% vim: tw=50 % 27/10/2022 11AM \myskip Can this theorem be generalised to any metric space? Obviously not: for example in $\RR \setminus \{0\}$ with usual metric, the set $[-1, 0) \cup (0, 1]$ is closed and bounded but the sequence $\left( \frac{1}{n} \right)_{n \ge 1}$ has no convergent subsequence. \myskip Problem: space not complete. Maybe complete + bounded could imply sequentially compact? \\ Even this doesn't work. Recall example from section 1: let \[ X = \{f \in B(\RR) \mid \sup_{x \in \RR} |f(x)| \le 1\} \] with uniform metric. Then $X$ is complete (closed subset of complete space $B(\RR)$) and bounded (if $f, g \in X$, $d(f, g) \le 2$). But consider \[ f_n(x) = \begin{cases} 1 & x = n \\ 0 & x \neq n \end{cases} \] Then $(f_n)$ is a sequence in $X$ but $\forall m, n$, $m \neq n$ implies $d(f_m, f_n) = 1$. So $(f_n)$ cannot have a convergent subsequence. So the problem is that $X$ is `too big'. So we need a stronger concept of boundedness. \begin{definition*} Let $(X, d)$ be a metric space. We say $X$ is \emph{totally bounded} if for all $\delta > 0$ we can find a \emph{finite} set $A \subset X$ such that $\forall x \in X$ $\exists a \in A$ with $d(x, a) < \delta$. \end{definition*} \begin{theorem} A metric space is sequentially compact if and only if it is complete and totally bounded. \end{theorem} \begin{proof} \begin{enumerate} \item[$\Leftarrow$] Suppose the metric space $(X, d)$ is complete and totally bounded. Let $(x_n)_{n \ge 1}$ be a sequence in $X$. \myskip As $X$ is totally bounded, can find finite $A_1 \subset X$ such that $\forall x \in X$ there exists $a \in A_1$ with $d(x, a) < 1$. In particular there is an infinite set $N_1 \subset \NN$ and a point $a_1 \in A_1$ such that $\forall n \in N_1$, $d(x_N, a_1) < 1$. Hence $\forall m, n \in N_1$, $d(x_m, x_n) < 2$. Similarly, we can find finite $A_2 \subset X$ such that $\forall x \in X$, $\exists a \in A_2$, $d(x, a) < \half$. In particular, there is an infinite $N_2 \subset N_1$ such that $\forall n \in N_2$, $d(x_n, a_2) < \half$ and thus $\forall m, n \in N_2$, $d(x_m, x_n) < 1$. \myskip Keep going. We get a sequence $N_1 \supset N_2 \supset N_3 \supset \cdots$ of infinite subsets of $\NN$ such that $\forall i$, $\forall m, n$, $m, n \in \NN_i \implies d(x_m, x_n) < \frac{2}{i}$. \myskip Now pick $n_1 \in N_2$. Then pick $n_2 \in N_2$ with $n_2 > n_1$. Then pick $n_3 \in N_3$ with $n_3 > n_2$ and so on. We obtain a subsequence $(x_{n_j})$ of $(x_n)$ such that for all $j$, $x_{n_j} \in N_j$. Thus if $i \le j$ then $x_{n_i}, x_{n_j} \in N_i$ and so \[ d(x_{n_i}, x_{n_j}) < \frac{2}{i} \] Hence $(x_{n_j})$ is a Cauchy sequence and hence, by completeness, converges. So $X$ is sequentially compact. \item[$\Rightarrow$] Suppose $X$ is not complete. Then $X$ has a Cauchy sequence $(x_n)$ which doesn't converge. Suppose we have a convergent subsequence, say $x_{n_j} \to x$. Then $x_n \to x$ (exercise). Contradiction. \myskip Suppose instead $X$ is not totally bounded. Then there is some $\delta > 0$ such that whenever $A \subset X$ is finite, there exists $x \in X$ such that $\forall a \in A$, $d(x, a) \ge \delta$. So pick $x_1 \in X$. Pick $x_2 \in X$ such that $d(x_1, x_2) \ge \delta$. Pick $x_3 \in X$ such that $d(x_1, x_3) \ge \delta$ and $d(x_2, x_3) \ge \delta$. Continue. Then we get a sequence $(x_n)$ in $X$ such that for all $i, j$ with $i \neq j$, $d(x_i, x_j) \ge \delta$. Then $(x_n)$ has no convergent subsequence. \end{enumerate} \end{proof} \myskip Exercise: A continuous function on a sequentially compact metric space is uniformly continuous. If the function is real-valued then it's bounded and attains its bounds. \subsection{The Topology of Metric Spaces} Theme of section 2: to generalise convergence / continuity, all we need is a distance. \myskip But, for example in $\RR^n$ we have the very different concepts of distance given by the Euclidean, $l_1$ and $l_\infty$ metrics. But all give same concept of convergence and continuity. \begin{definition*} Let $(X, d)$ and $(Y, e)$ be metric spaces. Let $f : X \to Y$. We say $f$ is a \emph{homeomorphism} and that $X, Y$ are \emph{homeomorphic} if $f$ is a continuous bijection with continuous inverse. \end{definition*} \begin{remark*} Homeomorphism is an equivalence `relation'. \end{remark*} \subsubsection*{Examples} \begin{enumerate}[(1)] \item If $x, y \in \RR^n$: \[ d_\infty(x, y) \le d_1(x, y) \le nd_\infty(x, y) \] So identity map $\RR^n \to \RR^n$ is continuous as map $(\RR^n, d_1) \to (\RR^n, d_\infty)$ and inverse map $(\RR^n, d_\infty) \to (\RR^n, d_1)$ is continuous. So it's a homeomorphism. Similarly, $\RR^n$ with Euclidean metric is homeomorphic to both these spaces. \item Same argument would show: if $(X, d)$ and $(Y, e)$ are metric spaces and $f : X \to Y$ is a bijection satisfying: \begin{enumerate}[(i)] \item $\exists A ~ \forall x, y \in X$, $e(f(x), f(y)) \le A d(x, y)$. \item and $\exists B ~\forall x, y \in X$, $d(x, y) \le B e(f(x), f(y))$. \end{enumerate} Then $f, f^{-1}$ are continuous so $X, Y$ are homeomorphic. \item Define $f : \left( -\frac{\pi}{2}, \frac{\pi}{2} \right) \to \RR$ by $f(x) = \tan x$. Then $f$ is a homeomorphism (usual metric in each case). But there is no constant $A$ such that \[ \forall x, y \in \left( -\frac{\pi}{2}, \frac{\pi}{2} \right) |\tan x - \tan y| \le A|x - y| \] \end{enumerate} \begin{proposition} Let $(V, b)$, $(W, c)$, $(X, d)$, $(Y, e)$ be metric spaces and $f : X \to V$, $g : Y \to W$ be homeomorphisms. Then \begin{enumerate}[(i)] \item In $X$, $x_n \to x$ if and only if in $V$, $f(x_n) \to f(x)$; \item and a function $g : X \to Y$ is continuous at $a \in X$ if and only if $g \circ h \circ f^{-1}$ is continuous at $f(a) \in V$. \end{enumerate} \end{proposition} \begin{proof} \begin{enumerate}[(i)] \item $x_n \to x$ implies $f(x_n) \to f(x)$ as $f$ is continuous. $f(x_n) \to f(x) \implies x_n = f^{-1}(f(x_n)) \to f^{-1}(f(x)) = x$ as $f^{-1}$ is continuous. \item $h$ continuous implies $g \circ h \circ f^{-1}$ because composition of continuous functions is continuous. \\ And $g \circ h \circ f^{-1}$ continuous implies $h = g^{-1} \circ (g \circ h \circ f^{-1}) \circ f$ is continuous because composition of continuous functions is continuous. \end{enumerate} \end{proof} \noindent We now have examples of metric spaces that look very different but behave identically with respect to convergence / continuity. \\ Thought: Could we disperse with distance altogether? \\ Another way to think about continuity: \begin{definition*} Let $(X, d)$ be a metric space let $a \in X$ and let $\eps > 0$. The \emph{open ball of radius $\eps$ about $a$} is the set \[ B_\eps(a) = \{x \in X \mid d(x, a) < \eps\} \] \end{definition*} \begin{remark*} Suppose $f : X \to Y$, $a \in X$. $d$ metric on $X$, $e$ metric on $Y$. Then \begin{align*} \text{$f$ continuous at $a$} &\iff \forall \eps > 0 ~ \exists \delta > 0 \quad d(x, a) < \delta \implies d(f(x), f(a)) < \eps \\ &\iff \forall \eps > 0~\exists \delta > 0 \quad x \in B_\delta(a) \implies f(x) \in B_\eps(f(a)) \\ &\iff \forall \eps > 0 ~ \delta > 0 \quad f(B_\delta(a)) \subset B_\eps(f(a)) \\ &\iff \forall \eps > 0 ~ \exists \delta > 0 \quad B_\delta(a) \subset f^{-1}(B_\eps(f(a)) \end{align*} So we have redefined continuity in terms of open balls. But open balls have radii, so this still uses a notion of distance. \end{remark*} \begin{definition*} Let $X$ be a metric space. A subset $G \subset X$ is \emph{open} if $\forall x \in G$, $\exists \eps > 0$ such that $B_\eps(x) \subset G$. \\ A subset $N \subset X$ is a \emph{neighbourhood} (\emph{nbd}) of a point $a \in X$ if there exists an open set $G \subset X$ such that $a \in G \subset N$. \end{definition*}