U)$ is a cube $(-\varepsilon,\varepsilon)^l$ with center $0=\varphi(m)$. Since $S$ is an orbit of $D_\Delta$, $$ m'\in S \Leftrightarrow \mathrm{Fl}^{X_1}_{t^1}\circ\dots \circ \mathrm{Fl}^{X_k}_{t^k}(m')\in S $$ for all $m'$ and $t^1,\dots,t^k$ where the right hand side is defined. Therefore, for any $m'=f(t^1,\dots,t^l)\in U$ we have \begin{equation}\label{locstru} m'=f(t^1,\dots,t^l)\in S \Leftrightarrow f(0,\dots,0,t^{k+1},\dots,t^l)\in S. \end{equation} This means that $U\cap S$ is the disjoint union of connected sets of the form $$ U_c:=\{m'\in U\mid x^{k+1}(m')=c_{k+1},\dots,x^l(m')=c_l\} $$ where $c=(c_{k+1},\dots,c_l)$ is constant. By assumption, $\Delta$ is integrable, so \ref{sussmain2} and \ref{sussmain1} show that any orbit is a maximal integral manifold of $\Delta$. Therefore, since $$ U_c = \{(\mathrm{Fl}^{X_1}_{t^1}\circ\dots \circ \mathrm{Fl}^{X_k}_{t^k})(\chi^{-1}(0,c))\mid (t^1,\dots,t^k)\in (-\varepsilon,\varepsilon)^k\}, $$ and $S$ is an integral manifold of $\Delta$, the proof of \ref{susslem3} demonstrates that $U_c$ is an open (and connected) submanifold of $S$. Now $S$, being a connected immersive submanifold, is contained in a connected component $C$ of $M$, and $C$ is second countable since $M$ is paracompact. Thus by \cite[14.9]{KLie}, $S$ is itself second countable. This shows that there can at most be countably many $U_c$ as above. If $\Delta$ is of constant rank $k$ then clearly the above construction works for this same $k$ for any orbit that intersects $U$. Finally, from \eqref{locstru} it follows that $m'\in C_m(U\cap S)$ if and only if $m'\in U_0$, i.e., $$ \varphi(C_m(U\cap S)) = \varphi(U)\cap (\mathbb R^k\times \{0\}), $$ so $S$ is indeed an initial submanifold of $M$. \ep \begin{thr} {\bf Remark. }\rm\label{alt34} The previous result provides an alternative proof of \ref{classfrob}, (iii) $\Leftrightarrow$ (iv). \et \begin{thr} {\bf Definition. } A chart as in \ref{locfol} is called a {\em distinguished chart}\index{distinguished chart} for $\Delta$. The connected components of $U\cap S$ are called plaques\index{plaque} (or slices). \et \begin{thr} {\bf Definition. } Let $D\subseteq {\mathfrak X}_{\mathrm{loc}}(M)$ be everywhere defined. $D$ is said to satisfy the {\em reachability condition}\index{reachability condition} if the $D$-orbits are exactly the connected components of $M$. \et A necessary and sufficient condition for reachability is given in the following result: \bt\label{reachth} Let $D\subseteq {\mathfrak X}_{\mathrm{loc}}(M)$ be everywhere defined. Then the following are equivalent: \begin{itemize} \item[(i)] $D$ satisfies the reachability condition. \item[(ii)] For every $m\in M$ we have $\dim P_D(m) = \dim(M)$. \end{itemize} \et \pr Let $n:=\dim(M)$. (i)$\Rightarrow$(ii) Suppose that for some $m\in M$ we have $k:=\dim P_D(m) < n$. Then (by \ref{sussmain1} (i)) the orbit $S$ of $D$ through $m$ is a $k$-dimensional connected immersive submanifold of $M$. It follows that the interior of $S$ in the topology of $M$ is empty. But any connected component of $M$ is open, so $S$ cannot be such a connected component, contradicting our assumption (i). (ii)$\Rightarrow$(i) By \cite[14.1]{KLie}, every maximal integral manifold of $P_D$ is open (being an immersive submanifold of the same dimension as $M$), and connected. Also, $M$ is the disjoint union of the orbits of $D$ which, by \ref{sussmain1} (ii), are exactly the maximal integral manifolds of $P_D$. Thus these orbits are the connected components of $M$. \ep Recall from \ref{dstar} that if $D$ is any subset of ${\mathfrak X}_{\mathrm{loc}}(M)$ then the smallest involutive subset of ${\mathfrak X}_{\mathrm{loc}}(M)$ that contains $D$ is denoted by $D^*$. \begin{thr} {\bf Corollary. } With $M$, $n$, $D$ as in \ref{reachth}, if $\Delta_{D^*}$ has dimension $n$ for every $m\in M$ then $D$ satisfies the reachability condition. \et \pr By \ref{incldist}, $\Delta_{D^*}\subseteq P_D$, so $\dim(P_D(m))=n$ for all $m\in M$, and the claim follows from \ref{reachth}. \ep To conclude this chapter we show how the classical results on the integrability of constant rank distributions can be derived from the results established above. \begin{thr} {\bf Definition. }\label{lftdef} Let $D\subseteq {\mathfrak X}_{\mathrm{loc}}(M)$ be everywhere defined. $D$ is called {\em locally of finite type}\index{distribution!locally finite type} if for every $m\in M$ there exist $X^1,\dots,X^k\in D$ such that \begin{itemize} \item[(i)] $\Delta_D(m) = \text{\rm span}(X^1(m),\dots,X^k(m))$, and \item[(ii)] for every $X\in D$ there exists a neighborhood $U$ of $m$ in $M$ and smooth functions $f^i_j\in {\mathcal C}^\infty(U)$ ($1\le i,j\le k$) such that $$ [X,X^i](m') = \sum_{j=1}^k f^i_j(m')X^j(m') $$ for $i=1,\dots,k$ and all $m'\in U$. \end{itemize} \et \bt\label{lftth} If $D$ is locally of finite type then $\Delta_D$ has the maximal integral manifold property. \et \pr By assumption, $\Delta=\Delta_D$ satisfies (v), and thereby also (ii) of \ref{sussmain2}. \ep \begin{thr} {\bf Corollary. } If $D^*$ is locally of finite type then $\Delta_{D^*}=P_D$ and the non-trivial $D$-orbits are precisely the maximal integral manifolds of $\Delta_{D^*}$. \et \pr We know from \ref{incldist} that $\Delta_{D^*}\subseteq P_D$. By \ref{lftth}, $\Delta_{D^*}$ has the maximal integral manifold property, so \ref{sussmain2} shows that $\Delta_{D^*}$ is ($D^*$-, hence) $D$-invariant. Therefore by definition of $P_D$ we get $P_D\subseteq \Delta_{D^*}$, i.e., $\Delta_{D^*}=P_D$. The claim then follows from \ref{sussmain1} (ii). \ep We can now derive the following version of the classical Frobenius theorem.\index{Frobenius Theorem} \bt\label{frob2} (Frobenius) Let $\Delta$ be a smooth distribution on $M$ of constant rank $k$. Then the following are equivalent: \begin{itemize} \item[(i)] $\Delta$ has the maximal integral manifold property. \item[(ii)] $\Delta$ is involutive. \end{itemize} \et \pr (i)$\Rightarrow$(ii) This is \ref{intinv}. (ii)$\Rightarrow$(i) Let $m\in M$ and pick $X^1,\dots,X^k\in \Delta$ such that $\{X^1(m),\dots,X^k(m)\}$ is a basis of $\Delta(m)$. Then also $\{X^1(m'),\dots,X^k(m')\}$ is linearly independent for all $m'$ in some neighborhood $U$ of $m$ in $M$. Since $\dim(\Delta(m'))=k$ for all $m'$, any local vector field on $U$ that belongs to $\Delta$ is a linear combination of $X^1,\dots,X^k$ with smooth coefficients. If $X$ belongs to $\Delta$ then since $\Delta$ is involutive, $[X,X^i]\in \Delta$, and so $D_\Delta$ is locally of finite type. Since $\Delta_{D_{\Delta}}=\Delta$ by definition, (i) follows from \ref{lftth}. \ep \begin{thr} {\bf Corollary. } Let $D\subseteq {\mathfrak X}_{\mathrm{loc}}(M)$ be everywhere defined, involutive and of constant rank. Then $\Delta_D$ has the maximal integral property. \et \pr By \ref{frob2} it suffices to show that $\Delta_D$ is involutive. Let $m\in M$ and pick $X^1,\dots,X^k\in \Delta$ such that $\{X^1(m'),\dots,X^k(m')\}$ is linearly independent for all $m'$ in some neighborhood $U$ of $m$ in $M$. If $X,\,X'\in \Delta_D$ on $U$ then both $X$ and $X'$ are linear combinations of $X^1,\dots,X^k$ with smooth coefficients. But then also $[X,X']$ is such a linear combination of the $X^j$ and of brackets of $X^j$ and $X^l$, which also belong to $\Delta_D$ due to the involutivity of $D$. Thus $[X,X']\in \Delta_D$. \ep \begin{thr} {\bf Remark. }\rm Collecting some of the results proved above we obtain an independent proof of the classical Frobenius theorem \ref{classfrob}. In fact, (iii)$\Leftrightarrow$(iv) follows from \ref{alt34}. \ref{frob2} shows that (i)$\Rightarrow$(iii), and (iii)$\Rightarrow$(i) is clear. (ii)$\Rightarrow$(i) is the easy part of \cite[17.13]{KLie}. Finally, to see that (iv)$\Rightarrow$(ii), let $(\varphi,U)$ be a chart as in (iv). Then for any point $m$ in $U$, $\partial_{x^1}|_m,\dots,\partial_{x^k}|_m$ span the tangent space at $m$ of the slice $U_a$ containing $m$. But $U_a$ is an integral manifold of $\Delta$, so this tangent space equals $\Delta(m)$. \et \chapter{Symmetry groups of differential equations} \section{Local transformation groups} In Chapter \ref{ltg_chapter} we studied Lie transformation groups on differentiable manifolds. Such group actions are always defined globally ($\Phi\colon G\times M\to M$). For the applications to symmetry groups of differential equations we have in mind, the natural actions will typically not be defined globally, however. In this section, following \cite{S,O}, we therefore study local transformation groups by means of the tools developed in Chapter \ref{nonconstchap}. Throughout, we will assume $M$ to be a connected paracompact (hence Hausdorff and second countable) ${\mathcal C}^\infty$-manifold. \begin{thr} {\bf Definition. }\label{ltgdef2} A local transformation group\label{local transformation group} (or local Lie transformation group) on $M$ consists of a Lie group $G$, an open subset ${\mathcal U}$ of $G\times M$ with $\{e\}\times M \subseteq {\mathcal U}$, and a smooth map $\Phi\colon{\mathcal U}\to M$ such that \begin{itemize} \item[(i)] If $(h,m)\in {\mathcal U}$, $(g,\Phi(h,m))\in{\mathcal U}$ and $(gh,m)\in {\mathcal U}$ then $$ \Phi(g,\Phi(h,m)) = \Phi(gh,m). $$ \item[(ii)] For all $m\in M$, $\Phi(e,m)=m$. \item[(iii)] If $(g,m)\in {\mathcal U}$ then $(g^{-1},\Phi(g,m))\in {\mathcal U}$ and (by (i), (ii)) $$ \Phi(g^{-1},\Phi(g,m)) = m. $$ \end{itemize} We will often abbreviate $\Phi(g,m)$ by $g\cdot m$. We set \begin{equation*} \begin{split} {\mathcal U}_g &:= \{m\in M\mid (g,m)\in {\mathcal U}\}\quad (g\in G)\\ {\mathcal U}_m &:= \{g\in G\mid (g,m)\in {\mathcal U}\}\quad (m\in M) \end{split} \end{equation*} and \begin{equation*} \begin{split} \Phi_g\colon {\mathcal U}_g\to M,\ m\mapsto \Phi(g,m)\\ \Phi_m: {\mathcal U}_m\to M,\ g\mapsto \Phi(g,m) \end{split} \end{equation*} \et For ${\mathcal U}=G\times M$ we obtain a global transformation group as in \ref{ltgdef}. \begin{thr} {\bf Definition. }\label{olvorb} A subset $\emptyset\not=S\subseteq M$ is called an orbit\index{orbit} of the local transformation group $\Phi\colon{\mathcal U}\subseteq G\times M \to M$, or a $\Phi$-orbit, if it is a minimal $G$-invariant subset of $M$, i.e., \begin{itemize} \item[(i)] If $m\in S$, $g\in G$ and $(g,m)\in {\mathcal U}$ then $\Phi(g,m)\in S$. \item[(ii)] If $S'\subseteq S$ is another subset of $M$ satisfying (i) then either $S'=\emptyset$ or $S' = S$. \end{itemize} For any $m\in M$ we denote by $S_{\Phi,m}$ the orbit of $m$ under $\Phi$. \et \begin{thr} {\bf Remark. }\rm If $\Phi$ is a global transformation group then $$ S_{\Phi,m}=\Phi_m(G)=G\cdot m = \{\Phi(g,m)\mid g\in G\}. $$ For a local transformation group $\Phi\colon{\mathcal U}\to M$ we obtain \begin{equation}\label{orbitform} \begin{split} S_{\Phi,m}=\{& p\in M\mid \exists k\in \N,\, \exists g_i\in G\ (1\le i\le k): g_k\in {\mathcal U}_m,\,\\ &g_i\in {\mathcal U}_{g_{i+1}\cdot\dots\cdot g_k\cdot m}\ (1\le i\le k-1)\ \text{and}\ g_1\cdot\dots\cdot g_k\cdot m = p\} \end{split} \end{equation} \et In what follows, we will illustrate many of the concepts we consider in the following example: \begin{thr}{\bf Example. }\rm\label{osex} Let $M=\mathbb R^2$, $G=(\mathbb R,+)$, and $$ \Phi(\varepsilon,(x,y)):=\left(\frac{x}{1-\varepsilon x},\frac{y}{1-\varepsilon x}\right). $$ The natural domain of $\Phi$ is $$ {\mathcal U}=\{(\varepsilon,(x,y))\mid \varepsilon<\frac{1}{x}\ \text{for}\ x>0,\ \varepsilon>\frac{1}{x}\ \text{for}\ x<0,\ \varepsilon\in \mathbb R \ \text{for}\ x=0\}\subseteq \mathbb R\times \mathbb R^2, $$ which is an open subset of $G\times M$, and $\Phi$ is ${\mathcal C}^\infty$ on ${\mathcal U}$. One easily checks that $$ \Phi(\varepsilon_1,\Phi(\varepsilon_2,(x,y))) = \Phi(\varepsilon_1+\varepsilon_2,(x,y)), $$ whenever both sides are defined. Thus we obtain a local transformation group. Its orbits are the points on the $y$-axis and the straight half-rays emanating from the origin (except for the positive and negative $y$-axis). Thus they are either single points or (regular) submanifolds of $M$. Note that $\Phi$ cannot be realized as the restriction to ${\mathcal U}$ of some global Lie transformation group on $\mathbb R^2$: in fact, for any $x\not=0$ we have $|\Phi(\varepsilon,(x,y))|\to \infty$ as $\varepsilon\to \frac{1}{x}$. \et \begin{thr} {\bf Definition. }\label{blanket} A local transformation group $\Phi\colon G\times M\supseteq {\mathcal U} \to M$ is called \begin{itemize} \item[(i)] {\em connected}\index{local transformation group!connected}, if \begin{itemize} \item[(a)] $M$ and $G$ are connected. \item[(b)] ${\mathcal U}$ is connected. \item[(c)] ${\mathcal U}_m$ is connected for each $m\in M$. \end{itemize} \item[(ii)] {\em semi-regular},\index{local transformation group!semi-regular} if all orbits can be endowed with a smooth structure as immersive submanifolds of $M$ of the same dimension. \item[(iii)] {\em regular},\index{local transformation group!regular} if it is semi-regular and every $m\in M$ possesses a neighborhood basis of open sets $U$ such that for every orbit $S$ of $G$ the set $U\cap S$ is connected in $S$. \end{itemize} \et {\bf Blanket assumption:} From now on we will assume all local transformation groups to be connected in the above sense. \begin{thr} {\bf Remark. }\rm\label{regmanrem} \begin{itemize} \item[(i)] Since $S$ (being a manifold) is locally pathwise connected, a semi-regular transformation group is regular if and only if every $m\in M$ possesses a neighborhood basis of open sets $U$ such that for every orbit $S$ of $G$ the set $U\cap S$ is pathwise connected in $S$. \item[(ii)] By \cite[1.1.12]{DG2}, every orbit of a regular transformation group is a regular submanifold of $M$. \end{itemize} \et \begin{thr}{\bf Example. }\rm (i) The map $\Phi$ from \ref{osex} defines a regular transformation group on $\mathbb R^2\setminus\{(0,y)\mid y\in \mathbb R\}$. (ii) (Cf.\ \cite[Ex.\ 18.7]{KLie}) Let $M=T^2=S^1\times S^1$ be the two-dimensional torus and $G=(\mathbb R,+)$. Fixing $\omega\in \mathbb R$ and using angular coordinates $(\theta,\rho)$ on $M$ we set $$ \Phi(\varepsilon,(\theta, \rho)) := (\theta+\varepsilon,\rho+\omega\varepsilon)\quad \text{mod}\ 2\pi. $$ Then the orbits of $\Phi$ are immersive submanifolds of dimension $1$, so $G$ acts semi-regularly on $M$. If $\omega\in \mathbb Q$ then the orbits are closed curves and $\Phi$ acts regularly. However, if $\omega$ is irrational then the orbits are dense in $M$ and therefore cannot be regular submanifolds by \cite[14.1]{KLie}. So in this case $\Phi$ does not act regularly on $M$. \et Our next aim is to show that the (non-discrete) orbits of any local transformation group can naturally be endowed with the structure of an immersive (indeed even initial) submanifold of $M$. \begin{thr} {\bf Remark. }\rm\label{globorbrem} If $\Phi\colon G\times M\to M$ is a global (connected) Lie transformation group on $M$ then we have already shown this in \ref{ltgiso} and \ref{ltgisocor}: Denoting by $G_m$ the isotropy group of $m$ in $G$ we have by \ref{isotropyrem} that the map $\Psi_m:G/G_m\to G\cdot m$, $gG_m\mapsto gm$ is a bijection. If the isotropy group $G_m$ of $m$ is open then by \ref{conrem} the orbit of $m$ is the singleton $\{m\}$. Otherwise, declaring $\Psi_m$ to be a diffeomorphism we may endow $G\cdot m$ with a smooth structure as an immersive submanifold of $M$. The fact that $G\cdot m$ is in fact an initial submanifold of $M$ will follow from \ref{orbitequ} and \ref{orbcoin} below. \et Turning now to the case of local transformation groups, we need to come up with a different construction since in general $G_m$ will no longer be a subgroup of $G$ in this case: in fact, if $g_1$, $g_2\in G_m$ then in general we will not have that $(g_1^{-1}\cdot g_2,m)\in {\mathcal U}$. But this fact was used in \ref{isotropyrem} to obtain injectivity of $\Psi_m$. The route we will take to finding a smooth structure on the orbits of $G$ goes via the infinitesimal generators of the action of $\Phi$. We first note that also for a local transformation group we may introduce the definitions of $\Phi(v)$ and ${\mathscr R}(G,M)$ exactly as in \eqref{phidef}. However, ${\mathscr R}(G,M)\subseteq {\mathfrak X}_{\mathrm{loc}}(M)$ now. Also, as in \ref{psihomo} it follows that the map $\Phi\colon v\mapsto \Phi(v)$ is a Lie algebra homomorphism from ${\mathfrak g}_R$ onto ${\mathscr R}(G,M)$, the Killing algebra\index{Killing algebra} of $\Phi$. Concerning \ref{flowaction} we have to be careful about the domain ${\mathcal U}_m$ and note that for a local transformation group the vector field $\Phi(v)$ need no longer be complete: \bp\label{flowaction2} Let $\Phi\colon{\mathcal U}\to M$ be a local transformation group, and let $m\in M$ and $v\in {\mathfrak g}$. Then for all $\varepsilon$ such that $\exp(\varepsilon v)\in {\mathcal U}_m$ we have $$ \mathrm{Fl}^{\Phi(v)}_\varepsilon(m) = \Phi(\exp(\varepsilon v),m) = \exp(\varepsilon v)\cdot m. $$ In particular, \begin{equation*} \Phi(v)(m) = \left.\frac{d}{d \eps}\right|_0 \Phi(\exp(\varepsilon v),m). \end{equation*} \et \pr For $\varepsilon$ in a sufficiently small interval around $0$, $\exp(\varepsilon v)\in {\mathcal U}_m$. On any such interval (hence on the maximal such interval), we can argue as in the proof of \ref{flowaction}. \ep For any $m\in M$ we set \begin{equation*} {\mathscr R}_m(G,M):= \{\Phi(v)(m)\mid v\in {\mathfrak g}\}. \end{equation*} Any ${\mathscr R}_m(G,M)$ is a linear subspace of $T_mM$. \begin{thr}{\bf Example. }\rm\label{osex2} For the example from \ref{osex} we have \begin{equation*} \begin{split} \Phi(\partial_\varepsilon|_0)(x,y)=T_0\Phi_{(x,y)}(\partial_\varepsilon|_0)&= \partial_\varepsilon|_0\left(\frac{x}{1-\varepsilon x}\right)\partial_x + \partial_\varepsilon|_0\left(\frac{y}{1-\varepsilon x}\right)\partial_y\\ &= x^2 \partial_x + xy\partial_y =:X(x,y). \end{split} \end{equation*} Therefore, ${\mathscr R}_{(x,y)}(G,M)$ is one-dimensional if $x\not=0$ and equals $\{0\}$ for $x=0$. To verify \ref{flowaction} in this example (cf.\ \ref{rflowex}) we first calculate the flow of $X$. The integral curve $c(\varepsilon)=(x(\varepsilon),y(\varepsilon))$ of $X$ through $(x_0,y_0)$ has to satisfy the following initial value problem: \begin{equation*} \begin{split} \partial_\varepsilon x(\varepsilon) &= x(\varepsilon)^2\\ \partial_\varepsilon y(\varepsilon) &= x(\varepsilon)y(\varepsilon)\\ (x(0),y(0)) &= (x_0,y_0). \end{split} \end{equation*} Indeed we obtain $$ (x(\varepsilon),y(\varepsilon))=\left(\frac{x_0}{1-\varepsilon x_0},\frac{y_0}{1-\varepsilon x_0}\right)=\Phi(\varepsilon,(x_0,y_0)) $$ with maximal domain $(-\infty,\frac{1}{x_0})$ if $x_0>0$, $(\frac{1}{x_0},\infty)$ if $x_0<0$, and $\mathbb R$, respectively, if $x_0=0$, i.e., the maximal domain of $c$ is precisely ${\mathcal U}_{(x_0,y_0)}$. In general, one can only expect that ${\mathcal U}_{(x_0,y_0)}$ is contained in the domain of the corresponding maximal integral curve. Also, the image of $c$ is the orbit of $(x_0,y_0)$. \et \bp\label{killaldist} Let $\Phi\colon G\times M\supseteq {\mathcal U} \to M$ be a local transformation group. Then the Killing algebra ${\mathscr R}(G,M)$ of $\Phi$ spans an integrable distribution $$ \Delta_\Phi := \Delta_{{\mathscr R}(G,M)} $$ on $M$. \et \pr By definition (see \eqref{rightvf}), ${\mathscr R}(G,M)$ consists of the smooth local vector fields $\Phi(v)$ for $v\in {\mathfrak g}$, hence $\Delta_\Phi$ is a smooth distribution on $M$. Pick any basis $\{v_1,\dots,v_k\}$ of ${\mathfrak g}$ and set $X_i:=\Phi(v_i)\in {\mathscr R}(G,M)$ for $1\le i \le k$. Then by \ref{psihomo}, $\{X_1,\dots, X_k\}$ is a Lie algebra of local vector fields spanning $\Delta_\Phi$. Consequently, $\Delta_\Phi$ satisfies condition (v) of \ref{sussmain2} and thereby is integrable. \ep Note that the dimension of $\Delta_\Phi$ may vary from point to point, so we really need the theory of distributions of non-constant rank from Chapter \ref{nonconstchap}. From \ref{killaldist} it follows by \ref{sussmain1} and \ref{locfol} that the non-trivial orbits $S_{\Delta_\Phi,m}$ of $\Delta_\Phi$ are initial submanifolds of $M$. It therefore remains to show that these orbits in fact coincide with those introduced in \ref{olvorb}, i.e., that $S_{\Phi,m}=S_{\Delta_\Phi,m}$. For this we need some auxiliary results. \blem\label{schmidtlem1} Let $\Phi\colon{\mathcal U}\to M$ be a local transformation group and as in \ref{ltgdef2}, for $m\in M$ let ${\mathcal U}_m = \{g\in G\mid (g,m)\in {\mathcal U}\}\quad (m\in M)$. Then for any $g\in {\mathcal U}_m$ there exist $n\in \N$ and $v_1,\dots,v_n\in{\mathfrak g}$, $t_1,\dots,t_n\in \mathbb R_{\ge 0}$ such that \begin{itemize} \item[(i)] $g=\exp(t_1v_1)\cdot\dots\cdot\exp(t_nv_n)$ \item[(ii)] $\forall s\in [0,t_n]:\ \exp(sv_n)\in {\mathcal U}_m$ \item[(iii)] $\forall i=1,\dots,n-1\ \forall s\in [0,t_i]:\ \exp(sv_i)\exp(t_{i+1}v_{i+1}) \dots \exp(t_nv_n)\in {\mathcal U}_m$. \end{itemize} \et \pr Denote by $\mathcal W$ the set of all $g\in {\mathcal U}_m$ that satisfy (i)--(iii). Then $\mathcal W\not=\emptyset$ since $e\in \mathcal W$. To show that $\mathcal W$ is open, let $g\in \mathcal W$ and pick an absolutely convex neighborhood $V$ of $0$ in ${\mathfrak g}$ with $\exp(V)g\subseteq {\mathcal U}_m$. Then $\exp(V)g\subseteq\mathcal W$: given $h\in \exp(V)g$ we can write $$ h = \exp(v)\exp(t_1v_1)\cdot\dots\cdot\exp(t_nv_n), $$ and since $V$ is absolutely convex it follows that also $\exp(tv)\exp(t_1v_1)\cdot\dots\cdot\exp(t_nv_n) \in {\mathcal U}_m$ for all $t\in [0,1]$. But also ${\mathcal U}_m\setminus \mathcal W$ is open: suppose that $g\in {\mathcal U}_m\setminus \mathcal W$ and again choose an absolutely convex neighborhood $V$ of $0$ in ${\mathfrak g}$ with $\exp(V)g\subseteq {\mathcal U}_m$. Suppose that $(\exp(V)g)\cap \mathcal W\not=\emptyset$, then there exists some $v\in V$ such that $h:=\exp(v)g\in \mathcal W$. Since $\exp(V)g\subseteq {\mathcal U}_m$ we also have $\exp(tv)g\in {\mathcal U}_m$ for all $|t|\le 1$. It follows that $$ g_t:=\exp(t(-v))h = \exp((1-t)v)g\in {\mathcal U}_m \quad (t\in [0,1]). $$ This, however, shows that $g=g_1$ is an element of $\mathcal W$, contradicting our assumption. Therefore, $\exp(V)g\subseteq {\mathcal U}_m\setminus \mathcal W$, implying that ${\mathcal U}_m\setminus \mathcal W$ is open. Since ${\mathcal U}_m$ is connected by our blanket assumption following \ref{blanket}, ${\mathcal W}={\mathcal U}_m$. \ep \blem\label{schmidtlem2} Let $\Phi\colon{\mathcal U}\to M$ be a local transformation group, let $m\in M$, $g\in {\mathcal U}_m$ and pick $n\in \N$, $v_1,\dots,v_n\in{\mathfrak g}$, $t_1,\dots,t_n\in \mathbb R_{\ge 0}$ such that (i)--(iii) of \ref{schmidtlem1} are satisfied. Define curves $$ \gamma_l:[0,\sum_{i=1}^n t_i] \to G \quad (l=1,2) $$ for $s\in [0,t_n]$ by $$ \gamma_1(s) := \exp(sv_n)m \ \text{ and } \ \gamma_2(s) := \mathrm{Fl}^{\Phi(v_n)}_s(m), $$ and for $s\in [\sum_{j=i+1}^n t_j,\sum_{j=i}^n t_j]$ $(i=1,\dots,n-1)$ by \begin{equation*} \begin{split} \gamma_1(s) &:= \Big(\exp\Big(\Big(s-\sum_{j=i+1}^n t_j\Big) v_i\Big)\exp(t_{i+1}v_{i+1})\dots \exp(t_n v_n)\Big)\cdot m\\ \gamma_2(s) &:= \mathrm{Fl}^{\Phi(v_i)}_{s-\sum_{j=i+1}^n t_j}(\mathrm{Fl}^{\Phi(v_{i+1})}_{t_{i+1}}(\dots \mathrm{Fl}^{\Phi(v_n)}_{t_n}(m))\dots). \end{split} \end{equation*} Then $\gamma_1(s)=\gamma_2(s)$ for all $s\in [0,\sum_{i=1}^n t_i]$. \et \pr For $s\in [0,t_n]$ this is immediate from \ref{schmidtlem1} and \ref{flowaction2}. Proceeding by induction, suppose that we already know that $\gamma_1=\gamma_2$ on $[0,\sum_{j=i+1}^{n}t_j]$. Now set \begin{equation*} \begin{split} \rho_1(s) &:= (\exp(sv_i)\exp(t_{i+1}v_{i+1})\dots \exp(t_n v_n))\cdot m\\ \rho_2(s) &:= \mathrm{Fl}^{\Phi(v_i)}_{s}(\mathrm{Fl}^{\Phi(v_{i+1})}_{t_{i+1}}(\dots \mathrm{Fl}^{\Phi(v_n)}_{t_n}(m))\dots). \end{split} \end{equation*} Then it suffices to show that $\rho_1=\rho_2$ on $[0,t_i]$. We do this by showing that $\rho_1$ satisfies the same initial value problem as $\rho_2$ on $[0,t_i]$. Set $\exp(t_{i+1}v_{i+1})\dots \exp(t_n v_n) =: h$. By \cite[8.2]{KLie}, $R_h(\exp(sv_i))=\mathrm{Fl}^{R^{v_i}}_s(h)$, so $$ \frac{d}{ds}(R_h(\exp(sv_i))) = R^{v_i}_s(R_h(\exp(sv_i))). $$ Using this, together with \eqref{olver1.46.2}, we calculate \begin{equation*} \begin{split} \frac{d}{ds}\rho_1(s) &= \frac{d}{ds}\Phi_m(R_h(\exp(sv_i))) = T\Phi_m(R^{v_i}_s(R_h(\exp(sv_i))))\\ &=\Phi(v_i)(\Phi_m(R_h(\exp(sv_i)))) = \Phi(v_i)(\rho_1(s)), \end{split} \end{equation*} which is precisely the defining ODE for $\rho_2$. Also, the initial values coincide since $\rho_1(0)=\rho_2(0)$ by our inductive assumption. \ep After these preparations we may now show the equality of the two kinds of orbits we introduced above: \bp\label{orbitequ} Let $\Phi\colon G\times M\supseteq {\mathcal U} \to M$ be a local transformation group and let $m\in M$. Then $$ S_{\Phi,m}=S_{\Delta_\Phi,m}. $$ In particular, each non-trivial orbit is an initial submanifold of $M$. \et \pr Since $G$ is connected, once we choose a basis $\{v_1,\dots,v_k\}$ of ${\mathfrak g}$, any element of $G$ is a product of certain $\exp(\varepsilon_iv_i)$. Thus by \ref{schmidtlem2} it follows that $S_{\Phi,m}$ is trivial if and only if $\Phi(v_i)(m)=0$ for all $i$, i.e., if and only if $\Delta_\Phi(m)=0$. Since $\Delta_\Phi$ is integrable by \ref{killaldist}, \ref{sussmain2} (vi) gives that $\Delta_\Phi = P_{{\mathscr R}(G,M)}$ (recall that $\Delta_\Phi=\Delta_{{\mathscr R}(G,M)}$ by definition). Consequently, \ref{trivorb} shows that the trivial orbits of $S_{\Phi,m}$ and $S_{\Delta_\Phi,m}$ coincide. Turning now to the case of non-trivial orbits, let $m'\in S_{\Delta_\Phi,m}$ and let $g\in {\mathcal U}_{m'}$. By \ref{schmidtlem1} there exist $n\in \N$, $v_1,\dots,v_n\in {\mathfrak g}$ and $t_1,\dots,t_n\in \mathbb R_{\ge 0}$ such that $g=\exp(t_1v_1)\dots\exp(t_nv_n)$ and (i)-(iii) of \ref{schmidtlem1} are satisfied (with $m'$ instead of $m$). Let $\xi:=(\Phi(v_1),\dots,\Phi(v_n))$ $\in$ ${\mathfrak X}_{\mathrm{loc}}(M)^n$, and $T:=(t_1,\dots,t_n)$. Then using the notation \eqref{xitdef}, \ref{schmidtlem2} shows that $g\cdot m' = \xi_T(m')$, and by definition $\xi_T(m')$ stays in the orbit $S_{\Delta_\Phi,m}$ of $m$, hence $g\cdot m'\in S_{\Delta_\Phi,m}$. From this, starting with $m'=m$ and then continuing inductively, it follows that any $g_1\dots g_l\cdot m$ stays in $S_{\Delta_\Phi,m}$, so by \eqref{orbitform} $S_{\Phi,m}\subseteq S_{\Delta_\Phi,m}$. Since in the above consideration we started out with any $m'\in S_{\Delta_\Phi,m}$ and since the $\Phi$-orbits are disjoint by definition we have even shown that $S_{\Delta_\Phi,m}$ is the disjoint union of certain $\Phi$-orbits. To conclude the proof we show that any $\Phi$-orbit that is contained in $S_{\Delta_\Phi,m}$ is in fact an open subset of $S_{\Delta_\Phi,m}$. Indeed, once we know this then due to $S_{\Delta_\Phi,m}$ being a disjoint union of such sets it will follow that any $\Phi$-orbit contained in $S_{\Delta_\Phi,m}$ is both open and closed in it, and since $S_{\Delta_\Phi,m}$ is connected the orbits must in fact coincide. Thus let $m'\in S_{\Phi,m}\subseteq S_{\Delta_\Phi,m}$ (since we may move points within orbits this is the only case we need to consider). Choose $X_i=\Phi(v_i)\in {\mathscr R}(G,M)$ ($1\le i\le n$, for a suitable $n$ and suitable elements $v_i$ of ${\mathfrak g}$) such that $X_1,\dots,X_n$ is a basis for $\Delta_\Phi(m')$. Set $\xi:=(X_1,\dots,X_n)$. Then by the proof of \ref{sussmain1} (i) and \ref{sussmain2} (vi) we have $$ \mathrm{rk}(T_0\rho_{\xi,m'}) = \dim(\Delta_\Phi(m')) = n. $$ As in \eqref{rhodia} it follows that there exists an open neighborhood $U$ of $0$ in $\mathbb R^n$ such that $\rho_{\xi,m'}(U)$ is the domain of a coordinate chart of $S_{\Delta_\Phi,m}$. There exists an absolutely convex open neighborhood $V\subseteq U$ of $0$ in $\mathbb R^n$ such that for any $T=(t_1,\dots,t_n)\in V$ we have $$ \exp(t_1v_1)\dots\exp(t_nv_n)\in {\mathcal U}_{m'}. $$ In particular, (i)-(iii) of \ref{schmidtlem1} are satisfied. Since $\rho_{\xi,m'}(V)$ is open in $S_{\Delta_\Phi,m}$ it suffices to show that $\rho_{\xi,m'}(V)\subseteq S_{\Phi,m}$. Also, because $V$ is absolutely convex we only need to show that $\rho_{\xi,m'}(T)\in S_{\Phi,m}$ for any $T=(t_1,\dots,t_n)\in V\cap \mathbb R_{\ge 0}^n$ (otherwise replace $v_i$ by $-v_i$). For such a $T$, \ref{schmidtlem2} (together with \eqref{xitdef}) gives \begin{equation}\label{fromproof} \begin{split} \rho_{\xi,m'}(T) &= \mathrm{Fl}^{X_1}_{t_1}(\mathrm{Fl}^{X_2}_{t_2}(\dots \mathrm{Fl}^{X_n}_{t_n}(m')\dots))\\ &= \exp(t_1v_1)\cdot\dots\cdot\exp(t_nv_n)\cdot m' \in S_{\Phi,m}, \end{split} \end{equation} as desired. The final claim was already shown in the remark following \ref{killaldist}. \ep If the transformation group $G$ is in fact global then at the moment we have two ways of endowing the non-trivial orbits $S_\Phi$ of $\Phi$ with a differentiable structure: the one from \ref{globorbrem} and the one from \ref{killaldist}. The following result shows that these approaches in fact coincide. \bp\label{orbcoin} Let $\Phi\colon G\times M \to M$ be a global Lie transformation group and let $m\in M$. Then the differentiable structures on any non-trivial orbit $S_{\Phi,m}$ introduced in \ref{globorbrem} and in \ref{killaldist}, \ref{orbitequ} coincide. \et \pr For brevity, we set $S:=S_{\Phi,m}$, with the smooth structure from \ref{killaldist}, and we write $S'$ for the smooth structure on $S_{\Phi,m}$ defined in \ref{globorbrem}. We know that the inclusions $i:S\hookrightarrow M$ and $i':S'\hookrightarrow M$ are injective immersions. Since $S$ is an initial submanifold of $M$ and $i' = i\circ \text{id}_{S'\to S}$ is smooth it follows from the universal property in \ref{initmanuniv} that $\mathrm{id}:S'\to S$ is smooth. Conversely, to see that $\mathrm{id}:S\to S'$ is smooth, let $\rho_{\xi,m'}(V)$ (with $m'\in S$) be as in the proof of \ref{orbitequ}. Then $\rho_{\xi,m'}: V\to \rho_{\xi,m'}(V)$ is a diffeomorphism onto an open submanifold (containing $m'$) of $S$, so we only need to show that $\mathrm{id}_{S\to S'}\circ \rho_{\xi,m'}$ is smooth on $V$. By \ref{globorbrem}, this is the case if and only if $\Psi_{m'}^{-1}\circ \rho_{\xi,m'}$ is smooth on $V$. Now since $\Phi$ is global, it follows from \ref{flowaction} that for any $T\in V$ we have \begin{equation*} \begin{split} \Psi_{m'}^{-1}\circ \rho_{\xi,m'}(T) &= \Psi_{m'}^{-1}(\exp(t_1v_1)\cdot\dots\cdot\exp(t_nv_n)\cdot m')\\ &= \exp(t_1v_1)\cdot\dots\cdot\exp(t_nv_n)\cdot G_{m'} \end{split} \end{equation*} and this map is indeed smooth as the composition of the smooth maps $T\mapsto \exp(t_1v_1)\cdot\dots\cdot\exp(t_nv_n)$, $V\to G$, and the quotient map $\pi: G\to G/G_{m'}$. \ep \begin{thr} {\bf Remark. }\rm\label{constrk} Suppose that $\Phi\colon G\times M \supseteq {\mathcal U} \to M$ is a semi-regular local transformation group. Then since all orbits of $\Phi$ are immersive submanifolds of the same dimension, say $k$, it follows that the integrable distribution $\Delta_\Phi$ is of constant rank $k\ge 1$. If $S$ is an orbit of $\Phi$ (i.e., by \ref{orbitequ}, an orbit of $D_{\Delta_\Phi}$) then by \ref{sussmain1} (ii) $S$ is a maximal integral manifold of $\Delta_\Phi$ and thereby a leaf of $\Delta_\Phi$ (cf.\ \cite[Def.\ 17.23]{KLie}). \et Recall from \cite[Def.\ 17.27]{KLie} that a flat chart for a distribution $\Delta$ is called regular if every leaf of $\Delta$ that intersects it does so in precisely one slice. Moreover, $\Delta$ is called regular if every point of $M$ lies in the domain of a regular chart. The following result shows that for a regular local transformation group $\Phi$, $\Delta_\Phi$ is regular in this sense. \blem\label{regslice} Let $\Phi\colon G\times M \supseteq {\mathcal U} \to M$ be a regular local transformation group.\label{local transformation group!regular} Then each point of $M$ lies in the domain of a regular chart for $\Delta_\Phi$, i.e., $\Delta_\Phi$ is a regular distribution. \et \pr Let $m\in M$ and pick a cubical chart $(U,\varphi)$ as in \ref{classfrob} centered at $m$, $\varphi(U)=[-c,c]^n$. Since $G$ acts regularly on $M$, there exists some neighborhood $V\subseteq U$ of $m$ such that the intersection of each orbit $S$ of $G$ (i.e., each maximal integral manifold of $\Delta_\Phi$) with $V$ is connected in $S$. This intersection is therefore a connected integral manifold of $\Delta_\Phi$, hence by \ref{classfrob} it is contained in a single slice $U_a$ of $\varphi$. Now pick $c'\in (0,c)$ so that $U':=\varphi^{-1}((-c',c')^n)\subseteq V$. If $S$ is any orbit of $G$ that intersects $U'$ then it also intersects $V$, hence $S\cap V\subseteq U_a$ for some $a$. Therefore, $S\cap U' \subseteq U'_a$. But since $S$ is a maximal integral manifold we also have $U'_a\subseteq S$, so altogether we obtain $S\cap U'=U'_a$. \ep \begin{thr}{\bf Example. }\rm Returning once more to the example from \ref{osex}, let \begin{equation*} \begin{split} U_+&:=\{(x,y)\in \mathbb R^2\mid x>0\}\\ U_-&:=\{(x,y)\in \mathbb R^2\mid x<0\} \end{split} \end{equation*} and set \begin{equation*} \begin{split} \varphi_\pm :U_\pm &\to \mathbb R\times\mathbb R_\pm\\ (x,y)&\mapsto \Big(x,\frac{y}{x}\Big) \end{split} \end{equation*} We show that $(U_+,\varphi_+)$ is regular chart for the transformation group $\Phi$ from \ref{osex} (and analogously for $(U_-,\varphi_-)$). Note that $\varphi^{-1}(r,s)=(r,rs)$. From \ref{osex2} we know that $\Delta_\Phi$ is spanned by the vector field $X=x^2 \partial_x + xy\partial_y$. In the chart $(U_+,\varphi_+)$, with coordinates $(r,s):=\varphi_+(x,y)$ we have $$ \varphi_*(\partial_x) = \frac{\partial r}{\partial x}\partial_r + \frac{\partial s}{\partial x}\partial_s = \partial_r - \frac{s}{r}\partial_s,\ \varphi_*(\partial_y) = \frac{\partial r}{\partial y}\partial_r + \frac{\partial s}{\partial y}\partial_s = \frac{1}{r}\partial_s. $$ Therefore, $X$ has the representation \begin{equation*} \begin{split} \varphi_*X &= (\varphi^{-1})_1^2 \varphi_*(\partial_x) + (\varphi^{-1})_1 (\varphi^{-1})_2 \varphi_*(\partial_y)\\ &= r^2\varphi_*(\partial_x) + rs\varphi_*(\partial_y) = r^2\partial_r. \end{split} \end{equation*} By \ref{osex} the non-trivial orbits, i.e., the maximal integral manifolds of $\Phi$ are half-rays emanating from $(0,0)$, except for those lying on the $y$-axis. Hence each such orbit $S$ is either contained in $U_+$ or in $U_-$. If $S\subseteq U_+$ then $$ S=\{(x,y)\in \mathbb R^2\mid s=\frac{y}{x}=c \text{ and } x>0\}, $$ so $S$ is a slice of $\varphi_+$. This shows that $(U_\pm,\varphi_\pm)$ are regular charts for $\Delta_\Phi$. Nevertheless, $\Phi$ is not regular because the points on the $y$-axis are trivial orbits, i.e., not every orbit of $\Phi$ has the same dimension. Removing the $y$-axis from $M=\mathbb R^2$ we obtain a regular local transformation group. \et Finally, we have the following fundamental result on the space of orbits of a regular local transformation group: \bt\label{quotient} Let $\Phi\colon G\times M \supseteq {\mathcal U} \to M$ be a regular local transformation group on an $n$-dimensional manifold $M$ with $k$-dimensional orbits. Then the set $M/G$ of orbits of $\Phi$ can be endowed with the structure of an $(n-k)$-dimensional manifold with the following properties: \begin{itemize} \item[(i)] The quotient map $\pi: M \to M/G$, $m\mapsto S_{\Phi,m}$ is a surjective submersion. \item[(ii)] $m$ and $m'$ belong to the same orbit if and only if $\pi(m)=\pi(m')$. \item[(iii)] For any $m\in M$, $\Delta_{\Phi}(m) = \ker T_m\pi$. \end{itemize} \et \pr Surjectivity of $\pi$ and (ii) are immediate from the definition. Since the orbits of $\Phi$ are precisely the leaves of the foliation induced by $\Delta_\Phi$, which is regular by \ref{regslice}, (i) is \cite[17.29]{KLie}. For (iii), note that since $\pi$ is constant on any orbit $S$ of $\Phi$, $T_m\Phi$ must vanish on the tangent space of $S$ at $m\in S$. But $S$ is an integral manifold of $\Delta_\Phi$, so $T_mS = \Delta_\Phi(m)$. Thus $\Delta_\Phi(m)\subseteq \ker(T_m\pi)$. Finally, since $\pi$ is a submersion, $\dim\ker(T_m\pi)=k=\dim(\Delta_\Phi(m))$, so we have equality. \ep \begin{thr} {\bf Remark. }\rm\label{quotientconst} For later use we recall the construction of \cite[17.29]{KLie}: Given a regular chart $(U,\varphi=(x^1,\dots,x^n))$, let $U':=\pi(U)$. Then the map $\varphi':U'\to \mathbb R^{n-k}$, $m'\mapsto \mathrm{pr}_2(\varphi(m))$, where $m$ is any element of $\pi^{-1}(m')\cap U$ (recall that $\pi^{-1}(m')\cap U = U_a$ for some $a$ and $\mathrm{pr}_2\circ\varphi|_{U_a}\equiv a$) is a typical chart for $M/G$. The local representation of $\pi$ with respect to the standard charts then is $$ \varphi'\circ\pi\circ \varphi^{-1}=\mathrm{pr}_2 = (x^1,\dots,x^n) \mapsto (x^{k+1},\dots,x^n): $$ \begin{equation*} \begin{CD} M\supseteq U @>\pi>> U'\subseteq M/G \\ @V\varphi VV @VV \varphi' V \\ \varphi(U) @> \mathrm{pr}_2>> \varphi'(U) \end{CD} \end{equation*} \et \section{Symmetries of algebraic equations} From this point of the course onwards we will closely follow Olver's work \cite{O}. Concerning notations, we will henceforth typically denote points in a smooth manifold $M$ by $x,\, y,\dots$, since the manifolds we are interested in will mainly be subsets of spaces of independent and dependent variables of differential equations. As already announced before \ref{flowaction}, we will from now on denote the group parameter with $\varepsilon$ instead of $t$ since we will often need $t$ as a variable in a differential equation. Moreover, given a local Lie group action $\Phi$ on $M$ we will often notationally suppress the Lie algebra homomorphism $v\mapsto \Phi(v)$ from \eqref{phidef}, i.e., we will often simply write $v$ instead of $\Phi(v)\in {\mathfrak X}_{\mathrm{loc}}(M)$. Also, we will usually only write $G$ instead of $\Phi$. By a system of algebraic equations\index{algebraic equation} we mean any system of equations \begin{equation}\label{algsys} F_\nu(x) = 0,\qquad \nu=1,\dots,l, \end{equation} where $F_1,\dots,F_l$ are smooth real-valued functions on $M$. The term `algebraic' is used to distinguish this situation from the case of differential equations to be considered later on. It does {\em not}, however, restrict the form of the $F_\nu$ (e.g., to polynomials). A {\em solution}\index{algebraic equation!solution} of \eqref{algsys} is any point $x\in M$ such that $F_\nu(x) = 0$ for $\nu=1,\dots,l$. A {\em symmetry group}\index{symmetry group} of \eqref{algsys} is any local transformation group $G$ on $M$ that transforms any solution of \eqref{algsys} into another solution. More generally, we define: \begin{thr} {\bf Definition. } Let $G$ be a local transformation group on a smooth manifold $M$. A subset $S\subseteq M$ is called $G$-invariant, and $G$ is called a symmetry group of $S$ if, whenever $x\in S$ and $g\in G$ is such that $g\cdot x$ is defined, then also $g\cdot x\in S$. \et \begin{thr}{\bf Example. }\rm\label{ase1} (i) Let $M=\mathbb R^2$ and consider the one-parameter group of translations $$ G_c: (x,y) \mapsto (x+c\varepsilon,y+\varepsilon) \qquad (c\in \mathbb R). $$ \et Then any line $\{x=cy+d\}\subseteq \mathbb R^2$ is an orbit of $G_c$, hence is invariant under $G_c$. Any $G_c$-invariant subset of $M$ is a union of such orbits. (ii) Again let $M=\mathbb R^2$ and for $\alpha\in \mathbb R$ let $$ G^\alpha: (x,y) \mapsto (\lambda x, \lambda^\alpha y) \qquad (\lambda>0). $$ Then $\{(0,0)\}$, as well as the positive and negative $x$- and $y$-axes are $G^\alpha$-invariant. Hence also the entire axes are invariant, being unions of invariant sets. Moreover, $\{xy=0\}$ and $\{y=k|x|^\alpha\}$ (for $x<0$ or $x>0$) are invariant. (iii) Most of the time we will be interested in invariant sets $S$ that are subvarieties,\index{subvariety} given by the common zero set of smooth functions $F=(F_1,\dots,F_l)$: $$ S=S_F = \{x \mid F_\nu(x)=0,\, \nu=1,\dots, l\}. $$ (iv) If $S_1$, $S_2$ are invariant subsets, so are $S_1\cup S_2$ and $S_1\cap S_2$. \begin{thr} {\bf Definition. } Let $G$ be a local transformation group acting on $M$ and let $F: M\to N$ be a map into a manifold $N$. $F$ is called $G$-invariant\index{invariant function} if for all $x\in M$ and all $g\in G$ such that $g\cdot x$ is defined we have $$ F(g\cdot x) = F(x). $$ If $N=\mathbb R$ then $F$ is simply called an invariant\index{invariant} of $G$. \et Obviously, $F=(F_1,\dots,F_l):M\to \mathbb R^l$ is $G$-invariant if and only if each $F_\nu$ is an invariant of $G$. \begin{thr}{\bf Example. }\rm\label{ase2} (i) For $G_c$ the group of translations from \ref{ase1} (i), the function $$ f(x,y) := x -cy $$ is an invariant: $f(x+c\varepsilon,y+\varepsilon) = f(x,y)$. Moreover, any invariant of $G_c$ must be of the form $g(x-cy)$ for some smooth $g$. (ii) As in \ref{ase1} (ii), let $G^1:(x,y)\mapsto (\lambda x,\lambda y)$ ($\lambda>0$). Then $f(x,y):=x/y$ is an invariant for $G^1$, defined on $\{y\not=0\}$. Another invariant is $(x,y)\mapsto xy/(x^2+y^2)$, defined on $\mathbb R^2\setminus \{(0,0)\}$. There is no smooth nonconstant invariant defined on all of $\mathbb R^2$. \et \begin{thr} {\bf Remark. }\rm\label{123} If $F:M\to \mathbb R^l$ is a $G$-invariant function then every level set of $F$ is a $G$-invariant subset of $M$: if $F(x)=c$ and $g\cdot x$ is defined then also $F(g\cdot x)=F(x)=c$. However, if the zero-set $\{F(x)=0\}$ of a smooth map $F$ is $G$-invariant then $F$ itself need not be invariant. For example, $\{(x,y)\mid xy=0\}$ is invariant under $G^1$ from \ref{ase1} (ii), but $F(x,y)=xy$ is not $G$-invariant since $F(\lambda x,\lambda y)=\lambda^2xy\not=F(x,y)$. To obtain a true statement we need to take all level sets into consideration: \et \blem Let $G$ be a local transformation group acting on $M$ and let $F\in {\mathcal C}^\infty(M,\mathbb R^l)$. The following are equivalent: \begin{itemize} \item[(i)] $F$ is invariant under $G$. \item[(ii)] Every level set $\{F(x)=c\}$ ($c\in \mathbb R^l$) of $F$ is invariant under $G$. \end{itemize} \et \pr (i)$\Rightarrow$(ii): See \ref{123}. (ii)$\Rightarrow$(i): Let $x$, $g$ be such that $g\cdot x$ is defined and set $c:= F(x)$. Then also $gx\in \{y\mid F(y)=c\}$, so $F(gx)=F(x)$. \ep The following result gives a simple, linear, criterion for a function to be an invariant of a group action. It is a first typical example of Lie theoretic methods in symmetry analysis. Recall from \ref{blanket} that we always assume transformation groups to be connected. \bt \label{finvcrit} Let $G$ be a local transformation group acting on $M$ and let $f\in {\mathcal C}^\infty(M,\mathbb R)$. The following are equivalent: \begin{itemize} \item[(i)] $f$ is an invariant of $G$. \item[(ii)] For every infinitesimal generator $v$ of $G$ we have $v(f)=0$. \end{itemize} \et \pr (i)$\Rightarrow$(ii): For clarity, we de-identify again and write $\Phi(g,x)=gx$ for the action of $G$. Let $v\in {\mathfrak g}$. Then since $f$ is an invariant, given any $x\in M$, by \ref{flowaction2} we obtain for $\varepsilon$ small: $$ f(x)=f(\exp(\varepsilon v)\cdot x) \equiv f(\Phi(\exp(\varepsilon v),x)) = f(\mathrm{Fl}^{\Phi(v)}_\varepsilon(x))\equiv f(\mathrm{Fl}^{v}_\varepsilon(x)). $$ Differentiating this expression with respect to $\varepsilon$ at $\varepsilon=0$ we get $$ v(f)\equiv \Phi(v)(f) = 0. $$ (ii)$\Rightarrow$(i): Since $v(f)$ vanishes identically on $M$ we have $$ \frac{d}{d\varepsilon} f(\exp(\varepsilon v)\cdot x) = \frac{d}{d\varepsilon} f(\mathrm{Fl}^{v}_\varepsilon(x)) = v(f)(\mathrm{Fl}^{v}_\varepsilon(x)) = 0, $$ wherever defined, so $f(\exp(\varepsilon v)\cdot x) = f(x)$ wherever defined. Since $G$ is connected, by \ref{schmidtlem1} any $g\in {\mathcal U}_x$ can be written as a product of certain $\exp(\varepsilon_iv_i)$ ($v_i\in {\mathfrak g}$) satisfying (i)--(iii) from that result. Then \ref{schmidtlem2} shows that we can iterate the above argument to conclude that $f(g\cdot x)= f(x)$ for all $g\in {\mathcal U}_x$. \ep It follows that, if $\{v_1,\dots,v_r\}$ is a basis of the local Lie algebra of infinitesimal generators of $G$ (i.e., if $\{\Phi(v_1),\dots,\Phi(v_r)\}$ is a local basis of $\Delta_\Phi$) then $f$ is an invariant of $G$ (on the open set where $\{v_1,\dots,v_r\}$ is a basis) if and only if $v_k(f)=0$ for $k=1,\dots,r$. If $G$ acts effectively, then by \ref{phiiso} we may take for $\{v_1,\dots,v_r\}$ any basis of ${\mathfrak g}$. Writing $v_k=\sum_{i=1}^n \xi^i_k \partial_{x_i}$ in local coordinates, this means that $f$ has to satisfy the homogeneous system of linear PDEs of first order \begin{equation}\label{invhomsys} v_k(f)(x) = \sum_{i=1}^n \xi^i_k(x) \frac{\partial f}{\partial x_i} = 0 \quad (k=1,\dots,r). \end{equation} \begin{thr}{\bf Example. }\rm We return to the translation group $G_c$ from \ref{ase1} (i). Its infinitesimal generator is $$ v = \left.\frac{d}{d \eps}\right|_0 (x+c\varepsilon,y+\varepsilon) = (c,1)\equiv c\partial_x + \partial_y. $$ We already know that $f(x,y)=x-cy$ is an invariant, and indeed $v(f)=0$. In the case of the scale group $G^\alpha$ from \ref{ase1} (ii) we have an action of the multiplicative group $(\mathbb R^+,\cdot)$ on $\mathbb R^2$ whose generator therefore is $v=\partial_\lambda|_{\lambda=1}(\lambda x,\lambda^\alpha y) = x\partial_x + \alpha y\partial_y$ and it is easily checked that the infinitesimal criterion is satisfied for the invariants of $G^\alpha$ given in \ref{ase2} (ii). \et Turning now to symmetries of systems of algebraic equations, we begin by deriving a general criterion for the local invariance of submanifolds under local group actions. \begin{thr} {\bf Definition. } Let $G$ be a local transformation group acting on $M$. A subset $S\subseteq M$ is called locally $G$-invariant\index{locally $G$-invariant} if for every $x\in S$ there exists a neighborhood $\tilde{\mathcal U}_x\subseteq {\mathcal U}_x$ (cf.\ \ref{ltgdef2}) of the identity in $G$ such that $g\cdot x\in S$ for all $g\in \tilde{\mathcal U}_x$. A smooth map $F:U\to N$ ($U$ open in $M$) is called locally $G$-invariant if for each $x\in U$ there exists some neighborhood $\tilde{\mathcal U}_x\subseteq {\mathcal U}_x$ of $e$ in $G$ such that $F(g\cdot x)=F(x)$ for all $g\in \tilde{\mathcal U}_x$ with $g\cdot x\in U$. $F$ is called globally $G$-invariant if $F(g\cdot x) = F(x)$ for all $x\in U$ and $g\in G$ such that $g\cdot x \in U$. \et \begin{thr}{\bf Example. }\rm Let $G$ be the group of translations $(x,y)\mapsto (x+\varepsilon,y)$ on $M=\mathbb R^2$. Then the set $S:=\{(x,0)\mid -1<x<1\}$ is locally $G$-invariant, but not $G$-invariant. Let \begin{equation*} f(x,y) := \left\{ \begin{array}{rl} 0 & y\le 0 \text{ or } y>0 \text{ and } x>0\\ e^{-1/y} & y>0 \text{ and } x< 0. \end{array} \right. \end{equation*} Then $f$ is smooth and locally $G$-invariant on $U:=\mathbb R^2\setminus\{(0,y)\mid y\ge 0\}$: in fact, $f(x+\varepsilon,y)=f(x,y)$ for $|\varepsilon| < |x|$. But $f$ is not globally $G$-invariant. \et \bt\label{locinvth} Let $N$ be an initial submanifold of $M$. The following are equivalent: \begin{itemize} \item[(i)] $N$ is locally $G$-invariant. \item[(ii)] The infinitesimal generators of $G$ are everywhere tangent to $N$, i.e.: $$ \forall x\in N\ \forall v\in {\mathfrak g}\colon v|_x \in T_xN $$ \end{itemize} \et \pr If $\Phi\colon G\times M\supseteq {\mathcal U} \to M$ is the action of $G$ then the above condition, written out, means (cf.\ \eqref{phidef}): $$ \forall x\in N\ \forall v\in {\mathfrak g}\colon \Phi(v)|_x \in T_xN $$ (i)$\Rightarrow$(ii): Let $v\in {\mathfrak g}$, $x\in N$. Since $N$ is locally $G$-invariant, for $|\varepsilon|$ small we have $\Phi(\exp(\varepsilon v),x)\in N$. As $N$ is initial, $\varepsilon\mapsto \Phi(\exp(\varepsilon v),x)$ is smooth as a map into $N$ as well. The derivative of this curve at $\varepsilon=0$ therefore is an element of $T_xN$. This implies (ii) since by \ref{flowaction2} we have $$ \Phi(v)(x) = \left.\frac{d}{d \eps}\right|_0 \Phi(\exp(\varepsilon v),x). $$ (ii)$\Rightarrow$(i): By assumption, any $\Phi(v)$ is tangent to $N$, hence by \cite[17.16]{KLie} it can be viewed as a vector field on $N$, and so the restriction of its flow to $N$ is a local diffeomorphism on $N$ (cf.\ \cite[7.2]{KLie}). Given $x\in N$, $\mathrm{Fl}^{\Phi(v)}_\varepsilon(x)$ exists for $|\varepsilon|$ small. Using \ref{flowaction2} we therefore have for $|\varepsilon|$ small \begin{equation}\label{singleeps} \exp(\varepsilon v)\cdot x = \Phi(\exp(\varepsilon v),x) = \mathrm{Fl}^{\Phi(v)}_\varepsilon(x \in N. \end{equation} Next, choose a basis $v_1,\dots,v_k$ of ${\mathfrak g}$. Then as in the last part of the proof of \ref{susslem3} it follows that there exists some $c>0$ such that $\mathrm{Fl}^{\Phi(v_1)}_{\varepsilon_1}\circ\dots\circ \mathrm{Fl}^{\Phi(v_k)}_{\varepsilon_k}(x)$ exists for all $|\varepsilon_i|<c$. In addition, we may assume $c$ so small that $\exp(\varepsilon_1 v_1)\cdot \dots\cdot \exp(\varepsilon_k v_k)\in {\mathcal U}_x$ for $|\varepsilon_i|<c$. Then (i)--(iii) of \ref{schmidtlem1} hold and so, by \ref{schmidtlem2}, iterating \eqref{singleeps} gives $$ \exp(\varepsilon_1 v_1)\cdot \dots\cdot \exp(\varepsilon_k v_k)\cdot x = \mathrm{Fl}^{\Phi(v_1)}_{\varepsilon_1}\circ\dots\circ \mathrm{Fl}^{\Phi(v_k)}_{\varepsilon_k}(x)\in N $$ for these $\varepsilon_i$. Therefore, defining the neighborhood $\tilde{\mathcal U}_x := \{\exp(\varepsilon_1v_1)\dots\exp(\varepsilon_kv_k)\mid |\varepsilon_i|<c\}$ of $e\in G$ (shrinking $c$ if necessary, cf.\ \cite[8.5]{KLie}) we obtain that $g\cdot x\in N$ for all $g\in \tilde{\mathcal U}_x$. \ep \begin{thr} {\bf Remark. }\rm As the proof of \ref{locinvth} shows, the implication (ii)$\Rightarrow$(i) remains correct even for $N$ an immersive submanifold of $M$. However, (i)$\Rightarrow$(ii) is not true in this generality: To see this, equip $M=\mathbb R^2$ with a new manifold structure $N$ whose charts are $\varphi_a:(x,a)\mapsto x$ ($a\in \mathbb R$), see \cite[Ex.\ 14.3]{KLie}. Then $N$ is a one-dimensional immersive submanifold of $M$ with underlying set $\mathbb R^2$, hence is (even globally) invariant under {\em any} local group action on $M$. But clearly not every group action has generators tangential to $N$ (i.e., horizontal). \et As the most important special case of \ref{locinvth} we consider zero sets of smooth maps: \begin{thr} {\bf Corollary. }\label{algsym} Let $G$ be a local transformation group acting on $M$, $\dim(M)=m$. Let $F:M\to \mathbb R^l$ ($l\le m$) be smooth and of maximal rank ($=l$) at every solution $x$ of the system $F_\nu(x)=0$ ($1\le \nu \le l$). Then $G$ is a symmetry group of this system if and only if \begin{equation}\label{algsymcrit} \forall v\in {\mathfrak g}\ \forall \nu=1,\dots l\ \forall x\in M \text{ with } F(x)=0:\ v(F_\nu)|_x = 0 \end{equation} \et \pr By \cite[1.1.23]{DG2}, $N:=\{x\in M\mid F(x)=0\}$ is a regular (hence in particular initial) submanifold of $M$. Thus by \ref{locinvth}, $N$ is locally invariant under $G$ if and only if $v|_x\in T_xN$ for all $x\in N$ and $v\in {\mathfrak g}$. By \cite[1.1.25]{DG2}, $$ T_xN = \ker(T_xF)=\bigcap_{\nu=1}^l \ker(T_xF_\nu) $$ for all $x\in N$. This implies that $N$ is locally $G$-invariant if and only if for each $x\in N$, i.e., for each $x\in M$ with $F(x)=0$, each $v\in {\mathfrak g}$, and each $\nu=1,\dots,l$ we have that $$ 0 = T_xF_\nu(v|_x) = v(F_\nu)|_x. $$ It remains to show that local invariance implies invariance of $N$ under $G$. Thus let $x\in N$ and $g\in {\mathcal U}_x$. By \ref{schmidtlem1}, $g= \exp(t_1v_1)\cdot \dots \cdot \exp(t_nv_n)$ with properties (i)--(iii) from that result. To show that $g\cdot x\in N$, we proceed by induction. Suppose we already know that $g_i\cdot x\in N$, where $g_i:=\exp(t_{i+1}v_{i+1})\cdot\dots\cdot \exp(t_nv_n)$ and set $A_i:=\{t\in [0,t_i]\mid (\exp(tv_i)\cdot g_i)\cdot x \in N\}$. Then $0\in A_i$ and $A_i$ is closed since $N$ is. If $t\in A_i$ then by local invariance of $N$ there exists some $\alpha>0$ such that for all $|s|<\alpha$ we have $\exp(sv_i)\cdot [(\exp(tv_i)\cdot g_i)\cdot x]\in N$. Here, $h_i:=\exp(tv_i)\cdot g_i\in {\mathcal U}_x$ and $\exp(sv_i)\in {\mathcal U}_{h_i\cdot x}$. Also, for $|s|<\alpha$ and $t+s\in [0,t_i]$, $\exp(sv_i)\cdot\exp(tv_i)\cdot g_i=\exp((s+t)v_i)\cdot g_i\in {\mathcal U}_x$ by \ref{schmidtlem1} (iii). Hence \ref{ltgdef2} (i) gives $[(\exp((s+t)v_i)\cdot g_i)]\cdot x= [\exp(sv_i)\cdot (\exp(tv_i)\cdot g_i)]\cdot x = \exp(sv_i)\cdot [(\exp(tv_i)\cdot g_i)\cdot x] \in N $ for these $s$, which establishes that $A_i$ is also open in $[0,t_i]$. Altogether, $A_i=[0,t_i]$, and our claim follows. \ep \begin{thr}{\bf Example. }\rm (i) Let $G=SO(2)$ be the rotation group acting on $M=\mathbb R^2$, with infinitesimal generator $$ v= \left.\frac{d}{d \eps}\right|_0(x\cos\varepsilon - y\sin\varepsilon, x\sin\varepsilon + y\cos\varepsilon)=-y\partial_x+x\partial_y $$ Then $S^1=\{x^2+y^2=1\}$ is clearly invariant under $G$. We can also see this using \ref{algsym}: let $F(x,y):=x^2+y^2-1$. Then $F$ is of maximal rank ($=1$) on its zero set $S^1$ and $v(F)= -2xy + 2xy = 0$. (ii) Let $H(x,y)=y^2-2y+1$. Then the zero set of $H$ is the horizontal line $\{y=1\}$, which is manifestly not invariant under $G$. Nevertheless, $v(H)=2xy-2x=2x(y-1)$ vanishes on this zero set. However, also $TH=(0,2y-2)$ vanishes there, so \ref{algsym} does not apply. This demonstrates that the maximal rank condition in \ref{algsym} cannot be dropped. \et The following criterion will turn out to be useful many times later on: \bt\label{hadamard} (Hadamard's Lemma)\index{Hadamard's Lemma} Let $F:M^m\to \mathbb R^l$ ($l\le m$) be of maximal rank on the subvariety $S_F:=\{x\in M\mid F(x)=0\}$. Then a smooth real-valued function $f\colon M\to \mathbb R$ vanishes on $S_F$ if and only if there exist $Q_1,\dots,Q_l\in {\mathcal C}^\infty(M,\mathbb R)$ such that $$ \forall x\in M:\ f(x) = Q_1(x)F_1(x)+\dots+Q_l(x)F_l(x). $$ \et \pr The condition is clearly sufficient. Conversely, suppose that $f$ vanishes on $S_F$ and let $x\in S_F$. Since $F$ is a submersion, by \cite[1.1.7]{DG2} there exists a chart $(\varphi=(y^1,\dots,y^m),U)$ centered at $x$ such that $F_\varphi(y):=F\circ \varphi^{-1}(y^1,\dots,y^m) = (y^1,\dots,y^l)$. It follows that $f_\varphi:=f\circ\varphi^{-1}$ vanishes on $\varphi(U)\cap (\{0\}\times \mathbb R^{m-l})$. We can suppose that $\varphi(U)$ is a ball around $0$ in $\mathbb R^m$. Then for $y=(y',y''):=(y^1,\dots,y^l,y^{l+1},\dots,y^m)$ we have \begin{equation*} \begin{split} f_\varphi(y',y'')&=f_\varphi(y',y'')-f_\varphi(0,y'')=\int_0^1 \partial_t f_\varphi(ty',y'')\,dt\\ &= \sum_{i=1}^l \int_0^1 \frac{\partial f_\varphi}{\partial y_i}(ty',y'')\cdot y^i\,dt =: \sum_{i=1}^l Q_\varphi^i(y)\cdot y^i = \sum_{i=1}^l Q_\varphi^i(y)\cdot (F_i)_\varphi(y). \end{split} \end{equation*} Therefore, setting $Q^i_U:=Q_\varphi^i\circ \varphi\in {\mathcal C}^\infty(U)$ we obtain $f|_U=\sum_{i=1}^l Q_U^i\cdot F_i|_U$. On the other hand, if $x\not\in S_F$ then there exists some open neighborhood $U$ of $x$ and some $i\in \{1,\dots,l\}$ such that $F_i(y)\not=0$ for all $y\in U$. Then we set $Q^i_U:=f/F_i$ and $Q^j_U:=0$ for $j\not=i$ to again obtain $f|_U=\sum_{i=1}^l Q_U^i\cdot F_i|_U$. Now pick a locally finite open cover $(U_\alpha)_{\alpha\in A}$ of neighborhoods as above and a subordinate partition of unity $(\chi_\alpha)_{\alpha\in A}$ with $\mathrm{supp}(\chi_\alpha)\subseteq U_\alpha$ for all $\alpha$. Then $$ f = \sum_\alpha \chi_\alpha f = \sum_\alpha \chi_\alpha \sum_{i=1}^l Q^i_{U_\alpha}F_i =: \sum_{i=1}^l Q_i F_i. $$ \ep It follows that an equivalent reformulation of \eqref{algsymcrit} is that for any $v\in {\mathfrak g}$ there exist smooth functions $Q_{\nu\mu}: M\to \mathbb R$ ($\mu,\nu=1,\dots,l$) such that \begin{equation}\label{algsymcrit2} v(F_\nu)(x) = \sum_{\mu=1}^l Q_{\nu\mu}(x) F_\mu(x) \quad \nu=1,\dots,l,\quad x\in M. \end{equation} \begin{thr}{\bf Example. }\rm The functions $Q_\nu$ in \ref{hadamard} are in general not unique: Let $M=\mathbb R^3$ and $F(x,y,z):=(x,y)$. Then the function $f(x,y,z)=xz+y^2$ vanishes on $S_F$ (which is the $z$-axis) and we have $$ f=zF_1 + yF_2 = (z-y)F_1 + (x+y)F_2. $$ \et If $f = \sum_\nu Q_\nu F_\nu = \sum_\nu \tilde Q_\nu F_\nu$, then the differences $R_\nu:=Q_\nu-\tilde Q_\nu$ satisfy the homogeneous system \begin{equation}\label{homsys} \sum_{\nu=1}^l R_\nu(x) F_\nu(x) = 0. \end{equation} For such functions we have: \bp\label{hadex} Let $F:M^m\to \mathbb R^l$ be of maximal rank on $S_F=\{x\in M\mid F(x)=0\}$. If $R_1,\dots,R_l\in {\mathcal C}^\infty(M,\mathbb R)$ satisfy \eqref{homsys}, then each $R_\nu$ vanishes identically on $S_F$. Equivalently, there exist $S^\mu_\nu\in {\mathcal C}^\infty(M,\mathbb R)$ ($\nu,\mu = 1,\dots,l$) such that, for all $x\in M$ and all $\nu=1,\dots,l$ \begin{equation}\label{homsysmn} R_{\nu} = \sum_{\mu=1}^l S^\mu_\nu F_\mu. \end{equation} In addition, the $S^\mu_\nu$ can be chosen to be anti-symmetric: $S^\mu_\nu = -S^\nu_\mu$. In this case, \eqref{homsysmn} is necessary and sufficient for \eqref{homsys}. \et \pr As in the proof of \ref{hadamard}, by using a partition of unity and suitable charts we may reduce the proof to neighborhoods of points with either $x\in S_F$ or $x\not\in S_F$. 1) If $x\in S_F$, using a chart as in \ref{hadamard}, we may suppose that $M$ is a ball in $\mathbb R^m$ and $F=(x^1,\dots,x^m)\mapsto (x^1,\dots,x^l)$. Then \eqref{homsys} reads \begin{equation}\label{homsys2} \sum_{\nu=1}^l R_\nu(x) x^\nu = 0. \end{equation} We show by induction that \eqref{homsys2} implies the existence of a skew-symmetric matrix $S^\mu_\nu$ of smooth functions satisfying \eqref{homsysmn}. For $l=1$, \eqref{homsys2} reduces to $R_1(x)x^1=0$ for all $x$, so by continuity $R_1(x)=0$ for all $x$, and we can choose $S^1_1=0$. $l-1 \rightarrow l$: Setting $x':=(x^{l+1},\dots,x^m)$, \eqref{homsys2} implies that $R_l(0,\dots,0,x^l,x')\cdot x^l=0$, so in fact $R_l(0,\dots,0,x^l,x')=0$ for all $(x^l,x')$. Applying \ref{hadamard} to $F:=(x^1,\dots,x^m) \mapsto (x^1,\dots,x^{l-1})$, it follows that there exist smooth functions $S_l^\mu$ ($\mu=1,\dots,l-1$) of $x$ with $R_l(x) = \sum_{\mu=1}^{l-1} S_l^\mu(x)x^\mu$. By \eqref{homsys2}, therefore, $$ 0 = \sum_{\mu=1}^{l-1} (R_\mu(x)x^\mu + S_l^\mu(x)x^\mu x^l) = \sum_{\mu=1}^{l-1} (R_\mu(x)+S_l^\mu(x)x^l)x^\mu =:\sum_{\mu=1}^{l-1} \tilde R_\mu x^\mu. $$ By our induction hypothesis there exist smooth functions $\tilde S^\mu_\nu$ ($1\le \mu,\nu\le l-1$) such that $\tilde R_\nu = \sum_{\mu=1}^{l-1}\tilde S^\mu_\nu x^\mu$ for $1\le \nu\le l-1$ and such that $\tilde S^\mu_\nu$ is skew-symmetric. By definition, $R_\mu = \tilde R_\mu - S_l^\mu x^l$ for $1\le \mu\le l-1$. Therefore, $$ \begin{pmatrix} R_1\\ \vdots \\ R_{l-1}\\ R_l \end{pmatrix} = \left( \begin{array}{cccc} \tilde S_1^1 &\dots &\tilde S_1^{l-1} &-S^l_1\\ \vdots & & &\vdots \\ \tilde S_{l-1}^1 &\dots &\tilde S_{l-1}^{l-1} &-S^l_{l-1} \\ S_l^1 &\dots &S_{l}^{l-1} & 0 \end{array} \right) \cdot \begin{pmatrix} x^1\\ \vdots \\ x^{l-1}\\ x^l \end{pmatrix} =: S\cdot x, $$ giving the desired skew-symmetric $l\times l$ matrix $S$. 2) Now suppose that $x\not\in S_F$. Again we proceed by induction. For $l=1$, $F_1(x)\not=0$ and $R_1\cdot F_1=0$ imply that we can write $R_1(x) = 0 = 0\cdot F_1=:S^1_1\cdot F_1$ on a neighborhood of $x$ where $F_1\not=0$. Next, assume that the result is already proved for $l-1$. Since $F(x)\not=0$, some component of $F(x)$ must be non-zero, and without loss of generality we may suppose that $F_{l-1}(x)\not=0$ (the proof below works the same way for any other component as well). Then on a neighborhood $U$ of $x$ where $F_{l-1}(x)\not=0$ we have $$ R_1F_1+\dots+R_{l-2}F_{l-2} + \Big(R_{l-1} + R_l\frac{F_{l}}{F_{l-1}}\Big)F_{l-1} = 0, $$ and so by our induction assumption (and shrinking $U$ if necessary) there exists a skew-symmetric $(l-1)\times (l-1)$ matrix $\hat S^\mu_\nu$ such that $$ \begin{pmatrix} R_1\\ \vdots \\ R_{l-2}\\ R_{l-1} + R_l\frac{F_{l}}{F_{l-1}} \end{pmatrix} = \left( \begin{array}{ccc} \hat S_1^1 &\dots & \hat S_1^{l-1} \\ \vdots & & \vdots \\ \hat S_{l-2}^1 &\dots &\hat S_{l-2}^{l-1} \\ \hat S_{l-1}^1 &\dots &\hat S_{l-1}^{l-1} \end{array} \right) \cdot \begin{pmatrix} F_1\\ \vdots \\ F_{l-2}\\ F_l \end{pmatrix}. $$ Consequently, $$ \begin{pmatrix} R_1\\ \vdots \\ R_{l-2}\\ R_{l-1}\\ R_l \end{pmatrix} = \left( \begin{array}{cccc} \hat S_1^1 &\dots & \hat S^{l-1}_1 & 0 \\ \vdots & & \vdots & \vdots \\ \hat S_{l-2}^1 &\dots & \hat S^{l-1}_{l-2} & 0 \\ \hat S^{1}_{l-1} & \dots & \hat S^{l-1}_{l-1} & -R_{l}/F_{l-1} \\ 0 &\dots & 0\ \ R_{l}/F_{l-1} & 0 \end{array} \right) \cdot \begin{pmatrix} F_1\\ \vdots \\ F_{l-2}\\ F_{l-1}\\ F_l \end{pmatrix}, $$ which gives the claim also in this case. Conversely, if \eqref{homsysmn} is satisfied, then denoting by $\langle \,,\,\rangle$ the standard scalar product on $\mathbb R^l$ we have $$ \langle R(x),F(x)\rangle = \langle S(x)F(x),F(x)\rangle = 0 $$ since $S^\top=-S$. This is \eqref{homsys}. \ep \section{Invariance and functional dependence} Our goal in this section is to determine `how many' invariants a given group action possesses. We first observe that if $\zeta^1,\dots,\zeta^k$ are invariants (either local or global) of a group $G$ and $F=F(z^1,\dots,z^k)$ is any smooth function then also $$ \zeta(x):=F(\zeta^1(x),\dots,\zeta^k(x)) $$ is an invariant. However, $\zeta$ is completely determined by $\zeta^1,\dots,\zeta^k$ and so adds no additional information. We therefore want to determine invariants up to this kind of relation. \begin{thr} {\bf Definition. }\label{funindef} Let $\zeta^1,\dots,\zeta^k\in {\mathcal C}^\infty(M,\mathbb R)$. Then \begin{itemize} \item[(i)] $\zeta^1,\dots,\zeta^k$ are called functionally dependent\index{functionally dependent} if for each $x\in M$ there exists a neighborhood $U$ of $x$ and a smooth map $F:\mathbb R^k\to \mathbb R$ that does not vanish identically on any open subset of $\mathbb R^k$ such that \begin{equation}\label{funcind} F(\zeta^1(x),\dots,\zeta^k(x)) = 0 \end{equation} for all $x\in U$. \item[(ii)] $\zeta^1,\dots,\zeta^k$ are called functionally independent\index{functionally independent} if they are not functionally dependent when restricted to any open subset $U$ of $M$. Equivalently, if $F$ as in (i) is such that \eqref{funcind} holds for all $x$ in some open subset $U$ of $M$ then $F(z_1,\dots,z_k)=0$ for all $z$ in some open subset of $\mathbb R^k$ (which contains $(\zeta^1,\dots,\zeta^k)(U')$ for some $U'\subseteq U$). \end{itemize} \et \begin{thr}{\bf Example. }\rm (i) The functions $x/y$ and $xy/(x^2+y^2)$ are functionally dependent on $\{(x,y)\mid y\not=0\}$ because $$ \frac{xy}{x^2+y^2} = \frac{x/y}{1+(x/y)^2} = f(x/y). $$ On the other hand, $x/y$ and $x+y$ are functionally independent where defined since if $F(x+y,x/y)\equiv 0$ for $(x,y)$ in any open subset of $\mathbb R^2$ then since $(x,y)\mapsto (x+y,x/y)$ is a local diffeomorphism, $F$ has to vanish identically on some open subset of the image of this set. (ii) The functions $$ \eta(x,y) = x, \quad \zeta(x,y) = \left\{ \begin{array}{cl} x & y\le 0\\ x+e^{-1/y} & y>0 \end{array} \right. $$ are functionally dependent on $\{y<0\}$, independent on $\{y>0\}$, but {\em neither} on the entire space $\mathbb R^2$. \et For a characterization of functional (in)dependence we need two auxiliary results, both of which are of independent interest as well. \bp\label{ccc} Let $M$ be a second countable smooth manifold and let $S$ be a closed subset of $M$. Then there exists a smooth function $f\colon M\to \mathbb R$ such that $S=\{x\in M\mid f(x)=0\}$. \et \pr Since $M\setminus S$ is open and $M$ is second countable, $M\setminus S$ can be written as a locally finite union of countably many sets $V_j$ ($j\in \N$) and there exist functions $\chi_j\in {\mathcal C}^\infty(M)$ with $\chi_j\ge 0$ and $\chi_j>0$ precisely on $V_j$ (see the proof of \cite[2.3.10]{DG1}). Then set $$ f:= \sum_{j\in\N} \chi_j. $$ Since $(V_j)_j$ is locally finite, locally around any point in $M$ only finitely many summands are nonzero, so $f\in {\mathcal C}^\infty(M)$. Moreover, for any $x\in M$ we have $$ f(x)=0\Leftrightarrow \forall j: \chi_j(x)=0 \Leftrightarrow x\in \bigcap_{j\in\N}(M\setminus V_j) = M\setminus \bigcup_{j\in \N}V_j = S. $$ \ep In the next theorem, by a {\em critical point}\index{critical point} of a smooth map $f\colon M\to N$ we mean a point $x\in M$ where $T_xf$ is not surjective. A point $y\in N$ is called a {\em critical value}\index{critical value} of $f$ if there exists a critical point $x\in M$ of $f$ such that $y=f(x)$. \bt\label{sard}\index{Sard's Theorem}{\em (Sard's Theorem)} Let $f\colon M\to N$ be a smooth map between smooth manifolds. Then the set of critical values of $f$ in $N$ has measure zero, in the sense that in any chart of $N$ the image of the set of critical values in that chart domain has Lebesgue measure zero. \et A proof of this result would take us too far afield, so we refer to \cite[3.2]{Kahn}. Based on the previous results we now have: \bt\label{funcindchar} Let $\zeta = (\zeta^1,\dots,\zeta^k)$ be a smooth map from $M$ to $\mathbb R^k$. Then the following are equivalent: \begin{itemize} \item[(i)] $\zeta^1,\dots,\zeta^k$ are functionally dependent on $M$. \item[(ii)] $\forall x\in M: \ \mathrm{rk}(T_x\zeta)<k$. \end{itemize} \et \pr (i)$\Rightarrow$(ii): Suppose that for some $x\in M$ we had $\mathrm{rk}(T_x\zeta)=k$. Since the rank can't fall locally, there would then exist some open set $U$ around $x$ with $\mathrm{rk}(T_{x'}\zeta)=k$ for all $x'\in U$. We may assume $U$ small enough that there is some $F$ as in \ref{funindef} with $F(\zeta^1(x'),\dots,\zeta^k(x')))\equiv 0$ on $U$. Since $\zeta$ is a submersion on $U$, it is an open map. Thus $F$ vanishes on the open set $\zeta(U)$, a contradiction. (ii)$\Rightarrow$(i): Let $U$ be open and relatively compact in $M$. By \ref{ccc} there exists some $F\in {\mathcal C}^\infty(\mathbb R^k)$ such that $F(z)=0$ if and only if $z\in \zeta(\overline U)$. Since $\zeta(\overline U) \subseteq \{\zeta(x)\mid \mathrm{rk}(T_x\zeta)<k\}$ and the latter set has measure zero by Sard's theorem \ref{sard}, $\zeta(\overline U)$ does not contain any non-empty open set. Thus $\zeta^1,\dots,\zeta^k$ are functionally dependent. \ep \bt\label{maininv} Let $G$ act semi-regularly on the $m$-dimensional manifold $M$ with $k$-dimensional orbits and let $x_0\in M$. Then there exists some open neighborhood $U$ of $x_0$ such that there are precisely $m-k$ functionally independent local invariants $\zeta^1,\dots,\zeta^{m-k}$ of $G$ on $U$. Any other invariant of $G$ on $U$ is of the form $$ f(x)=F(\zeta^1(x),\dots,\zeta^{m-k}(x)) $$ for some smooth function $F$. If the action of $G$ is regular then the invariants can be taken to be globally invariant on a neighborhood of $x_0$. \et \pr By \ref{constrk}, any orbit of $G$ is a leaf of the $k$-dimensional integrable distribution $\Delta_\Phi$. Thus the classical Frobenius theorem \ref{classfrob} yields a cubic chart $(\varphi=(x^1,\dots,x^m),U)$ centered at $x_0$, $\varphi(U)=(-c,c)^m$, such that any orbit of $G$ that intersects $U$ does so in a (by \ref{locfol} at most countable) union of slices $U_a=\varphi^{-1}(\mathbb R^k\times \{a\})$. For $1\le i\le m-k$, let $\zeta^i:U\to \mathbb R$, $\zeta^i:=x^{k+i} (=\varphi^{k+i})$. Then by definition, each $\zeta^i$ is constant on each slice, hence is locally constant on any orbit that intersects $U$, hence is a local invariant. Any other local invariant $f$ must be constant on each slice. In terms of the chart $\varphi$ this means that $f$ does in fact not depend on the variables $x^1,\dots,x^k$, and therefore is a function of the remaining variables $x^{k+i}$ ($1\le i\le k$) only, i.e., of the $\zeta^i$. Finally, if the action of $G$ is regular, then by \ref{regslice} the chart $(\varphi,U)$ can be chosen such that each $G$-orbit intersects $U$ in at most one slice. Then the $\zeta^i$ are constant on this slice, i.e., on $U\cap S$, so they are global invariants. \ep Since $\zeta=(\zeta^1,\dots,\zeta^{m-k})$ has maximal rank, the $\zeta^i$ are functionally independent by \ref{funcindchar}. Such a family of invariants (i.e., a functionally independent family such that any other invariant is a function of the members of the family) is called a {\em complete set of functionally independent invariants}. \bp\label{lociprop} Let $G$ act semi-regularly on the $m$-dimensional manifold $M$ with $k$-dimensional orbits. Then: \begin{itemize} \item[(i)] Any complete set of functionally independent local invariants of $G$ has $m-k$ elements. \item[(ii)] If $\eta^1,\dots,\eta^{m-k}$ is a set of functionally independent local invariants, then locally around any point in its domain it is complete. \end{itemize} \et \pr (i) Let $\eta=(\eta^1,\dots,\eta^l)$ and $(\theta^1,\dots,\theta^p)$ be two sets of functionally independent invariants of $G$ and assume that $l>p$. Since $\theta$ is complete, there exists some smooth map $F:\Omega\to \mathbb R^l$ with $\Omega\subseteq\mathbb R^p$ open such that $\eta = F\circ \theta$. But then for any $x$, $\mathrm{rk}(T_x\eta)\le \mathrm{rk} T_{\theta(x)}(F)\le p<l$, a contradiction to \ref{funcindchar}. By symmetry, $l=p$, and by \ref{maininv} it follows that in fact $l=m-k$. (ii) Suppose that $f$ is some invariant defined on the domain $\Omega$ of $\eta=(\eta^1,\dots,\eta^{m-k})$ and let $x\in \Omega$. Also, pick $\zeta = (\zeta^1,\dots,\zeta^{m-k})$ around $x$ as in \ref{maininv}. Then there exists some $F:\mathbb R^{m-k}\to\mathbb R^{m-k}$, such that $\eta = F\circ \zeta$ near $x$. Since $\mathrm{rk}(\eta)=\mathrm{rk}(\zeta)=m-k$, we also must have $\mathrm{rk}(F)=m-k$, i.e., $F$ is a local diffeomorphism. Since $f$ is a smooth function of $\zeta$ near $x$ it is therefore also a smooth function of $\eta$. \ep \bp\label{invvar} Let $G$ act semi-regularly on the $m$-dimensional manifold $M$ with $k$-dimensional orbits and let $\zeta^1,\dots,\zeta^{m-k}$ be a complete set of functionally independent invariants defined on an open subset $U$ of $M$. If a subvariety $S_F=\{x\in M\mid F(x)=0\}$ (with $F\in{\mathcal C}^\infty(M,\mathbb R^l)$) is $G$-invariant, then for each solution $x_0\in S_F\cap U$ there is a neighborhood $\tilde U\subseteq U$ of $x_0$ and a smooth function $f$ such that the solution set of the corresponding invariant $\tilde F = x\mapsto f(\zeta^1(x),\dots,\zeta^{m-k}(x))$ on $\tilde U$ coincides with that of $F$, i.e., $$ S_F\cap \tilde U = S_{\tilde F}\cap \tilde U = \{x\in \tilde U\mid f(\zeta^1(x),\dots,\zeta^{m-k}(x))=0\}. $$ \et \pr Since (by \ref{funcindchar}) the rank of $(\zeta^1,\dots,\zeta^{m-k})$ is $m-k$ on $U$ we may find coordinates $(y^1,\dots,y^m)$ on a neighborhood $\tilde U$ of $x_0$ such that $y^i=\zeta^i$ for $1\le i\le m-k$. In fact, the remaining $k$ coordinates can be determined by picking suitable $x^{i_j}$ (called {\em parametric variables}) from any given coordinates $(x^1,\dots,x^m)$. The change of coordinates is then of the form $y=\psi(x)=(\zeta(x),\hat x)$, with $\hat x$ the parametric variables. We show that these coordinates are then flat for $\Delta_\Phi$: Each $\zeta^j$ is constant on the orbits of $\Delta_\Phi$, hence if $v_1,\dots,v_k\in {\mathfrak g}$ are such that $\Delta_\Phi(x)=\text{span}(v_1|_x,\dots,v_k|_x)$ then $$ T_x{\zeta^j}(v_i|_x)=v_i(\zeta^j)|_x = \left.\frac{d}{d t}\right|_0\zeta^j(\mathrm{Fl}^{v_i}_t(x))=0 $$ for all $x\in \tilde U$. Therefore, $v_i|_x \in \bigcap_{j=1}^k \ker(T_x\zeta^j) = \ker T_x\zeta$, and since $\dim(\ker(T_x\zeta))=k$, $\Delta_\Phi(x)=\text{span}(v_1|_x,\dots,v_k|_x)=\ker T_x\zeta$. On the other hand, $\ker(T_x\zeta)$ is spanned by $\partial_{y^{m-k+1}},\dots,\partial_{y^{m}}$ because $T\zeta^i(\partial_{y^{j}})= \frac{\partial y^i}{\partial y^j}=0$ for $i\in \{1,\dots,m-k\}$ and $j\in \{m-k+1,\dots,m\}$. In terms of the new variables we can write $F(x) = F^*(y) = F^*(\zeta(x),\hat x)$, where $F^*:=F\circ\psi^{-1}$. Denoting by $\hat x_0$ the value of the parametric variables at $x_0$ we now set $f(z):=F^*(z,\hat x_0)$, and $$ \tilde F(x) := f(\zeta(x)) = f(\zeta^1(x),\dots,\zeta^k(x)). $$ By assumption, $S_F$ is $G$-invariant. Also, the orbits of $G$ intersect $U$ in the slices $\{\zeta(x)=c\}$. Therefore, $F(x)=F^*(\zeta(x),\hat x)=0$ if and only if $\tilde F(x) = F^*(\zeta(x),\hat x_0)=0$ since $(\zeta(x),\hat x)$ and $(\zeta(x),\hat x_0)$ lie in the same slice. \ep We now turn to the problem of actually calculating invariants of a local transformation group. To begin with, we consider the case of a one-parameter group $G$ with infinitesimal generator $v\in {\mathfrak X}_{\mathrm{loc}}(M)(M)$. In terms of local coordinates $x^1,\dots,x^n$ we can write $$ v = \xi^1(x)\partial_{x^1}+\dots+\xi^m(x)\partial_{x^m} $$ for certain smooth functions $\xi^i$. Then by \ref{finvcrit}, a local invariant $f$ of $G$ is a solution of the linear homogeneous first order PDE \begin{equation}\label{invpde} v(f) = \xi^1(x)\frac{\partial f}{\partial x^1} +\dots+\xi^m(x)\frac{\partial f}{\partial x^m} = 0. \end{equation} By \ref{maininv} we know that if $v|_x\not=0$ then locally around $x$ there exist $m-1$ functionally independent invariants, i.e., $m-1$ functionally independent solutions of \eqref{invpde}. In PDE theory, the method of choice for solving \eqref{invpde} is the {\em method of characteristics}.\index{method of characteristics}, cf.\ \cite[Sec.\ 3.2]{Ev}. This basically consists in solving the corresponding system of ODEs given by \begin{equation}\label{invode} \frac{dx^i}{dt} = \xi^i(x(t)) \quad (1\le i\le m). \end{equation} Solutions of \eqref{invpde} are then functions that are constant along solutions of \eqref{invode}. Such functions are also called {\em first integrals}\index{first integral} of \eqref{invode}. In this terminology, \ref{maininv} says that \eqref{invode} possesses $m-1$ functionally independent first integrals locally around any point. \begin{thr} {\bf Remark. }\rm Using the straightening-out theorem (cf.\ \cite[17.14]{KLie}) it is very easy to see all of the above directly. In fact, in an appropriate coordinate system $y^1,\dots,y^m$ we have $v=\partial_{y^1}$ near the point $x$. It is then immediate that the functions $y^2,\dots,y^m$ form a complete set of functionally independent invariants of $\mathrm{Fl}^v$. Furthermore, in these coordinates \eqref{invode} reads $\frac{dy^1}{dt} = 1$, and $\frac{dy^i}{dt} = 0$ for $i=2,\dots,m$, with general solution $y^1(t) = t + c_1$, $y^i=c_i$ for $i=2,\dots,m$. A complete set of functionally independent first integrals then of course is also given by $y^2,\dots,y^m$, because these functions are constant along this general solution of the ODE. \et \begin{thr}{\bf Example. }\rm Consider the rotation group $SO(2)$ on $\mathbb R^2\setminus \{(0,0)\}$ with infinitesimal generator $v=-y\partial_x +x\partial_y$. Then the characteristic system reads $$ \frac{dx}{dt} = -y \quad \frac{dy}{dt} =x. $$ A first integral is $f(x,y)=x^2+y^2$, since \begin{equation*} \begin{split} \frac{d}{dt}f(x(t),y(t)) &= \partial_x f(x(t),y(t))(-y(t)) + \partial_y f(x(t),y(t))(x(t))\\ &= -2x(t)y(t)+2x(t)y(t)=0. \end{split} \end{equation*} Any other first integral, i.e., any other invariant of $v$, is a function of $f$. \et In the remainder of this section we want to explore how invariance of functions or subvarieties under a local transformation group can be expressed in terms of quotient manifolds, using \ref{quotient}. The main result is as follows: \bt\label{quotob} Let $G$ be a local transformation group on $M$ that acts regularly on the $m$-dimensional manifold $M$ with $k$-dimensional orbits. \begin{itemize} \item[(i)] A smooth map $F: M\to \mathbb R^l$ is $G$-invariant if and only if there exists a smooth function $\tilde F: M/G \to \mathbb R^l$ such that $F(x)=\tilde F(\pi(x))$ for all $x\in M$. \item[(ii)] Let $F\in {\mathcal C}^\infty(M,\mathbb R^l)$. Then the corresponding subvariety ${\mathcal S}_F=\{x\in M\mid F(x)=0\}$ is $G$-invariant if and only if there exists a smooth map $\tilde F:M/G\to \mathbb R^l$ such that, for all $x\in M$, $F(x)=0$ if and only if $\tilde F(\pi(x))=0$. \item[(iii)] An $n$-dimensional regular submanifold $N$ of $M$ is $G$-invariant if and only if there exists a smooth $(n-k)$-dimensional regular submanifold $\tilde N = N/G$ of $M/G$ such that $N=\pi^{-1}(\tilde N)$ (and therefore $\tilde N=\pi(N)$). \end{itemize} \et \pr (i) By definition, $\pi$ is invariant under $G$, so any $\tilde F\circ\pi$ is $G$-invariant as well. Conversely, suppose that $F\in {\mathcal C}^\infty(M,\mathbb R^l)$ is $G$-invariant. Using a regular chart $(U,\varphi)$, the proof of \ref{maininv}, together with \ref{quotientconst} shows that on $\pi(U)$ there exists a smooth map $\tilde F_U$ such that $\tilde F_U\circ\pi= F$ on $U$. The sets $\pi(U)$ form a covering of $M/G$ by chart neighborhoods (since $\pi(U)=U'$), so we may find a partition of unity $(\chi_j)$ on $M/G$ subordinate to it, say with $\mathrm{supp} \chi_j\comp U_j'$ for all $j$. Then setting $\tilde F:= \sum_j \chi_j\cdot \tilde F_{U_j}$ it follows that, for any $x\in M$, $$ \tilde F(\pi(x)) = \sum_j \chi_j(\pi(x))\cdot \tilde F_{U_j}(\pi(x)) = \sum_j \chi_j(\pi(x))\cdot F(x) = F(x). $$ (ii) Suppose first that there exists some $\tilde F\in {\mathcal C}^\infty(M/G,\mathbb R^l)$ such that $F(x)=0$ if and only if $\tilde F(\pi(x))=0$. Then if $y$ is in the same orbit as $x$, $\pi(x)=\pi(y)$. Thus, if $F(x)=0$ then $0=\tilde F(\pi(x))=\tilde F(\pi(y))$, which in turn is equivalent to $F(y)=0$. It follows that ${\mathcal S}_F$ is $G$-invariant. Conversely, if ${\mathcal S}_F$ is $G$-invariant then by \ref{invvar} (together with \ref{quotientconst}) it follows that any point in $M$ has a neighborhood $U$ such that there exists some smooth map $\tilde F_U$ on $U'=\pi(U)$ with the property that, for all $x\in U$, $F(x)=0$ if and only if $\tilde F_U(\pi(x))=0$. Let $\tilde F_U(z)=(\tilde F_1(z),\dots,\tilde F_{l}(z))$. Then we may replace $\tilde F_i$ by $(\tilde F_i)^2$ for all $i\in \{1,\dots,l\}$ while retaining the property stated above. Hence without loss of generality we may assume that each $\tilde F_i$ is non-negative. As in (i), pick a partition of unity $(\chi_j)$ on $M/G$ subordinate to a covering by such neighborhoods $U_j'$ ($j\in \N$) and set $\tilde F:= \sum_j \chi_j\cdot \tilde F_{U_j}$. Let $x\in M$ with $F(x)=0$ and suppose that $\pi(x)\in U_j'=\pi(U_j)$. Then for some $x_j\in U_j$ we have $\pi(x)=\pi(x_j)$, so since $S_F$ is $G$-invariant we obtain $x_j\in S_F$ as well. Hence $0=\tilde F_{U_j}(\pi(x_j))= \tilde F_{U_j}(\pi(x))$ for any such $j$, implying $\tilde F(\pi(x))=0$. Conversely, suppose that $\tilde F(\pi(x))=0$ and pick $k$ such that $\chi_k(\pi(x))>0$. Then since each $\tilde F_{U_j}^ i$ is non-negative, it follows that $\tilde F_{U_k}(\pi(x))=0$. Since $U_k'=\pi(U_k)$, there exists some $x_k\in U_k$ with $\pi(x_k)=\pi(x)$. By construction of $\tilde F_{U_k}$, therefore, $F(x_k)=0$, i.e., $x_k\in S_F$. But then also $F(x)=0$ since $S_F$ is $G$-invariant. (iii) Suppose that $N = \pi^{-1}(\tilde N)$ and let $x\in N$ and $g\in G$ such that $y:=g\cdot x$ exists. Then $\pi(y)=\pi(x)\in \tilde N$, so $y\in \pi^{-1}(\tilde N)=N$. Thus $N$ is $G$-invariant. Conversely, if $N$ is a $G$-invariant regular submanifold of $M$ and $\Phi\colon{\mathcal U} \to M$ denotes the action of $G$ on $M$ (cf.\ \ref{ltgdef2}), then ${\mathcal U}':={\mathcal U}\cap (G\times N)$ is open in $G\times N$ and contains $\{e\}\times N$. Also, $\Phi\colon{\mathcal U}'\to M$ is smooth. Moreover, $\Phi({\mathcal U}')\subseteq N$ since $N$ is $G$-invariant and because $N$ is regular it follows that $\Phi\colon{\mathcal U}'\to N$ is smooth, hence is a local transformation group. Let $x\in N$ and let $S$ be the $G$-orbit of $x$. Then $S$ is a regular submanifold of $M$ (by \ref{regmanrem}) and is contained in $N$. Thus by \cite[13.6]{KLie} it is an immersive submanifold of $N$. Also, it carries the trace topology of $M$, as does $N$, so in fact it carries the trace topology of $N$, i.e., it is a regular submanifold of $N$. Furthermore, if $\mathcal B$ is a neighborhood basis of $x$ in $M$ such that $S'\cap U$ is connected in $S'$ for every orbit $S'$ and every $U\in \mathcal B$ then $\{U\cap N\mid U\in \mathcal B\}$ has the same property in $N$ (since it carries the trace topology of $M$), so $G$ acts regularly on $N$ as well. We are therefore in a position to apply \ref{quotient} to $\Phi\colon{\mathcal U}'\to N$ and obtain that $\pi|_N: N\to N/G$ is a surjective submersion and $N/G=\pi(N)=:\tilde N$ is an $(n-k)$-dimensional manifold. Since $N$ is $G$-invariant, if $y\in \pi^{-1}(\tilde N)$ then $\pi(y)=\pi(x)$ for some $x\in N$, so $y\in N$, i.e., $N=\pi^{-1}(\tilde N)$. It only remains to prove that $\tilde N$ is a regular submanifold of $M/G$. To see this, we will proceed analogously to the proof of \ref{locfol}. Let $x\in N$, let $S$ be the orbit of $x$ (so $S\subseteq N$), and pick $v_1,\dots,v_k\in{\mathfrak g}$ such that $\{v_1(x),\dots,v_k(x)\}$ is a basis of $\Delta_\Phi(x)$. Next, choose an adapted chart $(\chi=(y^1,\dots,y^m),W)$ around $x$ in $M$ such that $((y^1,\dots,y^n),W\cap N)$ is a chart of $N$, $\chi(x)=0\in \mathbb R^m$, and $\chi(W\cap N)=\{y^{n+1}=\dots=y^m=0\}$. Also, we may suppose that $v_1(x),\dots,v_k(x),\pdl{}{y^{k+1}}{x},\dots,\pdl{}{y^n}{x}$ is a basis of $T_xN$, and $v_1(x),\dots,v_k(x),\pdl{}{y^{k+1}}{x},$ $\dots,\pdl{}{y^m}{x}$ is a basis of $T_xM$. Let $$ f(t^1,\dots,t^m):=(\mathrm{Fl}^{v_1}_{t^1}\circ\dots \circ \mathrm{Fl}^{v_k}_{t^k})(\chi^{-1}(0,\dots,0,t^{k+1},\dots,t^m)). $$ Then $f$ is a diffeomorphism from some neighborhood $(-d,d)^m$ of $0\in \mathbb R^m$ onto a neighborhood of $x$ in $M$, and we take $\varphi:=f^{-1}$ on a suitable neighborhood $U$ of $x$ to obtain that $(U,\varphi)$ is a chart of $M$ around $x$. Furthermore, since $\chi$ is an adapted chart and $N$ is $G$-invariant, $$ (t^1,\dots,t^n)\mapsto f(t^1,\dots,t^n,0,\dots,0) $$ maps into $N$ and can (by making $d$ smaller if necessary) also be assumed to be a diffeomorphism. This means that $(\varphi,U)$ is also an adapted chart for $N$. Since $S$ is an orbit of $G$, $$ y\in S \Leftrightarrow \mathrm{Fl}^{v_1}_{t^1}\circ\dots \circ \mathrm{Fl}^{v_k}_{t^k}(y)\in S $$ for all $y$ and $t^1,\dots,t^k$ where the right hand side is defined. Therefore, for any $y=f(t^1,\dots,t^m)\in U$ we have \begin{equation}\label{locstru2} y=f(t^1,\dots,t^m)\in S \Leftrightarrow f(0,\dots,0,t^{k+1},\dots,t^m)\in S. \end{equation} This means that $U\cap S$ is the disjoint union of connected sets of the form $$ U_c:=\{y\in U\mid x^{k+1}(y)=c_{k+1},\dots,x^m(y)=c_m\} $$ where $c=(c_{k+1},\dots,c_m)$ is constant. Since $G$ acts regularly, we may suppose in addition that $U\cap S$ is connected. Then since $f(0,\dots,0)=x\in S$ it follows that the only non-empty $U_c$ is $U_0$, so $S\cap U = \{y\in U\mid x^{k+1}(y)=\dots=x^m(y)=0\}$. Summing up, we have that $(\varphi,U)$ is a flat chart for $G$ on $M$ and $(\varphi,U\cap N)$ is a flat chart for $G$ on $N$. Since $G$ acts regularly, the proof of \ref{regslice} shows that we may in addition assume that $U$ is so small that $(\varphi,U)$ is regular for $\Delta_\Phi$ on $M$ and $(\varphi,U\cap N)$ is regular for $\Delta_\Phi$ on $N$. Then by \ref{quotientconst}, on $U'=\pi(U)$ we obtain a chart $\varphi'$ for $M/G$ around $\pi(x)$ by $$ \varphi': y' \mapsto (x^{k+1}(y),\dots,x^m(y)), \quad U'\to \mathbb R^{m-k} $$ with $y$ any element of $\pi^{-1}(y')\cap U$. By the same reasoning, $$ y' \mapsto (x^{k+1}(y),\dots,x^n(y)), \quad U'\cap \tilde N \to \mathbb R^{n-k} $$ (with $y$ any element of $\pi^{-1}(y')\cap U\cap N$) is a chart for $\tilde N =N/G$ around $\pi(x)$. This shows that $(\varphi',U')$ is a chart adapted to $\tilde N$, i.e., $\tilde N$ is a regular submanifold of $M/G$. \ep \section{Dimensional analysis}\index{dimensional analysis} In a typical physics problem, there are certain fundamental physical quantities, such as length, time, mass, \dots, which can all be scaled independently of each other. Let $z^1,\dots,z^r$ denote these quantities, then this scaling can be described by the action of a scaling group $$ (z^1,\dots,z^r) \mapsto (\lambda_1z^1,\dots,\lambda_rz^r), $$ with fixed scaling factors $\lambda = (\lambda_1,\dots,\lambda_r)\in\mathbb R^r$. The underlying group therefore is the $r$-th power of the multiplicative group $\mathbb R^+$. Furthermore, there typically are certain derived quantities, such as velocity, acceleration, force, density, \dots, which also scale if the fundamental quantities are scaled. Call these quantities $x=(x^1,\dots,x^m)$. Then the action of the scaling group on the derived quantities takes the form \begin{equation}\label{derform} \lambda\cdot (x^1,\dots,x^m) = (\lambda_1^{\alpha_{11}}\lambda_2^{\alpha_{21}}\dots \lambda_r^{\alpha_{r1}}x^1,\dots, \lambda_1^{\alpha_{1m}}\lambda_2^{\alpha_{2m}}\lambda_r^{\alpha_{rm}}x^m), \end{equation} where $\alpha_{ij}$, $i=1,\dots,r$, $j=1,\dots,m$ is a matrix of real numbers determined by the problem at hand. For example, if $z^1$ is length, $z^2$ is time and $z^3$ is mass, then the action of scaling on velocity $x^1$ and force $x^2$ are given by $$ \lambda\cdot (x^1,x^2) = (\lambda_1\lambda_2^{-1}x^1,\lambda_1^1\lambda_2^{-2}\lambda_3 x^2). $$ If a derived quantity is invariant under the corresponding scaling then it is called {\em dimensionless}\index{dimensionless}. To describe some physical situation, one often has functional relations of the form $F(x^1,\dots,x^m)=0$ for the derived quantities. Such a relation is called {\em unit-free} if it remains unchanged under a rescaling of the fundamental quantities. It turns out that such unit-free relations are often of great physical significance. The Buckingham Pi-Theorem shows that any unit-free relation can be written solely in terms of dimensionless quantities. \bt\label{buckpi} (Buckingham Pi-theorem)\index{Buckingham Pi-theorem} Let $z^1,\dots,z^r$ be fundamental physical quantities that scale independently according to $z^i\mapsto \lambda_iz^i$. Let $x^1,\dots,x^m$ be derived quantities that scale according to \eqref{derform} for some $(r\times m)$-matrix of constants $A=(\alpha_{ij})$ and let $s$ be the rank of $A$. Then \begin{itemize} \item[(i)] There exist $m-s$ functionally independent dimensionless `power products' \begin{equation}\label{power} \pi^k = (x^1)^{\beta_{1k}}(x^2)^{\beta_{2k}}\dots(x^m)^{\beta_{mk}},\quad k=1,\dots, m-s \end{equation} such that any other dimensionless quantity can be written as a function of $\pi^1,\dots,$ $\pi^{m-s}$. The columns of the matrix $B=(\beta_{jk})$ in \eqref{power} can be taken to be any basis of $\ker A$. \item[(ii)] If $F(x^1,\dots,x^m)=0$ is any unit-free relation among the given derived quantities, then there is an equivalent relation $\tilde F=0$ which can be expressed solely in terms of the above dimensionless power products: $$ F(x)=0 \Leftrightarrow \tilde F(\pi^1(x),\dots,\pi^{m-s}(x))=0. $$ \end{itemize} \et \pr The proof will basically be an application of \ref{quotob}. As the underlying manifold $M$ we take the positive octant in $\mathbb R^m$, so $$ M = \{x=(x^1,\dots,x^m)\mid x^i>0,\, i=1,\dots, m\}. $$ Then the multiplicative group $G=(\mathbb R^+)^r$ acts globally on $M$ via \eqref{derform}. Since the action is multiplicative, its generators are found by differentiating \eqref{derform} with respect to $\lambda_i$ and then setting $\lambda_1=\dots=\lambda_r=1$: $$ v_i = \alpha_{i1}x^1\pd{}{x^1} + \alpha_{i2}x^2\pd{}{x^2}+\dots+\alpha_{im}x^m\pd{}{x^m}, \quad i=1,\dots,r. $$ It follows that the dimension of the span of $v_1,\dots,v_r$ is precisely the rank of $A$, namely $s$, so $G$ has $s$-dimensional orbits. For a function $f$ to be a global invariant of $G$, by \ref{finvcrit} it has to satisfy $$ v_i(f) = 0, \quad i=1,\dots,r. $$ In particular, if $f=\pi_k$ is given by \eqref{power}, then this condition is satisfied if and only if the exponents $\beta_{jk}$ satisfy the system of linear equations \begin{equation}\label{linsys} \sum_{j=1}^m \alpha_{ij}\beta_{jk}=0,\quad i=1,\dots,r. \end{equation} Since $\mathrm{rk}(A)=s$, there are $m-s$ linearly independent solutions to this system. Choosing these solutions for the $\beta_{jk}$, let $\pi:=(\pi^1,\dots,\pi^{m-s})$. Then the Jacobian of $\pi$ at $x=(1,\dots,1)$ is the transpose of the matrix $(\beta_{jk})$, $j=1,\dots,m$, $k=1,\dots,m-s$, and its rank therefore is $m-s$. For any $a>0$ we have (as matrix-multiplication) $$ \pi(ax^1,x^2,\dots,x^m) = \mathrm{diag}(a^{\beta_{11}},\dots,a^{\beta_{1(m-s)}})\cdot \pi(x^1,x^2,\dots,x^m), $$ so since $D\pi(1,1,\dots,1)$ has a non-zero $(m-s)\times (m-s)$-sub-determinant, the same is true of $D\pi(a,1,\dots,1)$. Analogously we can argue for varying the $x^2$-component, and so on. Altogether, it follows that the rank of the Jacobian of $\pi$ is $(m-s)$ everywhere on $M$. Thus by (the easy direction of) \ref{funcindchar}, $(\pi^1,\dots,\pi^{m-s})$ is a functionally independent system of invariants of $G$ on $M$. Next we show that the orbits of $G$ are precisely the level sets of the function $\pi$. To see this, let $x$, ${\tilde x}$ be any two points in $M$. Then by the definition of $M$ there exist exponents $t_j$, $j=1,\dots,m$ such that $x^j = e^{t_j}\cdot {\tilde x}^j$ for all $j$. Therefore \begin{equation*} \begin{split} \pi^k(x) &= e^{\beta_{1k}t_1}e^{\beta_{2k}t_2}\dots e^{\beta_{mk}t_m}({\tilde x}^1)^{\beta_{1k}} \dots ({\tilde x}^m)^{\beta_{mk}} = \pi^k({\tilde x})\\ &\Leftrightarrow \sum_{j=1}^m\beta_{jk}t_j = 0, \end{split} \end{equation*} so $x$ and ${\tilde x}$ lie in the same level set of $\pi$ if and only if the $t_j$ satisfy \begin{equation}\label{orbiteq} \sum_{j=1}^m\beta_{jk}t_j = 0, \quad k=1,\dots,m-s. \end{equation} By construction, the columns of the $m\times (m-s)$-matrix $B:=(\beta_{jk})$ form a basis of $\ker A$, so in particular $A\cdot B=0$. Moreover, \eqref{orbiteq} means that $B^\top\cdot t = 0$, where $t=(t^1,\dots,t^m)^\top$, i.e., $t\in \ker(B^\top)$. We claim that this, in turn, is equivalent to the existence of $s_1,\dots,s_r$ with \begin{equation}\label{orbiteq2} t_j=\sum_{i=1}^r s_i\alpha_{ij}, \quad j=1,\dots,m \end{equation} i.e., to $t\in \text{im}(A^\top)$. This means we have to show that $\text{im}(A^\top)=\ker(B^\top)$. Now if $w=A^\top u\in \text{im}(A^\top)$, then $B^\top w = B^\top A^\top u = (AB)^\top u =0$, so $\text{im}(A^\top)\subseteq\ker(B^\top)$. On the other hand, $\dim \text{im}(A^\top) = \dim \text{im}(A)=s$, and also $\dim\ker(B^\top)=m-\dim \text{im}(B^\top)= m-\dim \text{im}(B) = s$, so indeed we have equality. Hence two points $x$ and ${\tilde x}$ in $M$ have the same image under $\pi$ if and only if $x=\lambda {\tilde x}$, where $\lambda_i=e^{s_i}$, i.e., if and only if they lie in the same orbit under $G$. Since $\pi$ has maximal rank $m-s$ everywhere, its level sets are regular submanifolds of dimension $s$, so $G$ has $s$-dimensional regular submanifolds as orbits. Moreover, the proof of \ref{invvar} shows that, around any point, $\pi^1,\dots,\pi^{m-s}$ can be completed by a set of $s$ parametric variables to form a flat chart for the group action of $G$. In the present situation, such parametric variables can even be chosen globally on all of $M$: indeed, pick numbers $\beta_{jk}$, $j=1,\dots m$, $k=m-s+1,\dots,m$ that supplement the matrix $B$ from above to obtain an invertible $m\times m$ matrix $\tilde B$ and set $$ \tilde \pi(x) := \Big(\prod_{i=1}^m(x^i)^{\beta_{i1}},\dots,\prod_{i=1}^m(x^i)^{\beta_{im}}\Big). $$ Then the first $k$ components of $\tilde \pi$ are precisely $\pi$, and the remaining $m-s$ components are our new parametric variables. By the same reasoning as above for $\pi$ it follows that $\tilde \pi$ has rank $m$ everywhere on $M$, hence is a local diffeomorphism. In addition, $\tilde \pi$ is injective: to see this, note first that $\tilde \pi(x)=\tilde \pi(y)$ implies $x=y$ if and only if $\tilde\pi^i(x)=1$ implies $x^i=1$ for all $i$. Now if $\tilde \pi^i(x)=1$ for all $i$ then taking logarithms it follows that $\tilde B^\top \cdot (\mathrm{ln}(x^1),\dots,\mathrm{ln}(x^m))^\top=0$, which due to the invertibility of $\tilde B$ indeed implies $x^i=1$ for all $i$. Thus we obtain the desired global chart $\tilde \pi$ for $M$. Any orbit intersects this chart in a single slice ($\pi=$ const.), so the action of $G$ is regular. Moreover, by \ref{quotientconst}, $\pi$ induces global coordinates on $M/G$, hence can be identified with the quotient map $M\to M/G$ (and $M/G$ can be identified with $\mathbb R_+^{m-s}$). We are therefore in the position to apply \ref{quotob}. Since being $G$-invariant by definition is the same as being dimensionless, (i) follows from \ref{quotob} (i). Moreover, an equation is unit-free if and only if it is $G$-invariant, so (ii) follows from \ref{quotob} (ii). \ep \begin{thr}{\bf Example. }\rm The energy yield of the first atomic explosion. In 1947, when the amount of energy released by the first atomic explosion\index{atomic explosion} was still classified, G.\ Taylor calculated this energy using dimensional analysis (see \cite{Tay}, our presentation follows \cite{BCo}). We model the explosion by the radius $R$ of the spherical fireball emanating from the point of explosion. We assume that $R$ is a function \begin{equation}\label{nuke} R = F(x^1,x^2,x^3,x^4), \end{equation} where \begin{description} \item[]\hspace*{2em} $x^1=E$, the energy released by the explosion, \item[]\hspace*{2em} $x^2 = t$, the time elapsed since the explosion, \item[]\hspace*{2em} $x^3=\rho_0$, the initial ambient air density, \item[]\hspace*{2em} $x^4=P_0$, the initial ambient air pressure. \end{description} $R$ itself also is a derived quantity, so we set $x^5=R$. It follows that for this problem $m=5$. The fundamental physical quantities needed to describe the derived quantities are $z^1=$ length, $z^2 = $ mass, and $z^3=$ time, so $r=3$. The matrix $A=(\alpha_{ij})_{i=1,\dots,3}^{j=1,\dots,5}$ has the form $$ A = \begin{pmatrix} 2 & 0 & -3 & -1 & 1\\ 1 & 0 & 1 & 1 & 0\\ -2 & 1 & 0 & -2 & 0 \end{pmatrix} $$ E.g., energy has dimension $\text{length}^2\times \text{mass}/\text{time}^2$, hence the first column. The rank of this matrix is $s=3$, so $m-s=5-3=2$. Therefore, by \ref{buckpi} there are two functionally independent dimensionless power products $\pi^1$, $\pi^2$, in terms of which any dimensionless quantity can be described. By \eqref{power}, these power products are of the form \begin{equation*} \begin{split} \pi^1(x) &= (x^1)^{\beta_{11}}(x^2)^{\beta_{21}}(x^3)^{\beta_{31}}(x^4)^{\beta_{41}}(x^5)^{\beta_{51}}\\ \pi^2(x) &= (x^1)^{\beta_{12}}(x^2)^{\beta_{22}}(x^3)^{\beta_{32}}(x^4)^{\beta_{42}}(x^5)^{\beta_{52}}. \end{split} \end{equation*} From the proof of \ref{buckpi} we know that the $5\times 2$ matrix $B=(\beta_{jk})$ can be constructed by finding a basis of $\ker A$. It follows that $$ B= \begin{pmatrix} -2 & -1\\ 6 & -2\\ -3 & 1\\ 5 & 0\\ 0 & 5 \end{pmatrix}, $$ so \begin{equation*} \begin{split} \pi^1(x) &= (x^1)^{-2}(x^2)^{6}(x^3)^{-3}(x^4)^{5} = \frac{P_0^5 t^6}{E^2\rho_0^3}\\ \pi^2(x) &= (x^1)^{-1}(x^2)^{-2}(x^3)(x^5)^{5} = \frac{R^5\rho_0}{Et^2}. \end{split} \end{equation*} By \ref{buckpi} (ii), the relation \eqref{nuke} can equivalently be expressed in the form $\tilde F(\pi^1,\pi^2)=0$, and we assume that in fact we can solve for $\pi^2$. Thus there exists a smooth function $\tilde g$ of $\pi^1$ such that \eqref{nuke} is equivalent to $\pi^2=\tilde g(\pi^1)$, i.e., to \begin{equation}\label{nuke2} R = \left(\frac{Et^2}{\rho_0}\right)^{1/5}g(\pi^1). \end{equation} for some function $g$. By continuity, $g(\pi^1)\approx g(0)$ for $\pi^1$ small, so Taylor derived the approximative formula $$ R = At^{2/5}, \quad \text{ with } A = \left(\frac{E}{\rho_0}\right)^{1/5}g(0). $$ Here, $\rho_0$ is known, and to find $g(0)$ one can plot $\log_{10} R$ versus $\log_{10} t$ using data from experiments with conventional explosives (whose $E$ is known). Taylor then used motion picture records of the first atomic explosion, due to J.E.\ Mack, to plot $5/2 \log_{10} R$ versus $\log_{10} t$ (in c.g.s.-units, but as we know any other system of units would have given the same result!), to obtain an accurate estimate for $E$: $$ E \approx 7,14 \times 10^{20} \text{ ergs} = 7,14 \times 10^{13} J, $$ which corresponds to the energy release of about $16 800$ tons of TNT. \et \section{Groups and differential equations} We now want to start applying symmetry methods to differential equations. To this end we consider a system ${\mathcal S}$ of differential equations involving $p$ independent variables $x=(x^1,\dots,x^p)$ and $q$ dependent variables $u=(u^1,\dots,u^q)$. Solutions of ${\mathcal S}$ will be functions $u=f(x)$, or, in components, $$ u^\alpha = f^\alpha(x^1,\dots,x^p) \quad (\alpha=1,\dots, q). $$ Henceforth we will use the convention of using Latin indices for the independent variables and Greek indices for the dependent ones. We let $X=\mathbb R^p$ be the space of independent variables and $U=\mathbb R^q$ the space of dependent variables. Then basically a symmetry group of ${\mathcal S}$ will be a local transformation group $G$ acting on $M:=X\times U$ in such a way that it `transforms solutions of ${\mathcal S}$ into solutions of ${\mathcal S}$. We first need to clarify how such a group of transformations is to act on a function $f$. The key idea here is to identify $f\colon X\supseteq \Omega \to U$ with its graph $$ \Gamma_f = \{(x,f(x))\mid x\in\Omega\}. $$ Then $\Gamma_f$ is a $p$-dimensional submanifold of $M$. If $g\in G$ and $\Gamma_f\subseteq {\mathcal U}_g$, the domain of $g$, then $g$ acts on $\Gamma_f$ by $$ g\cdot \Gamma_f := \{({\tilde x},{\tilde u}) = g\cdot (x,u)\mid (x,u)\in \Gamma_f\}. $$ Note that $g\cdot \Gamma_f$ will in general no longer be the graph of a function. However, since $G$ acts smoothly and $e\cdot \Gamma_f = \Gamma_f$, by shrinking $\Omega$ and restricting to elements of $G$ near to $e$ we can always achieve that $g\cdot \Gamma_f = \Gamma_{{\tilde f}}$, where ${\tilde f}$ (corresponding to ${\tilde u}={\tilde f}({\tilde x})$) is a well-defined function. In this case we write ${\tilde f} = g\cdot f$ and call ${\tilde f}$ the transform of $f$ by $g$. \begin{thr}{\bf Example. }\rm\label{rotact2} Let $X=\mathbb R$, $U=\mathbb R$, so $p=q=1$ and again consider the rotation group $G=SO(2)$ on $M=X\times U=\mathbb R^2$. The action of $G$ is given by \begin{equation}\label{rotact} ({\tilde x}, {\tilde u}) = \varepsilon\cdot (x,u) = (x\cos\varepsilon -u\sin\varepsilon,x\sin\varepsilon + u\cos\varepsilon) \end{equation} Now if $f\colon x\mapsto u=f(x)$ is a function then $G$ acts on $f$ by rotating its graph, and in general this rotated graph will not be the graph of a well-defined function: think of a straight line or a parabola. The example of the parabola in particular demonstrates that it will in general not suffice to choose $g$ near $e$ (i.e., $\varepsilon$ near $0$) but that one has to also restrict the domain of $f$. But for $|\varepsilon|$ small and restricting $f$ to a sufficiently small domain we do obtain a well defined function $\varepsilon\cdot f = \tilde f$. To explicitly calculate the action of $G$ on a linear function, let $u=f(x)=ax+b$. Then any point in $\Gamma_f$ is of the form $(x,ax+b)$, which is rotated by $\varepsilon$ to the point $$ ({\tilde x}, {\tilde u}) = (x\cos\varepsilon - (ax+b)\sin\varepsilon,x\sin\varepsilon + (ax+b)\cos\varepsilon). $$ To determine ${\tilde u} ={\tilde f}({\tilde x})$ we have to eliminate $x$ from this equation, which is possible for $\varepsilon$ small (so that $\cot\varepsilon \not=a$). Then $$ x=\frac{{\tilde x} + b\sin\varepsilon}{\cos\varepsilon - a\sin\varepsilon}, $$ and therefore $$ {\tilde u} = {\tilde f}({\tilde x}) = \frac{\sin \varepsilon + a\cos\varepsilon}{\cos\varepsilon - a\sin\varepsilon}{\tilde x} + \frac{b}{\cos\varepsilon - a\sin\varepsilon}, $$ which, as expected, is again a linear function. \et The general procedure for calculating ${\tilde f}=g\cdot f$ from $f$ is as follows. Write \begin{equation}\label{xiphidef} ({\tilde x},{\tilde u}) = g\cdot (x,u) = (\Xi_g(x,u),\Phi_g(x,u)), \end{equation} with $\Xi_g,\Phi_g$ smooth. The graph $\Gamma_{\tilde f} = g\cdot \Gamma_f$ of ${\tilde f} = g\cdot f$ is then given parametrically by the equations (for $x\in\Omega$) \begin{equation* \begin{split} {\tilde x} = \Xi_g(x,f(x)) &= \Xi_g\circ (\mathrm{id}\times f)(x)\\ {\tilde u} = \Phi_g(x,f(x)) &= \Phi_g\circ (\mathrm{id}\times f)(x) \end{split} \end{equation*} To calculate ${\tilde f}$ we have to eliminate $x$ from these equations. For $g=e$, $\Xi_e\circ (\mathrm{id}\times f)=\mathrm{id}$, hence for $g$ in a neighborhood of $e$ the Jacobian of $\Xi_g\circ (\mathrm{id}\times f)$ is nonsingular and so by the inverse function theorem we can locally solve for $x$: $$ x=[\Xi_g\circ (\mathrm{id}\times f)]^{-1}({\tilde x}). $$ Consequently, \begin{equation}\label{gdotf} g\cdot f = [\Phi_g\circ (\mathrm{id}\times f)]\circ [\Xi_g\circ (\mathrm{id}\times f)]^{-1}. \end{equation} \blem\label{grcomp} Let $f$ be a local smooth map, $f\colon \Omega\to U$, and let $g$, $h\in G$. Then $h\cdot (g\cdot f) = (h\cdot g)\cdot f$, wherever defined. \et \pr Since any function is uniquely determined by its graph it suffices to note that $$ \Gamma_{(h\cdot g)\cdot f} = (h\cdot g) \Gamma_{f} = h\cdot (g\cdot \Gamma_{f}) = h\cdot \Gamma_{g\cdot f} = \Gamma_{h\cdot (g\cdot f)}. $$ \ep An important special case where ${\tilde f}$ is automatically defined unrestrictedly is that of a {\em projectable}\index{group action!projectable} group action. For such actions, $\Xi_g$ is a function of $x$ only, i.e., $$ ({\tilde x},{\tilde u}) = g\cdot (x,u) = (\Xi_g(x),\Phi_g(x,u)), $$ and so ${\tilde f}({\tilde x}) = [\Phi_g\circ (\mathrm{id}\times f)]\circ \Xi_g^{-1}({\tilde x})$. Using this terminology, we can now define: \begin{thr} {\bf Definition. }\label{dgsgdef} Let ${\mathcal S}$ be a system of differential equations. A symmetry group\index{symmetry group!of differential equation} of ${\mathcal S}$ is a local transformation group $G$ acting on an open subset of $M=X\times U$ such that whenever $f$ is a solution of ${\mathcal S}$ and $g\cdot f$ is defined for some $g\in G$ then also $g\cdot f$ is a solution of ${\mathcal S}$. (Here {\em solution} means any smooth solution defined on any subdomain $\Omega$ of $X$.) \et \begin{thr}{\bf Example. }\rm (i) Let ${\mathcal S}$ consist of the equation $u_{xx}=0$. The solutions are precisely the linear functions on $\mathbb R$. Since $G=SO(2)$ transforms linear functions to linear functions, it is a symmetry group of ${\mathcal S}$. (ii) Let ${\mathcal S}=\{u_t=u_{xx}\}$, the heat equation\index{heat equation}, and consider the group action $$ (x,t,u) \mapsto (x+\varepsilon a,t+\varepsilon b,u) \quad (\varepsilon\in \mathbb R). $$ This is a symmetry group of ${\mathcal S}$ since $f(x-\varepsilon a,t-\varepsilon b)$ is a solution of the heat equation whenever $f$ is. \et One immediate advantage of knowing the symmetry group of a differential equation is that it gives a straightforward way of constructing new solutions from known ones (sometimes even from trivial ones, e.g., constant solutions) simply by applying the group transformations. Our aim ultimately is to derive infinitesimal criteria for a local transformation group to be a symmetry group of a differential equation, which can then be used to explicitly calculate the full symmetry group of any given differential equation. For this purpose we first need to develop some more technical machinery. \section{Prolongation}\index{prolongation} In the previous section we identified any smooth function with its graph in order to define the action of a transformation group on it. Since we are now interested in differential equations we need to find a way of simultaneously considering a function and its derivatives up to a certain order. The mechanism we are going to employ is that of prolongation, a simplified version of the theory of jet bundles. Given a smooth function $f\colon \mathbb R^p\to\mathbb R$, $x\mapsto f(x^1,\dots,x^p)$ and any $k\in\N_0$, we let $J=(j_1,\dots,j_k)$ be an unordered $k$-tuple with $1\le j_l \le p$ for all $l$. The order of this tuple is defined as $\sharp J = k$, and the corresponding partial derivative of $f$ is $$ \partial_Jf(x) = \frac{\partial^k f}{\partial x^{j_1}\partial x^{j_2}\dots \partial x^{j_k}}. $$ There are $$ p_k := \begin{pmatrix} p+k-1\\ k \end{pmatrix} $$ different possible $k$-th order partial derivatives of $f$. To see this, note that since the order of derivatives is irrelevant we may first bring $J$ in ascending order, and then uniquely describe any possible choice by writing $*$ for each number that appears in $J$ and $|$ to signify that we increase the number. For example, if $p=5$ and $k=2$, the string $**||||$ corresponds to the tuple $(1,1)$, $|*|*||$ to $(2,3)$, and $(|||*|*)$ to $(4,5)$, etc. We need $k$ stars and $p-1$ vertical lines, so our problem is equivalent to determining the number of possible selections of $k$ elements from a $p-1+k$-element set, hence the above formula. If $f\colon X\to U$ with $X=\mathbb R^p$ and $U=\mathbb R^q$, so $u=f(x)=(f^1(x),\dots,f^q(x))$, then we need $q\cdot p_k$ numbers $u^\alpha_J = \partial_Jf^\alpha(x)$ to represent all possible $k$-th order derivatives of $f$ at any point $x$. Therefore, we set $U_k:=\mathbb R^{q\cdot p_k}$ and write $u^\alpha_J$, with $\alpha\in \{1,\dots,q\}$ and $J$ any unordered multi-index as above. Moreover, we let $$ U^{(n)} := U\times U_1\times \dots \times U_n. $$ The coordinates of $U^{(n)}$ can be used to represent all the derivatives of any map $f\colon X\to U$ of orders $0$ to $n$. The dimension of $U^{(n)}$ is $$ q+qp_1+\dots + qp_n = q \begin{pmatrix} p+n\\ n \end{pmatrix} =: qp^{(n)}. $$ We will denote elements of $U^{(n)}$ by $u^{(n)}$, with $q\cdot p^{(n)}$ components $u^\alpha_J$, with $\alpha\in \{1,\dots,q\}$ and $J=(j_1,\dots,j_k)$ an unordered multiindex, $1\le j_l\le p$, $0\le k \le n$. For $k=0$ there is only one such multiindex, denoted by $0$ and $u^\alpha_0$ is the component $u^\alpha$ of $u$. \begin{thr}{\bf Example. }\rm Let $p=2$, $q=1$. Then $X=\mathbb R^2$ has coordinates $(x^1,x^2)=(x,y)$, and $U=\mathbb R$ has the coordinate $u$. $U_1$ equals $\mathbb R^2$, with coordinates $(u_x,u_y)$, representing all first order derivatives of $u$. Also, $U_2=\mathbb R^3$ has coordinates $(u_{xx},u_{xy},u_{yy})$, representing all second order derivatives of $u$. In general, $U_k=\mathbb R^{k+1}$ since there are $k+1$ derivatives of $u$ of order $k$, namely $\frac{\partial^k u}{\partial x^i \partial y^{k-i}}$, $i=0,\dots,k$. Also, $U^{(2)} = U\times U_1\times U_2 = \mathbb R^6$, with coordinates $u^{(2)} = (u;u_x,u_y;u_{xx},u_{xy},u_{yy})$, representing all derivatives of $u$ with respect to $x$ and $y$ of order at most $2$. \et Now for any smooth function $f\colon X\to U$ we define its $n$-th prolongation\index{prolongation!of function} $\mathrm{pr}^{(n)} f\colon X\to U^{(n)}$, $u^{(n)}=\mathrm{pr}^{(n)} f(x)$ by $$ u^\alpha_J = \partial_J f^\alpha(x). $$ Thus for any $x$, $\mathrm{pr}^{(n)} f(x)$ is a vector with $q\cdot p^{(n)}$ entries representing the values of $f$ and all its derivatives up to order $n$ at $x$. In this sense we might also identify $\mathrm{pr}^{(n)} f(x)$ with the $n$-th Taylor polynomial of $f$ at $x$. \begin{thr}{\bf Example. }\rm For $p=2$, $q=1$, we have $u=f(x,y)$. Then the second prolongation of $f$, $u^{(2)}=\text{pr}^{(2)}f(x,y)$ is given by $$ (u;u_x,u_y;u_{xx},u_{xy},u_{yy}) = \left(f;\partial_x f,\partial_y f,\partial_x^2 f,\partial_{xy} f, \partial_y^2 f\right), $$ evaluated at $(x,y)$. \et The space $X\times U^{(n)}$ is also called the $n$-th order jet-space\index{jet space} of $X\times U$, and $\mathrm{pr}^{(n)} f(x)$ is called the $n$-jet of $f$ at $x$. If $M$ is an open subset of $X\times U$, then we set $$ M^{(n)} := M\times U_1\times \dots \times U_n. $$ If $u=f(x)$ is a function whose graph lies in $M$ then the $n$-th prolongation $\mathrm{pr}^{(n)} f$ is a function whose graph lies in $M^{(n)}$. \section{Systems of differential equations}\label{syssect} For a system ${\mathcal S}$ of $n$-th order differential equations in $p$ independent and $q$ dependent variables we write $$ P_\nu(x,u^{(n)}) = 0 \quad \nu=1,\dots,l, $$ where $x=(x^1,\dots,x^p)$, $u=(u^1,\dots,u^q)$. Here the functions $$ P(x,u^{(n)})=(P_1(x,u^{(n)}),\dots,P_l(x,u^{(n)})) $$ are assumed to be smooth, i.e., $P\in {\mathcal C}^\infty(X\times U^{(n)},\mathbb R^l)$. The differential equations in ${\mathcal S}$ determine where this map vanishes on $X\times U^{(n)}$, i.e., they determine a corresponding subvariety $$ {\mathcal S}_P = \{(x,u^{(n)})\mid P(x,u^{(n)})=0\}\subseteq X\times U^{(n)} $$ of the total jet space $X\times U^{(n)}$. We may therefore identify the system ${\mathcal S}$ with this set ${\mathcal S}_P$, and we shall do so at many places below. A solution of ${\mathcal S}$ is a smooth function $u=f(x)$ such that $$ P_\nu(x,\mathrm{pr}^{(n)} f(x)) = 0,\quad \nu =1,\dots,l, $$ for all $x$ in the domain of $f$. Geometrically, this means that the graph of the prolongation $\mathrm{pr}^{(n)} f$ is contained in ${\mathcal S}_P$: $$ \Gamma_f^{(n)} \equiv \{(x,\mathrm{pr}^{(n)} f(x))\}\subseteq {\mathcal S}_P = \{P(x,u^{(n)})=0\}. $$ \begin{thr}{\bf Example. }\rm Consider the Laplace equation in two variables: \begin{equation}\label{laplace} u_{xx} + u_{yy} = 0. \end{equation} Here, $p=2$, $q=1$, and $n=2$ since the equation is of second order. The coordinates of $X\times U^{(2)}$ are $(x,y;u;u_x,u_y;u_{xx},u_{xy},u_{yy})$, and \eqref{laplace} describes a hyperplane ${\mathcal S}_P$ in $X\times U^{(2)}$. A function $f$ is a solution of \eqref{laplace} if the graph of $\text{pr}^{(2)}f$ is contained in ${\mathcal S}_P$. For example, let $f(x,y)=x^3-3xy^2$. Then $$ (x,y,\text{pr}^{(2)}f(x,y)) = (x,y;x^3-3xy^2;3x^2-3y^2,-6xy;6x,-6y,-6x) \in {\mathcal S}_P. $$ \et \section{Prolongation of group actions} In this section, given a local transformation group $G$ on an open subset $M$ of $X\times U$, we will construct an induced action $\mathrm{pr}^{(n)} G$, the $n$-th prolongation of $G$, on $M^{(n)}$. The idea is to define $\mathrm{pr}^{(n)} G$ in such a way that it transforms the derivatives of any smooth function $u=f(x)$ into the corresponding derivatives of the transformed function ${\tilde u} = {\tilde f}({\tilde x})$. Thus let $(x_0,u^{(n)}_0)\in M^{(n)}$ and choose any smooth function $u=f(x)$ defined in a neighborhood of $x_0$ whose graph lies in $M$ and such that it has the given derivatives at $x_0$, i.e., such that $u^{(n)}_0=\mathrm{pr}^{(n)} f(x_0)$. To see that such an $f$ indeed exists, one may take it to be an appropriate Taylor polynomial: \begin{equation}\label{taylor} f^\alpha(x) := \sum_J \frac{u^\alpha_{J0}}{\tilde J!}(x-x_0)^J, \quad \alpha=1,\dots,q. \end{equation} Here the sum is over all $J=(j_1,\dots,j_k)$ with $0\le k\le n$, and $$ (x-x_0)^J:=(x^{j_1}-x_0^{j_1})(x^{j_2}-x_0^{j_2})\dots (x^{j_k}-x_0^{j_k}). $$ Also, for given $J$ we set $\tilde J:=(\tilde j_1,\dots,\tilde j_p)$, where $\tilde j_i$ equals the number of $j_\kappa$'s which equal $i$. For example, if $J=(1,1,1,2,4,4)$, $p=4$, $k=6$, then $\tilde J = (3,1,0,2)$. Finally, $\tilde J = \tilde j_1!\dots\tilde j_p!$. If $g$ is close to $e$ then the transformed function ${\tilde f} = g\cdot f$ (as in \eqref{gdotf}) is defined in a neighborhood of $({\tilde x}_0,{\tilde u}_0)=g\cdot (x_0,u_0)$, where $u_0=f(x_0)$. We then define $$ \mathrm{pr}^{(n)} g\cdot (x_0,u^{(n)}_0) = ({\tilde x}_0,{\tilde u}_0^{(n)}), $$ where \begin{equation}\label{groupprol} \tilde u^{(n)}_0 = \mathrm{pr}^{(n)}(g\cdot f)({\tilde x}_0). \end{equation} An important point to note here is that, due to the chain rule, the value of $\mathrm{pr}^{(n)} g\cdot (x_0,u^{(n)}_0)$ depends exclusively on the derivatives of $f$ at the point $x_0$ up to order $n$, i.e., only on $(x_0,u^{(n)}_0)$ itself. It is therefore independent of the choice of the auxiliary function $f$ from above and therefore provides a well-defined operation on $M^{(n)}$. \blem\label{prolloctrans} For any local transformation group $G$ on $M$ as above, $\mathrm{pr}^{(n)} G$ is a local transformation group on $M^{(n)}$. \et \pr We first note that $(g,(x_0,u^{(n)}_0))\mapsto \mathrm{pr}^{(n)} g\cdot (x_0,u^{(n)}_0)$ is smooth. Indeed, taking the $f$ from \eqref{taylor} this follows directly from \eqref{gdotf}. To see this, it suffices to note that the inverse function theorem yields an inverse that automatically depends smoothly on all parameters that the original function depends upon (in our case, the $u^{(n)}_0$): if $f=f(x,\alpha)$ ($\alpha$ representing any parameters), then $f^{-1}$ is the solution of the implicit equation $F(x,y,\alpha) = f(y,\alpha) - x = 0$, so the implicit function theorem yields the claim. It remains to verify (i) and (ii) from \ref{ltgdef2}. Here, (ii) is automatic because for $g=e$ we get $g\cdot f = f$. Thus we are left with proving that \begin{equation}\label{prolcomp} \mathrm{pr}^{(n)} h (\mathrm{pr}^{(n)} g \cdot (x_0,u^{(n)}_0)) = \mathrm{pr}^{(n)}(h\cdot g)(x_0,u^{(n)}_0). \end{equation} To see this, let $(x_0,u^{(n)}_0)\in M^{(n)}$ and pick a smooth function $f$ defined near $x_0$ with $u^{(n)}_0=\mathrm{pr}^{(n)} f(x_0)$. Then $$ ({\tilde x}_0,{\tilde u}_0^{(n)}) = g \cdot (x_0,u^{(n)}_0) = (\tilde x_0,\mathrm{pr}^{(n)}(g\cdot f)({\tilde x}_0)). $$ To determine the left hand side of \eqref{prolcomp} we have to find any smooth function $F$ such that $\mathrm{pr}^{(n)} F({\tilde x}_0)=\mathrm{pr}^{(n)}(g\cdot f)({\tilde x}_0)$, calculate $h\cdot F$ and then determine $\mathrm{pr}^{(n)} h\cdot F$ at the corresponding point. Naturally, we take $F:=g\cdot f$. Then using the terminology from \eqref{xiphidef} we obtain \begin{equation* \mathrm{pr}^{(n)} h (\mathrm{pr}^{(n)} g \cdot (x_0,u^{(n)}_0)) = (\Xi_h ({\tilde x}_0,(g\cdot f)({\tilde x}_0)), \mathrm{pr}^{(n)}(h\cdot F)(\Xi_h ({\tilde x}_0,(g\cdot f)({\tilde x}_0)))) \end{equation*} Here, \begin{equation*} \begin{split} \Xi_h ({\tilde x}_0,(g\cdot f)({\tilde x}_0)) &= \Xi_h(g\cdot(x_0,u_0)) = \mathrm{proj_1}(h\cdot (g\cdot(x_0,u_0)))\\ &= \mathrm{proj_1}((h\cdot g)\cdot(x_0,u_0)) = \Xi_{h\cdot g}(x_0,u_0), \end{split} \end{equation*} and, by \ref{grcomp}, $h\cdot F = h\cdot (g\cdot f) = (h\cdot g)\cdot f$, so altogether we get \begin{equation*} \begin{split} \mathrm{pr}^{(n)} h (\mathrm{pr}^{(n)} g \cdot (x_0,u^{(n)}_0)) &= (\Xi_{h\cdot g}(x_0,u_0), \mathrm{pr}^{(n)}((h\cdot g) f)(\Xi_{h\cdot g}(x_0,u_0)))\\ &= \mathrm{pr}^{(n)}(h\cdot g)(x_0,u^{(n)}_0). \end{split} \end{equation*} \ep \begin{thr}{\bf Example. }\rm\label{rotact3} Again let $G=SO(2)$ be the rotation group on $M=\mathbb R^2=X\times U$. We determine the first prolongation of $G$. We have $p=q=1$, so $X\times U^{(1)}=\mathbb R^3$, with coordinates $(x,u,u_x)$. If $u=f(x)$ is a local smooth function then $\mathrm{pr}^{(1)} f(x) = (f(x),f'(x))$. Now let $(x^0,u^0,u_x^0)\in X\times U^{(1)}$ and let $\varepsilon$ be an angle of rotation. We wish to determine $$ \mathrm{pr}^{(1)} \varepsilon \cdot (x^0,u^0,u_x^0) = (\tilde x^0,\tilde u^0,\tilde u_x^0). $$ To this end, let $f$ be the Taylor polynomial $$ f(x) = u^0 + u_x^0(x-x^0) = u^0_x\cdot x +(u^0-u^0_x x^0), $$ which satisfies $f(x^0)=u^0$, $f'(x^0)=u^0_x$. Then by \ref{rotact2} we get $$ {\tilde f}({\tilde x}) = (\varepsilon\cdot f)({\tilde x}) = \frac{\sin\varepsilon + u^0_x\cos\varepsilon}{\cos\varepsilon - u^0_x\sin\varepsilon}{\tilde x} + \frac{u^0-u^0_x x^0}{\cos\varepsilon - u^0_x\sin\varepsilon}, $$ which is well-defined for $u^0_x\not=\cot\varepsilon$. By \eqref{rotact} we have ${\tilde x}^0 = x^0\cos\varepsilon - u^0\sin\varepsilon$, so either again by \eqref{rotact} or by inserting we get $$ {\tilde u}^0 = {\tilde f}({\tilde x}^0) = x^0\sin\varepsilon + u^0\cos\varepsilon. $$ Finally, we get for the first derivative $$ {\tilde u}^0_x = {\tilde f}'({\tilde x}^0) = \frac{\sin\varepsilon + u^0_x\cos\varepsilon}{\cos\varepsilon - u^0_x\sin\varepsilon}. $$ Combining this and dropping the $0$-superscripts we obtain the action of $\mathrm{pr}^{(1)} SO(2)$ on $X\times U^{(1)}$: $$ \mathrm{pr}^{(1)}\varepsilon\cdot (x,u,u_x) = \left(x\cos\varepsilon - u\sin\varepsilon,x\sin\varepsilon + u\cos\varepsilon, \frac{\sin\varepsilon + u_x\cos\varepsilon}{\cos\varepsilon - u_x\sin\varepsilon}, \right), $$ defined for $|\varepsilon|<|\mathrm{arccot}(u_x)|$. The important point to note is that although the action of $SO(2)$ is linear and globally defined, its first prolongation is nonlinear and only locally defined. \et In the previous example we see that the first prolongation of $G$ acts on the original variables $(x,u)$ in exactly the same way as $G$ itself. This is actually a general phenomenon. In fact, it follows directly from the definition of $\mathrm{pr}^{(n)} G$ via evaluating derivatives of ${\tilde f}$ that the action of $\mathrm{pr}^{(n)} g$ on $(x,u^{(k)})$ with $k\le n$ coincides with that of $\mathrm{pr}^{(k)}$. In particular, $\mathrm{pr}^{(0)}G=G$. To give a concise formulation of this property we introduce the natural projections \begin{equation*} \begin{split} \pi^n_k: M^{(n)} &\to M^{(k)}\\ \pi^n_k(x,u^{(n)}) &= (x,u^{(k)}), \end{split} \end{equation*} where (for $k\le n$), $u^{(k)}$ consists of the components $u^\alpha_J$, $\sharp J\le k$ of $u^{(n)}$. For example, if $p=2$, $q=1$, then \begin{equation*} \begin{split} \pi^2_0(x,y;u;u_x,u_y;u_{xx},u_{xy},u_{yy}) &= (x,y;u)\\ \pi^2_1(x,y;u;u_x,u_y;u_{xx},u_{xy},u_{yy}) &= (x,y;u;u_x,u_y). \end{split} \end{equation*} Then for any $k\le n$ and any $g\in G$ we have \begin{equation}\label{graprol} \pi^n_k\circ \mathrm{pr}^{(n)} g = \mathrm{pr}^{(k)} g. \end{equation} In particular, this observation allows to calculate prolongations of group actions inductively. \section{Invariance of differential equations} In Section \ref{syssect} we have identified any given system ${\mathcal S}$ of differential equations $P_\nu(x,u^{(n)})=0$ ($\nu=1,\dots,l$) with a subvariety ${\mathcal S}_P$ of $M^{(n)}$. This identification will allow us to utilize the methods of symmetry group analysis of algebraic equations in the context of differential equations. The essential translation result between the two realms is the following theorem \bt\label{maintrans} Let $M$ be an open subset of $X\times U$ and let ${\mathcal S}$: $$ P_\nu(x,u^{(n)})=0\quad \nu=1,\dots,l $$ be an $n$-th order system of differential equations defined on $M$, with corresponding subvariety ${\mathcal S}_P$ of $M^{(n)}$. Suppose that $G$ is a local transformation group acting on $M$ whose prolongation leaves ${\mathcal S}_P$ invariant, i.e., $$ (x,u^{(n)})\in {\mathcal S}_P \Rightarrow \mathrm{pr}^{(n)} g\cdot (x,u^{(n)})\in {\mathcal S}_P, $$ whenever defined. Then $G$ is a symmetry group of the system ${\mathcal S}$ in the sense of \ref{dgsgdef}. \et \pr Let $u=f(x)$, $f\colon \Omega\to U$ be a local solution of $P(x,u^{(n)})=0$. This means that the graph $$ \Gamma^{(n)}_f = \{(x,\mathrm{pr}^{(n)} f(x))\mid x\in \Omega\} $$ of $\mathrm{pr}^{(n)} f$ is contained in ${\mathcal S}_P$. If $g\in G$ is such that $g\cdot f$ is well-defined then, by the very definition of $\mathrm{pr}^{(n)} g$, the graph of $\mathrm{pr}^{(n)} (g\cdot f)$ is given by the image of the graph of $\mathrm{pr}^{(n)} f$ under $\mathrm{pr}^{(n)} g$, i.e., $$ \Gamma^{(n)}_{g\cdot f} = (\mathrm{pr}^{(n)} g)\cdot \Gamma^{(n)}_f. $$ Since ${\mathcal S}_P$ is invariant under $\mathrm{pr}^{(n)} g$, it follows from this that the graph of $\mathrm{pr}^{(n)}(g\cdot f)$ is contained in ${\mathcal S}_P$ as well. But this just means that $g\cdot f$ is a local solution of the system ${\mathcal S}$. \ep \section{Prolongation of vector fields} To make the calculation of symmetry groups of differential equations accessible to the infinitesimal methods we developed for algebraic equations we need to be able to effectively calculate the infinitesimal generators of prolonged group actions. \begin{thr} {\bf Definition. }\label{vprolo} Let $M\subseteq X\times U$ be open and suppose that $v$ is the infinitesimal generator of a local one-parameter transformation group $\varepsilon\mapsto \mathrm{Fl}^v_\varepsilon$. Then the $n$-th prolongation of $v$, denoted by $\mathrm{pr}^{(n)} v$, is the local vector field on $M^{(n)}$ which is the infinitesimal generator of the corresponding prolonged one-parameter group $\mathrm{pr}^{(n)} \mathrm{Fl}^v_\varepsilon$: $$ \mathrm{pr}^{(n)} v|_{(x,u^{(n)})} = \left.\frac{d}{d \eps}\right|_0 \mathrm{pr}^{(n)}(\mathrm{Fl}^v_\varepsilon)(x,u^{(n)}), $$ for any $(x,u^{(n)})\in M^{(n)}$. \et Note that from this definition, together with \ref{flowaction2} it immediately follows that \begin{equation}\label{flowstuff} \mathrm{pr}^{(n)}(\mathrm{Fl}^v_\varepsilon) = \mathrm{Fl}^{\mathrm{pr}^{(n)} v}_{\varepsilon}. \end{equation} Any vector field $v$ on $M=X\times U$ can be written in the form $$ v|_{(x,u)} = \sum_{i=1}^p \xi^i(x,u)\partial_{x^i} + \sum_{\alpha=1}^q \phi_\alpha(x,u)\partial_{u^\alpha}. $$ Its $n$-th prolongation, being a vector field on $M^{(n)}$, therefore has to be of the form \begin{equation}\label{prolv} \mathrm{pr}^{(n)} v|_{(x,u^{(n)})} = \sum_{i=1}^p \xi^i(x,u)\partial_{x^i} + \sum_{\alpha=1}^q \sum_J\phi_\alpha^J(x,u^{(n)})\partial_{u^\alpha_J}, \end{equation} where the last sum is over all $J$ with $0\le \sharp J\le n$. Note that due to \eqref{graprol} the prolonged group action agrees with the original one on $M^{(0)}=M$, so the coefficients $\xi^i$ and $\phi^0_\alpha=\phi_\alpha$ agree with those of $v$. Moreover, for the same reason, if $\sharp J = k$ then the coefficient $\phi^J_\alpha$ of $\partial_{u^\alpha_J}$ will only depend on $k$-th and lower order derivatives of $u$, $\phi^J_\alpha = \phi^J_\alpha(x,u^{(k)})$, i.e., \begin{equation}\label{projproj} T\pi^n_k(\mathrm{pr}^{(n)} v) = \mathrm{pr}^{(k)}v \quad n\ge k, \end{equation} where $\mathrm{pr}^{(0)}=v$. Again this reflects the possibility of calculating the coefficients $\phi^J_\alpha$ recursively. Our main goal will be to derive a general formula for calculating the $\phi^J_\alpha$ given $\phi_\alpha$. \begin{thr}{\bf Example. }\rm\label{rotact4} Returning once more to the rotation group $G=SO(2)$ on $\mathbb R^2$ from \ref{rotact2}, \ref{rotact3}, with infinitesimal generator $v = -u\partial_x + x\partial_u$ and group action \begin{equation*} \mathrm{Fl}^v_\varepsilon(x,u) = \varepsilon\cdot (x,u) = (x\cos\varepsilon -u\sin\varepsilon,x\sin\varepsilon + u\cos\varepsilon) \end{equation*} we have $$ \mathrm{pr}^{(1)}[\mathrm{Fl}^v_\varepsilon] (x,u,u_x) = \left(x\cos\varepsilon - u\sin\varepsilon,x\sin\varepsilon + u\cos\varepsilon, \frac{\sin\varepsilon + u_x\cos\varepsilon}{\cos\varepsilon - u_x\sin\varepsilon}, \right), $$ By \ref{vprolo}, the first prolongation of $v$ is calculated by differentiating this expression at $\varepsilon=0$. This gives \begin{equation}\label{rotvprol} \mathrm{pr}^{(1)} v = -u\partial_x + x\partial_u + (1+u_x^2)\partial_{u_x} \end{equation} As predicted by \eqref{projproj}, the first two terms coincide with those of $v$ itself. \et In order to be able to apply the criterion \ref{algsym} to the present situation we need a suitable maximal rank condition for systems of differential equations. \begin{thr} {\bf Definition. } A system $$ P_\nu(x,u^{(n)}) = 0 \quad \nu=1,\dots,l $$ of differential equations is said to be of maximal rank\index{maximal rank} if the $l\times(p+qp^{(n)})$ Jacobian matrix $$ J_P(x,u^{(n)}) = \left(\frac{\partial P_\nu}{\partial x^i},\frac{\partial P_\nu}{\partial u^\alpha_J}\right) $$ of $P$ with respect to all its variables $(x,u^{(n)})$ is of rank $l$ whenever $P(x,u^{(n)})=0$. \et \begin{thr}{\bf Example. }\rm (i) The $2$-dimensional Laplace equation $P=u_{xx}+u_{yy}=0$ is of maximal rank since the Jacobian of $P$ with respect to $(x,y;u;u_x,u_y;u_{xx},u_{xy},u_{yy})\in X\times U^{(2)}$ is $$ J_P=(0,0;0;0,0;1,0,1), $$ which has rank $1$ everywhere. (ii) The equation $P=(u_{xx}+u_{yy})^2=0$ is not of maximal rank, because $$ J_P=(0,0;0;0,0;2(u_{xx}+u_{yy}),0,2(u_{xx}+u_{yy})) $$ vanishes on $P=0$. \et \begin{thr} {\bf Remark. }\rm In practice, the maximal rank condition is not much of a restriction, i.e., `most' systems automatically satisfy it. In fact, if ${\mathcal S}_P=\{(x,u^{(n)})\mid P(x,u^{(n)})=0\}$ is a regular ($l$-dimensional) submanifold of $M^{(n)}$ then locally around any point in $M^{(n)}$ we can pick an adapted coordinate system (cf.\ \cite[1.1.12]{DG2}), and thereby (extracting the relevant coordinates) find a smooth map $\tilde P$ of rank $l$ such that ${\mathcal S}_P={\mathcal S}_{\tilde P} = \{(x,u^{(n)})\mid \tilde P(x,u^{(n)})=0\}$. \et Using the above condition we can now formulate the main result on the calculation of symmetry groups of systems of differential equations: \bt\label{mainpde} Let $$ P_\nu(x,u^{(n)}) = 0 \quad \nu=1,\dots,l $$ be a system of differential equations of maximal rank defined on an open subset $M$ of $X\times U$. If $G$ is a local transformation group acting on $M$ and \begin{equation}\label{mainsymcrit} \mathrm{pr}^{(n)} v [P_\nu(x,u^{(n)})] = 0, \quad \nu=1,\dots,l, \text{ whenever } P(x,u^{(n)})=0 \end{equation} for every infinitesimal generator $v$ of $G$, then $G$ is a symmetry group of the system. \et \pr By \ref{maintrans} it suffices to show that ${\mathcal S}_P$ remains invariant under $\mathrm{pr}^{(n)} G$. By \ref{prolloctrans}, $\mathrm{pr}^{(n)} G$ is a local group of transformations on $M^{(n)}$, whose infinitesimal generators are exactly the $\mathrm{pr}^{(n)} v$, for $v$ the infinitesimal generators of $G$ (by \ref{vprolo}). Since $P$ is of maximal rank, the result now follows from \ref{algsym}. \ep We shall see in \ref{nesu} that if $P$ satisfies a certain local solvability condition then \eqref{mainsymcrit} is in fact necessary and sufficient for $G$ to be a symmetry group of the system. The Hadamard Lemma \ref{hadamard} allows the following equivalent reformulation of \eqref{mainsymcrit}: there exist smooth functions $Q_{\nu\mu}=Q_{\nu\mu}(x,u^{(n)})$ such that \begin{equation}\label{mainsymhadamard} \mathrm{pr}^{(n)} v [P_\nu(x,u^{(n)})] = \sum_{\mu=1}^l Q_{\nu\mu}(x,u^{(n)}) P_\mu(x,u^{(n)}) \end{equation} holds identically on $M^{(n)}$. \begin{thr}{\bf Example. }\rm\label{rotact5} Continuing our analysis of the rotation group $SO(2)$ on $\mathbb R^2$ from \ref{rotact4}, consider the first order ODE \begin{equation}\label{rotode} P(x,u,u_x) = (u-x)u_x + u +x =0. \end{equation} The Jacobian of $P$ is $$ J_P=(\partial_x P,\partial_u P,\partial_{u_x} P)=(1-u_x,1+u_x,u-x), $$ hence has rank $1$ everywhere. \et We now apply the infinitesimal generator $\mathrm{pr}^{(1)} v$ from \eqref{rotvprol} to $P$: \begin{equation*} \begin{split} \mathrm{pr}^{(1)} v(P) &= -u\partial_xP + x\partial_u P + (1+u_x^2)\partial_{u_x}P\\ &= -u(1-u_x) + x(1+u_x) + (1+u_x)^2(u-x)\\ &= u_x[(u-x)u_x + u +x] = u_x P. \end{split} \end{equation*} Therefore, \eqref{mainsymcrit} (or also \eqref{mainsymhadamard}) is satisfied, and we conclude that the rotation group transforms solutions of \eqref{rotode} to other solutions. The calculation of even the first prolongation of the infinitesimal generator of the rotation group in \ref{rotact3} demonstrates that the direct method of first determining the prolonged group action and then calculating the corresponding generators is not feasible in practice. We need to find an algorithmic way of directly calculating prolongations of vector fields, avoiding the detour of determining the prolonged group action along the way. Before deriving a general formula, let us start out by considering some special cases first. Let $$ v = \sum_{i=1}^p \xi^i(x)\partial_{x^i} $$ on $M\subseteq X\times U$, with $U=\mathbb R$. According to \eqref{xiphidef}, the corresponding group action $g_\varepsilon = \mathrm{Fl}^v_\varepsilon$ then is of the form $$ ({\tilde x},{\tilde u}) = g_\varepsilon\cdot (x,u) = (\Xi_\varepsilon(x),u), $$ where \begin{equation}\label{xiabl} \left.\frac{d}{d \eps}\right|_0 \Xi^i_\varepsilon(x) = \xi^i(x). \end{equation} On $M^{(1)}$ we have coordinates $(x,u^{(1)}) =(x^i,u,u_j)$ with $u_j:=\partial_{x^j}u$. By definition, to calculate $\mathrm{pr}^{(1)} v$ in $(x,u^{(1)})$, we may take any function $f$ with $u=f(x)$ and $u_j=\partial_{x^j}f(x)$ for $j=1,\dots,p$. Then $$ \mathrm{pr}^{(1)} g_\varepsilon\cdot (x,u^{(1)}) = ({\tilde x},{\tilde u}^{(1)}), $$ where ${\tilde x} = \Xi_\varepsilon(x)$, ${\tilde u} = u$, and ${\tilde u}_j$ are the derivatives of the transformed function ${\tilde f}_\varepsilon = g_\varepsilon\cdot f$, which, by \eqref{gdotf}, is given by $$ {\tilde u} = {\tilde f}_\varepsilon({\tilde x}) = f(\Xi_\varepsilon^{-1}({\tilde x})) = f(\Xi_{-\varepsilon}({\tilde x})). $$ Therefore, \begin{equation}\label{utildiff} {\tilde u}_j = \pd{{\tilde f}_\varepsilon}{{\tilde x}^j}({\tilde x}) = \sum_{k=1}^p\pd{f}{x^k}(\Xi_{-\varepsilon}({\tilde x}))\cdot \pd{\Xi^k_{-\varepsilon}}{{\tilde x}^j}({\tilde x}). \end{equation} Since $\Xi_{-\varepsilon}({\tilde x})=x$, this simplifies to $$ {\tilde u}_j = \sum_{k=1}^p \pd{\Xi^k_{-\varepsilon}}{{\tilde x}^j}(\Xi_\varepsilon(x)) u_k. $$ This is the explicit formula for the action of $g_\varepsilon$ on the $u_j$-variables. To find the infinitesimal generator of $\mathrm{pr}^{(1)} g_\varepsilon$, we have to differentiate this expression with respect to $\varepsilon$ at $\varepsilon=0$. Also, since $\mathrm{pr}^{(1)} g_\varepsilon$ acts on $(x,u)$ as $g_\varepsilon$, the coefficients of $\partial_{x^i}$ and $\partial_{u}$ in $\mathrm{pr}^{(1)} v$ stay the same. Hence \begin{equation}\label{pr1v} \mathrm{pr}^{(1)} v = \sum_{i=1}^p\xi^i(x)\partial_{x^i} + \sum_{j=1}^p \phi^j(x,u^{(1)})\partial_{u_j}, \end{equation} where $$ \phi^j(x,u^{(1)}) = \left.\frac{d}{d \eps}\right|_0 \sum_{k=1}^p \pd{\Xi^k_{-\varepsilon}}{{\tilde x}^j}(\Xi_\varepsilon(x)) u_k. $$ Here, using \eqref{xiabl}, we have \begin{equation*} \left.\frac{d}{d \eps}\right|_0\left[\pd{\Xi^k_{-\varepsilon}}{{\tilde x}^j}(\Xi_\varepsilon(x))\right] = -\pd{\xi^k}{x^j}(x) + \sum_{l=1}^p \left[\frac{\partial^2 \Xi^k_{-\varepsilon}}{\partial {\tilde x}^j \partial {\tilde x}^l}(\Xi_\varepsilon(x)) \frac{d}{d\varepsilon}\Xi^l_\varepsilon(x)\right]_{\varepsilon=0}, \end{equation*} and the second term vanishes since $\Xi_0=\mathrm{id}$. Therefore, \begin{equation}\label{phij} \phi^j(x,u,u_x) = - \sum_{k=1}^p \pd{\xi^k}{x^j} u_k. \end{equation} The second special case we look at is a local group action as above, only this time acting exclusively on the dependent variable: $$ ({\tilde x}, {\tilde u}) =g_\varepsilon\cdot(x,u) = (x,\Phi_\varepsilon(x,u)). $$ In this case, $v=\phi(x,u)\partial_u$, where $$ \phi(x,u) = \left.\frac{d}{d \eps}\right|_0 \Phi_\varepsilon(x,u). $$ Let $f$ be a local smooth function with $f(x)=u$. Then ${\tilde f}_\varepsilon=g_\varepsilon\cdot f$ is given by \begin{equation}\label{fepss} {\tilde u} = {\tilde f}_\varepsilon(x)=\Phi_\varepsilon(x,f(x)). \end{equation} To determine the prolonged group action, we need to differentiate ${\tilde f}_\varepsilon$: $$ {\tilde u}_j=\pd{{\tilde f}_\varepsilon}{x^j}(x) = \pd{\Phi_\varepsilon}{x^j}(x,f(x)) + \pd{\Phi_\varepsilon}{u}(x,f(x))\pd{f}{x^j}(x). $$ Consequently, $\mathrm{pr}^{(1)} g_\varepsilon\cdot(x,u^{(1)}) =(x,{\tilde u}^{(1)})$, where \begin{equation}\label{utilj} {\tilde u}_j = \pd{\Phi_\varepsilon}{x^j} + \pd{\Phi_\varepsilon}{u}u_j \end{equation} As before, to determine $\phi^j$ in $$ \mathrm{pr}^{(1)} v = v + \sum_{j=1}^p \phi^j(x,u^{(1)})\partial_{u_j}, $$ we have to differentiate \eqref{utilj} at $\varepsilon=0$: \begin{equation}\label{pr1spec} \phi^j(x,u^{(1)}) = \left.\frac{d}{d \eps}\right|_0 {\tilde u}_j = \pd{\phi}{x^j} + u_j\pd{\phi}{u}. \end{equation} We have here the first instance of a generally useful operation, namely that of total derivative\index{total derivative}: $$ \phi^j(x,u^{(1)}) = \pd{}{x^j}[\phi(x,f(x))]. $$ Thus $\phi^j(x,u^{(1)})$ is calculated from $\phi(x,u)$ by differentiating with respect to $x^j$ while treating $u$ as a function of $x$. Accordingly, the total derivative of $\phi$ with respect to $x^j$, $D_j$ is defined as $$ D_j\phi := \pd{\phi}{x^j} + u_j\pd{\phi}{u}. $$ More generally, we define: \begin{thr} {\bf Definition. } Let $(x,u^{(n)})\mapsto P(x,u^{(n)})$ be a smooth function of $x$, $u$, and derivatives of $u$ up to order $n$, defined on an open subset $M^{(n)}$ of $X\times U^{(n)}$. The total derivative\index{total derivative} of $P$ with respect to $x^i$ is the unique smooth function $D_iP(x,u^{(n+1)})$ defined on $M^{(n+1)}$ with the property that if $u=f(x)$ is any smooth function of $x$ then $$ D_iP(x,\mathrm{pr}^{(n+1)} f(x)) = \pd{}{x^i}[P(x,\mathrm{pr}^{(n)} f(x))]. $$ \et Thus $D_iP$ is calculated by differentiating $P$ with respect to $x^i$ while treating all $u^\alpha$ and their derivatives as functions of $x$. \blem For any $P(x,u^{(n)})$ we have \begin{equation}\label{totder1} D_iP = \pd{P}{x^i} + \sum_{\alpha=1}^q \sum_J u^\alpha_{J,i}\pd{P}{u^\alpha_J}, \end{equation} where, for $J=(j_1,\dots,j_k)$, \begin{equation}\label{totder2} u^\alpha_{J,i} = \pd{u^\alpha_J}{x^i} = \pd{^{k+1} u^\alpha}{x^i \partial x^{j_1}\dots \partial x^{j_k}}, \end{equation} and the sum in \eqref{totder1} extends over all $0\le \sharp J \le n$. \et For example, with $X=\mathbb R^2$, $U=\mathbb R$ we have \begin{equation*} \begin{split} D_xP &= \pd{P}{x} + u_x \pd{P}{u} + u_{xx} \pd{P}{u_x} + u_{xy} \pd{P}{u_y} + u_{xxx} \pd{P}{u_{xx}} + \dots\\ D_yP &= \pd{P}{y} + u_y \pd{P}{u} + u_{xy} \pd{P}{u_x} + u_{yy} \pd{P}{u_y} + u_{xxy} \pd{P}{u_{xx}} + \dots \end{split} \end{equation*} Higher order total derivatives are defined inductively: let $J=(j_1,\dots,j_k)$ be a $k$-th order multiindex with $1\le j_\kappa\le p$ for each $\kappa$, then we set $$ D_J := D_{j_1}D_{j_2}\dots D_{j_k}. $$ Also, note that the order of differentiation does not matter for total derivatives: $D_jD_l = D_lD_j$ for all $j$, $l$. In the proof of the general prolongation formula we will need the following auxiliary result: \blem\label{detdiff} Let $\varepsilon\mapsto M(\varepsilon)$ be a smooth map that takes values in the space of invertible $n\times n$-matrices. Then $$ \frac{d}{d\varepsilon}[M(\varepsilon)^{-1}] = -M(\varepsilon)^{-1}\frac{dM(\varepsilon)}{d\varepsilon}M(\varepsilon)^{-1}. $$ \et \pr It suffices to differentiate the identity $M(\varepsilon)M(\varepsilon)^{-1} = I$ by the product rule. \ep \begin{thr} {\bf Remark. }\rm\label{jetsub} As a final prerequisite for the proof of \ref{gpf} below we need the observation that given any open subset $M$ of $X\times U$, one can view the $(n+1)$-st jet space $M^{(n+1)}$ as a subspace of the first jet space $(M^{(n)})^{(1)}$ of the $n$-th jet space. In fact, any $(n+1)$-st derivative $u^\alpha_J$ can be viewed as a first order derivative of an $n$-th order derivative (which, in general, can be done in many different ways). For example, let $p=2$, $q=1$, so that the coordinates on $M^{(1)}$ can be written as $(x,y;u;u_x,u_y)$. Then we may view $(u_x,u_y)$ as new dependent variables, $u_x=v$, $u_y=w$. This turns $M^{(1)}$ into a subset of $X\times \tilde U$, where $X$ is still $\mathbb R^2$, but $\tilde U$ has three dependent variables $u,\,v,\,w$. Consequently, the first jet space $(M^{(1)})^{(1)}$ of $M^{(1)}$ turns into an open subset of $X\times \tilde U^{(1)}$, with coordinates $(x,y;u;v;w;u_x,u_y,v_x,v_y,w_x,w_y)$. Since we defined $v=u_x$, $w=u_y$ it follows that $M^{(2)}\subseteq (M^{(1)})^{(1)}$ is the subspace defined by the relations $$ v=u_x,\quad w=u_y,\quad v_y = w_x $$ in $X\times \tilde U^{(1)}$. Here, the third relation is a consequence of $u_{xy}=u_{yx}$. \et \bt\label{gpf} (The general prolongation formula)\index{prolongation formula} Let \begin{equation}\label{vfform} v = \sum_{i=1}^p \xi^i(x,u)\pd{}{x^i} + \sum_{\alpha=1}^q \phi_\alpha(x,u)\pd{}{u^\alpha} \end{equation} be a vector field on an open subset $M\subseteq X\times U$. The $n$-th prolongation of $v$ is the vector field \begin{equation}\label{theprol} \mathrm{pr}^{(n)} v = v + \sum_{\alpha=1}^q\sum_J \phi^J_\alpha(x,u^{(n)})\pd{}{u^\alpha_J} \end{equation} defined on $M^{(n)} \subseteq X\times U^{(n)}$, where $J$ runs over all multi-indices $J=(j_1,\dots,j_k)$ with $1\le j_\kappa\le p$, $1\le k\le n$. The coefficient functions $\phi_\alpha^J$ are given by \begin{equation}\label{phiform} \phi_\alpha^J(x,u^{(n)}) = D_J\left(\phi_\alpha - \sum_{i=1}^p \xi^iu^\alpha_i\right) + \sum_{i=1}^p \xi^iu^\alpha_{J,i}, \end{equation} where $u^\alpha_i = \partial_{x^i}u^\alpha$ and $u^\alpha_{J,i} = \partial_{x^i}u^\alpha_J$. \et \pr We start out with the case $n=1$. Let $g_\varepsilon = \mathrm{Fl}^v_\varepsilon$ and $$ ({\tilde x},{\tilde u}) = g_\varepsilon\cdot (x,u) = (\Xi_\varepsilon(x,u),\Phi_\varepsilon(x,u)). $$ Then \begin{equation}\label{xiphi} \begin{split} \xi^i(x,u) &= \left.\frac{d}{d \eps}\right|_0 \Xi^i_\varepsilon(x,u), \quad i=1,\dots,p,\\ \phi_\alpha(x,u) &= \left.\frac{d}{d \eps}\right|_0 \Phi^\alpha_\varepsilon(x,u), \quad \alpha=1,\dots,q. \end{split} \end{equation} Given any $(x,u^{(1)})\in M^{(1)}$, pick a local smooth map $f$ with $\mathrm{pr}^{(1)} f(x)=u^{(1)}$, i.e., with $u^\alpha = f^\alpha(x)$, $u^\alpha_i =\partial_{x^i} f^\alpha(x)$. By \eqref{gdotf}, for $\varepsilon$ small and restricting the domain of $f$ sufficiently, we have $$ {\tilde u} = {\tilde f}_\varepsilon({\tilde x}) = (g_\varepsilon\cdot f)({\tilde x}) = [\Phi_\varepsilon\circ (\mathrm{id}\times f)]\circ [\Xi_\varepsilon\circ (\mathrm{id}\times f)]^{-1}({\tilde x}). $$ By the chain rule, we conclude from this that for the Jacobian $J{\tilde f}_\varepsilon = (\partial {\tilde f}_\varepsilon^\alpha/\partial{\tilde x}^i)$ we have (using that $x=[\Xi_\varepsilon\circ (\mathrm{id}\times f)]^{-1}({\tilde x})$) \begin{equation}\label{jac} J{\tilde f}_\varepsilon({\tilde x}) = J[\Phi_\varepsilon\circ (\mathrm{id}\times f)](x)\cdot[J[\Xi_\varepsilon\circ (\mathrm{id}\times f)](x)]^{-1} \end{equation} The matrix entries of $J{\tilde f}_\varepsilon({\tilde x})$ then give formulae for $\mathrm{pr}^{(1)} g_\varepsilon$. To calculate $\mathrm{pr}^{(1)} v$ we have to differentiate \eqref{jac} at $\varepsilon=0$. In doing this, note that since $g_0=\mathrm{id}$, we have \begin{equation}\label{xiphi0} \Xi_0(x,f(x)) = x, \quad \Phi_0(x,f(x)) = f(x), \end{equation} so with $I$ the $p\times p$ unit matrix we get $$ J[\Xi_0\circ (\mathrm{id}\times f)](x) = I,\quad J[\Phi_0\circ (\mathrm{id}\times f)](x) = Jf(x). $$ Therefore, using \ref{detdiff}, \begin{equation*} \begin{split} \left.\frac{d}{d \eps}\right|_0 J{\tilde f}_\varepsilon({\tilde x}) &= \left.\frac{d}{d \eps}\right|_0 J[\Phi_\varepsilon\circ (\mathrm{id}\times f)](x) - Jf(x)\left.\frac{d}{d \eps}\right|_0 J [\Xi_\varepsilon\circ (\mathrm{id}\times f)](x)\\ &= J[\phi\circ (\mathrm{id}\times f)](x) - Jf(x)\cdot J[\xi\circ (\mathrm{id}\times f)](x). \end{split} \end{equation*} The matrix entries of this expression give the coefficient functions $\phi^k_\alpha$ of $\partial_{u^\alpha_k}$ in $\mathrm{pr}^{(1)} v$. The $(\alpha,k)$-th entry is $$ \phi^k_\alpha(x,\mathrm{pr}^{(1)} f(x)) = \pd{}{x^k}[\phi_\alpha(x,f(x))] -\sum_{i=1}^p \pd{f^\alpha}{x^i}\cdot \pd{}{x^k}[\xi^i(x,f(x))], $$ or, in terms of total derivatives, \begin{equation}\label{prol1form} \begin{split} \phi^k_\alpha(x,u^{(1)}) &= D_k\phi_\alpha(x,u^{(1)}) -\sum_{i=1}^p D_k\xi^i(x,u^{(1)})u^\alpha_i\\ &= D_k\left[\phi_\alpha -\sum_{i=1}^p \xi^i u^\alpha_i\right] + \sum_{i=1}^p \xi^i u^\alpha_{ki}, \end{split} \end{equation} with $u^\alpha_{ki}=\partial^2 u^\alpha/\partial x^k\partial x^i$. This is \eqref{phiform} for $n=1$. In the general case we proceed by induction. Here we make use of the key observation \ref{jetsub}, which allows us to view the $(n+1)$-st jet space $M^{(n+1)}$ as a subspace of the first jet space $(M^{(n)})^{(1)}$ of the $n$-th jet space. The strategy is to regard $\mathrm{pr}^{(n-1)} v$ as a vector field $M^{(n-1)}$ and then use the case $n=1$ above to prolong it to $(M^{(n-1)})^{(1)}$. Then we restrict the resulting vector field to the subspace $M^{(n)}$ of $(M^{(n-1)})^{(1)}$, and this will give the $n$-th prolongation of $v$. The fact that this is possible and gives a well-defined result will follow from the explicit formula below. The coordinates on $(M^{(n-1)})^{(1)}$ are given by $u^\alpha_{J,k}=\pd{u^\alpha_J}{x^k}$, where $J=(j_1,\dots,j_{n-1})$, $1\le k\le p$, and $1\le \alpha\le q$. By \eqref{projproj}, the only new coefficients we need to calculate are those of the highest order derivatives $\partial/\partial u^\alpha_{J,k}$. By \eqref{prol1form}, they are given by \begin{equation}\label{recursion} \phi^{J,k}_\alpha = D_k\phi^J_\alpha - \sum_{i=1}^p D_k\xi^i \cdot u^\alpha_{J,i}. \end{equation} To finish the proof, it suffices to show that the explicit formula \eqref{phiform} solves the recursion relation \eqref{recursion} in closed form. This follows by induction: supposing that $\phi^J_\alpha$ has already been shown to satisfy \eqref{phiform}, it follows from \eqref{recursion} that \begin{equation*} \begin{split} \phi^{J,k}_\alpha &= D_k\left[D_J\left(\phi_\alpha-\sum_{i=1}^p\xi^i u^\alpha_i\right) + \sum_{i=1}^p\xi^i u^\alpha_{J,i}\right] - \sum_{i=1}^p D_k\xi^i\cdot u^\alpha_{J,i}\\ &= D_kD_J\left(\phi_\alpha-\sum_{i=1}^p\xi^i u^\alpha_i\right) + \sum_{i=1}^p(D_k\xi^i\cdot u^\alpha_{J,i} + \xi^i u^\alpha_{J,ik}) - \sum_{i=1}^p D_k\xi^i\cdot u^\alpha_{J,i}\\ &=D_kD_J\left(\phi_\alpha-\sum_{i=1}^p\xi^i u^\alpha_i\right) + \sum_{i=1}^p \xi^i u^\alpha_{J,ik}, \end{split} \end{equation*} where $u^\alpha_{J,ik}=\partial^2 u^\alpha_J/\partial x^i\partial x^k$. Thus $\phi^{J,k}_\alpha$ is of the form \eqref{phiform} as well, which completes the induction step. \ep \begin{thr} {\bf Remark. }\rm Note that \eqref{recursion} provides a useful way of calculating prolongations of vector fields recursively. \et \begin{thr}{\bf Example. }\rm As an illustration of the relative ease with which one may calculate prolongations of vector fields using the general prolongation formula, we once more return to the rotation group $G=SO(2)$ acting on $\mathbb R^2$, cf.\ \ref{rotact5}. The infinitesimal generator is $v=-u\partial_x + x\partial_u$, so $\phi=x$ and $\xi=-u$. By \ref{gpf} we have $$ \mathrm{pr}^{(1)} v = v + \phi^x\partial_{u_x}, $$ where $$ \phi^x = D_x(\phi-\xi u_x) + \xi u_{xx} = D_x(x+uu_x)-uu_{xx} = 1 + u_x^2, $$ which confirms \eqref{rotvprol}. To calculate the coefficient $\phi^{xx}$ of $\partial_{u_{xx}}$ in $\mathrm{pr}^{(2)} v$, we can either use \eqref{phiform}: $$ \phi^{xx} = D_x^2(\phi-\xi u_x) + \xi u_{xxx} = D_x^2(x+uu_x)-uu_{xxx} = 3u_xu_{xx}, $$ or the recursion formula \eqref{recursion}: $$ \phi^{xx} = D_x\phi^x-u_{xx}D_x\xi = D_x(1+u_x^2)+u_xu_{xx} = 3u_xu_{xx}. $$ Consequently, $$ \mathrm{pr}^{(2)} v = -u\partial_x + x\partial_u + (1 + u_x^2)\partial_{u_x} + 3u_xu_{xx}\partial_{u_{xx}}. $$ Note that to derive this result via the second prolongation of the group action (as in \ref{rotact3}) would be much more involved. Based on this, and using the invariance criterion in \ref{mainpde}, it follows that the ODE $u_{xx}=0$ has $SO(2)$ as a symmetry group. In fact, $$ \mathrm{pr}^{(2)} v(u_{xx}) = 3u_xu_{xx} = 0 \quad\text{whenever }\ u_{xx}=0. $$ Since the solutions of $u_{xx}=0$ are linear functions, this is basically just the statement that rotations take straight lines to straight lines. Next, consider the function \begin{equation}\label{curv} \kappa(x,u^{(2)}) = \frac{u_{xx}}{(1+u_x^2)^{3/2}}, \end{equation} describing the Frenet-curvature of a planar curve. Then one easily checks that $\mathrm{pr}^{(2)} v(\kappa)=0$, so \ref{finvcrit} shows that $\kappa$ is an invariant of $\mathrm{pr}^{(2)} SO(2)$. Geometrically, this means that the curvature of a planar curve is invariant under rotations. \et \bt Let $v$, $w$ be smooth vector fields on the open subset $M$ of $X\times U$. Then \begin{itemize} \item[(i)] $\mathrm{pr}^{(n)}(cv+w) = c \cdot\mathrm{pr}^{(n)} v + \mathrm{pr}^{(n)} w \quad (c\in \mathbb R)$. \item[(ii)] $\mathrm{pr}^{(n)}[v,w] = [\mathrm{pr}^{(n)} v,\mathrm{pr}^{(n)} w]$. \end{itemize} \et \pr (i) This is immediate from \ref{gpf}. (ii) This also could be verified using \ref{gpf}, but we take a different route. First, by \ref{prolloctrans} we have $$ \mathrm{pr}^{(n)}(g\cdot h) = (\mathrm{pr}^{(n)} g)\cdot (\mathrm{pr}^{(n)} h). $$ Also, defining $(g+h)\cdot x := g\cdot x + h\cdot x$ we obtain $$ \mathrm{pr}^{(n)}(g + h) = \mathrm{pr}^{(n)} g + \mathrm{pr}^{(n)} h $$ directly from the definition \eqref{groupprol} (using sums of representing functions to represent sums of points in the jet bundle). Also, $\mathrm{pr}^{(n)} \mathrm{id}_M = \mathrm{id}_{M^{(n)}}$, and $v\mapsto \mathrm{pr}^{(n)} v$ is clearly continuous. Finally, recall from \eqref{flowstuff} that $$ \mathrm{pr}^{(n)}(\mathrm{Fl}^v_\varepsilon) = \mathrm{Fl}^{\mathrm{pr}^{(n)} v}_{\varepsilon}, $$ so we can use \ref{eduard} to calculate as follows: \begin{equation*} \begin{split} [\mathrm{pr}^{(n)} v,\mathrm{pr}^{(n)} w] &= \lim_{\varepsilon\to 0} \frac{ \mathrm{Fl}^{\mathrm{pr}^{(n)} w}_{-\sqrt{\varepsilon}}\circ\mathrm{Fl}^{\mathrm{pr}^{(n)} v}_{-\sqrt{\varepsilon}}\circ\mathrm{Fl}^{\mathrm{pr}^{(n)} w}_{\sqrt{\varepsilon}} \circ\mathrm{Fl}^{\mathrm{pr}^{(n)} v}_{\sqrt{\varepsilon}} - \mathrm{id}_{M^{(n)}}}{\varepsilon}\\ &=\lim_{\varepsilon\to 0} \frac{\mathrm{pr}^{(n)}[ \mathrm{Fl}^{w}_{-\sqrt{\varepsilon}}\circ\mathrm{Fl}^{v}_{-\sqrt{\varepsilon}}\circ\mathrm{Fl}^{w}_{\sqrt{\varepsilon}} \circ\mathrm{Fl}^{v}_{\sqrt{\varepsilon}} - \mathrm{id}_{M}]}{\varepsilon}\\ &=\mathrm{pr}^{(n)}\left[\lim_{\varepsilon\to 0} \frac{ \mathrm{Fl}^{w}_{-\sqrt{\varepsilon}}\circ\mathrm{Fl}^{v}_{-\sqrt{\varepsilon}}\circ\mathrm{Fl}^{w}_{\sqrt{\varepsilon}} \circ\mathrm{Fl}^{v}_{\sqrt{\varepsilon}} - \mathrm{id}_{M}}{\varepsilon}\right]\\ &=\mathrm{pr}^{(n)} [v,w]. \end{split} \end{equation*} \ep If we call a vector field $v$ an {\em infinitesimal symmetry}\index{infinitesimal symmetry} of a system ${\mathcal S}$ of differential equations if it satisfies \eqref{mainsymcrit}, then we obtain from the previous theorem: \begin{thr} {\bf Corollary. }\label{infliesym} Let ${\mathcal S}$ be a system of differential equations of maximal rank defined on $M\subseteq X\times U$. Then the set of all infinitesimal symmetries of ${\mathcal S}$ forms a Lie algebra of vector fields on $M$. \et To conclude this section we note the following equivalent way of writing the general prolongation formula \ref{gpf}. For $v$ a vector field on $M\subseteq X\times U$, let \begin{equation}\label{chardef} Q_\alpha(x,u^{(1)}) := \phi_\alpha(x,u) - \sum_{i=1}^p \xi^i(x,u)u^\alpha_i, \quad \alpha=1,\dots,q. \end{equation} The $q$-tuple $Q(x,u^{(1)})=(Q_1,\dots,Q_q)$ is called the {\em characteristic}\index{characteristic!of vector field} of the vector field $v$. Using this, \eqref{phiform} becomes \begin{equation}\label{phiform2} \phi^J_\alpha = D_JQ_\alpha + \sum_{i=1}^p \xi^i u^\alpha_{J,i} \end{equation} Substituting this into \eqref{theprol}, a brief calculation gives \begin{equation}\label{doch} \mathrm{pr}^{(n)} v = \sum_{\alpha=1}^q\sum_J D_JQ_\alpha \pd{}{u^\alpha_J} + \sum_{i=1}^p \xi^i\left(\pd{}{x^i} + \sum_{\alpha=1}^q\sum_J u^\alpha_{J,i} \pd{}{u^\alpha_J}\right), \end{equation} which, by \eqref{totder1}, means that \begin{equation}\label{charprol} \mathrm{pr}^{(n)} v = \mathrm{pr}^{(n)} v_Q + \sum_{i=1}^p \xi^iD_i, \end{equation} where \begin{equation}\label{charparts} v_Q:=\sum_{\alpha=1}^qQ_\alpha(x,u^{(1)})\pd{}{u^\alpha},\quad \mathrm{pr}^{(n)} v_Q = \sum_{\alpha=1}^q\sum_J D_JQ_\alpha \pd{}{u^\alpha_J}. \end{equation} Here, the last equality comes from applying \eqref{doch} to $v_Q$. In all of the above equations, the sums over $J$ extend over all $J$ with $0\le \sharp J\le n$. Note that the individual terms on the right hand side of \eqref{charprol} actually involve $(n+1)$-st order derivatives of $u$, and only their combination gives a genuine vector field on $M^{(n)}$. \section{Calculation of symmetry groups}\index{symmetry group!calculation of} In this section we want to illustrate how to explicitly calculate the symmetry group of any given system ${\mathcal S}$ of differential equations. From the previous sections, we obtain the following general strategy for tackling this problem: \begin{itemize} \item[(i)] Make an ansatz for an infinitesimal generator $v$ of a prospective one-parameter symmetry group of ${\mathcal S}$ in the form \eqref{vfform} with unknown functions $\xi_i$ and $\phi_\alpha$. \item[(ii)] Calculate $\mathrm{pr}^{(n)} v$ according to (\ref{theprol}), \eqref{phiform}, where $n$ is the order of ${\mathcal S}$. \item[(iii)] Insert the resulting expression into (\ref{mainsymcrit}). Since these equations only have to hold on ${\mathcal S}_P$, eliminate any dependencies of the derivatives of the $u$'s by using the equations in ${\mathcal S}$. The equations achieved in this way have to hold identically in $x$, $u$ and the remaining partial derivatives of the $u$'s. \item[(iv)] Solving the resulting so-called {\em defining equations}\index{defining equations} produces a number of (usually elementary) partial differential equations for the unknown functions $\xi_i$ and $\phi_\alpha$. \item[(v)] Compute the $\xi_i$ and $\phi_\alpha$ from these equations. By \ref{infliesym}, the vector fields gathered from this procedure form a Lie algebra of infinitesimal symmetries. \item[(vi)] The corresponding one-parameter symmetry groups are found by calculating the flows of the respective infinitesimal symmetries. \end{itemize} Let us systematically work through this list in a concrete example: \begin{thr}{\bf Example. }\rm\label{heatsym} Consider the one-dimensional heat equation \begin{equation}\label{heatequ} u_t=u_{xx} \end{equation} \et Thus we have $X=\mathbb R^2$ with variables $x$, $t$, and $U=\mathbb R$, and the equation $P(x,t,u^{(2)})=u_t-u_{xx}$ determines a subvariety in $X\times U^{(2)}$. For the infinitesimal generators of symmetries of ${\mathcal S}=\{P=0\}$ we make the ansatz \begin{equation}\label{ansatz} v = \xi(x,t,u) \pd{}{x} + \tau(x,t,u)\pd{}{t} + \phi(x,t,u)\pd{}{u} \end{equation} Our aim is to determine all possible coefficient functions $\xi,\,\tau,\,\phi$ such that $\mathrm{Fl}^v_\varepsilon$ is a symmetry group of the heat equation \eqref{heatequ}. Since we wish to apply \ref{mainpde}, we will need the second prolongation of $v$, $$ \mathrm{pr}^{(2)} v = v + \phi^x\pd{}{u_{x}} + \phi^t\pd{}{u_{t}} + \phi^{xx}\pd{}{u_{xx}} + \phi^{xt}\pd{}{u_{xt}} + \phi^{tt}\pd{}{u_{tt}}. $$ Inserting this into \eqref{mainsymcrit}, it follows that the only restriction on the coefficients of $v$ is that \begin{equation}\label{onlyres} \phi^t = \phi^{xx} \quad \text{ whenever }\quad u_t=u_{xx}. \end{equation} Using \eqref{phiform}, we calculate \begin{equation}\label{phit} \begin{split} \phi^t &= D_t(\phi-\xi u_x -\tau u_t) +\xi u_{xt} + \tau u_{tt} = D_t\phi -u_xD_t\xi-u_tD_t\tau\\ &= \phi_t - \xi_t u_x + (\phi_u-\tau_t)u_t - \xi_uu_xu_t -\tau_u u_t^2 \end{split} \end{equation} and \begin{equation}\label{phixx} \begin{split} \phi^{xx} &= D_x^2(\phi-\xi u_x -\tau u_t) +\xi u_{xxx} + \tau u_{xxt}\\ &= D_x^2\phi - u_xD_x^2\xi- u_tD_x^2\tau -2 u_{xx}D_x\xi - 2u_{xt}D_x\tau\\ &= \phi_{xx} + (2\phi_{xu}-\xi_{xx})u_x -\tau_{xx}u_t + (\phi_{uu} -2\xi_{xu})u_x^2 -2\tau_{xu}u_xu_t\\ &\hphantom{=} -\xi_{uu} u_x^3 -\tau_{uu}u_x^2u_t + (\phi_u-2\xi_x)u_{xx} - 2\tau_xu_{xt} - 3\xi_uu_xu_{xx}\\ &\hphantom{=}-\tau_uu_tu_{xx}-2\tau_uu_xu_{xt}. \end{split} \end{equation} If we insert these expressions into \eqref{onlyres} and replace $u_t$ by $u_{xx}$ wherever it occurs, we obtain an equation on the jet space $X\times U^{(2)}$ that has to hold identically in all the remaining variables. We may therefore `split' the equation by equating the coefficient functions of all monomials appearing in it. This gives the following table: \vskip2em \begin{center} \begin{tabular}{ccccc} \hline Monomial & & Coefficient & & \\ \hline $u_xu_{xt}$ & & $0=-2\tau_u$ & & (a) \\ $u_{xt}$ & & $0=-2\tau_x$ & & (b) \\ $u_{xx}^2$ & & $-\tau_u=-\tau_u$ & & (c) \\ $u_x^2u_{xx}$ & & $0=-\tau_{uu}$ & & (d) \\ $u_xu_{xx}$ & & $-\xi_u=-2\tau_{xu}-3\xi_u$ & & (e) \\ $u_{xx}$ & & $\phi_u-\tau_t=-\tau_{xx} +\phi_u-2\xi_x$ & & (f) \\ $u_{x}^3$ & & $0=-\xi_{uu}$ & & (g) \\ $u_{x}^2$ & & $0=\phi_{uu} - 2\xi_{xu}$ & & (h) \\ $u_{x}$ & & $-\xi_t=2\phi_{xu} -\xi_{xx}$ & & (j) \\ $1$ & & $\phi_t=\phi_{xx}$ & & (k) \\ \hline \end{tabular} \end{center} \vskip1em We first note that (a), (b) imply that $\tau$ is a function of $t$ only: $\tau=\tau(t)$. Using this, (e) shows that $\xi$ does not depend on $u$, and (f) gives $\tau_t=2\xi_x$. Therefore, $$ \xi(x,t) = \frac{1}{2}\tau_tx + \sigma(t), $$ where $\sigma$ is some function of $t$. (h) shows that $\phi$ is linear in $u$, so $$ \phi(x,t,u) = \beta(x,t)u + \alpha(x,t) $$ for some functions $\alpha$, $\beta$. By (j), $-2\beta_x = \xi_t = \frac{1}{2}\tau_{tt}x+\sigma_t$, so \begin{equation}\label{betae} \beta(x,t) = -\frac{1}{8}\tau_{tt}x^2-\frac{1}{2}\sigma_tx + \rho(t). \end{equation} Finally, (k) implies that both $\alpha$ and $\beta$ must be solutions of the heat equation: $$ \alpha_t = \alpha_{xx}, \qquad \beta_t=\beta_{xx}. $$ Combined with \eqref{betae} we get $$ \tau_{ttt} = 0, \qquad \sigma_{tt}=0, \qquad \rho_t = -\frac{1}{4}\tau_{tt}. $$ Hence $\tau$ is quadratic in $t$, $\sigma$ is linear in $t$, so for suitable constants $c_i$ we get: \begin{equation*} \begin{split} \tau &= c_2 +2c_4 t + 4c_6 t^2\\ \sigma &= c_1 + 2c_5 t\\ \rho_t &= -\frac{1}{4}\tau_{tt} = -2c_6 \Rightarrow \rho = -2c_6 t + c_3\\ \Rightarrow\beta &= -\frac{1}{8}\tau_{tt}x^2 -\frac{1}{2}\sigma_{t}x +\rho = -c_6x^2 -c_5x-2c_6t+c_3 \end{split} \end{equation*} Consequently, \begin{equation*} \begin{split} \xi &= \frac{1}{2}\tau_t + \sigma = c_1 +c_4 x + 2c_5 t + 4c_6xt\\ \tau &= c_2 + 2c_4 t + 4c_6t^2\\ \phi &= \beta u+\alpha = (c_3 -c_5 x - 2c_6 t -c_6x^2)u + \alpha(x,t), \end{split} \end{equation*} where $c_1,\dots,c_6$ are arbitrary constants and $\alpha$ is any solution of the heat equation. It follows that the Lie algebra of infinitesimal symmetries of the heat equation is spanned by the six vector fields \begin{equation*} \begin{split} v_1 &= \partial_x\\ v_2 &= \partial_t\\ v_3 &= u\partial_u\\ v_4 &= x\partial_x + 2t\partial_t\\ v_5 &= 2t\partial_x -xu\partial_u\\ v_6 &= 4tx\partial_x + 4t^2\partial_t - (x^2+2t)u\partial_u \end{split} \end{equation*} and the infinite-dimensional subalgebra $$ v_\alpha = \alpha(x,t)\partial_u, $$ where $\alpha$ is any solution of the heat equation. By \ref{infliesym} we know that these generators span a Lie algebra of vector fields. Thus with any two generators also their Lie bracket is an infinitesimal symmetry of the heat equation. The one-parameter symmetry group actions $\mathrm{Fl}^{v_i}_\varepsilon$ corresponding to the generators can be calculated by solving the ODE $$ \frac{d}{d\varepsilon}(x(\varepsilon),t(\varepsilon),u(\varepsilon)) = v_i(x(\varepsilon),t(\varepsilon),u(\varepsilon)) $$ with initial value $(x(0),t(0),u(0))=(x,t,u)$. This gives the groups \begin{equation*} \begin{split} G_1:\ & (x+\varepsilon,t,u)\\ G_2:\ & (x,t+\varepsilon,u)\\ G_3:\ & (x,t,e^\varepsilon u)\\ G_4:\ & (e^\varepsilon x,e^{2\varepsilon}t,u)\\ G_5:\ & (x+2\varepsilon t,t,ue^{-\varepsilon x -\varepsilon^2 t})\\ G_6:\ & \left(\frac{x}{1-4\varepsilon t},\frac{t}{1-4\varepsilon t},u\sqrt{1-4\varepsilon t}e^{\frac{-\varepsilon x^2}{1-4\varepsilon t}}\right)\\ G_\alpha:\ & (x,t,u+\varepsilon\alpha(x,t)) \end{split} \end{equation*} Using \eqref{gdotf} we conclude that, given any solution $f$ of the heat equation, so are the following functions (given by $g^{(i)}\cdot f$, for $g^{(i)}\in G_i$): \begin{equation*} \begin{split} u^{(1)} &= f(x-\varepsilon,t)\\ u^{(2)} &= f(x,t-\varepsilon)\\ u^{(3)} &= e^\varepsilon f(x,t)\\ u^{(4)} &= f(e^{-\varepsilon}x,e^{-2\varepsilon}t)\\ u^{(5)} &= e^{-\varepsilon x +\varepsilon^2 t}f(x-2\varepsilon t,t)\\ u^{(6)} &= \frac{1}{\sqrt{1+4\varepsilon t}}e^{\frac{-\varepsilon x^2}{1+4\varepsilon t}}f\left(\frac{x}{1+4\varepsilon t}, \frac{t}{1+4\varepsilon t}\right)\\ u^{(\alpha)} &= f(x,t) + \varepsilon \alpha(x,t). \end{split} \end{equation*} Here, $\varepsilon$ is a real number and $\alpha$ is any solution of the heat equation. $G_3$ and $G_\alpha$ are simply an expression of the linearity of the heat equation. $G_1$ and $G_2$ reflect the space- and time-invariance of the equation. $G_4$ is a scaling symmetry, and $G_5$ represents a kind of Galilean boost to a moving coordinate frame. $G_6$ is a genuinely local transformation group. Applying $G_6$ to a trivial solution $u=c$ ($c$ a constant) we obtain that also $$ u = \frac{c}{\sqrt{1+4\varepsilon t}}e^{\frac{-\varepsilon x^2}{1+4\varepsilon t}} $$ is a solution. Setting $c=\sqrt{\varepsilon/\pi}$ we obtain the fundamental solution to the heat equation at the point $(x_0,t_0)=(0,-1/(4\varepsilon))$. Translating in $t$ (using $G_2$) we arrive at the fundamental solution at $(0,0)$, $$ u = \frac{1}{\sqrt{4\pi t}}e^{\frac{-x^2}{4t}}. $$ The most general one-parameter group of symmetries has a generator of the form $c_1v_1+\dots +c_6v_6 + v_\alpha$. Also, an arbitrary element of the symmetry group of the heat equation (if sufficiently close to the identity element) can be expressed in the form $$ g = \mathrm{Fl}^{v_\alpha}_{\varepsilon_\alpha}\circ \mathrm{Fl}^{v_6}_{\varepsilon_6}\circ \dots \circ \mathrm{Fl}^{v_1}_{\varepsilon_1}, $$ and so the most general solution obtainable from any given solution $u$ by such a transformation is of the form $$ u=\frac{1}{\sqrt{1+4\varepsilon_6 t}}e^{\varepsilon_3-\frac{\varepsilon_5x + \varepsilon_6x^2-\varepsilon_5^2t}{1+4\varepsilon_6t}} f\left(\frac{e^{-\varepsilon_4}(x-2\varepsilon_5t)}{1+4\varepsilon_6t} - \varepsilon_1,\frac{e^{-2\varepsilon_4}t}{1+4\varepsilon_6t}-\varepsilon_2\right) +\alpha(x,t), $$ with $\varepsilon_1,\dots,\varepsilon_6$ real constants and $\alpha$ any solution of the heat equation. \section{Nondegeneracy conditions}\index{nondegeneracy conditions} The central result \ref{mainpde} only gives a sufficient condition for a local group of transformations to be a symmetry group of a given system of differential equations. In the present section we will see that under some mild condition on the system the infinitesimal criterion from \ref{mainpde} is in fact also sufficient. Let us first look at the difference between the criterion \ref{algsym} for the invariance of a system of algebraic equations, which is necessary and sufficient, and \ref{mainpde}. For a system of algebraic equations $F=0$, for each point $x_0$ on the subvariety ${\mathcal S}_F=\{x\mid F(x)=0\}$ there trivially exists a solution to the system, namely $x_0$ itself. In the case of a system $P(x,u^{(n)})=0$ of differential equations, however, if $(x_0,u^{(n)}_0)\in {\mathcal S}_P = \{(x,u^{(n)})\mid P(x,u^{(n)})=0\}$ there is no guarantee that there exists a solution $u=f(x)$ of the differential equation locally around $x_0$ such that $u^{(n)}_0=\mathrm{pr}^{(n)} f(x_0)$. Our definition \ref{dgsgdef} of a symmetry group of $P(x,u^{(n)})=0$ requires $\mathrm{pr}^{(n)} G$ to transform solutions into solutions, so we only can infer that it will move points in ${\mathcal S}_P$ for which there exists a solution as above to other points in ${\mathcal S}_P$. In general, these points need not constitute the entire set ${\mathcal S}_P$. Thus we define: \begin{thr} {\bf Definition. } A system $P(x,u^{(n)})=0$ of $n$-th order differential equations is called {\em locally solvable}\index{locally solvable} at the point $$ (x_0,u^{(n)}_0)\in{\mathcal S}_P = \{(x,u^{(n)})\mid P(x,u^{(n)})=0\} $$ if there exists a smooth function $u=f(x)$ defined near $x_0$ such that $u^{(n)}_0=\mathrm{pr}^{(n)} f(x_0)$. The system is called locally solvable if it is locally solvable at every point of ${\mathcal S}_P$. It is called {\em nondegenerate}\index{nondegenerate} if it is locally solvable and of maximal rank at every point $(x_0,u^{(n)}_0)\in{\mathcal S}_P$. \et \begin{thr} {\bf Remark. }\rm\label{odenon} Consider a non-singular ODE of order $n$, i.e., of the form \begin{equation}\label{oden1} u_n = P(x,u,u_x,\dots,u_{n-1}), \end{equation} with $u_k=\pd{^ku}{x^k}$ for any $k$ and let $(x^0,u^0,\dots,u_n^0)\in {\mathcal S}_P$. Then we need to find a local smooth function $u=f(x)$ such that $$ u^0=f(x^0),\ u^0_x = f'(x^0),\dots, u^0_{n-1}=f^{(n-1)}(x^0),\ u^0_{n}=f^{(n)}(x^0). $$ Here, the first $n$ conditions are the usual initial conditions for the ODE \eqref{oden1}, hence can always be satisfied by standard ODE solution theory. The last condition then simply follows from \eqref{oden1} itself (i.e., from the fact that $(x^0,u^0,\dots,u_n^0)\in {\mathcal S}_P$). We conclude that non-singular (systems of) ODEs are always locally solvable, hence are in fact nondegenerate. \et Note that even for PDEs, local solvability is usually a very mild restriction -- e.g., in Cauchy problems one usually is looking for solutions with prescribed values on an entire hypersurface. \begin{thr}{\bf Example. }\rm (i) The $2$-dimensional wave equation $$ P(x,t,u^{(2)}) = u_{tt}-u_{xx}=0 $$ is locally solvable. In fact, let $$ (x^0,t^0;u^0;u^0_x,u^0_t;u^0_{xx},u^0_{xt},u^0_{tt}) \in {\mathcal S}_P, $$ i.e., $u^0_{tt}=u^0_{xx}$. Then we need to find a solution $u=f(x)$ of the wave equation near $(x^0,t^0)$ with $\mathrm{pr}^{(2)} f(x^0,t^0) = (u^0_x,u^0_t;u^0_{xx},u^0_{xt},u^0_{tt})$. Since $u^0_{tt}=u^0_{xx}$, we can take for $f$ the polynomial solution \begin{equation*} \begin{split} f(x,t) = u^0 + &u^0_x(x-x^0)+u^0_t(t-t^0) + \frac{1}{2}u^0_{xx}[(x-x^0)^2+(t-t^0)^2]\\ &+u^0_{xt}(x-x^0)(t-t^0). \end{split} \end{equation*} (ii) The over-determined system \begin{equation}\label{over} u_x=yu,\quad u_y = 0 \end{equation} is not locally solvable: given $(x_0,y_0)$, let $$ u^0 = u(x_0,y_0)=1, \quad u^0_x = u_x(x_0,y_0)=y_0,\quad u^0_y = u_y(x_0,y_0)=0. $$ Then there is no local solution with the prescribed jet at $(x_0,y_0)$. In fact, by cross-differentiation we obtain $$ 0=u_{xy} = (yu)_y = yu_y + u = u, $$ which is incompatible with the above prescription. (iii) Apart from integrability conditions as in (ii), the second main reason why a PDE may fail to be locally solvable is that there are no smooth solutions, even locally. The most famous example is due to H.\ Lewy (\cite{L}), who showed that there exist smooth functions $h=h(x,y,z)$ such that the first order system \begin{equation*} \begin{split} u_x-v_y+2yu_z+2xv_z &= h(x,y,z)\\ u_y+v_x-2xu_z+2yv_z &= 0 \end{split} \end{equation*} has no smooth solutions even locally around any point. (iv) For analytic systems of PDEs (in Kovalevskaya-form), the Cauchy--Kovalevskaya theorem ensures local solvability. \et For nondegenerate systems of differential equations we can now show that indeed \eqref{mainsymcrit} is necessary and sufficient: \bt\label{nesu} Let $$ P_\nu(x,u^{(n)}) = 0 \quad \nu=1,\dots,l $$ be a nondegenerate system of differential equations defined on an open subset $M$ of $X\times U$. Let $G$ be a local transformation group acting on $M$. Then $G$ is a symmetry group of the system if and only if \begin{equation}\label{mainsymcrit1} \mathrm{pr}^{(n)} v [P_\nu(x,u^{(n)})] = 0, \quad \nu=1,\dots,l, \text{ whenever } P(x,u^{(n)})=0 \end{equation} for every infinitesimal generator $v$ of $G$. \et \pr The condition is sufficient by \ref{mainpde}. To see that it is also necessary, note that by \ref{algsym}, condition \eqref{mainsymcrit1} is equivalent to $\mathrm{pr}^{(n)} G$ mapping ${\mathcal S}_P=\{(x,u^{(n)})\mid P(x,u^{(n)})=0\}$ into itself. Thus we have to verify that any symmetry group $G$ of the system has this property. Let $(x_0,u^{(n)}_0)\in {\mathcal S}_P$. Then by local solvability we may find a smooth function $u=f(x)$ defined near $x_0$ such that $u^{(n)}_0=\mathrm{pr}^{(n)} f(x_0)$. If $g\in G$ is such that $\mathrm{pr}^{(n)} g\cdot (x_0,u^{(n)}_0)$ is defined, then by shrinking the domain of $f$ if necessary, we can obtain that ${\tilde f} = g\cdot f$ is a well-defined function near ${\tilde x}_0$, where $({\tilde x}_0,{\tilde u}_0)=g\cdot (x_0,u_0)$. Then by \eqref{groupprol} we have $$ \mathrm{pr}^{(n)} g \cdot (x_0,u^{(n)}_0) = ({\tilde x}_0,\mathrm{pr}^{(n)} (g\cdot f)({\tilde x}_0)) =: ({\tilde x}_0,{\tilde u}_0^{(n)}). $$ Since ${\tilde f}=g\cdot f$ is a solution, this shows that $({\tilde x}_0,{\tilde u}_0^{(n)})\in {\mathcal S}_P$, as claimed. \ep \section{Integration of ordinary differential equations}\index{ordinary differential equation} As an illustration of the power of the methods developed in this chapter, in the present section we investigate the general problem of solving ordinary differential equations (ODEs). It turns out that all the standard solution techniques for special types of ODEs are instances of symmetry methods. Moreover, knowledge of a sufficiently large group of symmetries of any given ODE allows to reduce the solution procedure to successive integrations (or quadratures,\index{quadrature} in classical terminology). Consider a first-order ODE \begin{equation}\label{firstode} \pd{u}{x} = F(x,u). \end{equation} We will show that if a one-parameter symmetry group of \eqref{firstode} is known, then it can be solved by integration. Let $G$ be a local one-parameter group of transformations acting on an open subset $M$ of $X\times U=\mathbb R^2$, with infinitesimal generator \begin{equation}\label{odeinf} v = \xi(x,u)\pd{}{x} + \phi(x,u)\pd{}{u}. \end{equation} By \ref{gpf}, its first prolongation is given by \begin{equation}\label{odeprol} \mathrm{pr}^{(1)} v = \xi(x,u)\pd{}{x} + \phi(x,u)\pd{}{u} + \phi^x(x,u,u_x)\pd{}{u_x}, \end{equation} where \begin{equation}\label{odeprol2} \phi^x = D_x\phi - u_x D_x\xi = \phi_x + (\phi_u - \xi_x)u_x - \xi_u u_x^2. \end{equation} The infinitesimal criterion \eqref{mainsymcrit} for $v$ to generate a symmetry of \eqref{firstode} then reads \begin{equation}\label{odesymcrit} \pd{\phi}{x} +\Big(\pd{\phi}{u} - \pd{\xi}{x}\Big)F - \pd{\xi}{u} F^2 = \xi \pd{F}{x} + \phi\pd{F}{u}, \end{equation} for all $(x,u)$. Thus any solution to the PDE \eqref{odesymcrit} generates a symmetry of the ODE \eqref{firstode}. Although at first sight, replacing the ODE \eqref{firstode} by the PDE \eqref{odesymcrit} may not seem to facilitate the problem, in practice one can often find symmetries directly, e.g.\ by geometric considerations. Suppose that $v$ generates a symmetry of \eqref{firstode} and let $v|_{(x_0,u_0)}\not=0$. Then by straightening out (see \cite[17.14]{KLie}) we may pick new coordinates \begin{equation}\label{newc} y = \eta(x,u),\quad w= \zeta(x,u) \end{equation} near $(x_0,u_0)$ such that in the new coordinates we have $v=\partial_{w}$, and thereby $$ \mathrm{pr}^{(1)} v = v = \pd{}{w} $$ The condition that \eqref{firstode} be invariant under $v$ then simply means that it must be independent of $w$, i.e., it can be written in the form $$ \pd{w}{y} = H(y) $$ for some smooth function $H$. This equation can then be solved trivially by integration: $$ w = \int H(y)\,dy + \text{ const}. $$ Transforming back to the $(x,u)$-coordinates gives the solution to \eqref{firstode}. To find the required coordinate transform, note that in the new coordinates we require (cf.\ \cite[2.5.3]{DG1}) that $$ v = v(\eta)\pd{}{y} + v(\zeta)\pd{}{w} \stackrel{!}{=} \pd{}{w}, $$ so in terms of $(x,u)$, we need to satisfy \begin{equation}\label{etaz} \begin{split} v(\eta) &= \xi\pd{\eta}{x} + \phi\pd{\eta}{u} = 0\\ v(\zeta) &= \xi\pd{\zeta}{x} + \phi\pd{\zeta}{u} = 1. \end{split} \end{equation} Here, the first equation means that $\eta$ is an invariant of $G$, hence can be found by the method of characteristics indicated in \eqref{invode}. Once $\eta$ is known, $\zeta$ can often be found by inspection. In any case, also the second equation can be solved by the method of characteristics. As indicated above, there is no guarantee that solving \eqref{etaz} is indeed easier than solving the original equation \eqref{firstode}. For example, if \begin{equation}\label{umsonst} \frac{\phi(x,u)}{\xi(x,u)} = F(x,u) \end{equation} then inserting $\phi=\xi\cdot F$ in \eqref{odesymcrit} gives an identity, i.e., finding a symmetry in this case is exactly the same problem as solving the original equation. \begin{thr}{\bf Example. }\rm\label{home} Consider the homogeneous equation\index{homogeneous ODE} $$ \pd{u}{x} = F\left(\frac{u}{x}\right). $$ This equation is invariant under the group of scaling transformations $$ G: (x,u)\mapsto (\lambda x,\lambda u),\quad \lambda>0. $$ Indeed, by \eqref{odeprol2}, the first prolongation of the generator $v=x\partial_x + u\partial_u$ is $v$ itself and so one readily checks that \eqref{odesymcrit} is satisfied. New coordinates satisfying \eqref{etaz} are given by $$ y=\frac{u}{x}, \quad w=\log x. $$ Then $$ \pd{u}{x} = \frac{du/dy}{dx/dy} = \frac{x(1+yw_y)}{xw_y} = \frac{1+yw_y}{w_y}, $$ so the equation in the new coordinates becomes $$ \frac{dw}{dy} = \frac{1}{F(y)-y}, $$ which can be solved by integration: $$ w = \int \frac{dy}{F(y)-y} + c. $$ \et \begin{thr}{\bf Example. }\rm Consider the equation $P\,dx + Q\,du = 0$. We show that if it possesses a one-parameter symmetry with generator $v=\xi\partial_x + \phi\partial_u$, then the function $$ R(x,u) = \frac{1}{\xi(x,u)P(x,u) + \phi(x,u)Q(x,u)} $$ is an integrating factor.\index{integrating factor} To see this, note first that the given ODE is of the form \eqref{firstode} with $F=-P/Q$. Inserting this into \eqref{odesymcrit} and recalling \ref{nesu} and \ref{odenon} it follows that $v$ is a symmetry if and only if \begin{equation}\label{intfac} \begin{split} \left(\xi\pd{P}{x} + \phi\pd{P}{u}\right)Q - &\left(\xi\pd{Q}{x} + \phi\pd{Q}{u}\right)P +\pd{\phi}{x}Q^2 \\ &-\left(\pd{\phi}{u} - \pd{\xi}{x}\right)PQ -\pd{\xi}{u}P^2 = 0. \end{split} \end{equation} On the other hand, that $R$ is an integrating factor means that $$ \pd{}{u}(RP) = \pd{}{x}(RQ), $$ and inserting $R$ from above here gives \begin{equation*} \begin{split} R^2 \Big( \phi \Big(Q\pd{P}{u} - P\pd{Q}{u}\Big) - &\pd{\xi}{u}P^2 - \pd{\phi}{u}PQ\Big) \\ &= R^2 \Big( \xi \Big(P\pd{Q}{x} - Q\pd{P}{x}\Big) - \pd{\xi}{x}PQ - \pd{\phi}{x}Q^2\Big), \end{split} \end{equation*} which is equivalent to \eqref{intfac}, as claimed. \et Consider now a non-singular ODE of order $n$: \begin{equation}\label{oden} P(x,u^{(n)}) = P(x,u,u_x,\dots,u_n), \end{equation} where $u_n=\frac{d^n u}{dx^n}$. We will show that knowledge of a one-parameter symmetry group $G$ of \eqref{oden} with generator $v$ allows to reduce the order of the equation by one. To see this, choose coordinates $y=\eta(x,u)$, $w=\zeta(x,u)$ as in \eqref{etaz} that straighten out $v$, so that $v=\partial_w$. Then by the chain rule, we can re-write the derivatives of $u$ with respect to $x$ in terms of $y$, $w$ and the derivatives of $w$ with respect to $y$: $$ \frac{d^ku}{dx^k} = \delta_k\left(y,w,\od{w}{y},\dots,\odk{w}{y}\right), $$ for suitable functions $\delta_k$. Substituting these expressions back into \eqref{oden}, we obtain the equivalent $n$-th order equation \begin{equation}\label{odene} \tilde P(y,w^{(n)}) = \tilde P(y,w,w_y,\dots,w_n) = 0 \end{equation} in the new coordinates. Now note that since \eqref{oden} has $G$ as a symmetry group, the same is true of the transformed system \eqref{odene}: this follows from the definition of a symmetry group and the fact that the prolongation of a group action according to \eqref{groupprol} is equivariant under changing variables (again by the chain rule). In the new variables, $\mathrm{pr}^{(n)} v = v = \partial_w$, and so the infinitesimal criterion \eqref{mainsymcrit} reduces to $$ \mathrm{pr}^{(n)} v(\tilde P) = \pd{\tilde P}{w} = 0 \ \text{ whenever } \tilde P(y,w^{(n)})=0. $$ Since $\tilde P$ is invariant under $G^{(n)}$, by \ref{invvar} and \ref{nesu} (and provided that $G$ acts semi-regularly with orbits of constant dimension) there exists an equivalent equation that depends exclusively on a complete set of functionally independent invariants of $G^{(n)}$. In the present situation, such a set of invariants is obviously given by $y,w_y,w_{yy},\dots,w_n$. Consequently, there is an equivalent equation $$ \hat P(y,w_y,\dots,w_n) = 0, $$ which is independent of $w$. Thus, setting $z=w_y$ we obtain an equivalent ODE of order $n-1$, as claimed: $$ \hat P\Big(y,z,\dots,\frac{d^{n-1}z}{dy^{n-1}}\Big) = \hat P(y,z^{(n-1)}) = 0. $$ Given any solution $z=h(y)$ of this ODE, $w=\int h(y)\,dy + c$ is a solution to \eqref{odene}, and transforming back to $(x,u)$-coordinates gives a solution to \eqref{oden}. \begin{thr}{\bf Example. }\rm Any homogeneous linear ODE of second order \begin{equation}\label{hom2} u_{xx} + p(x)u_x + q(x)u = 0 \end{equation} is invariant under the scaling group $G: (x,u)\mapsto (x,\lambda u)$. In fact, the generator of $G$ is $v=u\partial_u$, so $\mathrm{pr}^{(2)} v = u\partial_u + u_x\partial_{u_x} + u_{xx}\partial_{u_{xx}}$, and so the infinitesimal criterion \eqref{mainsymcrit} is satisfied. Coordinates satisfying \eqref{etaz} are given by $$ y = x,\quad w=\log u, $$ so $v=\partial_w$. To re-express the equation in the new coordinates, note that $$ u=e^w,\quad u_x=w_x e^w,\quad u_{xx} = (w_{xx}+w_x^2)e^w, $$ giving $$ w_{xx} + w_x^2 + p(x)w_x + q(x) = 0, $$ which, as expected, is independent of $w$. Setting $z=w_x=u_x/u$ we therefore obtain the well-known transformation of \eqref{hom2} into the Riccati equation\index{Riccati equation} $$ z_x = - z^2 - p(x)z -q(x). $$ \et \section{Differential invariants}\index{differential invariant} In this section we look at a converse to the problem of determining the symmetry group of a differential equation, namely: Given a local transformation group, what is the most general type of differential equation that admits it as a symmetry group? We will give an answer to this question in the setting of ordinary differential equations. Recall from \ref{nesu} that a nondegenerate differential equation $P(x,u^{(n)})=0$ admits a local transformation group as a symmetry group if and only if the corresponding subvariety ${\mathcal S}_P$ of $M^{(n)}$ is invariant under $\mathrm{pr}^{(n)} G$. Provided that $\mathrm{pr}^{(n)} G$ acts semi-regularly with orbits of constant dimension, \ref{invvar} then implies that there is an equivalent equation $\tilde P=0$ describing the subvariety ${\mathcal S}_P$ and depending only on the functionally independent invariants of $\mathrm{pr}^{(n)} G$. We therefore give a name to the invariants of $\mathrm{pr}^{(n)} G$: \begin{thr} {\bf Definition. } Let $G$ be a local transformation group acting on $M\subseteq X\times U$ and let $n\ge 1$. An $n$-th order {\em differential invariant}\index{differential invariant} of $G$ is a smooth function $\eta:M^{(n)}\to \mathbb R$ such that $\eta$ is an invariant of $\mathrm{pr}^{(n)} G$: $$ \eta(\mathrm{pr}^{(n)} g \cdot (x,u^{(n)})) = \eta(x,u^{(n)}) $$ for all $(x,u^{(n)})\in M^{(n)}$ and $g\in G$ such that $\mathrm{pr}^{(n)} g \cdot (x,u^{(n)})$ is defined. \et \begin{thr}{\bf Example. }\rm\label{rotdi} For the rotation group $G=SO(2)$ on $X\times U=\mathbb R^2$ we have $v=-u\partial_x+x\partial_u$, and by \eqref{rotvprol} we have $$ \mathrm{pr}^{(1)} v = -u\partial_x + x\partial_u + (1+u_x^2)\partial_{u_x} $$ By definition, the first order differential invariants of $G$ are the invariants of $\mathrm{pr}^{(1)} G$, which can be calculated by the method of characteristics, cf.\ \eqref{invode}. This gives the following complete set of first order invariants of $G$: $$ y= \sqrt{x^2+u^2}, \quad w=\frac{xu_x-u}{x+uu_x}. $$ Any other first order invariant must be a function of these (by \ref{lociprop}). \et Our next aim is to show that one can always calculate higher order differential invariants inductively from known ones of lower order. To show this we need an auxiliary result first. \blem Let $v=\xi\partial_x+\phi\partial_u$ be a vector field on $M\subseteq X\times U=\mathbb R^2$ and let $\zeta = \zeta(x,u^{(n)})$ be smooth. Then \begin{equation}\label{prnp1} \mathrm{pr}^{(n+1)} v (D_x\zeta) = D_x[\mathrm{pr}^{(n)} v(\zeta)] - D_x\xi\cdot D_x\zeta. \end{equation} \et \pr By \eqref{charprol}, $$ \mathrm{pr}^{(n+1)} v(D_x\zeta) = \mathrm{pr}^{(n+1)} v_Q(D_x\zeta) + \xi D_x^2\zeta, $$ and $$ D_x[\mathrm{pr}^{(n)} v(\zeta)] = D_x[\mathrm{pr}^{(n)} v_Q(\zeta)] + D_x(\xi D_x\zeta). $$ It therefore remains to show that $$ \mathrm{pr}^{(n+1)} v_Q(D_x\zeta) = D_x[\mathrm{pr}^{(n)} v_Q(\zeta)]. $$ Using \eqref{charparts}, we calculate: \begin{equation*} \begin{split} \mathrm{pr}^{(n+1)} v_Q(D_x\zeta) &= \sum_{j=0}^{n+1} D_jQ \pd{}{u_j}\left(\pd{\zeta}{x}+ \sum_{k=0}^{n}u_{k+1}\pd{\zeta}{u_k}\right)\\ &= \sum_{j=0}^{n+1} D_jQ \left(\pd{^2\zeta}{u_j\partial x} +\sum_{k=0}^{n}\left(\delta^{k+1}_j\pd{\zeta}{u_k}+u_{k+1}\pd{^2\zeta}{u_j\partial u_k}\right)\right)\\ &= \sum_{j=0}^n\left(D_{j+1}Q\pd{\zeta}{u_j} + D_jQ\left( \pd{^2\zeta}{u_j\partial x} + \sum_{k=0}^{n}u_{k+1}\pd{^2\zeta}{u_j\partial u_k}\right)\right)\\ &= D_x\left(\sum_{j=0}^{n} D_jQ \pd{\zeta}{u_j}\right) = D_x[\mathrm{pr}^{(n)} v_Q(\zeta)], \end{split} \end{equation*} where in the penultimate line we used that $\zeta=\zeta(x,u^{(n)})$, hence its derivative with respect to $u_{n+1}$ vanishes. \ep \bp\label{nextderp} Let $G$ be a local one-parameter group of transformations acting on $M\subseteq X\times U=\mathbb R^2$. Let $y=\eta(x,u^{(n)})$ and $w=\zeta(x,u^{(n)})$ be $n$-th order differential invariants of $G$ ($n\ge 1$). Then the derivative \begin{equation}\label{nextder} \od{w}{y} = \frac{dw/dx}{dy/dx} \equiv \frac{D_x\zeta}{D_x\eta} \end{equation} is a differential invariant of $G$ of order $n+1$. \et \pr Note first that for any vector field $Z$ and any smooth functions $f$, $g$ we have the usual quotient rule $Z(f/g)=g^{-2}(Z(f)g-fZ(g))$, as follows from the derivation description of vector fields. Using this and \eqref{prnp1}, we obtain from \eqref{nextder}: \begin{equation*} \begin{split} \mathrm{pr}^{(n+1)} v \left(\od{w}{y}\right) &= \frac{1}{(D_x\eta)^2}\big(\mathrm{pr}^{(n+1)} v(D_x\zeta)\cdot D_x\eta -D_x\zeta\cdot \mathrm{pr}^{(n+1)} v(D_x\eta)\big)\\ &= \frac{1}{(D_x\eta)^2}\big(D_x[\mathrm{pr}^{(n)} v(\zeta)]\cdot D_x\eta - D_x\xi\cdot D_x\zeta\cdot D_x\eta\\ &\hphantom{xxxxxxxxxx}-D_x\zeta\cdot D_x[\mathrm{pr}^{(n)} v(\eta)] + D_x\zeta\cdot D_x\xi\cdot D_x\eta\big)=0 \end{split} \end{equation*} since $\zeta$ and $\eta$ are $n$-th order invariants, so $\mathrm{pr}^{(n)} v(\zeta) = \mathrm{pr}^{(n)}(\eta) = 0$. \ep \begin{thr} {\bf Corollary. }\label{oi} Let $G$ be a local one-parameter group of transformations acting on $M\subseteq X\times U=\mathbb R^2$. Let $\{y=\eta(x,u),\, w=\zeta(x,u,u_x)\}$ be a complete set of functionally independent invariants of $\mathrm{pr}^{(1)} G$. Then (locally around any point) the derivatives $$ y,w,\od{w}{y},\dots,\od{^{n-1}w}{y^{n-1}} $$ form a complete set of functionally independent invariants for $\mathrm{pr}^{(n)} G$ for $n\ge 1$. \et \pr Successively applying \ref{nextderp} it follows that the derivatives up to order $k$ are invariants of $\mathrm{pr}^{(k)}G$, hence also of any higher prolongation of $G$ (by \eqref{graprol}). So it only remains to show that they are functionally independent. But this is immediate because $\odk{w}{y}$ depends explicitly on $u_{k+1}$, hence is independent of the previously constructed invariants $y,w,\pd{w}{y},\dots,\od{^{k-1}w}{y^{k-1}}$, since those only depend on $x,u,\dots,u_k$. By \ref{lociprop}, since $G$ has one-dimensional orbits and $\dim X\times U^{(n)} = n+2$, completeness follows. \ep \begin{thr}{\bf Example. }\rm\label{rotsec} Applying the previous result to the first order invariants of the rotation group given in \ref{rotdi} it follows that $y$, $w$, and $$ \od{w}{y} = \frac{dw/dx}{dy/dx} = \frac{\sqrt{x^2+u^2}}{(x+uu_x)^3}[(x^2+u^2)u_{xx} -(1+u_x^2)(xu_x-u)] $$ together form a complete set of functionally independent invariants for $\mathrm{pr}^{(2)} G$. Thus any other second order differential invariant can be written as a smooth function of these invariants. For example, the curvature $\kappa$ from \eqref{curv} is an invariant of $\mathrm{pr}^{(2)} G$ (as one checks by verifying that $\mathrm{pr}^{(2)} v(\kappa) = 0$), and in fact $$ \kappa =\frac{u_{xx}}{(1+u_x^2)^{3/2}} = \frac{w_y}{(1+w^2)^{3/2}} + \frac{w}{y(1+w^2)^{1/2}}. $$ \et The important point to note now is that once the differential invariants of a local group of transformations $G$ on $M\subseteq X\times U=\mathbb R^2$ are known, we can determine all ODEs that admit $G$ as a symmetry group, i.e., all ODEs that can be integrated using $G$: \bt\label{it} Let $G$ be a connected local transformation group acting on $M\subseteq X\times U=\mathbb R^2$ and let $\eta^1(x,u^{(n)}),\dots,\eta^k(x,u^{(n)})$ be a complete set of functionally independent $n$-th order differential invariants of $G$. \begin{itemize} \item[(i)] If $\mathrm{pr}^{(n)} G$ acts semi-regularly on $M^{(n)}$, then an $n$-th order nondegenerate differential equation $P(x,u^{(n)})=0$ admits $G$ as a symmetry group if and only if, locally around any point in ${\mathcal S}_P=\{(x,u^{(n)})\mid P(x,u^{(n)})=0\}\subseteq M^{(n)}$, there is an equivalent equation $$ \tilde P(\eta^1(x,u^{(n)}),\dots,\eta^k(x,u^{(n)}))=0 $$ involving only the differential invariants of $G$. \item[(ii)] If $G$ is a one-parameter group of transformations, any nondegenerate $n$-th order differential equation having $G$ as a symmetry group is equivalent to an $(n-1)$-st order equation $$ \tilde P\Big(y,w,\od{w}{y},\dots,\od{^{n-1}w}{y^{n-1}}\Big) = 0, $$ where $y=\eta(x,u),\, w=\zeta(x,u,u_x)$ form a complete set of functionally independent invariants of $\mathrm{pr}^{(1)} G$. \end{itemize} \et \pr (i) Clearly if there is an equivalent equation depending only on invariants then ${\mathcal S}_P$ itself is locally invariant. Since $P$ is non-degenerate, ${\mathcal S}$ is a closed regular submanifold of $M^{(n)}$, so the proof of \ref{algsym} shows that this implies that ${\mathcal S}$ is invariant under $\mathrm{pr}^{(n)} G$, which in turn yields that $G$ is a symmetry group of $P$. Conversely, if $G$ is a symmetry group then the proof of \ref{nesu} shows that $\mathrm{pr}^{(n)} G$ leaves ${\mathcal S}_P$ invariant. The claim then follows from \ref{invvar}. (ii) This is immediate from (i) and \ref{oi}. \ep \begin{thr}{\bf Example. }\rm Continuing our analysis of the rotation group from \ref{rotsec}, we can use \ref{it} to find the most general first and second order ODEs that admit the rotation group as a symmetry group. By \ref{it} (ii), any such first order equation is equivalent to one involving only the first order invariants given in \ref{rotdi}. If we solve for $w$ it follows that any such equation is of the form $$ \frac{xu_x-u}{x+uu_x} = F(\sqrt{x^2+u^2}). $$ Analogously, any second order equation invariant under rotations is equivalent to one only involving $y$, $w$, and, say, the curvature $\kappa =\frac{u_{xx}}{(1+u_x^2)^{3/2}}$, hence is of the general form $$ u_{xx} = (1+u_x^2)^{3/2} F\Big(\sqrt{x^2+u^2},\frac{xu_x-u}{x+uu_x}\Big). $$ \et We may also employ \ref{it} to obtain an alternative way of reducing the order of a given differential equation $P(x,u^{(n)})=0$ once we know a one-parameter symmetry group. In fact, we know from \ref{it} (ii) that the equation must be equivalent to one involving only $y,w,\od{w}{y},\dots,\od{^{n-1}w}{y^{n-1}}$. But this latter equation automatically is of order $n-1$. Thus simply by re-expressing the equation in terms of the differential invariants already reduces the order by $1$. Furthermore, once the solution $w=h(y)$ of the reduced equation is known, the solution of the original equation is found by solving the auxiliary first order equation \begin{equation}\label{auxi} \zeta(x,u,u_x) = h(\eta(x,u)) \end{equation} obtained by substituting for $y$ and $w$ their expression in terms of $x$ and $u$. As \eqref{auxi} depends only on the invariants $\eta$, $\zeta$ of $\mathrm{pr}^{(1)} G$, it has $G$ as a one-parameter symmetry group, hence can be integrated using the methods described above for first order equations. \begin{thr}{\bf Example. }\rm The second order equation \begin{equation}\label{secco} x^2u_{xx} + xu_x^2 = uu_x \end{equation} is invariant under the scaling group $G: (x,u)\mapsto (\lambda x,\lambda u)$. In fact, the infinitesimal generator of $G$ is $v=\left.\od{}{\lambda}\right|_{\lambda=1} (\lambda x,\lambda u)= x\partial_x + u\partial_u$. Hence \eqref{phiform} gives $$ \mathrm{pr}^{(2)} v = x\partial_x + u\partial_u - u_{xx}\partial_{u_{xx}}, $$ so $$ \mathrm{pr}^{(2)} v(x^2u_{xx} + xu_x^2 - uu_x) = 2x^2u_{xx} + xu_x^2 -uu_x -x^2u_{xx} = 0 $$ for any $(x,u,u_x,u_{xx})\in {\mathcal S}_P$. By \ref{oi} and \ref{home}, a complete set of functionally independent invariants of second order is given by $$ y=\frac{u}{x},\quad w=u_x,\quad \od{w}{y} = \frac{dw/dx}{dy/dx}=\frac{x^2u_{xx}}{xu_x-u}. $$ Inserting this into \eqref{secco} gives the transformed equation $$ (w-y)\od{w}{y} + w^2 = yw, $$ which, as expected, is of first order. This has two families of solutions: either $w=y$ (which is a singular solution), or $\od{w}{y}=-w$, which implies $w=ce^{-y}$. Re-substituting gives $$ \od{u}{x}=\frac{u}{x}, \quad\text{ or }\quad \od{u}{x}=ce^{-u/x}. $$ This results in the one-parameter family of singular solutions $u=kx$, and the implicit solutions $$ \int\frac{dy}{ce^{-y}-y} = \log x + k, $$ with $y=u/x$, which constitutes the general solution to the original equation \eqref{secco}. \et \begin{thr}{\bf Example. }\rm We return to the two-dimensional heat equation $u_t=u_{xx}$, whose symmetry group we calculated in \ref{heatsym}, and use the methods of this section to determine the general form of {\em traveling wave solutions}\index{traveling wave solution} to the heat equation. In general, a traveling wave solution to a differential equation is one which is invariant under a translation group. Here we consider the symmetry of the heat equation given by the translation group $$ (x,t,u) \mapsto (x+c\varepsilon,t+\varepsilon,u),\quad \varepsilon\in \mathbb R, $$ with generator $v=\partial_t + c\partial_x$, for $c$ some fixed constant. Global invariants are given by \begin{equation}\label{heati} y=x-ct,\quad v=u. \end{equation} Thus any group-invariant solution is of the form $v=h(y)$, or $u=h(x-ct)$, which is a function that doesn't change its shape along the straight line $x-ct=$ const. If we express the derivatives of $u$ in terms of the new variables we obtain $$ u_t = -cv_y,\quad u_x=v_y,\quad u_{xx} = v_{yy}. $$ Inserting into the heat equation, we arrive at the ODE determining all traveling wave solutions of the above form: $$ -cv_y=v_{yy}, $$ with general solution $$ v(y) = ke^{-cy} + l $$ ($k$, $l$ arbitrary constants). If we now substitute back into the heat equation we obtain the general form of traveling wave solutions: $$ u(x,t) = ke^{-c(x-ct)} + l. $$ \et Much more can be said about group-invariant solutions of PDEs. In fact, there is an entire theory that allows to calculate solutions invariant under any subgroup of the full symmetry group of any given system of differential equations. While these methods also basically rest on choosing new variables from invariants of the given sub-group, the case of higher (than one) dimensional symmetry groups is significantly more involved than what we considered in this section. For a precise formulation, the theory of extended jet bundles and group invariant prolongation is required. We refer to \cite[Ch.\ 3]{O} for an in-depth study. \chapter{Variational symmetries} Conservation laws (like conservation of energy, momentum, etc.) play a central role in physics. In contemporary theoretical physics one usually formulates the governing principles of a theory in the form of a variational principle -- roughly, this says that some relevant quantity is to take an optimal value. It is a deep and far-reaching discovery of Emmy Noether (1918) that for such systems, every conservation law comes from a corresponding symmetry property. In the present chapter we give an introduction to this field. \section{Calculus of variations} Let $X=\mathbb R^p$, with coordinates $(x^1,\dots,x^p)$, and $U=\mathbb R^q$ with coordinates $(u^1,\dots,$ $u^q)$. Also, let $\Omega$ be a connected open subset of $X$ with smooth boundary $\partial \Omega$. By a {\em variational problem}\index{variational problem} we mean the problem of finding extrema (maxima, minima, or stationary points) of the functional $$ {\mathcal L}[u] = \int_\Omega L(x,u^{(n)}(x))\,dx $$ in some class of admissible functions $u=f(x)$ on $\Omega$. The integrand $L:X\times U^{(n)}\to \mathbb R$ is called the {\em Lagrangian}\index{Lagrangian} of the variational problem ${\mathcal L}$. The class of admissible functions varies with the problem under consideration (e.g., through the imposition of various boundary or regularity conditions). We will confine ourselves here to studying smooth variational problems. The calculus of variations is a vast subject and we will here barely be able to scratch the surface, with a focus on symmetry methods. For a gentle introduction see, e.g., \cite{Sa}, a rather comprehensive treatise is \cite{GH1,GH2}. \begin{thr}{\bf Example. }\rm\label{first} (i) To find the shortest curve $u=f(x)$ connecting two points $(a,b)$, $(c,d)$ in the plane, one needs to minimize the length of $u=f(x)$: $$ {\mathcal L}[u] = \int_a^c \sqrt{1+u_x^2}\,dx. $$ (ii) More generally, if $(M,g)$ is a smooth Riemannian manifold and $x$, $y\in M$ then the problem of finding the shortest curve from $x$ to $y$ is a variational problem, namely that of minimizing the arc-length of curves $u\colon [0,1]\to M$ connecting $x$ and $y$: $$ {\mathcal L}[u] = \int_0^1 \sqrt{g(u'(t),u'(t))}\,dt. $$ Solutions to this problem are called geodesics of $(M,g)$. \et The basic approach to finding extremals of variational problems is very similar to that of extremizing smooth functions in basic real analysis. There, given a function $f\colon \Omega\to \mathbb R$, one notes that if $x$ is an extremum of $f$ then for any $y$, the one-dimensional function $\varepsilon\mapsto f(x+\varepsilon y)$ must have an extremal. Therefore, we must have $$ 0= \left.\frac{d}{d \eps}\right|_0 f(x+\varepsilon y) = \langle \nabla f(x),y\rangle \quad (y\in \mathbb R^p), $$ and so the gradient $\nabla f$ of $f$ at $x$ vanishing is a necessary condition for $x$ being an extremum of $f$. For functionals ${\mathcal L}[u]$, the role of the gradient is taken over by what is called the {\em variational derivative}.\index{variational derivative} Moreover, the inner product on $\mathbb R^n$ gets replaced by the $L^2$-inner product of functions $f,\,g\colon \Omega\to \mathbb R^q$: $$ \langle f,g\rangle = \int_\Omega f(x)\cdot g(x)\,dx = \int_\Omega \sum_{\alpha=1}^q f^\alpha(x)g^\alpha(x)\,dx. $$ With these notations we can define: \begin{thr} {\bf Definition. } Let ${\mathcal L}[u]$ be a variational problem. The variational derivative\index{variational derivative} of ${\mathcal L}$ is the unique $q$-tuple $$ \delta{\mathcal L}[u] = (\delta_1{\mathcal L}[u],\dots,\delta_q{\mathcal L}[u]), $$ $\delta{\mathcal L}:{\mathcal C}^\infty(\Omega)\to {\mathcal C}^\infty(\Omega)^q$ with the property that \begin{equation}\label{varder} \left.\frac{d}{d \eps}\right|_0{\mathcal L}[f+\varepsilon\eta] = \int_\Omega \delta{\mathcal L}[f](x)\cdot \eta(x)\,dx \end{equation} for every smooth function $f$ on $\Omega$ and any compactly supported smooth function $\eta\in {\mathcal D}(\Omega)^q$ such that $f+\varepsilon\eta$ is itself admissible. $\delta_\alpha{\mathcal L}$ is called the variational derivative of ${\mathcal L}$ with respect to $u^\alpha$. \et Here, by ${\mathcal D}(\Omega)$ we denote the space of test functions on $\Omega$, ${\mathcal D}(\Omega)=\{f\in {\mathcal C}^\infty(\Omega)\mid \mathrm{supp}(f)\comp \Omega\}$. \bp If $u=f(x)$ is an extremal of ${\mathcal L}[u]$, then \begin{equation}\label{cv0} \delta{\mathcal L}[f](x)=0 \quad \forall x\in \Omega. \end{equation} \et \pr For any $\eta\in {\mathcal D}(\Omega)^q$ and for all $\varepsilon$ small, $f+\varepsilon\eta$ is admissible, and by assumption, $\varepsilon\mapsto {\mathcal L}[f+\varepsilon\eta]$ must have an extremum at $\varepsilon=0$. Hence by classical real analysis, $$ 0=\left.\frac{d}{d \eps}\right|_0{\mathcal L}[f+\varepsilon\eta] = \int_\Omega \delta{\mathcal L}[f](x)\cdot \eta(x)\,dx. $$ As this has to hold for any $\eta$, the claim follows. \ep Using the notion of total derivative, cf. \eqref{totder1}, we now wish to derive an explicit formula for the variational derivative. First, \begin{equation*} \begin{split} \left.\frac{d}{d \eps}\right|_0{\mathcal L}[f+\varepsilon\eta] &= \int_\Omega \left.\frac{d}{d \eps}\right|_0 L(x,\mathrm{pr}^{(n)}(f+\varepsilon\eta)(x))\,dx\\ &= \int_\Omega \Big( \sum_{\alpha,J} \pd{L}{u^\alpha_J}(x,\mathrm{pr}^{(n)} f(x))\cdot \partial_J\eta^\alpha(x)\Big)\,dx. \end{split} \end{equation*} Since $\eta$ has compact support, we may apply the divergence theorem to this expression, with all boundary terms vanishing. Note also that when a partial derivative $\partial_{x_j}$ is applied to $\partial L/\partial u^\alpha_J$ then the result can be expressed by a total derivative, so \begin{equation} \left.\frac{d}{d \eps}\right|_0{\mathcal L}[f+\varepsilon\eta] = \int_\Omega \sum_{\alpha=1}^q \big[\sum_J (-D)_J\pd{L}{u^\alpha_J} (x,\mathrm{pr}^{(n+\sharp J)} f(x)) \big]\eta^\alpha(x)\,dx. \label{mcl} \end{equation} Here, for $J=(j_1,\dots,j_k)$ we set $(-D)_J:=(-D_{j_1})(-D_{j_2})\dots(-D_{j_k})$. This formula is the first instance of the following operator that will play a central role below. \begin{thr} {\bf Definition. } For $1\le \alpha\le q$, the $\alpha$-th Euler operator\index{Euler operator} is defined by \begin{equation}\label{eulerop} E_\alpha = \sum_J (-D)_J \pd{}{u^\alpha_J}, \end{equation} where the sum extends over all $J=(j_1,\dots,j_k)$ with $1\le j_\kappa\le p$, $k\ge 0$. Moreover, we set $E(L):=(E_1(L),\dots,E_q(L))$. \et Although the above sum is formally infinite, whenever $E_\alpha$ is applied to some concrete $L(x,u^{(n)})$ only finitely many terms are required. In terms of the Euler operator, \eqref{mcl} says that $$ \delta{\mathcal L} = E(L) $$ Together with \ref{cv0}, this gives the classical necessary condition for extremals of a variational problem: \bt If $u=f(x)$ is a smooth extremal of the variational problem ${\mathcal L}[u]=\int_\Omega L(x,u^{(n)})\,dx$ then it must be a solution of the {\em Euler--Lagrange equations}\index{Euler--Lagrange equations} $$ E_\alpha(L)=0, \quad \alpha=1,\dots,q. $$ \hspace*{\fill}$\Box$ \et \begin{thr}{\bf Example. }\rm Let $p=q=1$, i.e., we consider a variational problem for a real-valued function of one variable. In this case, $$ E = \sum_{j=0}^\infty (-D_x)^j\pd{}{u_j} = \pd{}{u} - D_x\pd{}{u_x} + D_x^2\pd{}{u_{xx}}-\dots $$ Thus the Euler--Lagrange equation for an $n$-th order variational problem $$ {\mathcal L}[u] = \int_a^b L(x,u^{(n)})\,dx $$ is given by (setting $u_j=\pd{^j u}{x^j}$) $$ 0=E(L) = \pd{L}{u} - D_x\pd{L}{u_x} + D_x^2\pd{L}{u_{xx}} -\dots + (-1)^n D_x^n\pd{L}{u_n}. $$ This is an ODE of order (at most) $2n$. The most common case of a first order variational problem $L=L(x,u,u_x)$ gives $$ 0=E(L) = \pd{L}{u} - D_x\pd{L}{u_x} = \pd{L}{u} - \frac{\partial^2 L}{\partial x\partial u_x} - u_x \frac{\partial^2 L}{\partial u\partial u_x} - u_{xx} \pd{^2L}{u_x^2}. $$ In particular, for the curve length problem \ref{first} the Euler--Lagrange equation reads $$ -D_x\left(\frac{u_x}{\sqrt{1+u_x^2}}\right) = - \frac{u_{xx}}{(1+u_x^2)^{3/2}} = 0, $$ i.e., $u_{xx}=0$. As geometrically expected, the solutions are straight lines $u=kx+d$. \et \begin{thr}{\bf Example. }\rm The Dirichlet principle:\index{Dirichlet principle} Consider the variational problem of minimizing the total energy (kinetic plus potential) of some system in the form $$ {\mathcal L}[u] = \int_\Omega \frac{1}{2}|\nabla u|^2 - u h\,dx $$ with some external potential $h$. Then the Euler--Lagrange equation reads $$ 0=E(L)=\pd{L}{u} - \sum_{i=1}^p D_i\pd{L}{u_i} = \pd{L}{u}-\sum_{i=1}^p D_i u_i = -h -\Delta u, $$ i.e., we obtain the Poisson equation\index{Poisson equation} $-\Delta u = h$. \et \section{Variational symmetries} In this section we want to develop a notion of symmetry that applies to variational problems, similar to the symmetry groups of differential equations studied in the previous chapter. Consider a variational problem \begin{equation}\label{vp} {\mathcal L}[u] = \int_{\Omega_0} L(x,u^{(n)})\,dx. \end{equation} Let $G$ be a local transformation group on an open subset $M$ of $\Omega_0\times U\subseteq X\times U$. If $u=f(x)$ is a smooth function on a sufficiently small subdomain $\Omega\subseteq\Omega_0$ such that the graph of $f$ lies in $M$ then any transformation $g\in G$ sufficiently close to the identity will transform $f$ into another smooth function ${\tilde u}={\tilde f}({\tilde x}) =(g\cdot f) ({\tilde x})$ defined on some $\tilde\Omega\subseteq\Omega_0$. Therefore, we define: \begin{thr} {\bf Definition. } A local transformation group $G$ acting on $M\subseteq \Omega_0\times U$ is called a {\em variational symmetry}\index{variational symmetry} of the functional \eqref{vp} if whenever $\Omega$ is a subdomain with $\overline{\Omega}\subseteq\Omega_0$, $u=f(x)$ is a function defined over $\Omega$ with graph contained in $M$, and $g\in G$ is such that ${\tilde u}={\tilde f}({\tilde x}) =(g\cdot f) ({\tilde x})$ is a well-defined function defined on $\tilde\Omega\subseteq \Omega_0$, then \begin{equation}\label{varsym} \int_{\tilde\Omega} L({\tilde x},\mathrm{pr}^{(n)} {\tilde f}({\tilde x}))\,d{\tilde x} = \int_\Omega L(x,\mathrm{pr}^{(n)} f(x))\,dx. \end{equation} \et Our first aim, in line with our previous considerations on symmetries of differential equations, is to derive an infinitesimal criterion for variational symmetries. For this, we need a few preparations. To begin with, we need to study how a variational problem transforms under the action of a local transformation group. Let \begin{equation}\label{chvar} {\tilde x} = \Xi(x,u), \qquad {\tilde u} = \Phi(x,u) \end{equation} be any change of variables. Then there is an induced change of variables $$ {\tilde u}^{(n)} = \Phi^{(n)}(x,u^{(n)}) $$ for the derivatives, given by prolongation (i.e., by differentiating \eqref{chvar} by the chain rule). If the conditions of the inverse function theorem are satisfied then \eqref{chvar} determines a transformed function ${\tilde u}={\tilde f}({\tilde x})$. Thereby, any functional $$ {\mathcal L}[f] = \int_\Omega L(x,\mathrm{pr}^{(n)} f(x))\,dx $$ is transformed into a new functional $$ \tilde {\mathcal L}[{\tilde f}] = \int_{\tilde\Omega} \tilde L({\tilde x},\mathrm{pr}^{(n)} {\tilde f}({\tilde x}))\,d{\tilde x}. $$ Here, the new domain $\tilde\Omega = \{{\tilde x}=\Xi(x,f(x))\mid x\in \Omega\}$ depends both on $\Omega$ and on $f$. By the change of variables formula we get \begin{equation}\label{trvar} L(x,\mathrm{pr}^{(n)} f(x)) = \tilde L({\tilde x},\mathrm{pr}^{(n)} {\tilde f}({\tilde x}))\det J(x,\mathrm{pr}^{(1)} f(x)), \end{equation} where $J$ is the Jacobian matrix with entries \begin{equation}\label{jacob} J^{ij}(x,\mathrm{pr}^{(1)} f(x)) = \pd{}{x^j}\Xi^i(x,f(x)) = D_j\Xi^i(x,\mathrm{pr}^{(1)} f(x)) \quad i,j=1,\dots,p, \end{equation} where, for simplicity, we assume that $\det J(x)>0$, otherwise we have to add absolute value signs. \begin{thr} {\bf Definition. }\label{totdiv} Let $x=(x^1,\dots,x^p)$ and let $P=P(x,u^{(n)})=(P_1(x,u^{(n)}),\dots,$ $P_p(x,u^{(n)}))$ be a $p$-tuple of smooth functions. Then the {\em total divergence}\index{total divergence} of $P$ is $$ \mathrm{Div}\, P = D_1P_1+\dots +D_pP_p, $$ where $D_j$ denotes the total derivative with respect to $x^j$. \et With these notations, we have: \blem Let $v=\sum_i\xi^i\partial_{x^i} +\sum_\alpha \phi^\alpha\partial_{u^\alpha}$ be a vector field on $M\subseteq \Omega_0\times U$ and set $g_\varepsilon:=\mathrm{Fl}^{v}_\varepsilon = (\Xi_{g_\varepsilon},\Phi_{g_\varepsilon})$. Denoting by $J_{g_\varepsilon}$ the Jacobian \eqref{jacob} corresponding to $\Xi_{g_\varepsilon}$, we have \begin{equation}\label{jacobform} \frac{d}{d\varepsilon}(\det J_{g_\varepsilon}(x,u^{(1)})) = (\mathrm{Div}\,\xi)(\mathrm{pr}^{(1)} g_\varepsilon\cdot (x,u^{(1)}))\cdot\det J_{g_\varepsilon}(x,u^{(1)}). \end{equation} \et \pr Recall first that for matrices $A$, $B$ with $A$ invertible, the derivative of the determinant function at $A$ in the direction $B$ is given by $\det'(A)(B) = \det(A)\cdot\mathrm{tr}(A^{-1}B)$. Consequently, for any fixed function $u$ we get \begin{equation}\label{jumf} \frac{d}{d\varepsilon}\det J_{g_\varepsilon} = \det{}'(J_{g_\varepsilon})\Big(\frac{d}{d\varepsilon}J_{g_\varepsilon}\Big) = \det J_{g_\varepsilon}\cdot\mathrm{tr}(J_{g_\varepsilon}^{-1}\frac{d}{d\varepsilon}J_{g_\varepsilon}). \end{equation} By \eqref{jacob}, $J_{g_\varepsilon}(x,u^{(1)}(x)) = \partial_x(\Xi_{g_\varepsilon}(x,u(x)))$ (with $\partial_x$ abbreviating the $x$-gradient, and similar for $\partial_1$, $\partial_2$ below), so \begin{equation*} \begin{split} &\frac{d}{d\varepsilon}(J_{g_\varepsilon}(x,u^{(1)})(x)) = \partial_x\Big(\frac{d}{d \eps}\Xi_{g_\varepsilon}(x,u(x))\Big) =\partial_x(\xi(\Xi_{g_\varepsilon}(x,u(x)),\Phi_{g_\varepsilon}(x,u(x))))\\ &=(\partial_1\xi)(\Xi_{g_\varepsilon}(x,u(x)),\Phi_{g_\varepsilon}(x,u(x)))\partial_x(\Xi_{g_\varepsilon}(x,u(x)))\\ & \hphantom{=}+ (\partial_2\xi)(\Xi_{g_\varepsilon}(x,u(x)),\Phi_{g_\varepsilon}(x,u(x)))\partial_x(\Phi_{g_\varepsilon}(x,u(x)))\\ &=[(\partial_1\xi)(\dots) + \partial_2\xi(\dots) \cdot \partial_x(\Phi_{g_\varepsilon}(x,u(x)))\cdot \partial_x(\Xi_{g_\varepsilon}(x,u(x)))^{-1}] \partial_x(\Xi_{g_\varepsilon}(x,u(x))) \end{split} \end{equation*} Now let $\tilde u$ be the transformed function $g_\varepsilon\cdot u$, so that with ${\tilde x}(x)=\Xi_{g_\varepsilon}(x,u(x))$ we have ${\tilde u}({\tilde x}(x))=\Phi_{g_\varepsilon}(x,u(x))$. Then $$ \partial_x(\Phi_{g_\varepsilon}(x,u(x))) = \partial_x({\tilde u}\circ{\tilde x})(x) = \partial_{\tilde x}{\tilde u}({\tilde x}(x))\cdot \partial_x{\tilde x}(x) =\partial_{\tilde x}{\tilde u}({\tilde x}(x))\cdot \partial_x(\Xi_{g_\varepsilon}(x,u(x))), $$ so, together with the above calculations we obtain \begin{equation*} \begin{split} \frac{d}{d\varepsilon}(J_{g_\varepsilon}(x,u^{(1)})(x)) &= [(\partial_1\xi)({\tilde x},{\tilde u}({\tilde x}))+(\partial_2\xi)({\tilde x},{\tilde u}({\tilde x})) \cdot \partial_{\tilde x}{\tilde u}({\tilde x}(x))]\cdot J_{g_\varepsilon}(x,u^{(1)}(x))\\ &= D_{\tilde x} \xi(\mathrm{pr}^{(1)} g_\varepsilon(x,u^{(1)}(x)))\cdot J_{g_\varepsilon}(x,u^{(1)}(x)). \end{split} \end{equation*} Finally, using \eqref{jumf}, and the fact that $\mathrm{tr}(AB)=\mathrm{tr}(BA)$, we arrive at \begin{equation*} \begin{split} \frac{d}{d\varepsilon}(\det J_{g_\varepsilon}(x,u^{(1)}(x))) &= \det(J_{g_\varepsilon}(x,u^{(1)}(x)))\cdot \mathrm{tr}(D_{\tilde x} \xi(\mathrm{pr}^{(1)} g_\varepsilon(x,u^{(1)}(x))))\\ &= \det(J_{g_\varepsilon}(x,u^{(1)}(x)))\cdot (\mathrm{Div}\, \xi)(\mathrm{pr}^{(1)} g_\varepsilon(x,u^{(1)}(x))). \end{split} \end{equation*} \ep The infinitesimal criterion for variational symmetries now is given by the following result. \bt Let $G$ be a connected local transformation group acting on $M\subseteq \Omega_0\times U$. Then $G$ is a variational symmetry group of the variational problem \eqref{vp} if and only if \begin{equation}\label{varcrit} \mathrm{pr}^{(n)} v(L) + L\cdot\mathrm{Div}\,\xi = 0 \end{equation} for all $(x,u^{(n)})\in M^{(n)}$ and every infinitesimal generator $$ v = \sum_{i=1}^p \xi^i(x,u)\pd{}{x^i} + \sum_{\alpha=1}^q \phi_\alpha(x,u)\pd{}{u^\alpha} $$ of $G$. \et \pr As usual, for $g\in G$ we write $({\tilde x},{\tilde u})=g\cdot(x,u)=(\Xi_g(x,u),\Phi_g(x,u))$. By definition of a variational symmetry, see \eqref{varsym}, we get \begin{equation*} \begin{split} \int_\Omega L(x,\mathrm{pr}^{(n)} f(x))\,dx &= \int_{\tilde\Omega} L({\tilde x},\mathrm{pr}^{(n)} {\tilde f}({\tilde x}))\,d{\tilde x} \\ &\stackrel{\eqref{trvar}}{=} \int_{\Omega} L({\tilde x}(x),\mathrm{pr}^{(n)} {\tilde f}({\tilde x}(x)))\det J_g(x,\mathrm{pr}^{(1)} f(x))\,dx \end{split} \end{equation*} Since this equality has to hold for all subdomains $\Omega$ and all functions $f$, the integrands must agree pointwise, so $g$ is a variational symmetry if and only if \begin{equation}\label{pointa} L(\mathrm{pr}^{(n)} g \cdot(x,u^{(n)})) \det J_g(x,u^{(1)}) = L(x,u^{(n)}) \quad \forall (x,u^{(n)})\in M^{(n)} \end{equation} for all $g$ such that both sides are defined. We now insert $g=g_\varepsilon=\mathrm{Fl}^v_\varepsilon$ and differentiate with respect to $\varepsilon$, which by \eqref{jacobform} and the product rule yields \begin{equation}\label{jacobdiff} (\mathrm{pr}^{(n)} v(L) + L\cdot \mathrm{Div}\,\xi)|_{(\mathrm{pr}^{(n)} g_\varepsilon\cdot(x,u^{(n)}))}\cdot \det J_{g_\varepsilon}(x,u^{(1)}) = 0 \end{equation} Setting $\varepsilon=0$, i.e., $g_\varepsilon=\mathrm{id}$, this formula reduces to \eqref{varcrit}, proving necessity. Conversely, if \eqref{varcrit} is satisfied then also \eqref{jacobdiff} holds for $\varepsilon$ sufficiently small. Therefore, the derivative of \eqref{pointa} (with $g=g_\varepsilon$) with respect to $\varepsilon$ vanishes identically. Integrating from $0$ to $\varepsilon$ then shows that \eqref{pointa} holds for any $g_\varepsilon$. Since $G$ is connected the result then follows since any $g\in G$ can be written as a product of such $g_\varepsilon$. \ep \begin{thr}{\bf Example. }\rm Returning to \eqref{first} (i), where ${\mathcal L}[u] = \int_a^c \sqrt{1+u_x^2}\,dx$, we expect that rotations should be variational symmetries, as they leave curve lengths invariant. To verify this, let $v=-u\partial_x + x\partial_u$ be a generator of the rotation group on $\mathbb R^2$. Then by \eqref{rotvprol}, $$ \mathrm{pr}^{(1)} v = -u\partial_x + x\partial_u + (1+u_x^2)\partial_{u_x}. $$ In particular $\xi=-u$, and we calculate $$ \mathrm{pr}^{(1)} v(L) +L\cdot D_x\xi = (1+u_x^2)\pd{}{u_x}\sqrt{1+u_x^2} - \sqrt{1+u_x^2}\cdot u_x =0, $$ so indeed \eqref{varcrit} is satisfied. \et Next, we want to clarify the relationship between variational symmetries of some variational problem and symmetries of the corresponding Euler--Lagrange equations. To this end, we first need to understand how the Euler--Lagrange equations of variational problems transform under a change of variables. \bp Let $L(x,u^{(n)})$, $\tilde L({\tilde x},{\tilde u}^{(n)})$ be two Lagrangians related by a change of variables given by \eqref{chvar}, \eqref{trvar}, i.e., $$ {\tilde x} = \Xi(x,u), \qquad {\tilde u} = \Phi(x,u) $$ and $$ L(x,u^{(n)}) = \tilde L({\tilde x},{\tilde u}^{(n)})\det J(x,u^{(1)}). $$ Then \begin{equation}\label{eulertrans} E_{u^\alpha}(L)(x,u^{(2n)}) = \sum_{\beta=1}^q F_{\alpha\beta}(x,u^{(1)})E_{{\tilde u}^\beta}(\tilde L)({\tilde x},{\tilde u}^{(2n)}), \quad \alpha=1,\dots,q, \end{equation} where $F_{\alpha\beta}$ is the determinant of the following $(p+1)\times (p+1)$-matrix: \begin{equation}\label{fmatrix} F_{\alpha\beta} = \det \begin{pmatrix} D_1\Xi^1 & \dots & D_p\Xi^1 & \pd{\Xi^1}{u^\alpha}\\ \vdots & & \vdots & \vdots\\ D_1\Xi^p & \dots & D_p\Xi^p & \pd{\Xi^p}{u^\alpha}\\ D_1\Phi^\beta & \dots & D_p\Phi^\beta & \pd{\Phi^\beta}{u^\alpha}\\ \end{pmatrix}. \end{equation} \et \pr Let $(x_0,u_0^{(2n)})$ be any fixed point in $X\times U^{(2n)}$ and let $f$ be a smooth function defined near $x_0$ with $(x_0,u_0^{(2n)})=(x_{0},\mathrm{pr}^{(2n)}f(x_0))$. By the inverse function theorem, on an open ball $\Omega$ around $x_0$, $f$ is transformed by our change of variables into a function ${\tilde u}={\tilde f}({\tilde x})$. If $\eta\in {\mathcal D}(\Omega)^q$ and $\varepsilon$ is sufficiently small, then the perturbations $u_\varepsilon = f(x,\varepsilon) = f(x) +\varepsilon\eta(x)$, required for calculating variational derivatives, are transformed into functions ${\tilde u} = {\tilde f}({\tilde x},\varepsilon)$, which are determined implicitly by \begin{equation}\label{impldet} {\tilde x} = \Xi(x,f(x)+\varepsilon\eta(x))=:\Xi_\varepsilon(x),\quad {\tilde u}=\Phi(x,f(x)+\varepsilon\eta(x))=:\Phi_\varepsilon(x). \end{equation} Again the inverse function theorem determines the domain of each ${\tilde f}(\,.\,,\varepsilon)$. More precisely, by \cite[Suppl.\ 2.5 A]{AMR}, the minimal radius of a ball around $\tilde x_\varepsilon:= \Xi(x,f(x_0)+\varepsilon\eta(x_0))$ where ${\tilde f}(\,.\,,\varepsilon)$ is defined can be estimated from below in terms of the second derivatives of the maps $\Xi_\varepsilon$ from \eqref{impldet} in a neighborhood of $x_0$ as well as the norm of the inverse of their Jacobians at $x_0$. Since $\eta$ is compactly supported, $\overline \Omega$ is compact, and ${\tilde x}_\varepsilon\to {\tilde x}_0$ as $\varepsilon\to 0$, by choosing $\varepsilon$ small we can therefore achieve that all these balls contain a fixed neighborhood of $\tilde x_0$, which we then take as a common domain $\tilde \Omega$ of all ${\tilde f}(\,.\,,\varepsilon)$, independently of $\varepsilon$. In fact, since the radius of the ball $\Omega$ around $x_0$ is itself determined by the mentioned estimates (with $\varepsilon=0$), we may in addition achieve that $\mathrm{supp}(\eta)$ is contained in $\Xi_\varepsilon^{-1}(\tilde\Omega)$ for $\varepsilon$ small. Now set $\tilde \eta({\tilde x},\varepsilon) := {\tilde f}({\tilde x},\varepsilon)-{\tilde f}({\tilde x})$ and note that $\Xi_\varepsilon(x)=\Xi_0(x)$ as well as $\Phi_\varepsilon(x)=\Phi_0(x)$, and thereby ${\tilde f}(\Xi_\varepsilon(x),\varepsilon)$ $ = $ $ {\tilde f}(\Xi_0(x))$ for $x\not\in \mathrm{supp}(\eta)$. This implies that $\tilde\eta(\tilde y,\varepsilon)=0$ for $\tilde y \not\in \Xi_\varepsilon(\mathrm{supp}(\eta))$. Again due to the compactness of $\mathrm{supp}(\eta)$ this shows that, for $\varepsilon$ sufficiently small, all $\tilde\eta(\,.\,,\varepsilon)$ have compact support contained in $\tilde \Omega$. The same is therefore true of $\pd{{\tilde f}}{\varepsilon}\big|_{\varepsilon=0} = \pd{\tilde\eta}{\varepsilon}\big|_{\varepsilon=0}$. Using this, exactly as in \eqref{mcl} we obtain \begin{equation}\label{ltifti} \left.\frac{d}{d \eps}\right|_0 \tilde{\mathcal L}[{\tilde f}] = \int_{\tilde\Omega} E_{\tilde u}(\tilde L)\cdot \left.\pd{{\tilde f}}{\varepsilon}\right|_{\varepsilon=0}\,d{\tilde x}, \end{equation} where $E_{\tilde u}(\tilde L)$ is evaluated at ${\tilde u}={\tilde f}$. To continue, we need to determine $\pd{{\tilde f}}{\varepsilon}$. To calculate ${\tilde f}({\tilde x},\varepsilon)$, we need to solve the first equation in \eqref{impldet} for $x$ and then insert into the second equation. Thus we have $x=x({\tilde x},\varepsilon)$ with $$ {\tilde x}=\Xi(x({\tilde x},\varepsilon),f(x({\tilde x},\varepsilon))+\varepsilon\eta(x({\tilde x},\varepsilon))). $$ When computing variations of $\tilde L$, the base variables are not allowed to depend on $\varepsilon$, so for $i=1,\dots,p$ we must have $$ \pd{{\tilde x}^i}{\varepsilon} = 0 = \sum_{j=1}^p D_j\Xi^i\pd{x^j}{\varepsilon} + \sum_{\alpha=1}^q \pd{\Xi^i}{u^\alpha}\eta^\alpha. $$ Using Cramer's rule we can solve for $\pd{x^j}{\varepsilon}$ (and insert $\varepsilon=0$), to obtain $$ \left.\pd{x^j}{\varepsilon}\right|_{\varepsilon=0} = -\frac{1}{\det J}\sum_{i=1}^p K_{ij}\sum_{\alpha=1}^q \pd{\Xi^i}{u^\alpha}\eta^\alpha, $$ where $K_{ij}$ is the $(i,j)$-cofactor (i.e., signed minor) of the Jacobian $J(x)$ from \eqref{jacob}. Consequently, \begin{equation*} \begin{split} \left.\pd{{\tilde f}^\beta}{\varepsilon}\right|_{\varepsilon=0} &= \sum_{\al=1}^q \pd{\Phi^\beta}{u^\alpha}\eta^\alpha + \sum_{j=1}^p D_j\Phi^\beta \left.\pd{x^j}{\varepsilon}\right|_{\varepsilon=0}\\ &= \frac{1}{\det J} \sum_{\al=1}^q \Big[\pd{\Phi^\beta}{u^\alpha}\det J - \sum_{i,j=1}^p D_j\Phi^\beta\cdot K_{ij} \pd{\Xi^i}{u^\alpha}\Big]\eta^\alpha. \end{split} \end{equation*} Here, the term in brackets is the expansion of the determinant \eqref{fmatrix} along the last column, wherein $\sum_j D_j\Phi^\beta\cdot K_{ij}$ is the row expansion of the $(i,p+1)$-st minor along its last row (recall that the necessary signs are already contained in the $K_{ij}$). Thus we have shown that $$ \left.\pd{{\tilde f}^\beta}{\varepsilon}\right|_{\varepsilon=0} = \frac{1}{\det J}\sum_{\al=1}^q F_{\alpha\beta}\eta^\alpha. $$ If we insert this into \eqref{ltifti} and change variables we get $$ \left.\frac{d}{d \eps}\right|_0 \tilde{\mathcal L}[{\tilde f}] = \int_{\Omega}\Big(\sum_{\alpha,\beta=1}^q F_{\alpha\beta}E_{\tilde u^\beta}(\tilde L)\cdot \eta^\alpha\Big)\,dx, $$ which, on the other hand, must equal $$ \left.\frac{d}{d \eps}\right|_0 {\mathcal L}[f+\varepsilon\eta] = \int_{\Omega}\Big(\sum_{\alpha=1}^q E_{u^\alpha}(L)\cdot \eta^\alpha\Big)\,dx. $$ Since $\eta$ was arbitrary, this proves that \eqref{eulertrans} holds at $(x_0,u_0^{(2n)})$. \ep Based on this result, we can now show: \bt If $G$ is a variational symmetry group of the functional $$ {\mathcal L}[u]=\int_{\Omega_0}L(x,u^{(n)})\,dx, $$ then $G$ is also a symmetry group of the corresponding Euler--Lagrange equations $E(L)=0$. \et \pr Let $g\in G$ and let $f$ be a solution to the Euler--Lagrange equations. Then since $G$ is a variational symmetry group, \eqref{pointa} shows that $L({\tilde x},\mathrm{pr}^{(n)} {\tilde f}({\tilde x}))$ and $L(x,\mathrm{pr}^{(n)} f(x))$ are related by the change of variables formula \eqref{trvar} with $L=\tilde L$, where $g^{-1}\cdot(x,u)=:(\Xi(x,u),\Phi(x,u))$. Interchanging the roles of $u$ and ${\tilde u}$ in \eqref{eulertrans} we therefore have $$ E_{\tilde u^\alpha}(L)({\tilde x},\tilde u^{(n)}) = \sum_{\beta=1}^q F_{\alpha\beta}({\tilde x},\tilde u^{(1)}) E_{u^\beta}(L)(x,u^{(n)}), \quad \alpha=1,\dots,q. $$ Since $f$ is a solution to the Euler--Lagrange equations, $E_{u^\beta}(L)(x,\mathrm{pr}^{(n)} f(x))=0$ for all $\beta$ and all $x$. Thus also $E_{\tilde u^\alpha}(L)({\tilde x},\mathrm{pr}^{(n)} {\tilde f}({\tilde x}))=0$, so $g\cdot f$ is a solution as well, i.e., $G$ is a symmetry group of $E(L)=0$. \ep \section{Conservation laws}\index{conservation law} \begin{thr} {\bf Definition. } Let $P(x,u^{(n)})=0$ be a system of differential equations. A {\em conservation law} is a divergence expression \begin{equation}\label{cl} \mathrm{Div}\, F = 0, \end{equation} which vanishes for all solutions $u=f(x)$ of the system. Here, for some $m$, $$ F=(F_1(x,u^{(m)}),\dots,F_p(x,u^{(m)})) $$ is a $p$-tuple of smooth functions and $\mathrm{Div}\, F$ is the total divergence from \ref{totdiv}. $F$ then is called a conserved current.\index{conserved current} \et \begin{thr}{\bf Example. }\rm (i) Let $P=\Delta(u)$, where $\Delta$ is the Laplace-operator. Then $\Delta$ itself is a conservation law because $$ \Delta u = \mathrm{Div}\,(\mathrm{grad} u) = 0 $$ for every solution. Further conservation laws are obtained by multiplying the equation with $u_i = \pd{u}{x^i}$: $$ 0 = u_i\Delta u = \sum_{j=1}^p D_j\Big(u_iu_j - \frac{1}{2}\delta^j_i\sum_{k=1}^p u_k^2\Big). $$ (ii) If $P$ is any system of ODEs, with independent variable $x$, then any conservation law is of the form $D_xF=0$ for all solutions $u=f(x)$ of the system, i.e., $F(x,u^{(n)})$ has to be constant along solutions, so $F$ is a first integral\index{first integral} of the system (cf.\ the remarks following \eqref{invode}). In light of this, \eqref{cl} can be viewed as a generalization of the concept of a first integral to PDEs. \et In applications to physics, often there is a distinguished time-variable $t$, while the remaining variables $(x^1,\dots,x^p)$ are spatial variables. Then any conservation law takes the form $$ D_tT + \mathrm{Div}\, X = 0, $$ where $\mathrm{Div}\,$ now is the spatial total divergence (with respect to $(x^1,\dots,x^p)$). Then $T$ is called the (conserved) {\em density}\index{density} and $X=(X_1,\dots,X_p)$ is called the {\em flux}\index{flux} of the conservation law. They are functions of $x,t,u$ and the derivatives of $u$ with respect to $x$ and $t$. Suppose now that $\Omega\subseteq\mathbb R^p$ is a spatial domain and let $u=f(x,t)$ be a solution defined for all $x\in \Omega$ and $a\le t\le b$. Consider the functional \begin{equation}\label{tofdef} {\mathcal T}_\Om[f](t) = \int_\Omega T(x,t,\mathrm{pr}^{(m)} f(x,t))\,dx, \end{equation} which, for $f$, $\Omega$ fixed, depends only on $t$. Then we have: \bp $T$ and $X$ are the density and the flux of a conservation law of a system of differential equations if and only if, for any bounded domain $\Omega\subseteq\mathbb R^p$ with smooth boundary $\partial\Omega$ and for any solution $u=f(x,t)$ defined for $x\in\Omega$ and $t\in [a,b]$ we have \begin{equation}\label{tofcons} {\mathcal T}_\Om[f](t) - {\mathcal T}_\Om[f](a) = -\int_a^t\int_{\partial\Omega} X(x,\tau,\mathrm{pr}^{(m)} f(x,\tau))\cdot n\,dS\,d\tau, \end{equation} with $n$ the outward-pointing unit normal vector field on $\partial\Omega$. \et \pr Differentiating \eqref{tofdef}, by the divergence theorem we have \begin{equation}\label{divth} \frac{d}{dt}{\mathcal T}_\Om[f](t) = \int_\Omega D_tT(x,t,\mathrm{pr}^{(m+1)} f)\,dx = - \int_{\partial\Omega} X(x,t,\mathrm{pr^{(m-1)}} f)\cdot n\,dS, \end{equation} so \eqref{tofcons} follows by integration. Conversely, differentiating \eqref{tofcons} with respect to $t$ and applying the divergence theorem we obtain $$ \int_\Omega (D_tT(x,t,\mathrm{pr}^{(m+1)} f) + \mathrm{Div}\, X(x,t,\mathrm{pr}^{(m+1)} f))\,dx = 0. $$ As this is to hold for any $\Omega$ and $f$, the claim follows. \ep \begin{thr}{\bf Example. }\rm Consider the motion of an incompressible, inviscid fluid. With $x\in \mathbb R^3$ representing the spatial coordinates, let $u=u(x,t)\in\mathbb R^3$ denote the velocity of a fluid particle at position $x$ and time $t$. Let $\rho(x,t)$ be the density and $p(x,t)$ be the pressure of the fluid. We assume that the flow is isentropic (has constant entropy), then $p$ depends only on $\rho$. The equation of continuity then takes the form \begin{equation}\label{cons1} \rho_t + \mathrm{Div}\,(\rho\cdot u) = 0, \end{equation} where $\mathrm{Div}\,(\rho u) = \sum_j \pd{(\rho u^j)}{x^j}$ is the spatial total divergence. Also, momentum balance gives the three equations \begin{equation}\label{cons2} \pd{u^i}{t} + \sum_{j=1}^3 u^j\pd{u^i}{x^j} = -\frac{1}{\rho}\pd{p}{x^i}, \quad i=1,2,3. \end{equation} Here, the equation of continuity is itself a conservation law with density $T=\rho$ and flux $X=\rho u$. Using \eqref{divth} for the functional \eqref{tofdef} corresponding to the flux we get $$ \frac{d}{dt}\int_\Omega \rho\,dx = - \int_{\partial\Omega} \rho u\cdot n dS. $$ This equation has an immediate physical interpretation: $\int_\Omega\rho\,dx$ is the total mass of the fluid within $\Omega$, and $\rho u\cdot n$ is the flux of fluid out of a point on the boundary of $\Omega$. This means that the net change of mass inside $\Omega$ equals the influx of fluid into $\Omega$ via $\partial\Omega$. In particular, if the normal component $u\cdot n$ of the velocity on $\partial \Omega$ vanishes then there is no change in the mass in $\Omega$ and we obtain the law of conservation of mass: $$ \int_\Omega \rho\,dx = \text{ const}. $$ Furthermore, combining \eqref{cons1} and \eqref{cons2} and re-arranging we get three further conservation laws: $$ D_t(\rho u^i) +\sum_{j=1}^3 D_j(\rho u^iu^j + p\delta^j_i) = 0,\quad i=1,2,3. $$ Again using \eqref{divth} we obtain the laws of conservation of momentum: $$ \frac{d}{dt}\int_\Omega \rho u^i\,dx = - \int_{\partial\Omega}(\rho u^i(u\cdot n) + p n_i)\,dS, \quad i=1,2,3. $$ Here, the first term on the right hand side is the transport of momentum $\rho u^i$ due to the flow across the surface $\partial\Omega$, and the second term is the net change in momentum due to the pressure across $\partial\Omega$. Thus $X_j=\rho u^iu^j + p\delta^j_i$ represents the components of the momentum flux. \et \section{Noether's theorem}\index{Noether's theorem} In this section we prove a fundamental result on the connection between variational symmetries and conservation laws, established by Emmy Noether in 1918. It is of central importance in mathematical physics as the source of many fundamental laws of nature. \bt\label{noetherth} Let $G$ be a local one-parameter group of variational symmetries of the variational problem ${\mathcal L}[u]=\int L(x,u^{(n)})\,dx$. Let \begin{equation}\label{vfform2} v = \sum_{i=1}^p \xi^i(x,u)\pd{}{x^i} + \sum_{\alpha=1}^q \phi_\alpha(x,u)\pd{}{u^\alpha} \end{equation} be the infinitesimal generator of $G$, and as in \eqref{chardef} let $$ Q_\alpha(x,u^{(1)}) := \phi_\alpha(x,u) - \sum_{i=1}^p \xi^i(x,u)u^\alpha_i, \quad \alpha=1,\dots,q. $$ be the characteristic of $v$. Then there exists a $p$-tuple $$ F(x,u^{(m)})=(F_1(x,u^{(m)}),\dots,F_p(x,u^{(m)})) $$ such that \begin{equation}\label{noether} \mathrm{Div}\, F = Q\cdot E(L) = \sum_{\al=1}^q Q_\alpha\cdot E_\alpha(L). \end{equation} Thus $\mathrm{Div}\, F=0$ is a conservation law for the Euler--Lagrange equations $E(L)=0$. \et \pr Since $v$ is the infinitesimal generator of a variational symmetry, by \eqref{varcrit} and \eqref{charprol} we have \begin{equation*} \begin{split} 0 &= \mathrm{pr}^{(n)} v(L) + L\,\mathrm{Div}\,\xi = \mathrm{pr}^{(n)} v_Q(L) +\sum_{i=1}^p\xi^iD_iL + L\,\mathrm{Div}\,\xi\\ &= \mathrm{pr}^{(n)} v_Q(L) +\mathrm{Div}\,(L\xi), \end{split} \end{equation*} where $L\xi:=(L\xi^1,\dots,L\xi^p)$. We now `integrate by parts' the first term, using \eqref{charparts}: \begin{equation*} \begin{split} \mathrm{pr}^{(n)} v_Q(L) &= \sum_{\alpha,J}D_JQ_\alpha \pd{L}{u^\alpha_J} = \sum_{\alpha,J} Q_\alpha\cdot (-D)_J \pd{L}{u^\alpha_J} +\mathrm{Div}\, A\\ &= \sum_{\al=1}^q Q_\alpha E_\alpha(L) + \mathrm{Div}\, A, \end{split} \end{equation*} where $A=(A_1,\dots,A_p)$ is some $p$-tuple of functions depending on $Q$, $L$ and their derivatives, but whose concrete form is not important for us. It follows that \begin{equation}\label{noether1} \mathrm{pr}^{(n)} v_Q(L) = Q\cdot E(L)+\mathrm{Div}\, A. \end{equation} Therefore, \begin{equation}\label{noether2} 0=Q\cdot E(L) + \mathrm{Div}\,(A+L\xi), \end{equation} which means that \eqref{noether} holds with $F=-(A+L\xi)$. \ep To explicitly calculate $F$ one may in principle follow the calculation of the previous proof. In applications, in most cases one encounters first order variational problems. For these, an explicit formula is given in the following result: \begin{thr} {\bf Corollary. } Let ${\mathcal L}[u]=\int L(x,u^{(1)})\,dx$ be a first order variational problem, and let $v$ as in \eqref{vfform2} be a variational symmetry. Then \begin{equation}\label{expnoether} F_i = \sum_{\al=1}^q \phi_\alpha \pd{L}{u^\alpha_i} + \xi^iL -\sum_{\al=1}^q\sum_{j=1}^p \xi^j u^\alpha_j\pd{L}{u^\alpha_i}, \quad i=1,\dots,p \end{equation} form the components of a conservation law $\mathrm{Div}\, F = 0$ for the corresponding Euler--Lagrange equations $E(L)=0$. \et \pr Since $L$ depends only on derivatives up to order one, \eqref{charparts} gives $$ \mathrm{pr}^{(1)} v_Q(L) = \sum_{\al=1}^q \Big( Q_\alpha\pd{L}{u^\alpha} + \sum_{i=1}^p D_iQ_\alpha\pd{L}{u^\alpha_i}\Big). $$ Now setting $A_i:=\sum_{\al=1}^q Q_\alpha \pd{L}{u^\alpha_i}$ we have $$ \mathrm{Div}\, A = \sum_{\alpha,i} \Big( D_iQ_\alpha\pd{L}{u^\alpha_i} + \sum_{\alpha,i} Q_\alpha D_i\pd{L}{u^\alpha_i}\Big). $$ On the other hand, \eqref{eulerop} implies $$ Q\cdot E(L) = \sum_{\al=1}^q Q_\alpha \pd{L}{u^\alpha} -\sum_{\al=1}^q Q_\alpha D_i \pd{L}{u^\alpha_i}, $$ so altogether we obtain $$ \mathrm{pr}^{(1)} v_Q(L) = Q\cdot E(L)+\mathrm{Div}\, A, $$ verifying \eqref{noether1}. \ep \begin{thr}{\bf Example. }\rm\label{partmech} Perhaps the most prominent application of the Noether theorem is to the mechanics of particles. Consider a system of $n$ particles moving in $\mathbb R^3$ in some force field given by a potential. The system is described by the position of its particles, where $\mathbf{x}^\alpha = (x^\alpha,y^\alpha,z^\alpha)$ is the position of the $\alpha$-th particle, and $\mathbf{x}=(\mathbf{x}^1,\dots,\mathbf{x}^n)$ is a vector containing the information about all positions of the particles. Assuming that the $\alpha$-th particle has mass $m_\alpha$, the total kinetic energy of the system is $$ K(\dot\mathbf{x}) = \frac{1}{2}\sum_{\alpha=1}^n m_\alpha |\dot \mathbf{x}^\alpha|^2. $$ Also, we assume that the force field is determined by a potential $U=U(\mathbf{x},t)$. Newton's equation of motion then gives $$ m_\alpha\ddot \mathbf{x}^\alpha = -\nabla_\alpha U = -(U_{x^\alpha},U_{y^\alpha},U_{z^\alpha}), \quad \alpha=1,\dots,n. $$ The important point to note here is that these equations are the Euler--Lagrange equations of the Lagrange function (or action integral) \begin{equation}\label{lagrangian} {\mathcal L}[\mathbf{x}]=\int_{-\infty}^\infty (K-U)\,dt. \end{equation} Indeed, $$ E_{x^\alpha}(L) = \pd{L}{x^\alpha} - D_t\pd{L}{\dot x^\alpha} = -U_{x^\alpha} - m_\alpha \ddot x^\alpha, $$ and analogously for $y^\alpha$ and $z^\alpha$. By \eqref{varcrit}, a vector field $$ v = \tau(t,\mathbf{x})\pd{}{t} + \sum_\alpha {\boldsymbol\xi}^\alpha(t,\mathbf{x})\cdot \pd{}{\mathbf{x}^\alpha} \equiv \tau\pd{}{t} +\sum_\alpha \Big(\xi^\alpha \pd{}{x^\alpha} + \eta^\alpha \pd{}{y^\alpha} + \zeta^\alpha \pd{}{z^\alpha} \Big) $$ generates a variational symmetry of the Lagrangian \eqref{lagrangian} if and only if \begin{equation}\label{newtoncrit} \mathrm{pr}^{(1)} v(K-U) + (K-U)D_t\tau = 0 \quad \forall (t,\mathbf{x}). \end{equation} Given a variational symmetry, by Noether's theorem \ref{noether} we obtain a corresponding conservation law (or first integral) with single component (since $p=1$) \begin{equation}\label{newcons} \begin{split} F &= \sum_{\alpha=1}^n m_\alpha {\boldsymbol\xi}^\alpha\cdot \dot\mathbf{x}^\alpha + \tau (K-U) - \sum_{\alpha=1}^n \tau \dot \mathbf{x}^\alpha\cdot (m_\alpha \dot\mathbf{x}^\alpha)\\ &= \sum_{\alpha=1}^n m_\alpha {\boldsymbol\xi}^\alpha\cdot \dot\mathbf{x}^\alpha -\tau (K+U) = \sum_{\alpha=1}^n m_\alpha {\boldsymbol\xi}^\alpha\cdot \dot\mathbf{x}^\alpha -\tau E, \end{split} \end{equation} where $E=K+U$ is the total energy. Since $D_tF=0$ it follows that $F$ has to be constant for any solution of Newton's equations of motion. We now analyze some examples of variational symmetries of \eqref{lagrangian} that lead to conservation laws of physical interest. \begin{itemize} \item $v=\partial_t$: In this case, $\mathrm{pr}^{(1)} v = v$, so \eqref{newtoncrit} holds if and only $\partial_t U = 0$, i.e., if and only if $U$ does not explicitly depend on the time $t$. Since $\tau=1$ and ${\boldsymbol \xi}=0$, the conserved quantity given in \eqref{newcons} then is the total energy $E$. We obtain that {\em invariance of a physical system under time translations implies conservation of energy}.\index{conservation!of energy} \item $v=\sum_{\alpha=1}^n \mathbf{a}\cdot\pd{}{\mathbf{x}^\alpha}$: This is the generator of the translation $\mathbf{x}\mapsto \mathbf{x}+\mathbf{a}$, so all particles are simultaneously translated in the same fixed direction $\mathbf{a}\in \mathbb R^3$. Also in this case, $\mathrm{pr}^{(1)} v = v$, so \eqref{newtoncrit} holds if and only $v(U)=0$, i.e., if and only if the potential is translationally invariant in the direction $\mathbf{a}$. As $\tau=0$ and ${\boldsymbol \xi}^\alpha =\mathbf{a}$ for all $\alpha$, the corresponding conserved quantity from \eqref{newcons} is the momentum $$ \sum_{\alpha=1}^n m_\alpha \mathbf{a}\cdot \dot\mathbf{x}^\alpha = \text{ const}. $$ In particular, if $U$ is invariant under all translations then the total momentum $\sum_{\alpha=1}^n m_\alpha \dot\mathbf{x}^\alpha$ is conserved.\index{conservation!of momentum} \item $v = \sum_{\alpha=1}^n \big(x^\alpha\pd{}{y^\alpha} - y^\alpha\pd{}{x^\alpha}\big)$: This is the generator of a simultaneous rotation of all the masses in the system about some fixed axis, in this case the $z$-axis. By \eqref{phiform} we have $$ \mathrm{pr}^{(1)} v = v +\sum_{\alpha=1}^n \Big(\dot x^\alpha\pd{}{\dot y^\alpha} - \dot y^\alpha\pd{}{\dot x^\alpha}\Big). $$ Note that $\mathrm{pr}^{(1)} v(K)=0$ (since $K$ is invariant under rotations of velocities), and since $\tau=0$, \eqref{newtoncrit} shows that $v$ generates a variational symmetry if and only if $v(U)=0$, i.e., if and only if $U$ is invariant under rotations around the $z$-axis. In this case, the conserved quantity corresponding to this symmetry via \eqref{newcons} is the angular momentum around the $z$-axis $$ \sum_{\alpha=1}^n m_\alpha (x^\alpha \dot y^\alpha - y^\alpha \dot x^\alpha) = \text{ const.} $$ Thus {\em rotational invariance implies conservation of angular momentum}.\index{conservation!of angular momentum} \end{itemize} In particular, if we assume that the particles only interact through their mutual gravitational (or electrostatic, \dots) attraction, then the potential energy is of the form $$ U(t,\mathbf{x}) = \sum_{\alpha\not=\beta} \gamma_{\alpha\beta} |\mathbf{x}^\alpha-\mathbf{x}^\beta|^{-1}. $$ In this case, all of the above assumptions are satisfied, so we obtain conservation of energy, momentum, and angular momentum. \et To conclude this chapter we introduce a straightforward generalization of Noether's theorem that also leads to conservation laws of physical interest. \begin{thr} {\bf Definition. } Let ${\mathcal L}[u]=\int L(x,u^{(n)})\,dx$ be a variational problem. A vector field $v$ on $M\subseteq X\times U$ is called an {\em infinitesimal divergence symmetry}\index{infinitesimal divergence symmetry} of ${\mathcal L}$ if there exists a $p$-tuple $B(x,u^{(m)}) = (B_1,\dots,B_p)$ of functions of $x$, $u$ and the derivatives of $u$ such that \begin{equation}\label{divsym} \mathrm{pr}^{(n)} v(L) + L\cdot \mathrm{Div}\,\xi = \mathrm{Div}\, B \end{equation} for all $(x,u)\in M$. \et The point to note is that the Noether theorem \ref{noetherth} remains valid if in the hypothesis we replace `variational symmetry' by `divergence symmetry'. Indeed, the only thing that has to be changed in the proof is that we have to incorporate the new term $\mathrm{Div}\, B$ into \eqref{noether2}, which now becomes $$ Q\cdot E(L) + \mathrm{Div}\,(A+L\xi) = \mathrm{Div}\, B, $$ but we still obtain a conservation law of the form \eqref{noether}, namely $F=B-A-L \xi$. In particular, in the case of a variational problem of first order, the conserved quantity corresponding to \eqref{expnoether} now becomes \begin{equation}\label{expnoether2} F_i = \sum_{\al=1}^q \phi_\alpha \pd{L}{u^\alpha_i} + \xi^iL -\sum_{\al=1}^q\sum_{j=1}^p \xi^j u^\alpha_j\pd{L}{u^\alpha_i} - B_i. \quad i=1,\dots,p \end{equation} \begin{thr}{\bf Example. }\rm Returning to the setup of \ref{partmech}, consider a Galilean boost $$ (t,\mathbf{x}^\alpha)\mapsto (t,\mathbf{x}^\alpha + \varepsilon t\mathbf{a}), $$ for some $\mathbf{a}\in\mathbb R^3$. By\eqref{phiform}, the infinitesimal generator $v=\sum_{\alpha=1}^n t\mathbf{a} \partial_{\mathbf{x}^\alpha}$ then has prolongation $$ \mathrm{pr}^{(1)} v = \sum_{\alpha=1}^n \Big(t\mathbf{a} \pd{}{\mathbf{x}^\alpha} + \mathbf{a} \pd{}{\dot\mathbf{x}^\alpha}\Big), $$ so $$ \mathrm{pr}^{(1)} v(L) = \mathrm{pr}^{(1)}(K-U) = \sum_{\alpha=1}^n m_\alpha \mathbf{a}\cdot \dot\mathbf{x}^\alpha - t \sum_{\alpha=1}^n \mathbf{a}\cdot \nabla_\alpha U. $$ Since $\tau = 0$, \eqref{newtoncrit} demands that this expression should vanish identically, but this is never the case for $\mathbf{a}\not=0$. However, since $$ \sum_{\alpha=1}^n m_\alpha \mathbf{a}\cdot \dot\mathbf{x}^\alpha = D_t\Big(\sum_{\alpha=1}^n m_\alpha \mathbf{a}\cdot\mathbf{x}^\alpha \Big), $$ $v$ is an infinitesimal divergence symmetry if $\mathbf{a}\cdot \nabla_\alpha U=0$, i.e., if $U$ is translationally invariant in the direction $\mathbf{a}$. The conserved quantity according to \eqref{expnoether2} now reads $$ t \sum_{\alpha=1}^n m_\alpha \mathbf{a}\cdot\dot\mathbf{x}^\alpha - \sum_{\alpha=1}^n m_\alpha \mathbf{a}\cdot\mathbf{x}^\alpha. $$ Here the first sum, when divided by the total mass $\sum_\alpha m_\alpha$ is the position of the center of mass of the system in the direction $\mathbf{a}$, and the second is the momentum in that direction. It follows that if $U$ is translationally invariant in a given direction then on the one hand, as shown in \ref{partmech}, the momentum in that direction is constant. In addition it now follows that the center of mass in that direction is a linear function of $t$: $$ \text{position of center of mass =} \ \text{initial position} + t(\text{momentum})/\text{mass}. $$ In particular, if $U$ is invariant under all translations, then the center of mass of any such system moves linearly in one fixed direction. \et \section{Trivial conservation laws and characteristics} There are certain types of conservation laws that do not yield any useful information on the system under consideration and hence are called trivial.\index{conservation law!trivial} A conservation law $F=(F_1,\dots,F_p)$ can be trivial for one of two reasons: triviality of the first kind holds if $F$ itself vanishes on every solution of the system. This kind of triviality\index{trivial conservation law!first kind} is usually easy to eliminate by solving the system itself and its prolongations for certain of the variables $u^\alpha_J$ and then substituting for these variables wherever they occur. For example, in the case of an evolution equation $u_t=P(x,u^{(n)})$, any time derivative of $u$, e.g., $u_{tt},\,u_{xt}$, etc.\ can be expressed in terms of $x$, $u$ and spatial derivatives of $u$. \begin{thr}{\bf Example. }\rm\label{heatc} Consider the system of first order evolution equations $$ u_t=v_x,\quad v_t=u_x, $$ which is equivalent to the one-dimensional wave equation $u_{tt}=u_{xx}$. Then $$ D_t\Big(\frac{1}{2}u_t^2+\frac{1}{2}u_x^2\Big) - D_x(u_tu_x) = u_t(u_{tt}-u_{xx}) = 0 $$ is a conservation law for this system. As explained above, we can replace the density and the flux of this conservation law by ones that depend only on spatial derivatives. This gives $$ D_t\Big(\frac{1}{2}u_x^2+\frac{1}{2}v_x^2\Big) - D_x(u_xv_x) = 0. $$ The two conservation laws differ by the trivial conservation law $$ D_t\Big(\frac{1}{2}u_t^2-\frac{1}{2}v_x^2\Big) + D_x(v_xu_x-u_tu_x) = 0, $$ for which both density and flux vanish on any solution of the system. In the same way, for any conservation law of an evolution equation there is, up to addition of a trivial conservation law of the first kind, one where density and flux depend only on spatial derivatives. \et A conservation law is called trivial of the second kind\index{trivial conservation law!second kind} if the total divergence $\mathrm{Div}\, F$ in fact vanishes on {\em all} smooth functions $f$, whether or not they solve the equation. An example of this kind of triviality is the relation \begin{equation}\label{triv} D_xu_y - D_yu_x=0, \end{equation} which obviously holds for any smooth function $u=f(x,y)$. Hence \eqref{triv} is a trivial conservation law of the second kind for any partial differential equation for functions of $x$ and $y$. The underlying $p$-tuples $(F_1,\dots,F_p)$ of conservation laws of the second kind are also called null divergences.\index{null divergence} There is in fact a complete characterization of null divergences that is a direct analogue of the characterization of the kernel of the usual divergence operator via the Poincar\'e lemma. Since the proof of this result, building on the so-called variational complex, is quite involved, we only state the result and refer to \cite[Sec.\ 5.4]{O} for a proof. \bt\label{tcc} Let $F=(F_1,\dots,F_p)$ be a $p$-tuple of smooth functions depending on $x=(x^1,\dots,x^p)$, $(u^1,\dots,u^q)$ and derivatives of $u$, and defined on all of $X\times U^{(n)}$. Then the following are equivalent: \begin{itemize} \item[(i)] $\mathrm{Div}\, F \equiv 0$. \item[(ii)] There exist smooth functions $G_{jk}$, $j,k=1,\dots,p$, depending on $x$, $u$, and derivatives of $u$, such that $G_{jk}=-G_{kj}$ for all $j$, $k$, and \begin{equation}\label{tcce} F_j = \sum_{k=1}^p D_k G_{jk}, \quad j=1,\dots,p \end{equation} \end{itemize} \et For example, if $p=3$, then $$ \mathrm{Div}\, F = D_1F_1+D_2F_2+D_3F_3\equiv 0 $$ if and only if $F$ is a `total curl': $F=\mathrm{Curl}(G)$, i.e., $$ F_1 = D_2G_3 - D_3G_2,\quad F_2 = D_3G_1 - D_1G_3,\quad F_3 = D_1G_2 - D_2G_1 $$ (where we have identified $G_{12}=-G_{21}$ from the theorem with $G_3$, etc.). We define a trivial conservation law to be any conservation law that is a linear combination of trivial conservation laws of the first and second kind. Thus $\mathrm{Div}\, F = 0$ is a trivial conservation law if and only if there exist functions $G_{ij}$ as in \ref{tcc} (ii) such that \eqref{tcce} holds for all solutions of the system. Two conservation laws are called {\em equivalent} if they differ only by a trivial conservation law.\index{conservation law!trivial} The interesting objects in the study of conservation laws therefore are equivalence classes in this sense. For the following considerations on characteristics of conservation laws we need some preparations. \begin{thr} {\bf Definition. } Let $P_\nu(x,u^{(n)})=0$, $\nu=1,\dots,l$, be a system of differential equations, with $P:M^{(n)}\to \mathbb R^l$. The $k$-th prolongation of this system is the $(n+k)$-th order system of differential equations $$ P^{(k)}(x,u^{(n+k)}) = 0 $$ obtained by differentiating in all possible ways $k$ times. Thus the new system consists of the $\begin{pmatrix} p+k-1\\ k \end{pmatrix}\cdot l $ equations $$ D_J P_\nu(x,u^{(n+k)}) = 0, $$ where $\nu=1,\dots,l$ and $0\le\sharp J \le k$. \et \begin{thr}{\bf Example. }\rm If $P$ is the heat equation $u_t=u_{xx}$, then the first prolongation $P^{(1)}$ is the system $$ u_t=u_{xx},\quad u_{xt} = u_{xxx}, \quad u_{tt}=u_{xxt}, $$ and the second one contains, in addition, the equations $$ u_{xxt}=u_{xxxx},\quad u_{xtt} = u_{xxxt}, \quad u_{ttt}=u_{xxtt}. $$ \et We then clearly have: \blem\label{sowieso} If $u=f(x)$ is a solution of the system $P(x,u^{(n)})=0$, then it is also a solution of every prolongation $P^{(k)}(x,u^{(n+k)})=0$, $k=0,1,2,\dots$. \et \begin{thr} {\bf Definition. }\label{conspr} A system of differential equations is called {\em totally nondegenerate}\index{nondegenerate!totally} if it and all its prolongations are nondegenerate (i.e., of maximal rank and locally solvable). \et \blem\label{consprlem} Let $P_\nu(x,u^{(n)})=0$, $\nu=1,\dots,l$, be a totally nondegenerate system of differential equations. Let $Q=Q(x,u^{(m)})$ be a smooth function. Then the following are equivalent: \begin{itemize} \item[(i)] $Q$ vanishes for all solutions $u=f(x)$ of the system. \item[(ii)] There exist differential operators ${\mathcal D}_\nu = \sum_J Q^J_\nu(x,u^{(m)})D_J$, $\nu=1,\dots,l$, such that $$ Q =\sum_{\nu=1}^l {\mathcal D}_\nu P_\nu $$ for all $(x,u^{(m)})$. \end{itemize} \et \pr (ii)$\Rightarrow$(i) is obvious. (i)$\Rightarrow$(ii): Clearly we can assume that $m\ge n$. By \ref{sowieso}, (i) is equivalent to $Q$ vanishing on all solutions of the prolonged system to $P^{(m-n)}$, which is locally solvable by assumption. Thus it follows that $Q$ vanishes on the zero set of $P^{(m-n)}$. Since, moreover, $P^{(m-n)}$ is nondegenerate, by \ref{hadamard} there exist smooth functions $Q^J_\nu$ ($\sharp J\le m-n$, $1\le \nu\le l$) such that $$ Q(x,u^{(m)}) = \sum_J \sum_\nu Q^J_\nu(x,u^{(m)}) D_JP_\nu(x,u^{(m)}) $$ \ep Let now $\mathrm{Div}\, F=0$ be a conservation law of a totally nondegenerate system $$ P(x,u^{(n)})=0 $$ of differential equations. Then by \ref{consprlem}, $\mathrm{Div}\, F$ vanishes on every solution of the system if and only if there exist smooth functions $Q^J_\nu(x,u^{(m)})$ such that \begin{equation}\label{divfg} \mathrm{Div}\, F = \sum_{\nu,J} Q^J_\nu D_JP_\nu. \end{equation} We now note that each term in \eqref{divfg} can be `integrated by parts': for example, if $1\le j\le p$ then $$ Q^j_\nu D_j P_\nu = D_j(Q^j_\nu P_\nu) - D_j(Q^j_\nu)P_\nu. $$ Proceeding in this way, we can cast \eqref{divfg} in the form $$ \mathrm{Div}\, F = \mathrm{Div}\, G + \sum_{\nu=1}^l Q_\nu P_\nu \equiv \mathrm{Div}\, G + Q\cdot P, $$ where $Q=(Q_1,\dots,Q_l)$ has entries \begin{equation}\label{qe} Q_\nu = \sum_J (-D)_J Q^J_\nu, \end{equation} and all we need to know about $G=(G_1,\dots,G_p)$ is that it depends linearly on the components $P_\nu$ of $P$ and their total derivatives. This means that $G$ defines a trivial conservation law of the first kind. Consequently, replacing $F$ by $F-G$ we obtain an equivalent conservation law of the form \begin{equation}\label{cc} \mathrm{Div}\, F = Q\cdot P. \end{equation} This is called the {\em characteristic form}\index{conservation law!characteristic form} of the conservation law \eqref{divfg}, and $Q=(Q_1,\dots,Q_l)$ is called the characteristic of the conservation law.\index{conservation law!characteristic of} Unless $l=1$, the $Q_\nu$ from \eqref{cc} are in general not uniquely determined. If $Q$ and $\tilde Q$ are two $l$-tuples both satisfying \eqref{cc} for the same $F$, then $(Q-\tilde Q)\cdot P = 0$. Since $P$ is nondegenerate, by \ref{hadex} it follows from this that $Q-\tilde Q$ vanishes on all solutions. Based on this observation, we call a characteristic {\em trivial}\index{characteristic!trivial} if it vanishes on all solutions of the system, and we call two characteristics equivalent if they differ by a trivial one. In this terminology, characteristics are in general only determined up to equivalence. \begin{thr}{\bf Example. }\rm To obtain the characteristic form of the conservation law for the heat equation given in \ref{heatc}, we re-write it in the form \eqref{divfg}: $$ D_t(\frac{1}{2}u_t^2+\frac{1}{2}u_x^2) - D_x(u_tu_x) = u_tD_t(u_t-v_x) + u_tD_x(v_t-u_x). $$ Then \eqref{qe} shows that the characteristic is given by $$ Q = (-D_t(u_t),-D_x(u_t)) = (-u_{tt},-u_{xt}), $$ and an equivalent conservation law in characteristic form can be found through integration by parts: $$ D_t\Big(\frac{1}{2}u_t^2+\frac{1}{2}u_x^2\Big) + D_x(-u_tu_x) \cong -u_{tt}(u_t-v_x) - u_{xt}(v_t-u_x). $$ \et One can show that for systems of differential equations that are totally nondegenerate and normal (i.e., possessing a noncharacteristic direction at every point) the two notions of equivalence we have introduced above actually coincide: \bt Let $P(x,u^{(n)})=0$ be a normal and totally nondegenerate system of differential equations. Let the $p$-tuples $F$ and $\tilde F$ determine conservation laws with characteristics $Q$ and $\tilde Q$, respectively. Then $F$ and $\tilde F$ are equivalent as conservation laws if and only if $Q$ and $\tilde Q$ are equivalent as characteristics. \et For a proof we refer to \cite[Sec.\ 4.3]{O}. Finally, we may apply the terminology developed in this chapter to re-formulate Noether's theorem \ref{noetherth} more precisely: \bt\label{noetherth2} Let $G$ be a local one-parameter group of variational symmetries of the variational problem ${\mathcal L}[u]=\int L(x,u^{(n)})\,dx$. Let \begin{equation*}\label{vfform3} v = \sum_{i=1}^p \xi^i(x,u)\pd{}{x^i} + \sum_{\alpha=1}^q \phi_\alpha(x,u)\pd{}{u^\alpha} \end{equation*} be the infinitesimal generator of $G$, and as in \eqref{chardef} let $$ Q_\alpha(x,u^{(1)}) := \phi_\alpha(x,u) - \sum_{i=1}^p \xi^i(x,u)u^\alpha_i, \quad \alpha=1,\dots,q. $$ be the characteristic of $v$. Then $Q=(Q_1,\dots,Q_q)$ is also the characteristic of a conservation law for the corresponding Euler--Lagrange equations $E(L)=0$, i.e., there exists a $p$-tuple $$ F(x,u^{(m)})=(F_1(x,u^{(m)}),\dots,F_p(x,u^{(m)})) $$ such that \begin{equation*}\label{noether3} \mathrm{Div}\, F = Q\cdot E(L) = \sum_{\al=1}^q Q_\alpha\cdot E_\alpha(L) \end{equation*} is a conservation law in characteristic form for the Euler--Lagrange equations $$ E(L)=0. $$ \et \addcontentsline{toc}{part}{Bibliography}
{ "timestamp": "2015-06-24T02:13:53", "yymm": "1506", "arxiv_id": "1506.07131", "language": "en", "url": "https://arxiv.org/abs/1506.07131", "abstract": "These are lecture notes of a course on symmetry group analysis of differential equations, based mainly on P. J. Olver's book 'Applications of Lie Groups to Differential Equations'. The course starts out with an introduction to the theory of local transformation groups, based on Sussman's theory on the integrability of distributions of non-constant rank. The exposition is self-contained, pre-supposing only basic knowledge in differential geometry and Lie groups.", "subjects": "Differential Geometry (math.DG)", "title": "Lie Transformation Groups - An Introduction to Symmetry Group Analysis of Differential Equations", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9766692271169855, "lm_q2_score": 0.8376199653600371, "lm_q1q2_score": 0.8180776441859436 }
https://arxiv.org/abs/1805.08375
Counting partitions inside a rectangle
We consider the number of partitions of $n$ whose Young diagrams fit inside an $m \times \ell$ rectangle; equivalently, we study the coefficients of the $q$-binomial coefficient $\binom{m+\ell}{m}_q$. We obtain sharp asymptotics throughout the regime $\ell = \Theta (m)$ and $n = \Theta (m^2)$. Previously, sharp asymptotics were derived by Takács only in the regime where $|n - \ell m /2| = O(\sqrt{\ell m (\ell + m)})$ using a local central limit theorem. Our approach is to solve a related large deviation problem: we describe the tilted measure that produces configurations whose bounding rectangle has the given aspect ratio and is filled to the given proportion. Our results are sufficiently sharp to yield the first asymptotic estimates on the consecutive differences of these numbers when $n$ is increased by one and $m, \ell$ remain the same, hence significantly refining Sylvester's unimodality theorem.
\section{Introduction} A partition $\lambda$ of $n$ is a sequence of weakly decreasing nonnegative integers $\lambda=(\lambda_1\geq \lambda_2\geq \ldots)$ whose sum $|\lambda|=\lambda_1+\lambda_2+\cdots$ is equal to $n$. The study of integer partitions is a classic subject with applications ranging from number theory to representation theory and combinatorics, and integer partitions with various restrictions on properties, such as part sizes or number of parts, occupy the field of partition theory~\cite{Andrews1976}. The generating functions of integer partitions play a role in number theory and the theory of modular forms. In representation theory, integer partitions index the conjugacy classes and irreducible representations of the symmetric group $S_n$; they are also the signatures of the irreducible polynomial representation of $GL_n$ and give a basis for the ring of symmetric functions. More recently, partitions have appeared in the study of interacting particle systems and other statistical mechanics models. The number of partitions of $n$, typically denoted by $p(n)$ but here unconventionally\footnote{We use the notation $N_n$ to distinguish scenarios of probability with those of enumeration, both of which occur in the present manuscript.} by $N_n$, was implicitly determined by Euler via the generating function \[ \sum_{n=0}^{\infty} N_n q^n = \prod_{i=1}^\infty \frac{1}{1-q^i} \, . \] There is no exact explicit formula for the numbers $N_n$. The asymptotic formula \begin{align} \label{eq:hardy-ram} N_n:=\#\{ \lambda \vdash n \} \sim \frac{1}{4n \sqrt{3}} \exp\left( \pi \sqrt{\frac{2n}{3}} \right)\,, \end{align} obtained by Hardy and Ramanujan~\cite{HardyRamanujan1918}, is considered to be the beginning of the use of complex variable methods for asymptotic enumeration of partitions (the so-called circle method). Our goal is to obtain asymptotic formulas similar to~\eqref{eq:hardy-ram} for the number of partitions $\lambda$ of $n$ whose Young diagram fits inside an $m \times \ell$ rectangle, denoted \[ N_n(\ell,m):=\# \{ \lambda \vdash n: \lambda_1 \, \leq \, \ell, \quad \text{length}(\lambda) \, \leq \, m\}\, .\] These numbers are also the coefficients in the expansion of the $q$-binomial coefficient \[ \binom{\ell +m}{m}_q = \frac{ \prod_{i=1}^{\ell+m} (1-q^i)} {\prod_{i=1}^\ell (1-q^i) \prod_{i=1}^m (1-q^i)} = \sum_{n=0}^{\ell m} N_n(\ell,m)q^n \, . \] The $q$-binomial coefficients are themselves central to enumerative and algebraic combinatorics. They are the generating functions for lattice paths restricted to rectangles and taking only north and east steps under the area statistic, given by the parameter $n$. They are also the number of $\ell$-dimensional subspaces of $\mathbb{F}_q^{\ell+m}$ and appear in many other generating functions as the $q$-analogue generalization of the ubiquitous binomial coefficients. Notably, the numbers $N_n(\ell,m)$ form a symmetric unimodal sequence \[ 1 = N_0(\ell,m)\leq N_1(\ell,m) \leq \cdots \leq N_{\lfloor m\ell/2\rfloor}(\ell,m) \geq \cdots \geq N_{m\ell}(\ell,m)=1, \] a fact conjectured by Cayley in 1856 and proven by Sylvester in 1878 via the representation theory of $sl_2$~\cite{Sylvester1878}. One hundred forty years later, no previous asymptotic methods have been able to prove this unimodality. \subsection*{Asymptotics of $N_n(\ell,m)$} Our first result is an asymptotic formula for $N_n (\ell,m)$ in the regime $\ell/m \to A$ and $n/m^2 \to B$ for any fixed $A > B > 0$. This is the regime in which a limit shape of the partitions exists: $\ell/m \to A$ implies the aspect ratio has a limit and $n/m^2 \to B \in (0,A)$ implies the portion of the $m \times \ell$ rectangle that is filled approaches a value that is neither zero nor one. By ``asymptotic formula'' we mean a formula giving $N_n(\ell,m)$ up to a factor of $1 + o(1)$; such asymptotic equivalence is denoted with the symbol $\sim$. By the obvious symmetry $N_{n}(\ell,m) = N_{m\ell -n}(\ell,m)$ it suffices to consider only the case $A \geq 2B > 0$. To state our results, given $A \geq 2B > 0$ we define three quantities $c, d$ and $\Delta$. The quantities $c$ and $d$ are the unique positive real solutions (see Lemma~\ref{lem:uniquecd}) to the simultaneous equations \begin{align} A &= \int_0^1 \frac{1}{1-e^{-c-dt}}dt - 1 = \frac{1}{d} \log \left( \frac{e^{c+d}-1}{e^c-1}\right) - 1 \, , \label{eq:integrals A} \\ B &= \int_0^1 \frac{t}{1-e^{-c-dt}}dt - \frac{1}{2} = \frac{d\log(1-e^{-c-d})+{\rm dilog}\,(1-e^{-c})-{\rm dilog}\,(1-e^{-c-d})}{d^2} \label{eq:integrals B} \, , \end{align} where we recall the dilogarithm function \[ {\rm dilog}\,(x) = \int_1^x \frac{\log t}{1-t}dt = \sum_{k=1}^{\infty} \frac{(1-x)^k}{k^2} \] for $|x-1|<1$. The quantity $\Delta$, which will be seen to be strictly positive, is defined by \begin{equation} \label{eq:Delta} \Delta = \frac{2Be^c(e^d-1) + 2A(e^c-1)-1}{d^2(e^{d+c}-1)(e^c-1)} - \frac{A^2}{d^2}\, . \end{equation} \begin{thm} \label{th:main} Given $m, \ell$ and $n$, let $A := \ell/m$ and $B := n/m^2$ and define $c, d$ and $\Delta$ as above. Let $K$ be any compact subset of $\{ (x,y) : x \geq 2y > 0 \}$. As $m \to \infty$ with $\ell$ and $n$ varying so that $(A,B)$ remains in $K$, \begin{equation} \label{eq:N} N_n(\ell,m) \sim \frac{e^{m \left[cA + 2dB - \log(1-e^{-c-d}) \right ]}} {2\pi m^2 \sqrt{\Delta \, (1-e^{-c})\, (1-e^{-c-d})}} \, , \end{equation} where $c$ and $d$ vary in a Lipschitz manner with $(A,B) \in K$. \end{thm} \begin{unremark} In the special case $B = A/2$, the parameters take on the elementary values $$ d = 0 \, , \qquad c = \log \left(\frac{A+1}{A}\right) \, , \qquad \mbox{ and } \qquad \Delta = \frac{A^2 (A+1)^2}{12} \, . $$ In this case we understand the exponent and leading constant to be their limits as $d \to 0$, giving $$ N_{Am^2/2} (Am,m) \sim \frac{\sqrt{3}}{A\pi m^2} \left [ \frac{(A+1)^{A+1}}{A^A} \right ]^m \, . $$ The special case when $A \to \infty$, so that $N_n(\ell,m) = N_n(m)$ and the restriction on partition sizes is removed corresponds to $c=0$ and $d$ is a solution to an explicit equation given in Lemma~\ref{lem:uniquecd}. In this case the result matches the one obtained first by Szekeres~\cite{Szekeres1953} using complex analysis, then by Canfield~\cite{Canfield1997} using a recursion, and most recently by Romik~\cite{Romik} using probabilistic methods based on Fristedt's ensemble~\cite{Fristedt1993}. These works and others are further explained in Section~\ref{sec:history}. \end{unremark} \subsection*{Unimodality} Our second result gives an asymptotic estimate of the consecutive differences of $N_n$. In fact our motivation for deriving more accurate asymptotics for $N_n (\ell, m)$ was to be able to analyze the sequence $\{ N_{n+1} (\ell,m) - N_n(\ell , m) : n \geq 1 \}$. Sylvester's proof of unimodality of $N_n (\ell,m)$ in $n$~\cite{Sylvester1878}, and most subsequent proofs~\cite{Stanley1983,Stanley1985,Proctor1982}, are algebraic, viewing $N_n(\ell,m)$ as dimensions of certain vector spaces, or their differences as multiplicities of representations. While there are also purely combinatorial proofs of unimodality, notably O'Hara's~\cite{OHara1990} and the more abstract one in~\cite{PouzetRosenberg}, they do not give the desired symmetric chain decomposition of the subposet of the partition lattice. These methods do not give ways of estimating the asymptotic size of the coefficients or their difference. It is now known that $N_n (\ell , m)$ is strictly unimodal~\cite{PakPanova2013}, and the following lower bound on the consecutive difference was obtained in~\cite[Theorem 1.2]{PakPanova2017} using a connection between integer partitions and Kronecker coefficients: \begin{equation} \label{eq:pak-panova} N_n(\ell,m)-N_{n-1}(\ell,m) \geq 0.004 \frac{2^{\sqrt{s}}}{s^{9/4}}, \end{equation} where $n\leq\ell m/2$ and $s=\min\{2n,\ell^2,m^2\}$. In particular, when $\ell=m$ we have $s=2n$. Any sharp asymptotics of the difference appears to be out of reach of the algebraic methods in this previous series of papers. Refining Theorem~\ref{th:main} we are able to obtain the following estimate. \begin{thm} \label{th:difference} Given $m, \ell$ and $n$, let $A := \ell/m$ and $B := n/m^2$ and define $d$ as above. Suppose $m, \ell$ and $n$ go to infinity so that $(A,B)$ remains in a compact subset $K$ of $\{ (x,y) : x \geq 2y > 0 \}$ and $$ m^{-1} \, |n-lm/2| \rightarrow \infty . $$ Then $$ N_{n+1}(\ell,m) - N_n(\ell,m) \sim \frac{d}{m}N_n(\ell,m). $$ \end{thm} \begin{unremark} The condition $m^{-1} \, |n-lm/2| \rightarrow \infty$ is equivalent to $m \, \left|A-B/2\right| \rightarrow \infty$ and is satisfied if and only if $d$, which depends on $m$, is not $O(m^{-1})$. It is automatically satisfied whenever the compact set $K$ is a subset of $\{ (x,y) : x > 2y > 0 \}$. \end{unremark} \subsection*{Corollary: Asymptotics of Kronecker coefficients} Recent developments in the representation theory of the symmetric and general linear groups, motivated by applications in Computational Complexity theory, have realized the consecutive differences $N_{n+1}(\ell,m)-N_n(\ell,m)$ as specific Kronecker coefficients for the tensor product of irreducible $S_{\ell m}$ representations (see, for instance,~\cite{PakPanova2013} which is also one of the unimodality proofs). The Kronecker coefficient $g(\lambda,\mu,\nu) = \dim {\rm Hom} (\mathbb{S}_\lambda, \mathbb{S}_\mu \otimes \mathbb{S}_\nu)$ is the multiplicity of the irreducible $S_{|\lambda|}$ Specht module $\mathbb{S}_\lambda$ in the tensor product of two other irreducible representations. It is a notoriously hard problem to determine the values of these coefficients, and their combinatorial interpretation has been an outstanding open problem in Algebraic Combinatorics (see Stanley~\cite{StanleyOP}) since their definition by Murnaghan in 1938. In general, determining even whether they are nonzero is an NP-hard problem and it is not known whether computing them lies in NP. See~\cite{IP17} and the literature therein for some recent developments on the relevance of Kronecker coefficients in distinguishing complexity classes on the way towards ${\rm P} \neq {\rm NP}$. Being able to estimate particular values of Kronecker coefficients is crucial to the Geometric Complexity Theory approach towards these problems. Because it is known (see~\cite{PakPanova2013}) that the consecutive difference $N_n(m,\ell) -N_{n-1}(m,\ell)$ equals the Kronecker coefficient $g((m\ell -n,n), (m^\ell), (m^\ell)),$ Theorem~\ref{th:difference} gives the first tight asymptotic estimate on this family of Kronecker coefficients. \begin{cor} The Kronecker coefficient of $S_{m\ell}$ for the (rectangle, rectangle, two-row) case is asymptotically given by $$g( (m^\ell), (m^\ell), (m\ell-n-1,n+1)) = N_{n+1}(\ell,m) - N_{n}(\ell,m) \sim \frac{d}{m} N_n(\ell,m) \sim \frac{de^{m \left[cA + 2dB - \log(1-e^{-c-d}) \right ]}} {2\pi m^3 \sqrt{\Delta \, (1-e^{-c})\, (1-e^{-c-d})}},$$ with constants and ranges as in Theorems~\ref{th:main} and~\ref{th:difference}. \end{cor} \section{Review of previous results and description of methods} \label{sec:history} \subsection{Combinatorial Enumeration} Work on this problem has developed in two streams. First, there have been combinatorial results aimed at asymptotic enumeration in various regimes. After Hardy and Ramanujan obtained an asymptotic formula for $N_n$ in~\cite{HardyRamanujan1918}, enumerative work focused on $N_n (m)$, the number of partitions with part sizes bounded by $m$, or equivalently, partitions of $n$ that fit in an $m \times \infty$ strip. In~1941, Erd{\"o}s and Lehner~\cite{ErdosLehner1941} showed that $N_n(m) \sim \frac{n^{m-1}}{m!(m-1)!}$ for $m = o(n^{1/3})$. This was generalized by Szekeres and others, ultimately leading to asymptotics of $N_n(m)$ for all $m$ in 1953~\cite{Szekeres1953}. Szekeres simplified his arguments a number of times, ultimately giving asymptotics using only a saddle-point analysis, without needing results on modular functions; his argument has been referred to as the Szekeres circle method. Canfield~\cite{Canfield1997} gave a completely elementary proof (no complex analysis) of asymptotics for $N_n(m)$ using a recursive formula satisfied by these numbers. The combinatorial stream contains a few results on the asymptotics of $N_n (m , \ell)$ but only in the regime where $m$ and $\ell$ are greater than $\sqrt{n}$ by at least a factor of $\log n$. This is a natural regime to study because the typical values of the maximum part (equivalently the number of parts) of a partition of size $n$ was shown by Erd{\"o}s and Lehner~\cite{ErdosLehner1941} to be of order $\sqrt{n \log n}$. Szekeres~\cite[Theorem 1]{Szekeres1990} used saddle-point techniques to express $N_n(\ell,m)$ in terms of $N_n$, $\lambda := \frac{\pi \ell}{\sqrt{6n}}$ and $\mu := \frac{\pi m}{\sqrt{6n}}$. If, in fact, $$\frac{\sqrt{6n}}{\pi}\left(\frac{1}{4} + \varepsilon \right) \log n < \ell,m < \frac{\sqrt{6n} \, \log n}{\pi} $$ for some $\varepsilon > 0$, then the distributions defined by $\ell$ and $m$ are independent and equal, and Szekeres' formula simplifies to $$ N_n(\ell,m) \sim N_n \, \exp\left[-(\lambda + \mu) - \sqrt{\frac{6n}{\pi}}\left(e^{-\lambda}+e^{-\mu}\right)\right]. $$ The Szekeres circle method was recently revisited by Richmond~\cite{Richmond2018}. In~\cite{JiangWang} the authors, independently and concurrently with our paper, used the generating function for $q$-binomial coefficients and a saddle point analysis to derive the asymptotics for $N_n (m , \ell)$ in the cases when $m,\ell \geq 4\sqrt{n}$, corresponding to $B \leq \min\{1,A^2\}/16$ in our notation. Those authors express their result using the root of a hypergeometric identity similar to~\eqref{eq:integrals B}, however their methods give weaker error bounds and consequently cannot answer questions of unimodality. \subsection{Probabilistic limit theorems} The second strand of work on this problem has been probabilistic. The goal in this strand has been to determine properties of a random partition or Young diagram, picked from a suitable probability measure. This approach goes back at least to Mann and Whitney~\cite{MannWhitney1947}, who showed that the size of a uniform random partition contained in an $\ell \times m$ rectangle satisfies a normal distribution. Frisdtedt~\cite{Fristedt1993} defined a distribution on partitions of all sizes, weighted with respect to a parameter $q < 1$. The key property of the measure employed is that it makes the number $X_k(\lambda)$ of parts of size $k$ in the partition $\lambda$ drawn under this distribution independent as $k$ varies; the distributions of the $X_k$ are reduced geometrics with respective parameter $1 - q^k$, so that their mean is $q^k / (1 - q^k)$. Fristedt is chiefly concerned with the limiting behavior of $k X_k$ for $k = o(\sqrt{n})$, which rescales, on division by $\sqrt{n}$, to an exponential distribution. Much of the work following Fristedt's is concerned with a description of the limiting shape of the random partition, and fluctuations around that shape. The limit shape of an unrestricted partition was posed as a problem by Vershik and first answered in~\cite{SzalayTuran1,SzalayTuran2}. In 2001, Vershik and Yakubovich~\cite{VershikYakubovich2001} describe the limit shape for singly restricted partitions: those with $m \leq c \sqrt{n}$. They obtain both main (strong law) results and fluctuation (CLT) results. It is in this paper that the probability measures $\mathbb{P}_m$ used in our analysis below first arose, although we were unaware of this when we first derived them from large deviation principles. The limit shape for doubly restricted partitions in the regime $m, \ell = \Theta (\sqrt{n})$ was first described by Petrov~\cite{Petrov2009}. It is identified there with a portion of the curve $e^{-x} + e^{-y} = 1$, which represents the limit shape of unrestricted partitions. More recently, Beltoft et al.~\cite{BeltoftBoutillierEnriquez2012} obtained fluctuation results in the doubly restricted regime. The limiting fluctuation process is an Ornstein-Uhlenbeck bridge, generalizing the two-sided stationary Ornstein-Uhlenbeck process that gives the limiting fluctuations in the unrestricted case~\cite{VershikYakubovich2001}. \subsection{Enumeration via probability} Strangely, we know of only one paper combining these two streams. Tak{\'a}cs~\cite{Takacs1986} observed the following consequence of the work of Fristedt and others. Begin a discrete walk at $(\ell,0)$ and randomly choose steps in the $(0,-1)$ or $(-1,0)$ directions by making independent fair coin flips. If this walk goes from $(\ell,0)$ to $(0,-m)$ it takes precisely $m + \ell$ steps and encloses a Young diagram fitting in an $m \times \ell$ rectangle: see Figure~\ref{fig:RW}. Let $G(m,\ell)$ denote the event that a walk of length $m+\ell$ ends at $(0,-m)$ and let $H(m,n)$ denote the event that the resulting Young diagram has area $n$. Under the IID fair coin flip probability measure on paths, all paths of length $m + \ell$ have the same probability $2^{-(m+\ell)}$. Therefore, $\mathbb{P} [G(m,\ell) \cap H(m,n)] = 2^{-(m+\ell)} N_n(\ell , m)$ and the problem of counting $N_n(\ell,m)$ is reduced to determining the probability $\mathbb{P} [G(m,\ell) \cap H(m,n)]$. \begin{figure}[!ht] \centering \includegraphics[height=1.6in]{RW} \caption{The red arrows are the steps in a South and West directed simple random walk} \label{fig:RW} \end{figure} Tak{\'a}cs observed that this probability is computable by a two-dimensional local central limit theorem, ultimately obtaining bounds on the relative error that are of order $(m + \ell)^{-3}$. These error bounds are meaningful when $n$ differs from $m \ell / 2$ by up to a few multiples of $\log (m+\ell)$ standard deviations: if $\ell = \theta (m)$ this means that $|B - A/2|m^2 = \Theta (m^{3/2} \log m)$. When $|B - A/2| \gg m^{-1/2} \log m$ the error is much bigger than the main term of the Gaussian estimate provided by the LCLT and one cannot recover meaningful information about $N_n(\ell , m)$. This is where Tak{\'a}cs left off and the present manuscript picks up. \subsection{Description of our methods} We use a local large deviation computation in place of a local central limit theorem: this is possible because the restriction to an $m \times \ell$ rectangle is a linear constraint. Indeed, consider now a partition $\lambda=(\lambda_1,\dots,\lambda_m)$ with at most $m$ parts (so some $\lambda_j$ may be zero) and define $\lambda_0 := \ell$ and $\lambda_{m+1} := 0$. It is convenient to encode a partition with respect to its \emph{gaps} $x_j := \lambda_j - \lambda_{j+1}$, so the condition that $\lambda$ be a partition of $n$ of size at most $\ell$ is equivalent to $x_j \geq 0$ and \begin{equation} \label{eq:diagram} \sum_{j=0}^m x_j = \ell \,, \qquad\qquad \sum_{j=0}^m j x_j = n \, . \end{equation} Figure~\ref{fig:1} gives a pictorial proof. \begin{figure}[!ht] \centering \includegraphics[width=4in]{diagram1.pdf} \caption{The total area $n$ of a partition is composed of rectangles of area $j x_j$} \label{fig:1} \end{figure} Solving the large deviation problem produces a ``tilted measure'' in which the gaps $X_j$ are no longer IID reduced geometrics with parameter $1/2$ but are instead given by independent reduced geometric variables whose parameters $q_j = 1 - p_j$ vary in a log-linear manner. Log-linearity is dictated by the variational large deviation problem and leads to the same simplification as before. Not all partitions have the same probability under the tilted measure, but all those resulting in a given value of $\ell$ and $n$ do have the same probability. Lastly, one must choose the particular linear function $\log q_j = -c - d (j/m)$ to ensure that $\lambda$ being a partition of $n$ with parts of size at most $\ell$ will again be in the central part of the tilted measure, so that asymptotics can be read off from a local CLT for the tilted measure. The tilted measures $\mathbb{P}_m$ that we employ are denoted $\mu_{x,y}$ in~\cite{VershikYakubovich2001} and referred to there as the grand ensemble of partitions. That paper, however, was not concerned with enumeration, only with limit shape results. For this reason they do not state or prove enumeration results. In fact~\cite{Petrov2009} is able to prove the shape result by estimating exponential rates only, showing rather elegantly that an $\varepsilon$ error in the rescaled shape leads to an exponential decrease in the number of partitions. The present manuscript combines the idea of the grand ensemble with some precise central limit estimates and some algebra inverting the relation between the log-linear parameters and the parameters $A$ and $B$ defining the respective limits of $\ell/m$ and $n/m^2$ to give estimates on $N_n(\ell , m)$ precise enough also to yield asymptotic estimates on $N_{n+1} (\ell , m) - N_n (\ell , m)$. The first step of carrying this out necessarily recovers the leading exponential behavior for $N_n (\ell , m)$, which is implicit in~\cite{VershikYakubovich2001} and~\cite{Petrov2009} though Petrov only states it as an upper bound. Interestingly, Tak{\'a}cs did not seem to be aware of the ease with which the exponential rate may be obtained. His result states a Gaussian estimate and an error term. As noted above, it is nontrivial only when the $(m+\ell)^{-3}$ relative error term does not swamp the main terms, which occurs when $n$ is close to $\ell m/2$ (see also~\cite{AlmkvistAndrews1991}). Figure~\ref{fig:expgrow} shows Tak{\'a}cs' predicted exponential growth rate on a family of examples compared to the actual exponential growth rate that follows from Theorem~\ref{th:main}. \begin{figure}[!ht] \centering \includegraphics[width=0.4\linewidth]{ExpGrowth.pdf} \caption{Exponential growth of $N_{Bm^2}(m,m)$ predicted by Tak{\'a}cs' formula (blue, above) compared to the actual exponential growth given by Theorem~\ref{th:main} (red, below).} \label{fig:expgrow} \end{figure} \section{A discretized analogue to Theorem~\protect{\ref{th:main}}} \label{sec:discrete} We now implement this program to derive asymptotics. With $c_m$ and $d_m$ to be specified later, let $q_j := e^{-c_m - j d_m/m}$, let $p_j := 1-q_j$ and let \[ L_m := \sum_{j=0}^m \log p_j.\] Let $\mathbb{P}_m$ be a probability law making the random variables $\{ X_j : 0 \leq j \leq m \}$ independent reduced geometrics with respective parameters $p_j$. Define random variables $S_m$ and $T_m$ by \begin{equation} \label{eq:S and T} S_m := \sum_{i=0}^m X_i \, ; \qquad T_m:=\sum_{i=1}^m i X_i \, , \end{equation} corresponding to the unique partition $\lambda$ satisfying $X_j = \lambda_j - \lambda_{j+1}$. We first prove a result similar to Theorem~\ref{th:main}, except that the parameters $c$ and $d$ that solve integral Equations~\eqref{eq:integrals A} and~\eqref{eq:integrals B} are replaced by $c_m$ and $d_m$ satisfying the discrete summation Equations~\eqref{eq:mu_m} and~\eqref{eq:nu_m} below. These equations say that $\mathbb{E} S_m = \ell$ and $\mathbb{E} T_m = m$. Writing this out, using $\displaystyle \mathbb{E} X_j = 1/p_j - 1 = 1 / \left ( 1 - e^{-c_m - d_m j/m} \right ) - 1$, gives \begin{align} \ell & = \sum_{j=0}^m \frac{1}{1 - e^{-c_m - d_m j/m}} - (m+1) \label{eq:mu_m} \\ n & = m \sum_{j=0}^m \frac{j/m}{1 - e^{-c_m - d_m j/m}} - \frac{m(m+1)}{2} \label{eq:nu_m} \, . \end{align} Let $M_m$ denote the covariance matrix for $(S_m , T_m)$. The entries may be computed from the basic identity ${\rm Var} \, (X_j) = q_j / p_j^2$, resulting in \begin{align} {\rm Var} \, (S_m) & = \sum_{j=0}^m \frac{e^{-c_m - d_m j/m}} {\left ( 1 - e^{-c_m - d_m j/m} \right )^2} \label{eq:alpha_m} \\ {\rm Cov} \, (S_m , T_m) & = \sum_{j=0}^m j \frac{e^{-c_m - d_m j/m}} {\left ( 1 - e^{-c_m - d_m j/m} \right )^2} \label{eq:beta_m} \\ {\rm Var} \, (T_m) & = \sum_{j=0}^m j^2 \frac{e^{-c_m - d_m j/m}} {\left ( 1 - e^{-c_m - d_m j/m} \right )^2} \label{eq:gamma_m} \, . \end{align} \begin{thm}[discretized analogue] \label{th:1'} Let $c_m$ and $d_m$ satisfy~\eqref{eq:mu_m}~--~\eqref{eq:nu_m}. Define $\alpha_m, \beta_m$ and $\gamma_n$ to be the normalized entries of the covariance matrix \[ \alpha_m := m^{-1} {\rm Var} \,(S_m) \; ; \qquad \beta_m := m^{-2} {\rm Cov} \,(S_m , T_m) \; ; \qquad \gamma_m := m^{-3} {\rm Var} \,(T_m) \; , \] which are $O(1)$ as $m \rightarrow\infty$. Again, let $A := \ell / m$ and $B := n/m^2$ and $\Delta_m := \alpha_m \gamma_m - \beta_m^2$. Then as $m \to \infty$ with $\ell$ and $n$ varying so that $(A,B)$ remains in a compact subset of $\{ (x,y) : x \geq 2y > 0 \}$, \begin{equation} \label{eq:N'} N_n(\ell,m) \sim \frac{1}{2\pi m^2 \sqrt{\Delta_m}} \exp \left \{ m \left( -\frac{L_m}{m} + c_m A + d_m B \right ) \right \} \, . \end{equation} \end{thm} \begin{proof} The atomic probabilities $\mathbb{P}_m ({\bf X} = {\bf x})$ depend only on $S_m$ and $T_m$ as \begin{align*} \log \mathbb{P}_m ({\bf X} = {\bf x}) & = \sum_{j=0}^m \left(\log p_j + x_j \log q_j\right) \\ & = L_m - \sum_{j=0}^m \left ( c_m + j \frac{d_m}{m} \right ) x_j \\ & = L_m - c_m \left ( \sum_{j=0}^m x_j \right ) - \frac{d_m}{m} \left ( \sum_{j=0}^m j x_j \right ). \end{align*} In particular, for any ${\bf x}$ satisfying~\eqref{eq:diagram}, \begin{equation} \label{eq:xx} \log \mathbb{P} ({\bf X} = {\bf x}) = L_m - c_m \ell - \frac{d_m}{m} n \, . \end{equation} Three things are equivalent: $(i)$ the vector ${\bf X}$ satisfies the identities~\eqref{eq:diagram}; $(ii)$ the pair $(S_m , T_m)$ is equal to $(\ell,n)$; $(iii)$ the partition $\lambda= (\lambda_1 , \ldots , \lambda_m)$ defined by $\lambda_j - \lambda_{j+1} = X_j$ for $2 \leq j \leq m-1$, together with $\lambda_1 = \ell - X_0$ and $\lambda_m = X_m$, is a partition of $n$ fitting inside a $m \times \ell$ rectangle. Thus, \begin{align} N_n(\ell , m) &= \mathbb{P}_m \left [ (S_m,T_m) = (\ell , n) \right ] \, \exp \left(- L_m + c_m \ell + \frac{d_m}{m} n\right) \nonumber \\ &= \mathbb{P}_m \left [ (S_m,T_m) = (\ell , n) \right ] \, \exp \left [ m \left ( -\frac{L_m}{m} + c_m A + d_m B \right ) \right ] \label{eq:P_m} \, . \end{align} Comparing~\eqref{eq:N'} to~\eqref{eq:P_m}, the proof is completed by an application of the LCLT in Lemma~\ref{lem:lclt}. \end{proof} Lemma~\ref{lem:lclt} is stated for an arbitrary sequence of parameters $p_0, \ldots , p_m$ bounded away from 0 and 1, though we need it only for $p_j = 1 - e^{-c_m - d_m j/m}$. For a $2 \times 2$ matrix $M$, denote by $M(s,t):= [s\, , \,\, t] \, M\, [s\,,\,\,t]^T$ the corresponding quadratic form. \begin{lem}[LCLT]\label{lem:lclt} Fix $0 < \delta < 1$ and let $p_0, \ldots , p_m$ be any real numbers in the interval $[\delta , 1 - \delta]$. Let $\{ X_j \}$ be independent reduced geometrics with respective parameters $\{ p_j \}$, $S_m := \sum_{j=0}^m X_j,$ and $T_m := \sum_{j=0}^m j X_j$. Let $M_m$ be the covariance matrix for $(S_m , T_m)$, written \[ M_m = \displaystyle \left ( \begin{array}{cc} \alpha_m m & \beta_m m^2 \\ \beta_m m^2 & \gamma_m m^3 \end{array} \right ) \, ,\] $Q_m$ denote the inverse matrix to $M_m$, and $\Delta_m = m^{-4} \det M_m = \alpha_m \gamma_m - \beta_m^2$. Let $\mu_m$ and $\nu_m$ denote the respective means $\mathbb{E} S_m$ and $\mathbb{E} T_m$. Denote $p_m(a,b) := \mathbb{P} ((S_m , T_m) = (a,b))$. Then \begin{equation} \label{eq:LCLT} \sup_{a,b \in \mathbb{Z}} m^2 \left | p_m (a,b) - \frac{1}{2 \pi (\det M_m)^{1/2}} e^{-\frac{1}{2} Q_m (a - \mu_m , b - \nu_m)} \right | \to 0 \end{equation} as $m \to \infty$, uniformly in the parameters $\{ p_j \}$ in the allowed range. In particular, if the sequence $(a_m , b_m)$ satisfies $Q_m (a_m - \mu_m , b_m - \nu_m) \to 0$ then \[ \mathbb{P}(S_m = a_m , T_m = b_m) = \frac{1}{2\pi \sqrt{\Delta_m} m^2} \left(1+O\left(m^{-3/2}\right)\right) \,. \] \end{lem} The following consequence will be used to prove Theorem~\ref{th:difference}. \begin{cor}[LCLT consecutive differences]\label{cor:lclt_error} Let $\mathcal{N}_m(a,b) := \frac{1}{2 \pi (\det M_m)^{1/2}} e^{-\frac{1}{2} Q_m (a - \mu_m , b - \nu_m)}$ be the normal approximation in Equation~\eqref{eq:LCLT}. Using the notation of Lemma~\ref{lem:lclt}, \[ \sup_{a,b\in \mathbb{Z}} \bigg|p_m(a,b+1)-p_m(a,b) - \big(\mathcal{N}_m(a,b+1)-\mathcal{N}_m(a,b)\big) \bigg| = O(m^{-4}). \] \end{cor} The technical but unsurprising proofs of Lemma~\ref{lem:lclt} and Corollary~\ref{cor:lclt_error} are given in the Appendix at the end of this article. \section{Limit shape} \label{sec:shape} Suppose a Young diagram is chosen uniformly from among all partitions of $n$ fitting in a $m \times \ell$ rectangle. To simplify calculations, we imagine this Young diagram outlining a compact set in the fourth quadrant of the plane and rotate $90^\circ$ counterclockwise to obtain a shape in the first quadrant. Let $\Xi_{n,m,\ell}$ denote the random set obtained in this manner after rescaling by a factor of $1/m$, so that the length in the positive $x$-direction is bounded by~1. Fix $A > 2B > 0$ and metrize compact sets of $\mathbb{R}^2$ by the Hausdorff metric. As $m \to \infty$ with $\ell/m \to A$ and $n/m^2 \to B$, the random set $\Xi_{n,m,\ell}$ converges in distribution to a deterministic set $\Xi^{A,B}$. See Figure~\ref{fig:limtcurve} for some examples. Our methods immediately recover the distributional convergence result $\Xi_{n,m,\ell} \to \Xi^{A,B}$. As previously mentioned, this limit shape was known to Petrov~\cite{Petrov2009} and others. Petrov identifies it as a portion of the limit curve for unrestricted partitions, which itself was posed as a problem by Vershik and answered in~\cite{SzalayTuran1,SzalayTuran2} (see also~\cite{Vershik96}). Because this result is already known, along with precise fluctuation information which we do not derive, we give only the short argument here for distributional convergence. We do not determine the best possible fluctuation results following from this method. The shape $\Xi_{n,m,\ell}$ is determined by its boundary, a polygonal path obtained from a partition $\lambda$ by filling in unit vertical connecting lines in the step function $x \mapsto m^{-1} \lambda_{\lfloor mx \rfloor}$. Recall that the probability measure $\mathbb{P}_m$ restricted to the event $\{ (S_m, T_m) = (\ell , n) \}$ gives all partitions counted by $N_n(m , \ell)$ equal probability and that $\mathbb{P}_m$ gives the event $\{ (S_m, T_m) = (\ell , n) \}$ probability $\Theta (m^{-2})$. Distributional convergence of $\Xi_{n,m,\ell}$ to $\Xi^{A,B}$ then follows from the following. \begin{pr} \label{pr:shape} Fix $A > 2B > 0$. Define the maximum discrepancy by $$\mathcal{M} := \max_{0 \leq j \leq m} \left | \sum_{i=0}^j \left ( X_i - \frac{q_i}{p_i} \right ) \right | \, .$$ Then for any $\varepsilon > 0$, $$\mathbb{P}_m \left [ \mathcal{M} \geq \varepsilon m \right ] = o(m^{-2})$$ as $m \to \infty$ with $\ell/m \to A$ and $n/m \to B$. \end{pr} \begin{proof} This is a routine application of exponential moment bounds. By our definition of $p_i$, in this regime there exists $\delta>0$ such that $p_i \in [\delta , 1 - \delta]$ for all $i$. Therefore, there are $\eta,K>0$ such that for $\lambda < \eta$, the mean zero variables $X_i - q_i / p_i$ all satisfy $\mathbb{E} \exp (\lambda (X_i - q_i / p_i)) \leq \exp (K \lambda^2)$. Independence of the family $\{ X_i \}$ then gives $$\mathbb{E} \exp \left [ \lambda \sum_{i=0}^j (X_i - p_i / q_i) \right ] \leq e^{K m \lambda^2}$$ for all $j \leq m$. By Markov's inequality, $$\mathbb{P} ( |X_i - p_i / q_i| \geq \varepsilon m) \leq e^{K m \lambda^2 - \lambda m} \, .$$ Fixing $\lambda = 1/(2K)$ shows that this probability is bounded above by $\exp (- m/(4K))$. Hence, $\mathbb{P} (\mathcal{M} \geq \varepsilon m) \leq m e^{-m/(4K)} = o(m^{-2})$ as desired. \end{proof} To see that Proposition~\ref{pr:shape} implies the limit shape statement, let $\lambda_i := \ell - (X_0 + \cdots + X_{i-1})$ so that \[ y^{(m)}(i) := \mathbb{E}_m \lambda_i = \ell - \sum_{j=0}^{i-1} q_j / p_j.\] Proposition~\ref{pr:shape} shows the boundary of $\Xi_m$ to be within $o(m)$ of the step function $y^{(m)}(\cdot)$ except with probability $o(m^{-2})$. Since $\mathbb{P}_m$ restricted to the event $\{ (S_m, T_m) = (\ell , n) \}$ gives all partitions counted by $N_n(m , \ell)$ equal probability and $\mathbb{P}_m$ gives the event $\{ (S_m, T_m) = (\ell , n) \}$ probability $\Theta (m^{-2})$, the conditional law $(\mathbb{P}_m \, | \, (S_m , T_m) = (\ell , n))$ gives the event $\{ \mathcal{M} > \varepsilon m \}$ probability $o(1)$ as $m \to \infty$ with $\ell/m \to A$ and $n/m \to B$. Thus, the boundary of $\Xi_m$ converges in distribution to the limit \begin{equation} \label{eq:y(x)} y(x) := \lim_{m \to \infty} m^{-1} y^{(m)} (\lfloor mx \rfloor ) \, . \end{equation} Figure~\ref{fig:limtcurve} shows examples of two families of the limit curve as well as a plot of the limit curve against uniformly generated restricted partitions for several values of $m$ in the range $[120,300]$. \begin{figure}[!ht] \centering \begin{subfigure}{0.315\textwidth} \centering \includegraphics[width=\linewidth]{limcurves} \subcaption*{$(A,B)=(1,1/k)$ \\ $k=2,\dots,15$} \end{subfigure} \begin{subfigure}{0.315\textwidth} \centering \includegraphics[width=\linewidth]{limmul} \subcaption*{$(A,B)=(5/k,1/k)$ \\ $k=2,\dots,15$} \end{subfigure} \begin{subfigure}{0.35\textwidth} \centering \includegraphics[width=0.9\linewidth]{limsample} \subcaption*{Limit curve of $(A,B)=(1,1/3)$ and random partitions of size {\color{red}120}, {\color{darkgreen}201} and {\color{blue}300}.} \end{subfigure} \caption{Limit shapes of scaled partitions as $m\rightarrow \infty$.} \label{fig:limtcurve} \end{figure} Substituting the definition of $y^{(m)}(i)$ into~\eqref{eq:y(x)} and evaluating the limit as an integral gives $$y(x) = A + x - \int_0^x \frac{1}{1-e^{-c-dt}}dt = A+x - \frac{1}{d}\ln\left(\frac{e^{xd+c}-1}{e^c-1}\right) \, .$$ After expressing $c$ in terms of $d$, this may be written implicitly as $$e^{(A+1)d}-1 = (e^d-1) e^{d(A-y)} + (e^{Ad}-1)e^{d(1-x)}$$ which simplifies to \begin{equation} \label{eq:lim_curve} (1-e^{-c})e^{d(A-y)} + e^{-c}e^{-dx} = 1 \end{equation} as long as $A > 2B$; in the special case $A = 2B$ one obtains simply $y = A \cdot (1-x)$. It is worth comparing this result with the limit shape derived in~\cite{Petrov2009}. There the limit shape of the boxed partitions is identified as the portion of the curve $\{ e^{-x} + e^{-y} = 1 \}$, which is the limit shape of unrestricted partitions. The portion is determined implicitly by the restriction that the endpoints of the curve are the opposite corners of a $1 \times A$-proportional rectangle and that the area under the curve has the desired proportion, that is $B/A$ of the total rectangular area. To see that this matches~\eqref{eq:lim_curve} we can calculate the given portion explicitly. Let $x=s_1,s_2$ be the starting and ending points of the bounding rectangle. The side ratio and the area requirement are respectively equivalent to \begin{align*} \frac{\log(1-e^{s_1})-\log(1-e^{-s_2})}{s_2-s_1} & = A \\ \mbox{and} \hspace{4in} & \hspace{4in} \\ \int_{s_1}^{s_2} - \log(1-e^{-t})dt + (s_2-s_1) \log(1-e^{-s_2}) & = B (s_2-s_1)^2 \end{align*} which simplify to \begin{eqnarray} A & = & \frac{1}{s_2-s_1} \log\left(\frac{e^{s_2}-1}{e^{s_2} - e^{s_2-s_1}}\right) \label{eq:compare A} \, , \\ B & = & \frac{-{\rm dilog}\,(1-e^{-s_2}) + {\rm dilog}\,(1-e^{-s_1}) + (s_2-s_1)\log(1-e^{-s_2})}{(s_2-s_1)^2} \, . \label{eq:compare B} \end{eqnarray} Comparing these equations with equations~\eqref{eq:integrals A} and~\eqref{eq:integrals B} it is immediate that the solutions are given by $s_1=c$ and $s_2=c+d$. Finally, to match the curve in the second line of equation~\eqref{eq:lim_curve} we need the coordinate transform from the curve $\gamma$ in the segment $x = [c,c+d]$ given by $$x \to x_1 = \frac{(x-c)}{d}, \qquad y \to y_1-A = \frac{y+\log(1-e^{-c})}{d}$$ whence $x = d x_1 + c$ and $y = -d (A-y_1) - \log(1-e^{-c})$ and the curves match. \section{Existence and Uniqueness of \texorpdfstring{$c,d$}{c,d}} \label{sec:constants} We now show that for any $A \geq 2B > 0$ there exists unique positive constants $c$ and $d$ satisfying Equations~\eqref{eq:integrals A} and~\eqref{eq:integrals B}. If $A=B/2$ then $d=0$ and $c$ can be determined uniquely, so we may assume $A > 2B > 0$. The following lemma will be used to show uniqueness. \begin{lem} \label{lem:J} Let $\psi$ denote the map taking the pair $(c,d)$ to $(A,B)$ defined by the two integrals in Equations~\eqref{eq:integrals A} and~\eqref{eq:integrals B}, and let $K$ be a compact subset of $\{ (x,y) : x > 2y > 0 \}$. The Jacobian matrix $J := D[\psi]$ is negative definite for all $(c,d) \in (0,\infty)^2$, and all entries of $\psi$ and $J$ (respectively $\psi^{-1}$ and $J^{-1}$) are Lipshitz continuous on $\psi^{-1}[K]$ (respectively $K$). \end{lem} \begin{proof} Differentiating under the integral sign shows that the partial derivatives comprising the entries of $D[\psi]$ are given by \begin{eqnarray*} J_{A,c} & = & \int_0^1 \frac{- e^{-(c+dt)}}{(1 - e^{-(c+dt)})^2} \, dt \\ J_{A,d} & = & \int_0^1 \frac{- t \, e^{-(c+dt)}}{(1 - e^{-(c+dt)})^2} \, dt \\ J_{B,c} & = & \int_0^1 \frac{- t \, e^{-(c+dt)}}{(1 - e^{-(c+dt)})^2} \, dt \\ J_{B,d} & = & \int_0^1 \frac{- t^2 \, e^{-(c+dt)}}{(1 - e^{-(c+dt)})^2} \, dt \, ; \end{eqnarray*} note that each term is negative. Let $\rho$ denote the finite measure on $[0,1]$ with density $e^{-(c+dt)} / (1 - e^{-(c+dt)})^2$ and let $\mathbb{E}_\rho$ denote expectation with respect to $\rho$. Then \[ J_{A,c} = \mathbb{E}_\rho [-1], \qquad J_{A,d} = J_{B,c} = \mathbb{E}_\rho [-t], \qquad J_{B,d} = \mathbb{E}_\rho [-t^2], \] and \[ \det J = \mathbb{E}_\rho [1] \cdot \mathbb{E}_\rho [t^2] - \left(\mathbb{E}_\rho [t]\right)^2 = \mathbb{E}_{\rho}[1]^2 \cdot {\rm Var} \,_{\sigma}[t], \] where ${\rm Var} \,_{\sigma}[t]$ denotes the variance of $t$ with respect to the normalized measure $\sigma = \rho/\mathbb{E}_{\rho}[1]$. In particular, $\det J$ is positive, and bounded above and below when $c$ and $d$ are bounded away from 0, implying the stated results on Lipshitz continuity. As $J$ is real and symmetric, it has real eigenvalues. Since the trace of $J$ is negative while its determinant is positive, the eigenvalues of $J$ have negative sum and positive product, meaning both are strictly negative and $J$ is negative definite for any $c,d>0$. \end{proof} \begin{lem} \label{lem:uniquecd} For any $A>0$ and $B \in (0,A/2)$ there exist unique $c,d>0$ satisfying Equations~\eqref{eq:integrals A} and~\eqref{eq:integrals B}. Moreover, for a fixed $A$, when $B$ decreases from $A/2$ to $0$ then $d$ increases strictly from $0$ to $\infty$ and $c$ decreases strictly from $\log\left(\frac{A+1}{A}\right)$ to $1$. When $B>0$ is fixed and $A$ goes to $\infty$ then $c$ goes to $0$ and $d$ goes to the root of $$d^2 =B \left( d\log (1-e^{-d}) -{\rm dilog}\,(1-e^{-d}) \right).$$ \end{lem} \begin{proof} Solving Equation~\eqref{eq:integrals A} for $c$ (assuming $d\geq 0$) gives \[ c=\log \left( \frac{e^{(A+1)d}-1}{e^{(A+1)d}-e^d}\right). \] Substituting this into Equation~\eqref{eq:integrals B} gives an explicit expression for $B$ in terms of $A$ and $d$, and shows that for fixed $A>0$ as $d$ goes from 0 to infinity $B$ goes from $A/2$ to 0. By continuity, this implies the existence of the desired $c$ and $d$. It also shows that, for a fixed $A$, $c$ is a decreasing function of $d$ with the given maximal and minimal values as $d$ goes from $0$ to $\infty$. To prove uniqueness, we note that for ${\bf x},{\bf y} \in \mathbb{R}^2$ Stokes' theorem implies \[ \psi({\bf y}) - \psi({\bf x}) = \int_0^1 D[\psi]\left( t{\bf x} + (1-t){\bf y} \right) \cdot ({\bf x} - {\bf y}) \, dt \] so that \[ ({\bf x}-{\bf y})^T \cdot \left(\psi({\bf y}) - \psi({\bf x})\right) = \int_0^1 \left[({\bf x}-{\bf y})^T \cdot D[\psi]\left( t{\bf x} + (1-t){\bf y} \right) \cdot ({\bf x} - {\bf y})\right] dt. \] When ${\bf x} \neq {\bf y}$, negative-definiteness of $D[\psi]$ implies that the last integrand is strictly negative on $[0,1]$, and $\psi({\bf y}) \neq \psi({\bf x})$. Thus, distinct values of $c$ and $d$ give distinct values of $A$ and $B$. To see the monotonicity, let $A$ be fixed and let $F_B(d)=B$ be the equation obtained after substituting $c=c(A,d)$ above in Equation~\eqref{eq:integrals B}, i.e. $F_B(d) = \psi_2(c(A,d),d)$. Then $d$ is a decreasing function of $B$ and vice versa since \[ \frac{\partial F_B(d)}{\partial d} = \frac{J_{B,d}J_{A,c} - J_{A,d}J_{B,c}}{J_{A,c}} = \frac{\det D[\psi]}{J_{A,c}} <0\, . \] For the last part, the explicit formula for $c$ in terms of $A$ and $d$ shows that $c\to 0$. Substitution in Equation~\eqref{eq:integrals B} gives the desired equation. \end{proof} \section{Proof of Theorem~\protect{\ref{th:main}} from the discretized result} \label{sec:implication} Here we show how $c_m$ and $d_m$ from the discretized result are related to $c,d$ defined independently of $m$. The proof below also shows that $c_m$ and $d_m$ exist and are unique. The Euler-MacLaurin summation formula~\cite[Section 3.6]{Bruijn1958} gives an expansion \begin{align} \frac{L_m}{m} &= \int_0^1 \log(1-e^{-c_m-d_mt})\, dt + \frac{\log(1-e^{-c_m})+\log(1-e^{-c_m-d_m})}{2m} + O(m^{-2}) \notag \\[+2mm] &= \frac{{\rm dilog}\,(1 - e^{-c_m-d_m}) - {\rm dilog}\,(1-e^{-c_m})}{d_m} + \frac{\log(1-e^{-c_m})+\log(1-e^{-c_m-d_m})}{2m} + O(m^{-2}) \label{eq:Lm1} \end{align} of the sum $L_m$ in terms of $c_m$ and $d_m$. Assume that there is an asymptotic expansion \begin{eqnarray} c_m & = & c + u m^{-1} + O(m^{-2}) \label{eq:c+} \\ d_m & = & d + v m^{-1} + O(m^{-2}) \label{eq:d+} \end{eqnarray} as $m\rightarrow\infty$, where $u$ and $v$ are constants depending only on $A$ and $B$. Under such an assumption, substitution of Equations~\eqref{eq:c+} and~\eqref{eq:d+} into Equation~\eqref{eq:Lm1} implies \begin{align} \label{eq:L+} \frac{L_m}{m} &= \frac{{\rm dilog}\,(1 - e^{-c-d}) - {\rm dilog}\,(1-e^{-c})}{d} + \frac{u A + v B}{m} + O(m^{-2}) \notag \\ &= \log(1-e^{-c-d}) - dB + \frac{u A + v B}{m} + O(m^{-2}). \end{align} Substituting Equations~\eqref{eq:c+}--\eqref{eq:L+} into Equation~\eqref{eq:N'} of Theorem~\ref{th:1'} and taking the limit as $m \rightarrow \infty$ then gives Theorem~\ref{th:main}, as \[ \Delta_m\rightarrow \left(\int_0^1 \frac{e^{-c-dt}}{(1-e^{-c-dt})^2}\,dt\right)\left(\int_0^1 \frac{t^2e^{-c-dt}}{(1-e^{-c-dt})^2}\,dt\right) - \left(\int_0^1 \frac{te^{-c-dt}}{(1-e^{-c-dt})^2}\,dt\right)^2 = \Delta. \] It remains to show the expansions in Equations~\eqref{eq:c+} and~\eqref{eq:d+}. For $x,y>0$, define \begin{align*} {\overline{S}}_m(x,y) & := \frac{1}{m} \sum_{j=0}^m \frac{1}{1 - e^{-(x+yj/m)}} - 1 \, , \\ {\overline{T}}_m(x,y) & := \frac{1}{m} \sum_{j=0}^m \frac{j/m}{1 - e^{-(x+yj/m)}} - \frac{1}{2} \, . \end{align*} Another application of the Euler-MacLaurin summation formula implies \begin{align} {\overline{S}}_m(c,d) &= A + A_1 (c,d) m^{-1} + O(m^{-2}) \, , \label{eq:A1} \\ {\overline{T}}_m(c,d) &= B + B_1 (c,d) m^{-1} + O(m^{-2}) \, , \label{eq:B1} \end{align} with \[ A_1 = \frac{1}{2}\left(\frac{1}{1-e^{-c}} + \frac{1}{1-e^{-c-d}}\right) \qquad \text{and} \qquad B_1 = \frac{1}{2(1-e^{-c-d})}. \] Let $\mathcal{J}$ denote the Jacobian $D[\psi]$ of the map $\psi$, introduced in Lemma~\ref{lem:J}, with respect to $c$ and $d$, and let \[ (c_m' , d_m') = (c,d) - m^{-1} \mathcal{J}^{-1} \cdot (A_1-1 , B_1-1/2)^T.\] A Taylor expansion around the point $(c,d)$ gives \begin{align*} \begin{pmatrix} {\overline{S}}_m(c_m',d_m') \\ {\overline{T}}_m(c_m',d_m') \end{pmatrix} &= \begin{pmatrix} {\overline{S}}_m(c,d) \\ {\overline{T}}_m(c,d) \end{pmatrix} - \left(\mathcal{J} + O\left(m^{-1}\right) \right) \cdot \left(m^{-1} \mathcal{J}^{-1} \begin{pmatrix} A_1 \\ B_1 \end{pmatrix} \right) + O(m^{-2}) \\ &= \begin{pmatrix} A-1/m \\ B-1/2m \end{pmatrix} + O\left(m^{-2}\right) \\ &= \begin{pmatrix} {\overline{S}}_m(c_m,d_m) \\ {\overline{T}}_m(c_m,d_m) \end{pmatrix} + O\left(m^{-2}\right), \end{align*} where Equations~\eqref{eq:A1} and~\eqref{eq:B1} were used to approximate the Jacobian of $\psi_m: (x,y) \mapsto \left({\overline{S}}_m(x,y),{\overline{T}}_m(x,y)\right)$ with respect to $x$ and $y$. The map $\psi_m$ is Lipschitz for a similar reason as its continuous analogue. Namely, consider the partial derivatives \begin{eqnarray*} J_{S,x} & = & \frac{1}{m}\sum_{j=0}^m \ -\frac{e^{-x-yj/m}}{(1-e^{-x-yj/m})^2}\, \\ J_{S,y} & = & \frac{1}{m^2}\sum_{j=0}^m \ -\frac{j\ e^{-x-yj/m}}{(1-e^{-x-yj/m})^2}\, \\ J_{T,x} & = & \frac{1}{m^2}\sum_{j=0}^m \ -\frac{j\ e^{-x-yj/m}}{(1-e^{-x-yj/m})^2}\, \\ J_{S,y} & = & \frac{1}{m^3}\sum_{j=0}^m \ -\frac{j^2\ e^{-x-yj/m}}{(1-e^{-x-yj/m})^2}. \end{eqnarray*} Let $\rho_m$ be a discrete finite measure on $R_m:=\{0,1/m,2/m,\ldots,1\}$ with density $e^{-x-yt}/(1-e^{-x-yt})$ for $t\in R_m$ and 0 otherwise, and let $\mathbb{E}_{\rho_m}$ be the expectation with respect to $\rho_m$. Then \[ J_{S,x} = \mathbb{E}_{\rho_m}[-1] \, , \qquad J_{T,x}=J_{S,y} = \mathbb{E}_{\rho_m}[-t] \, , \qquad J_{T,y} =\mathbb{E}_{\rho_m}[-t^2]\] and \[ \det D[\psi_m] = \mathbb{E}_{\rho_m}[1]\mathbb{E}_{\rho_m}[t^2] - \mathbb{E}_{\rho_m}[t]^2 = \mathbb{E}_{\rho_m}[1]^2 {\rm Var} \,_{\sigma_m} [t]\, ,\] where $\sigma_m$ is the probability function $\rho_m/\mathbb{E}_{\rho_m}[1]$. For any fixed $m$ and $(x,y)$ in a compact neighborhood of $(A,B)$, both the variance and the expectation are finite and bounded away from 0, as is the Jacobian determinant. Moreover, the trace ${\rm Tr} D[\psi] = -\mathbb{E}_{\rho_m}[1+t^2]$ is bounded away from 0 and infinity, so the Jacobian is negative definite with locally bounded eigenvalues, and hence $\psi_m$ is locally Lipschitz. Since the norm of the Jacobian is bounded away from 0 and infinity, we have that the inverse map $\psi_m^{-1}$ is also locally Lipschitz in a neighborhood of $\psi^{-1}(A,B)$. Moreover, similarly to proof of existence and uniqueness of $c$ and $d$ in Section~\ref{sec:constants}, we have that there indeed are $c_m$ and $d_m$ as unique solutions of Equations~\eqref{eq:mu_m} and~\eqref{eq:nu_m} since the Jacobian is negative semi-definite. The trapezoid formula implies $\left|J_{S,c}-J_{A,c}\right| = O(m^{-1})$, and similar bounds for the other differences of partial derivatives in the continuous and discrete settings. Hence, the bounds for the norms and eigenvalues of $D[\psi_m]$ are within $O(m^{-1})$ of the ones for $D[\psi]$, and $\psi_m$ (and its inverse) is Lipschitz with a constant independent of $m$. Thus, \[ O(m^{-2})=\|\psi_m(c'_m,d'_m) -\psi_m(c_m,d_m)\|\geq C^{-1} \| (c_m'-c_m,d'_m-d_m)\| \] for some constant $C$, so that the expansions~\eqref{eq:c+} and~\eqref{eq:d+} hold. $\hfill \Box$ \section{Proof of Theorem~\protect{\ref{th:difference}}} \label{sec:proof} We will prove Theorem~\ref{th:difference} from Equation~\eqref{eq:P_m} and Corollary~\ref{cor:lclt_error}. Let $p_m(\ell,n) = \mathbb{P}_m \left [ (S_m,T_m) = (\ell , n) \right ]$ and let \begin{align}\label{eq:discrete_funcs} L_m(x,y) &:= \sum_{j=0}^m \log(1-e^{-x-yj/m}) \ ,\\ A_m(x,y) &:= \sum_{j=0}^m \frac{1}{1-e^{-x-yj/m}} - (m+1) \ ,\\ B_m(x,y) &:= \sum_{j=0}^m \frac{j/m}{1-e^{-x-yj/m}} - \frac{m+1}{2} \ . \end{align} Then $c_m$ and $d_m$ are the solutions to \[ A_m(c_m,d_m) = \ell= Am \, , \qquad B_m(c_m,d_m) = n/m = Bm \, . \] Let $c'_m,d'_m$ be the solutions to $A_m(c'_m,d'_m)=\ell$ and $B_m(c'_m,d'_m)=(n+1)/m$, and let $\Delta x = c'_m-c_m=O(m^{-2})$ and $\Delta y = d'_m-d_m=O(m^{-2})$ by the Lipschitz properties proven in Section~\ref{sec:constants}. Observe that \begin{equation}\label{eq:derivatives} \frac{\partial L_m(x,y) }{\partial x} = A_m(x,y) \quad \text{and} \quad \frac{\partial L_m(x,y)}{\partial y} = B_m(x,y). \end{equation} Using the Taylor expansion for $L_m(c'_m,d'_m)$ around $(c_m,d_m)$ and the $L_m$ partial derivatives from Equation~\eqref{eq:derivatives}, \[ -L_m(c_m', d_m') = -L_m(c_m+\Delta x, d_m +\Delta y) = -L_m(c_m,d_m) - \Delta x \, A_m(c_m,d_m) - \Delta y \, B_m(c_m,d_y) + O(m^{-3}), \] so that \begin{align*} -L_m(c_m', d_m') + (c_m+\Delta x) \ell + (d_m + \Delta y) (n + 1) \, m^{-1} =-L_m(c_m,d_m) + c_m\ell + d_m(n+1) \, m^{-1} + O(m^{-3}). \end{align*} To lighten notation, we now write $L_m := L_m(c_m,d_m)$ and $L_m' := L_m(c_m',d_m')$. Then \begin{align} N_{n+1}(\ell,m) - N_n(\ell,m) &= p_m(\ell,n+1)\exp \left[ -L_m' + c_m' \ell + \frac{d_m'}{m}(n+1) \right] - p_m(\ell,n)\exp \left[ -L_m + c_m \ell + \frac{d_m}{m}n \right] \notag \\[+2mm] &= p_m(\ell,n)\exp \left[ -L_m + c_m \ell + \frac{d_m}{m}n \right]\left[e^{d_m/m}-1\right] \label{eq:diffE1} \\[+1mm] & \quad + \, \left[p_m(\ell,n+1)-p_m(\ell,n)\right]\exp \left[ -L_m + c_m \ell + \frac{d_m}{m}(n+1) \right] \label{eq:diffE2} \\[+1mm] & \quad + \, p_m(\ell,n+1)\left( e^{-L_m' + c_m' \ell +d_m'(n+1)/m} - e^{-L_m + c_m \ell + d_m(n+1)/m} \right). \label{eq:diffE3} \end{align} We now bound each of these summands. \begin{itemize} \item Since $d_m = d + O(m^{-1})$, Equation~\eqref{eq:P_m} implies that the quantity on line~\eqref{eq:diffE1} equals \[ N_n(\ell,m)\left(\frac{d}{m} + O(m^{-2}) \right) \] as long as $d \notin O(m^{-1})$. This holds when $|A-B/2| \notin O(m^{-1})$ as $d=0$ when $A=B/2$ and the map taking $(A,B)$ to $(c,d)$ is Lipschitz. \item By Corollary~\ref{cor:lclt_error}, \begin{align*} \left[p_m(\ell,n+1)-p_m(\ell,n)\right] &\leq \left|\mathcal{N}_m(\ell,n+1)-\mathcal{N}_m(\ell,n)\right| + O(m^{-4}) \\ &= O\left(m^{-2} \cdot \left|1 - e^{\frac{1}{2} Q_m (0,1)} \right| \right) + O(m^{-4}) \\ &= O(m^{-4}), \end{align*} where $Q_m$ is the inverse of the covariance matrix of $(S_m,T_m)$. Thus, the quantity on line~\eqref{eq:diffE2} is $O(m^{-4} \cdot m^2 N_n(\ell,m)) = O(m^{-2}N_n(\ell,m))$. \item Let \[ \psi_m := \exp\bigg[ -L_m' + c_m' \ell +d_m'(n+1) \, m^{-1} - \left(-L_m + c_m \ell +d_m(n+1) \, m^{-1}\right) \bigg] - 1 = O(m^{-3}). \] As $p_m(\ell,n+1) = p_m(\ell,n) + O(m^{-4})$, it follows that the quantity on line~\eqref{eq:diffE3} is \begin{align*} p_m(\ell,n+1) \, e^{-L_m + c_m \ell + d_m(n+1)/m} \, \psi_m &= N_n(\ell,m) \, \psi_m \, e^{d_m/m} + O(m^{-4} \, e^{d_m/m} \, e^{-L_m + c_m \ell + d_m n/m} \, \psi_m) \\ &= O(m^{-3} N_n(\ell,m)). \end{align*} \end{itemize} Putting everything together, \[ N_{n+1}(\ell,m) - N_n(\ell,m) = N_n(\ell,m)\left(\frac{d}{m} + O(m^{-2}) \right), \] as desired. $\hfill \Box$ \section*{Appendix: Proof of the Local Central Limit Theorem} Throughout this section, $1/2 \geq \delta > 0$ is fixed and $\{ p_j : 0 \leq j \leq m \}$ are arbitrary numbers in $[\delta , 1-\delta]$. The variables $\{ X_j \}$ and $(S_m, T_m)$ are as in Lemma~\ref{lem:lclt}; we drop the index $m$ on the remaining quantities $\alpha_m, \beta_m, \gamma_m, \Delta_m, \mu_m, \nu_m, p_m(a,b)$ and the matrices $M_m$ and $Q_m$. Recall the quadratic form notation $M(s,t):= [s\, , \,\, t] \, M\, [s\,,\,\,t]^T$. \begin{lem} \label{lem:delta} The constants $\alpha, \beta, \gamma$ and $\Delta$ are bounded below and above by positive constants depending only on $\delta$. \end{lem} \noindent{\sc Proof:} Upper and lower bounds on $\alpha, \beta$ and $\gamma$ are elementary: $\displaystyle \alpha \in \left [ \frac{\delta}{(1-\delta)^2}, \frac{(1-\delta)}{\delta^2} \right ] , \beta \in \left [ \frac{\delta}{2(1-\delta)^2}, \frac{(1-\delta)}{2\delta^2} \right ]$ and $\displaystyle \gamma \in \left [ \frac{\delta}{3(1-\delta)^2}, \frac{(1-\delta)}{3\delta^2} \right ]$. The upper bound on $\Delta$ follows from these. For the lower bound on $\Delta$, let $\displaystyle {\tilde{M}} = \left ( \begin{array}{cc} \alpha_n & \beta_n \\ \beta_n & \gamma_n \end{array} \right )$ denote $M$ without the factors of $m$. We show $\Delta$ is bounded from below by the positive constant $(4 - \sqrt{13}) \delta / 6$. A lower bound for the determinant $\Delta$ of $\tilde{M}$ is $|\lambda|^2$ where $\lambda$ is the least modulus eigenvalue of $\tilde{M}$; note that $|\lambda|^2 = \inf_\theta \tilde{M} (\cos \theta , \sin \theta)$. We compute \begin{align*} \tilde{M} (\cos \theta , \sin \theta) & = m^{-1} \mathbb{E} \left (\cos \theta S + m^{-1} \sin \theta T \right )^2 \\ & \geq \delta m^{-1} \sum_{k=0}^m \left ( \cos \theta + \frac{k}{m} \sin \theta \right )^2 \\ & > \delta \cdot \left ( \cos^2 \theta + \cos \theta \sin \theta + \frac{1}{3} \sin^2 \theta \right ) \, . \end{align*} This is at least $\displaystyle \frac{4 - \sqrt{13}}{6} \delta$ for all $\theta$, proving the lemma. $\hfill \Box$ \begin{lem} \label{lem:unif} Let $X_p$ denote a reduced geometric with parameter $p$. For every $\delta \in (0,1/2)$ there is a $K$ such that simultaneously for all $p \in [\delta , 1-\delta]$, \[ \left | \log \mathbb{E} \exp(i \lambda X_p) - \left ( i \frac{q}{p} \lambda - \frac{q^2}{2 p^2} \lambda^2 \right ) \right | \leq K \lambda^3 \, . \] \end{lem} \noindent{\sc Proof:} For fixed $p$ this is Taylor's remainder theorem together with the fact that the characteristic function $\phi_p (\lambda)$ of $X_p$ is thrice differentiable. The constant $K(p)$ one obtains this way is continuous in $p$ on the interval $(0,1)$, therefore bounded on any compact sub-interval. $\hfill \Box$ \noindent{\sc Proof of the LCLT:} The proof of Lemma~\ref{lem:lclt} comes from expressing the probability as an integral of the characteristic function, via the inversion formula, and then estimating the integrand in various regions. Let $\phi(s,t) := \mathbb{E} e^{i (s S + t T)}$ denote the characteristic function of $(S , T)$. Centering the variables at their means, denote $\ov{S} := S - \mu$, $\ov{T} := T-\nu$, and $\ov{\phi}(s,t) := \mathbb{E} e^{i(s\ov{S} +t \ov{T})}$ so that $\phi(s,t) = \ov{\phi}(s,t) e^{is\mu + it\nu}$. Then \begin{align} p(a,b) & = \frac{1}{(2 \pi)^2} \int_{-\pi}^\pi \int_{-\pi}^\pi e^{-i s a - i t b} \phi (s,t) \, ds \, dt \nonumber \\ & = \frac{1}{(2 \pi)^2} \int_{-\pi}^\pi \int_{-\pi}^\pi e^{-i s (a-\mu) - i t (b-\mu)} {\ov{\phi}} (s,t) \, ds \, dt \, . \label{eq:inversion} \end{align} Following the proof of the univariate LCLT for IID variables found in~\cite{Durrett2010}, we observe that \begin{equation} \label{eq:normal} \frac{1}{2 \pi (\det M)^{1/2}} e^{-\frac{1}{2} Q (a - \mu , b - \nu)} = \frac{1}{(2 \pi)^2} \int_{-\infty}^{\infty} \int_{-\infty}^{\infty} e^{-i s (a-u) - i t (b-v)} \exp \left ( - \frac{1}{2} M(s,t) \right ) \, ds \, dt \, . \end{equation} Hence, comparing this to~\eqref{eq:inversion} and observing that $e^{-is(a-\mu) - it(b-\nu)}$ has unit modulus, the absolute difference between $p(a,b)$ and the left-hand side of~\eqref{eq:normal} is bounded above by \begin{equation} \label{eq:diff} \frac{1}{(2 \pi)^2}\int_{-\infty}^{\infty} \int_{-\infty}^{\infty} \left | {\bf 1}_{(s,t) \in [-\pi,\pi]^2} \ov{\phi} (s,t) - e^{-(1/2) M(s,t)} \right | ds \, dt \, . \end{equation} Fix positive constants $L$ and $\varepsilon$ to be specified later and decompose the region ${\mathcal R} := [-\pi,\pi]^2$ as the disjoint union ${\mathcal R}_1 + {\mathcal R}_2 + {\mathcal R}_3$, where \begin{align*} {\mathcal R}_1 & = [-L m^{-1/2} , L m^{-1/2}] \times [-L m^{-3/2} , L m^{-3/2}] \\ {\mathcal R}_2 & = [-\varepsilon , \varepsilon] \times [-\varepsilon m^{-1} , \varepsilon m^{-1}] \;\; \setminus \; {\mathcal R}_1 \\ {\mathcal R}_3 & = {\mathcal R} \setminus ({\mathcal R}_1 \cup {\mathcal R}_2) \, ; \end{align*} see Figure~\ref{fig:rectangles} for details. \begin{figure} \centering \includegraphics[width=0.3\linewidth]{rectangles} \caption{The regions ${\mathcal R}_1 \subseteq {\mathcal R}_2 \subseteq {\mathcal R}$ in the proof of the LCLT.} \label{fig:rectangles} \end{figure} As $\int_{{\mathcal R}_2^c} e^{- (1/2) M(s,t)} \, ds \, dt$ decays exponentially with $m$, it suffices to obtain the following estimates \begin{align} \int_{{\mathcal R}_1} \left | {\ov{\phi}}(s,t) - e^{-(1/2) M(s,t)} \right | \, ds \, dt & = O\left ( m^{-5/2} \right ) \label{eq:i}\\ \int_{{\mathcal R}_2} \left | {\ov{\phi}}(s,t) - e^{-(1/2) M(s,t)} \right | \, ds \, dt & = O(m^{-5/2}) \label{eq:ii} \\ \int_{{\mathcal R}_3} \left | {\ov{\phi}}(s,t) \right | \, ds \, dt & = o \left ( m^{-3} \right ). \label{eq:iii} \end{align} By independence of $\{ X_j \}$, \[ \log {\ov{\phi}} (s,t) = \sum_{j=0}^m \log \mathbb{E} e^{i (s+jt) (X_j - \mu_j)} \, .\] Using Lemma~\ref{lem:unif} with $p = p_j$ gives \[ \left | \log \mathbb{E} e^{i (s+jt) (X_j - q_j/p_j)} + \frac{q_j}{2 p_j^2} (s+jt)^2 \right | \leq K |s+jt|^3 \, .\] The sum of $(q_j / p_j^2) (s+jt)^2$ is $M(s,t)$, therefore summing the previous inequalities over $j$ gives \begin{equation} \label{eq:R1} \left | \log {\ov{\phi}}(s,t) + \frac{1}{2} M(s,t) \right | \leq K \sum_{j=0}^m |s + jt|^3 \, . \end{equation} On ${\mathcal R}_1$ we have the upper bound $|s + jt| \leq |s| + m |t| \leq 2 L m^{-1/2}$. Thus, \[ \sum_{j=0}^m |s+jt|^3 \leq (m+1) (8 L^3) m^{-3/2} = O \left ( m^{-1/2} \right ) \, .\] Plugging this into~\eqref{eq:R1} and exponentiating shows that the left hand side of~\eqref{eq:i} is at most $|{\mathcal R}_1| \cdot O(m^{-1/2}) = O(m^{-5/2})$. To bound the integral on ${\mathcal R}_2$, we define the sub-regions \[ S_k := \left\{(x,y) : k \leq \max\left(m^{1/2}|x|,m^{3/2}|y|\right) \leq k+1 \right\}.\] As the area of $S_k$ is $(8k+4)m^{-2}$, \begin{align} \int_{{\mathcal R}_2} \left | {\ov{\phi}}(s,t) - e^{-(1/2) M(s,t)} \right | \, ds \, dt &\leq \sum_{k=L}^{\lceil \, \epsilon \sqrt{m} \, \rceil} \int_{S_k}\left|{\ov{\phi}}(s,t)-e^{-M(s,t)/2}\right|dsdt \notag \\ &\leq m^{-2} \sum_{k=L}^{\lceil \, \epsilon \sqrt{m} \, \rceil} (8k+4) \max_{(s,t) \in S_k} \left|{\ov{\phi}}(s,t)-e^{-M(s,t)/2}\right| \,. \label{eq:R2m} \end{align} We break this last sum into two parts, and bound each part. For $(s,t) \in {\mathcal R}_2$, we have $|s + j t| \leq |s| + m |t| \leq 2 \varepsilon$ so that \[ \sum_{j=0}^m |s + jt|^3 \leq 2 \varepsilon \sum_{j=0}^m (|s| + j|t|)^2\leq (2 \varepsilon \Delta^{-1}) M(|s|,|t|).\] Comparing this to~\eqref{eq:R1} shows we may choose $\varepsilon$ small enough to guarantee that \[\left | \log {\ov{\phi}}(s,t) + \frac{1}{2} M(s,t) \right | \leq \frac{1}{4} M(|s|,|t|) \, , \] so $|{\ov{\phi}}(s,t)| \leq e^{-(1/4) M(s,t)}$. Lemma~\ref{lem:delta} shows there is a positive constant $c$ such that the minimum value of $M(s,t)$ on $S_k$ is at least $ck^2$. Thus, for $(s,t) \in S_k$, \[ \left|{\ov{\phi}}(s,t)-e^{-M(s,t)/2}\right| \leq \left|e^{-M(s,t)/4}\right| + \left|e^{-M(s,t)/2}\right| \leq 2e^{-ck^2}.\] If $r_m := \left\lceil \sqrt{(\log m)/c} \, \right\rceil$ then \begin{align} \sum_{k=r_m}^{\infty} (8k+4)(k+1) \max_{(s,t) \in S_k} \left|{\ov{\phi}}(s,t)-e^{-M(s,t)/2}\right| &\leq 2\sum_{k=r_m}^{\infty} (8k+4)(k+1) e^{-ck^2} \notag \\ &= O(m^{-1} \, \text{polylog}(m)) \notag \\ &= O(m^{-1/2}), \label{eq:R2m2} \end{align} where $\text{polylog}(m)$ denotes a quantity growing as an integer power of $\log m$. Furthermore, for $(s,t) \in S_k$ there exist constants $C$ and $C'$ such that \[ \left|\log {\ov{\phi}}(s,t) + M(s,t)/2 \right| \leq C \sum_{j=0}^m \left|s+jt\right|^3 \leq C \left(2(k+1)m^{-1/2}\right)^3(m+1) = C'k^3m^{-1/2}.\] This implies the existence of a constant $K>0$ such that for $0 \leq k \leq r_m$ and $(s,t) \in S_k$, \begin{align*} \left|{\ov{\phi}}(s,t)-e^{-M(s,t)/2}\right| &= \left|e^{-M(s,t)/2}\right|\left|1-e^{\log {\ov{\phi}}(s,t)+ M(s,t)/2}\right| \\ &\leq K e^{-ck^2} k^3m^{-1/2}. \end{align*} Thus, \begin{align} \sum_{k=L}^{r_m} (8k+4)(k+1) \max_{(s,t) \in S_k} \left|{\ov{\phi}}(s,t)-e^{-M(s,t)/2}\right| &\leq K m^{-1/2} \sum_{k=L}^{r_m} (8k+4)(k+1)k^3e^{-ck^2} \notag \\ &= O(m^{-1/2}). \label{eq:R2m3} \end{align} Combining~\eqref{eq:R2m}--\eqref{eq:R2m3} gives~\eqref{eq:ii}. Finally, for~\eqref{eq:iii}, we claim there is a positive constant $c$ for which $|{\ov{\phi}}(s,t)| \leq e^{-cm}$ on ${\mathcal R}_3$. To see this, observe (see~\cite[p. 144]{Durrett2010}) that for each $p$ there is an $\eta > 0$ such that $|\phi_p (\lambda)| < 1 - \eta$ on $[-\pi , \pi] \setminus [-\varepsilon/2,\varepsilon/2]$. Again, by continuity, we may choose one such $\eta$ valid for all $p \in [\delta , 1-\delta]$. It suffices to show that when either $|s|$ or $m |t|$ is at least $\varepsilon$, then at least $m/3$ of the summands $\log \mathbb{E} e^{i(s+jt)(X_j - \mu_j)}$ have real part at most $-\eta$. Suppose $s \geq \varepsilon$ (the argument is the same for $s \leq -\varepsilon$). Interpreting $s+jt$ modulo $2\pi$ always to lie in $[-\pi,\pi]$, the number of $j \in [0,m]$ for which $s + jt \in [-\varepsilon/2,\varepsilon/2]$ is at most twice the number for which $s + jt \in [\varepsilon/2,\varepsilon]$, hence at most twice the number for which $s + jt \notin [-\varepsilon/2,\varepsilon/2]$; thus at least $m/3$ of the $m+1$ values of $s+jt$ lie outside $[-\varepsilon/2,\varepsilon/2]$ and these have real part of $\log \mathbb{E} e^{i(s+jt)(X_j-\mu_j)} \leq -\eta$ by choice of $\eta$. Lastly, if instead one assumes $\pi \geq t \geq \varepsilon/m$, then at most half of the values of $s+jt$ modulo $2\pi$ can fall inside any interval of length $\varepsilon/2$. Choosing $\eta$ such that the real part of $\log \mathbb{E} e^{i(s+jt)(X_j - \mu_j)}$ is at most $-\eta$ outside of $[-\varepsilon/4,\varepsilon/4]$ finishes the proof of~\eqref{eq:iii} and the LCLT. $\hfill \Box$ \noindent{\sc Proof of Corollary~\ref{cor:lclt_error}.} In order to estimate the error terms in the approximation of $p(a,b)$ we will consider the partial differences and repeat the approximation arguments above. Changing $b$ to $b+1$ in Equations~\eqref{eq:inversion} and~\eqref{eq:normal} implies \begin{equation} \bigg|p(a,b+1)-p(a,b) - \big(\mathcal{N}(a,b+1)-\mathcal{N}(a,b)\big) \bigg| = \int_{[-\pi,\pi]^2} \left|1-e^{-it}\right| \left|{\ov{\phi}}(s,t) - e^{-1/2M(s,t)}\right|dsdt. \label{eq:pdiff} \end{equation} For $(s,t) \in {\mathcal R}_3$, the proof of the LCLT shows that the integral in Equation~\eqref{eq:pdiff} decays exponentially with $m$. As $\left|1-e^{-it}\right| = \sqrt{2-2\cos(t)} \leq |t| = O(m^{-3/2})$ for $(s,t) \in {\mathcal R}_1$, the proof of the LCLT shows that the integral in Equation~\eqref{eq:pdiff} grows as $O(m^{-3/2} \cdot m^{-5/2}) = O(m^{-4})$. Finally, since $\left|1-e^{-it}\right| \leq |t| \leq (k+1)m^{-3/2}$ for $(s,t) \in S_k$ following the proof of the LCLT shows that \begin{align*} \int_{{\mathcal R}_2} \left|1-e^{-it}\right| \left|{\ov{\phi}}(s,t) - e^{-1/2M(s,t)}\right|dsdt & \leq m^{-7/2} \sum_{k=L}^{\lceil \, \epsilon \sqrt{m} \, \rceil} (8k+4)(k+1) \max_{(s,t) \in S_k} \left|{\ov{\phi}}(s,t)-e^{-M(s,t)/2}\right| \\ & = O(m^{-4}). \end{align*} $\hfill \Box$ \section*{Acknowledgments} We are very thankful to an anonymous referee of an earlier version. This referee pointed out a body of literature that we had missed, in which the statistical mechanical ideas drawn on in this paper are already present. \bibliographystyle{amsalpha}
{ "timestamp": "2019-02-05T02:21:48", "yymm": "1805", "arxiv_id": "1805.08375", "language": "en", "url": "https://arxiv.org/abs/1805.08375", "abstract": "We consider the number of partitions of $n$ whose Young diagrams fit inside an $m \\times \\ell$ rectangle; equivalently, we study the coefficients of the $q$-binomial coefficient $\\binom{m+\\ell}{m}_q$. We obtain sharp asymptotics throughout the regime $\\ell = \\Theta (m)$ and $n = \\Theta (m^2)$. Previously, sharp asymptotics were derived by Takács only in the regime where $|n - \\ell m /2| = O(\\sqrt{\\ell m (\\ell + m)})$ using a local central limit theorem. Our approach is to solve a related large deviation problem: we describe the tilted measure that produces configurations whose bounding rectangle has the given aspect ratio and is filled to the given proportion. Our results are sufficiently sharp to yield the first asymptotic estimates on the consecutive differences of these numbers when $n$ is increased by one and $m, \\ell$ remain the same, hence significantly refining Sylvester's unimodality theorem.", "subjects": "Combinatorics (math.CO); Number Theory (math.NT); Probability (math.PR)", "title": "Counting partitions inside a rectangle", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9867771805808551, "lm_q2_score": 0.8289387998695209, "lm_q1q2_score": 0.8179778918093235 }
https://arxiv.org/abs/2101.06583
Powers of sums and their associated primes
Let $A, B$ be polynomial rings over a field $k$, and $I\subseteq A, J\subseteq B$ proper homogeneous ideals. We analyze the associated primes of powers of $I+J\subseteq A\otimes_k B$ given the data on the summands. The associated primes of large enough powers of $I+J$ are determined. We then answer positively a question about the persistence property of $I+J$ in many new cases.
\section{Introduction} \label{sect_intro} Inspired by work of Ratliff \cite{R}, Brodmann \cite{Br} proved that in any Noetherian ring, the set of associated primes of powers of an ideal is eventually constant for large enough powers. Subsequent work by many researchers have shown that important invariants of powers of ideals, for example, the depth and the Castelnuovo--Mumford regularity also eventually stabilize in the same manner. For a recent survey on associated primes of powers and related questions, we refer to Hoa's paper \cite{Hoa20}. Our work is inspired by the aforementioned result of Brodmann, and recent studies about powers of sums of ideals \cite{HNTT, HTT, NgV}. Let $A, B$ be standard graded polynomial rings over a field $k$, and $I\subseteq A, J\subseteq B$ proper homogeneous ideals. Denote $R=A\otimes_k B$ and $I+J$ the ideal $IR+JR$. Taking sums of ideals this way corresponds to the geometric operation of taking fiber products of projective schemes over the field $k$. In \cite{HNTT,HTT, NgV}, certain homological invariants of powers of $I+J$, notably the depth and regularity, are computed in terms of the corresponding invariants of powers of $I$ and $J$. In particular, we have exact formulas for $\depth R/(I+J)^n$ and $\reg R/(I+J)^n$ if either $\chara k=0$, or $I$ and $J$ are both monomial ideals. It is therefore natural to ask: \begin{quest} \label{quest_Assformula} Is there an exact formula for $\Ass(R/(I+J)^n)$ in terms of the associated primes of powers of $I$ and $J$? \end{quest} The case $n=1$ is simple and well-known: Using the fact that $R/(I+J)\cong (A/I)\otimes_k (B/J)$, we deduce (\cite[Theorem 2.5]{HNTT}): \[ \Ass R/(I+J)=\mathop{\bigcup_{{\mathfrak p} \in \Ass_A(A/I)}}_{{\mathfrak q} \in \Ass_B(B/J)} \Min_R(R/{\mathfrak p}+{\mathfrak q}). \] Unexpectedly, in contrast to the case of homological invariants like depth or regularity, we do not have a complete answer to Question \ref{quest_Assformula} in characteristic zero yet. One of our main results is the following partial answer to this question. \begin{thm}[Theorem \ref{thm_Asscontainments}] \label{thm_main1} Let $I$ be a proper homogeneous ideal of $A$ such that $\Ass(A/I^n)=\Ass(I^{n-1}/I^n)$ for all $n\ge 1$. Let $J$ be any proper homogeneous ideal of $B$. Then for all $n\ge 1$, there is an equality \[ \Ass_R \frac{R}{(I+J)^n} = \bigcup_{i=1}^n \mathop{\bigcup_{{\mathfrak p} \in \Ass_A(A/I^i)}}_{{\mathfrak q} \in \Ass_B(J^{n-i}/J^{n-i+1})} \Min_R(R/{\mathfrak p}+{\mathfrak q}). \] If furthermore $\Ass(B/J^n)=\Ass(J^{n-1}/J^n)$ for all $n\ge 1$, then for all such $n$, there is an equality \[ \Ass_R \frac{R}{(I+J)^n} = \bigcup_{i=1}^n \mathop{\bigcup_{{\mathfrak p} \in \Ass_A(A/I^i)}}_{{\mathfrak q} \in \Ass_B(B/J^{n-i+1})} \Min_R(R/{\mathfrak p}+{\mathfrak q}). \] \end{thm} The proof proceeds by filtering $R/(I+J)^n$ using exact sequences with terms of the form $M\otimes_k N$, where $M,N$ are nonzero finitely generated modules over $A,B$, respectively, and applying the formula for $\Ass_R(M\otimes_k N)$. Concerning Theorem \ref{thm_main1}, the hypothesis $\Ass(A/I^n)=\Ass(I^{n-1}/I^n)$ for all $n\ge 1$ holds in many cases, for example, if $I$ is a monomial ideal of $A$, or if $\chara k=0$ and $\dim(A/I)\le 1$ (see Theorem \ref{thm_specialcase_ass} for more details). We are not aware of any ideal in a polynomial ring which does not satisfy this condition (over non-regular rings, it is not hard to find such an ideal). In characteristic zero, we show that the equality $\Ass(A/I^n)=\Ass(I^{n-1}/I^n)$ holds for all $I$ and all $n$ if $\dim A\le 3$. If $\chara k=0$ and $A$ has Krull dimension four, using the Buchsbaum--Eisenbud structure theory of perfect Gorenstein ideals of height three and work by Kustin and Ulrich \cite{KU}, we establish the equality $\Ass(A/I^2)=\Ass(I/I^2)$ for all $I\subseteq A$ (Theorem \ref{thm_dim4_nequals2}). Another motivation for this work is the so-called persistence property for associated primes. The ideal $I$ is said to have the \emph{persistence property} if $\Ass(A/I^n)\subseteq \Ass(A/I^{n+1})$ for all $n\ge 1$. Ideals with this property abound, including for example complete intersections. The persistence property has been considered by many people; see, e.g., \cite{FHV, HQu, KSS, SW}. As an application of Theorem \ref{thm_main1}, we prove that if $I$ is a monomial ideal satisfying the persistence property, and $J$ is any ideal, then $I+J$ also has the persistence property (Corollary \ref{cor_persistence}). Moreover, we generalize previous work due to I. Swanson and R. Walker \cite{SW} on this question: If $I$ is an ideal such that $I^{n+1}:I=I^n$ for all $n\ge 1$, then for any ideal $J$ of $B$, $I+J$ has the persistence property (see Corollary \ref{cor_persistence}(ii)). In \cite[Corollary 1.7]{SW}, Swanson and Walker prove the same result under the stronger condition that $I$ is normal. It remains an open question whether for any ideal $I$ of $A$ with the persistence property and any ideal $J$ of $B$, the sum $I+J$ has same property. The paper is structured as follows. In Section \ref{sect_assofquotient}, we provide large classes of ideals $I$ such that the equality $\Ass(A/I^n)=\Ass(I^{n-1}/I^n)$ holds true for all $n\ge 1$. An unexpected outcome of this study is a counterexample to \cite[Question 3.6]{AM}, on the vanishing of the map $\Tor^A_i(k,I^n)\to \Tor^A_i(k,I^{n-1})$. Namely in characteristic 2, we construct a quadratic ideal $I$ in $A$ such that the natural map $\Tor^A_*(k,I^2)\to \Tor^A_*(k,I)$ is not zero (even though $A/I$ is a Gorenstein Artinian ring, see Example \ref{ex_non-Torvanishing}). This example might be of independent interest, for example, it gives a negative answer to a question in \cite{AM}. Using the results in Section \ref{sect_assofquotient}, we give a set-theoretic upper bound and a lower bound for $\Ass(R/(I+J)^n)$ (Theorem \ref{thm_Asscontainments}). Theorem \ref{thm_Asscontainments} also gives an exact formula for the asymptotic primes of $I+J$ without any condition on $I$ and $J$. In the last section, we apply our results to the question on the persistence property raised by Swanson and Walker. \section{Preliminaries} \label{sect_prelim} For standard notions and results in commutative algebra, we refer to the books \cite{BH, Eis}. Throughout the section, let $A$ and $B$ be two commutative Noetherian algebras over a field $k$ such that $R = A\otimes_k B$ is also Noetherian. Let $M$ and $N$ be two nonzero finitely generated modules over $A$ and $B$, respectively. Denote by $\Ass_A M$ and $\Min_A M$ the set of associated primes and minimal primes of $M$ as an $A$-module, respectively. By a filtration of ideals $(I_n)_{n\ge 0}$ in $A$, we mean the ideals $I_n, n\ge 0$ satisfies the conditions $I_0=A$ and $I_{n+1} \subseteq I_n$ for all $n\ge 0$. Let $(I_n)_{n\ge 1}$ and $(J_n)_{n\ge 1}$ be filtrations of ideals in $A$ and $B$, respectively. Consider the filtration $(W_n)_{n\ge 0}$ of $A\otimes_k B$ given by $W_n= \sum_{i=0}^n I_iJ_{n-i}$. The following result is useful for the discussion in Section \ref{sect_asspr_powers}. \begin{prop} [{\cite[Lemma 3.1, Proposition 3.3]{HNTT}}] \label{prop_decomposition} For arbitrary ideals $I\subseteq A$ and $J\subseteq B$, we have $I\cap J=IJ$. Moreover with the above notation for filtrations, for any integer $n \ge 0$, there is an isomorphism $$\displaystyle W_n/W_{n+1} \cong \bigoplus_{i=0}^n\big(I_i/I_{i+1} \otimes_k J_{n-i}/J_{n-i+1}\big).$$ \end{prop} We recall the following description of the associated primes of tensor products; see also \cite[Corollary 3.7]{STY}. \begin{thm}[{\cite[Theorem 2.5]{HNTT}}] \label{thm_asso} Let $M$ and $N$ be nonzero finitely generated modules over $A$ and $B$, respectively. Then there is an equality \[ \Ass_R(M \otimes_k N) = \displaystyle \bigcup_{\begin{subarray}{l} {\mathfrak p}\in \Ass_A(M)\\ {\mathfrak q} \in \Ass_B(N)\end{subarray}} \Min_R(R/{\mathfrak p}+{\mathfrak q}). \] \end{thm} The following simple lemma turns out to be useful in the sequel. \begin{lem} \label{lem_colon_containment} Assume that $\chara k=0$. Let $A=k[x_1,\ldots,x_r]$ be a standard graded polynomial ring over $k$, and ${\mathfrak m}$ its graded maximal ideal. Let $I$ be proper homogeneous ideal of $A$. Denote by $\partial(I)$ the ideal generated by partial derivatives of elements in $I$. Then there is a containment $I:{\mathfrak m} \subseteq \partial(I).$ In particular, $I^n:{\mathfrak m} \subseteq I^{n-1}$ for all $n\ge 1$. If for some $n\ge 2$, ${\mathfrak m}\in \Ass(A/I^n)$ then ${\mathfrak m} \in \Ass(I^{n-1}/I^n)$. \end{lem} \begin{proof} Take $f\in I:{\mathfrak m}$. Then $x_if\in I$ for every $i=1,\ldots,r$. Taking partial derivatives, we get $f+x_i(\partial f/\partial x_i) \in \partial(I)$. Summing up and using Euler's formula, $(r+\deg f)f\in \partial(I)$. As $\chara k=0$, this yields $f\in \partial(I)$, as claimed. The second assertion holds since by the product rule, $\partial(I^n)\subseteq I^{n-1}$. If ${\mathfrak m} \in \Ass(A/I^n)$ then there exists an element $a\in (I^n:{\mathfrak m})\setminus I^n$. Thus $a\in I^{n-1}\setminus I^n$, so ${\mathfrak m} \in \Ass(I^{n-1}/I^n)$. \end{proof} The condition on the characteristic is indispensable: The inclusion $I^2:{\mathfrak m} \subseteq I$ may fail in positive characteristic; see Example \ref{ex_non-Torvanishing}. The following lemma will be employed several times in the sequel. Denote by $\gr_I(A)$ the associated graded ring of $A$ with respect to the $I$-adic filtration. \begin{lem} \label{lem_sufficient_forpersistence} Let $A$ be a Noetherian ring, and $I$ an ideal. Then the following are equivalent: \begin{enumerate}[\quad \rm (i)] \item $I^{n+1}:I=I^n$ for all $n\ge 1$, \item $(I^{n+1}:I)\cap I^{n-1}=I^n$ for all $n\ge 1$, \item $\depth \gr_I(A)>0$, \item $I^n=\widetilde{I^n}$ for all $n\ge 1$, where $\widetilde{I}=\bigcup\limits_{i\ge 1}(I^{i+1}:I^i)$ denotes the Ratliff-Rush closure of $I$. \end{enumerate} If one of these equivalent conditions holds, then $\Ass(A/I^n)\subseteq \Ass(A/I^{n+1})$ for all $n\ge 1$, namely $I$ has the \textup{persistence property}. \end{lem} \begin{proof} Clearly (i) $\Longrightarrow$ (ii). We prove that (ii) $\Longrightarrow$ (i). Assume that $(I^{n+1}:I)\cap I^{n-1}=I^n$ for all $n\ge 1$. We prove by induction on $n\ge 1$ that $I^n:I=I^{n-1}$. If $n=1$, there is nothing to do. Assume that $n\ge 2$. By the induction hypothesis, $I^n:I \subseteq I^{n-1}:I=I^{n-2}$. Hence $I^n:I=(I^n:I)\cap I^{n-2}=I^{n-1}$, as desired. That (i) $\Longleftrightarrow$ (iii) $\Longleftrightarrow$ (iv) follows from \cite[(1.2)]{HLS} and \cite[Remark 1.6]{RS}. The last assertion follows from \cite[Section 1]{HQu}, where the property $I^{n+1}:I=I^n$ for all $n\ge 1$, called the \emph{strong persistence property} of $I$, was discussed. \end{proof} \section{Associated primes of quotients of consecutive powers} \label{sect_assofquotient} The following question is quite relevant to the task of finding the associated primes of powers of sums. \begin{quest} \label{quest_ass} Let $A$ be a standard graded polynomial ring over a field $k$ (of characteristic zero), and $I$ a proper homogeneous ideal. Is it true that \[ \Ass(A/I^n)=\Ass(I^{n-1}/I^n) \quad \text{for all $n\ge 1$?} \] \end{quest} We are not aware of any ideal not satisfying the equality in Question \ref{quest_ass} (even in positive characteristic). In the first main result of this paper, we provide some evidence for a positive answer to Question \ref{quest_ass}. Denote by $\Rees(I)$ the Rees algebra of $I$. The ideal $I$ is said to be \emph{unmixed} if it has no embedded primes. It is called \emph{normal} if all of its powers are integrally closed ideals. \begin{thm} \label{thm_specialcase_ass} Question \ref{quest_ass} has a positive answer if any of the following conditions holds: \begin{enumerate}[\quad \rm(1)] \item $I$ is a monomial ideal. \item $\depth \gr_I(A) \ge 1$. \item $\depth \Rees(I)\ge 2$. \item $I$ is normal. \item $I^n$ is unmixed for all $n\ge 1$, e.g. $I$ is generated by a regular sequence. \item All the powers of $I$ are primary, e.g. $\dim(A/I)=0$. \item $\chara k=0$ and $\dim(A/I)\le 1$. \item $\chara k=0$ and $\dim A\le 3$. \end{enumerate} \end{thm} \begin{proof} (1): See \cite[Lemma 4.4]{MV}. (2): By Lemma \ref{lem_sufficient_forpersistence}, since $\depth \gr_I(A)\ge 1$, $I^n:I=I^{n-1}$ for all $n\ge 1$. Induce on $n\ge 1$ that $\Ass(A/I^n)=\Ass(I^{n-1}/I^n)$. Let $I=(f_1,\ldots,f_m)$. For $n\ge 2$, as $I^n:I=I^{n-1}$, the map $$ I^{n-2} \to \underbrace{I^{n-1}\oplus \cdots \oplus I^{n-1}}_{m \, \text{times}}, a\mapsto (af_1,\ldots,af_m), $$ induces an injection \[ \frac{I^{n-2}}{I^{n-1}} \hookrightarrow \left(\frac{I^{n-1}}{I^n}\right)^{\oplus m}. \] Hence $\Ass(A/I^{n-1})=\Ass(I^{n-2}/I^{n-1}) \subseteq \Ass(I^{n-1}/I^n)$. The exact sequence \[ 0\to I^{n-1}/I^n \to A/I^n \to A/I^{n-1} \to 0 \] then yields $\Ass(A/I^n)\subseteq \Ass(I^{n-1}/I^n)$, which in turn implies the desired equality. Next we claim that (3) and (4) all imply (2). (3) $\Longrightarrow$ (2): This follows from a result of Huckaba and Marley \cite[Corollary 3.12]{HuMa} which says that either $\gr_I(A)$ is Cohen-Macaulay (and hence has depth $A= \dim A$), or $\depth \gr_I(A)=\depth \Rees(I)-1$. (4) $\Longrightarrow$ (2): If $I$ is normal, then $I^n:I=I^{n-1}$ for all $n\ge 1$. Hence we are done by Lemma \ref{lem_sufficient_forpersistence}. (5): Take $P\in \Ass(A/I^n)$, we show that $P\in \Ass(I^{n-1}/I^n)$. Since $A/I^n$ is unmixed, $P \in \Min(A/I^n)= \Min(I^{n-1}/I^n)$. Observe that (6) $\Longrightarrow$ (5). (7): Because of (6), we can assume that $\dim(A/I)=1$. Take $P\in \Ass(A/I^n)$, we need to show that $P\in \Ass(I^{n-1}/I^n)$. If $\dim(A/P)=1$, then as $\dim(A/I)=1$, $P\in \Min(A/I^n)$. Arguing as for case (5), we get $P\in \Ass(I^{n-1}/I^n)$. If $\dim(A/P)=0$, then $P={\mathfrak m}$, the graded maximal ideal of $A$. Since ${\mathfrak m} \in \Ass(A/I^n)$, by Lemma \ref{lem_colon_containment}, ${\mathfrak m} \in \Ass(I^{n-1}/I^n)$. (8) It is harmless to assume that $I\neq 0$. If $\dim(A/I)\le 1$ then we are done by (7). Assume that $\dim(A/I)\ge 2$, then the hypothesis forces $\dim A=3$ and $\htt I=1$. Thus we can write $I=xL$ where $x$ is a form of degree at least 1, and $L=R$ or $\htt L\ge 2$. The result is clear when $L=R$, so it remains to assume that $L$ is proper of height $\ge 2$. In particular $\dim(A/L)\le 1$, and by (7), for all $n\ge 1$, \[ \Ass(A/L^n)=\Ass(L^{n-1}/L^n). \] Take ${\mathfrak p} \in \Ass(A/I^n)$. Since $A/I^n$ and $I^{n-1}/I^n$ have the same minimal primes, we can assume $\htt {\mathfrak p} \ge 2$. From the exact sequence \[ 0\to A/L^n \xrightarrow{\cdot x^n} A/I^n \to A/(x^n) \to 0 \] it follows that ${\mathfrak p} \in \Ass(A/L^n)$. Thus ${\mathfrak p} \in \Ass(L^{n-1}/L^n)$. There is an exact sequence \[ 0\to L^{n-1}/L^n \xrightarrow{\cdot x^n} I^{n-1}/I^n \] so ${\mathfrak p} \in \Ass(I^{n-1}/I^n)$, as claimed. This concludes the proof. \end{proof} \begin{ex} Here is an example of a ring $A$ and an ideal $I$ not satisfying any of the conditions (1)--(8) in Theorem \ref{thm_specialcase_ass}. Let $I=(x^4+y^3z,x^3y,x^2t^2,y^4,y^2z^2) \subseteq A=k[x,y,z,t]$. Then $\depth \gr_I(A)=0$ as $x^2y^3z\in (I^2:I)\setminus I$. So $I$ satisfies neither (1) nor (2). Note that $\sqrt{I}=(x,y)$, so $\dim(A/I)=2$. Let ${\mathfrak m}=(x,y,z,t)$. Since $x^2y^3zt\in (I:{\mathfrak m})\setminus I$, $\depth(A/I)=0$, hence $A/I$ is not unmixed. Thus $I$ satisfies neither (5) nor (7). By the proof of Theorem \ref{thm_specialcase_ass}, $I$ satisfies none of the conditions (3), (4), (6). Unfortunately, experiments with Macaulay2 \cite{GS} suggest that $I$ satisfies the conclusion of Question \ref{quest_ass}, namely for all $n\ge 1$, \[ \Ass(A/I^n)=\Ass(I^{n-1}/I^n)=\{(x,y),(x,y,z),(x,y,t),(x,y,z,t)\}. \] \end{ex} \begin{rem} In view of Lemma \ref{lem_colon_containment} and Question \ref{quest_ass}, one might ask whether if $\chara k=0$, then $\Ass(A/I)=\Ass(\partial(I)/I)$ for any homogeneous ideal $I$ in a polynomial ring $A$? Unfortunately, this has a negative answer. Let $A=\mathbb{Q}[x,y,z], f=x^5+x^4y+y^4z, L=(x,y)$ and $I=fL$. Then we can check with Macaulay2 \cite{GS} that $\partial(I):f=L$. In particular, \[ \Ass (\partial(I)/I) =(f) \neq \Ass(A/I)=\{(f),(x,y)\}. \] Indeed, if $L=(x,y)\in \Ass(\partial(I)/I)$ then $\Hom_R(R/L,\partial(I)/I)=(\partial(I)\cap (I:L))/I=(\partial(I)\cap (f))/I\neq 0$, so that $\partial(I):f\neq L$, a contradiction. \end{rem} \subsection{Partial answer to Question \ref{quest_ass} in dimension four} We prove that if $\chara k=0$ and $\dim A=4$, the equality $\Ass(A/I^2)=\Ass(I/I^2)$ always holds, in support of a positive answer to Question \ref{quest_ass}. The proof requires the structure theory of perfect Gorenstein ideals of height three and their second powers. \begin{thm} \label{thm_dim4_nequals2} Assume $\chara k=0$. Let $(A,{\mathfrak m})$ be a four dimensional standard graded polynomial ring over $k$. Then for any proper homogeneous ideal $I$ of $A$, there is an equality $\Ass(A/I^2)=\Ass(I/I^2)$. \end{thm} \begin{proof} It is harmless to assume $I$ is a proper ideal. If $\htt I\ge 3$ then $\dim(A/I)\le 1$, and we are done by Theorem \ref{thm_specialcase_ass}(7). If $\htt I=1$, then $I=fL$, where $f\in A$ is a form of positive degree and $\htt L\ge 2$. The exact sequence \[ 0 \to \frac{A}{L} \xrightarrow{\cdot f} \frac{A}{I} \to \frac{A}{(f)} \to 0, \] yields $\Ass(A/I)=\Ass(A/L) \bigcup \Ass(A/(f))$, as $\Min(I) \supseteq \Ass(A/(f))$. An analogous formula holds for $\Ass(A/I^2)$, as $I^2=f^2L^2$. If we can show that $\Ass(A/L^2) \subseteq \Ass(L/L^2)$, then from the injection $L/L^2 \xrightarrow{\cdot f^2} I/I^2$ we have \begin{align*} \Ass(A/I^2) &=\Ass(A/L^2) \bigcup \Ass(A/(f)) \\ & =\Ass(L/L^2)\bigcup \Ass(A/(f)) \subseteq \Ass(I/I^2). \end{align*} Hence it suffices to consider the case $\htt I=2$. Assume that there exists ${\mathfrak p} \in \Ass(A/I^2) \setminus \Ass(I/I^2)$. The exact sequence \[ 0\to I/I^2 \to A/I^2 \to A/I \to 0 \] implies ${\mathfrak p} \in \Ass(A/I)$. By Lemma \ref{lem_colon_containment}, ${\mathfrak p} \neq {\mathfrak m}$. Since $\Min(I) = \Min(I/I^2)$, ${\mathfrak p} \notin \Min(I)$, we get $\htt {\mathfrak p}=3$. Localizing yields ${\mathfrak p} A_{\mathfrak p} \in \Ass(A_{\mathfrak p}/I_{\mathfrak p}^2) \setminus \Ass(I_{\mathfrak p}/I_{\mathfrak p}^2)$. Then there exists $a\in (I_{\mathfrak p}^2:{\mathfrak p} A_{\mathfrak p}) \setminus I_{\mathfrak p}^2$. On the other hand, since $A_{\mathfrak p}$ is a regular local ring of dimension 3 containing one half, Lemma \ref{lem_colon_dim3} below implies $I_{\mathfrak p}^2:{\mathfrak p} A_{\mathfrak p} \subseteq I_{\mathfrak p}$, so $a \in I_{\mathfrak p} \setminus I_{\mathfrak p}^2$. Hence ${\mathfrak p} A_{\mathfrak p} \in \Ass(I_{\mathfrak p}/I_{\mathfrak p}^2)$. This contradiction finishes the proof. \end{proof} To finish the proof of Theorem \ref{thm_dim4_nequals2}, we have to show the following \begin{lem} \label{lem_colon_dim3} Let $(R,{\mathfrak m})$ be a three dimensional regular local ring such that $1/2 \in R$. Then for any ideal $I$ of $R$, there is a containment $I^2:{\mathfrak m} \subseteq I$. \end{lem} We will deduce it from the following result. \begin{prop} \label{prop_Gorenstein_height3_Torvanishing} Let $(R,{\mathfrak m})$ be a regular local ring such that $1/2\in R$. Let $J$ be a perfect Gorenstein ideal of height $3$. Then for all $i\ge 0$, the maps \[ \Tor^R_i(J^2,k) \to \Tor^R_i(J,k) \] is zero. In particular, there is a containment $J^2:{\mathfrak m} \subseteq J$. \end{prop} \begin{proof} We note that the second assertion follows from the first. Indeed, the hypotheses implies that $\dim(R)=d\ge 3$. Using the Koszul complex of $R$, we see that \[ \Tor^R_{d-1}(J,k) \cong \Tor^R_d(R/J,k) \cong \frac{J:{\mathfrak m}}{J}. \] Since the map $\Tor^R_i(J^2,k) \to \Tor^R_i(J,k)$ is zero for $i=d-1$, the conclusion is $J^2:{\mathfrak m} \subseteq J$. Hence it remains to prove the first assertion. We do this by exploiting the structure of the minimal free resolution of $J$ and $J^2$, and constructing a map between these complexes. Since $J$ is Gorenstein of height three, it has a minimal free resolution \[ P: 0\to R \xrightarrow{\delta} F^* \xrightarrow{\rho} F \to 0. \] Here $F=Re_1 \oplus \cdots \oplus Re_g$ is a free $R$-module of rank $g$ -- an odd integer. The map $\tau: F\to J$ maps $e_i$ to $f_i$, where $J=(f_1,\ldots,f_g)$. The free $R$-module $F^*$ has basis $e_1^*,\ldots,e_g^*$. The map $\rho: F^* \to F$ is alternating with matrix $(a_{i,j})_{g\times g}$, namely $a_{i,i}=0$ for $1\le i\le g$ and $a_{i,j}=-a_{j,i}$ for $1\le i<j\le g$, and \[ \rho(e_i^*)=\sum_{j=1}^g a_{j,i}e_j \quad \text{for all $i$}. \] The map $\delta: R\to F^*$ has the matrix $(f_1 \ldots f_g)^T$, i.e. it is given by $\delta(1)=f_1e_1^*+\cdots+f_g e_g^*$. It is known that if $J$ is Gorenstein of height three, then $J\otimes_R J \cong J^2$, and by constructions due to Kustin and Ulrich \cite[Definition 5.9, Theorems 6.2 and 6.17]{KU}, $J^2$ has a minimal free resolution $Q$ as below. Note that in the terminology of \cite{KU} and thanks to the discussion after Theorem 6.22 in that work, $J$ satisfies $\text{SPC}_{g-2}$, hence Theorem 6.17, parts (c)(i) and (d)(ii) in \emph{ibid.} are applicable. The resolution $Q$ given in the following is taken from (2.7) and Definition 2.15 in Kustin and Ulrich's paper. \[ Q: 0 \to \wedge^2 F^* \xrightarrow{d_2} (F\otimes F^*)/\eta \xrightarrow{d_1} S_2(F) \xrightarrow{d_0} J^2 \to 0. \] Here $S_2(F)=\bigoplus_{1\le i\le j\le g} R(e_i\otimes e_j)$ is the second symmetric power of $F$, $\eta=R(e_1\otimes e_1^*+\cdots+e_g\otimes e_g^*) \subseteq F\otimes F^*$, and $\wedge^2 F^*$ is the second exterior power of $F^*$. The maps $d_0: S_2(F)\to J^2$, $d_1: (F\otimes F^*)/\eta \to S_2(F)$, $d_2: \wedge^2 F^* \to (F\otimes F^*)/\eta$ are given by: \begin{align*} d_0(e_i\otimes e_j) &=f_if_j \quad \text{for $1\le i,j\le g$},\\ d_1(e_i\otimes e_j^*+\eta) &=\sum_{l=1}^g a_{l,j} (e_i\otimes e_l) \quad \text{for $1\le i,j\le g$},\\ d_2(e_i^*\wedge e_j^*) &= \sum_{l=1}^g a_{l,i}(e_l\otimes e_j^*) - \sum_{v=1}^g a_{v,j}(e_v\otimes e_i^*)+\eta \quad \text{for $1\le i<j\le g$}. \end{align*} We construct a lifting $\alpha: Q\to P$ of the natural inclusion map $J^2\to J$ such that $\alpha(Q)\subseteq {\mathfrak m} P$. \[ \xymatrix{ Q: & 0 \ar[r] & \wedge^2 F^* \ar[d]^{\alpha_2} \ar[r]^{d_2} & (F\otimes F^*)/\eta \ar[r]^{d_1} \ar[d]^{\alpha_1} & S_2(F) \ar[r]^{d_0} \ar[d]^{\alpha_0} & J^2 \ar[r] \ar@{^{(}->}[d]_{\iota} & 0 \\ P: & 0 \ar[r] & R \ar[r]^{\delta} & F^* \ar[r]^{\rho} & F \ar[r]^{\tau} & J \ar[r] & 0. } \] In detail, this lifting is \begin{itemize} \item $\alpha_0(e_i\otimes e_j) = \dfrac{f_ie_j+f_je_i}{2} \quad \text{for $1\le i,j\le g$},$ \item $\alpha_1(e_i\otimes e_j^*+\eta) = \begin{cases} \dfrac{f_ie_j^*}{2}, & \text{if $(i,j) \neq (g,g)$},\\ \dfrac{-\sum_{v=1}^{g-1}f_v e_v^*}{2}, &\text{if $(i,j)=(g,g)$}, \end{cases}$ \item $\alpha_2(e_i^*\wedge e_j^*) = \begin{cases} 0, &\text{if $1\le i<j\le g-1$},\\ \dfrac{-a_{g,i}}{2}, &\text{if $1\le i\le g-1, j=g$}. \end{cases}$ \end{itemize} Note that $\alpha_1$ is well-defined since $$ \alpha_1(e_1\otimes e_1^*+\cdots+e_g\otimes e_g^*+\eta) = \dfrac{\sum_{v=1}^{g-1}f_v e_v^*}{2}\dfrac{-\sum_{v=1}^{g-1}f_v e_v^*}{2}=0. $$ Observe that $\alpha(Q)\subseteq {\mathfrak m} P$ since $f_i,a_{i,j} \in {\mathfrak m}$ for all $i,j$. It remains to check that the map $\alpha: Q\to P$ is a lifting for $J^2\hookrightarrow J$. For this, we have: \begin{itemize} \item $\tau(\alpha_0(e_i\otimes e_j))=\tau\left(\dfrac{f_ie_j+f_je_i}{2}\right)=f_if_j=\iota(d_0(e_i\otimes e_j))$. \end{itemize} Next we compute \begin{align*} \alpha_0(d_1(e_i\otimes e_j^*+\eta)) &=\alpha_0\left(\sum_{l=1}^g a_{l,j} (e_i\otimes e_l)\right)= \sum_{l=1}^g a_{l,j}\dfrac{f_ie_l+f_le_i}{2} \\ &= \dfrac{f_i\left(\sum_{l=1}^ga_{l,j}e_l\right)}{2} \quad \text{(since $\sum_{l=1}^g a_{l,j}f_l=0$)}. \end{align*} \begin{itemize} \item If $(i,j)\neq (g,g)$ then \begin{align*} \rho(\alpha_1(e_i\otimes e_j^*+\eta)) &= \rho(f_ie_j^*/2)= \dfrac{f_i\left(\sum_{l=1}^ga_{l,j}e_l\right)}{2}. \end{align*} \item If $(i,j)=(g,g)$ then \begin{align*} \rho(\alpha_1(e_g\otimes e_g^*+\eta)) &= \rho\left(\dfrac{-\sum_{v=1}^{g-1}f_ve_v^*}{2}\right)= \dfrac{-\sum_{v=1}^{g-1}f_v(\sum_{l=1}^g a_{l,v}e_l)}{2}\\ &= \dfrac{\sum\limits_{l=1}^g(\sum\limits_{v=1}^{g-1} a_{v,l}f_v)e_l}{2} \quad (\text{since $a_{v,l}=-a_{l,v}$})\\ &= \dfrac{-\sum\limits_{l=1}^g(a_{g,l}f_g)e_l}{2} \quad \text{(since $\sum_{v=1}^g a_{v,l}f_v=0$)}\\ &= \dfrac{f_g\left(\sum_{l=1}^ga_{l,g}e_l\right)}{2} \quad (\text{since $a_{g,l}=-a_{l,g}$}) \end{align*} \item Hence in both cases, $\alpha_0(d_1(e_i\otimes e_j^*+\eta))=\rho(\alpha_1(e_i\otimes e_j^*+\eta))$. \end{itemize} Next, for $1\le i< j \le g-1$, we compute \begin{align*} \alpha_1(d_2(e_i^*\wedge e_j^*)) &= \alpha_1\left(\sum_{l=1}^g a_{l,i}(e_l\otimes e_j^*) - \sum_{v=1}^g a_{v,j}(e_v\otimes e_i^*)+\eta\right)\\ &= \dfrac{\left(\sum_{l=1}^g a_{l,i}f_l\right)e_j^*}{2}-\dfrac{\left(\sum_{v=1}^g a_{v,j}f_v\right)e_i^*}{2}\\ & \qquad \text{(since neither $(l,j)$ nor $(v,i)$ is $(g,g)$)}\\ &=0 \quad \text{(since $\sum_{v=1}^g a_{v,l}f_v=0$)}\\ & =\delta(\alpha_2(e_i^*\wedge e_j^*)). \end{align*} Finally, for $1\le i\le g-1, j=g$, we have \begin{align*} \alpha_1(d_2(e_i^*\wedge e_g^*)) &= \alpha_1\left(\sum_{l=1}^g a_{l,i}(e_l\otimes e_g^*) - \sum_{v=1}^g a_{v,g}(e_v\otimes e_i^*)+\eta\right)\\ &= \dfrac{\left(\sum_{l=1}^{g-1} a_{l,i}f_l\right)e_g^*}{2}-\dfrac{\sum_{v=1}^{g-1}a_{g,i}f_v e_v^*}{2} -\dfrac{\left(\sum_{v=1}^g a_{v,g}f_v\right)e_i^*}{2}\\ & \text{\footnotesize{(the formula for $\alpha_1(e_l\otimes e_g^*)$ depends on whether $l=g$ or not)}} \\ &= \dfrac{-a_{g,i}f_ge_g^*}{2}-\dfrac{\sum_{v=1}^{g-1}a_{g,i}f_v e_v^*}{2} \quad \text{(since $\sum_{v=1}^g a_{v,l}f_v=0$)} \\ &= \dfrac{-a_{g,i}\left(\sum_{v=1}^gf_v e_v^*\right)}{2} \end{align*} We also have \[ \delta(\alpha_2(e_i^*\wedge e_g^*))=\delta(-a_{g,i}/2)=\dfrac{-a_{g,i}\left(\sum_{v=1}^gf_v e_v^*\right)}{2}. \] Hence $\alpha: Q\to P$ is a lifting of the inclusion map $J^2\to J$. Since $\alpha(Q) \subseteq {\mathfrak m} P$, it follows that $\alpha\otimes (R/{\mathfrak m})=0$. Hence $\Tor^R_i(J^2,k) \to \Tor^R_i(J,k)$ is the zero map for all $i$. The proof is concluded. \end{proof} \begin{proof}[Proof of Lemma \ref{lem_colon_dim3}] It is harmless to assume that $I\subseteq {\mathfrak m}$. We can write $I$ as a finite intersection $I_1\cap \cdots \cap I_d$ of irreducible ideals. If we can show the lemma for each of the components $I_j$, then \[ I^2:{\mathfrak m} \subseteq (I_1^2:{\mathfrak m}) \cap \cdots \cap (I_d^2:{\mathfrak m}) \subseteq \bigcap_{j=1}^d I_j=I. \] Hence we can assume that $I$ is an irreducible ideal. Being irreducible, $I$ is a primary ideal. If $\sqrt{I} \neq {\mathfrak m}$, then $I^2:{\mathfrak m}\subseteq I:{\mathfrak m}=I$. Therefore we assume that $I$ is an ${\mathfrak m}$-primary irreducible ideal. Let $k=R/{\mathfrak m}$. It is a folklore and simple result that any ${\mathfrak m}$-primary irreducible ideal must satisfy $\dim_k (I:{\mathfrak m})/I=1$. Note that $R$ is a regular local ring, so being a Cohen-Macaulay module of dimension zero, $R/I$ is perfect. Hence $I$ is a perfect Gorenstein ideal of height three. It then remains to use the second assertion of Proposition \ref{prop_Gorenstein_height3_Torvanishing}. \end{proof} In view of Lemma \ref{lem_colon_dim3}, it seems natural to ask the following \begin{quest} \label{quest_colon_dim3} Let $(R,{\mathfrak m})$ be a three dimensional regular local ring containing $1/2$. Let $I$ be an ideal of $R$. Is it true that for all $n\ge 2$, $I^n:{\mathfrak m} \subseteq I^{n-1}$? \end{quest} For regular local rings of dimension at most two, Ahangari Maleki has proved that Question \ref{quest_colon_dim3} has a positive answer regardless of the characteristic \cite[Proof of Theorem 3.7]{AM}. Nevertheless, if $\dim A$ is not fixed, Question \ref{quest_colon_dim3} has a negative answer in positive characteristic in general. Here is a counterexample in dimension 9(!). \begin{ex} \label{ex_non-Torvanishing} Choose $\chara k=2$, $A=k[x_1,x_2,x_3,\ldots,z_1,z_2,z_3]$ and \[ M=\begin{pmatrix} x_1 & x_2 & x_3 \\ y_1 & y_2 & y_3 \\ z_1 & z_2 & z_3 \end{pmatrix}. \] Let $I_2(M)$ be the ideal generated by the 2-minors of $M$, and $$ I=I_2(M)+\sum_{i=1}^3(x_i,y_i,z_i)^2+(x_1,x_2,x_3)^2+(y_1,y_2,y_3)^2+(z_1,z_2,z_3)^2. $$ Denote ${\mathfrak m}=A_+$. The Betti table of $A/I$, computed by Macaulay2 \cite{GS}, is \begin{verbatim} 0 1 2 3 4 5 6 7 8 9 total: 1 36 160 315 404 404 315 160 36 1 0: 1 . . . . . . . . . 1: . 36 160 315 288 116 . . . . 2: . . . . 116 288 315 160 36 . 3: . . . . . . . . . 1 \end{verbatim} Therefore $I$ is an ${\mathfrak m}$-primary, binomial, quadratic, Gorenstein ideal. Also, the relation $x_1y_2z_3+x_2y_3z_1+x_3y_1z_2\in (I^2:{\mathfrak m})\setminus I$ implies $I^2:{\mathfrak m} \not\subseteq I$. This means that the map $\Tor^A_8(k,I^2)\to \Tor^A_8(k,I)$ is not zero. In particular, this gives a negative answer to \cite[Question 3.6]{AM} in positive characteristic. \end{ex} \section{Powers of sums and associated primes} \label{sect_asspr_powers} \subsection{Bounds for associated primes} The second main result of this paper is the following. Its part (3) generalizes \cite[Lemma 3.4]{HM}, which deals only with squarefree monomial ideals. \begin{thm} \label{thm_Asscontainments} Let $A, B$ be commutative Noetherian algebras over $k$ such that $R=A\otimes_k B$ is Noetherian. Let $I,J$ be proper ideals of $A,B$, respectively. \begin{enumerate}[\quad \rm(1)] \item For all $n\ge 1$, we have inclusions \begin{align*} \bigcup_{i=1}^n \mathop{\bigcup_{{\mathfrak p} \in \Ass_A(I^{i-1}/I^i)}}_{{\mathfrak q} \in \Ass_B(J^{n-i}/J^{n-i+1})} \Min_R(R/{\mathfrak p}+{\mathfrak q}) & \subseteq \Ass_R \frac{R}{(I+J)^n},\\ \Ass_R \frac{R}{(I+J)^n} & \subseteq \bigcup_{i=1}^n \mathop{\bigcup_{{\mathfrak p} \in \Ass_A(A/I^i)}}_{{\mathfrak q} \in \Ass_B(J^{n-i}/J^{n-i+1})} \Min_R(R/{\mathfrak p}+{\mathfrak q}). \end{align*} \item If moreover $\Ass(A/I^n)=\Ass(I^{n-1}/I^n)$ for all $n\ge 1$, then both inclusions in \textup{(1)} are equalities. \item In particular, if $A$ and $B$ are polynomial rings and $I$ and $J$ are monomial ideals, then for all $n\ge 1$, we have an equality \[ \Ass_R \frac{R}{(I+J)^n} = \bigcup_{i=1}^n \mathop{\bigcup_{{\mathfrak p} \in \Ass_A(A/I^i)}}_{{\mathfrak q} \in \Ass_B(B/J^{n-i+1})} \{{\mathfrak p}+{\mathfrak q}\}. \] \end{enumerate} \end{thm} \begin{proof} (1) Denote $Q=I+J$. By Proposition \ref{prop_decomposition}, we have \[ Q^{n-1}/Q^n = \bigoplus_{i=1}^n (I^{i-1}/I^i \otimes_k J^{n-i}/J^{n-i+1}). \] Hence \begin{equation} \label{eq_inclusion_Ass1} \bigcup_{i=1}^n \Ass_R (I^{i-1}/I^i \otimes_k J^{n-i}/J^{n-i+1}) = \Ass_R (Q^{n-1}/Q^n) \subseteq \Ass_R(R/Q^n). \end{equation} For each $1\le i\le n$, we have $J^{n-i}Q^i\subseteq J^{n-i}(I^i+J)=J^{n-i}I^i+J^{n-i+1}$. We claim that $(J^{n-i}I^i+J^{n-i+1})/J^{n-i}Q^i \cong J^{n-i+1}/J^{n-i+1}Q^{i-1}$, so that there is an exact sequence \begin{equation} \label{eq_exactseq} 0\longrightarrow \frac{J^{n-i+1}}{J^{n-i+1}Q^{i-1}} \longrightarrow \frac{J^{n-i}}{J^{n-i}Q^i} \longrightarrow \frac{J^{n-i}}{J^{n-i+1}+J^{n-i}I^i} \cong \frac{A}{I^{i}}\otimes_k \frac{J^{n-i}}{J^{n-i+1}} \longrightarrow 0. \end{equation} For the claim, we have \begin{align*} (J^{n-i}I^i+J^{n-i+1})/J^{n-i}Q^i &= \frac{J^{n-i}I^i+J^{n-i+1}}{J^{n-i}(I^i+JQ^{i-1})}= \frac{J^{n-i}I^i+J^{n-i+1}}{J^{n-i}I^i+J^{n-i+1}Q^{i-1}}\\ &= \frac{(J^{n-i}I^i+J^{n-i+1})/J^{n-i}I^i}{(J^{n-i}I^i+J^{n-i+1}Q^{i-1})/J^{n-i}I^i}\\ &\cong \frac{J^{n-i+1}/J^{n-i+1}I^i}{J^{n-i+1}Q^{i-1}/J^{n-i+1}I^i} \cong \frac{J^{n-i+1}}{J^{n-i+1}Q^{i-1}}. \end{align*} In the display, the first isomorphism follows from the fact that \[ J^{n-i+1} \cap J^{n-i}I^i=J^{n-i+1}I^i=J^{n-i}I^i\cap J^{n-i+1}Q^{i-1}, \] which holds since by Proposition \ref{prop_decomposition}, \[ J^{n-i+1}I^i \subseteq J^{n-i}I^i\cap J^{n-i+1}Q^{i-1} \subseteq J^{n-i}I^i \cap J^{n-i+1} \subseteq I^i\cap J^{n-i+1}=J^{n-i+1}I^i. \] Now for $i=n$, the exact sequence \eqref{eq_exactseq} yields \[ \Ass_R (R/Q^n) \subseteq \Ass_R(J/JQ^{n-1}) \cup \Ass_R (A/I^n\otimes_k B/J). \] Similarly for the cases $2\le i\le n-1$ and $i=1$. Putting everything together, \begin{equation} \label{eq_inclusion_Ass2} \Ass_R (R/Q^n) \subseteq \bigcup_{i=1}^n \Ass_R (A/I^i \otimes_k J^{n-i}/J^{n-i+1}). \end{equation} Combining \eqref{eq_inclusion_Ass1}, \eqref{eq_inclusion_Ass2} and Theorem \ref{thm_asso}, we finish the proof of (1). (2) If $\Ass_A(A/I^n)=\Ass_A(I^{n-1}/I^n)$ for all $n\ge 1$, then clearly the upper bound and lower bound for $\Ass(R/(I+J)^n)$ in part (1) coincide. The conclusion follows. (3) In this situation, every associated prime of $A/I^i$ is generated by variables. In particular, ${\mathfrak p}+{\mathfrak q}$ is a prime ideal of $R$ for any ${\mathfrak p} \in \Ass(A/I^i), {\mathfrak q}\in \Ass_B(B/J^j)$ and $i, j\ge 1$. The conclusion follows from part (2). \end{proof} \begin{rem} If Question \ref{quest_ass} has a positive answer, then we can strengthen the conclusion of Theorem \ref{thm_Asscontainments}: Let $A, B$ be standard graded polynomial rings over $k$. Let $I,J$ be proper homogeneous ideals of $A,B$, respectively. Then for all $n\ge 1$, there is an equality \[ \Ass_R \frac{R}{(I+J)^n} = \bigcup_{i=1}^n \mathop{\bigcup_{{\mathfrak p} \in \Ass_A(A/I^i)}}_{{\mathfrak q} \in \Ass_B(B/J^{n-i+1})} \Min(R/({\mathfrak p}+{\mathfrak q})). \] \end{rem} \begin{ex} In general, for singular base rings, each of the inclusions of Theorem \ref{thm_Asscontainments} can be strict. First, take $A=k[a,b,c]/(a^2,ab,ac), I=(b)$, $B=k, J=(0)$. Then $R=A, Q=I=(b)$ and $I^2=(b^2)$. Let ${\mathfrak m}=(a,b,c)$. One can check that $a \in (I^2:{\mathfrak m})\setminus I^2$ and $I/I^2\cong A/(a,b)\cong k[c]$, whence $\depth(A/I^2)=0 <\depth(I/I^2)$. In particular, ${\mathfrak m}\in \Ass_A(A/I^2) \setminus \Ass_A(I/I^2)$. Thus the lower bound for $\Ass_R(R/Q^2)$ is strict in this case. Second, take $A,I$ as above and $B=k[x,y,z]$, $J=(x^4,x^3y,xy^3,y^4,x^2y^2z).$ In this case $Q=(b,x^4,x^3y,xy^3,y^4,x^2y^2z) \subseteq k[a,b,c,x,y,z]/(a^2,ab,ac)$. Then $c+z$ is $(R/Q^2)$-regular, so $\depth R/Q^2 >0=\depth A/I^2+\depth B/J$. Hence $(a,b,c,x,y,z)$ does not lie in $\Ass_R(R/Q^2)$, but it belongs to the upper bound for $\Ass_R(R/Q^2)$ in Theorem \ref{thm_Asscontainments}(1). \end{ex} \subsection{Asymptotic primes} Recall that if $I\neq A$, $\grade(I,A)$ denotes the maximal length of an $A$-regular sequence consisting of elements in $I$; and if $I=A$, by convention, $\grade(I,A)=\infty$ (see \cite[Section 1.2]{BH} for more details). Let $\astab^*(I)$ denote the minimal integer $m\ge 1$ such that both $\Ass_A(A/I^i)$ and $\Ass_A(I^{i-1}/I^i)$ are constant sets for all $i\ge m$. By a result due to McAdam and Eakin \cite[Corollary 13]{McE}, for all $i\ge \astab^*(I)$, $\Ass_A(A/I^i)\setminus \Ass_A(I^{i-1}/I^i)$ consists only of prime divisors of $(0)$. Hence if $\grade(I,A)\ge 1$, i.e. $I$ contains a non-zerodivisor, then $\Ass_A(A/I^i)= \Ass_A(I^{i-1}/I^i)$ for all $i\ge \astab^*(I)$. Denote $\Ass_A^*(I)=\bigcup_{i\ge 1} \Ass_A(A/I^i)=\bigcup_{i=1}^{\astab^*(I)}\Ass_A(A/I^i)$ and $$ \Ass_A^{\infty}(I)=\Ass_A(A/I^i) \quad \text{for any $i\ge \astab^*(I)$.} $$ The following folklore lemma will be useful. \begin{lem} \label{lem_unionAssequal} For any $n\ge 1$, we have \[ \bigcup_{i=1}^n \Ass_A(A/I^i)=\bigcup_{i=1}^n \Ass_A(I^{i-1}/I^i). \] In particular, if $\grade(I,A)\ge 1$ then $$\Ass_A^*(I)=\bigcup_{i=1}^{\astab^*(I)} \Ass_A(I^{i-1}/I^i)=\bigcup_{i\ge 1} \Ass_A(I^{i-1}/I^i).$$ \end{lem} \begin{proof} For the first assertion: Clearly the left-hand side contains the right-hand one. Conversely, we deduce from the inclusion $\Ass_A(A/I^i) \subseteq \Ass_A(I^{i-1}/I^i) \cup \Ass_A(A/I^{i-1})$ for $2\le i\le n$ that the other containment is valid as well. The remaining assertion is clear. \end{proof} Now we describe the asymptotic associated primes of $(I+J)^n$ for $n\gg 0$ and provide an upper bound for $\astab^*(I+J)$ under certain conditions on $I$ and $J$. \begin{thm} \label{thm_asymptoticAss} Assume that $\grade(I,A)\ge 1$ and $\grade(J,B)\ge 1$, e.g. $A$ and $B$ are domains and $I, J$ are proper ideals. Then for all $n\ge \astab^*(I)+\astab^*(J)-1$, we have \begin{align*} \Ass_R \frac{R}{(I+J)^n}&=\Ass_R \frac{(I+J)^{n-1}}{(I+J)^n} \\ &= \mathop{\bigcup_{{\mathfrak p} \in \Ass^*_A(I)}}_{{\mathfrak q} \in \Ass^{\infty}_B(J)} \Min_R(R/{\mathfrak p}+{\mathfrak q}) \bigcup \mathop{\bigcup_{{\mathfrak p} \in \Ass^{\infty}_A(I)}}_{{\mathfrak q} \in \Ass^*_B(J)} \Min_R(R/{\mathfrak p}+{\mathfrak q}). \end{align*} In particular, $\astab^*(I+J)\le \astab^*(I)+\astab^*(J)-1$ and \[ \Ass^{\infty}_R(I+J)= \mathop{\bigcup_{{\mathfrak p} \in \Ass^*_A(I)}}_{{\mathfrak q} \in \Ass^{\infty}_B(J)} \Min_R(R/{\mathfrak p}+{\mathfrak q}) \bigcup \mathop{\bigcup_{{\mathfrak p} \in \Ass^{\infty}_A(I)}}_{{\mathfrak q} \in \Ass^*_B(J)} \Min_R(R/{\mathfrak p}+{\mathfrak q}). \] \end{thm} \begin{proof} Denote $Q=I+J$. It suffices to prove that for $n\ge \astab^*(I)+\astab^*(J)-1$, both the lower bound (which is nothing but $\Ass_R(Q^{n-1}/Q^n)$) and upper bound for $\Ass_R(R/Q^n)$ in Theorem \ref{thm_Asscontainments} are equal to \[ \mathop{\bigcup_{{\mathfrak p} \in \Ass^*_A(I)}}_{{\mathfrak q} \in \Ass^{\infty}_B(J)} \Min_R(R/{\mathfrak p}+{\mathfrak q}) \bigcup \mathop{\bigcup_{{\mathfrak p} \in \Ass^{\infty}_A(I)}}_{{\mathfrak q} \in \Ass^*_B(J)} \Min_R(R/{\mathfrak p}+{\mathfrak q}). \] First, for the lower bound, we need to show that for $n\ge \astab^*(I)+\astab^*(J)-1$, \begin{align} &\bigcup_{i=1}^n \mathop{\bigcup_{{\mathfrak p} \in \Ass_A(I^{i-1}/I^i)}}_{{\mathfrak q} \in \Ass_B(J^{n-i}/J^{n-i+1})} \Min_R(R/{\mathfrak p}+{\mathfrak q}) \label{eq_asymptoticlowerboundofAss}\\ &= \mathop{\bigcup_{{\mathfrak p} \in \Ass^*_A(I)}}_{{\mathfrak q} \in \Ass^{\infty}_B(J)} \Min_R(R/{\mathfrak p}+{\mathfrak q}) \bigcup \mathop{\bigcup_{{\mathfrak p} \in \Ass^{\infty}_A(I)}}_{{\mathfrak q} \in \Ass^*_B(J)} \Min_R(R/{\mathfrak p}+{\mathfrak q}). \nonumber \end{align} If $i\le \astab^*(I)$, $n-i+1\ge \astab^*(J)$, hence $\Ass_B(J^{n-i}/J^{n-i+1})=\Ass^{\infty}_B(J)$. In particular, \begin{align*} &\bigcup_{i=1}^{\astab^*(I)} \mathop{\bigcup_{{\mathfrak p} \in \Ass_A(I^{i-1}/I^i)}}_{{\mathfrak q} \in \Ass_B(J^{n-i}/J^{n-i+1})} \Min_R(R/{\mathfrak p}+{\mathfrak q}) \\ &=\bigcup_{i=1}^{\astab^*(I)} \mathop{\bigcup_{{\mathfrak p} \in \Ass_A(I^{i-1}/I^i)}}_{{\mathfrak q} \in \Ass^{\infty}_B(J)} \Min_R(R/{\mathfrak p}+{\mathfrak q})=\mathop{\bigcup_{{\mathfrak p} \in \Ass^*_A(I)}}_{{\mathfrak q} \in \Ass^{\infty}_B(J)} \Min_R(R/{\mathfrak p}+{\mathfrak q}), \end{align*} where the second equality follows from Lemma \ref{lem_unionAssequal}. If $i\ge \astab^*(I)$ then $\Ass_A(A/I^i)=\Ass^{\infty}_A(I)$, $1\le n+1-i \le n+1-\astab^*(I)$. Hence \begin{align*} &\bigcup_{i=\astab^*(I)}^{n} \mathop{\bigcup_{{\mathfrak p} \in \Ass_A(I^{i-1}/I^i)}}_{{\mathfrak q} \in \Ass_B(J^{n-i}/J^{n-i+1})} \Min_R(R/{\mathfrak p}+{\mathfrak q}) \\ &=\bigcup_{i=1}^{n+1-\astab^*(I)} \mathop{\bigcup_{{\mathfrak p} \in \Ass^{\infty}_A(I)}}_{{\mathfrak q} \in \Ass_B(J^{i-1}/J^i)} \Min_R(R/{\mathfrak p}+{\mathfrak q})=\mathop{\bigcup_{{\mathfrak p} \in \Ass^{\infty}_A(I)}}_{{\mathfrak q} \in \Ass^*_B(J)} \Min_R(R/{\mathfrak p}+{\mathfrak q}). \end{align*} The second equality follows from the inequality $n+1-\astab^*(I)\ge \astab^*(J)$ and Lemma \ref{lem_unionAssequal}. Putting everything together, we get \eqref{eq_asymptoticlowerboundofAss}. The argument for the equality of the upper bound is entirely similar. The proof is concluded. \end{proof} \section{The persistence property of sums} \label{sect_persistence} Recall that an ideal $I$ in a Noetherian ring $A$ has the \emph{persistence property} if $\Ass(A/I^n)\subseteq \Ass(A/I^{n+1})$ for all $n\ge 1$. There exist ideals which fail the persistence property: A well-known example is $I=(a^4,a^3b,ab^3,b^4,a^2b^2c)\subseteq k[a,b,c]$, for which $I^n=(a,b)^{4n}$ and $(a,b,c) \in \Ass(A/I) \setminus \Ass(A/I^n)$ for all $n\ge 2$. (For the equality $I^n=(a,b)^{4n}$ for all $n\ge 2$, note that $$ U=(a^4,a^3b,ab^3,b^4) \subseteq I \subseteq (a,b)^4. $$ Hence $U^n\subseteq I^n\subseteq (a,b)^{4n}$ for all $n$, and it remains to check that $U^n=(a,b)^{4n}$ for all $n\ge 2$. By direct inspection, this holds for $n\in \{2,3\}$. For $n\ge 4$, since $U^n=U^2U^{n-2}$, we are done by induction.) However, in contrast to the case of monomial ideals, it is still challenging to find a homogeneous prime ideal without the persistence property (if it exists). Swanson and R. Walker raised the question \cite[Question 1.6]{SW} whether given two ideals $I$ and $J$ living in different polynomial rings, if both of them have the persistence property, so does $I+J$. The third main result answers in the positive \cite[Question 1.6]{SW} in many new cases. In fact, its case (ii) subsumes \cite[Corollary 1.7]{SW}. \begin{cor} \label{cor_persistence} Let $A$ and $B$ be standard graded polynomial rings over $k$, $I$ and $J$ are proper homogeneous ideals of $A$ and $B$, respectively. Assume that $I$ has the persistence property, and $\Ass(A/I^n)=\Ass(I^{n-1}/I^n)$ for all $n\ge 1$. Then $I+J$ has the persistence property. In particular, this is the case if any of the conditions hold: \begin{enumerate}[\quad \rm(i)] \item $I$ is a monomial ideal satisfying the persistence property; \item $I^{n+1}:I=I^n$ for all $n\ge 1$. \item $I^n$ is unmixed for all $n\ge 1$. \item $\chara k=0$, $\dim(A/I)\le 1$ and $I$ has the persistence property. \end{enumerate} \end{cor} \begin{proof} The hypothesis $\Ass(A/I^n)=\Ass(I^{n-1}/I^n)$ for all $n\ge 1$ and Theorem \ref{thm_Asscontainments}(2), yields for all such $n$ an equality \begin{equation} \label{eq_Ass_formula} \Ass_R \frac{R}{(I+J)^n} = \bigcup_{i=1}^n \mathop{\bigcup_{{\mathfrak p} \in \Ass_A(A/I^i)}}_{{\mathfrak q} \in \Ass_B(J^{n-i}/J^{n-i+1})} \Min_R(R/{\mathfrak p}+{\mathfrak q}). \end{equation} Take $P \in \Ass_R \dfrac{R}{(I+J)^n}$, then for some $1\le i\le n,{\mathfrak p} \in \Ass_A(A/I^i)$ and ${\mathfrak q} \in \Ass_B(J^{n-i}/J^{n-i+1})$, we get $P\in \Min_R(R/{\mathfrak p}+{\mathfrak q})$. Since $I$ has the persistence property, it follows that $\Ass(A/I^i)\subseteq \Ass(A/I^{i+1})$, so ${\mathfrak p}\in \Ass(A/I^{i+1})$. Hence thanks to \eqref{eq_Ass_formula}, \[ P \in \mathop{\bigcup_{{\mathfrak p}_1 \in \Ass_A(A/I^{i+1})}}_{{\mathfrak q}_1 \in \Ass_B(J^{n-i}/J^{n-i+1})} \Min_R(R/{\mathfrak p}_1+{\mathfrak q}_1) \subseteq \Ass_R \frac{R}{(I+J)^{n+1}}. \] Therefore $I+J$ has the persistence property. The second assertion is a consequence of the first, Theorem \ref{thm_specialcase_ass} and Lemma \ref{lem_sufficient_forpersistence}. \end{proof} \section*{Acknowledgments} The first author (HDN) and the second author (QHT) are supported by Vietnam National Foundation for Science and Technology Development (NAFOSTED) under grant numbers 101.04-2019.313 and 1/2020/STS02, respectively. Nguyen is also grateful to the support of Project CT0000.03/19-21 of the Vietnam Academy of Science and Technology. Tran also acknowledges the partial support of the Core Research Program of Hue University, Grant No. NCM.DHH.2020.15. Finally, part of this work was done while the second author was visiting the Vietnam Institute for Advance Study in Mathematics (VIASM). He would like to thank the VIASM for its hospitality and financial support. The authors are grateful to the anonymous referee for his/her careful reading of the manuscript and many useful suggestions.
{ "timestamp": "2021-06-29T02:12:22", "yymm": "2101", "arxiv_id": "2101.06583", "language": "en", "url": "https://arxiv.org/abs/2101.06583", "abstract": "Let $A, B$ be polynomial rings over a field $k$, and $I\\subseteq A, J\\subseteq B$ proper homogeneous ideals. We analyze the associated primes of powers of $I+J\\subseteq A\\otimes_k B$ given the data on the summands. The associated primes of large enough powers of $I+J$ are determined. We then answer positively a question about the persistence property of $I+J$ in many new cases.", "subjects": "Commutative Algebra (math.AC)", "title": "Powers of sums and their associated primes", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9814534316905262, "lm_q2_score": 0.8333245870332531, "lm_q1q2_score": 0.8178692756558769 }
https://arxiv.org/abs/1007.0198
Reconstruction of Bandlimited Functions from Unsigned Samples
We consider the recovery of real-valued bandlimited functions from the absolute values of their samples, possibly spaced nonuniformly. We show that such a reconstruction is always possible if the function is sampled at more than twice its Nyquist rate, and may not necessarily be possible if the samples are taken at less than twice the Nyquist rate. In the case of uniform samples, we also describe an FFT-based algorithm to perform the reconstruction. We prove that it converges exponentially rapidly in the number of samples used and examine its numerical behavior on some test cases.
\section{Introduction\label{SecIntro}} In a series of recent papers \cite{BBCE09,BCE06,BCE07}, Balan, Casazza and Edidin have investigated the possibility of reconstructing finite-dimensional signals using measurements that do not contain sign or phase information. Motivated by an application in the denoising of speech signals, they studied \textit{$M$-element frames} for $\mathbb{R}^{n}$, i.e. collections of $M$ vectors that span $\mathbb{R}^{n}$. They considered $M$-element frames, $\{f_{k}\}_{1\leq k\leq M}$, $f_{k}\in\mathbb{R}^{n}$, such that any vector $x\in\mathbb{R}^{n}$ can be uniquely determined from the inner products $\left\{ \left|\left\langle f_{k},x\right\rangle \right|\right\} _{1\leq k\leq M}$ up to an ambiguity of a sign factor. Using frame theory and combinatorial methods, they showed that such frames exist if and only if $M\geq2n+1$. In \cite{BBCE09}, a computational method to carry out this reconstruction was described in the case where $M\geq\tfrac{n(n+1)}{2}$, using a special class of frames.\\ It is natural to ask if there are analogous results for continuous-domain signals, namely in the context of samples of bandlimited functions. The well-known Whittaker-Shannon-Kotelnikov (WSK) sampling theorem \cite{Br04} shows that if a bandlimited function $f$ is sampled at a rate greater than its Nyquist frequency, then it can be uniquely reconstructed from the samples. However, if the signs of the samples are unknown, this condition may no longer be sufficient. An example illustrating this is given by the functions $\sin\left(\pi\left(z+\frac{1}{4}\right)\right)$ and $\cos\left(\pi\left(z+\frac{1}{4}\right)\right)$, which agree in absolute value at $z=\frac{k}{2}$, $k\in\mathbb{Z}$. On the other hand, it might be expected that if we oversample $f$ at a sufficiently high rate, the samples may contain enough redundancy that can we afford to lose their signs and still recover $f$ up to a global sign factor. It turns out that this is indeed the case.\\ In this paper, we use a complex variable approach to show that if a real-valued bandlimited function $f$ is sampled at more than twice its Nyquist rate, then $f$ can be uniquely determined from the absolute values of its samples up to a sign factor. Conversely, we find that if $f$ is sampled at less than twice its Nyquist rate, then it is not always possible to uniquely determine it in this way. We present an algorithm to perform this reconstruction, and show that it converges exponentially rapidly in the number of samples used. We consider a fairly general class of nonuniformly spaced samples in this paper, although our numerical approach is developed with uniformly spaced samples in mind for reasons of computational efficiency.\\ We review some existing theory on nonuniform sampling and bandlimited functions in Section \ref{SecBG}, and then state and prove our main theoretical results in Section \ref{SecMain}. We describe our algorithm and study its convergence properties in Section \ref{SecAlgo}, and apply it to two test cases in Section \ref{SecNum}. \section{Background Material\label{SecBG}} We normalize the Fourier transform as $\hat{f}(\omega)=\int_{-\infty}^{\infty}f(t)e^{-2\pi i\omega t}dt$ for Schwartz functions $f$ and extend it to tempered distributions in the usual way. For $0<p\leq\infty$, we define the \textit{Paley-Wiener spaces} of bandlimited functions \cite{Se04} by \noindent \[ PW_{b}^{p}=\left\{ f\in L^{p}:\mathrm{supp}(\hat{f})\subset[-\tfrac{b}{2},\tfrac{b}{2}]\right\} .\] \noindent An entire function $g$ is said to be of \textit{exponential type $b$} if \noindent \[ b=\inf\left(\beta:|g(z)|\leq e^{\beta|z|},z\in\mathbb{C}\right),\] and we denote this by writing $\mathrm{type}(g)=b$. By the Paley-Wiener-Schwartz theorem \cite{Ho03}, $PW_{b}^{p}$ can be equivalently described as the space of all entire functions $f$ with $\mathrm{type}(f)\leq\pi b$ whose restrictions to $\mathbb{R}$ are in $L^{p}$. It also follows that $PW_{b}^{p}\subset PW_{b}^{q}$ for $p<q$. Functions $f\in PW_{b}^{p}$ satisfy the classical estimates $\left\Vert f'\right\Vert _{L^{p}}\leq\pi b\left\Vert f\right\Vert _{L^{p}}$ and $\left\Vert f(\cdot+ic)\right\Vert _{L^{p}}\leq e^{\pi b|c|}\left\Vert f\right\Vert _{L^{p}}$, respectively known as the \textit{Bernstein} and \textit{Plancherel-Polya inequalities} \cite{Le96,Se04}.\\ We now consider a sequence of points $X=\{x_{k}\}\subset\mathbb{R}$, indexed so that $x_{k}<x_{k+1}$. For any set $B$, we denote the number of $x_{k}$ in $B$ by $N(X,B)$. We say that $X$ is \textit{separated} if $\inf_{k}|x_{k+1}-x_{k}|>0$. Following \cite{DS52}, $X$ is also said to be \textit{uniformly dense} if it satisfies\begin{equation} \sup_{k}\left|x_{k}-\frac{k}{d}\right|<\infty\label{Density}\end{equation} \noindent for some finite $d>0$. We denote this by writing $D(X)=d$, and by $D(X)=\infty$ if $X$ is not uniformly dense. We will mainly deal with separated, uniformly dense sequences in this paper. It is worth mentioning that $D(X)$ is not directly related to Beurling's upper and lower densities (see \cite{Se04}), and there are sequences $X$ with finite Beurling densities but for which $D(X)=\infty$.\\ \noindent The \textit{generating function }of a sequence $X\subset\mathbb{R}$ is given by\begin{equation} S(z)=z^{\delta_{X}}\lim_{r\rightarrow\infty}\prod_{0<|x_{k}|<r}\left(1-\dfrac{z}{x_{k}}\right),\label{GenFunc}\end{equation} where $\delta_{X}=1$ if $0\in X$ and $\delta_{X}=0$ otherwise. For a uniform sequence $x_{k}=\tfrac{k}{b},$ $S(z)=\tfrac{\sin(\pi bz)}{\pi b}$. If $X$ is separated and $D(X)=b<\infty$, then the limit in (\ref{GenFunc}) is finite and the function $S$ lies in the \textit{Cartwright class} $\, CW_{b}$ \cite{Ko96}, the set of all entire functions $f$ with $\mathrm{type}(f)\leq\pi b$ that satisfy the growth condition\\ \noindent \[ \int_{-\infty}^{\infty}\dfrac{\max(\log|f(t)|,0)}{t^{2}+1}dt<\infty.\] \noindent In fact, functions in $CW_{b}$ satisfy the apparently stronger condition \cite{Ko98}\begin{equation} \int_{-\infty}^{\infty}\dfrac{\left|\log|f(t)|\right|}{t^{2}+1}dt<\infty.\label{LogInt}\end{equation} \noindent The following result shows that an arbitrary function in $CW_{b}$ can be expanded in the form (\ref{GenFunc}) and gives a useful geometric description of its zeros. \cite{Le96} \begin{thm*} \noindent (Cartwright-Levinson) Let $W^{+}(\theta,r)$, $W^{-}(\theta,r)$ and $W'(\theta,r)$ respectively be the wedges $\{z:|z|<r,|\arg z|\leq\theta\}$, $\{z:|z|<r,|\pi-\arg z|\leq\theta\}$ and $\{z:|z|<r,|\arg z|>\theta,|\pi-\arg z|>\theta\}$. Suppose \textup{$f\in CW_{b}$, $f\not\equiv0$,} and let $U=\{u_{k}\}$ be the set of its zeros. Define $b'=\frac{1}{\pi}\inf(\mathrm{type}(e^{i\omega z}f(z)),\omega\in\mathbb{R})$.\\ \noindent 1: For any $\theta\in(0,\tfrac{\pi}{2})$, as $r\rightarrow\infty$,\[ \dfrac{N(U,W^{+}(\theta,r))}{r}\rightarrow\dfrac{b'}{2},\,\dfrac{N(U,W^{-}(\theta,r))}{r}\rightarrow\dfrac{b'}{2},\textnormal{\,\ and\,\,}\dfrac{N(U,W'(\theta,r))}{r}\rightarrow0.\] \\ 2: $U$ satisfies \textup{$\sum_{|u_{k}|>0}|\mathrm{Im}(\frac{1}{u_{k}})|<\infty.$} \\ \noindent 3:\begin{equation} f(z)=f(0)e^{iqz}z^{\delta_{U}}\lim_{r\rightarrow\infty}\prod_{0<|u_{k}|<r}\left(1-\dfrac{z}{u_{k}}\right),\label{CanonProd}\end{equation} \noindent for some constant $q$. \end{thm*} \noindent If $f$ is real-valued on $\mathbb{R}$, then $b'=b$ and $q=0$, so in particular, the sequence of real zeros $V\subset U$ of $f$ satisfies $\lim_{r\to\infty}\frac{N(V,[-r,r])}{r}\leq b$. The expression on the right side of (\ref{CanonProd}) is called a \textit{canonical product}.\\ \noindent We will also need a second, deeper theorem on $CW_{b}$. \begin{thm*} \noindent (Beurling-Malliavin) Let \textup{$f\in CW_{b}$}. Then for any $\epsilon>0$, there exists $h\in CW_{\epsilon}$, $h\not\equiv0$, such that $fh\in L^{\infty}$. \end{thm*} \noindent A discussion of this result and its significance can be found in \cite{HMN06}. We can actually choose the above $h$ so that $fh\in L^{p}$ for any $p>0$, by replacing $h$ with $h\phi$ where $\phi$ is a function for which $\hat{\phi}$ is smooth and has sufficiently small support. By considering the function $h(z)\overline{h(\overline{z})}$, $h$ can also be taken to be real and nonnegative.\\ For $1\leq p\leq\infty$, $X$ is called a \textit{sampling sequence} for $PW_{b}^{p}$ if there are constants $\alpha_{p,b}$ and $\beta_{p,b}$ such that for all $f\in PW_{b}^{p}$, $\alpha_{p,b}||f(X)||_{l^{p}}\leq\left\Vert f\right\Vert _{L^{p}}\leq\beta_{p,b}\left\Vert f(X)\right\Vert _{l^{p}}$. In particular, this condition implies that any $f\in PW_{b}^{p}$ is uniquely determined by its samples at $X$. Precise geometric characterizations of sampling sequences can be very complicated (see \cite{Br04,Se97,Se04}), but for real, separated and uniformly dense sequences $X$, it is necessary that $D(X)\geq b$ and sufficient that $D(X)>b$ for $X$ to be a sampling sequence. If $D(X)>b$ and $S$ is the generating function of $X$, a consequence of the Beurling-Malliavin theorem is that there is an $h\in CW_{\epsilon}$ with zero set $U$, $\epsilon=D(X)-b$, such that any $f\in PW_{b}^{\infty}$ can be expressed in terms of its samples,\\ \begin{equation} f(z)=\sum_{k=-\infty}^{\infty}f(x_{k})\,\dfrac{S(z)h(x_{k})}{h(z)S'(x_{k})(z-x_{k})},\label{SamplingSeries}\end{equation} \noindent with uniform convergence on compact subsets of $\mathbb{C}\backslash U$ \cite{Ko96,Se97}. $D(X)$ can be thought of as a generalization of the {}``sampling rate'' to uniformly dense sequences $X$, and when $X=\{\frac{k}{s}\}$ is a uniform sequence, the condition $b<D(X)=s$ simply says that $f$ is being oversampled beyond its Nyquist rate. $h$ can be taken as constant for uniform sequences, and the expansion (\ref{SamplingSeries}) reduces to the classical WSK sampling theorem. \section{Main Results\label{SecMain}} \noindent We fix $p\in(0,\infty]$ for the rest of this section. We first show that if a bandlimited function is sampled at more than twice its Nyquist rate, then we can reconstruct it up to a sign factor from the absolute values of the samples. \begin{thm} \label{ThmMain}Let $f\in PW_{b}^{p}$ be real-valued on $\mathbb{R}$, and let $X\subset\mathbb{R}$ be a separated, uniformly dense sequence with $D(X)>2b$. Then f can be uniquely determined from $a_{k}=|f(x_{k})|$, up to a sign factor.\end{thm} \begin{proof} \noindent We normalize $b=1$ without loss of generality. Let $S$ be the generating function of $X$ and suppose $h$ is given as in (\ref{SamplingSeries}). The zeros of $f$ and $h$ are countable, so we can choose $c>0$ so that $f$ and $h$ have no zeros on the line $L=\{z:\mathrm{Im}(z)=c\}$. We can write $g=f(\cdot+ic)^{2}\in PW_{2}^{p/2}$ as \noindent \begin{equation} g(z)=\sum_{k=-\infty}^{\infty}a_{k}^{2}\,\dfrac{S(z+ic)h(x_{k})}{h(z+ic)S'(x_{k})(z+ic-x_{k})}.\label{MainSamp1}\end{equation} \noindent From Bernstein's inequality, $g'\in PW_{2}^{p/2}$ and differentiating (\ref{MainSamp1}) gives a similar expansion for $g'(z)$. Let $a_{k}^{*}=f(x_{k}+ic)$ and choose a point $x_{l}\in X$. Since $f$ has no zeros on $L$, there is a branch of $\arg f$, which we denote by $\arg_{0}f$, that is continuous on $L$ and satisfies $\arg_{0}f(x_{l}+ic)\in(-\pi,\pi]$. We define $\arg_{0}g$ in the same way, and we then have \noindent \begin{eqnarray} a_{k}^{*} & = & \exp\left(\frac{1}{2}\log|g(x_{k})|+i\arg_{0}f(x_{k}+ic)\right)\\ & = & \eta|g(x_{k})|^{1/2}\exp\left(\dfrac{i}{2}\left(\int_{x_{l}}^{x_{k}}\mathrm{Im}\left(\frac{g'(t)}{g(t)}\right)dt+\arg_{0}g(x_{l})\right)\right)\label{fSamp}\end{eqnarray} \noindent where $\eta=\exp\left(i\left(\frac{1}{2}\arg_{0}g(x_{l})-\arg_{0}f(x_{l}+ic)\right)\right)=\pm1$. We have now determined samples of $f$, which we can use to express $f$ as \noindent \begin{equation} f(z)=\eta\sum_{k=-\infty}^{\infty}a_{k}^{*}\,\dfrac{S(z-ic)h(x_{k})}{h(z-ic)S'(x_{k})(z-ic-x_{k})}.\label{MainSamp2}\end{equation} \end{proof} \begin{rem*} The proof of Theorem \ref{ThmMain} suggests a three-step procedure to recover $f$ from $\{a_{k}\}$. We can determine $f^{2}$ from $\{a_{k}\}$ and take its square root by unwrapping its phase. Since $f$ will typically have zeros on the real axis, we first move up in the complex plane using (\ref{MainSamp1}), unwrap the phase there with (\ref{fSamp}) and then move back to the real axis with (\ref{MainSamp2}). We will use this approach in Section \ref{SecAlgo}. \end{rem*} The next result shows that Theorem \ref{ThmMain} is in a sense sharp. If we sample a function at less than twice its Nyquist rate, it may or may not be uniquely determined by the absolute values of the samples, essentially depending on {}``how many'' of the samples are zero. \begin{thm} \label{ThmUnder}Let $X\subset\mathbb{R}$ be a separated sequence with $D(X)<2b$. Then there is a real-valued function $f\in PW_{b}^{p}$ that cannot be uniquely determined from $|f(x_{k})|$ up to a sign factor. If in addition $D(X)>\frac{4}{3}b$, then there is another real-valued function $\tilde{f}\in PW_{b}^{p}$ that can be uniquely determined from \textup{$|\tilde{f}(x_{k})|$} up to a sign factor, but for which there is no subsequence $Y\subset X$ with $\tilde{f}(y_{k})=0$ and $\lim_{r\to\infty}\frac{N(Y,[-r,r])}{r}\geq b$.\end{thm} \begin{proof} As before, we normalize $b=1$. Suppose $D(X)=2-\epsilon$ for some $\epsilon>0$, and let $X_{1}=\{x_{2k}\}$ and $X_{2}=\{x_{2k+1}\}$. It follows from the definition (\ref{Density}) that $D(X_{1})=D(X_{2})=1-\tfrac{\epsilon}{2}$. Now let $S_{1}$ and $S_{2}$ be the generating functions of $X_{1}$ and $X_{2}$. By the Beurling-Malliavin theorem, we can find $h_{1},h_{2}\in CW_{\epsilon/2}$ such that $S_{1}h_{1}$ and $S_{2}h_{2}$ are in $PW_{1}^{p}$. Then the functions $f_{1}=S_{1}h_{1}+S_{2}h_{2}$ and $f_{2}=$$S_{1}h_{1}-S_{2}h_{2}$ are also in $PW_{1}^{p}$ and satisfy $|f_{1}(X)|=|f_{2}(X)|$.\\ For the other part of Theorem \ref{ThmUnder}, suppose $\epsilon$ above satisfies $\epsilon<\tfrac{2}{3}$. By countability, we can choose $c$ so that $h_{3}(z)=h_{1}(z+c)$ has no zeros on $X$. Let $\tilde{f}=S_{1}h_{3}$, so that among the samples of $\tilde{f}$ at $X$, only the ones at $X_{1}$ are zero. If there is any real-valued $g\in PW_{1}^{p}$ with $|\tilde{f}(X)|=|g(X)|$, then $\tilde{f}{}^{2}-g^{2}$ is in $PW_{2}^{p/2}$ and has zeros at $X$, so by considering canonical product expansions, $\tilde{f}{}^{2}-g^{2}=S_{1}S_{2}h_{4}$ for some $h_{4}\in CW_{\epsilon}$. Since $g^{2}=S_{1}(S_{1}h_{3}^{2}-S_{2}h_{4})$ has an analytic square root, it can only have double zeros, so in particular, $S_{1}h_{3}^{2}-S_{2}h_{4}$ has to be zero on $X_{1}$. This implies that $h_{4}$ must have zeros on $X_{1}$, which contradicts the Cartwright-Levinson theorem because $D(X_{1})=1-\tfrac{\epsilon}{2}>\epsilon$. So $h_{4}\equiv0$ and $\tilde{f}=\pm g$. \end{proof} \begin{rem*} The nonexistence of $Y$ in Theorem \ref{ThmUnder} is what makes the result interesting, as it means that $\tilde{f}$ cannot be determined from $Y$ by just using a canonical product. In other words, the nonzero samples at $X_{2}$ play a role in the uniqueness of $\tilde{f}$. However, there appears to be no simple characterization of all such functions $\tilde{f}$ or a numerically useful method of computing $\tilde{f}$ from $|\tilde{f}(X)|$. \end{rem*} $\,$ \begin{rem*} We have not considered the border case of $D(X)=2b$ in the above results, in which case the conditions required on the sequence $X$ would become more subtle and depend on the value of $p$. However, in the elementary case where $x_{k}=\frac{k}{2b}$ is a uniform sequence and $p=2$, the conclusion of Theorem \ref{ThmMain} still holds by just using the WSK sampling theorem in place of (\ref{MainSamp1}) and (\ref{MainSamp2}). \end{rem*} $\,$ There are no simple analogs of these results if we allow $f\in PW_{b}^{p}$ to be complex-valued. In Theorem \ref{ThmMain}, we used the fact that when $f$ is real-valued, $f^{2}$ has the same samples as $|f|^{2}$, but this is no longer the case for complex-valued $f$. In general, such an $f$ will have complex zeros $u_{k}$ and complex-valued functions of the form $Bf$, where $B$ is a Blaschke product formed from any subset of $\{\overline{u_{k}}\}$, will be in $PW_{b}^{p}$ and satisfy $|Bf|=|f|$ identically on $\mathbb{R}$. If we require all the zeros of $f$ to be real, then since $f$ can be written as a canonical product over them, it is simply a modulation of a real-valued function $g$, i.e. it has the form $f(z)=e^{i(cz+d)}g(z)$ for $c,d\in\mathbb{R}$. This situation is in contrast to the findings in \cite{BCE06}, where the types of frames the authors studied had results for complex vectors comparable to those outlined in Section \ref{SecIntro} for real vectors. \section{A Reconstruction Algorithm\label{SecAlgo}} \noindent We now describe how to computationally implement the technique in the proof of Theorem \ref{ThmMain}. We restrict our attention to uniform sampling sequences here, as they lead to convolution-type sampling series that can be calculated efficiently by Fast Fourier Transform (FFT) methods, but the same ideas can be adapted to the nonuniform case. We first define the following functions:\begin{eqnarray*} G(z,M) & = & \frac{\sin(\pi z)}{\pi z}e^{-\frac{\pi}{2M}z^{2}}\\ G'(z,M) & = & \left(\frac{\cos(\pi z)}{\pi z}-\frac{\sin(\pi z)}{M}-\frac{\sin(\pi z)}{\pi z}\right)e^{-\frac{\pi}{2M}z^{2}}\\ G^{*}(z,M) & = & \frac{1}{M}\int_{z-M}^{z}G(t,M)dt\end{eqnarray*} We also denote the strip $\{z:|\mathrm{Im}(z)|<\delta\}$ by $T_{\delta}$. Then we have the following results from \cite{SS03}, reproduced here in slightly different forms for our purposes. \begin{thm} \noindent \label{ThmConvFact1}(Schmeisser-Stenger, 2002) Let $f\in PW_{b}^{\infty}$, $s>b$ and $d<1$. Then for $\mathrm{Re}(z)\in[-dM,dM]$, $f(\frac{z}{s})=\sum_{k=-M}^{M}f(\frac{k}{s})G(z-k,M)+E(z)$, where $|E(z)|\leq C_{1}M^{-1/2}e^{-\frac{\pi(1-d)}{2}\left(1-\frac{b}{s}\right)M+2\pi|Im(z)|}\cdot\left\Vert f\right\Vert _{L^{\infty}}$ for some constant $C_{1}=C_{1}(\tfrac{b}{s})$. \end{thm} \noindent $\,$ \begin{thm} \noindent \label{ThmConvFact2} (Schmeisser-Stenger, 2002) Suppose $f$ is analytic in $T_{\delta}$ and $|f(z)|\leq K|z|$ for $z\in T_{\delta}$ and some constant $K$. Then for $z\in\mathbb{R}$,\textup{$-\frac{d}{2}\leq z\leq\frac{d}{2}$,} $f(z)=\sum_{k=-dM}^{dM}f(\frac{k}{M})G(Mz-k,M)+E(z)$, where $|E(z)|\leq C_{2}K(M/\delta)^{1/2}e^{-\frac{\pi\delta M}{4}}\left\Vert f\right\Vert _{L^{\infty}}$ for some constant $C_{2}$. \end{thm} \noindent Theorem \ref{ThmConvFact1} is a form of (\ref{SamplingSeries}) with a non-bandlimited $h$. It converges very rapidly in practice for about $s\geq1.3b$. Theorem \ref{ThmConvFact2} is a version of Theorem \ref{ThmConvFact1} as $s\to\infty$.\\ These results lead to the following numerical approach. We impose a mild restriction to rule out some pathological functions, and assume for notational convenience that an odd number of samples are used. \begin{algorithm} \noindent \label{Algo} Suppose $f\in PW_{b}^{\infty}$ is nonzero in some strip $\{z:|\mathrm{Im}(z)-c|\leq\delta\}$, $\delta>0$. Let $s>2b$ and $a_{k}=|f(\frac{k}{s})|$ for $-M\leq k\leq M$.\\ \noindent \textup{1: Compute $g_{M}(\frac{z}{s})=\sum_{k=-M}^{M}a_{k}^{2}\, G(z-k+ic,M)$ at $z=\frac{n}{M}$, $-M^{2}\leq n\leq M^{2}$.} \noindent \textup{2: Compute $g_{M}^{\prime}(\frac{z}{s})=\sum_{k=-M}^{M}a_{k}^{2}\, G'(z-k+ic,M)$ for $z$ as above.} \noindent \textup{3: Compute $Q_{M}(n)=\frac{1}{s}\sum_{k=(n-2)M}^{(n+1)M}\mathrm{Im}\left(\dfrac{g_{M}^{\prime}(\frac{k}{Ms})}{g_{M}(\frac{k}{Ms})}\right)G^{*}(Mn-k,M)$ for $-(M-2)\leq n\leq M-1$.} \noindent \textup{4: Compute $R_{M}(n)$ given by $R_{M}(0)=0$, $R_{M}(n)=R_{M}(n-1)+Q_{M}(n-1)$ for $-(M-1)\leq n\leq M-1$.} \noindent \textup{5: Compute $f_{M}(\frac{z}{s})=\sum_{k=-(M-1)}^{M-1}\sqrt{|g_{M}(\frac{k}{s})|}e^{\frac{i}{2}(R_{M}(k)+\arg g_{M}(0))}\, G(z-k-ic,M)$.}\\ \end{algorithm} The convolutions in steps 1 and 2 can be performed by $2M$ FFTs of size $2M+1$ each. Steps 3 and 4 are most efficiently done by direct computations that respectively involve $O(M^{2})$ and $O(M)$ operations. Step 5 involves $2N$ FFTs of size $2M+1$ for some integer $N$, depending on how finely we want to compute $f$. This gives an overall complexity of $O(M^{2}\log M)$.\\ Theorem \ref{ThmConvFact2} is used above to calculate the integral in (\ref{fSamp}). The advantage of this approach is that the error bound in Theorem \ref{ThmConvFact2} does not depend on the derivatives of $f$. If we instead used conventional methods of integration such as Gauss quadrature, we would need additional restrictions on $\delta$ to ensure that the algorithm actually converges. There is no simple closed-form expression for $G^{*}$ but its numerical calculation only depends on $M$, so we can tabulate its values by a standard quadrature method using $O(M^{2})$ operations and reuse them for different $f$ with a lookup table.\\ \noindent We can establish the following convergence result for Algorithm \ref{Algo}. \begin{thm} \noindent \label{ThmAlgo}Let $f$ and $f_{M}$ be defined as in Algorithm \ref{Algo}. As $M\rightarrow\infty$, $f_{M}\rightarrow\eta f$ uniformly on compact subsets of $\mathbb{R}$, where $\eta=\pm1$ and $|f_{M}(z)-\eta f(z)|=O\left(e^{-\min\left(\frac{\pi}{16}\left(1-\frac{2b}{s}\right),\frac{\pi\delta}{8}\right)M+4\pi c}\right)$. \end{thm} This estimate is somewhat conservative and larger exponents of convergence are possible if we make more assumptions on $f$, but it shows how $\delta$ and $c$ affect the rate of convergence. Many real-world bandlimited signals have most of their zeros on or near the real axis, so choosing $c$ too small will result in a small $\delta$ while choosing $c$ too large will sharply increase the constant in the error bound. Values of $c$ between about $0.01$ to $0.25$ appear to work well in practice.\\ \noindent Before we prove Theorem \ref{ThmAlgo}, we will need an auxiliary lemma. \begin{lem} \label{hDh}Let $h\in PW_{b}^{\infty}$ be nonzero in $T_{\delta}$. Then $\left|\frac{h'(z)}{h(z)}\right|\leq K(|z|^{2}+1)$ in $T_{\delta/2}$ for some constant $K$.\end{lem} \begin{proof} \noindent Let $U=\{u_{k}\}$ be the zeros of $h(z-i\delta)$ lying in the upper half plane $\mathbb{C}^{+}=\{z:\mathrm{Im}(z)>0\}$, and suppose $\frac{\delta}{2}<\mathrm{Im}(z)<\frac{3\delta}{2}$. The Plancherel-Polya inequality shows that $h(\cdot-i\delta)\in PW_{b}^{\infty}$ and that the function $e^{\pi ibz}h(z-i\delta)$ is bounded and analytic on $\mathbb{C}^{+}$, so it has the inner-outer factorization \cite{Ga07} \[ \log h(z-i\delta)=\frac{1}{\pi i}\int_{-\infty}^{\infty}\frac{1+zt}{(t-z)(t^{2}+1)}\log\left|h(t-i\delta)\right|dt+\log\left(\prod_{k}\frac{|u_{k}^{2}+1|}{u_{k}^{2}+1}\frac{z-u_{k}}{z-\overline{u_{k}}}\right)+Az+B\] \noindent for some constants $A$ and $B$. We differentiate this to find that\begin{eqnarray*} \left|\frac{h'(z-i\delta)}{h(z-i\delta)}\right| & = & \left|\frac{1}{\pi i}\int_{-\infty}^{\infty}\frac{\log\left|h(t-i\delta)\right|}{(t-z)^{2}}dt+\sum_{k}\frac{2i\mathrm{Im}(u_{k})}{(z-u_{k})(z-\overline{u_{k}})}+A\right|\\ & \leq & (2/\delta)^{2}\left(|\mathrm{Re}(z)|^{2}+1\right)\left(\frac{1}{\pi}\int_{-\infty}^{\infty}\frac{\log\left|h(t-i\delta)\right|}{t^{2}+1}dt+\sum_{k}\frac{2\mathrm{Im}(u_{k})}{|u_{k}|^{2}+1}+|A|\right).\end{eqnarray*} \noindent Since $PW_{b}^{\infty}\subset CW_{b}$, condition (\ref{LogInt}) shows that the integral in the first term is finite. The Cartwright-Levinson theorem implies that the sum in the second term also converges, which finishes the proof. \end{proof} \noindent It is possible to obtain sharper results than this, but this lemma is good enough for our purposes.\\ \begin{proof}[Proof of Theorem \ref{ThmAlgo}] \noindent We proceed by establishing several intermediate bounds and then combine them all at the end. To simplify the notation, we will always use $z\in\mathbb{R}$ and $k\in\mathbb{Z}$ to denote function arguments within norms, e.g. the norm of $F\in L^{\infty}(\mathbb{R})$ will be denoted by $\left\Vert F(z)\right\Vert $.\\ \noindent Let $g$ and $\eta$ be defined as in the proof of Theorem \ref{ThmMain}. Define $I_{r,M}=\big[-\left(\lfloor r(M-1)\rfloor-1\right),\lfloor r(M-1)\rfloor-1\big]$, $i_{r,M}=\mathbb{Z}\bigcap I_{r,M}$ and $j_{r,M}=\mathbb{Z}\bigcap\left[-M\lfloor rM\rfloor,M\lfloor rM\rfloor\right]$. Theorem \ref{ThmConvFact1} shows that\[ E_{1}:=\left\Vert g_{M}(\tfrac{k}{Ms})-g(\tfrac{k}{Ms})\right\Vert _{l^{\infty}(j_{3/4,M})}\leq\frac{C_{1}}{\sqrt{M}}e^{-\frac{\pi}{8}\left(1-\frac{2b}{s}\right)M+2\pi c}\left\Vert f\right\Vert _{L^{\infty}(\mathbb{R})}^{2}.\] \noindent The error term $E$ in Theorem \ref{ThmConvFact1} is an entire function and it satisfies the classical Cauchy estimate\[ \left\Vert E'(z+ic)\right\Vert _{L^{\infty}(I_{3/4,M+2})}\leq2\pi\max\left(\left|E(z+ic)\right|,\left\{ z:|\mathrm{Re}(z)|\leq\frac{3}{4}(M+1),|\mathrm{Im}(z)|\leq\frac{1}{2\pi}\right\} \right).\] \noindent This shows that\[ E_{2}:=\left\Vert g_{M}^{\prime}(\tfrac{k}{Ms})-g^{\prime}(\tfrac{k}{Ms})\right\Vert _{l^{\infty}(j_{3/4,M})}\leq\frac{2\pi C_{1}}{\sqrt{M}}e^{-\frac{\pi}{8}\left(1-\frac{2b}{s}\right)(M-1)+2\pi c+1}\left\Vert f\right\Vert _{L^{\infty}(\mathbb{R})}^{2}.\] \\ \noindent A standard identity for Jacobi theta functions (\cite{WW27}, p. 475) gives the bound\[ E_{3}:=\left\Vert \sum_{l=-M}^{M}G(z-l-ic,M)\right\Vert _{L^{\infty}(I_{1/4,M})}\leq\left\Vert \sum_{l=-\infty}^{\infty}e^{\frac{\pi}{2M}(c^{2}-(z-l)^{2})}\right\Vert _{L^{\infty}(\mathbb{R})}\leq\frac{3}{2}e^{\frac{\pi c^{2}}{2M}}\sqrt{2M},\] \noindent and similarly,\[ E_{4}:=\left\Vert \sum_{l=(k-2)M}^{(k+1)M}|G^{*}(Mk-l,M)|\right\Vert _{l^{\infty}(i_{3/4,M+2})}\leq\frac{3}{2}\sqrt{2M}.\] \\ \noindent Now by Lemma \ref{hDh}, $\left|\frac{g'(z)}{g(z)}\right|\leq K(|z|^{2}+1)$ in $T_{\delta/2}$ for a constant $K$. Since $\left\Vert g_{M}-g\right\Vert _{L^{\infty}(I_{3/4,M+2})}\to0$ and $g$ has no zeros on $\mathbb{R}$, we have $\left\Vert 1/g_{M}\right\Vert _{L^{\infty}(I_{3/4,M+2})}<\infty$ for sufficiently large $M$. We also have the bound\begin{eqnarray*} \left\Vert \dfrac{g_{M}'(\frac{k}{Ms})}{g_{M}(\frac{k}{Ms})}-\dfrac{g'(\frac{k}{Ms})}{g(\frac{k}{Ms})}\right\Vert _{l^{\infty}(j_{3/4,M})} & \leq & \left\Vert \frac{1}{g_{M}}\right\Vert _{L^{\infty}(I_{3/4,M+2})}\left(E_{2}+\left\Vert g'\right\Vert _{L^{\infty}(I_{3/4,M+2})}\left\Vert 1-\dfrac{g_{M}}{g}\right\Vert _{L^{\infty}(I_{3/4,M+2})}\right)\\ & \leq & \left\Vert \frac{1}{g_{M}}\right\Vert _{L^{\infty}(I_{3/4,M+2})}\left(K(M^{2}+1)+1\right)\max(E_{1},E_{2}).\end{eqnarray*} We can use Theorem \ref{ThmConvFact2} with this to find that\begin{eqnarray*} E_{5} & := & \left\Vert R_{M}(k)-\int_{0}^{k}\mathrm{Im}\left(\frac{g'(\frac{t}{s})}{g(\frac{t}{s})}\right)dt\right\Vert _{l^{\infty}(i_{3/4,M})}\\ & \leq & M\left\Vert Q_{M}(k)-\int_{k-1}^{k}\mathrm{Im}\left(\frac{g'(\frac{t}{s})}{g(\frac{t}{s})}\right)dt\right\Vert _{l^{\infty}(i_{3/4,M}\backslash\{1-\lfloor\frac{3}{4}(M-1)\rfloor\})}\\ & \leq & M\left(\left\Vert \dfrac{g_{M}'(\frac{k}{Ms})}{g_{M}(\frac{k}{Ms})}-\dfrac{g'(\frac{k}{Ms})}{g(\frac{k}{Ms})}\right\Vert _{l^{\infty}(j_{3/4,M})}E_{4}+C_{2}(2M/\delta)^{1/2}e^{-\frac{\pi\delta M}{8}}\left\Vert \dfrac{g'(\frac{k}{Ms})}{g(\frac{k}{Ms})}\right\Vert _{l^{\infty}(j_{3/4,M})}\right)\\ & \leq & \left\Vert \frac{1}{g_{M}}\right\Vert _{L^{\infty}(I_{3/4,M+2})}\left(K(M^{2}+1)+1\right)M^{3/2}E_{2}+C_{2}KM^{3/2}(2/\delta)^{-1/2}e^{-\frac{\pi\delta M}{8}}.\end{eqnarray*} We now put everything together. Using the elementary inequality $||u|^{1/2}e^{i\theta}-|v|^{1/2}|\leq|u-v|^{1/2}+|1-e^{i\theta}||v|^{1/2}$ along with (\ref{fSamp}) gives {\allowdisplaybreaks \noindent \begin{eqnarray*} E_{6} & := & \left\Vert f_{M}\left(\frac{z}{s}\right)-\eta f\left(\frac{z}{s}\right)\right\Vert _{L^{\infty}(I_{1/4,M})}\\ & \leq & \left\Vert f_{M}\left(\frac{k}{s}+ic\right)-\eta f\left(\frac{k}{s}+ic\right)\right\Vert _{l^{\infty}(i_{3/4,M})}E_{3}+C_{1}e^{-\frac{\pi}{8}\left(1-\frac{b}{s}\right)(M-1)+2\pi c}\left\Vert f\right\Vert _{L^{\infty}(\mathbb{R})}\\ & & +\left\Vert \sum_{l\in i_{1,M}\backslash i_{3/4,M}}e^{-\frac{\pi}{2(M-1)}\left(z-l\right)^{2}+\pi c}\right\Vert _{L^{\infty}(\mathbb{R})}\left\Vert f_{M}\left(\frac{k}{s}+ic\right)\right\Vert _{l^{\infty}(i_{1,M}\backslash i_{3/4,M})}\\ & \leq & \left\Vert \left|g_{M}\left(\frac{k}{s}\right)\right|^{1/2}e^{\frac{i}{2}(R_{M}(k)+\arg g_{M}(0))}-\eta f\left(\frac{k}{s}+ic\right)\right\Vert _{l^{\infty}(i_{3/4,M})}E_{3}\\ & & +2C_{1}e^{-\frac{\pi}{8}\left(1-\frac{b}{s}\right)(M-1)+4\pi c}\left\Vert f\right\Vert _{L^{\infty}(\mathbb{R})}\\ & \leq & \Bigg(E_{1}^{1/2}+\left\Vert 1-\exp\left(R_{M}(k)-\int_{0}^{k}\mathrm{Im}\left(\frac{g'(t)}{g(t)}\right)dt+\arg g_{M}(0)-\arg g(0)\right)\right\Vert _{l^{\infty}(i_{3/4,M})}\\ & & \cdot\left\Vert f\right\Vert _{L^{\infty}(\mathbb{R})}E_{3}\Bigg)+2C_{1}e^{-\frac{\pi}{8}\left(1-\frac{b}{s}\right)(M-1)+4\pi c}\left\Vert f\right\Vert _{L^{\infty}(\mathbb{R})}\\ & \leq & \left(E_{1}^{1/2}+\frac{1}{2}(E_{5}+E_{1})\left\Vert f\right\Vert _{L^{\infty}(\mathbb{R})}\right)E_{3}+2C_{1}e^{-\frac{\pi}{8}\left(1-\frac{b}{s}\right)(M-1)+4\pi c}\left\Vert f\right\Vert _{L^{\infty}(\mathbb{R})}\\ & = & C_{3}(f)M^{7/2}\exp\left(-\min\left(\frac{\pi}{16}\left(1-\frac{2b}{s}\right),\frac{\pi\delta}{8}\right)M+\max\left(\frac{\pi c^{2}}{2M},4\pi c\right)\right)\end{eqnarray*} \noindent } for some $C_{3}(f)<\infty$ and sufficiently large $M$, which establishes the result. \end{proof} \section{Numerical Experiments\label{SecNum}} We illustrate how Algorithm \ref{Algo} works on two test cases. We consider the translated Bessel function (see \cite{WW27}) $f(z)=J_{1}(z+20)$ sampled at $z=k$, $-M\leq k\leq M$, and a collection of $2M+1$ samples taken from an 8-bit, 44khz audio file. To show the effect of the parameter $c$, we take $c=0.1$ in the first example and $c=0.04$ in the second one. Values of $G^{*}$ are tabulated using the built-in Gauss-Kronrod algorithm in MATLAB. We measure the worst-case error over $I_{1/2,M+1}$ as defined in Section \ref{SecAlgo}, in order to avoid influence from inaccuracies close to the boundary of the full domain $I_{1,M+1}$. \begin{figure}[H] \begin{raggedright} $\,$ \par\end{raggedright} \begin{raggedright} $\,$ \par\end{raggedright} \begin{raggedright} $\,$ \par\end{raggedright} \begin{raggedright} \begin{tabular}{|c|c|} \hline $M$ & Error over $I_{1/2,M+1}$\tabularnewline \hline 10 & $3.7490\cdot10^{-2}$\tabularnewline \hline 20 & $5.9513\cdot10^{-4}$\tabularnewline \hline 30 & $4.0158\cdot10^{-5}$\tabularnewline \hline 40 & $3.8732\cdot10^{-6}$\tabularnewline \hline 50 & $3.8362\cdot10^{-7}$\tabularnewline \hline \end{tabular}\includegraphics[bb=0bp 400bp 0bp 200bp,scale=0.5]{Fig1.png} \par\end{raggedright} \begin{raggedright} $\,$ \par\end{raggedright} \begin{raggedright} $\,$ \par\end{raggedright} $\,$ \caption{The graph of $f(z)=J_{1}(z+20)$ and the reconstruction errors with $c=0.1$. It is clear that the error decays exponentially as $M$ increases.} \end{figure} \begin{figure}[H] $\,$ \begin{raggedright} $\,$ \par\end{raggedright} \begin{tabular}{|c|c|} \hline $M$ & Error over $I_{1/2,M+1}$\tabularnewline \hline 10 & $1.9752\cdot10^{-2}$\tabularnewline \hline 20 & $5.1622\cdot10^{-3}$\tabularnewline \hline 30 & $1.5710\cdot10^{-4}$\tabularnewline \hline 40 & $1.1563\cdot10^{-4}$\tabularnewline \hline 50 & $8.8637\cdot10^{-5}$\tabularnewline \hline \end{tabular}\includegraphics[bb=0bp 400bp 0bp 200bp,scale=0.5]{Fig2.png} $\,$ \begin{raggedright} $\,$ \par\end{raggedright} $\,$\caption{The graph of a section of audio data and the resulting reconstruction errors with $c=0.04$. The convergence here is slower than it was in the preceding example. This is likely due to the presence of complex zeros with imaginary parts very close to $c$, as well as the proximity of the real zeros.} \end{figure} \begin{acknowledgement*} The author would like to thank Professor Ingrid Daubechies for many valuable discussions in the course of this work. \end{acknowledgement*} \bibliographystyle{amsplain}
{ "timestamp": "2010-07-02T02:01:40", "yymm": "1007", "arxiv_id": "1007.0198", "language": "en", "url": "https://arxiv.org/abs/1007.0198", "abstract": "We consider the recovery of real-valued bandlimited functions from the absolute values of their samples, possibly spaced nonuniformly. We show that such a reconstruction is always possible if the function is sampled at more than twice its Nyquist rate, and may not necessarily be possible if the samples are taken at less than twice the Nyquist rate. In the case of uniform samples, we also describe an FFT-based algorithm to perform the reconstruction. We prove that it converges exponentially rapidly in the number of samples used and examine its numerical behavior on some test cases.", "subjects": "Numerical Analysis (math.NA); Complex Variables (math.CV)", "title": "Reconstruction of Bandlimited Functions from Unsigned Samples", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9865717436787543, "lm_q2_score": 0.8289388040954683, "lm_q1q2_score": 0.8178076013594475 }
https://arxiv.org/abs/2111.01359
Unfoldings and Nets of Regular Polytopes
Over a decade ago, it was shown that every edge unfolding of the Platonic solids was without self-overlap, yielding a valid net. We consider this property for regular polytopes in arbitrary dimensions, notably the simplex, cube, and orthoplex. It was recently proven that all unfoldings of the $n$-cube yield nets. We show this is also true for the $n$-simplex and the $4$-orthoplex but demonstrate its surprising failure for any orthoplex of higher dimension.
\section{Introduction} \label{s:intro} The study of unfolding polyhedra was popularized by Albrecht D\"urer in the early 16th century who first recorded examples of polyhedral \emph{nets}, connected edge unfoldings of polyhedra that lay flat on the plane without overlap. Motivated by this, Shephard \cite{sh1} conjectures that every convex polyhedron can be cut along certain edges to admit a net. This claim remains tantalizingly open and has resulted in numerous areas of exploration; see \cite{gho} for a survey. We consider this question for higher-dimensional \emph{polytopes}: The codimension-one faces of a polytope are \emph{facets} and its codimension-two faces are \emph{ridges}. The analog of an edge unfolding of polyhedron is the \emph{ridge unfolding} of an $n$-dimensional polytope: the process of cutting the polytope along a collection of its ridges so that the resulting (connected) arrangement of its facets develops isometrically into an $\R^{n-1}$ hyperplane. There is a rich history of higher-dimensional unfoldings of polytopes, with the collected works of Alexandrov \cite{alex1} serving as seminal reading. In particular, Buekenhout and Parker \cite{bupa} enumerate the ridge unfoldings of the six regular convex 4-polytopes. In our work, instead of trying to find one valid net for each convex polyhedron (as posed by Shephard), we consider a more aggressive property: \begin{defn} A polytope $P$ is \emph{all-net} if every ridge unfolding of $P$ yields a valid net.\footnote{This nomenclature comes from Joe O'Rourke: akin to a basketball ``all-net'' shot that scores by not touching the rim, all unfoldings become successful nets by facets not overlapping and touching each other.} \end{defn} \noindent A decade ago, Horiyama and Shoji \cite{hosh} showed that the five Platonic solids are all-net. This is easy to check for the tetrahedron, cube, and octahedron, for which there are few unfoldings, but it is significantly nontrivial for the dodecahedron and icosahedron, for which there are 43,380 distinct unfoldings. Figure~\ref{f:3dnet} shows the 11 different unfoldings (up to symmetry) of the octahedron, all of which are nets. Recent work \cite{pnas} has shown applications in protein science: polyhedral nets are used to find a balance between entropy loss and energy gain for the folding propensity of a given shape. \begin{figure}[h] \includegraphics[width=.95\textwidth]{11-3d-ortho-unfold} \caption{The 11 nets of the octahedron, also known as the 3-orthoplex.} \label{f:3dnet} \end{figure} The higher-dimensional analogs of the Platonic solids are the regular polytopes. Three classes of regular polytopes exist for all dimensions: $n$-simplex, $n$-cube, and $n$-orthoplex (sometimes called the \emph{cross-polytope}). Three additional regular polytopes appear only in four-dimensions: the 24-cell, 120-cell, and 600-cell. It was recently shown that the $n$-cube is all-net \cite{ddrw}; in particular, the 261 different ridge unfoldings of the $4$-cube do not overlap. In Section~\ref{s:simplex}, we show that the $n$-simplex is all-net as well. Section~\ref{s:orthoplex} establishes combinatorial and geometric attributes of the orthoplex which Section~\ref{s:orthonet} uses to demonstrate the all-net property of the $4$-orthoplex. That is, the 110,912 ridge unfoldings of the 4-orthoplex enumerated in \cite{bupa} are without overlap. Surprisingly, for all $n >4$, the $n$-orthoplex fails to be all-net; we provide explicit counterexamples to a conjecture in \cite{ddrw}. Thus only three regular polytopes stand with unresolved all-net traits: the 24-cell, 120-cell, and 600-cell. \begin{rem} We encourage the reader to explore a lovely interactive software \cite{zhang} by Sam Zhang that creates every net of the 4-cube, 4-simplex, and 4-orthoplex by drawing on its dual 1-skeleton. \end{rem} \begin{ack} We thank Nick Bail, Zihan Miao, Andy Nelson, and Joe O'Rourke for helpful conversations. The first author was partially supported from an endowment by the Fletcher Jones Foundation. \end{ack} \section{Unfolding the Simplex} \label{s:simplex} \subsection{Lists and Chains} We explore ridge unfoldings of a convex polytope $P$ by focusing on the combinatorics of the arrangement of its facets in the unfolding. In particular, a ridge unfolding induces a spanning tree in the 1-skeleton of the dual of $P$ : a tree whose nodes are the facets of the polytope and whose edges are the uncut ridges between the facets \cite{sh2}. Our focus throughout this paper will be on the $n$-simplex and the $n$-orthoplex, both of whose facets are $(n-1)$-simplices. We begin by studying paths in the 1-skeleton, corresponding to a chain of unfolded simplicial facets. \begin{defn} A \emph{list} $\lst=\langle a_1, a_2, \dots, a_k \rangle$ is a sequence of numbers from $\{1,\ldots, n\}$ (possibly with repeats) where with no number is listed twice in a row. \end{defn} Label the vertices of the $(n-1)$-simplex $S$ with the numbers $1,\ldots, n$. Given a list $\lst$ with $k$ elements, we construct a chain $C(\lst$) of $k+1$ simplices from the list as follows: Starting with $S = S_1$, attach a simplex $S_2$ to $S_1$ on the facet of $S_1$ that is opposite vertex $a_1$. Note that all but one of the vertices of $S_2$ will inherit a label from $S_1$ and we label the remaining one $a_1$. Attach a third simplex $S_3$ to $S_2$ on the facet opposite vertex $a_2$, and extend the labeling from $S_2$ to $S_3$ as before, and continue in this matter until the list is exhausted. Figure~\ref{f:lists} shows this process in action for the list $\langle 3,2,3 \rangle$, creating a chain of four 2-simplices. \begin{figure}[h] \includegraphics[width=.9\textwidth]{lists} \caption{The chain of simplices assembled from the list $\langle 3,2,3\rangle$.} \label{f:lists} \end{figure} \subsection{Matrix Coordinates} Beyond this combinatorial formation, we introduce a coordinate system to capture the geometry. Begin by placing the $n$ vertices of the $(n-1)$-simplex $S$ at the standard basis vectors $e_i$ of $\R^n$. Note that the coordinates of its vertices are recorded as the column vectors of the $n\times n$ identity matrix. The rest of the chain is then placed in the hyperplane $x_1 + \cdots +x_n=1$ by a sequence of reflections. Let $\rho$ denote the reflection of $S$ across its facet opposite the vertex (say $v$) labeled with number $a_1$. Thus, $\rho$ fixes all vertices except for $v$; see Figure \ref{f:reflection}. \begin{figure}[h] \includegraphics[width=.9\textwidth]{reflection} \caption{The reflection of the vertex across the opposite face.} \label{f:reflection} \end{figure} To calculate the coordinate of $\rho(v)$ in $\R^n$, we first find the center $\sigma$ of the facet opposite $v$: $$\sigma =\left (\frac{1}{n-1}, \ldots, \ 0,\ \ldots, \frac{1}{n-1} \right ),$$ where $0$ occurs in the $a_1$-{th} coordinate. Since $\sigma$ bisects the segment from $v$ to $\rho(v)$, \begin{align*} \rho(v) & = \ v + 2 \ \overrightarrow{v \, \rho(v)} \\ & = \ \left ( 0, \ldots, 1, \ldots 0 \right ) \ + \ 2 \left ( \frac{1}{n-1}, \ldots, -1, \ldots, \frac{1}{n-1} \right ) \\ &= \ \left ( \frac{2}{n-1}, \ldots, -1, \ldots, \frac{2}{n-1} \right ), \end{align*} where the $-1$ occurs in the $a_1$-{th} coordinate. Hence the reflection $\rho$ is given by a matrix $M_{a_1}$, which is the identity except for $\rho(v)$ in the $a_1$-{th} column. Thus the coordinates of the $i$-th vertex of $S_2$ are recorded in the $i$-th column $v_i$ of $N_1 = M_{a_1}$. By change of coordinates, its image under the reflection from $S_2$ to $S_3$ is \[ N_1 M_{a_2} N_1^{-1} v_i = N_1 M_{a_2} e_i \,,\] and thus, the coordinates of the $i$-th vertex of $S_3$ are recorded in the $i$-th column of $N_2 = N_1 M_{a_2}$. Note that because $M_{a_2}$ affects only the $a_2$ column, $N_1$ and $N_2$ differ only in the $a_2$ column. Continuing in this way, the vertices of $S_{k+1}$ are recorded as the columns of $N_k = N_{k-1} M_{a_k}$. \begin{exmp} For two-dimensional chains (where $n=3$), the reflection matrices are \[ M_1 = \begin{bmatrix} -1 & 0 & 0 \\ 1 & 1 & 0 \\ 1 & 0 & 1 \end{bmatrix}, \quad M_2 = \begin{bmatrix} 1 & 1 & 0 \\ 0 & -1 & 0 \\ 0 & 1 & 1 \end{bmatrix}, \quad M_3 =\begin{bmatrix} 1 & 0 & 1 \\ 0 & 1 & 1 \\ 0 & 0 & -1 \end{bmatrix}. \] For the chain generated by $L=\langle 3,2,3 \rangle$, the coordinates of the vertices of the second, third, and fourth simplices are the column vectors of the matrices \[N_1 = M_3 = \begin{bmatrix} 1 & 0 & 1 \\ 0 & 1 & 1 \\ 0 & 0 & -1 \end{bmatrix}, \quad N_2 = N_1M_2= \begin{bmatrix} 1 & 2 & 1 \\ 0 & 0 & 1 \\ 0 & -1 & -1 \end{bmatrix}, \quad N_3 = N_2M_3 = \begin{bmatrix} 1 & 2 & 2 \\ 0 & 0 & -1 \\ 0 & -1 & 0 \end{bmatrix}. \] We remark that these are the barycentric coordinates of the vertices relative to the first simplex. \end{exmp} \subsection{Simplex Unfolding} An $n$-simplex has $n+1$ facets, and each is adjacent to every other. Thus, any listing of the facets (without repeat) describes a chain. However, because the full symmetric group acts transitively on the simplex, there is only one possibility: \begin{lem} \label{l:congruent} A chain of length $k+1$ in an unfolding of an $n$-simplex is congruent to $C(\langle 1,2,\ldots, k\rangle)$. \end{lem} \hide \begin{proof} Let $C$ be a chain with length $k+1$. Number the facets of the $n$-simplex: $0$ for the first facet in the chain, $1$ for the second, and so on. If $k<n$, the remaining facets can be numbered arbitrarily. Then label the vertices so that vertex $i$ is opposite facet $i$. Those vertex labels descend to the unfolded chain. They agree with the vertex labeling of the $0^{th}$ facet of $C(\langle 1\ldots, k \rangle)$, however not necessarily with subsequent facets, as shown in fig ..?.. \xxx{S: I think we probably do need a picture here to show what is going on.} The process of of unfolding the chain and the process of attaching facets to form $C(\langle 1,\ldots,k\rangle)$ are the same as long as the labels on the vertices that govern the procedure agree. Since each number in the chain occurs only once, the only vertices that are referenced are in the $0^{th}$ facet, so the two chains are the same. \end{proof} When $C(\langle 1,\ldots,k\rangle)$ is embedded in $\R^n$ as described above, the coordinates of the vertices of its facets are the columns of the matrices $N_i = N_{i-1} M_i$, where $N_0=I$. Furthermore, the points in the interior of a facet are linear combinations of its vertex coordinates, $c_1 v_1 + \cdots + c_n v_n$, where each $c_i$ is positive, and $c_1+\cdots +c_n=1$. Therefore, to see whether this chain overlaps itself, we need to look carefully at the values in those matrices. \begin{lem} \label{l:positive} The first row of $N_k$ contains no positive terms when $k>0$. \end{lem} \begin{proof} We prove the more precise statement that the $(1,j)$-entry is negative if $j\le k$ and zero if $j>k$. The proof is by induction and the base case is apparent. Assuming the result for $N_{l-1}$, the product $N_l=N_{l-1} M_l$ has all the same entries as $N_{l-1}$ except in the $l$ column. Hence the $(1,j)$-entries remain negative when $j<l$ and zero when $j>l$. If $a_{ij}$ denotes the $(i,j)$-entry of $N_{l-1}$, then the $(1,l)$-entry of $N_l$ (from the dot product of row 1 of $N_{l-1}$ and column $l$ of $M_l$) is given by $$ \bigg( a_{1,1}, \ \ldots, \ a_{1,l-1}, \ 0, \ \ldots, \ 0 \bigg) \ \cdot \ \left ( \frac{2}{n-1}, \ \ldots, \ -1, \ \ldots, \ \frac{2}{n-1} \right ), $$ where the $-1$ occurs in the $l$ coordinate. This becomes $$ a_{1,1} \ \bigg(\frac{2}{n-1}\bigg) \ + \ \cdots \ + \ a_{1,l-1} \ \bigg(\frac{2}{n-1}\bigg) \ + \ 0 \ (-1) \ + \ 0 \ \bigg(\frac{2}{n-1}\bigg) \ + \ \cdots \ + \ 0 \ \bigg(\frac{2}{n-1}\bigg) \,, $$ which is clearly negative. \end{proof} \begin{thm} Every unfolding of the $n$-simplex yields a net. \end{thm} \begin{proof} Assume otherwise, where there exists an unfolding of the simplex with two overlapping simplicial facets. The path (in dual 1-skeleton) connecting these two facets forms a chain of simplices. By Lemma~\ref{l:congruent}, that chain is congruent to $C(\langle 1, 2,\ldots, k \rangle)$ for some $k$, where facet $f_0$ and facet $f_k$ must overlap. By construction, the interior points of facet $f_0$ have all positive coordinates. For facet $f_k$, Lemma~\ref{l:positive} guarantees that there are no positive entries in the first row; hence, all interior points of $f_k$ must have a negative first coordinate. Thus, the two facets cannot intersect, resulting in a contradiction. \end{proof} \section{Orthoplex Combinatorics and Geometry} \label{s:orthoplex} \subsection{Valid Lists} In contrast to the simplex, both unfoldings of the $n$-orthoplex and the chains within these unfoldings exhibit considerable variety. Unfoldings of the $n$-orthoplex are in bijection with spanning trees of the 1-skeleton of the $n$-cube. Consider the following approach to record paths on this skeletal structure: Position the $n$-cube with antipodal vertices at $(0, \dots, 0)$ and $(1, \dots, 1)$. A path along the edges of this cube is encoded as a list of binary numbers where exactly one digit changes from one entry to the next.\footnote{This list of binary numbers is called a \emph{Gray code}.} For our work, our list $\lst$ simply records the digit entry that changes in moving from one vertex to another. Thus by duality, the ridges of the orthoplex inherit these labels and the process of unfolding the chain corresponds to the construction of $C(\lst)$. \begin{exmp} Consider the Gray code $\langle 101, 100, 110, 111\rangle$ associated to the list $\langle 3, 2, 3\rangle$. Figure~\ref{f:codes}(a) shows the path on four vertices of the cube, (b) corresponding to four adjacent facets of the octahedron, (c) resulting in a partial chain unfolding. Compare to Figure~\ref{f:lists}. \end{exmp} \begin{figure}[h] \includegraphics[width=.9\textwidth]{codes} \caption{Path on the 3-cube and a partial unfolding of the octahedron.} \label{f:codes} \end{figure} \begin{rem} Up to symmetry, there are just three spanning paths on the 1-skeleton of the $3$-cube: $\langle 1, 2, 1, 3, 1, 2, 1\rangle$, $\langle 1,2,1,3,2,1,2 \rangle$, and $\langle 1,2,3,2,1,2,3\rangle$, corresponding to the first three highlighted nets shown in Figure~\ref{f:3dnet}. The situation escalates rapidly as $n$ increases: there are 238 spanning paths on the $4$-cube and 48,828,036 on the $5$-cube \cite{oeis}. \end{rem} \begin{defn} A list of numbers from $\{1, \dots, n\}$ is \emph{valid} if it corresponds to a path on the $n$-cube. \end{defn} \noindent Note that every path naturally yields a valid list, and given the starting position of the path, every valid list yields a path. The following provides a combinatorial check for list validity: \begin{lem}\label{lem:validCheck} A list is valid if and only if it contains no sublist of consecutive entries in which each entry occurs an even number of times. \end{lem} \begin{proof} For this proof, we position the $n$-cube with antipodal vertices at $(-1, \dots, -1)$ and $(1, \dots, 1)$. Consider a list $\lst = \langle a_1, \cdots, a_k\rangle$ and label the vertices of the path it generates as $v_1, \ldots, v_{k+1}$, starting from $v_1=(1,1,\ldots,1)$. Then $v_{i+1}$ is the image of $v_i$ under the reflection across the hyperplane $x_{a_i}=0$. This reflection is given by the $n \times n$ diagonal matrix $M_i$ whose $(i,i)$-entry is $-1$, and whose remaining diagonal entries are $1$. Suppose $\lst$ contains a sublist $a_j, a_{j+1}, \ldots, a_l$ in which all entries occur an even number of times. Then, because each $M_i$ is its own inverse and they all commute, $$M_l \cdot M_{l-1} \cdots M_j \cdot v_{j-1} \ = \ I \cdot v_{j-1} \ = \ v_{j-1} \,.$$ Thus, the path is truly a loop and the list is invalid. Conversely, if the sequence describes a loop, so that $v_j=v_k$ for some $j \ne k$, then $v_j$ is a fixed point of $$N \ = \ M_{l-1} \cdot M_{l-2} \cdot \ \cdots \ \cdot \ M_{j+1}\, .$$ Since $N$ is a diagonal matrix, it cannot have fixed points of the form $(\pm 1, \pm 1, \ldots, \pm 1)$ unless $N$ is the identity matrix. Thus, $M_i$ must occur an even number of times in the product. \end{proof} \begin{rem} With the characterization given by this lemma, it is straightforward to create an algorithm to build valid lists: recursively append numbers $\{1,\dots, n\}$ and check whether any of the new consecutive sublists have entries that occur an even number of times. \end{rem} \subsection{Centroids} The question of whether two facets overlap depends on how close they are to each other, which can be estimated by calculating the distance between their centroids. If the vertices are $v_i = (a_{i1}, \ldots a_{in})$, the centroid is found by averaging their coordinates: $$\left (\ \frac{1}{n}\sum a_{1j}, \ \ldots, \ \frac{1}{n}\sum a_{nj} \ \right ) \,.$$ \begin{lem} \label{l:centroid} Let $d$ denote the distance between the centroids of two $(n-1)$-simplex facets of the $n$-orthoplex in an unfolding. If $d<2/\sqrt{n(n-1)}$, the facets must intersect. If $d>2\sqrt{(n-1)/n}$, the facets cannot intersect. \end{lem} \begin{proof} The closest point to the centroid on the surface of the facet is the midpoint of a face. Since all the facets are identical, we use the initially placed facet to compute the distance: The centroid of the $(n-1)$-simplex is at $(1/n,\ldots,1/n)$ and the center of one of the faces is at $$(1/(n-1), \ \ldots, \ 1/(n-1), \ 0) \,.$$ The distance between them is \[ \sqrt{\left (\frac{1}{n}-\frac{1}{n-1}\right )^2(n-1)+\left (\frac{1}{n}-0\right )^2} =\frac{1}{\sqrt{n(n-1)}} \ .\] If centroids are separated by less than double that amount, the facets must overlap. The farthest point from the centroid to the surface of the facet is a vertex, one of which is $(0,\ldots,0,1)$. The distance between them is \[ \sqrt{ \left ( \frac{1}{n}-0 \right )^2(n-1) + \left ( \frac{1}{n}-1 \right )^2} = \sqrt{\frac{n-1}{n}} \ . \] If centroids are separated by more than double that amount, the facets cannot overlap. \end{proof} \section{Orthoplex Unfolding} \label{s:orthonet} \subsection{Path Extensions} This section proves that the $4$-orthoplex is all-net. We do this by extending paths on the skeleton of the $4$-cube. While any path along a $3$-cube can always be extended to a spanning path, this is not true for $n \geq 4$. For example, Figure~\ref{f:4-path}(a) shows the 1-skeleton of the 4-cube, and the blue path shown in (b) cannot be extended further. However, there is a partial result. \begin{figure}[h] \includegraphics[width=.9\textwidth]{4cube-path} \caption{A path in the 4-cube that cannot be further extended.} \label{f:4-path} \end{figure} \begin{lem} \label{l:8less} A path on skeleton of the $4$-cube can be extended to connect at least nine vertices. \end{lem} \begin{proof} Let $s$ be the starting point and $e$ the ending point of a path that cannot be extended further. Let $V_s$ be the set of the four vertices adjacent to $s$, and $V_e$ be the set of the four vertices adjacent to $e$. Since the path cannot be extended, it must already pass through all the vertices in $V_e$ and $V_s$. If $s$, $e$, and the vertices of $V_s$ and $V_e$ are all distinct, then the path connects at least ten vertices. Even if a pair of listed vertices coincide, the path connects at least nine. However, there are two ways in which it is possible for $\{s\} \cup \{e\} \cup V_s \cup V_e$ to contain only eight vertices. First, if $s$ and $e$ are adjacent, then $e \in V_s$ and $s \in V_e$. In this case, without loss of generality, we may assume the path starts at $(0, 0, 0, 0)$ and ends at $(1, 0, 0, 0)$. Then \begin{align*} V_e &= \{(0,1,0,0), (0,0,1,0), (0,0,0,1)\}, \\ V_s &= \{(1,1,0,0), (1,0,1,0), (1,0,0,1)\}. \end{align*} However, a simple check shows that there is no path containing only these eight points. In fact, no such path can contain more than four of them. Second, if $s$ and $e$ are opposite corners of a codimension two (square) face, then $V_s$ and $V_e$ share two vertices. Color a vertex red if the sum of its coordinates is even, and blue if it is odd. Any path connecting $s$ and $e$ will pass through nodes that alternate in color. Since $s$ and $e$ are opposite corners of a square, they must be the same color (say, red). Then all six vertices of $V_s$ and $V_e$ must be blue. Thus, a path that passes through all six of them must also pass through at least five red vertices, yielding a path with length greater than nine. \end{proof} \subsection{Unfolding the $4$-orthoplex} Rephrasing Lemma~\ref{l:8less} differently, we can say that any valid list can be extended to a valid list with at least eight entries. We use this to begin unfolding the $4$-orthoplex. \begin{lem} \label{l:eight} Every valid list containing exactly eight entries unfolds to form a partial net of the $4$-orthoplex. \end{lem} \begin{proof} Using the check for valid lists described in Lemma~\ref{lem:validCheck}, there are 128 valid lists. Taking advantage of reverse symmetry\footnote{The unfolding given by $\langle 1,2,1,3\rangle$ is congruent to that given by $\langle 3,1,2,1\rangle$, which can be renumbered $\langle 1,2,3,2\rangle$.} reduces the number to 66 (where four lists are their own reverses). One such is shown in Figure~\ref{f:8length}. By direct inspection, none of them self-intersect. \end{proof} \begin{figure}[h] \includegraphics[width=.5\textwidth]{path-ortho} \caption{The partial unfolding corresponding to a valid list with length eight.} \label{f:8length} \end{figure} \begin{lem}\label{l:8more} If two facets of the $4$-orthoplex are separated by eight or more facets, they cannot overlap. \end{lem} \begin{proof} This is verified computationally by generating all valid lists with length nine or more facets and then computing distances between centroids of facets that have eight or more facets between them. By Lemma~\ref{l:centroid}, two facets do not overlap when the distance between their centroids is greater than $\sqrt{3}$, which happens in every case. \end{proof} Buekenhout and Parker \cite{bupa} enumerate 110,912 ridge unfoldings of the 4-orthoplex. The following guarantees all of them to be valid nets. \begin{thm} The 4-orthoplex is all-net. \end{thm} \begin{proof} If there were an unfolding of the 4-orthoplex that did not yield a net, then there would be a path between two of its overlapping facets. By Lemma~\ref{l:8more}, those facets must be separated by fewer than eight intervening facets along the path, corresponding to a valid list $\lst$ whose length is at most eight. By Lemma~\ref{l:8less}, that list can be extended to one whose length is exactly eight. As described in Lemma~\ref{l:eight}, none of the unfolds generated by these lists exhibit overlap. \end{proof} \subsection{Higher dimensions} Although the $n$-cube is all-net \cite{ddrw}, it is surprising that its dual does not inherit this property: \begin{thm} For each $n>4$, the $n$-orthoplex is not all-net. \end{thm} \begin{proof} We first consider case-by-case analysis of dimensions 5 through 9 using centroid arguments. Consider the valid list $$\langle 1, 2, 1, 3, 1, 2, 1, 4, 1, 2, 3, 5, 1, 2, 3, 2, 5, 3, 2, 1, 5, 4, 3, 4, 2, 4, 1, 2, 3, 2, 1\rangle$$ that captures the path unfolding of all 32 facets of the 5-orthoplex. The centroid of its last facet is approximately \[ ( 0.22687086, 0.04417632, 0.36540937, 0.12942886, 0.23411458) \,,\] which is a distance of approximately $0.24188305539$ from the centroid of the first facet, $$(0.2,0.2,0.2,0.2,0.2) \,.$$ By Lemma~\ref{l:centroid}, since the centroids are separated by less than $1/\sqrt{5}$, their facets overlap. In dimension 6, there is an even shorter chain (16 facets) that fails the centroid test, given by \[ \langle 1,2,3,1,4,5,4,3,5,4,1,3,2,1,4 \rangle. \] In dimensions 7 and 8, the shortest chain (13 facets) that fails the centroid test is given by \[ \langle 1,2,3,4,1,5,3,5,4,3,2,1 \rangle. \] For dimension 9, there is an even shorter chain (10 facets) given by $\langle 1,2,3,4,2,4,1,2,3 \rangle.$ For dimensions $n>9$, we focus on the unfolding corresponding to the list $\langle 1,2,3,4,2,4,1,2,3 \rangle$ above but the centroid argument will not be utilized. Recall that the unfolding of the orthoplex occurs in the hyperplane $\sum x_i=1$. The first facet is the intersection of this hyperplane with the positive orthant. Let $$v= \Bigl< \ \frac{1}{n-1} \ ,\ 0 \ ,\ \frac{1}{n-1}, \ \ldots, \ \frac{1}{n-1} \ \Bigr> \,,$$ the midpoint of a ridge of the first facet. We show that its image under the unfolding, which is on a ridge of the tenth facet, has all positive coordinates. Therefore it is a point of intersection of the first and tenth facets. To do this, let $x=2/(n-1)$. Then each $M_i$ can be written in terms of $x$, and $$v= \Bigl< \ \frac{x}{2} \ ,\ 0 \ ,\ \frac{x}{2}, \ \ldots, \ \frac{x}{2} \ \Bigr> \,.$$ The matrix product $$M_1 \cdot M_2 \cdot M_3 \cdot M_4 \cdot M_2 \cdot M_4 \cdot M_1 \cdot M_2 \cdot M_3 \cdot v$$ is then a vector $\langle p_1(x), p_2(x), \ldots, p_n(x) \rangle$ whose entries are polynomials in $x$. Note that although the variable $x$ depends on $n$, the polynomials themselves are the same for all $n$. Using {\tt Sympy} \cite{sympy}, these were calculated to be \begin{align*} p_1(x) &= 0.5 x^9 - 3 x^8 - 7 x^7 - 8 x^6 - 5 x^5 - 1.5 x^4 + x^3 + x^2 + 0.5x \\ p_2(x) &= 0.5x^{10} + 3 x^9 + 6.5 x^8 + 5.5 x^7 + 0.5 x^6 - 1.5 x^5 - 0.5 x^4 + x^3 + x^2 \\ p_3(x) &= 0.5 x^{10} + 3.5 x^9 + 9.5 x^8 + 12 x^7 + 6 x^6 - 2 x^5 - 5.5 x^4 - 4 x^3 - x^2 + 0.5x \\ p_4(x) &= 0.5 x^{10} + 3.5 x^9 + 10 x^8 + 14.5 x^7 + 10.5 x^6 + 2 x^5 - 2.5 x^4 - 1.5 x^3 + 0.5x \end{align*} and for $i \ge 5$, \[ p_i(x) = 0.5 x^{10} + 3.5 x^9 + 10 x^8 + 15 x^7 + 13 x^6 + 6.5 x^5 + x^4 - 0.5 x^3 + 0.5x. \] The lowest degree term, which determines the behavior near zero, is positive in each. Thus, for sufficiently small $x$, each is positive. In fact, their graphs show that they are all positive when $0 \le x \le 0.2278$. Since $x=2/(n-1)$, this corresponds to all $n>9$. \end{proof} \begin{rem} The shortest chains that fail the centroid test in dimension 5 have length 20. One is given by the list $\langle 1,2,3,4,2,1,5,4,2,4,5,4,2,1,5,4,3,1,5 \rangle$. It is possible that there are shorter chains that overlap even though their centroids are not sufficiently close to guarantee it. \end{rem} \begin{rem} For dimensions $n>9$, the list $\langle 1,2,3,4,2,4,1,2,3 \rangle$ provides a 10-length chain of facets that overlaps. It would be interesting to discover if this is the shortest length chain with this property. \end{rem} \subsection{Closing} Horiyama and Shoji \cite{hosh} showed that the five regular polytopes in 3D (the Platonic solids) are all-net. Three classes of regular polytopes exist for all dimensions: the $n$-simplex, $n$-cube, and $n$-orthoplex. The $n$-simplex and $n$-cube are all-net, but the $n$-orthoplex fails (except for $n=4$). There are only three additional regular polytopes whose all-net property has not been studied, all of which are four-dimensional: the 24-cell, 120-cell, and 600-cell. The number of distinct unfoldings of these three polytopes are enumerated in \cite{bupa}: \begin{align*} {\text{24-cell}} & \ : \ \ 6 \ (2^{19} \cdot 5688888889 \ + \ 347) \\ {\text{120-cell}} & \ : \ \ 2^7 \cdot 5^2 \cdot 7^3 \ (2^{114} \cdot 3^{78} \cdot 5^{20} \cdot 7^{33} \ + \ 2^{47} \cdot 3^{18} \cdot 5^2 \cdot 7^{12} \cdot 53^{5} \cdot 2311^3 \ + \ 239^2 \cdot 3931^2) \\ {\text{600-cell}} & \ : \ \ 2^{188} \cdot 3^{102} \cdot 5^{20} \cdot 7^{36} \cdot 11^{48} \cdot 23^{48} \cdot 29^{30} \end{align*} Unlike the other three regular polytopes of this dimension (4-simplex, 4-cube, 4-orthoplex), these enormous unfolding numbers point us towards the following claim: \begin{conj} The 24-cell, the 120-cell, and the 600-cell fail to be all-net. \end{conj} Easing the condition of regularity for 3D polyhedra move us away from Platonic solids to the Archimedean solids. However, it is known that several of these polyhedra (such as the snub dodecahedron, truncated icosahedron, truncated dodecahedron, rhombicosidodecahedron, and truncated icosidodecahedron) fail to be all-net \cite{hosh}. It would be interesting to find conditions for polyhedra, and polytopes in general, that guarantee the all-net property. \bibliographystyle{amsplain}
{ "timestamp": "2021-11-03T01:10:23", "yymm": "2111", "arxiv_id": "2111.01359", "language": "en", "url": "https://arxiv.org/abs/2111.01359", "abstract": "Over a decade ago, it was shown that every edge unfolding of the Platonic solids was without self-overlap, yielding a valid net. We consider this property for regular polytopes in arbitrary dimensions, notably the simplex, cube, and orthoplex. It was recently proven that all unfoldings of the $n$-cube yield nets. We show this is also true for the $n$-simplex and the $4$-orthoplex but demonstrate its surprising failure for any orthoplex of higher dimension.", "subjects": "Computational Geometry (cs.CG); Combinatorics (math.CO)", "title": "Unfoldings and Nets of Regular Polytopes", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9891815535796247, "lm_q2_score": 0.8267117940706734, "lm_q1q2_score": 0.8177680568214275 }
https://arxiv.org/abs/2003.06340
On Alignment in Deep Linear Neural Networks
We study the properties of alignment, a form of implicit regularization, in linear neural networks under gradient descent. We define alignment for fully connected networks with multidimensional outputs and show that it is a natural extension of alignment in networks with 1-dimensional outputs as defined by Ji and Telgarsky, 2018. While in fully connected networks, there always exists a global minimum corresponding to an aligned solution, we analyze alignment as it relates to the training process. Namely, we characterize when alignment is an invariant of training under gradient descent by providing necessary and sufficient conditions for this invariant to hold. In such settings, the dynamics of gradient descent simplify, thereby allowing us to provide an explicit learning rate under which the network converges linearly to a global minimum. We then analyze networks with layer constraints such as convolutional networks. In this setting, we prove that gradient descent is equivalent to projected gradient descent, and that alignment is impossible with sufficiently large datasets.
\section{Introduction} \label{sec: Introduction} Although overparameterized deep networks can interpolate randomly labeled training data \cite{DuOptimizesOverparameterizedNetworks, DuAdaptiveMethods}, training overparameterized networks with modern optimizers often leads to solutions that generalize well. This suggests that there is a form of \textit{implicit regularization} occurring through training \cite{RethinkingGeneralization}. As an example of implicit regularization, the authors in \cite{LogisticAlignment} proved that the layers of linear neural networks used for binary classification on linearly separable datasets become aligned in the limit of training. That is, for a linear network parameterized by the matrix product $W_d, W_{d-1}, \ldots W_1$, the top left/right singular vectors $u_i$ and $v_i$ of layer $W_i$ satisfy $|v_{i+1}^T u_i| \rightarrow 1$ as the number of gradient descent steps goes to infinity. Alignment of singular vector spaces between adjacent layers allows for the network representation to be drastically simplified (see Equation \eqref{eq: aligned_network}); namely, the product of all layers becomes a product of diagonal matrices with the exception of the outermost unitary matrices. If alignment is an invariant of training, then optimization over the set of weight matrices reduces to optimization over the set of singular values of weight matrices. Thus, importantly, alignment of singular vector spaces allows for the gradient descent update rule to be simplified significantly, which was used in \cite{LogisticAlignment} to show convergence to a max-margin solution. In this work, we generalize the definition of \emph{alignment} to the multidimensional setting. We study when alignment can occur and moreover, under which conditions it is an invariant of training in linear neural networks under gradient descent. Our main contributions are as follows: \begin{enumerate} \item We extend the definition of alignment from the 1-dimensional classification setting to the multi-dimensional setting (Definition~\ref{def: Alignment for 2 matrices}) and characterize when alignment is an invariant of training in linear fully connected networks with multi-dimensional outputs (Theorem~\ref{thm: Alignment for fully connected networks with multi-dimensional output}). \item We demonstrate that alignment is an invariant for fully connected networks with multidimensional outputs only in special problem classes including autoencoding, matrix factorization and matrix sensing. This is in contrast to networks with 1-dimensional outputs, where there exists an initialization such that adjacent layers remain aligned throughout training under \textit{any} real-valued loss function and any training dataset (Proposition~\ref{prop: Alignment for 1 dimensional outputs}). \item Alignment largely simplifies the analysis of training linear networks: We provide an explicit learning rate under which gradient descent converges linearly to a global minimum under alignment in the squared loss setting (Proposition~\ref{prop: linear convergence}). \item We prove that alignment cannot occur, let alone be invariant, in networks with constrained layer structure (such as convolutional networks), when the amount of training data dominates the dimension of the layer structure (Theorem \ref{thm: No Alignment in constrained layers}). \item We support our theoretical findings via experiments in Section \ref{sec: Experiments}. \end{enumerate} As a consequence, our characterization of the invariance properties of alignment provides settings under which the gradient descent dynamics can be simplified and the implicit regularization properties can be fully understood, yet also shows that further results are required to explain implicit regularization in linear neural networks more generally. \section{Related Work} \label{sec: Related Work} Implicit regularization in overparameterized networks has become a subject of significant interest \cite{ImplictRegularizationBach, GunasekarImplicitBiasOptimizationGeometry, GunasekarImplicitRegularizationLinearConvNetworks, ImplicitSelfRegularization, neyshabur2014search}. In order to characterize the specific form of implicit regularization, several works have focused on analyzing deep \emph{linear} networks \cite{DeepMatrixFactorization, GunasekarImplicitRegularizationLinearConvNetworks, MatrixFactorizationNIPS2017, GunasekarLinearlySeparableData}. Even though such networks can only express linear maps, parameter optimization in linear networks is non-convex and is studied in order to obtain intuition about optimization of deep networks more generally. One such form of implicit regularization is \emph{alignment}, identified by \cite{LogisticAlignment} in their analysis of linear fully connected networks with 1-dimensional outputs trained on linearly separable data. They proved that in the limit of training, each layer, after normalization, approaches a rank $1$ matrix, i.e. \begin{align*} \lim_{t\rightarrow \infty} \frac{W_i^{(t)}} { \|W_i^{(t)}\|_F} = u_i v_i^T \end{align*} and that adjacent layers, $W_{i+1}$ and $W_i$ become \textit{aligned}, i.e. $|v_{i+1}^T u_i| \rightarrow 1.$ In addition, \cite{LogisticAlignment} proved that alignment in this setting occurs concurrently with convergence to the max-margin solution. Follow-up work mainly focused on this convergence phenomenon and gave explicit convergence rates for overparameterized networks trained with gradient descent \cite{AroraTwoLayerNetwork, SGDOptimization}. While the connection to alignment was not mentioned in their work, the authors in ~\cite{ImplictRegularizationBach} begin to generalize alignment to the case of multidimensional outputs. In particular, they consider two-layer networks initialized so that the two layers are aligned with each other and where both of the layers are aligned with the data. We generalize this to networks of any depth, showing that our definition of alignment corresponds to the initialization considered in \cite{ImplictRegularizationBach}. Moreover, we establish necessary and sufficient conditions for when alignment is an invariant of training in Theorem \ref{thm: Alignment for fully connected networks with multi-dimensional output} instead of assuming these conditions as in \cite{ImplictRegularizationBach}. Furthermore, their result on sequential learning of components is one of a variety of results which can be derived via our singular value update rule presented in Corollary \ref{cor:singular-update}. Balancedness is another closely related form of implicit regularization in linear neural networks. It was introduced in~\cite{Balancedness} and defined as the property that if $W_i^TW_i = W_{i+1}W_{i+1}^T$ for all $i$ at initialization, then this property is \emph{invariant} through gradient flow. \cite{BalancednessJasonLee} presented a more general form, which is that the difference $W_i^TW_i - W_{i+1}W_{i+1}^T$ is constant through gradient flow. In practice, however, analyses are based on this quantity being close to or exactly zero (in Frobenius norm). In this exact setting, the connection between balancedness and alignment becomes clear since balancedness implies alignment of singular vector spaces between consecutive layers. To study gradient descent, slightly more general notions such as approximate balancedness \cite{ApproximateBalancedness} and $\epsilon$-balancedness have been introduced. \cite{BalancednessJasonLee} also defined balancedness with respect to convolutional networks, showing that under gradient flow, the difference in the norm of the weights of consecutive layers is an invariant. Generally, the goal of identifying invariants of training such as balancedness or alignment is to help understand both the dynamics of training and properties of solutions at the end of training. \section{Definition of Alignment in the Multi-dimensional Setting} \label{sec: Definition of alignment in multi-dimensional setting} In this section, we first define alignment for linear neural networks with multi-dimensional outputs. We then define when alignment is an invariant of training. We consider \emph{linear} neural networks. Let $f: \mathbb{R}^{k_0} \rightarrow{} \mathbb{R}^{k_{d}}$ denote such a $d$-layer network, i.e. \begin{align} \label{eq: Network Representation} f(x) &= W_d W_{d-1}\ldots W_1 x, \end{align} where $W_i \in \mathbb{R}^{k_{i} \times k_{i-1}}$ for $i \in [d]$, where we follow the convention that $[d]=\{1, 2, \ldots d \}$. Let $(X, Y) \in \mathbb{R}^{k_0 \times n} \times \mathbb{R}^{k_{d} \times n}$ denote the set of training data pairs $\{(x^{(i)}, y^{(i)})\}$ for $i \in [n]$. Gradient descent with learning rate $\gamma$ is used to find a solution to the following optimization problem: \begin{align} \arg\min_{f \in \mathcal{F}} \; \frac{1}{2n} \displaystyle\sum\limits_{i=1}^{n} \ell(f(x^{(i)}), y^{(i)}), \end{align} where $\mathcal{F}$ is the set of linear functions represented by $f$ and $\ell$ is a real-valued loss function. When not stated otherwise, we assume $\ell(f(x^{(i)}), y^{(i)}) = \|y^{(i)} - f(x^{(i)})\|^2_2$, which is the squared loss (MSE). In addition, we denote by $W_i^{(t)}$ for ${t \in \mathbb{Z}_{\geq 0}}$ the weight matrix $W_i$ after $t$ steps of gradient descent. When there are no additional constraints on the matrices $W_i$, then $f$ is a fully connected network. We next introduce a generalized form of the singular value decomposition: \begin{definition} An \textbf{unsorted, signed singular value decomposition (usSVD)} of a matrix $A \in \mathbb{R}^{m \times n}$ is a triple $U \in \mathbb{R}^{m \times m},\Sigma \in \mathbb{R}^{m \times n}, V \in \mathbb{R}^{n \times n}$ such that $U, V$ are orthonormal matrices, $\Sigma$ is diagonal, and $A = U \Sigma V^T$. \end{definition} In contrast to the usual definition of singular value decomposition (SVD) of a matrix, the diagonal entries of $\Sigma$ may be in any order and take negative values. Throughout, we will refer to the entries of $\Sigma$ in a usSVD as singular values and the vectors in $U, V$ as singular vectors. Using the usSVD, we now generalize the notion of alignment from~\cite{LogisticAlignment} to the multi-dimensional setting. \begin{definition} \label{def: Alignment for 2 matrices} Let $f= W_d W_{d-1} \ldots W_1$ be a linear network. We say that $f$ is \textbf{aligned} if there exists a usSVD $W_i = U_i\Sigma_iV^T_i$ with $U_i = V_{i+1}$ for all $i \in [d-1]$. (We also say that a matrix $A$ is aligned with another matrix $B$ if there exist usSVD's $A = U_A\Sigma_Av_A^T, B = U_B\Sigma_BV_B^T$ such that $V_A = U_B$.) \end{definition} Note that if $W_i$ and $W_{i+1}$ are rank $1$ matrices in an aligned network $f$, then the inner product of the first columns of $V_{i+1}$ and $U_i$ is $1$ in absolute value. Hence Definition~\ref{def: Alignment for 2 matrices} is consistent with alignment in the 1-dimensional setting from \cite{LogisticAlignment}. We next define when alignment is an invariant of training for deep linear networks. Again, such invariants are of interest since they may provide insights into properties of trained networks and significantly simplify the dynamics of gradient descent. \begin{definition} \label{def: Invariance of Alignment} \textbf{Alignment is an invariant of training} for a linear neural network $f$ if there exists an initialization $\{W_j^{(0)}\}_{j=1}^{d}$ such that $W_1^{(\infty)}, W_{2}^{(\infty)}, \ldots , W_d^{(\infty)}$ achieves zero training error \footnote{The interpolation condition in this definition (i.e., achieving zero training error) is important in ruling out several architectures where the layers are trivially aligned. For example, if all layers are constrained to be diagonal matrices throughout training, then the layers are all trivially aligned, but cannot interpolate datasets where the target is not the product of a diagonal matrix with the input.} and for all gradient descent steps $t \in \mathbb{Z}_{\geq 0} \vspace{-0.2cm} \begin{enumerate} \item[(a)] the network $f$ is aligned; \item[(b)] $W_i^{(t)} = U_i \Sigma_i^{(t)} V_i^T$ for all $i \in \{2, \ldots d-1\}$, that is, $U_i, V_i$ are not updated; \vspace{-0.2cm} \item[(c)] $W_1^{(t)} = U_1 \Sigma_1^{(t)} {V_1^{(t)}}^T$ and $W_d^{(t)} = U_d^{(t)} \Sigma_d^{(t)} V_d^T$, that is, $U_1$ and $V_d$ are not updated. \end{enumerate} \vspace{-0.2cm} If additionally, $V_1$ and $U_d$ are not updated for any $t\in\mathbb{Z}_{\geq 0}$, then we say that \textbf{strong alignment is an invariant of training}. \end{definition} When alignment is an invariant of training, there are important consequences for training. In particular, note that when the network $f$ is aligned with usSVDs $W_i = U_i \Sigma_i V_i^T$ for all $1\leq i \leq d$, then \begin{align} \label{eq: aligned_network} f(x) = W_d\cdots W_1x = U_d \left(\prod\limits_{i=0}^{d-1}\Sigma_{d-i}\right) V_1^Tx. \end{align} Hence if alignment is an invariant of training, then the singular vectors of layers $2$ through $d-1$ are never updated and the analysis of gradient descent can be limited to the singular values of the layers and the matrices $V_1$ and $U_d$. \textbf{Remarks.} For the remainder of the paper, we assume that the gradient of the loss function at initialization $\{W_i^{(0)}\}_{i=1}^{d}$ is non-zero. Otherwise, training with gradient descent would not proceed. We also only consider datasets $(X,Y)$ for which there is a linear network that achieves loss zero. This is consistent with the assumptions in~\cite{LogisticAlignment}. \section{Alignment in Fully Connected Networks} \label{sec:Alignment in Fully Connected Networks} In this section, we first characterize when alignment is an invariant of training for fully connected networks (Theorem \ref{thm: Alignment for fully connected networks with multi-dimensional output}). In particular, we show that this is not the case in general. We then present special classes of problems for which alignment is an invariant of training, namely autoencoding, matrix factorization, and matrix sensing. In contrast, for a linear neural network with 1-dimensional outputs, we demonstrate that there exists an initialization for which the layers remain aligned throughout training given any dataset and any real-valued loss function. Finally, we discuss various consequences of alignment, including a proof of linear convergence of gradient descent to an interpolating solution. \subsection{Characterization of Alignment with Multi-dimensional Outputs} \label{subsec: Data Conditions for Alignment in FC Networks} Theorem~\ref{thm: Alignment for fully connected networks with multi-dimensional output} is one of our main results and characterizes when alignment is an invariant of training in a fully connected network with multi-dimensional outputs. To simplify notation, we consider the case when the layers are square matrices, i.e. $k_i = k_j$ for all $0\leq i,j \leq d$. The general result for non-square matrices is provided in Appendix~\ref{proof strong alignment non square}. \begin{theorem} \label{thm: Alignment for fully connected networks with multi-dimensional output} Let $f: \mathbb{R}^{k} \rightarrow \mathbb{R}^{k}$ be a linear fully connected network with $d \geq 3$ square layers of size $k>1$. Alignment is an invariant of training under the squared loss on a dataset $(X, Y)\in\mathbb{R}^{k\times n}\times \mathbb{R}^{k\times n}$ if and only if there exist orthonormal matrices $U, V \in \mathbb{R}^{k \times k}$ such that $U^TYX^TV$ and $V^TXX^TV$ are diagonal. \end{theorem} The full proof of this result is presented in Appendices \ref{sec: appendix outline}-\ref{sec: completing the proof}; here, we provide a proof sketch. \vspace{-0.2cm} \begin{proof}[Proof Sketch] The proof essentially follows by induction. For the base case, we initialize the layers $\{W_i\}_{i=1}^{n}$ to satisfy the conditions for alignment given in Definition \ref{def: Invariance of Alignment}. Assuming that these conditions hold at gradient descent step $t$, we prove that they hold at step $t + 1$. After substituting the alignment conditions into the gradient descent update equation for the squared loss at step $t+1$ and cancelling terms, we obtain that alignment is an invariant of training~if~and~only~if \begin{equation} \label{eq: Condition for alignment in fully connected matrices} {U_d^{(t)}}^T \displaystyle\sum\limits_{k=1}^{n} (y^{(k)} - f(x^{(k)})) {x^{(k)}}^T V_{1}^{(t)} \end{equation} is a diagonal matrix. By considering the update for $W_1^{(t)}$ and $W_d^{(t)}$, one sees that alignment implies strong alignment and so $U_d, V_1$ are also invariant across updates. Thus, let $U_d = U$ and $V_1 = V$. By expanding $f(x^{(k)})$ using \eqref{eq: aligned_network}, and considering the update across multiple timesteps, we obtain that the matrix in \eqref{eq: Condition for alignment in fully connected matrices} is diagonal if and only if $U^TYX^TV$ and $V^TXX^TV$ are diagonal. To complete the proof, we show in Appendix D that under strong alignment, gradient descent converges to a solution~with~zero~training~error. \end{proof} Theorem~\ref{thm: Alignment for fully connected networks with multi-dimensional output} implies that invariance of alignment throughout training holds only for special classes of problems. In particular, the above implies that alignment is an invariant of training when $X$ and $Y$ have the same right singular vectors, a very special condition on the data. Note that this corresponds to the $\epsilon=0$ data condition with the initialization considered in \cite{ImplictRegularizationBach}. In Section \ref{sec: Experiments}, we also provide empirical support showing that alignment is not an invariant of training for important tasks that violate the data condition presented here, such as multi-class classification. \subsection{Classes of Problems with Alignment} We next discuss classes of problems for which alignment is an invariant of training.\\\\ \textbf{Autoencoding:} In the case when $X = Y$, it holds that $U^TYX^TV = U^TXX^TV$. Taking $U = V$ to be the left singular vectors of $X$ satisfies the conditions of Theorem \ref{thm: Alignment for fully connected networks with multi-dimensional output}.\\\\ \textbf{Matrix Factorization and Inversion}: In the case of matrix factorization, we have that $X = I$. Hence taking $U$ and $V$ to be the left and right singular vectors of $Y$ respectively satisfies the conditions of Theorem \ref{thm: Alignment for fully connected networks with multi-dimensional output}. For matrix inversion, we have that $Y = I$ and we proceed analogously. \\\\ \textbf{Matrix Sensing.} Given pairs of observations $\{(M_i, y_i)\}_{i=1}^{n}$ with $M_i\in \mathbb{R}^{k \times k}$ and $y_i = \text{Tr}(M_i^T X^*)$ for some unobserved matrix $X^* \in \mathbb{R}^{k \times k}$, gradient descent on $\{W_i\}_{i=1}^n$ is used to solve \begin{align*} \arg\min_{\{W_i\} }\frac{1}{2n} \sum\limits_{i=1}^{n}\| y_i - \text{Tr}(M_i^T W_d W_{d-1} \ldots W_1) \|_2^2. \end{align*} Implicit regularization of linear networks in the matrix sensing setting has been analyzed extensively~\cite{DeepMatrixFactorization, BalancednessJasonLee, MatrixFactorizationNIPS2017, MatrixSensing}. Theorem \ref{thm: Alignment for fully connected networks with multi-dimensional output} shows that alignment is an invariant of training for this problem if and only if $M_i = U \Lambda_i V^T$ for all $i \in [n]$, and $U_d = U, V_1 = V$. \\\\ \textbf{1-dimensional Outputs.} In the following proposition, we show that alignment is an invariant of training for fully connected networks with 1-dimensional outputs for any real-valued loss function provided that gradient descent converges to zero training error. The proof is given in Appendix \ref{alignment 1d proof}. \begin{prop} \label{prop: Alignment for 1 dimensional outputs} Alignment is an invariant of training for any linear fully connected network $f: \mathbb{R}^{k_0} \rightarrow \mathbb{R}$, any real-valued loss function, and data $(X,Y)\in \mathbb{R}^{k_0 \times n} \times \mathbb{R}^{1 \times n}$ for which gradient descent minimizes the loss to zero. \end{prop} \subsection{Consequences of Alignment} \label{subsec: Consequences of Alignment} \vspace{-0.1cm} We next discuss various consequences of the invariance of alignment for the analysis of training. Our explicit characterization of alignment as an invariant is significant as it allows us to greatly simplify the convergence analysis of gradient descent, which is a main goal of defining an invariant of training. The following corollary (proof in Appendix~\ref{sec: singular update proof}) follows from the proof of Theorem \ref{thm: Alignment for fully connected networks with multi-dimensional output}, and shows that under alignment the gradient descent update rule is simplified significantly. \begin{corollary} \label{cor:singular-update} Let $r = \min(k_0, k_1, \ldots, k_d) > 1$ and let the top left $r \times r$ submatrix of $U^TY X^TV$ be $\Lambda'$ and that of $V^TX X^TV$ be $\Lambda$. Under the invariance of strong alignment (i.e., when $\Lambda'$ and $\Lambda$ are diagonal), we can express the partial derivative with respect to $W_i$ as follows: \begin{align} \label{eq: gradient formula} \frac{\partial L}{\partial W_i} = -\frac{1}{n}U_i \left(\prod\limits_{j={i+1}}^{d}{\Sigma_j}^T (U^T YX^TV_1 - \Sigma_d\cdots\Sigma_1V^TXX^T V) \prod\limits_{j={1}}^{i-1}{\Sigma_j}^T\right)V_i^T. \end{align} As a result, gradient descent only updates the first $r$ values of $\Sigma_i$. Let ${\Sigma'}_i^{(t)}$ be the top left $r \times r$ matrix of $\Sigma_i^{(t)}$. The updates are then given by: \vspace{-0.1cm} \begin{align} \label{eq: Updating diagonal elements under alignment} {\Sigma'}_i^{(t+1)} = {\Sigma'_i}^{(t)} + \frac{\gamma}{n} \prod\limits_{j = 1}^d {{\Sigma'}_j^{(t)}} (\Lambda' - \prod_{j=1}^{d}{\Sigma'}_j^{(t)} \Lambda). \vspace{-1.2cm} \end{align} The other entries of $\Sigma_i^{(t+1)}$ are not updated. \end{corollary} We can use this corollary to provide an explicit learning rate under which gradient descent converges linearly to a global minimum. The proof of the following proposition is given in Appendix \ref{linear convergence proof}. \begin{prop} \label{prop: linear convergence} For $k \in [r]$, let $\sigma_k(W_i)$ denote the $k$th entry of $\Sigma_i$ in the usSVD of $W_i$, and let $\lambda_k$, $\lambda'_k$ denote the $k$th entries of $\Lambda$, $\Lambda'$ respectively. Under the conditions of Corollary~\ref{cor:singular-update} and assuming that $\sigma_k(W_i^{(0)}) > 0\,$ and $\prod\limits_{i=1}^{d}\sigma_k(W_i^{(0)}) < \frac{\lambda'_k}{\lambda_k}\,$ for all $k \in [r]$, if the learning rate satisfies $\gamma \le \frac{n \ln 2}{d} \cdot \min_{k}\frac{ \sigma_k(W_i^{(0)})^2\lambda_k}{\lambda'^2_k}$ then gradient descent only updates the top $r$ singular values of the solution and converges linearly to the global minimum. \end{prop} \subsection{Alignment in the Limit of Training} \label{sec: Alignment in Limit} While the previous section was primarily concerned with the invariance properties of alignment, we briefly comment on understanding whether alignment will occur in the limit of training. We first present the following proposition, which states that for a 2-layer network, an aligned solution achieves the minimum $\ell_2$-norm. The proof is given in Appendix \ref{min norm proof}. \begin{prop} \label{prop: aligned min norm} Let $W_1, W_2$ be matrices such that $W_2W_1 = P$, for a fixed matrix $P$. Then, $\|W_1\|^2_F + \|W_2\|^2_F$ achieves a minimum at the solution where $W_1$ and $W_2$ are aligned \emph{and} 0-balanced, i.e. there exist usSVD's $W_1 = W\Sigma V^T, W_2 = U\Sigma W^T$. \end{prop} It has also been shown that SGD in the overparameterized setting for a network initialized close to zero will converge to a solution close in $\ell_2$-norm to the minimum $\ell_2$-norm solution~\cite{AzizanMirrorDescent}. Therefore we expect such networks to converge to a solution which is close to an aligned solution. \section{Alignment Under General Layer Structure} \label{sec: Alignment in Networks with Layer Structures} In the previous section, we analyzed fully connected networks, where parameters of each weight matrix are optimized independently. The most commonly used deep learning models, however, rely on convolutional layers or layers with other forms of constraints. In this section, we analyze alignment in the setting of linear networks with layer constraints. In particular, we show that when the dimension of the subspace induced by the layer constraints is small compared to the number of training samples, alignment cannot happen, let alone be an invariant of training. \subsection{Linear Neural Networks with Layer Structure} We start by setting up mathematical terminology to describe different layer structures. \begin{definition} \label{def_layer} Let $\mathcal{S} \subset \mathbb{R}^{m \times n}$ be a linear subspace of matrices and let $\{A_i\}_{i=1}^{r}$ be an orthogonal\footnote{Orthogonality is w.r.t the inner product $\langle A, B \rangle = \text{Tr}(A^TB)$, or equivalently the dot product in $\mathbb{R}^{mn}$} basis for $\mathcal{S}$. Layer $W_i$ has \textit{layer structure} $\mathcal{S}$ if $W_i\in\mathcal{S}$, i.e., there exist coefficients $\{c_j^i\}_{j=1}^r \subset \mathbb{R}$ such that $W_i = \sum_{j=1}^{r}c_j^i A_j$, and gradient descent operates on the $\{c_j^i\}_{i,j=1}^r$. \end{definition} Definition~\ref{def_layer} encompasses layer structures commonly used in practice, such as: \begin{itemize} \vspace{-0.1cm} \item\textit{Convolutional layers:} treating a $p\times p$ image as a vector in $\mathbb{R}^{p^2}$, a single $s \times s$ convolutional filter with stride 1 and padding $\frac{s - 1}{2}$ maps the image to another $p\times p$ image; this linear transformation can be represented as a matrix in $\mathbb{R}^{p^2 \times p^2}$ and the set of all such transformations forms an $s^2$-dimensional subspace. The parameters of the filter are coefficients of an orthogonal basis of this subspace, and hence performing gradient descent on the parameters is equivalent to optimizing over the basis coefficients; see Appendix~\ref{conv example} for an example. \vspace{-0.1cm} \item\textit{Layers with Sparse Connections:} Consider a fixed connection pattern between layers such that the $j^{th}$ hidden unit in layer $i$ depends only on a subset of units in layer $i-1$. In this case, the subspace $\mathcal{S}$ consists of matrices where particular entries are forced to be zero corresponding to missing connections between features in consecutive layers. \end{itemize} \vspace{-0.1cm} The following theorem provides, in closed-form, the gradient descent update rules for linear networks with layer structure. The proof is provided in Appendix~\ref{projected gd proof}. \vspace{0.1cm} \begin{theorem} \label{thm:projected gd} Performing gradient descent on the basis coefficients $\{c_j^i\}_{j=1}^r$ leads to the following weight matrix updates: $$W_i^{(t+1)} = W_i^{(t)} - \eta\cdot\pi_\mathcal{S}\left( \frac{\partial l}{\partial W_i^{(t)}} \right),$$ where $\pi_\mathcal{S}$ denotes the projection operator onto $\mathcal{S}$. \end{theorem} Theorem \ref{thm:projected gd} shows that gradient descent in networks with layer structure is equivalent to projected gradient descent\footnote{$\pi_\mathcal{S}$ is a projection in the traditional sense if and only if the $A_j$ form an orthonormal basis; otherwise, $\pi_\mathcal{S}$ is a projection onto $\mathcal{S}$ followed by an appropriate scaling in each basis direction.}. Hence alignment is an invariant of training if and only if it holds throughout the projected gradient descent updates and leads to an aligned solution~with~zero~training~loss. \subsection{Necessary Condition for Alignment} Motivated by the above characterization via projected gradient descent, we now show that for layer structures with constrained dimension, aligned networks generally cannot achieve zero training error under the squared loss, given sufficient data (Proposition \ref{prop: No alignment with constrained layers}). This is the case even when there is a solution with the desired layer structure that achieves zero training error. Hence, if loss is minimized to zero, gradient descent must lead to a non-aligned network. We first show that for an aligned network which interpolates the data, the first and last layer must align with the pseudoinverse. The proof of this result is presented in Appendix~\ref{conv alignment lemma}. \begin{prop} \label{prop: Interpolation with layer structure} Let $(X,Y) \in \mathbb{R}^{k_0 \times n} \times \mathbb{R}^{k_d \times n}$ such that $n \ge k_0$ and $X$ is full-rank (ensuring that $XX^T$ is invertible). If an aligned network $f = W_d W_{d-1} \ldots W_1$ achieves zero error under squared loss (i.e. if $Y = f(X)$), then $W_d^T$ aligns with $YX^T(XX^T)^{-1}$, which in turn aligns with $W_1^T$. \end{prop} The following result tells us that when a linear space $\mathcal{S}$ of matrices is sufficiently low-dimensional, the set of matrices that align with an element of $\mathcal{S}$ has measure zero. While we are mainly interested in the setting where $n\geq k$, we state it in full generality using $\binom{m}{2}=0$, when $m<2$. \begin{prop}\label{prop: No alignment with constrained layers} Let $\mathcal{S}$ be an $r$-dimensional linear subspace of $k\times k$ matrices. If $r < k-1-\binom{k-n}{2}$ then the set of matrices of size $k\times n$ that can align with an element of $\mathcal{S}$, excluding scalar multiples of the identity, has Lebesgue measure zero. \end{prop} The proof of Proposition~\ref{prop: No alignment with constrained layers} is provided in Appendix~\ref{conv alignment bound}. Taken together, Propositions \ref{prop: Interpolation with layer structure} and \ref{prop: No alignment with constrained layers} directly imply Theorem~\ref{thm: No Alignment in constrained layers}, which states that alignment does not occur in linear networks with constrained layer structures given enough training samples. To simplify notation, we let $k=k_0=\cdots =k_d$ and let all layers have the same structure, $\mathcal{S}$. The statement can trivially be extended to the general setting without these assumptions. \begin{theorem} \label{thm: No Alignment in constrained layers} Let $n \ge k$, let $X,Y \in \mathbb{R}^{k\times n}$ be generic, let $\mathcal{S} \subset \mathbb{R}^{k \times k}$ be a linear subspace of dimension $r < k-1$, and let $W_1,\dots,W_d \in \mathcal{S}$ such that at least one $W_i$ is not a scalar multiple of the identity\footnote{This is not a serious restriction; modulo scalar multiplication, the only case in which such a network could achieve zero loss is autoencoding, in which case the latent space would be a scalar multiple of the data itself.}. If the network $f = W_d \cdots W_1$ satisfies $Y = f(X)$, then $f$ is not aligned. \end{theorem} Theorem \ref{thm: No Alignment in constrained layers} is in contrast to fully connected networks (i.e., no layer constraints), where we showed that alignment is possible for particular classes of problems including autoencoders. An explicit example of a convolutional linear autoencoder, where alignment is ruled out by Theorem~\ref{thm: No Alignment in constrained layers}, is discussed next. \begin{example} If $m \ge 4$, then a generic dataset consisting of $n \ge m^2$ $m\times m$ images cannot be aligned by any convolutional linear autoencoder with filter size $3$, aside from the trivial case where all layers are scalar multiples of the identity. This follows from letting $k = m^2$, $r = 9$ in Proposition~\ref{prop: No alignment with constrained layers}. \end{example} \section{Empirical Support} \label{sec: Experiments} In this section, we provide experimental results to validate our theoretical findings in the settings where alignment is not an invariant of training\footnote{Hyperparameter settings are detailed in Appendix~\ref{experimental setup}}. We measure two properties: (1) invariance of alignment from initialization, and (2) alignment between layers. Invariance of alignment at time $t$ is measured by the average dot product between corresponding columns of $U_i^{(t)}$ and $U_i^{(0)}$, as well as $V_i^{(t)}$ and $V_i^{(0)}$. Alignment is measured by the average dot product between corresponding columns of $U_i^{(t)}$ and $V_{i+1}^{(t)}$. For both, a value of 1 is perfect alignment / invariance. \begin{figure*}[!t] \centering \begin{subfigure}[t]{.33 \textwidth} \centering \includegraphics[scale=0.3]{Figures/rand_data_final.png} \caption{\centering Multi-dimensional regression on\newline random data with squared loss.} \end{subfigure}% \begin{subfigure}[t]{.33 \textwidth} \centering \includegraphics[scale=0.3]{Figures/mse_multiclass_final.png} \caption{\centering Multi-class classification on\newline MNIST with squared loss.} \end{subfigure}% \begin{subfigure}[t]{.33 \textwidth} \centering \includegraphics[scale=0.3]{Figures/cross_entropy_multiclass_final.png} \caption{\centering Multi-class classification on\newline MNIST with cross entropy loss.} \end{subfigure}% \vspace{-0.1cm} \caption{Examples of fully connected networks with multi-dimensional outputs where alignment is not an invariant of training.} \vspace{-0.3cm} \label{fig: Fully connected networks do not align} \end{figure*} \begin{figure*}[!t] \centering \begin{subfigure}[t]{.4 \textwidth} \centering \includegraphics[scale=0.3]{Figures/toep_final.png} \vspace{-0.12cm} \caption{\centering Matrix factorization with layers\newline constrained to be Toeplitz matrices.} \end{subfigure}% \begin{subfigure}[t]{.4 \textwidth} \centering \includegraphics[scale=0.3]{Figures/conv_1sample_final.png} \vspace{-0.12cm} \caption{\centering Autoencoding a single MNIST example\newline using a convolutional network.} \end{subfigure} \vspace{-0.12cm} \caption{Examples of layer constrained networks, where alignment is not an invariant of training.} \label{fig: General networks do not align.} \vspace{-0.5cm} \end{figure*} We begin with examples demonstrating that alignment is not an invariant of training for fully connected networks when the data conditions of Theorem \ref{thm: Alignment for fully connected networks with multi-dimensional output} are violated. Figure \ref{fig: Fully connected networks do not align}a shows an example where alignment is not an invariant for multi-dimensional regression with random data under squared loss. We used standard normal inputs $X \in \mathbb{R}^{9\times 9}$ and targets $Y \in \mathbb{R}^{9 \times 9}$, and a 2-hidden layer network initialized so that alignment holds at the start of training. Since $X$ and $Y$ do not have the same right singular vectors, the conditions of Theorem \ref{thm: Alignment for fully connected networks with multi-dimensional output} are violated, and hence alignment is not an invariant of training, which is reflected in Figure~\ref{fig: Fully connected networks do not align}a. In Figures \ref{fig: Fully connected networks do not align}b and c, we show that alignment is also not an invariant in standard classification settings. In particular, we trained a 2-hidden layer fully connected network to classify a linearly separable subset of $256$ MNIST examples under MSE loss and cross entropy loss. Figure \ref{fig: Fully connected networks do not align}b is consistent with the generalization of Theorem \ref{thm: Alignment for fully connected networks with multi-dimensional output} to non-square layers (see Appendix~\ref{proof strong alignment non square}). It is interesting that this result transfers to the case of cross entropy loss, at least empirically, suggesting that our theoretical results may also be relevant for other loss functions. In networks with constrained layer structure, Theorem~\ref{thm: No Alignment in constrained layers} shows that given a sufficient amount of data, alignment cannot occur. We now present empirical evidence that alignment is not an invariant of training, even when the number of training samples is much smaller than the output dimension of the network or the dimensionality of the layer structure is much larger than the output dimension. We provide an example from matrix factorization ($Y \in \mathbb{R}^{k \times k}$, $X = I$). Here, $k = n$, so Theorem~\ref{thm: No Alignment in constrained layers} states that alignment is impossible when the linear structure has dimension $r < k - 1$. In Figure \ref{fig: General networks do not align.}a, we show that alignment does not occur also when $r \geq k - 1$. In particular, alignment is not an invariant when training a 2-hidden layer Toeplitz network to factorize a $4 \times 4$ matrix. Our network has $4$ hidden units per layer and thus $r = 7, k = 4, n = 4$. Even when $n < r < k$, we observe that alignment is not an invariant. In Figure \ref{fig: General networks do not align.}b, we show that alignment is not an invariant of training when autoencoding a single MNIST example using a 2-hidden layer linear convolutional network (i.e. $n = 1, r = 9, k = 784$). \section{Discussion} \vspace{-1mm} We generalized the definition of alignment to linear networks with multi-dimensional outputs. We then analyzed the invariance properties of alignment, showing that under particular data conditions alignment is an invariant for fully connected networks, which allows us to significantly simply the convergence analysis of gradient descent. We then extended our analysis of alignment to networks with constrained layer structures, such as convolutions, and proved that alignment cannot be an invariant of training in such networks when the dimension of the layer structure $r$ is small compared to the number of training samples $n$. While the simplification of gradient descent convergence analysis in the fully connected setting shows that our alignment definition is useful in understanding such networks, the fact that it does not generalize as an invariant to the constrained layer structure setting suggests that other approaches may be necessary to fully understand implicit regularization, such as studying how architecture influences the function classes that can be represented by deep networks~\cite{savarese2019infiniteWidth, IdentityCrisis, RadhaAutoencoders}. \section*{Acknowledgements} A. Radhakrishnan and C. Uhler thank the Simons Institute at UC Berkeley for hosting them during the summer 2019 program on ``Foundations of Deep Learning'', which facilitated this work. A.~Radhakrishnan and C.~Uhler were partially supported by the National Science Foundation (DMS-1651995), Office of Naval Research (N00014-17-1-2147 and N00014-18-1-2765), IBM, and a Simons Investigator Award to C.~Uhler. Daniel Irving Bernstein was supported by an NSF Mathematical Sciences Postdoctoral Research Fellowship (DMS-1802902). The Titan Xp used for this research was donated by the NVIDIA Corporation. \bibliographystyle{plain}
{ "timestamp": "2020-06-18T02:06:32", "yymm": "2003", "arxiv_id": "2003.06340", "language": "en", "url": "https://arxiv.org/abs/2003.06340", "abstract": "We study the properties of alignment, a form of implicit regularization, in linear neural networks under gradient descent. We define alignment for fully connected networks with multidimensional outputs and show that it is a natural extension of alignment in networks with 1-dimensional outputs as defined by Ji and Telgarsky, 2018. While in fully connected networks, there always exists a global minimum corresponding to an aligned solution, we analyze alignment as it relates to the training process. Namely, we characterize when alignment is an invariant of training under gradient descent by providing necessary and sufficient conditions for this invariant to hold. In such settings, the dynamics of gradient descent simplify, thereby allowing us to provide an explicit learning rate under which the network converges linearly to a global minimum. We then analyze networks with layer constraints such as convolutional networks. In this setting, we prove that gradient descent is equivalent to projected gradient descent, and that alignment is impossible with sufficiently large datasets.", "subjects": "Machine Learning (cs.LG); Machine Learning (stat.ML)", "title": "On Alignment in Deep Linear Neural Networks", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9838471670723236, "lm_q2_score": 0.8311430436757312, "lm_q1q2_score": 0.8177177289522366 }
https://arxiv.org/abs/2008.08718
Minimum discrepancy principle strategy for choosing $k$ in $k$-NN regression
We present a novel data-driven strategy to choose the hyperparameter $k$ in the $k$-NN regression estimator. We treat the problem of choosing the hyperparameter as an iterative procedure (over $k$) and propose using an easily implemented in practice strategy based on the idea of early stopping and the minimum discrepancy principle. This model selection strategy is proven to be minimax-optimal, under the fixed-design assumption on covariates, over some smoothness function classes, for instance, the Lipschitz functions class on a bounded domain. The novel method often improves statistical performance on artificial and real-world data sets in comparison to other model selection strategies, such as the Hold-out method and 5-fold cross-validation. The novelty of the strategy comes from reducing the computational time of the model selection procedure while preserving the statistical (minimax) optimality of the resulting estimator. More precisely, given a sample of size $n$, assuming that the nearest neighbors are already precomputed, if one should choose $k$ among $\left\{ 1, \ldots, n \right\}$, the strategy reduces the computational time of the generalized cross-validation or Akaike's AIC criteria from $\mathcal{O}\left( n^3 \right)$ to $\mathcal{O}\left( n^2 (n - k) \right)$, where $k$ is the proposed (minimum discrepancy principle) value of the nearest neighbors. Code for the simulations is provided atthis https URL.
\section{Introduction} \label{sec:intro} Nonparametric regression estimation is a fundamental problem in statistics and machine learning. The $k$-NN regression estimator (\cite{gyorfi2006distribution, biau2015lectures}) is a very simple and popular choice in practice. For this estimator, the central issue is choosing properly the number of neighbors $k$. The theoretical performance of the $k$-NN regression estimator has been widely studied since the 1970s (\cite{devroye1978uniform, collomb1979estimation, devroye1981almost, bhattacharya1987weak, biau2010rates, kpotufe2011k, biau2015lectures, zhao2021minimax}). For example, in (\cite[Chapter 12]{biau2015lectures}) the uniform consistency of the $k$-NN estimator is proved under the condition that $k(n) / n \to 0$ as $n \to \infty$, where $n$ is the sample size. However, as it was shown in \cite{gyorfi2006distribution}, the nearest neighbor estimator ($k = 1$) is proved to be consistent only in the noiseless case. Recently, researchers started to be interested in choosing $k$ from the data (\cite{gyorfi2006distribution, arlot2009data, kpotufe2011k, azadkia2019optimal}). Apparently, the most common and simplest strategy to choose $k$ is to assume some smoothness assumption on the regression function (e.g., the Lipschitz condition \cite{gyorfi2006distribution}) and to find $k$ that makes an upper bound on the bias and the variance of the $k$-NN regression estimator equal. This method has a clear lack: one needs to know the smoothness of the regression function (e.g., the Lipschitz constant). \cite{2009arXiv0909.1884A} gave a \textit{data-driven strategy} for choosing a hyperparameter for different linear estimators (e.g., the $k$-NN estimator) based on the idea of minimal penalty, introduced previously in \cite{birge2007minimal}. The main inconvenience of this strategy is that one needs to compute all the estimators $\mathbf{F}_n = \{ f^k, \ k = 1, \ldots, n \}$ of the regression function in order to choose the optimal one among them by comparing them via a special criterion that involves the empirical error (least-squares loss). To list other (similar) strategies, one can think about the Akaike's AIC (\cite{akaike1998information}), Mallows's $C_p$ (\cite{mallows2000some}) criteria or generalized cross-validation (\cite{li1987asymptotic, hastie2009elements}), where one has to compute the empirical risk error plus a penalty term for any $k = 1, \ldots, n$. Often it is computationally expensive and restricts the use in practice. This gives rise to the problem of choosing the hyperparameter "in real-time", meaning that the practitioner should compute iteratively $f^k \in \mathbf{F}_n$. Eventually, this iterative process has to be stopped. This problem can be solved by applying the \textit{early stopping rule} (\cite{engl1996regularization, zhang2005boosting, Yao2007, raskutti2014early, blanchard2018optimal, wei2017early}). The first early stopping rule that could be potentially data-driven was proposed by (\cite{blanchard2018optimal, blanchard2018early, celisse2021analyzing}) for spectral filter iterative algorithms (see, e.g., \cite{bauer2007regularization, gerfo2008spectral} for examples of such algorithms). The idea behind the construction of this early stopping rule is the so-called \textit{minimum discrepancy principle} (MDP) that is based on finding a first iteration for which a learning algorithm starts to fit the noise. The key quantity for the analysis of the minimum discrepancy principle is the \textit{empirical risk error} (the train error in the terminology of the machine learning community \cite{goodfellow2016deep}), which is monitored throughout the whole learning process. The process thus is stopped if the empirical risk starts to fit the noise. \textbf{Contribution.} In the present paper, we propose applying the minimum discrepancy principle stopping rule for the $k$-NN regression estimator in order to select $k$. We prove a non-asymptotic upper bound on the performance of the minimum dicrepancy principle stopping rule measured in the empirical $L_2(\mathbb{P}_n)$ norm. This bound implies that, under a quite mild assumption on the regression function, the minimum discrepancy principle stopping rule provides a statistically (minimax) optimal functional estimator, in particular, over the class of Lipschitz functions on a bounded domain. In Section 5, we show that our method improves the performances of classical selection procedures, such as 5-fold cross-validation and Hold-out, on artificial and real data. Besides that, the proposed strategy lowers the computational time of the selection procedure compared to some well-known model selection strategies, such as the generalized cross-validation and Akaike's AIC, i.e., $\mathcal{O}\left( n^2 \left( n - k \right) \right)$ for MDP($k$) and $\mathcal{O}\left( n^3 \right)$ for Akaike's AIC and generalized cross-validation, respectively. We emphasize that the proposed strategy for choosing $k$ is \textit{easy-to-implement} in practice since it involves only computing the empirical risk and estimating the noise level of the regression model. \textbf{Outline of the paper.} The organization of the paper is as follows. Section \ref{sec:model} describes the statistical model, its main assumption and introduces the notation that will be used along the paper. In Section \ref{sec:estimator_stopping_rule}, we introduce the $k$-NN estimator and explain how to compute the minimum discrepancy early stopping rule. Section \ref{sec:optimality_result} provides the main theoretical result that shows that the proposed rule is statistically optimal for some classes of functions (e.g., the well-known class of Lipschitz functions on a bounded domain). In Section \ref{sec:simulations}, one can find simulation results for the proposed stopping rule. To be precise, we compare this rule to the generalized cross-validation estimator, Akaike's AIC criterion, $V$--fold and Hold-out cross-validation stopping rules (\cite{arlot2010survey}) tested on some artificial and real-world data sets. Section \ref{sec:conc} concludes the paper. All the technical proofs are in the supplementary material. \section{Statistical model, main assumption, and notation} \label{sec:model} In the nonparametric regression setting, one works with a sample $(x_1, y_1), \ldots, \\ (x_n, y_n) \in \mathcal{X}^n \times \mathbb{R}^n$ that satisfies the statistical model \begin{equation} \label{nonparametric_regression_model} y_i = f^*(x_i) + \varepsilon_i, \ \ i = 1,\ldots, n, \end{equation} where $f^*:\mathcal{X} \mapsto \mathbb{R}$, $\mathcal{X} \subseteq \mathbb{R}^d$, is a measurable function on some set $\mathcal{X}$, and $\{ \varepsilon_i \}_{i=1}^n$ are i.i.d. Gaussian noise variables $\mathcal{N}(0, \sigma^2)$. On the one hand, in the theoretical analysis in Section \ref{sec:optimality_result}, we assume that the parameter $\sigma^2 > 0$ is fixed and known. On the other hand, from the practical perspective in Sections \ref{artificial_data} and \ref{real_data}, we estimate $\sigma^2$ and provide experimental results on artificial and real-world data sets. One should point out here that the assumption of known variance $\sigma^2$ is quite typical in the (theoretical) model selection literature with nonparametric regression (see, e.g., \cite{li1986asymptotic, li1987asymptotic, yang2005can, cao2006oracle}). In addition to that, we assume that $\{ x_i \in \mathcal{X} \}_{i=1}^n$ are \textit{fixed} covariates (corresponds to the so-called fixed design setting), thus we observe noise only in the responses $\{ y_i \}_{i=1}^n$. The goal of the present paper is to estimate optimally the regression function $f^*$. The term "optimally" will be explained in Section \ref{sec:estimator_stopping_rule}. In the context of the \textit{fixed design} setting, the performance of an estimator $\widehat{f}$ of $f^*$ is measured in terms of the so-called \textit{empirical norm} defined as \begin{equation} \lVert \widehat{f} - f^* \rVert_n^2 \coloneqq \frac{1}{n}\sum_{i=1}^n \left[ \widehat{f}(x_i) - f^*(x_i) \right]^2, \end{equation} where $\lVert h \rVert_n \coloneqq \sqrt{1/n\sum_{i=1}^n h(x_i)^2}$ for any bounded on $\mathcal{X}$ function $h$. We denote the empirical norm as $L_2(\mathbb{P}_n)$. For each bounded over $\mathcal{X}$ functions $h_1, h_2$, $\langle h_1, h_2 \rangle_n$ denotes the related inner product defined as $\langle h_1, h_2 \rangle_n \coloneqq 1/n \sum_{i=1}^n h_1(x_i)h_2(x_i)$. Further, $\mathbb{P}_{\varepsilon}$ and $\mathbb{E}_{\varepsilon}$ denote the probability and expectation with respect to $\{ \varepsilon_i \}_{i=1}^n$. \textbf{Notation.} Throughout the paper, $\lVert \cdot \rVert$ and $\langle \cdot, \cdot \rangle$ are the usual Euclidean norm and related inner product. $\lVert M \rVert_2$ and $\lVert M \rVert_{F}$ signify the operator and Frobenius norms of the matrix $M \in \mathbb{R}^{n \times n}$, respectively. We denote the trace of the matrix $M$ by $\textnormal{tr}(M)$. In addition to that, $\mathbb{I}\left\{ \mathcal{E} \right\}$ is equal to $1$ if the probabilistic event $\mathcal{E}$ holds true, otherwise it is equal to $0$. For $a \geq 0$, we denote by $\floor*{a}$ the largest natural number that is smaller than or equal to $a$. We denote by $\ceil*{a}$ the smallest natural number that is greater than or equal to $a$. Along the paper, $I_n$ is the identity matrix of size $n \times n$. We make the following assumption on the regression function $f^*$ introduced earlier in Eq. (\ref{nonparametric_regression_model}). \begin{assumption}[Boundness of the r.f.] \label{assumption_boundness} $f^*$ is bounded on $\mathcal{X}$, meaning that there exists a constant $\mathcal{M} > 0$ such that \begin{equation} \left| f^*(x) \right| \leq \mathcal{M} \qquad \textnormal{ for all }x \in \mathcal{X}. \end{equation} \end{assumption} Assumption \ref{assumption_boundness} is quite standard in the nonparametric regression literature (\cite{gyorfi2006distribution, zhao2021minimax}). In particular, Assumption \ref{assumption_boundness} holds when the set $\mathcal{X}$ is bounded, and the regression function $f^*$ is $L$-Lipschitz with some positive constant $L$ (cf. \cite{gyorfi2006distribution}). Along the paper, we use the notation $c, c_1, C, \widetilde{c}, \widetilde{C}, \ldots$ to show that numeric constants $c, c_1, C, \widetilde{c}, \widetilde{C}, \ldots$ can depend only on $d, \sigma$, and $\mathcal{M}$. The values of all the constants may change from line to line or even in the same line. \section{$k$-NN estimator and minimum discrepancy stopping rule} \label{sec:estimator_stopping_rule} \subsection{$k$-NN regression estimator} Let us transform the initial nonparametric regression model (\ref{nonparametric_regression_model}) into its vector form \begin{equation} \label{nonparametric_regression_vector} Y \coloneqq \left[y_1, \ldots, y_n\right]^\top = F^* + \varepsilon \in \mathbb{R}^n, \end{equation} where the vectors $F^* \coloneqq [f^*(x_1), \ldots, f^*(x_n)]^\top$ and $\varepsilon \coloneqq [\varepsilon_1, \ldots, \varepsilon_n]^\top$. Define a $k$-nearest neighbor estimator $f^k$ of $f^*$ from (\ref{nonparametric_regression_model}) at the point $x_i, \ i = 1, \ldots, n,$ as \begin{equation} \label{estimator} f^k(x_i) \coloneqq F^k_i = \frac{1}{k} \sum_{j \in \mathcal{N}_k(i)} y_j, \qquad k = 1, \ldots, n, \end{equation} where $\mathcal{N}_k(i)$ denotes the indices of the $k$ nearest neighbors of $x_i$ among $\{1, \ldots, n \}$ in the usual Euclidean norm in $\mathbb{R}^d$, where ties are broken at random. In words, in Eq. (\ref{estimator}) one weights by $1/k$ the response $y_j$ if $x_j$ is a $k$ nearest neighbor of $x_i$ measured in the Euclidean norm. Note that other adaptive metrics (instead of the Euclidean one) have been also considered in the literature (\cite[Chap. 14]{hastie2009elements}). One can notice that the $k$-NN regression estimator (\ref{estimator}) belongs to the class of linear estimators (\cite{2009arXiv0909.1884A, hastie2009elements}), meaning that $F^k \in \mathbb{R}^n$ estimates the vector $F^*$ as it follows. \begin{equation} \label{main_estimator} F^k \coloneqq \left( f^k(x_1), \ldots, f^k(x_n) \right)^\top = A_k Y, \end{equation} where $A_k \in \mathbb{R}^{n \times n}$ is the matrix described below. \begin{equation} \label{a_k_matrix} \begin{cases} \forall 1 \leq i, j \leq n, \ \left( A_k \right)_{ij} \in \{0, 1/k\} \textnormal{ with } k \in \{1, \ldots, n \},\\ \forall 1 \leq i \leq n, \ \left( A_k \right)_{ii} = 1/k \textnormal{ and } \sum_{j=1}^n \left( A_k \right)_{ij} = 1. \end{cases} \end{equation} \vspace{0.05cm} Saying differently, $(A_k)_{ij} = 1/k$ if $x_j$ is a $k$ nearest neighbor of $x_i$, otherwise $(A_k)_{ij} = 0, \ i, j \in \{1, \ldots, n\}$. Define the mean-squared error (the risk error) of the estimator $f^k$ as \begin{equation} \label{mean-squared-error} \textnormal{MSE}(k) \coloneqq \mathbb{E}_{\varepsilon} \lVert f^k - f^* \rVert_n^2 = \frac{1}{n} \mathbb{E}_{\varepsilon} \sum_{i=1}^n \Big( \frac{1}{k}\sum_{j \in \mathcal{N}_k(i)}y_j - f^*(x_i) \Big)^2. \end{equation} Further, we will introduce the (squared) bias and variance of the functional estimator $f^k$ (see, e.g., \cite[Eq. (7)]{2009arXiv0909.1884A}), \begin{equation} \textnormal{MSE}(k) = B^2(k) + V(k), \end{equation} where \begin{equation*} B^2(k) = \lVert (I_n - A_k)F^* \rVert_n^2, \qquad V(k) = \frac{\sigma^2}{n} \textnormal{tr}\left( A_k^\top A_k \right) = \frac{\sigma^2}{k}. \end{equation*} Thus, the variance term $\sigma^2/k$ is a decreasing function of $k$. Note that $B^2(1) = 0, \ V(1) = \sigma^2$, and $B^2(n) = (1 - 1/n)^2\lVert f^* \rVert_n^2, \ V(n) = \sigma^2/n$. Importantly, the bias term $B^2(k)$ can have \textit{arbitrary} behavior on the interval $[1, n]$. \vspace{0.05cm} Ideally, we would like to minimize the mean-squared error (\ref{mean-squared-error}) as a function of $k$. However, since the bias term is not known (it contains the unknown regression function), one should introduce other quantities that will be related to the bias. In our case, this quantity will be the \textit{empirical risk} at $k$: \begin{equation} \label{empirical_risk} R_k \coloneqq \lVert (I_n - A_k) Y \rVert_n^2. \end{equation} $R_k$ measures how well the estimator $f^k$ fits $Y$. Remark that $R_1 = 0$ (corresponds to the "overfitting" regime) and $R_n = (1 - 1/n)^2 \frac{1}{n} \sum_{i=1}^n y_i^2$ (corresponds to the "underfitting" regime), but there is no information about the monotonicity of $R_k$ on the interval $[1, n]$. Furthermore, some information about the bias is contained in the expectation (over the noise $\{ \varepsilon_i \}_{i=1}^n$) of the empirical risk. To be precise, for any $k \in \{ 1, \ldots, n\}$, \begin{align} \label{exp_empirical_risk} \begin{split} \mathbb{E}_{\varepsilon}R_k &= \sigma^2 + B^2(k) - \frac{\sigma^2(2 \textnormal{tr}(A_k) - \textnormal{tr}(A_k^\top A_k))}{n}\\ &= \sigma^2 + B^2(k) - V(k). \end{split} \end{align} Let us illustrate all the mentioned quantities in one example in Fig \ref{fig:bvr}. We take the regression function equal to $f^*(x) = \lVert x - 0.5 \rVert / \sqrt{3} - 0.5$ and the noise variance $\sigma^2 = 0.01$. We take $n = 50$, the uniform covariates $x_i \overset{\text{i.i.d.}}{\sim} \mathbb{U}[0, 1]^3$, and plot the bias term $B^2(k)$, the variance term $V(k)$, risk error $\textnormal{MSE}(k)$, empirical risk $R_k$, and its expectation $\mathbb{E}_{\varepsilon}R_k$ versus the number of neighbours $k$. Note that among all defined quantities, only the variance term $V(k)$ can be proved monotonic (without an additional assumption on the smoothness of $f^*$). Importantly, Fig \ref{fig:bvr} indicates that choosing $k = 6$ will provide the user with the global optimum of the risk (the mean-squared error) curve. Thus, for instance, it would be meaningless (according to the risk curve) to compute all the estimators $f^k$ (\ref{estimator}) for $k = 1, \ldots, 6$. \begin{figure} \centering \begin{tikzpicture}[scale=0.85] \tikzstyle{every node}=[font=\normalsize] \begin{axis}[ title={}, xlabel={Number of neighbors}, ylabel={Value}, xmin=1, xmax=12, ymin=0, ymax=0.02, xtick={1,3,5,7, 10}, ytick={0.002, 0.006, 0.01, 0.02}, legend pos=north west, ymajorgrids=true, xmajorgrids=true, grid style=dashed, legend style={nodes={scale=0.6, transform shape}}, ] \addplot[ color=blue, mark=square, ] coordinates { (1,0)(2,6.29080327e-03)(3,8.47616891e-03)(4,1.08716416e-02)(5,1.12560012e-02)(6,1.14496210e-02)(7,1.37171039e-02)(8,1.57263962e-02)(9,1.83168233e-02)(10,1.78368289e-02)(11,1.94323757e-02)(12,1.98591401e-02) }; \addlegendentry{Empirical risk} \addplot[ color=red, mark=+, ] coordinates { (1,0)(2,1.48163079e-03)(3,1.91217442e-03)(4,1.90325965e-03)(5,2.31901908e-03)(6,2.42449703e-03)(7,3.02981110e-03)(8,3.27183370e-03)(9,4.00873281e-03)(10,4.53268739e-03)(11,5.57342952e-03)(12,6.21032204e-03) }; \addlegendentry{Bias} \addplot[ color=violet, mark=star, ] coordinates { (1,0.01)(2,0.005)(3,0.00333333)(4,0.0025)(5,0.002)(6,0.00166667)(7,0.00142857)(8,0.00125)(9,0.00111111)(10,0.001)(11,0.00090909)(12,0.00083333) }; \addlegendentry{Variance} \addplot[ color=orange, mark=x, ] coordinates { (1,0.01)(2,0.00648163)(3,0.00524551)(4,0.00440326)(5,0.00431902)(6,0.00409116)(7,0.00445838)(8,0.00452183)(9,0.00511984)(10,0.00553269)(11,0.00648252)(12,0.00704366) }; \addlegendentry{Risk} \addplot[ color=black, mark=*, ] coordinates { (1,0)(2,6.48163079e-03)(3,8.57884109e-03)(4,9.40325965e-03)(5,1.03190191e-02)(6,1.07578304e-02)(7,1.16012397e-02)(8,1.20218337e-02)(9,1.28976217e-02)(10,1.35326874e-02)(11,1.46643386e-02)(12,1.53769887e-02) }; \addlegendentry{Exp. empirical risk} \end{axis} \end{tikzpicture} \caption{Sq. bias, variance, risk and (expected) empirical risk behaviour.} \label{fig:bvr} \end{figure} Our main concern is to design a data-driven strategy to choose $\widehat{k} \in \{1, \ldots, n \}$, which can be seen as a mapping from the data $\{ (x_i, y_i) \}_{i=1}^n$ to a positive integer so that the $L_2(\mathbb{P}_n)$-error $\lVert f^{\widehat{k}} - f^* \rVert_n^2$ is as small as possible. More precisely, the goal is to define a data-driven $\widehat{k}$ such that it satisfies the following non-asymptotic upper bound ("oracle-type inequality" \cite{wainwright2019high}): \begin{equation} \label{or_type_ineq} \lVert f^{\widehat{k}} - f^* \rVert_n^2 \leq C_n \mathbb{E}_{\varepsilon} \lVert f^{k_{\textnormal{opt}}} - f^* \rVert_n^2 + r_n \end{equation} with high (exponential) probability over $\{ \varepsilon_i \}_{i=1}^n$, where $f^{k_{\textnormal{opt}}}$ is a minimax-optimal estimator of the regression function $f^* \in \mathcal{F}$, $\mathcal{F}$ is some a priori chosen function space. The leading constant $C_n$ should be bounded, and the remainder term $r_n$ is negligible (smaller) with respect to the "optimal risk error" $\mathbb{E}_{\varepsilon} \lVert f^{k_{\textnormal{opt}}} - f^* \rVert_n^2$. \subsection{Related work} \label{related_work} The idea of choosing the hyperparameter $k$ from the data has been already considered in the literature. For example, the classical procedures such as generalized cross-validation (\cite{craven1978smoothing, li1987asymptotic, cao2006oracle}), penalized estimators (\cite{li1987asymptotic, mallows2000some, arlot2009data, arlot2009data_}), and different cross-validation methods (\cite{arlot2010survey}) are popular choices for linear estimators. Let us consider them in more detail. \textbf{Generalized CV (\cite{li1987asymptotic, cao2006oracle, hastie2009elements}).} This model selection method has been widely studied in the case of (kernel) ridge regression (\cite{craven1978smoothing}) and smoothing splines (\cite{cao2006oracle}). In particular, \cite{cao2006oracle} proved a non-asymptotic oracle inequality for the generalized CV estimator when the variance $\sigma^2$ is known. However, in a more general case, GCV estimates $\sigma^2$ implicitly, which is an advantage of the method. In addition to that, GCV for $k$-NN regression is proved by \cite{li1987asymptotic} to be an asymptotically optimal model selection criterion (i.e., $\lVert f^{k_{\textnormal{GCV}}} - f^* \rVert_n^2 \ / \ \underset{k}{\inf}\lVert f^{k} - f^* \rVert_n^2 \to 1$ in probability when $n \to + \infty$) under the assumption $\lVert A_k \rVert_2 \leq c, \ k = 1, \ldots, n,$ for some constant $c$. It is worth to mention that generalized cross-validation provides an approximation to the so-called leave-one-out cross-validation (\cite{arlot2010survey, celisse2018theoretical}), which is an exhaustive model selection procedure. In the case of GCV, if the nearest neighbors' matrices are already precomputed, its computational time is $\mathcal{O}\left( n^3 \right)$ elementary operations. The GCV strategy will be later considered in our simulations (see Section \ref{sec:simulations}). \textbf{Penalized estimators} date back to the work on AIC (\cite{akaike1998information}) or Mallow's $C_p$ (\cite{mallows2000some}) criteria, where a penalty proportional to the dimension of the model is added to the quadratic loss (i.e., \textit{the empirical risk error} in our notation (\ref{empirical_risk})) when the noise level $\sigma^2$ is known. As it was for the GCV strategy, the computational time of AIC and Mallow's $C_p$ are $\mathcal{O}\left( n^3 \right)$. After that, a new approach was developed by \cite{birge2007minimal}, where the authors introduced the so-called "slope heuristics" for projection matrices. This notion was based on the introduction of the penalty $\textnormal{pen}(k) = K \textnormal{tr}(A_k)$, where $\textnormal{tr}(A_k)$ is the dimension of the model, and $K$ is a constant that can depend on $\sigma^2$, in particular. It appeared that there exists a constant $K_{\textnormal{min}}$ such that $2 K_{\min} \textnormal{tr}(A_k)$ yields an asymptotically optimal model selection procedure. This gives rise to some strategies for the estimation of constant $K_{\textnormal{min}}$ from the data, as it was done, for instance, by \cite{2009arXiv0909.1884A} for some linear estimators when $\sigma^2$ is unknown. \textbf{Cross-validation methods (\cite{arlot2010survey}).} These model selection methods are the most used in practice. Compared to generalized cross-validation, for instance, $V$--fold cross-validation method (\cite{geisser1975predictive, arlot2010survey}) incurs a large computational cost (with $V$, which is not too small). To be precise, the $V$--fold cross-validation requires the model selection procedure to be performed $V$ times for each value of $k \in \{1, \ldots, n\}$. Another alternative could be the Hold-out method (\cite{wegkamp2003model, arlot2010survey}), which consists in randomly splitting the data into two parts for each value $k \in \{1, \ldots, n\}$: one is dedicated for training the estimator (\ref{estimator}) and the other one is dedicated for testing (see, e.g., Section \ref{sec:simulations} for more details in a simulated example). \subsection{Bias-variance trade-off and minimum discrepancy principle rule} We are at the point to define our first "reference rule". Based on the nonparametric statistics literature (\cite{wasserman2006all, tsybakov2008introduction}), the bias-variance trade-off usually provides an optimal estimator in the minimax sense: \begin{equation} \label{k_star} k^* = \inf \left\{ k \in \{ 1, \ldots, n \} \mid B^2(k) \geq V(k) \right\}. \end{equation} The bias-variance trade-off stopping rule $k^*$ does not always exist due to arbitrary behavior of the bias term $B^2(k)$. Thus, if no such $k^*$ exists, set $k^* = n$. If it exists, then $k^* \geq 2$ since $V(1) > B^2(1)$. Notice that the stopping rule $k^*$ is \textit{not computable} in practice, since it depends on the unknown bias $B^2(k)$. Nevertheless, we can create a data-driven version of $k^*$ by means of the empirical risk $R_k$ from Eq. (\ref{empirical_risk}). Eq. (\ref{exp_empirical_risk}) gives us that the event $\{ B^2(k) \geq V(k) \}$ is equivalent to the event $\{ \mathbb{E}_{\varepsilon}R_k \geq \sigma^2 \}$, so we conclude that $k^* = \inf \{ k \in \{ 1, \ldots, n \} \mid \mathbb{E}_{\varepsilon}R_k \geq \sigma^2 \}$. This gives rise to an estimator of $k^*$ that we denote as $k^{\tau}$. This stopping rule is called \textit{the minimum discrepancy principle} stopping rule and is defined as \begin{equation} \label{k_tau} k^{\tau} = \sup \left\{ k \in \{ 1, \ldots, n \} \mid R_k \leq \sigma^2 \right\}. \end{equation} \textbf{Remarks.} Note that in Eq. (\ref{k_tau}), we introduced a supremum instead of the infimum from Eq. (\ref{k_star}). That was done on purpose since there could be several points of the bias-variance trade-off, and apparently, the bias (and the empirical risk) could behave badly in the area "in-between". In order to calculate $k^{\tau}$, the user should, first, compute the empirical risk $R_k$ at $k = n$ (thus, the matrix $A_n$ of $n$ nearest neighbors). After that, one needs to decrease $k$ until the event $\{ R_k \leq \sigma^2 \}$ holds true. It is worth mentioning that it is not necessary to compute explicitly all the matrices $A_k, \ k = n, \ n-1, \ldots$, since, for instance, the matrix $A_{n-1}$ could be easily derived from the matrix $A_n$ (assuming that one has already arranged the neighbors and removed the $n^{\text{th}}$ neighbors from the matrix $A_n$), i.e., \begin{equation} \label{iteration_for_matrix} [A_{n-1}]_{ij} = \frac{n}{n-1}[A_n]_{ij}, \ \forall i, j \in \{1, \ldots, n\}. \end{equation} It is one of the main computational advantages of the proposed rule (\ref{k_tau}). For more details on the efficient computation of the nearest neighbors' matrices, see, e.g., \cite{bentley1975multidimensional, omohundro1989five}. In addition to all of that, we emphasize that the definition (\ref{k_tau}) of $k^{\tau}$ does not require the knowledge of the constant $\mathcal{M}$ from Assumption \ref{assumption_boundness}, and $k^{\tau}$ does not require computing the empirical risk $R_k$ for all values $k = 1, \ldots, n$, while it is the case, for instance, for generalized cross-validation or Mallow's $C_p$ (see Section \ref{sec:simulations}). Moreover, we need to point out that the stopping rule (\ref{k_tau}) depends on the noise level $\sigma^2$, which should be estimated in practice, as for the AIC or Mallow's $C_p$ criteria (\cite{akaike1974new, mallows2000some, hastie2009elements}). We will provide a consistent estimator of $\sigma^2$ in Section \ref{real_data} while dealing with real-world data. Regarding the computational time of $k^{\tau}$, if the nearest neighbors' matrices are already computed, it is of the order $\mathcal{O}\left( n^2 \left( n - k^{\tau} \right) \right)$, which is less than $\mathcal{O}\left( n^3 \right)$ for the AIC/Mallow's $C_p$ criteria, or GCV. \vspace{0.2cm} There is a large amount of literature (\cite{engl1996regularization, bauer2007regularization, blanchard2018optimal, blanchard2018early, celisse2021analyzing}) on the minimum discrepancy principle for spectral filter algorithms such as gradient descent, ridge regularization, and spectral cut-off regression, and providing an exhaustive review on this strategy is out of the scope of the paper (e.g., \cite{blanchard2018optimal, celisse2021analyzing} provide a thorough review). We should, however, emphasize that intuitively the minimum discrepancy principle determines the first time at which a learning algorithm starts to fit noise, which is measured by $\sigma^2$ in the present context. Moreover, one is able to notice that, if the empirical risk is close to its expectation, $k^{\tau}$ should produce an optimal estimator in some sense. The main question that should be asked is "In which setting is it possible to quantify this gap between $R_k$ and $\mathbb{E}_{\varepsilon}R_k$ that will not be statistically large?". This question is the main technical obstacle of the present paper. In what follows, we show that for a quite large class of regression functions, $k^{\tau}$ is optimal in the sense of Ineq. (\ref{or_type_ineq}). \section{Theoretical optimality result} \label{sec:optimality_result} Let us start to describe the main theoretical result of the present paper. The following theorem applies to the estimator defined in Eq. (\ref{main_estimator}). \begin{theorem} \label{main_th} Under Assumption \ref{assumption_boundness}, for arbitrary $u \geq 0$, \begin{equation} \label{main_ineq} \lVert f^{k^{\tau}} - f^* \rVert_n^2 \leq 8 V(k^*) + C_1 \left( \frac{u}{n} + \frac{\sqrt{u}}{\sqrt{n}} \right) + C_2 \sqrt{\frac{\log n}{n}} \end{equation} with probability at least $1 - 16 \exp \left(- u \right)$, where positive constants $C_1, C_2$ can depend on $d$, $\sigma$, and $\mathcal{M}$. Moreover, if $k^*$ from Eq. (\ref{k_star}) exists, then for arbitrary $u \geq 0$, \begin{equation} \label{corollary_from_the_main} \lVert f^{k^{\tau}} - f^* \rVert_n^2 \leq \underbrace{4 \ \textnormal{MSE}(k^*)}_{\textnormal{Main term}} + \underbrace{C_1 \left( \frac{u}{n} + \frac{\sqrt{u}}{\sqrt{n}} \right) + C_2 \sqrt{\frac{\log n}{n}}}_{\textnormal{Remainder term}} \end{equation} with probability at least $1 - 16 \exp (-u)$, where constants $C_1, C_2$ are from Ineq. (\ref{main_ineq}). \end{theorem} \begin{proof} The full proof is deferred to the supplementary material. Let us provide a sketch of the proof here. The main ingredients of the proof are two deviation inequalities (cf. Corollary \ref{corollary_for_variance} and Lemma \ref{deviation_bias_lemma} in the supplementary material): for any $x \geq 0$, \begin{equation} \mathbb{P}_{\varepsilon} \left( V(k^{\tau}) > 2 V(k^*) + x \right) \leq 2 \exp \left( -c_d n \min \left( \frac{x}{\sigma^2}, \frac{x^2}{\sigma^4} \right) \right), \end{equation} and \begin{equation} \label{bias_k_tau_variance_k_star} B^2(k^{\tau}) \leq 2 V(k^*) + c_1 \sqrt{\frac{\log n}{n}} + 2x, \end{equation} where Ineq. (\ref{bias_k_tau_variance_k_star}) holds with probability at least $1 - 12 \exp \left( -c n \min \left( x^2, x \right) \right)$. After that, one can split the $L_2(\mathbb{P}_n)$-error at $k^{\tau}$ into two parts: \begin{equation*} \lVert f^{k^{\tau}} - f^* \rVert_n^2 \leq 2 B^2(k^{\tau}) + 2 \lVert A_{k^{\tau}}\varepsilon \rVert_n^2. \end{equation*} It is sufficient to derive high probability control of $\underset{k}{\sup} \left| \lVert A_{k}\varepsilon \rVert_n^2 - V(k) \right|$ for $k = 1, \ldots, n$ (see Appendix \ref{var_control} in the supplementary material). That was the reason why the term $\mathcal{O}\left(\sqrt{\frac{\log n}{n}}\right)$ appeared in Eq. (\ref{main_ineq}). Finally, one can apply $V(k^*) \leq \frac{1}{2}\textnormal{MSE}(k^*)$, if $k^*$ exists, and $u = c n \min \left( x^2, x \right)$. The claim follows. \end{proof} In order to gain some intuition of the claim of Theorem \ref{main_th}, let us make some comments. First of all, Ineq. (\ref{corollary_from_the_main}) is non-asymptotic, meaning that it holds true for any sample size $n \geq 1$. Second, Ineq. (\ref{corollary_from_the_main}) holds "with high probability", which is a stronger result than in expectation since (\cite{li1987asymptotic}) there are model selection procedures that are asymptotically optimal (when $n \to +\infty$) with high probability but not in expectation. Third, the main term in Ineq. (\ref{corollary_from_the_main}) is the risk error at the bias-variance trade-off times 4 (this constant could be improved). Ideally, one should rather introduce the oracle risk $\underset{k=1, \ldots, n}{\inf} \left[ \mathbb{E}_{\varepsilon} \lVert f^k - f^* \rVert_n^2 \right]$ and compare $\lVert f^{k^{\tau}} - f^* \rVert_n^2$ to it. However, to the best of our knowledge, a smoothness assumption is needed to connect the bias-variance trade-off risk and the oracle risk. That was the reason to keep the main term as it was stated. Fourth, the right hand side term of Ineq. (\ref{corollary_from_the_main}) is of the order $\mathcal{O}\left(\sqrt{\frac{\log n}{n}}\right)$. Notice that the same rate for this term was achieved in \cite{azadkia2019optimal} but in terms of the expectation over the noise. A natural question would be to understand if the rate $\mathcal{O}\left(\sqrt{\frac{\log n}{n}}\right)$ in Ineq. (\ref{main_ineq}) or (\ref{corollary_from_the_main}) is sufficiently fast. In order to do that, one should precise the function space $\mathcal{F}$, where $f^*$ lies in. In what follows, we will mention one famous example (among others) of a such function space $\mathcal{F}$. \begin{example} Consider the class of functions \begin{equation} \label{lipschitz} \mathcal{F}_{\textnormal{Lip}}(L) \coloneqq \left\{ f: [0, 1]^d \mapsto \mathbb{R} \mid f(0) = 0, \ f \textnormal{ is } L-\textnormal{Lipschitz} \right\}, \end{equation} where $f$ is $L$-Lipschitz means that $|f(x) - f(x^\prime)| \leq L \lVert x - x^{\prime} \rVert$ for all $x, x^{\prime} \in [0, 1]^d$. In this case (see, e.g., \cite[Theorem 3.2]{gyorfi2006distribution} with $p = 1$), \begin{equation} \label{minimax_lipschitz} \mathbb{E}_{\varepsilon} \lVert \widehat{f} - f^* \rVert_n^2 \geq c_l n^{-\frac{2}{2 + d}}, \end{equation} for some positive constant $c_l$, for any measurable of the input data $\widehat{f}$. \end{example} Therefore, for the class of $L$-Lipshitz functions, the rate $\mathcal{O}(\sqrt{\log n / n})$ is faster than the minimax-optimal rate $\mathcal{O}(n^{-\frac{2}{2+d}})$ for any $d > 2$. As for the main term $8 V(k^*)$ in Ineq. (\ref{main_ineq}), it should be of a minimax-optimal order since the common strategy for obtaining optimal rates for the $k$-NN regression estimator is two-fold. First, one should derive a uniform (over $k$) upper bound on the bias term (knowing the smoothness of the regression function), which is a non-decreasing function of $k$. After that, this upper bound is made equal to the variance term, which results in the optimal $k^{\textnormal{b/v}}$. Following this argument, one can conclude that $k^{\textnormal{b/v}} \leq k^*$, which implies $V(k^*) \leq V(k^{\textnormal{b/v}})$. We summarize our findings in the theorem and corollary below. \begin{theorem}[Theorem 6.2 in \cite{gyorfi2006distribution}] \label{theorem_gyorfi} Under the Lipschitz condition (\ref{lipschitz}) on the regression function $f^*$, for any $k \in \{ 1, \ldots, n \}$, \begin{equation} \label{upper_bound_bias_ineq} \mathbb{E}_{\varepsilon} \lVert f^k - f^* \rVert_n^2 \leq C \left( \frac{k}{n} \right)^{2/d} + \frac{\sigma^2}{k}, \end{equation} where $C$ may depend on $d$ and $L$. Thus, Ineq. (\ref{upper_bound_bias_ineq}) gives $k^{\textnormal{b/v}} = \ceil*{\left( \frac{\sigma^2}{C} \right)^{d/(2+d)} n^{\frac{2}{2 + d}}}$. \end{theorem} \begin{corollary} \label{main_corollary} Set $u = \log n$ in Ineq. (\ref{main_ineq}), then, under the $L$-Lipschitz condition (\ref{lipschitz}) on the regression function $f^*$, early stopping rule $k^{\tau}$ from Eq. (\ref{k_tau}) satisfies \begin{equation} \mathbb{E}_{\varepsilon} \lVert f^{k^{\tau}} - f^* \rVert_n^2 \leq c_u n^{-\frac{2}{2+d}}, \end{equation} where positive constant $c_u$ depends on $d, \sigma$, and $L$; $d > 2$. \end{corollary} \begin{proof} First, by taking the expectation of Ineq. (\ref{main_ineq}), it gives \begin{align} \label{to_exp_from_proba} \begin{split} \mathbb{E}_{\varepsilon} \lVert f^{k^{\tau}} - f^* \rVert_n^2 &= \mathbb{E}_{\varepsilon} \left[ \lVert f^{k^{\tau}} - f^* \rVert_n^2 \mathbb{I}\left\{ \lVert f^{k^{\tau}} - f^* \rVert_n^2 \leq 8 V(k^*) + C_1 \frac{\sqrt{\log n}}{\sqrt{n}} + C_2 \frac{\log n}{n} \right\} \right] \\ &+ \mathbb{E}_{\varepsilon} \left[ \lVert f^{k^{\tau}} - f^* \rVert_n^2 \mathbb{I}\left\{ \lVert f^{k^{\tau}} - f^* \rVert_n^2 > 8 V(k^*) + C_1 \frac{\sqrt{\log n}}{\sqrt{n}} + C_2 \frac{\log n}{n} \right\} \right]. \end{split} \end{align} After that, due to Lemma \ref{operator_norm_lemma} from the supplementary material, $\lVert I_n - A_k \rVert_2 \leq c$ for any $k \in \{1, \ldots, n\}$, and $| f^*(x_i) | \leq \mathcal{M}$ for $i \in \{1, \ldots, n\}$ due to the Lipschitz condition (\ref{lipschitz}), which implies that \begin{align*} \lVert f^{k^{\tau}} - f^* \rVert_n^2 &\leq 2 \lVert (I_n - A_{k^{\tau}})F^* \rVert_n^2 + 2 \lVert A_{k^{\tau}}\varepsilon \rVert_n^2\\ &\leq 2 \lVert I_n - A_{k^{\tau}} \rVert_2^2 \lVert f^* \rVert_n^2 + 2 \lVert A_{k^{\tau}} \rVert_2^2 \lVert \varepsilon \rVert_n^2\\ &\leq c_1 + c_2 \lVert \varepsilon \rVert_n^2, \end{align*} where constants $c_1$ and $c_2$ depend only on $\mathcal{M}$ and $d$. Thus, \begin{equation} \label{aux_inequal} \Vert f^{k^{\tau}} - f^* \rVert_n^4 \leq c_1 + c_2 \lVert \varepsilon \rVert_n^4 + c_3 \lVert \varepsilon \rVert_n^2. \end{equation} From Ineq. (\ref{to_exp_from_proba}) and Cauchy-Schwarz inequality, it comes \begin{align*} &\mathbb{E}_{\varepsilon} \lVert f^{k^{\tau}} - f^* \rVert_n^2 \leq 8V(k^*) + C_1 \frac{\sqrt{\log n}}{\sqrt{n}} + C_2 \frac{\log n}{n} \\ &+ \sqrt{\mathbb{E}_{\varepsilon} \lVert f^{k^{\tau}} - f^* \rVert_n^4} \sqrt{\mathbb{P}_{\varepsilon} \left( \lVert f^{k^{\tau}} - f^* \rVert_n^2 > 8 V(k^*) + C_1 \frac{\sqrt{\log n}}{\sqrt{n}} + C_2 \frac{\log n}{n} \right) }. \end{align*} Applying Ineq. (\ref{main_ineq}) and Ineq. (\ref{aux_inequal}), we obtain \begin{equation*} \mathbb{E}_{\varepsilon} \lVert f^{k^{\tau}} - f^* \rVert_n^2 \leq 8 V(k^*) + C_1 \frac{\sqrt{\log n}}{\sqrt{n}} + C_2 \frac{\log n}{n} + \sqrt{c_1 + c_3 \sigma^4 + c_2 \sigma^2}\frac{4}{\sqrt{n}}. \end{equation*} The claim follows from $V(k^*) \leq V(k^{\textnormal{b/v}}) = \sigma^2 / k^{b/v}$, for $k^{\textnormal{b/v}}$ defined in Theorem \ref{theorem_gyorfi}. \end{proof} Therefore, the function estimator $f^{k^{\tau}}$ achieves (up to a constant) the minimax bound presented in Eq. (\ref{minimax_lipschitz}), thus non-improvable in general for the class of Lipschitz functions on a bounded domain. \section{Empirical comparison to other model selection rules} \label{sec:simulations} The present section aims at comparing the practical behaviour of our stopping rule $k^{\tau}$ from Eq. (\ref{k_tau}) with other existing and the most used in practice model selection rules. We split the section into three parts: Subsection \ref{description_early_stopping_rules} defines the competitive stopping rules. Subsection \ref{artificial_data} presents experiments on some artificial data sets, while Subsection \ref{real_data} presents experiments on some real data sets. \subsection{Description of the model selection rules to compare} \label{description_early_stopping_rules} In what follows, we will briefly describe five competitive model selection rules. \subsection*{Akaike's AIC criterion} The Akaike's information criterion (\cite{akaike1974new, hastie2009elements}) estimates the risk error by means of a log-likelihood loss function. In the case of $k$-NN regression with Gaussian noise, the maximum likelihood and least-squares are essentially the same things. This gives the AIC criterion as \begin{equation*} R_{\textnormal{AIC}}(f^k) = \frac{1}{n \widehat{\sigma}^2} \left( \lVert Y - A_k Y \rVert^2 + 2 \textnormal{tr}(A_k)\widehat{\sigma}^2 \right), \qquad k = 1, \ldots, n, \end{equation*} where $\widehat{\sigma}^2$ is an estimator of $\sigma^2$ obtained from a low-bias model (i.e. with $k = 2$). Using this criterion, we adjust the training error by a factor proportional to the "degree of freedom". Then, the AIC choice for the optimal $k$ is \begin{equation} \label{k_aic} k_{\textnormal{AIC}} \coloneqq \underset{k = 2, \ldots, n}{\textnormal{argmin}} \left\{ R_{\textnormal{AIC}}(f^k) \right\} - 1. \end{equation} Notice that in the mentioned case, AIC criterion is equivalent to Mallow's $C_p$ criterion (\cite{mallows2000some}). AIC criterion has been widely criticized in the literature, especially for the constant $2$ in the definition and/or its asymptotic nature. That is why some authors proposed corrections for the criterion (\cite{schwarz1978estimating, yang1999model}). Nevertheless, AIC, $C_p$, and other related criteria have been proved to satisfy non-asymptotic oracle-type inequalities (see \cite{birge2007minimal} and references therein). Notice that the computational time of the AIC criterion is $\mathcal{O}\left( n^3 \right)$, which is higher than $\mathcal{O}\left( n^2 \left( n - k^{\tau} \right) \right)$ for the minimum discrepancy principle stopping rule $k^{\tau}$ (\ref{k_tau}). \subsection*{Generalized cross-validation.} The generalized (GCV) cross-validation strategy (\cite{craven1978smoothing, arlot2010survey}) was introduced in least-squares regression as a rotation-invariant version of the leave-one-out cross-validation procedure. The GCV estimator of the risk error of the linear estimator $A_k Y, \ k = 1, \ldots, n,$ is defined via \begin{equation*} R_{\textnormal{GCV}}(f^k) = \frac{n^{-1}\lVert Y - A_k Y \rVert^2}{(1 - n^{-1}\textnormal{tr}(A_k))^2}, \end{equation*} The final model selection rule is \begin{equation} \label{k_1_out} k_{\text{GCV}} \coloneqq \underset{k = 2, \ldots, n}{\textnormal{argmin}} \left\{ R_{\textnormal{GCV}}(f^k) \right\} - 1. \end{equation} GCV should be close to the AIC model selection procedure when the sample size $n$ is large. The asymptotic optimality of GCV, meaning that $\lVert f^{k_{\textnormal{GCV}}} - f^* \rVert_n^2 \ / \ \underset{k}{\inf}\lVert f^{k} - f^* \rVert_n^2 \to 1$ in probability, has been proved for the $k$-NN estimator in (\cite{li1987asymptotic}) under mild assumptions. As for the AIC criterion, the computational time of generalized cross-validation is $\mathcal{O}\left( n^3 \right)$ elementary operations. \subsection*{Hold-out cross-validation stopping rule.} The Hold-out cross-validation strategy (\cite{geisser1975predictive, arlot2010survey}) is described as follows. The data $\{ x_i, y_i \}_{i=1}^n$ are randomly split into two parts of equal size: the training sample $S_{\textnormal{train}} = \{ x_{\textnormal{train}}, y_{\textnormal{train}} \}$ and the test sample $S_{\textnormal{test}} = \{ x_{\textnormal{test}}, y_{\textnormal{test}} \}$ so that the training and test samples represent a half of the whole data set $\approx n/2$. For each $k = 1, \ldots, n$, one trains the $k$-NN estimator (\ref{estimator}) and evaluates its performance by $R_{\text{HO}}(f^k) = \frac{1}{n}\sum_{i \in S_{\textnormal{test}}}(f^k(x_i)- y_i)^2$, where $f^k(x_i)$ denotes the output of the algorithm trained for $k$ and evaluated at the point $x_i \in x_{\textnormal{test}}$. Then, the Hold-out CV stopping rule is defined via \begin{equation} \label{k_ho} k_{\text{HO}} \coloneqq \underset{k = 2, \ldots, n}{\textnormal{argmin}} \left\{ R_{\text{HO}}(f^k) \right\} - 1. \end{equation} The main inconvenience of this model selection rule is the fact that a part of the data is lost, which increases the risk error. Besides that, the Hold-out strategy is not stable (\cite{arlot2010survey}), which often requires some aggregation of it. The (asymptotic) computational time of the Hold-out strategy is $\mathcal{O}\left( n^3 \right)$ elementary operations. \subsection*{$V$--fold cross-validation} $V$--fold cross-validation is certainly the most used cross-validation procedure: the data $\{ (x_i, y_i) \}_{i=1}^n$ are randomly split into $V=5$ equal sized blocks, and at each round (among the $V$ ones), $V-1$ blocks are devoted to training $S_{\text{train}} = (x_{\text{train}}, y_{\text{train}})$, and the remaining one is used for the evaluation $S_{\text{test}} = (x_{\text{test}}, y_{\text{test}})$. The risk error of the $k$-NN estimator is estimated by $R_{\text{VFCV}}(f^k) = \frac{1}{V}\sum_{j=1}^{V} \frac{1}{n/V}\sum_{i \in S_{\text{test}}(j)}\left( f^k(x_i) - y_i \right)^2$, where $f^k(x_i)$ denotes the output of the algorithm trained for $k$ and evaluated at the point $x_i \in S_{\text{test}}(j)$, thus \begin{equation} \label{k_vfcv} k_{\text{VFCV}} \coloneqq \underset{k = 2, \ldots, n}{\textnormal{argmin}} \left\{ R_{\text{VFCV}}(f^k) \right\} - 1. \end{equation} $V$--fold cross-validation is a more computationally tractable solution than other splitting-based model selection methods, such as the leave-one-out or leave-$p$-out (\cite{hastie2009elements, arlot2010survey}). Usually, the optimal $V$ is equal to $5$ or $10$ due to the fact that the statistical error does not increase a lot for larger values of $V$ whereas averaging over more than $10$ folds becomes infeasible. To the best of our knowledge, there are no theoretical results for the $V$--fold cross validation model selection strategy with the $k$-NN regression estimator. \subsection*{Theoretical bias-variance trade-off stopping rule} The fourth stopping rule is the one introduced in Eq. (\ref{k_star}). This stopping rule is the classical bias-variance trade-off stopping rule that provides minimax-optimal rates (see the monographs \cite{wasserman2006all, tsybakov2008introduction}): \begin{equation} k^* = \inf \{ k \in \{ 1, \ldots, n \} \mid B^2(k) \geq V(k) \}. \end{equation} The stopping rule $k^*$ is introduced for comparison purposes only because it \textit{cannot be computed} in practice. One can say that this stopping rule is minimax-optimal if $f^*$ belongs, for instance, to the class of Lipschitz functions on a bounded domain (\ref{lipschitz}). Therefore, it could serve as a (lower bound) reference in the present simulated experiments with artificial data. \subsection{Artificial data} \label{artificial_data} First, the goal is to perform some simulated experiments (a comparison of mentioned stopping rules) on artificial data. \vspace{0.2cm} \textbf{Description of the simulation design} The data in this case is generated according to the regression model $y_j = f^*(x_j) + \varepsilon_j$, where $\varepsilon_j \overset{i.i.d.}{\sim} \mathcal{N}(0, \sigma^2)$ (Gaussian), $j = 1, \ldots, n$. We choose the uniform covariates $x_j \overset{i.i.d.}{\sim} \mathbb{U}[0, 1]^3$ (uniform), $j = 1, \ldots, n,$ and $\sigma = 0.15$. Consider two regression functions with different smoothness: a "smooth" $f_1^*(x) = 1.5 \cdot \left[ \lVert x - 0.5 \rVert / \sqrt{3} - 0.5 \right]$ and a "sinus" $f_2^*(x) = 1.5 \cdot \sin (\lVert x \rVert / \sqrt{3})$ for any $x \in [0, 1]^3$. Notice that both functions belong to the class of Lipschitz functions (\ref{lipschitz}) on $[0, 1]^3$. The sample size $n$ varies from $50$ to $250$. The $k$-NN algorithm (\ref{estimator}) is trained first for $k = \floor*{\sqrt{n}}$, after that we decrease the value of $k$ until $k = 1$ such that at each step of the iteration procedure we increase the variance of the $k$-NN estimator $V(k) = \sigma^2/k$ (cf. Fig. \ref{fig:bvr}). In other words, the model becomes more complex successively due to the increase of its "degree of freedom" measured by $\textnormal{tr}(A_k) = n/k$. If the condition in Eq. (\ref{k_tau}) is satisfied, the learning process is stopped, and it outputs the stopping rule $k^{\tau}$. The performance of the stopping rules is measured in terms of the empirical $L_2(\mathbb{P}_n)$-norm $\lVert f^k - f^* \rVert_n^2$ averaged over $N = 1000$ repetitions (over the noise $\{ \varepsilon_j \}_{j=1}^n$). For our simulations, we use a consistent (low-bias) estimator of $\sigma^2$ described in Section \ref{real_data} (see Eq. (\ref{var_est})). \begin{figure} \begin{subfigure}[b]{0.45\textwidth} \begin{tikzpicture}[scale=0.7] \tikzstyle{every node}=[font=\normalsize] \begin{axis}[ title={$k$-NN regressor, $\sigma = 0.15$, smooth r.f.}, xlabel={Sample size}, ylabel={Average loss}, xmin=50, xmax=250, ymin=0.004, ymax=0.0115, ytick={0.005,0.007,0.009}, xtick={50, 100, 150, 200, 250}, legend pos=north east, ymajorgrids=true, xmajorgrids=true, grid style=dashed, ] \addplot[ color=blue, mark=star, ] coordinates { (50,0.00959589)(80,0.00723882)(100,0.0064927)(160,0.00538029)(200,0.00527075)(250,0.00440348) }; \addlegendentry{$k^{\tau}$} \addplot[ color=red, mark=+, ] coordinates { (50, 0.01117475)(80, 0.00820779)(100, 0.00692115)(160, 0.00578596)(200, 0.00556723)(250,0.00480027) }; \addlegendentry{$k_{\textnormal{HO}}$} \addplot[ color=black, mark=x, ] coordinates { (50, 0.00960814)(80, 0.00708104)(100, 0.0064098)(160, 0.00534147)(200, 0.005275)(250,0.00434834) }; \addlegendentry{$k_{\textnormal{GCV}}$} \addplot[ color=green, mark=star, ] coordinates { (50, 0.00934164)(80, 0.00708591)(100, 0.0062164)(160, 0.00527903)(200, 0.00512319)(250,0.00420769) }; \addlegendentry{$k^{*}$} \end{axis} \end{tikzpicture} \caption{} \label{smooth} \end{subfigure} \qquad \begin{subfigure}[b]{0.45\textwidth} \begin{tikzpicture}[scale=0.7] \tikzstyle{every node}=[font=\normalsize] \begin{axis}[ title={$k$-NN regressor, $\sigma = 0.15$, sinus r.f.}, xlabel={Sample size}, ylabel={Average loss}, xmin=50, xmax=250, ymin=0.0025, ymax=0.0095, xtick={50,100,150,200,250}, ytick={0.003,0.005, 0.008}, legend pos=north east, ymajorgrids=true, xmajorgrids=true, grid style=dashed, ] \addplot[ color=blue, mark=star, ] coordinates { (50,0.00856475)(80,0.00527245)(100,0.00520583)(160,0.00414462)(200,0.00360486)(250,0.0029507) }; \addlegendentry{$k^{\tau}$} \addplot[ color=red, mark=+, ] coordinates { (50, 0.00932031)(80, 0.00640974)(100, 0.00562799)(160, 0.00452315)(200, 0.00401955)(250,0.00365806) }; \addlegendentry{$k_{\textnormal{HO}}$} \addplot[ color=black, mark=x, ] coordinates { (50, 0.00863128)(80, 0.00534549)(100, 0.00504782)(160, 0.00411817)(200, 0.00362582)(250,0.00289884) }; \addlegendentry{$k_{\textnormal{GCV}}$} \addplot[ color=green, mark=star, ] coordinates { (50, 0.00861029)(80, 0.00477626)(100, 0.00499522)(160, 0.00389686)(200, 0.00349439)(250,0.00275263) }; \addlegendentry{$k^{*}$} \end{axis} \end{tikzpicture} \caption{} \label{sinus} \end{subfigure} \caption{$k$-NN estimator (\ref{estimator}) with two noised regression functions: smooth $f_1^*(x) = 1.5 \cdot \left[ \lVert x - 0.5 \rVert / \sqrt{3} - 0.5 \right]$ for panel (a) and "sinus" $f_2^*(x) = 1.5 \cdot \textnormal{sin}(\lVert x \rVert / \sqrt{3})$ for panel (b), with uniform covariates $x_j \overset{i.i.d.}{\sim} \mathbb{U}[0, 1]^3$. Each curve corresponds to the $L_2(\mathbb{P}_n)$ squared norm error for the stopping rules (\ref{k_tau}), (\ref{k_star}), (\ref{k_ho}), (\ref{k_1_out}), averaged over $1000$ independent trials, versus the sample size $n = \{50, 80, 100, 160, 200, 250 \}$.} \label{fig:comp_knn_art} \end{figure} \textbf{Results of the simulation experiments.} Figure \ref{fig:comp_knn_art} displays the resulting (averaged over $1000$ repetitions) $L_2(\mathbb{P}_n)$-error of $k^{\tau}$ (\ref{k_tau}), $k^*$ (\ref{k_star}), $k_{\text{HO}}$ (\ref{k_ho}), and $k_{\text{GCV}}$ (\ref{k_1_out}) versus the sample size $n$. In particular, Figure \ref{smooth} shows the results for the "smooth" regression function, whereas Figure \ref{sinus} provides the results for the "sinus" regression function. First, from all the graphs, (almost) all curves do not increase as the sample size $n$ grows. One can notice that in both graphs, the prediction error of the Hold-out strategy is the worst among the model selection criteria. In more detail, Figure \ref{smooth} indicates that the best performance is achieved by $k^*$ (non-computable in practice bias-variance trade-off). Besides that, the minimum discrepancy principle rule $k^{\tau}$ is uniformly better than $k_{\textnormal{HO}}$ and has the same performance as the one of $k_{\textnormal{GCV}}$. Moreover, the gap between $k^{\tau}$ and $k^*$ is getting smaller for the sample sizes $n \geq 160$. This behavior supports the theoretical part of the present paper because $k^{\tau}$ should serve as an estimator of $k^*$. Since $k^*$ is the well-known bias-variance trade-off, the minimum discrepancy principle stopping rule seems to be a meaningful model selection method. In Figure \ref{sinus}, the best performance is achieved again by $k^*$ -- a non-computable in practice stopping rule. As for the data-driven model selection methods, the stopping rules $k^{\tau}$ and $k_{\text{GCV}}$ (an asymptotically optimal model selection strategy) perform almost equivalently. \subsection{Real data} \label{real_data} Here, we tested the performance (prediction error and runtime) of the early stopping rule $k^{\tau}$ (\ref{k_tau}) for choosing the hyperparameter in the $k$-NN estimator on four different data sets mostly taken from the UCI repository (\cite{Dua:2019}). \textbf{Data sets description} The housing data set "Boston Housing Prices" concerns the task of predicting housing values in areas of Boston (USA), the input points are $13$-dimensional. The "Diabetes" data set consists of 10 columns that measure different patient's characteristics (age, sex, body mass index, etc), the output is a quantitative measure of disease progression one year after the baseline. The "Power Plants" data set contains 9568 data points collected from a Combined Cycle Power Plant over 6 years (2006-2011), when the plant was set to work with the full load. "California Houses Prices" data set (\cite{pace1997sparse}) contains information from the 1990's California census. The input variables are "total bedrooms", "total rooms", etc. The output variable is the median house value for households within a block (measured in US Dollars). Notice that for "California Houses Prices" and "Power Plants" data sets we take the first 3000 data points in order to speed up the calculations. \vspace{0.2cm} \textbf{Description of the simulation design} Assume that we are given one of the data sets described above. Let us rescale each variable of this data set $\widetilde{x} \in \mathbb{R}^n$ such that all the components $\widetilde{x}_i, \ i = 1, \ldots, n$, belong to $[0, 1]$: \begin{equation*} \widetilde{x}_i = \frac{\widetilde{x}_i - \min(\widetilde{x})}{\max(\widetilde{x}) - \min(\widetilde{x})}, \ i = 1, \ldots, n, \end{equation*} where $\min(\widetilde{x})$ and $\max(\widetilde{x})$ denote the minimum and the maximum component of the vector $\widetilde{x}$. After that, we split the data set into two parts: one is denoted $S_{\textnormal{train}} = \{ x_{\textnormal{train}}, y_{\textnormal{train}} \}$ (70 \% of the whole data) and is made for training and model selection (model selection rules $k^{\tau}$, $k_{\text{GCV}}$, $k_{\text{5FCV}}$, and $k_{\text{AIC}}$), the other one (30 \% of the whole data) is denoted $S_{\textnormal{test}} = \{ x_{\textnormal{test}}, y_{\textnormal{test}} \}$ and is made for making prediction on it. We denote $n_{\textnormal{train}}$ and $n_{\textnormal{test}}$ as the sample sizes of $S_{\textnormal{train}}$ and $S_{\textnormal{test}}$, respectively. Then, our experiments' design is divided into four parts. \vspace{0.15cm} At the beginning, we create a grid of sub-sample size for each data set: \begin{equation} \label{subsamples} n_s \in \Big\{\floor*{n_{\textnormal{train}}/5}, \floor*{n_{\textnormal{train}}/4}, \floor*{n_{\textnormal{train}}/3}, \\ \floor*{n_{\textnormal{train}}/2}, n_{\textnormal{train}}\Big\}, \end{equation} and a grid of the maximum number of neighbors $k_{\textnormal{max}} = 3 \floor*{\log(n_s)}$, where $n_{\textnormal{train}} = \ceil*{0.7n}$ and $n$ is the sample size of the whole data. Further, for each data set and sub-sample size from Eq. (\ref{subsamples}), we estimate the noise variance $\sigma^2$ from the regression model (\ref{nonparametric_regression_model}). There is large amount of work on the estimation of $\sigma^2$ in nonparametric regression (\cite{rice1984bandwidth, hall1990variance}). In our simulated experiments, we take the estimator from \cite[Eq. (5.86)]{wasserman2006all}, which is a consistent estimator of $\sigma^2$ under an assumption that $f^*$ is "sufficiently smooth". It satisfies our simulation experiments' purposes. \begin{equation} \label{var_est} \widehat{\sigma}^2 \coloneqq \frac{\lVert (I_{n_s} - A_k)y_{s} \rVert^2}{n_s (1 - 1/k)} \ \ \ \textnormal{ with } \ \ \ k = 2, \end{equation} where $y_s$ corresponds to the vector of responses from the chosen sub-samples. After that, we compute our stopping rule $k^{\tau}$ and other model selection strategies from Section \ref{description_early_stopping_rules}. To do that, for each data set and each integer $n_s$ from Eq. (\ref{subsamples}), we randomly sample $n_s$ data points from $S_{\textnormal{train}}$, compute the $k$-NN estimator (\ref{estimator}) and the empirical risk (\ref{empirical_risk}) for $k_{\textnormal{max}}$, and at each step of the iteration process we reduce the value of $k$ by one. Remark that one does not have to calculate the neighborhood matrix $A_k$ for each $k \in \{1, \ldots, k_{\textnormal{max}}\}$, since it is sufficient to do only for $k_{\textnormal{max}}$ (cf. Eq. (\ref{iteration_for_matrix})). This process is repeated until the empirical risk crosses the threshold $\widehat{\sigma}^2$. Fig. \ref{fig:process_learning} provides two illustrations of the minimum discrepancy strategy $k^{\tau}$ for two data sets: "Diabetes" and "Boston Housing Prices". After that, the AIC criterion (\ref{k_aic}), $5$--fold cross-validation (\ref{k_vfcv}), and the generalized cross-validation $k_{\textnormal{GCV}}$ are calculated. Let us describe how we do that in two steps. We start by defining a grid of values for $k: \{ 1, 2, \ldots, k_{\textnormal{max}} \}$. Further, one should compute $k_{\textnormal{AIC}}$, $k_{\textnormal{GCV}}$, and $k_{\textnormal{5FCV}}$ from Eq. (\ref{k_aic}), Eq. (\ref{k_1_out}), and Eq. (\ref{k_vfcv}) over the mentioned grid. In the final part, given $k^{\tau}$, $k_{\text{AIC}}$, $k_{\text{5FCV}}$, and $k_{\text{GCV}}$, the goal is to make a prediction on the test data set $x_{\textnormal{test}}$. This can be done as follows. Assume that $x_0 \in x_{\textnormal{test}}$, then the prediction of the $k$-NN estimator on this point can be defined as \begin{equation*} f^k(x_0) = a_k(x_0)^\top y_{s}, \end{equation*} where $a_k(x_0) = [a_k(x_0, x_1), \ldots, a_k(x_0, x_{n_s})]^\top$ and $x_{s} = [x_1^{\top}, \ldots, x_{n_{s}}^\top]^{\top}$, with $a_k(x_0, x_i) = 1/k$ if $x_i, \ i \in \{1, \ldots, n_{\textnormal{train}} \}$, is a nearest neighbor of $x_0$, otherwise $0$. Further, one can choose $k$ to be equal $k^{\tau}$, $k_{\text{AIC}}$, $k_{\text{5FCV}}$ or $k_{\text{GCV}}$ that are already computed. Combining all the steps together, one is able to calculate the least-squares prediction error $\lVert f^k - y_{\textnormal{test}} \rVert$. For each sub-sample size $n_s$ from Eq. (\ref{subsamples}) and data set, the procedure has to be performed $25$ times (via new sub-samples from the data set). \vspace{0.2cm} \textbf{Results of the simulation experiments.} Figures \ref{small_datasets_results} and \ref{large_datasets_results} display the averaged (over 25 repetitions) runtime (in seconds) and the prediction error of the model selection rules $k^{\tau}$ (\ref{k_tau}), $k_{\textnormal{AIC}}$ (\ref{k_aic}), 5-fold cross-validation (\ref{k_vfcv}), and generalized cross-validation (\ref{k_1_out}) for "Boston Housing Prices", "Diabetes" (in Figure \ref{small_datasets_results}), and "California Houses Prices", "Power Plants" data sets (in Figure \ref{large_datasets_results}). Figures \ref{runtime_boston}, \ref{runtime_diabetes} indicate that the minimum discrepancy principle rule $k^{\tau}$ has the smallest runtime among the model selection criteria. At the same time, Figure \ref{prediction_error_boston} shows that the prediction error of $k^{\tau}$ has better performance than that of $k_{\textnormal{AIC}}$ and $k_{\textnormal{5FCV}}$, and is similar to that of $k_{\textnormal{GCV}}$. Figure \ref{pred_error_diabetes} indicates that the performance of the minimum discrepancy stopping rule $k^{\tau}$ is better than that of $k_{\textnormal{AIC}}$, $k_{\textnormal{GCV}}$, and $k_{\textnormal{5FCV}}$ for the sub-sample size $n_s \geq 100$. Let us turn to the results for the "California Houses Prices" and "Power Plants" data sets. Figures \ref{prediction_error_california}, \ref{prediction_error_power} display the prediction performance of the model selection rules: for the "California Houses Prices" data set, the prediction performance of $k^{\tau}$ is comparable to the performance of $k_{\textnormal{GCV}}$ and is uniformly better than the $5$FCV rule $k_{\textnormal{5FCV}}$ and $k_{\textnormal{AIC}}$; for the "Power Plants" data set, the prediction error of the minimum discrepancy principle is similar to that of $k_{\textnormal{AIC}}, k_{\textnormal{GCV}}$, and $k_{\textnormal{5FCV}}$ for the sub-sample sizes $n_s \leq 1000$, and is little worse for $n_s = 2100$. This deterioration of the performance could be due to the estimation of the variance $\sigma^2$ from Eq. (\ref{var_est}). Figures \ref{runtime_california}, \ref{runtime_power} show the runtime of the stopping rules: one can conclude that the computational time of the minimum discrepancy rule $k^{\tau}$ is less than the computational time of the generalized cross-validation, AIC criterion, and $5$ fold cross-validation. The overall conclusion from the simulation experiments is that the prediction error of the MDP stopping rule $k^{\tau}$ is often better than for standard model selection strategies, such as the AIC criterion or $5$--fold cross-validation, while its computational time is lower. \begin{figure} \begin{subfigure}[b]{0.45\textwidth} \begin{tikzpicture}[scale=0.7] \tikzstyle{every node}=[font=\normalsize] \begin{axis}[ title={Diabetes data (sub-sample)}, xlabel={Number of neighbors}, ylabel={Value}, xmin=0, xmax=10, ymin=0, ymax=6114.66446281, xtick={1,3,5,7, 10}, ytick={1500, 4500}, legend pos=south east, ymajorgrids=true, xmajorgrids=true, grid style=dashed, ] \addplot[ color=blue, mark=square, ] coordinates { (1,0)(2,0)(3,3664.18333333)(4,4731.1125)(5,5358.508)(6,5680.28472222)(7,5452.15408163)(8,5851.84765625)(9,5631.3308642)(10,6114.66446281) }; \addlegendentry{Empirical risk} \addplot[ color=red, mark=+, ] coordinates { (1,4215.1)(2,4215.1)(3,4215.1)(4,4215.1)(5,4215.1)(6,4215.1)(7,4215.1)(8,4215.1)(9,4215.1)(10,4215.1) }; \addlegendentry{Threshold} \addplot +[mark=none, color=orange] coordinates {(3, 0)(3, 6114.66446281)}; \addlegendentry{$k^{\tau}$} \end{axis} \end{tikzpicture} \caption{} \label{a} \end{subfigure} \qquad \begin{subfigure}[b]{0.45\textwidth} \begin{tikzpicture}[scale=0.7] \tikzstyle{every node}=[font=\normalsize] \begin{axis}[ title={Boston Housing data (sub-sample)}, xlabel={Number of neighbors}, ylabel={Value}, xmin=0, xmax=10, ymin=0, ymax=68.4, xtick={1,3,5,10}, ytick={20, 45, 65}, legend pos=south east, ymajorgrids=true, xmajorgrids=true, grid style=dashed, ] \addplot[ color=blue, mark=square, ] coordinates { (1,0)(2,0)(3,0)(4,49.69747283)(5,63.26987826)(6,60.0285628)(7,60.71023957)(8,67.93981658)(9,67.28213634)(10,68.37776957) }; \addlegendentry{Empirical risk} \addplot[ color=red, mark=+, ] coordinates { (1, 53.79369565)(2, 53.79369565)(3, 53.79369565)(4, 53.79369565)(5, 53.79369565)(6,53.79369565)(7,53.79369565)(8,53.79369565)(9,53.79369565)(10,53.79369565) }; \addlegendentry{Threshold} \addplot +[mark=none, color=orange] coordinates {(4, 0)(4, 6114.66446281)}; \addlegendentry{$k^{\tau}$} \end{axis} \end{tikzpicture} \caption{} \label{b} \end{subfigure} \caption{Stopping the learning process based on the rule (\ref{k_tau}) applied to two data sets: a) "Diabetes" and b) "Boston Housing Prices". "Threshold" horizontal line corresponds to the estimated variance from Eq. (\ref{var_est}).} \label{fig:process_learning} \end{figure} \begin{figure} \begin{subfigure}[b]{0.45\textwidth} \begin{tikzpicture}[scale=0.65] \tikzstyle{every node}=[font=\normalsize] \begin{axis}[ title={Boston H.P., runtime}, xlabel={Sub-sample size}, ylabel={Average time (sec)}, xmin=70, xmax=354, ymin=0.003, ymax=0.11, xtick={70,100,140,200,300,400}, ytick={0.01,0.1}, legend pos=north west, ymajorgrids=true, xmajorgrids=true, grid style=dashed, ] \addplot[ color=blue, mark=square, ] coordinates { (70,0.00855612)(88,0.01127354)(118,0.01240211)(177,0.01705173)(354,0.05850043) }; \addlegendentry{$k^{\tau}$} \addplot[ color=red, mark=+, ] coordinates { (70, 0.01082651)(88, 0.01258839)(118, 0.01454256)(177, 0.01753586)(354, 0.06210802) }; \addlegendentry{$k_{\textnormal{AIC}}$} \addplot[ color=black, mark=x, ] coordinates { (70, 0.00830978)(88, 0.02092349)(118, 0.01145007)(177, 0.01990505)(354, 0.06201828) }; \addlegendentry{$k_{\textnormal{GCV}}$} \addplot[ color=violet, mark=*, ] coordinates { (70, 0.02412378)(88, 0.0344868)(118, 0.0204067)(177, 0.03076374)(354, 0.09219963) }; \addlegendentry{$k_{\textnormal{5FCV}}$} \end{axis} \end{tikzpicture} \caption{} \label{runtime_boston} \end{subfigure} \qquad \begin{subfigure}[b]{0.45\textwidth} \begin{tikzpicture}[scale=0.65] \tikzstyle{every node}=[font=\normalsize] \begin{axis}[ title={Boston H.P., prediction error}, xlabel={Sub-sample size}, ylabel={Average loss}, xmin=70, xmax=354, ymin=50, ymax=85, xtick={70,100,140,200,300,400}, ytick={60, 70, 80}, legend pos=north east, ymajorgrids=true, xmajorgrids=true, grid style=dashed, ] \addplot[ color=blue, mark=square, ] coordinates { (70,74.2853426)(88,73.28846512)(118,67.86918275)(177,62.26424591)(354,50.59923418) }; \addlegendentry{$k^{\tau}$} \addplot[ color=red, mark=+, ] coordinates { (70, 75.8850821)(88, 73.72479664)(118, 67.90502476)(177, 62.45029884)(354, 53.17254931) }; \addlegendentry{$k_{\textnormal{AIC}}$} \addplot[ color=black, mark=x, ] coordinates { (70, 75.6347345)(88, 72.5862216)(118, 68.21207572)(177, 63.64873349)(354, 50.59923418) }; \addlegendentry{$k_{\textnormal{GCV}}$} \addplot[ color=violet, mark=*, ] coordinates { (70, 75.20374274)(88, 72.83081082)(118, 68.20018387)(177, 63.40840655)(354, 52.9666841) }; \addlegendentry{$k_{\textnormal{5FCV}}$} \end{axis} \end{tikzpicture} \caption{} \label{prediction_error_boston} \end{subfigure} \\ \begin{subfigure}[b]{0.45\textwidth} \begin{tikzpicture}[scale=0.65] \tikzstyle{every node}=[font=\normalsize] \begin{axis}[ title={Diabetes, runtime}, xlabel={Sub-sample size}, ylabel={Average time (sec)}, xmin=60, xmax=309, ymin=0.003, ymax=0.09, xtick={61,100,140,200,300}, ytick={0.01,0.05}, legend pos=north west, ymajorgrids=true, xmajorgrids=true, grid style=dashed, ] \addplot[ color=blue, mark=square, ] coordinates { (61,0.00528295)(77,0.006069)(103,0.00993328)(154,0.01101583)(309,0.04544517) }; \addlegendentry{$k^{\tau}$} \addplot[ color=red, mark=+, ] coordinates { (61, 0.0069945)(77, 0.00938141)(103, 0.00800133)(154, 0.01413276)(309, 0.05715765) }; \addlegendentry{$k_{\textnormal{AIC}}$} \addplot[ color=black, mark=x, ] coordinates { (61, 0.00631852)(77, 0.0088942)(103, 0.00800728)(154, 0.01413908)(309, 0.0594244) }; \addlegendentry{$k_{\textnormal{GCV}}$} \addplot[ color=violet, mark=*, ] coordinates { (61, 0.01438442)(77, 0.01810751)(103, 0.01518609)(154, 0.02330532)(309, 0.09792542) }; \addlegendentry{$k_{\textnormal{5FCV}}$} \end{axis} \end{tikzpicture} \caption{} \label{runtime_diabetes} \end{subfigure} \qquad \begin{subfigure}[b]{0.45\textwidth} \begin{tikzpicture}[scale=0.65] \tikzstyle{every node}=[font=\normalsize] \begin{axis}[ title={Diabetes, prediction error}, xlabel={Sub-sample size}, ylabel={Average loss}, xmin=61, xmax=309, ymin=630, ymax=740, xtick={61,100,140,200,300}, ytick={650, 700, 725}, legend pos=north east, ymajorgrids=true, xmajorgrids=true, grid style=dashed, ] \addplot[ color=blue, mark=square, ] coordinates { (61,735.32410221)(77,698.07531817)(103,675.96322628)(154,653.35361124)(309,637.9835621) }; \addlegendentry{$k^{\tau}$} \addplot[ color=red, mark=+, ] coordinates { (61, 729.27694327)(77, 694.41295062)(103, 676.86668009)(154, 658.35153366)(309, 649.31011705) }; \addlegendentry{$k_{\textnormal{AIC}}$} \addplot[ color=black, mark=x, ] coordinates { (61, 715.16104155)(77, 694.23348373)(103, 679.64844807)(154, 656.60010857)(309, 649.31011705) }; \addlegendentry{$k_{\textnormal{GCV}}$} \addplot[ color=violet, mark=*, ] coordinates { (61, 730.96038738)(77, 716.52927512)(103, 679.82796817)(154, 674.75239488)(309, 645.99680041) }; \addlegendentry{$k_{\textnormal{5FCV}}$} \end{axis} \end{tikzpicture} \caption{} \label{pred_error_diabetes} \end{subfigure} \caption{Runtime (in seconds) and $L_2(\mathbb{P}_n)$ prediction error versus sub-sample size for different model selection methods: MD principle (\ref{k_tau}), AIC (\ref{k_aic}), GCV (\ref{k_1_out}), and $5$--fold cross-validation (\ref{k_vfcv}), tested on the "Boston Housing Prices" and "Diabetes" data sets. In all cases, each point corresponds to the average of $25$ trials. (a), (c) Runtime verus the sub-sample size $n \in \{70, 88, 118, 177, 354 \}$. (b), (d) Least-squares prediction error $\lVert f^k - y_{\textnormal{test}} \rVert$ versus the sub-sample size $n \in \{70, 88, 118, 177, 354 \}$.} \label{small_datasets_results} \end{figure} \begin{figure} \begin{subfigure}[b]{0.45\textwidth} \begin{tikzpicture}[scale=0.65] \tikzstyle{every node}=[font=\normalsize] \begin{axis}[ title={California H.P., runtime}, xlabel={Sub-sample size}, ylabel={Average time (sec)}, xmin=420, xmax=2100, ymin=0.02, ymax=1.65, xtick={500,700,1000,2000}, ytick={0.1,0.5, 1}, legend pos=north west, ymajorgrids=true, xmajorgrids=true, grid style=dashed, ] \addplot[ color=blue, mark=square, ] coordinates { (420,0.06766566)(525,0.1230032)(700,0.14770074)(1050,0.40171222)(2100,1.35318375) }; \addlegendentry{$k^{\tau}$} \addplot[ color=red, mark=+, ] coordinates { (420,0.0669849)(525,0.11980861)(700,0.21817273)(1050,0.43286912)(2100,1.48403368) }; \addlegendentry{$k_{\textnormal{AIC}}$} \addplot[ color=black, mark=x, ] coordinates { (420,0.06887476)(525,0.12885856)(700,0.17433068)(1050,0.42578732)(2100,1.45269584) }; \addlegendentry{$k_{\textnormal{GCV}}$} \addplot[ color=violet, mark=*, ] coordinates { (420,0.11478419)(525,0.16037658)(700,0.26191102)(1050,0.63462547)(2100,1.92028911) }; \addlegendentry{$k_{\textnormal{5FCV}}$} \end{axis} \end{tikzpicture} \caption{} \label{runtime_california} \end{subfigure} \qquad \begin{subfigure}[b]{0.45\textwidth} \begin{tikzpicture}[scale=0.65] \tikzstyle{every node}=[font=\normalsize] \begin{axis}[ title={California H.P., prediction error}, xlabel={Sub-sample size}, ylabel={Average loss}, xmin=420, xmax=2100, ymin=11.2, ymax=13.5, xtick={500,700,1000,2000}, ytick={12 ,13 , 13.5}, legend pos=north east, ymajorgrids=true, xmajorgrids=true, grid style=dashed, ] \addplot[ color=blue, mark=square, ] coordinates { (420,13.19666715)(525,13.02057484)(700,12.61806828)(1050,12.12776051)(2100,11.43766597) }; \addlegendentry{$k^{\tau}$} \addplot[ color=red, mark=+, ] coordinates { (420,13.17638071)(525,13.1502115)(700,12.76413759)(1050,12.18476347)(2100,11.43766597) }; \addlegendentry{$k_{\textnormal{AIC}}$} \addplot[ color=black, mark=x, ] coordinates { (420,13.15911354)(525,12.98821598)(700,12.54897291)(1050,12.10185179)(2100,11.46773943) }; \addlegendentry{$k_{\textnormal{GCV}}$} \addplot[ color=violet, mark=*, ] coordinates { (420,13.3404094)(525,13.27833772)(700,12.66432926)(1050,12.27531468)(2100,11.57245704) }; \addlegendentry{$k_{\textnormal{5FCV}}$} \end{axis} \end{tikzpicture} \caption{} \label{prediction_error_california} \end{subfigure} \\ \begin{subfigure}[b]{0.45\textwidth} \begin{tikzpicture}[scale=0.65] \tikzstyle{every node}=[font=\normalsize] \begin{axis}[ title={Power Plants, runtime}, xlabel={Sub-sample size}, ylabel={Average time (sec)}, xmin=350, xmax=2100, ymin=0.03, ymax=1.8, xtick={420,700,1000,2000}, ytick={0.1, 1, 2}, legend pos=north west, ymajorgrids=true, xmajorgrids=true, grid style=dashed, ] \addplot[ color=blue, mark=square, ] coordinates { (420,0.05278417)(525,0.09394525)(700,0.14946687)(1050,0.32476645)(2100,1.31787278) }; \addlegendentry{$k^{\tau}$} \addplot[ color=red, mark=+, ] coordinates { (420, 0.05410801)(525, 0.11433973)(700, 0.18229741)(1050, 0.38104916)(2100, 1.55990878) }; \addlegendentry{$k_{\textnormal{AIC}}$} \addplot[ color=black, mark=x, ] coordinates { (420, 0.05847129)(525, 0.09902878)(700, 0.16085543)(1050, 0.36054396)(2100, 1.41506519) }; \addlegendentry{$k_{\textnormal{GCV}}$} \addplot[ color=violet, mark=*, ] coordinates { (420, 0.08910254)(525, 0.17428734)(700, 0.21586569)(1050, 0.49328215)(2100, 1.84269897) }; \addlegendentry{$k_{\textnormal{5FCV}}$} \end{axis} \end{tikzpicture} \caption{} \label{runtime_power} \end{subfigure} \qquad \begin{subfigure}[b]{0.45\textwidth} \begin{tikzpicture}[scale=0.65] \tikzstyle{every node}=[font=\normalsize] \begin{axis}[ title={Power Plants, prediction error}, xlabel={Sub-sample size}, ylabel={Average loss}, xmin=350, xmax=2100, ymin=115, ymax=136, xtick={420,700,1000,2000}, ytick={120, 125, 130}, legend pos=north east, ymajorgrids=true, xmajorgrids=true, grid style=dashed, ] \addplot[ color=blue, mark=square, ] coordinates { (420,134.9333523)(525,131.49142363)(700,128.50263383)(1050,124.02441083)(2100,117.00074146) }; \addlegendentry{$k^{\tau}$} \addplot[ color=red, mark=+, ] coordinates { (420, 136.59589921)(525, 131.38593144)(700, 129.14665047)(1050, 124.20131175)(2100, 115.65787105) }; \addlegendentry{$k_{\textnormal{AIC}}$} \addplot[ color=black, mark=x, ] coordinates { (420, 135.36142759)(525, 131.05128992)(700, 128.72729084)(1050, 124.05305432)(2100, 115.65787105) }; \addlegendentry{$k_{\textnormal{GCV}}$} \addplot[ color=violet, mark=*, ] coordinates { (420, 135.40563633)(525, 131.06600903)(700, 128.2952843)(1050, 124.04019418)(2100, 116.08168959) }; \addlegendentry{$k_{\textnormal{5FCV}}$} \end{axis} \end{tikzpicture} \caption{} \label{prediction_error_power} \end{subfigure} \caption{Runtime (in seconds) and $L_2(\mathbb{P}_n)$ prediction error versus sub-sample size for different model selection methods: MDP (\ref{k_tau}), AIC (\ref{k_aic}), GCV (\ref{k_1_out}), and $5$--fold cross-validation (\ref{k_vfcv}), tested on the "California Houses Prices" and "Power Plants" data set. In all cases, each point corresponds to the average of $25$ trials. (a), (c) Runtime verus the sub-sample size $n \in \{420, 525, 700, 1050, 2100 \}$. (b), (d) Least-squares prediction error $\lVert f^k - y_{\textnormal{test}} \rVert$ versus the sub-sample size $n \in \{420, 525, 700, 1050, 2100 \}$.} \label{large_datasets_results} \end{figure} \section{Conclusion} \label{sec:conc} In the present paper, we tackled the problem of choosing the tuning parameter $k$ in the $k$-NN regression estimator. A strategy based on early stopping and the minimum discrepancy principle was proposed. In Section \ref{sec:optimality_result}, it was shown that the minimum discrepancy stopping rule $k^{\tau}$ (\ref{k_tau}) provides a minimax-optimal estimator, in particular, over the class of Lipschitz functions on a bounded domain. Besides that, this theoretical result was confirmed empirically on artificial and real data sets: the stopping rule has better statistical performance than other model selection rules, such as Hold-out or $5$--fold cross-validation, while reducing the computational time of the model selection procedure. The main inconvenience of the proposed strategy is that one has to estimate the variance $\sigma^2$ of the regression model (as it is the case for the AIC or Mallow's $C_p$ criteria), thus a plug-in estimator is needed. We have constructed such an estimator for simulated experiments with artificial and real data in Section \ref{real_data}. As for perspectives of this work, we are interested in the theoretical performance of the Nadaraya-Watson regressor (\cite{nadaraya1964estimating, wasserman2006all}). Apparently, it should be close to the one of the k-NN regression estimator since these two non-parametric estimators are closely related (see the monographs \cite{gyorfi2006distribution, tsybakov2008introduction}). The main difficulty should come from the fact that, if $h$ is the bandwidth parameter and $A_h$ is the smoothing matrix of the Nadaraya-Watson estimator, then $\textnormal{tr}(A_h^{\top}A_h) \neq \textnormal{tr}(A_h)$. This fact implies that the expectation of the empirical risk minus the noise variance will not be equal to the difference between the bias and variance terms (see Eq. (\ref{exp_empirical_risk})). Therefore, there should be another concentration result that deals with this problem. Besides that, we should emphasize that the early stopping rules in this work were estimating the famous bias-variance trade-off (\cite[Chapter 7]{hastie2009elements}). However, recently, (\cite{belkin2019reconciling, belkin2020two}) the bias-variance balancing paradigm was rethought by discovering some settings (exact fit to the data) for which a phenomenon of the "double descent" of the risk curve appeared. It would be interesting to understand if early stopping can work for these settings. The reader can look at a recent paper (\cite{dwivedi2020revisiting}) and references therein for another reexamination of the paradigm. \printbibliography \newpage
{ "timestamp": "2021-05-06T02:15:58", "yymm": "2008", "arxiv_id": "2008.08718", "language": "en", "url": "https://arxiv.org/abs/2008.08718", "abstract": "We present a novel data-driven strategy to choose the hyperparameter $k$ in the $k$-NN regression estimator. We treat the problem of choosing the hyperparameter as an iterative procedure (over $k$) and propose using an easily implemented in practice strategy based on the idea of early stopping and the minimum discrepancy principle. This model selection strategy is proven to be minimax-optimal, under the fixed-design assumption on covariates, over some smoothness function classes, for instance, the Lipschitz functions class on a bounded domain. The novel method often improves statistical performance on artificial and real-world data sets in comparison to other model selection strategies, such as the Hold-out method and 5-fold cross-validation. The novelty of the strategy comes from reducing the computational time of the model selection procedure while preserving the statistical (minimax) optimality of the resulting estimator. More precisely, given a sample of size $n$, assuming that the nearest neighbors are already precomputed, if one should choose $k$ among $\\left\\{ 1, \\ldots, n \\right\\}$, the strategy reduces the computational time of the generalized cross-validation or Akaike's AIC criteria from $\\mathcal{O}\\left( n^3 \\right)$ to $\\mathcal{O}\\left( n^2 (n - k) \\right)$, where $k$ is the proposed (minimum discrepancy principle) value of the nearest neighbors. Code for the simulations is provided atthis https URL.", "subjects": "Machine Learning (stat.ML); Machine Learning (cs.LG); Statistics Theory (math.ST)", "title": "Minimum discrepancy principle strategy for choosing $k$ in $k$-NN regression", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9838471613889295, "lm_q2_score": 0.831143045767024, "lm_q1q2_score": 0.8177177262860357 }
https://arxiv.org/abs/1810.09250
Optimal terminal dimensionality reduction in Euclidean space
Let $\varepsilon\in(0,1)$ and $X\subset\mathbb R^d$ be arbitrary with $|X|$ having size $n>1$. The Johnson-Lindenstrauss lemma states there exists $f:X\rightarrow\mathbb R^m$ with $m = O(\varepsilon^{-2}\log n)$ such that $$ \forall x\in X\ \forall y\in X, \|x-y\|_2 \le \|f(x)-f(y)\|_2 \le (1+\varepsilon)\|x-y\|_2 . $$ We show that a strictly stronger version of this statement holds, answering one of the main open questions of [MMMR18]: "$\forall y\in X$" in the above statement may be replaced with "$\forall y\in\mathbb R^d$", so that $f$ not only preserves distances within $X$, but also distances to $X$ from the rest of space. Previously this stronger version was only known with the worse bound $m = O(\varepsilon^{-4}\log n)$. Our proof is via a tighter analysis of (a specific instantiation of) the embedding recipe of [MMMR18].
\section{Introduction}\label{sec:intro} {\it Metric embeddings} may play a role in algorithm design when input data is geometric, in which the technique is applied as a pre-processing step to map input data living in some metric space $(\mathcal X, d_{\mathcal X})$ into some target space $(\mathcal Y, d_{\mathcal Y})$ that is algorithmically friendlier. One common approach is that of {\it dimensionality reduction}, in which $\mathcal X$ and $\mathcal Y$ are both subspaces of the same normed space, but where $\mathcal Y$ is of much lower dimension. Working with lower-dimensional embedded data then typically results in efficiency gains, in terms of memory, running time, and/or other resources. A cornerstone result in this area is the Johnson-Lindenstrauss (JL) lemma \cite{JohnsonL84}, which provides dimensionality reduction for Euclidean space. \begin{lemma}{\cite{JohnsonL84}} Let $\varepsilon\in(0,1)$ and $X\subset\mathbb{R}^d$ be arbitrary with $|X|$ having size $n>1$. There exists $f:X\rightarrow\mathbb{R}^m$ with $m = O(\varepsilon^{-2}\log n)$ such that \begin{equation} \forall x,y\in X, \|x-y\|_2 \le \|f(x)-f(y)\|_2 \le (1+\varepsilon)\|x-y\|_2 . \label{eqn:jl2} \end{equation} \end{lemma} It has been recently shown that the dimension $m$ of the target Euclidean space achieved by the JL lemma is best possible, at least for $\varepsilon \gg 1/\sqrt{\min(n,d)}$ \cite{LarsenN17} (see also \cite{AlonK17}). The multiplicative factor on the right hand side of Eqn.~\eqref{eqn:jl2}, in this case $1+\varepsilon$, is referred to as the {\em distortion} of the embedding $f$. Recent work of Elkin et al.\ \cite{ElkinFN17} showed a stronger form of Euclidean dimensionality reduction for the case of constant distortion. Namely, they showed that in Eqn.~\eqref{eqn:jl2}, whereas $x$ is taken as an arbitrary point in $X$, $y$ may be taken as an arbitrary point in $\mathbb{R}^d$. They called such an embedding a {\it terminal embedding}\footnote{More generally, a terminal embedding from $(\mathcal X, d_{\mathcal X})$ into $(\mathcal Y, d_{\mathcal Y})$ with terminal set $K\subset \mathcal X$ and {\it terminal distortion} $\alpha$ is a map $f:\mathcal X\rightarrow\mathcal Y$ s.t.\ $\exists\ c>0$ satisfying $d_{\mathcal X}(u,w) \le c\cdot d_{\mathcal Y}(f(u), f(w)) \le \alpha\cdot d_{\mathcal X}(u, w)$ for all $u\in K, w \in \mathcal X$.}. Though rather than achieving terminal distortion $1+\varepsilon$, their work only showed how to achieve constant terminal distortion with $m = O(\log n)$ for a constant that could be made arbitrarily close to $\sqrt{10}$ (see \cite[Theorem 1]{ElkinFN17}). Terminal embeddings can be useful in static high-dimensional computational geometry data structural problems. For example, consider nearest neighbor search over some finite database $X\subset\mathbb{R}^d$. If one builds a data structure over $f(X)$ for a terminal embedding $f$, then {\it any} future query is guaranteed to be handled correctly (or, at least, the embedding will not be the source of failure). Contrast this with the typical approach where one uses a randomized embedding oblivious to the input (e.g.\ random projections) that preserves the distance between any fixed pair of vectors with probability $1-1/\mathop{poly}(n)$. One can verify that the embedding preserves distances {\it within} $X$ during pre-processing, but for any later query there is some non-zero probability that the embedding will fail to preserve the distance between the query point $q$ and some points in $X$. Subsequent to \cite{ElkinFN17}, work of Mahabadi et al.\ \cite{MahabadiMMR18} gave a construction for terminal dimensionality reduction in Euclidean space achieving terminal distortion $1+\varepsilon$, with $m = O(\varepsilon^{-4}\log n)$. They asked as one of their main open questions (see \cite[Open Problem 3]{MahabadiMMR18}) whether it is possible to achieve this terminal embedding guarantee with $m = O(\varepsilon^{-2}\log n)$, which would be optimal given the JL lower bound of \cite{LarsenN17}. Our contribution in this work is to resolve this question affirmatively; the following is our main theorem. \begin{theorem} Let $\varepsilon\in(0,1)$ and $X\subset\mathbb{R}^d$ be arbitrary with $|X|$ having size $n>1$. There exists $f:X\rightarrow\mathbb{R}^m$ with $m = O(\varepsilon^{-2}\log n)$ such that \begin{equation*} \forall x\in X,\forall y\in \mathbb{R}^d,\ \|x-y\|_2 \le \|f(x)-f(y)\|_2 \le (1+\varepsilon)\|x-y\|_2 . \end{equation*} \end{theorem} The embedding we analyze in this work is in fact one that fits within the family introduced in \cite{MahabadiMMR18}; our contribution is to provide a sharper analysis. We note that unlike in \cite{MahabadiMMR18}, which provided a deterministic polynomial time construction of the terminal embedding, we here only provide a Monte Carlo polynomial time construction algorithm. Our error probability though only comes from mapping the points in $X$. If the points in $X$ are mapped to $\mathbb{R}^m$ well (with low ``convex hull distortion'', which we define later), which happens with high probability, then our final terminal embedding is guaranteed to have low terminal distortion as map from all of $\mathbb{R}^d$ to $\mathbb{R}^m$. \bigskip \begin{remark}\label{rem:nonlinear} Unlike in the JL lemma for which the embedding $f$ may be linear, the terminal embedding we analyze here (as was also in the case in \cite{ElkinFN17,MahabadiMMR18}) is nonlinear, as it must be. To see that it must be nonlinear, consider that any linear embedding with constant terminal distortion, $x\mapsto \Pi x$ for some matrix $\Pi\in\mathbb{R}^{m\times d}$, must have $\Pi (x-y) \neq 0$ for any $x\in X$ and any $y\in\mathbb{R}^d$ on the unit sphere centered at $x$. In other words, one needs $\mathop{ker}(\Pi)\neq \emptyset$, which is impossible unless $m\ge d$. \end{remark} \subsection{Overview of approach}\label{sec:approach} We outline both the approach of previous works as well as our own. In all these approaches, one starts with an embedding $f:X\rightarrow\ell_2^m$ with good distortion, then defines an {\it outer extension} $f_\mathrm{Ext}$ as introduced in \cite{ElkinFN17} and defined explicitly in \cite{MahabadiMMR18}. \begin{defn} For $f:X\rightarrow \mathbb{R}^m$ and $Z\supsetneq X$, we say $g:Z\rightarrow\mathbb{R}^{m'}$ is an {\it outer extension} of $f$ if $m'\ge m$, and $g(x)$ for $x\in X$ has its first $m$ entries equal to $f(x)$ and last $m'-m$ entries all $0$. \end{defn} In \cite{ElkinFN17,MahabadiMMR18} and the current work, a terminal embedding is obtained by, for each $u\in \mathbb{R}^d\backslash X$, defining an outer extension $f^{(u)}_\mathrm{Ext}$ for $Z = X\cup \{u\}$ with $m' = m+1$. Since $f^{(u)}_\mathrm{Ext}$ and $f^{(u')}_\mathrm{Ext}$ act identically on $X$ for any $u,u'\in \mathbb{R}^d$, we can then define our final terminal embedding by $\tilde f(u) = (f(u), 0)$ for $u\in X$ and $\tilde f(u) = f^{(u)}_\mathrm{Ext}(u)$ for $u\in \mathbb{R}^d\backslash X$, which will have terminal distortion $\sup_{u\in \mathbb{R}^d\backslash X} \mathsf{Dist}(f^{(u)}_\mathrm{Ext})$, where $\mathsf{Dist}(g)$ denotes the distortion of $g$. The main task is thus creating an outer extension with low distortion for a set $Z = X \cup \{u\}$ for some $u$. In all the cases that follow, we will have $f_\mathrm{Ext}(x) = (f(x),0)$ for $x\in X$, and we will then specify how to embed points $u\notin X$. The construction of Elkin et al.\ \cite{ElkinFN17}, the ``EFN extension'', is as follows. Suppose $f:X\rightarrow\ell_p^m$ is an $\alpha$-distortion embedding. The EFN extension $f_\mathrm{EFN}:X\cup\{u\}\rightarrow\ell_p^{m+1}$ is then defined by $f_\mathrm{EFN}(u) = (f(\rho_{\ell_p}(u)), d(\rho_{\ell_p}(u), u))$, where $\rho_d(u) = \mathop{argmin}_{x\in X} d(x, u)$ for some metric $d$. In the case that we view $X\cup\{u\}$ as living in the metric space $\ell_p^d$, Elkin et al.\ showed that the EFN extension has terminal distortion at most $2^{(p-1)/p} \cdot ((2\alpha)^p + 1)^{1/p}$ \cite[Theorem 1]{ElkinFN17}. If $p=2$ and $f$ is obtained via the JL lemma to have distortion $\alpha\le 1+\varepsilon$, this implies the EFN extension would have terminal distortion at most $\sqrt{10} + O(\varepsilon)$. The EFN extension does not in general achieve terminal distortion $1+\varepsilon$, even if starting with $f$ a perfect isometry. In fact, the bound of \cite{ElkinFN17} showing distortion at least $\sqrt{10}$ is sharp. Consider for example $X = \{-1,0,2\}\subset \mathbb{R}$. Consider the identity map $f(x) = x$, which has distortion $1$ as a map from $(X, \ell_2)$ to $(\mathbb{R}, \ell_2)$. Then $f_\mathrm{EFN}(-1) = (-1,0)$, $f_\mathrm{EFN}(0) = (0,0)$, $f_\mathrm{EFN}(2) = (2,0)$, and $f_\mathrm{EFN}(1) = (0, 1)$, and thus $\|f_\mathrm{EFN}(1) - f_\mathrm{EFN}(-1)\|_2 = \sqrt 2$ (distance shrunk by a $\sqrt 2$ factor) and $\|f_\mathrm{EFN}(1) - f_\mathrm{EFN}(2)\|_2 = \sqrt 5$ (distance increased by a $\sqrt 5$ factor). Thus $f_\mathrm{EFN}$ has terminal distortion at least (in fact exactly equal to) $\sqrt{10}$. This example in fact shows sharpness for $\ell_p$ for all $p\ge 1$. Thus to achieve terminal distortion $1+\varepsilon$, the work of \cite{MahabadiMMR18} had to develop a new outer extension, the ``MMMR extension'', which they based on the following core lemma. \begin{lemma}[{\cite[Lemma 3.1 (rephrased)]{MahabadiMMR18}}]\label{lem:mmmr} Let $X$ be a finite subset of $\ell_2^d$, and suppose $f:X\rightarrow\ell_2^m$ has distortion $1+\gamma$. Fix some $u\in \mathbb{R}^d$, and define $x_0 := \rho_{\ell_2}(u)$. Then $\exists u'\in\mathbb{R}^m$ s.t. \begin{itemize} \item $\|u' - f(x_0)\|_2 \le \|u - x_0\|_2$, and \item $\forall x\in X$, $|\langle u' - f(x_0), f(x) - f(x_0)\rangle - \langle u - x_0, x - x_0\rangle| \le 3\sqrt{\gamma}(\|x - x_0\|_2^2 + \|u - x_0\|_2^2)$ \end{itemize} \end{lemma} Mahabadi et al.\ then used the $u'$ promised by Lemma~\ref{lem:mmmr} as part of a construction that takes a $(1+\gamma)$-distortion embedding $f:X\rightarrow \ell_2^m$ and uses it in a black box way to construct an outer extension $f_\mathrm{MMMR}:X\cup\{u\}\rightarrow\ell_2^{m+1}$. In particular, they define $f_\mathrm{MMMR}(u) = (u', \sqrt{\|u - x_0\|_2^2 - \|u' - f(x_0)\|_2^2})$. It is clear this map perfectly preserves the distance from $u$ to $x_0$; in \cite[Theorem 1.5]{MahabadiMMR18}, it is furthermore shown that $f_\mathrm{MMMR}$ preserves distances from $u$ to all of $X$ up to a $1+O(\sqrt{\gamma})$ factor. Thus one should set $\gamma = \Theta(\varepsilon^2)$ so that $f_\mathrm{MMMR}$ has distortion $1+\varepsilon$, which is achieved by starting with an $f$ guaranteed by the JL lemma with $m = \Theta(\varepsilon^{-4}\log n)$. They then showed that this loss is {\it tight}, in the sense that there exist $X$, $u$, $f:X\rightarrow\mathbb{R}^m$, where $f$ has distortion $1+\gamma$, such that {\it any} outer extension $f_\mathrm{Ext}$ to domain $X\cup\{u\}$ has distortion $1+\Omega(\sqrt{\gamma})$ \cite[Section 3.2]{MahabadiMMR18}. Thus, seemingly a new approach is needed to achieve $m = O(\varepsilon^{-2}\log n)$. One may be discouraged by the above-mentioned tightness of the $\gamma\rightarrow\Omega(\sqrt{\gamma})$ loss, but in this work we show that, in fact, the MMMR extension can be made to provide $1+\varepsilon$ distortion with the optimal $m = O(\varepsilon^{-2}\log n)$! The tightness result mentioned in the last paragraph is only an obstacle to using the low-distortion property of $f$ {\it in a black box way}, as it is only shown that there {\it exist} $f$ where the $\gamma\rightarrow\Omega(\sqrt{\gamma})$ loss is necessary. However, the $f$ we are using is not an arbitrary $f$, but rather is the $f$ obtained via the JL lemma. A standard way of proving the JL lemma is to choose $\Pi\in\mathbb{R}^{m\times d}$ with i.i.d.\ subgaussian entries, scaled by $1/\sqrt m$ for $m = \Theta(\varepsilon^{-2}\log n)$ \cite[Exercise 5.3.3]{Vershynin18}. The low-distortion embedding is then $f(x) = \Pi x$. We show in this work that by not just using that this $f$ is a low-distortion embedding for $X$, but rather that it satisfies a stronger property we dub {\em convex hull distortion} (which we show $x\mapsto\Pi x$ does satisfy with high probability), one can achieve the desired terminal embedding result with optimal $m$. \begin{defn}\label{def:chull} For $T\subset S^{d-1}$ a subset of the unit sphere in $\mathbb{R}^d$, and $\varepsilon\in(0,1)$, we say for $\Pi\in\mathbb{R}^{m\times d}$ that $\Pi$ {\it provides $\varepsilon$-convex hull distortion for $T$} if $$ \forall x\in \mathrm{conv}(T),\ |\|\Pi x\|_2 - \|x\|_2 | < \varepsilon $$ where $\mathrm{conv}(T) := \{\sum_i \lambda_i t_i : \forall i\ t_i\in T, \lambda_i \ge 0, \sum_i \lambda_i = 1\}$ denotes the convex hull of $T$. \end{defn} We show that a random $\Pi$ with subgaussian entries provides $\varepsilon$-convex hull distortion for $T$ with probability at least $1-\delta$ as long as $m = \Omega(\varepsilon^{-2}\log(|T|/(\varepsilon\delta)))$. We then replace Lemma~\ref{lem:mmmr} with a new lemma that shows that as long as $f(x) = \Pi x$ does not just have $1+\gamma$ distortion for $X$, but rather provides $\gamma$-convex hull distortion for $T = \{(x-y) / \|x - y\|_2 : x,y\in X\}$, then the $3\sqrt{\gamma}$ term on the RHS of the second bullet of Lemma~\ref{lem:mmmr} can be replaced with $O(\gamma)$ (plus one other technical improvement; see Lemma~\ref{ReductionLemma1}). Note $|T| = \binom n2$, so $\log|T| = O(\log n)$. We then show how to use the modified lemma to achieve an outer extension with $m = O(\varepsilon^{-2}\log(n/\varepsilon)) = O(\varepsilon^{-2}\log n)$. This last equality holds since we may assume $\varepsilon = \Omega(1/\sqrt n)$, since otherwise there is a trivial terminal embedding with $m = n = O(\varepsilon^{-2})$ with no distortion: if $d \le n$, take the identity map. Else, translate $X$ so $0\in X$; then $E:=\mathop{span}(X)$ has $\mathop{dim}(E) \le n-1$. By rotation, we can assume $E=\mathop{span}\{e_1,\ldots,e_{n-1}\}$ so that every $x\in E$ can be written as $\sum_{i=1}^{n-1} \alpha(x)_i e_i$ for some vector $\alpha(x)\in\mathbb{R}^{n-1}$. We can then define a terminal embedding $\tilde{f}:\mathbb{R}^d\rightarrow\mathbb{R}^n$ with $\tilde f(x) = (\alpha(\mathop{proj}_E(x)), \|\mathop{proj}_{E^\perp}(x)\|_2)$ for all $x\in\mathbb{R}^d$. Here $\mathop{proj}_E$ denotes orthogonal projection onto $E$. \section{Preliminaries} For our optimal terminal embedding analysis, we rely on two previous results. The first result is the von Neumann Minimax theorem \cite{vonNeumann28}, which was also used in the terminal embedding analysis in \cite{MahabadiMMR18}. The theorem states the following: \begin{theorem} \label{Minimax} Let $X \subset \mathbb{R}^n$ and $Y \subset \mathbb{R}^m$ be compact convex sets. Suppose that $f: X \times Y \to \mathbb{R}$ is a continuous function that satisfies the following properties: \begin{enumerate} \item $f(\cdot, y): X \to \mathbb{R}$ is convex for any fixed $y \in Y$, \item $f(x, \cdot): Y \to \mathbb{R}$ is concave for any fixed $x \in X.$ \end{enumerate} Then, we have that \[\min\limits_{x \in X} \max\limits_{y \in Y} f(x, y) = \max\limits_{y \in Y} \min\limits_{x \in X} f(x, y).\] \end{theorem} The second result is a result of Dirksen \cite{Dirksen15, Dirksen16} that provides a uniform tail bound on subgaussian empirical processes. To explain the result, we first make the following definitions: \begin{defn} A semi-metric $d$ on a space $X$ is a function $X \times X \to \mathbb{R}_{\ge 0}$ such that $d(x, x) = 0,$ $d(x, y) = d(y, x),$ and $d(x, y)+d(y, z) \ge d(x, z)$ for all $x, y, z \in X$. \end{defn} Note a semi-metric may have $d(x,y) = 0$ for $x\neq y$. \begin{defn} Given a semi-metric $d$ on $T$, and subset $S \subset T,$ define $d(t, S) = \inf_{s \in S} d(t, s)$ for any point $t \in T$. \end{defn} \begin{defn} Given a semi-metric $d$ on $T$, define \[\gamma_2(T, d) = \inf\limits_{\{S_r\}_{r=0}^\infty} \sup\limits_{t \in T} \sum\limits_{r \ge 0} 2^{r/2} d(t, S_r),\] where the first infimum runs over all subsequences $S_0 \subset S_1 \subset \ldots \subset T$ where $|S_0| = 1$ and $|S_r| \le 2^{2^r}$. \end{defn} \begin{defn} For any random variable $X$, we define the \textit{subgaussian norm} of $X$ as \[\|X\|_{\psi_2} = \inf \left\{C \ge 0: \mathbb{E}\left(e^{X^2/C^2}\right) \le 2\right\}.\] \end{defn} It is well known that the subgaussian norm as defined above indeed defines a norm \cite{BoucheronLM13}. We also note the following proposition: \begin{proposition} \label{IIDSubgaussianAddition} \cite{BoucheronLM13} There exists some constant $k$ such that if $X_1, ..., X_n$ are i.i.d. subgaussians with mean $0$ and subgaussian norm $C$, then for any $a_1, ..., a_n \in \mathbb{R},$ $a_1X_1+...+a_nX_n$ is subgaussian with subgaussian norm $\le k \|a\|_2 C,$ where $\|a\|_2 = \sqrt{a_1^2+...+a_n^2}.$ \end{proposition} \begin{theorem} \label{DirksenInequality} \cite[Theorem 3.2]{Dirksen16} Let $T$ be a set, and suppose that for every $t \in T$ and every $1 \le i \le m,$ $X_{t, i}$ is a random variable with finite expected value and variance. For each $t \in T,$ let \[A_t = \frac{1}{m} \sum\limits_{i = 1}^{m} \left(X_{t, i}^2 - \mathbb{E} X_{t, i}^2\right).\] Consider the following semi-metric on $T$: for $s, t \in T$, define \[d_{\psi_2}(s, t) := \max\limits_{1 \le i \le m} \|X_{s, i} - X_{t, i}\|_{\psi_2}\] and define \[\overline{\Delta}_{\psi_2}(T) := \sup\limits_{t \in T} \max\limits_{1 \le i \le m} \|X_{t, i}\|_{\psi_2}.\] Then, there exist constants $c, C > 0$ such that for all $u \ge 1,$ \[\mathbb{P}\left(\sup\limits_{t \in T} |A_t| \ge C \left(\frac{1}{m} \gamma_2^2 (T, d_{\psi_2}) + \frac{1}{\sqrt{m}} \overline{\Delta}_{\psi_2}(T) \gamma_2(T, d_{\psi_2})\right) + c\left(\sqrt{u} \frac{\overline{\Delta}_{\psi_2}^2(T)}{\sqrt{m}} + u \frac{\overline{\Delta}_{\psi_2}^2(T)}{m}\right)\right) \le e^{-u}.\] \end{theorem} \section{Construction of the Terminal Embedding} \subsection{Universal Dimensionality Reduction with an additional $\ell_1$ condition} \label{L1JL} Here we show that for all $X = \{x_1, \dots, x_n\}\subset\mathbb{R}^d$ a set of unit norm vectors, there exists $\Pi\in\mathbb{R}^{m\times d}$ for $m = O(\varepsilon^{-2} \log n)$ providing $\varepsilon$-convex hull distortion for $X$, as defined in Definition~\ref{def:chull}. If $\varepsilon^{-2} > n,$ this construction follows by projecting onto a spanning subspace of $x_1, \dots, x_n$ and choosing an orthonormal basis. For $\varepsilon^{-2} < n$, our goal is to show that if $\Pi \in \mathbb{R}^{m \times d}$ is a normalized random matrix with i.i.d.\ subgaussian entries (normalized by $1/\sqrt m$), then $\Pi$ provides $\varepsilon$-convex hull distortion with high probability. Define $T = \mathrm{conv}(X)$. We apply Theorem~\ref{DirksenInequality} as follows. For some $m$ which we will choose later, let $\Pi^0$ be a matrix in $\mathbb{R}^{m \times d}$ with i.i.d.\ subgaussian entries with mean $0$, variance $1$, and some subgaussian norm $C_1.$ Let $\Pi$ be the scaled matrix, i.e. $\frac{1}{\sqrt{m}} \cdot \Pi^0.$ Let $\Pi_i$ denote the $i$th row of $\Pi^0$ and for any $t \in \mathbb{R}^d,$ let $X_{t, i} = \Pi_i t.$ Finally, let $T_k$ be the subset of $T$ of points with norm at most $2^{-k},$ i.e., $T_k = \{x \in T: \|x\|_2 \le 2^{-k}\}$. First, note that \[A_t := \frac{1}{m} \sum\limits_{i = 1}^{m} \left((\Pi_i t)^2 - \mathbb{E}(\Pi_i t)^2\right) = \frac{1}{m} \left(\| \Pi^0 t\|_2^2 - m \cdot \|t\|_2^2\right) = \|\Pi t\|_2^2 - \|t\|_2^2.\] For any $t \in T$ and $1 \le i \le m,$ define $X_{t, i} = \Pi_i t$. Then, $A_t$ corresponds to the definition in Theorem~\ref{DirksenInequality}. Note that for any $s, t \in T$ and for any $1 \le i \le m,$ $X_{s, i} - X_{t, i} = X_{s-t, i},$ which is a subgaussian with mean $0$, variance $\|s-t\|_2^2,$ and subgaussian norm at most $k \|s-t\|_2 C_1$ by Proposition~\ref{IIDSubgaussianAddition}. Therefore, if we define $d_{\psi_2}(s, t) = \max_{1 \le i \le m} \|X_{s, i} - X_{t, i}\|_{\psi_2},$ as in Theorem \ref{DirksenInequality}, $d_{\psi_2}(s, t) \le k \|s-t\|_2 C_1.$ As a result, we have the following: \begin{proposition} $\overline{\Delta}_{\psi_2}(T_k) = O(2^{-k})$. \end{proposition} \begin{proof} Since any point $t \in T_k$ has Euclidean norm at most $2^{-k}$, the conclusion is immediate. \end{proof} We also note the following: \begin{proposition}\label{prop:majorizing-measures} $\gamma_2(T_k, \ell_2) = O(\sqrt{\log n}),$ where $\ell_2$ represents the standard Euclidean distance. \end{proposition} \begin{proof} The Majorizing Measures theorem \cite{Talagrand14} gives $\gamma_2(T_k, \ell_2) = \Theta\left(\mathbb{E}_g (\sup_{x \in T_k} \langle g, x \rangle \right)),$ where $g$ is a $d$-dimensional vector of i.i.d.\ standard normal random variables. Also, $\mathbb{E}_g (\sup_{x \in T_k} \langle g, x \rangle) \le \mathbb{E}_g (\sup_{x \in T} \langle g, x \rangle)$ because $T_k \subset T$. However, since $T = \mathrm{conv}(X)$ and since $\langle g, x \rangle$ is a linear function of $x$, we have $\sup_{x \in T} \langle g, x \rangle = \sup_{x\in X} \langle g, x\rangle.$ Since the $x_i$'s are unit norm vectors, then each $g_i := \langle g, x_i\rangle$ is standard normal, which implies (even if they are dependent) that $\mathbb{E} \sup_i |g_i| = O(\sqrt{\log n}).$ This means that $\gamma_2(T_k, \ell_2) = O(\mathbb{E}_g (\sup_{x \in T} \langle g, x \rangle)) = O(\sqrt{\log n}),$ so the proof is complete. \end{proof} This allows us to bound $\gamma_2(T, d_{\psi_2})$ as follows. \begin{corollary} $\gamma_2(T_k, d_{\psi_2}) = O(\sqrt{\log n}).$ \end{corollary} \begin{proof} As $d_{\psi_2}(s, t) = O(\|s-t\|_2)$ for any points $s, t,$ the result follows from Proposition~\ref{prop:majorizing-measures}. \end{proof} Therefore, using $T = T_k$ in Theorem \ref{DirksenInequality} and varying $k$ gives us the following. \begin{theorem} \label{MainIdea} Suppose that $\varepsilon, \delta < 1$ and $m = \Theta\left(\frac{1}{\varepsilon^2} \log \frac{n \log (2/\varepsilon)}{\delta}\right).$ Then, there exists some constant $C_2$ such that for all $k \ge 0,$ \[\mathbb{P}\left(\sup\limits_{t \in T_k} \left|\| \Pi t\|_2^2 - \|t\|_2^2\right| \ge C_2\left(\varepsilon^2 + \varepsilon \cdot 2^{-k}\right)\right) \le \frac{\delta}{n \log (2/\varepsilon)}.\] Consequently, with probability at least $1 - \delta/n,$ we have that for all $t \in T$, \[\left|\|\Pi t\|_2 - \|t\|_2\right| = O(\varepsilon).\] \end{theorem} \begin{proof} The first part follows from Theorem \ref{DirksenInequality} and the previous propositions. Note that \[\sup\limits_{t \in T_k} \left|\|\Pi t\|_2^2 - \|t\|_2^2\right| = \sup\limits_{t \in T_k} |A_t|.\] Moreover, $\frac{1}{m} \gamma_2^2(T_k, d_{\psi_2}) = O(\varepsilon^2)$ and $\frac{1}{\sqrt{m}} \overline{\Delta}_{\psi_2}(T_k) \gamma_2(T_k, d_{\psi_2}) = O(\varepsilon \cdot 2^{-k}).$ If we let $u = \ln \frac{n \log (2/\varepsilon)}{\delta},$ then $\sqrt{u} \frac{\overline{\Delta}_{\psi_2}^2(T_k)}{\sqrt{m}} = O(\varepsilon \cdot 2^{-2k}),$ and $u \frac{\overline{\Delta}_{\psi_2}^2(T)}{m} = O(\varepsilon^2 \cdot 2^{-2k}).$ The first result now follows immediately from Theorem \ref{DirksenInequality}, as $e^{-u} = \frac{\delta}{n \log (2/\varepsilon)}.$ Note that it is possible for $T_k$ to be empty, but in this case we see that $\sup_{t \in T_k} |A_t| = 0$ so the first result is immediate. For the second part, assume WLOG that $\varepsilon^{-1} = 2^\ell$ for some $\ell.$ Define $T_0', T_1', \dots, T_{\ell}'$ such that $T_{\ell}' = T_{\ell}$ and for all $k < \ell,$ $T_k' = T_k \backslash T_{k+1}.$ Then, $T_0', \dots, T_{\ell}'$ forms a partition of $T$, since $\|x\|_2 \le 1$ if $x \in T$. Note that if $T \in T_{\ell},$ then with probability at least $1 - \frac{\delta}{n\log(2/\varepsilon)}$, $\left|\|\Pi t\|_2^2 - \|t\|_2^2\right| = O(\varepsilon^{2})$ and thus $\|\Pi t\|_2^2 = O(\varepsilon^2)$ and $\left|\|\Pi t\|_2 - \|t\|_2\right| = O(\varepsilon).$ If $T \in T_k'$ for some $k < \ell,$ then with probability at least $1 - \frac{\delta}{n\log(2/\varepsilon)}$, $\left|\|\Pi t\|_2^2 - \|t\|_2^2\right| = \left|\|\Pi t\|_2 + \|t\|_2\right| \cdot \left|\|\Pi t\|_2 - \|t\|_2\right| = O(\varepsilon \cdot 2^{-k})$. This means that since $\left|\|\Pi t\|_2 + \|t\|_2\right| \ge 2^{-(k+1)}$, we must have that $\left|\|\Pi t\|_2 - \|t\|_2\right| = O(\varepsilon).$ Therefore, by union bounding over all $0 \le k \le \ell$, with probability at least \[1 - (\ell+1) \frac{\delta}{n \log (2/\varepsilon)} = 1 - \frac{(\ell+1)\delta}{n (\ell+1)} = 1 - \frac{\delta}{n},\] for all $t \in T,$ $\left|\|\Pi t\|_2 - \|t\|_2\right| = O(\varepsilon).$ Thus, we are done. \end{proof} We therefore have the following immediate corollary. \begin{corollary} \label{MainL1Result} For $1 \le \varepsilon^{-2} < n$ and for any $X = \{x_1, ..., x_n\} \subset S^{d-1}$, with probability at least $1 - poly(n)^{-1}$, a randomly chosen $\Pi$ with $m = \Omega(\varepsilon^{-2} \log n)$ provides $\varepsilon$-convex hull distortion for $X$. \end{corollary} \iffalse Now, for any vector $x \in \mathbb{R}^d$ and some $m$ (we will attempt to perform dimensionality reduction to $m$ dimensions), consider $\Pi x,$ where $\Pi \in \mathbb{R}^{m \times d}$ where each entry of $\Pi$ is an i.i.d. $\mathcal{N}(0, 1/m)$ Normal. Then, one can verify that if $A_x \in \mathbb{R}^{m \times md}$ is the matrix \[\frac{1}{\sqrt{m}} \cdot \left[\begin{matrix} -x^T- & 0\cdots 0 & \cdots & 0\cdots 0\\ 0\cdots 0 & -x^T- & \cdots & 0\cdots 0\\ \vdots & \vdots & \ddots & \vdots \\ 0\cdots 0 & 0\cdots 0 & \cdots & -x^T-\\ \end{matrix}\right]\] and $g$ is an $md$-dimensional column vector of i.i.d. Gaussians, then $\Pi x = A_x g.$ We will use the following theorem: \begin{theorem} For any set $\mathcal{A}$ of matrices in $\mathbb{R}^{m \times md}$, and for $g \in \mathbb{R}^{md}$ a vector of i.i.d. Gaussians, \[\mathbb{E}_g \left[\sup\limits_{A \in \mathcal{A}} \left|\|Ag\|_2^2 - \mathbb{E}_g\|Ag\|_2^2\right|\right] \lesssim \gamma_2^2(\mathcal{A}, \|\cdot\|) + \gamma_2(\mathcal{A}, \|\cdot\|) \cdot d_F (\mathcal{A}) + d_F(\mathcal{A}) \cdot d_{\ell_2 \to \ell_2}(\mathcal{A}),\] where $\|\cdot\|$ represents the operator norm metric. \end{theorem} Define $T_k = T \cap \{x \in \mathbb{R}^d: 2^{-k} \ge \|x\|_2\}$ and $\mathcal{A}_k$ as the according set of $m \times md$ matrices. Our goal is to bound three things: $\gamma_2(\mathcal{A}_k, \|\cdot\|),$ $d_F(\mathcal{A}_k),$ and $d_{\ell_2 \to \ell_2}(\mathcal{A}_k)$. \begin{proposition} $d_{\ell_2 \to \ell_2}(\mathcal{A}_k) = O(2^{-k}/\sqrt{m})$ \end{proposition} \begin{proof} Note that $\|A_x-A_y\| = \|A_{x-y}\|,$ but one can check that $A_{x-y}A_{x-y}^T$ equals $\frac{1}{m} \cdot \|x-y\|_2^2 \cdot I_m,$ and thus $\|A_x-A_y\| = \frac{1}{\sqrt{m}} \|x-y\|_2.$ But $\sup_{x, y \in T_k} \|x-y\|_2 \le 2 \cdot \sup_{x \in T_k} \|x\|_2 \le 2 \cdot 2^{-k},$ so we are done. \end{proof} \begin{proposition} $d_F(\mathcal{A}_k) = O(2^{-k})$ \end{proposition} \begin{proof} Similar to the previous proposition, $\|A_x-A_y\|_F = \|A_{x-y}\|_F = \frac{1}{\sqrt{m}} \cdot \|x-y\|_2 \cdot \|I_m\|_F = \|x-y\|_2.$ But $\sup_{x, y \in T_k} \|x-y\|_2 \le 2 \cdot \sup_{x \in T_k} \|x\|_2 \le 2 \cdot 2^{-k}.$ \end{proof} Note that $\|A_x-A_y\| = \|A_{x-y}\|,$ but one can check that $A_{x-y}A_{x-y}^T$ equals $\frac{1}{m} \cdot \|x-y\|_2^2 \cdot I_m,$ and thus $\|A_x-A_y\| = \frac{1}{\sqrt{m}} \|x-y\|_2.$ Similarly, $\|A_x-A_y\|_F = \|A_{x-y}\|_F = \frac{1}{\sqrt{m}} \cdot \|x-y\|_2 \cdot \|I_m\|_F = \|x-y\|_2.$ \begin{proposition} $\gamma_2(\mathcal{A}_k, \|\cdot\|) \le \gamma_2(\mathcal{A}, \|\cdot\|) \lesssim \sqrt{\frac{\log n}{m}}.$ \end{proposition} \begin{proof} The first inequality is trivial since $\mathcal{A}_k \subset \mathcal{A}.$ For the second inequality, note that since $\|A_x-A_y\| = \frac{1}{\sqrt{m}} \|x-y\|_2,$ $\gamma_2(\mathcal{A}, \|\cdot\|) = \frac{1}{\sqrt{m}} \gamma_2(T, \ell_2).$ Thus, it suffices to show that $\gamma_2(T, \ell_2) \lesssim \sqrt{\log n}.$ However, the Majorizing Measures theorem tells us that $\gamma_2(T, \ell_2) = \Theta\left(\mathbb{E}_g (sup_{x \in T} \langle g, x \rangle \right)).$ However, since $\langle g, x \rangle$ is a linear function on $x$ and since $T = Conv(x_1, \dots, x_n),$ we have that $sup_{x \in T} \langle g, x \rangle = \sup_{x_i} \langle g, x_i\rangle.$ Since the $x_i$'s are unit normed vectors, if we define $g_i = \langle g, x_i\rangle,$ each $g_i$ is a $\mathcal{N}(0, 1)$ distribution, which implies (even if they are dependent) that $\mathbb{E} \sup_i g_i = O(\sqrt{\log n}).$ This means that $\gamma_2(T, \ell_2) = O(\sqrt{\log n}),$ so we are done. \end{proof} These three imply the following: \begin{lemma} Suppose that $\Pi$ is chosen from $\mathbb{R}^{m \times d}$ where $m = O(\varepsilon^{-2} \log^2 \varepsilon^{-1} \log n).$ Then, for any $A_{k}$ where $k \le \log \frac{1}{\varepsilon},$ \[\mathbb{E}_{\Pi}\left[\sup\limits_{x \in T_k} \left|\|\Pi x\|_2^2 - \|x\|_2^2\right|\right] \lesssim \frac{\varepsilon}{2^k \log \varepsilon^{-1}}.\] \end{lemma} \begin{proof} Recall that $\Pi x = A_x g$ and thus $\|x\|_2^2 = \mathbb{E} \|\Pi x\|_2^2 = \mathbb{E} \|A_x g\|_2^2.$ Therefore, this equals \[\mathbb{E}_g\left[\sup\limits_{A \in \mathcal{A}_k} \left|\|A g\|_2^2 - \mathbb{E}_g \|A g\|_2^2\right|\right]\lesssim \gamma_2^2(\mathcal{A}, \|\cdot\|) + \gamma_2(\mathcal{A}, \|\cdot\|) \cdot d_F (\mathcal{A}) + d_F(\mathcal{A}) \cdot d_{\ell_2 \to \ell_2}(\mathcal{A})\] \[\lesssim \frac{\log n}{m} + \sqrt{\frac{\log n}{m}}\frac{1}{2^k} + \frac{1}{2^{2k}} \cdot \frac{1}{\sqrt{m}} = \frac{\varepsilon^2}{\log^2 \varepsilon^{-1}} + \frac{1}{2^k} \cdot \frac{\varepsilon}{\log \varepsilon^{-1}} + \frac{\varepsilon}{2^k \log \varepsilon^{-1} \sqrt{\log m}},\] and since $\varepsilon \le 2^{-k},$ this is at most \[3 \cdot \frac{1}{2^k} \cdot \frac{\varepsilon}{\log \varepsilon^{-1}},\] as desired. \end{proof} Therefore, by Markov's inequality, with probability at least $2/3$ over a randomly selected $\Pi$ if we pick $m$ appropriately, every $k \le \log \frac{1}{\varepsilon}$ satisfies \[\sup\limits_{x \in T_k} \left| \|\Pi x\|_2^2 - \|x\|_2^2 \right| \le \frac{\varepsilon}{2^k}.\] Therefore, for every $\|x\|_2 \le \varepsilon,$ we have that $\left|\|\Pi x\|_2^2 - \|x\|_2^2\right| \le \varepsilon^2,$ and thus $\left|\|\Pi x\|_2 - \|x\|_2\right| \le \varepsilon.$ For any $x$ between $2^{-k}$ and $2^{-(k-1)}$ for $1 \le k \le \log \frac{1}{\varepsilon},$ we know that \[\frac{\varepsilon}{2^{k-1}} \ge \left|\|\Pi x\|_2^2 - \|x\|_2^2\right| = \left|\|\Pi x\|_2 - \|x\|_2\right| \cdot \left(\|\Pi x\|_2^2 + \|x\|_2^2\right) \ge \left|\|\Pi x\|_2 - \|x\|_2\right| \cdot 2^{-k},\] and thus $\left|\|\Pi x\|_2 - \|x\|_2\right| \le 2\varepsilon.$ Therefore, this implies the following: \begin{theorem} Fix $x_1, \dots, x_n \in \mathbb{R}^d$ with unit norm and let $m = O(\varepsilon^{-2} \log^2 \varepsilon^{-1} \log n).$ Then, with probability at least $2/3$, if we pick a random Gaussian sketch matrix $\Pi \in \mathbb{R}^{m \times d},$ for all $x$ that can be written as $\lambda_1 x_1 + \cdots + \lambda_n x_n$ where $\|\lambda\|_1 = 1,$ \[\left|\|\Pi x\|_2 - \|x\|_2\right| \le \varepsilon.\] Therefore, we have that for any $\lambda = (\lambda_1, \dots, \lambda_n)$ by linearity, \[\left|\|\Pi x\|_2 - \|x\|_2\right| \le \varepsilon \|\lambda\|_1.\] \end{theorem} \fi \iffalse Ok so here I am going to show that the following two problems are equivalent for a fixed $d$. \begin{enumerate} \item For any $n$-point set $X = \{x_1, \dots, x_n\} \in \mathbb{R}^d,$ there exists a terminal embedding $f$ of distortion $\delta$ from $X$ to $\mathbb{R}^{O(\log n/\varepsilon^2)}$ with distortion $\delta.$ \item For any $n$-point set $X = \{v_1, \dots, v_n\}$ where each $\|v_i\| = 1,$ there exists a map $g$ from $X$ to $\mathbb{R}^{O(\log n/\varepsilon^2)}$ with distortion at most $\delta,$ such that for all $\lambda \in \mathbb{R}^n$ with $\|\lambda_1\| = 1,$ \[\left|\left|\sum \lambda_i v_i\right|\right| - \left|\left|\sum \lambda_i g(v_i)\right|\right| \le \delta.\] \end{enumerate} \begin{proof} We first show 2 implies 1. To prove this, we first show that for any $u$ and any $X = \{x_1, \dots, x_n\}$, there exists some $u' \in \mathbb{R}^m$ such that $\|u'\| \le \|u\|$ and $\left|\langle u', f(x_i)\rangle - \langle u, x_i \rangle\right| \le \delta \cdot \|u\| \cdot \|x_i\|$ for all $x_i$. To see why, let $v_i = \frac{x_i}{\|x_i\|}$, and let $f(x_i) = \|x_i\| g(v_i).$ Then, it suffices to show that $|\langle u', g(v_i)\rangle - \langle u, v_i\rangle| \le \delta$ for all $v_i.$ Now, let $B$ be the ball in $\mathbb{R}^{O(\log n/\varepsilon^2)}$ of radius $\|u\|$ and define for $u' \in B, \lambda \in \Lambda,$ \[\Phi(u', \lambda) := \sum\limits_{i = 1}^{n} \left(\lambda_i(\langle u, v_i\rangle - \langle u', g(v_i)\rangle) - \delta |\lambda_i| \cdot \|u\|\right).\] It suffices to show that $\min_{u' \in B} \max_{\lambda \in \Lambda} \Phi(u', \lambda) \le 0.$ This is because then there is some $u' \in B$ such that if $\lambda_i = \pm 1$ for some $i$ and $\lambda_j = 0$ for all $j \neq i,$ then $|\langle u, v_i\rangle - \langle u', g(v_i)\rangle| = O(\delta)$. Now, note that $\Lambda, B$ are compact and convex, $\Phi$ is linear in $u',$ and concave in $\lambda.$ Therefore, the von Neumann minimax theorem tells us that $\min_{u' \in B} \max_{\lambda \in \Lambda} \Phi(u', \lambda) = \max_{\lambda \in \Lambda} \min_{u' \in B} \Phi(u', \lambda).$ Therefore, it suffices to show that for all $\lambda,$ there exists some $u'$ such that \[\sum\limits_{i = 1}^{n} \left(\lambda_i(\langle u, v_i\rangle - \langle u', g(v_i)\rangle) - \delta |\lambda_i| \cdot \|u\|\right) \le 0.\] Letting $P = \sum \lambda_i v_i$ and $P' = \sum \lambda_i g(v_i),$ then if we let $u' = \|u\| \cdot \frac{P'}{\|P'\|}$ we get it suffices to show that \[\langle u, P \rangle - \|u\| \cdot \|P'\| - \delta \|u\| \cdot \|\lambda\|_1 \le 0\] for all $\|\lambda\|_1 \le 1.$ But $\langle u, P \rangle \le \|u\| \cdot \|P\|,$ so we just have to see that $\|P\|-\|P'\|-\delta \|\lambda\|_1 \le 0.$ This follows from condition 2. Ok the rest follows from the paper. Now, we prove the other direction. Suppose we have points $v_1, \dots, v_n$ of unit norm, as well as $v_0 = 0.$ We can assume WLOG that $f(v_0) = 0.$ Then, $v_1, \dots, v_n$ are sent to points such that $\|f(v_i)\|_2 = (1 \pm \delta)\|v_i\|_2$ and $\|f(v_i)-f(v_j)\|_2 = (1 \pm \delta) \|v_i-v_j\|_2.$ Moreover, for any $u$ with unit norm, there exists some $f(u)$ such that $\|f(u)-f(v_i)\|_2^2 = (1 \pm 3\delta)\|u-v_i\|_2^2$ and $\|f(u)\|_2^2 = (1 \pm 3\delta) \|u\|_2^2.$ Therefore, we have that \[\left|\langle f(u), f(v_i)\rangle - \langle u, v_i\rangle\right| \le 3\delta\left(\|u\|_2^2 + \|u-v_i\|_2^2 + \|v_i\|_2^2 \right) \le 18 \delta.\] Now, consider $P(\lambda) = \sum \lambda_i v_i$ and $P'(\lambda) = \sum \lambda_i f(v_i)$ for $\|\lambda\|_1 = 1$. Then, $\langle f(u), P(\lambda)\rangle - \langle u, P(\lambda)\rangle$ is a linear function of $\lambda$ and thus reaches its maximum and minimum at some endpoints of the simplex $\Lambda.$ Therefore, we have that $|\langle f(u), P'(\lambda) \rangle - \langle u, P(\lambda)\rangle| \le 18\delta$. Since this is a linear function of $\lambda,$ we actually have that \[\left|\sum \lambda_i \left(\langle u, v_i\rangle - \langle f(u), f(v_i)\rangle\right)\right| \lesssim 18\delta,\] and therefore, for any $u$, \[\max_{\lambda \in \Lambda} \Phi(\lambda, f(u)) = \max_{\lambda \in \Lambda} \sum \left(\lambda_i \left(\langle u, v_i\rangle - \langle f(u), f(v_i)\rangle\right) - 18\delta|\lambda_i|\right) \le 0.\] Thus, applying the minimax theorem again tells us that $\max_{\lambda \in \Lambda} \min_{f(u) \in B_{1+\delta}} \Phi(\lambda, f(u)) \le 0,$ which means for any $\|\lambda\|_1 = 1,$ there exists a $u' \in B_{1+\delta}$ such that \[\sum \left(\lambda_i \left(\langle u, v_i\rangle - \langle u', f(v_i)\rangle\right)\right) \le 18\delta.\] Therefore, letting $\sum \lambda_i v_i =: P$ and $\sum \lambda_i f(v_i) =: P'$, we have that for any $\lambda, u$ with $\|u\| = 1, \|\lambda\|_1 = 1,$ there exists $u' \in B_{1+\delta}$ such that $\langle u, P\rangle - \langle u', P'\rangle \le 18 \delta.$ Now, choose $u = \frac{P}{\|P\|}$ and this means that $\|P\| - (1+\delta)\|P'\| \le 18\delta$. But $P' = \sum \lambda_i v_i$ so $\|P'\| \le 1,$ so we also have $\|P\|-\|P'\| \le 19\delta.$ \end{proof} \fi \subsection{Completion of the Terminal Embedding} \label{TerminalJL} We note that the methods for completing the terminal embedding in this section are very similar to those in \cite{MahabadiMMR18}. Specifically, our proofs for Lemmas \ref{ReductionLemma1} and \ref{ReductionLemma2} are based on the proofs of \cite[Lemma 3.1]{MahabadiMMR18} and \cite[Theorem 1.5]{MahabadiMMR18}, respectively. \begin{lemma} \label{ReductionLemma1} Let $x_1, ..., x_n$ be nonzero points in $\mathbb{R}^d$ and let $v_i = \frac{x_i}{\|x_i\|_2}.$ Suppose that $\Pi$ provides $\varepsilon$-convex hull distortion for $V = \{v_1, -v_1, ..., v_n, -v_n\}.$ Then, for any $u \in \mathbb{R}^d$, there exists a $u' \in \mathbb{R}^m$ such that $\|u'\|_2 \le \|u\|_2$ and $|\langle u', \Pi x_i \rangle - \langle u, x_i\rangle| \le \varepsilon \|u\|_2 \cdot \|x_i\|_2$ for every $x_i.$ \end{lemma} \begin{proof} This statement is trivial for for $u = 0,$ so assume $u \neq 0.$ It suffices to show that there always exists a $u'$ such that $\|u'\|_2 \le \|u\|_2$ and $|\langle u', \Pi v_i\rangle - \langle u, v_i \rangle| \le \varepsilon \|u\|_2$ for all $v_i.$ We will have that $|\langle u', \Pi x_i \rangle - \langle u, x_i \rangle| \le \varepsilon \|u\|_2 \cdot \|x_i\|_2$ by scaling. Now, let $B$ be the ball in $\mathbb{R}^m$ of radius $\|u\|_2$ and $\Lambda$ be the the unit $\ell_1$-ball in $\mathbb{R}^n,$ where we write $\lambda \in \Lambda$ as $(\lambda_1, \dots, \lambda_n).$ Furthermore, define for $u' \in B, \lambda \in \Lambda,$ \[\Phi(u', \lambda) := \sum\limits_{i = 1}^{n} \left(\lambda_i (\langle u, v_i\rangle - \langle u', \Pi v_i\rangle) - \varepsilon |\lambda_i| \cdot \|u\|_2\right).\] We wish to show that there exists a $u' \in B$ such that for all $\lambda \in \Lambda,$ $\Phi(u', \lambda) \le 0.$ This clearly suffices by looking at $\lambda = \pm v_j$ for $1 \le j \le n.$ Note that $\Phi$ is linear in $u'$ and concave in $\lambda,$ and that $B, \Lambda$ are compact and convex. Then, by the von Neumann minimax theorem \cite{vonNeumann28}, \[\min\limits_{u' \in B} \max\limits_{\lambda \in \Lambda} \Phi(u', \lambda) = \max\limits_{\lambda \in \Lambda} \min\limits_{u' \in B} \Phi(u', \lambda).\] Thus it suffices to show the right hand side is nonpositive, i.e.\ for any $u, \lambda$, there exists $u' \in B$ s.t.\ $\Phi(u', \lambda) \le 0.$ Defining $P = \sum \lambda_i v_i$, then for $u' = \|u\|_2 \cdot \frac{\Pi P}{\|\Pi P\|_2},$ it suffices to show for all $u, \lambda$ \[\langle u, P\rangle - \langle u', \Pi P\rangle = \langle u, P\rangle - \|u\|_2 \cdot \|\Pi P\|_2 \le \varepsilon \|\lambda\|_1 \|u\|_2.\] But in fact $\langle u, P \rangle \le \|u\|_2 \cdot \|P\|_2,$ so having that $\|P\|_2 - \|\Pi P\|_2 \le \varepsilon \|\lambda\|_1$ for all $\lambda \in \Lambda$ is sufficient. For $\|\lambda\|_1 = 1$ this follows from $\Pi$ providing $\varepsilon$-convex hull distortion for $V$, and for $\|\lambda\|_1 < 1$, it follows because we can scale $\lambda$ so that $\|\lambda\|_1 = 1.$ \end{proof} \begin{lemma} \label{ReductionLemma2} Let $x_1, ..., x_n \in \mathbb{R}^d$ be distinct. Let $Y = \left\{\frac{x_i-x_j}{\|x_i-x_j\|_2}: i \neq j \right\}$. Moreover, suppose that $\Pi\in\mathbb{R}^{m\times d}$ provides $\varepsilon$-convex hull distortion for $Y$. Then, for any $u \in \mathbb{R}^d$, there exists an outer extension $f: \{x_1, \dots, x_n, u\} \to \mathbb{R}^{m+1}$ with distortion $1 + O(\varepsilon),$ where $f(x_i) = \Pi x_i$. \end{lemma} \begin{proof} Note that the map $x\mapsto \Pi x$ yields a $1 + \varepsilon$-distortion embedding of $x_1, \dots, x_n$ into $\mathbb{R}^m$ as it approximately preserves the norm of all points in $Y$. Therefore, we just have to verify that there exists $f(u) \in \mathbb{R}^{m+1}$ such that $\|f(u) - \Pi x_i\|_2 = (1 \pm O(\varepsilon)) \|u-x_i\|_2$ for all $i$ and any $u \in \mathbb{R}^d$. Fix $u \in \mathbb{R}^d,$ and let $x_k$ be the closest point to $u$ among $\{x_1, \dots, x_n\}.$ By Lemma \ref{ReductionLemma1}, there exists $u' \in \mathbb{R}^{m}$ such that $\|u'\|_2 \le \|u - x_k\|_2$ and for all $i$, \begin{equation*} |\langle u', \Pi (x_i-x_k)\rangle - \langle u - x_k, x_i-x_k\rangle| \le \varepsilon \|u-x_k\|_2 \|x_i-x_k\|_2. \end{equation*} Next, let $f(u) \in \mathbb{R}^{m+1}$ be the point $(\Pi x_k + u', \sqrt{\|u-x_k\|_2^2 - \|u'\|_2^2})$. If $w_i := x_i-x_k,$ then \begin{equation} \|f(u) - f(x_i)\|_2^2 = \|u-x_k\|_2^2 - \|u'\|_2^2 + \|u' - \Pi w_i\|_2^2 = \|u-x_k\|_2^2 + \|\Pi w_i\|_2^2 - 2\langle u', \Pi w_i\rangle \label{eqn:bash1} \end{equation} and \begin{equation} \|u-x_i\|_2^2 = \|u-x_k\|_2^2 + \|w_i\|_2^2 - 2\langle u-x_k, w_i\rangle. \label{eqn:bash2} \end{equation} Since $\|u-x_i\|_2^2 \ge \|u-x_k\|_2^2$ and $\|u-x_i\|_2^2 \ge (\|w_i\|_2 - \|u-x_k\|_2)^2,$ we have that $\|u-x_i\|_2^2 \ge (\|u-x_k\|_2^2 + \|w_i\|_2^2)/5.$ This follows from the fact that $\max(1, (x-1)^2) \ge (x^2+1)/5$ for all $x \ge 0.$ Since $\|\Pi(x_i-x_k)\|_2 = (1 \pm \varepsilon) \|x_i-x_k\|_2$, we also have that \begin{equation*} \left|\|\Pi w_i\|_2^2 - \|w_i\|_2^2\right| = \left|\|\Pi(x_i-x_k)\|_2^2 - \|x_i-x_k\|_2^2\right| \le 3 \varepsilon \|x_i-x_k\|_2^2 = 3 \varepsilon \|w_i\|_2^2 \end{equation*} for all $i, j,$ assuming $\varepsilon \le 1.$ Therefore, by subtracting Eq.~\eqref{eqn:bash2} from Eq.\eqref{eqn:bash1}, we have that \begin{equation*} \left|\|f(u) - f(x_i)\|_2^2 - \|u-x_i\|_2^2\right| \le 3 \varepsilon \|w_i\|_2^2 + 2\left|\langle u', \Pi w_i\rangle - \langle u-x_k, w_i\rangle\right| \end{equation*} \begin{equation*} \le 3\varepsilon \|w_i\|_2^2 + 2 \varepsilon \|u-x_k\|_2 \|w_i\|_2 \le 4\varepsilon\left(\|w_i\|_2^2 + \|u-x_k\|_2^2\right) \le 20 \varepsilon \|u-x_i\|_2^2, \end{equation*} as desired. \end{proof} We summarize the previous results, which allows us to prove our main result. \begin{theorem} \label{MainThm} For all $\varepsilon < 1$ and $x_1, \dots, x_n \in \mathbb{R}^d,$ there exists $m = O(\varepsilon^{-2} \log n)$ and a (nonlinear) map $f: \mathbb{R}^d \to \mathbb{R}^m$ such that for all $1 \le i \le n$ and $u \in \mathbb{R}^d,$ \[(1 - O(\varepsilon)) \|u-x_i\|_2 \le \|f(u)-f(x_i)\|_2 \le (1 + O(\varepsilon)) \|u -x_i\|_2.\] \end{theorem} \begin{proof} By Corollary~\ref{MainL1Result}, there exists a $\Pi \in \mathbb{R}^{m \times d}$ with $m = \Theta(\varepsilon^{-2} \log n)$ that provides $\varepsilon$-convex hull distortion for $Y$, where $Y$ is defined in Lemma \ref{ReductionLemma2}. By Lemma \ref{ReductionLemma2}, for each $u\in\mathbb{R}^d$ there exists an outer extension $f^{(u)}: \{x_1, \dots, x_n, u\} \to \mathbb{R}^{m+1}$ with distortion $1+O(\varepsilon)$ sending $x_i$ to $(\Pi x_i, 0)$. Therefore, we map $x_i \mapsto (\Pi x_i, 0)$ and map each $u \not\in \{x_1, ..., x_n\}$ to $f^{(u)}(u)$, which gives us a terminal embedding to $m = O(\varepsilon^{-2} \log n)$ dimensions with distortion $1+O(\varepsilon).$ \end{proof} \subsection{Algorithm to Construct Terminal Embedding} We briefly note how one can produce a terminal embedding into $\mathbb{R}^{m}$ with a Monte Carlo randomized polynomial time algorithm, where $m = O(\varepsilon^{-2} \log n).$ By choosing a random $\Pi$ from Subsection \ref{L1JL}, we get with at least $1 - n^{-\Theta(1)}$ probability a matrix providing $\varepsilon$-convex hull distortion for our set $Y$ in Lemma \ref{ReductionLemma2}. To map any point in $\mathbb{R}^d$ into $\mathbb{R}^{m+1}$ dimensions, for any $u \in \mathbb{R}^d$, it suffices to find a $u' \in \mathbb{R}^m$ such that if $x_k$ is the point in $X$ closest to $u$, $||u'||_2 \le ||u - x_k||_2$ and for all $i$, \[|\langle u', \Pi (x_i-x_k)\rangle - \langle u - x_k, x_i-x_k\rangle| \le \varepsilon \|u-x_k\|_2 \|x_i-x_k\|_2.\] Assuming that $\Pi$ provides an $\varepsilon$-convex hull distortion for $Y$, such a $u'$ exists for all $u$, which means that $u'$ can be found with semidefinite programming in polynomial time, as noted in \cite{MahabadiMMR18}.
{ "timestamp": "2018-10-23T02:26:47", "yymm": "1810", "arxiv_id": "1810.09250", "language": "en", "url": "https://arxiv.org/abs/1810.09250", "abstract": "Let $\\varepsilon\\in(0,1)$ and $X\\subset\\mathbb R^d$ be arbitrary with $|X|$ having size $n>1$. The Johnson-Lindenstrauss lemma states there exists $f:X\\rightarrow\\mathbb R^m$ with $m = O(\\varepsilon^{-2}\\log n)$ such that $$ \\forall x\\in X\\ \\forall y\\in X, \\|x-y\\|_2 \\le \\|f(x)-f(y)\\|_2 \\le (1+\\varepsilon)\\|x-y\\|_2 . $$ We show that a strictly stronger version of this statement holds, answering one of the main open questions of [MMMR18]: \"$\\forall y\\in X$\" in the above statement may be replaced with \"$\\forall y\\in\\mathbb R^d$\", so that $f$ not only preserves distances within $X$, but also distances to $X$ from the rest of space. Previously this stronger version was only known with the worse bound $m = O(\\varepsilon^{-4}\\log n)$. Our proof is via a tighter analysis of (a specific instantiation of) the embedding recipe of [MMMR18].", "subjects": "Data Structures and Algorithms (cs.DS); Functional Analysis (math.FA); Machine Learning (stat.ML)", "title": "Optimal terminal dimensionality reduction in Euclidean space", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.98901305931244, "lm_q2_score": 0.8267118004748677, "lm_q1q2_score": 0.8176287669573444 }
https://arxiv.org/abs/2009.10873
Dyson's Crank and the Mex of Integer Partitions
Andrews and Newman have recently introduced the notion of the mex of a partition, the smallest positive integer that is not a part. The concept has been used since at least 2011, though, with connections to Frobenius symbols. Recently the parity of the mex has been associated to the crank statistic named by Dyson in 1944. In this note, we extend and strengthen the connection between the crank and mex (along with a new generaliztion of the mex) by proving a number of properties that naturally relate these partition statistics.
\section{Introduction} Given a positive integer $n$, a partition of $n$ is a collection of positive integers $\lambda = (\lambda_1, \ldots, \lambda_j)$ with $\sum \lambda_i = n$. The $\lambda_i$, called parts, are ordered so that $\lambda_1 \ge \cdots \ge \lambda_j$. Write $p(n)$ for the number of partitions of $n$. The crank is a partition statistic requested by Dyson \cite{d} in 1944 to combinatorially prove a congruence of $p(n)$ values found by Ramanujan using analytic methods. The interested reader can learn more about the crank, including its definition, from the 1988 work of Andrews and Garvan \cite{ag}. The primary tool that we need in this paper is the generating function for the crank. Let $M(m,n)$ denote the number of partitions of $n$ with crank $m$. We use the $q$-series notation \[(a;q)_n = (1-aq)\cdots(1-aq^n) \text{ and } (a;q)_\infty = \lim_{n \rightarrow \infty} (a;q)_n.\] When $a=q$, we write $(q)_n$ and $(q)_\infty$ instead. The following is \cite[Theorem 7.19]{g}. \begin{theorem}[Garvan] For a given integer $m$, the generating function for the number of partitions of $n$ with crank $m$ is \begin{equation} \label{cgf} \sum_{n=0}^\infty M(m,n) q^n = \frac{1}{(q)_\infty} \sum_{n=1}^\infty (-1)^{n-1} q^{n(n-1)/2 + n|m|} (1-q^n). \end{equation} \end{theorem} While introducing work of Karl Mahlburg, George Andrews and Ken Ono \cite{ao} discussed ``the central role of the crank in the theory of partitions.'' Our goal is to further elucidate Dyson's crank by way of a newer partition statistic, the mex. The minimal excludant of a set of integers, abbreviated mex, is the smallest positive integer not in the set. Some examples of the mex applied to partitions follow. \[\mex(3,2) = 1, \quad \mex(3,1,1) = 2, \quad \mex(2,2,1) = 3.\] Andrews first used ``the smallest part that is \textit{not} a summand'' in 2011 \cite{a}; he and Newman started using the term mex, from combinatorial game theory, in 2019 \cite{an19}. Here is an outline of the remainder of the paper. In Section \ref{s2}, we introduce a generalization of the mex statistic and connect it to the partitions with a given minimum crank. This, along with a formula for the number of partitions with a given mex, leads to a formula for the number of partitions with a given crank. Section \ref{s3} establishes connections between the crank and a classic partition representation, the Frobenius symbol. Finally, in Section \ref{s4}, we refine the partitions with odd mex by considering the mex modulo 4; this unveils a surprising identity in which the function $q(n)$, the number of partitions of $n$ with distinct parts, arises. \section{Crank and a generalized mex} \label{s2} We introduce a generalization of the mex in order to find results on $M(m,n)$ for arbitrary $m$. We will focus on cases with $m \ge 0$. Results for negative $m$ follow from the relation $M(m,n) = M(-m,n)$ \cite[(1.23)]{g}. \begin{definition} Given a partition $\lambda$ and a positive integer $j$ that is a part of $\lambda$, define $\mex_j(\lambda)$ to be the least integer greater than $j$ that is not a part of $\lambda$. If $j$ is not among the parts of $\lambda$, then $\mex_j(\lambda)$ is not defined. \end{definition} If we consider 0 to be a part of every partition, then $\mex_j$ for $j=0$ reduces to the mex described above. Our first result gives the number of partitions with a specified minimum crank in terms of $\mex_j$. \begin{theorem} \label{jcrank} Given a nonnegative integer $j$, the number of partitions of $n$ with $\crank(\lambda) \ge j$ equals the number of partitions of $n$ such that $\mex_j(\lambda) - j$ is odd. \end{theorem} The core of the proof is in modifying \eqref{cgf} to address partitions with a specified minimum crank. We state this as a lemma since it leads to additional results. \begin{lemma} \label{crankj+} For a given integer $j$, the generating function for the number of partitions of $n$ with $\crank(\lambda) \ge j$ is \[ \frac{1}{(q)_\infty} \sum_{k=0}^\infty q^{j(2k+1) + k(2k+1)} (1-q^{2k+j+1}). \] \end{lemma} \begin{proof} Summing \eqref{cgf} over $m\geq j$ yields \begin{align} \sum_{m= j}^\infty \sum_{n=0}^\infty M(m,n) q^n & = \frac{1}{(q)_\infty} \sum_{n=1}^\infty (-1)^{n-1} q^{n(n-1)/2} (1-q^n) \sum_{m= j}^\infty q^{nm} \nonumber \\ & = \frac{1}{(q)_\infty} \sum_{n=1}^\infty (-1)^{n-1} q^{n(n-1)/2} (1-q^n) \left( \frac{q^{nj}}{1-q^n} \right) \nonumber \\ & = \frac{1}{(q)_\infty} \sum_{n=1}^\infty (-1)^{n-1} q^{n(n-1)/2}q^{nj} \label{later} \\ & = \frac{1}{(q)_\infty} \sum_{n=1}^\infty (-1)^{n-1} q^{n(n-1+2j)/2} \nonumber \\ & = \frac{1}{(q)_\infty} \left( \sum_{k= 1}^\infty q^{(2k-1)((2k-1)-1+2j)/2} - \sum_{k= 1}^\infty q^{2k(2k-1+2j)/2} \right) \nonumber \\ & = \frac{1}{(q)_\infty} \left( \sum_{k= 0}^\infty q^{(2k+1)(k+j)} - \sum_{k= 0}^\infty q^{(k+1)(2k+1+2j)} \right) \nonumber \\ & = \frac{1}{(q)_\infty} \sum_{k=0}^\infty q^{j(2k+1)+k(2k+1)} (1 - q^{2k+j+1}). \nonumber \qedhere \end{align} \end{proof} \begin{proof}[Proof of Theorem \ref{jcrank}.] The partition condition $\mex_j - j$ odd means, for some nonnegative integer $k$, that $j, j+1, \dots, 2k+j$ are parts and $2k+j+1$ is not. The generating function for such partitions of $n$ is \begin{align*} \sum_{k=0}^\infty \frac{q^{j+(j+1)+\cdots+(j+2k)}}{\prod_{i=1, \, i \ne 2k+j+1}^\infty (1-q^i)} & = \frac{1}{(q)_\infty} \sum_{k=0}^\infty q^{j(2k+1) + (2k)(2k+1)/2} (1-q^{2k+j+1}) \\ & = \frac{1}{(q)_\infty} \sum_{k=0}^\infty q^{j(2k+1) + k(2k+1)} (1-q^{2k+j+1}) \end{align*} which matches the generating function in Lemma \ref{crankj+} for the number of partitions of $n$ with $\crank(\lambda) \ge j$. \end{proof} Let $o(n)$ be the number of partitions of $n$ with odd mex and $e(n)$ the number with even mex. The $j=0$ case of Theorem \ref{jcrank}, i.e., that $o(n)$ equals the number of partitions of $n$ with nonnegative crank, was proven independently by Andrews--Newman \cite[Thm. 2]{an20} and Hopkins--Sellers \cite[Thm. 1]{hs} in 2020. For the $j=1$ case, the conditions of Theorem \ref{jcrank} are equivalent to having even mex, thus $e(n)$ equals the number of partitions of $n$ with positive crank \cite[Cor. 2]{hs}. Next, we derive a formula for the number of partitions with a given crank in terms of $p(n)$. We will then connect a special case of this with the mex. \begin{corollary} \label{crankrecur} Given nonnegative $j$, the number of partitions of $n$ with crank $j$ is \begin{align*} M(j,n) & = \sum_{k=1}^\infty (-1)^{k+1} \left(p\!\left( n - \frac{k(k+2j-1)}{2} \right) - p\!\left( n - \frac{k(k+2j+1)}{2}\right)\right) \\ & = p(n-j) - p(n-j-1) - p(n-2j-1) + p(n-2j-3) + \cdots. \end{align*} \end{corollary} Note that such a formula involves finitely many terms since $p(n) = 0$ for $n<0$. \begin{proof} Using \eqref{later} and interpreting $1/(q)_\infty$ as $\sum p(n) q^n$, the number of partitions of $n$ with minimum crank $j$ is \[\sum_{k=1}^\infty (-1)^{k+1} p\!\left( n - \binom{k}{2}-kj \right).\] Subtracting from this the analogous expression for the number of partitions of $n$ with minimum crank $j+1$ gives \[M(j,n) = \sum_{k=1}^\infty (-1)^{k+1} \left(p\!\left( n - \binom{k}{2}-kj \right) - p\!\left( n - \binom{k}{2}-k(j+1) \right)\right)\] which is equivalent to the stated result. \end{proof} One could also prove Corollary \ref{crankrecur} from \eqref{cgf}. We want to highlight the $m=0$ case of Corollary \ref{crankrecur}, which is Corollary \ref{0crank}. While it needs no additional proof, we provide a second verification in order to apply the following result which we will use again in the final section. Let $x(m,n)$ be the number of partitions of $n$ with mex $m$. Write $t_k = 1 + \cdots + k$ for the $k$th triangular number. \begin{prop} Given positive integers $n$ and $m$, the number of partitions of $n$ with mex $m$ is \[x(m,n) = p(n-t_{m-1}) - p(n-t_m).\] \label{mexform} \end{prop} \begin{proof} For a partition to have mex $m$, it must include $1, \ldots, m-1$ as parts and exclude $m$. The number of partitions of $n$ with the required included parts is $p(n - t_{m-1})$, since removing one 1, one 2, \dots, one $m-1$ leaves a partition of $n-t_{m-1}$. The number of partitions of $n$ with $m$ also included as a part is \[p(n-t_{m-1}-m) = p(n-t_m),\] so the number with $m$ excluded is $p(n - t_{m-1}) - p(n-t_m)$. \end{proof} One can find the ideas for this result in the second proof of \cite[Thm 1.1]{an19}. We can use Proposition \ref{mexform} and the earlier results about odd and even mex to verify the formula for $M(0,n)$ in terms of $p(n)$. \begin{corollary} \label{0crank} The number of crank 0 partitions of $n$ is \begin{align*} M(0,n) & = p(n) + 2 \sum_{k=1}^\infty (-1)^k p\!\left(n-\frac{k(k+1)}{2}\right) \\ & = p(n) - 2p(n-1) + 2p(n-3) - 2p(n-6) + \cdots. \end{align*} \end{corollary} \begin{proof} By the $j=0$ case of Theorem \ref{jcrank}, there are $o(n)$ partitions of $n$ with nonnegative crank. As discussed above, there are $e(n)$ partitions of $n$ with positive crank. Therefore, \begin{align*} M(0,n) & = o(n) - e(n) \\ &= \sum_{k=0}^\infty x(2k+1,n) - \sum_{k=0}^\infty x(2k+2,n) \\ & = \sum_{k=0}^\infty (p(n-t_{2k}) - p(n-t_{2k+1})) - \sum_{k=0}^\infty (p(n-t_{2k+1}) - p(n-t_{2k+2})) \\ & = \sum_{k=0}^\infty (p(n-t_{2k}) - 2p(n-t_{2k+1}) + p(n-t_{2k+2})) \\ & = p(n) + 2 \sum_{k=1}^\infty (-1)^k p\!\left(n-\frac{k(k+1)}{2}\right). \qedhere \end{align*} \end{proof} This formula for $M(0,n)$ was recorded by Vaclav Kotesovec in 2016 \cite[A064410]{o}. In 2019, Yuefei Shen showed that $o(n) \ge e(n)$ for $n > 1$ \cite[Thm. 1]{s}. Hopkins--Sellers improved this to $o(n) > e(n)$ for $n>2$ \cite[Cor. 3]{hs}. Corollary \ref{0crank} provides a precise formula for $o(n) - e(n)$. \section{Frobenius symbol and the crank} \label{s3} Frobenius developed a two-row notation for partitions \cite{f} now called the Frobenius symbol. Each row consists of strictly decreasing nonnegative integers; see \cite[p. 78]{ae} for more information. In 2011, Andrews \cite[Thm. 4]{a} proved that the number of partitions of $n$ whose Frobenius symbol has no 0 in the top row equals the number of partitions of $n$ with odd mex (in the current terminology). Thus, by Andrews--Newman \cite[Thm. 2]{an20} and Hopkins--Sellers \cite[Thm. 1]{hs}, the number of partitions of $n$ whose Frobenius symbol has no 0 in the top row equals the number of partitions of $n$ with nonnegative crank. In this section, we extend this result of Andrews in two ways. First, we connect the crank 0 partitions with partitions whose Frobenius symbols have no 0 in either row. Let $[q^k]f(q)$ denote the coefficient of $q^k$ in the polynomial $f(q)$. \begin{prop} \label{noF0} The number of partitions of $n$ with crank 0 equals the number of partitions of $n$ whose Frobenius symbol has no 0 minus the number of partitions of $n-1$ whose Frobenius symbol has no 0. \end{prop} \begin{proof} The generating function for Frobenius symbols with no occurrence of 0 is, letting $s$ be the number of parts, \[ \sum_{s=0}^\infty q^s [t^s w^s] (-tq;q)_\infty (-wq;q)_\infty = \sum_{s=0}^\infty q^{s^2} \frac{q^{2s}}{(q)_s^2}.\] Thus the number of partitions of $n$ whose Frobenius symbol has no 0 minus the number of partitions of $n-1$ of the same type is \[\sum_{s=0}^\infty q^{s^2} \frac{q^{2s}}{(q)_s^2} - q\sum_{s=0}^\infty q^{s^2} \frac{q^{2s}}{(q)_s^2} = (1-q) \sum_{s=0}^\infty \frac{q^{s^2+2s}}{(q)_s^2}.\] Now the $a=b=0$ case of Heine's transformation as given in \cite[Eq. III.3]{gr} with $c=q$ and $z=q^2$ gives \[ \sum_{s=0}^\infty \frac{q^{s^2-s}(q^3)^s}{(q)_s^2} = (q^2;q)_\infty \sum_{k=0}^\infty \frac{q^{2k}}{(q)_k^2}.\] Since $(q^2;q)_\infty = (q)_\infty/(1-q)$, we have \[ (1-q) \sum_{s=0}^\infty \frac{q^{s^2+2s}}{(q)_s^2} = (q)_\infty \sum_{k=0}^\infty \frac{q^{2k}}{(q)_k^2}.\] The right-hand side of this last equation is an alternative expression for the generating function for the number of crank 0 partitions of $n$ \cite[p. 5]{an20}. \end{proof} The sequence of the number of partitions of $n$ whose Frobenius symbol has no 0 is \cite[A188674]{o}. By Proposition \ref{noF0}, the first differences of this sequence give the number of partitions with crank 0, \cite[A064410]{o}. For our second result in this section, we connect partitions with minimal crank $j$ to certain partitions whose Frobenius symbols have no $j$ in the top row. Andrews's \cite[Thm. 4]{a} is the $j=0$ case of the following theorem. \begin{theorem} The number of partitions of $n$ with $\crank(\lambda) \ge j$ equals the number of partitions of $n - j$ whose Frobenius symbol has no $j$ in its top row. \end{theorem} \begin{proof} The generating function for Frobenius symbols with no $j$ in the top row is, letting $s$ be the number of parts, \begin{align*} \sum_{s=0}^\infty q^s [t^s w^s] \frac{(-t;q)_\infty (-w;q)_\infty}{1+tq^j} & =\sum_{s=0}^\infty q^s \frac{q^{\binom{s}{2}}}{(q)_s} \left(\sum_{b=0}^s (-1)^b q^{jb} \frac{q^{\binom{s-b}{2}}}{(q)_{s-b}}\right) \\ & =\sum_{s,b \ge 0} q^{s+b} \frac{q^{\binom{s+b}{2}}}{(q)_{s+b}} (-1)^b q^{jb} \frac{q^{\binom{s}{2}}}{(q)_s} \\ & = \sum_{b=0}^\infty q^{b+\binom{b}{2}} (-1)^b {q^{jb}} \left(\sum_{s=0}^\infty \frac{q^{s^2+bs}}{(q)_s (q)_{s+b}}\right) \\ & = \frac{1}{(q)_\infty} \sum_{b=0}^\infty (-1)^b q^{\binom{b+1}{2}} {q^{jb}} \end{align*} where the last line follows from \cite[(4.1)]{a} based on the decomposition of all partitions by Durfee rectangles of size $s \times (s+b)$ \cite[p. 92]{gh}. Next, by \eqref{later}, we know that the generating function for partitions with minimum crank $j$ is \begin{align*} \frac{1}{(q)_\infty} \sum_{n=1}^\infty (-1)^{n-1} q^{n(n-1)/2}q^{nj} & = \frac{1}{(q)_\infty} \sum_{n=0}^\infty (-1)^{n} q^{(n+1)n/2}q^{(n+1)j} \\ & = \frac{1}{(q)_\infty} \sum_{n=0}^\infty (-1)^{n} q^{\binom{n+1}{2}}q^{nj+j} \\ & = q^j \frac{1}{(q)_\infty} \sum_{n=0}^\infty (-1)^{n} q^{\binom{n+1}{2}}q^{nj}. \end{align*} Comparing the corresponding coefficients in the two expressions gives the desired result. \end{proof} \section{Splitting the odd mexes} \label{s4} Recall that $o(n)$ is the number of partitions of $n$ with odd mex. Write $o_1(n)$ for the number of partitions of $n$ with mex congruent to $1 \bmod 4$, similarly $o_3(n)$ for $3 \bmod 4$, so that $o(n) = o_1(n) + o_3(n)$. Let $q(n)$ denote the number of partitions of $n$ with distinct parts. We prove an interesting relationship between $o_1(n)$, $o_3(n)$, and $q(n)$ which depends on the parity of $n$. \begin{prop} \label{o13} For $n \ge1$, \[o_1(n) = \begin{cases} o_3(n) & \text{if $n$ is odd,} \\ o_3(n) + q(n/2) & \text{if $n$ is even.} \end{cases} \] \label{comb} \end{prop} To prove Proposition \ref{o13}, we use the following 1973 result of John Ewell \cite[Thm. 2]{e}. Recall that $t_k$ denotes the $k$th triangular number $1+\cdots+k$. \begin{theorem}[Ewell] For each nonnegative integer $k$, \begin{gather} \sum_{j=0}^\infty (-1)^{t_j} p(2k-t_j) = q(k), \label{eweven} \\ \sum_{j=0}^\infty (-1)^{t_j} p(2k+1-t_j) = 0. \label{ewodd} \end{gather} \end{theorem} \begin{proof}[Proof of Proposition \ref{o13}] From Proposition \ref{mexform}, \begin{align*} o_1(n) & = \sum_{i=0}^\infty x(n,4i+1) = \sum_{i=0}^\infty (p(n-t_{4i}) - p(n-t_{4i+1})), \\ o_3(n) & = \sum_{i=0}^\infty x(n,4i+3) = \sum_{i=0}^\infty (p(n-t_{4i+2}) - p(n-t_{4i+3})). \end{align*} Consider the two cases of the theorem separately. First, suppose $n = 2k$. Write \eqref{eweven} as $$\sum_{i=0}^\infty (p(2k-t_{4i}) - p(2k-t_{4i+1}) - p(2k-t_{4i+2}) + p(2k-t_{4i+3})) = q(k).$$ Then we have $$\sum_{i=0}^\infty (p(2k-t_{4i}) - p(2k-t_{4i+1})) = q(k) + \sum_{i=0}^\infty (p(2k-t_{4i+2}) - p(2k-t_{4i+3}))$$ which, by Proposition \ref{mexform} and the comment above, can be written $$o_1(2k) = q(k) + o_3(2k).$$ Second, suppose $n = 2k+1$. Write \eqref{ewodd} as $$\sum_{i=0}^\infty (p(2k+1-t_{4i}) - p(2k+1-t_{4i+1}) - p(2k+1-t_{4i+2}) + p(2k+1-t_{4i+3})) = 0.$$ Then we have $$\sum_{i=0}^\infty (p(2k+1-t_{4i}) - p(2k+1-t_{4i+1})) = \sum_{i=0}^\infty (p(2k+1-t_{4i+2}) - p(2k+1-t_{4i+3}))$$ which, by Proposition \ref{mexform} and the comment above, can be written \[ o_1(2k+1) = o_3(2k+1). \qedhere \] \end{proof} The surprising relationship between $o_1(n)$ and $o_3(n)$ allows us to give a short proof of \cite[Thm. 1.2]{an19}. (Unfortunately, the terms even and odd were reversed in that original theorem statement.) \begin{theorem}[Andrews--Newman] $o(n)$ is almost always even and is odd exactly when $n=j(3j\pm1)$ for some $j$. \end{theorem} \begin{proof} By Proposition \ref{o13}, \begin{align*} o(n) = o_1(n) + o_3(n) &= \begin{cases} 2o_3(n) & \text{if $n$ is odd,} \\ 2o_3(n) + q(n/2) & \text{if $n$ is even,} \end{cases} \\ &\equiv \begin{cases} 0 \bmod 2 & \text{if $n$ is odd,} \\ q(n/2) \bmod 2 & \text{if $n$ is even.} \end{cases} \end{align*} That is, the parity of $o(n)$ for even $n$ reduces to the parity of $q(n/2)$. The result follows from Euler's well-known pentagonal number theorem. \end{proof} We conclude with some ideas for further exploration. Proposition \ref{o13} suggests a bijective proof perhaps along the lines of Franklin's combinatorial proof of Euler's pentagonal number theorem \cite[p. 25]{ae}. Also, which partitions of $n$ with nonnegative crank correspond to $o_1(n)$ rather than $o_3(n)$? Similarly, which partitions of $n$ with Frobenius symbols having no 0 in the top row correspond to $o_1(n)$? Regarding Section \ref{s2}, while there are combinatorial understandings of the crank \cite{bg}, what tools allow for combinatorial proofs involving the mex and $\mex_j$?
{ "timestamp": "2020-09-24T02:06:52", "yymm": "2009", "arxiv_id": "2009.10873", "language": "en", "url": "https://arxiv.org/abs/2009.10873", "abstract": "Andrews and Newman have recently introduced the notion of the mex of a partition, the smallest positive integer that is not a part. The concept has been used since at least 2011, though, with connections to Frobenius symbols. Recently the parity of the mex has been associated to the crank statistic named by Dyson in 1944. In this note, we extend and strengthen the connection between the crank and mex (along with a new generaliztion of the mex) by proving a number of properties that naturally relate these partition statistics.", "subjects": "Combinatorics (math.CO); Number Theory (math.NT)", "title": "Dyson's Crank and the Mex of Integer Partitions", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9916842217682944, "lm_q2_score": 0.8244619306896956, "lm_q1q2_score": 0.8176058881135964 }
https://arxiv.org/abs/1409.3600
Selection Algorithms with Small Groups
We revisit the selection problem, namely that of computing the $i$th order statistic of $n$ given elements, in particular the classic deterministic algorithm by grouping and partition due to Blum, Floyd, Pratt, Rivest, and Tarjan (1973). Whereas the original algorithm uses groups of odd size at least $5$ and runs in linear time, it has been perpetuated in the literature that using smaller group sizes will force the worst-case running time to become superlinear, namely $\Omega(n \log{n})$. We first point out that the usual arguments found in the literature justifying the superlinear worst-case running time fall short of proving this claim. We further prove that it is possible to use group size smaller than $5$ while maintaining the worst case linear running time. To this end we introduce three simple variants of the classic algorithm, the repeated step algorithm, the shifting target algorithm, and the hyperpair algorithm, all running in linear time.
\section{Introduction} \label{sec:intro} Together with sorting, selection is one of the most widely used procedure in computer algorithms. Indeed, it is easy to find hundreds if not thousands of algorithms (documented in at least as many research articles) that use selection as a subroutine. A classical example is~\cite{Me85}. Given a sequence $A$ of $n$ numbers (usually stored in an array), and an integer (target) parameter $1\leq i\leq n$, the selection problem asks to find the $i$th smallest element in $A$. Trivially sorting solves the selection problem, but if one aims at a linear time algorithm, a higher level of sophistication is needed. A now classical approach for selection~\cite{BFP+73,FR75,Hy76,SPP76,Y76} from the 1970s is to use an element in $A$ as a pivot to partition $A$ into two smaller subsequences and recurse on one of them with a (possibly different) selection parameter $i$. The time complexity of this kind of algorithms is sensitive to the pivots used. For example, if a good pivot is used, many elements in $A$ can be discarded; while if a bad pivot is used, in the worst case, the size of the problem may be only reduced by a constant, leading to a quadratic worst-case running time. But choosing a good pivot can be time consuming. Randomly choosing the pivots yields a well-known randomized algorithm with expected linear running time (see {e.g.},~\cite[Ch.~9.2]{CLR+09},~\cite[Ch.~13.5]{KT06}, or~\cite[Ch.~3.4]{MU05}), however its worst case running time is quadratic in $n$. The first deterministic linear time selection algorithm \textsc{select } (called \textsc{pick} by the authors), in fact a theoretical breakthrough at the time, was introduced by Blum~{et~al.}~\cite{BFP+73}. By using the median of medians of small (constant size) disjoint groups of $A$, good pivots that guarantee reducing the size of the problem by a constant fraction can be chosen with low costs. The authors~\cite[page~451, proof of Theorem~1]{BFP+73} required the group size to be at least $5$ for the \textsc{select } algorithm to run in linear time. It has been largely perpetuated in the literature the idea that \textsc{select } with groups of $3$ or $4$ does not run in linear time. Even an exercise of the book by Cormen~{et~al.}~\cite[page~223, exercise~{9.3-1}]{CLR+09} asks the readers to argue that ``\textsc{select } does not run in linear time if groups of $3$ are used''. We first point out that the argument used in the solution to this exercise~\cite[page~23]{CLL09} is invalid failing to provide an input sequence with one third of the elements being discarded in each recursive call in both the current sequence and its sequence of medians; the difficulty (and the flaw in the argument) lies in the fact that these two sequences are not disjoint thus cannot be constructed or controlled independently. The question whether the original \textsc{select } algorithm runs in linear time with groups of $3$ remains open at the time of this writing. Further, we show that this restriction on the group size is unnecessary, namely that group sizes 3 or 4 can be used to obtain a deterministic linear time algorithm for the selection problem. Since selecting the median in smaller groups is easier to implement and requires fewer comparisons ({e.g.}, $3$ comparisons for group size $3$ versus $6$ comparisons for group size $5$), it is attractive to have linear time selection algorithms that use smaller groups. Our main result concerning selection with small group size is summarized in the following theorem. \begin{theorem} \label{thm:thm1} There exist suitable variants of \textsc{select } with groups of $3$ and $4$ running in $O(n)$ time. \end{theorem} \paragraph{Historical background.} The interest in selection algorithms has remained high over the years with many exciting developments ({e.g.}, lower bounds, parallel algorithms, etc) taking place; we only cite a few here~\cite{AKSS89,BJ85,CM89,DHU+01,DZ96,DZ99,FR75,FG79,Ho61,J88,Ki81,Pa96,YY82,Y76}. We also refer the reader to the dedicated book chapters on selection in~\cite{AHU83,Ba88,CLR+09,DPV08,KT06,Kn98} and the recent article~\cite{Ki13}. \paragraph{Outline.} In Section~\ref{sec:prelim}, the classical \textsc{select } algorithm is introduced (rephrased) under standard simplifying assumptions. In Section~\ref{sec:repeat}, we introduce a variant of \textsc{select}, the \emph{repeated step} algorithm, which runs in linear time with both group size $3$ and $4$. With groups of $3$, the algorithm executes a certain step, ``group by $3$ and find the medians of the groups'', twice in a row. In Section~\ref{sec:shift}, we introduce another variant of \textsc{select}, the \emph{shifting target} algorithm, a linear time selection algorithm with group size~$4$. In each iteration, upper or lower medians are used based on the current rank of the target, and the shift in the target parameter $i$ is controlled over three consecutive iterations. In Section~\ref{sec:remark}, we conclude with some remarks and a conjecture on the running time of the original \textsc{select } algorithm from~\cite{BFP+73} with groups of $3$ and~$4$. \section{Preliminaries} \label{sec:prelim} Without affecting the results, the following two standard simplifying assumptions are convenient: (i)~the input sequence $A$ contains $n$ distinct numbers; and (ii)~the floor and ceiling functions are omitted in the descriptions of the algorithms and their analyses. Under these assumptions, \textsc{select } with groups of $5$ (from~\cite{BFP+73}) can be described as follows (using this group size has become increasingly popular, see {e.g.},~\cite[Ch.~9.2]{CLR+09}): \begin{enumerate} \itemsep 2pt \item \label{alg5:step1} If $n\leq 5$, sort $A$ and return the $i$th smallest number. \item Arrange $A$ into groups of size 5. Let $M$ be the sequence of medians of these $n/5$ groups. Select the median of $M$ recursively, let it be $m$. \item Partition $A$ into two subsequences $A_1=\{x|x<m\}$ and $A_2=\{x|x>m\}$ (order of other elements is preserved). If $|A_1|=i-1$, return $m$. Otherwise if $|A_1|>i-1$, go to step~\ref{alg5:step1} with $A \leftarrow A_1$ and $n\leftarrow |A_1|$. Otherwise go to step~\ref{alg5:step1} with $A \leftarrow A_2$, $n\leftarrow |A_2|$ and $i\leftarrow i-|A_1|-1$. \end{enumerate} Denote the worst case running time of the recursive selection algorithm on an $n$-element input by $T(n)$. As shown in Figure~\ref{fig:group5}, at least $(n/5)/2*3=3n/10$ elements are discarded at each iteration, which yields the recurrence $$T(n)\leq T(n/5)+T(7n/10)+O(n).$$ Since the coefficients sum to $1/5+7/10=9/10<1$, the recursion solves to $T(n)=\Theta(n)$ (as it is well-known). \medskip \begin{figure}[htbp] \centering \includegraphics[scale=0.6]{group5.pdf} \vspace*{0.05in} \caption{One iteration of the \textsc{select } algorithm with group size 5. At least $3n/10$ elements can be discarded.} \label{fig:group5} \end{figure} \section{The Repeated Step Algorithm} \label{sec:repeat} Using group size $3$ directly in the \textsc{select } algorithm in \cite{BFP+73} yields \begin{equation} \label{eq:4} T(n)\leq T(n/3)+T(2n/3)+O(n), \end{equation} which solves to $T(n)=O(n\log n)$. Here a large portion (at least one third) of $A$ is discarded in each iteration but the cost of finding such a good pivot is too high, namely $T(n/3)$. The idea of our \emph{repeated step} algorithm, inspired by the algorithm in~\cite{BCC+00}, is to find a weaker pivot in a faster manner by performing the operation ``group by $3$ and find the medians'' twice in a row (as illustrated in Figure~\ref{fig:group3}). \paragraph{Algorithm} \begin{enumerate} \itemsep 2pt \item \label{alg3:step1} If $n\leq 3$, sort $A$ and return the $i$t smallest number. \item \label{alg3:step2} Arrange $A$ into groups of size $3$. Let $M$ be the sequence of medians of these $n/3$ groups. \item Arrange $M$ into groups of size $3$. Let $M'$ be the sequence of medians of these $n/9$ groups. \item \label{alg3:step3} Select the median of $M'$ recursively, let it be $m$. \item Partition $A$ into two subsequences $A_1=\{x|x<m\}$ and $A_2=\{x|x>m\}$. If $|A_1|=i-1$, return $m$. Otherwise if $|A_1|>i-1$, go to step~\ref{alg3:step1} with $A\leftarrow A_1$ and $n\leftarrow |A_1|$. Otherwise go to step~\ref{alg3:step1} with $A\leftarrow A_2$, $n\leftarrow |A_2|$ and $i\leftarrow i-|A_1|-1$. \end{enumerate} \begin{figure}[htbp] \centering \includegraphics[scale=0.6]{group3.pdf} \caption{One iteration of the \emph{repeated step} algorithm with groups of $3$. Empty disks represent elements that are guaranteed to be smaller or equal to $m$.} \label{fig:group3} \end{figure} \paragraph{Analysis.} Since elements are discarded if and only if they are too large or too small to be the $i$th smallest element, the correctness of the algorithm follows. Regarding the time complexity of this algorithm, we have the following lemma: \begin{lemma} \label{lem:1} The repeated step algorithm with groups of $3$ runs in $\Theta(n)$ time on an $n$-element input. \end{lemma} \begin{proof} By finding the median of medians of medians instead of the median of medians, the cost of selecting the pivot $m$ reduces from $T(n/3)+O(n)$ to $T(n/9)+O(n)$. We need to determine how well $m$ partitions $A$ in the worst case. In step~\ref{alg3:step3}, $m$ is guaranteed to be greater or equal to $(n/9)/2*2=n/9$ elements in $M$. Each element in $M$ is a median of a group of size $3$ in $A$, so it is greater or equal to 2 elements in its group. All the groups of $A$ are disjoint, thus $m$ is at least greater or equal to $2n/9$ elements in $A$. Similarly, $m$ is at least smaller or equal to $2n/9$ elements in $A$. Thus, in the last step, at least $2n/9$ elements can be discarded. The recursive call in step~\ref{alg3:step3} takes $T(n/9)$ time. So the resulting recurrence is $$T(n)\leq T(n/9)+T(7n/9)+O(n),$$ and since the coefficients on the right side sum to $8/9<1$, we have $T(n)=\Theta(n)$, as required. \end{proof} \section{The Shifting Target Algorithm} \label{sec:shift} In the \textsc{select } algorithm introduced in \cite{BFP+73}, the group size is restricted to odd numbers in order to avoid the calculation of the average of the upper and lower median. For group size of $4$, depending on the choice of upper, lower or average median, there are three possible partial orders to be considered (see Figure~\ref{fig:partial}). \begin{figure}[htbp] \centering \includegraphics[scale=0.6]{partialOrder.pdf} \caption{Three partial orders of $4$ elements based on the upper (left), lower (middle) and average (right) medians. The empty square represents the average of the upper and lower median which is not necessarily part of the $4$-element sequence.} \label{fig:partial} \end{figure} If the upper (or lower) median is always used, only $(n/4)/2*2=n/4$ elements are guaranteed to be discarded in each iteration (see Figure~\ref{fig:group4_lower}) which gives the recurrence \begin{equation} \label{eq:1} T(n)\leq T(n/4)+T(3n/4)+O(n). \end{equation} The term $T(n/4)$ is for the recursive call to find the median of all $n/4$ medians. This recursion solves to $T(n)=O(n\log n)$. Even if we use the average of the two medians, the recursion remains the same since only $2$ elements from each of the $(n/4)/2=n/8$ groups are guaranteed to be discarded. Observe that if the target parameter satisfies $i \leq n/2$ (resp., $i \geq n/2$), using the lower (resp., upper) median gives a better chance to discard more elements and thus obtain a better recurrence; detailed calculations are given in the proof of Lemma~\ref{lem:2}. Inspired by this idea, we propose the \emph{shifting target} algorithm as follows: \paragraph{Algorithm} \begin{enumerate} \itemsep 2pt \item \label{alg4:step1} If $n\leq 4$, sort $A$ and return the $i$th smallest number. \item Arrange $A$ into groups of size $4$. Let $M$ be the sequence of medians of these $n/4$ groups. If $i\leq n/2$, the lower medians are used; otherwise the upper medians are used. Select the median of $M$ recursively, let it be $m$. \item Partition $A$ into two subsequences $A_1=\{x|x<m\}$ and $A_2=\{x|x>m\}$. If $|A_1|=i-1$, return $m$. Otherwise if $|A_1|>i-1$, go to step~\ref{alg4:step1} with $A\leftarrow A_1$ and $n\leftarrow |A_1|$. Otherwise go to step~\ref{alg4:step1} with $A\leftarrow A_2$, $n\leftarrow |A_2|$ and $i\leftarrow i-|A_1|-1$. \end{enumerate} \begin{figure}[htbp] \centering \includegraphics[scale=0.6]{group4.pdf} \caption{Group size $4$ with lower medians used.} \label{fig:group4_lower} \end{figure} \paragraph{Analysis.} Regarding the time complexity, we have the following lemma: \begin{lemma} \label{lem:2} The shifting target algorithm with group size $4$ runs in $\Theta(n)$ time on an $n$-element input. \end{lemma} \begin{proof} Assume first that $i\leq n/4$ in some iteration so the lower medians are used. Recall that $m$ is guaranteed to be greater or equal to $(n/4)/2*2=n/4$ numbers in $A$. So either $m$ is the $i$th smallest element in $A$ or at least $(n/4)/2*3=3n/8$ largest numbers are discarded (see Figure~\ref{fig:group4_lower}), hence the worst-case running time recurrence is \begin{equation} \label{eq:2} T(n)\leq T(n/4)+T(5n/8)+O(n). \end{equation} Observe that in this case the coefficients on the right side sum to $7/8<1$, yielding a linear solution, as required. Now consider the case $n/4<i\leq n/2$, so the lower medians are used. If $|A_1| \geq i$, {i.e.}, the rank of $m$ is higher than $i$, again at least $(n/4)/2*3=3n/8$ largest numbers are discarded and~\eqref{eq:2} applies. Otherwise, suppose that only $t =|A_1| \geq (n/4)/2*2=n/4$ smallest numbers are discarded. Then in the next iteration, $i'=i-t$, $n'=n-t$. If $i'\leq n'/4$, at least $3n'/8$ numbers are discarded. The first iteration satisfies recurrence~\eqref{eq:1} and we can use recurrence~\eqref{eq:2} to bound the term $T(3n/4)$ from above. We deduce that in two iterations the worst case running time satisfies the recurrence: \begin{align} \label{eq:3} T(n)&\leq T(n/4)+T(3n/4)+O(n) \nonumber \\ &\leq T(n/4)+ T((3n/4)/4) + T((3n/4)*5/8) + O(n) \nonumber \\ &= T(n/4)+T(3n/16)+T(15n/32)+O(n). \end{align} Observe that the coefficients on the right side sum to $29/32<1$, yielding a linear solution, as required. Subsequently, we can therefore assume that $i' \geq n'/4$. We have \begin{align*} i'/n'&=(i-t)/(n-t)\\ &\leq(i-n/4)/(n-n/4)\\ &\leq(n/2-n/4)/(n-n/4)\\ &=1/3. \end{align*} Since $1/4<i'/n'\leq 1/3 \leq 1/2$, the lower medians will be used. As described above, if at least $3n'/8$ largest numbers are discarded, in two iterations, the worst case running time satisfies the same recurrence~\eqref{eq:3}. So suppose that only $t'\geq (n'/4)/2*2=n'/4$ smallest numbers are discarded. Let $i''=i'-t'$, $n''=n'-t'$. We have \begin{align*} i''/n''&=(i'-t')/(n'-t')\\ &\leq(i'-n'/4)/(n'-n'/4)\\ &\leq(n'/3-n'/4)/(n'-n'/4)\\ &=1/9. \end{align*} Since $i''/n''<1/4$, in the next iteration, at least $3n''/8$ numbers will be discarded. The first two iterations satisfy recurrence~\eqref{eq:1} and we can use recurrence~\eqref{eq:2} to bound the term $T(9n/16)$ from above. We deduce that in three iterations the worst case running time satisfies the recurrence: \begin{align*} T(n)&\leq T(n/4)+T(3n/4)+O(n)\\ &\leq T(n/4)+ T((3n/4)/4) + T((3n/4)*3/4) + O(n)\\ &= T(n/4)+T(3n/16)+T(9n/16)+O(n)\\ &\leq T(n/4)+T(3n/16)+ T((9n/16)/4) + T((9n/16)*5/8) + O(n)\\ &= T(n/4)+T(3n/16)+T(9n/64)+T(45n/128)+O(n). \end{align*} The sum of the coefficients on the right side is $119/128<1$, so again the solution is $T(n)=\Theta(n)$. By symmetry, the analysis also holds for the case $i \geq n/2$, and the proof of Lemma~\ref{lem:2} is complete. \end{proof} \section{Concluding Remarks}\label{sec:remark} A similar idea of repeating the group step also applies to the case of groups of $4$ and yields $$T(n)\leq T(n/16)+T(7n/8)+O(n),$$ which gives another linear time selection algorithm with group size $4$. Another variant of \textsc{select } with group size $4$ can be obtained by using the ideas of both algorithms together, {i.e.}, repeat the grouping by $4$ step twice in a row while $M$ contains the lower medians and $M'$ contains the upper medians (or vice versa). Recursively selecting the median $m$ of $M'$ takes time $T(n/16)$. Notice that $m$ is greater or equal to at least $(n/16)/2*3=3n/32$ elements in $M$ of which each is greater or equal to 2 elements in its group in $A$. So $m$ is greater or equal to at least $3n/16$ elements of $A$. Also, $m$ is smaller or equal to at least $(n/16)/2*2=n/16$ elements in $M$ of which each is smaller or equal to $3$ elements in its group of $A$. So $m$ is smaller or equal to at least $3n/16$ elements of $A$, thus the resulting recurrence is $$T(n)\leq T(n/16)+T(13n/16)+O(n), $$ again with a linear solution, as desired. \paragraph{Final comment.} The question whether the original selection algorithm introduced in~\cite{BFP+73} (outlined in Section~\ref{sec:prelim}) runs in linear time with group size 3 and $4$ remains open. Although the recurrences % \begin{align*} T(n)&\leq T(n/3)+T(2n/3)+O(n)\text{, and}\\ T(n)&\leq T(n/4)+T(3n/4)+O(n) \end{align*} (see~\eqref{eq:4} and~\eqref{eq:1}) for its worst-case running time with these group sizes both solve to $T(n)=O(n\log n)$, we believe that they only give non-tight upper bounds on the worst case scenarios. In any case, and against popular belief we think that $\Theta(n \log{n})$ is \emph{not} the answer: \begin{conjecture} The \textsc{select } algorithm introduced by Blum~{et~al.}~\cite{BFP+73} runs in $o(n \log{n})$ time with groups of $3$ or $4$. \end{conjecture}
{ "timestamp": "2014-09-15T02:01:28", "yymm": "1409", "arxiv_id": "1409.3600", "language": "en", "url": "https://arxiv.org/abs/1409.3600", "abstract": "We revisit the selection problem, namely that of computing the $i$th order statistic of $n$ given elements, in particular the classic deterministic algorithm by grouping and partition due to Blum, Floyd, Pratt, Rivest, and Tarjan (1973). Whereas the original algorithm uses groups of odd size at least $5$ and runs in linear time, it has been perpetuated in the literature that using smaller group sizes will force the worst-case running time to become superlinear, namely $\\Omega(n \\log{n})$. We first point out that the usual arguments found in the literature justifying the superlinear worst-case running time fall short of proving this claim. We further prove that it is possible to use group size smaller than $5$ while maintaining the worst case linear running time. To this end we introduce three simple variants of the classic algorithm, the repeated step algorithm, the shifting target algorithm, and the hyperpair algorithm, all running in linear time.", "subjects": "Data Structures and Algorithms (cs.DS); Combinatorics (math.CO)", "title": "Selection Algorithms with Small Groups", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9664104904802132, "lm_q2_score": 0.8459424373085146, "lm_q1q2_score": 0.8175276457573486 }
https://arxiv.org/abs/1611.02455
On the maximum dual volume of a canonical Fano polytope
We give an upper bound on the volume vol(P*) of a polytope P* dual to a d-dimensional lattice polytope P with exactly one interior lattice point, in each dimension d. This bound, expressed in terms of the Sylvester sequence, is sharp, and is achieved by the dual to a particular reflexive simplex. Our result implies a sharp upper bound on the volume of a d-dimensional reflexive polytope. Translated into toric geometry, this gives a sharp upper bound on the anti-canonical degree $(-K_X)^d$ of a d-dimensional toric Fano variety X with at worst canonical singularities.
\section{Introduction}\label{sec:intro} \subsection{Background and results} Let $N\cong\mathbb{Z}^d$ be a lattice of rank $d$. A convex polytope $P\subsetN_\R$, where $N_\R:=N\otimes_\mathbb{Z}\mathbb{R}\cong\mathbb{R}^d$, is called a \emph{lattice polytope} if the vertices $\operatorname{vert}(P)$ of $P$ are contained in $N$. Two lattice polytopes $P,Q\subsetN_\R$ are said to be \emph{unimodular equivalent} if there exists an affine lattice automorphism $\varphi\in\operatorname{GL}_d(\mathbb{Z})\ltimes\mathbb{Z}^d$ of $N$ such that $\varphi_\mathbb{R}(P)=Q$. Unless stated otherwise, we regard lattice polytopes as being defined only up to unimodular equivalence. Let $P\subsetN_\R$ be a lattice polytope of dimension $d$ (that is, $P$ is of maximum dimension in $N_\R$) containing exactly one lattice point in its (strict) interior,~i.e.\ $\abs{\operatorname{int}(P)\cap N}=1$. We can assume that this interior point is the origin $\boldsymbol{0}\in N$. For reasons that are explained in~\S\ref{subsec:alg_geom} below, we call $P$ a \emph{canonical Fano polytope}. As a consequence of results by Hensley~\cite[Theorem~3.6]{Hen83} and Lagarias--Ziegler~\cite[Theorem~2]{LZ91}, there are finitely many canonical Fano polytopes (up to unimodular equivalence) in each dimension $d$. Canonical Fano polytopes in dimensions $d\leq 3$ have been classified~\cite{Kas10}, and we find that $\operatorname{vol}(P)\leq12$. For $d\geq 4$ it is conjectured that the volume of a $d$-dimensional canonical Fano polytope is bounded by \begin{equation}\label{eq:vol_P_bound} \operatorname{vol}(P)\leq\frac{1}{d!}2(s_d-1)^2, \end{equation} where $s_i$ denotes the $i$-th term of the \emph{Sylvester sequence}: \[ s_1:=2,\quad s_{i+1}:=s_1\cdots s_i+1\text{ for } i\in\mathbb{Z}_{\ge 1}. \] Moreover, the case of equality in~\eqref{eq:vol_P_bound} is expected to be attained only by the canonical Fano simplex \[ R_{(d)}:=S_{(d)}-\sum_{i=1}^d e_i,\qquad\text{ where } S_{(d)} :=\operatorname{conv}\{\boldsymbol{0}, 2(s_d-1)e_d, s_{d-1} e_{d-1},\ldots, s_1 e_1\}. \] Here $\{e_1,\ldots,e_d\}$ is a basis of $N$. This conjecture is hinted at in~\cite{Reid82,ZPW82,LZ91}, explicitly stated in~\cite[Conjecture~1.7]{Nil07}, and proved by Averkov--Kr\"{u}mpelmann--Nill~\cite{AKN14} for the case when $P$ is a canonical Fano simplex. The conjecture remains open for a general canonical Fano polytope. The currently best upper bound on the volume of a canonical Fano polytope that is not a simplex is established in~\cite[Theorem~2.7]{AKN14} (improving upon a result by Pikhurko~\cite{Pik01}), however this is presumed to be far from sharp: \[ \operatorname{vol}(P)\leq (s_{d+1}-1)^d. \] Instead of bounding $\operatorname{vol}(P)$, it is also natural to consider the volume of the dual polytope $P^*$ (see~\S\ref{notationnow} for the definition of the dual polytope). The main result of this paper is: \begin{thm}\label{thm:main} Let $P\subsetN_\R$ be a $d$-dimensional canonical Fano polytope, where $d\geq 4$. Then \[ \operatorname{vol}(P^*)\leq\frac{1}{d!}2(s_d-1)^2, \] with equality if and only if $P=R_{(d)}^*$. \end{thm} In three dimensions, the expected bound $\operatorname{vol}(P^*)\leq 12$ is proved in~\cite[Theorem~4.6]{Kas10}. In this case, however, equality is obtained by the duals of two distinct simplices: \begin{equation}\label{eq:maximal_degree_dim_3} P_{1,1,1,3}=\operatorname{conv}\{e_1,e_2,e_3,-e_1-e_2-3e_3\}\qquad\text{ and }\qquad P_{1,1,4,6}=R_{(3)}^*. \end{equation} The analogue of Theorem~\ref{thm:main} is proved in~\cite[Theorem~2.5(b)]{AKN14} for $d$-dimensional canonical Fano simplices. Probably one of the most studied class of canonical Fano polytopes are the \emph{reflexive polytopes}, consisting of those $P\subsetN_\R$ such that the dual $P^*$ is also a canonical Fano polytope (for a brief survey see~\cite{KN13}). Note that $R_{(d)}$ is a reflexive simplex~\cite{Nil07}. An immediate consequence of Theorem~\ref{thm:main} is a proof of the conjectured inequality~\eqref{eq:vol_P_bound} in the case of reflexive polytopes: \begin{cor}\label{cor:vol_P_bound_reflexive} Let $P\subsetN_\R$ be a $d$-dimensional reflexive polytope, where $d\geq 4$. Then \[ \operatorname{vol}(P)\leq\frac{1}{d!}2(s_d-1)^2, \] with equality if and only if $P=R_{(d)}$. \end{cor} \noindent The analogue of Corollary~\ref{cor:vol_P_bound_reflexive} in the case of reflexive simplices is proved in~\cite[Theorem~A]{Nil07}. \subsection{Toric geometry and Fano varieties}\label{subsec:alg_geom} Canonical Fano polytopes arise naturally in algebraic geometry. To each $d$-dimensional canonical Fano polytope $P\subsetN_\R$ we can associate a $d$-dimensional projective toric variety $X_P$ whose fan is given by the cones in $N_\R$ spanning the faces of $P$ (here we require that the unique interior point of $P$ is taken to be the origin $\boldsymbol{0}$ of $N$). This variety is Fano -- recall that a variety $X$ is Fano if its anti-canonical divisor $-K_X$ is ample -- and has at worst canonical singularities. In fact this construction is reversible, and there exists a one-to-one correspondence between (unimodular equivalence classes of) canonical Fano polytopes and (isomorphism classes of) toric Fano varieties with at worst canonical singularities. For details on canonical singularities and their importance in algebraic geometry, see~\cite{Rei87}; for details on toric geometry, see~\cite{Dan78}; and for additional background material see the survey~\cite{KN13}. The classification of Fano varieties is a long-standing open problem. An important advance would be to bound the degree $(-K_X)^d$. In the case when $X$ is non-singular the bound \begin{equation}\label{eq:KMM} (-K_X)^d\leq\left(3(2^d - 1)(d + 1)^{(d + 1)(2^d - 1)}\right)^d \end{equation} was established by Koll\'{a}r--Miyaoka--Mori~\cite{KMM92}, although this is almost certainly not sharp. Very little is known when $X$ has canonical singularities, however Prokhorov~\cite{Pro05} proved that if $X$ is a three-dimensional Fano with Gorenstein canonical singularities then the degree is bounded by $(-K_X)^3\leq 72$. In this case the maximum degree is obtained by the two weighted projective spaces $\mathbb{P}(1,1,1,3)$ and $\mathbb{P}(1,1,4,6)$, and these two toric varieties correspond to the two canonical Fano simplices in~\eqref{eq:maximal_degree_dim_3}. It is tempting to conjecture that, in higher dimensions, the maximum degree is obtained by a toric Fano variety. Recalling that $(-K_{X_P})^d=d!\operatorname{vol}(P^*)$, Theorem~\ref{thm:main} provides a sharp bound on the degree when $X$ is toric: \begin{cor}\label{cor:bound_canonical_Fano_degree} Let $X$ be a $d$-dimensional toric Fano variety with at worst canonical singularities, where $d\ge 4$. Then \begin{equation}\label{eq:canonical_degree_bound} (-K_X)^d\leq 2 (s_d-1)^2, \end{equation} with equality if and only if $X$ is isomorphic to the weighted projective space \[ \mathbb{P}\left(1,1,2 (s_d-1) / s_{d-1},\ldots, 2 (s_d-1)/s_1\right). \] \end{cor} \noindent This extends~\cite[Theorem~A]{Nil07} and~\cite[Theorem~2.11]{AKN14}, where analogous results are stated when $X$ is a Gorenstein fake weighted projective space, and when $X$ is a fake weighted projective space with at worst canonical singularities, respectively. Corollary~\ref{cor:bound_canonical_Fano_degree} also generalises the three-dimensional bound of~\cite[Theorem~4.6]{Kas10}. Finally, Corollary~\ref{cor:bound_canonical_Fano_degree} also has implications for current attempts to classify non-singular Fano varieties via Mirror Symmetry~\cite{CCGK16}. Here the hope is that a non-singular Fano variety $X$ with $-K_X$ very ample has a $\mathbb{Q}$-Gorenstein deformation to a Gorenstein canonical toric Fano variety $X_P$. Since this deformation would leave the degree unchanged, so the bound of Corollary~\ref{cor:bound_canonical_Fano_degree} would apply to $X$. It is interesting to note that, in this case, the bound~\eqref{eq:canonical_degree_bound} is significantly smaller that the bound~\eqref{eq:KMM} of Koll\'{a}r--Miyaoka--Mori. \subsection{Overview of the proof} Our strategy to prove Theorem~\ref{thm:main} is as follows. In~\S\ref{sec:2} we reduce the problem to canonical Fano polytopes satisfying some minimality condition. We observe that such polytopes admit a decomposition into canonical Fano simplices (following~\cite{Kas10}; compare also with the decomposition used in~\cite{KS97}), for which the statement is already known~\cite{AKN14}. In~\S\ref{sec:monotonicity} we use this decomposition, together with the monotonicity of the normalised volume, to prove Theorem~\ref{thm:main} in the majority of cases (Corollary~\ref{cor:almostallcases}). Finally, the remaining cases are proved in~\S\ref{sec:final} using a mixture of integration techniques (developed in~\S\S\ref{sec:slicing}--\ref{sec:integration}) and explicit classification. \subsection{Notation and terminology}\label{notationnow} Let $P\subsetN_\R$ be a lattice polytope of maximum dimension in a rank $d$ lattice $N\cong\mathbb{Z}^d$, and let $M:=\operatorname{Hom}_{\mathbb{Z}}(N,\mathbb{Z})\cong\mathbb{Z}^d$ be the lattice dual to $N$. The \emph{dual} (or \emph{polar}) polyhedron of $P$ is: \[ P^*:=\{y\in M_\mathbb{R}\,\colon\langle y,x\rangle\geq-1\text{ for every }x\in P\}. \] If $\boldsymbol{0}\in P$ then $P^*$ is a convex polytope, although typically $P^*$ has rational vertices, and so is not a lattice polytope. Let $P$ and $Q$ be two maximum-dimensional polytopes in $(N_P)_\mathbb{R}\cong\mathbb{R}^p$ and $(N_Q)_\mathbb{R}\cong\mathbb{R}^q$, respectively. Suppose that $P$ and $Q$ contain the origin $\boldsymbol{0}_{P}\in N_P$ and $\boldsymbol{0}_{Q}\in N_Q$ of their respective ambient space. The \emph{free sum} (or \emph{direct sum}) is the maximum-dimensional polytope \[ P\oplus Q:=\operatorname{conv}((P\times\{\boldsymbol{0}_Q\})\cup(\{\boldsymbol{0}_P\}\times Q))\subset\mathbb{R}^{p+q}. \] The \emph{product} is the polytope \[ P\times Q:=\{(x_p,x_q)\,\colon x_p\in P,\, x_q\in Q\}\subset\mathbb{R}^{p+q}. \] Free sums and products of polytopes are related, via duality, by: \[ (P\oplus Q )^*=P^*\times Q^*. \] On the affine hull $\operatorname{aff}(P)$ there exists a volume form, called the \emph{relative lattice volume}, that is normalised by setting the volume of a fundamental parallelepiped of $\operatorname{aff}_\mathbb{Z}(P)$ equal to $1$. We denote the relative lattice volume of $P$ by $\operatorname{vol}_N(P)$. The volume $\operatorname{Vol}_N(P):=\dim(P)!\operatorname{vol}_N(P)$ is often called the \emph{normalised lattice volume} of P. If $N'\subseteq N$ is a sublattice of $N$ then, for $S\subseteq\operatorname{lin}(N')$, we have $\operatorname{vol}_{N'}(S)\leq\operatorname{vol}_N(S)$. If in addition we have that $N'\to N$ splits over $\mathbb{Z}$, then $\operatorname{vol}_{N'}(S)=\operatorname{vol}_N(S)$. \section{Decomposition of minimal polytopes}\label{sec:2} The case of canonical Fano simplices is already considered in~\cite{AKN14}. Our focus is on the case when $P$ is not a simplex. Notice that if $P\subsetneq Q$ then $Q^*\subsetneq P^*$, and hence $\operatorname{vol}(Q^*)<\operatorname{vol}(P^*)$. It is therefore sufficient to prove Theorem~\ref{thm:main} for ``small'' polytopes $P$; that is, for the \emph{minimal} canonical Fano polytopes: \begin{defn}[\!\!{\cite[Definition~2.2]{Kas10}}] A $d$-dimensional canonical Fano polytope $P\subsetN_\R$ is \emph{minimal} if, for each vertex of $P$, the polytope obtained by removing this vertex is not a $d$-dimensional canonical Fano polytope; that is, if $\operatorname{conv}(P\cap N\setminus\{v\})$ is not a $d$-dimensional canonical Fano polytope, for each $v\in\operatorname{vert}(P)$. \end{defn} \noindent Each canonical Fano polytope $Q$ can, via successive removal of vertices, be reduced to a minimal polytope $P\subset Q$. Of course $P$ need not be uniquely determined. Minimal canonical Fano polytopes admit a decomposition in terms of lower-dimensional minimal canonical Fano simplices: \begin{prop}[\!\!{\cite[Proposition~3.2]{Kas10}}]\label{prop:Kasp} Let $P$ be a minimal canonical Fano $d$-polytope that is not a simplex. Then there exists a minimal canonical Fano $k$-simplex $S$ contained in $P$ with $\operatorname{vert}(S)\subset\operatorname{vert}(P)$, for some $1\le k<d$. For any such $S$ there exists a minimal canonical Fano $(d-k+s)$-polytope $P'$ with $\operatorname{vert}(P')\subset\operatorname{vert}(P)$ such that $P=\operatorname{conv}(S\cup P')$, $s=\abs{\operatorname{vert}(S)\cap\operatorname{vert}(P')}$, and $0\leq s < k$. \end{prop} \noindent For brevity we write ``$d$-polytope'' rather than ``polytope of dimension $d$'', and ``$k$-simplex'' rather than ``simplex of dimension $k$''. \begin{cor}\label{cor:Kasp} Let $P$ be a minimal canonical Fano $d$-polytope that is not a simplex. Then, for some $2\le t\le d$, there exist minimal canonical Fano simplices $S_1,\ldots,S_t$ such that $P=\operatorname{conv}(S_1\cup\ldots\cup S_t)$, where $\dim(S_i)=d_i\geq 1$ and $\operatorname{vert}(S_i)\subset\operatorname{vert}(P)$, for each $1\leq i\leq t$. Set $r_1:=0$ and, for each $2\leq i\leq t$, set $r_i:=\abs{\operatorname{vert}(S_i)\cap\operatorname{vert}(P^{(i-1)})}$, where $P^{(i-1)}:=\operatorname{conv}(S_1\cup\ldots\cup S_{i-1})$. Then: \begin{align} \label{eq:dr} d_1+\cdots+d_t &= d + r,&\quad\text{ where }r:=r_1 +\cdots + r_t;\\ \label{eq:di} r_i < d_i &\leq d -t +1,&\quad\text{ for each }1\leq i\leq t;\\ \label{vertices} \abs{\operatorname{vert}(P)}&=d+t.& \end{align} \end{cor} \noindent An example of this decomposition is illustrated in Figure~\ref{fig:1}. \begin{proof} We apply Proposition~\ref{prop:Kasp} iteratively, at each step choosing $S$ to be of smallest possible dimension. Thus $P$ can be written as $P=\operatorname{conv}(S_1\cup\ldots\cup S_t)$ for some $t\geq 1$, where the $S_i$ are minimal canonical Fano simplices of dimension $d_i\ge 1$ with $\operatorname{vert}(S_i)\subseteq\operatorname{vert}(P)$ having $r_i$ common vertices with $P^{(i-1)}$, such that $d_t\le d_{t-1}\le\cdots\le d_1$. The case $P^{(0)}$ is taken to be the empty set, giving $r_1=0$. At each step, the dimension of $P^{(i)}$ can be obtained from Proposition~\ref{prop:Kasp}: $\dim(P^{(i)})=\dim(P^{(i-1)})+\dim(S_i)-r_i$. Hence $d=\sum_{i=1}^t(d_i-r_i)$, and so~\eqref{eq:dr} holds. Once again using Proposition~\ref{prop:Kasp}, since $\dim(S_i)>r_i$, so $\dim(P^{(i)})\geq\dim(P^{(i-1)})+1$. It follows that $t\leq d$ and so $d_1\leq d-t+1$. Hence our choice of simplices implies~\eqref{eq:di}. Finally, the number of vertices of $P^{(i)}$ is $\abs{\operatorname{vert}(P^{(i-1)})}+\abs{\operatorname{vert}(S_i)}-r_i$. This implies that $\abs{\operatorname{vert}(P)}=\sum_{i=1}^t(d_i+1)-r$, and from~\eqref{eq:dr} we deduce that equation~\eqref{vertices} holds. \end{proof} \noindent Notice that equality~\eqref{vertices}, combined with the bound $t\leq d$, implies that a minimal canonical Fano polytope $P$ satisfies $\abs{\operatorname{vert}(P)}\leq 2 d$ (this is known as \emph{Steinitz's inequality}). \begin{figure}[tb] \centering \begin{tikzpicture}[scale=0.8] \draw[fill,black!15] (2,-1)--(3.5,0)--(2.5,2.5)--cycle; \draw[fill,black!35] (2.8,0.4)--(5.5,0.3)--(2.5,2.5)--cycle; \draw[fill,black!35] (0,0.5)--(2.2,0.4)--(2.5,2.5)--cycle; \draw[thick] (0,0.5)--(2,-1)--(5.5,0.3); \draw[dashed] (2,-1)--(3.5,0)--(5.5,0.3); \draw[thick] (2.5,2.5)--(2,-1); \draw[thick] (2.5,2.5)--(0,0.5); \draw[dashed] (2.5,2.5)--(3.5,0)--(0,0.5); \draw[thick] (2.5,2.5)--(5.5,0.3); \draw[fill=black] (2.73,1) circle (0.2em) node[below left]{$\boldsymbol{0}$}; \draw (1.3,2.2) node{$P$}; \draw (1.4,0.9) node{$S_1$}; \draw (2.5,-0.25) node{$S_2$}; \draw (2.5,2.5) node[above]{$v$}; \end{tikzpicture} \caption{An example of a three-dimensional minimal canonical Fano polytope $P$ which decomposes into two canonical Fano simplices $S_1$ and $S_2$ sharing a common vertex $v$. In the notation of Corollary~\ref{cor:Kasp}, $d=3$, $t=2$, $d_1=d_2=2$, and $r_2=1$.}\label{fig:1} \end{figure} \section{Bounding the volume of $P^*$ via monotonicity of the normalised volume}\label{sec:monotonicity} As noted above, it is sufficient to prove Theorem~\ref{thm:main} for minimal canonical Fano polytopes that are not simplices. Let $P\subsetN_\R$ be such a polytope of dimension $d\geq 4$. Fix a decomposition of $P$, and use the notation $t,S_i,d_i,r_i,r$ as defined in Corollary~\ref{cor:Kasp}. In this section we prove Theorem~\ref{thm:main} for the majority of decompositions. The decompositions \emph{not} addressed in this section, and whose proof is the focus of~\S\S\ref{sec:slicing}--\ref{sec:integration} below, are listed in Corollary~\ref{cor:almostallcases}. \begin{cor}\label{cor:almostallcases} In order to prove Theorem~\ref{thm:main} it is enough to verify that the inequality \[ \operatorname{vol}(P^*)\leq\frac{1}{d!}2(s_d-1)^2 \] holds for all minimal canonical Fano polytopes $P\subsetN_\R$ of dimension $d\ge 4$ whose decomposition into minimal canonical Fano simplices falls into one of the following five cases: \begin{enumerate} \item $t=2$ and $d_1=d_2=d-1$; or \item $t=2$, $d=4$, $d_1=3$, and $d_2=2$; or \item $t=2$, $d=5$, $d_1=4$, and $d_2=3$; or \item $t=3$, $d=4$, and $d_1=d_2=d_3=2$; or \item $t=3$, $d=5$, and $d_1=d_2=d_3=3$. \end{enumerate} \end{cor} \noindent In order to prove Corollary~\ref{cor:almostallcases} we use the monotonicity of the normalised volume. Let $N_i:=\operatorname{lin}_{\mathbb{R}}(S_i)\cap N$ be the sublattice of lattice points in the linear hull of $S_i$ (recall that $\boldsymbol{0}\in\operatorname{int}(S_i)$, so this really is a sublattice), for each $1\leq i\le t$. Define the map \[ \varphi\colon N_1\oplus\cdots\oplus N_t\to N,\qquad(x_1,\ldots,x_t)\mapsto\sum_{i=1}^{t} x_i . \] Notice that $\varphi$ may not be surjective, however since its image has the same rank as $N$, the extension $\varphi_\mathbb{R}$ of $\varphi$ to a map of vector spaces is surjective. Moreover, $\varphi_\mathbb{R}$ gives the following representation of $P$: \[ P=\varphi_\mathbb{R}(S_1\oplus\cdots\oplus S_t). \] Let $M,M_1,\ldots,M_t$ denote the lattices dual to $N,N_1,\ldots,N_t$, respectively. The map $\varphi_\mathbb{R}^*$ dual to $\varphi_\mathbb{R}$ is an injection, and in particular \[ P^*\cong\varphi^*_\mathbb{R}(P^*)\subset(S_1\oplus\cdots\oplus S_t)^*=S_1^*\times\cdots\times S_t^*, \] where $M$ is naturally embedded via $\varphi^*$ into $M_1\oplus\cdots\oplus M_t$. This situation will be studied in more detail in~\S\ref{sec:slicing}. Using the monotonicity of the normalised volume, finding an upper bound for the normalised volume of $S_1^*\times\cdots\times S_t^*$ yields an upper bound for the normalised volume of $P^*$. Specifically, we know that: \begin{align}\label{eq:geq3} \begin{split} \operatorname{Vol}_M(P^*)&\leq\operatorname{Vol}_{M_1\oplus\cdots\oplus M_t}(S_1^*\times\cdots\times S_t^*)\\ &=(d_1+\cdots+d_t)!\operatorname{vol}_{M_1\oplus\cdots\oplus M_t}(S_1^*\times\cdots\times S_t^*)\\ &=(d_1+\cdots+d_t)!\prod_{i=1}^t\operatorname{vol}_{M_i}(S_i^*)\\ &=\frac{(d_1+\cdots+d_t)!}{d_1!\cdots d_t!}\prod_{i=1}^t\operatorname{Vol}_{M_i}(S_i^*). \end{split} \end{align} The normalised volume of $S_i^*$ is bounded from above (see~\cite{Kas10} and~\cite[Theorem~2.5(b)]{AKN14}): \[ \operatorname{Vol}_{M_i}(S^*_i)\leq B_i,\qquad\text{with}\; B_i:=\left\{ \begin{array}{ll} 9 &\text{if}\;d_i=2,\\ 2(s_{d_i}-1)^2 &\text{if}\;d_i\neq 2. \end{array} \right. \] Hence inequality~\eqref{eq:geq3} becomes: \[ \operatorname{Vol}_M(P^*)\leq\frac{(d_1+\cdots+d_t)!}{d_1!\cdots d_t!}\,\prod_{i=1}^t B_i. \] At this point, Theorem~\ref{thm:main} would follow from \begin{equation}\label{eq:provet3} \frac{(d_1+\cdots+d_t)!}{d_1!\cdots d_t!}\,\prod_{i=1}^t B_i< B_d. \end{equation} Unfortunately inequality~\eqref{eq:provet3} does not always hold: for example, it fails when $t=2$ and $d_1=d_2=d-1$, for any $d\geq 3$. Nevertheless, this technique is sufficient to prove Theorem~\ref{thm:main} for a large number of cases: \begin{lem}\label{lem:provet3_cases} Inequality~\eqref{eq:provet3} -- and therefore Theorem~\ref{thm:main} -- holds whenever: \begin{enumerate} \item\label{item:first_case} $t\geq 3$, with the exception of the following six cases: \begin{enumerate} \item $t=3$, $d=4$, and $d_1=d_2=d_3=2$; or \item $t=3$, $d=5$, and $d_1=d_2=d_3=3$; or \item $t=3$, $d=4$, $d_1=d_2=2$, and $d_3=1$; or \item $t=3$, $d=5$, $d_1=d_2=3$, and $d_3=2$; or \item $t=3$, $d=6$, and $d_1=d_2=d_3=4$; or \item $t=4$, $d=5$, and $d_1=d_2=d_3=d_4=2$; \end{enumerate} \item\label{item:second_case} $t=2$, with the exceptions of the following three cases: \begin{enumerate} \item $d_1=d_2=d-1$; or \item $d=4$, $d_1=3$, and $d_2=2$; or \item $d=5$, $d_1=4$, and $d_2=3$. \end{enumerate} \end{enumerate} \end{lem} \begin{proof} We prove~\eqref{item:first_case} and~\eqref{item:second_case} separately, but by the same general technique: first we show that the statement is true for large values of $d$; then we check the finite number of remaining values. \begin{enumerate} \item[\eqref{item:first_case}] Since the quantity \[ \frac{(d_1 +\cdots + d_t)!}{d_1!\cdots d_t!}\,\prod_{i=1}^t B_i \] increases as the $d_i$ increase, by~\eqref{eq:di} it is enough to prove inequality~\eqref{eq:provet3} when $d_i=d-t+1$ for all $i$. That is, it is sufficient to show that \begin{equation}\label{eq:naive} \frac{(t(d-t+1))!}{(d-t+1)!^t}\, (B_{d-t+1})^{t} < B_d. \end{equation} From $n!\leq 2\cdot 2^2\cdots 2^{n-1} = 2^{n(n-1)/2}$ (which is strict when $n\ge 3$) we obtain: \[ \frac{(t(d-t+1))!}{(d-t+1)!^t}\le (t(d-t+1))! < 2^{\frac{1}{2}t(d-t+1)(t(d-t+1)-1)}. \] Therefore, if the inequality \begin{equation}\label{eq:toprove} 2^{\frac{1}{2}t(d-t+1)(t(d-t+1)-1)}2^t(B_{d-t+1})^{t}\leq B_d \end{equation} holds, so too does inequality~\eqref{eq:naive}. To prove~\eqref{eq:toprove} we make use of the well-known description due to Aho--Sloane~\cite[Example~2.5]{AS73} and Vardi of the Sylvester sequence in terms of the constant $c\approx 1.2640847353\dots$: \[ s_n=\left\lfloor c^{2^n}+\frac{1}{2}\right\rfloor. \] Notice that $B_d=2(s_d-1)^2>(s_d + 1)^2$ whenever $d\geq 3$. Since $s_d+1>c^{2^d}$, the right-hand side of~\eqref{eq:toprove} is bounded from below: \[ B_d=2(s_d-1)^2 > (s_d + 1)^2 > c^{2^{d+1}}. \] Moreover $B_{d-t+1}/2 < c^{2^{d-t+2}}$. Since $c^3 > 2$, the left-hand side of~\eqref{eq:toprove} is bounded from above: \[ 2^{\frac{1}{2}t(d-t+1)(t(d-t+1)-1)} 2^t\left(\frac{B_{d-t+1}}{2}\right)^{t} <c^{\frac{3}{2}t(d-t+1)(t(d-t+1)-1)} c^{3t} c^{2^{d-t+2}t}. \] We shall show that $c^{\frac{3}{2}t(d-t+1)(t(d-t+1)-1)} c^{3t} c^{2^{d-t+2}t}\leq c^{2^{d+1}}$, from which we conclude that inequality~\eqref{eq:toprove} holds. Taking $\log_c$, we have to verify that the inequality \begin{equation*} \frac{3}{2}t(d-t+1)(t(d-t+1)-1) + 3t + 2^{d-t+2} t\leq 2^{d+1} \end{equation*} is satisfied. Rewrite this inequality as: \begin{equation}\label{eq:exp} 3t(d-t+1)(t(d-t+1)-1) + 6t\leq 2^{d+2}\left(1-\frac{t}{2^{t-1}}\right). \end{equation} Since $t\ge 3$, by setting $t=3$ in the right-most factor it is enough to prove that: \begin{equation*} 3t(d-t+1)(t(d-t+1)-1) + 6t\leq 2^{d}. \end{equation*} Since $t(d-t+1)$ is maximised when $t=(d+1)/2$, and since $6t\leq 6d$, the above inequality is valid when \[ \frac{3(d+1)}{2}\left(d-\frac{d+1}{2}+1\right)\left(\frac{d+1}{2}\left(d-\frac{d+1}{2}+1\right)-1\right) + 6d\leq 2^{d}. \] This holds when $d\geq 13$. Recalling that $d$ bounds the quantities $t,d_1,\ldots,d_t$, we are left with finitely many cases to verify. Inequality~\eqref{eq:provet3} holds in all but six cases, as listed in the statement. \item[\eqref{item:second_case}] By the same monotonicity argument used at the beginning of the previous case, we choose $d_1$ and $d_2$ as great as possible, i.e.~we fix $d_1=d-1$ and $d_2=d-2$ (we noted above that inequality~\eqref{eq:provet3} is not satisfied when $d_1=d_2=d-1$). Inequality~\eqref{eq:provet3} becomes \begin{equation}\label{eq:eq2} \frac{(2d-3)!}{(d-2)! (d-1)!}\, B_{d-2} B_{d-1} < B_d. \end{equation} Proceeding as above, we reduce the problem to proving the inequality \[ 3(2d^2-7d+8)+2^{d-1}+2^d\le 2^{d+1}. \] This holds when $d\geq 10$. Removing the assumptions $d_1=d-1$ and $d_2=d-2$ on $d_1$ and $d_2$, the finitely many cases for $4\leq d\leq 9$ can be directly verified against inequality~\eqref{eq:provet3}. We find the exceptional cases listed in the statement of the Lemma.\qedhere \end{enumerate} \end{proof} \begin{proof}[Proof of Corollary~\ref{cor:almostallcases}] By Lemma~\ref{lem:provet3_cases} we need to show that proving Theorem~\ref{thm:main} for all deompositions listed in the statement of Corollary~\ref{cor:almostallcases} also proves it in the four cases: \begin{enumerate} \item $t=3$, $d=4$, $d_1=d_2=2$, and $d_3=1$; or \item $t=3$, $d=5$, $d_1=d_2=3$, and $d_3=2$; or \item $t=3$, $d=6$, and $d_1=d_2=d_3=4$; or \item $t=4$, $d=5$, and $d_1=d_2=d_3=d_4=2$. \end{enumerate} In each cases we have that either $t=3$ or $t=4$. By Corollary~\ref{cor:Kasp} we can express $P$ as $P = P'\cup S_t$, where $P' = S_1\cup\ldots\cup S_{t-1}$ is a minimal polytope of dimension $d'$ decomposed into $t'=t-1$ minimal simplices. Note that in all four cases $d'=d-1$. We now proceed exactly as in the first part of this section. Let $N':=\operatorname{lin}_\mathbb{R}(P')\cap N$ be the sublattice of $N$ of lattice points in the linear hull of $P'$. We define the map $\varphi'\colon N'\oplus N_{t}\to N$ by $(x_1,x_2)\mapsto x_1 + x_2$, whose extension $\varphi_\mathbb{R}$ to a map of vector spaces is surjective and gives the following representation of $P$: \[ P=\varphi_\mathbb{R}(P'\oplus S_{t}). \] Let $M'$ denote the lattice dual to $N'$. The map $(\varphi')_\mathbb{R}^*$ dual to $(\varphi')_\mathbb{R}$ is an injection, and in particular \[ P^*\cong(\varphi')_\mathbb{R}^*(P^*)\subset(P'\oplus S_{t})^*=(P')^*\times S_{t}^*. \] As in~\eqref{eq:geq3}, by the monotonicity of the normalised volume, \begin{align}\label{eq:geq3cor} \begin{split} \operatorname{Vol}_M(P^*)&\leq\frac{(d'+d_{t})!}{d'!d_{t}!}\operatorname{Vol}_{M'}((P')^*)\operatorname{Vol}_{M_{t}}(S_{t}^*). \end{split} \end{align} By our assumption and Lemma~\ref{lem:provet3_cases}, Theorem~\ref{thm:main} holds for $t'=2$ and for $t'=3$, $d'=4$, $d_1=d_2=d_3=2$. Hence, in all four cases, Theorem~\ref{thm:main} holds for $P'$; i.e.~$\operatorname{Vol}_{M'}((P')^*) < B_{d-1}$. Since $\operatorname{Vol}_{M_{t}}(S_{t}^*)\leq B_{d_{t}}$ and $d'=d-1$, \[ \operatorname{Vol}_M(P^*) <\frac{(d-1+d_{t})!}{(d-1)!d_{t}!} B_{d-1} B_{d_{t}}. \] Hence it is enough to prove that \[ \frac{(d-1+d_{t})!}{(d-1)!d_{t}!} B_{d-1} B_{d_{t}} < B_d. \] This inequality can be directly checked in all four cases. \end{proof} \section{Slicing minimal polytopes}\label{sec:slicing} We now develop the foundations for a finer technique that we use in~\S\ref{sec:final} to help prove the remaining cases of Theorem~\ref{thm:main}. In particular, we shall explain how minimal polytopes can be described as a particular union of slices which are products of slices of simplices (see Figure~\ref{fig:slice}). Using this construction, in~\S\ref{sec:integration} we give a better estimate of the dual volume via integration. \subsection{Embedding the dual polytope}\label{subsec:slicing1} As above, we are in the setup of Corollary~\ref{cor:Kasp}: $P\subsetN_\R$ is a $d$-dimensional minimal canonical Fano polytope decomposed into minimal canonical Fano simplices $S_1,\ldots,S_t$, for some $t\geq 2$. We define \[ \mathcal{V}:=\!\!\bigcup_{1\leq i_1 < i_2\leq t}\!\!\operatorname{vert}(S_{i_1})\cap\operatorname{vert}(S_{i_2}) \] to be the set of those vertices of $P$ which occur multiple times amongst the vertices of the $S_i$, and define $\mathcal{V}_i:=\mathcal{V}\cap\operatorname{vert}(S_{i})$. For example, in Figure~\ref{fig:1} we have $\mathcal{V}=\mathcal{V}_1=\mathcal{V}_2=\{v\}$. \begin{figure}[tb] \centering \begin{minipage}[t]{0.3\textwidth} \begin{tikzpicture}[scale=0.6] \draw[dashed] (-1,0)--(2,1)--(1.7,2.8); \draw[dashed] (2,1)--(5,1); \draw[fill,black!20] (0.25,1.3)--(2.36,1.3)--(3.94,1.86)--(1.86,1.86); \draw[thick,black] (1.7,2.8)--(2.8,2.8); \draw[thick,black] (-1,0)--(2,0); \draw[thick,black] (2,0)--(5,1)--(2.8,2.8)--cycle; \draw[thick,black] (1.7,2.8)--(-1,0); \draw[fill=black] (2.15,1.6) circle (0.2em) node[above]{$\boldsymbol{0}$}; \end{tikzpicture} \end{minipage} \begin{minipage}[t]{0.25\textwidth} \begin{tikzpicture}[scale=0.6] \draw[very thick,black!50] (1.07,1.58)--(3.17,1.58); \draw[thick,black!30] (3.5,0.5)--(2.4,4.1)--(0.5,0.5)--cycle; \draw[very thick,black!50] (3.5,0.5)--(0.5,0.5); \draw[fill=black] (2.15,1.6) circle (0.2em) node[above]{$\boldsymbol{0}$}; \draw (4.1,0.8) node[below]{{\tiny $F^*_1$}}; \draw (3,1.6) node[right]{{\tiny $H_{1,\boldsymbol{0}}$}}; \end{tikzpicture} \end{minipage} \begin{minipage}[t]{0.25\textwidth} \begin{tikzpicture}[scale=0.6] \draw[very thick,black!50] (1.30,1.3)--(2.88,1.86); \draw[thick,black!30] (3.5,1)--(2.25,2.8)--(0.5,0)--cycle; \draw[very thick,black!50] (3.5,1)--(0.5,0); \draw[fill=black] (2.15,1.6) circle (0.2em) node[above]{$\boldsymbol{0}$}; \draw (1.2,0.7) node[right]{{\tiny $F^*_2$}}; \draw (0.6,1.6) node[right]{{\tiny $H_{2,\boldsymbol{0}}$}}; \end{tikzpicture} \end{minipage} \caption{The dual $P^*$ of the polytope $P$ from Figure~\ref{fig:1}, together with the dual triangles $(S'_1)^*\subset(M'_1)_\mathbb{R}$ and $(S'_2)^*\subset(M'_2)_\mathbb{R}$. In the first picture, the grey slice is $H_{1,\boldsymbol{0}}\times H_{2,\boldsymbol{0}}$. We refer to~\S\ref{subsec:slicing2} for the precise definitions.\label{fig:slice}} \end{figure} It will be convenient to coarsen the lattice $N$. We note that coarsening the ambient lattice $N$ to a lattice $N'$ is an assumption we can make. Indeed, if $P_M^*$ and $P_{M'}^*$ denote the duals of $P$ with respect to the lattices $M=N^*$ and $M'=(N')^*$, respectively, then the volume of $P_{M'}^*$ is equal to the volume of $P_M^*$ multiplied by the index of $M'$ as a subgroup of $M$ (which is a positive integer). Let $N'_i$ denote some sublattice of $N_i=\operatorname{lin}_{\mathbb{R}}(S_i)\cap N$ of same rank $d_i$ with $\mathcal{V}_i\subset N'_i$ (a specific choice of $N'_i$ will be given in~\S\ref{subsec:slicing2}). Notice that $S_i$ may no longer be a \emph{lattice} simplex with respect to $N'_i$. Therefore, in order to avoid any confusion, we denote by $S'_i\subseteq (N'_i)_\mathbb{R}=(N_i)_\mathbb{R}$ the \emph{rational} simplex with vertices $\operatorname{vert}(S_i)$ with respect to the lattice $N'_i$. Now, by possibly coarsening the lattice $N$ we may suppose that $N$ is the image of the lattice $N'_1\oplus\cdots\oplus N'_t$ via the map \begin{equation}\label{phimap} \begin{array}{r@{\ }c@{\ }l} \varphi\colon N'_1\oplus\cdots\oplus N'_t &\to&N\\ (x_1,\ldots,x_t) &\mapsto&{\displaystyle\sum_{i=1}^{t}x_i}. \end{array} \end{equation} Hence we can assume that this map is surjective. Notice that the polytope $P$ may no longer be a lattice polytope with respect to this ambient lattice. We extend the map $\varphi$ to the map of real vector spaces $\varphi_\mathbb{R}\colon(N'_1)_\mathbb{R}\oplus\cdots\oplus(N'_t)_\mathbb{R}\toN_\R$. As in the previous section we can describe $P$ as \[ P =\varphi_\mathbb{R} (S'_1\oplus\cdots\oplus S'_t). \] By definition $\varphi$ is a surjective map, so we have the exact sequence \[ 0\to\ker\varphi\hookrightarrow N'_1\oplus\cdots\oplus N'_t\twoheadrightarrow N\to 0, \] which splits over $\mathbb{Z}$. From~\eqref{eq:dr} we have that $N'_1\oplus\cdots\oplus N'_t$ splits in parts of rank $d$ and $r$. As a consequence, the dual sequence \[ 0\to M\hookrightarrow M'_1\oplus\cdots\oplus M'_t\twoheadrightarrow (\ker\varphi)^*\to 0 \] is exact and splits too. Here we used the notation $M'_1,\ldots,M'_t$ for the dual lattices of $N'_1,\ldots,N'_t$, respectively. Let $(\ker\varphi)^\perp$ denote the elements of $M'_1\oplus\cdots\oplus M'_t$ vanishing on $\ker\varphi$. By the exactness of the dual sequence, $\varphi^*(M)=(\ker\varphi)^\perp$; that is, the lattices $M$ and $(\ker\varphi)^\perp$ are isomorphic via $\varphi^*$. In particular, $(\ker\varphi)^\perp = (M'_1\oplus\cdots\oplus M'_t)\cap (\ker\varphi)^\perp_\mathbb{R}$ is a direct summand of $M'_1\oplus\cdots\oplus M'_t$ of rank $d$. By tensoring by $\mathbb{R}$ to extend the maps to the ambient real vector spaces, it follows that the following polytopes are isomorphic as rational polytopes with respect to their respective lattices: \begin{align}\label{eq:P*} \begin{split} P^* &\cong\varphi_\mathbb{R} ^* (P^*)\\ &=(S'_1\oplus\cdots\oplus S'_t)^*\cap (\ker\varphi)_\mathbb{R}^\perp\\ &=((S'_1)^*\times\cdots\times (S'_t)^*)\cap (\ker\varphi)_\mathbb{R}^\perp. \end{split} \end{align} We now describe a set of generators of $(\ker\varphi)_\mathbb{R}$. For this, let us identify $N'_i$ with the corresponding direct summand in $N'_1\oplus\cdots\oplus N'_t$. In this way, we can identify $v\in\operatorname{vert}(S'_i)$ with $e_{i,v}\in N'_1\oplus\cdots\oplus N'_t$, i.e., $(e_{i,v})_i = v\in N'_i$ and $(e_{i,v})_j =\boldsymbol{0}_{N'_j}$ for $j\not=i$. Recall that $\dim_\mathbb{R} (\ker\varphi)_\mathbb{R} = r$. Let $1\leq i_1 < i_2\leq t$, and $v\in\mathcal{V}_{i_1}\cap\mathcal{V}_{i_2}$. We denote by $w_{v,i_1,i_2}$ the element $e_{i_2,v}-e_{i_1,v}\in N'_1\oplus\cdots\oplus N'_t$. \begin{lem}\label{lem:generators} With notation as above, $\ker\varphi_\mathbb{R}$ is generated by the set \[ \Omega :=\{w_{v,i_1,i_2}\in N'_1\oplus\cdots\oplus N'_t\,\colon 1\leq i_1 < i_2\leq t,\, v\in\mathcal{V}_{i_1}\cap\mathcal{V}_{i_2}\}. \] \end{lem} \begin{proof} We prove that the subset \[ \Omega':=\left\{w_{v,i_1,i_2}\in\Omega\,\colon i_1=\max\{i\,\colon v\in\mathcal{V}_i, i<i_2\}\right\} \] of $\Omega$ is a basis of $\ker\varphi_\mathbb{R}$. Since for $2\le i\leq t$ we have $\abs{\{w_{v,i_1,i}\in\Omega'\,\colon i_2=i\}}=r_i$, this implies that $\abs{\Omega'}=\sum_{i=2}^t r_i = r$. Hence it is enough to prove that the elements of $\Omega'$ are linearly independent. Denote the elements of $\Omega'$ by $\boldsymbol{x}_1,\ldots,\boldsymbol{x}_r$, where $\boldsymbol{x}_j=((\boldsymbol{x}_j)_1,\ldots,(\boldsymbol{x}_j)_t)\in N'_1\oplus\cdots\oplus N'_t$. Assume there exists a nontrivial relation $\mu_1\boldsymbol{x}_1 +\ldots +\mu_r\boldsymbol{x}_r =\boldsymbol{0}$ with $\boldsymbol{\mu} := (\mu_1,\ldots,\mu_r)\in\mathbb{R}^r\setminus\{(0,\ldots,0)\}$. Let us define \[ \operatorname{supp}(\boldsymbol{\mu}) :=\{j\in\{1,\ldots, r\}\,\colon\mu_j\neq 0\}. \] Let $i\in\{1,\ldots, t\}$ be the largest such integer such that there exists an integer $j\in\operatorname{supp}(\boldsymbol{\mu})$, an index $1\le i_1 < i$, and a vertex $v\in\mathcal{V}_{i_1}\cap\mathcal{V}_i$, with $w_{v,i_1,i}=\boldsymbol{x}_j$. By definition of $i$ and $\Omega'$, all elements in $\{(\boldsymbol{x}_j)_i\,\colon j\in\operatorname{supp}(\boldsymbol{\mu}), (\boldsymbol{x}_j)_i\not=\boldsymbol{0}_{N'_i}\}\neq\emptyset$ are pairwise different vertices in $\mathcal{V}_i\cap\operatorname{vert}(P^{(i-1)})$. Hence \[ \sum_{j\in\operatorname{supp}(\boldsymbol{\mu})}\mu_j\, (\boldsymbol{x}_j)_i =\boldsymbol{0}_{N'_i} \] implies a nontrivial relation of a non-empty subset of the vertices in $\mathcal{V}_i\cap\operatorname{vert}(P^{(i-1)})$. However, as $S_i$ contains the origin in its interior, any proper subset of the set of vertices of $S_i$ is linearly independent, so $\mathcal{V}_i\cap\operatorname{vert}(P^{(i-1)}) =\operatorname{vert}(S_i)$. Hence, $r_i=d_i+1$, a contradiction to~\eqref{eq:di}. \end{proof} \noindent We now apply Lemma~\ref{lem:generators} to~\eqref{eq:P*}: \begin{align} \label{eq:slicing1} \begin{split} P^*&\cong\varphi^*_\mathbb{R} (P^*)\\ &=((S'_1)^*\times\cdots\times (S'_t)^*)\cap (\ker\varphi)_\mathbb{R}^\perp\\ &=\{(y_1,\ldots , y_t)\in (S'_1)^*\times\cdots\times (S'_t)^*\,\colon\langle (y_1,\ldots,y_t) ,\omega\rangle = 0\text{ for each }\omega\in (\ker\varphi)_\mathbb{R}\}\\ &=\{(y_1,\ldots , y_t)\in (S'_1)^*\times\cdots\times (S'_t)^*\,\colon\langle y_{i_1} , e_{i_1,v}\rangle =\langle y_{i_2} , e_{i_2,v}\rangle\text{ for each } w_{v,i_1,i_2}\in\Omega\}\\ &=\{(y_1,\ldots , y_t)\in (S'_1)^*\times\cdots\times (S'_t)^*\,\colon\langle y_{i_1} , e_{i_1,v}\rangle =\langle y_{i_2} , e_{i_2,v}\rangle\text{ for each } v\in\mathcal{V}_{i_1}\cap\mathcal{V}_{i_2}\}.\\ \end{split} \end{align} \subsection{The integration map}\label{subsec:slicing2} From here onwards we will assume that the decomposition of $P$ into the simplices $S_i$ is \emph{irredundant}, i.e.~$\mathcal{V}_i\subsetneq\operatorname{vert}(S_i)$ for $i=1,\ldots, t$. Under this assumption, we describe a specific choice for $N'_i$. For this, we choose a vertex $v_i\in\operatorname{vert}(S_i)\setminus\mathcal{V}_i$, and set \[ \widehat{\V}_i:=\operatorname{vert}(S_i)\setminus\{v_i\}. \] We have $\mathcal{V}_i\subset\widehat{\V}_i$. We define $N'_i$ to be the lattice spanned by $\widehat{\V}_i$, that is, \[ N'_i:=\langle v\in\widehat{\V}_i\rangle_\mathbb{Z}. \] By construction, the $d_i$ vertices in $\widehat{\V}_i$ form a lattice basis \[ \{e_{i,v}\}_{v\in\widehat{\V}_i} \] of $N'_i$ (as a sublattice of $N'_1\oplus\cdots\oplus N'_t$). Note that the vertex $v_i$ need not be a lattice point in $N'_i$. We again assume that $N$ is given as the image of $\varphi$, see~\eqref{phimap}, and we will refer to $S_i$ as $S'_i$ when referring to it with respect to the lattice $N'_i$. This choice of lattice will allow us to prove Lemma~\ref{lem:psisurj} which simplifies the considerations in~\S\ref{sec:integration}. In particular, it will yield a convenient explicit description of $(S'_i)^*$ (see Lemma~\ref{lem:S*}). Set $q :=\abs{\mathcal{V}}$ and $q_i :=\abs{\mathcal{V}_i}$ for $i=1,\ldots, t$. We define $\Psi$ to be the map \begin{align*} \Psi: (\ker\varphi)^\perp &\to\bigoplus_{v\in\mathcal{V}}\mathbb{Z}\cong\mathbb{Z}^q\\ (y_1,\ldots,y_t) &\mapsto (\langle y_{i_v} , e_{i_v,v}\rangle)_{v\in\mathcal{V}}, \end{align*} where, for each $v$, $i_v$ is any index such that $v\in\mathcal{V}_{i_v}$. Since $\langle y_{i_1} , e_{i_1,v}\rangle=\langle y_{i_2} , e_{i_2,v}\rangle$ whenever $v\in\mathcal{V}_{i_1}\cap\mathcal{V}_{i_2}$, $\Psi$ is a well defined map. In an analogous fashion to the definition of $\Psi$, for each $i\in\{1,\ldots,t\}$ we define the map \begin{align*} \Psi_i: M'_i &\to\bigoplus_{v\in\mathcal{V}_i}\mathbb{Z}\cong\mathbb{Z}^{q_i}\\ y &\mapsto (\langle y , e_{i,v}\rangle)_{v\in\mathcal{V}_i}. \end{align*} \begin{lem}\label{lem:psisurj} The maps $\Psi,\Psi_1,\ldots,\Psi_t$ are surjective. \end{lem} \begin{proof} Let $\{\epsilon_{i,v}\}_{v\in\mathcal{V}_i}$ be the standard basis of $\bigoplus_{v\in\mathcal{V}_i}\mathbb{Z}$, and $\{e_{i,v}^*\}_{v\in\widehat{\V}_i}$ the lattice basis of $M'_i$ dual to the lattice basis $\{e_{i,v}\}_{v\in\widehat{\V}_i}$ of $N'_i$. The maps $\Psi_i$ are surjective, since each element $e_{i,v}^*$ is mapped into $\epsilon_{i,v}$, for $v\in\mathcal{V}_i$. We now prove that $\Psi$ is surjective. Since the codomains of the maps $\Psi_i$ span the codomain of $\Psi$, it is enough to check that for each $i\in\{1,\ldots,t\}$ and for each $v\in\mathcal{V}_i$, there exists an element $(y_1,\ldots,y_t)\in (\ker\varphi)^\perp\subset M'_1\oplus\cdots\oplus M'_t$, such that $y_i=e_{i,v}^*$. This is true, since it suffices to choose $(y_1,\ldots,y_t)$ as \[\sum_{j\text{ such that }v\in V_j}\!\!e_{j,v}^*\in M'_1\oplus\cdots\oplus M'_t.\qedhere \] \end{proof} \noindent As a consequence of Lemma~\ref{lem:psisurj} the extensions of $\Psi,\Psi_1,\ldots,\Psi_t$ to the real vector space maps \[ \Psi_\mathbb{R},(\Psi_1)_\mathbb{R},\ldots,(\Psi_t)_\mathbb{R} \] are linear surjective maps. We define natural projections \[ p_i:\bigoplus_{v\in\mathcal{V}}\mathbb{R}\to\bigoplus_{v\in\mathcal{V}_i}\mathbb{R} \] as the identity over $\bigoplus_{v\in\mathcal{V}_i}\mathbb{R}$ and the zero map over $\bigoplus_{v\in\mathcal{V}\setminus\mathcal{V}_i}\mathbb{R}$. Let $\mathcal{D}$ be the set of parameters \[ \mathcal{D}:=\Psi_\mathbb{R}(\varphi^*_\mathbb{R} (P^*))\subset\bigoplus_{v\in\mathcal{V}}\mathbb{R}. \] Given a point $\boldsymbol{\lambda}=(\lambda_v)_{v\in\mathcal{V}}\in\mathcal{D}$, define the fibre \[ H_{i,\boldsymbol{\lambda}} := (\Psi_i)_\mathbb{R}^{-1}(p_i(\boldsymbol{\lambda}))\cap (S'_i)^* =\{y\in (S'_i)^*\,\colon\langle y , v\rangle =\lambda_v\text{ for all }v\in\mathcal{V}_i\}\subset (M'_i)_\mathbb{R}. \] Denote by $F^*_i$ the $(d_i-q_i)$-dimensional face of $(S'_i)^*$ given by \begin{equation} \label{eq:Fi} F^*_i:=H_{i,(-1,\ldots,-1)}. \end{equation} From~\eqref{eq:slicing1} we obtain the desired decomposition of $P^*$: \begin{align} \label{eq:slicing2} \begin{split} P^* &\cong\bigsqcup_{(\lambda_v)_{v\in\mathcal{V}}\in\mathcal{D}}\{(y_1,\ldots , y_t)\in (S'_1)^*\times\cdots\times (S'_t)^*\,\colon\langle y_i , e_{i,v}\rangle =\lambda_v\text{ for all } v\in\mathcal{V}_i,\, i = 1,\ldots, t\}\\ &=\bigsqcup_{\boldsymbol{\lambda}\in\mathcal{D}} H_{1,\boldsymbol{\lambda}}\times\cdots\times H_{t,\boldsymbol{\lambda}}\\ \end{split} \end{align} In other words, $P^*$ is sliced into a disjoint union of sections (see Figure~\ref{fig:slice}). \section{Bounding the volume of $P^*$ via integration}\label{sec:integration} In this section we apply~\eqref{eq:slicing2} to obtain a finer bound on the volume of $P^*$ in the case when $P$ decomposes into just two simplices. From here onwards we assume we are in the setup of Corollary~\ref{cor:Kasp} with $t=2$,~i.e.\ $P$ decomposes in two minimal canonical simplices $S_1$ and $S_2$ of dimensions $d_1$ and $d_2$ respectively. As $P$ is not a simplex, clearly this decomposition is irredundant, so the results of~\S\ref{subsec:slicing2} apply. We will continue to use the notation introduced in~\S\ref{sec:slicing}, and in particular the choice of $N'_i,N,S'_i$ in~\S\ref{subsec:slicing2}. Note that $q=r_2=r = |\mathcal{V}| = |\mathcal{V}_1| = |\mathcal{V}_2|$ is the number of common vertices of $S_1$ and $S_2$. Equality~\eqref{eq:slicing2} and Lemma~\ref{lem:psisurj} allow us to calculate the volume $\operatorname{vol}_M(P^*)$ by integrating the sections over the possible values of $\boldsymbol{\lambda}$. In particular: \begin{equation}\label{eq:int1} \operatorname{vol}_M(P^*) =\int_{\boldsymbol{\lambda}\in\mathcal{D}}\operatorname{vol}_{M'_1}(H_{1,\boldsymbol{\lambda}})\operatorname{vol}_{M'_2}(H_{2,\boldsymbol{\lambda}})\,d\boldsymbol{\lambda}. \end{equation} Before attempting to bound such a value, we present an alternative description of $\mathcal{D}$. For $i=1,2$, we define $\mathcal{D}_i$ as \[ \mathcal{D}_i:=(\Psi_i)_\mathbb{R} ((S'_i)^*), \] and we note that (since the maps $p_i$ defined in the previous section correspond to the identity maps here), \begin{equation} \label{eq:lambda} \mathcal{D}=\mathcal{D}_1\cap\mathcal{D}_2. \end{equation} Recall that a lattice basis $\{e_{i,v}\}$ for $N'_i$ is given by all the elements of $\widehat{\V}_i=\operatorname{vert}(S_i)\setminus\{v_i\}$. Denote by $(\beta_{i,v})_{v\in\operatorname{vert}(S_i)}$ the barycentric coordinates of the origin in the simplex $S_i$,~i.e.\ $\sum_{v\in\operatorname{vert}(S_i)}\beta_{i,v} v =\boldsymbol{0}$, where $\sum_{v\in\operatorname{vert}(S_i)}\beta_{i,v} = 1$. Note that $\beta_{i,v} > 0$ for any $v\in\operatorname{vert}(S_i)$. Hence, we can express $v_i$ as \[v_i= -\sum_{v\in\widehat{\V}_i}\frac{\beta_{i,v}}{\beta_{i,v_i}}e_{i,v}.\] Let us denote by $\{\epsilon_{i,v}\}_{v\in\mathcal{V}_i}$ the standard basis of $\bigoplus_{v\in\mathcal{V}_i}\mathbb{Z}$. Lemma~\ref{lem:S*} below gives an explicit description for $(S'_i)^*$ and $\mathcal{D}_i$ in terms of our chosen lattice bases. We omit the straightforward proof. \begin{lem}\label{lem:S*} With notation as above, for $i=1,2$ \[ (S'_i)^* =\operatorname{conv}\left(\left\{-\sum_{v\in\mathcal{V}} e^*_{i,v}\right\}\cup\left\{\left(\frac{1}{\beta_{i,w}}-1\right)e^*_{i,w}-\sum_{v\in\mathcal{V}\setminus\{w\}} e^*_{i,v}\right\}_{w\in\widehat{\V}_i}\right), \] \[ \mathcal{D}_i =\operatorname{conv}\left(\left\{-\sum_{v\in\mathcal{V}}\epsilon^*_{i,v}\right\}\cup\left\{\left(\frac{1}{\beta_{i,w}}-1\right)\epsilon^*_{i,w}-\sum_{v\in\mathcal{V}\setminus\{w\}}\epsilon^*_{i,v}\right\}_{w\in\mathcal{V}}\right). \] \end{lem} By using the inequality $f_1 f_2\leq\frac{f_1^2+f_2^2}{2}$ we can bound~\eqref{eq:int1} via \begin{align} \begin{split} \operatorname{vol}_{M}(P^*) &\leq\int_{\boldsymbol{\lambda}\in\mathcal{D}}\frac{\operatorname{vol}_{M'_1}(H_{1,\boldsymbol{\lambda}})^2 +\operatorname{vol}_{M'_2}(H_{2,\boldsymbol{\lambda}})^2}{2}\,d\boldsymbol{\lambda}\\ &=\frac{1}{2}\int_{\boldsymbol{\lambda}\in\mathcal{D}}\operatorname{vol}_{M'_1}(H_{1,\boldsymbol{\lambda}})^2\,d\boldsymbol{\lambda} +\frac{1}{2}\int_{\boldsymbol{\lambda}\in\mathcal{D}}\operatorname{vol}_{M'_2}(H_{2,\boldsymbol{\lambda}})^2\,d\boldsymbol{\lambda}\\ &\leq\frac{1}{2}\int_{\boldsymbol{\lambda}\in\mathcal{D}_1}\operatorname{vol}_{M'_1}(H_{1,\boldsymbol{\lambda}})^2\,d\boldsymbol{\lambda} +\frac{1}{2}\int_{\boldsymbol{\lambda}\in\mathcal{D}_2}\operatorname{vol}_{M'_2}(H_{2,\boldsymbol{\lambda}})^2\,d\boldsymbol{\lambda},\\ \end{split} \label{nextiq} \end{align} where the final inequality follows from~\eqref{eq:lambda}. It is convenient to perform a change of variables for $i=1,2$, via the maps \[ \boldsymbol{\alpha}=(\alpha_v)_{v\in\mathcal{V}}\xmapsto{f_i} (\frac{1}{\beta_{i,v}}\alpha_v - 1)_{v\in\mathcal{V}}. \] By Lemma~\ref{lem:S*}, the integration domain $\mathcal{D}_i$ becomes the unimodular $q$-dimensional simplex $\Delta_{(q)}$; that is, the convex hull of the origin and the standard basis of $\mathbb{Z}^q$. Hence~\eqref{nextiq} can be rewritten as: \begin{equation}\label{eq:int2} \operatorname{vol}_{M}(P^*)\leq\frac{1}{2}\prod_{v\in\mathcal{V}}\frac{1}{\beta_{1,v}}\int_{\boldsymbol{\alpha}\in\Delta_{(q)}}\!\!\operatorname{vol}_{M'_1}(H_{1,f_1(\boldsymbol{\alpha})})^2\,d\boldsymbol{\alpha} +\frac{1}{2}\prod_{v\in\mathcal{V}}\frac{1}{\beta_{2,v}}\int_{\boldsymbol{\alpha}\in\Delta_{(q)}}\!\!\operatorname{vol}_{M'_2}(H_{2,f_2(\boldsymbol{\alpha})})^2\,d\boldsymbol{\alpha}.\\ \end{equation} \begin{lem}[\!\!{\cite[Lemma~3.5~III]{Ave12}}] With notation as above, for $i=1,2$, \[ \operatorname{vol}_{M'_i}(H_{i,f_i(\boldsymbol{\alpha})})=\operatorname{vol}_{M'_i}(F^*_i)\left( 1 -\sum_{v\in\mathcal{V}_i}\alpha_v\right) ^{d_i-q}, \] where $F_i$ is the $(d_i-q)$-dimensional face of $(S'_i)^*$ defined in~\eqref{eq:Fi}. \end{lem} \noindent Inequality~\eqref{eq:int2} can now be rewritten as \begin{align}\label{eq:int3} \begin{split} \operatorname{vol}_{M}(P^*)\leq &\frac{1}{2}\prod_{v\in\mathcal{V}}\frac{1}{\beta_{1,v}}\operatorname{vol}_{M'_1}(F^*_1)^t\int_{\boldsymbol{\alpha}\in\Delta_{(q)}}\left( 1 -\sum_{v\in\mathcal{V}}\alpha_v\right) ^{2(d_1-q)}\,d\boldsymbol{\alpha}\\ +& \frac{1}{2}\prod_{v\in\mathcal{V}}\frac{1}{\beta_{2,v}}\operatorname{vol}_{M'_2}(F^*_2)^t\int_{\boldsymbol{\alpha}\in\Delta_{(q)}}\left( 1 -\sum_{v\in\mathcal{V}}\alpha_v\right) ^{2(d_2-q)}\,d\boldsymbol{\alpha}.\\ \end{split} \end{align} The following Lemma derives from a special case of a well-known representation of the beta function (see, for example,~\cite[Representation~4.3-2]{Car77}). \begin{lem}\label{lem:beta-representation} \[ \int_{\boldsymbol{\alpha}\in\Delta_{(a)}}( 1 -\alpha_1 -\ldots -\alpha_a )^b\,d\boldsymbol{\alpha} =\frac{b!}{(a+b)!}. \] \end{lem} \noindent Applying Lemma~\ref{lem:beta-representation} to~\eqref{eq:int3} yields: \begin{equation}\label{eq:int4} \operatorname{vol}_{M}(P^*)\leq\frac{1}{2}\prod_{v\in\mathcal{V}}\frac{1}{\beta_{1,v}}\operatorname{vol}_{M'_1}(F^*_1)^2\frac{(2(d_1-q))!}{(q+2(d_1-q))!} +\frac{1}{2}\prod_{v\in\mathcal{V}}\frac{1}{\beta_{2,v}}\operatorname{vol}_{M'_2}(F^*_2)^2\frac{(2(d_2-q))!}{(q+(2(d_2-q))!}.\\ \end{equation} The volume of $F^*_i$ is computed in Lemma~\ref{lem:vol_F_dual} below. Its proof is omitted, since it is a straightforward consequence of the description of $(S'_i)^*$ given in Lemma~\ref{lem:S*}. \begin{lem}\label{lem:vol_F_dual} With notation as above, for $i=1,2$, \[ \operatorname{vol}_{M'_i}(F_i^*)=\frac{1}{(d_i-q)!}\prod_{v\in\widehat{\V}_i\setminus\mathcal{V}}\frac{1}{\beta_{i,v}}. \] \end{lem} \noindent Finally, applying Lemma~\ref{lem:vol_F_dual} to~\eqref{eq:int4} gives the following bound for $\operatorname{vol}_{M}(P^*)$: \begin{align}\label{eq:int5} \begin{split} \operatorname{vol}_{M}(P^*)\leq\frac{1}{2}&\frac{(2(d_1-q))!}{(q+2(d_1-q))!((d_1-q)!)^2}\prod_{v\in\mathcal{V}}\frac{1}{\beta_{1,v}}\prod_{v\in\widehat{\V}_1\setminus\mathcal{V} }\frac{1}{\beta_{1,v}^2}\\ &+\frac{1}{2}\frac{(2(d_2-q))!}{(q+2(d_2-q))!((d_2-q)!)^2}\prod_{v\in\mathcal{V}}\frac{1}{\beta_{2,v}}\prod_{v\in\widehat{\V}_2\setminus\mathcal{V} }\frac{1}{\beta_{2,v}^2} \end{split} \end{align} \section{Final cases}\label{sec:final} In this final section we address the remaining cases of Corollary~\ref{cor:almostallcases}. That is, we prove that the decompositions \begin{enumerate} \item $t=2$, $d_1=d_2=d-1$, for $d\geq 4$ \item $t=2$, $d_1=d-1$, $d_2=d-2$, $d\in\{4,5\}$ \item $t=3$, $d_1=d_2=d_3=d-2$, $d\in\{4,5\}$ \end{enumerate} satisfy Theorem~\ref{thm:main}. \subsection{The case $t=2$, $d_1=d_2=d-1$}\label{sec:t2_d} By~\eqref{eq:dr} we have $q=d-2$. Hence, inequality~\eqref{eq:int5} can be rewritten as \begin{equation}\label{eq:lastcase} \operatorname{vol}_{M}(P^*)\leq\frac{1}{d!}\left(\prod_{v\in\mathcal{V}}\frac{1}{\beta_{1,v}}\prod_{v\in\widehat{\V}_1\setminus\mathcal{V} }\frac{1}{\beta_{1,v}^2} +\prod_{v\in\mathcal{V}}\frac{1}{\beta_{2,v}}\prod_{v\in\widehat{\V}_2\setminus\mathcal{V} }\frac{1}{\beta_{2,v}^2}\right) \end{equation} We focus on the product \[ \prod_{v\in\mathcal{V}}\frac{1}{\beta_{i,v}}\prod_{v\in\widehat{\V}_i\setminus\mathcal{V} }\frac{1}{\beta_{i,v}^2} \] for each $i=1,2$. Note that in~\S\ref{subsec:slicing2} we made a choice to exclude one of the vertices (called $v_i$) of $\operatorname{vert}(S_i)\setminus\mathcal{V}$ from appearing in $\widehat{\V}_i$. As there are two such vertices (say, $\operatorname{vert}(S_i)\setminus\mathcal{V}=\{v_i, u_i\}$), we can exclude the one whose corresponding barycentric coordinate is smaller; that is, $\beta_{i,{v_i}}\le\beta_{i,{u_i}}$. This yields \begin{equation}\label{eq:last} \prod_{v\in\mathcal{V}}\frac{1}{\beta_{i,v}}\prod_{v\in\widehat{\V}_i\setminus\mathcal{V} }\frac{1}{\beta_{i,v}^2} =\left(\prod_{v\in\mathcal{V}}\frac{1}{\beta_{i,v}}\right)\frac{1}{\beta_{i,{u_i}}^2}\le\left(\prod_{v\in\mathcal{V}}\frac{1}{\beta_{i,v}}\right)\frac{1}{\beta_{i,{u_i}}}\frac{1}{\beta_{i,{v_i}}} =\frac{1}{\beta_{i,0},\ldots,\beta_{i,d-1}}, \end{equation} where $\{\beta_{i,v}\,\colon v\in\operatorname{vert}(S_i)\} =\{\beta_{i,j}\,\colon j=0,\ldots, d-1\}$. Notice that equality in~\eqref{eq:last} is attained if and only if $\beta_{i,u_i} =\beta_{i,v_i}$. For each $i=1,2$, let us sort the barycentric coordinates such that $\beta_{i,0}\ge\beta_{i,1}\ge\cdots\ge\beta_{i,d-1}$. \begin{lem}[\!\!{\cite[Lemma~4.2(d)]{AKN14}}]\label{lem:bary_iff} With notation as above, \[ \frac{1}{\beta_{i,0}\cdots\beta_{i,d-1}}\leq (s_d-1)^2 \] with equality if and only if \begin{equation}\label{eq:bary_iff} \left(\beta_{i,0},\ldots,\beta_{i,d-1}\right) =\left(\frac{1}{s_1},\ldots,\frac{1}{s_{d-1}},\frac{1}{s_d-1}\right). \end{equation} \end{lem} \noindent Applying Lemma~\ref{lem:bary_iff} and~\eqref{eq:last} to~\eqref{eq:lastcase} we obtain \[ \operatorname{vol}_{M}(P^*) <\frac{2(s_d-1)^2}{d!}. \] This inequality is strict, since the condition that $\beta_{i,u_i} =\beta_{i,v_i}$ from~\eqref{eq:last} and the condition~\eqref{eq:bary_iff} from Lemma~\ref{lem:bary_iff} cannot hold simultaneously. \subsection{The cases $t=2$, $d_1=d-1$, $d_2=d-2$, $d\in\{4,5\}$}\label{sec:t2_45} The barycentric coordinates of the canonical Fano simplices up to and including dimension four are classified in~\cite{Kas13}. Hence we can verify that, in this situation, the right hand side of~\eqref{eq:int5} is always strictly less than $2(s_d-1)^2/d!$. \subsection{The cases $t=3$, $d_1=d_2=d_3=d-2$, $d\in\{4,5\}$}\label{sec:t3_45} To prove the inequality in these final cases we explicitly construct every minimal polytope $P$ of dimension four or five that admits a decomposition into three minimal simplices of dimensions two or three, respectively. Moreover, we insist that the vertices of $P$ generate the ambient lattice $N$. Indeed, if $P$ is a minimal polytope then $P$ restricted to the lattice generated by the vertices of $P$ is also minimal, and the volume of the dual polytope will have increased. Under this setting we note that $P$ is uniquely determined by: \begin{enumerate} \item the barycentric coordinates of the simplices $S_1,S_2,S_3$ in the decomposition; and \item the choice of $d-3$ vertices in common with $S_2$ and $S_1$, together with the choice of $d-3$ vertices in common with $S_3$ and $S_1\cup S_2$. \end{enumerate} This follows from the following general construction. The \emph{(reduced) weights} of a canonical Fano simplex $S$ of dimension $n$ are the positive integers $(k\beta_0,\ldots, k\beta_n)$ given by the barycentric coordinates $(\beta_0,\ldots,\beta_n)$ of the origin (with respect to the vertices of $S$), where $k$ is the smallest positive integer such that the $k\beta_i$ are all integral. In particular, the weights of a canonical Fano simplex are coprime. Moreover, since the vertices of a canonical Fano simplex are primitive lattice points, the weights are also \emph{well-formed}; that is, any $n$ of them are also coprime. For the construction we use the fact that any minimal polytope $P$ has a decomposition into $t$ minimal simplices. We proceed invariantly, since we do not know the embedding of these simplices into the lattice $N$. Let $\underline{\lambda}^{(n)}=(\lambda_0,\ldots,\lambda_n)$ denote the (reduced, well-formed) weights of a minimal canonical Fano simplex of dimension $n$. Fix weights $\underline{\lambda}^{(d_1)},\ldots,\underline{\lambda}^{(d_t)}$. For each pair $(i,j)$ with $1\leq i < j\leq t$ we pick a (possibly empty) subset $V_{ij}\subset\{0,\ldots,d_i\}\times\{0,\ldots,d_j\}$ such that $V_{ij}:\pi_1(V_{ij})\to\pi_2(V_{ij})$ is a bijection (here $\pi_k$ denotes the projection on the $k$-th factor). Let $\iota_j:\mathbb{Z}^{d_j+1}\to\bigoplus_{i=1}^t\mathbb{Z}^{d_i+1} $, $1\leq j\leq t$, be the natural inclusion on the $j$-th factor. Define: \begin{align*} W&:=\left\langle\iota_i(\underline{\lambda}^{(d_i)})\mid 1\leq i\leq d \right\rangle,\\ V&:=\left\langle\iota_i(e_{\pi_1(v)}) -\iota_j(e_{\pi_2(v)})\mid v\in V_{ij}, 1\leq i < j\leq t\right\rangle. \end{align*} Applying $-\otimes\mathbb{R}$ ensures torsion-freeness of the quotient $\left(\bigoplus_{i=1}^t\mathbb{Z}^{d_i+1}\right)/(W + V)$, therefore we get the exact sequence \[ 0\to (W + V)\otimes\mathbb{R}\to\left(\bigoplus_{i=1}^t\mathbb{Z}^{d_i+1}\right)\otimes\mathbb{R}\xrightarrow{\varphi_\mathbb{R}} N\otimes\mathbb{R}\to 0, \] where $N$ is the lattice obtained as the quotient $\left(\bigoplus_{i=1}^t\mathbb{Z}^{d_i+1}\right)/K$, where $K$ is the direct summand defined by $\left(\bigoplus_{i=1}^t\mathbb{Z}^{d_i+1}\right)\cap\left((W + V)\otimes\mathbb{R}\right)$. We now define \[ Q:=\varphi_\mathbb{R}\left( \bigoplus_{i=1}^t\iota_i\left(\operatorname{conv}\{e_0,\ldots,e_{d_i}\}\right)\right)\subset N\otimes\mathbb{R} \] which by construction is a polytope whose vertices generate its ambient lattice $N$. $Q$ in general may not be a minimal polytope, however, if $P$ is a minimal lattice polytope of dimension $d$ whose vertices generate its ambient lattice then there exists a choice of integers $t,d_1,\ldots,d_t$, weights $\underline{\lambda}^{(d_1)},\ldots,\underline{\lambda}^{(d_t)}$ of minimal Fano simplices $S_1,\ldots,S_t$ of dimensions $d_1,\ldots,d_t$, and subsets $V_{ij}$ (for $1\leq i < j\leq t$) such that the polytope $Q$ constructed above is equal to $P$. The fact that we can recover $P$ from the construction of $Q$ is a consequence of Lemma~\ref{lem:generators}, while existence of the parameters $t,d_1,\ldots,d_t$ and the weights follows from Corollary~\ref{cor:Kasp}. We now specialise this construction to the case $t=3$, $d_1=d_2=d_3=d-2$, for $d\in\{4,5\}$. The weights of the minimal canonical Fano simplices of dimension two and three have been classified in~\cite[Figure~1 and Proposition~4.3]{Kas10}. There are two possible weights in dimension two: $(1,1,1)$ and $(1,1,2)$. In dimension three there are $13$ possible weights\footnote{\cite[Proposition~4.3]{Kas10} incorrectly lists $(2,2,3,5)$ as the weight of a minimal canonical Fano simplex, however any such simplex will contain a canonical Fano sub-simplex with weights $(1,1,1,3)$.}, recorded in Table~\ref{tab:dim3}. Since the choices for the common vertices (encoded in the sets $V_{ij}$, $1\leq i < j\leq 3$) are finite, so all the minimal canonical Fano polytopes $P$ admitting such a decomposition and whose vertices generate the ambient lattice $N$ can be classified. \begin{table}[tb] \centering \begin{tabular}{ccccc} $( 1, 1, 1, 1 )$&$( 1, 1, 1, 2 )$&$( 1, 1, 1, 3 )$&$( 1, 1, 2, 2 )$&$( 1, 1, 2, 3 )$\\ $( 1, 1, 2, 4 )$&$( 1, 1, 3, 4 )$&$( 1, 1, 3, 5 )$&$( 1, 1, 4, 6 )$&$( 1, 2, 3, 5 )$\\ $( 1, 3, 4, 5 )$&$( 2, 3, 5, 7 )$&$( 3, 4, 5, 7 )$&& \end{tabular}\vspace{0.5em} \caption{The weights of the minimal canonical Fano simplices in dimension three.} \label{tab:dim3} \end{table} We use the computer algebra system \textsc{Magma}~\cite{BCP97} to derive the classification. Source code and output can be downloaded from the Graded Ring Database~\cite{GRDb}. In the first case ($d_1=d_2=d_3=2$), there are exactly four such four-dimensional polytopes, and in each case the inequality of Theorem~\ref{thm:main} holds. In order to solve the second case ($d_1=d_2=d_3=3$), we first build all possible four-dimensional minimal polytopes $P'$ whose vertices generate the ambient lattice, and admitting a decomposition into two three-dimensional minimal canonical Fano simplices $S_1$ and $S_2$. We then verify that any five-dimensional polytope $P$ decomposing as $S_1$, $S_2$, and $S_3$ satisfies inequality~\eqref{eq:geq3cor} for each choice of three-dimensional minimal canonical Fano simplex $S_3$; that is, we verify that \[ \operatorname{Vol}(P^*)\leq\frac{7!}{4!\,3!}\operatorname{Vol}(P'^*)\cdot 2(s_3-1)^2 < 2(s_5-1)^2 \] holds in each case. There are $147$ minimal four-dimensional polytopes with a decomposition into two three-dimensional minimal canonical Fano simplices and whose vertices generate the lattice $N$, and in each case the inequality holds. This completes the proof of Theorem~\ref{thm:main}. \subsection*{Acknowledgements} GB is supported by the Stiftelsen GS Magnusons Fund and by a Jubileumsfond grant from the Knut and Alice Wallenbergs Foundation. In addition, both GB and BN are partially supported by Vetenskapsr{\aa}det grant~NT:2014-3991. BN is an affiliated researcher of Stockholm University; he would like to thank the Fields Institute for the financial support to participate in the thematic program ``Combinatorial Algebraic Geometry''. AK is supported by EPSRC grant~EP/N022513/1. \bibliographystyle{amsplain} \providecommand{\bysame}{\leavevmode\hbox to3em{\hrulefill}\thinspace} \providecommand{\MR}{\relax\ifhmode\unskip\space\fi MR } \providecommand{\MRhref}[2]{% \href{http://www.ams.org/mathscinet-getitem?mr=#1}{#2} } \providecommand{\href}[2]{#2}
{ "timestamp": "2016-11-09T02:04:53", "yymm": "1611", "arxiv_id": "1611.02455", "language": "en", "url": "https://arxiv.org/abs/1611.02455", "abstract": "We give an upper bound on the volume vol(P*) of a polytope P* dual to a d-dimensional lattice polytope P with exactly one interior lattice point, in each dimension d. This bound, expressed in terms of the Sylvester sequence, is sharp, and is achieved by the dual to a particular reflexive simplex. Our result implies a sharp upper bound on the volume of a d-dimensional reflexive polytope. Translated into toric geometry, this gives a sharp upper bound on the anti-canonical degree $(-K_X)^d$ of a d-dimensional toric Fano variety X with at worst canonical singularities.", "subjects": "Combinatorics (math.CO); Algebraic Geometry (math.AG); Metric Geometry (math.MG)", "title": "On the maximum dual volume of a canonical Fano polytope", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9888419694095656, "lm_q2_score": 0.8267117962054049, "lm_q1q2_score": 0.817487320693872 }
https://arxiv.org/abs/1007.1095
The number of unit distances is almost linear for most norms
We prove that there exists a norm in the plane under which no n-point set determines more than O(n log n log log n) unit distances. Actually, most norms have this property, in the sense that their complement is a meager set in the metric space of all norms (with the metric given by the Hausdorff distance of the unit balls).
\section{Introduction} What is the maximum possible number $u(n)$ of unit distances determined by an $n$-point set in the Euclidean plane? This tantalizing question, raised by Erd\H{o}s \cite{e-sdnp-46} in 1946, has motivated an extensive research (see, e.g., Brass, Moser, and Pach \cite{BMP-problems} for a survey), but it remains wide open. Erd\H{o}s \cite{e-sdnp-46} proved a lower bound $u(n)=\Omega(n^{1+c/\log\log n})$ for a constant $c>0$, attained for the $\sqrt n\times\sqrt n$ grid, and he conjectured that it has the right order of magnitude (and in particular, that $u(n)=O(n^{1+\eps})$ for every fixed $\eps>0$). However, the current best upper bound is only $O(n^{4/3})$. It was first proved by Spencer, Szemer\'edi, and Trotter \cite{sst-udep-84}, based on the method of Szemer{\'e}di and Trotter \cite{st-cdbep-83}, and several simpler proofs are available by now (by Clarkson et al.~\cite{cegsw-ccbac-90}, by Aronov and Sharir \cite{AronovSharir-newincid}, and the simplest one by Sz\'ekely \cite{s-cnhep-97}). The problem of unit distances has also been considered for norms other than the Euclidean one. For a norm\footnote{We recall that a (real) \emph{norm} on a real vector space $Z$ is a mapping that assigns a nonnegative real number $\|\makevec{x}\|$ to each $\makevec{x}\in Z$ so that $\|\makevec{x}\|=0$ implies $\makevec{x}=0$, $\|\alpha \makevec{x}\|=|\alpha|\cdot\|\makevec{x}\|$ for all $\alpha\in{\mathbb{R}}$, and the triangle inequality holds: $\|\makevec{x}+\makevec{y}\|\leq \|\makevec{x}\|+\|\makevec{y}\|$. The \emph{unit ball} of the norm $\|.\|$ is the set $B_{\|.\|}=\{\makevec{x}\in Z: \|\makevec{x}\|\le 1\}$. The unit ball of any norm is a closed bounded convex body $B$ that is symmetric about $\makevec{0}$ and contains $\makevec{0}$ in the interior. Conversely, every $B\subset Z$ with the listed properties is the unit ball of a (uniquely determined) norm. } $\|.\|$ on ${\mathbb{R}}^2$, let $u_{\|.\|}(n)$ denote the maximum possible number of unit distances determined by $n$ points in $({\mathbb{R}}^2,\|.\|)$. If the boundary of the unit ball $B_{\|.\|}$ of $\|.\|$ contains a straight segment, then it is easy to construct $n$-point sets with $\Omega(n^2)$ unit distances. On the other hand, if $\|.\|$ is \emph{strictly convex}, meaning that the boundary of $B_{\|.\|}$ contains no straight segment, then $u_{\|.\|}(n)=O(n^{4/3})$, as can be shown by a straightforward generalization of the known proofs for the Euclidean case. Valtr \cite{valtr-norm-ud}, strengthening an earlier result of Brass, constructed a strictly convex norm $\|.\|$ in the plane with $u_{\|.\|}=\Omega(n^{4/3})$, thus showing that the upper bound cannot be improved in general for strictly convex norms. A simple construction shows that $u_{\|.\|}(n)=\Omega(n\log n)$ holds for every norm $\|.\|$ (see, e.g., \cite{BMP-problems}). Here we will show that there exists a norm $\|.\|$ with $u_{\|.\|}(n)=O(n\log n\log\log n)$, almost matching the lower bound. Actually, we show that most norms, in the sense of Baire category, have this property. To formulate this result, we recall the relevant notions. Let $\BB$ be the set of all unit balls of norms in ${\mathbb{R}}^2$, i.e., of all closed bounded $\makevec{0}$-symmetric convex sets containing $\makevec{0}$ in the interior. Endowed with the Hausdorff metric\footnote{We recall that the Hausdorff distance $d_H(A,B)$ of two sets in the Euclidean plane is defined as $\min(h(A,B),h(B,A))$, where $h(A,B)= \sup_{a\in A}\inf_{b\in B}\|a-b\|_2$, with $\|.\|_2$ denoting the Euclidean distance.} $d_H$, the set $\BB$ forms a Baire space, meaning that each meager set\footnote{A set $S$ in a metric (or topological) space $X$ is \emph{nowhere dense} if every nonempty open set $U\subseteq X$ contains a nonempty open set $V$ with $V\cap S=\emptyset$. A \emph{meager set} is a countable union of nowhere dense sets.} has a dense complement; see, e.g., Gruber \cite[Chapter~13]{Gruber-book}. If $P$ is some property that a norm on ${\mathbb{R}}^2$ may or may not have, we say that \emph{most norms have property $P$} if the (unit balls of the) norms not having property $P$ form a meager set in $\BB$. A similar terminology is commonly used for convex bodies. If most norms have property $P_1$ and most norms have property $P_2$, then most norms have both $P_1$ and $P_2$ (and similarly for countably many properties), which makes this approach a powerful tool for proving existence results. Starting with a paper of Klee \cite{klee-mostsmooth}, who proved that most norms are smooth and strictly convex, there have been many papers establishing that most norms or most convex bodies have various properties (see \cite{Gruber-book}). We add the following item to this collection. \begin{theorem}\label{t:} There exists a constant $C_0$ such that most norms $\|.\|$ on ${\mathbb{R}}^2$ satisfy $$u_{\|.\|}(n) < C_0n\log n\log\log n $$ for all $n\ge 3$ (\,$\log$ stands for logarithm in base~$2$ everywhere in this paper). In particular, there exists a smooth and strictly convex norm $\|.\|$ with this property. \end{theorem} Since, as was mentioned above, $u_{\|.\|}(n)=\Omega(n\log n)$ for all norms, the bound in the theorem is tight up to the $O(\log\log n)$ factor. This factor comes out of a graph-theoretic result, Proposition~\ref{p:} below, and I have no good guess whether it is really needed. The proof of the theorem has two main parts. We begin with the first, purely graph-theoretic part in Section~\ref{s:edgecol}. The result needed for the rest of the proof is Proposition~\ref{p:}, asserting the existence of a certain subgraph in every sufficiently dense graph with a given proper edge-coloring. Its proof relies heavily on a similar result of P\v{r}\'{\i}v\v{e}tiv\'y, {\v{S}}kovro\v{n}, and the author \cite{MatPrivSko} (but the presentation below is self-contained). Then, in Section~\ref{s:udg} we continue with the second, geometric part of the proof of Theorem~\ref{t:}. Very roughly speaking, using the graph-theoretic result from the first part of the proof, we show that if there is a set $P$ with many unit distances, under any norm, and if $\makevec{u}_1,\ldots,\makevec{u}_k$ are all the mutually non-parallel unit vectors defined by pairs of points of $P$, then there are ``many'' linear dependences among the $\makevec{u}_i$. Namely, there is an integer $\ell$, such that some $\ell+1$ vectors among the $\makevec{u}_i$ can be expressed as linear functions of some other $\ell$ of the $\makevec{u}_i$ (where the linear functions don't depend on the norm). Finally, we show that most norms don't admit such linear dependences---this is done by approximating the unit ball of the considered norm by a convex polygon, and employing a linear-algebraic perturbation argument to the lines bounding the polygon. It would be interesting to prove a similar result for some narrower class of norms. For example, one might hope to prove that the $\ell_p$ norms admit only a near-linear number of unit distances for most $p$ (in the Baire category sense or even for almost all $p$ w.r.t.\ the Lebesgue measure). For that, the idea of polygonal approximations seems unusable, but perhaps more powerful tools from algebraic geometry might help. Finally, of course, it might be possible to use some pieces from the method of this paper for attacking the Euclidean case. However, since the number of unit distances for the Euclidean case can be much larger than $n\log n\log \log n$, additional ideas are certainly needed. \section{Connected subgraphs with few colors in edge-colored graphs} \label{s:edgecol} Let $G=(V,E)$ be a (simple, undirected) graph. An \emph{edge coloring} of $G$ is a mapping $c\:E\to{\mathbb{N}}=\{1,2,3,\ldots\}$. The edge coloring $c$ is called \emph{proper} if $c(e)\ne c(e')$ whenever the edges $e$ and $e'$ share a vertex. Let $G$ be a graph with a given edge coloring. For a subset $W\subseteq V$ of vertices we let $G[W]$ stand for the subgraph of $G$ induced by $W$, with the edge coloring inherited from that of $G$. Further, if $I\subseteq {\mathbb{N}}$ is a set of colors, we write $G[I,W]$ for the subgraph induced by $W$ on the edges with colors in $I$, that is, $$ G[I,W]=\Bigl(W,\{\{u,v\}\in E: u,v\in W,c(e)\in I\}\Bigr) $$ (the coloring is not explicitly mentioned in the notation). \begin{prop}\label{p:} Let $q>1$ be a real parameter. Let $G=(V,E)$ be a graph on $n\ge 4$ vertices, with at least $Cq n\log n\log\log n$ edges (where $C$ is a suitable absolute constant), and with a given proper edge coloring. Then there exist a nonempty subset $W\subseteq V$ of vertices, $|W|\ge 2$, and a subset $I\subset{\mathbb{N}}$ of colors such that the subgraph $G[I,W]$ is connected and the edges of $G[W]$ have at least $q|I|$ distinct colors. \end{prop} As was mentioned in the introduction, this proposition is similar to a result from \cite{MatPrivSko}, and the proof is also quite similar to the one in \cite{MatPrivSko}. I still consider it worth presenting in full, since describing the required modifications would be clumsy, and moreover, the proof below is significantly simpler than that in \cite{MatPrivSko}, mainly because the required result is weaker (in Proposition~\ref{p:} we obtain a single connected subgraph, while in \cite{MatPrivSko} several color-disjoint connected subgraphs on the same vertex set were needed). \medskip At the beginning of the proof, we use a well-known observation stating that every graph of average degree $\delta$ has a subgraph whose minimum degree is at least $\delta/2$ (this follows by repeatedly deleting vertices of degree below $\delta/2$ and checking that the average degree can't decrease). So we may assume that $G$ has minimum degree at least $Cq n\log n\log\log n$. Let $W\subseteq V$ be a subset of vertices of $G$ (so far arbitrary). An \emph{edge cut} in $G[W]$ is a partition $(A,B)$ of $W$ into two nonempty subsets. We define the \emph{maximum degree} $\Delta(A,B)$ of such an edge cut as the maximum number of neighbors of a vertex from $A$ in $B$ or of a vertex from $B$ in $A$; formally, $$ \Delta(A,B):=\max\left\{\max_{a\in A}|\{\{a,b\}\in E:b\in B\}|, \max_{b\in B}|\{\{b,a\}\in E:a\in A\}|\right\}. $$ The proof of Proposition~\ref{p:} proceeds in two stages. In the first stage, we forget about the edge colors; we select the set $W$ so that every edge cut in $G[W]$ has a sufficiently large maximum degree. In order to get the (almost tight) quantitative result in the proposition, we need to quantify the ``sufficiently large maximum degree'' of a cut depending on the \emph{imbalance} of the cut, which is defined by $$ {\rm imb}(A,B) := \frac{|A|+|B|}{\min(|A|,|B|)}. $$ \begin{lemma}\label{l:coloredcuts} Let $r\ge 1$ be a parameter (which we will later set to $Cq\log\log n$ in the application of the lemma), and let $G=(V,E)$ be a graph on $n\ge 2$ vertices of minimum degree at least $ r\log n$. Then there exists $W\subseteq V$, $|W|\ge 2$, such that every edge cut $(A,B)$ in $G[W]$ satisfies $$ \Delta(A,B)\ge r \log {\rm imb}(A,B). $$ \end{lemma} \heading{Proof. } The proof proceeds by a recursive partitioning: As long as we can find an edge cut $(A,B)$ of small maximum degree in the current graph, we discard the \emph{larger} of the sets $A,B$. More formally, we set $V_1:=V$. If $G[V_j]$ has already been constructed and if there is an edge cut $(A_j,B_j)$ in $G[V_j]$ with $\Delta(A_j,B_j)<r\log{\rm imb}(A_j,B_j)$, we let $V_{j+1}$ be the smaller of the sets $A_j$ and $B_j$ (ties broken arbitrarily) and iterate. If there is no such edge cut, we set $W:=V_j$, $t:=j$, and finish. It remains to show that the resulting $W$ is nontrivial, i.e., $|W|\ge 2$. This is clear for $t=1$ (no partition step was made), so we assume $t\ge 2$. We show that $G[W]=G[V_t]$ has minimum degree at least~$1$, and thus $W$ can't consist of a single vertex. Initially, in $G$, each vertex has degree at least $r\log n$, and by passing from $V_j$ to $V_{j+1}$, each vertex of $V_{j+1}$ loses at most $\Delta(A_j,B_j)<r\log{\rm imb}(A_j,B_j)$ neighbors. Thus, the minimum degree in $G[V_t]$ is strictly larger than \begin{eqnarray*} r\log n - r\sum_{j=1}^{t-1} \log{\rm imb}(A_j,B_j) &=&r\log n - r\sum_{j=1}^{t-1}\log\frac{|V_j|}{|V_{j+1}|}\\ &=&r\log n - r\Bigl(\log |V_1| - \log |V_{t}|\Bigr)\ge 0.\\ \end{eqnarray*} The lemma is proved. \ProofEndBox\smallskip \medskip Now we continue with the \emph{second stage of the proof of Proposition~\ref{p:}}. Only here we start considering the edge colors. According to Lemma~\ref{l:coloredcuts}, we now assume that $W\subseteq V$, $|W|\ge 2$, is such that every edge cut $(A,B)$ in $G[W]$ has maximum degree at least $r\log {\rm imb}(A,B)$, with $r=Cq\log\log n$. Consequently, the edges of every edge cut $(A,B)$ have at least $r\log {\rm imb}(A,B)$ distinct colors (since the edge coloring is proper), and this is the only property of $G[W]$ we will use. Let $k$ denote the number of colors occurring on the edges of $G[W]$. We note that $k\ge r$ (this follows by using the condition above for an arbitrary cut). It remains to show that $G[W]$ has a connected subgraph that uses at most $k/q$ colors. We select the colors greedily one by one, as follows. We set $I_0:=\emptyset$, and for $j=0,1,2,\ldots,$ we do the following: If $G[I_{j},W]$ is connected, we set $I:=I_{j}$ and finish. Otherwise, we let $i_j$ be a color $i$ minimizing the number of connected components of $G[I_j\cup\{i\},W]$. Then we set $I_{j+1}:=I_j\cup\{i_j\}$, and we continue with the next step. We need to show that we obtain a connected graph before exhausting more than $k/q$ colors. Let $m_j$ be the number of connected components of $G[I_j,W]$. We want an upper bound on the smallest $j$ with $m_j=1$. First we observe that $m_{j+1}\le m_j-1$ for all $j$, since every edge cut contains at least one color. In the sequel, we will actually estimate the smallest $j$ such that $m_j\le 3$. Then at most two more steps suffice to get down to $m_j=1$. We now want to bound $m_{j+1}$ in terms of $m_j$. Essentially, we will see that adding a random color to $I_j$ is likely to connect up many components. Let $K_1,\ldots,K_{m_j}$ be the vertex sets of the connected components of $G[I_j,W]$. The average number of vertices in a component is $m/m_j$; we call a component \emph{small} if it has at most $2m/m_j$ vertices. By Markov's inequality, there are at least $m_j/2$ small components. Let $i$ be one of the colors occurring on the edges of $G[W]$ but not belonging to $I_j$ (so there are $k-j$ possible choices for $i$). We say that a component $K_s$ \emph{gets connected} by $i$ if there is an edge of color $i$ connecting a vertex of $K_s$ to a vertex outside $K_s$. By the condition on the edge cuts of $G[W]$, if $K_s$ is a small component, then the number of colors $i$ by which $K_s$ gets connected is at least $$ r\log {\rm imb}\left(K_s,W\setminus K_s\right)\ge r\log \frac m{2m/m_j}= r\log (m_j/2). $$ Thus, the expected number of small components that get connected by a random color is at least $$ \frac{m_j}{2}\cdot \frac{r\log (m_j/2)}{k-j}\ge \frac{m_j}{2}\cdot \frac{r\log (m_j/2)}{k}. $$ So at least this many components get connected by the color $i_{j+1}$. It is easy to check that the number of components always decreases at least by half of the number of components that get connected (an extremal case being components merged in pairs). Thus, we have $$ m_{j+1}\le m_j - \frac{m_j}{4}\cdot \frac{r\log (m_j/2)}{k} \le m_j\left(1-\frac{r\log (m_j/2)}{4k}\right) \le m_je^{-r\log (m_j/2)/4k} $$ (we used $1-x\le e^{-x}$ in the last step). Assuming, as we may, that $m_j\ge 4$, we have $\log (m_j/2)\ge \frac 12\log m_j\ge \frac12\ln m_j$, and so $$ \ln m_{j+1}\le \ln m_j-r(\ln m_j)/8k=(1-r/8k)\ln m_j\le e^{-r/8k}\ln m_j. $$ Since $m_1\le n$, we can see that $m_j$ drops below $4$ in at most $O((k/r)\log\log n)=O(k/Cq)$ steps. We need at most two extra colors to get all the way to $m_j=1$, so altogether the number of colors needed to build a connected graph is $O(k/Cq+2)=O(k/Cq)$ (since $k\ge r$, and thus $k/Cq\ge \log\log n\ge 1$). The implicit constant in the $O(.)$ notation is independent of $C$, and thus we can set $C$ so large that the number of colors is at most $k/q$. Proposition~\ref{p:} is proved. \ProofEndBox\smallskip \section{Unit-distance graphs}\label{s:udg} Let $\|.\|$ be a norm in the plane, and let $P=(\makevec{p}_1,\ldots,\makevec{p}_n)$ be a sequence of $n$ distinct points in the plane. With these objects we associate a finite combinatorial object, which we will call the \emph{decorated unit-distance graph}. First, we define the \emph{unit-distance graph} $G=G(\|.\|,P)$ as the (undirected) graph $(V,E)$ with vertex set $V:=[n]$ (where we use the notation $[n]=\{1,2,\ldots,n\}$) and with edges corresponding to the pairs of points of unit distance; that is, $E=\{\{a,b\}:\|\makevec{p}_b-\makevec{p}_a\|=1\}$. To every edge $e=\{a,b\}\in E$ we assign a vector $\makevec{u}(e)$, in such a way that $\makevec{u}(e)=\pm(\makevec{p}_b-\makevec{p}_a)$, and the sign is chosen using some globally consistent rule, so that parallel edges get the same $\makevec{u}(e)$. For example, we may require that $\makevec{u}(e)$ lie in the closed upper halfplane minus the negative $x$-axis. Let $U:=\{\makevec{u}(e):e\in E\}$ be the \emph{unit direction set} of $P$, and we let $\makevec{u}_1,\makevec{u}_2,\ldots,\makevec{u}_k$ be an enumeration of all distinct elements of $U$, say in the lexicographic order. We call $\makevec{u}_1,\ldots,\makevec{u}_k$ the \emph{unit directions} of $P$ (under $\|.\|$). Then we define a coloring $c\:E\to[k]$ of the edges of the unit-distance graph, setting $c(e)=i$ if $\makevec{u}(e)=\makevec{u}_i$. (We note that $c$ need \emph{not} be a proper edge coloring, since there can be two edges with the same direction incident to a single vertex.) Finally, we record the geometric orientation of each edge. Namely, we define a mapping $\sigma\:E\to\{-1,+1\}$: For an edge $\{a,b\}\in E$ with $a<b$ we set $$ \sigma(\{a,b\})=\alterdef{+1&\mbox{ if $\makevec{u}(e)=\makevec{p}_b-\makevec{p}_a$},\\ -1&\mbox{ if $\makevec{u}(e)=\makevec{p}_a-\makevec{p}_b$.}} $$ The \emph{decorated unit-distance graph} of $P$ under $\|.\|$ is defined as the triple $\makevec{G}=\makevec{G}(\|.\|,P):=(G,c,\sigma)$. Now we define an \emph{abstract decorated unit-distance graph} as expected, i.e., as a triptuple $\makevec{G}=(G,c,\sigma)$, where $G$ is a graph with vertex set $[n]$ for some $n$, $c$ is a mapping $E(G)\to[k]$ for some $k$, and $\sigma$ is a mapping $E\to\{-1,+1\}$. We say that a sequence $P$ of distinct points in ${\mathbb{R}}^2$ is a \emph{realization} of an abstract decorated unit-distance graph $\makevec{G}$ under $\|.\|$ if $\makevec{G}$ is equal to the decorated unit-distance graph of $P$ under $\|.\|$. (We require equality to keep the definitions simple; we could as well introduce a suitable notion of isomorphism, but there is no need.) Here is the main result of this section. Roughly speaking, it tells us that if $\makevec{G}$ is a sufficiently dense abstract decorated unit-distance graph, then for every realization, the unit directions satisfy certain fixed linear dependences---some $\ell+1$ of the unit directions can be expressed using some other $\ell$ of the unit directions. \begin{lemma}\label{l:lindep} The following holds for a sufficiently large constant $C_0$. Let $\makevec{G}$ be an abstract decorated unit-distance graph with $n\ge 4$ vertices, at least $f(n):=C_0n\log n\log\log n$ edges, and $k$ colors. Then there exists an integer $\ell\ge 1$, a sequence $(i(1),i(2),\ldots,i(2\ell+1))$ of distinct indices in $[k]$, and linear maps $L_1,L_2,\ldots,L_{\ell+1}\:({\mathbb{R}}^2)^\ell\to{\mathbb{R}}^2$ such that for every realization $P$ of $\makevec{G}$ (under any norm), we have $$ \makevec{u}_{i(\ell+j)}=L_j(\makevec{u}_{i(1)},\makevec{u}_{i(2)},\ldots,\makevec{u}_{i(\ell)}), \ \ \ j=1,2,\ldots,\ell+1, $$ where $\makevec{u}_1,\ldots,\makevec{u}_k$ are the unit directions of~$P$. \end{lemma} \heading{Proof. } Let $\makevec{G}=(G,c,\sigma)$. In order to apply Proposition~\ref{p:}, we may need to prune the graph so that $c$ becomes a proper edge coloring. If $\makevec{G}$ has any realization at all, then, for geometric reasons, no color occurs on more than two edges incident to each vertex. Hence, for each $i$, the subgraph made of edges of color $i$ consists of paths and cycles, and so by deleting at most $\frac 23$ of the edges, we can turn this subgraph into a matching, and hence obtain a subgraph $\tilde G$ of $G$ with at least $\frac1{3}f(n)$ edges for which $c$ is a proper edge coloring. (By using more geometry, it is easily seen that it even suffices to delete only at most $\frac12$ of the edges, rather than $\frac23$.) Now we are ready to apply Proposition~\ref{p:} on the graph $\tilde G$ with the proper edge coloring $c$, and with $q=2.001$, say. This yields a subset $W\subseteq V(\tilde G)$ and a subset $I\subset[k]$ of colors, such that the subgraph $\tilde G[I,W]$ is connected, and $\tilde G[W]$ uses at least $2|I|+1$ colors. Let $J$ be a set of $|I|+1$ colors used on the edges of $W$ but not belonging to $I$. Now we can define the objects whose existence is claimed in the lemma. We set $\ell:=|I|$, let $(i(1),\ldots,i(\ell))$ be an enumeration of $I$, and let $i(\ell+1),\ldots,i(2\ell+1)$ be an enumeration of~$J$. Let us consider some color $j\in J$, and let $\{a,b\}$ be an edge of color $j$ in $\tilde G[W]$. Then there is a path $\pi$ from $a$ to $b$ in $\tilde G[W]$ whose edges have only colors in $I$, and for every realization $P$ of $\makevec{G}$, $\makevec{u}_j$ is a signed sum of the unit directions along this path. An example is given in Fig.~\ref{f:sigsum}: If $a=1$, $b=6$, $j=1$, the edge $\{1,6\}$ has sign $-1$, the path $\pi$ goes through the vertices $2,3,4,5$ in this order, and its edges have colors $2,2,3,4,2$ and signs $+1,+1,-1,+1,+1$, then $\makevec{u}_1=-3\makevec{u}_2+\makevec{u}_3-\makevec{u}_4$. This yields the desired linear maps $L_1,\ldots,L_{\ell+1}$, and the lemma is proved. \ProofEndBox\smallskip \labepsfig{sigsum}{Expressing $\makevec{u}_j$ in terms of the $\makevec{u}_i$, $i\in I$.} \section{Proof of Theorem~\ref{t:}} Let us call a norm $\|.\|$ on ${\mathbb{R}}^2$ \emph{bad} if $u_{\|.\|}(n)\ge f(n)=C_0n\log n \log\log n$ for some $n\ge 3$, and let $\MM\subseteq\BB$ be the set of all bad norms. We want to show that $\MM$ is meager, and thus we want to cover it by countably many nowhere dense sets. In our proof, the nowhere dense sets $\MM_{\makevec{G},\eta}$ are indexed by two parameters: $\makevec{G}$, which runs through all abstract decorated unit-distance graphs with $n$ vertices and at least $f(n)$ edges, $n=3,4,\ldots$, and $\eta$, which runs through all positive numbers of the form $\frac 1m$, $m$ an integer. To define $\MM_{\makevec{G},\eta}$, we first define that a realization $P$ of $\makevec{G}$ under a norm $\|.\|$ is \emph{$\eta$-separated} if for every two unit direction vectors $\makevec{u}_i,\makevec{u}_j$ of this realization, the lines spanned by $\makevec{u}_i$ and $\makevec{u}_j$ have angle at least $\eta$. Now $\MM_{\makevec{G},\eta}$ consists of all norms $\|.\|$ under which $\makevec{G}$ has an $\eta$-separated realization. It is easily checked that the $\MM_{\makevec{G},\eta}$ cover all of $\MM$. Indeed, for every bad norm $\|.\|$ we can choose $n$ and an $n$-point sequence $P$ with at least $f(n)$ unit distances. We define $\makevec{G}$ as the decorated unit-distance graph of $P$ under $\|.\|$. It remains to observe that, trivially, every realization of $\makevec{G}$ under some norm is $\eta$-separated for some $\eta>0$. Thus $\|.\|\in\MM_{\makevec{G},\eta}$. The main part of the proof consists of showing that each $\MM_{\makevec{G},\eta}$ is nowhere dense. Explicitly, this is expressed in the following lemma; once we prove it, we will be done with Theorem~\ref{t:} (the smoothness and strict convexity asserted in the theorem follows from Klee's result \cite{klee-mostsmooth} mentioned in the introduction, namely, that most norms are smooth and strictly convex). \begin{lemma}\label{l:nowd} Let $\makevec{G}$ be an abstract decorated unit-distance graph with $n$ vertices and at least $f(n)$ edges, let $B_0\in \BB$ be the unit ball of some norm, and let $\eta,\eps>0$. Then there exist $B\in \BB$ with $d_H(B,B_0)<\eps$ (where $d_H$ denotes the Hausdorff distance) and $\delta>0$ such that no $B'\in \BB$ with $d_H(B',B)<\delta$ belongs to $\MM_{\makevec{G},\eta}$. \end{lemma} \heading{Proof. } First we approximate $B_0$ by a $\makevec{0}$-symmetric convex polygon $B_1$ within Hausdorff distance at most $\frac\eps2$ from $B_0$. We make sure that all sides of $B_1$ are sufficiently short, so short that two lines through $\makevec{0}$ with angle at least $\eta$ never meet the same side of $B_1$. (If $B_0$ has straight segments in the boundary, we need to to ``bulge'' $B_1$ slightly; see Fig.~\ref{f:polapprox}.) \labepsfig{polapprox}{Approximating the unit ball $B_0$ by a convex polygon. } Let $s_1,s_2,\ldots,s_{2m}$ be the sides of $B_1$ listed in clockwise order, say, so that $s_{i}$ and $s_{m+i}$ are opposite (i.e., $s_{m+i}=-s_i$). Let $\lambda_i$ be the line spanned by $s_i$, and for a real parameter $t$, let $\lambda_i(t)$ be the line obtained by a parallel translation of $\lambda_i$ by distance $t$, where $t>0$ means translation away from the origin and $t<0$ translation towards the origin. We have $\lambda_{m+i}(t)=-\lambda_i(t)$. Let us consider an $m$-tuple $\makevec{t}=(t_1,\ldots,t_m)\in T_0:=[-\delta_0,\delta_0]^m$. For $\delta_0>0$ sufficiently small, the lines $\lambda_1(t_1),\ldots,\lambda_m(t_m),\lambda_{m+1}(t_1), \ldots,\lambda_{2m}(t_m)$ bound a symmetric convex polygon with $2m$ sides, which we denote by $B_1(\makevec{t})$. Moreover, for $\delta_0$ sufficiently small, $d_H(B_1(\makevec{t}),B_0)<\eps$, and the sides of $B_1(\makevec{t})$ are still short in the same sense as those of~$B_1$. Now we digress from geometry for a moment and we apply Lemma~\ref{l:lindep} to the abstract decorated unit-distance graph $\makevec{G}$. This yields an integer $\ell$, indices $i(1),\ldots,i(2\ell+1)$, and linear maps $L_1,\ldots,L_{\ell+1}$ as in the lemma. In order to make the notation slightly simpler, let us pretend that $i(j)=j$ for all $j=1,\ldots,2\ell+1$. Thus, for every realization of $\makevec{G}$, the unit directions $\makevec{u}_1,\ldots,\makevec{u}_{2\ell+1}$ satisfy the linear relations $\makevec{u}_{\ell+i}=L_i(\makevec{u}_1,\ldots,\makevec{u}_\ell)$, $i=1,2,\ldots,\ell+1$. Next, let us consider a particular realization of $\makevec{G}$ under the norm induced by $B_1(\makevec{t})$ for some $\makevec{t}\in T_0$. Each of the unit directions $\makevec{u}_1,\ldots,\makevec{u}_{2\ell+1}$ lies on the boundary of $B_1(\makevec{t})$, and thus on some line $\lambda_\alpha(t_\alpha)$. (Here we abuse the notation slightly, since the range of $\alpha$ is $[2m]$, while $\makevec{t}$ is indexed only by $[m]$, in order to preserve the symmetry of the polygon. So we make the convention that $t_{m+i}$ is the same as $t_i$.) Let $\alpha(i)\in [2m]$ be the index such that $\makevec{u}_i$ lies on $\lambda_{\alpha(i)}(t_{\alpha(i)})$ (if $\makevec{u}_i$ is a vertex of the polygon and thus lies on two of the lines, we pick one arbitrarily). Since the sides of $B_1(\makevec{t})$ are short, we have $\alpha(i)\ne\alpha(i')$ whenever $i\ne i'$, and also $\alpha(i)+m)\ne\alpha(i')$ (where $\alpha(i)+m$ is to be understood modulo $2m$). Let us call a mapping $\alpha\:[2\ell+1]\to[2m]$ an \emph{admissible assignment of lines} if it satisfies the condition in the previous sentence. Let us define a \emph{box} $T\subseteq T_0$ as a product of closed intervals with a nonempty interior; each box can be written as an $m$-dimensional ``interval'' $[\makevec{t}_{\rm min},\makevec{t}_{\rm max}]$. Our next goal is establishing the following claim. \begin{claim}\label{c:laim} There exists a box $\tilde T\subseteq T_0$ such that and for every admissible assignment of lines $\alpha$ and for every $\makevec{t}\in \tilde T$ there are no vectors $\makevec{u}_1,\ldots,\makevec{u}_{2\ell+1}\in{\mathbb{R}}^2$ such that each $\makevec{u}_i$ lies on the appropriate line, i.e., $\makevec{u}_i\in\lambda_{\alpha(i)}(t_{\alpha(i)})$, and the $\makevec{u}_i$ satisfy the linear relations $\makevec{u}_{\ell+i}=L_i(\makevec{u}_1,\ldots,\makevec{u}_\ell)$, $i=1,2,\ldots,\ell+1$. \end{claim} \heading{Proof of the claim. } We will kill all admissible assignments $\alpha$ one by one inductively, progressively shrinking the current box. The following statement allows us to make an inductive step: \emph{Let $T\subseteq T_0$ be a box, and let $\alpha$ be an admissible assignment of lines. Then there exists a box $T'\subseteq T$ such that for every $\makevec{t}\in T'$ there are no vectors $\makevec{u}_1,\ldots,\makevec{u}_{2\ell+1}\in{\mathbb{R}}^2$ with $\makevec{u}_i\in\lambda_{\alpha(i)}(t_{\alpha(i)})$ for all $i$ and with $\makevec{u}_{\ell+i}=L_i(\makevec{u}_1,\ldots,\makevec{u}_\ell)$, $i=1,2,\ldots,\ell+1$.} To prove this, let us consider a vector $\makevec{x}\in{\mathbb{R}}^{2\ell}$, which we think of as a concatenation of $\makevec{u}_1,\ldots,\makevec{u}_{\ell}$, and let us think of its components $x_i$ as unknowns. For each $i=1,2,\ldots,\ell$, the condition $\makevec{u}_i\in \lambda_{\alpha(i)}(t_{\alpha(i)})$ translates to a single linear equation for $\makevec{x}$, of the form $\makevec{a}_i^T\makevec{x}=b_i$, where the coefficient vector $\makevec{a}_i$ on the left-hand side doesn't depend on $\makevec{t}$, while $b_i=b_i(t_{\alpha(i)})$ is a \emph{nonconstant} linear function of $t_{\alpha(i)}$. Similarly, for $i=1,2,\ldots,\ell+1$, the condition $\makevec{u}_{\ell+i}\in \lambda_{\alpha(\ell+i)}(t_{\alpha(\ell+i)})$ together with $\makevec{u}_{\ell+i}=L_i(\makevec{u}_1,\ldots,\makevec{u}_\ell)$ translate to a similar linear equation $\makevec{a}_{\ell+i}^T\makevec{x}=b_{\ell+i}$, again with $\makevec{a}_{\ell+i}$ independent of $\makevec{t}$ and with $b_{\ell+i}=b_{\ell+i}(t_{\alpha(\ell+i)})$ a nonconstant linear function of $t_{\alpha(\ell+i)}$. Since the $\alpha(i)$ are all distinct, altogether we get that if the appropriate $\makevec{u}_1,\ldots,\makevec{u}_{2\ell+1}$ exist, then $\makevec{x}$ satisfies the system $A\makevec{x}=\makevec{b}$ of $2\ell+1$ linear equations with $2\ell$ unknowns, where $A$ is a fixed matrix and the right-hand side $\makevec{b}=\makevec{b}(\makevec{t})$ is a \emph{surjective} linear function ${\mathbb{R}}^{m}\to{\mathbb{R}}^{2\ell+1}$. Since we have more equations than unknowns, the system $A\makevec{x}=\makevec{b}$ has a solution only for $\makevec{b}$ contained in a proper linear subspace of ${\mathbb{R}}^{2\ell+1}$. Hence, by the surjectivity of $\makevec{b}(\makevec{t})$, the set of all $\makevec{t}\in{\mathbb{R}}^{m}$ for which $A\makevec{x}=\makevec{b}(\makevec{t})$ is unsolvable is a dense open subset of ${\mathbb{R}}^{m}$. From this the existence of the desired box $T'$ follows, and the Claim~\ref{c:laim} is proved. \ProofEndBox\smallskip \heading{Finishing the proof of Lemma~\ref{l:nowd}. } Let us consider the box $\tilde T=[\makevec{t}_{\rm min},\makevec{t}_{\rm max}]$ as in Claim~\ref{c:laim}. We set $\makevec{t}_{\rm mid}:= (\makevec{t}_{\rm min}+\makevec{t}_{\rm max})/2$, and we consider the polygons $B_{\rm in}:=B_1(\makevec{t}_{\rm min})$, $B_{\rm out}:=B_1(\makevec{t}_{\rm max})$, and $B:=B_1(\makevec{t}_{\rm mid})$; see Fig.~\ref{f:polygs}. We claim that $B$ is as in the lemma, i.e., no $B'\in\BB$ sufficiently close to $B$ belongs to $\MM_{\makevec{G},\eta}$. \labepsfig{polygs}{The polygons $B_{\rm in}$, $B_{\rm out}$, and $B$.} To see this, we note that every $B'$ sufficiently close to $B$ satisfies $B_{\rm in}\subseteq B'\subseteq B_{\rm out}$. For contradiction, we assume that there is an $\eta$-separated realization of $\makevec{G}$ under $B'$. Then the unit directions $\makevec{u}_1,\ldots,\makevec{u}_{2\ell+1}$ lie on the boundary of $B'$. The region $B_{\rm out}\setminus B_{\rm in}$ is naturally divided into $2m$ trapezoids $R_1,\ldots,R_{2m}$ belonging to the sides, as in Fig.~\ref{f:polygs1}. Each of $\makevec{u}_i$, $i=1,2,\ldots,2\ell+1$, lies in one of these trapezoids, let us call it $R_{\alpha(i)}$ (border disputes resolved arbitrarily). Since the considered realization is $\eta$-separated, no two of the $\makevec{u}_i$ share the same trapezoid, and also no two of these trapezoids are opposite to one another. So $\alpha$ defines an admissible assignment of sides. \labepsfig{polygs1}{Dividing the region $B_{\rm out}\setminus B_{\rm in}$ into trapezoids.} Let us consider the trapezoid $R_{\alpha(i)}$. As the line $\lambda_{\alpha(i)}(t)$ moves from the inner position (with $t=(\makevec{t}_{\rm min})_{\alpha(i)}$) to the outer position (with $t=(\makevec{t}_{\rm max})_{\alpha(i)}$), it sweeps the whole of $R_{\alpha(i)}$, and hence for some $t$ it contains $\makevec{u}_i$; let us denote this value of $t$ by $\bar t_{\alpha(i)}$. This defines $2\ell+1$ of the components of a vector $\bar\makevec{t}\in{\mathbb{R}}^m$. Let us set the remaining components to the corresponding components of $\makevec{t}_{\rm mid}$, say. Then $\bar\makevec{t}$ lies in the box $\tilde T$, and hence, by Claim~\ref{c:laim}, $\makevec{u}_1,\ldots,\makevec{u}_{2\ell+1}$ cannot lie on the corresponding lines. The resulting contradiction proves the lemma, and this also finishes the proof of Theorem~\ref{t:}. \ProofEndBox\smallskip \newcommand{\etalchar}[1]{$^{#1}$}
{ "timestamp": "2010-07-08T02:01:05", "yymm": "1007", "arxiv_id": "1007.1095", "language": "en", "url": "https://arxiv.org/abs/1007.1095", "abstract": "We prove that there exists a norm in the plane under which no n-point set determines more than O(n log n log log n) unit distances. Actually, most norms have this property, in the sense that their complement is a meager set in the metric space of all norms (with the metric given by the Hausdorff distance of the unit balls).", "subjects": "Combinatorics (math.CO)", "title": "The number of unit distances is almost linear for most norms", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9888419677654415, "lm_q2_score": 0.8267117962054048, "lm_q1q2_score": 0.8174873193346551 }
https://arxiv.org/abs/0812.0476
On approximations by shifts of the Gaussian function
The paper study the discrete sets of translations of the Gaussian function that span the spaces L1(R) and L2(R).
\section{Introduction.} The study of the sets of translations of functions that span $L^p(\mathbb R)$ spaces is a classical topic in harmonic analysis. One wants to determine under what conditions a sequence $\{\varphi(t-\lambda)_{\lambda\in\Lambda}$ span these spaces for a function $\varphi\in L^p(\mathbb R)$ and a set $\Lambda\subset\mathbb R$. Wiener's Tauberian theorem \cite{wiener} asserts that $\{\varphi(t-\lambda)\}_{\lambda\in\mathbb R}$ span $L^2(\mathbb R)$ if and only if the set of points at which the Fourier transform of $\varphi$ vanishes has measure $0$, and span $L^1(\mathbb R)$ if this set is empty. It is natural to consider this problem when $\Lambda$ is a discrete set. We give an overview of the results on this topic in the last section. This paper deals with the spanning properties of a discrete set of translations of the Gaussian function $\phi(t)=e^{-\pi t^2}$. This problem was first studied by Zalik in \cite{Zal}. He considered a more general point of view (functions like the Gaussian), and he obtained necessary and sufficient conditions for the set $\Lambda$ that depend on \begin{equation*} S(\varepsilon)=\sum_{\lambda\in\Lambda}\frac{1}{|\lambda|^{2+\varepsilon}}. \end{equation*} He proved that a necessary condition for $\{\phi(t-\lambda)\}_{\lambda\in\Lambda}$ to span $L^2(\mathbb R)$ is the divergence of $S(0)$ and that a necessary condition is the divergence of $S(\varepsilon)$ for some $\varepsilon>0$. The question is what happens with the sets such that $S(0)$ is divergent but $S(\varepsilon)$ is convergent for all positive $\varepsilon$. These kind of sets are known as sets of order $2$. The density of such a set is defined as: \begin{equation*} \Delta(\Lambda)=\lim_{r\to\infty}\frac{|\Lambda\cap(-r,r)|}{r^2}=\lim_{r\to\infty}\frac{n_{\Lambda}(r)}{r^2}. \end{equation*} In this expression $n_{\Lambda}$ is the counting function of $\Lambda$. If we write $\Lambda=\{\lambda_n\}_{n\in\mathbb N}$ with $|\lambda_n|\leqslant|\lambda_{n+1}|$, the density of $\Lambda$ coincides with $\lim_{n\to\infty}n/|\lambda_n|^2$. $S(0)$ is divergent and $S(\varepsilon)$ is convergent for those sets $\Lambda$ such that $0<\Delta(\Lambda)<\infty$. If $S(0)<\infty$ then the density will be $0$ and the divergence of $S(\varepsilon)$ implies infinite density. For sets without density (oscillation of the limit) we can define upper and lower versions, but we will not use this generalization in this note. Our principal result is the next one: \begin{thm}\label{teorema} Let $\Lambda\subset\mathbb R$ be a discrete set and $\phi(t)=e^{-\pi t^2}$ the Gaussian function. Let $\Lambda^+=\Lambda\cap[0,\infty)$ and $\Lambda^-=\Lambda\cap(-\infty,0]$ the positive and negative parts of $\Lambda$. \begin{enumerate} \item If $\Delta(\Lambda^+)<1/2$ and $\Delta(\Lambda^-)< 1/2$ then $\{\phi(t-\lambda)\}_{\lambda\in\Lambda}$ does not span $L^2(\mathbb R)$. \item If $\Delta(\Lambda^+)>1/2$ or $\Delta(\Lambda^-)>1/2$ then $\{\phi(t-\lambda)\}_{\lambda\in\Lambda}$ spans $L^2(\mathbb R)$. \end{enumerate} Both statements apply also to $L^1(\mathbb R)$. \end{thm} The proof for $L^2(\mathbb R)$ is in section \ref{seccio2} the one for $L^1(\mathbb R)$ in section \ref{seccio3}. In section \ref{seccio4} we find generalizations (that improve Zalik's result) and a brief discussion of related problems. \section{The $L^2(\mathbb R)$ case.}\label{seccio2} The key tool to study the problem is the Bargmann transform: \begin{equation*} Bf(z)=2^{\frac{1}{4}}\int_{\mathbb R}f(t)e^{2\pi tz-\pi t^2-\frac{\pi}{2}z^2}\,dt, \end{equation*} that gives (see \cite{Fol,Charly}) an isomorphism between $L^2(\mathbb R)$ and the Fock space: \begin{equation*} \mathcal{F}=\left\{f: f \text{ is entire and } \|f\|_{\mathcal{F}}^2=\int_{\mathbb C}|f(z)|^2e^{-\pi|z|^2}\, dm(z)<\infty\right\}, \end{equation*} where $dm$ is the area measure of the plane. This isomorphism allows us to transform our problem into a question of uniqueness sets of the Fock space, so we will be able to use theory of entire functions. Let us precise this statement. \begin{lem} Set $\phi(t)=e^{-\pi t^2}$ and let $\Lambda\subset\mathbb R$ be a discrete set. Then $\{\phi(t-\lambda)\}_{\lambda\in\Lambda}$ is a complete set for $L^2(\mathbb R)$ if and only if $\Lambda$ is an uniqueness set for the Fock space. \end{lem} \begin{proof} We observe that, for $x\in\mathbb R$, we have that \begin{equation*} Bf(x)=2^{\frac{1}{4}}e^{\frac{\pi}{2}x^2}\langle f,\phi_x\rangle \end{equation*} for any $f\in L^2(\mathbb R)$. Since any function of the Fock space can be represented in this way in the real line, the values there determine the function and the factor $2^{\frac{1}{4}}e^{\frac{\pi}{2}x^2}\neq 0$, the result follows by duality. \end{proof} In fact, the uniqueness sets are sets that can not be included in any zero set of the Fock space. There is no description of the zero sets of the Fock space. We will obtain some properties of these sets using standard theory of entire functions (Hadamard decomposition and Phragm\'{e}n--Lindel\"{o}f principle). Good references for this material are \cite{Lev} or \cite{Boas}. First one has to observe that functions of the Fock space have order at most $2$, and type at most $\pi/2$. Also, it is easy to check that any entire function of order less or equal to $2$ and type less than $\pi/2$ is in this space (see \cite{Zhu}). To prove the theorem one uses the properties of the indicator of an entire function $F$ of order $2$: \begin{equation*} h_F(\theta)=\limsup_{r\to\infty}\frac{\log|F(re^{i\theta})|}{r^2}. \end{equation*} The main properties that we will use are that the indicator (of an order $2$ function) is a $2$-trigonometrically convex function (and therefore continuous) and that the $\sup h_F$ is the type of $F$. We will also use properties of functions of completely regular growth ($\log |F(re^{i\theta})|=h_F(\theta)r^2+o(r^2)$ except for a $C^0$-set). For details and precise statements of all this see \cite{Boas,Lev}. \begin{proof}[Proof of (1)] The idea is to construct a function $F$ in the Fock space such that $F(\lambda)=0$ for all $\lambda\in\Lambda$. We suppose for simplicity that $\Delta(\Lambda^+)=\Delta(\Lambda^-)=\Delta$ (if not, one can add points to $\Lambda$, or alternatively make the same arguments but the calculations would not be so easy). We define the set $\Gamma=\Lambda\cup i\Lambda$. This set has angular density \begin{equation*} \Delta(\theta)=\lim_{r\to\infty}\frac{|\Gamma\cap\{|z|<r,\arg z<\theta\}|}{r^2}=\Delta\sum_{k=0}^3\chi_{[\frac{k\pi}{2},2\pi)}(\theta). \end{equation*} Moreover, \begin{equation*} \sum_{\gamma\in\Gamma,|\gamma|<r}\frac{1}{\gamma^2}=0\qquad\forall r. \end{equation*} In this conditions one can calculate the indicator function of the Weierstrass canonical product $\Pi_{\Gamma}(z)$ that vanishes in this set: \begin{equation*} h_{\Pi}(\theta)=\pi\Delta|\sin 2\theta|, \end{equation*} and $\Pi_{\Gamma}$ will be a function of completely regular growth. If $\Delta<1/2$ then $h_{\Pi}<\frac{\pi}{2}$ and $\Pi_{\Gamma}$ is in the Fock space. \end{proof} \begin{proof}[Proof of (2)] We can assume without loss of generality that $\Lambda\subset(0,\infty)$. Let $\Delta=\Delta(\Lambda)$. We define the Weierstrass function $\Pi(z)=\Pi_{\lambda\in\Lambda}(1-z^4/\lambda^4)$. This product defines an entire function ($\Lambda^4$ is a set of order $1/2$) with completely regular growth and its indicator function is $h_{\Pi}(\theta)=\Delta\pi|\sin 2\theta|$. We will compare this function with any other that vanishes in $\Lambda$. Let $F(z)$ be an entire function of the Fock space such that $F(\Lambda)=0$. As $\Lambda$ has order $2$, Hadamard theorem ensures us that $F$ has order at least $2$. We define: \begin{equation*} \varphi(z)=\frac{F(z)}{\Pi(z)}. \end{equation*} $\varphi$ is an holomorphic function in any sector of the form $-\pi/2<-\alpha\leqslant\arg z\leqslant\alpha<\pi/2$, and, as $h_{\pi}(\pm\alpha)>0$, in this sector it has order $2$ and mean type (see theorem 5 of section 11.3 in \cite{Lev}). We fix $\alpha>\pi/4$. In the sector $\arg z\in[-\alpha,\alpha]$ we write $F(z)=\varphi(z)\Pi(z)$. Using that the indicator function of a product of two functions is the sum of their indicator functions if one of them has completely regular growth, we have that: \begin{equation*} h_F(\theta)=h_{\varphi}(\theta)+h_{\Pi}(\theta)=h_{\varphi}(\theta)+\pi\Delta|\sin 2\theta|. \end{equation*} As $h_{\varphi}$ is a $2$-trigonometrically convex function (because it is the indicator of an order $2$ function), we have that $h_{\varphi}(\theta)+h_{\varphi}(\theta+\pi/2)\geqslant 0$. Then $h_{\varphi}(\pi/4)\geqslant 0$ or $h_{\varphi}(-\pi/4)\geqslant 0$. Thus: \begin{equation*} h_{F}(-\frac{\pi}{4})\geqslant \pi\Delta \qquad\text{ or }\qquad h_{F}(\frac{\pi}{4})\geqslant \pi\Delta. \end{equation*} If $F$ is in the Fock space we can deduce from the former equation that $\Delta\leqslant1/2$, or equivalently, that $F$ cannot be in the Fock space if $\Delta>1/2$. \end{proof} \section{The $L^1(\mathbb R)$ case.}\label{seccio3} By the duality principle, $\{\phi(t-\lambda)\}_{\lambda\in\Lambda}$ span $L^1(\mathbb R)$ if and only if given $f\in L^{\infty}(\mathbb R)$, if $\langle f,\phi(t-\lambda)\rangle=0$ for all $\lambda\in\Lambda$ then $f=0$. Then, as $2^{\frac{1}{4}}e^{\frac{\pi}{2}x^2}\langle f,\phi_x\rangle=Bf(x)$, one has to look at the image of $L^{\infty}(\mathbb R)$ for the Bargmann transform. However, the Bargmann transform is an isomorphism only in the space $L^2(\mathbb R)$. We will obtain the proof of (1) using convolution methods, because one can not ensure that the function $F$ constructed in the former section will belong to the image of $L^{\infty}(\mathbb R)$. The proof of (2) will use that $B\bigl(L^{\infty}(\mathbb R)\bigr)$ is included in an appropriate space. \begin{lem}\label{lemma1} Assume $h\in L^1(\mathbb R)$ and that $\widehat{h}(\xi)\neq 0$ for all $\xi$. If $f\in L^p(\mathbb R), 1\leq p\leq \infty$ and the convolution $f*h$ is zero then $f=0$. The same holds if $h\in L^2(\mathbb R)$ and $\widehat{h}(\xi)\neq 0$ almost everywhere. \end{lem} \begin{proof} For $1\leq p\leq 2$, the Fourier transform of $f$ is a function in $L^{q}, \frac 1p+\frac 1q=1$, and the Fourier transform of $f*h$ is $\widehat{f}\,\widehat{h}$, so the lemma follows. In the general case, we consider the closed subspace $E$ of $L^1(\mathbb R)$ consisting of functions $g$ such that $f*g=0$. Since $E$ is translation invariant and contains $h$, Wiener's Tauberian theorem implies that $E$ is the whole $L^1(\mathbb R)$, and this implies $f=0$. When $h\in L^2(\mathbb R)$ we use Beurling's theorem describing all closed translation-invariant subspaces of $L^2(\mathbb R)$ to reach the same result. \end{proof} \begin{lem}\label{generaciopertotselsp} Assume $h\in L^1(\mathbb R)\cap L^{\infty}(\mathbb R)$ and $\widehat{h}(\xi)\neq 0 $ for every $\xi$. Then if $\{\varphi(t-\lambda), \lambda\in\Lambda \}$ span $L^1(\mathbb R)$ then $\{(\varphi\ast h)(t-\lambda), \lambda\in\Lambda\}$ span $L^p(\mathbb R)$ for $1\leqslant p<\infty$. \end{lem} \begin{proof} By duality, $\{\varphi(t-\lambda), \lambda\in\Lambda \}$ span $L^p(\mathbb R)$ if and only if \begin{equation*} \widetilde{\varphi}*f(\lambda)=\int_{\mathbb R}f(t) \varphi(t-\lambda) \,dt=0 \quad \forall \lambda\in\Lambda \end{equation*} implies $f=0$ for $f\in L^q(\mathbb R)$. Here $\widetilde{\varphi}(t)=\varphi(-t)$. For $f\in L^q(\mathbb R)$, \begin{equation*} \int_{\mathbb R} f(t)\, (\varphi\ast h)(t-\lambda)\, dt=\int_{\mathbb R}(\widetilde{h}\ast f)(x)\varphi(x-\lambda)\, dx, \end{equation*} whence the result follows from lemma \ref{lemma1} (Hausdorff-Young theorem ensures us that $\varphi\ast h\in L^p(\mathbb R)$ and $\widetilde{h}\ast f \in L^{\infty}(\mathbb R)$). \end{proof} \begin{proof}[Proof of (1) for $L^1(\mathbb R)$] Define $\phi_a(t)=2^{1/4}e^{-a\pi t^2}$ for $a>1$ and observe that if $\frac{1}{a}+\frac{1}{b}=1$ then \begin{equation*} \phi_a\ast\phi_b(t)=\frac{(a-1)^{\frac{1}{2}}}{a}\phi(t). \end{equation*} If $\{\phi(t-\lambda)\}_{\lambda\in\Lambda}$ span $L^1(\mathbb R)$ then $\{\phi_a(t-\sigma)\}_{\sigma\in\frac{1}{\sqrt{a}}\Lambda}$ span it too, hence using lemma \ref{generaciopertotselsp} with $h=\phi_b$ we have that $\{\phi(t-\sigma)\}_{\sigma\in\frac{1}{\sqrt{a}}\Lambda}$ span $L^2(\mathbb R)$ for all $a>1$. As $\Delta(\frac{1}{\sqrt{a}}\Lambda^+)=\frac{1}{a}\Delta(\Lambda^+)$ (respectively for $\Lambda^-$), the statement follows from the $L^2(\mathbb R)$ case of theorem \ref{teorema}. \end{proof} \begin{lem} Let $f\in L^{p}(\mathbb R)$, $1\leqslant p\leqslant\infty$. The function: \begin{equation*} Bf(z)=2^{\frac{1}{4}}\int_{\mathbb R}f(t)e^{2\pi tz-\pi t^2-\frac{\pi}{2}z^2}\,dt \end{equation*} is an entire function and \begin{equation*} |Bf(z)|\leqslant\|f\|_p\|\phi\|_q e^{\frac{\pi}{2}|z|^2},\qquad \frac{1}{p}+\frac{1}{q}=1. \end{equation*} \end{lem} \begin{proof} We write the Bargmann transform in the following way: \begin{equation*} Bf(z)=2^{\frac{1}{4}}e^{\frac{\pi}{2}}\int_{\mathbb R}f(t)e^{-\pi(t-z)^2}\,dt. \end{equation*} The integral is well defined and defines an entire function for any $f\in L^p(\mathbb R)$. As \begin{equation*} e^{2\pi t z-\pi t^2-\frac{\pi}{2}z^2}=e^{\frac{\pi}{2}|z|^2}e^{-\pi(t-x)^2}e^{2\pi ity-\pi ixy}, \end{equation*} we can bound \begin{equation*} |Bf(z)|\leqslant e^{\frac{\pi}{2}|z|^2}\int_{\mathbb R}|f(t)||e^{-\pi(t-x)^2|}\,dt, \end{equation*} thus obtaining the statement by using H\"older inequality. \end{proof} \begin{proof}[Proof of (2) for $L^1(\mathbb R)$.] The proof for $L^2(\mathbb R)$ applies here without modification because $\Lambda$ is a uniqueness set for entire functions of order $2$ and type $\frac{\pi}{2}$. \end{proof} We observe that the theorem can be extended to $L^p(\mathbb R)$ for $1<p<2$ and the sufficient condition also for $p>2$. It is an open question whether the discrete sets that can be used to generate $L^p(\mathbb R)$ are independent of $p$. \section{Generalizations and comments.}\label{seccio4} Using the same ideas we can also prove \begin{thm} Fix $n,m\in \mathbb N$ and $0<a\leqslant b<\infty$. Let $\phi$ be a function such that \begin{equation*} \frac{A}{1+\xi^{2n}}e^{-a\pi\xi^2}\leqslant|\widehat{\varphi}(\xi)|\leqslant B(1+\xi^{2m})e^{-b\pi\xi^2}, \end{equation*} and $\Lambda$ a discrete set. \begin{enumerate} \item If $\Delta(\Lambda^+)<\frac{a}{2}$ and $\Delta(\Lambda^-)<\frac{a}{2}$ then $\{\varphi(t-\lambda)\}_{\lambda\in\Lambda}$ do not span $L^2(\mathbb R)$. \item If $\Delta(\Lambda^+)>\frac{b}{2}$ or $\Delta(\Lambda^-)>\frac{b}{2}$ then $\{\varphi(t-\lambda)\}_{\lambda\in\Lambda}$ span $L^2(\mathbb R)$. \end{enumerate} This implications also applies to $L^1(\mathbb R)$ if we assume $|\widehat{\varphi}'(\xi)|\leqslant C(1+\xi^{2m})e^{-b\pi\xi^2}$. \end{thm} To prove this theorem one must use theorem \ref{teorema} and \ref{generaciopertotselsp} with an appropriate $h$. The condition on $\widehat{\varphi}'$ ensures that $\varphi$ and $h$ will be in $L^1(\mathbb R)$. This result improves Zalik's one, i. e. the conditions of theorem $2$ of \cite{Zal} are stronger than the ones here. Zero sets of the Fock space have been studied in \cite{Zhu}. One can also find information about uniqueness sets of zero excess \cite{ALS,LSe99} and the description of sampling \cite{Jan94,Lyu,SW1} and interpolating sets \cite{S1} of this space, with applications to the Gabor transform (details and more information of this transform in \cite{Charly}). In \cite{Br} and \cite{BOU} one can find the characterization of the discrete sets $\Lambda\subseteq\mathbb R$ for which there exists a function $\varphi\in L^1(\mathbb R)$ with the property that $\{\varphi(t-\lambda)\}_{\lambda\in\Lambda}$ span $L^1(\mathbb R)$ as those having infinite Beurling-Malliavin density. In \cite{Ole} and \cite{OlU} one can find results proving that in $L^2(\mathbb R)$ there are more sets with this property, and that a characterization in terms of densities is not possible. The same problem studied here but for the Poisson function $1/(1+t^2)$ is totally solved in \cite{BrM07}, and generalized in \cite{AB}, where one can found the convolution lemmas stated here. The author does not know any other examples of functions for which such a study has been made. It would be interesting to study this problem for functions such as $e^{-|t|^n}$.
{ "timestamp": "2008-12-02T12:48:42", "yymm": "0812", "arxiv_id": "0812.0476", "language": "en", "url": "https://arxiv.org/abs/0812.0476", "abstract": "The paper study the discrete sets of translations of the Gaussian function that span the spaces L1(R) and L2(R).", "subjects": "Classical Analysis and ODEs (math.CA); Complex Variables (math.CV)", "title": "On approximations by shifts of the Gaussian function", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.986151391412456, "lm_q2_score": 0.82893881677331, "lm_q1q2_score": 0.8174591675567946 }
https://arxiv.org/abs/1912.07205
The Last Temptation of William T. Tutte
In 1999, at one of his last public lectures, Tutte discussed a question he had considered since the times of the Four Color Conjecture. He asked whether the 4-coloring complex of a planar triangulation could have two components in which all colorings had the same parity. In this note we answer Tutte's question to the contrary of his speculations by showing that there are triangulations of the plane whose coloring complexes have arbitrarily many even and odd components. We end up with a closely related conjecture, which is based on an extensive computation, and which claims that for every planar triangulation whose 4-coloring complex is disconnected has a component of even parity and one of odd parity.
\section{Introduction} In \cite{TutteNotes}, Tutte asked if there existed a triangulation of the plane whose 4-coloring complex had more than one component of the same parity. He based this question on numerous examples he examined over the course of many years. One obvious possibility is the icosahedron, whose 4-colorings form ten Kempe equivalence classes. However, Tutte found out that the icosahedron has a connected colouring complex. Nonetheless, he was resistant to making this question into a conjecture that no such examples exist, since, in his view, ``the data are too few to justify a Conjecture.'' However, he asked: ``If anyone knows of any case of two components of the same parity, I would be glad to hear of it.'' In this short note, we provide examples of 4-connected triangulations of the plane whose 4-coloring complexes have arbitrarily many components with odd colorings and arbitrarily many components with even colorings. Although this seems to resolve the Tutte problem, we end up with a closely related conjecture, which is based on an extensive computation, and which claims that for every planar triangulation whose 4-coloring complex is disconnected has a component of even parity and one of odd parity (see Conjecture \ref{conj:new}). \section{Coloring Complexes and the Parity of 4-Colorings} Given a 4-colorable graph $G$, an independent set of vertices $A \subseteq V(G)$ is said to be a \emph{color class} if it is one of the color classes for some 4-coloring of $G$. The \emph{$4$-coloring complex} $B(G)$ is the graph which has all the color classes of $G$ as its vertices and two vertices $C,D \in V(B(G))$ joined by an edge if the color classes $C$ and $D$ appear together in a 4-coloring of $G$. These graphs were introduced by Tutte \cite{Tutte1}, who discussed their basic structure. See Figure 2 for an example of a coloring complex. A coloring complex may be disconnected. To see which 4-colorings have their color classes in the same component of $B(G)$, the following notion can be used. We say that two 4-colorings $f$ and $f^{\prime}$ of $G$ are adjacent, $f \sim f^{\prime}$, if they have a common color class. Then we take the transitive closure of the adjacency relation and we say that two colorings are \emph{connected} if they are in the same equivalence class of the transitive closure. Since each coloring corresponds to a 4-clique in $B(G)$, colorings are connected if and only if their cliques are in the same component of the graph $B(G)$. In the same paper \cite{Tutte1}, Tutte introduced the notion of even and odd colorings, whose explanation requires a bit more background. Given a 4-colorable graph $G$, consider a 4-coloring $f$ of $G$, whose color classes are $A, B, C, D$. An \emph{AB-Kempe chain} in $G$ is a connected component of $G[A \cup B]$. We will denote by $p(A,B)$ the number of $AB$-Kempe chains in $G$. Notice that one can derive $2^{p(A,B)-1}-1$ new colorings of $G$ from $f$ by taking an arbitrary subset of $AB$-Kempe chains and swapping the colors on each $AB$-Kempe chain in the subset. This operation is called an \emph{$AB$-Kempe change}. Given these definitions, Tutte defines \begin{equation}\label{eq:1} J_{A}(f) = p(A,B) + p(A,C) + p(A,D) - p(B,C) - p(B,D) - p(C,D). \end{equation} Observe that, for a given 4-coloring $f$, the parity of $J_{A}(f)$ equals the parity of $J_{X}(f)$, for each $X \in \{ B,C,D \}$. This motivates calling a 4-coloring $f$ an \emph{even 4-coloring} if $J_{A}(f)$ is even and calling $f$ an \emph{odd 4-coloring} if $J_{A}(f)$ is odd. Moreover, following Tutte \cite{Tutte1}, we can easily show that every coloring in the same connected component of the 4-coloring complex $B(G)$ has the same parity if $G$ is a triangulation of the plane. We include a short proof for the sake of completeness. Given a particular 4-colouring $f$ of a planar triangulation $T$, we will denote by $A,B,C,D$ its color classes, by $p(A,B)$ the number of $AB$-Kempe chains, and by $e(A,B)$ the number of edges in $G[A\cup B]$. Finally, we will denote by $\deg(A)$ the sum of the degrees of all the vertices in $A$. The same notation applies for other choices of color classes. \begin{theorem}[Tutte \cite{Tutte1}] \label{thm: Tutte} Let $T$ be a triangulation of the plane with $n$ vertices and $A$ a color class of its $4$-coloring $f$. Then \begin{equation} \label{eq:2} J_{A}(f) = 2 \vert A \vert - \deg(A) + n - 3. \end{equation} \end{theorem} \begin{proof} Let $t=2n-4$ denote the number of facial triangles of $T$ and observe that the edges of the dual graph $T^*$ of $T$ that correspond to $AB$-edges and $CD$-edges form a perfect matching in $T^*$, which shows that \begin{equation} \label{eq:3} e(A,B) + e(C,D) = t/2 = n-2. \end{equation} Let us now consider the subgraph $T[A\cup B]$ and its faces in the plane. Since it has $|A|+|B|$ vertices, $e(A,B)$ edges, and $p(A,B)$ components, it is easy to see that it has precisely $e(A,B) - (|A|+|B|-p(A,B)) + 1$ faces. Each of the faces contains precisely one $CD$-Kempe chain, so we conclude that $p(C,D) = e(A,B) - (|A|+|B|-p(A,B))+1$. By exchanging the roles of $AB$ vs. $CD$, we get $p(A,B) = e(C,D) - (|C|+|D|-p(C,D)) +1$. By combining these two equations, using that $|A|+|B|+|C|+|D|=n$, and using (\ref{eq:3}), we obtain that \begin{equation} \label{eq:4} p(A,B)-p(C,D) = |A| + |B| - e(A,B) - 1. \end{equation} Finally, by using the same equations for $p(A,C)-p(B,D)$ and $p(A,D)-p(B,C)$ and taking their sum, we obtain: \begin{equation} \label{eq:5} J_{A}(f) = 3|A| +|B|+|C|+|D| - (e(A,B)+e(A,C)+e(A,D)) - 3 = 2 \vert A \vert - \deg(A) + n - 3. \end{equation} This completes the proof. \end{proof} The following corollary then immediately follows from the equation (\ref{eq:2}). \begin{corollary} \label{cor: Parity} If $f$ and $g$ are two 4-colourings of $T$ with a common colour class $A$, then $J_{A}(f) = J_{A}(g)$. \end{corollary} Now, by Corollary~\ref{cor: Parity}, any two 4-colorings in the same connected component of $B(T)$ will have the same parity. Consequently, we will say that components of $B(T)$ are \emph{even} or \emph{odd} according to the parity of their colourings. \section{4-Colorings of Triangulated Surfaces} Fisk \cite{Fisk1} introduced an alternative view of the parity of 4-colorings. We show how to reconcile this approach with Tutte's definition. Given a triangulation $T$ of an orientable surface with one of its orientations fixed, we can view a 4-coloring $f$ of $T$ as a simplicial mapping onto the boundary of the tetrahedron, which will be denoted as $K_{4}$. For each triangle $T_{i,j,l}$ of $K_{4}$, we consider the facial triangles in $T$ that are mapped onto $T$ in such a way that their orientation is preserved and those facial triangles whose orientation is reversed. Let $t_{ijl}^{+}$ $(t_{ijl}^{-})$ be the number of triangular faces in $T$ that are mapped onto $T_{ijl}$ with positive (negative) orientation. Then the value \[ \deg(f) = t_{ijl}^{+} - t_{ijl}^{-} \] is independent of the choice of the $ijl$ and is called the (\emph{homology}) \emph{degree of the mapping} $f : T \rightarrow K_{4}$. Of course, modulo 2, $\deg(f)$ is equal to $t_{ijl} = t_{ijl}^{+} + t_{ijl}^{-}$, which is the number of triangular faces of $T$ with colors $i, j, l$. If $T$ is a triangulation of a non-orientable surface, then we cannot define the values $t_{ijl}^{+}$ and $t_{ijl}^{-}$, but we can still define the parity (modulo 2) value of $t_{ijl}$ as $\deg(f)$. Now, reinterpreting Tutte in the context of higher genus surfaces, we first define the \emph{Euler genus} of a surface triangulation $T$ to be equal to twice the genus of the surface if it is orientable, and be equal to the crosscap number of the surface if it is non-orientable. Then we define \[ J_{A}(f) = 2 \vert A \vert - \deg(A) + n - 3 + g \] where $n=|V(T)|$ and $g$ is the Euler genus of $T$. \begin{theorem} \label{thm: genusparity} Let $T$ be a triangulation of a surface of Euler genus $g$ with $n=|V(T)|$ vertices. If $f$ is a $4$-coloring of $T$ and $A$ is one of its color classes, then \begin{equation*} J_{A}(f) \equiv \deg(f) + n - 3 + g \pmod 2, \end{equation*} Consequently, the homology degree $\deg(f)$ has the same parity as $\deg(A)$. \end{theorem} \begin{proof} We may assume that the color class $A$ corresponds to the color 4. Let $T_{123}$ be the triangle of $K_{4}$ corresponding to the other 3 colors. Then $\deg(f) \equiv t_{123} \pmod 2$. On the other hand, \begin{equation*} \deg(A) = \sum_{v \in A} \deg(v) = t_{124} + t_{134} + t_{234} = \vert F(T) \vert - t_{123}, \end{equation*} where $F(T)$ denotes the set of the triangular faces of $T$. Since $T$ is a triangulation, Euler's formula implies that $|F(T)| = \tfrac{2}{3}|E(T)| = 2n-4+2g$. Thus, \begin{flalign*} J_{A}(f) & = 2 \vert A \vert - \deg(A) + n - 3 + g \\ & = 2 \vert A \vert + t_{123} - \vert F(T) \vert + n - 3 + g \\ & \equiv t_{123} + n - 3 + g \equiv \deg(f) + n - 3 + g \pmod 2. \end{flalign*} \end{proof} \section{Coloring Complexes with Many Components} We have found two triangulations of the plane on 12 vertices which answer Tutte's question in the affirmative. Their 4-coloring complexes each have three components, from which it follows that two of these components must have the same parity. These triangulations are drawn in Figure 1. \begin{figure}[htb] \centering \subfloat[][Example 1]{\includegraphics[width=0.42\textwidth]{TutteCounterexample1.pdf}\label{<figure1>}} \qquad \subfloat[][Example 2]{\includegraphics[width=0.42\textwidth]{TutteCounterexample2.pdf}\label{<figure2>}} \caption{Triangulations whose 4-coloring complexes have three components. The half-edges sticking out form the edges 9-11 and 10-12, respectively.} \end{figure} We will now determine the 4-coloring complex of Example 1, and show that it has 3 connected components (the argument for Example 2 is similar). \begin{theorem} \label{thm: Example 1} The 4-coloring complex of the first example $G$ given in Figure 1(a) is the graph drawn in Figure 2. \end{theorem} \begin{proof} We need to determine all the 4-colorings $f$ of $G$. We start by precoloring the 4-cycle with vertices labelled 5, 6, 7 and 8. Up to relabelling, this 4-cycle can be precolored in 4 ways: \\[1mm] (a) $f(5) = 1, f(6) = 2, f(7) = 3, f(8) = 4$. \\ (b) $f(5) = 1, f(6) = 2, f(7) = 1, f(8) = 2$. \\ (c) $f(5) = 1, f(6) = 2, f(7) = 1, f(8) = 3$. \\ (d) $f(5) = 1, f(6) = 2, f(7) = 3, f(8) = 2$. \\[1mm] We consider these cases separately. (a) Suppose that $f(1) = 3$. Then $f(4) = 2$, $f(3) = 1$ and $f(2) = 4$. Alternatively, if $f(1) = 4$, then $f(2) = 1$, $f(3) = 2$, and $f(4) = 3$. Now, as $G$ is symmetric about the precolored 4-cycle, we also have exactly 2 ways of coloring the exterior of this 4-cycle. As a consequence, we find that there are exactly 4 colorings satisfying (a). Their color classes are:\\ $\{ \{ 3, 5, 11 \}, \{ 4, 6, 10 \}, \{ 1, 7, 9 \}, \{ 2, 8, 12 \} \}$, $\{ \{ 3, 5, 12 \}, \{ 4, 6, 11 \}, \{ 1, 7, 10 \}, \{ 2, 8, 9 \} \}$, \\ $\{ \{ 2, 5, 11 \}, \{ 3, 6, 10 \}, \{ 4, 7, 9 \}, \{ 1, 8, 12 \} \}$, $\{ \{ 2, 5, 12 \}, \{ 3, 6, 11 \}, \{ 4, 7, 10 \}, \{ 1, 8, 9 \} \}$. (b) We may assume that $f(1) = 3$. Then $f(2) = 4$, and so vertex 3 has no available color. Consequently, no 4-colorings of $G$ satisfy (b). (c) Suppose that $f(1) = 3$. Then $f(2) = 4$, so $f(3) = 2$. Thus, $f(4) = 4$. On the other hand, if $f(1) = 4$, then $f(2) = 3$. Consequently, $f(3) = 2$, and vertex 4 cannot be colored. Hence, this coloring does not occur. Thus, we find that there is exactly one coloring satisfying (c): $\{ \{ 5, 7 \}, \{ 3, 6, 11 \}, \{ 1, 8, 9 \}, \{ 2, 4, 10, 12 \} \}$. (d) This case is mirror-symmetric to (c). We obtain exactly one coloring satisfying (d): \\ $\{ \{ 3, 5, 11 \}, \{ 6, 8 \}, \{ 1, 7, 9 \}, \{ 2, 4, 10, 12 \} \}$. The six colorings of $G$ clearly make the coloring complex exhibited in Figure \ref{fig:2}. \end{proof} \begin{figure}[htb] \centering {\includegraphics[width=0.8\textwidth]{BEX1.pdf}} \caption{The coloring complex $B(G)$} \label{fig:2} \end{figure} The coloring complex $B(G)$ from Theorem~\ref{thm: Example 1} certainly has three components, two of which have even parity, while the third one has odd parity. This provides an affirmative answer to Tutte's question. The triangulation shown in Figure 1(b) is not isomorphic to the previous example (this is easy to see by noting that the first graph has two degree 5 vertices with two common neighbors of degree 6, while the the second graph does not) but their coloring complexes are isomorphic. It can be shown through computation that there are no such examples on 11 or fewer vertices. (Let us observe that using our Lemma \ref{lem: sum}, the computation can be reduced to 4-connected triangulations.) We can now easily construct infinitely many examples by successively adding vertices to a triangle in $G$. For example, consider the triangle 123 in $G$. If we add vertex 13 in the centre of this triangle, and join it to vertices 1, 2 and 3 by edges, then we obtain a new triangulation $G^{\prime}$ of the plane such that $B(G) \cong B(G^{\prime}$). As we can repeat this procedure indefinitely, this approach yields infinitely many examples of graphs which have two components with even parity. Indeed, with a bit more effort, we can obtain a stronger result in this direction. \begin{theorem} \label{thm: infinite} Let $k \in \mathbb{N}$. Then there are infinitely many triangulations of the plane, whose coloring complex has more than $k$ components with even colorings and more than $k$ components with odd colorings. \end{theorem} This theorem follows directly from Theorem~\ref{thm: Example 1} and the following lemma. However, this lemma requires some further notation. Let $G$ and $H$ be triangulations of surfaces of Euler genus $g_1$ and $g_2$, respectively. Let $T_1 = xyz$ ($T_2 = x'y'z'$) be a face in $G$ ($H$). By identifying $x$ with $x'$, $y$ with $y'$, and $z$ with $z'$, we obtain a new triangulation $G \bigtriangleup H$ with $\vert V(G) \vert + \vert V(H) \vert - 3$ vertices and with Euler genus $g_1+g_2$. Any 4-coloring of $G \bigtriangleup H$ induces a 4-coloring $f$ on $G$ and a 4-coloring $h$ on $H$, and then we use $f \bigtriangleup h$ to denote the coloring of $G \bigtriangleup H$. This gives a bijection between colorings of $G \bigtriangleup H$ and pairs $(f,h)$ of colorings of $G$ and $H$. \begin{lemma} \label{lem: sum} (a) If two 4-colorings $f \bigtriangleup h$ and $f^{\prime} \bigtriangleup h^{\prime}$ of $G \bigtriangleup H$ are adjacent in $B(G \bigtriangleup H)$, then $f \sim f^{\prime}$ in $B(G)$, and $h \sim h^{\prime}$ in $B(H)$. Conversely, if $f \sim f^{\prime}$ in $B(G)$, and $h \sim h^{\prime}$ in $B(H)$, then $f \bigtriangleup h \sim f^{\prime} \bigtriangleup h \sim f^{\prime} \bigtriangleup h^{\prime}$ in $B(G \bigtriangleup H)$. \\ (b) If $A$ is a color class of $f \bigtriangleup h$, then $J_{A}(f \bigtriangleup h) = J_{A \cap V(G)}(f) + J_{A \cap V(H)}(h)$. \end{lemma} \begin{proof} (a) By definition, $f \bigtriangleup h \sim f^{\prime} \bigtriangleup h^{\prime}$ means that $f \bigtriangleup h$ and $f^{\prime} \bigtriangleup h^{\prime}$ have a color class in common, say $A$. Then $A \cap V(G)$ is a color class that is common to $f$ and $f^{\prime}$, so $f \sim f^{\prime}$. Similarly, $h \sim h^{\prime}$. The reverse statement is equally clear. (b) Consider the color class $A$ of $f \bigtriangleup h$ which is disjoint from the new triangle we created by identifying vertices in $G$ and $H$. Recall that, by Corollary~\ref{cor: Parity}, computing $J_{A}(f \bigtriangleup h)$ for this color class is equivalent to computing $J_{C}(f \bigtriangleup h)$ for any color class $C$ of $f \bigtriangleup h$. We write $A_{G}$ for $A \cap V(G)$ and $A_{H}$ for $A \cap V(H)$. Then: \[ J_{A_{G}}(f) = 2 \vert A_{G} \vert - \deg_G(A_{G}) + n_1 - 3 + g_1, \] \[ J_{A_{H}}(h) = 2 \vert A_{H} \vert - \deg_H(A_{H}) + n_2 - 3 + g_2. \] Now, as $A_{G}$ and $A_{H}$ are disjoint, $\deg(A) = \deg_G(A_{G})+\deg_H(A_{H})$. We can use the above expressions to compute $J_{A}(f \bigtriangleup h)$: \begin{flalign*} J_{A}(f \bigtriangleup h) & = 2 \vert A \vert - \deg(A) + (n_1+n_2-3) - 3 + (g_1+g_2) \\ & = 2 (\vert A_{G} \vert + \vert A_{H} \vert) - \deg_G(A_{G}) - \deg_H(A_{H}) + (n_1-3+g_1) + (n_2-3+g_2) \\ & = J_{A_{G}}(f) + J_{A_{H}}(f). \end{flalign*} If the color class $A$ contains one of the identified vertices, the calculation is the same except that $|A| = |A_{G}| + |A_{H}| - 1$ and $\deg(A) = \deg_G(A_{G}) + \deg_H(A_{H}) - 2$, which yields the same final outcome. \end{proof} \section{4-connected Examples} Identification over a triangle produces triangulations that are not 4-connected. However, it may be that in the back of Tutte's mind was an implicit condition that the triangulations should be 4-connected. The following generalized examples show that one can also find 4-connected examples whose coloring complexes have many components (of the same parity). The two graphs $Q_1$ and $Q_1'$ in Figure 1 are just smallest examples in two infinite families of 4-connected triangulations. We define $Q_0,Q_1,Q_2,\dots$ as follows. To obtain $Q_k$ we take $k+2$ nested 4-cycles $D_i=a_ib_ic_id_i$ ($i=0,1,\dots,k+1$). We connect $D_i$ with $D_{i+1}$ ($i=0,1,\dots,k$) as shown in Figure \ref{fig:Qk}. To obtain $Q_k$, we add the edges $a_0c_0$ and $a_{k+1}c_{k+1}$. We also define $Q_k'$, which is obtained in the same way except that we add the edge $b_{k+1}d_{k+1}$ instead of $a_{k+1}c_{k+1}$. \begin{figure}[htb] \centering {\includegraphics[width=0.42\textwidth]{Qk.pdf}} \caption{Nested 4-cycles $D_0,D_1,D_2,D_3$. To obtain $Q_2$, we add the edges $a_0c_0$ and $a_3c_3$.} \label{fig:Qk} \end{figure} We say that a 4-coloring of $Q_k$ is of \emph{type I} if it uses all four colors on $D_0$, and of \emph{type II} if it uses only 3 colors on $Q_0$. Note that a coloring $f$ of type II has $f(b_0)=f(d_0)$ (and necessarily $f(a_0)\ne f(c_0)$). As we saw in the proof of Theorem \ref{thm: Example 1}, every 4-coloring of type I on $D_i$ can be extended in two different ways to a 4-coloring on $D_{i+1}$, and both extensions are of type I (use all four colors on $D_{i+1}$ in the same clockwise order as on $D_i$). Similarly, every 3-coloring of type II on $D_i$ can be extended in two different ways to a 3-coloring on $D_{i+1}$, and both extensions are of type II. Of course, when $i=k$, only one of these 3-colorings of type II yields a 4-coloring in $Q_k$. This implies that $Q_k$ has precisely $2^{k+1}$ 4-colorings of type I and has precisely $2^k$ 4-colorings of type II. Next, we can easily show that two 4-colorings of $Q_k$ of type I that have a color class in common must be identical (have all four color classes the same). In colorings of type I all color classes have a vertex in each $D_i$. A coloring of $Q_k$ of type II has two color classes that meet every $D_i$, while the other two color classes have two vertices in every second 4-cycle $Q_i$. Thus, if a color class of a type II coloring coincides with a color class of a type I coloring, then the color class is contained in precisely one coloring of type I and one coloring of type II. In this way, each coloring of type II has 2 color classes that coincide with two color classes of a 4-coloring of type I, and this correspondence is 1-1. The colorings $f$ of type I under this correspondence are precisely those fulfilling the following condition: either $\{f(b_1),f(d_1)\} = \{f(b_{k+1}),f(d_{k+1})\}$ (if $k$ is even) or $\{f(b_0),f(d_0)\} = \{f(b_{k+1}),f(d_{k+1})\}$ (if $k$ is odd). All of them are of odd parity and of even homology degree $\deg(f)$ (by Theorem \ref{thm: Tutte}, since the only vertices of odd degree are $a_0,c_0,a_{k+1}$ and $c_{k+1}$). Finally, there are precisely $2^k$ 4-colorings of type I that do not fulfill the condition stated in the previous paragraph, and we can show that each of them forms a separate component in the coloring complex $B(Q_k)$. To see this, consider a color class $C$ of a coloring $f$ of type I. It has exactly one vertex in each of the nested 4-cycles $D_i$. By fixing the other three colors on vertices of $D_0$, it is easy to see that having a coloring on $D_i$ and knowing the vertex in $D_{i+1}\cap C$, the colors uniquely extend to $D_{i+1}$, for $i=0,1,\dots,k$. This shows that $f$ cannot share a color class with another coloring of type I. Thus, excluding the $2^{k+1}$ color classes in 4-colorings of type I which are Kempe equivalent to colorings of type II, we obtain $2^{k+2}$ color classes in the $2^{k}$ remaining 4-colorings of type I, each of which then forms a separate component in the coloring complex $B(Q_k)$. Observe that all such colorings have even parity (odd homology degree). It is also easy to see that all colorings of type II (together with their mates of type I) form a single component in $B(Q_k)$. To prove this, observe that two colorings of type II with a common class that intersects each $Q_i$ must be identical. However, if they have a common color class containing $b_0$ and $d_0$, then there are 2 ways to color each $D_i$ with $i$ odd, so there are $2^t$ colorings of type II with the same color class, where $t = \lfloor (k+1)/2 \rfloor$. And if they share a color class containing $a_1$ and $c_1$, or $b_1$ and $d_1$, then the same holds with $t = \lfloor k/2 \rfloor$. In conclusion, $B(Q_k)$ has one (large) component containing all $2^{k+1}$ odd 4-colorings (they all have even homology degree) and $2^k$ components, each of which corresponds to a single coloring of even degree that is of type I. The same holds for $Q_k'$. \begin{theorem} \label{thm:Qk} The $4$-coloring complex of the $4$-connected triangulation $Q_k$ (and that of $Q_k'$) has $2^k + 1$ connected components. One of them corresponds to $2^{k+1}$ odd colorings, and each of the other $2^k$ components corresponds to a single even $4$-coloring (odd homology degree). \end{theorem} \section{Next Steps} At present, we have not found 5-connected triangulations of the plane whose 4-coloring complexes would have arbitrarily many components. However, we have found some 5-connected examples whose 4-coloring complexes have more than one component of the same parity. The three smallest triangulations of this kind are illustrated in Figure \ref{fig: Three 5-Connected Triangulations on 20 Vertices}. These are triangulations of the plane on 20 vertices, and their minimality follows from an extensive computation that we performed on 5-connected triangulations with at most 23 vertices. \begin{figure}[htp] \begin{minipage}{0.33\textwidth} \centering \includegraphics[width=0.8\textwidth]{Tutte5ConnCounterEx1Interior20.pdf} \end{minipage {\hfill\color{black}\vrule\hfill}% \begin{minipage}{0.33\textwidth} \centering \includegraphics[width=0.8\textwidth]{Tutte5ConnCounterEx2Interior20.pdf} \end{minipage {\hfill\color{black}\vrule\hfill}% \begin{minipage}{0.33\textwidth} \centering \includegraphics[width=0.8\textwidth]{Tutte5ConnCounterEx3Interior20.pdf} \end{minipage \medskip \begin{minipage}{0.33\textwidth} \centering \includegraphics[width=0.8\textwidth]{Tutte5ConnCounterEx1Exterior20.pdf} \end{minipage {\hfill\color{black}\vrule\hfill}% \begin{minipage}{0.33\textwidth} \centering \includegraphics[width=0.8\textwidth]{Tutte5ConnCounterEx2Exterior20.pdf} \end{minipage {\hfill\color{black}\vrule\hfill}% \begin{minipage}{0.33\textwidth} \centering \includegraphics[width=0.8\textwidth]{Tutte5ConnCounterEx3Exterior20.pdf} \end{minipage \caption{Three 5-connected triangulations on 20 vertices. Each one is represented by two triangulated octagons. In order to obtain the full graph from each pair of subgraphs, identify the vertices on the outer cycle of each subgraph according to their labeling.} \label{fig: Three 5-Connected Triangulations on 20 Vertices} \end{figure} The 4-coloring complex of our first example has three components, two odd and one even. The second example's 4-coloring complex has three even components and one odd component. Our final example's 4-coloring complex has two even and one odd component. There are also 6 non-isomorphic 5-connected examples on 21 vertices, 33 on 22 vertices and 66 on 23 vertices. Unfortunately, we do not have an infinite family. However, there is another interesting question about parity we can ask. If the 4-coloring complex of a triangulation of the plane has at least two components, must it have one component of even parity and one component of odd parity? Based on a large number of computations, we formulate the following conjecture. \begin{conjecture}\label{conj:new} Suppose that $T$ is a triangulation of the plane, and that its 4-coloring complex $B(T)$ has at least two components. Then $B(T)$ has a component of even parity and a component of odd parity. \end{conjecture} \section*{References} \bibliographystyle{plain}
{ "timestamp": "2019-12-17T02:19:14", "yymm": "1912", "arxiv_id": "1912.07205", "language": "en", "url": "https://arxiv.org/abs/1912.07205", "abstract": "In 1999, at one of his last public lectures, Tutte discussed a question he had considered since the times of the Four Color Conjecture. He asked whether the 4-coloring complex of a planar triangulation could have two components in which all colorings had the same parity. In this note we answer Tutte's question to the contrary of his speculations by showing that there are triangulations of the plane whose coloring complexes have arbitrarily many even and odd components. We end up with a closely related conjecture, which is based on an extensive computation, and which claims that for every planar triangulation whose 4-coloring complex is disconnected has a component of even parity and one of odd parity.", "subjects": "Combinatorics (math.CO)", "title": "The Last Temptation of William T. Tutte", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9861513897844354, "lm_q2_score": 0.8289388062084421, "lm_q1q2_score": 0.8174591557887059 }
https://arxiv.org/abs/1304.2488
Log-Quadratic Bounds for the Gaussian Q-function
We present bounds of quadratic form for the logarithm of the Gaussian Q-function. We also show an analytical method for deriving log-quadratic approximations of the Q-function and give an approximation with absolute error less than $10^{-3}$.
\section{Introduction} Approximations and bounds for the Gaussian Q-function have been studied extensively, motivated by the fact that the Q-function is not an elementary function \cite{mills26, gordon41, birnbaum42, tate53, boyd59, borj79, chiani03, deabreu09, chang11}\footnote{The Q-function is the tail probability for a standard normal random variable: \begin{eqnarray} Q(x) \triangleq \frac{1}{\sqrt{2\pi}} \int_x^\infty e^{-\frac{u^2}{2}} \mathrm{d} u. \end{eqnarray}}. There has been specific effort in deriving bounds that remain tractable under multiplication and exponentiation, especially those with a single exponential term. The classical example is the Chernoff bound \cite{chernoff52}, \begin{equation} Q(x) \le \frac{1}{2}e^{-\frac{x^2}{2}}, \quad x \ge 0. \end{equation} In recent work, C{\^o}t{\'e} et al. \cite{cote12} showed strong Chernoff-type lower bounds for the Q-function of the form \begin{equation} Q(x) \ge \alpha e^{-\beta x^2}, \quad x \in \mathbb{R}. \end{equation} Chernoff-type bounds are sufficient for many applications but may be loose for small values of $x$. L{\'o}pez-Ben{\'\i}tez and Casadevall \cite{lopez11} considered Q-function approximations of the log-quadratic form \begin{eqnarray} Q(x) &\approx& e^{-ax^2-bx-c}, \quad x \ge 0, \end{eqnarray} and determined the parameters $a,b,c$ using numerical fitting techniques for different ranges of $x$. We consider bounds and approximations of this form from an analytical perspective. The bounds are stated by the following theorem. \begin{thm} \begin{eqnarray} Q(x) &\ge& \frac{1}{2} e^{-\frac{1}{2}x^2 - \sqrt{\frac{2}{\pi}}x}, \quad x \ge 0, \\ Q(x) &\le& \frac{1}{2} e^{-\frac{1}{\pi}x^2 - \sqrt{\frac{2}{\pi}}x}, \quad x \ge 0. \end{eqnarray} \end{thm} \noindent We derive the following approximation, which has absolute error less than $10^{-3}$. \begin{eqnarray} Q(x) &\approx& \frac{1}{2} e^{-0.374 x^2 - 0.777 x}, \quad x \ge 0. \end{eqnarray} \section{Bounds and Approximations} We show the proof of Theorem 1 and then the derivation of the approximation. ~\\ ~\\ \textit{Proof of Theorem 1.} Let the standard normal density be given by \begin{equation} \phi(x) \triangleq \frac{e^{-\frac{x^2}{2}}}{\sqrt{2\pi}}. \end{equation} Define \begin{eqnarray} L(x) &\triangleq& \log(Q(x)), \\ \ell(x) &\triangleq& \frac{\partial L(x)}{\partial x} = -\frac{\phi(x)}{Q(x)}, \\ \ell'(x) &\triangleq& \frac{\partial^2 L(x)}{\partial^2 x} =\frac{\phi(x)Q(x)x-\phi(x)^2}{Q(x)^2} , \\ L_l(x) &\triangleq& -\frac{1}{2}x^2-\sqrt{\frac{2}{\pi}}x-\log(2), \\ \ell_l(x) &\triangleq& \frac{\partial L_l(x)}{\partial x} = -x-\sqrt{\frac{2}{\pi}}, \\ \ell_l'(x) &\triangleq& \frac{\partial^2 L_l(x)}{\partial x^2} = -1. \end{eqnarray} For the lower bound, we show that $L_l(x) \le L(x)$ for $x\ge0$. It is clear that $L_l(0) = L(0)$, so it is sufficient to show that $\ell_l(x) \le \ell(x)$ for $x \ge 0$. Again, we have $\ell_l(0) = \ell(0) = -\sqrt{\frac{2}{\pi}}$, so we are left to show that $\ell_l'(x) \le \ell'(x)$, or equivalently $\ell'(x) \ge -1$, for $x \ge 0$. The desired property \begin{eqnarray} \ell'(x) &=& \frac{\phi(x)Q(x)x-\phi(x)^2}{Q(x)^2} \ge -1 \end{eqnarray} is equivalent to the statement \begin{eqnarray} Q(x)^2+Q(x)\phi(x)x-\phi(x)^2 \ge 0. \end{eqnarray} The above inequality follows from a simple bound on Mill's ratio (e.g. \cite{sampford53}). The upper bound is found with similar reasoning. It is sufficient to show that $\ell'(x) \le -\frac{2}{\pi}$, which is equivalent to the property \begin{eqnarray} \frac{2}{\pi}Q(x)^2+Q(x)\phi(x)x-\phi(x)^2 \le 0. \end{eqnarray} An upper bound referred to in \cite{borj79} and \cite{dasgupta10} (as the Mitrinovi{\'c} inequality) is \begin{eqnarray} Q(x) \le \frac{2\phi(x)}{x+\sqrt{x^2+\frac{8}{\pi}}}. \end{eqnarray} The substitution yields \begin{eqnarray} \frac{2}{\pi}Q(x)^2+Q(x)\phi(x)x-\phi(x)^2 &\le& \frac{\left(\frac{8}{\pi}\right)\phi(x)^2}{\left(x+\sqrt{x^2+\frac{8}{\pi}}\right)^2} + \frac{2x\phi(x)^2}{x+\sqrt{x^2+\frac{8}{\pi}}} - \phi(x)^2 = 0. \end{eqnarray}\qed ~\\ \clearpage \noindent\textit{Q-function approximation}. The approximation is of the form \begin{eqnarray} Q(x) &\approx& e^{-ax^2-bx-c} ~\triangleq~ \hat{Q}(x). \end{eqnarray} for $x \ge 0$. We use the function \begin{eqnarray} \hat{L}(x) &\triangleq& \log(\hat{Q}(x)) ~=~ -ax^2-bx-c \end{eqnarray} and select parameters to approximately minimize the mean squared error between derivatives of $L(x)$ and $\hat{L}(x)$ for $0 \le x \le t$. Define \begin{eqnarray} \hat{\ell}(x,a,b) &\triangleq& \frac{\partial \hat{L}(x)}{\partial x} = -2ax-b, \\ \hat{\ell}'(x,a) &\triangleq& \frac{\partial^2 \hat{L}(x)}{\partial^2 x} = -2a. \end{eqnarray} The mean squared value for the difference of second derivatives of the two terms is proportional to \begin{eqnarray} \int_0^t (\ell'(x)-\hat{\ell}'(x,a))^2 \mathrm{d} x. \end{eqnarray} Differentiating with respect to $a$ and equating to zero gives an optimal value of $a$ for a given $t$: \begin{eqnarray} \int_0^t \ell'(x) \mathrm{d} x &=& \int_0^t \hat{\ell'}(x,a) \mathrm{d} x, \\ \int_0^t\ell'(x) \mathrm{d} x &=& \frac{\phi(0)}{Q(0)} - \frac{\phi(t)}{Q(t)} ~=~ \sqrt{\frac{2}{\pi}} - \frac{\phi(t)}{Q(t)}, \\ \int_0^t \hat{\ell'}(x,a) \mathrm{d} x &=& -2at, \\ a^*(t) &=& -\frac{1}{2t}\left (\sqrt{\frac{2}{\pi}} - \frac{\phi(t)}{Q(t)} \right ). \end{eqnarray} The mean squared value for the difference of first derivatives is proportional to \begin{eqnarray} \int_0^t (\ell(x)-\hat{\ell}(x,a,b))^2 \mathrm{d} x. \end{eqnarray} Differentiating with respect to $b$, equating to zero, and substituting $a = a^*(t)$ gives the optimal value of $b$ as a function of $t$, \begin{eqnarray} \int_0^t \ell(x) \mathrm{d} x &=& \int_0^t \hat{\ell}(x,a,b) \mathrm{d} x, \\ \int_0^t \ell(x)\mathrm{d} x &=& \log(Q(t)) - \log(Q(0)) ~=~ \log \left(2Q(t) \right), \\ \int_0^t \hat{\ell}(x,a,b)\mathrm{d} x &=& -at^2-bt, \\ b^*(t) &=& -a^*(t) t - \frac{1}{t} \log(2Q(t)), \\ b^*(t)&=& \frac{1}{2}\left (\sqrt{\frac{2}{\pi}} - \frac{\phi(t)}{Q(t)} \right ) - \frac{1}{t} \log(2Q(t)). \end{eqnarray} We simply use $c = \log(2)$ to match $Q(0) = \frac{1}{2}$. This gives the general approximation for a selected $t$ \begin{eqnarray} Q(x) \approx \frac{1}{2} \exp \left ({\frac{1}{2t}\left (\sqrt{\frac{2}{\pi}} - \frac{\phi(t)}{Q(t)} \right )x^2 + \left ( \frac{1}{t} \log(2Q(t)) - \frac{1}{2}\left (\sqrt{\frac{2}{\pi}} - \frac{\phi(t)}{Q(t)} \right ) \right )x} \right),\quad x \ge 0. \end{eqnarray} For a given $t$, an initial evaluation $Q(t)$ is required. By making the parameter $t$ large, the approximation becomes increasingly more accurate for large values of $x$ and less accurate for small values of $x$. Selecting $t = 1.295$ approximately minimizes the maximum absolute error $|Q(x) - \hat{Q}(x)|$ for all $x \ge 0$; the corresponding error is $9.485\times 10^{-4}$. This gives the approximation \begin{equation} Q(x) \approx \frac{1}{2} e^{-0.374 x^2 - 0.777 x}, \quad x \ge 0. \end{equation} \qed \bibliographystyle{jota}
{ "timestamp": "2013-04-10T02:02:19", "yymm": "1304", "arxiv_id": "1304.2488", "language": "en", "url": "https://arxiv.org/abs/1304.2488", "abstract": "We present bounds of quadratic form for the logarithm of the Gaussian Q-function. We also show an analytical method for deriving log-quadratic approximations of the Q-function and give an approximation with absolute error less than $10^{-3}$.", "subjects": "Probability (math.PR)", "title": "Log-Quadratic Bounds for the Gaussian Q-function", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9861513897844354, "lm_q2_score": 0.8289388040954684, "lm_q1q2_score": 0.817459153704994 }
https://arxiv.org/abs/2104.12373
Detecting random sets by samplings from their values
We examine the extent to which random samplings from the values of a random set, determine the distribution of the random set itself. We also comment on how, given the statistics of the sampling, to detect the distribution. Several methods are displayed, leading to positive results, counter-examples, and open problems.
\section{Introduction}\label{S1} Random sets are random variable with the variables being subsets of a given, say metric, space. The mathematics of random sets offers both geometrical and combinatorial challenges, emerging, at times, from applications, including in morphology, pattern analysis, data analysis, econometrics, and more. For the theory and applications of random sets, consult the comprehensive presentations, and references therein, in Chui, Stoyan, Kendall and Mecke \cite{ChStKeMe13}, Molchanov \cite{Mol17}, Molchanov and Molinari \cite{MolMol18}, Nguyen \cite{Ngu06}, and the workshop proceedings Goutsias, Mahler and Nguyen (editors) \cite{GoMaNg97}. The mathematical question we address in this paper has not been looked at, so it seems, in the literature. It alludes to a situation where samples from the values of the random set are given, but not the random set itself. Eventually, one may wish to exploit the samples from the values of the random sets for the inference of properties of the random set, or even to detect the distribution of the random set. To this end, the correspondence between the distribution of the random set and the distribution of the samples from its values, should be one-to-one. This is not always the case. A challenge, therefore, is to identify conditions under which the distribution of samples, which we refer to as the statistics of samples, determines the distribution of the random set. When the samples determine the distribution, algorithms are sought, that identify, or at least approximate, the distribution. In the next section we display, via some easy to follow examples and comments, what are the issues. In particular, we show that in order to achieve the goals, one-point sample may not be enough, while $m$-point statistics may determine the distribution. Then the challenge is to find the smallest such $m$. The inference issue itself is not address in the present paper. The results introduced here, in addition to the mathematical interest they offer, should help create a reliable inference procedures. In Section 3 we display the general technical setting, and mention some conditions that will be of help when reaching general positive results. In Section 4 we concentrate on the case of random intervals in the real line. The geometrical structure of intervals yields quite general positive results, yet, the general question concerning random intervals is still open. A procedure, named the peeling property, is introduced in Section 5. When holds, it helps in detecting the distribution of the random set. Finitely supported random sets are examined in Section 6. Combinatorial arguments help to state a general criterion, and to identify, in some specific families of random sets, the smallest $m$ for which the $m$-point statistics determines the distribution. In the closing section we examine the possibility to find a generalization of the interval-valued results, in convex-valued random sets in higher dimensions. A counter-example is given, yet, for some families of random convex-valued sets, an $m$ for which $m$-point statistics determines the distribution exists, and can be computed. A general criterion is still not available, \section{Comments and examples}\label{S2} We offer in this section some easy to follow examples, and comments, demonstrating the mathematical issues and challenges examined in the paper. The first example is concerned with a random compact interval in the real line. \begin{example}\label{Exa2.1} Consider two random intervals. One with distribution being the probability measures $\hbox{\boldmath\char'026}_1$ supported on the interval $[1,3]$ (namely, a random set with a constant value). The second, with distribution $\hbox{\boldmath\char'026}_2$ equally split between the intervals $[1,2]$ and $[2,3]$. Suppose that once an interval is chosen according to one of the distributions, the sampling from that interval is taken according to the normalized Lebesgue measure. \end{example} In the previous example, the distributions of the $1$-point sampling of the two distributions coincide. It is the uniform distribution over $[1,3]$. In fact, the same statistics results from many random intervals, e.g., any convex combination of the two distributions, namely $\hbox{\boldmath\char'026} = \alpha \hbox{\boldmath\char'026}_1 + (1 - \alpha)\hbox{\boldmath\char'026}_2 $, with $0 \le \alpha \le 1$. We conclude that the $1$-point sampling in this example does not determine that distribution of the random interval. The situation is different when the statistics of the $2$-point sampling is revealed, namely, at each stage two points are sampled, independently, from the interval. Then the statistics of the $2$-point sample for $\hbox{\boldmath\char'026}_1$, is the uniform distribution over the square $[1,3]^2$ in $R^2$, while the corresponding statistics for $\hbox{\boldmath\char'026}_2$ is the uniform distribution over the union of $[1,2]^2$ and $[2,3]^2$, see Figure 1. \begin{figure}[h] \centering \includegraphics[width=0.5\textwidth]{Figure1} \caption{} \end{figure} We conclude that in this example, between $\hbox{\boldmath\char'026}_1$ and $\hbox{\boldmath\char'026}_2$, the $2$-point statistics determines the distribution of the random interval. In fact, it is easy to see that the $2$-point statistics of any random interval with values being the three intervals in Example \ref{Exa2.1}, determines the distribution of the random interval. In Section 4 we verify, the not so obvious claim, that when the sampling is done according to the normalized Lebesgue measure, the statistics of the $2$-point sampling determines the distribution of any compact-valued random interval. A similar result holds for other sampling rules, as follows. \begin{example}\label{Exa2.2} Consider two random intervals with distributions being probability measures $\hbox{\boldmath\char'026}_1$ and $\hbox{\boldmath\char'026}_2$, where $\hbox{\boldmath\char'026}_1$ is equally distributed between the two intervals $[1,2]$ and $[3,4]$, and $\hbox{\boldmath\char'026}_2$ is equally split between $[1,3]$ and $[2,4]$. Suppose that the sampling rule is that given an interval, the two end points of it are chosen with equal probabilities. \end{example} The $1$-point statistics for both $\hbox{\boldmath\char'026}_1$ and $\hbox{\boldmath\char'026}_2$ in Example \ref{Exa2.2} is equally distributed among the points $1, 2, 3$ and $4$. In particular, this statistics does not determine the distribution. The statistics of the $2$-point sample, however, reveals the distribution of the random set. Indeed, the statistics of $\hbox{\boldmath\char'026}_1$ is equally distributed among the points $\{(x,y): x = 1, 2, y = 1,2\}$ and $\{(x,y): x = 3, 4, y = 3, 4\}$, and the statistics of $\hbox{\boldmath\char'026}_2$ is equally distributed among the points $\{(x,y): x = 1, 3, y = 1,3\}$ and $\{(x,y): x = 2, 4, y = 2, 4\}$. In Section 4 we prove a general result, relating to Example \ref{Exa2.2}, namely, that when the sampling is among the end points of the interval, the $2$-point statistics determines the distribution. Thus, the Examples \ref{Exa2.1} and \ref{Exa2.2} give rise to general positive results. A general characterization of random intervals for which the $2$-point statistics determines the random interval, is still not available. Once an $m$-point statistics determines a distribution of a random set, one may seek an algorithm to actually construct or at least to approximate the distribution. It is easy to construct this distribution in the two previous examples. We do display algorithms that approximate the distribution for a general random interval with the sampling rules as in the two example. It is clear that in both Examples \ref{Exa2.1} and \ref{Exa2.2}, the fact that the intervals are $1$-dimensional, plays a role (and the same will probably be for the aforementioned characterization). One may ask whether a similar general result holds for random $2$-dimensional compact convex-valued, random sets. We give a counter-example in Section 7, namely, a random set with convex and compact values, where for no $m$ the $m$-point statistics determines the distribution. Under further restrictions on the geometry of the values, one may establish positive results. We demonstrate it, yet, a general result for families in higher-dimensional spaces is still not known. \section{The setting}\label{S3} We display here, and comment on, the general technical framework we work in. Specific assumptions may be needed when specific results are presented. The notions we use are standard. Definitions, explanations, and elaborations, can be found in Molchanov \cite{Mol17}. Relevant information on probability measures can also be found in Ash \cite{Ash72}, and Billingsley \cite{Bil99}. We may abuse, slightly, rigorous terminology, as we refer to the notions of probability measure, distribution, and, at times, statistics, as synonymous terms. No confusion should arise. For convenience, we recall some definitions. Weak convergence of measures $\nu$, on a complete separable metric space $Y$, is characterized by the continuity of the mapping $\int_Y b(y) \nu(dy)$, into the real line, this for every $b(y)$, real valued, bounded, and continuous mapping. Another characterization, of the weak convergence of, say, $\nu_i$ to $\nu_0$, is that $\nu_i(B)$ converge to $\nu_0(B)$, whenever $B$ is a Borel set with $\nu_0(\partial B) =0$ (here $\partial B$ is the boundary of $B$). Although we are concerned primarily with probability measures, the weak convergence is applicable to general finite measures, and we shall use it that way. The weak convergence is metrizable, for instance, with the Prokhorov metric. We shall allude to a metric, since we do not need to adhere to a specific metric. The family of probability measures supported on a prescribed compact set in $Y$, is compact with respect to the weak convergence. The Hausdorff distance of two compact subsets, say $A$ and $B$, of a metric space $Y$, is denoted by $haus(A,B)$, and given by \begin{equation}\label{Eq3.1} haus(A,B) = \max (\max_{a \in A}\min_{b \in B} d(a,b), \max_{b \in B}\min_{a \in A} d(a,b)), \end{equation} where $d(a,b)$ is the distance in $Y$ between $a$ and $b$. The family of compact subsets of a compact set in $Y$, is compact with respect to the Hausdorff metric. A random set is a measurable mapping from a probability measure space into a collection of sets. In this paper, the values of the random sets are compact subsets of a complete separable metric space $X$ (often taken as $R^n$, the $n$-dimensional Euclidean space). We denote by $\Kscr$ the family of compact subsets of $X$, endowed with the Hausdorff metric. We consider $\Kscr$ with its Borel field. In fact, what matters to us is the distribution of the values of the random set, which is represented by a probability measure, say $\hbox{\boldmath\char'026}$, on $\Kscr$. As mentioned, our goal is to find out whether the distribution, say $\hbox{\boldmath\char'026}$, of a random set, is determined by samplings from the values of the random set. To that end, for each $K \in \Kscr$, a probability measure $P_K$ on $X$, supported on $K$ is prescribed, according to which, points in $K$ are sampled. We refer to $P_K$ as the {\sl sampling rule.} The following are two useful properties. \begin{property}\label{Prty3.1} The measures $P_K$ depend continuously on $K$. \end{property} \begin{property}\label{Prty3.2} For every compact set $C$, the mapping $P_K(C)$ is measurable in the variable $K$. \end{property} The continuity and measurability alluded to in the previous properties is with respect to the Hausdorff metric on $\Kscr$. They are useful as they allow a variety of analytical tools. The continuity of $P_K$ implies, via the aforementioned characterization of weak convergence, that for every compact set $C$ the mapping $P_K(C)$ is upper semi-continuous, hence measurable. Hence, the first property implies the second one. The continuity property may not hold even in natural examples. For instance, the Lebesgue measure (or rather, the normalized Lebesgue measure) on a set is not continuous with respect to the Hausdorff distance. It is easy to see, though, that the measurability property holds. Clearly, the continuity and measurability properties hold trivially when the distribution of the random set is supported on a finite number of sets. A useful class of sampling rules is as follows. \begin{definition}\label{Def3.3} We say that the sampling rule $P_K$ for the subsets of $X$ is {\sl regular}, if there exists a positive measure $\gamma$ on $X$ (not necessarily a probability measure), such that when $C$ is a subset of $K$ with $\gamma(K) > 0$, then $P_K(C) = {\gamma(C) \over \gamma(K)}$. \end{definition} A regular sampling rule $P_K$ may not be continuous. It is easy to see, however, that it posses the measurability Property \ref{Prty3.2}. The samples in our model are formed by a random choice of, say $m$, points in $K$, sampled consequently and independently according to the distribution $P_K$. Equivalently, for $K \in \Kscr$, denote by $K^m$ the $m$-times product $K^m = K \times K \times ... \times K$, and by $P_K^m$ the $m$-times product of $P_K$ on $K^m$. Then the $m$-point sample amounts to choosing a point in $K^m$, randomly with respect to $P_K^m$. Thus, by first choosing $K$ randomly according to $\hbox{\boldmath\char'026}$, then choosing an $m$-sample randomly according to $P_K^m$, we get a point in $X^m$. The distribution of these points on $X^m$, which is a probability measure on $X^m$, is the {\sl statistics of the $m$-sampling}, denoted $S_m(\hbox{\boldmath\char'026})$. Thus, to say that the statistics determines the distribution, on a given domain of distributions, means that $S_m(\hbox{\boldmath\char'026})$ is one-to-one in the domain. In general, given the distribution $\hbox{\boldmath\char'026}$ of a random set, and the sampling rules $P_K$, the statistics of the random $m$-samples from $K \in \Kscr$ is the probability measure $S_m(\hbox{\boldmath\char'026})$ on $X^m$, given by integrating the measures $P_K^m$ on $X^m$, according to $\hbox{\boldmath\char'026}$. Namely, for every Borel set $H \subseteq X^m$, the value $S_m(\hbox{\boldmath\char'026})(H)$ is given by \begin{equation}\label{Eq3.2} S_m(\hbox{\boldmath\char'026})(B) = \int_\Kscr P_K^m(H)~\hbox{\boldmath\char'026}(dK). \end{equation} For (\ref{Eq3.2}) to make sense, we need that for each Borel set $H$ the real-valued map $P_K^m(H)$ be measurable. For the results in this paper, it is enough that the Borel field on $X^m$ be generated by sets $H$ of the form $H = B_1 \times ... \times B_m$, where each $B_i$ is a compact subset of $X$, and since each measure on $X^m$ is determined by its values on such sets, it is enough to consider (\ref{Eq3.2}) on such sets (then the measurability for every Borel $H$ can be easily deduced). The statistics $S_m$ depends, of course, on the probability measures $P_K$. We suppress this dependence from the notation. No confusion is likely to arise. As stated, the $m$-point statistics is a probability measure on $X^m$, namely, on ordered strings of points in $X$. It is also possible to consider the sample as a subset of $X$ of size $m$. Simple combinatorial adjustments to the arguments should then be done. We do not address this possibility here. In this paper we address primarily the determination issue. The possibility to construct the random set, is displayed in comments that follow the positive results below. The following observations are listed for further reference. \begin{lemma}\label{Lem3.4} Let $\Kscr_0$ be a compact collection of compact subsets of $X$. Then, \\ $(1)$ If for each compact set $C$ the real valued mapping $P_K(C)$ is measurable, then for every $m$, the mapping $P_K^m(H)$ is measurable for every set $H$ which is the product of compact sets (namely, of the form $H = C_1 \times ... \times C_m$, where each $C_i$ is a compact subset of $X$). \\ $(2)$ If $P_K$ is continuous on $\Kscr_0$, then $P_K^m$ is continuous on the space of $K^m$, for $K \in \Kscr_0$. Consequently, for each compact set $C$, the real valued mapping $P_K(C)$ is measurable, and, for every set $H$ of the form $H = C_1 \times ... \times C_m$, where each $C_i$ is a compact subset of $X$, $P_K^m(H)$ is measurable. \\ $(3)$ If $P_K$ is continuous on $\Kscr_0$, then $S_m(\cdot)$ is continuous (as a mapping from probability measures on $\Kscr_0$, to the space of probability measures on $X^m$). \end{lemma} \begin{proof} Item $(1)$ is trivial, as $P_K^m(H)$ is the product of $m$ measurable functions. When $P_K$ is continuous, the continuity of $P_K^m$ can be deduced from an iterated Fubini's Theorem. Also, it is easy to see that for each compact set $B$ the function $P_k(B)$ is upper semicontinous, hence measurable, and likewise for $P_K^m (H)$. For verifying $(3)$, we show that if $\mu_j$ converge weakly to $\mu_0$, then $S_m(\mu_j)$ converge weakly to $S_m(\mu_0)$. For convenience, we show it first for $m = 1$. The latter convergence holds if $S_1(\mu_j)(B)$ converge to $S_1(\mu_0)(B)$ whenever $S_1(\mu_0)(\partial B) = 0$. For such a set $B$, the equality, in view of (\ref{Eq3.2}), implies that $P_K(\partial B) = 0$, except possibly on a set in $\Kscr_0$ of $\mu_0$-measure zero. Denote the latter set by $\Kscr_1$. The continuity of $P_K$ implies then that $P_K(B)$ is continuous in $K$, except possibly at the set $\Kscr_1$. Now, cover the set $\Kscr_1$ by a compact neighborhood, say $\Uscr_\varepsilon$ (i.e., a closure of an open set), whose $\mu_0$-measure is small, say $\varepsilon$. The mapping $P_K(B)$ can now be changed on $\Uscr_\varepsilon$ to make a continuous function bounded by $1$, say $P_K(B)_\varepsilon$ on $\Kscr_0$. The convergence of $\mu_j$ to $\mu_0$ implies then that $\int_{\Kscr_0} P_K(B)_\varepsilon \mu_j(dK)$ converge to $\int_{\Kscr_0} P_K(B)_\varepsilon \mu_0(dK)$. Since $\varepsilon$ is arbitrarily small, and since $\mu_j$ converge weakly to $\mu_0$ and $\mu_0(\Uscr_\varepsilon) \le \varepsilon$, it follows that $\int_{\Kscr_0} P_K(B) \mu_j(dK)$ converge to $\int_{\Kscr_0} P_K(B) \mu_0(dK)$. The latter convergence is what we need to verify for $S_1(\mu_j)$ to converge weakly to $S_1(\mu_0)$. The arguments for the case $m > 1$ are the same. This completes the proof. \end{proof} \begin{corollary}\label{Cor3.5} Let $\Kscr_0$ be a compact collection of compact subsets of $X$, and let $\Mscr_0$ be a compact set of distributions of random sets, all supported on $\Kscr_0$. Suppose that for each $\hbox{\boldmath\char'026} \in \Mscr_0$ the statistics $S_m(\hbox{\boldmath\char'026})$ determines $\hbox{\boldmath\char'026}$, and that $S_m(\hbox{\boldmath\char'026})$ is continuous (in $\hbox{\boldmath\char'026}$). Then the inverse of $S_m(\cdot)$ is continuous (from the space of probability measures on $X^m$ to probability measures on $\Kscr_0$). \end{corollary} \begin{proof} It is an easy exercise in point-set topology to show that the inverse of a one-to-one continuous mapping on a compact set, is continuous, \end{proof} \section{Random intervals} \label{S4} In this section we examine random sets that take values in the collection $\Iscr$ of intervals $[x,y]$, with $x \le y$, in the real line. Here are some useful observations. It is convenient to identify the space $\Iscr$ of intervals $[x,y]$, with the subset of the plane $R^2$ given by $\{(x,y): x \le y\} $. Notice that then the Hausdorff distance between intervals coincides with the $\max$ distance on $R^2$, namely, the distance derived from the norm $\Vert (x,y) \Vert_\infty = \max\{\vert x \vert, \vert y \vert \}$. Employing this isometry, the distribution of a random interval is now a probability measure on the subset of $R^2$ above the diagonal. Such a probability measure is determined by its values on rectangles in the plane, see Figure 2. Furthermore, it is enough to consider rectangles in the plain where the statistics $S_2(\hbox{\boldmath\char'026})$ assigns zero probability to the lines that define them (i.e., the dashed lines in Figure 2, given by $\{(x,y) : x = x_1\},~\{(x,y) : x = x_2\}$, etc.). \begin{figure}[h] \centering \includegraphics[width=0.5\textwidth]{Figure2} \caption{} \end{figure} Given an interval $I_0 = [x_0,y_0]$, the family of intervals which strictly contain it, is the set in $R^2$ given by $\{(x,y) : x < x_0, y > y_0 \}$, namely, the open upper left quadrant from $(x_0,y_0)$. Denote the open upper left quadrant from the point $(x,y)$ by $U(x,y)$. Then, given a distribution $\hbox{\boldmath\char'026}$ of a random interval, and a rectangular $B$ determined by $x_1, x_2, y_1, y_2$, where the statistics $S_2(\hbox{\boldmath\char'026})$ assigns zero probability to the lines that define $B$, the measure $\hbox{\boldmath\char'026}(B)$ satisfies \begin{equation}\label{Eq4.1} \hbox{\boldmath\char'026}(B) = \hbox{\boldmath\char'026}(U(x_2,y_1)) - \hbox{\boldmath\char'026}(U(x_1,y_1)) - \hbox{\boldmath\char'026}(U(x_2,y_2)) + \hbox{\boldmath\char'026}(U(x_1,y_2)). \end{equation} Hence, we have verified the following claim. \begin{observation}\label{Obs4.1} Let $\hbox{\boldmath\char'026}$ be the distribution of a random interval, and let $S_2(\hbox{\boldmath\char'026})$ be its $2$-point statistics. The $2$-point statistics determines the distribution if for every interval $[x,y]$ it determines $\hbox{\boldmath\char'026}(U(x,y))$. \end{observation} Here is our first positive result. Notice that, here, the sampling rule given by the normalized Lebesgue measure has the continuity property \ref{Prty3.1}. \begin{theorem}\label{The4.2} Suppose that the sampling rule from intervals is the normalized Lebesgue measure. Let $\Mscr_0$ be a compact family of distributions of random intervals, all supported on one compact collection of intervals. Then, for $\hbox{\boldmath\char'026}$ in $\Mscr_0$, the statistics of the $2$-point sample $S_2(\hbox{\boldmath\char'026})$ determines $\hbox{\boldmath\char'026}$, and the inverse of the one-to-one map $S_2(\cdot)$, is continuous. \end{theorem} \begin{proof} Once we verify that $S_2(\hbox{\boldmath\char'026})$ determines $\hbox{\boldmath\char'026}$, namely, $S_2(\cdot)$ on $\Mscr_0$ is invertible, the continuity of the inverse follows from Lemma \ref{Lem3.4} and Corollary \ref{Cor3.5}. In view of Observation \ref{Obs4.1}, in order to verify that $S_2(\hbox{\boldmath\char'026})$ determines $\hbox{\boldmath\char'026}$, it is enough to show that for every interval $[x_0,y_0]$, the statistics $S_2(\hbox{\boldmath\char'026})$ determines $\hbox{\boldmath\char'026}(U(x_0,y_0))$. It is enough to consider the case $y_ 0 > x_0$. Indeed, the definition (\ref{Eq3.2}) of the statistics $S_2(\hbox{\boldmath\char'026})$ implies that the contribution to $S_2(\hbox{\boldmath\char'026})$ of the restriction of $\hbox{\boldmath\char'026}$ to the line $\{(x,x): x \in R\}$ (namely, the degenerate intervals), coincides with the distribution $\hbox{\boldmath\char'026}$ itself. Now, given such a point $(x_0,y_0)$, consider the rectangular in $R^2$ given by \begin{equation}\label{Eq4.2} B_\varepsilon = [x_0 - \varepsilon, x_0] \times [y_0 ,y_0 + \varepsilon]. \end{equation} Denote by $\lambda(I)$ the Lebesgue measure, i.e. the length, of the interval $I$. According to (\ref{Eq3.2}), the contribution of an interval $I = [x,y]$ in $U(x_0,y_0)$ to the value of the statistics $S_2(\hbox{\boldmath\char'026})$ of $B_\varepsilon$, is $\varepsilon^2 \lambda(I)^{-2} \hbox{\boldmath\char'026}(dI)$, this unless $x_0 - x < \varepsilon$ or $y - y_0 < \varepsilon$. The latter cases may introduce an error $e_\varepsilon(I)$. This error is not greater than $\varepsilon^2 \lambda(I)^{-2} \hbox{\boldmath\char'026}(dI)$, and recall that $\lambda(I)$ is bounded away from $0$ on $U(x_0,y_0)$. Thus, we get that \begin{equation}\label{Eq4.3} S_2(\hbox{\boldmath\char'026})(B_\varepsilon) = \int_{U(x_0,y_0)} (\varepsilon^2 \lambda(I)^{-2} + e_\varepsilon(I)) \hbox{\boldmath\char'026}(dI). \end{equation} Dividing by $\varepsilon^2$, and realizing that $e_\varepsilon(I)$ is zero except being bounded on a set with $\hbox{\boldmath\char'026}$-measure converging to $0$ as $\varepsilon \to 0$, we get \begin{equation}\label{Eq4.4} \lim_{\varepsilon \to 0}\varepsilon^{-2} S_2(\hbox{\boldmath\char'026})(B_\varepsilon) = \int_{U(x_0,y_0)} \lambda(I)^{-2} \hbox{\boldmath\char'026}(dI). \end{equation} The left hand side of (\ref{Eq4.4}) is a function of the $2$-point statistics of the random interval. Therefore, this statistics determines the measure of $U(x_0,y_0)$ according, however, to $\lambda(I)^{-2} \hbox{\boldmath\char'026}(dI)$. Since this holds for any $U(x_0,y_0)$, it follows that the $2$-point statistics determines the measure $\lambda(I)^{-2} \hbox{\boldmath\char'026}(dI)$ on the entire set of intervals satisfying $x < y$. The latter measure clearly determines $\hbox{\boldmath\char'026}(dI)$. This completes the proof. \end{proof} \begin{comment}\label{Com4.3} The proof of the preceding theorem suggests a way to compute an approximation of the random interval, given its $2$-point statistics. It is, however, a cumbersome way. More efficient avenues are welcome. \end{comment} Here is our second positive result. It is related to Example \ref{Exa2.2}. Notice that, also here, the continuity Property \ref{Prty3.1} holds. \begin{theorem}\label{The4.4} Suppose that the sampling from the values of the random interval is carried out such that one of the two ends of the interval $[x,y]$ is chosen, with probabilities $\alpha(x,y)$ and $\beta(x,y) = 1 - \alpha(x,y)$, with $0 < \alpha(x,y) < 1$ and continuous. Let $\Mscr_0$ be a compact family of distributions of random intervals, all supported on one compact collection of intervals. Then, for $\hbox{\boldmath\char'026}$ in $\Mscr_0$, the statistics of the $2$-point sample $S_2(\hbox{\boldmath\char'026})$ determines $\hbox{\boldmath\char'026}$, and the inverse of the one-to-one map $S_2(\cdot)$ on $\Mscr_0$, is continuous. \end{theorem} \begin{proof} Once we verify that $S_2(\hbox{\boldmath\char'026})$ determines $\hbox{\boldmath\char'026}$, namely, $S_2(\cdot)$ is invertible, the continuity follows from Lemma \ref{Lem3.4} and Corollary \ref{Cor3.5}. Consider an interval $I = [x,y]$ in the support of $\hbox{\boldmath\char'026}$, such that $x < y$. Its contribution to the statistics $S_2(\hbox{\boldmath\char'026})$ is supported on the four points in $R^2$ given by $(x,x), (x,y), (y,y)$ and $(y,x)$, with weights $\alpha(x,y)^2$, $\alpha(x,y)\beta(x,y)$, $\beta(x,y)^2$ and $\alpha(x,y)\beta(x,y)$ respectively, weighted, however, by $\hbox{\boldmath\char'026}$, namely, by the infinitesimal $\hbox{\boldmath\char'026}(dI)$ (which becomes $\hbox{\boldmath\char'026}(I)$ if $I$ is an atom of the distribution). Furthermore, since only the end points of the intervals are sampled, the contribution to the statistics of a point $(x,y)$ with $x < y$, comes only from the infinitesimal measure of the interval $[x,y]$. Hence, for a Borel subset $B$ in the region $\{(x,y): x < y\}$, we have \begin{equation}\label{Eq4.5} S_2(\hbox{\boldmath\char'026})(B) = \int_B \alpha(x,y)\beta(x,y) \hbox{\boldmath\char'026}(dI). \end{equation} The latter measure on $\{(x,y): x < y\}$ gives rise to computing $\hbox{\boldmath\char'026}$ on $\{(x,y): x < y\}$, namely \begin{equation}\label{Eq4.6} \hbox{\boldmath\char'026}(B) = \int_B (\alpha(x,y)\beta(x,y))^{-1} S_2(\hbox{\boldmath\char'026})(d(x,y)). \end{equation} In particular, the statistics $S_2(\hbox{\boldmath\char'026})$ determines the distribution $\hbox{\boldmath\char'026}$ on $\{(x,y): x < y\}$. Now, each interval $I = [x,y]$ with $x < y$, contributes to the statistics on $\{(x,y): x = y\}$, namely, the values $\alpha(x,y)^2 \hbox{\boldmath\char'026}(dI)$ on $(x,x)$ and the value $\beta(x,y)^2 \hbox{\boldmath\char'026}(dI)$ on $(y,y)$. By removing the latter statistics from the diagonal $\{(x,y): x = y\}$, we are left on the latter line with the statistics emanating from the value of $\hbox{\boldmath\char'026}$ on the diagonal, namely on degenerate intervals. Clearly, the latter value coincides with the, given, statistics on the diagonal. This completes the proof. \end{proof} \begin{comment}\label{Com4.5} For the former sampling rule, (\ref{Eq4.6}) provides a way to compute the distribution of the random interval on non-degenerate intervals, given the $2$-point statistics. The only information needed is the sampling rule. In particular, there is no need to know beforehand what the support of the random interval is. \end{comment} We introduce now a useful, and natural, assumption concerning random intervals. \begin{assumption}\label{Ass4.6} For every $I =[a.b]$, the support of the probability measure $P_I$ contains the end points $a$ and $b$. \end{assumption} The role for the assumption is clear. If samples cannot get close to the end points of the interval, then (unless additional assumptions are made) it is not possible to distinguish between the interval and a shorter one. The sampling rules in the previous two positive results satisfy the assumption. A natural question arises, namely, is the generalization of the previous theorems to arbitrary sampling rules possessing the continuity property, true? Likewise in regard to general regular sampling rules, that satisfy Assumption \ref{Ass4.6}. We verify now a general result for random intervals with finite support. Another result toward the general one is provided in the next section (see Proposition \ref{Pro5.5}). The general case is, however, left open, \begin{theorem}\label{The4.7} Let $\hbox{\boldmath\char'026}$ be a probability measure on $\Iscr$, with finite support, say, supported on $\{I_1, ..., I_l\}$, with corresponding weights $\{\alpha_1, ..., \alpha_l\}$ (i.e., $\alpha_j \ge 0$ and the sum of the $\alpha_j$ is $1$). Then the statistics of the $2$-point sample $S_2(\hbox{\boldmath\char'026})$, determines $\hbox{\boldmath\char'026}$. \end{theorem} \begin{proof} The union of the sets $I_j^2$ is a union of squares in $R^2$, symmetric with respect to the line $x = y$ (where $R^2$ is the space of pairs $(x,y)$, Figure 1 portrays the union of squares in the case of Example \ref{Exa2.1}). Denote the union of the finite number of squares on which $S_2(\hbox{\boldmath\char'026})$ is supported by $V(\hbox{\boldmath\char'026})$. We claim that in any such $V(\hbox{\boldmath\char'026})$, there is a point in $R^2$, say $(x_0,y_0)$, such that a neighborhood of $(x_0,y_0)$ in $V(\hbox{\boldmath\char'026})$, say $N(x_0,y_0)$, belongs to only one of the squares $I_j^2$, say $I_1^2$, and that $P_{I_1}^2(N(a_0,b_0)) >0$. To see this, just pick $x_0$ to be the smallest $x$ among the finite number of intervals $[x,y]$ in the support of $\hbox{\boldmath\char'026}$, and if $x_0$ belongs to more than one interval, pick $y_0$ to be the largest $y$ among these intervals. (In Figure 1, if all three squares participate in the union, this point would be $(1,3)$.) The condition $P_{I_1}^2(N(a_0,b_0)) >0$ is met since both $x_0$ and $y_0$ are in the support of $P_{I_1}$. Also notice that the point $(x_0,y_0)$ in $R^2$ determines the square $I_1^2$. The point $(x_0,y_0)$ can be detected using only the statistics $S_2(\hbox{\boldmath\char'026})$, with no need to know beforehand on what intervals the distribution $\hbox{\boldmath\char'026}$ is supported. Now, since the neighborhood $N(x_0,y_0)$ is part of only the square $I_1^2$, the value of the statistics $S(\hbox{\boldmath\char'026})$ on this neighborhood is given by $\alpha_1 P_{I_1}^2(N(a_0,b_0))$. The coefficient $\alpha_1$ is now defined by the value the statistics assigns to $N(x_0,y_0)$ and the sampling rule $P_{I_1}$. We can consider now the random set and its $2$-point statistics, when $I_1$, and the corresponding components of $\hbox{\boldmath\char'026}$ and $S_2(\hbox{\boldmath\char'026})$, are removed. We are left with a random set supported on $l - 1$ intervals. Repeating the argument $l-1$ times, determines the entire distribution $\hbox{\boldmath\char'026}$. \end{proof} \begin{comment}\label{Com4.8} The proof of the preceding result provides an algorithm to compute the distribution of the random interval, given its $2$-point statistics. The only information needed is the sampling rule, and that the random interval is supported on a finite set of intervals. Indeed, given Assumption \ref{Ass4.6}, the set $V(\hbox{\boldmath\char'026})$ reveals, one step at the time, the support of $\hbox{\boldmath\char'026}$. \end{comment} We may refer to the main tool in the previous proof as {\sl peeling}. We have used peeling in a previous proof. It can be considered in a more general framework, as suggested in the next section. \section{The general peeling property}\label{S5} Let $\Kscr$ be a collection of compact subsets of $X$, and let $P_K$ be the associated sampling rule. For a finite collection $\Kscr_0$ of compact sets in $\Kscr$, denote by $V_m(\Kscr_0)$ the union, in $X^m$, of the sets $K^m$ for $K \in \Kscr_0$. \begin{definition}\label{Def5.1} We say that a collection $\Kscr_1$ of compact sets in $\Kscr$ has the {\sl $m$-point peeling property}, if for any finite sub-collection $\Kscr_0$ of $\Kscr_1$, there is a subset $H$ of $V_m(\Kscr_0)$, such that $H$ is included in only one of the sets $K^m$ for $K \in \Kscr_0$, and $P_K^m(H) > 0$. \end{definition} A collection of intervals has the $2$-point peeling property, as demonstrated and used in Section 4. We encounter more examples later in the paper. \begin{proposition}\label{Pro5.2} Let $\hbox{\boldmath\char'026}$ be the distribution of a random set, supported on a finite collection $\Kscr_1$ of compact sets. If $\Kscr_1$ has the $m$-point peeling property, then $S_m(\hbox{\boldmath\char'026})$ determines the distribution $\hbox{\boldmath\char'026}$. \end{proposition} \begin{proof} Suppose that the random set assigns the weight $\alpha_i$ to $K_i \in \Kscr_1$. Suppose that $K_1$ is a set in the support of $\hbox{\boldmath\char'026}$, such that a subset, say $H$, of $K_1^m$ has a positive $P_{K_1}^m$ measure, and $H$ does not intersect any of the sets $K_j^m$, for $j \ne 1$. Let $h = S_m(\hbox{\boldmath\char'026})(H)$. The choice of $H$ and (\ref{Eq3.2}) imply that $h = \alpha_1 P_{K_1}^m(H)$. This determines the value of $\alpha_1$. Consider now $\Kscr_1$ without the set $K_1$, with its weight $\alpha_1$. Then the $m$-point statistics is the former one without $\alpha_1 P_{K_1}^m$. We are left then with a smaller collection of sets, still possessing the $m$-point peeling property. After repeating the process a finite number of times, all the weights that determine $\hbox{\boldmath\char'026}$ are revealed. This completes the proof. \end{proof} \begin{comment}\label{Com5.3} The preceding proof suggests an algorithm to compute the distribution of $\hbox{\boldmath\char'026}$, subject, however, to the conditions set in Comment \ref{Com4.8}. \end{comment} Here is a result on general distributions with the peeling property. \begin{proposition}\label{Pro5.4} Let $\Mscr_0$ be a compact family of distributions of random sets, all supported on a compact collection of compact sets $\Kscr_1$. Suppose $\Kscr_1$ has the $m$-point peeling property. Also suppose that each $\hbox{\boldmath\char'026} \in \Mscr_0$ can be approximated, arbitrarily close, by a finitely supported random set in $\Mscr_0$. Suppose that the inverse of $S_m^{-1}(\hbox{\boldmath\char'027})$ on the image of distributions with finite support $\hbox{\boldmath\char'027}$ in $\Mscr_0$ (the existence of the inverse is guaranteed by the previous proposition), is uniformly continuous. Then the $m$-point statistics $S_m(\hbox{\boldmath\char'026})$ of any distribution in $\Mscr_0$ on $\Kscr_1$ determines $\hbox{\boldmath\char'026}$, and its inverse is continuous. \end{proposition} \begin{proof} Suppose two distributions $\hbox{\boldmath\char'026}_1$ and $\hbox{\boldmath\char'026}_2$ yield the same $m$-point statistics, say $S_m$. The distributions can be approximated arbitrarily close by distributions, say $\hbox{\boldmath\char'027}_1$ and $\hbox{\boldmath\char'027}_2$, each supported on a finite collection of compact sets. The continuity of the $m$-point statistics (see Lemma \ref{Lem3.4}) implies that $S_m(\hbox{\boldmath\char'027}_1)$ and $S_m(\hbox{\boldmath\char'027}_2)$ are both arbitrarily close to $S_m$. This contradicts the assumed uniform continuity. Hence the $m$-point statistics on $\Kscr_1$ is a one-to-one mapping. The continuity of its inverse follows, see Corollary \ref{Cor3.5}. \end{proof} Notice that in the previous proof, the uniform continuity of the inverse of the statistics of the finitely supported distributions, is needed only for approximations of the distributions in $\Mscr_0$. The previous result is relevant to the open problem suggested in the previous section, concerning random intervals. Indeed, the problem will be resolved if the inverse of the $2$-point statistics of random intervals with a finite number of values (whose existence is verified in Theorem \ref{The4.7}), is uniformly continuous. Here is a partial result in this direction. \begin{proposition}\label{Pro5.5} Let $\Iscr_0$ be a compact family of intervals with the property that for no interval, say $[x,y]$, in $\Iscr_0$, a neighborhood $[x-\varepsilon, y+\varepsilon]$ is included in another interval in $\Iscr_0$. Let the sampling rule $P_I$ on $\Iscr_0$ satisfy the continuity property \ref{Prty3.1} and the end points Assumption \ref{Ass4.6}. Then the $2$-point statistics $S_2(\hbox{\boldmath\char'026})$ of a distribution $\hbox{\boldmath\char'026}$ on $\Iscr_0$ determines $\hbox{\boldmath\char'026}$, and the inverse of $S_2(\cdot)$ is continuous. \end{proposition} \begin{proof} We utilize the observations made in Section 4. An interval $[x, y]$ can be viewed as the point $(x,y)$ in $R^2$, and its contribution to the statistics $S_2(\hbox{\boldmath\char'026})$ is via the rectangle $[x, y]^2$ in $R^2$. The non-inclusion property among intervals, implies that for each interval $[x, y]$ in the support of $\hbox{\boldmath\char'026}$, the point $(x, y)$ is on the line that forms the upper and left boundary of the support of $S_2(\hbox{\boldmath\char'026})$. In particular, $\hbox{\boldmath\char'026}$ is supported on this line. A finitely supported approximation of $\hbox{\boldmath\char'026}$ exists, supported on the same line. It is clear then that the inverse of the statistics of the finitely supported approximations are uniformly continuous. In view of Proposition \ref{Pro5.4}, the statistics is a one-to-one mapping on $\Mscr_0$ and its inverse is continuous. \end{proof} \section{Finitely supported random sets}\label{S6} We start with a general combinatorial criterion for the determination of a finitely supported random set, by the statistics of samplings. Additional structure will help to compute the smallest $m$ for specific families of random sets. Let $K_1, K_2, ..., K_L$ be $L$ compact sets in a metric space $X$. The distribution $\hbox{\boldmath\char'026}$ of a random set with values in this finite collection is given by a vector $< \alpha >$ with $L$ coordinates, all greater or equal to $0$, sum up to $1$. We think of $<\alpha>$ as a row vector (and denote its transpose by $<\alpha>^T$). Thus, the $i$-th coordinate of $< \alpha >$ is the probability that $K_i$ will be sampled We examine the $m$-point sampling of $<\alpha>$. The mapping $S_m(\hbox{\boldmath\char'026})$ is linear in $<\alpha>$, as is evident from $(\ref{Eq3.2})$. Furthermore, the latter formula reveals the role of $\hbox{\boldmath\char'026}$ in the generation of the $m$-point statistics, as follows. The probability measure $S_m(\hbox{\boldmath\char'026})$ is determined by its values on sets in $X^m$ of the form $B_1 \times ... \times B_m$, where each $B_i$ is a Borel subset in $X$ such that if $B_i$ intersects one of the sets, say $K$, in the support, then it is included in that $K$. Indeed, any probability measure on a product space is determined by its values on product of sets, and any such product is determined by products of the form just described. Now, given such a product, we have \begin{equation}\label{Eq6.1} S_m(\hbox{\boldmath\char'026})(B_1 \times ... \times B_m) = \sum_{j =1}^L \alpha_j P_{K_j}(B_1)\cdot ... \cdot P_{K_j}(B_m). \end{equation} Suppose now that the sampling rule is regular (see Definition \ref{Def3.3}), and let $\gamma$ be the positive measure that governs the regularity. It follows that (\ref{Eq6.1}) can be written as \begin{equation}\label{Eq6.2} S_m(\hbox{\boldmath\char'026})(B_1 \times ... \times B_m) = \gamma(B_1)\cdot ... \cdot \gamma(B_m) <\beta> <\alpha>^T, \end{equation} where $<\beta>$ is a row vector that depends on the sets $\{B_1, ..., B_m\}$. The vector $<\beta>$ has $L$ entries. The $j$-th coordinate of $<\beta>$ is equal to $0$ if one of the $B_i$ is not included in $K_j$. Otherwise, the entry is equal to $\gamma(K_j)^{-m}$. The number of distinct vectors $<\beta>$ may be large, depending on the combinatorial structure of intersections among the different sets in the support of $\hbox{\boldmath\char'026}$. The number of linearly independent vectors among them does not exceed $L$, since each vector has $L$ coordinates. Let $l_0$ be the maximal number of linearly independent vectors, and let $<\beta>_k$ for $k = 1, ..., l_0$ be $l_0$ linearly independent vectors. Arrange the $l_0$ vectors as an $l_0 \times L$ matrix, denoted ${\bf B}$. Denote by ${\bf B}_1$ the matrix obtained when a row vector with entries equal to $1$ is added to ${\bf B}$. \begin{theorem}\label{The6.1} Let $\Mscr_0$ be the family of probability measures supported on the sets $\{K_1, ..., K_L\}$. Suppose that the sampling is carried out according to a regular sampling rule $\gamma$. The statistics $S_m(\hbox{\boldmath\char'026})$ determines $\hbox{\boldmath\char'026}$ if and only if the matrix ${\bf B}_1$ has rank $L$. \end{theorem} \begin{proof} Suppose ${\bf B}_1$ has rank less than $L$. Then there exists a non-zero vector $<\eta>$ such that ${\bf B}_1<\eta>^T $ is the origin. In particular, since all the entries in the last row of ${\bf B}_1$ are equal to $1$, the sum of the coordinates of $<\eta>$ is $0$. Choose now a probability vector $<\alpha>$ with positive entries. Then, for $\varepsilon > 0$ small enough, both $<\alpha> +\varepsilon <\eta>$ and $<\alpha> -\varepsilon <\eta>$ are positive vectors. Both are probability vectors (since the sum of the coordinates of $<\eta>$ is $0$). The operation of ${\bf B}$ on both probability vectors yields the same outcome, hence yields the same $m$-point statistics. This verifies the ``only if'' part. If ${\bf B}_1$ has rank equal to $L$, the image by ${\bf B}_1$ of two distinct probability vectors cannot be identical. Since for any probability vector, the image by ${\bf B}_1$ has its last coordinate equal to $1$, the image by ${\bf B}$ of two distinct probability vectors cannot be identical. Since distinct products of ${\bf B}_1 <\alpha>^T$ result in distinct statistics, the proof is complete. \end{proof} \begin{comment}\label{Com6.2} The proof of Proposition \ref{The6.1} is not constructive. It is possible, however, to derive the distribution $<\alpha>$ from the $m$-point statistics is case ${\bf B}$ has full rank, or when ${\bf B}_1$ is $L \times L$ and has full rank. Indeed, Choose $l_0$ strings of the form $B_1, ..., B_m$, each related to a row in ${\bf B}$, or, to a row in ${\bf B}_1$, except the last one, respectively. Consider the diagonal matrix ${\bf D}_1$, with entries $T_J = \gamma(B_1)\cdot ... \cdot \gamma(B_m)$ in the diagonal, or in the first $L-1$ positions of the diagonal, and $1$ in the last, respectively. The vector $S_{n-1}(\hbox{\boldmath\char'026})(T_J)$ is now equal to ${\bf B}_1 {\bf D}_1 <\alpha>^T $. Inverting ${\bf B}_1 {\bf D}_1 $, yields $<\alpha>$ in terms of the $m$-point statistics of $\hbox{\boldmath\char'026}$. \end{comment} A trivial example of the previous considerations is the case where the sets $\{K_1, ..., K_L\}$ are disjoint. Then, clearly, the $1$-point statistics suffices to determine the distribution, as can also be deduced from the construction of the matrix ${\bf B}$. We now work out a more involved example, for which the exact bound $m$ for the statistics that determines the distribution is computed. The example will serve in the next section, in the study of random convex-valued sets. \begin{example}\label{Exa6.3} Let $G_0$, and $G_1, ..., G_n$ be $n+1$ compact subsets of a metric space, mutually disjoint. We consider the family of random sets that take as values the $2^n$ subsets, marked as $K_1, ..., K_{2^n}$, each is the union of $G_0$ and a subset of $G_1, ..., G_n$. Thus, a set $K$ in the support of the distribution of the random set, corresponds to a subset, say $J$, of $\{1, ..., n\}$ (including the empty set). Then $K$ is the union of $G_0$ and the sets $G_j, j \in J$ (in particular, when $J$ is the empty set, the value of the random set is $G_0$). We refer to such $K$ as $K_J$. The distribution of such a random set is determined by $2^n$ non-negative numbers $\alpha_1, ..., \alpha_{2^n}$, whose sum is $1$. The sampling from the value $K$ of the random set, is done according to a probability measure $P_K$. We assume that if $i \in J$ then $P_{K_J}(G_i) > 0$, and that $P_{K_J}(G_0) > 0$ for all $J$. \end{example} The results that follow pertain to Example \ref{Exa6.3}. \begin{proposition}\label{Pro6.4} Let $\Mscr_0$ be the family of probability measures supported on $\{K_1, ..., K_{2^n}\}$. Then the statistics of the $n$-point sample $S_n(\hbox{\boldmath\char'026})$ of $\hbox{\boldmath\char'026} \in \Mscr_0$, determines $\hbox{\boldmath\char'026}$, and the inverse of $S_n(\cdot)$ is continuous. \end{proposition} \begin{proof} We use the method of peeling. The statistics $S_n(\hbox{\boldmath\char'026})$ is a probability measure on $X^n$, namely, it is the integral (in our case, the sum), of $\alpha_J P_J^n$, with $J$ running over all subsets of $\{1, ..., n\}$. Consider within the support of $S_n(\hbox{\boldmath\char'026})$, the subset where the $n$-sample includes points in all the sets $\{G_1, ..., G_n\}$. This part of the statistics is generated by only one value of the random set, namely, the union of all $\{G_1, ..., G_n\}$ union $G_0$. The probability of this subset of $X^n$, according to $P_{K_J}^n$ is positive. The value of the statistics on this set is given. The ratio of the two numbers is the coefficient $\alpha_{J_1}$ where $J_1$ represents the said subset. Once this $\alpha_{J_1}$ is revealed, it can be removed from the distribution of the random set, and the corresponding part of the statistics can be peeled from the statistics. We are left with a random set (actually, with part of a random set as the underlying probability measure may have probability less than $1$, but we still refer to it as a random set), with values as the initial one, without the eliminated set. Now consider within the support of the new $S_n(\hbox{\boldmath\char'026})$, the subset where the $n$-sample includes points in $n-1$ sets $G_i$, say the first sets $\{G_1, ..., G_{n-1}\}$. Once the set of all indices was eliminated, this part of the statistics is generated by only one value of the random set, namely, the union of $\{G_0, G_1, ..., G_{n-1}\}$. The same argument as the previous one can be applied, so the relevant $\alpha_J$ is determined, and the respective parts can be peeled from both the random set and the statistics. Repeating the argument $n-1$ times will determine the coefficients $\alpha_J$ associated with points in the subsets of size $n-1$ of $\{G_1, ..., G_n\}$. Then the respective contributions to the random set and the statistics can be removed. Next stage will be to consider subsets of $\{1, ..., n\}$ of size $n-2$. Since all subsets of size $n-1$ were peeled, the peeling argument can be applied at this stage as well. Continuing with the peeling argument through subsets of $\{1, ..., n\}$ of sizes $n-3, ..., 2$, will reveal all the coefficients $\alpha_J$. The continuity of the inverse mapping is obvious. This completes the proof. \end{proof} \begin{comment}\label{Com6.5} As in a previous case of peeling, the proof of the preceding result provides an algorithm to compute the distribution of the random set. This time, however, the information needed is the sampling rule, and the support of the distribution of the random set (indeed, in general, the statistics may not determine the peeled set). \end{comment} There may be sampling rules for which the $m$-point statistics of a random set depicted in Example \ref{Exa6.3}, determines the distribution of the random set with $m < n$. Under more structured sampling rules, we can get exact estimates. One such rule is the regular sampling rule, introduced in Definition \ref{Def3.3}. The following is the corresponding result, where the regular sampling rule is generated by a measure $\gamma$ on $X$. \begin{proposition}\label{Pro6.6} Let $\Mscr_0$ be the family of probability measures supported on $\{K_1, ..., K_{2^n}\}$. Suppose that the sampling rule is regular, governed by a measure $\gamma$. Then the statistics of $n$-point sample $S_n(\hbox{\boldmath\char'026})$ of $\hbox{\boldmath\char'026} \in \Mscr_0$, determines $\hbox{\boldmath\char'026}$, and the inverse of $S_n(\cdot)$ on $\Mscr_0$ is continuous. The statistics of the $(n-1)$-point sample $S_{n-1}(\hbox{\boldmath\char'026})$, does not determine $\hbox{\boldmath\char'026}$. \end{proposition} \begin{proof} The case of $n$-point samples is a particular case of the previous result. We proceed to prove that the $(n-1)$-point sampling does not determine the random set. We follow the derivations of the general case, introduced earlier in the section. Let $<\alpha>$ denote the vector of weights that $\hbox{\boldmath\char'026}$ assigns to each set in its support, namely, the coordinates of the vector $<\alpha>$ are indexed by $J$, running over all $2^n$ subsets of $\{1, ..., n\}$. We think of $<\alpha>$ as a row vector. The probability measure $S_{(n-1)}(\hbox{\boldmath\char'026})$ is determined by its values on sets in $X^{(n-1)}$ of the form $B_1 \times ... \times B_{(n-1)}$, where each $B_i$ is a Borel subset in $X$ such that if $B_i$ intersects one of the $G_l$, then it is included in that $G_l$. Indeed, any probability measure on a product space is determined by its values on product of sets, and any such product can be separated into products of the form just described. Now, given such a product, we have (compare with (\ref{Eq6.1})) \begin{equation}\label{Eq6.3} S_{(n-1)}(\hbox{\boldmath\char'026})(B_1 \times ... \times B_{(n-1)}) = \sum_J \alpha_J P_J(B_1)\cdot ... \cdot P_J(B_{(n-1)}), \end{equation} where $J$ runs over all subsets of $\{1, ..., n\}$. Since the sampling rule is regular, it follows (compare with (\ref{Eq6.2})) that (\ref{Eq6.3}) can be written as \begin{equation}\label{Eq6.4} S_{(n-1)}(\hbox{\boldmath\char'026})(B_1 \times ... \times B_{(n-1)}) = \gamma(B_1)\cdot ... \cdot \gamma(B_{(n-1)}) <\beta> <\alpha>^T, \end{equation} where $<\beta>$ is a row vector that depends on the sets $\{B_1, ..., B_{(n-1)}\}$, with $2^n$ entries according to $J \subseteq \{1, ..., n\}$. The $J$-th coordinate of $<\beta>$ is equal to $0$ if one of the $B_i$ is not included in $K_J$. Otherwise, the entry is equal to $\gamma(K_J)^{-(n-1)}$. This structure implies that the number of distinct vectors $<\beta>$ is given by \begin{equation}\label{Eq6.5} \sum_{l = 0}^{n-1} {n \choose l}, \end{equation} which is equal to $2^n - 1$. Now, arrange the distinct vectors $<\beta>$ as a matrix, denoted ${\bf B}$. It has $2^n - 1$ rows and $2^n$ columns. We add to ${\bf B}$ a row vector with all its coordinates equal to $1$. We denote the new matrix by ${\bf B}_1$. Furthermore, (for visual convenience) we arrange the coordinates of $<\alpha>$ in a non-decreasing order of the size of $J$, and, accordingly, we arrange the columns of ${\bf B}_1$ in the same order. The result is a matrix with a bottom row of entries equal to $1$, and the rest being an upper diagonal matrix, with entries in the $J$-th column being either equal to $0$ or equal to $\gamma(K_J)^{-(n-1)}$, depending whether the subset that is associated with the row is included in $J$ or not. (To make it easier to follow the arguments, see Example \ref{Exa6.7} below, with $n = 3$). All in all, the value of the $(n-1)$-point statistics $S_{(n-1)}(\hbox{\boldmath\char'026})$ of $B_1 \times ... \times B_{(n-1)}$ is then given by (\ref{Eq6.4}) for the appropriate $<\beta>$. We claim that the matrix ${\bf B}_1$ just constructed is singular, namely, its rank is less than $2^n$. Once this claim is verified, Theorem \ref{The6.1} implies that the $(n-1)$-point statistics does not determine the distribution of the random set. It remains to verify that ${\bf B}_1$ is singular. To that end recall that a nonzero entry in the column of ${\bf B}_1$ associated with a subset $J$ of $\{1,2, ..., n\}$, is equal to $\gamma(K_J)^{-(n-1)}$, except on the last row, where the entry is $1$. For each $J$ we multiply the $J$-th column by $\gamma(K_J)^{(n-1)}$. This operation does not change the rank of the matrix. Now, the entries of the last row in the matrix are $\gamma(K_J)^{(n-1)}$ with $J$ running over all subsets of $\{1,2, ..., n\}$. The first $(2^n-1)$ rows of ${\bf B}_1$ form an inclusion matrix, namely, its $(J_1,J_2)$ entry is $1$ if $J_1$ is a subset of $J_2$, and $0$ otherwise. Now multiply by $-1$ each column of the matrix which is associated with a subset $J$, i.e., a coordinate of $<\alpha>$, that has an odd number of elements. Denote the new matrix by ${\bf C}$. The rank of ${\bf C}$ is the same as the rank of ${\bf B}_1$. We claim that ${\bf C}$ is such that the sum of the entries in each of its rows is $0$. This would verify that ${\bf C}$ has rank less than $2^n$. To verify the claim, consider first the top row. Counting the number of entries associated with sets with even number of elements, adding to that the number of entries associated with an odd number of elements, multiplied however by $-1$, results in \begin{equation}\label{Eq6.6} \sum_{l = 0}^n (-1)^l {n \choose l}. \end{equation} The value of the latter term is $( 1-1)^n = 0$ The computation of any other row except the last one is similar. Indeed, given a row in ${\bf B}$, it has entries equal to $0$ until the diagonal, where the column is associated with a set $J_0$. The non-zero entries in the rest of the row, are those associated with sets $J$ that include $J_0$. Counting those with even number of entries, and adding the number of those with odd number of elements multiplied by $-1$, results in a term like (\ref{Eq6.6}), with $n$ replaced by $n - j_0$, where $j_0$ is the number of elements in $J_0$ (in particular, $j_0 < n$). Hence the sum of the entries of any of the first $2^n-1$ rows in ${\bf C}$ is $0$. The last step needed for the completion of the proposition is to show that the sum of entries of the last row of ${\bf C}$ is $0$. To this end we need to spell out the exact form of these entries, namely, \begin{equation}\label{Eq6.7} \gamma(K_J)^{(n-1)} = (\gamma(G_0) + \sum_{j \in J} \gamma(G_j))^{(n-1)}. \end{equation} Expanding each of the latter terms, we get that each entry in the last row in ${\bf C}$ is a sum of monomials, of the form $ \gamma(G_0)^{k_0} \gamma(G_1))^{k_1}\cdot \cdot \cdot \gamma(G_n))^{k_{(n-1)}}$, with the sum of the $k_i$ is $n-1$, namely of order $n-1$, multiplied by the corresponding binomial (rather, multinomial) coefficient ${(n-1)! \over k_1! k_2! \cdot \cdot \cdot k_{n-1}!}$. Let $J_0$ be the subset associated with the indices with $k_j > 0$. It is clear that this monomial appears exactly in those columns associated with $J$ which includes $J_0$. Now the observations made for the $2^n-1$ rows, where the entries are equal to $1$, is valid also for the summation of the constant entries forming the identified monomials. Applying this to all the monomials appearing in the expansion verifies the claim. This completes the proof. \end{proof} Here is the example promised in the proof of the previous result. \begin{example}\label{Exa6.7} We consider a specific case of the example examined so far, in the case $n=3$, namely, the a value of the random set is the union of a given set $G_0$ with the sets that form subset of $(G_1, G_2, G_3)$. The sampling rule regular, generated by a measure $\gamma$. Suppose that the measure $\gamma$ assigns to the sets $G_i$ the values $\gamma(G_0) = 3$, $\gamma(G_1) = 2$, $\gamma(G_2) = 1$, and $\gamma(G_3) = 4$. Arrange $<\alpha>$ according to the column vector as appears in the displayed form. Here the index reflects the sets $G_i$ the union of which forms a value of the random set. With these data, the matrix ${\bf B}_1$ invoked in the proof, here for a sample size $2$, is as follows. \vskip 2pt \begin{equation}\label{Eq6.8} \begin{pmatrix} 3^{-2} &5^{-2} &4^{-2} &7^{-2} &6^{-2} &8^{-2} &9^{-2} &10^{-2} \\ 0 &5^{-2} &0 &0 &6^{-2} &8^{-2} &0 &10^{-2} \\ 0 &0 &4^{-2} &0 &6^{-2} &0 &9^{-2} &10^{-2} \\ 0 &0 &0 &7^{-2} &0 &8^{-2} &9^{-2} &10^{-2} \\ 0 &0 &0 &0 &6^{-2} &0 &0 &10^{-2} \\ 0 &0 &0 &0 &0 &8^{-2} &0 &10^{-2} \\ 0 &0 &0 &0 &0 &0 &9^{-2} &10^{-2} \\ 1 &1 &1 &1 &1 &1 &1 &1 \end{pmatrix} \begin{pmatrix} \alpha_0 \\ \alpha_{01} \\ \alpha_{02} \\ \alpha_{03} \\ \alpha_{012} \\ \alpha_{013} \\ \alpha_{023} \\ \alpha_{0123} \end{pmatrix}. \end{equation} \vskip 2pt To demonstrate the way the expression (\ref{Eq6.8}) is used in the generation of the statistics, let $<\alpha>$ be the distribution of the random set. The value of the $(n-1)$-point statistics of, say $G_1 \times G_3$, is as follows. Invoking (\ref{Eq6.4}), and realizing that the string $G_1 \times G_3$ relates to the $7$-th row, say ${\bf b}_7$, the value of the statistics is $\gamma (G_1)\gamma(B_3) {\bf b}_7 <\alpha> $, namely, $2 \cdot 3 ( 8^{-2}\alpha_{13} + 10^{-2}\alpha_{0123})$. \end{example} \begin{remark}\label{Rem6.8} We may apply the procedure in Proposition \ref{Pro6.6} to $n$-point sampling. It would yield another proof (under, however, the additional conditions), that the $n$-point statistics determines the distribution. Indeed, the resulting matrix ${\bf B}$ is then upper triangular with non-zero entries on the diagonal, hence with rank $2^n$. In particular, the statistics of distinct distributions cannot coincide. This also leads to a constructive way to compute the distribution, as noted in Comment \ref{Com6.2}. \end{remark} We exploit the previous computations to examine two small variations of the former one, demonstrating, however, a change in the determination parameters. \begin{example}\label{Exa6.9} Consider Example \ref{Exa6.3}, with the following change. The set $G_0$ is excluded from being a value of the random set. Thus, the number of sets in the support of the distribution of a random set is $2^n -1$, and the distribution of such a random set is given by the vector $\alpha_1, ..., \alpha_{2^n-1}$. \end{example} The, seemingly minor change of the former example, affects the size of sampling needed to determine the distribution, as follows. \begin{proposition}\label{Pro6.10} Let $\Mscr_0$ be the family of probability measures supported on $\{K_1, ..., K_{2^n-1}\}$ in Example \ref{Exa6.9}. Suppose that the sampling rule is regular, governed by a measure $\gamma$. Then the statistics of $(n-1)$-point sample $S_{n-1}(\hbox{\boldmath\char'026})$ of $\hbox{\boldmath\char'026} \in \Mscr_0$, determines $\hbox{\boldmath\char'026}$, and the inverse of $S_{n-1}(\cdot)$ on $\Mscr_0$ is continuous. The statistics of $(n-2)$-point sample $S_{n-2}(\hbox{\boldmath\char'026})$, does not determine $\hbox{\boldmath\char'026}$. \end{proposition} \begin{proof} First we verify that the $(n-1)$-point sample $S_{n-1}(\hbox{\boldmath\char'026})$ determines $\hbox{\boldmath\char'026}$. Consider the matrix ${\bf B}$ as previously constructed. It is clear that it is identical to the matrix ${\bf B}$ in Proposition \ref{Pro6.6}, without, however, the first row (since the set $G_0$ is not allowed to be a value of the random set). We argue that the, now $(2^n-1) \times (2^n -1)$, matrix ${\bf B}$ has rank $2^n -1$. To see that, you may first multiply each column of ${\bf B}$ by $\gamma(K_J)^{(n-1)}$ where the column is associated with the set $K_J$. This makes ${\bf B}$ a matrix of zeroes and ones, without changing the rank. Now multiply each column associated with $K_J$ with the size of $J$ being an odd number, by $-1$. Again, the rank has not changed. Adding all columns to the last one results in the vector $(1, 0, 0, ..., 0)^T$. This was verified in the computations in the proof of Proposition \ref{Pro6.6}. Indeed, adding up the columns of ${\bf B}$ in the said proof results in the zero vector. Here, however, we have a non-zero result since the first column in the latter sum does not take part. Moving the last column to be the first one yields an upper diagonal matrix with non-zero entries on the diagonal, hence having full $2^n -1$ rank. Invoking Proposition \ref{The6.1} completes the first part of the proof, To verify that the $(n-2)$-point statistics does not determine the distribution, construct the relevant matrix ${\bf B}$. Its number of rows is given by $\sum_{l = 0}^{n-2} {n \choose l}$, which is less than $2^n -1$ even after adding the row of ones, to get ${\bf B}_1$. Hence the latter is singular. Invoking Theorem \ref{The6.1} completes the proof. \end{proof} \begin{example}\label{Exa6.11} Let $\{G_1, G_2, ..., G_n\}$ be a finite collection of mutually disjoint sets in a space $X$. Let $K_1, ..., K_{(2^n-1)}$ be the $2^n-1$ unions of non-empty sub-collections. Let $P_K$ be the sampling rule from the set $K$. A random set with values in the family of such unions is determined by a probability vector $\alpha_1, ..., \alpha_{2^n-1}$. \end{example} \begin{proposition}\label{Pro6.12} Let $\Mscr_0$ be the family of probability measures supported on the sets $\{K_1, ..., K_{2^n-1}\}$ in Example \ref{Exa6.11}. Suppose that the sampling rule is regular, with regularity governed by a measure $\gamma$. Then the statistics of $(n-1)$-point sample $S_{n-1}(\hbox{\boldmath\char'026})$ of $\hbox{\boldmath\char'026} \in \Mscr_0$, determines $\hbox{\boldmath\char'026}$, and the inverse of $S_{n-1}(\cdot)$ on $\Mscr_0$ is continuous. The statistics of $(n-2)$-point sample $S_{n-2}(\hbox{\boldmath\char'026})$, does not determine $\hbox{\boldmath\char'026}$. \end{proposition} \begin{proof} First we verify that the $(n-1)$-point sample $S_{n-1}(\hbox{\boldmath\char'026})$ determines $\hbox{\boldmath\char'026}$. Consider the matrix ${\bf B}$ as previously constructed. It is clear that it is identical to the matrix ${\bf B}$ in Proposition \ref{Pro6.6}, without, however, the first row and the first column. (since the set $G_0$ does not take part in this example). The matrix ${\bf B}$ is, now, $(2^n-2) \times (2^n -1)$. We add to it a row of $1$s, forming the matrix ${\bf B}_1$. We claim that ${\bf B}_1$ has full $2^n-1$ rank. To see that, you may first multiply each column of ${\bf B}$ by $\gamma(K_J)^{(n-1)}$ where the column is associated with the set $K_J$. This makes ${\bf B}$ a matrix of zeroes and ones, without changing the rank. Now multiply each column associated with $K_J$ with the size of $J$ being an odd number, by $-1$. Again, the rank has not changed. Adding all columns to the last one result in the vector $(0, 0, 0, ..., c)^T$ with $c$ not equal to $0$. This was verified in the computations in the said proof. Indeed, adding up the columns of ${\bf B}$ in that proof results in the zero vector. Here, however, we have a non-zero result since the first column in the latter sum does not take part. Simple operations on the columns of the matrix will now lead to an upper diagonal matrix with non-zero entries on the diagonal. Invoking Theorem \ref{The6.1} completes the first part of the proof, To verify that the $(n-2)$-point statistics does not determine the distribution, construct the relevant matrix ${\bf B}$. Its number of rows is given by $\sum_{l = 0}^{n-2} {n \choose l}$, which is less than $2^n -1$ even after adding the row of ones, to get ${\bf B}_1$. Hence the latter is singular. Invoking Theorem \ref{The6.1} completes the proof. \end{proof} The two possibilities mentioned in the Comment \ref{Com6.2}, namely, that ${\bf B}$ is $L \times L$ and has full rank, and that ${\bf B}$ is $(L-1) \times L$ and ${\bf B}_1$ has full rank, may occur. For instance, the first possibility occurs in Example \ref{Exa6.9}, Another example is when the sets $K_j$ are disjoint. Then the ${\bf B}$ matrix generated by the $1$-point statistics is the identity matrix. An example for the second possibility is the one in Example \ref{Exa6.11}. The matrix ${\bf B}$ is then of order $(2^n -2) \times (2^n - 1)$, and the addition of the line to make ${\bf B}_1$, results in a matrix of full rank. \section{On random convex sets}\label{S7} One may seek a generalization of Theorem \ref{The4.2} to random convex sets in, say, the two-dimensional plain, replacing the $2$-point sample by an $m$-point sample for a certain $m$. We give counter-examples to such a possibility. First consider a convex polygon with $n$ edges, whose vertices touch the edges of another convex $n$-polygon. The case of a regular hexagon is depicted in Figure 3. Denote the set included in the smaller $n$-polygon by $G_0$, and the $n$ triangles between the smaller polygon and the larger one by $G_1, ...,G_n$. \begin{figure}[h] \centering \includegraphics[width=0.5\textwidth]{Figure3} \caption{} \end{figure} Notice that the union of $G_0$ and the elements in any subset of $\{G_1, ..., G_n\}$, is a convex set (we include in this count the empty set union $G_0$). There are $2^n$ such sets. Notice that if the sampling rule for the $2^n$ convex sets just described is regular, say governed by the measure $\gamma$ on $R^2$, and if $\gamma$ assigns zero probability to the boundaries of all $G_i$, say it is the Lebesgue measure, then the considerations of Proposition \ref{Pro6.6} apply. Then the $m$-point statistics of a random convex set supported on the $2^n$ sets determines the distribution if and only if $m \ge n$. If $G_0$ is excluded from being a value, then Proposition \ref{Pro6.10} applies. Then the $m$-point statistics of a random convex set supported on the $2^n-1$ sets determines the distribution if and only if $m \ge n-1$. In particular, there is no bound on the least number $m$ that implies determination. Furthermore, a variation of the example shows that if a denumerable number of convex values is allowed, no $m$-point statistics is enough to determine the random set. Indeed, consider a ``polygon'' like the one in Figure 3, with, however, a denumerable number of edges. Additional geometrical restrictions on the support of the random convex set, may yield generalizations of the results for random intervals (homothetic polygons may yield such estimates). The extent of such restrictions is not clear at this point.
{ "timestamp": "2021-07-22T02:17:06", "yymm": "2104", "arxiv_id": "2104.12373", "language": "en", "url": "https://arxiv.org/abs/2104.12373", "abstract": "We examine the extent to which random samplings from the values of a random set, determine the distribution of the random set itself. We also comment on how, given the statistics of the sampling, to detect the distribution. Several methods are displayed, leading to positive results, counter-examples, and open problems.", "subjects": "Probability (math.PR)", "title": "Detecting random sets by samplings from their values", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9783846653465639, "lm_q2_score": 0.8354835452961425, "lm_q1q2_score": 0.8174242888671271 }
https://arxiv.org/abs/1702.01885
Some bounds for commuting probability of finite rings
Let $R$ be a finite ring. The commuting probability of $R$ is the probability that any two randomly chosen elements of $R$ commute. In this paper, we obtain some bounds for commuting probability of $R$.
\section{Introduction} Throughout the paper $R$ denotes a finite ring. The commuting probability of $R$, denoted by $\Pr(R)$, is the probability that a randomly chosen pair of elements of $R$ commute. That is \begin{equation}\label{com_prob_eq1} \Pr(R) = \frac{|\{(s, r) \in R \times R : sr = rs\}|}{|R \times R|}. \end{equation} The study of $\Pr(R)$ was initiated by MacHale \cite{dmachale} in the year 1976. After the works of Erd$\ddot{\rm o}$s and Tur$\acute{\rm a}$n \cite{pEpT68}, many papers have been written on commuting probability of finite groups in the last few decades, for example see \cite{DN11} and the references therein. However, people did not work much on commuting probability of finite rings. We have only few papers \cite{BM, BMS, dmachale} on $\Pr(R)$ in the literature. In this paper, we obtain some bounds for $\Pr(R)$. Recall that $[s, r]$ to denote the additive commutator $sr - rs$ for any two elements $s, r \in R$. By $K(R, R)$ we denote the set $\{[s, r] : s, r \in R\}$ and $[R, R]$ denotes the subgroup of $(R, +)$ generated by $K(R, R)$. Note that $[R, R]$ is the commutator subgroup of $(R, +)$ (see \cite{BMS}). Also, for any $x \in R$, we write $[x, R]$ to denote the subgroup of $(R, +)$ consisting of all elements of the form $[x, y]$ where $y \in R$. \section{Main Results} Let $C_R(r)$ denote the subset $\{s \in R : sr = rs\}$ of $R$, where $r$ is an element of $R$. Then $C_R(r)$ is a subring of $R$ known as centralizer of $r$ in $R$. Note that the center $Z(R)$ of $R$ is the intersection of all the centralizers in $R$. By \eqref{com_prob_eq1}, we have \[ \Pr(R) = \frac{1}{|R|^2} \underset{r \in R}{\sum}|C_R(r)| \] and hence \begin{equation}\label{com_prob_eq2} \Pr(R) = \frac{|Z(R)|}{|R|} + \frac{1}{|R|^2} \underset{r \in R \setminus Z(R)}{\sum}|C_R(r)|. \end{equation} If $p$ is the smallest prime dividing the order of a finite non-commutative ring $R$ then, by \cite[Theorem 2]{dmachale}, we have \begin{equation}\label{com_prob_eq3} \Pr(R) \leq \frac{p^2 + p - 1}{p^3}. \end{equation} In the following theorem, we give two bounds for $\Pr(R)$. We shall see that the upper bound for $\Pr(R)$ in Theorem \ref{theorem001} is better than \eqref{com_prob_eq3}. \begin{theorem}\label{theorem001} Let $R$ be a finite non-commutative ring. If $p$ is the smallest prime dividing $|R|$ then \begin{enumerate} \item[\rm (a)] $\Pr(R) \geq \frac{|Z(R)|}{|R|} + \frac{p(|R| - |Z(R)|)}{|R|^2}$ with equality if and only if $|C_R(r)| = p$ for all $r \notin Z(R)$. \item[\rm (b)] $\Pr(R) \leq \frac{(p - 1)|Z(R)| + |R|}{p|R|}$ with equality if and only if $|R: C_R(r)| = p$ for all $r \notin Z(R)$. \end{enumerate} \end{theorem} \begin{proof} By \eqref{com_prob_eq2}, we have \begin{equation}\label{boundeq-3} |R|^2 \Pr(R) = |R||Z(R)| + \underset{r \in R\setminus Z(R)}{\sum}|C_R(r)|. \end{equation} (a) If $r \notin Z(R)$ then $|C_R(r)| \geq p$. Therefore \[ \underset{r \in R\setminus Z(R)}{\sum}|C_R(r)| \geq p(|R| - |Z(R)|) \] with equality if and only if $|C_R(r)| = p$ for all $r \notin Z(R)$. Hence, the result follows from \eqref{boundeq-3}. (b) If $r \notin Z(R)$ then $|C_R(r)| \leq \frac{|R|}{p}$. Therefore \[ \underset{r \in R\setminus Z(R)}{\sum}|C_R(r)| \leq \frac{|R|(|R| - |Z(R)|)}{p} \] with equality if and only if $|R: C_R(r)| = p$ for all $r \notin Z(R)$. Hence, the result follows from \eqref{boundeq-3}. \end{proof} \noindent If $R$ is a non-commutative ring and $p$ the smallest prime dividing $|R|$ then $|R : Z(R)| \geq p^2$. Therefore \[ \frac{(p - 1)|Z(R)| + |R|}{p|R|} \leq \frac{p^2 + p - 1}{p^3}. \] Thus the bound obtained in Theorem \ref{theorem001}(b) is better than \eqref{com_prob_eq3}. If $S$ is a subring of $R$ then MacHale \cite[Theorem 4]{dmachale} showed that \begin{equation}\label{MachaleThm4} \Pr(R) \leq \Pr(S). \end{equation} We now proceed to derive an improvement of \eqref{MachaleThm4}. We require the following lemma. \begin{lemma}\label{lemma2} Let $N$ be an ideal of a finite non-commutative ring $R$. Then \[ \frac{C_R(x) + N}{N} \subseteq C_{R/N}(x + N) \; \text{for all} \; x \in R. \] The equality holds if $N \cap [R, R] = \{0\}$. \end{lemma} \begin{proof} For any element $s \in C_R(x) + N$, where $s = r + n$ for some $r \in C_R(x)$ and $n \in N$, we have $s + N = r + N \in R/N$. Also, \[ (s + N)(x + N) = rx + N = xr + N = (x + N)(s + N), \] as $r \in C_R(x)$. This proves the first part. Let $N \cap [R, R] = \{0\}$ and $y + N \in C_{R/N}(x + N)$. Then $y \in R$ and $(y + N)(x + N) = (x + N)(y + N)$. This gives $yx - xy \in N \cap [R, R] = \{0\}$ and so $y \in C_R(x)$. Therefore, $y + N \in \frac{C_R(x) + N}{N}$. Hence the equality holds. \end{proof} The following result which is an improvement of \eqref{MachaleThm4} also gives a relation between $\Pr(R), \Pr(R/N)$ and $\Pr(N)$, where $N$ is an ideal of $R$. \begin{theorem}\label{theorem3} Let $N$ be an ideal of a finite non-commutative finite ring $R$. Then \[ \Pr(R) \leq \Pr (R/N) \Pr(N). \] The equality holds if $N \cap [R, R] = \{0\}$. \end{theorem} \begin{proof} We have that \allowdisplaybreaks{ \begin{align*} |R|^2 \Pr(R) = &\underset{x \in R}{\sum} |C_R(x)| \\ = &\underset{S \in \frac{R}{N}}{\sum}\underset{y \in S}{\sum}\frac{|C_R(y)|}{|N \cap C_R(y)|} |C_N(y)| \\ = &\underset{S \in \frac{R}{N}}{\sum}\underset{y \in S}{\sum}\frac{|C_R(y) + N|}{|N|} |C_N(y)| \\ \leq &\underset{S \in \frac{R}{N}}{\sum}\underset{y \in S}{\sum}|C_{\frac{R}{N}}(y + N)| |C_N(y)| \;\;(\text{using Lemma \ref{lemma2}}) \\ = &\underset{S \in \frac{R}{N}}{\sum} |C_{\frac{R}{N}}(S)| \underset{y \in S}{\sum}|C_N(y)| \\ = &\underset{S \in \frac{R}{N}}{\sum} |C_{\frac{R}{N}}(S)| \underset{n \in N}{\sum}|C_R(n) \cap S|. \end{align*} } Let $a + N = S$ where $a \in R\setminus N$. If $C_R(n) \cap S = \phi$ then $|C_R(n) \cap S| < |C_N(n)|$. If $C_R(n) \cap S \neq \phi$ then there exists $x_0 \in C_R(n) \cap S$ such that $x_0 = a + n_0$ for some $a \in R\setminus N$ and $n_0 \in N$. Therefore $x_0 + N = a + N = S$ and so $S \cap C_R(n) = (x_0 + N) \cap (x_0 + C_R(n)) = x_0 + (N \cap C_R(n)) = x_0 + C_N(n)$. Hence $|S \cap C_R(n)| \leq |C_N(n)|$. This gives \begin{align*} |R|^2 \Pr(R) \leq &\underset{S \in \frac{R}{N}}{\sum} |C_{\frac{R}{N}}(S)| \underset{n \in N}{\sum}|C_N(n)| \\ = & |R/N|^2 \Pr(R/N) |N|^2 \Pr(N) \\ = & |R|^2 \Pr(R/N)\Pr(N). \end{align*} Hence the bound follows. Let $N \cap [R, R] = \{0\}$. Then, by Lemma \ref{lemma2}, we have \[ \frac{C_R(x) + N}{N} = C_{R/N}(x + N) \; \text{for all} \; x \in R. \] If $S = a + N$ then it can be seen that $a + n \in C_R(n) \cap S$ for all $n \in N$. Therefore, $C_R(n) \cap S \ne \phi$ for all $n \in N$ and for all $S \in R/N$. Thus all the inequalities above become equalities if $N \cap [R, R] = \{0\}$. This completes the proof. \end{proof} In Lemma 2.3 of \cite{BMS}, Buckley et al. showed that \begin{equation}\label{buckley_eq} \Pr(R) > \frac{1}{|[R, R]|}. \end{equation} The following two results give some improvements of \eqref{buckley_eq}. \begin{theorem}\label{newlb1} Let $R$ be a finite ring $R$. Then \[ \Pr(R) \geq \frac{1}{|K(S, R)|}\left(1 + \frac{|K(R, R)| - 1}{|R : Z(R)|} \right) \] with equality if and only if $|K(R, R)| = |[r, R]|$ for all $r \in R\setminus Z(R)$. In particular, if $R$ is non-commutative then $\Pr(R) > \frac{1}{|K(R, R)|}$. \end{theorem} \begin{proof} By \eqref{com_prob_eq2}, we have \[ \Pr(R) = \frac{|Z(R)|}{|R|} + \frac{1}{|R|}\underset{r \in R\setminus Z(R)}{\sum}\frac{1}{|R : C_R(r)|}. \] Since $|K(R, R)| \geq |[r, R]| = |R : C_R(r)|$ for all $r \in R\setminus Z(R)$, we have \begin{align*} \Pr(R) \geq & \frac{|Z(R)|}{|R|} + \frac{1}{|R|} \underset{r \in R\setminus Z(R)}{\sum}\frac{1}{|K(R, R)|}\\ = & \frac{|Z(R)|}{|R|} + \frac{|R| - |Z(R)|}{|R||K(R, R)|} \end{align*} from which the result follows. \end{proof} \noindent We also have the following lower bound. \begin{theorem}\label{newlb2} Let $S$ be a subring of a finite ring $R$. Then \[ \Pr(R) \geq \frac{1}{|[R, R]|}\left(1 + \frac{|[R, R]| - 1}{|R : Z(R)|} \right) \] with equality if and only if $|[R, R]| = |[r, R]|$ for all $r \in R\setminus Z(R)$. In particular, if $R$ is non-commutative then $\Pr(R) > \frac{1}{|[R, R]|}$. \end{theorem} \begin{proof} By \eqref{com_prob_eq2}, we have \[ \Pr(R) = \frac{|Z(R)|}{|R|} + \frac{1}{|R|}\underset{r \in R\setminus Z(R)}{\sum}\frac{1}{|R : C_R(r)|}. \] Since $|[R, R]| \geq |R : C_R(r)|$ for all $r \in R\setminus Z(R)$, we have \begin{align*} \Pr(R) \geq & \frac{|Z(R)|}{|R|} + \frac{1}{|R|} \underset{r \in R\setminus Z(R)}{\sum}\frac{1}{|[R, R]|}\\ = & \frac{|Z(R)|}{|R|} + \frac{|R| - |Z(R)|}{|R||[R, R]|} \end{align*} from which the result follows. \end{proof} Let $p$ be the smallest prime dividing $|R|$. If $R$ is non-commutative and $[R, R] \ne R$ then it is easy to see that \[ \frac{1}{|[R, R]|}\left(1 + \frac{|[R, R]| - 1}{|R : Z(R)|} \right) \geq \frac{|Z(R)|}{|R|} + \frac{p(|R| - |Z(R)|)}{|R||R|} \] with equality if and only if $|R : [R , R]| = p$. Also, \[ \frac{1}{|K(R, R)|}\left(1 + \frac{|K(R, R)| - 1}{|R : Z(R)|} \right) \geq \frac{1}{|[R, R]|}\left(1 + \frac{|[R, R]| - 1}{|R : Z(R)|} \right) \] with equality if and only if $K(R, R) = [R, R]$. Hence, the lower bound obtained in Theorem \ref{newlb1} is better than the lower bounds obtained in Theorem \ref{theorem001} and Theorem \ref{newlb2}. We conclude the paper noting that Theorem \ref{newlb1} and Theorem \ref{newlb2} are analogous to \cite[Theorem A]{nY15} and \cite[Theorem 1]{ND10} respectively.
{ "timestamp": "2017-02-08T02:03:08", "yymm": "1702", "arxiv_id": "1702.01885", "language": "en", "url": "https://arxiv.org/abs/1702.01885", "abstract": "Let $R$ be a finite ring. The commuting probability of $R$ is the probability that any two randomly chosen elements of $R$ commute. In this paper, we obtain some bounds for commuting probability of $R$.", "subjects": "Rings and Algebras (math.RA)", "title": "Some bounds for commuting probability of finite rings", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.988668248138067, "lm_q2_score": 0.8267118004748677, "lm_q1q2_score": 0.8173437074905546 }
https://arxiv.org/abs/1611.03129
Regular bipartite graphs and intersecting families
In this paper we present a simple unifying approach to prove several statements about intersecting and cross-intersecting families, including the Erd\H os--Ko--Rado theorem, the Hilton--Milner theorem, a theorem due to Frankl concerning the size of intersecting families with bounded maximal degree, and versions of results on the sum of sizes of non-empty cross-intersecting families due to Frankl and Tokushige. Several new stronger results are also obtained.Our approach is based on the use of regular bipartite graphs. These graphs are quite often used in Extremal Set Theory problems, however, the approach we develop proves to be particularly fruitful.
\section{Intersecting families}\label{sec1} Let $[n]:=\{1,\ldots,n\}$ denote the standard $n$-element set. The family of all $k$-element subsets of $[n]$ we denote by ${[n]\choose k}$, and the set of all subsets of $[n]$ we denote by $2^{[n]}$. Any subset of $2^{[n]}$ we call a \textit{family}. We say that a family is \textit{intersecting}, if any two sets from the family intersect. Probably, the first theorem, devoted to intersecting families, was the famous theorem of Erd\H os, Ko and Rado: \begin{thrm}[Erd\H os, Ko, Rado, \cite{EKR}]\label{ekr} Let $n\ge 2k>0.$ Then for any intersecting family $\mathcal F\subset {[n]\choose k}$ one has $|\mathcal F|\le {n-1\choose k-1}$. \end{thrm} It is easy to give an example of an intersecting family, on which the bound from Theorem \ref{ekr} is attained: take the family of all $k$-element sets containing element $1$. Any family, in which all sets contain a fixed element, we call \textit{trivially intersecting}. What size can an intersecting family have, provided that it is not trivially (nontrivially) intersecting? For $n=2k$ it is easy to construct many intersecting families of size ${2k-1\choose k-1}$ by choosing exactly one $k$-set out of each two complementary sets. For $n>2k$ the answer is given by the Hilton-Milner theorem. \begin{thrm}[Hilton, Milner, \cite{HM}]\label{hm} Let $n> 2k$ and $\mathcal{F}\subset {[n]\choose k}$ be a nontrivially intersecting family. Then $|\mathcal{F}|\le {n-1\choose k-1}-{n-k-1\choose k-1}+1$. \end{thrm} Let us mention that the upper bound is attained on the family that contains exactly one $k$-element set $A$, such that $1\notin A$, and all $k$-element sets containing $1$ and intersecting $A$. Hilton--Milner theorem shows in a quite strong sense that the example for the Erd\H os--Ko--Rado theorem is unique, moreover, that Erd\H os--Ko--Rado theorem is stable in the sense that if any intersecting family is large enough, then it must be a trivially intersecting family. A theorem due to Frankl, discussed in Section \ref{sec3}, provides us with a much stronger stability result. In this section we illustrate our new technique by giving a unified proof of Theorems \ref{ekr} and \ref{hm}. The results on cross-intersecting families are discussed in Section~\ref{sec2}, and Frankl's theorem is discussed in Section~\ref{sec3}.\\ The first part of the proof is fairly standard. We show that we may simplify the structure of the families using \textit{shifting}.\\ For a given pair of indices $1\le i<j\le n$ and a set $A \in 2^{[n]}$ define its {\it $(i,j)$-shift} $S_{i,j}(A)$ as follows. If $i\in A$ or $j\notin A$, then $S_{i,j}(A) = A$. If $j\in A, i\notin A$, then $S_{i,j}(A) := (A-\{j\})\cup \{i\}$. That is, $S_{i,j}(A)$ is obtained from $A$ by replacing $j$ with $i$. Next, define the $(i,j)$-shift $S_{i,j}(\mathcal F)$ of a family $\mathcal F\subset 2^{[n]}$: $$S_{i,j}(\mathcal F) := \{S_{i,j}(A): A\in \mathcal F\}\cup \{A: A,S_{i,j}(A)\in \mathcal F\}.$$ We call a family $\mathcal F$ \textit{shifted}, if $S_{i,j}(\mathcal F) = \mathcal F$ for all $1\le i<j\le n$. Note that shifting does not change the cardinality of the family. Note also that we always shift to the left, replacing the larger element with the smaller one.\\ Consider a family $\mathcal F$ of maximal size, satisfying the conditions of Theorem \ref{hm} or \ref{ekr}. We prove that we may suppose that $\mathcal{F}$ is shifted. It is easy to check that an intersecting family stays intersecting after shifts. This is already sufficient for the Erd\H os--Ko--Rado theorem. In the case of the Hilton--Milner theorem we must make sure that the shifted family we obtain is nontrivially intersecting. For that it is sufficient to do several shifts of $\mathcal{F}$ and renumberings of the ground set, that result in ${[k+1]\choose k}\subset \mathcal{F}$. The subfamily ${[k+1]\choose k}$ is invariant under shifts, thus its presence guarantees that $\mathcal F$ will stay nontrivially intersecting after any subsequent shifts. Do the shifts of $\mathcal{F}$ until either the family is shifted or any new shift will result in $\mathcal F$ becoming trivially intersecting. If the former happens, we end up having a nontrivially intersecting shifted family. And if the latter happens, then after the last shift all the sets of $\mathcal F$ intersect a certain two-element subset $\{x, y\}$. Renumber the elements of $[n]$, so that this pair becomes a pair $\{1, 2\}$. Since $\mathcal F$ has maximal size, it contains all the $k$-element sets, containing both $1$ and $2$. Due to this, the $(i,j)$-shifts, where $i,j>2$, do not affect the property of $\mathcal F$ to be notrivially intersecting. Therefore, we may assume that $\mathcal F$ is invariant under these shifts. Since not all sets in $\mathcal F$ contained $1$ (and not all sets contained $2$), then $\mathcal F$ must contain the sets $\{2,\ldots, k+1\}$ and $\{1, 3, \ldots, k+1\}$. The condition ${[k+1]\choose k}\subset \mathcal{F}$ is now fulfilled, and thus no shift can affect the nontriviality of the family. We note that, after preparing an initial version of the manuscript, we were told by Peter Frankl that exactly the same reduction to a shifted family was used by him to prove the Hilton--Milner theorem in \cite{Fra3}. The remaining part of his proof is completely different from our proof. A somewhat similar use of shifting technique to prove the Hilton--Milner theorem appeared in \cite{FF1}. \\ Next we pass to the part of the proof where the bipartite graphs are employed. \begin{defi} For given $A,B\in {[n]\choose k}$ we say that $A$ {\it lexicographically precedes} $B$, or $A\prec B$, if the smallest element of $A\setminus B$ is smaller than that of $B\setminus A$. \end{defi} The \textit{order} of a $k$-element set is its number in the lexicographic order on ${[n]\choose k}$. The \textit{order} of a family $\mathcal F\subset {[n]\choose k}$ is the maximal order among its elements. Among the families of maximal size satisfying the conditions of Theorem \ref{ekr} or \ref{hm} choose a shifted family $\mathcal{F}$ of minimal order. Consider the set $F\in \mathcal{F}$ of maximal order. By the definition of order, if $\mathcal F$ is non-trivial, then $F$ does not contain $1$. Note that for some $l\ge 1$ one has $|F\cap [2l-1]|=l$, otherwise one of the images of $F$ under several shifts is disjoint from $F$ (this property was first noted by Frankl in \cite{Fra2}). Take maximal such $l$, and put $ L=[2l]\setminus F$. Note that $|L|=l$. We remark that, unless $\mathcal F$ is trivial, we have $l\ge 2$. Consider the families $\mathcal{A}:=\{A\in\binom{[n]}{k}:A\cap [2l]=F\cap [2l]\}$ and $\mathcal{B}:=\{B\in\binom{[n]}{k}:B\cap [2l]=L\}$. The following lemma is the key to the proof. \begin{lem}\label{lembp} Let $n>2k>0$ and $\mathcal F$ be a nontrivially intersecting shifted family. Suppose that $F\in \mathcal F$ has maximal order. Then, in the notations above, the family $\mathcal F':=(\mathcal F\setminus {\mathcal A})\cup {\mathcal B}$ is intersecting and has smaller order than $\mathcal F$. Moreover, $|\mathcal F'|\ge |\mathcal F|$. \end{lem} \begin{proof} Let us first show that for any $A,B$, where $B\in \mathcal{B},\ A\in \mathcal{F}\setminus \mathcal{A},$ we have $ A\cap B\not = \emptyset$. The contrary may happen only for $A\in \mathcal{F}$ satisfying $A\cap L = \emptyset$. Since $F$ has the maximal order in $\mathcal F$, we have $F\cap [2l]\subset A$, and, consequently, $A\in \mathcal{A}$. Therefore, $\mathcal{F}'$ is an intersecting family. Since all sets from ${\mathcal B}$ contain $1$, it is clear that the order of $\mathcal F'$ is smaller than that of $\mathcal F$. Finally, we have $|\mathcal{A}|=\binom{n-2l}{k-l}=|\mathcal{B}|$. Consider a bipartite graph with parts $\mathcal A,\mathcal B$, and edges connecting disjoint sets. Independent sets in this graph correspond to intersecting subfamilies of $\mathcal A\cup\mathcal B$. This graph is regular, and, therefore, it contains a perfect matching. Consequently, the largest independent set has size $|\mathcal{B}|$. Therefore $|\mathcal{F}'|\ge|\mathcal{F}|$. \end{proof} In the case of the Erd\H os--Ko--Rado theorem, since the family $\mathcal F$ had the smallest order among the families of maximal size, we conclude that $\mathcal F$ must be trivial.\\ In the case of the Hilton--Milner theorem we obtain a contradiction between the properties of $\mathcal F$ and Lemma \ref{lembp}, unless $F = \{2,\ldots, k+1\}$. Indeed, if $F \ne \{2,\ldots, k+1\}$, then $\{2,\ldots,k+1\}\in \mathcal F'$, that is, $\mathcal F'$ is nontrivially intersecting. Therefore, $F=\{2,\ldots, k+1\}$, we have $l=k$, and all sets in $\mathcal F$, different from $F$, contain $1$. Both theorems are proved.\\ We remark that bipartite graphs have been used many times in Extremal Set Theory, in particular, in the proof the famous Sperner theorem. The Sperner theorem states that the size of the largest family on $[n]$ with no two sets containing each other is ${n\choose \lceil n/2\rceil}$. Sperner used matchings in biregular bipartite graphs between ${[n]\choose k}$ and ${[n]\choose k+1}$, where the edges connected sets, one of which contained the other. With the proof above, both cornerstones of Extremal Set Theory have proofs based on matchings in regular bipartite graphs. \section{Cross-intersecting families}\label{sec2} We call two families $\mathcal A,\mathcal B$ \textit{cross-intersecting}, if for any $A\in \mathcal A,B\in \mathcal B$ we have $A\cap B\ne \varnothing$. Cross-intersecting families are very useful in the problems on intersecting families, as we will illustrate in Section \ref{sec3}. In this section we discuss and prove some important properties of pairs of cross-intersecting families. \begin{defi} Let $0<k<n$ and $0\le m\le {n\choose k}$. Denote $\mathcal L(k,m)$ the family of the first $m$ sets from ${[n]\choose k}$ in the lexicographic order. \end{defi} The following theorem, proved independently by Kruskal \cite{Kr} and Katona \cite{Ka}, is central in Extremal Set Theory. \begin{thrm}[Kruskal \cite{Kr}, Katona \cite{Ka}]\label{thkk} If the families $\mathcal A\subset{[n]\choose a},\mathcal B\subset {[n]\choose b}$ are cross-intersecting, then $\mathcal L(a,|\mathcal A|),\mathcal L(b,|\mathcal B|)$ are cross-intersecting as well. \end{thrm} We remark that it implies immediately that for any such cross-intersecting families either $|{\mathcal A}|\le {n-1\choose a-1}$, or $|{\mathcal B}|\le {n-1\choose b-1}$. The Kruskal--Katona theorem is ubiquitous in studying cross-intersecting families. One of the types of problems concerning cross-intersecting families is to bound the sum of the cardinalities of the two families. In this paper we give a proof for the following general theorem of this type. \begin{thm}\label{lem2} Let $n \ge a+b$, and suppose that families $\mathcal F\subset{[n]\choose a},\mathcal G\subset{[n]\choose b}$ are cross-intersecting. Fix an integer $j\ge 1$. Then the following inequality holds in three different assumptions, listed below: \begin{equation}\label{eqlem1}|\mathcal F|+|\mathcal G|\le {n\choose b}+{n-j\choose a-j}-{n-j\choose b}.\end{equation} 1. If $a< b$ and we have $|\mathcal F|\ge {n-j\choose a-j}$;\\ 2. If $a\ge b$ and we have ${n-j\choose a-j}\le |\mathcal F|\le {n+b-1-a\choose b-1}+{n+b-2-a\choose b-1}$;\\ 3. If $a\ge b$ and we have ${n-a+b-3\choose b-3}+{n-a+b-4\choose b-3}\le |\mathcal F|\le {n-j\choose a-j}$. \end{thm} It is not difficult to come up with an example of a cross-intersecting pair on which the bound \eqref{eqlem1} is attained: take $\mathcal F :=\{F\in{[n]\choose a}: [j]\subset F\}, \mathcal G:=\{ G\in {[n]\choose b}: G\cap [j]\ne \emptyset\}$. We also give the intuition for the bounds on $|\mathcal F|$ in Points 2 and 3. If $\mathcal F=\mathcal L(a,|\mathcal F|)$, then the family $\mathcal F$ of size attaining the upper bound in Point 2 consists of all sets containing $[a-b+1]$ and all the sets that contain $[a-b]$ and $\{a-b+2\}$. Similarly, such family $\mathcal F$ attaining the lower bound in Point 3 consists of all sets containing $[a-b+3]$ and all sets containing $[a-b+2]$ and $\{a-b+4\}$. We give the proof of Theorem~\ref{lem2} in the end of this section. We note that another characteristic that is well-studied for intersecting families is the product of the cardinalities. We refer the reader to the result due to Pyber \cite{P} and a recent refinement \cite{FK3}. Returning to the sum of cardinalities, the following theorem including Point 1 and some cases of Point 2 of Theorem \ref{lem2} was proven in \cite{FT}. \begin{thrm}[Frankl, Tokushige, \cite{FT}]\label{lemft} Let $n > a+b$, $a\le b$, and suppose that families $\mathcal F\subset{[n]\choose a},\mathcal G\subset{[n]\choose b}$ are cross-intersecting. Suppose that for some real number $\alpha \ge 1$ we have ${n-\alpha \choose n-a}\le |\mathcal F|\le {n-1\choose n-a}$. Then \begin{equation}\label{eqft}|\mathcal F|+|\mathcal G|\le {n\choose b}+{n-\alpha \choose n-a}-{n-\alpha \choose b}.\end{equation} \end{thrm} On the one side, in Theorem~\ref{lemft} the parameter $\alpha$ may take real values, while in Theorem~\ref{lem2} the parameter $j$ is bound to be integer. On the other side, Theorem~\ref{lem2} deals with the case $a>b$ and, more importantly, even for $a=b$ the restriction on the size of $\mathcal F$ is weaker. This plays a crucial role in the proof of Frankl's theorem in the next section. Let us obtain a useful corollary of Theorem~\ref{lemft}, overlooked by Frankl and Tokushige, which allows to extend it to the case $a>b$. \begin{cor}\label{corft} Let $n > a+b$, $a> b$, and suppose that $\mathcal F\subset{[n]\choose a},\mathcal G\subset{[n]\choose b}$ are cross-intersecting. Suppose that for some real $\alpha \ge a-b+1$ we have ${n-\alpha \choose n-a}\le |\mathcal F|\le {n-a+b-1\choose n-a}$. Then (\ref{eqft}) holds. \end{cor} \begin{proof} Applying Theorem \ref{thkk}, we may suppose that $\mathcal F = \mathcal L(a,|\mathcal F|), \mathcal G=\mathcal L(b,|\mathcal G|)$. Because of the restriction on the size of $\mathcal F$ we have $[a-b+1]\subset F$ for any $F\in \mathcal F$. Consider two families $\mathcal F':=\{F'\in {[a-b+1,n]\choose b}: F'\cup [a-b]\in \mathcal F\}$, \- $\mathcal G':=\{G'\in {[a-b+1,n]\choose b}: G\in \mathcal G\}$. We have $|\mathcal F'| = |\mathcal F|$, and we may apply Theorem~\ref{lemft} for $\mathcal F', \mathcal G'$ with $n':= n-a+b, \alpha':=\alpha-a+b$. We get that $|\mathcal F|+|\mathcal G'|\le {n-a+b\choose b}+{n-\alpha \choose n-a}-{n-\alpha \choose b}$. We have $|\mathcal G\setminus\mathcal G'|\le {n\choose b}-{n-a+b\choose b}$, and summing this inequality with the previous one we get the statement of the corollary. \end{proof} \vskip+0.1cm \begin{proof}[Proof of Theorem~\ref{lem2}] The proof of Point 1 of the theorem is a simplified version of the proof of Point 2, and thus we first give the proof of Point 2, and then specify the parts which are different in the proof of Point 1. The proof of the theorem uses bipartite graph considerations similar to the ones embodied in Lemma \ref{lembp}. We give two slightly different versions of it in the proof of Point 2 and Point 3.\\ \textbf{Point 2.} Let $\mathcal F,\mathcal G$ be a cross-intersecting pair with {\it minimal} cardinality of $\mathcal F$ among the pairs satisfying the conditions of Point 2 and having maximal possible sum of cardinalities. Due to Theorem~\ref{thkk} we may suppose that $\mathcal F = \mathcal L(a,|\mathcal F|), \mathcal G=\mathcal L(b,|\mathcal G|)$. Let $F\in \mathcal F$ be the set with the largest order in $\mathcal F$. The bound on the cardinality of $|\mathcal F|$ implies $F\supset \{1,\ldots a-b\}$ and $F\cap \{a-b+1,a-b+2\}\ne \emptyset$. Take the largest $l\ge 1$, for which $|F\cap [a-b+2l]|= a-b+l$ is satisfied. Put $L:= [a-b+2l]\setminus F$. Consider the families $\mathcal{A}:=\{A\in\binom{[n]}{a}:A\cap [a-b+2l]= F\cap [a-b+2l]\}$ and $\mathcal{B}:=\{B\in\binom{[n]}{b}:B\cap [a-b+2l]=L\}$. The considerations in this and the next paragraph follow closely the argument from Lemma \ref{lembp}. For any $B\in \mathcal{B},\ F'\in \mathcal{F}\setminus \mathcal{A}$ we have $ F'\cap B\not = \emptyset$. Indeed, take $F''\in \mathcal{F}, F''\cap L = \emptyset$. Since $F$ has the largest order in $\mathcal F$, we have $F''\cap [a-b+2l]\subset F$, and, consequently, $F\in \mathcal{A}$. This implies that the pair $\mathcal{F}\setminus \mathcal{A}, \mathcal G\cup \mathcal{B}$ is cross-intersecting. Next, $|\mathcal{A}|=\binom{n-a+b-2l}{b-l}=|\mathcal{B}|$. Consider a bipartite graph with parts $\mathcal A,\mathcal B$ and edges connecting disjoint sets. Independent sets in this graph correspond to the cross-intersecting pairs of subfamilies of $\mathcal A$ and $ \mathcal B$. This is a regular graph, and, consequently, the maximal independent set has the size $|\mathcal{B}|$. Therefore $|\mathcal{F}\setminus \mathcal{A}|+|\mathcal G\cup \mathcal{B}|\ge |\mathcal{F}|+|\mathcal G|$. Finally, if $|\mathcal F|>{n-j\choose a-j},$ then $F\nsupseteq [j]$, and thus all the sets containing $[j]$ belong to $\mathcal F\setminus {\mathcal A}$. Therefore, ${n-j\choose a-j}\le |\mathcal{F}\setminus \mathcal{A}|<|\mathcal F|$, and we obtain a contradiction with the minimality of $\mathcal F$.\\ \textbf{Point 1.} Let $\mathcal F,\mathcal G$ be a cross-intersecting pair with {\it minimal} cardinality of $\mathcal F$ among the pairs satisfying the conditions of Point 2 and having maximal possible sum of cardinalities. The bound on the cardinality of $|\mathcal F|$ implies $F\supset [1,j]$. Take the largest $l\ge 1$, for which $|F\cap [2l]|= l$ is satisfied. Put $L:= [2l]\setminus F$. Consider the families $\mathcal{A}:=\{A\in\binom{[n]}{a}:A\cap [2l]= F\cap [2l]\}$ and $\mathcal{B}:=\{B\in\binom{[n]}{b}:B\cap [2l]=L\}$. The rest of the proof is the same. \\ \textbf{Point 3.} Let $\mathcal F,\mathcal G$ be a cross-intersecting pair with \textit{maximal} cardinality of $\mathcal F$ among the pairs of maximal sum of cardinalities satisfying the conditions of Point 3. We again w.l.o.g. suppose that $\mathcal F = \mathcal L(a,|\mathcal F|), \mathcal G=\mathcal L(b,|\mathcal G|)$. Let $F\in {[n]\choose a}\setminus \mathcal F$ have the smallest order. Due to the lower bound on $|\mathcal F|$, one of the following conditions hold: either $F\cap [a-b+4]=[a-b+2]$ or for some integer $1\le i\le a-b+1$ one has $F\cap [i+1] = [i]$. Indeed, if the latter condition does not hold, then $F\supset [a-b+2]$. But all the sets containing $[a-b+2]$ and at least one of $\{a-b+3,a-b+4\}$ are in $\mathcal F$, and so $F\cap [a-b+4]=[a-b+2]$. Denote by $t$ the expression $a-b+4$ in the former case, and $i+1$ in the latter case. Put $L:=[t]\setminus F$ and consider two families: $\mathcal{A}:=\{A\in\binom{[n]}{a}:A\cap [t]= F\cap [t]\}$ and $\mathcal{B}:=\{B\in\binom{[n]}{b}:B\cap [t]=L\}$. As in the previous point, the pair $\mathcal{F}\cup\mathcal{A}, \mathcal G\setminus \mathcal{B}$ is cross-intersecting. Moreover, it is easy to see that $a-|F\cap [t]|\ge b-|L|$ and $n-t\ge a-|F\cap [t]|+b-|L|=a+b-t$. Consequently, $|\mathcal A| = {n-t\choose a-|F\cap [t]|}\ge {n-t\choose b-|L|} = |\mathcal B|$. Therefore, arguing as in the previous point, we conclude that $|\mathcal F| = {n-j\choose a-j}$. \end{proof} \section{The diversity of intersecting families}\label{sec3} For a family $\mathcal F$ the \textit{diversity} $\gamma(\mathcal F)$ is the quantity $|\mathcal F|-\Delta(\mathcal F)$, where $\Delta(\mathcal F):=\max_{i\in[n]}\big|\{F:i\in F\in \mathcal F\}\big|$. This notion was studied, in particular, in \cite{LP}. In this section we discuss the connection between the diversity and the size of an intersecting family. The following theorem was de-facto proved by Frankl \cite{Fra1} for \textit{integer} $u$ (with the difference that Frankl uses $\Delta(\mathcal F)$ instead of $\gamma(\mathcal F)$ to bound the cardinality of the family): \begin{thm}\label{thm1} Let $n>2k>0$ and $\mathcal F\subset {[n]\choose k}$ be an intersecting family. Then, if $\gamma(\mathcal F)\ge {n-u-1\choose n-k-1}$ for some real $3\le u\le k$, then \begin{equation}\label{eq01}|\mathcal F|\le {n-1\choose k-1}+{n-u-1\choose n-k-1}-{n-u-1\choose k-1}.\end{equation} \end{thm} The bound from Theorem \ref{thm1} is sharp for integer $u$, as witnessed by the example $\mathcal A:=\{A\in {[n]\choose k}:[2,u]\subset A\}\cup\{A\in{[n]\choose k}: 1\in A, [2,u]\cap A\ne \emptyset\}.$ We note that the case $u = k$ of Theorem~\ref{thm1} is precisely the Hilton--Milner theorem. Below we give a new proof of Theorem~\ref{thm1}, which is much simpler than the proof of the integral case in \cite{Fra1}. The proof makes use of the Kruskal--Katona theorem and the methods developed in the previous sections. For any disjoint $I,J\subset [n]$ define $$\mathcal F(\bar I J):=\{F\setminus J: F\in \mathcal F, J\subset F, F\cap I=\varnothing\} \subset {[n]\setminus (I\cup J)\choose k-|J|}.$$ \begin{proof}[Proof of Theorem~\ref{thm1}] Let $\mathcal F$ have maximal cardinality among the families satisfying the condition of the theorem.\\ \textbf{Case 1:} $\pmb{\gamma(\mathcal F)\le {n-4\choose k-3}}$. Let $1$ be the most popular element in $\mathcal F$. Consider the families $\mathcal F(1)$ and $\mathcal F(\bar 1)$. It is clear that $|\mathcal F| = |\mathcal F(1)|+|\mathcal F(\bar 1)|$. These two families are cross-intersecting and, moreover, we have ${n-u-1\choose n-k-1}\le |\mathcal F(\bar 1)|\le {n-4\choose n-k-1}$. Using Corollary~\ref{corft} for $n':=n-1, a := k, b:=k-1,$ we get the statement of the theorem.\\ \textbf{Case 2:} $ \pmb{{n-4\choose k-3}<\gamma(\mathcal F)\le {n-3\choose k-2}+{n-4\choose k-2}}$. This case is treated analogously, with the only difference that instead of Corollary~\ref{corft} we use Point 2 of Theorem~\ref{lem2} with $n':=n-1, a := k, b:=k-1, j=3$.\\ \textbf{Case 3:} $\pmb{\gamma(\mathcal F)> {n-3\choose k-2}+{n-4\choose k-2}}$. When the diversity is large, we cannot proceed as in Part 2 of Theorem~\ref{lem2}, since we cannot guarantee that, after passing to the lexicographically minimal elements, the set $L$ of maximal order will satisfy the necessary condition $|L\cap [2l]|=l$ for some $l\ge 1$. (We cannot neither apply nor imitate the proof of Part 2 of Theorem~\ref{lem2} in this case.) We will proceed differently. First we show that we may assume that $\mathcal F$ is shifted. \begin{lem}\label{lemdiv} For any intersecting family $\mathcal F\subset{[n]\choose k}$ and any $i,j$, satisfying $|\mathcal F(i)|\ge |\mathcal F(j)|$, we have $\gamma(\mathcal F)-\gamma(S_{i,j}(\mathcal F))\le {n-3\choose k-2}$. \end{lem} \begin{proof} Indeed, it is easy to see that the families $\mathcal F(i\bar j),\mathcal F(j\bar i)$ are cross-intersecting. The assumption of the lemma implies $|\mathcal F(i\bar j)|\ge |\mathcal F(j\bar i)|$, and thus by Theorem~\ref{thkk} we have $|\mathcal F(j\bar i)|\le {n-3\choose k-2}$ (see the remark after the theorem). On the other hand, after the $(i,j)$-shift the degree of $i$ cannot increase by more than $|\mathcal F(j\bar i)|$, since it is only in the sets from $\mathcal F(j\bar i)$ that the element $j$ may be replaced by $i$. Therefore, $\gamma(\mathcal F)-\gamma(S_{i,j}(\mathcal F))\le |\mathcal F(j\bar i)|\le {n-3\choose k-2}$. \end{proof} Rearrange the elements in the order of decreasing degree, and do consecutively the $(1,j)$-shifts, $j=2,3,\ldots$, then the $(2,j)$-shifts, $j=3,4,\ldots$ etc., until either the family becomes shifted, or the diversity of the family becomes at most ${n-3\choose k-2}+{n-4\choose k-2}$. We denote the obtained family by $\mathcal F$ again. In the latter case by Lemma~\ref{lemdiv} we have $\gamma(\mathcal F)\ge {n-4\choose k-2}=\frac{n-k-1}{k-2}{n-4\choose k-3}>{n-4\choose k-3}.$ Therefore, this case is reduced to Case 2.\\ Finally, what if $\mathcal F$ is shifted? We may suppose that\\ 1. $\gamma(\mathcal F)>{n-3\choose k-2}+{n-4\choose k-2}$, \\ 2. $\mathcal F$ is shifted,\\ 3. $\mathcal F$ has the smallest diversity among the families $\mathcal F'$ of maximal size with $\gamma(\mathcal F')\ge{n-3\choose k-2}$.\\ The last inequality may look strange, since it does not coincide with the inequality defining Case 3. However, since we know that the theorem holds for families with diversity between ${n-3\choose k-2}$ and ${n-3\choose k-2}+{n-4\choose k-2}$ with $u=3$, we may include all potential families of maximal size with such diversity to the class of families in question. The element $1$ is the most popular among the sets in $\mathcal F$. Find the set $F\in \mathcal F$ of maximal order and, in terms of the proof from Section~\ref{sec1}, apply Lemma~\ref{lembp} to get a family $\mathcal F'=(\mathcal F\setminus {\mathcal A})\cup {\mathcal B}$ of smaller diversity and at least as large as $\mathcal F$. Note that, again in terms of Lemma~\ref{lembp}, $l\ge 2$ and thus $|{\mathcal A}|={n-2l\choose k-l}\le {n-4\choose k-2}.$ Therefore, we have $\gamma(\mathcal F')\ge \gamma(\mathcal F)- |\mathcal A| \ge {n-3\choose k-2}$, which is a contradiction with the choice of $\mathcal F$.\end{proof} \section{Conclusion} Results of the type presented in this paper has proven to be useful in other questions concerning intersecting and cross-intersecting families. In particular, new results concerning the structure of intersecting and cross-intersecting families were obtained in \cite{FK8}, and degree versions of results on intersecting families were obtained in \cite{Kup10}. There are several questions that remain and that are important for applications. The following problem was proposed by Frankl: is it true that $\gamma(\mathcal F)\le {n-3\choose k-2}$ for any $n\ge 3k$? This is easy to verify for shifted families. More generally, how large the diversity can be for different values of $n, 2k\le n\le 3k$? The following families are the natural candidates of families with the largest diversity for different ranges of $n$: $\mathcal F_i:=\{F\in {[n]\choose k}: |F\cap [2i+1]|\ge i+1\}$. Another problem driven by the applications is the following strengthening of the Erd\H os--Ko--Rado theorem: \begin{pro} Given an intersecting family $\mathcal F$ of $k$-sets of $[n]$ for $n > 2k$, prove that $\Delta(\mathcal F)+C\gamma(\mathcal F) \le {n-1\choose k-1}$ for the largest possible $C> 1$. \end{pro} \textsc{Acknowledgements. } We would like to thank Peter Frankl and the anonymous referee for bringing several references to our attention, and for useful comments on the manuscript that helped to improve the presentation.
{ "timestamp": "2017-10-20T02:02:39", "yymm": "1611", "arxiv_id": "1611.03129", "language": "en", "url": "https://arxiv.org/abs/1611.03129", "abstract": "In this paper we present a simple unifying approach to prove several statements about intersecting and cross-intersecting families, including the Erd\\H os--Ko--Rado theorem, the Hilton--Milner theorem, a theorem due to Frankl concerning the size of intersecting families with bounded maximal degree, and versions of results on the sum of sizes of non-empty cross-intersecting families due to Frankl and Tokushige. Several new stronger results are also obtained.Our approach is based on the use of regular bipartite graphs. These graphs are quite often used in Extremal Set Theory problems, however, the approach we develop proves to be particularly fruitful.", "subjects": "Combinatorics (math.CO); Discrete Mathematics (cs.DM)", "title": "Regular bipartite graphs and intersecting families", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9886682458008671, "lm_q2_score": 0.8267117983401363, "lm_q1q2_score": 0.8173437034478227 }
https://arxiv.org/abs/1604.04028
Partitions with fixed largest hook length
Motivated by a recent paper of Straub, we study the distribution of integer partitions according to the length of their largest hook, instead of the usual statistic, namely the size of the partitions. We refine Straub's analogue of Euler's Odd-Distinct partition theorem, derive a generalization in the spirit of Alder's conjecture, as well as a curious analogue of the first Rogers-Ramanujan identity. Moreover, we obtain a partition theorem that is the counterpart of Euler's pentagonal number theory in this setting, and connect it with the Rogers-Fine identity. We concludes with some congruence properties.
\section{Introduction}\label{sec1} A \emph{partition} \cite{Andr} $\pi$ of an integer $n\in\mathbb{N}$ is a finite sequence of positive integers $(\pi_{1}, \pi_{2}, \ldots, \pi_{r})$ such that $\pi_{1}\geq \pi_{2}\geq\cdots\geq \pi_{r}\geq 1$ and $\pi_{1}+\pi_{2}+\cdots+\pi_{r}=n$. When $n=0$, we consider the empty partition as the only partition of $0$, and for the most part of this paper, we choose to neglect the empty partition. By convention \cite{Andr}, the integers $\pi_{1}, \pi_{2}, \cdots, \pi_{r}$ are called the \emph{parts} of $\pi$, with $\pi_{1}$ being its largest part, $\ell(\pi):=r$ the number of parts and $|\pi|:=n$ the \emph{size} of $\pi$. Such a partition $\pi$ is frequently represented by its \emph{Young diagram} (or \emph{Ferrers graph}) \cite[Chapter 1.3]{Andr}, which we take to be a left-justified array of square boxes with $r$ rows such that the $i$-th row consists of $\pi_{i}$ boxes (see Figure~\ref{yd}). This graphical representation of partitions naturally gives rise to further statistics defined on each partition. A notable one is the notion of \emph{hook length}. Each box $\nu$ is assigned a \emph{hook}, which is composed of the box $\nu$ itself as well as boxes to the right of $\nu$ (which is called the \emph{arm}, with its length denoted as $a(\nu)$) and below $\nu$ (which is called the \emph{leg}, with its length denoted as $l(\nu)$). The \emph{hook length} of $\nu$ is the number of boxes the hook consists of, i.e., $h(\nu)=a(\nu)+l(\nu)+1$. In particular, the largest hook of $\pi$, with length denoted as $\Gamma(\pi)$, traverses the leftmost column as well as the topmost row of cells, i.e., \begin{align}\label{Gamma} \Gamma(\pi)=\pi_1+\ell(\pi)-1. \end{align} \begin{figure} \begin{tikzpicture}[scale=0.8] \draw (0,0) grid (1,3); \draw (1,1) grid (2,3); \draw (0.5,0.5) node{$1$}; \draw (0.5,1.5) node{$3$}; \draw (0.5,2.5) node{$4$}; \draw (1.5,1.5) node{$1$}; \draw (1.5,2.5) node{$2$}; \end{tikzpicture} \caption{Young diagram for $\pi=(2,2,1)$ with hook lengths filled in} \label{yd} \end{figure} In 1748, Euler \cite{Eu} proved arguably the most celebrated partition theorem. When phrased in ``partition language'', it goes as follows. \begin{theorem}\label{od} Given any $n\in\mathbb{N}$, there are as many partitions of $n$ into distinct parts as into odd parts. \end{theorem} In the following some three hundred years, numerous partition theorems have been discovered and stated in the form of ``there are as many partitions of $n$ satisfying condition $A$ as those satisfying condition $B$''. All of these results keep the number being partitioned, i.e., the size of the partition fixed. In a recent paper by Straub \cite{Str}, however, he chose to fix the largest hook length instead, and obtained among other things, the following nice analogue of Theorem~\ref{od}. \begin{theorem}[Theorem 1.4 in \cite{Str}] \label{ana-od} The number of partitions into distinct parts with perimeter $M$ is equal to the number of partitions into odd parts with perimeter $M$. Both are enumerated by the Fibonacci number $F_M$. \end{theorem} Note that Straub's use of the term \emph{perimeter} was following Corteel and Lovejoy \cite[Section 4.2]{CL} (up to a shift by $1$), and it is equivalent to our largest hook length $\Gamma(\pi)$ for a given partition $\pi$. Straub commented on his own analogue as ``missing from the literature on partitions'', and it certainly motivates us to explore further down the road and reveal more analogues with $\Gamma(\pi)$ replacing $|\pi|$, which have been unfortunately overlooked in the past. We present one of our results below, which can be viewed as an analogue of Euler's famous pentagonal number theorem \cite{Eu1}, see also \cite[Corollary~1.7]{Andr}. \begin{Def} For all positive integer $n$, let $\mathcal{H}(n)$ (resp. $h(n)$) be the set (resp. the number) of partitions $\pi$ with $\Gamma(\pi)=n$. Moreover, we consider the following subsets with respect to further restrictions, and their cardinalities will be denoted as $h_{\star}(n)$ respectively. \begin{itemize} \item $\mathcal{H}_{\mathcal{D}}(n):$ the set of partitions in $\mathcal{H}(n)$ with distinct parts; \item $\mathcal{H}_{\mathcal{O}}(n):$ the set of partitions in $\mathcal{H}(n)$ with purely odd parts; \item $\mathcal{H}_{\mathcal{D},\mathcal{O}}(n):$ the set of partitions in $\mathcal{H}(n)$ with distinct parts and the number of parts being odd; \item $\mathcal{H}_{\mathcal{D},\mathcal{E}}(n):$ the set of partitions in $\mathcal{H}(n)$ with distinct parts and the number of parts being even. \end{itemize} \end{Def} \begin{theorem}\label{ana-pent} For all positive integer $n$, \begin{align}\label{id:ana-pent} e(n) &:=h_{\mathcal{D},\mathcal{E}}(n)-h_{\mathcal{D},\mathcal{O}}(n)= \begin{cases} 0\, &n\equiv 0,3\pmod {6};\cr -1\, &n\equiv 1,2\pmod {6};\cr 1\, &n\equiv 4,5\pmod {6}.\end{cases} \end{align} \end{theorem} The rest of the paper is organized as follows. In section~\ref{sec:ana-od} we study function $h(n)$ and its various relatives from scratch, deduce their generating functions combinatorially. This eventually leads to three refinements of Theorem~\ref{ana-od} as well as two generalizations. We continue to consider a natural analogue of Euler's pentagonal number theorem in section~\ref{sec:ana-pent} and establish Theorem~\ref{ana-pent}. Then Subbarao's crucial observation \cite{Sub} on Franklin's involution leads us to identity \eqref{id: Andrews} that encompasses both the pentagonal number theorem and Theorem~\ref{ana-pent}. In the last section, some congruence properties are obtained for $h_{\mathcal{D}}(n)$ and we conclude with some remarks. \section{Straub's analogue of Euler's theorem}\label{sec:ana-od} \subsection{Refinement of Straub's analogue} We use $\mathcal{P}$ (resp. $\mathcal{D}$, $\mathcal{O}$) to denote the set of non-empty partitions with parts being unrestricted (resp. distinct, odd). \begin{Def} Let $H(q):=\sum_{\pi\in\mathcal{P}}q^{\Gamma(\pi)}=\sum_{n=1}^{\infty}h(n)q^n$ be the generating function for $h(n)$. Similarly we define $H_{\mathcal{D}}(q)$ (resp. $H_{\mathcal{O}}(q)$) as the generating function for $h_{\mathcal{D}}(n)$ (resp. $h_{\mathcal{O}}(n)$). And we shall also consider the generating functions for the following refinements. \begin{itemize} \item $H(x,y,q):=\sum\limits_{\pi\in\mathcal{P}}x^{\pi_1}y^{\ell(\pi)}q^{\Gamma(\pi)}=\sum\limits_{n=1}^{\infty}\sum\limits_{m=1}^{n}h(m,n)x^my^{n+1-m}q^n$; \item $H_{\mathcal{D}}(x,y,q):=\sum\limits_{\pi\in\mathcal{D}}x^{\pi_1}y^{\ell(\pi)}q^{\Gamma(\pi)}=\sum\limits_{n=1}^{\infty}\sum\limits_{m=1}^{n}h_{\mathcal{D}}(m,n)x^my^{n+1-m}q^n$; \item $H_{\mathcal{O}}(x,y,q):=\sum\limits_{\pi\in\mathcal{O}}x^{\pi_1}y^{\ell(\pi)}q^{\Gamma(\pi)}=\sum\limits_{n=1}^{\infty}\sum\limits_{m=1}^{n}h_{\mathcal{O}}(m,n)x^my^{n+1-m}q^n$. \end{itemize} \end{Def} \begin{remark} Note that the relationship between the exponents of $x,y$ and $q$ is a direct application of \eqref{Gamma}. And clearly $H(1,1,q)=H(q)$, $H_{\mathcal{D}}(1,1,q)=H_{\mathcal{D}}(q)$ and $H_{\mathcal{O}}(1,1,q)=H_{\mathcal{O}}(q)$, with $h(n)=\sum_{m=1}^{n}h(m,n)$, $h_{\mathcal{D}}(n)=\sum_{m=1}^{n}h_{\mathcal{D}}(m,n)$ and $h_{\mathcal{O}}(n)=\sum_{m=1}^{n}h_{\mathcal{O}}(m,n)$. \end{remark} Now we proceed to derive the main results of this subsection, namely the closed form for $H(x,y,q)$, $H_{\mathcal{D}}(x,y,q)$ and $H_{\mathcal{O}}(x,y,q)$, by analysing the so-called \emph{profile} of a given partition. Following \cite{KN}, for a given partition, we use profile to denote the set of southmost and eastmost edges of the boxes in its Young diagram, as depicted in Figure~\ref{profile}. And when we label the horizontal (resp. vertical) edges in the profile with $E$ (resp. $N$), we see that the profile of any non-empty partition in $\mathcal{H}(n)$ is in bijection with a word consists of two letters $E$ and $N$, whose length is $n+1$ and always begins with $E$, ends with $N$. This observation essentially serves as a quick bijective proof of the fact that $h(n)=2^{n-1}$, see Corollary~\ref{h(n)}. And it sets us up for our main theorem. \begin{figure} \begin{tikzpicture}[scale=1] \draw (0,0) grid (1,3); \draw (1,1) grid (2,3); \draw (0.5,-0.3) node{$E$}; \draw (1.3,0.3) node{$N$}; \draw (1.6,0.8) node{$E$}; \draw (2.3,1.5) node{$N$}; \draw (2.3,2.5) node{$N$}; \draw (4,1.2) node{$\Longleftrightarrow$}; \draw (6,1.2) node{$ENENN$}; \end{tikzpicture} \caption{The profile of $\pi=(2,2,1)$ labelled with $E$ and $N$} \label{profile} \end{figure} \begin{theorem}\label{gfxyq} \begin{align} H(x,y,q) &= xyq(1+(xq+yq)+(xq+yq)^2+\cdots)=\dfrac{xyq}{1-(xq+yq)},\label{gfH}\\ H_{\mathcal{D}}(x,y,q) &= xyq(1+(xq+xyq^2)+(xq+xyq^2)^2+\cdots)=\dfrac{xyq}{1-(xq+xyq^2)},\label{gfHD}\\ H_{\mathcal{O}}(x,y,q) &= xyq(1+(yq+x^2q^2)+(yq+x^2q^2)^2+\cdots)=\dfrac{xyq}{1-(yq+x^2q^2)}\label{gfHO}. \end{align} \end{theorem} \begin{proof} We show \eqref{gfH} first, and explain the changes need to be made for deriving \eqref{gfHD} and \eqref{gfHO}. Given any partition in $\mathcal{P}$, consider the word corresponding to its profile. As noted earlier, this word must begin with $E$ and end with $N$, together they contribute the factor $xyq$ (apply \eqref{Gamma} to see why it is $xyq$, not $xyq^2$). And all the remaining letters in between could be either an $E$ that contributes $xq$, or an $N$ that contributes $yq$, this explains the factor $(1+(xq+yq)+(xq+yq)^2+\cdots)$ and thus establishes \eqref{gfH}. Next for partitions in $\mathcal{D}$, the same restriction applies to the first and the last letters, so we also have the factor $xyq$. But now for the letters in the middle, ``having distinct parts'' translates to ``having no consecutive $N$''. So we could have either an $E$ still contributing $xq$, or a pair $NE$ contributing $xyq^2$. This justifies the factor $(1+(xq+xyq^2)+(xq+xyq^2)^2+\cdots)$ and proves \eqref{gfHD}. Finally for partitions in $\mathcal{O}$, all the parts are odd, which forces the corresponding word to have an even number of $E$s between two consecutive $N$, and an even number of $E$s after the first $E$. This explains the factor $(1+(x^2q^2+yq)+(x^2q^2+yq)^2+\cdots)$ and completes the proof. \end{proof} The next two corollaries follow immediately from Theorem~\ref{gfxyq}. \begin{corollary}\label{h(n)} For positive integer $n$, $h(n)=2^{n-1}$. \end{corollary} \begin{proof} Put $x=y=1$ in \eqref{gfH} to get $H(q)=q/(1-2q)=\sum_{n=1}^{\infty}2^{n-1}q^n$. \end{proof} \begin{corollary} Theorem~\ref{ana-od} is true. \end{corollary} \begin{proof} Put $x=y=1$ in both \eqref{gfHD} and \eqref{gfHO} to get $H_{\mathcal{D}}(q)=H_{\mathcal{O}}(q)=q/(1-q-q^2)$, which is well-known as the generating function for the Fibonacci numbers. \end{proof} \begin{remark} Euler's original proof of Theorem~\ref{od} used generating function. Interested readers are referred to \cite[Chapter~2,5]{AE} for a gentle introduction and \cite{Andr} for more serious discussions. Straub \cite{Str} gave two proofs of his analogue, but neither of them used generating function. So our Theorem~\ref{gfxyq} together with the above corollary can be viewed as the first generating function proof of Theorem~\ref{ana-od}. \end{remark} In his paper, Straub recalled two refinements \cite[Example 1.7, 1.8]{Str} of Euler's theorem due to Fine \cite{Fine}, and raised the question of seeking similar refinements for his analogue. We supply three candidates here, each of them is a direct result of evaluating $x,y$ appropriately in \eqref{gfHD} and \eqref{gfHO}. \begin{theorem}\label{ref1} For integers $n\geq 1$, $1\leq k \leq \lceil n/2\rceil$. The number of partitions $\pi$ into exactly $k$ parts, all distinct, and with $\Gamma(\pi)=n$ is equal to the number of partitions $\lambda$ into odd parts with $\lambda_1=2k-1$ and $\Gamma(\lambda)=n$. Both are enumerated by $\binom{n-k}{k-1}$. \end{theorem} \begin{proof} Take $x=1$ in \eqref{gfHD} and $y=1$ in \eqref{gfHO}, then we see $H_{\mathcal{D}}(1,y,q)=\dfrac{yq}{1-(q+yq^2)}$ and $H_{\mathcal{O}}(x,1,q)=\dfrac{xq}{1-(q+x^2q^2)}$, both are specializations of $\dfrac{aq}{1-(q+bq^2)}$. Now we simply compare the coefficients of $ab^{k-1}q^n$ from both cases to see they are equal as claimed. Next for the exact enumeration, note that to resemble a term $ab^{k-1}q^n$, besides the numerator $aq$, we need to extract from $(k-1)+(n-1-2(k-1))=n-k$ copies of the factor $q+bq^2$, wherein $k-1$ copies we must choose $bq^2$ to guarantee $b^{k-1}$, and the remaining $n-2k+1$ copies all choose $q$ to make the total power of $q$ exactly $n$, hence the count $\binom{n-k}{k-1}$. \end{proof} \begin{example} Table~\ref{tab1} lists all partitions $\pi$ into distinct parts with $\Gamma(\pi)=9, \ell(\pi)=4$, and the odd partitions $\lambda$ they get matched with such that $\Gamma(\lambda)=9, \lambda_{1}=7$. Both columns contain $\binom{9-4}{4-1}=\binom{5}{3}=10$ partitions in total. \end{example} \begin{table}[htbp] \centering \caption{} \begin{tabular}{|c|c|} \hline distinct partitions & odd partitions\\ \hline (6,5,4,3) &(7,7,7)\\ \hline (6,5,4,2) &(7,7,5)\\ \hline (6,5,4,1) &(7,7,3)\\ \hline (6,5,3,2) &(7,7,1)\\ \hline (6,5,3,1) &(7,5,5)\\ \hline (6,5,2,1) &(7,5,3)\\ \hline (6,4,3,2) &(7,5,1)\\ \hline (6,4,3,1) &(7,3,3)\\ \hline (6,4,2,1) &(7,3,1)\\ \hline (6,3,2,1) &(7,1,1)\\ \hline \end{tabular} \label{tab1} \end{table} \begin{theorem} For integers $n\geq 1$, $\lceil (n+1)/2\rceil\leq k \leq n$. The number of partitions $\pi$ into distinct parts, with $\pi_1=k$ and $\Gamma(\pi)=n$ is equal to the number of partitions $\lambda$ into odd parts with $\lambda_1/2+\ell(\lambda)=k+1/2$ and $\Gamma(\lambda)=n$. Both are enumerated by $\binom{k-1}{n-k}$. \end{theorem} \begin{proof} Take $y=1$ in \eqref{gfHD} and $x=x^{1/2},y=x$ in \eqref{gfHO}, then we see $H_{\mathcal{D}}(x,1,q)=\dfrac{xq}{1-(xq+xq^2)}$ and $H_{\mathcal{O}}(x^{1/2},x,q)=\dfrac{x^{3/2}q}{1-(xq+xq^2)}$, which implies the statement in the theorem immediately. And we can make the exact count similarly as in the proof of Theorem~\ref{ref1}. \end{proof} \begin{example} Table~\ref{tab2} lists all partitions $\pi$ into distinct parts with $\Gamma(\pi)=8, \pi_{1}=6$, and the odd partitions $\lambda$ they get matched with such that $\Gamma(\lambda)=8, \lambda_1/2+\ell(\lambda)=13/2$. Both columns contain $\binom{6-1}{8-6}=\binom{5}{2}=10$ partitions in total. \end{example} \begin{table}[htbp]\caption{} \centering \begin{tabular}{|c|c|} \hline distinct partitions & odd partitions\\ \hline (6,5,4) &(5,5,5,5)\\ \hline (6,5,3) &(5,5,5,3)\\ \hline (6,5,2) &(5,5,5,1)\\ \hline (6,5,1) &(5,5,3,3)\\ \hline (6,4,3) &(5,5,3,1)\\ \hline (6,4,2) &(5,5,1,1)\\ \hline (6,4,1) &(5,3,3,3)\\ \hline (6,3,2) &(5,3,3,1)\\ \hline (6,3,1) &(5,3,1,1)\\ \hline (6,2,1) &(5,1,1,1)\\ \hline \end{tabular} \label{tab2} \end{table} \begin{theorem} For integers $n\geq 1$, $0\leq k \leq n-1$. The number of partitions $\pi$ into distinct parts, with $\mathrm{rank}(\pi)=k$ and $\Gamma(\pi)=n$ is equal to the number of partitions $\lambda$ into odd parts with $\ell(\lambda)=k+1$ and $\Gamma(\lambda)=n$, where $\mathrm{rank}(\pi):=\pi_1-\ell(\pi)$. Both are enumerated by $\binom{(n+k-1)/2}{k}$. Note that $n-1$ and $k$ always have the same parity due to their definition. \end{theorem} \begin{proof} Take $y=x^{-1}$ in \eqref{gfHD} and $x=1$ in \eqref{gfHO}, then we see $H_{\mathcal{D}}(x,x^{-1},q)=\dfrac{q}{1-(xq+q^2)}$ and $H_{\mathcal{O}}(1,y,q)=\dfrac{yq}{1-(yq+q^2)}$, which implies the statement in the theorem immediately. The exact count follows analogously as the previous two theorems. \end{proof} \begin{example} Table~\ref{tab3} lists partitions $\pi$ into distinct parts with $\Gamma(\pi)=7, \mathrm{rank}(\pi)=2$, and the odd partitions $\lambda$ they get matched with such that $\Gamma(\lambda)=7, l(\lambda)=3$. Both columns contain $\binom{(7+2-1)/2}{2}=6$ partitions in total. \end{example} \begin{table}[htbp]\caption{} \centering \begin{tabular}{|c|c|} \hline distinct partitions & odd partitions\\ \hline (5,4,3) &(5,5,5)\\ \hline (5,4,2) &(5,5,3)\\ \hline (5,4,1) &(5,5,1)\\ \hline (5,3,2) &(5,3,3)\\ \hline (5,3,1) &(5,3,1)\\ \hline (5,2,1) &(5,1,1)\\ \hline \end{tabular} \label{tab3} \end{table} \begin{remark} All the above three refinements could be given direct bijective proofs as Straub did for Theorem~\ref{ana-od}. The idea is to correspond each $E$ step (weighted as $xq$) in distinct partition with each $N$ step (weighted as $yq$) in odd partition, while each $NE$ step (weighted as $xyq^2$) in distinct partition corresponds to each $EE$ step (weighted as $x^2q^2$) in odd partition, then keep track of how many steps of these types in them. We leave the details to interested readers. \end{remark} \subsection{Generalization to $d$-distinct partitions} Following \cite{AE}, we call a partition $d$-distinct if the difference between any two of its parts is at least $d$. In 1956, Alder~\cite{Ald} investigated $q_d(n)$ and $Q_d(n)$, the number of partitions of $n$ into $d$-distinct parts and into parts $\equiv \pm 1\pmod {d+3}$, respectively, and he conjectured that \begin{align}\label{AConj} q_d(n)\geq Q_d(n). \end{align} See \cite[Chapter 4.3]{AE} for more information on this conjecture, and see \cite{AJO, Andr1, Yee} for its proof. Alder's pre-conjecture concerns us here because when $d=1,2$ we actually get ``equality'' instead of ``inequality'' in \eqref{AConj}. Indeed, the $d=1$ case is Euler's identity (Theorem~\ref{od}), and $d=2$ produces the first Rogers-Ramanujan identity. While $q_d(n)=Q_d(n)$ is too good to be true for all $d$, in our case however, we can generalize Straub's analogue (Theorem~\ref{ana-od}) to $d$-distinct partitions and get equalities for all $d$. \begin{Def} For positive integers $n$ and $d$, we denote $h_d(n)$ the number of $d$-distinct partitions $\pi$ with $\Gamma(\pi)=n$, and $f_d(n)$ the number of partitions $\lambda$ into parts $\equiv 1 \pmod {d+1}$ with $\Gamma(\lambda)=n$. Similarly, we denote $\mathcal{H}_d$ the set of all $d$-distinct partitions, and $\mathcal{F}_d$ the set of all partitions into parts $\equiv 1 \pmod{d+1}$. \end{Def} Note that with this new definition, we have $h_1(n)=h_{\mathcal{D}}(n), f_1(n)=h_{\mathcal{O}}(n)$ and $\mathcal{H}_1=\mathcal{D},\mathcal{F}_1=\mathcal{O}$. \begin{theorem} For all $d\geq 1$, we have \begin{align} H_d(x,y,q) &:=\sum\limits_{\pi\in\mathcal{H}_d}x^{\pi_1}y^{\ell(\pi)}q^{\Gamma(\pi)}\label{gfHd}\\ &=xyq(1+(xq+x^dyq^{d+1})+(xq+x^dyq^{d+1})^2+\cdots)=\dfrac{xyq}{1-(xq+x^dyq^{d+1})},\nonumber \\ F_d(x,y,q) &:=\sum\limits_{\pi\in\mathcal{F}_d}x^{\pi_1}y^{\ell(\pi)}q^{\Gamma(\pi)}\label{gfMd}\\ &=xyq(1+(yq+x^{d+1}q^{d+1})+(yq+x^{d+1}q^{d+1})^2+\cdots)=\dfrac{xyq}{1-(yq+x^{d+1}q^{d+1})}.\nonumber \end{align} In particular, we have for all $n\geq 1$ \begin{align}\label{d-distinct-id} h_d(n)=f_d(n). \end{align} \end{theorem} \begin{proof} It will suffice to show \eqref{gfHd} and \eqref{gfMd}, since \eqref{d-distinct-id} can be easily deduced from them by taking $x=y=1$ and then extracting the coefficient of $q^n$ from both. The proof of \eqref{gfHd} and \eqref{gfMd} is analogous to that of \eqref{gfHD} and \eqref{gfHO}. Given $\pi\in\mathcal{H}_d$, the condition of being $d$-distinct leads to the restriction on $\pi$'s profile. Indeed, the first and the last letter of the corresponding word still produces $xyq$, while the middle letters could be either an $E$ contributing $xq$, or a single $N$ followed by $d$ copies of $E$, contributing $(yq)(xq)\cdots(xq)=x^dyq^{d+1}$, this gives us \eqref{gfHd}. Similarly, for $\mathcal{F}_d$, to keep the parts always $\equiv 1 \pmod{d+1}$, we must have either an $N$ contributing $yq$, or consecutive $d+1$ copies of $E$ contributing $x^{d+1}q^{d+1}$, which explains the change in \eqref{gfMd} and completes the proof. \end{proof} When $d=2$, $2$-distinct partitions are exactly those generated by the series side of the first Rogers-Ramanujan identity, but partitions into parts $\equiv 1 \pmod 3$ (or equivalently $\equiv 1,4 \pmod 6$) do not agree with the product side of the first Rogers-Ramanujan identity, which generates partitions into parts $\equiv 1,4 \pmod 5$. Actually, a quick look at Table~\ref{tab4} reveals that with $\Gamma(\pi)\leq 5$, there are as many partitions into parts $\equiv 1,4 \pmod 6$ as there are into parts $\equiv 1,4 \pmod 5$. But when $\Gamma(\pi)\geq 6$, the former set of partitions is smaller. For instance, partition $(6)$ belongs to the latter but not to the former. This observation suggests that if one wants to force modular $5$ instead of modular $6$, some further restrictions need to be added, this is reminiscent of Schur's fix for $q_3(n)>Q_3(n)$ (see \cite[Section 4.4]{AE}). We are led to another partition theorem that involves $d$-distinct partitions. \begin{Def}\label{defgd} For positive integers $n$ and $d$, we denote $\mathfrak{G}_d$ the set of all partitions $\pi=(\pi_1,\pi_2,\ldots,\pi_r)$ satisfying the following two conditions, and we denote $g_d(n)$ the number of partitions $\pi\in \mathfrak{G}_d$ with $\Gamma(\pi)=n$. \begin{enumerate}[i.] \item $\pi_i\equiv 1 \text{ or } d+2 \pmod{2d+1}$, for $i=1,2,\ldots,r$; \item $\pi_i-\pi_{i+1}\leq 2d+1$, for $i=1,2,\ldots,r$, where $\pi_{r+1}=0$, with strict sign taken whenever $\pi_i\equiv 1\pmod{2d+1}$. \end{enumerate} \end{Def} \begin{theorem}\label{gen2} For all $d\geq 1$, we have \begin{align} G_d(x,y,q) &:=\sum\limits_{\pi\in\mathfrak{G}_d}x^{\pi_1}y^{\ell(\pi)}q^{\Gamma(\pi)}=\dfrac{xyq(1-yq+x^{d+1}q^{d+1})}{1-2yq+y^2q^2-x^{2d+1}yq^{2d+2}}. \label{gfGd} \end{align} In particular, we have for all $n\geq 1$ \begin{align}\label{h=f=g} h_d(n)=f_d(n)=g_d(n). \end{align} \end{theorem} \begin{proof} Take $x=y=1$ in \eqref{gfGd} gives \begin{align*} \sum_{n\geq 1}g_d(n)q^n=G_d(1,1,q)=\dfrac{q(1-q+q^{d+1})}{(1-q)^2-q^{2d+2}}=\dfrac{q}{1-q-q^{d+1}}=H_d(1,1,q)=F_d(1,1,q), \end{align*} and we get \eqref{h=f=g}. To prove \eqref{gfGd} we need to translate conditions i and ii for a given partition $\pi\in\mathfrak{G}_d$ into restrictions on the word, say $w_{\pi}$, that corresponds to the profile of $\pi$. We claim that we can always divide $w_{\pi}$ uniquely into blocks of the following types, with all the middle blocks having ``type I, type II, type I, type II,$\ldots$'' alternately. In the following, $j=0,1,2,\ldots$ can be any natural number. \begin{itemize} \item Initial block: A single $E$ followed by $j$ consecutive $N$s. \item Middle block of type I: A ($d+1$)-tuple $EE\cdots E$ followed by $j$ consecutive $N$s. \item Middle block of type II: A ($d+1$)-tuple $NE\cdots E$ followed by $j$ consecutive $N$s. \item Terminal block: consists of a single $N$. \end{itemize} Conversely, given any word that consists of these four types of blocks such that the middle blocks, if any, start with type I and switch between type I and type II, we can uniquely realize it as the profile of certain partition in $\mathfrak{G}_d$. To prove this claim, one can trace the profile of any $\pi\in\mathfrak{G}_d$, from the southwest-most edge to the northeast-most edge to see how $\pi$ is ``outlined'' edge by edge, and at the same time go through the word $w_{\pi}$ letter by letter from left to right. Then we see \begin{itemize} \item The initial block in $w_{\pi}$ corresponds to $j$ copies of $1$ as parts in $\pi$. \item Having a middle block of type I corresponds to making the current largest part (which is necessarily $\equiv 1 \pmod{2d+1}$), say $k(2d+1)+1$, becomes $k(2d+1)+d+2$, and adding $j$ copies of it on top of the current sub-partition. \item Having a middle block of type II corresponds to keeping the current largest part (which is necessarily $\equiv d+2 \pmod{2d+1}$), say $k(2d+1)+d+2$, and adding $j$ copies of $(k+1)(2d+1)+1$ on top of the current sub-partition. \item The terminal block ends the process and we have a complete $\pi$, corresponding to the completed word $w_{\pi}$. \end{itemize} A moment of reflection reveals that in the above process of traversing the profile of $\pi$, the largest part for each sub-partition must be $\equiv 1 \pmod{2d+1}$ (resp. $\equiv d+2 \pmod{2d+1}$) after each block of type II (resp. type I), and the two middle types alternate will force condition ii, vice versa. We provide one example of $d=2$ in Figure~\ref{block}. Next we only need to figure out the generating function for each type of block. \begin{itemize} \item Initial$+$terminal: $xyq(1+yq+y^2q^2+\cdots)=\dfrac{xyq}{1-yq}$. \item Type I: $x^{d+1}q^{d+1}(1+yq+y^2q^2+\cdots)=\dfrac{x^{d+1}q^{d+1}}{1-yq}$. \item Type II: $x^dyq^{d+1}(1+yq+y^2q^2+\cdots)=\dfrac{x^{d}yq^{d+1}}{1-yq}$. \end{itemize} Together we have \begin{align*} G_d(x,y,q) &=\dfrac{xyq}{1-yq}(1+\dfrac{x^{d+1}q^{d+1}}{1-yq}+\dfrac{x^{d+1}q^{d+1}}{1-yq}\dfrac{x^{d}yq^{d+1}}{1-yq}+\dfrac{x^{d+1}q^{d+1}}{1-yq}\dfrac{x^{d}yq^{d+1}}{1-yq}\dfrac{x^{d+1}q^{d+1}}{1-yq}+\cdots)\\ &=\dfrac{\dfrac{xyq}{1-yq}(1+\dfrac{x^{d+1}q^{d+1}}{1-yq})}{1-\dfrac{x^{d+1}q^{d+1}}{1-yq}\dfrac{x^{d}yq^{d+1}}{1-yq}}=\dfrac{xyq(1-yq+x^{d+1}q^{d+1})}{1-2yq+y^2q^2-x^{2d+1}yq^{2d+2}}. \end{align*} This completes the proof. \end{proof} \begin{figure} \begin{tikzpicture}[scale=0.5] \draw (0,0) grid (1,9); \draw (1,3) grid (4,9); \draw (4,4) grid (6,9); \draw (6,7) grid (9,9); \path[draw, line width=2pt] (0,0) -- (1,0) -- (1,3) -- (4,3) -- (4,4) -- (6,4) -- (6,7) -- (9,7) -- (9,9); \draw (10,4.5) node{$\Longleftrightarrow$}; \draw (20,4.5) node{$ENNN\mid EEE\mid NEENNN\mid EEEN\mid N$}; \draw (14,6) node{{\small initial}}; \draw (17,6) node{{\small type I}}; \draw (21,6) node{{\small type II}}; \draw (25,6) node{{\small type I}}; \draw (28,6) node{{\small terminal}}; \end{tikzpicture} \caption{Decomposition of $w_{\pi}$ into blocks for $\pi=(9,9,6,6,6,4,1,1,1)$} \label{block} \end{figure} When we take $d=2$ in \eqref{h=f=g}, we get the following intriguing analogue of the first Rogers-Ramanujan identity. \begin{corollary} For any positive integer $n$, the number of $2$-distinct partitions with perimeter $n$ is equal to the number of partitions into parts $\equiv 1,4 \pmod 5$ with perimeter $n$ and satisfying condition ii with $d=2$ in Definition~\ref{defgd}. \end{corollary} \begin{example} The partitions listed in Table \ref{tab4} all satisfy $\Gamma(\pi)\leq 7$. More precisely, we have $2$-distinct partitions in the first column, paired with partitions into parts $\equiv 1\pmod{3}$ in the second column, and with partitions into parts $\equiv 1,4 \pmod{5}$ satisfying the extra condition ii listed in the third column. \end{example} \begin{table}[htbp]\caption{} \centering \begin{tabular}{|c|c|c|} \hline $2$-distinct & $\equiv 1 \pmod 3$ & condition i, ii\\ \hline (1) & (1) & (1)\\ \hline (2) & (1,1) & (1,1)\\ \hline (3) &( 1,1,1) & (1,1,1)\\ \hline (4) & (1,1,1,1) & (1,1,1,1)\\ (3,1) & (4) & (4)\\ \hline (5) & (1,1,1,1,1) & (1,1,1,1,1)\\ (4,2) & (4,1) & (4,1)\\ (4,1) & (4,4) & (4,4)\\ \hline (6) & (1,1,1,1,1,1) & (1,1,1,1,1,1)\\ (5,3) & (4,1,1) & (4,1,1)\\ (5,2) & (4,4,1) & (4,4,1)\\ (5,1) & (4,4,4) & (4,4,4)\\ \hline (7) & (1,1,1,1,1,1,1) & (1,1,1,1,1,1,1)\\ (6,4) & (4,1,1,1) & (4,1,1,1)\\ (6,3) & (4,4,1,1) & (4,4,1,1)\\ (6,2) & (4,4,4,1) & (4,4,4,1)\\ (6,1) & (4,4,4,4) & (4,4,4,4)\\ (5,3,1) & (7) & (6,4)\\ \hline \end{tabular} \label{tab4} \end{table} \section{An analogue of Euler's pentagonal number theorem}\label{sec:ana-pent} \subsection{Pentagonal number theorem and a new analogue} Straub's novel decision to fix $\Gamma(\pi)$ instead of $|\pi|$ proves successful in Theorem~\ref{ana-od}. This encourages us to seek for other analogues of the classic partition theorems with this new mindset. As mentioned in the introduction, Theorem~\ref{ana-pent} is a new partition theorem that parallels Euler's Pentagonal Number Theorem nicely. We present two proofs of it in this subsetion, and we continue to discuss yet another, purely combinatorial proof and its implications in next subsection. \begin{proof}[1st proof of Theorem~\ref{ana-pent}] For all partitions in $\mathcal{H}_{\mathcal{D}}(n)$, we want to calculate the number of those with even number of parts in excess of those with odd number of parts. To this end, we simply take $x=1,y=-1$ in \eqref{gfHD} and get \begin{align*} \sum\limits_{n=1}^{\infty} e(n)q^n &=H_{\mathcal{D}}(1,-1,q) = \dfrac{-q}{1-q+q^2}= \dfrac{-q}{(1-\omega q)(1-q/\omega)}\\ &=\dfrac{-1}{\sqrt{3}\:i}(\dfrac{1}{1-\omega q}-\dfrac{1}{1-q/\omega})= \dfrac{-1}{\sqrt{3}\:i}\sum\limits_{n=0}^{\infty}(\omega^n-\omega^{-n})q^n, \end{align*} where $\omega=e^{\pi i/3}$ is a sixth root of unity. Now a simple calculation verifies \eqref{id:ana-pent}. \end{proof} In preparation for our second proof of Theorem~\ref{ana-pent}, we first derive the recurrence relation for $h_{\mathcal{D},\mathcal{E}}(n)$ and $h_{\mathcal{D},\mathcal{O}}(n)$. \begin{proposition}\label{rr-hDEO} The numbers $h_{\mathcal{D, E}}(n)$ and $h_{\mathcal{D, O}}(n)$ satisfy the following system of recurrence relations, for all $n\geq 3$: \begin{align} h_{\mathcal{D, E}}(n) &=h_{\mathcal{D, E}}(n-1)+h_{\mathcal{D, O}}(n-2),\label{rrhDE}\\ h_{\mathcal{D, O}}(n) &=h_{\mathcal{D, O}}(n-1)+h_{\mathcal{D, E}}(n-2),\label{rrhDO} \end{align} with the initial values \begin{gather}\label{iv} h_{\mathcal{D},\mathcal{E}}(1)=h_{\mathcal{D, E}}(2) = 0, h_{\mathcal{D, O}}(1)=h_{\mathcal{D, O}}(2)=1.\notag \end{gather} \end{proposition} \begin{proof} We show \eqref{rrhDE} combinatorially by investigating the smallest part, say $\pi_r$, in a given partition $\pi\in\mathcal{H}_{\mathcal{D},\mathcal{E}}(n)$. \begin{itemize} \item $\pi_r=1$. Then $\pi_{r-1}\geq 2$ and we can remove the leftmost column in the Ferrers graph of $\pi$, or equivalently, subtract $1$ from each part of $\pi$. This results in a new partition $\hat{\pi}$ with one fewer parts, thus $\ell(\hat{\pi})$ is odd and $\Gamma(\hat{\pi})=n-2$. This operation is clearly seen to be invertible. Indeed, given a partition $\hat{\pi}\in\mathcal{H}_{\mathcal{D},\mathcal{O}}(n-2)$, we simply add $1$ to each part of $\hat{\pi}$, as well as a new part of $1$ to arrive at a partition $\pi\in\mathcal{H}_{\mathcal{D},\mathcal{E}}(n)$. \item $\pi_r>1$. We also subtract $1$ from each part of $\pi$ to get $\hat{\pi}$. But this time $\hat{\pi}$ has the same number of parts as $\pi$, so $\ell(\hat{\pi})$ remains even and $\Gamma(\hat{\pi})=n-1$. This operation is also invertible. \end{itemize} Combining these two cases gives us \eqref{rrhDE}. \eqref{rrhDO} can be deduced similarly and thus omitted. \end{proof} It is now a routine exercise to check the following formulas for $h_{\mathcal{D},\mathcal{E}}(n)$ and $h_{\mathcal{D},\mathcal{O}}(n)$ using Proposition~\ref{rr-hDEO} and the Pascal relation for the binomial coefficients. So we supply here a direct combinatorial proof that do not need recurrence relations \eqref{rrhDE} and \eqref{rrhDO}. \begin{proposition}\label{formula-hDEO} For any positive integer $n\geq 1$, \begin{align} h_{\mathcal{D},\mathcal{E}}(n) &=\sum_{k\geq0}\binom{n-2k-2}{2k+1},\label{f-hDE}\\ h_{\mathcal{D},\mathcal{O}}(n) &=\sum_{k\geq0}\binom{n-2k-1}{2k}.\label{f-hDO} \end{align} \end{proposition} \begin{proof} We show proof of \eqref{f-hDE} only, since \eqref{f-hDO} can be proved similarly. Recall from the proof of \eqref{gfHD} that for a distinct partition $\pi$, the profile word $w_{\pi}$ can only have $E$s or couples $NE$ in the middle, with the fixed initial $E$ and the terminal $N$. And between $E$ and $NE$, only $NE$ can affect the value of $\ell(\pi)$. Therefore, if $\pi\in\mathcal{H}_{\mathcal{D},\mathcal{E}}(n)$, then with the terminal $N$ accounting for the largest part, we should have an odd number of $NE$s in $w_{\pi}$, say $2k+1$ for some $k\geq 0$. Since the total length of $w_{\pi}$ is $n+1$ (see \eqref{Gamma}), we must have $n+1-2(2k+1)-1-1=n-4k-3$ copies of $E$s remained in the middle, excluding the initial $E$. Now these $n-4k-3$ copies of $E$s and $2k+1$ copies of $NE$s can be arranged in any possible order, which gives rise to $\binom{n-2k-2}{2k+1}$. We sum over all $k\geq 0$ to finish the proof. \end{proof} Our next proof builds on Proposition~\ref{rr-hDEO} and is free of generating function \eqref{gfHD}. \begin{proof}[2nd proof of Theorem~\ref{ana-pent}] For all positive integer $n$, define $e(n):=h_{\mathcal{D},\mathcal{E}}(n)-h_{\mathcal{D},\mathcal{O}}(n)$. Now we can subtract \eqref{rrhDO} from \eqref{rrhDE} to get $e(n)=e(n-1)-e(n-2)$, then iterate it twice we have \begin{align*} e(n)=-e(n-3),\quad \forall n\geq 4. \end{align*} Therefore $e(n)$ has a period of $6$, and it now suffice to calculate $e(1)=e(2)=-1,e(3)=0$ so as to establish \eqref{id:ana-pent} and completes the proof. \end{proof} \begin{remark} Not surprisingly, we are not the first to stumble on the sequence $\{h_{\mathcal{D},\mathcal{E}}(n)\}_{n\geq 1}$, see \href{http://oeis.org/A024490}{\tt{oeis:A024490}}. This is where we tracked Munarini and Salvi's paper \cite{MS}, in which both our Proposition~\ref{rr-hDEO} and \ref{formula-hDEO} have been derived in a quite different setting. \end{remark} \subsection{Franklin's involution} Readers who are unfamiliar with Franklin's ingenious proof \cite{Fran} of Euler's pentagonal number theorem, are referred to \cite[Section 3.5]{AE} for an enlightening exposition. Now if one takes another look at Theorem~\ref{ana-pent} with Franklin's involution in mind, it should not take long before she/he realizes that, Franklin's involution between distinct partitions with even number of parts and those with odd number of parts not only preserve the size of the two partitions that get paired up, but also keep their perimeter unchanged! According to Andrews \cite{Andr3}, this crucial observation was first made by Subbarao \cite{Sub}, then Andrews \cite{Andr2} extended the idea considerably to give a combinatorial proof of the following Rogers--Fine identity: \begin{align}\label{R-F} \sum_{n=0}^{\infty}\dfrac{(\alpha)_n}{(\beta)_n}\tau^n &=\sum_{n=0}^{\infty}\dfrac{(\alpha)_n(\alpha\tau q/\beta)_n\beta^n\tau^nq^{n^2-n}(1-\alpha\tau q^{2n})}{(\beta)_n(\tau)_{n+1}}. \end{align} Here and in the sequel, we employ the customary notation $(a;q)_0:=1, \; (a)_n=(a;q)_n:=\prod_{k=0}^{n-1}(1-aq^k), \; \forall n\geq 1,\; (a)_{\infty}=(a;q)_{\infty}:=\prod_{k=0}^{\infty}(1-aq^k), \; |q|<1$. One special case of \eqref{R-F} that concerns us here is the following identity we take from Andrews' paper \cite[(3.2)$\sim$(3.4)]{Andr2}: \begin{align}\label{id: Andrews} \sum\limits_{n=0}^{\infty}\dfrac{(-1)^ny^{2n}q^{n(n+1)/2}}{(yq)_n}&= 1+\sum\limits_{n=1}^{\infty}\sum\limits_{r=1}^{\infty}(Q_{e}(r,n)-Q_{o}(r,n))y^rq^n\\ &=1+\sum\limits_{n=1}^{\infty}(-1)^n(q^{n(3n-1)/2}y^{3n-1}+q^{n(3n+1)/2}y^{3n}),\nonumber \end{align} where $Q_e(r,n)$ (resp. $Q_o(r,n)$) denotes the number of distinct partitions of $n$, say $\pi$, into an even (resp. odd) number of parts such that $\pi_1+\ell(\pi)=r$. Note that now both our Theorem~\ref{ana-pent} and Euler's pentagonal number theorem are special cases of \eqref{id: Andrews}, modulo a nuance due to \eqref{Gamma}. Indeed, given a partition $\pi$, Subbarao's original observation was on $\pi_1+\ell(\pi)$, while our perimeter $\Gamma(\pi)$ is always one less in value, and we discard entirely the empty partition that corresponds to the isolated summand $1$ in \eqref{id: Andrews}. This new connection raises a natural question, is there a generalization of \eqref{id: Andrews} in the spirit of Sylvester's generalization \cite[(3.3)]{Andr3} of Euler's pentagonal number theorem? We give an affirmative answer with the following identity: \begin{align}\label{Ramaphi} \sum\limits_{n=1}^{\infty}\dfrac{x^ny^nq^{n(n+1)/2}}{(xq)_n}&= \sum\limits_{\pi\in\mathcal{D}}x^{\pi_1}y^{\ell(\pi)}q^{|\pi|}= \sum\limits_{n=1}^{\infty}\dfrac{(-yq)_{n-1}x^{2n-1}y^nq^{n(3n-1)/2}(1+xyq^{2n})}{(xq)_{n}}. \end{align} Note that upon taking $x=y, y=-y$ in \eqref{Ramaphi}, we get back to \eqref{id: Andrews}. \begin{remark} Three remarks on \eqref{Ramaphi} are in order. \begin{itemize} \item The first three entries from page 41 of Ramanujan's lost notebook \cite{Rama} deal with the function \begin{align*} \phi(a)&:=\sum_{n=0}^{\infty}\dfrac{a^nq^{n(n+1)/2}}{(bq)_n}. \end{align*} It is a simple matter of variable change to see the equivalence between $\phi(a)$ and the left extreme of \eqref{Ramaphi}. Consequently, \eqref{Ramaphi} is equivalent to Entry 2.6 in \cite{Rama}. \item If we take $\beta=xq, \tau=-xyq/\alpha$, and let $\alpha\rightarrow \infty$ in \eqref{R-F}, we immediately recover the left extreme of \eqref{Ramaphi}, but the right hand side does not ``look right''. Actually this specialization gives rise to Entry 2.5 in \cite{Rama}. We recommend \cite{BY} for combinatorial proofs of all three aforementioned entries, and of many more related identities. \item If we take $x=xq,y=yq,q=1$ in \eqref{Ramaphi} and divide both sides by $q$ to account for the ``$-1$'' in \eqref{Gamma}, we see that both extremes reduce to the same rational $\dfrac{xyq}{1-(xq+xyq^2)}$ and we recover \eqref{gfHD}. \end{itemize} \end{remark} \section{Congruence properties for $h_{\mathcal{D}}(n)$ and final remarks}\label{sec:cong} The last statement in Theorem~\ref{ana-od} links $h_{\mathcal{D}}(n)=h_{\mathcal{O}}(n)$ to the $n$-th Fibonacci number $F_n$, which we haven't explored so far. Let us first recall two properties of the Fibonacci numbers, and then use them to give congruence properties for $h_{\mathcal{D}}(n)$. \begin{proposition}\label{Fm+n} Let $m\geq 0$,$n\geq 1$ be integers, then \begin{align}\label{id: Fm+n} F_{m+n}=F_{m+1}F_{n}+F_{m}F_{n-1}. \end{align} \end{proposition} \begin{proposition}\label{Fm|Fn} Let $m$,$n$ be positive integers, if $m|n$, then $F_{m}|F_{n}$. \end{proposition} \begin{remark} Proposition~\ref{Fm+n} can be easily proved by induction, then Proposition~\ref{Fm|Fn} follows from \eqref{id: Fm+n} combined with induction. It is worth recommending the award-winning book \cite{BQ}, where Benjamin and Quinn give an interesting tiling proof of Proposition~\ref{Fm+n}, among their masterful treatment of tons of other identities involving the Fibonacci numbers. \end{remark} Now the following congruence properties for $h_{\mathcal{D}}(n)$ should not come as surprise. \begin{align} h_{\mathcal{D}}(3n) &\equiv0\pmod{2}\label{3n}\\ h_{\mathcal{D}}(4n) &\equiv0\pmod{3}\\ h_{\mathcal{D}}(5n) &\equiv0\pmod{5}\\ h_{\mathcal{D}}(6n) &\equiv0 \pmod{8}\label{6n}\\ h_{\mathcal{D}}(6n+3) &\equiv2 \pmod{16}\label{6n+3}\\ h_{\mathcal{D, O}}(6n) &=h_{\mathcal{D, E}}(6n)\equiv0 \pmod{4}\label{DOE6n}\\ h_{\mathcal{D, O}}(6n+3) &=h_{\mathcal{D, E}}(6n+3)\equiv1 \pmod{8}\label{DOE6n+3} \end{align} Note that \eqref{3n} through \eqref{6n} require only Theorem~\ref{ana-od}, Proposition~\ref{Fm|Fn} and initial values, while \eqref{6n+3} needs more patience to wait until a complete period for $\{F_n \pmod{16}\}_{n\geq 1}$ emerges. Finally \eqref{DOE6n} and \eqref{DOE6n+3} are direct results of \eqref{6n} and \eqref{6n+3}, upon applying Theorem~\ref{ana-pent}. These results are by no means the complete list, but they raise the natural question of seeking partition statistic analogous to the rank (or crank) that would provide a combinatorial interpretation of the above congruences. Lastly, it would certainly be interesting to also try and generalize both \eqref{gfH} and \eqref{gfHO} to the extent of \eqref{Ramaphi}. \section*{Acknowledgement} The authors are grateful to Jiang Zeng for bringing Straub's paper \cite{Str} to their attention and for some initial discussions that motivated this work. Thanks also go to Zichen Yang for suggesting a better form of Theorem~\ref{ref1}. Both authors were supported by the Fundamental Research Funds for the Central Universities (No.~CQDXWL-2014-Z004) and the National Science Foundation of China (No.~115010\\61).
{ "timestamp": "2016-04-15T02:05:06", "yymm": "1604", "arxiv_id": "1604.04028", "language": "en", "url": "https://arxiv.org/abs/1604.04028", "abstract": "Motivated by a recent paper of Straub, we study the distribution of integer partitions according to the length of their largest hook, instead of the usual statistic, namely the size of the partitions. We refine Straub's analogue of Euler's Odd-Distinct partition theorem, derive a generalization in the spirit of Alder's conjecture, as well as a curious analogue of the first Rogers-Ramanujan identity. Moreover, we obtain a partition theorem that is the counterpart of Euler's pentagonal number theory in this setting, and connect it with the Rogers-Fine identity. We concludes with some congruence properties.", "subjects": "Combinatorics (math.CO)", "title": "Partitions with fixed largest hook length", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9833429609670702, "lm_q2_score": 0.8311430562234877, "lm_q1q2_score": 0.8172986738940246 }
https://arxiv.org/abs/2004.01677
Generalization of Kimberling's concept of triangle center for other polygons
In this article we introduce a general definition of the concept of center of an $n$-gon, for $n\geq 3$, generalizing the idea of C. Kimberling for triangle. We define centers associated to functions instead of to geometrical properties. We discuss the definition of those functions depending on both, the vertices of the polygons or the lengths of sides and diagonals. We explore the problem of characterization of regular polygons in terms of these $n$-gon center functions and we study the relation between our general definition of center of a polygon and other approaches arising from Applied Mathematics.
\section{Introduction} C. Kimberling \cite{K.O, K.FE} in the second half of the 20th century decided to give an unified definition for triangle center, including the classical centers (incenter, barycenter, circumcenter and orthocenter) and much others (Steiner Point, Fermat point,\ldots). Moreover, he created an encyclopedia \cite{K.E} trying to contain all known triangle centers. His new idea was to consider triangle centers as functions of the set of lengths, instead of loci. \medskip Following this spirit, we provide a definition of center of a polygon as a function ($n$-gon center functions). We also provide a geometric interpretation of this center functions as points in the plane. Although we still can find some works in the literature studying ``centers'' of polygons (see for example \cite{AS.Q, K.CCP, MM.E}), as far as we know there is no general definition for this concept. \medskip The main obstacle is that $n$-gons are not determined by their sidelegths for $n\geq 4$, so we have defined \emph{centers} as functions of the vertices (Definition \ref{def.main}). To connect with the definition given by Kimberling for triangles, we also provide an alternative and equivalent definition for the concept of \emph{center} involving, not only sidelengths, but also the lengths of the diagonals, which seems also to be fruitful to encompass some of the well known examples of polygon centers (Definition \ref{def.main2}). \medskip In \cite{AS.Q} the authors already studied the problem of exploring \emph{``the degree of regularity implied by the coincidence of two or more''} centers for quadrilaterals. In that article only the center of mass of the four vertices, the center of mass of the four sides, the center of mass of the whole figure considered as a lamina of uniform density and the Fermat–Torricelli center are considered. This problem is related to the fact that for squares those four centers coincide. In our new general setting all the centers (in their geometric interpretation as points in the plane) also coincide for regular polygons (Proposition \ref{prop.reg}). In relation to this, we study the problem of \emph{characterization} of equiangular, equilateral and regular polygons by means of one or more centers (Theorem \ref{th.equiangular}, Theorem \ref{th.equilateral} and Corollary \ref{coro.reg}). \medskip We also explore other possibilities for defining \emph{center of a polygon}, that would include some interesting examples (Definition \ref{defi.implicit} and Definition \ref{defi.optimization}) arising from other areas of Mathematics. We also briefly study their relation to our concept of center of a polygon. \medskip This article is structured as follows. In Section \ref{section.tc} we review the main aspects about Kimberling's definition of triangle center. Our definition of $n$-gon center function and its geometric interpretation appear in Section \ref{section.defi}, together with a discussion of its suitability: we justify that it satisfies some properties that we would expect for a point to be called \emph{center}. In Section \ref{section.equiv} we provide an equivalent defintion of center function in tems of sidelegths and lengths of diagonals, as announce before. Section \ref{section.exam} includes some examples of classic polygon centers that are subsumed in our new definition of \emph{center}. The problem of characterization of equiangular, equilateral and regular polygons by means of one or more centers is studied in Section \ref{section.characterization}. The other possibilities for a definition of \emph{center} are discussed in Section \ref{section.othercenter}. Finally in Section \ref{section.final} we briefly comment some topics related to this concept of polygon center and propose some future worklines (in relation to equivariant maps, computational geometry and others). We also include some Open Questions to close up this work. \section{Basics about triangle centers} \label{section.tc} As we explained in the introduction, C. Kimberling in his works \cite{K.O, K.FE, K.E} decided to define \emph{triangle centers} as functions (the so-called \emph{triangle center functions}) instead of loci in the plane. These functions also allow a geometric interpretation as points in the plane, via trilinear coordinates, as it is described below. \medskip Unlike what happens for $n$-gons for $n\geq 4$, triangles are determined by their sidelengths: \begin{rem}\label{rk:identification} Every triangle $T$ can be identified with the tuple of its three sidelengths $(a,b,c)$ (placed in clockwise order), up to congruences. \end{rem} \noindent According to this identification, we denote the set of all triangles as $$\mathcal T = \{(a,b,c) \in \mathbb{R}^3_+ \, :\, a +b > c,\ a+c > b, \ b+c >a\} . $$ \medskip Before proceeding with the definition, let us say that we denote by $[a:b:c]$ the points in the real projective plane $P_2\mathbb R$ with the usual convention $[a:b:c]=[a':b':c']\Leftrightarrow \frac{a}{a'}=\frac{b}{b'}=\frac{c}{c'}$. Then we have: \begin{defi}[Kimberling's definition of triangle center \cite{K.O, K.FE, K.E}] \label{defi.TC} A real-valued function $f$ of three real variables $a,b,c$ is a \emph{triangle center function} if it satisfies the following properties: \begin{enumerate} \item[(i)] Homogeneity: there exists some constant $n\in\mathbb N$ such that for all $t\in\mathbb R_{\geq 0}$ we have $f(ta,tb,tc)=t^nf(a,b,c)$. \item[(ii)] Bisymmetry in the second and third variables: for all $(a,b,c)\in\mathcal T$, we have $f(a,b,c)=f(a,c,b)$. \end{enumerate} \noindent Define also the \emph{coordinate map} $\varphi:\mathcal T\to P_2\mathbb R$ given by: $$\varphi_f(a,b,c)=[f(a,b,c):f(b,c,a):f(c,a,b)] .$$ \end{defi} $\varphi_f(a,b,c)$ can be also interpreted as trilinear coordinates (the first one corresponding to the side of length $a$, the second one to the side of lenght $b$ and the last one to the one of length $c$). Thanks to these two interpretations, we can think of a center as, actually, a point in the plane. \begin{figure}[h] \centering \includegraphics[width=50mm]{circuncentrotrilinear} \caption{There is a unique point $P$ in the plane such that the relative distances from $P$ to each one of the sides are $[t_1:t_2:t_3]$. $[t_1:t_2:t_3]$ are then said to be the trilinear coordinates of $P$.} \end{figure} \begin{exam}[circumcenter, see \cite{K.E}] The triangle center function corresponding to the circumcenter is $f(a,b,c)=a(b^2+c^2-a^2)$, since the trilinear coordinates corresponding to the circumcenter are: $$[t_1:t_2:t_3]=[a(b^2+c^2-a^2):b(c^2+a^2-b^2):c(a^2+b^2-c^2)]. $$ \end{exam} Kimberling's definition of \emph{triangle center function} ensures that: \begin{property}[the coordinate map is well defined] \label{property.wd} The trilinear coordinates of the center are associated to the triangle independently of the labelling of the sides (but obviously re-ordered). \end{property} \begin{property}[the definition is coherent with respect to similarities] \label{property.cs} Let $f$ be a triangle center function. Let $T_1=(a,b,c)$, $T_2=(a',b',c')$ be two similar triangles such that $S$ is the similarity $T_1=S(T_2)$. Let $P_1,P_2$ be the points with trilinear coordinates given by $\varphi_f(a,b,c)$ and $\varphi_f(a',b',c')$, with respect to each of the triangles. Then $P_1=S(P_2)$. \end{property} In this setting, we see that if we have a triangle $(a_1,a_2,a_3)$ and a permutation $\sigma$ of the set $\{1,2,3\}$, then if: $$\varphi_f(a_1,a_2,a_3)=[t_1:t_2:t_3]=[f(a_1,a_2,a_3):f(a_2,a_3,a_1): f(a_3,a_1,a_2)], $$ \noindent we have that: $$\varphi_f(a_{\sigma(1)},a_{\sigma( 2)},a_{\sigma(3)})=[t_{\sigma(1)}:t_{\sigma(2)}:t_{\sigma(3)}] .$$ There is a correspondence between trilinear coordinates and barycentric coordinates. We are interested in this second setting. Kimberling decided to use the first option, although he also claimed the existence of this correspondence: \begin{rem} If $[t_1:t_2:t_3]$ are the trilinear coordinates of a point $P$ (with respect to the sides of a triangle), then $[at_1:bt_2:ct_3]$ are the barycentric coordinates of this point (with respect to the vertices of this triangle). \end{rem} \section{Polygon centers} \label{section.defi} The concept of triangle center can be generalized to polygons. First we will fix the notation. For us, a polygon is a finite number of straight line segments connected in a closed chain. We say that the polygon is \emph{non-degenerated} if none of the vertex coincide and that it is \emph{simple} if those segments only intersect with the adjacent element of the chain in a vertex. We will also introduce the following convention, which we will use along the paper: \begin{rem} \label{rem.conv} Any $n$-gon can be identified with an $n$-uple $(V_1,\ldots,V_n)$ after a labelling of its vertices. This labelling is chosen in such a way that for $1\leq i\leq n$ the segments $V_i,V_{(i+1\mod n)}$ are edges of the polygon, and the rest of segments joining vertices are diagonals. \end{rem} Consider the dihedral group $D_n=\langle\rho,\sigma:\rho^n=id,\sigma^2=id,\sigma\rho\sigma=\rho^{-1}\rangle$ (it has $2n$ elements). It can be viewed as a subset of the permutation group of the set $\{1,\ldots,n\}$, determined by \begin{equation} \label{eq.sigma}\rho(i)=i+1\mod n \, , \qquad\qquad \sigma(i)=2+n-i \mod n. \end{equation} \noindent But it can also be viewed as a rellabeling (in the sense explained in Remark \ref{rem.conv}) in the set of all $n$-gons: $$\forall \alpha\in D_n,\qquad \alpha(V_1,\ldots,V_n)=(V_{\alpha(1)},\ldots,V_{\alpha(n)}) .$$ As in the case of triangles, we want to define the \emph{$n$-gon center function} as a function of the vertices $f(V_1,\ldots,V_n)$ (then defined in $(\mathbb R^2)^n$) and the \emph{coordinate map}. Trilinear coordinates are not a good option to provide a geometric interpretation since they do not extend in a natural way from triangles to $n$-gons for $n\geq 4$. So, we will use barycentric coordinates instead. \medskip For a fixed $n$, we denote the set of all $n$-gons by $\mathcal{P}_n\approx (\mathbb R^2)^n$. See that: \begin{rem}[other domains] \label{rem.domain} Sometimes, we may restrict ourselves to convex $n$-gons (whose vertices satisfy (1) $V_i\neq V_j$ if $i\neq j$ and (2) for any $i$, all vertices (except $V_i,V_{i+1\mod n}$) lie on the same side of the line defined by $V_i, V_{i+1\mod n}$) or to non-degenerated $n$-gons. \end{rem} Now we are ready for generalizing the definition of \emph{triangle center function} by C. Kimberling: \begin{defi}[main definition of $n$-gon center function] \label{def.main} We say that a real-valued function \break $f(V_1,\ldots,V_n)$ is a \emph{$n$-gon center function} if it satisfies the following properties: \begin{enumerate} \item[(1)] Preservation with respect to relabellings: for the symmetry $\sigma \in D_n$: $$f(V_1,V_2,\ldots,V_n)=f(V_1,V_{\sigma(2)},\ldots,V_{\sigma(n)}) .$$ \item[(2)] Homogeneity: there exists some $k\in\mathbb N$ such that, for all $t\in\mathbb R_{\geq 0}$ we have that $f(tV_1,\ldots,tV_n)=t^kf(V_1,\ldots,V_n)$. \item[(3)] Preservation with respect to motions: for every rigid motion $T$ in the plane, $f(T(V_1),\ldots,T(V_n))=f(V_1,\ldots,V_n)$. \end{enumerate} \noindent Define also the \emph{coordinate map} $\varphi : \mathcal{P}_n \to P_{n-1} \mathbb R$ given by: $$\varphi_f(V_1,\ldots,V_n)=[f(V_1,\ldots,V_n): f(V_2,\ldots,V_n,V_1):\ldots: f(V_n,V_1,\ldots,V_{n-1})] .$$ \noindent Note that it is not possible to define the coordinate map for $n$-gons such that $$f(V_1,\ldots,V_n)=f(V_2,\ldots,V_n,V_1)=\ldots=f(V_n,V_1,\ldots,V_{n-1})=0.$$ \end{defi} Coordinates $\varphi_f(V_1,\ldots,V_n)$ (when defined) are interpreted as barycentric coordinates with respect to the vertices. Then, the \emph{geometric interpretation} of the center of a given $n$-gon $(V_1,\ldots,V_n)$ associated to the $n$-gon center function $f$ is the point: \begin{equation} \label{eq.gi} P=\widetilde f(V_1,\ldots,V_n)V_1+\ldots+\widetilde f(V_n,\ldots,V_{n-1})V_n \end{equation} \noindent where $\widetilde f(V_1,\ldots,V_n),\ldots ,\widetilde f(V_n,\ldots,V_1)$ are the normalized coordinates \begin{equation} \label{eq.norm}\begin{cases} \varphi_f(V_1,\ldots,V_n)=[\widetilde f(V_1,\ldots,V_n):\ldots: \widetilde f(V_n,V_1,\ldots,V_{n-1})],\\ \widetilde f(V_1,\ldots,V_n)+\ldots+ \widetilde f(V_n,V_1,\ldots,V_{n-1})=1 \end{cases}\end{equation} We want the coordinate map to satisfy an analogue of Properties \ref{property.wd} and \ref{property.cs}. The next result ensures that, and may help us to clarify the notation and ideas in this paper. \begin{theo} Definition \ref{def.main} and the geometric interpretation described in \eqref{eq.gi} provide an analogue to Properties \ref{property.wd} and \ref{property.cs} for $n$-gon center functions, i.e., \begin{itemize} \item[P1] \emph{(the coordinate map is well defined)} The coordinates given by the coordinate map $\varphi_f$ are associated to each polygon independently of the labelling (but obviously re-ordered). \item[P2] \emph{(the definition is coherent with respect to similarities)} Let $f$ be a $n$-gon center function. Let $N_1=(V_1,\ldots,V_n)$ and $N_2=(V_1',\ldots,V_n')$ be two similar $n$-gons such that $S$ is the similarity $N_1=S(N_2)$. Let $P_1,P_2$ be the points with barycentric coordinates given by $\varphi_f(V_1,\ldots,V_n)$ and $\varphi_f(V_1',\ldots,V_n')$, with respect to each of the $n$-gons. Then $P_1=S(P_2)$. \end{itemize} \end{theo} \begin{proo} To prove P1, see that for every $\alpha\in D_n$ $$\varphi_f(\alpha(V_1,\ldots,V_n))=\alpha(\varphi_f(V_1,\ldots,V_n)), $$ \noindent where $\alpha([t_1:\ldots:t_n])=[t_{\alpha(1)}:\ldots:t_{\alpha(n)}]$. \medskip To prove P2 see that any similarity $S$ can be obtained as a composition $T\circ H$ of a rigid motion $T$ and an homotethy $H=\lambda\cdot id$ fixing the origin. Hence, \begin{align*} \varphi_f(S(V_1),\ldots,S(V_n)) & =\lambda^k \varphi_f(T(V_1),\ldots,T(V_n)) =\varphi_f(T(V_1),\ldots,T(V_n))= \\ &=\varphi_f(V_1,\ldots,V_n). \end{align*} \hspace{15cm}\end{proo} We conclude this section with the following result, which states an important property. Recall that we say that an $n$-gon is \emph{regular} if it is equiangular and equilateral. Regular $n$-gons can be either convex or star. \begin{prop}\label{prop.reg} For any center function $f$, a regular $n$-gon (convex or star) $(V_1,\ldots,V_n)$ satisfies $$f(V_1,\ldots,V_n)=f(V_2,\ldots, V_n,V_1)=\ldots=f(V_n,V_1,\ldots,V_{n-1}), $$ \noindent so, if defined, $\varphi_f(V_1,\ldots,V_n)=[1:\ldots:1]$. \end{prop} \begin{proo} Let $(V_1, \ldots, V_n)$ be a regular $n$-gon and $f$ a center function for this $n$-gon. Then, the $n$-gon $(V_i, V_{i+1}, \ldots , V_n, V_1, \ldots , V_{i-1})$ is also regular and corresponds to a rotation $T_i$ of $\frac{2\pi (i-1)}{n}$rad of $(V_1, \ldots, V_n)$, i.e., $$T_i(V_1, \ldots, V_n) = (T_i(V_1), \ldots, T_i(V_n)) = (V_i, V_{i+1}, \ldots , V_n, V_1, \ldots , V_{i-1}). $$ \noindent Since a rotation is a rigid motion in the plane, by property (3) of Definition \ref{def.main} we have that \break $f(T_i(V_1), \ldots, T_i(V_n))= f(V_1, \ldots, V_n)$. Thus, if defined, \begin{align*} \varphi_f & (V_1,\ldots,V_n) =[f(V_1,\ldots,V_n): f(V_2,\ldots,V_n,V_1):\ldots: f(V_n,V_1,\ldots,V_{n-1})] \\ & =[f(V_1,\ldots,V_n): f(T_2(V_1), \ldots, T_2(V_n)):\ldots: f(T_n(V_1), \ldots, T_n(V_n))] \\ &= [f(V_1,\ldots,V_n): f(V_1,\ldots,V_n):\ldots: f(V_1,\ldots,V_n)] = [1:1: \ldots :1] . \end{align*} \hspace{15cm} \end{proo} \section{An equivalent definition of polygon center function in terms of lengths} \label{section.equiv} The definition of center above is, in some sense, not satisfactory. The first reason is that it may be not inmediate to verify condition (3). And the second one is that, traditionally, some of the most useful centers are described in terms of the sidelengths, not in terms of the coordinates of the vertices. \medskip We need again to stablish some conventions. Let $(V_1,\ldots,V_n)$ be an $n$-gon. We will denote by $d_{ij}$ the length of the segment with endpoints $V_i,V_j$. It is obvious that $d_{ij} = d_{ji}$. If $j=i+1\mod n$, then $d_{ij}$ is a sidelength. We will write $e_{ij}$ instead of $d_{ij}$ when we want to emphasize that we are referring to sidelengths. \medskip An $n$-gon is not completely determined by its sidelengths, some of the lenths of the diagonals are required to determine it up to congruence. The set of all the sidelengths and of the lengths of the diagonals of an $n$-gon must satisfy some compatibility conditions. For example, consider a quadrilateral with sidelengths $e_{12},e_{23},e_{34},e_{41}$ and diagonals $d_{13},d_{24}$. According to the Cayley-Menger determinant formula for the volume of a 3-dimensional tetrahedron (see \cite{S.NDIM}) we have that: $$\left|\begin{array}{c c c c c}0 & 1 & 1 & 1 & 1 \\ 1 & 0 & e_{12}^2 & d_{13}^2 & e_{14}^2 \\ 1 & e_{21}^2 & 0 & e_{23}^2 & d_{24}^2 \\ 1 & d_{31}^2 & e_{32}^2 & 0 & e_{34}^2 \\ 1 & e_{41}^2 & d_{42}^2 & e_{43}^2 & 0 \end{array}\right|=0 .$$ So, sometimes we will identify an $n$-gon with the $(n+(n-3))$-uple $$(e_{12},\ldots,e_{n1},d_{13},\ldots, d_{1,n-1})$$ \noindent and sometimes, if it is more simple for the corresponding formulas, with the $n\cdot (n-1)$-uple $(d_{ij})$ with $i\neq j$, taking in mind that some of the entries $d_{ij}$ are redundant. \medskip In this setting we can define the $n$-gon center function as a real-valued function $g$ depending on the sidelengths and the lengths of the diagonals (defined then in $\mathbb{R}^{(n(n-1))}$), instead of the vertices, as follows. \begin{defi}[definition of $n$-gon center function in terms of lengths] \label{def.main2} We say that a real-\break valued function $g(d_{ij})$, $i,j= 1, \ldots,n$, is a \emph{$n$-gon center function} if it satisfies the following properties: \begin{enumerate} \item[(1')] Preservation with respect to relabellings: for the symmetry $\sigma \in D_n$: $$g(d_{ij})=g(d_{\sigma(i),\sigma(j)}). $$ \item[(2')] Homogeneity: there exists some $k\in\mathbb N$ such that, for all $t\in\mathbb R_{\geq 0}$, we have that $g(t\cdot d_{ij})=t^k\cdot g(d_{ij})$. \end{enumerate} \noindent Define also the \emph{coordinate map} $\varphi : \mathcal{P}_n \to P_{n-1} \mathbb R$ given by: $$\varphi_g(d_{ij})=[g(d_{ij}):g(d_{\rho(i),\rho(j)}):\ldots:g(d_{\rho^{n-1}(i),\rho^{n-1}(j)})] ,$$ for $\rho\in D_n$. \noindent Note that it is not possible to define the coordinate map for $n$-gons such that: $$g(d_{ij})=g(d_{\rho(i),\rho(j)}))=\ldots=g(d_{\rho^{n-1}(i),\rho^{n-1}(j)})=0. $$ \end{defi} In this context, the geometric interpretation of the \emph{center} is again (in those cases where the coordinate map is defined): $$\widetilde g(d_{ij})V_1+\widetilde g(d_{\rho(i),\rho(j)}))V_2+\ldots+\widetilde g(d_{\rho^{n-1}(i),\rho^{n-1}(j)})V_n$$ \noindent where $\widetilde g(d_{ij}),\ldots,\widetilde g(d_{\rho^{n-1}(i),\rho^{n-1}(j)})$ are the normalized coordinates as done in \eqref{eq.norm}. \medskip The next result ensures that both definitions of center function are compatible. \begin{theo}[equivalence between definitions \ref{def.main} and \ref{def.main2}] \label{th.equiv} Given a $n$-gon center function \break $f(V_1,\ldots,V_n)$, it is possible to find an $n$-gon center function $g(d_{ij})$ such that the geometric interpretations (if they exist) of the centers corresponding to $f$ and $g$ coincide for every element in $dom(f)\subset \mathcal P_n$, and viceversa. \end{theo} \begin{proo} Supose that we have an $n$-gon with vertices $(V_1,\ldots,V_n)$. The vertices determine univocally the lengths $d_{ij}=\left|V_i-V_j\right|$. On the other hand, the lengths $d_{ij}$ determine modulo congruence the vertices $(V_1,\ldots,V_n)$. So we can consider the vertices $V_k$ as functions of the lengths $V_k(d_{ij})$, if we impose $V_1=(0,0)$, $V_2=(d_{12},0)$ and that the rest of the vertices are ordered clockwise. \medskip First, see that if $f(V_1,\ldots,V_n)$ is a center function in the sense of Definition \ref{def.main}, then it is easy to find an associated center function $g(d_{ij})$ in the sense of Definition \ref{def.main2} of the form: $$g(d_{ij})=f(V_1(d_{ij}),\ldots,V_n(d_{ij})) .$$ We just have to prove that if $f$ satisfies properties (1), (2) and (3), then $g$ satisfies properties (1') and (2'). For $\sigma\in D_n$ as defined in \eqref{eq.sigma} we have: \begin{align*} g(d_{\sigma(i),\sigma(j)})&=f(V_1(d_{\sigma(i),\sigma(j)}),V_{\sigma(2)}(d_{\sigma(i),\sigma(j)}),\ldots,V_{\sigma(n)}(d_{\sigma(i),\sigma(j)})) =\\ &=f(V_1(d_{ij}),\ldots,V_n(d_{ij})) = g(d_{ij}), \end{align*} \noindent by properties (1) and (3). So, property (1') holds. On the other hand: \begin{align*} g(t\cdot d_{ij})& =f(V_1(t\cdot d_{ij}),\ldots,V_n(t\cdot d_{ij}))=f(t\cdot V_1(d_{ij}),\ldots,t\cdot V_n(d_{ij}))= \\ & =t^n\cdot f(V_1(d_{ij}),\ldots,t_n(V_n(d_{ij})))=t^n\cdot g(d_{ij}) , \end{align*} and so property (2') also holds. \medskip Next, see that if $g(d_{ij})$ is a center function in the sense of Definition \ref{def.main2}, then it is easy to find an associated center function $f(V_1,\ldots,V_n)$ in the sense of Definition \ref{def.main} of the form: $$f(V_1,\ldots,V_n)=g(\left|V_i-V_j\right|) .$$ \noindent We just have to prove that if $g$ satisfies properties (1') and (2'), then this $f$ satisfies properties (1), (2), and (3). First see that, for $\sigma \in D_n$ as defined in \eqref{eq.sigma} we have that: $$f(V_1,V_{\sigma(2)},\ldots,V_{\sigma(n)})=g(\left|V_{\sigma(i)}-V_{\sigma(j)}\right|)=g(\left|V_i-V_j\right|)=f(V_1,\ldots,V_n) ,$$ by property (1'). Hence, property (1) holds. Now, see that: \begin{align*} f(t\cdot V_1,\ldots,t\cdot V_n)&=g(\left|t\cdot V_i-t\cdot V_j\right|)=g(t\cdot \left|V_i-V_j\right|)=t^n\cdot g(\left|V_i-V_j\right|)\\ &=t^n\cdot f(V_1,\ldots,V_n) ,\end{align*} by property (2'). So, property (2) also holds. Finally, the proof of property (3) is inmediate: all congruent $n$-gons have the same sidelengths and diagonals. \end{proo} \section{Some examples} \label{section.exam} In this section we present some of the more relevant centers for polygons. Most of them arise from important problems in Applied Mathematics, and can be naturally defined as an affine combination of the vertices. The coefficients of this combinatio are functions of either the vertices or the sidelengths and lengths of the diagonals. Some of those examples can be found in \cite{AS.Q} in the particular case $n=4$ (quadrilaterals). \begin{exam}[centroid, barycenter or center of mass of the vertices] The barycenter of a polygon with vertices $V_1,\ldots,V_n$, or the center of mass of the vertices (provided that all the vertices have the same weight) is the point: \begin{equation} \label{eq.bary} G_0(V_1,\ldots,V_n)=\frac{1}{n}V_1+\ldots+\frac{1}{n}V_n . \end{equation} \noindent So, the associated center function can be chosen to be $f_0(V_1,\ldots,V_n)=1$ and the coordinate map is $\varphi_{f_0}(V_1, \ldots, V_n) = [1:\ldots :1]$ (recall that the coefficients for the affine combination are the normalized ones). \end{exam} \begin{exam}[center of mass of the perimeter of a convex polygon] The center of mass of the perimeter of a convex polygon (provided that all the points in the perimeter have the same weight) with vertices $V_1,\ldots,V_n$ is the point (see \cite{K.CCP}): $$ G_1(V_1,\ldots,V_n) = \sum_{i=1}^n \dfrac{e_{i-1(\modulo n),i}+e_{i,i+1(\modulo n)} }{2 \sum_{j=1}^n e_{j,j+1 (\modulo n)} } \cdot V_i . $$ So, the associated center function can be chosen as $$g_1(d_{ij})= {e_{n1}+e_{12} }$$ \noindent and the coordinate map is $\varphi_{g_1}(V_1, \ldots, V_n) =[(e_{n1}+e_{12}):\ldots:(e_{n-1,n}+e_{n1})]$. \end{exam} \begin{exam}[centroid of the polygonal lamina] The centroid of a polygonal lamina with vertices $V_1,\ldots,V_n$ is the point (see \cite{B.CPL}): \begin{equation} \label{eq.exmalo} G_2(V_1,\ldots,V_n) = \sum_{i=1}^n\left(\frac{ V_{i-1 (\modulo n)}\wedge V_i+V_i\wedge V_{i+1 (\modulo n)}}{3\sum_{j=1}^n V_j\wedge V_{j+1 (\modulo n)}}\right) \cdot V_i ,\end{equation} \noindent where $V_m\wedge V_{k} = x_m y_k - x_k y_m$, for $V_m = (x_m, y_m)$ and $V_k = (x_k, y_k)$. This is not an affine combination since $$\sum_{i=1}^n\left(\frac{ V_{i-1 (\modulo n)}\wedge V_i+V_i\wedge V_{i+1 (\modulo n)}}{3\sum_{j=1}^n V_j\wedge V_{j+1 (\modulo n)}}\right)=\frac{2}{3}\neq 1,$$ \noindent so this is not the geometric interpretation of a center in our sense. \medskip To include this important center in the setting of our definition we are going to modify \eqref{eq.exmalo}. $G_2(V_1,\ldots,V_n)$, if the $n$-gon $(V_1,\ldots,V_n)$ is convex, can be computed via ``geometric decomposition'' as $$\frac{\sum_{i=1}^n (G_2(V_i,V_{i+1 (\modulo n)},B))\cdot AREA(V_i,V_{i+1 (\modulo n)},B)}{\text{AREA}(V_1,\ldots,V_n)}, $$ \noindent for $B=G_0(V_1,\ldots,V_n)$ (see \eqref{eq.bary}), which, according to the \emph{Shoelace Formula} (see \cite{P.SF}) and to the formula of the centroif of a triangular lamina (the classical centroid of the triangle), equals to $$\frac{\sum_{i=1}^n (\frac{1}{3}V_i+\frac{1}{3}V_{i+1 (\modulo n)}+\frac{1}{3}B)\cdot \frac{1}{2}\left|(B-V_i)\wedge (B-V_{i+1 (\modulo n)})\right|}{\frac{1}{2}\sum_{i=1}^n\left|(B-V_i)\wedge (B-V_{i+1 (\modulo n)})\right|} =$$ \begin{equation}\label{eq.exbueno}=\frac{\sum_{i=1}^n(C_1(i)+C_2(i)+C_3(i))}{3\sum_{i=1}^n \left|(B-V_i)\wedge(B-V_{i+1})\right|}\end{equation} \noindent where: \begin{align*} C_1(i)&= \left|(B-V_i)\wedge(B-V_{i+1(\modulo n)})\right|,\\ C_2(i)&=\left|(B-V_{i-1 (\modulo n)})\wedge (B-V_i)\right|,\\ C_3(i)&=\frac{1}{n}\sum_{j=1}^n\left|(B-V_j)\wedge(B-V_{j+1 (\modulo n)})\right|.\end{align*} \noindent Expression \eqref{eq.exbueno} does correspond to the geometric interpretation of a center with center function \begin{align*} f_2(V_1,\ldots,V_n)=&\left|(B-V_1)\wedge(B-V_{2})\right|+\left|(B-V_{n})\wedge (B-V_1)\right|+\\ &+\frac{1}{n}\sum_{j=1}^n\left|(B-V_j)\wedge(B-V_{j+1 (\modulo n)})\right|.\end{align*} \end{exam} \begin{exam}[medoid] \label{ex:medoid} The medoid of the set of vertices $V_1,\ldots,V_n$ is the point $G_3$ such that (see, for example, the recent work \cite{B.M}): $$G_3 ={\arg \min}_{V \in\{V_1,\ldots,V_n\}}\sum_{i=1}^n \left|V-V_i \right| .$$ \noindent The medoid is not well defined for any $n$-gon (this minimum may be reached by two or more of the vertices). The medoid can also be considered as a center in our sense. In this case the center function is: $$f_3(V_1,\ldots,V_n)=\begin{cases}1 & \text{if }V_1=\min_{V\in\{V_1,\ldots,V_n\}}\sum_{i=1}^n \left|V-V_i \right| , \\ 0 & \text{in other case.} \end{cases}$$ \end{exam} \section{Characterization of $n$-gons using centers (specially mention to quadrilaterals)} \label{section.characterization} The idea of characterizing regular polygons using $n$-gon center functions was one of the main reasons of our interest in this topic, in connection to other geometric problems. This study was already started in \cite{AS.Q} for quadrilaterals. We say that: \begin{defi} A set of center functions $f_1,\ldots,f_k$ with associated coordinate maps $\varphi_{f_1},\ldots,\varphi_{f_k}$ characterizes a family $\mathcal F$ of $n$-gons if $(V_1,\ldots,V_n)\in\mathcal F$ if and only if: $$\begin{cases} f_1(V_1,\ldots,V_n)=f_1(V_2,\ldots,V_n,V_{1})=\ldots=f_1(V_{n},V_1,\ldots,V_{n-1}) ,\\ \vdots \\ f_k(V_1,\ldots,V_n)=f_k(V_2,\ldots,V_n,V_{1})=\ldots=f_k(V_{n},V_1,\ldots,V_{n-1}) . \end{cases}$$ \end{defi} If a family $\mathcal F$ is characterized by a set of center functions then it must be closed under congruences and it must contain regular $n$-gons (convex and star, see Proposition \ref{prop.reg}). \medskip Regular triangles (for triangles equilaterality and equiangularity are equivalent properties) are characterized by just one center function. Take for example: $f(V_1,V_2,V_3)=\|V_2-V_3\|$. Equiangular quadrilaterals (rectangles and their non-simple version called \emph{crossed rectangles}), provided that they are non-degenerated, are also characterized by one center function: $$f (V_1,V_2,V_3, V_4) = \frac{\vec{V_1V_2}\cdot \vec{V_1V_4}}{\|\vec{V_1V_2}\|\cdot \|\vec{V_1V_4}\|} .$$ This is not so trivial: the cosine of two angles being equal does not imply the angles are equal but complementary. But in this case this is not a problem since the sum of the angles of a non-degenerated quadrilateral must be less or equal to $2\pi$rad. However, there is no center or set of centers characterizing either equilateral quadrilateral (rhombi) or regular quadrilaterals (squares). The following results formalize this idea: \begin{theo} \label{th.equiangular} Equiangular $n$-gons can be characterized by one center function, provided that they are non-degenerated and convex. \end{theo} \begin{proo} The $n$-gon center function that characterizes equiangular $n$-gons (provided that they are convex and so the angle between two adjacent sides is less than $\pi$ rad) is again \begin{equation} \label{eq.centercos} f_1 (V_1,\ldots, V_n) = \frac{\vec{V_1V_2}\cdot \vec{V_1V_n}}{\|\vec{V_1V_2}\|\cdot \|\vec{V_1V_n}\|}.\end{equation} \hspace{15cm}\end{proo} In Figure \ref{fig.mountain} we show a pentagon which, despite not being equiangular, is also included in a family of polygons characterized by the center \eqref{eq.centercos}. \begin{figure}[h] \centering \includegraphics[width=50mm]{mountain.jpg} \caption{This pentagon is obtained as the union of two equilateral triangles. So all the angles equal to $\pi/3$, except the one over $V_3$, which equals $2\pi-\pi/6$.} \label{fig.mountain} \end{figure} \begin{theo} \label{th.equilateral} For $n\geq 3$ being an odd number, equilateral $n$-gons can be characterized by one center function. For $n\geq 4$ being an even number, equilateral $n$-gons cannot be characterized using $n$-gon center functions. This is a consequence of the fact that any $n$-gon center function $f$ must satisfy: $$(V_1,\ldots,V_n)\in \mathcal S_n \Rightarrow f(V_1,\ldots,V_n)=\ldots=f(V_n,V_1,\ldots,V_{n-1}) ,$$ \noindent for $\mathcal S_n$ being the family of equiangular $n$-gons $(V_1,\ldots,V_n)$ such that $\|V_\rho(i)-V_\rho(j)\|=\|V_{\sigma(i)}-V_{\sigma(j)}\|$ (generalization of rectangles for $n\geq 6$). This family can be characterized by one $n$-gon center function. \end{theo} \begin{proo} The $n$-gon center function that characterizes equilateral $n$-gon for $n$ odd is $$f_2(V_1,\ldots,V_n)=\|V_{(n+1)/2}-V_{(n+3)/2}\| .$$ The fact that any $n$-gon center function must characterize $\mathcal S_n$ is inmediate from the ``preservation with respect to relabellings property''. The center function that characterizes $\mathcal S_n$ is: $$f_3(V_1,\ldots,V_n)=\|V_{n/2}-V_{(n/2)+2}\| . $$ \hspace{15cm} \end{proo} Theorems \ref{th.equiangular} and \ref{th.equilateral} together imply: \begin{coro}\label{coro.reg} For $n\geq 3$ odd, regular $n$-gons can be characterized by two center functions, provided that they are convex. For $n\geq 4$ even, they cannot. \end{coro} See that Theorems 3.1, 3.2, 3.3, 3.4, 3.5 in \cite{AS.Q} are compatible with the results proved here, although the authors are only interested in some particular quadrilateral centers. \section{Other possible definitions of center} \label{section.othercenter} Some of the centers arising from Applied Mathematics are defined by an implicit equation involving the vertices, or as solution of a problem of optimization (see Example \ref{ex:medoid}). Besides, the term ``center'' appears in a different setting for compact length spaces in \cite{MM.E}, from a totally different approach. This leads to the two following alternative definitions of center: \begin{defi}\label{defi.implicit} Let $F(P,V_1,\ldots,V_n)$ be a map satisfying the following properties: \begin{itemize} \item[(a)] For every $(V_1,\ldots,V_n)$ in $\mathcal P_n$ (additionally non-degeneration property can be required) \break $F(P,V_1,\ldots,V_n)=0$ defines univocally $P$. \item[(b)] Preservation with repect to labellings: if $F(P,V_1,\ldots,V_n)=0$, then for any element $\alpha\in D_n$ we have $$F(P,V_{\alpha(1)},\ldots,V_{\alpha(n)})=0 .$$ \item[(c)] Homogeneity: for every $\lambda\in\mathbb R_{\geq 0}$, $$F(P,V_1,\ldots,V_n)=0\Rightarrow F(\lambda P,\lambda V_1,\ldots,\lambda V_n)=0.$$ \item[(d)] Preservation by motions: for every rigid motion $T$ in the plane, $$F(P,V_1,\ldots,V_n)=0\Rightarrow F(T(P),T(V_1),\ldots,T(V_n))=0.$$ \end{itemize} \noindent We say that the point $P$ ensured by (a) is an \emph{implicit center} of $(V_1,\ldots,V_n)$. \end{defi} \begin{defi} \label{defi.optimization} Let $G(P,V_1,\ldots,V_n)$ be a real function defined in $\mathcal P_n$ (additionally non-degeration property can be required for the domain) such that: \begin{itemize} \item[(a')] For every $(V_1,\ldots,V_n)$ in $\mathcal P_n$ there exists a unique $$X(V_1,\ldots,V_n)={\arg\min}_{P\in\mathbb R^2}(G(P,V_1,\ldots,V_n)).$$ \item[(b')] Preservation with repect to labellings: for any element $\alpha\in D_n$ we have $${\arg\min}_{P\in\mathbb R^2}(G(P,V_1,\ldots,V_n))={\arg\min}_{P\in\mathbb R^2}(G(P,V_{\alpha(1)},\ldots,V_{\alpha(n)})).$$ \item[(c')] Homogeneity: for every $\lambda\in\mathbb R_{\geq 0}$, $$\lambda\cdot{\arg\min}_{P\in\mathbb R^2}(G(P,V_1,\ldots,V_n))= {\arg\min}_{P\in\mathbb R^2}(G(P,\lambda V_{1},\ldots,\lambda V_{n})).$$ \item[(d')] Preservation by motions: for every rigid motion $T$ in the plane, $$T\left({\arg\min}_{P\in\mathbb R^2}(G(P,V_1,\ldots,V_n))\right)= {\arg\min}_{P\in\mathbb R^2}(G(P,T(V_{1}),\ldots,T(V_{n}))).$$ \end{itemize} \noindent We will say that the point $X(V_1,\ldots,V_n)$ ensured by (a') is a \emph{minimal center} of $(V_1,\ldots,V_n)$. \end{defi} We have that: \begin{theo} Any implicit center in the sense of Definition \ref{defi.implicit} is a minimal center in the sense of Definition \ref{defi.optimization} and viceversa. Moreover, any center in the sense of Definition \ref{def.main} is an implicit center in the sense of Definition \ref{defi.implicit} (and therefore, a minimal center in the sense of Definition \ref{defi.optimization}). \end{theo} \begin{proo} To show that definitions \ref{defi.implicit} and \ref{defi.optimization} are equivalent take $$F(P,V_1,\ldots,V_n)=G(P,V_1,\ldots,V_n)-X(V_1,\ldots,V_n) $$ \noindent and, for $X$ the unique point satisfying $F(X,V_1,\ldots,V_n)=0$, $$G(P,V_1,\ldots,V_n)=\min_{P\in\mathbb R^2}(\text{dist}(P,X)) .$$ The proof of the second statement is inmediate taking $$F(P,V_1,\ldots,V_n)=(\widetilde f(V_1,\ldots,V_n)V_1+\ldots+\widetilde f(V_n,V_1,\ldots,V_{n-1})V_n)-P . $$ \hspace{15cm} \end{proo} Some examples of well-known points usually called ``centers'' that could be naturally included in this different definitions could be: \begin{exam}[geometric median of the vertices] \label{ex.gm} The geometric median of the set of vertices \break $V_1,\ldots,V_n$ of an $n$-gon is the point $X$ minimizing the sum of distances to the vertices. Thus, it could be naturally considered as a \emph{minimal center} defined by (see \cite{D.GM}): $$ X= {\arg \min}_{P \in \mathbb{R}^2} \sum_{j=1}^n \left|V_j-P\right| . $$ \noindent Provided that $X$ is distinct from any vertex, it can be also described as an \emph{implicit center} by the formula: $$\sum_{j=1}^n\frac{V_j-X}{\left|V_j-X\right|} =0 .$$ \noindent It is known that there is no explicit ``simple'' formula for $G$ or its coordinates (see \cite{B.NP}). \end{exam} \begin{exam}[Chebyshev center] The Chebyshev center of a bounded set $Q$ is the center $X$ of the minimal-radius ball enclosing the entire set $Q$ (see \cite{B.CC}). It is described as a minimal center by the formula: $$ X= {\arg \min}_{P \in \mathbb{R}^2} \left (\max_{V \in Q} \|V - P\|^2 \right ) . $$ \end{exam} \section{Final comments} \label{section.final} During the development of this article, some questions have arisen: \begin{open} Can regular $n$-gons, for $n$-odd, be characterized by only one center function? \end{open} \begin{open} What de we know about the Characterization Problem when we do not have the restriction of the polygons being convex? \end{open} \begin{open} Is any \emph{implicit center} in the sense of Definition \ref{defi.implicit} a \emph{center} in the sense of Definition \ref{def.main}? See that Example \ref{ex.gm} shows that the corresponding \emph{center function} may not be trivial at all to find. \end{open} $\mathcal P_n$ is naturally a $D_n$-space (a topological space endowed with a group of symmetries, see \cite{M.BU}). In this context, coordinate maps can be understood as $G$-maps. It could be interesting to explore this point of view. In particular, this may connect $n$-gon centers with interesting problems in Plane Geometry such as as the Square Peg Problem and its variants \cite{M.S}. \medskip Finally we would like to remark that the study of centers for $k$-dimensional polyhedra ($k\geq 3$, but specially $k=3$) would be of great interest in different areas (computational geometry and computer vision, for instance), and is a problem still to be explored. \section*{Acknowledgements} The second author is supported by a postdoctoral grant (PEJD-2018-POST/TIC-9490) from UNED, co-financed by the Regional Government of Madrid with funds from the Youth Employment Initiative (YEI) of the European Union.
{ "timestamp": "2020-04-14T02:25:26", "yymm": "2004", "arxiv_id": "2004.01677", "language": "en", "url": "https://arxiv.org/abs/2004.01677", "abstract": "In this article we introduce a general definition of the concept of center of an $n$-gon, for $n\\geq 3$, generalizing the idea of C. Kimberling for triangle. We define centers associated to functions instead of to geometrical properties. We discuss the definition of those functions depending on both, the vertices of the polygons or the lengths of sides and diagonals. We explore the problem of characterization of regular polygons in terms of these $n$-gon center functions and we study the relation between our general definition of center of a polygon and other approaches arising from Applied Mathematics.", "subjects": "Metric Geometry (math.MG)", "title": "Generalization of Kimberling's concept of triangle center for other polygons", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9859363741964572, "lm_q2_score": 0.8289388019824946, "lm_q1q2_score": 0.8172809168573757 }
https://arxiv.org/abs/1910.14454
Orthogonal Colourings of Cayley Graphs
Two colourings of a graph are orthogonal if they have the property that when two vertices are coloured with the same colour in one colouring, then those vertices receive distinct colours in the other colouring. In this paper, orthogonal colourings of Cayley graphs are discussed. Firstly, the orthogonal chromatic number of cycle graphs are completely determined. Secondly, the orthogonal chromatic number of certain circulant graphs is explored. Lastly, orthogonal colourings of product graphs and Hamming graphs are studied.
\section{Introduction} Two colourings of a graph are \textit{orthogonal} if they have the property that when two vertices are coloured with the same colour in one colouring, then those vertices must have distinct colours in the other colouring. A $k$-\textit{orthogonal colouring} is a collection of $k$ mutually orthogonal colourings. For simplicity, a $2$-orthogonal colouring is called an \textit{orthogonal colouring}. The \textit{$k$-orthogonal chromatic number} of a graph $G$, denoted $O\chi_k(G)$, is the minimum number of colours required for a $k$-orthogonal colouring. The $2$-orthogonal chromatic number is simply called the \textit{orthogonal chromatic number}, and is denoted $O\chi(G)$. Orthogonal colourings were first defined in 1985 by Archdeacon, Dinitz, and Harary, under the context of edge colourings \cite{archdeacon1985orthogonal}. Later in 1999, Caro and Yuster revisited orthogonal colourings, this time under the context of vertex colourings \cite{caro1999orthogonal}. Then in 2013, Ballif studied upper bounds on sets of orthogonal vertex colourings \cite{ballif2013upper}. Let $(G,\circ)$ denote a group $G$ with group operation $\circ$ and let $S$ be a generating set of $G$. Then, the associated \textit{Cayley graph}, denoted $\Gamma(G,S)$, has a vertex for each element of $G$ and there is a directed edge between two elements $u$ and $v$ if and only if $u\circ v^{-1}\in S$. It is assumed that $S$ is closed under inverses and that the additive identity is not in $S$. This is so that the Cayley graphs considered in this paper are simple, and undirected. To start, the orthogonal chromatic number of cycle graphs are determined. Then, orthogonal colourings of more general circulant graphs are studied. Lastly, orthogonal colourings of Cartesian product graphs and Hamming graphs are explored. \section{Orthogonal Colourings of Cycle Graphs} The Cayley graphs of cyclic groups are called circulant graphs. For example, consider $\mathbb{Z}_n$, the integers modulo $n$ with group operation addition modulo $n$. Then, the Cayley graph $\Gamma(\mathbb{Z}_n,\{1,-1\})\cong C_n$, the cycle graph on $n$ vertices. For a graph $G$ with $n$ vertices, a lower bound for $O\chi(G)$ is $\lceil\sqrt{n}\,\rceil$. In this section, it is shown that $O\chi(C_n)=\lceil\sqrt{n}\,\rceil$ if and only if $n>4$. The following lemma establishes that this is the correct orthogonal chromatic number for most cases. \begin{lem}\label{Lemma:Cycle Part 1} For $n>4$, if $\lceil\sqrt{n}\,\rceil\nmid (n-1)$, and $\lceil\sqrt{n}\,\rceil\nmid \Big( n-1+\Big\lfloor\frac{n-1}{\lceil\sqrt{n}\,\rceil}\Big\rfloor\Big)$, then $O\chi(C_n)=\lceil\sqrt{n}\,\rceil$. \end{lem} \begin{pf} Let $v_i\in \mathbb{Z}_n$ where $0\leq i <n$ and let $N=\lceil\sqrt{n}\,\rceil$. Set $c_1(v_i)=i(\textrm{mod}~N)$ and $c_2(v_i)=\big(i+\big\lfloor\frac{i}{N}\big\rfloor\big)(\textrm{mod}~N)$. For illustration, these two colourings are applied to $C_9$ in Figure \ref{Figure: Orthogonal colouring of C9}. Displayed next to each vertex $v_i$ is the colour pair $(c_1(v_i),c_2(v_i))$. \begin{figure}[h!] \centering \begin{tikzpicture}[scale=.6,line cap=round,line join=round,>=triangle 45,x=1.0cm,y=1.0cm] \draw [line width=2.pt] (0.52,2.82)-- (1.22,4.18); \draw [line width=2.pt] (1.22,4.18)-- (2.96,4.66); \draw [line width=2.pt] (2.96,4.66)-- (4.7,4.18); \draw [line width=2.pt] (4.7,4.18)-- (5.5,2.7); \draw [line width=2.pt] (5.5,2.7)-- (5.16,1.26); \draw [line width=2.pt] (5.16,1.26)-- (4.,0.); \draw [line width=2.pt] (4.,0.)-- (2.,0.); \draw [line width=2.pt] (2.,0.)-- (0.8,1.38); \draw [line width=2.pt] (0.8,1.38)-- (0.52,2.82); \draw (2.5,4.5) node[anchor=north west] {$v_0$}; \draw (3.9,4.15) node[anchor=north west] {$v_1$}; \draw (4.4,3.2) node[anchor=north west] {$v_2$}; \draw (4.2,2) node[anchor=north west] {$v_3$}; \draw (3.2,0.9) node[anchor=north west] {$v_4$}; \draw (1.7,0.9) node[anchor=north west] {$v_5$}; \draw (.7,2) node[anchor=north west] {$v_6$}; \draw (0.55,3.2) node[anchor=north west] {$v_7$}; \draw (1.1,4.2) node[anchor=north west] {$v_8$}; \draw (2,5.8) node[anchor=north west] {(0,0)}; \draw (4.1,5.35) node[anchor=north west] {(1,1)}; \draw (5.5,3.25) node[anchor=north west] {(2,2)}; \draw (5,1.4) node[anchor=north west] {(0,1)}; \draw (3.5,0) node[anchor=north west] {(1,2)}; \draw (.9,0) node[anchor=north west] {(2,0)}; \draw (-.8,1.4) node[anchor=north west] {(0,2)}; \draw (-1.3,3.35) node[anchor=north west] {(1,0)}; \draw (-.1,5.35) node[anchor=north west] {(2,1)}; \begin{scriptsize} \draw [fill=black] (2.,0.) circle (2.5pt); \draw [fill=black] (4.,0.) circle (2.5pt); \draw [fill=black] (2.96,4.66) circle (2.5pt); \draw [fill=black] (0.8,1.38) circle (2.5pt); \draw [fill=black] (5.16,1.26) circle (2.5pt); \draw [fill=black] (0.52,2.82) circle (2.5pt); \draw [fill=black] (5.5,2.7) circle (2.5pt); \draw [fill=black] (1.22,4.18) circle (2.5pt); \draw [fill=black] (4.7,4.18) circle (2.5pt); \end{scriptsize} \end{tikzpicture} \caption{Orthogonal Colouring of $C_9$} \label{Figure: Orthogonal colouring of C9} \end{figure} It is now shown that $c_1$ and $c_2$ are both proper colourings. For $0\leq i \leq n-2$, $c_1(v_i)=i(\textrm{mod}~N)\not\equiv (i+1)(\textrm{mod}~N)=c_1(v_{i+1})$. Then, since $N\nmid (n-1)$ by assumption, $c_1(v_{n-1})=(n-1)(\textrm{mod}~N)\neq 0 =c_1(v_0)$. Thus, $c_1$ is a proper colouring. To show that $c_2$ is a proper colouring, notice that $\big\lfloor\frac{i}{N}\big\rfloor\leq\big\lfloor\frac{i+1}{N}\big\rfloor\leq \big\lfloor\frac{i}{N}\big\rfloor+1 $. Therefore, for $0\leq i \leq n-2$, it follows that $1\leq c_2(v_{i+1})-c_2(v_i)\leq 2$. Thus, since $N>2$, $c_2(v_i)\neq c_2(v_{i+1})$. Now, since $N\nmid \big(n-1+\big\lfloor\frac{n-1}{N}\big\rfloor\big)$ by assumption, it follows that $c_2(v_{n-1})=\big(n-1+\big\lfloor\frac{n-1}{N}\big\rfloor\big)(\textrm{mod}~N)\neq 0 =c_2(v_0)$. Therefore, $c_2$ is a proper colouring. It remains to show that $c_1$ and $c_2$ are orthogonal colourings. Suppose otherwise, that is, $c_1(v_i)=c_1(v_j)$ and $c_2(v_i)=c_2(v_j)$ where $i\neq j$. Since $c_1(v_i)=c_1(v_j)$, this implies that $i=j+mN$ where $0<m<N$. Therefore: \begin{align*} c_2(v_i)&=\Big(j+\Big\lfloor\frac{j+mN}{N}\Big\rfloor\Big)(\textrm{mod}~N)\\ &=\Big(j+m+\Big\lfloor\frac{j}{N}\Big\rfloor\Big)(\textrm{mod}~N)\\ &=(c_2(v_j)+m) (\textrm{mod}~N). \end{align*} Since $c_2(v_i)=c_2(v_j)$, this gives that $m\equiv 0(\textrm{mod}~N)$, contradicting $0<m<N$. Therefore, $c_1$ and $c_2$ are orthogonal colourings of $C_n$. Since $c_1$ and $c_2$ both used $N$ colours, the fewest possible, $O\chi(C_n)=N$. \end{pf} Notice that the orthogonality property of $c_1$ and $c_2$ in Lemma \ref{Lemma:Cycle Part 1} did not depend on the assumed divisibility conditions. Therefore, the problem with using $c_1$ and $c_2$ in the cases where $N\mid (n-1)$ and $N\mid \big(n-1+\big\lfloor\frac{n-1}{N}\big\rfloor\big)$, is that there is some sort of colour conflict. The following lemma shows that this conflict can be resolved by assigning $v_{n-1}$ a different colour pair. \begin{lem}\label{Lemma: Cycle Part 2} For $n> 16$, if $\lceil\sqrt{n}\,\rceil\mid (n-1)$ or $\lceil\sqrt{n}\,\rceil\mid \Big(n-1+\Big\lfloor\frac{n-1}{\lceil\sqrt{n}\,\rceil}\Big\rfloor\Big)$, then $O\chi(C_n)=\lceil\sqrt{n}\,\rceil$. \end{lem} \begin{pf} Let $N=\lceil\sqrt{n}\,\rceil$, $c_1(v_i)=i(\textrm{mod}~N)$ and $c_2(v_i)=\big(i+\big\lfloor\frac{i}{N}\big\rfloor\big)(\textrm{mod}~N)$. There are four cases to consider: Case 1: $N\mid (n-1)$ and $N\nmid\big(n+\big\lfloor\frac{n}{N}\big\rfloor\big)$. Case 2: $N\nmid n$ and $N\mid\big(n-1+\big\lfloor\frac{n-1}{N}\big\rfloor\big)$. Case 3: $N\mid (n-1)$ and $N\mid\big(n+\big\lfloor\frac{n}{N}\big\rfloor\big)$. Case 4: $N\mid\big(n-1+\big\lfloor\frac{n-1}{N}\big\rfloor\big)$ and $N\mid n$. For Case 1 and Case 2, the orthogonal colouring $(\bar c_1, \bar c_2)$ is used, where $\bar c_1$ and $\bar c_2$ are defined as follows: \begin{align*} \bar c_1(v_i)&= \begin{cases} c_1(v_i) & 0\leq i\leq n-2 \\ n(\textrm{mod}~N) & i=n-1 \\ \end{cases}\quad\\ \bar c_2(v_i)&= \begin{cases} c_2(v_i) & 0\leq i\leq n-2 \\ \big(n+\big\lfloor\frac{n}{N}\big\rfloor\big)(\textrm{mod}~N) & i=n-1 \\ \end{cases} \end{align*} It is shown that $\bar c_1$ and $\bar c_2$ are proper colourings in both cases. For $0\leq i\leq n-2$, $\bar c_1(v_i)=c_1(v_i)$ and $\bar c_2(v_i)=c_2(v_i)$. Therefore, by the proof of Lemma \ref{Lemma:Cycle Part 1}, there are no colour conflicts between these vertices. Note that in both Case 1 and Case 2, $n\not\equiv 0(\textrm{mod}~N)$ and $\big(n+\big\lfloor\frac{n}{N}\big\rfloor\big)(\textrm{mod}~N)\not\equiv 0(\textrm{mod}~N)$. Therefore, $\bar c_1(v_{n-1})\neq 0=\bar c_1(v_0)$, and $\bar c_2(v_{n-1})\neq 0=\bar c_2(v_0)$. Notice that $\big\lfloor\frac{n-2}{N}\big\rfloor\leq \big\lfloor\frac{n}{N}\big\rfloor\leq \big\lfloor\frac{n-2}{N}\big\rfloor+1$. Therefore, $2\leq \bar c_2(v_{n-1})-\bar c_2(v_{n-2})\leq 3$. Since $N>4$ by assumption, this implies that $\bar c_2(v_{n-1})\neq \bar c_2(v_{n-2})$. Also, since $N>4$, $\bar c_1(v_{n-2})=(n-2)(\textrm{mod}~N)\neq n(\textrm{mod}~N)=\bar c_1(v_{n-1})$. Therefore, $\bar c_1$ and $\bar c_2$ are proper colourings of $C_n$. It remains to show that $\bar c_1$ and $\bar c_2$ are orthogonal colourings. For $0\leq i \leq n-2$, there are no orthogonal conflicts on the vertices $v_i$ by the proof of Lemma \ref{Lemma:Cycle Part 1}. In Case 1, since $n\equiv 1(\textrm{mod}~N)$, the colour pair $(1,(1+\lfloor\frac{n}{N}\rfloor)(\textrm{mod}~N))$ is assigned to $v_{n-1}$. Let $i\equiv 1 (\textrm{mod}~N)$ and $i<n$. Let $m_1$ and $m_2$ be integers so that $i=m_1N+1$ and $n=m_2N+1$. Since $i<n$, this gives that $m_1<m_2$. Therefore, $\bar c_2(v_i)=(1+\lfloor\frac{m_1N+1}{N}\rfloor)(\textrm{mod}~N))=(1+m_1)(\textrm{mod}~N)\neq (1+m_2)(\textrm{mod}~N)=\bar c_2(v_{n-1})$. In Case 2, since $N\mid\big(n-1+\big\lfloor\frac{n-1}{N}\big\rfloor\big)$, the colour pair $(n(\textrm{mod}~N),1)$ is assigned to $v_{n-1}$. A similar argument as in Case 1 shows that there are no orthogonal conflicts. Therefore, in both Case 1 and Case 2, $\bar c_1$ and $\bar c_2$ are proper orthogonal colourings. For Case 3 and Case 4, a different orthogonal colouring $(\hat c_1,\hat c_2)$ is used, where $\hat c_1$ and $\hat c_2$ are defined as follows: \begin{align*} \hat c_1(v_i)&= \begin{cases} c_1(v_i) & 0\leq i\leq n-2 \\ (n+1)(\textrm{mod}~N) & i=n-1 \\ \end{cases}\quad\\ \hat c_2(v_i)&= \begin{cases} c_2(v_i) & 0\leq i\leq n-2 \\ \big(n+1+\big\lfloor\frac{n+1}{N}\big\rfloor\big)(\textrm{mod}~N) & i=n-1 \\ \end{cases} \end{align*} It is shown that $\hat c_1$ and $\hat c_2$ are proper colourings in both cases. For $0\leq i \leq n-2$, $\hat c_1(v_i)=c_1(v_i)$ and $\hat c_2(v_i)=c_2(v_i)$. Therefore, by the proof of Lemma \ref{Lemma:Cycle Part 1}, there are no colour conflicts between these vertices. Note that in both Case 3 and Case 4, $(n+1)\not\equiv 0(\textrm{mod}~N)$ and $\big(n+1+\big\lfloor\frac{n+1}{N}\big\rfloor\big)\not\equiv 0(\textrm{mod}~N)$ . Thus, $\hat c_1(v_{n-1})\neq 0=\hat c_1(v_0)$, and $\hat c_2(v_{n-1})\neq 0 =\hat c_2(v_0)$. Notice that $\big\lfloor\frac{n-2}{N}\big\rfloor\leq \big\lfloor\frac{n+1}{N}\big\rfloor\leq \big\lfloor\frac{n-2}{N}\big\rfloor+1$. Therefore, $3\leq \hat c_2(v_{n-1})-\hat c_2(v_{n-2})\leq 4$. Since $N>4$ by assumption, this implies that $\hat c_2(v_{n-2})\neq \hat c_2(v_{n-1})$. Also, since $N>4$, $\hat c_1(v_{n-2})=(n-2)(\textrm{mod}~N)\not\equiv (n+1)(\textrm{mod}~N)=\hat c_1(v_{n-1})$. Therefore, $\hat c_1$ and $\hat c_2$ are proper colourings of $C_n$. It remains to show that $\hat c_1$ and $\hat c_2$ are orthogonal colourings. For $0\leq i \leq n-2$, there are no orthogonal conflicts on the vertices $v_i$ by the proof of Lemma \ref{Lemma:Cycle Part 1}. In Case 3, since $n\equiv 1(\textrm{mod}~N)$, the colour pair $(2,(2+\lfloor\frac{n+2}{N}\rfloor)(\textrm{mod}~N))$ is assigned to $v_{n-1}$. Let $i\equiv 2(\textrm{mod}~N)$ and $i<n$. Let $m_1$ and $m_2$ be integers so that $i=m_1N+2$ and $n=m_2N+1$. Since $i<n$, this gives that $m_1<m_2$. Therefore, $\hat c_2(v_i)=(2+\lfloor\frac{m_1N+1}{N}\rfloor)(\textrm{mod}~N)=(2+m_1)(\textrm{mod}~N)\neq (2+m_2)(\textrm{mod}~N)=\hat c_2(v_{n-1})$. In Case 4, since $N\mid\big(n-1+\big\lfloor\frac{n-1}{N}\big\rfloor\big)$, the colour pair $((n+2)(\textrm{mod}~N),2)$ is assigned to $v_{n-1}$. A similar argument as in Case 3 shows that there are no orthogonal conflicts. Therefore, in both Case 3 and Case 4, $\hat c_1$ and $\hat c_2$ are proper orthogonal colourings. \end{pf} For $n>4$, the remaining cases left to consider are $n=6,7,8,11,13,14$. These can be orthogonally coloured with $\lceil\sqrt{n}\,\rceil$ colours, shown in Figure \ref{Figure: Case 1} and Figure \ref{Figure: Case 2}. For $n=3,4$, $C_n$ needs $3$ colours. The following theorem summarizes these results. \begin{figure}[h!] \centering \begin{tikzpicture}[scale=.75,line cap=round,line join=round,>=triangle 45,x=1.0cm,y=1.0cm] \draw [line width=2.pt] (-1,3.)-- (1,3.94); \draw [line width=2.pt] (1,3.94)-- (3.,3.); \draw [line width=2.pt] (3.,3.)-- (3.,1.); \draw [line width=2.pt] (3.,1.)-- (1,0.); \draw [line width=2.pt] (1,0.)-- (-1,1.); \draw [line width=2.pt] (-1,1.)-- (-1,3.); \draw (.3,4.85) node[anchor=north west] {(0,0)}; \draw (.3,-.1) node[anchor=north west] {(0,1)}; \draw (1.6,3.15) node[anchor=north west] {(1,1)}; \draw (1.6,1.7) node[anchor=north west] {(2,2)}; \draw (-1,1.7) node[anchor=north west] {(1,2)}; \draw (-1,3.15) node[anchor=north west] {(2,1)}; \draw [line width=2.pt] (5.,1.)-- (6.,0.); \draw [line width=2.pt] (6.,0.)-- (8.,0.); \draw [line width=2.pt] (8.,0.)-- (9.,1.); \draw [line width=2.pt] (9.,1.)-- (9.,3.); \draw [line width=2.pt] (9.,3.)-- (7.,4.); \draw [line width=2.pt] (7.,4.)-- (5.,3.); \draw [line width=2.pt] (5.,3.)-- (5.,1.); \draw (6.3,4.85) node[anchor=north west] {(0,0)}; \draw (7.6,3.15) node[anchor=north west] {(1,1)}; \draw (7.6,1.7) node[anchor=north west] {(2,2)}; \draw (7.3,-0.1) node[anchor=north west] {(0,1)}; \draw (5,1.7) node[anchor=north west] {(1,2)}; \draw (5.3,-0.1) node[anchor=north west] {(2,0)}; \draw (5,3.15) node[anchor=north west] {(2,1)}; \draw [line width=2.pt] (11.,3.)-- (12.,4.); \draw [line width=2.pt] (12.,4.)-- (14.,4.); \draw [line width=2.pt] (14.,4.)-- (15.,3.); \draw [line width=2.pt] (15.,3.)-- (15.,1.); \draw [line width=2.pt] (15.,1.)-- (14.,0.); \draw [line width=2.pt] (14.,0.)-- (12.,0.); \draw [line width=2.pt] (12.,0.)-- (11.,1.); \draw [line width=2.pt] (11.,1.)-- (11.,3.); \draw (11.3,4.85) node[anchor=north west] {(0,0)}; \draw (13.3,4.85) node[anchor=north west] {(1,1)}; \draw (13.6,3.15) node[anchor=north west] {(2,2)}; \draw (13.6,1.7) node[anchor=north west] {(0,1)}; \draw (13.3,-0.1) node[anchor=north west] {(1,2)}; \draw (11.3,-0.1) node[anchor=north west] {(2,0)}; \draw (11,1.7) node[anchor=north west] {(0,2)}; \draw (11,3.15) node[anchor=north west] {(2,1)}; \begin{scriptsize} \draw [fill=black] (1,0.) circle (2.5pt); \draw [fill=black] (-1.,1.) circle (2.5pt); \draw [fill=black] (-1.,3.) circle (2.5pt); \draw [fill=black] (1,3.94) circle (2.5pt); \draw [fill=black] (3.,1.) circle (2.5pt); \draw [fill=black] (3.,3.) circle (2.5pt); \draw [fill=black] (6.,0.) circle (2.5pt); \draw [fill=black] (8.,0.) circle (2.5pt); \draw [fill=black] (5.,1.) circle (2.5pt); \draw [fill=black] (5.,3.) circle (2.5pt); \draw [fill=black] (9.,1.) circle (2.5pt); \draw [fill=black] (9.,3.) circle (2.5pt); \draw [fill=black] (7.,4.) circle (2.5pt); \draw [fill=black] (12.,0.) circle (2.5pt); \draw [fill=black] (14.,0.) circle (2.5pt); \draw [fill=black] (11.,1.) circle (2.5pt); \draw [fill=black] (11.,3.) circle (2.5pt); \draw [fill=black] (12.,4.) circle (2.5pt); \draw [fill=black] (14.,4.) circle (2.5pt); \draw [fill=black] (15.,3.) circle (2.5pt); \draw [fill=black] (15.,1.) circle (2.5pt); \end{scriptsize} \end{tikzpicture} \caption{Orthogonal Colourings of $C_6,C_7$, and $C_8$.} \label{Figure: Case 1} \end{figure} \begin{figure}[h!] \centering \begin{tikzpicture}[scale=.75,line cap=round,line join=round,>=triangle 45,x=1.0cm,y=1.0cm] \draw [line width=2.pt] (0.,3.)-- (0.,2.); \draw [line width=2.pt] (0.,2.)-- (0.,1.); \draw [line width=2.pt] (0.,1.)-- (1.,0.); \draw [line width=2.pt] (1.,0.)-- (2.,0.); \draw [line width=2.pt] (2.,0.)-- (3.,0.); \draw [line width=2.pt] (3.,0.)-- (4.,1.); \draw [line width=2.pt] (4.,1.)-- (4.,2.); \draw [line width=2.pt] (4.,2.)-- (4.,3.); \draw [line width=2.pt] (0.,3.)-- (1.,4.); \draw [line width=2.pt] (1.,4.)-- (3.,4.); \draw [line width=2.pt] (3.,4.)-- (4.,3.); \draw [line width=2.pt] (6.48,0.48)-- (6.,1.); \draw [line width=2.pt] (6.,1.)-- (6.,2.); \draw [line width=2.pt] (6.,2.)-- (6.,3.); \draw [line width=2.pt] (6.,3.)-- (7.,4.); \draw [line width=2.pt] (7.,4.)-- (8.,4.); \draw [line width=2.pt] (8.,4.)-- (9.,4.); \draw [line width=2.pt] (9.,4.)-- (10.,3.); \draw [line width=2.pt] (10.,3.)-- (10.,2.); \draw [line width=2.pt] (10.,2.)-- (10.,1.); \draw [line width=2.pt] (10.,1.)-- (9.,0.); \draw [line width=2.pt] (9.,0.)-- (8.,0.); \draw [line width=2.pt] (8.,0.)-- (7.,0.); \draw [line width=2.pt] (7.,0.)-- (6.48,0.48); \draw [line width=2.pt] (12.46,0.4)-- (12.,1.); \draw [line width=2.pt] (12.,1.)-- (12.,2.); \draw [line width=2.pt] (12.,2.)-- (12.,3.); \draw [line width=2.pt] (12.,3.)-- (13.,4.); \draw [line width=2.pt] (13.,4.)-- (14.,4.); \draw [line width=2.pt] (14.,4.)-- (15.,4.); \draw [line width=2.pt] (15.,4.)-- (15.62,3.5); \draw [line width=2.pt] (15.62,3.5)-- (16.,3.); \draw [line width=2.pt] (16.,2.)-- (16.,3.); \draw [line width=2.pt] (16.,2.)-- (16.,1.); \draw [line width=2.pt] (16.,1.)-- (15.,0.); \draw [line width=2.pt] (15.,0.)-- (14.,0.); \draw [line width=2.pt] (14.,0.)-- (13.,0.); \draw [line width=2.pt] (13.,0.)-- (12.46,0.4); \draw (2.3,4.85) node[anchor=north west] {(0,0)}; \draw (2.6,3.15) node[anchor=north west] {(1,1)}; \draw (2.6,2.4) node[anchor=north west] {(2,2)}; \draw (2.6,1.7) node[anchor=north west] {(3,3)}; \draw (2.5,-0.1) node[anchor=north west] {(0,1)}; \draw (1.3,-0.1) node[anchor=north west] {(1,2)}; \draw (0.1,-0.1) node[anchor=north west] {(2,3)}; \draw (0,1.7) node[anchor=north west] {(3,0)}; \draw (0,2.4) node[anchor=north west] {(0,2)}; \draw (0,3.15) node[anchor=north west] {(1,3)}; \draw (.3,4.85) node[anchor=north west] {(3,1)}; \draw (8.5,4.85) node[anchor=north west] {(0,0)}; \draw (8.6,3.15) node[anchor=north west] {(1,1)}; \draw (8.6,2.4) node[anchor=north west] {(2,2)}; \draw (8.6,1.7) node[anchor=north west] {(3,3)}; \draw (8.5,-0.1) node[anchor=north west] {(0,1)}; \draw (7.3,-0.1) node[anchor=north west] {(1,2)}; \draw (6.1,-0.1) node[anchor=north west] {(2,3)}; \draw (5.2,0.6) node[anchor=north west] {(3,0)}; \draw (6,1.7) node[anchor=north west] {(0,2)}; \draw (6,2.4) node[anchor=north west] {(1,3)}; \draw (6.1,4.85) node[anchor=north west] {(2,0)}; \draw (6,3.15) node[anchor=north west] {(3,1)}; \draw (7.3,4.85) node[anchor=north west] {(3,2)}; \draw (14.5,4.85) node[anchor=north west] {(0,0)}; \draw (15.6,4.2) node[anchor=north west] {(1,1)}; \draw (14.6,3.15) node[anchor=north west] {(2,2)}; \draw (14.6,2.4) node[anchor=north west] {(3,3)}; \draw (14.6,1.7) node[anchor=north west] {(0,1)}; \draw (14.5,-0.1) node[anchor=north west] {(1,2)}; \draw (13.3,-0.1) node[anchor=north west] {(2,3)}; \draw (12,3.15) node[anchor=north west] {(3,1)}; \draw (12.1,-0.1) node[anchor=north west] {(3,0)}; \draw (11.2,0.6) node[anchor=north west] {(0,2)}; \draw (12,1.7) node[anchor=north west] {(1,3)}; \draw (12,2.4) node[anchor=north west] {(2,0)}; \draw (12.1,4.85) node[anchor=north west] {(0,3)}; \draw (13.3,4.85) node[anchor=north west] {(2,1)}; \begin{scriptsize} \draw [fill=black] (1.,0.) circle (2.5pt); \draw [fill=black] (0.,1.) circle (2.5pt); \draw [fill=black] (0.,3.) circle (2.5pt); \draw [fill=black] (1.,4.) circle (2.5pt); \draw [fill=black] (3.,0.) circle (2.5pt); \draw [fill=black] (4.,1.) circle (2.5pt); \draw [fill=black] (4.,3.) circle (2.5pt); \draw [fill=black] (0.,2.) circle (2.5pt); \draw [fill=black] (4.,2.) circle (2.5pt); \draw [fill=black] (2.,0.) circle (2.5pt); \draw [fill=black] (3.,4.) circle (2.5pt); \draw [fill=black] (8.,0.) circle (2.5pt); \draw [fill=black] (7.,0.) circle (2.5pt); \draw [fill=black] (6.,1.) circle (2.5pt); \draw [fill=black] (6.,2.) circle (2.5pt); \draw [fill=black] (6.,3.) circle (2.5pt); \draw [fill=black] (7.,4.) circle (2.5pt); \draw [fill=black] (8.,4.) circle (2.5pt); \draw [fill=black] (9.,4.) circle (2.5pt); \draw [fill=black] (10.,3.) circle (2.5pt); \draw [fill=black] (10.,2.) circle (2.5pt); \draw [fill=black] (10.,1.) circle (2.5pt); \draw [fill=black] (9.,0.) circle (2.5pt); \draw [fill=black] (6.48,0.48) circle (2.5pt); \draw [fill=black] (12.,1.) circle (2.5pt); \draw [fill=black] (12.,2.) circle (2.5pt); \draw [fill=black] (12.,3.) circle (2.5pt); \draw [fill=black] (13.,4.) circle (2.5pt); \draw [fill=black] (14.,4.) circle (2.5pt); \draw [fill=black] (15.,4.) circle (2.5pt); \draw [fill=black] (16.,3.) circle (2.5pt); \draw [fill=black] (16.,2.) circle (2.5pt); \draw [fill=black] (16.,1.) circle (2.5pt); \draw [fill=black] (15.,0.) circle (2.5pt); \draw [fill=black] (14.,0.) circle (2.5pt); \draw [fill=black] (13.,0.) circle (2.5pt); \draw [fill=black] (12.46,0.4) circle (2.5pt); \draw [fill=black] (15.62,3.5) circle (2.5pt); \end{scriptsize} \end{tikzpicture} \caption{Orthogonal Colouring of $C_{11},C_{13}$, and $C_{14}$.} \label{Figure: Case 2} \end{figure} \begin{thrm} $O\chi(C_n)=\lceil\sqrt{n}\,\rceil$ if and only if $n>4$. \end{thrm} \begin{pf} Follows from Lemma \ref{Lemma:Cycle Part 1}, Lemma \ref{Lemma: Cycle Part 2}, Figure \ref{Figure: Case 1}, and Figure \ref{Figure: Case 2}. \end{pf} Now that the orthogonal chromatic number of $C_n$ is determined, the goal is to construct multiple mutual orthogonal colourings. This is accomplished if restrictions are put upon the size of the vertex set. The following theorem gives a method to construct multiple mutual orthogonal colourings of $C_n$ when $\lceil\sqrt{n}\,\rceil=p$, where $p$ is a prime number. \begin{thrm} If $\lceil\sqrt{n}\,\rceil=p$ is a prime number, then $O\chi_{p-2}(C_n)=p$. \end{thrm} \begin{pf} Let $v_i\in \mathbb{Z}_n$ where $0\leq i <n$. It is now shown that there are $p-1$ orthogonal assignments, but only $p-2$ of these assignments are proper colourings for a given $n$. Consider the following $p-1$ colourings: $$c_k(v_i)=\Big(i+k\Big\lfloor\frac{i}{p}\Big\rfloor\Big)(\textrm{mod}~p)$$ where $0\leq k < p$. It is now shown that any two are mutually orthogonal. Suppose otherwise, that is, $c_t(v_i)=c_t(v_j)$ and $c_s(v_i)=c_s(v_j)$, where $0\leq t<s<p$ and $0\leq i<j<n$. Then: \begin{align} (i-j)&\equiv t\Big(\Big\lfloor\frac{j}{p}\Big\rfloor-\Big\lfloor\frac{i}{p}\Big\rfloor\Big)(\textrm{mod}~p). \label{Equation: 1}\\ (i-j)&\equiv s\Big(\Big\lfloor\frac{j}{p}\Big\rfloor-\Big\lfloor\frac{i}{p}\Big\rfloor\Big)(\textrm{mod}~p). \label{Equation: 2} \end{align} Note that $t,s\in \mathbb{Z}_p$, which is a field because $p$ is a prime. Thus, if $t\neq 0$, then $t^{-1}$ and $s^{-1}$ exist. Then, Equation \ref{Equation: 1} and Equation \ref{Equation: 2} can be rearranged to give that $(i-j)(s^{-1}-t^{-1})\equiv 0(\textrm{mod}~p)$. However, $0\leq t<s<p$, so $t^{-1}\not\equiv s^{-1}(\textrm{mod}~p)$. Therefore, it must be that $i\equiv j(\textrm{mod}~p)$, or equivalently, $j=i+mp$ where $0<m<p$. Since $s,m\neq 0$, $sm\not\equiv 0(\textrm{mod}~p)$ because $\mathbb{Z}_p$ has no zero divisors. This implies that: \begin{align*} \Big(j+s\Big\lfloor\frac{j}{p}\Big\rfloor\Big)(\textrm{mod}~p)&\equiv \Big(i+s\Big\lfloor\frac{i+mp}{p}\Big\rfloor\Big)(\textrm{mod}~p)\\ &\equiv \Big(i+sm+s\Big\lfloor\frac{i}{p}\Big\rfloor\Big)(\textrm{mod}~p)\\ &\not\equiv \Big(i+s\Big\lfloor\frac{i}{p}\Big\rfloor\Big)(\textrm{mod}~p). \end{align*} This contradicts Equation \ref{Equation: 2} however, hence $t=0$. Putting $t=0$ into Equation \ref{Equation: 1} gives that $i\equiv j(\textrm{mod}~p)$, and the same contradiction arises. Therefore, the colourings are all mutually orthogonal. It remains to show that $p-2$ of the colourings are proper. Notice that $k\Big\lfloor\frac{i}{p}\Big\rfloor\leq k\Big\lfloor\frac{i+1}{p}\Big\rfloor\leq k\Big\lfloor\frac{i}{p}\Big\rfloor+k $. Therefore, $1\leq c_k(v_{i+1})-c_k(v_i)\leq k$ where $1\leq i \leq n-2$. Since $k<p$, this gives that $c_k(v_{i+1})\neq c_k(v_i)$. It remains to show that there is no colour conflict between the vertices $v_0$ and $v_{n-1}$. Note that the colour $0$ is assigned to $v_0$ in all of the colourings. Therefore, by the mutual orthogonality of the colourings, at most one colouring has $0$ assigned to the vertex $v_{n-1}$. Therefore, choosing the $k-1=p-2$ colourings that don't assign the colour $0$ to the vertex $v_{n-1}$ gives a $(p-2)$-orthogonal colouring of $C_n$ using $p$ colours. \end{pf} \section{Orthogonal Colourings of Circulant Graphs} Cayley graphs on the additive group $\mathbb{Z}_{p^2}$ are now considered, where $p$ is a prime. Also, the size of the generating set $S$ is varied throughout this section. To start, a method for orthogonally colouring $\Gamma(\mathbb{Z}_{p^2},S)$ is constructed when $|S|$ is sufficiently small. Let $\alpha\in\mathbb{Z}_p\backslash\{0\}$ and consider the function $\hat F_{\alpha,p}:\mathbb{Z}_p\times \mathbb{Z}_p\to \mathbb{Z}_{p^2}$ defined by $$\hat F_{\alpha,p}(i,j)=(((\alpha(j-i)(\textrm{mod}~p))+p(2i-j))(\textrm{mod}~p^2).$$ This function assigns colour pairs from $\mathbb{Z}_p\times\mathbb{Z}_p$ to the vertices of $\Gamma(\mathbb{Z}_{p^2},S)$. Therefore, the goal is to show that the inverse of this function is an orthogonal colouring. The following lemma shows that this assignment is injective. \begin{lem}\label{Lemma: Circulant Map 2} For every $\alpha\in\mathbb{Z}_p\backslash\{0\}$, $\hat F_{\alpha,p}(i,j)$ is a bijection. \end{lem} \begin{pf} Since $|\mathbb{Z}_p\times \mathbb{Z}_p|=|\mathbb{Z}_{p^2}|=p^2$, it is sufficient to show that $\hat F_{\alpha,p}$ is an injective function. Suppose $\hat F_{\alpha,p}(i,j)=\hat F_{\alpha,p}(r,s)$. For this equality to be true, the two modular components of $\hat F_{\alpha,p}$ must be equal. That is: \begin{align} \alpha(j-i)(\textrm{mod}~p)&\equiv \alpha(s-r)(\textrm{mod}~p). \label{Equation: 3}\\ p(2i-j)(\textrm{mod}~p^2)&\equiv p(2r-s)(\textrm{mod}~p^2). \label{Equation: 4} \end{align} Note that $\alpha$ has a multiplicative inverse in $\mathbb{Z}_p$, which is denoted by $\alpha^{-1}$. Therefore, Equation (\ref{Equation: 3}) can be rewritten as $(i-j)(\textrm{mod}~p)\equiv (r-s)(\textrm{mod}~p)$. Multiplying by $p$ gives that $p(i-j)(\textrm{mod}~p^2)\equiv p(r-s)(\textrm{mod}~p^2)$. Substituting this into Equation (\ref{Equation: 4}) gives that $pi(\textrm{mod}~p^2)\equiv pr(\textrm{mod}~p^2)$. Note that this implies that $i(\textrm{mod}~p)\equiv r(\textrm{mod}~p)$. Since $i$ and $r$ are elements of $\mathbb{Z}_p$, it follows that $i=r$. Lastly, substituting $i=r$ into $(i-j)(\textrm{mod}~p)\equiv (r-s)(\textrm{mod}~p)$ gives that $j(\textrm{mod}~p)=s(\textrm{mod}~p)$. Similarly, since $j$ and $s$ are elements of $\mathbb{Z}_p$, it follows that $j=s$. Therefore, $(i,j)=(r,s)$. \end{pf} In particular, $\hat F_{\alpha,p}^{-1}:\mathbb{Z}_{p^2} \to \mathbb{Z}_p\times \mathbb{Z}_p$ is an injective function. Since $\hat F_{\alpha,p}^{-1}$ is injective into $\mathbb{Z}_p\times\mathbb{Z}_p$, $\hat F_{\alpha,p}^{-1}$ will orthogonally assign the vertices to colour pairs in $\mathbb{Z}_p\times \mathbb{Z}_p$. Therefore, if it can be shown that the components of $\hat F_{\alpha,p}^{-1}$ are both proper colourings, then $\hat F_{\alpha,p}^{-1}$ will be an orthogonal colouring. For a general generating set $S$ and $\alpha$, this is not always the case. However, the following theorem shows that if the size of the generating set $S$ is sufficiently small, then there is an $\alpha$ for which $\hat F_{\alpha,p}^{-1}$ is a proper colouring of $\Gamma(\mathbb{Z}_{p^2},S)$. \begin{thrm}\label{Theorem: Circulant 1} For $p$ a prime, if $|S|<\frac{p-1}{2}$, then $O\chi(\Gamma(\mathbb{Z}_{p^2},S))=p$. \end{thrm} \begin{pf} The goal is to show that $\hat F_{\alpha,p}^{-1}$ is an orthogonal colouring for some $\alpha\in \mathbb{Z}_{p}\backslash\{0\}$. By Lemma \ref{Lemma: Circulant Map 2}, $\hat F_{\alpha,p}^{-1}$ is an orthogonal assignment of the vertices. It remains to show that there is an $\alpha$ such that $\hat F_{\alpha,p}^{-1}$ is a proper colouring. Suppose two vertices $k$ and $l$ receive the same colour in the first colouring. That is, $\hat F_{\alpha,p}^{-1}(k)=(i,(j+x)(\textrm{mod}~p))$ and $\hat F_{\alpha,p}^{-1}(l)=(i,j)$ for some $i,j\in \mathbb{Z}_p$ and $x\in\mathbb{Z}_p\backslash\{0\}$. Then $$k-l=\hat F_{\alpha,p}(i,(j+x)(\textrm{mod}~p))-\hat F_{\alpha,p}(i,j)=(((\alpha x)(\textrm{mod}~p))-px)(\textrm{mod}~p^2).$$ Similarly, suppose two vertices $k$ and $l$ receive the same colour in the second colouring. That is, $\hat F_{\alpha,p}^{-1}(k)=((i+x)(\textrm{mod}~p),j)$ and $\hat F_{\alpha,p}^{-1}(l)=(i,j)$ for some $i,j\in \mathbb{Z}_p$ and $x\in\mathbb{Z}_p\backslash\{0\}$. Then $$k-l=\hat F_{\alpha,p}((i+x)(\textrm{mod}~p),j)-\hat F_{\alpha,p}(i,j)=(((-\alpha x)(\textrm{mod}~p))+2px)(\textrm{mod}~p^2).$$ Let $A_\alpha$ be the set of differences in the first colouring and $B_\alpha$ be the set of differences in the second colouring. That is, $A_{\alpha}=\{(((\alpha x)(\textrm{mod}~p))-px)(\textrm{mod}~p^2)|x\in\mathbb{Z}_p\backslash\{0\}\}$ and $B_{\alpha}=\{(((-\alpha x)(\textrm{mod}~p))+2px)(\textrm{mod}~p^2)|x\in\mathbb{Z}_p\backslash\{0\}\}$. Therefore, there is a colour conflict in the first colouring if and only if $S\cap A_\alpha\neq \emptyset$. Similarly, there is a colour conflict in the second colouring in and only if $S\cap B_\alpha \neq \emptyset$. Properties of $A_{\alpha}$ and $B_{\alpha}$ are now discussed. The first is that $A_\alpha = B_{2\alpha(\textrm{mod}~p)}$. Since $\mathbb{Z}_p$ is a field, $2$ has a multiplicative inverse in $\mathbb{Z}_p$, which is denoted $2^{-1}$. Then, since $\mathbb{Z}_p$ has no zero divisors, $\mathbb{Z}_p\backslash\{0\}=-2\mathbb{Z}_p\backslash\{0\}$. Therefore, it follows that \begin{align*} B_{2\alpha(\textrm{mod}~p)}&=\{(((-2\alpha x)(\textrm{mod}~p))+2px)(\textrm{mod}~p^2)|x\in\mathbb{Z}_p\backslash\{0\}\}\\ &=\{(((-2\alpha (-2^{-1}x)(\textrm{mod}~p))+2p(-2^{-1}x)(\textrm{mod}~p^2)|x\in -2\mathbb{Z}_p\backslash\{0\}\}\\ &=\{(((\alpha x)(\textrm{mod}~p))-px)(\textrm{mod}~p^2)|x\in -2\mathbb{Z}_p\backslash\{0\}\}\\ &=\{(((\alpha x)(\textrm{mod}~p))-px)(\textrm{mod}~p^2)|x\in\mathbb{Z}_p\backslash\{0\}\}\\ &=A_\alpha. \end{align*} A second property of $A_\alpha$ is the following. The $A_\alpha$'s together with $\{m| m\in\mathbb{Z}_p\backslash\{0\}\}$ and $\{mp| m\in\mathbb{Z}_p\backslash\{0\}\}$ are disjoint. First, notice that $\alpha x(\textrm{mod}~p)\not\equiv 0(\textrm{mod}~p)$ for any $x\in\mathbb{Z}_p\backslash\{0\}$. Therefore, it follows that $A_{\alpha}\cap \{mp|m\in\mathbb{Z}_p\backslash\{0\}\}=\emptyset$. Next, notice that $-px(\textrm{mod}~p^2)\not\equiv 0(\textrm{mod}~p^2)$ for any $x\in\mathbb{Z}_p\backslash\{0\}$. Therefore, $A_\alpha\cap\{m| m\in\mathbb{Z}_p\backslash\{0\}\}=\emptyset$. It is also the case that $\{mp| m\in\mathbb{Z}_p\backslash\{0\}\}\cap\{m| m\in\mathbb{Z}_p\backslash\{0\}\}=\emptyset$. This is because the first set is multiples of $p$ and the second set is not. Therefore, it remains to show that the $A_\alpha$'s are all mutually disjoint. Let $\alpha_1,\alpha_2\in \mathbb{Z}_p\backslash\{0\}$ where $\alpha_1\neq \alpha_2$. Suppose that $A_{\alpha_1}\cap A_{\alpha_2}\neq\emptyset$. Then, there exists some $c=((\alpha_1 x(\textrm{mod}~p))-px)(\textrm{mod}~p^2)$ and $c=((\alpha_2 y(\textrm{mod}~p))-py)(\textrm{mod}~p^2)$ for some $x,y\in\mathbb{Z}_p\backslash\{0\}$. Note that $c$ can be uniquely written as $r-px$, where $r\in\mathbb{Z}_p\backslash\{0\}$. This means that $-px(\textrm{mod}~p^2)=-py(\textrm{mod}~p^2)$. This implies that $x(\textrm{mod}~p)=y(\textrm{mod}~p)$. Since $x$ and $y$ are elements of $\mathbb{Z}_p$, its follows that $x=y$. However, since $c$ can be uniquely written as $r-px$, it follows that $\alpha_1 x(\textrm{mod}~p)=\alpha_2 y(\textrm{mod}~p)$. Substituting $x=y$ into this gives that $\alpha_1(\textrm{mod}~p)=\alpha_2(\textrm{mod}~p)$. Then, since $\alpha_1$ and $\alpha_2$ are elements of $\mathbb{Z}_p\backslash\{0\}$, it follows that $\alpha_1=\alpha_2$, a contradiction. Therefore, the $A_\alpha$'s along with $\{mp| m\in\mathbb{Z}_p\backslash\{0\}\}$ and $\{m| m\in\mathbb{Z}_p\backslash\{0\}\}$ are all mutually disjoint. Now to show that there is a choice of $\alpha$ for which $\hat F_{\alpha,p}^{-1}$ is a proper colouring. Recall that there is a colour conflict in the first colouring if and only if $S\cap A_\alpha\neq \emptyset$ and there is a colour conflict in the second colouring if and only if $S\cap B_{\alpha}\neq \emptyset$. Let $c\in S$ and suppose that $c\in \{mp| m\in\mathbb{Z}_p\backslash\{0\}\}$ or $c\in\{m| m\in\mathbb{Z}_p\backslash\{0\}\}$. Then, since the $A_\alpha$'s, along with $\{mp| m\in\mathbb{Z}_p\backslash\{0\}\}$ and $\{m| m\in\mathbb{Z}_p\backslash\{0\}\}$ are all mutually disjoint, $c\not\in A_\alpha$ for any $\alpha\in\mathbb{Z}_p\backslash\{0\}$. Since $B_{2\alpha(\textrm{mod}~p)}=A_\alpha$, then $c\not\in B_\alpha$ for any $\alpha$ as well. Therefore, $c$ will not result in any colouring conflicts in the first colouring or in the second colouring. Suppose now that $c\in A_{\alpha_1}$ for some $\alpha_1\in\mathbb{Z}_p\backslash\{0\}$. Then, since the $A_\alpha$'s are disjoint, for any $\alpha_2\neq \alpha_1$, $c\not\in A_{\alpha_2}$. Thus, since $B_{2\alpha(\textrm{mod}~p)}=A_{\alpha}$, it follows that $c\in B_{2\alpha_1}$ but $c\not\in B_{\alpha_3}$ for any $\alpha_3\neq 2\alpha_1$. Hence, if $\alpha_1$ is chosen, then $c\in A_{\alpha_1}$, and there is a colour conflict in the first colouring. Similarly, if $2\alpha_1$ is chosen, then $c\in B_{2\alpha_1}$, and there is a colour conflict in the second colouring. However, any other choice of $\alpha$ will result in $c\not\in A_\alpha$ and $c\not\in B_\alpha$. Thus each $c\in S$ will result in at most $2$ restrictions on the choice of $\alpha$. Since that are at most $|S|<\frac{p-1}{2}$ elements in $S$, there are fewer than $p-1$ restrictions on the choice of $\alpha$. Since $\alpha\in\mathbb{Z}_p\backslash\{0\}$, there are $p-1$ choices for $\alpha$. Therefore, there are more choices than restrictions. \end{pf} Theorem \ref{Theorem: Circulant 1} says that if the size of the generating set is sufficiently small, then an orthogonal colouring can be constructed using only $p$ colours. This leads to the following question. \begin{ques} What is the largest $m$, such that if $|S|< m$, then $O\chi(\Gamma(\mathbb{Z}_{p^2},S))=p$? \end{ques} Theorem \ref{Theorem: Circulant 1} gives a lower bound of $m> \frac{p-1}{2}$. However, on the other hand, the following theorem by Klotz and Sander gives an upper bound on the value of $m$. \begin{thrm}[Klotz and Sander \cite{klotz2016uniquely}]\label{Theorem: Walter} If $S=\{\pm 1,\pm 2,\dots, \pm (p-1)\}$, then $\Gamma(\mathbb{Z}_{p^2},S)$ is uniquely $p$-colourable. \end{thrm} Note that uniquely $p$-colourable graphs can not be orthogonally coloured with $p$ colours, unless the graph is $K_p$. This is because the first and second colouring are the same, and thus if any colour is used twice in the first colouring, then that colour pair occurs twice. Therefore, since $|S|$ in Theorem \ref{Theorem: Walter} is of size $2p-2$, this gives an upper bound of $m<2p-2$. It is now shown that if no multiples of $p$ are in $S$, then the lower bound on $m$ can be increased to $p-1$. Let $\alpha\in\mathbb{Z}_{p^2}$ be such that $gcd(p^2,\alpha)=1$. Then, consider the following function: $F_{\alpha,p}:\mathbb{Z}_p\times \mathbb{Z}_p\to \mathbb{Z}_{p^2}$ defined by $$F_{\alpha,p}(i,j)=(ip+j\alpha)(\textrm{mod}~p^2).$$ This function will be used to assign colour pairs to $\Gamma(\mathbb{Z}_{p^2},S)$, when there are no multiples of $p$ in $S$. The following lemma shows that this assignment is injective. \begin{lem}\label{Lemma: Circulant Map 1} For any $\alpha\in\mathbb{Z}_{p^2}\backslash\{xp|0\leq x<p\}$, $F_{\alpha,p}(i,j)$ is a bijection. \end{lem} \begin{pf} Since $|\mathbb{Z}_p\times \mathbb{Z}_p|=|\mathbb{Z}_{p^2}|=p^2$, it is sufficient to show that $F_{\alpha,p}$ is surjective. Let $x\in \mathbb{Z}_{p^2}$. Since $gcd(p^2,\alpha)=1$, the multiplicative inverse $\alpha^{-1}\in \mathbb{Z}_{p^2}$. Now, by the Division Algorithm, there exist unique integers $q$ and $r$ such that $x=qp+r$ where $0\leq q,r <p$. Let $i=q$ and $j=r\alpha^{-1}(\textrm{mod}~p^2)$. Substituting gives that $F_{\alpha,p}(i,j)=(qp+r)(\textrm{mod}~p^2)=x(\textrm{mod}~p^2)$. Therefore, $F_{\alpha,p}$ is surjective, and thus bijective. \end{pf} In particular, $F_{\alpha,p}^{-1}:\mathbb{Z}_{p^2} \to \mathbb{Z}_n\times \mathbb{Z}_n$ is an injective function. Therefore, $F_{\alpha,p}^{-1}$ will orthogonally assign the vertices to colour pairs in $\mathbb{Z}_p\times\mathbb{Z}_p$. The following lemma shows that if the size of the generating set $S$ is sufficiently small and $S$ contains no multiples of $p$, then there is an $\alpha$ for which $F_{\alpha,p}^{-1}$ is a proper colouring of $\Gamma(\mathbb{Z}_{p^2},S)$. \begin{thrm}\label{Theorem: Circulant 1} For $p$ a prime, if $|S|<p$ and $xp\not\in S$ for all $x$ where $x\in \mathbb{Z}_p\backslash\{0\}$, then $O\chi(\Gamma(\mathbb{Z}_{p^2},S))=p$. \end{thrm} \begin{pf} It is shown that $F_{\alpha,p}^{-1}$ is an orthogonal colouring for some $\alpha\in \mathbb{Z}_{p^2}$ where $\gcd(p^2,\alpha)=1$. By Lemma \ref{Lemma: Circulant Map 1}, $F_{\alpha,p}^{-1}$ is an orthogonal assignment of the vertices. It remains to show that there is a choice for $\alpha$ such that $F_{\alpha,p}^{-1}$ is a proper colouring. Suppose two vertices $k$ and $l$ receive the same colour in the first colouring. That is, $F^{-1}_{\alpha,p}(k)=(i,(j+x)(\textrm{mod}~p))$ and $F^{-1}_{\alpha,p}(l)=(i,j)$ for some $i,j\in \mathbb{Z}_p$ and $x\in \mathbb{Z}_p\backslash\{0\}$. Note that $F_{\alpha,p}(i,(j+x)(\textrm{mod}~p))=ip+(j+x)\alpha=k$ and $F_{\alpha,p}(i,j)=ip+j\alpha=l$. Therefore, there is a colour conflict in the first colouring if and only if $k$ and $l$ are adjacent, which occurs when $k-l=x\alpha \in S$. Suppose two vertices $k$ and $l$ receive the same colour in the second colouring. That is, $F^{-1}_{\alpha,p}(k)=(i,j)$ and $F^{-1}_{\alpha,p}(l)=((i+x)(\textrm{mod}~p),j)$ for some $i,j\in \mathbb{Z}_p$ and $x\in\mathbb{Z}_p\backslash\{0\}$. Note that $F_{\alpha,p}(i,j)=ip+j\alpha=k$ and $F_{\alpha,p}((i+x)(\textrm{mod}~p),j)=(i+x)p+j\alpha=l$. Therefore, there is a colour conflict in the second colouring if and only if $k$ and $l$ are adjacent, so when $l-k=xp \in S$. By assumption, $xp\not\in S$ for all $x$, so there are no colour conflicts in the second colouring. So the only conflict that can occur is in the first colouring, which happens when $x\alpha \in S$. Since $x\in \mathbb{Z}_p\backslash\{0\}$, which is a field, $x^{-1}$ exists. Therefore, if $\alpha\not\in\bigcup_{x}\{x^{-1}S\}$, then there will be no colour conflicts. Note that since $p$ is a prime, there are $p(p-1)$ choices for $\alpha$ so that $\gcd(p^2,\alpha)=1$. Then, since there are $p-1$ choices for $x$, there are at most $(p-1)|S|<(p-1)p$ elements in $\bigcup_x\{x^{-1}S\}$. Therefore, there is an $\alpha$ such that $\alpha\not\in\bigcup_{x}\{x^{-1}S\}$. \end{pf} This concludes our study of circulant graphs with varying generating sets. Certain circulant graphs that have large generating sets are now studied. This is illustrated with the Paley graphs. To talk about this graph, relevant properties of finite fields are now summarized. \section{Paley Graphs} Finite fields of order $q$ exist if and only if $q=p^k$ where $p$ is a prime and $k\in\mathbb{Z}^+$. These fields are unique up to isomorphism, so they are denoted $\mathbb{F}_q$. The multiplicative group, $\mathbb{F}_q^*$, is cyclic, so all non-zero elements can be expressed as powers of a single element, called a primitive element of the field. Finite fields can be explicitly constructed as such. If $q=p^k$, then $\mathbb{F}_q\cong \mathbb{Z}_p[X]/(P)$ where $(P)$ is the ideal generated by an irreducible polynomial $P$ of degree $k$ in $\mathbb{Z}_p[X]$. The Paley graph, denoted $QR(q)$, can be constructed as a Cayley graph of $\mathbb{F}_q$. Let $\alpha$ be a primitive element of $\mathbb{F}_q$ and let $S=\{\alpha^{2m}:1\leq m\leq \frac{q-1}{2}\}$. That is, $S$ is the set of all quadratic residues in $\mathbb{F}_q$. Then the Paley graph $QR(q)=\Gamma(\mathbb{F}_q,S)$. Paley graphs have a variety of interesting properties, but the one that is used is that they are self-complementary \cite{jones2016paley}. This allows the cliques in the Paley graphs to be considered as colour classes, since they can be turned into independent sets by taking the complement. The following lemma, which is an exercise in \cite{fraleigh2003first}, describes a relation between cosets of two different subgroups. \begin{lem}\label{Lemma: Coset Intersection} Let $G$ be a group, $H\leq G$, and $K\leq G$. Then for any $a,b,c\in G$, either $(a+H)\cap (b+K)=\emptyset$ or $(a+H)\cap (b+K)= c+(H\cap K)$. \end{lem} The following theorem gives a method for constructing multiple orthogonal colourings of Paley graphs by utilizing the structure of the finite field. The idea is to find particular cosets that share at most one element. Then, Lemma \ref{Lemma: Coset Intersection} gives that their intersection also contains at most one element. These cosets can then be used as the colour classes. Since they intersect in at most one element, they will be an orthogonal assignment of the colours. The following theorem formalizes this argument. \begin{thrm} For $p$ a prime and an integer $r\geq 1$, $O\chi_{\frac{p^r+1}{2}}(QR(p^{2r}))=p^r$. \end{thrm} \begin{pf} Let $G=\mathbb{F}_{p^{2r}}$. Since $p^r\mid p^{2r}$, there exists a subfield $H\subset G$ where $H\cong \mathbb{F}_{p^r}$. Also, notice that $H^*$ is a multiplicative subgroup of $G^*$ with index $\frac{p^{2r}-1}{p^r-1}=p^r+1$. Thus, $H=\{0\}\bigcup\{\alpha^{m(p^r+1)}|1\leq m < p^r-1\}$ where $\alpha$ is a primitive element of $G$. Next, for $0\leq i < \frac{p^r+1}{2}$, consider the sets $H_i=\{0\}\bigcup\{\alpha^{m(p^r+1)+2i}|1\leq m < p^r-1\}$. The sets $H_i$ are additive subgroups of $G$. To see this, it suffices to show for $x,y\in H_i$ that $x-y\in H_i$. \begin{align*} x-y&=\alpha^{m(p^r+1)+2i}-\alpha^{n(p^r+1)+2i} &&\text{by the definition of $H_i$}\\ &=\alpha^{2i}(\alpha^{m(p^r+1)}-\alpha^{n(p^r+1)})\\ &=\alpha^{2i}(\alpha^{t(p^r+1)}) &&\text{because $\mathbb{F}_{p^{2r}}$ is a field}\\ &=\alpha^{t(p^r+1)+2i}\\ &\in H_i &&\text{by the definition of $H_i$.} \end{align*} Therefore, the sets $H_i$ are additive subgroups of $G$. Since $\alpha^{m(p^r+1)+2i}$ is an even power of $\alpha$, $H_i\subset S$ and so $H_i$ are cliques in $QR(p^{2r})$. Similarly, the $p^r-1$ cosets of $H_i$ are cliques in $QR(p^{2r})$. Therefore, $H_i$ and its cosets are independent sets in the complement graph. For $1\leq j \leq p^r-1$, let $H_i+k_j$ be the $j$th coset of $H_i$. Then, define the colourings of the complement as $c_i(x)=j$ for $x\in H_i+k_j$. Since these are independent sets, the $c_i$ are proper colourings. It remains to show that these colourings are mutually orthogonal. Notice that $H_i\bigcap H_j=\{0\}$. Therefore, by Lemma \ref{Lemma: Coset Intersection}, any coset of $H_i$ and any coset of $H_j$ will intersect in at most 1 element. Therefore, the colour pair $(i,j)$ is only assigned to at most one vertex. Hence, there is no orthogonal conflict. \end{pf} \section{Orthogonal Colourings of Product Graphs} The Cartesian product of two graphs $G$ and $H$, $G\square H$, has vertex set $V(G)\times V(H)$, and two vertices $(u_1,v_1)$ and $(u_2,v_2)$ in $G\square H$ are adjacent if and only if either $u_1=u_2$ and $v_1v_2\in E(H)$ or $v_1=v_2$ and $u_1u_2\in E(G)$. A relationship between Cayley graphs and the Cartesian graph product was given by Theron \cite{theron1989extension}. He showed that if $G_1$ and $G_2$ are groups with Cayley graphs $\Gamma(G_1,S_1)$ and $\Gamma(G_2,S_2)$ respectively, then a Cayley graph of $G_1\times G_2$ is $\Gamma(G_1\times G_2,(S_1\times\{0\})\cup(\{0\}\times S_2))=\Gamma(G_1,S_1)\square \Gamma(G_2,S_2)$. Let $S=\mathbb{Z}_q\backslash\{0\}$ and consider the group $\mathbb{Z}_q^d$ with addition (component-wise) as the group operation. Notice that $\Gamma(\mathbb{Z}_q,S)=K_q$. Therefore, by the above result, a possible Cayley graph for $\mathbb{Z}_q^d$ is the Cartesian graph product of $d$ complete graphs $K_q$. This is what is defined to be the Hamming graph $H(d,q)$. The following theorem shows how the Cartesian graph product affects orthogonal colourings. \begin{thrm}\label{Theorem: Cartesian Product} If $G$ has $n^2$ vertices, $H$ has $m^2$ vertices, and $O\chi(G)=n\geq m$, then $O\chi(G\square H)=nm$. \end{thrm} \begin{pf} Label $V(G)=\{v_k:0\leq k <n^2\}$ and $V(H)=\{(u_i,u_j):0\leq i,j<m \}$ where $n\geq m$. Let $f=(f_1,f_2)$ be an orthogonal colouring of $G$ using the colours $0,1,\dots, n-1$. It is shown that $g=(g_1,g_2)$ is an orthogonal colouring of $G\square H$ where: \begin{align*} g_1((v_k,(u_i,u_j)))&=(f_1(v_k)+j)(\textrm{mod}~n)+in.\\ g_2((v_k,(u_i,u_j)))&=(f_2(v_k)+i)(\textrm{mod}~n)+jn. \end{align*} First, it is shown that $g$ has no orthogonal conflicts. Let $v_{k_1},v_{k_2}\in V(G)$ and let $(u_{i_1},u_{j_1}),(u_{i_2},u_{j_2})\in V(H)$. If $g((v_{k_1},(u_{i_1},u_{j_1})))=g((v_{k_2},(u_{i_2},u_{j_2})))$, then: \begin{align} (f_1(v_{k_1})+j_1)(\textrm{mod}~n)+i_1n&=(f_1(v_{k_2})+j_2)(\textrm{mod}~n)+i_2n. \label{Equation: ONE}\\ (f_2(v_{k_1})+i_1)(\textrm{mod}~n)+j_1n&=(f_2(v_{k_2})+i_2)(\textrm{mod}~n)+j_2n. \label{Equation: TWO} \end{align} Without loss of generality, suppose that $i_1<i_2$. Then: \begin{align*} (f_1(v_{k_1})+j_1)(\textrm{mod}~n)+i_1n &<n+i_1n\\ &=n(1+i_1)\\ &\leq i_2 n\\ &\leq (f_1(v_{k_2})+j_2)(\textrm{mod}~n)+i_2n. \end{align*} Therefore, by transitivity, $(f_1(v_{k_1})+j_1)(\textrm{mod}~n)+i_1n<(f_1(v_{k_2})+j_2)(\textrm{mod}~n)+i_2n$, which contradicts Equation \eqref{Equation: ONE}, thus $i_1=i_2$. A similar argument shows that $j_1=j_2$. Substituting $i_1=i_2$ and $j_1=j_2$ into Equations \eqref{Equation: ONE} and \eqref{Equation: TWO}, gives $f_1(v_{k_1})=f_1(v_{k_2})$ and $f_2(v_{k_1})=f_2(v_{k_2})$. Hence, $v_{k_1}=v_{k_2}$ because $f$ is an orthogonal colouring. Thus, $(v_{k_1},(u_{i_1},u_{j_1}))=(v_{k_2},(u_{i_2},u_{j_2}))$. It remains to show that $g_1$ and $g_2$ are proper colourings of $G\square H$. First, consider adjacencies of the form $(v_{k_1},(u_i,u_j))\sim(v_{k_2},(u_i,u_j))$. Since $v_{k_1}\sim v_{k_2}$ in $G$ and $f_1$ is a proper colouring of $G$, $f_1(v_{k_1})\neq f_1(v_{k_2})$. Thus \begin{align*} g_1((v_{k_1},(u_i,u_j)))&=(f_1(v_{k_1})+j)(\textrm{mod}~n)+in \\ &\neq (f_1(v_{k_2})+j)(\textrm{mod}~n)+in \\ &= g_1((v_{k_2},(u_i,u_j)). \end{align*} Next, consider adjacencies of the form $(v_k,(u_{i_1},u_{j_1}))\sim (v_k,(u_{i_2},u_{j_2}))$. Suppose that $i_1=i_2$ and $j_1\neq j_2$, then: \begin{align*} g_1((v_k,(u_{i_1},u_{j_1})))&=(f_1(v_k)+ j_1)(\textrm{mod}~n) +i_1n\\ &\neq (f_1(v_k)+ j_2)(\textrm{mod}~n) +i_1n &&\text{because $m\leq n$}\\ &=(f_1(v_k)+ j_2)(\textrm{mod}~n) +i_2n\\ &=g_1((v_k,(u_{i_2},u_{j_2}))). \end{align*} Thus, there are no colour conflicts in the case. If $i_1\neq i_2$, then the argument used to prove the orthogonality of $g$ shows that $g_1((v_k,(u_{i_1},u_{j_1})))\neq g_1((v_k,(u_{i_2},u_{j_2})))$. Hence, there are no conflicts in this case either. Therefore, $O\chi(G\square H)=nm$. \end{pf} Note that $H(d+2,q)=H(d,q)\square H(2,q)$. Therefore, if $O\chi(H(2,q))=q$, then Theorem \ref{Theorem: Cartesian Product} allows the use of induction to orthogonally colour Hamming graphs of the form $H(2d,q)$. Notice that a pair of orthogonal Latin squares corresponds to an orthogonal colouring of $K_q\square K_q= H(2,q)$. It is known that orthogonal Latin squares of size $n\neq 2,6$ exist \cite{bose1960further,tarry1900probleme}, giving the following lemma. \begin{lem}\label{Lemma: Hamming Graph} If $q\neq 2,6$ then $O\chi(H(2,q))=q$. \end{lem} \begin{cor}\label{Corollary: Case 1} If $q\neq 2,6$, and $d\geq 1$, then $O\chi(H(2d,q))=q^d$. \end{cor} \begin{pf} Proceed by induction on $d$. If $d=1$, then $O\chi(H(2,q))=q$ by Lemma \ref{Lemma: Hamming Graph}. Assume true for $d\leq k$, $k\geq 1$. Then, $H(2(k+1),q)=H(2k,q)\square H(2,q)$. By the induction hypothesis, $O\chi(H(2k,q))=q^k$. Therefore, by Theorem \ref{Theorem: Cartesian Product} it follows that, $O\chi(H(2(k+1),q)=q^{k+1}$. \end{pf} It remains to find orthogonal colourings of Hamming graphs when $q= 2,6$. In the case when $q=2$, $O\chi(H(4,2))=4$, as shown in Figure \ref{Figure: Orthogonal Q4}. Theorem \ref{Theorem: Cartesian Product} and induction gives the following corollary. \begin{figure}[h!] \centering \begin{tikzpicture}[line cap=round,line join=round,>=triangle 45,x=1cm,y=1cm] \draw [line width=2pt] (0,4)-- (1,0); \draw [line width=2pt] (0,4)-- (2,0); \draw [line width=2pt] (0,4)-- (3,0); \draw [line width=2pt] (1,4)-- (0,0); \draw [line width=2pt] (1,4)-- (2,0); \draw [line width=2pt] (1,4)-- (3,0); \draw [line width=2pt] (2,4)-- (1,0); \draw [line width=2pt] (2,4)-- (0,0); \draw [line width=2pt] (2,4)-- (3,0); \draw [line width=2pt] (3,4)-- (2,0); \draw [line width=2pt] (3,4)-- (1,0); \draw [line width=2pt] (3,4)-- (0,0); \draw [line width=2pt] (4,4)-- (5,0); \draw [line width=2pt] (4,4)-- (6,0); \draw [line width=2pt] (4,4)-- (7,0); \draw [line width=2pt] (5,4)-- (4,0); \draw [line width=2pt] (5,4)-- (6,0); \draw [line width=2pt] (5,4)-- (7,0); \draw [line width=2pt] (6,4)-- (7,0); \draw [line width=2pt] (6,4)-- (5,0); \draw [line width=2pt] (6,4)-- (4,0); \draw [line width=2pt] (7,4)-- (6,0); \draw [line width=2pt] (7,4)-- (5,0); \draw [line width=2pt] (7,4)-- (4,0); \draw [line width=2pt] (0,4)-- (4,0); \draw [line width=2pt] (1,4)-- (5,0); \draw [line width=2pt] (2,4)-- (6,0); \draw [line width=2pt] (3,4)-- (7,0); \draw [line width=2pt] (0,0)-- (4,4); \draw [line width=2pt] (1,0)-- (5,4); \draw [line width=2pt] (2,0)-- (6,4); \draw [line width=2pt] (3,0)-- (7,4); \draw (-0.6,4.7) node[anchor=north west] {$(0,0)$}; \draw (0.4,4.7) node[anchor=north west] {$(0,1)$}; \draw (1.4,4.7) node[anchor=north west] {$(0,2)$}; \draw (2.4,4.7) node[anchor=north west] {$(0,3)$}; \draw (6.4,4.7) node[anchor=north west] {$(1,0)$}; \draw (5.4,4.7) node[anchor=north west] {$(1,1)$}; \draw (4.4,4.7) node[anchor=north west] {$(1,2)$}; \draw (3.4,4.7) node[anchor=north west] {$(1,3)$}; \draw (-0.6,0) node[anchor=north west] {$(2,0)$}; \draw (0.4,0) node[anchor=north west] {$(2,1)$}; \draw (1.4,0) node[anchor=north west] {$(2,2)$}; \draw (2.4,0) node[anchor=north west] {$(2,3)$}; \draw (6.4,0) node[anchor=north west] {$(3,0)$}; \draw (5.4,0) node[anchor=north west] {$(3,1)$}; \draw (4.4,0) node[anchor=north west] {$(3,2)$}; \draw (3.4,0) node[anchor=north west] {$(3,3)$}; \begin{scriptsize} \draw [fill=black] (0,4) circle (2.5pt); \draw [fill=black] (1,4) circle (2.5pt); \draw [fill=black] (2,4) circle (2.5pt); \draw [fill=black] (3,4) circle (2.5pt); \draw [fill=black] (4,4) circle (2.5pt); \draw [fill=black] (5,4) circle (2.5pt); \draw [fill=black] (6,4) circle (2.5pt); \draw [fill=black] (7,4) circle (2.5pt); \draw [fill=black] (0,0) circle (2.5pt); \draw [fill=black] (1,0) circle (2.5pt); \draw [fill=black] (2,0) circle (2.5pt); \draw [fill=black] (3,0) circle (2.5pt); \draw [fill=black] (4,0) circle (2.5pt); \draw [fill=black] (5,0) circle (2.5pt); \draw [fill=black] (6,0) circle (2.5pt); \draw [fill=black] (7,0) circle (2.5pt); \end{scriptsize} \end{tikzpicture} \caption{Orthogonal Colouring of $H(4,2)$} \label{Figure: Orthogonal Q4} \end{figure} \begin{cor} If $d\geq 1$, then $O\chi(H(4d,2))=2^{2d}$ \end{cor} There are some open problems left to be explored for orthogonal colourings of Hamming graphs. First, to complete the categorization of even Hamming graphs, an orthogonal colouring of $H(4,6)$ using 36 colours would need to be found. Secondly, orthogonal colourings of odd Hamming graphs, $H(2d+1,q)$, have not been studied. Lastly, multiple orthogonal colourings of Hamming graphs have not been studied. In the case of $H(2,q)$, this would correspond to a collection of mutually orthogonal Latin squares. \section*{Acknowledgements} This research was supported by an NSERC Discovery grant. \bibliographystyle{amsplain}
{ "timestamp": "2020-07-06T02:19:27", "yymm": "1910", "arxiv_id": "1910.14454", "language": "en", "url": "https://arxiv.org/abs/1910.14454", "abstract": "Two colourings of a graph are orthogonal if they have the property that when two vertices are coloured with the same colour in one colouring, then those vertices receive distinct colours in the other colouring. In this paper, orthogonal colourings of Cayley graphs are discussed. Firstly, the orthogonal chromatic number of cycle graphs are completely determined. Secondly, the orthogonal chromatic number of certain circulant graphs is explored. Lastly, orthogonal colourings of product graphs and Hamming graphs are studied.", "subjects": "Combinatorics (math.CO)", "title": "Orthogonal Colourings of Cayley Graphs", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9912886152849367, "lm_q2_score": 0.824461928533133, "lm_q1q2_score": 0.817279723490758 }
https://arxiv.org/abs/2208.04666
Nilpotent probability of compact groups
Let $k$ be any positive integer and $G$ a compact (Hausdorff) group. Let $\mf{np}_k(G)$ denote the probability that $k+1$ randomly chosen elements $x_1,\dots,x_{k+1}$ satisfy $[x_1,x_2,\dots,x_{k+1}]=1$. We study the following problem: If $\mf{np}_k(G)>0$ then, does there exist an open nilpotent subgroup of class at most $k$? The answer is positive for profinite groups and we give a new proof. We also prove that the connected component $G^0$ of $G$ is abelian and there exists a closed normal nilpotent subgroup $N$ of class at most $k$ such that $G^0N$ is open in $G$.
\section{Introduction and Results} Let $G$ be a compact (Hausdorff) group and denote by ${\mf m}_G$ the (unique) Haar measure on $G$. The following question is proposed in \cite[Question 1.20.]{MTVV}: \begin{qu}\label{Question} Let $G$ be a compact group, and suppose that $${\mathcal N}_k(G):=\{(x_1,\dots, x_{k+1}) \in G^{k+1} \;|\; [x_1, \dots, x_{k+1}] = 1\},$$ has positive Haar measure in $G^{k+1}$. Does $G$ have an open $k$-step nilpotent subgroup? \end{qu} Here $[x_1,x_2]=x_1^{-1}x_2^{-1}x_1 x_2$ and $[x_1,\dots,x_{k+1}]$ is inductively defined by $[[x_1,\dots,x_k],x_{k+1}]$ for all $k>1$. The Haar measure of $\mathcal N_k(G)$ is called the degree of $k$-nilpotence of the group $G$, and denoted in \cite{MTVV} by $\mf{dc}^k(G)$. \par Denote by $\mathsf{PA}_k$ the class of compact groups giving positive answer to Question \ref{Question}. In \cite{LP} it is proved that $\mathsf{PA}_1$ contains profinite groups and in \cite{HR} it is proved that all compact groups includes $\mathsf{PA}_1$. In \cite{AS0} it is shown that $\mathsf{PA}_2$ contains all compact groups. Martino et al. have shown that $\mathsf{PA}_k$ contains all profinite groups for any $k$ (see \cite[Theorem 1.19.]{MTVV}). Here in Theorem \ref{anotherproof} below, we give another proof of the latter result without using submultiplicativity of the $k$-nilpotence degree of a finite group i.e. for a normal subgroup $N$ of a finite group $G$, $\mf{dc}^k(G)\leq\mf{dc}^k(G/N)\mf{dc}^k(N)$ (see \cite[Theorem 1.21.]{MTVV}).\\ Our main result is the following. \begin{thm}\label{main} Let $G$ be a compact group with $\mf{np}_k(G)>0$. Then the connected component $G^0$ of $G$ is abelian and there exists a closed normal nilpotent subgroup $N$ of class at most $k$ such that $G^0N$ is open in $G$. \end{thm} \section{Relative nilpotent probability and submultiplicativity} We denote the set of continuous functions on a compact group $G$ by $\mathcal C(G)$. If $N$ is a closed normal subgroup of $G$, then $G/N$ is a compact group, equipped with quotient topology. There exists a projection-like map $P:\mathcal C(G)\rightarrow\mathcal C(G/N)$ defined by \[ Pf(\overline{g}):=\int_Nf(gy)\,\text dy,\] where $\text dy$ stands for the Haar measure on $N$. We will use the following two simple lemmas repeatedly: \begin{lem}\label{Gotosubgroups} Let $G$ be a compact group and $N$ be a closed normal subgroup of $G$. Then, for all $f\in\mathcal C(G)$ \[ \int_Gf(g)\,\text dg=\int_{G/N}Pf(\overline{g})\,\text d\overline{g}.\] \end{lem} \begin{proof} The functional \[ I:\mathcal C(G)\rightarrow\mathbb C, \quad f\mapsto\int_{G/N}Pf(\overline{g})\,\text d\overline{g}\] is nonzero, positive and left-invariant, also $I(1)=1$, so by the uniqueness of Haar integral the result holds. \end{proof} \begin{rem} We have an analog of Lemma \ref{Gotosubgroups} whenever $H$ is a closed not necessarily normal subgroup, see \cite[(2.49) Theorem]{folland}. \end{rem} The Uniqueness of Harr measure has another achievement; consider a continuous epimorphism $\phi:G\rightarrow H$ between compact groups. Then \[I:\mathcal C(H)\rightarrow\mathbb C, \quad f\mapsto\int_Gf\circ\phi(g)\,\text dg\] defines a nonzero, positive and left-invariant functional. Aslo, $I(1)=1$, so we have \begin{lem}\label{themeasureinquoitient} Let $\phi:G\rightarrow H$ be a continuous epimorphism between compact groups. Then \[\int_Hf(h)\,\text dh=\int_Gf\circ\phi(g)\,\text dg\quad(f\in\mathcal C(H)).\] \end{lem} \begin{lem} Let $H$ be a closed subgroup of a group $G$. Then $yC_G(x)\cap H$ is either empty or a coset of $C_H(x)$, for all $x,y\in G$. In particular, we have $\mf m_H(yC_G(x)\cap H)\leq\mf m_H(C_H(x))$, for all $x,y\in G$. \end{lem} \begin{proof} If $z\in C_G(x)\cap y^{-1}H$, then one can easily check that $C_G(x)\cap y^{-1}H=zC_H(x)$. \end{proof} \begin{prop}\label{npleqcp} Let $H$ be a closed subgroup of a compact group $G$. Then $\mf{np}(H;x,y)\leq\mf{cp}(H)$ for all $x$ and $y$ in $G$. \end{prop} \begin{proof} By the above lemma, we have \begin{align*} \mf{np}(H;x,y)&=\int_H\int_H1_{\mathcal N(H;x,y)}(\xi,\eta)\,\text d\eta\,\text d\xi\\ &=\int_H\mf m_H(x^{-1}C_G(y\eta)\cap N)\,\text d\eta\\ &\leq\int_H\mf m_H(C_H(y\eta))\,\text d\eta\\ &=\int_H\mf m_H(y^{-1}C_G(\xi)\cap N)\,\text d\xi\\ &\leq\int_H\mf m_H(C_H(\xi))\,\text d\xi=\mf{cp}(H). \end{align*} \end{proof} \begin{rem}\label{npleqnp} In \cite[Lemma 6.2.]{MTVV}, using combinatorics methods, the authors have shown that for a normal subgroup $N$ of a \textit{finite} group $G$, \[\mf{np}(N; x_1,\dotsc, x_n)\leq\mf{np}(N;\overbrace{1,\dotsc,1}^{n\text{-times}})\] for all $x_k\in G$, $1\leq k\leq n$. We could not implement these methods in the case of infinite compact groups to have the above inequality for a closed normal subgroup $N$, and of course we do not need this generalization. \end{rem} Now, we are going to prove that $\mf{np}(N; x_1,\dotsc, x_n)$ is bounded away from 1, for all closed normal subgroup $N$ of a compact group $G$ which is not nilpotent of class at most $k$ and all $x_k\in G$, $1\leq k\leq n$. \begin{prop}\label{2.4n} Let $H$ be a closed subgroup of a compact group $G$. Then \[ \mf{np}(H; x_1,\dotsc,x_{n+1})\leq\frac12\pr{1+ \mf{np}\pr{\frac H{\mf Z\cap H}; \overline{x_1},\dotsc,\overline{x_n}}},\] for all $x_1,\dotsc,x_n\in G$, where $\mf Z$ is the center of $G$, and $\overline x$ denotes the image of $x$ in $H/(\mf Z\cap H)$. \end{prop} \begin{proof} Put $K:=\mf Z\cap H$, \[ X=\set{\mf y=(y_1,\dotsc,y_n)\in H^n: [x_1y_1,\dotsc,x_ny_n]\in K}\] and \[ \mathcal N=\mathcal N\pr{\frac HK; \overline{x_1},\dotsc,\overline{x_n}}.\] Then one can easily verify that \begin{align}\label{1} 1_X(\mf y\mf z)=1_X(\mf y)\quad (\mf y\in H^n, \mf z\in K^n) \end{align} ($1_X$ is the characteristic function of $X$) and \begin{align}\label{2} 1_X(\mf y)=1\quad\text{if and only if}\quad 1_{\mathcal N}(\bar{\mf y})=1. \end{align} Now, we have \begin{align*} \mf{np}(H; x_1,\dotsc,x_{n+1})&=\int_{H^n}\mf m_H\pr{x_{n+1}^{-1}C_G([x_1y_1,\dotsc,x_ny_n])\cap H}\,\text d\mf y\\ &\leq\int_{H^n}\mf m_H\pr{C_H([x_1y_1,\dotsc,x_ny_n])}\,\text d\mf y\\ &=\int_{X}\mf m_H\pr{C_H([x_1y_1,\dotsc,x_ny_n])}\,\text d\mf y\\ &\quad\,+\int_{H^n\setminus X}\mf m_H\pr{C_H([x_1y_1,\dotsc,x_ny_n])}\,\text d\mf y\\ &\leq\mf m_{H^n}(X)+\frac12(1-\mf m_{H^n}(X))\\ &=\frac12\pr{1+\mf m_{H^n}(X)}. \end{align*} The latter inequality holds because on $X^c$, $C_G([x_1y_1,\dots,x_ny_n])$ has index at least 2. We compute $\mf m_{H^n}(X)$: \begin{align*} \mf m_{H^n}(X)&=\int_{H^n}1_{X}(\mf y)\,\text d\mf y\\ &=\int_{H^n/K^n}\int_{K^n}1_{X}(\mf y\mf z)\,\text d\mf z\,\text d\bar{\mf y} , \quad \text{by Lemma \ref{Gotosubgroups},} \\ &=\int_{H^n/K^n}1_{X}(\mf y)\,\text d\bar{\mf y}, \quad \text{by (\ref{1})},\\ &=\int_{(H/K)^n}1_{\mathcal N}(\bar{\mf y})\,\text d\bar{\mf y}, \quad\text{by Lemma \ref{themeasureinquoitient}},\\ &= \mf{np}\pr{\frac H{\mf Z\cap H}; \overline{x_1},\dotsc,\overline{x_n}}. \end{align*} \end{proof} By Theorem \ref{2.4n} we obtain the following characterization of closed and nilpotent normal subgroups of a compact group: \begin{prop}\label{nocamn} Let $H$ be a closed subgroup of a compact group $G$. Then $H$ is nilpotent of class at most $n$ if and only if \[\mf{np}(H;x_1,\dotsc,x_{n+1})=1\] for some $x_k\in G$, $1\leq k\leq n+1$. \end{prop} \begin{proof} Only one direction needs proof.\par Using induction on $n$, we prove the result. For $n=1$, assume that $\mf{np}(H; x,y)=1$ for some $x,y\in G$, so by Proposition \ref{npleqcp}, $\mf{cp}(H)=1$, whence implies that $H$ is an abelian group. Assuming that the result is valid for $n-1$, we prove its correctness for $n$. If \[\mf{np}(H;x_1,\dotsc,x_{n+1})=1\] for some $x_k\in G$, $1\leq k\leq n+1$, then by Proposition \ref{2.4n}, we get $\mf{np}\pr{\frac H{\mf Z\cap H}; \overline{x_1},\dotsc,\overline{x_n}}=1$. Now, by induction hypothesis, $\frac H{\mf Z\cap H}$ is nilpotent of class at most $n-1$, so, $H$ is nilpotent of class at most $n$. \end{proof} \begin{defn}\label{rknil} Let $H$ be a closed subgroup of a compact group $G$. The \textit{the relative $k$-nilpotence of $H$ in $G$}, denoted by $\mf{np}_{k,G}(H)$, defined to be \[\sup\set{\mf{np}(H, x_1,\dotsc, x_{k+1}): x_1,\dotsc, x_k\in G}.\] When $H=G$, then we remove $G$ from the index and simply write $\mf{np}_k(G)$. \end{defn} \begin{rem}\label{indepen} Since the Haar measure is left-invariant, one can easily verify that $\mf{np}_k(G)=\mf{np}(G;x_1,\dotsc,x_{k+1})$ for all $x_1,\dotsc,x_{k+1}\in G$. So, $\mf{np}_k(G)$ is nothing but the degree of $k$-step nilpotence of $G$ defined in \cite{MTVV}. Also, by Lemma \ref{npleqcp}, we have $\mf{np}_{1,G}(H)=\mf{cp}(H)$ for a closed subgroup $H$ of a compact group $G$. \end{rem} The following result follows inductively from Proposition \ref{nocamn}, \ref{npleqcp} and \cite[Theorem]{Gus}. \begin{thm}\label{3/2} Let $H$ be closed subgroup of a compact group $G$ which is not nilpotent of class at most $n$. Then $\mf{np}_{k,G}(H)\leq 1-\frac{3}{2^{k+1}}$, for all positive integer $k$. \end{thm} By Theorem \ref{3/2} in hand, we give another (simple) proof for the following statement. \begin{thm}(\cite[Theorem 1.19.]{MTVV})\label{anotherproof} Let $G$ is a profinite group with $\mf{np}_k(G)>0$. Then $G$ has an open $k$-step nilpotent subgroup. \end{thm} \begin{proof} By \cite[Lemma 2.4.]{AS}, there is an open normal subgroup $N$ along with elements $x_1,\dotsc, x_{k+1}\in G$ such that $\mf{np}(N;x_1,\dotsc, x_{k+1})> 1-\frac{3}{2^{k+1}}$, so by Theorem \ref{3/2}, $N$ should be nilpotent of class at most $k$. \end{proof} If $B$ is a measurable subgroup of a compact group $A$, then $\mf m_A(B)=[A:B]^{-1}$, by convention on ``$ \frac1\infty=0 $''. We have \begin{lem}\label{A=(A/B)B} Let $G$ be a compact group, $H$ be a closed subgroup of $G$ and $N$ be a normal closed subgroup of $G$. Then \[ \mf m_G(H)=\mf m_N(N\cap H)\mf m_{G/N}\pr{NH/N}.\] \end{lem} \begin{proof} The following is well known \[ [G:H]=[G:NH][NH:H]=[G/N:NH/N][N:N\cap H].\] This along with the explanation before the lemma implies the result. \end{proof} \begin{thm}\label{increase wrt N} Let $G$ be a compact group with closed subgroups $H$ and closed normal subgroup $N$ contained in $H$. Then $\mf{np}_{k,G}(H)\leq\mf{np}_{k,G/N}(H/N)\mf{np}_{k,G}(N)$. \end{thm} \begin{proof} For $x_1,\dotsc,x_{k+1}\in G$ we have \begin{align*} \mf{np}(H;x_1,\dotsc,x_{k+1})&\leq\int_{H^k}\mf m_H(C_H([x_1y_1,\dotsc,x_ky_k]))\,\text d\mf y. \end{align*} By Lemmas \ref{Gotosubgroups} and \ref{A=(A/B)B}, \begin{footnotesize} \begin{align*} \int_{H^k}&\mf m_H(C_H([x_1y_1,\dotsc,x_ky_k]))\,\text d\mf y=\int_{H^k/N^k}\int_{N^k} \mf m_H(C_H([x_1y_1z_1,\dotsc,x_ky_kz_k]))\,\text d\mf z\,\text d\bar{\mf y}\\ &=\int_{H^k/N^k}\int_{N^k}\mf m_N(C_N([x_1y_1z_1,\dotsc,x_ky_kz_k]))\mf m_{G/N}\pr{NC_H([x_1y_1z_1,\dotsc,x_ky_kz_k])/N}\,\text d\mf z\,\text d\bar{\mf y}\\ &\overset{(*)}{\leq}\int_{H^k/N^k}\int_{N^k}\mf m_N(C_N([x_1y_1z_1,\dotsc,x_ky_kz_k]))\mf m_{H/N}\pr{C_{H/N}([\overline{x_1y_1},\dotsc,\overline{x_ky_k}])}\,\text d\mf z\,\text d\bar{\mf y}\\ &=\int_{H^k/N^k}\mf m_{H/N}\pr{C_{H/N}([\overline{x_1y_1},\dotsc,\overline{x_ky_k}])} \pr{\int_{N^k}\mf m_N(C_N([x_1y_1z_1,\dotsc,x_ky_kz_k]))\,\text d\mf z}\,\text d\bar{\mf y}\\ &=\int_{H^k/N^k}\mf m_{G/N}\pr{C_{G/N}([\overline{x_1y_1},\dotsc,\overline{x_ky_k}])} \mf{np}(N;x_1y_1,\dotsc,x_ky_k,1)\,\text d\bar{\mf y}\\ &\leq\mf{np}_{k,G}(N)\int_{(G/N)^k}\mf m_{G/N}\pr{C_{G/N}([\overline{x_1}\overline{y_1},\dotsc,\overline{x_k}\overline{y_k}])}\,\text d\bar{\mf y}\\ &\leq\mf{np}_{k,G}(N)\mf{np}_{k,G/N}(H/N). \end{align*} \end{footnotesize} The inequality (*) obtained by the inclusion \[NC_H([x_1y_1,\dotsc,x_ky_k])/N\subset C_{H/N}([\overline{x_1y_1},\dotsc,\overline{x_ky_k}]).\] \end{proof} As a corollary of the above theorem, we state: \begin{cor}\label{increase wrt Nv2} Let $N$ be a closed normal subgroup of a compact group $G$. Then $\mf{np}_k(G)\leq\mf{np}_k(G/N)\mf{np}_{k,G}(N)$. In particular, $\mf{cp}(G)\leq\mf{cp}(G/N)\mf{cp}(N)$. \end{cor} \section{Proofs of Main Results} Using Corollaries \ref{increase wrt Nv2} and \ref{3/2}, we prove the following. \begin{thm}\label{finitelength} Let $G$ be a compact group and assume that $\mf{np}_k(G)>0$. Then the length of all closed normal series of $G$ which its factors are not nilpotent of class at most $k$ should be bounded. \end{thm} \begin{proof} Let \[ 1=G_{r+1}\leq G_{r}\leq\dots\leq G_1\leq G_0=G\] be a closed normal series such that $G_i/G_{i+1}$, $0\leq i\leq r$, is not nilpotent of class at most $k$. Using the Theorem \ref{increase wrt N} repeatedly, we have \begin{align*} \mf{np}_k(G)&\leq\mf{np}_{k,G_0}(G_r)\mf{np}_k(G_0/G_r)\\ &\leq\mf{np}_{k,G_0}(G_r)\mf{np}_{k,G_0/G_r}(G_{r-1}/G_r)\mf{np}_k(G_0/G_{r-1})\\ &\vdots\\ &\leq\prod_{i=2}^{r+1}\mf{np}_{k,G_0/G_i}(G_{i-1}/G_i). \end{align*} The Corollary \ref{3/2} implies that the right hand side of the above inequality is at most $\pr{1-\frac{3}{2^{k+1}}}^r$. All in all, we get \[ r<\frac{\ln\mf{np}_k(G)}{\ln\pr{1-\frac{3}{2^{k+1}}}}.\] \end{proof} \begin{thm}\label{Decresingnet} Let $G$ be a compact group with $\mf{np}_k(G)>0$. Assume that there exists a decreasing net $(H_\alpha)$ of closed normal subgroups of $G$ such that $\bigcap_\alpha H_\alpha=1$. Then $H_\alpha$ is nilpotent of class at most $k$ for sufficiently large $\alpha$. \end{thm} \begin{proof} If none of $H_\alpha$'s are nilpotent of class at most $k$, then we can construct an infinite closed normal series of $G$ which its factors are not nilpotent of class at most $k$, violates Theorem \ref{finitelength}; indeed, suppose we have constructed a closed normal series, \begin{equation}\label{ons} G=H_0\geq H_{\alpha_1}\geq\dots\geq H_{\alpha_n} \end{equation} such that its factors are not nilpotent of class at most $k$. Since $H_{\alpha_n}$ is not $k$-step nilpotent, there exists $(x_1,\dotsc,x_{k+1})$ in $H_{\alpha_n}^{k+1}$ such that $[x_1,\dotsc,x_{k+1}]\neq1$. By the assumption, we can choose $H_{\alpha_{n+1}}$ such that $[x_1,\dotsc,x_{k+1}]\not\in H_{\alpha_{n+1}}$, so $H_{\alpha_{n+1}}$ can be added to (\ref{ons}). \end{proof} Since $\bigcap\set{N: N\,\text{is an open normal subgroup of } G}=1$, for a profinite group $G$, Theorem \ref{Decresingnet} gives another proof for the following well-known fact: \begin{cor}(\cite[Theorem 1.19.]{MTVV}) Let $G$ be a profinite group with $\mf{np}_k(G)>0$. Then $G$ has a $k$-step nilpotent open subgroup. \end{cor} Theorem \ref{finitelength} has also another implication: \begin{cor}\label{gotofinite} Let $(G_\alpha)_{\alpha}$ be a collection of compact groups. If $$\mathbf{np}_k\left(\prod_\alpha G_\alpha\right)>0,$$ then $G_\alpha$'s are nilpotent of class at most $k$ for all but finitely many $\alpha$. \end{cor} Now we're ready to show that the class $\mathsf{PA}_k$ contains connected compact groups. A little preparation is needed before proving the later. \begin{lem}\label{zariskiclosed} Let $G$ be a compact group. Then $G$ contains no nontrivial closed semi simple connected compact Lie subgroup $H$ with {\small $\mf{np}_{k,G}(H)>0$}. \end{lem} \begin{proof} Suppose $\mathcal N=\mathcal N(H;x_1,\dotsc,x_{k+1})$ has positive measure for some $x_1,\dotsc, x_{k+1}\in G$. Since $\mathcal N$ is Zariski-closed and has positive measure, it should be equal to $G$, so $H$ should be nilpotent by Proposition \ref{nocamn} and hence trivial by semi simplicity. \end{proof} \begin{thm}\label{CCG} Let $G$ be a connected compact group with $\mf{np}_k(G)>0$. Then $G$ is an abelian group. \end{thm} \begin{proof} Note first that simply connected compact groups are semi simple (\cite[Theorem 9.29.]{HM}). Also, Corollary \ref{increase wrt Nv2} implies that $\mf{np}_k(G/\mf Z)>0$, where $\mf Z$ stands for the center of $G$. Therefore by the proof of \cite[Theorem 9.24]{HM} and Corollary \ref{gotofinite}, there is a finite family $\{L_1,\dotsc,L_n\}$ of simple simply connected compact Lie groups such that $\frac{G}{\mf Z}\cong\prod_iL_i$, whence implies $\prod_i\mf{np}_k(L_i)>0$. Now the result follows from Lemma \ref{zariskiclosed}. \end{proof} Also compact Lie groups belong to $\mathsf{PA}_k$. Next we use $G^0$ to denote the connected component of a compact group $G$. \begin{lem}\label{CLG} Let $G$ be a compact Lie group with $\mf{np}_k(G)>0$. Then $G^0$ is torus. In particular $G$ has an open $k$-step nilpotent subgroup. \end{lem} \begin{proof} Denote by $\mf Z_0$ the center of $G^0$. By Corollary \ref{increase wrt Nv2} and Theorem \ref{increase wrt N}, we obtain \[ 0<\mf{np}_k(G)\leq\mf{np}_{k,G}(G^0)\leq\mf{np}_{k,G/\mf Z_0}(G^0/\mf Z_0).\] Now the result obtained with Corollary \ref{zariskiclosed}, because $G^0/\mf Z_0$ is a semi simple connected compact Lie subgroup of $G/\mf Z_0$.The last claim of Lemma follows from the fact that $G^0$ is a closed subgroup of $G$ of finite index (\cite[Theorem 6.36.]{HM}), so it should be open. \end{proof} \begin{rem}\label{YCor} The above lemma is attributed to Yves de Cornulier (\cite{Ycor}), which is given in response to a question posed by the second author (\cite{msol}). Of course, the above proof is somewhat different from the proof presented in \cite{Ycor}. \end{rem} The main result of the Lemma \ref{CLG} is true for arbitrary compact groups. Denote by $\mathfrak{N}(G)$ the set of all closed normal subgroups $N$ of $G$ such that $G/N$ is a Lie group. \begin{proof}[Proof of Theorem \ref{main}] By Corollary \ref{increase wrt Nv2}, $\mf{np}_k(G/N)>0$ for all $N\in\mathfrak{N}(G)$. Lemma \ref{CLG}, now implies that $(G/N)^0=G^0N/N$ (\cite[Lemma 9.18.]{HM}) is an open abelian group and so, $G^0N$ becomes open and normal in $G$. If none of $G^0N$'s is $k$-step nilpotent, $N\in\mathfrak{N}(G)$, then we can construct a closed normal series of $G$ whose factors are not nilpotent of class at most $k$, violates Theorem \ref{finitelength}; indeed, suppose we have constructed a closed normal series, \begin{equation}\label{ons} G\geq G^0N_1\geq\dots\geq G^0N_s \end{equation} such that its factors are not nilpotent of class at most $k$ and $N_1,\dotsc, N_s\in\mathfrak{N}(G)$. Since $G^0N_s$ is not $k$-step nilpotent, there exists $(x_1,\dotsc,x_{k+1})$ in $(G^0N_s)^{k+1}$ such that $[x_1,\dotsc,x_{k+1}]\neq1$. But, $[x_1,\dotsc,x_{k+1}]$ is equal 1 modulus $N_s$, for $G^0N_s/N_s$ is abelian. Therefore $[x_1,\dotsc,x_{k+1}]\in N_s$, and by the fact that $\bigcap\{N: N\in \mathfrak{N}(G)\}=1$ we obtain $N_{s+1}$ in $\mathfrak{N}(G)$, which does not contain $[x_1,\dotsc,x_{k+1}]$. \end{proof} The following question is proposed in \cite{AS0} in order to resolve Question \ref{Question}. \begin{qu}[Question 3.2 of \cite{AS0}] Are there finitely many words $w_{i,j}$ ($1 \leq i \leq n$, $1 \leq j \leq k + 1$) in the free group on $k + 1$ generators such that if $G$ is a compact group, $(x_1, \dots , x_{k+1}), \mf u = (u_1, \dots , u_{k+1}) \in G^{k+1}$ and $[x_1 w_{i,1}(\mf u), \dots , x_{k+1} w_{i,k+1}(\mf u)] = 1$ for all $i \in \{1, \dots , n\}$ then $[u_1, \dots , u_{k+1}] = 1$? \end{qu} To settle Question \ref{Question}, we will prove that using Theorem \ref{main}, we may study the following slightly simpler question on finite groups. \begin{qu}\label{qusnew} Are there finitely many words $w_{i,j}$ ($1 \leq i \leq n$, $1 \leq j \leq k + 1$) in the free group on $k + 1$ generators such that if $G$ is a finite group, $(x_1, \dots , x_{k+1}), \mf u = (u_1, \dots , u_{k+1}) \in G^{k+1}$, $\langle u_1, \dots , u_{k+1} \rangle$ is abelian-by-nilpotent and $[x_1 w_{i,1}(\mf u), \dots , x_{k+1} w_{i,k+1}(\mf u)] = 1$ for all $i \in \{1, \dots , n\}$ then $[u_1, \dots , u_{k+1}] = 1$? \end{qu} \begin{thm} If the answer of Question \ref{qusnew} is positive, the answer of Question \ref{Question} is also affirmative. \end{thm} \begin{proof} First note that by Theorem \ref{main} together with Hall Theorem, \cite[15.4.1]{Rob}, all finitely generated subgroups of $G$ are residually finite.\par Let $n$ be the positive integer in Question \ref{qusnew}. By \cite[Theorem 3.1]{AS0} there exists an open subset $V$ of $G$ containing $1$ such that \begin{align}\label{cap} \mathbf{m}\pr{\bigcap_{i=1}^n \mathcal{N}_k(G)\mf y_j}>0, \end{align} where $\mf y_i\in V^{k+1}$, $i=1,\dotsc,n$. By continuity of word maps, we can find an open neighborhood $U$ of $1$, such that $w_{i,j}(\mf u)\in V$ for $$\mf u=(u_1,\dotsc,u_{k+1})\in U^{k+1},$$ and $1 \leq i \leq n$ , $1 \leq j \leq k + 1$. Now, for $\mf u=(u_1,\dotsc,u_{k+1})\in U^{k+1}$, (\ref{cap}) ensures the existence of $x_1, \dots , x_{k+1}\in G$, such that \begin{align} \label{EQ} [x_1 w_{i,1}(\mf u), \dots , x_{k+1} w_{i,k+1}(\mf u)] = 1,\quad (1 \leq i \leq n). \end{align} But $\valu{x_i, u_i: i=1,\dotsc, k+1}$ is residually finite, so by passing to its finite quotients and the correctness of the Question \ref{qusnew}, we get from (\ref{EQ}) that $[u_1, \dots , u_{k+1}] = 1$, whence the open subgroup $\valu{U}$ is a $k$-step open subgroup of $G$. \end{proof}
{ "timestamp": "2022-08-10T02:12:58", "yymm": "2208", "arxiv_id": "2208.04666", "language": "en", "url": "https://arxiv.org/abs/2208.04666", "abstract": "Let $k$ be any positive integer and $G$ a compact (Hausdorff) group. Let $\\mf{np}_k(G)$ denote the probability that $k+1$ randomly chosen elements $x_1,\\dots,x_{k+1}$ satisfy $[x_1,x_2,\\dots,x_{k+1}]=1$. We study the following problem: If $\\mf{np}_k(G)>0$ then, does there exist an open nilpotent subgroup of class at most $k$? The answer is positive for profinite groups and we give a new proof. We also prove that the connected component $G^0$ of $G$ is abelian and there exists a closed normal nilpotent subgroup $N$ of class at most $k$ such that $G^0N$ is open in $G$.", "subjects": "Group Theory (math.GR)", "title": "Nilpotent probability of compact groups", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.988491848562538, "lm_q2_score": 0.8267117983401363, "lm_q1q2_score": 0.8171978737697015 }