Anda di halaman 1dari 16

WMS

MA222
Metric Spaces

Revision Guide

Written by Gareth Speight

WMS
ii MA222 Metric Spaces

Contents
1 Metric Spaces 1

2 Topological Spaces 4

3 Compactness 7

4 Connectedness 10

5 Completeness 12

Introduction
This revision guide for MA222 Metric Spaces has been designed as an aid to revision, not a substitute
for it. As can be seen from the length of this revision guide, Metric Spaces is a fairly long course with
lots of confusing definitions and big theorems; hopefully this guide will help you make sense of how all
the abstraction fits together. Only some proofs are included; the inclusion of a particular proof is no
indication of how likely it is to appear on the exam. Note that this guide is omits material on connections
between compact metric spaces and the Cantor set (Every compact metric space is a continuous image
of the Cantor set).

Disclaimer: Use at your own risk. No guarantee is made that this revision guide is accurate or
complete, or that it will improve your exam performance. Use of this guide will increase entropy,
contributing to the heat death of the universe. Contains no GM ingredients. Your mileage may vary.
All your base are belong to us.

Authors
Originally by Gareth Speight (g.speight@warwick.ac.uk).
Based upon lectures given by David Preiss at the University of Warwick, 2007.
Any corrections or improvements should be entered into our feedback form at http://tinyurl.com/WMSGuides
(alternatively email revision.guides@warwickmaths.org).

History
First Edition: May 23, 2007
Current Edition: May 6, 2013.
MA222 Metric Spaces 1

1 Metric Spaces
We recall that the modulus function on R allows us to calculate the distance between real numbers. The
idea of a metric on a set generalises this and allows us to have notions of distances on any set.

Definition 1.1. A metric on a set M is a function d : M × M → R such that for all x, y, z ∈ M ,

• d(x, y) ≥ 0; and d(x, y) = 0 ⇐⇒ x = y (positive definiteness).

• d(x, y) = d(y, x) (symmetry).

• d(x, z) ≤ d(x, y) + d(y, z) (triangle inequality).

The pair (M, d) is called a metric space. When it is clear which metric we are using we may refer to “the
metric space M ”.

Recall the definition of a norm.

Definition 1.2. A norm on a vector space V over R (or C) is a function k · k : V → R such that for
every x, y ∈ V ,

• kxk ≥ 0; kxk = 0 ⇐⇒ x = 0 (positive definiteness).

• kcxk = |c|kxk for every c ∈ R (or C) (homogeneity).

• kx + yk ≤ kxk + kyk (triangle inequality).

The pair (V, k · k) is called a normed vector space.

Every normed vector space is in fact a metric space with metric d(x, y) = kx − yk and we can obtain
several important metrics from normed vector spaces.

Examples 1.3. Here are some familiar examples of norms or metrics.


n n
pPn
• The standard Euclidean 2
pPn norm on R or C is given by kxk = i=1 |xi | and the corresponding
metric is d(x, y) = 2
i=1 |xi − yi | . Assume in this guide that, unless otherwise stated, whenever
Rn or Cn are mentioned we have in mind the Euclidean norm or metric.
Pn 1/p
• The lp (p ≥ 1) norm on Rn or Cn is given by kxkp = ( i=1 |xi |p ) and the corresponding metric
p 1/p
Pn
is dp (x, y) = ( i=1 |xi − yi | ) .

• The l∞ norm on Rn or Cn is given by kxk∞ = maxni=1 |xi | and the corresponding metric is
d∞ (x, y) = maxni=1 |xi − yi |.

• kf k = supx∈[a,b] |f (x)| is a norm on the set C([a, b]) of continuous functions [a, b] → R and the
corresponding metric is d(f, g) = kf − gk = supx∈[a,b] |f (x) − g(x)|.

• Let M be any set. Then we can define a metric on M by d(x, y) = 1 whenever x 6= y and d(x, y) = 0
when x = y. We call d the discrete metric on M and (M, d) a discrete space.

Definition 1.4. The open ball centred at a ∈ M with radius r is the set B(a, r) = {x ∈ M : d(x, a) < r}
and the closed ball centred at a ∈ M with radius r is {x ∈ M : d(x, a) ≤ r}.

Definition 1.5. A subset S of a metric space M is bounded if there exists a ∈ M and r > 0 such that
S ⊂ B(a, r).

Equivalently S ⊂ M is bounded if the distance function is bounded on S, i.e. if there exists r > 0
such that d(x, y) ≤ r for all x, y ∈ S.
We can define new metric spaces from existing ones by considering subsets or taking products.

Definition 1.6. Let (M, dM ) be a metric space and let H ⊂ M . Define a metric dH on H by dH (x, y) :=
dM (x, y) for all x, y ∈ H. Then (H, dH ) is a metric space and we refer to it as a subspace of M .
2 MA222 Metric Spaces

Definition 1.7. If Mi , 1 ≤ i ≤ n, are metric spaces then the product M1 × . . . × Mn becomes a


1/p
metric space, which we call a product space with either of the metrics d(x, y) = ( ni=1 (di (xi , yi ))p )
P
or
d(x, y) = maxni=1 di (xi , yi ) where x = (x1 , . . . , xn ) and y = (y1 , . . . , yn ) for xi , yi ∈ Mi .
Care should be taken with subspaces. If we are working in the metric space R the ball B(0, 1) is the
interval (−1, 1), while if the metric space is [0, 2] the ball B(0, 1) is the interval [0, 1).
Proposition 1.8. |d(x, y) − d(x, z)| ≤ d(y, z) for all x, y, z ∈ M .
Proof. The triangle inequality gives d(x, y) ≤ d(x, z) + d(z, y) and hence d(x, y) − d(x, z) ≤ d(z, y) =
d(y, z). Interchanging y and z gives d(x, z) − d(x, y) ≤ d(y, z). Hence |d(x, y) − d(x, z)| ≤ d(y, z).
Replacing the modulus in the definition of continuity for a function R → R by distances defined on
the domain and codomain gives the definition of continuity for a map between metric spaces.
Definition 1.9. Let f : M1 → M2 be a map between metric spaces (M1 , d1 ) and (M2 , d2 ). Then we say:

• f is (d1 , d2 )-continuous at a ∈ M1 (or just continuous at a ∈ M1 ) if ∀ ε > 0, ∃ δ > 0 such that


x ∈ M1 and d1 (x, a) < δ =⇒ d2 (f (x), f (a)) < ε.
• f is continuous if it is continuous at every point of M1 .
• f is Lipschitz if there exists C ∈ R≥0 such that d2 (f (x), f (y)) ≤ Cd1 (x, y) for all x, y ∈ M1 .

It is easy to see a Lipschitz map is continuous by letting δ = ε/C. Sums, products and quotients
(where the quotient function makes sense) of continuous functions M → R are all continuous with similar
proofs to those in the case R → R. The previous proposition shows that, for any a ∈ M , the map M → R,
x 7→ d(a, x) is Lipschitz with constant C = 1.
Definition 1.10. Let M be a metric space. We say U ⊂ M is open in M if for all x ∈ U there exists
δ > 0 such that B(x, δ) ⊂ U . We say F ⊂ M is closed in M if M \ F is open.
Informally U ⊂ M is open if given a point in U you can move some small distance δ > 0 in any
direction and still remain in U . It is important to stress that a set could be both open and closed or
neither open nor closed. In any metric space M both ∅ and M are open and closed while in R the
interval [0, 1) is neither open nor closed.
We now prove a few elementary properties about open balls and open sets.
Proposition 1.11. An open ball B(a, r) is open.
Proof. Take x ∈ B(a, r) and define δ = r−d(x, a) > 0. Let y ∈ B(x, δ). Then d(y, a) ≤ d(y, x)+d(x, a) <
δ + d(x, a) = r. Hence B(x, δ) ⊂ B(a, r).

Proposition 1.12. Let U1 , . . . Uk be open in M . Then ki=1 Ui is open in M .


T

Tk
Proof. Take x ∈ i=1 Ui . Then, for each i, x ∈ Ui and Ui is open so there is δi > 0 such that B(x, δi ) ⊂
Tk
Ui . Let δ = min(δ1 , . . . , δk ). Then B(x, δ) ⊂ B(x, δi ) ⊂ Ui for each i and hence B(x, δ) ⊂ i=1 Ui .
T
Note that k being finite really is necessary here. n∈N (−1/n, 1/n) = 0 is a countable intersection of
open sets which is not open.
Proposition 1.13. The union of any collection of sets open in M is open in M .
S
Proof. Write U = i∈I Ui . Take x ∈ U . Then x ∈ Ui for some i. Since Ui is open there is δ > 0 such
that B(x, δ) ⊂ Ui ⊂ U .
Proposition 1.14. H ⊂ M is open if and only if it is a union of open balls.
Proof. By the above proposition a union of open balls is open. Suppose
S H ⊂ M is open. Then for all
x ∈ H there is δx > 0 such that B(x, δx ) ⊂ H. We can write H = x∈H B(x, δx ).
Definition 1.15. A sequence xn ∈ M converges to x ∈ M if d(xn , x) → 0 as n → ∞.
MA222 Metric Spaces 3

This definition of convergence can be reformulated to provide an equivalent definition in the language
of open sets.
Proposition 1.16. xn → x if and only if for every open set U containing x there is N ∈ N such that
xn ∈ U for all n ≥ N .1
Proof. (=⇒) Suppose xn → x. Since U is open and contains x there is δ > 0 such that B(x, δ) ⊂ U .
As xn → x there is N ∈ N such that n ≥ N =⇒ d(xn , x) < δ =⇒ xn ∈ B(x, δ) ⊂ U . So
n ≥ N =⇒ xn ∈ U .
(⇐=) Let ε > 0. B(x, ε) is open so there is N such that n ≥ N =⇒ xn ∈ B(x, ε) =⇒ d(xn , x) < ε.
Convergence of sequences and closed sets are also closely related.
Theorem 1.17. Let F be a subset of a metric space M , and let xn ∈ F be a sequence with xn → x ∈ M .
Then F is closed if and only if x ∈ F ; that is, a subset F of a metric space M is closed if and only if for
every sequence xn ∈ F that converges to some x ∈ M we have x ∈ F .
Proof. (=⇒) Suppose F is closed, xn ∈ F and xn → x ∈ M . Suppose x ∈ M \ F which is open. Then
there is N such that n ≥ N =⇒ xn ∈ M \ F , contradicting xN ∈ F .
(⇐=) Suppose F is not closed. Then M \ F is not open and consequently there is x ∈ M \ F such
that for every n > 0 we have B(x, 1/n) not contained in M \ F , i.e. B(x, 1/n) ∩ F is nonempty. Choose
xn ∈ B(x, 1/n) ∩ F . Then d(xn , x) < 1/n → 0 so xn → x with x ∈ / F , giving a contradiction.
As in R we have the notion of sequential continuity and its relation to continuity.
Theorem 1.18. A function f : M1 → M2 where M1 , M2 are metric spaces, is continuous at x ∈ M1 if
and only if for every sequence xn ∈ M1 with xn → x ∈ M1 we have f (xn ) → f (x).2
Recall that for a function f : X → Y we define the preimage of a set U under f by f −1 (U ) =
{x : f (x) ∈ U }. The following theorem reformulates continuity in terms of open sets, motivating our
progression to topological spaces.
Theorem 1.19. A function f : M1 → M2 is continuous if and only if for every open set U ⊂ M2 , the
preimage f −1 (U ) is open in M1 . In other words, f is continuous if and only if the preimage of every
open set is open.
Proof. (=⇒) Let U be open in M2 and take x ∈ f −1 (U ). Then f (x) ∈ U so, as U is open, there
is ε > 0 such that B(f (x), ε) ⊂ U . Since f is continuous at x there exists δ > 0 such that (y ∈
M1 , d1 (y, x) < δ) =⇒ d2 (f (y), f (x)) < ε. In other words y ∈ B(x, δ) =⇒ f (y) ∈ B(f (x), ε) ⊂ U .
Hence B(x, δ) ⊂ f −1 (U ). Hence f −1 (U ) is open.
(⇐=) Let x ∈ M1 and ε > 0. B(f (x), ε) is open in M2 so f −1 (B(f (x), ε)) is open in M1 . Furthermore
x ∈ f −1 (B(f (x), ε)) so there is δ > 0 such that B(x, δ) ⊂ f −1 (B(f (x), ε)). In other words (y ∈
M1 , d1 (y, x) < δ) =⇒ d2 (f (y), f (x)) < ε so f is continuous at x. Since x ∈ M1 was arbitrary f is
continuous.
As well as providing a concise formulation of continuity the above theorem allows us to show subsets
are open by using continuous maps.
Example 1.20. Define U = {(x, y) ∈ R2 :x2 − y 2 > 1,y > 0}. Define f (x, y) = x2 − y 2 and g(x, y) = y.
Then f and g are continuous so f −1 (1, ∞) , g −1 (0, ∞) are open and hence f −1 (1, ∞) ∩g −1 (0, ∞) =
U is open.
Theorem 1.21. Let d1 , d2 be two metrics on a set M . Then the following statements are equivalent:
• For every metric space (N, d), every f : M → N is (d1 , d)-continuous if and only if it is (d2 , d)-
continuous.
• For every metric space (N, d), every g : N → M is (d, d1 )-continuous if and only if it is (d, d2 )-
continuous.
1 This is the idea behind convergence of sequences in general topological spaces.
2 Sequential continuity and continuity are not equivalent in general topological spaces.
4 MA222 Metric Spaces

• d1 -open and d2 -open sets coincide.


So informally two metrics on M give the same answers to questions of continuity if and only if they
give the same collection of open sets.
Definition 1.22. Two metrics d1 , d2 on the same set M are called topologically equivalent if d1 -open
and d2 -open sets coincide3 .
Proposition 1.23. If there exist c, C ∈ (0, ∞) such that cd1 (x, y) ≤ d2 (x, y) ≤ Cd1 (x, y) for all x,
y ∈ M then d1 , d2 are topologically equivalent.
Proof. Let U be d2 open and x ∈ U . There is δ > 0 such that Bd2 (x, δ) ⊂ U . Then,

Bd1 (x, δ/C) = {y ∈ M : d1 (y, x) < δ/C} ⊂ {y ∈ M : d2 (y, x) < δ} = Bd2 (x, δ) ⊂ U

so U is d1 open. The converse follows from the inequality (1/C)d2 (x, y) ≤ d1 (x, y) ≤ (1/c)d2 (x, y).
Example 1.24. The norms dp on Rn are topologically equivalent for p = 1, 2, . . . , ∞. To see this note
that d∞ (x, y) ≤ dp (x, y) ≤ n1/p d∞ (x, y) for all x, y ∈ Rn , so dp is equivalent to d∞ for any p ≥ 1.
Definition 1.25. A bijection f : M1 → M2 between two metric spaces is called an isometry if
d2 (f (x), f (y)) = d1 (x, y) for every x, y ∈ M1 . We say M1 , M2 are isometric if there is an isometry
between them.
Isometric metric spaces are essentially the same as metric spaces, just with points relabelled.
Definition 1.26. A bijection f : M1 → M2 between two metric spaces is called a homeomorphism if
U is open in M1 if and only if f (U ) is open in M2 . We say M1 , M2 are homeomorphic if there is a
homeomorphism between them.
Equivalently a homeomorphism is a bijection f : M1 → M2 where both f and f −1 are continuous.
Homeomorphic metric spaces give the same answers for questions involving continuity and open sets.
We say a property is a topological property or a topological invariant of metric spaces if whenever
M has the property so does every metric space homeomorphic to it. In other words the property is
a topological invariant if it is preserved by homeomorphisms. If one space has a certain topological
property and another doesn’t then the two spaces cannot be homeomorphic.

2 Topological Spaces
Definition 2.1. A topology on a set T is a collection T of subsets of T , whose members are called open
sets, such that

• ∅ and T ∈ T (i.e. ∅ and T are open in T ).


• The intersection of any finite collection of open sets is open.
• The union of any collection of open sets is open.

The pair (T, T ) is called a topological space; if the topology is implicit from the context, we often
write “the topological space T”. We say a subset F of T is closed if its complement is open.
There is symmetry between open and closed sets. In particular the intersection of any collection of
closed sets is closed and the union of any finite collection of closed sets is closed.4 We could, equivalently,
define a topology in terms of closed sets and call subsets open if their complements are closed.
Examples 2.2. Some examples of topologies:
3 It is not difficult to see that being topologically equivalent gives an equivalence relation on the set of metrics on M .

Any metric is equivalent to itself. If d1 and d2 are equivalent then d2 and d1 are equivalent. If d1 , d2 and d2 , d3 are
equivalent pairs then d1 , d3 are equivalent.S T T S
4 To see this, recall De Morgan’s Laws ( c c c c
i∈I Ai ) = i∈I Ai , ( i∈I Ai ) = i∈I Ai for sets Ai , i ∈ I, where I is some
index set.
MA222 Metric Spaces 5

• If d is a metric on T then the collection of d-open sets is a topology on T . We often refer to this
as the topology induced by the metric.
• The discrete topology on T is the topology in which all subsets of T are open.
• The indiscrete topology on T is the topology in which only ∅ and T are open.
We use our motivation from continuous functions between metric spaces to define a continuous func-
tion between topological spaces.
Definition 2.3. A map f : T1 → T2 between two topological spaces is said to be continuous if for every
open set U ⊂ T2 the set f −1 (U ) is open in T1 .
Continuous functions are useful because they preserve important properties that a topological space
might have. We will see later that continuous functions preserve compactness and connectedness.
Examples 2.4. Any constant map f : T1 → T2 is continuous (the preimage of an open set is either T1
or ∅. The identity map f : T → T is continuous (the preimage of an open set is the same open set in the
same space). Continuous maps between metric spaces are continuous maps between the corresponding
topological spaces.
Theorem 2.5. If T1 , T2 , T3 are topological spaces and f : T1 → T2 and g : T2 → T3 are continuous then
g ◦ f : T1 → T3 is continuous.
Proof. Let U be open in T3 and apply the formula (g ◦ f )−1 (U ) = f −1 (g −1 (U )).
To describe a topology on a set it isn’t necessary to list every element. As a topology is closed under
unions and finite intersections it is enough to list the elements that generate the topology on the set.
Definition 2.6. A basis for a topology T on T is a collection B ⊂ T such that every set from T is a
union of sets from B.
A sub-basis for a topology T on T is a collection B ⊂ T such that every set from T is a union of
finite intersections of sets from B.5
Example 2.7. In a metric space M a set is open if and only if it is a union of open balls. Hence open balls
form a basis for the topology induced by the metric. The collection of open intervals {(a, ∞), (−∞, b) :
a, b ∈ R} is a sub-basis for the Euclidean topology of R.
Proposition 2.8. Let B be a collection of subsets of a set T . Suppose T is the union of sets from B
and the intersection of any two sets from B is a union of sets from B. Then there is a unique topology
on T with basis B. It can also be described as the weakest (or smallest) topology containing B. Its open
sets are exactly the unions of sets from B.
Proposition 2.9. If B is any collection of subsets of a set T then there is a unique topology on T with
sub-basis B. Its open sets are exactly the unions of finite intersections of sets from B.
Just as in metric spaces, we can define new topological spaces from old ones using subsets and
products.
Definition 2.10. If (T, T ) is a topological space and S ⊂ T we define the subspace topology on S as
TS = {(U ∩ S : U ∈ T }. We call (S, TS ) a (topological ) subspace of T .
We say subsets of a space have a particular property (such as compactness or connectedness which
we see later) if they do when regarded as subspaces under the subspace topology.
Definition 2.11. If (T1 , T1 ), (T2 , T2 ) are topological spaces, the product topology on T1 × T2 is the
topology with basis B = {U1 × U2 : U1 ∈ T1 , U2 ∈ T2 }. We call T1 × T2 the topological product of T1 , T2 .
Notice that the open sets of T1 × T2 are not simply the products of open sets from T1 and T2 . We
have to take the unions of all such products to obtain the topology on T1 × T2 . This is because the set
of products of open sets isn’t necessarily closed under unions (though it is under finite intersections).
5 We use the sensible convention that the union of an empty collection of sets is the empty set while the intersection of

an empty collection of sets is T .


6 MA222 Metric Spaces

Proposition 2.12. Define the first projection of T1 ×T2 onto T1 by π1 (x, y) = x and the second projection
by π2 (x, y) = y. Then the projections of a topological product to its factors are continuous.

A particular use of a sub-basis is that it suffices to check continuity only for the sets in a sub-basis.

Proposition 2.13. Suppose that T1 , T2 are topological spaces and B is a sub-basis for T2 . Then
f : T1 → T2 is continuous if and only if f −1 (U ) is open for every U ∈ B.

Proof. (=⇒) This direction is obvious.


(⇐=) An open set is a union of finite intersections of sets from B. Taking the preimage preserves
unions and intersections so we have a union of finite intersections of open sets, which is open.

Proposition 2.14. Let f1 : T → T1 , f2 : T → T2 be maps between topological spaces. Then f =


(f1 , f2 ) : T → T1 × T2 is continuous if and only if f1 and f2 are continuous.

Proof. (=⇒) fi = πi ◦ f is the composition of two continuous functions, hence continuous.


(⇐=) If Ui are open in Ti then f −1 (U1 × U2 ) = f1−1 (U1 ) ∩ f2−1 (U2 ) is open in T . We have checked
continuity for a basis of T1 × T2 so the result follows by applying the above proposition.

Examples 2.15. If f, g : T → R are continuous and c ∈ R, then cf , |f |, max(f, g), min(f, g), f + g, f g
and, if defined, f /g are all continuous. Proof for f + g:
[  [
(f + g)−1 (a, ∞) = f −1 (r, ∞) ∩ g −1 (a − r, ∞),
 
{x : f (x) > r} ∩ {x : g(x) > a − r} =
r∈R r∈R

which is open. Similarly (f + g)−1 (−∞, b) is open and since (a, ∞), (−∞, b) constitute a sub-basis of
R we see that f + g is continuous.

Definition 2.16. A bijection f : T1 → T2 is called a homeomorphism if U is open in T1 if and only if


f (U ) is open in T2 . If there is a homeomorphism f : T1 → T2 we say T1 and T2 are homeomorphic.
A topological invariant (or topological property) of topological spaces is a property preserved by
homeomorphisms. In other words, if a space T has the property then so does any space homeomorphic
to T .

Homeomorphic topological spaces behave the same for questions phrased in terms of open sets and
continuity.
The idea of topological invariants is useful when we want to distinguish between different spaces. If
one space has a particular topological property but another does not, then they cannot be homeomorphic.

Remark 2.17. A bijection f : T1 → T2 is a homeomorphism if and only if f and f −1 are both continuous.
The inverse of a homeomorphism is a homeomorphism and the composition of two homeomorphisms is
a homeomorphism.

Example 2.18. f : (−1, 1) → R, f (x) = x/(1 − |x|) is a homeomorphism.

Definition 2.19. Let T be a topological space.

• The interior H ◦ of a set H ⊂ T is the union of all open sets contained in H.

• A neighbourhood of x is a set H such that x ∈ H ◦ .

• The closure H of a set H ⊂ T is the set of points x such that every neighbourhood of x meets H,
i.e. every neighbourhood of x has nonempty intersection with H.

Proposition 2.20. H ◦ = T \ (T \ H) and H = T \ (T \ H)◦ .

Proof. We prove the first. Suppose x ∈ H ◦ . Then H is a neighbourhood of x not meeting T \ H, so


x ∈/ (T \ H) and hence x ∈ T \ (T \ H). Suppose x ∈ T \ (T \ H). Then x ∈ / (T \ H) so there is a
neighbourhood of x not meeting T \ H. This neighbourhood is contained in H and so x ∈ H ◦ . The
second follows upon replacing H by H ′ := T \ H.
MA222 Metric Spaces 7

Examples 2.21. On R we have (a, b) = [a, b], [a, b]◦ = (a, b), Q = R, Q◦ = ∅. Note though that the
space we are working in is important. In [a, b] we have [a, b]◦ = [a, b] ([a, b] is open in [a, b]) while in (a, b)
we have (a, b) = (a, b) ((a, b) is closed in (a, b)).

Remark 2.22. The following follow easily from the definitions and the proposition:

• H ◦ is open and H is closed.

• H ◦ is the largest open subset contained in H and H is the smallest closed set containing H.

• H is open if and only if H = H ◦ and H is closed if and only if H = H.

• (H ◦ )◦ = H ◦ and H = H.

• H ⊂ K =⇒ H ◦ ⊂ K ◦ and H ⊂ K.

• (H ∩ K)◦ = H ◦ ∩ K ◦ and H ∪ K = H ∪ K.

Definition 2.23. The boundary ∂H of H is the set of points x whose every neighbourhood meets both
H and its complement.

Remark 2.24. ∂H = H ∩ (T \ H) is closed.

Examples 2.25. In R, ∂(a, b) = ∂[a, b] = {a, b}, ∂Q = R.

Example 2.26 (Cantor set). Define C0 = [0, 1], C1 = [0, 1/3] ∪ [2/3, 1], . . . so that at each stage we
remove the open middle third of each of the closed intervals in the union. We define the Cantor set to
be the intersection C = ∞
T
n=0 Cn . Since each Cn is closed it follows C is closed. The only open set
contained in C is ∅ so C has empty interior. C also has uncountably many points.

In many cases the notion of a topological space is too general and we might wish to impose some
additional axioms. Separation axioms allow a way to distinguish between points and sets. One of
the most used separation axioms is the Hausdorff condition. Informally we say a topological space is
Hausdorff if any two points can be separated by neighbourhoods.

Definition 2.27. A topological space T is said to be Hausdorff if for any two distinct points x, y ∈ T
there are disjoint open sets U , V containing x, y respectively.

Proposition 2.28. Every metric space is Hausdorff.

Proof. If x 6= y let r = d(x, y) > 0. Then the open balls B(x, r/2), B(y, r/2) are disjoint.

Remark 2.29. The property of being Hausdorff is a topological invariant since it is phrased in terms
of open sets.

3 Compactness
Definition 3.1. We say a topological space T is compact if every open cover of T has a finite subcover.
To be precise,

• A cover of a set A is a collection U of sets whose union contains A.

• A subcover is a subcollection of U which still covers A.

• We say a cover is open if all of its members are open.

The word ‘every’ in the definition of compactness is important. T is always covered by {T } which is
an open cover with the finite subcover {T }.

Example 3.2. {(a, 1] : 0 < a < 1} is an open cover of (0, 1] which has no finite subcover.
8 MA222 Metric Spaces

Remark 3.3. Suppose T is a subspace of another topological space S. Then the notion of compactness
is the same regardless of whether we think of a cover of T by open subsets of T or by sets open in S.

Theorem 3.4 (Heine–Borel). Any closed bounded interval [a, b] in R is compact6 .

Proof. Let U be a cover of [a, b] by sets open in R. Let A denote the set of x ∈ [a, b] such that [a, x] can
be covered by a finite subfamily of U. Then a ∈ A so A 6= ∅ and is bounded above by b. Let c = sup A.
Then a ≤ c ≤ b so c ∈ U for some U ∈ U. Since U is open there is δ > 0 such that (c − δ, c + δ) ⊂ U .
Since c = sup A there is x ∈ A with x > c − δ. So,

[a, c + δ) = [a, c − δ] ∪ (c − δ, c + δ) ⊂ [a, x] ∪ (c − δ, c + δ)

can be covered by a finite subfamily of U since [a, x] can be covered by a finite subfamily and (c − δ, c + δ)
is covered by the single set U .
So any element of [a, b] that is also contained in (c, c + δ) belongs to A but is strictly bigger than c,
contradicting c = sup A. Consequently (c, c + δ) ∩ [a, b] = ∅, so c = b and [a, b] = [a, c] can be covered
by a finite subfamily of U.

Proposition 3.5. Any closed subset C of a compact space T is compact.

Proof. Let U be a cover of C by sets open in T . Adding to this collection the open set T \ C gives an
open cover of T . This has a finite subcover of T . Remove T \ C if it appears in this collection. We are
then left with a finite subcover of C.

It is sometimes said that compactness is the next best thing to finiteness since we can use open covers
to move from working with infinite collections of sets to finite collections. Try to see this in some of the
proofs.

Proposition 3.6. Any compact subspace C of a Hausdorff space T is closed in T .

Proof. This is trivial if T = C. Suppose not and let a ∈ T \ C. Since T is Hausdorff for each x ∈ C
there are disjoint T -open sets Ux containing x and TnVx containing a. Ux form an open cover of C. Let
Ux1 , . . . , Uxn be a finite subcover of C. Then V = i=1 Vxi is an open set containing a and disjoint from
C. Hence a is contained in an open set contained in T \ C, so a ∈ (T \ C)◦ . Hence T \ C ⊂ (T \ C)◦ .
But (T \ C)◦ ⊂ T \ C always holds so (T \ C)◦ = T \ C. Thus T \ C is open and C is closed.

Proposition 3.7. Any compact subspace C of a metric space M is bounded.

Proof.S Let a ∈ M . Then the balls B(a, r), r > 0 form an open cover of C so there are r1 , . . . , rn with
n
C ⊂ i=1 B(a, ri ) = B(a, max(r1 , . . . , rn )).

Theorem 3.8. Let F be a collection of non empty closed subsets of a compact space T such that every
finite subcollection of F has a non empty intersection. Then the intersection of all sets from F is non
empty.
T
Proof. Write F = {F Si : i c∈ I} for some cindex set I. Suppose i∈I Fi = ∅. Then applying De
Morgan’s laws gives i∈I Fi S= T . Hence Fi , i ∈ I is an open cover of T and so has a T finite subcover
{Ficj : 1 ≤ j ≤ n}. Thus nj=1 Ficj = T so applying De Morgan’s laws again gives nj=1 Fij = ∅,
contradicting the statement of the theorem.

Corollary
T∞3.9. Let F1 ⊃ F2 ⊃ F3 ⊃ . . . be a sequence of non empty closed subsets of a compact space
T . Then k=1 Fk 6= ∅.

Theorem 3.10 (Tychonov). The product T × S of compact spaces T and S is compact.

A useful property of continuous functions is that they allow us to easily show spaces are compact.

Proposition 3.11. A continuous image of a compact space is compact.


6 Allow [x, x] = {x}.
MA222 Metric Spaces 9

Proof. Let f : T → S be continuous and T be compact. Let U be an open cover of f (T ). Since f is


continuous the sets f −1 (U ), U ∈ U are open and form a cover of T . By compactness of T there is a finite
subcover f −1 (U1 ), . . . , f −1 (Un ) of T . Hence U1 , . . . , Un is a finite subcover of f (T ).
Theorem 3.12. If T is compact and S is Hausdorff then any continuous bijection f : T → S is a
homeomorphism.
Proof. We apply the above proposition, translating between closed and compact sets twice. If U is open
in T then T \ U is closed, hence compact. By the above proposition f (T \ U ) is compact and hence
closed. Thus (f −1 )−1 (U ) = f (U ) = S \ f (T \ U ) is open. Hence f −1 is continuous.
Corollary 3.13. If T is compact and f : T → R is continuous then f is bounded and attains its maximum
and minimum.
Proof. Let c = supx∈T f (x) (possibly c = ∞). Suppose f does not  attain the value c. Then f (x) < c
for every x and so the open sets S {x : f (x) < r} = f −1 (−∞, r) , r < c cover T . Use compactness to
n
find r1 , . . . , rn < c so that T ⊂ i=1 {x : f (x) < ri }. Then f (x) < max(r1 , . . . , rn ) for all x. Hence
c = supx∈T f (x) ≤ max(r1 , . . . , rn ) < c, which is a contradiction. Similarly, f is bounded below and
attains its minimum.
The following theorem allows us an easy characterisation of compact subsets of Rn .
Theorem 3.14 (Heine–Borel). A subset of Rn is compact if and only if it is closed and bounded.
Proof. Since metric spaces are Hausdorff we have already proved that any compact subset of Rn is closed
and bounded. If C ⊂ Rn is bounded then there is an interval [a, b] such that C ⊂ [a, b] × · · · × [a, b].
The space on the right is compact by Tychonov’s theorem. If C is also closed it is a closed subset of a
compact space hence compact.
Example 3.15. (0, 1) is a closed and bounded subset of the space (0, 1) but is not compact. In R with
the metric d(x, y) = min(|x − y|, 1) the interval [0, ∞) is closed and bounded but non compact. So the
above theorem does not hold in general metric or topological spaces.
Theorem 3.16. A metric space M is compact if and only if every sequence in M has a subsequence
converging to a point of M .
The two statements aren’t equivalent in general topological spaces and we sometimes say a space is
sequentially compact if every sequence has a convergent subsequence.
Definition 3.17. Given a cover U of a metric space M , a number δ > 0 is called a Lebesgue number of
U if for any x ∈ M there is U ∈ U such that B(x, δ) ⊂ U .
Example 3.18. The sets (x/2, x), 0 < x < 1 form an open cover of (0, 1) which has no Lebesgue number.
Proposition 3.19. Every open cover of a compact metric space has a Lebesgue number.
We extended our definition of continuity from real functions to metric spaces by replacing the modulus
with suitable distance functions. We can do exactly the same with uniform continuity.
Definition 3.20. Let M , N be metric spaces. Then we say a map f : M → N is uniformly continuous
if ∀ ε > 0 ∃ δ > 0 such that x, y ∈ M, dM (x, y) < δ =⇒ dN (f (x), f (y)) < ε.
Recall that in R a continuous function [a, b] → R is necessarily uniformly continuous. This in fact
holds for any continuous map from a compact metric space.
Theorem 3.21. A continuous map from a compact metric space (M, dM ) to a metric space (N, dN ) is
uniformly continuous.
Proof. Let ε > 0. Then the sets Uz = f −1 (BdN (f (z), ε/2)), z ∈ M form an open cover of M . Let δ be its
Lebesgue number. If x, y ∈ M and dM (x, y) < δ then y ∈ B(x, δ). B(x, δ) is contained in a single set Uz
so dN (f (a), f (z)) < ε/2 for any a ∈ B(x, δ). Hence dN (f (x), f (y)) ≤ dN (f (x), f (z))+dN (f (y), f (z)) < ε
and so f is uniformly continuous.
10 MA222 Metric Spaces

4 Connectedness
Definition 4.1. A topological space T is connected if for every decomposition T = A ∪ B, A ∩ B = ∅
into open subsets A and B, either A or B is empty. Otherwise T is said to be disconnected.
Remark 4.2. It is important to remember that here open means open in T , even if T is a subspace of
some other space.
To get around this difficulty there are other equivalent definitions.
Definition 4.3. A subset T of a set S is separated by U, V ⊂ S if

T ⊂ U ∪ V, U ∩ V ∩ T = ∅, U ∩ T 6= ∅, V ∩ T 6= ∅.

Proposition 4.4. A subset T of a topological space S is disconnected if and only if it is separated by


some open subsets U , V ⊂ S.
Proof. (=⇒) If T is disconnected there are non empty A, B ⊂ T open in T such that T = A ∪ B and
A ∩ B = ∅. Since T is a subspace of S, there are U, V open in S such that A = U ∩ T and B = V ∩ T .
Then U, V separate T .
(⇐=) If U, V separate T the definition of connectedness fails with A = U ∩ T and B = V ∩ T .
Proposition 4.5. T is disconnected if and only if

(i) T has a subset which is open, closed and different from ∅ and T .
(ii) There exists a non-constant continuous function from T to a two point discrete space.

Proof. Disconnected =⇒ (i): If T is disconnected then there is a decomposition T = A ∪ B with A, B


open and nonempty. Hence A = T \ B is open, closed and different from ∅, T .
(i) =⇒ (ii): If A ⊂ T is open, closed and different from ∅, T define f : T → {0, 1} by f (x) = 0 if
/ A. Then the possible preimages are T , A, Ac , ∅ which are all open so f is
x ∈ A and f (x) = 1 if x ∈
continuous. Since A and Ac are nonempty f is non-constant.
(ii) =⇒ disconnected: If f : T → {0, 1} is non constant and continuous define A = f −1 (0), B =
−1
f (1). Then A, B are open, nonempty, have empty intersection and their union is T . Hence T is
disconnected.
Remark 4.6. A set I ⊂ R is an interval if and only if for every x, y ∈ I we have z ∈ R, x < z < y =⇒
z ∈ I.
Lemma 4.7. If I ⊂ R is connected then it is an interval.
Proof. Suppose I is not an interval. Then there are x, y ∈ I and z ∈ R such that x < z < y and z ∈
/ I.
Let A = (−∞, z) ∩ I, B = (z, ∞) ∩ I. Then A, B are disjoint, open (by definition of topology on I),
nonempty and, since z ∈/ I, A ∪ B = I. Hence I is not connected.
Lemma 4.8. Any interval I ⊂ R is connected.
Proof. Let I be an interval and suppose it is partitioned into non empty open sets A, B. Choose
a ∈ A, b ∈ B and assume a < b. Then A, B form an open cover of [a, b]. [a, b] is compact so let δ
be its Lebesgue number. Then, invoking use of the Lebesgue number at a, a + δ/2, a + 2δ/2, . . . gives
[a, a+δ/2] ⊂ A, [a+δ/2, a+2δ/2] ⊂ A, [a+2δ/2, a+3δ/2] ⊂ A . . . until we come to an interval containing
b. Hence b ∈ A and A, B are not disjoint.
The following proposition is very useful and allows us to easily show things are connected.
Proposition 4.9. The continuous image of a connected space is connected.
Proof. Let f : T → S be continuous with T connected. Suppose f (T ) is disconnected. Then there are
open U, V ⊂ S separating f (T ). i.e f (T ) ⊂ U ∪ V , U ∩ V ∩ f (T ) = ∅, U ∩ f (T ) 6= ∅, V ∩ f (T ) 6= ∅.
Then f −1 (U ), f −1 (V ) are non empty, open in T (since U, V are open in S and f is continuous), disjoint
(since U and V are) and cover T (since U , V cover f (T )). This contradicts T being connected.
MA222 Metric Spaces 11

Remark 4.10. Together with the description of connected subsets of R this implies the Intermediate
Value Theorem: the image of a continuous function on an interval is another interval.
Example 4.11. Curves in the plane, and graphs of continuous functions on intervals, are connected
subsets of R2 .
The following proposition is very useful for showing connectedness.
Proposition 4.12 (Octopus). If C and Cj , S
j ∈ J, are connected subspaces of a topological space T and
if Cj ∩ C 6= ∅ for each j ∈ J then K = C ∪ j∈J Cj is connected.

Corollary 4.13. If C ⊂ T is connected and C ⊂ K ⊂ C then K is connected.


S
Proof. K = C ∪ x∈K {x} and {x} ∩ C 6= ∅ for each x ∈ K.
Proposition 4.14. The product of two connected spaces is connected.
Proof. Let T, S be connected and s0 ∈ S. Denote C = T × {s0 } and Ct = {t} × S for each t ∈ T . Then
for each t ∈ T , C and Ct , being homeomorphic
S to T and S respectively, are connected. Since Ct ∩ C 6= ∅
(it contains (t, s0 )) and T × S = C ∪ t∈T Ct , we see that T × S is connected.
Example 4.15. R2 = R × R is connected. Circles are connected because they are continuous images of
intervals. R2 \ {0} is connected because it is the union of circles about (0, 0) each of which meets the
positive x axis.
Example 4.16. The topologist’s sine curve, the graph of sin(1/t) together with the vertical segment I
from (0, −1) to (0, 1), is connected.
Definition 4.17. Let T be a topological space. Write x ∼ y if x and y belong to a common connected
subspace of T . Then ∼ is an equivalence relation on T and the equivalence classes are called the connected
components of T .
Remark 4.18. A component containing x is the union of all connected subsets of T that contain x and
is the largest connected subset containing x. If C is connected then C is connected so the connected
components are closed. Components may also be defined as maximal connected components of X.
Connectedness is a topological invariant and hence the number of components is also a topological
invariant. The property that T \ {x} is connected for every x ∈ T is also a topological invariant.
Example 4.19. The components of (0, 1)∪(1, 2) are (0, 1) and (1, 2). The sets (0, 1), (0, 1)∪(1, 2), (0, 1)∪
(1, 2) ∪ (2, 3) are mutually non-homeomorphic as they have different numbers of connected components.
[0, 1] is not homeomorphic to a circle since removing any point from [0, 1], e.g. [0, 1] \ { 21 } = [0, 21 ) ∪ ( 12 , 1]
makes it disconnected, while removing any point from a circle leaves it connected. R is not homeomorphic
to R2 since removing the origin from R makes it disconnected while R2 remains connected after removing
a point.
We introduce a sufficient, but not necessary, condition for a topological space T to be connected.
It is more easily visualised than the previous definition and can be useful for showing a given space is
connected.
Definition 4.20. Let a, b ∈ T . A path from a to b in T is a continuous map φ : [0, 1] → T such that
φ(0) = a and φ(1) = b.
Definition 4.21. T is called path connected if any two points in T can be joined by a path in T .
Proposition 4.22. A path connected topological space T is connected.
Proof. Let a ∈ T . For every x ∈ T the image Cx of aSpath from a to x is connected (continuous image
of an interval), and all the Cx contain a. Hence T = x∈T Cx is connected.
Example 4.23. The topologist’s sine curve S is not path connected.
The following shows connectedness and path connectedness are actually equivalent notions in open
subsets of Rn .
12 MA222 Metric Spaces

Theorem 4.24. Any two points of a connected open subset U of Rn may be joined by a (piecewise
linear) path in U . That is, any open subset U is path-connected.

Theorem 4.25. All connected components of an open subset U of Rn are open.

Proof. Let C be a component of U and x ∈ C. Find δ > 0 such that B(x, δ) ⊂ U . Since B(x, δ) is
connected and C is the union of all connected subsets of U that contain x we have B(x, δ) ⊂ C. Hence
C is open.

Theorem 4.26. A subset U of R is open if and only if it is a union of countably many disjoint open
intervals.

Proof. (⇐=) Any union of open intervals is open.


(=⇒) Let U ⊂ R be open and let Cj , j ∈ J be its components. Since each Cj is connected and open,
these are open intervals. As components they are mutually disjoint. Hence it suffices to show that J can
be ordered into a sequence. For each j there is a rational number rj ∈ Cj . Since the Cj are disjoint, the
map j 7→ rj is an injection J → Q. Hence there is a surjection Q → J. Since Q is countable, it follows
that J is also countable, as required.

5 Completeness
Definition 5.1. A sequence (xn ) in a metric space M is Cauchy if ∀ ε > 0, ∃ k ∈ N such that n, m ≥
k =⇒ d(xn , xm ) < ε.

Definition 5.2. We say a metric space M is complete if every Cauchy sequence in M converges (to a
point of M ).

Intuitively a metric space is complete if it doesn’t have any holes. Q is not complete because 2 (for
example) is missing, though we can choose a Cauchy sequence which converges (in R) to it.

Remark 5.3. Completeness is not a topological invariant. R is complete while (0, 1), homeomorphic to
it, is not. Completeness is a property of the metric, not of the topology.

Proposition 5.4. A convergent sequence is Cauchy.

Proof. Let ε > 0. Then there is nε such that n ≥ nε =⇒ d(xn , x) < ε/2. Hence m, n ≥ nε =⇒
d(xm , xn ) ≤ d(xm , x) + d(xn , x) < ε.

Recall that a subset S of a metric space M is closed if and only if for every sequence xn ∈ S converging
to a point x ∈ M we have x ∈ S.

Proposition 5.5. A complete subspace S of any metric space M is closed.

Proof. Suppose xn ∈ S and xn → x ∈ M . (xn ) is a Cauchy sequence in M , hence also in S. Thus xn


converges in S to some y ∈ S. But xn → y also in M . Hence x = y by uniqueness of limits, and so
x ∈ S. Therefore S is closed.

Proposition 5.6. A closed subset S of a complete metric space M is complete.

Proof. Let (xn ) be a Cauchy sequence in S. Then it is Cauchy in M so converges to a point of M , which,
since S is closed, belongs to S. Hence xn → x ∈ S.

Proposition 5.7. A sequence (xn ) in a metric space M is Cauchy if and only if there is a sequence
εn ≥ 0 such that εn → 0 and d(xm , xn ) ≤ εn for m > n.

Proposition 5.8. Let (xn ) be a sequence in a metric space M for which there are τn ≥ 0 such that
P ∞
n=1 τn < ∞ and d(xn , xn+1 ) ≤ τn for each n. Then the sequence xn is Cauchy.

Proposition 5.9. For any set S, the space B(S) of bounded real-valued functions on S with the norm
kf k = supx∈S |f (x)| is complete.
MA222 Metric Spaces 13

Proof. Let (fn ) be a Cauchy sequence. Given ε > 0 there is nε such that m, n ≥ nε =⇒ supx∈S |fm (x)−
fn (x)| < ε. Hence for each fixed x ∈ S there sequence (fn (x)) is Cauchy in R so converges to some f (x) ∈
R. Let n ≥ nε . Then taking the limit m → ∞ in supx∈S |fm (x) − fn (x)| < ε gives kf (x) − fn (x)k ≤ ε
for every x ∈ S, n ≥ nε . Hence (fn ) converges (in supremum norm) to some f ∈ B(S).
Example 5.10. If K is a compact topological space then the space C(K) of continuous functions on K
with the norm kf k = maxx∈K |f (x)| is complete.
Definition 5.11. Let M be a metric space. A set S ⊂ M is dense in M if S = M .
Definition 5.12. A completion of M is a complete metric space N such that M is a dense subspace of
N.
The more modern definition is that a completion of M is a complete metric space N together with
an isometry i of M onto a subset of N such that i(M ) is dense in N .
Example 5.13. Q is dense in R so R is a completion of Q.
Theorem 5.14. Any metric space can be isometrically embedded into a complete metric space. In other
words, for any metric space M , there is an isometry from M onto a subset of a complete metric space.
Corollary 5.15. Any metric space has a completion.
Proof. Embed M into a complete metric space N . Then M (taking the closure in N ) is complete and
M is dense in M . Hence M is a completion of M .

Definition 5.16. Let f : S → S. A point x ∈ S such that f (x) = x is called a fixed point of f .
Definition 5.17. Let M be a metric space. A map f : M → M is a contraction if there is a constant
κ < 1 such that d(f (x), f (y)) ≤ κd(x, y) for every x, y ∈ M .
Theorem 5.18 (Banach’s Contraction Mapping Theorem). If f is a contraction on a complete metric
space M then f has a unique fixed point.

Proof. If f (x) = x and f (y) = y then d(x, y) = d(f (x), f (y)) ≤ κd(x, y) implies d(x, y) = 0, so x = y,
establishing uniqueness.
Choose x0 ∈ M and set xn+1 = f (xn ). Then

d(xj , xj+1 ) = d(f (xj−1 ), f (xj )) ≤ κd(xj−1 , xj ) ≤ . . . ≤ κj d(x0 , x1 )


P∞
Since j=1 κj d(x0 , x1 ) < ∞ the sequence (xn ) is Cauchy. As M is complete, xn converges to some x ∈
M . Since contractions are Lipschitz, hence continuous, we see that f (xn ) → f (x) and f (xn ) = xn+1 → x
so x = f (x) by uniqueness of limits.

The Contraction Mapping Theorem is useful in proving the existence of solutions for complicated
sets of equations, including differential equations, as well as solving them numerically.

Definition 5.19. The diameter of a non empty subset S of a metric space is defined by diam(S) =
supx,y∈S d(x, y).
Theorem 5.20 (Cantor’s Theorem). Let (Fn ) be a decreasing sequence of nonempty
T∞ closed subsets of
a complete metric space, i.e. F1 ⊃ F2 ⊃ · · · , such that diam(Fn ) → 0. Then n=1 Fn 6= ∅.
Proof. Choose xn ∈ Fn . For i ≥ n, xi ∈ Fi ⊂ Fn . For i, j ≥ n we have xi , xj ∈ Fn so d(xi , xj ) ≤
diam(Fn ). Since diam(Fn ) → 0 the sequence (xn ) is Cauchy so, by completeness of M T, converges to
some x ∈ X. Again xi ∈ Fn for i ≥ n so since Fn is closed, we have x ∈ Fn . Hence x ∈ ∞
n=1 Fn .

Definition 5.21. A subset S of a metric space M is said to be

• (everywhere) dense in M if S = M .
• nowhere dense in M if M \ S is dense in M .
14 MA222 Metric Spaces

• meagre in M (or of the first category in M ) if it is a union of countably many nowhere dense sets.

Example 5.22. In R one point sets are nowhere dense. Hence the countable set Q is meagre in R.
R \ Q = ∅ so Q is not nowhere dense in R.
Theorem 5.23 (Baire’s Category Theorem). A non empty complete metric space is not meagre in
itself.S In other words, if Sn are nowhere dense subsets of a non empty complete metric space M then

M \ n=1 Sn 6= ∅.
Proof. Let Gk = M \ Sk . Gk are dense in M (since Sk are nowhere dense), open (since Sk are closed) so
G1 6= ∅. Choose x1 ∈ G1 and use that G1 is open to find δ1 > 0 such that B(x1 , δ1 ) ⊂ G1 . In general
use the fact that Gk is dense to find xk ∈ Gk ∩ B(xk−1 , δk−1 /2) and then use that Gk is open to find
0 < δk < δk−1 /2 so that B(xk , δk ) ⊂ Gk . T∞
The choice of δk implies that δk → 0 and B(xk , δk ) ⊂ B(xk−1 , δk−1 ). So k=1 B(xk , δk ) 6= ∅ by
Cantor’s Theorem. Let x belong to this set. Then x ∈ B(xk+1 , δk+1 ) ⊂ B(xk , δk ) for each k, so x ∈ Gk
for each k and so x belongs to no Sk .
Baire’s theorem is very useful and can be used in proving existence. Generally if we want to show
that something exists then it suffices to show that the set of objects which are not of the required type
can be given as a union of countably many nowhere dense subsets of a complete metric space. This idea
can, for example, be used to show there is a continuous but nowhere differentiable function [0, 1] → R.
Definition 5.24. A metric space M is totally bounded if for any ǫ > 0 there is a finite set F ⊂ M such
that {B(f, ǫ) : f ∈ F } covers M .
Theorem 5.25. A metric space M is totally bounded iff every sequence in M has a Cauchy subsequence.

Theorem 5.26. A subspace C of a complete metric space M is compact iff closed and totally bounded.
Definition 5.27. A subset S of a metric space C(M ) = {continuous functionsf : M → R} is said to be:
• equicontinuous at x if for every ǫ > 0 there is δ > 0 such that |f (x) − f (y)| < ǫ whenever f ∈ S
and y ∈ B(x, δ).
• equicontinuous if it is equicontinuous at every x ∈ M .
• uniformly equicontinuous if for every ǫ > 0 there is δ > 0 such that |f (x) − f (y)| < ǫ whenever
f ∈ S and d(x, y) < δ.
Theorem 5.28 (Arzelà-Ascoli). Let M be a compact metric space. A subset of C(M ) is totally bounded
iff it is bounded and equicontinuous. proof: examinable, in notes.
Corollary 5.29. Consider a sequence of real-valued continuous functions fn defined on a closed and
bounded interval [a, b]. If this sequence of functions is uniformly bounded and equicontinuous, then there
exists a subsequence fnk which converges uniformly

Closing Remarks
As you can see, there’s quite a lot of material in Metric Spaces, and it’s easy to get confused with all
the abstract definitions. Being able to accurately state definitions and key theorems is important (and
easy marks in the exam), so learn them well. While a smattering of short proofs might be examined, it’s
more likely that you’ll be asked questions about simple examples to test your understanding, so know
how to apply the key theorems as well as what they state.
Above all though, practice is the best medicine, especially through past exam questions; the format
isn’t likely to change all that much compared to previous years. For even more practice questions, you
can turn to W. A. Sutherland’s excellent Introduction to Metric and Topological Spaces, which is very
reasonably priced and contains many examples as well as lots of questions at the right level; there’s
barely anything irrelevant in it, and even if it is irrelevant it will undoubtedly help your understanding.

Anda mungkin juga menyukai