Complete Metrics

In real analysis, we have a rather useful property of \mathbb R.

Theorem 1. Let \{x_n\} be an \mathbb R-sequence. Then \{x_n\} converges in the usual topology if and only if \{x_n\} is Cauchy.

In this case, we call \mathbb R a complete metric space. What other metric spaces are complete?

Definition 1. We say that K is a complete if every Cauchy K-sequence converges in K.

In terms of the metric d, recall that \{x_n\} is Cauchy if for any \epsilon > 0, there exists N \in \mathbb N such that for m \geq n \geq N, d(x_m, x_n) < \epsilon.

Lemma 1. If K is compact, then any K-sequence contains a convergent subsequence (i.e. K is sequentially compact).

Proof. For the direction (\Rightarrow), suppose K is compact. Let \{x_n\} be any K-sequence. If \{x_n\} is finite, then there exists some x \in K such that x_k = x for infinitely many indices k. Thus, the subsequence \{x\} trivially converges to x.

Suppose instead X := \{x_n : n \in \mathbb N\} is infinite and non-repeating. We claim that there exists x \in K such that for any k \in \mathbb N, B_d(x,1/k)\backslash \{x\} \cap X \neq \emptyset. In this instance, choose x_{n_k} \in B_d(x, 1/k) so that x_{n_k} \to x.

Suppose for a contradiction that for any x \in K, there exists \epsilon_x > 0 such that B_d(x, \epsilon_x) \backslash \{x\} \cap X = \emptyset. We first claim that X is closed. For any x \in K \backslash X, find \epsilon_x > 0 such that

B_d(x, \epsilon_x) \backslash \{x\} \cap X = \emptyset.

In particular, B_d(x, \epsilon_x) \cap X = \emptyset so that x \in B_d(x, \epsilon_x) \subseteq K \backslash X. Therefore, K \backslash X is open and X is closed.

By hypothesis, for any n \in \mathbb N, there exists k_n \in \mathbb N such that B_d(x_n, 1/k_n) \backslash \{x_n\} \cap X = \emptyset. Defining V_0 := K \backslash X and V_n := B_d(x_n, 1/k_n), the collection \{V_0\} \cup \{V_n : n \in \mathbb N\} forms an open cover for K. Since K is compact, there exists some N \in \mathbb N such that without loss of generality, \{V_0,V_1,\dots,V_N\} covers K.

Hence, there exists i = 1, \dots, N such that x_{N+1} \in B_d(x_i, 1/k_i) \backslash \{x_i\}. Therefore, B_d(x_i, 1/k_i) \backslash \{x_i\} \cap X \neq \emptyset, a a contradiction, as required.

For the direction (\Leftarrow), suppose K is sequentially compact. Let \{V_\alpha\} be any open cover of K.

Lemma 2. If K is sequentially compact, then it is complete.

Proof. Let \{x_n\} be any Cauchy sequence in K. Since K is sequentially compact, \{x_n\} contains a convergent subsequence \{x_{n_k}\} such that x_{n_k} \to x. We claim that x_n \to x.

Fix \epsilon > 0. Since \{x_{n_k}\} converges, for any k_1 > 0, there exists N_1 \in \mathbb N such that for k \geq N_1, n_k \geq N_1 and d(x_{n_k}, x) < k_1 \cdot \epsilon. Since \{x_n\} is Cauchy, for any k_2 > 0, there exists N_2 \in \mathbb N such that for m \geq n \geq N_2, d(x_m, x_n) < k_2 \cdot \epsilon. Therefore, for n > \max \{N_1,N_2\} =: N_3,

d(x_n,x) \leq d(x_n,x_{n_{N_3}}) + d(x_{n_{N_3}},x) < (k_1+k_2)\cdot \epsilon.

Setting k_1 = k_2 = 1/2 yields the desired result.

Theorem 2. If K is compact, then it is complete.

Proof. Since K is compact, it is sequentially compact by Lemma 1. Thus, K is complete by Lemma 2.

Theorem 3. Suppose K is a compact metric space. Let \mathcal C(K) \subseteq \mathcal F(K, \mathbb R) denote the collection of continuous real-valued functions equipped with the supremum norm and supremum metric respectively:

\displaystyle \| f \|_\infty := \sup_{x \in K} |f(x)|,\quad d_\infty(f,g) := \|f - g \|_\infty.

Then (\mathcal C(K), d_\infty) is a complete metric space.

Proof. We note that convergence in (\mathcal C(K), d_\infty) corresponds to convergence in the supremum norm. Thus, f_n \to f means that f_n converges uniformly to f. Let \{f_n\} be a Cauchy sequence. This implies that for each x \in K,

|f_m(x) - f_n(x)| = |(f_m - f_n)(x)| \leq \| f_m - f_n \|_\infty = d_\infty(f_m,f_n).

Thus, \{f_n(x)\} is Cauchy and converges in \mathbb R to some \{ f(x) \}. We claim that f_n \to f uniformly. Fix \epsilon > 0. Find N \in \mathbb N such that for any m, n \geq N, for any x \in K,

\displaystyle |f_m(x) - f_n(x)| < \frac{ \epsilon }2.

Since | \cdot | is continuous in the usual topology on \mathbb R, taking m \to \infty,

\displaystyle |f(x) - f_n(x)| = \lim_{m \to \infty}|f_m(x) - f_n(x)| \leq \lim_{m \to \infty} \frac{ \epsilon }2 = \frac{\epsilon}2.

Since this inequality holds for all x \in K, d_\infty (f_n,f) \leq \epsilon/2 < \epsilon. Thus, f_n \to f uniformly, as required.

Remarkably, however, \mathcal C(K) is not always compact.

Example 1. Define the sequence \{f_n\} \subseteq \mathcal C([0,1]) by f_n(x) = nx. Then \{f_n(x)\} diverges at every x \in [0, 1] pointwise, so that \{f_n\} does not have any convergent subsequence.

When then, is a subset of functions \mathcal F \subseteq \mathcal C(K) compact? The answer to this question is the famous Arzelà-Ascoli theorem, and we will explore this theorem in the future. For now, we want to use completeness to deduce Banach’s fixed-point theorem, which helps us establish the existence of solutions to first-order differential equations.

—Joel Kindiak, 1 Apr 25, 2008H

,

Published by


Leave a comment