Datasets:
Tasks:
Text Generation
Modalities:
Text
Sub-tasks:
language-modeling
Languages:
English
Size:
100K - 1M
License:
\chapter{Compactness} | |
One of the most important notions of topological spaces is that of \emph{compactness}. | |
It generalizes the notion of ``closed and bounded'' in Euclidean space | |
to any topological space | |
(e.g.\ see \Cref{thm:bzw}). | |
For metric spaces, there are two equivalent ways of formulating compactness: | |
\begin{itemize} | |
\ii A ``natural'' definition using \emph{sequences}, called sequential compactness. | |
\ii A less natural definition using open covers. | |
\end{itemize} | |
As I alluded to earlier, sequences in metric spaces are super nice, | |
but sequences in general topological spaces \emph{suck} (to the point where | |
I didn't bother to define convergence of general sequences). | |
So it's the second definition that will be used for general spaces. | |
\section{Definition of sequential compactness} | |
\prototype{$[0,1]$ is compact, but $(0,1)$ is not.} | |
To emphasize, compactness is one of the | |
\emph{best} possible properties that a metric space can have. | |
\begin{definition} | |
A \vocab{subsequence} of an infinite sequence | |
$x_1, x_2, \dots$ is exactly what it sounds like: | |
a sequence $x_{i_1}, x_{i_2}, \dots$ | |
where $i_1 < i_2 < \cdots$ are positive integers. | |
Note that the sequence is required to be infinite. | |
\end{definition} | |
Another way to think about this is ``selecting infinitely many terms'' | |
or ``deleting some terms'' of the sequence, depending on whether | |
your glass is half empty or half full. | |
\begin{definition} | |
A metric space $M$ is \vocab{sequentially compact} if | |
every sequence has a subsequence which converges. | |
\end{definition} | |
This time, let me give some non-examples before the examples. | |
\begin{example} | |
[Non-examples of compact metric spaces] | |
\listhack | |
\begin{enumerate}[(a)] | |
\ii The space $\RR$ is not compact: consider the sequence $1,2,3,4,\dots$. | |
Any subsequence explodes, hence $\RR$ cannot possibly be compact. | |
\ii More generally, if a space is | |
not bounded it cannot be compact. | |
(You can prove this if you want.) | |
\ii The open interval $(0,1)$ is bounded but not compact: | |
consider the sequence $\frac12, \frac13, \frac14, \dots$. | |
No subsequence can converge to a point in $(0,1)$ because the sequence ``converges to $0$''. | |
\ii More generally, any space which is not complete cannot be compact. | |
\end{enumerate} | |
\end{example} | |
Now for the examples! | |
\begin{ques} | |
Show that a finite set is compact. | |
(Pigeonhole Principle.) | |
\end{ques} | |
\begin{example}[Examples of compact spaces] | |
Here are some more examples of compact spaces. | |
I'll prove they're compact in just a moment; | |
for now just convince yourself they are. | |
\begin{enumerate}[(a)] | |
\ii $[0,1]$ is compact. Convince yourself of this! | |
Imagine having a large number of dots in the unit interval\dots | |
\ii The surface of a sphere, $S^2 = \left\{ (x,y,z) \mid x^2+y^2+z^2=1 \right\}$ is compact. | |
\ii The unit ball $B^2 = \left\{ (x,y) \mid x^2+y^2 \le 1 \right\}$ is compact. | |
\ii The \vocab{Hawaiian earring} living in $\RR^2$ is compact: | |
it consists of mutually tangent circles of radius $\frac 1n$ for each $n$, | |
as in \Cref{fig:hawaiian}. | |
\end{enumerate} | |
\end{example} | |
\begin{figure}[ht] | |
\centering | |
\begin{asy} | |
size(4cm); | |
for (int n=1; n<=15; ++n) draw(CP(dir(0)/n, origin)); | |
fill(CP(dir(0)/15, origin), black); | |
\end{asy} | |
\caption{Hawaiian Earring.} | |
\label{fig:hawaiian} | |
\end{figure} | |
To aid in generating more examples, we remark: | |
\begin{proposition}[Closed subsets of compacts] | |
Closed subsets of sequentially compact sets are compact. | |
\end{proposition} | |
\begin{ques} | |
Prove this. (It should follow easily from definitions.) | |
\end{ques} | |
We need to do a bit more work for these examples, which we do in the next section. | |
\section{Criteria for compactness} | |
%Quick note: right now I've only defined compactness for metric spaces. | |
%In the next section I'll define compactness for general spaces, but | |
%all the results in this section will still remain true. | |
%However the proofs become much harder (in particular, \Cref{thm:tychonoff} | |
%becomes notoriously difficult). | |
%So you should assume all spaces in this section are metric spaces. | |
\begin{theorem} | |
[Tychonoff's theorem] | |
\label{thm:tychonoff} | |
If $X$ and $Y$ are compact spaces, then so is $X \times Y$. | |
\end{theorem} | |
\begin{proof} | |
\Cref{prob:tychonoff}. | |
\end{proof} | |
We also have: | |
\begin{theorem}[The interval is compact] | |
\label{thm:interval_compact} | |
$[0,1]$ is compact. | |
\end{theorem} | |
\begin{proof} | |
Killed by \Cref{thm:bzw}; | |
however, here is a sketch of a direct proof. | |
Split $[0,1]$ into $[0,\half] \cup [\half,1]$. | |
By Pigeonhole, infinitely many terms of the sequence lie in the left half (say); | |
let $x_1$ be the first one and then keep only the terms in the left half after $x_1$. | |
Now split $[0, \half]$ into $[0,\frac14] \cup [\frac14,\half]$. | |
Again, by Pigeonhole, infinitely many terms fall in some half; pick one of them, call it $x_2$. | |
Rinse and repeat. | |
In this way we generate a sequence $x_1$, $x_2$, \dots which is Cauchy, | |
implying that it converges since $[0,1]$ is complete. | |
\end{proof} | |
Now we can prove the main theorem about Euclidean space: | |
in $\RR^n$, compactness is equivalent to being ``closed and bounded''. | |
\begin{theorem}[Bolzano-Weierstra\ss] | |
A subset of $\RR^n$ is compact if and only if it is closed and bounded. | |
\label{thm:fakeBW} | |
\end{theorem} | |
\begin{ques} | |
Why does this imply the spaces in our examples are compact? | |
\end{ques} | |
\begin{proof} | |
Well, look at a closed and bounded $S \subseteq \RR^n$. | |
Since it's bounded, it lives inside some box $[a_1, b_1] \times [a_2, b_2] \times \dots \times [a_n, b_n]$. | |
By Tychonoff's theorem, since each $[a_i, b_i]$ is compact the entire box is. | |
Since $S$ is a closed subset of this compact box, we're done. | |
\end{proof} | |
One really has to work in $\RR^n$ for this to be true! | |
In other spaces, this criterion can easily fail. | |
\begin{example}[Closed and bounded but not compact] | |
Let $S = \{s_1, s_2, \dots\}$ be any infinite set equipped with the discrete metric. | |
Then $S$ is closed (since all convergent sequences are constant sequences) | |
and $S$ is bounded (all points are a distance $1$ from each other) | |
but it's certainly not compact since the sequence $s_1, s_2, \dots$ doesn't converge. | |
\end{example} | |
The Bolzano-Weierstrass theorem, which is \Cref{thm:bzw}, tells you exactly | |
which sets are compact in metric spaces in a geometric way. | |
\section{Compactness using open covers} | |
\prototype{$[0,1]$ is compact.} | |
There's a second related notion of compactness which I'll now define. | |
The following definitions might appear very unmotivated, but bear with me. | |
\begin{definition} | |
An open cover of a topological space $X$ | |
is a collection of open sets $\{U_\alpha\}$ | |
(possibly infinite or uncountable) which \emph{cover} it: | |
every point in $X$ lies in at least one of the $U_\alpha$, | |
so that \[ X = \bigcup U_\alpha. \] | |
Such a cover is called an \vocab{open cover}. | |
A \vocab{subcover} is exactly what it sounds like: | |
it takes only some of the $U_\alpha$, | |
while ensuring that $X$ remains covered. | |
\end{definition} | |
Some art: | |
\begin{center} | |
\begin{asy} | |
size(12cm); | |
path blob = (-8,3)..(-10,1)..(-9.4,0)..(-8.2,-3)..(-8,-4) | |
..(-2,-4.3)..(2,-4.2)..(8,-4) | |
..(8.6,-2)..(8.2,0.5)..(8,3) | |
..(4,3.3)..(0,3.1)..(-6,2.9)..cycle; | |
void open_ball(pair O, real r, pen p) { | |
dot(O, p); | |
filldraw(CR(O, r), p+opacity(0.1), p+dashed); | |
} | |
filldraw(blob, cyan+opacity(0.2), blue+0.7); | |
open_ball((-7,0.8), 3.5, red); | |
open_ball((-4.7,-1.2), 4.6, orange); | |
open_ball((-1.7,-0.3), 3.9, red); | |
open_ball((1.4,0.9), 2.9, yellow); | |
open_ball((2.5,-0.9), 3.7, orange); | |
open_ball((0.4,-3.7), 1.3, yellow); | |
open_ball((4.3,0.8), 2.7, red); | |
open_ball((6.3,1.8), 2.4, orange); | |
open_ball((6.0,-2.4), 4.0, yellow); | |
label("$X$", (8,3), dir(5), blue); | |
label(scale(2)*"$X = \bigcup_\alpha U_\alpha$", (0,4), blue); | |
\end{asy} | |
\end{center} | |
\begin{definition} | |
A topological space $X$ is \vocab{quasicompact} | |
if \emph{every} open cover has a finite subcover. | |
It is \vocab{compact} if it is also Hausdorff. | |
\end{definition} | |
\begin{remark} | |
The ``Hausdorff'' hypothesis that I snuck in | |
is a sanity condition which is not worth worrying about unless you're | |
working on the algebraic geometry chapters, | |
since all the spaces you will deal with are Hausdorff. | |
(In fact, some authors don't even bother to include it.) | |
For example all metric spaces are Hausdorff | |
and thus this condition can be safely ignored | |
if you are working with metric spaces. | |
\end{remark} | |
What does this mean? Here's an example: | |
\begin{example}[Example of a finite subcover] | |
Suppose we cover the unit square $M = [0,1]^2$ by | |
putting an open disk of diameter $1$ centered at every point | |
(trimming any overflow). | |
This is clearly an open cover because, | |
well, every point lies in \emph{many} of the open sets, | |
and in particular is the center of one. | |
But this is way overkill -- we only need about four | |
of these circles to cover the whole square. | |
That's what is meant by a ``finite subcover''. | |
\begin{center} | |
\begin{asy} | |
size(4cm); | |
draw(shift( (-0.5,-0.5) )*unitsquare, black+1); | |
real d = 0.4; | |
real r = 0.5; | |
draw(CR(dir( 45)*d, r), dotted); | |
draw(CR(dir(135)*d, r), dotted); | |
draw(CR(dir(225)*d, r), dotted); | |
draw(CR(dir(315)*d, r), dotted); | |
\end{asy} | |
\end{center} | |
\end{example} | |
Why do we care? | |
Because of this: | |
\begin{theorem}[Sequentially compact $\iff$ compact] | |
A metric space $M$ is sequentially compact if and only if it is compact. | |
\label{thm:compactness_metric} | |
\end{theorem} | |
We defer the proof to the last section. | |
This gives us the motivation we wanted for our definition. | |
Sequential compactness was a condition that made sense. | |
The open-cover definition looked strange, | |
but it turned out to be equivalent. | |
But we now prefer it, because we have seen that | |
whenever possible we want to resort to open-set-only based definitions: | |
so that e.g.\ they are preserved under homeomorphism. | |
\begin{example}[An example of non-compactness] | |
The space $X = [0,1)$ is not compact in either sense. % chktex 9 | |
We can already see it is not sequentially compact, because it is not even complete (look at $x_n = 1 - \frac 1n$). | |
To see it is not compact under the covering definition, consider the sets | |
\[ U_m = \left[0, 1 - \frac{1}{m+1} \right) \] % chktex 9 | |
for $m = 1, 2, \dots$. Then $X = \bigcup U_i$; hence the $U_i$ are indeed a cover. | |
But no finite collection of the $U_i$'s will cover $X$. | |
\end{example} | |
\begin{ques} | |
Convince yourself that $[0,1]$ \emph{is} compact; | |
this is a little less intuitive than it being sequentially compact. | |
\end{ques} | |
\begin{abuse} | |
Thus, we'll never call a metric space ``sequentially compact'' again | |
--- we'll just say ``compact''. | |
(Indeed, I kind of already did this in the previous few sections.) | |
\end{abuse} | |
\section{Applications of compactness} | |
Compactness lets us reduce \emph{infinite} open covers to finite ones. | |
Actually, it lets us do this even if the open covers are \emph{blithely stupid}. | |
Very often one takes an open cover consisting | |
of an open neighborhood of $x \in X$ for every single point $x$ in the space; | |
this is a huge number of open sets, | |
and yet compactness lets us reduce to a finite set. | |
To give an example of a typical usage: | |
\begin{proposition}[Compact $\implies$ totally bounded] | |
Let $M$ be compact. Then $M$ is totally bounded. | |
\end{proposition} | |
\begin{proof}[Proof using covers] | |
For every point $p \in M$, take an $\eps$-neighborhood of $p$, say $U_p$. | |
These cover $M$ for the horrendously stupid reason that each point $p$ is | |
at the very least covered by its open neighborhood $U_p$. | |
Compactness then lets us take a finite subcover. | |
\end{proof} | |
Next, an important result about maps between compact spaces. | |
\begin{theorem}[Images of compacts are compact] | |
Let $f \colon X \to Y$ be a continuous function, where $X$ is compact. | |
Then the image \[ f\im(X) \subseteq Y \] is compact. | |
\end{theorem} | |
\begin{proof}[Proof using covers] | |
Take any open cover $\{V_\alpha\}$ in $Y$ of $f\im(X)$. | |
By continuity of $f$, | |
it pulls back to an open cover $\{U_\alpha\}$ of $X$. | |
Thus some finite subcover of this covers $X$. | |
The corresponding $V$'s cover $f\im(X)$. | |
\end{proof} | |
\begin{ques} | |
Give another proof using the sequential definitions | |
of continuity and compactness. | |
(This is even easier.) | |
\end{ques} | |
Some nice corollaries of this: | |
\begin{corollary} | |
[Extreme value theorem] | |
Let $X$ be compact and consider a continuous function $f : X \to \RR$. | |
Then $f$ achieves a \emph{maximum value} at some point, | |
i.e.\ there is a point $p \in X$ such that $f(p) \ge f(q)$ for any | |
other $q \in X$. | |
\end{corollary} | |
\begin{corollary} | |
[Intermediate value theorem] | |
Consider a continuous function $f: [0,1] \to \RR$. | |
Then the image of $f$ is of the form $[a,b]$ for some real numbers $a \le b$. | |
\end{corollary} | |
\begin{proof}[Sketch of Proof] | |
The point is that the image of $f$ is compact in $\RR$, | |
and hence closed and bounded. | |
You can convince yourself that the closed sets are just unions of closed intervals. | |
That implies the extreme value theorem. | |
When $X=[0,1]$, the image is also connected, | |
so there should only be one closed interval in $f\im([0,1])$. | |
Since the image is bounded, we then know it's of the form $[a,b]$. | |
(To give a full proof, you would use the so-called \emph{least upper bound} | |
property, but that's a little involved for a bedtime story; | |
also, I think $\RR$ is boring.) | |
\end{proof} | |
\begin{example} | |
[$1/x$] | |
The compactness hypothesis is really important here. | |
Otherwise, consider the function | |
\[ (0,1) \to \RR \quad \text{ by } \quad | |
x \mapsto \frac 1x. \] | |
This function (which you plot as a hyperbola) is not bounded; | |
essentially, you can see graphically that the issue | |
is we can't extend it to a function on $[0,1]$ because it explodes near $x=0$. | |
\end{example} | |
One last application: if $M$ is a compact metric space, | |
then continuous functions $f : M \to N$ | |
are continuous in an especially ``nice'' way: | |
\begin{definition} | |
A function $f : M \to N$ of metric spaces | |
is called \vocab{uniformly continuous} | |
if for any $\eps > 0$, there exists a $\delta > 0$ | |
(depending only on $\eps$) such that | |
whenever $d_M(x,y) < \delta$ we also have $d_N(fx, fy) < \eps$. | |
\end{definition} | |
The name means that for $\eps > 0$, | |
we need a $\delta$ that works for \emph{every point} of $M$. | |
\begin{example}[Uniform continuity] | |
\listhack | |
\begin{enumerate}[(a)] | |
\ii The functions $\RR$ to $\RR$ of the form | |
$x \mapsto ax+b$ are all uniformly continuous, | |
since one can always take $\delta = \eps/|a|$ (or $\delta=1$ if $a=0$). | |
\ii Actually, it is true that a differentiable function $\RR \to \RR$ | |
with a bounded derivative is uniformly continuous. | |
(The converse is false for the reason that uniformly continuous | |
doesn't imply differentiable at all.) | |
\ii The function $f : \RR \to \RR$ by $x \mapsto x^2$ | |
is \emph{not} uniformly continuous, since for large $x$, | |
tiny $\delta$ changes to $x$ lead to fairly large changes in $x^2$. | |
(If you like, you can try to prove this formally now.) | |
Think $f(2017.01) - f(2017) > 40$; | |
even when $\delta = 0.01$, one can still cause large changes in $f$. | |
\ii However, when restricted to $(0,1)$ or $[0,1]$ | |
the function $x \mapsto x^2$ becomes uniformly continuous. | |
(For $\eps > 0$ one can now pick for example $\delta = \min\{1,\eps\}/3$.) | |
\ii The function $(0,1) \to \RR$ by $x \mapsto 1/x$ is \emph{not} | |
uniformly continuous (same reason as before). | |
\end{enumerate} | |
\end{example} | |
Now, as promised: | |
\begin{proposition}[Continuous on compact $\implies$ uniformly continuous] | |
If $M$ is compact and $f \colon M \to N$ is continuous, | |
then $f$ is uniformly continuous. | |
\end{proposition} | |
\begin{proof}[Proof using sequences] | |
Fix $\eps > 0$, and assume for contradiction that for every $\delta = 1/k$ | |
there exists points $x_k$ and $y_k$ within $\delta$ of each | |
other but with images $\eps > 0$ apart. | |
By compactness, take a convergent subsequence $x_{i_k} \to p$. | |
Then $y_{i_k} \to p$ as well, since the $x_k$'s and $y_k$'s are close to each other. | |
So both sequences $f(x_{i_k})$ and $f(y_{i_k})$ should converge to $f(p)$ by sequential continuity, | |
but this can't be true since the two sequences are always $\eps$ apart. | |
\end{proof} | |
\section{(Optional) Equivalence of formulations of compactness} | |
We will prove that: | |
\begin{theorem} | |
[Heine-Borel for general metric spaces] | |
For a metric space $M$, the following are equivalent: | |
\begin{enumerate}[(i)] | |
\ii Every sequence has a convergent subsequence, | |
\ii The space $M$ is complete and totally bounded, and | |
\ii Every open cover has a finite subcover. | |
\end{enumerate} | |
\end{theorem} | |
We leave the proof that (i) $\iff$ (ii) as \Cref{thm:bzw}; | |
the idea of the proof is much in the spirit of \Cref{thm:interval_compact}. | |
\begin{proof} | |
[Proof that (i) and (ii) $\implies$ (iii)] | |
We prove the following lemma, which is interesting in its own right. | |
\begin{lemma} | |
[Lebesgue number lemma] | |
Let $M$ be a compact metric space and $\{U_\alpha\}$ an open cover. | |
Then there exists a real number $\delta > 0$, | |
called a \vocab{Lebesgue number} for that covering, | |
such that the $\delta$-neighborhood of any point | |
$p$ lies entirely in some $U_\alpha$. | |
\end{lemma} | |
\begin{subproof}[Proof of lemma] | |
Assume for contradiction that for every $\delta = 1/k$ | |
there is a point $x_k \in M$ | |
such that its $1/k$-neighborhood isn't contained in any $U_\alpha$. | |
In this way we construct a sequence $x_1$, $x_2$, \dots; | |
thus we're allowed to take a subsequence which converges to some $x$. | |
Then for every $\eps > 0$ we can find an integer $n$ such that $d(x_n, x) + 1/n < \eps$; | |
thus the $\eps$-neighborhood at $x$ isn't contained in any $U_\alpha$ for every $\eps > 0$. | |
This is impossible, because we assumed $x$ was covered by some open set. | |
\end{subproof} | |
Now, take a Lebesgue number $\delta$ for the covering. | |
Since $M$ is totally bounded, finitely many $\delta$-neighborhoods cover the space, | |
so finitely many $U_\alpha$ do as well. | |
\end{proof} | |
\begin{proof} | |
[Proof that (iii) $\implies$ (ii)] | |
One step is immediate: | |
\begin{ques} | |
Show that the covering condition $\implies$ totally bounded. | |
\end{ques} | |
The tricky part is showing $M$ is complete. | |
Assume for contradiction it isn't and thus that the sequence $(x_k)$ is Cauchy, | |
but it doesn't converge to any particular point. | |
\begin{ques} | |
Show that this implies for each $p \in M$, there is an $\eps_p$-neighborhood $U_p$ | |
which contains at most finitely many of the points of the sequence $(x_k)$. | |
(You will have to use the fact that $x_k \not\to p$ and $(x_k)$ is Cauchy.) | |
\end{ques} | |
Now if we consider $M = \bigcup_p U_p$ we get a | |
finite subcover of these open neighborhoods; | |
but this finite subcover can only cover finitely | |
many points of the sequence, by contradiction. | |
\end{proof} | |
\section{\problemhead} | |
The later problems are pretty hard; | |
some have the flavor of IMO 3/6-style constructions. | |
It's important to draw lots of pictures so one can tell what's happening. | |
% I'd be happy to learn of any easier ones. | |
Of these \Cref{thm:bzw} is definitely my favorite. | |
\begin{problem} | |
Show that the closed interval $[0,1]$ and | |
open interval $(0,1)$ are not homeomorphic. | |
\begin{hint} | |
$[0,1]$ is compact. | |
\end{hint} | |
\begin{sol} | |
Compactness is preserved under homeomorphism, | |
but $[0,1]$ is compact while $(0,1)$ is not. | |
\end{sol} | |
\end{problem} | |
\begin{problem} | |
Let $X$ be a topological space with the discrete topology. | |
Under what conditions is $X$ compact? | |
\begin{hint} | |
If and only if it is finite. | |
\end{hint} | |
\end{problem} | |
\begin{problem} | |
[The cofinite topology is quasicompact only] | |
We let $X$ be an infinite set and equip it with the | |
\vocab{cofinite topology}: | |
the open sets are the empty set and complements of finite sets. | |
This makes $X$ into a topological space. | |
Show that $X$ is quasicompact but not Hausdorff. | |
\end{problem} | |
\begin{problem} | |
[Cantor's intersection theorem] | |
\label{prob:cantor_intersect} | |
Let $X$ be a compact topological space, and suppose | |
\[ X = K_0 \supseteq K_1 \supseteq K_2 \supseteq \dots \] | |
is an infinite sequence of nested nonempty closed subsets. | |
Show that $\bigcap_{n \ge 0} K_n \neq \varnothing$. | |
\end{problem} | |
%\begin{sproblem}[Compact Implies Bounded] | |
% Let $f : X \to \RR$ be a continuous function, | |
% where $X$ is a compact topological space. | |
% Show that $f$ is bounded. | |
% \begin{hint} | |
% Immediate by the fact that the image of $f$ is compact, | |
% and hence bounded. | |
% Remember this! | |
% \end{hint} | |
%\end{sproblem} | |
\begin{problem} | |
[Tychonoff's theorem] | |
Let $X$ and $Y$ be compact metric spaces. Show that $X \times Y$ is compact. | |
(This is also true for general topological spaces, | |
but the proof is surprisingly hard, | |
and we haven't even defined $X \times Y$ in general yet.) | |
\label{prob:tychonoff} | |
\begin{hint} | |
Suppose $p_i = (x_i, y_i)$ is a sequence in $X \times Y$ ($i=1,2,\dots$). | |
Take a sub-sequence such that the $x$-coordinate converges | |
(throwing out some terms). | |
Then take a sub-sequence of \emph{that} sub-sequence | |
such that $y$-coordinate converges (throwing out more terms). | |
\end{hint} | |
\begin{sol} | |
Suppose $p_i = (x_i, y_i)$ is a sequence in $X \times Y$ ($i=1,2,\dots$). | |
Looking on the $X$ side, some subsequence converges: | |
for the sake of illustration say it's $x_1, x_4, x_9, x_{16}, \dots \to x$. | |
Then look at the corresponding sequence $y_1, y_4, y_9, y_{16}, \dots$. | |
Using compactness of $Y$, it has a convergent subsequence, say | |
$y_1, y_{16}, y_{81}, y_{256}, \dots \to y$. | |
Then $p_1, p_{16}, p_{81}, \dots$ will converge to $(x,y)$. | |
One common mistake is to just conclude | |
that $(x_n)$ has a convergent subsequence | |
and that $(y_n)$ does too. | |
But these sequences could be totally unrelated. | |
For this proof to work, | |
you do need to apply compactness of $X$ first, | |
and then compactness of $Y$ on the resulting \emph{filtered} | |
sequence like we did here. | |
\end{sol} | |
\end{problem} | |
\begin{dproblem} | |
[Bolzano-Weierstra\ss\ theorem for general metric spaces] | |
\gim | |
Prove that a metric space $M$ is sequentially compact | |
if and only if it is complete and totally bounded. | |
\label{thm:bzw} | |
\begin{hint} | |
Mimic the proof of \Cref{thm:interval_compact}. | |
The totally bounded condition lets you do Pigeonhole. | |
\end{hint} | |
\end{dproblem} | |
\begin{problem} | |
[Almost Arzel\`a-Ascoli theorem] | |
\gim | |
Let | |
$f_1, f_2, \ldots \colon [0,1] \to [-100,100]$ | |
be an \vocab{equicontinuous} sequence of functions, meaning | |
\[ | |
\forall \eps > 0 \quad | |
\exists \delta > 0 \quad | |
\forall n \; | |
\forall x,y \quad | |
\left( \left\lvert x-y \right\rvert <\delta | |
\implies \left\lvert f_n(x) - f_n(y) \right\rvert < \eps \right) | |
\] | |
Show that we can extract a subsequence $f_{i_1}, f_{i_2}, \dots$ | |
of these functions such that for every $x \in [0,1]$, | |
the sequence $f_{i_1}(x)$, $f_{i_2}(x)$, \dots\ converges. | |
\end{problem} | |
\begin{problem} | |
\gim | |
Let $M = (M,d)$ be a bounded metric space. | |
Suppose that whenever $d'$ is another metric on $M$ | |
for which $(M,d)$ and $(M,d')$ are homeomorphic | |
(i.e.\ have the same open sets), then $d'$ is also bounded. | |
Prove that $M$ is compact. | |
\end{problem} | |
\begin{problem} | |
\yod | |
In this problem a ``circle'' | |
refers to the boundary of a disk with \emph{nonzero} radius. | |
\begin{enumerate}[(a)] | |
\ii Is it possible to partition the plane $\RR^2$ | |
into disjoint circles? | |
\ii From the plane $\RR^2$ we delete two distinct points $p$ and $q$. | |
Is it possible to partition the remaining points into disjoint circles? | |
\end{enumerate} | |
\begin{hint} | |
The answer to both parts is no. | |
For (a) use \Cref{prob:cantor_intersect}. | |
For (b), color each circle in the partition | |
based on whether it contains $p$ but not $q$, | |
$q$ but not $p$, or both. | |
\end{hint} | |
\begin{sol} | |
Part (a) follows by the Cantor intersection theorem | |
(\Cref{prob:cantor_intersect}). | |
Assume for contradiction such a partition existed. | |
Take any of the circles $C_0$, and let $K_0$ denote the closed disk | |
with boundary $C_0$. | |
Now take the circle $C_1$ passing through the center of $C_0$, | |
and let $K_1$ denote the closed disk with boundary $C_1$. | |
If we repeat in this way, | |
we get a nested sequence $K_0 \supseteq K_1 \supseteq \dots$ | |
and the radii of $C_i$ approach zero | |
(since each is at most half the previous once). | |
Thus some point $p$ lies in $\bigcap_n K_n$ which is impossible. | |
Now for part (b), | |
again assume for contradiction a partition into circles exists. | |
Color a circle magenta if it contains $p$ but not $q$ | |
and color a circle cyan if it contains $q$ but not $p$. | |
Color $p$ itself magenta and $q$ itself cyan as well. | |
Finally, color a circle neon yellow if it contains both $p$ and $q$. | |
(When we refer to coloring a circle, | |
we mean to color all the points on it.) | |
By repeating the argument in (a) there are no circles | |
enclosing neither $p$ nor $q$. | |
Hence every point is either magenta, cyan, or neon yellow. | |
Now note that given any magenta circle, | |
its interior is completely magenta. | |
Actually, the magenta circles can be totally ordered | |
by inclusion (since they can't intersect). | |
So we consider two cases: | |
\begin{itemize} | |
\ii If there is a magenta circle which is maximal by inclusion | |
(i.e.\ a magenta circle not contained in any other magenta circle) | |
then the set of all magenta points is just a closed disk. | |
\ii If there is no such magenta circle, | |
then the set of magenta points can also be expressed | |
as the union over all magenta circles of their interiors. | |
This is a union of open sets, so it is itself open. | |
\end{itemize} | |
We conclude the set of magenta points is | |
either a closed disk or an open set. | |
Similarly for the set of cyan points. | |
Moreover, the set of such points is convex. | |
To finish the problem: | |
\begin{itemize} | |
\ii Suppose there are no neon yellow points. | |
If the magenta points form a closed disk, | |
then the cyan points are $\mathbb R^2$ minus a disk which is not convex. | |
Contradiction. So the magenta points must be open. | |
Similarly the cyan points must be open. | |
But $\mathbb R^2$ is connected, | |
so it can't be written as the union of two open sets. | |
\ii Now suppose there are neon yellow points. | |
We claim there is a neon yellow circle minimal by inclusion. | |
If not, then repeat the argument of (a) to get a contradiction, | |
since any neon yellow circle must have diameter the distance from $p$ to $q$. | |
So we can find a neon yellow circle $\mathscr C$ whose | |
interior is all magenta and cyan. | |
Now repeat the argument of the previous part, | |
replacing $\mathbb R^2$ by the interior of $\mathscr C$. | |
\end{itemize} | |
\end{sol} | |
\end{problem} | |