Metrizability of Topological Spaces

43
Metrizability of Topological Spaces Narcisse Roland Loufouma Makala ([email protected]) African Institute for Mathematical Sciences (AIMS) Supervised by Hans-Peter K¨unzi University of Cape Town (UCT) May 24, 2007

description

metrizability

Transcript of Metrizability of Topological Spaces

Page 1: Metrizability of Topological Spaces

Metrizability of Topological Spaces

Narcisse Roland Loufouma Makala ([email protected])African Institute for Mathematical Sciences (AIMS)

Supervised by Hans-Peter KunziUniversity of Cape Town (UCT)

May 24, 2007

Page 2: Metrizability of Topological Spaces

Abstract

It is well known that the distance function or metric defined on a metric space X induces atopology on that space X. The goal of this essay is to study well-known characterizations of theclass of topologies that can be obtained from metrics in this way. We present some importantcriteria, which are necessary and sufficient, that topological spaces must possess in order to bemetrizable.

i

Page 3: Metrizability of Topological Spaces

Contents

Abstract i

1 Introduction 1

2 Metric and Topological Spaces 2

2.1 Metric Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2

2.2 Topological Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

2.3 Covering, Compact and Paracompact Spaces . . . . . . . . . . . . . . . . . . . 7

2.4 Continuity and Homeomorphism . . . . . . . . . . . . . . . . . . . . . . . . . . 8

3 Metrizability of Topological Spaces 10

3.1 Urysohn’s Metrization Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . 10

3.1.1 Urysohn’s Lemma. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

3.1.2 Tychonoff’s Theorem. . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

3.2 Nagata-Smirnov Metrization Theorem. . . . . . . . . . . . . . . . . . . . . . . 22

3.3 Bing’s Metrization Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

3.4 Pseudometrization Theorem. . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

3.5 Neighborhood Metrization Theorem . . . . . . . . . . . . . . . . . . . . . . . . 28

3.6 Frink’s Metrization Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

3.7 Alexandroff-Urysohn’s Metrization Theorem . . . . . . . . . . . . . . . . . . . . 29

3.8 Bing’s Metrization Theorem for Moore Spaces . . . . . . . . . . . . . . . . . . 30

4 Conclusion 32

A The Separation Axioms 33

Bibliography 40

ii

Page 4: Metrizability of Topological Spaces

1. Introduction

In any metric space (X, d), one can define open sets as follows :

O is open in X if and only if for all x ∈ O, there exists ε > 0 such that B(x, ε) ⊂ O. (∗)

If we put τ = {O| O is open in X in the sense given by (∗)}, then τ is a topology on X calledmetric topology or the topology induced by the metric d, and (X, τ) is called metric topologicalspace. This means that every metric space is a topological space. In comparaison with theobviousness of the last statement, in this essay, we study how is it possible to reverse thatstatement. This enables us to ask the question that given a topological space (X, τ), can wedefine a metric d on X such that the topology induced by d is τ ? if so, the space X is said tobe metrizable.

The main purpose of this paper is to give the answer to that question, since unfortunately, notall topological spaces are metrizable.

The organisation of this paper is as follows. We begin with preliminaries in Chapter 2, in whichwe give an overview of some background on metric and topological spaces that is pertinent toour purposes, that is to say notions we use in the next chapter.

In Chapter 3, which is the main part of this essay, we present some main results on metrizabilityof topological spaces, particularly the results due to Urysohn [4], Nagata and Smirnov [14], R.H.Bing [4], H.A. Frink [14], and Alexandroff and Urysohn [14]. we conclude in chapter 4.

1

Page 5: Metrizability of Topological Spaces

2. Metric and Topological Spaces

In this chapter, we only give a brief overview of some topological notions in metric and topologicalspaces which will be relevant to this essay. Thus, notions such as interior, derived set, boundaryof the set,. . . , and complete and connected spaces for example will not be presented. The readerwho is interested to these notions can find them in almost any book of general topology. (Seefor example [8], [6], [13] and [2]).

2.1 Metric Spaces

Definition 2.1.1. Let X be any non-empty set. A function d : X × X −→ R+ is said to be ametric or a distance on X if the following conditions are satisfied.

(O1) d(x, y) ≥ 0 and d(x, y) = 0 ⇐⇒ x = y for all x, y ∈ X (Positivity),

(O2) d(x, y) = d(y, x) for all x, y ∈ X (Symmetry),

(O3) d(x, y) ≤ d(x, z) + d(z, y) for all x, y, z ∈ X (Triangle inequality).

The pair (X, d) is called a metric space and the elements of X are called points.

Example 2.1.2.

• Let X = Rn. One can define on X the following distances: for any two points x = (x1, x2, ..., xn)and y = (y1, y2, ..., yn) of X;

d1(x, y) =

n∑

i=1

|xi − yi|, d2(x, y) =

n∑

i=1

(xi − yi)2, and d∞(x, y) = sup1≤i≤n

|xi − yi|.

These three distances are equivalent, i.e. there are real constants k1, k2, k3 such that

d∞(x, y) ≤ k1d2(x, y) ≤ k2d1(x, y) ≤ k3d∞(x, y).

• Let X = C the set of complex numbers. For z1, z2 in C,

d(z1, z2) = |z1 − z2| is a distance in C.

• Let X be any arbitrary non-empty set and let us define for x, y ∈ X

d(x, y) =

{

1 if x 6=y

0 if x=y .

One can see that d is a metric in X, called discrete metric and (X, d) the discrete metric space.

2

Page 6: Metrizability of Topological Spaces

Section 2.1. Metric Spaces Page 3

• If (X, d) is any metric space and Y any subset of X, then (Y, d) can also be considered as ametric space using the same distance as X. Y is then called a subspace of the metric space(X, d).

We define the distance from a point x of a metric space (X, d) to a subspace A of X by:

d(x, A) = infy∈A

d(x, y).

There are particularly interesting subsets in a metric space, namely the balls. One can define theopen ball and the closed ball with center x and radius r > 0 respectively by:

B(x, r) = {y ∈ X : d(x, y) < r} and B(x, r) = {y ∈ X : d(x, y) 6 r} .

Proposition 2.1.3. Given any ball B(x, r) in a metric space and a point y in B(x, r), thereexists δ > 0 such that B(y, δ) ⊆ B(x, r).

Proof. Let B(x, r) be a ball of a metric space X, y ∈ B(x, r) and suppose δ = r − d(x, y) > 0,then B(y, δ) ⊆ B(x, r).

The ball B(x, r) is called r-neighborhood of the point x in X.

Proposition 2.1.4. Let (Xn, dn), n ∈ N, be a countable family of metric spaces. Then theproduct of Xn countably many times,

n∈N

Xn, is a metric space.

Proof. Let (Xn, dn), n ∈ N, be a countable family of metric spaces. We define a metric on theproduct

n∈N

Xn as follow: let x and y be any points of∏

n∈N

Xn; let us denote by xn and yn the

nth coordinates of x and y respectively. Let us define

d′n (xn, yn) = min (dn (xn, yn) , 1) .

The metric d′n thus defined on Xn is bounded by 1 and is equivalent to the metric dn. The

distance defined by

d(x, y) =∑

n∈N

d′n (xn, yn)

2n(2.1)

is a metric on∏

n∈N

Xn.

Remark 2.1.5. The assumption that the metrics d′n are bounded by 1 and the factors 1

2n areneeded only to guarantee that the series given by (2.1) is convergent. For a finite sequenceX1, X2,...,Xn of metric spaces, one can define the distance between two points x = (xk) andy = (yk), 1 ≤ k ≤ n, on the set X = X1 × X2 × ... × Xn by

d(x, y) = d1 (x1, y1) + d2 (x2, y2) + ... + dn (xn, yn) .

Corollary 2.1.6. The Hilbert cube∏

n∈N

[0, 1], which is the product of [0, 1] countably many times,

is a metric space.

Page 7: Metrizability of Topological Spaces

Section 2.2. Topological Spaces Page 4

Open and Closed sets in metric spaces

Definition 2.1.7. A subset U of a metric space X is open if and only if it is union of open balls,or U is open in X if given any y ∈ U , there exists ε > 0 such that B(y, ε) ⊂ U .

Definition 2.1.8. A subset F of a metric space X is closed if and only if its complement1 X rFis open.

2.2 Topological Spaces

Definition 2.2.1. A topological space is a pair (X, τ) consisting of a non-empty set X and afamily τ of subsets of X satisfying the following conditions,

(T1) ∅ ∈ τ and X ∈ τ ,

(T2) The intersection of any two (and hence any finite number of) sets of τ is again in τ ,

(T3) The union of any family of sets of τ is again in τ .

The family τ is called a topology for X, and its members are called open sets of X. Hencethe statements “U ∈ τ” and “U is open in τ” mean the same thing. Elements of X are calledpoints.

Example 2.2.2.

• Let X be any non-empty space and τ1 = {∅, X}. Clearly, the axioms for a topology (T1), (T2), (T3)given above hold. τ1 = {∅, X} is a topology in X called the indiscrete topology.

• Let X be any non-empty set and let τ2 = P(E). Then τ2 is a topology in X called the discretetopology.

• Let X = {a, b, c} and τ = {∅, {b} , X}. It is easily verified that τ is a topology in X.

• The real line. Let X = R. Define the family τ as follows:

U ∈ τ ⇐⇒ for any x ∈ U, ∃δx > 0 such that u ∈ U if |x− u| < δx. τ is a topology called usualtopology in R.

Remark 2.2.3. A subset U of a topological space X is a neighborhood of the point x if andonly if there is an open set V such that x ∈ V ⊂ U . A neighborhood of a point need not tobe an open set, but every open set is a neighborhood of each of its points. Thus we have thefollowing characterization of open sets, like in the case of metric spaces.

Proposition 2.2.4. A subset U of a topological space X is open if and only if it is a neighborhoodof all its points. That is for all x ∈ U , we can always find an open set Ox such that x ∈ Ox ⊂ U .

1The complement of a subspace A of a metric space X is given by:X r A = {y ∈ X : y /∈ A}

Page 8: Metrizability of Topological Spaces

Section 2.2. Topological Spaces Page 5

The neighborhood system of a point is the family of all neighborhoods of that point.

Definition 2.2.5. A subset F of a topological space (X, τ) is said to be closed if its complementX \ F is open.

Therefore, since closed sets are complement of open sets, then taking the complements of prop-erties (T1),(T2) and (T3) from Definition 2.2.1, we have the following topological space axiomsof a family F of subsets of X called closed sets.

Proposition 2.2.6. Let X be any set and suppose that F is a family of closed subsets of X.The F has the following properties,

(F1) ∅ ∈ F and X ∈ F ,

(F2) the union of any two (and hence any finite number of) sets of F is again in F ,

(F3) The intersection of any family of sets of F is again in F .

Remark 2.2.7. From properties (T1), (T2), (T3) of open sets and (F1), (F2), (F3) of closed sets,we have the following result: in any topological space, the whole space X and the empty set ∅are both open and closed.

Definition 2.2.8. Subspace of a topological space and subspace topology.

Let (X, τ) be a topological space and G ⊂ X. Then the family O of all sets G⋂

U where Uis open in X satisfies axioms for a topology (T1), (T2), (T3) of the definition 2.2.1 given above.Indeed,

(T1) Since X and ∅ are in τ , then G⋂

X = G and G⋂ ∅ = ∅ are open in G.

(T2) Suppose U and V are open in G, then U = G⋂

U ′ and V = G⋂

V ′ where U ′ and V ′ areopen subsets of X. We have

U⋂

V =(

G⋂

U ′)

(

G⋂

V ′)

= G⋂

(

U ′⋂

V ′)

.

But U ′⋂

V ′ is an open subset of X; hence U⋂

V is an open subset of G.

(T3) Suppose {Ui}, i ∈ I, is a family of open subsets of G, then Ui = G⋂

U ′i where U ′

i is anopen subset of X for each i ∈ I, then

i∈I

Ui =⋃

i∈I

(

G⋂

U ′i

)

= G⋂

(

i∈I

U ′i

)

.

Since⋃

i∈I

U ′i is the union of open sets, it is open, and therefore

i∈I

Ui is open in G.

The family O = {G⋂U : U is open in X} of open sets in G is a topology on G called subspacetopology or induced topology. The set G with this topology is called a subspace of the topologicalspace X.

Page 9: Metrizability of Topological Spaces

Section 2.2. Topological Spaces Page 6

Remark 2.2.9. We can define by the same way a closed subset F of a subspace G of a topologicalspace X as the intersection of G with a closed set F ’ in X, i.e.

X is a topological space, G a subspace of X. Then F ⊂ G is closed in G ⇐⇒ F = G⋂

F ′,

where F ′ is closed in X.

We shall define now some other families of sets in topological spaces.

Definition 2.2.10. Base of a topological space (X, O).

A family B ⊂ O is called a base of the topological space (X, O) if every non-empty open subsetof X can be writen as the union of the members of B i.e.

B base of (X, O) if for every open set U ⊂ X, U =⋃

i∈I

Oi , where Oi ∈ O .

Definition 2.2.11. Subbase of a topological space (X, O).

Let (X, O) be a topological space. A subcollection S of O is said to be a subbase for O if theset

B = {B |B is the intersection of finitely many members of S }is a base for O .

Example 2.2.12. In R, the collection of all open intervals is a base for the usual topology.For a subbase, one can take the open intervals of the form (−∞, a) and (b,∞), since anyopen interval is either one of these or else the intersection of the two of them, i.e. if a < b,(a, b) = (−∞, b)

(a,∞); and any open set in R is a union of open intervals.

Definition 2.2.13. A collection of subsets B of a topological space X is said to be locally finite(respectively discrete) if for any x ∈ X, there is an open set U containing x which meets onlyfinetely many sets (respectively at most one set) in B i.e.

{

B ∈ B | B⋂

U 6= ∅}

is a finite subcollection of B (respectively is either empty or contains exactly one subset fromthe collection).

A collection of subsets B is said to be σ-locally finite (respectively σ-discrete) if it is countableunion of locally finite (respectively discrete) collection, i.e.

B =∞⋃

n=1

Bn,

where each Bn is locally finite (respectively discrete).

Page 10: Metrizability of Topological Spaces

Section 2.3. Covering, Compact and Paracompact Spaces Page 7

Definition 2.2.14. Closure of a subset A of X

For any A ∈ X, let us consider the family FA of all closed sets containing A. By axiom (F1),FA 6= ∅, and from (F3) it follows that the intersection

FA is closed. It is the smallest closedset containing A called the closure of A and represented by A.

Obviously, the subset F of X is closed if and only if F = F .

The closure of a subset F satifies the following properties;

Property 2.2.15. For any subsets A, B, C of a space X, we have;

(P1) A ⊂ A.

(P2) A ⊂ B =⇒ A ⊂ B.

(P3)¯A = A, A

B = A⋃

B, A⋂

B ⊂ A⋂

B.

Definition 2.2.16. Density

A subspace B of a topological space X is said to be dense in X if B = X, in other words, everypoint of B is limit of a net2 in X.

Example 2.2.17. The set Q of rational numbers if dense in the set R of real numbers. Indeed,each real number is a limit of some sequence of rational numbers.

2.3 Covering, Compact and Paracompact Spaces

Definition 2.3.1. Covering

A collection C = {Gα : α ∈ I} of subsets of a set X is said to be a cover of X if

α∈I

Gα = X.

If C1 and C2 are two covers of X such that C2 ⊂ C1, then C2 is called a subcover of C1.

Example 2.3.2. The collection C1 = {]−n, n[ : n ∈ N} is a cover of the set R of real numbersand C2 = {]−2n, 2n[ : n ∈ N} is a subcover of C1.

• A cover C of (X, τ) is said to be τ -open cover of X if every member of C is a τ -open set,

• A cover C of X is said to be finite if C has only a finite number of members.

2A net in a topological space X is a function from some directed set D into X

Page 11: Metrizability of Topological Spaces

Section 2.4. Continuity and Homeomorphism Page 8

Definition 2.3.3. Refinement.

Let V be a cover of X. The collection U is called a refinement of the cover V , or U refinesV , and we write U < V , if and only if U is a cover of X and each member of U is a subsetof a member of V i.e.

U < V ⇐⇒ for all U ∈ U , U ⊂ V where V ∈ V .

Notice that any open subcover is a refinement, but a refinement is not necessary an open subcover.If U is a cover of X and A ⊂ X, the star of A with respect to U is the set

St(A, U ) =⋃

{

U ∈ U | A⋂

U 6= ∅}

.

We say that U star-refines V , or U is a star-refinement of V , written U ∗ < V , if and only iffor each U ∈ U , there is some V ∈ V such that St (U, U ) ⊂ V . Let us recall that the star ofa one-point set {x} with respect to a cover U is called the star of the point x with respect toU , and is denoted by St(x, U ).

Finally, U is a barycentic-refinement of U , written U ∆V , provided the sets St(x, U ), for eachx ∈ X, refine V . Clearly, every star-refinement is a barycentric refinement and every barycentricrefinement is a refinement.

Definition 2.3.4. Compactness

A topological space (X, τ) is said to be compact if every τ -open cover of X has a finite subcover.

Example 2.3.5. For all real numbers a and b such that a < b, the closed interval [a, b] iscompact, while the open interval ]a, b[ is not (indeed, R itself is not compact).

Definition 2.3.6. Paracompactness.

A topological space (X, τ) is said to be paracompact if X is T2 and every open cover of X hasa locally finite open refinement.

2.4 Continuity and Homeomorphism

Let (X, O1) and (Y, O2) be two topological spaces and f a map from (X, O1) into (Y, O2).

Definition 2.4.1. The map f :(X, O1) −→ (Y, O2) is continuous if the inverse image of anyopen subset of Y is open in X, i.e.

f is continuous if f−1 (U) ∈ O1 for any U ∈ O2.

Since closed sets are complements of open sets, we have the following proposition.

Proposition 2.4.2. The inverse image of a closed subset of Y through a continuous map is aclosed set in X.

Page 12: Metrizability of Topological Spaces

Section 2.4. Continuity and Homeomorphism Page 9

Example 2.4.3.

∗ If X is a discrete space, then any map from X into a topological space Y is continuous.

∗ Any map from a topological space into an indiscrete space is continuous.

∗ For any set A, x 7−→ d(x, A) is a continuous function.

Definition 2.4.4. Homeomorphism

A map f : X −→ Y is called a homeomorphism if the following conditions hold,

∗ f is a continuous bijective map,

∗ the inverse map f−1 : Y −→ X is also continuous.

If there exists a homeomorphism from X into Y , we say that X and Y are homeomorphic ortopologically equivalent. Hence, the statement ‘X and Y are homeomorphic’ is an equivalencerelation. In other words, the homeomorphisms preserve all the structures that topological spacespossess3.

Proposition 2.4.5. If f is a bijective map from (X, O1) to (Y, O2), then the following statementsare equivalent;

(S1) f is a homeomorphism,

(S2) The subset F is open (respectively closed) in X if and only if f (F ) is open (respectivelyclosed) in Y ,

(S3) The subset G is open (respectively closed) in Y if and only if f−1 (B) is open (respectivelyclosed) in X.

Proof. The proposition is proved by using definitions 2.4.1 and 2.4.4, and the fact that if f is ahomeomorphism, then its inverse f−1 is also a homeomorphism and we have

A = f−1 (f (A)) and f (A) =(

f−1)−1

(A) .

3Because of that property, we will later see that if a topological space X is homeomorphic to a metric space,

then X is metrizable.

Page 13: Metrizability of Topological Spaces

3. Metrizability of Topological Spaces

Any metric space (X, d) gives rise to a topological space (X, τ), where τ is defined to be thecollection of all subsets which are open in the sense of Definition 2.1.7, Chapter 2. τ is then atopology on X called a metric topology induced by d. Suppose now instead of starting with ametric space, we are given a topological space (X, τ). We would like to know if there exits ametric d which can be defined on X such that the topology induced by d is τ? If so, the spaceX is said to be metrizable. Otherwise X is nonmetrizable.

The question now is; when is a topological space metrizable? A theorem which answers to thatquestion is called a Metrization Theorem. One of the most important Metrization Theorems isthe Urysohn’s Metrization Theorem which gives the criteria which suffice for metrizability.

3.1 Urysohn’s Metrization Theorem

Theorem 3.1.1. [4]. Let X be a normal space which is second countable. Then X is metrizable.

Let us first define the second axiom of countability.1

Definition 3.1.2. A space X is said to be second countable if it has a countable base, i.e. thereis a countable collection of open sets such that any open set can be expressed as a union of setsfrom this collection.

Example 3.1.3. Rn is second countable. We may take as a countable base, the collection of allballs of rational radius, centered at points having all coordinates rational.

However not all metrizable spaces are second countable.

Example 3.1.4. Let us consider an uncountable set with the discrete topology. This topologyarises from a metric (e.g. take the distance between distinct points to be always 1). Howeverany base for the discrete topology must contain all the singleton sets, and there are uncountablymany of them.

Proposition 3.1.5. Any subspace of a second countable space is second countable.

Proof. Let (X, τ) be a second countable space, B = {Bi}i∈Z+a countable base for τ , and B a

subspace of X. Then the family

BB ={

B⋂

O| O ∈ B

}

is a countable base for the subspace topology on B. Then B is second countable.

1A first-countable space is a topological space which has a countable base at each of its points.

10

Page 14: Metrizability of Topological Spaces

Section 3.1. Urysohn’s Metrization Theorem Page 11

The idea in the proof of Urysohn’s Metrization Theorem is to construct a homeomorphism fbetween the given topological space X, which is normal and second countable, and a subspaceof the Hilbert cube, which is metrizable (see Corolary 2.1.6, Chapter 2). Then f(X) will bemetrizable as a subspace of a metrizable space (see Example 2.1.2, Chapter 2) and so is X sincehomeomorphic to a metrizable space. Now constructing a map into a product is equivalent toconstructing a map into each factor. Thus the following lemma plays a central role in the proofof Urysohn’s Metrization Theorem:

3.1.1 Urysohn’s Lemma.

Lemma 3.1.6. Let A and B be disjoint closed subsets of a normal space X. Then we can finda continuous function

f : X −→ [0, 1]

such that f (a) = 0 for all a ∈ A and f (b) = 1 for all b ∈ B.

Proof. Let X be a normal space, then given any closed subset F of X and an open set H whichcontains F , we can always find an open set G (See the proposition A.0.22 in the Appendix) suchthat

F ⊂ G ⊂ Cl(G) ⊂ H. (3.1)

Let D denote the set of dyadic rational numbers, i.e. all numbers which can be written in theform

p =m

2n, where m and n are positive integers such that 0 ≤ m ≤ 2n.

We shall associate to each such dyadic rational number p an open set Up ⊆ X such that:

(i) A ⊂ Up,

(j) B⋂

Up = ∅, (3.2)

(k) if p < q, then Cl(Up) ⊆ Uq.

We will construct the Up’s by induction by using the characterization (3.1) of normality.

We start the induction by taking U1 = X \ B and applying (3.1) to the inclusion A ⊆ U1, weobtain an open set U0 satisfying

A ⊆ U0 ⊆ Cl(U0) ⊆ U1.

Similarly, using again (3.1) to the inclusions A ⊆ U0 and Cl(U0) ⊆ U1, there exist open sets U 1

2

and U 3

4

satisfying

A ⊆ U 1

2

⊆ Cl(U 1

2

) ⊆ U0,

Cl(U0) ⊆ U 3

4

⊆ Cl(U 3

4

) ⊆ U1.

Page 15: Metrizability of Topological Spaces

Section 3.1. Urysohn’s Metrization Theorem Page 12

We now have the following inclusions

A ⊆ U 1

2

⊆ Cl(U 1

2

) ⊆ U0 ⊆ Cl(U0) ⊆ U 3

4

⊆ Cl(U 3

4

) ⊆ U1.

We will continue by finding the Up’s by induction on n, the exponent of 2 in p = m2n . Let us

notice that we have already defined Up for n = 1 and n = 2. Now assume we have defined Up forn. We now define Up for n + 1 (and thus for m = 1, 3, . . . , 2n+1 − 1). Note that the definitionof Up needs to be given only for odd n; because if n were even, the numerator and denominatorof p could be divided by 2. Because the Up’s have already been constructed for p = m

2n , m odd,and since

Cl(U m−1

2n+1) = Cl(U m−1

22n

) , which has already been defined,

we haveCl(U m−1

2n+1) ⊆ U m+1

2n+1, because of the property (k) of (3.2).

Applying (3.1), we obtain an open set, which we let be U m

2n+1, such that

Cl(U m−1

2n+1) ⊆ U m

2n+1⊆ Cl(U m

2n+1) ⊆ U m+1

2n+1.

We thus have an inductive definition of Up for each p in D⋂

[0, 1] as described. By construction,the collection of Up have properties (i) through (k) given by (3.2).

We extend the definition to dyadic rationals outside the interval [0, 1] by defining

Up = ∅ if p < 0 Up = X if p > 1,

and (3.2) obviously still holds.

We now define the function f : X −→ [0, 1] by f(x) = inf {p ∈ D| x ∈ Up}.Note that since A ⊆ Up for all p, then if a ∈ A,

f(a) = inf {p ∈ D| p ≥ 0} = 0.

Similarly, since Up ⊆ U1 for all p and Up

B = ∅ for all p, then if b ∈ B

f(b) = inf {p ∈ D| p > 1} = 1.

All that remains to be proved is that f is continuous, and to do that, we will use the followinglemma.

Lemma 3.1.7. Let P be a subset of the real numbers and let S be a set. Suppose that for eachp ∈ P , there is associated a subset Fp ⊆ S and that

p∈P

Fp = S.

If we define a function f : S −→ R by

f(s) = inf {p ∈ P | s ∈ Fp} ,

then for any real number c

f−1((−∞, c)) =⋃

p<c

Fp.

If we assume in addition that

Page 16: Metrizability of Topological Spaces

Section 3.1. Urysohn’s Metrization Theorem Page 13

(a) P is dense in R,

(b) Fp ⊂ Fq whenever p < q,

then we also have f−1((−∞, c]) =⋂

p>c

Fp.

Proof. Since for any set of real numbers T

inf T < x ⇐⇒ ∃t ∈ T such that t < x,

we have

f−1((−∞, c)) = {s| inf {p ∈ P |s ∈ Fp} < c}= {s|∃p such that s ∈ Fp and p < c}=⋃

p<c

Fp.

For the second part, we have:

f−1((−∞, c]) = f−1

(

d>c

(−∞, d)

)

=⋂

d>c

f−1 ((−∞, d))

=⋂

d>c

(

p<d

Fp

)

.

Thus it remains to show that⋂

d>c

(

p<d

Fp

)

=⋂

q>c

Fq.

First of all given any d > c, by density of P , we can find a q0 ∈ P such that c < q0 < d. Thus

q>c

Fq ⊆ Fq0⊂⋃

p<d

Fp.

Since this is true for all d, it follows that

q>c

Fq ⊆⋂

d>c

(

p<d

Fp

)

.

On the other hand, given any q ∈ P such that q > c, we have by (b):

d>c

(

p<d

Fp

)

⊆⋃

p<q

Fp ⊆ Fq.

Page 17: Metrizability of Topological Spaces

Section 3.1. Urysohn’s Metrization Theorem Page 14

The two inclusions thus establish

d>c

(

p<d

Fp

)

=⋂

q>c

Fq.

This conclude the proof of the lemma.

Let us go back to the proof of Urysohn’s Lemma. To check that f : X −→ [0, 1] is continuouswe think of f as a function f : X −→ R. Thus we have to check that the inverse images underf of the subbasic open sets (−∞, c) and (c,∞) are open in X for all c. By the preceding lemma,we have

f−1((−∞, c)) =⋃

p<c

Up,

which is obviously open.

We also havef−1((c,∞)) = f−1(R \ (−∞, c]) = X \ f−1((−∞, c]).

Now by the preceding lemma

f−1((−∞, c]) =⋂

p>c

Up,

and it suffices to show that this is closed in X. This will follow if we can show that⋂

p>c

Up =⋂

p>c

Cl(Up).

Clearly the left hand side is contained in the right hand side, i.e.⋂

p>c

Up ⊂ ⋂

p>c

Cl(Up). To

show the reverse inclusion, let us choose for each p > c, another dyadic rational λ(p) such thatc < λ(p) < p. Then we have

Cl(Uλ(p)) ⊆ Up,

from which follows⋂

p>c

Cl(Up) ⊆⋂

p>c

Cl(Uλ(p)) ⊆⋂

p>c

Up.

This conclude the proof of Urysohn’s Lemma.

Remark 3.1.8.

• It is a common error to read Urysohn’s Lemma as saying that

A = f−1({0}) B = f−1({1}).

This however is wrong2. All Urysohn’s Lemma says is that

A ⊆ f−1({0}) B ⊆ f−1({1}).2In order to have equalities, the subsets A and B have to satisfy some additional conditions given in the next

Lemma 3.1.9.

Page 18: Metrizability of Topological Spaces

Section 3.1. Urysohn’s Metrization Theorem Page 15

• Conversely, if a T1-space X has the property that for each pair A and B of disjoint closedsubsets there exists such a function, then X is normal. Indeed the sets U0 =

{

x| 0 ≤ x < 12

}

and V0 ={

x| 12

< x ≤ 1}

are two disjoint open subsets of [0, 1]; therefore, since f is continuous,then the sets U = f−1(U0) and V = f−1(V0) are two disjoint open subsets of X such thatA ⊂ U and B ⊂ V .

Let us give another lemma which will be useful in the proof of Urysohn’s Metrization Theorem.

Lemma 3.1.9. If S is a subset of a normal space X, then S is the preimage of a point under acontinuous map f : X −→ R if and only if S is both closed and Gδ.

Proof. First off all, let us suppose that X is normal, f : X −→ R is continuous and S a subsetof X such that S = f−1({c}), c ∈ R. Then S being the continuous preimage of the closed set{c} is also closed. Furthermore, we have

S = f−1({c})

= f−1

(

∞⋂

n=1

(

c − 1

n, c +

1

n

)

)

=

∞⋂

n=1

f−1

((

c − 1

n, c +

1

n

))

.

i.e. S is Gδ. Conversely let us suppose that S is closed and Gδ. Then

S =

∞⋂

n=1

On,

for some open sets On. By Urysohn’s Lemma, we can find continuous functions gn : X −→ [0, 1],such that

S ⊆ g−1n ({0}) and X \ On ⊆ g−1

n ({1}). (see Remark 3.1.8).

Then we can define a continuous map into the Hilbert cube:

g :X −→∞∏

n=1

[0, 1]

x 7−→ (g1(x), g2(x), g3(x), . . .) .

It is clear that the inverse image under g of (0, 0, 0, . . .) is S.

Now let us compose g with the continuous map

τ :∞∏

n=1

[0, 1] −→ R,

where τ denotes the distance from the point (0, 0, 0, . . .). Then

f = τ ◦ g

Page 19: Metrizability of Topological Spaces

Section 3.1. Urysohn’s Metrization Theorem Page 16

and clearly

f−1({0}) = (τ ◦ g)−1({0})= g−1

[

τ−1({0})]

= g−1(0, 0, 0, . . .)

= S.

Remark 3.1.10. In the proof of Urysohn’s Metrization Theorem, we can choose f : X −→ [0, 1]rather than R, with f−1 ({0}) = S.

Remark 3.1.11. The Urysohn’s Metrization Theorem, which says that if the space X is T4 andsecond countable then X is metrizable can now be stated in a stronger form and we have thefollowing theorem.

Theorem 3.1.12. If the space X is T3 and second countable, then X is metrizable.

The connection between these two forms of Urysohn’s Metrization Theorem is provided by thefollowing theorem, due to Tychonoff.

3.1.2 Tychonoff’s Theorem.

Theorem 3.1.13. If the space X is T3 and second countable, then X is T4.

Proof. Let X be a T3-space which is second countable with a countable base B = {Bi}i∈Z+.

Let A and B be two disjoint closed subsets of X. If x ∈ A, then X \ B is an open set whichcontains x. Then, there is a neighborhood U of x such that x ∈ U ⊂ X \ B. Since X is T3,then we can find a neighborhood V of x such that Cl(V ) ⊂ U (see Proposition A.0.17 in theAppendix). Finally, there is a member Bi of B containing x such that Bi ⊂ V . Now

Cl(Bi) ⊂ Cl(V ) ⊂ U ⊂ X \ B. so Cl(Bi)⋂

B = ∅.

If we repeat this procedure for each x ∈ A, we will have a countable subcollection C = {Cj}j∈Z+

of B which covers A and which satisfies

Cl(Cj)⋂

B = ∅ for all j.

Similarly, if x ∈ B, we can find a countable subcollection D = {Dk}k∈Z+of B which covers B

and which satisfiesCl(Dk)

A = ∅ for all k.

It is clear thatA ⊂

j∈Z+

Cj and B ⊂⋃

k∈Z+

Dk.

Page 20: Metrizability of Topological Spaces

Section 3.1. Urysohn’s Metrization Theorem Page 17

Let us notice that⋃

j∈Z+

Cj and⋃

k∈Z+

Dk can be considered as the neighborhoods of A and B

respectively, but the problem is that their intersections may not be disjoint. For this reason, weneed to construct new collections whose intersections will be disjoint. Thus, for each j and k,let us define

C ′j = Cj −

j⋃

k=1

Cl(Dk) and D′k = Dk −

k⋃

j=1

Cl(Cj).

which are open since equal to an open set minus a closed set. Let us consider

W =⋃

j∈Z+

C ′j and Z =

k∈Z+

D′k.

Then W and Z are open sets since they are unions of open sets.

Let us suppose a ∈ A, then a ∈ Cj for some j since A ⊂ ⋃

j∈Z+

Cj, and

C ′j = Cj −

j⋃

k=1

Cl(Dk).

But for all k, Cl(Dk)⋂

A = ∅, this implies a /∈ Cl(Dk) for all k i.e. a /∈j⋃

k=1

Cl(Dk).

Finally, if a ∈ A, then a ∈ Cj and a /∈j⋃

k=1

Cl(Dk) i.e. we have

a ∈ Cj −⋃j

k=1 Cl(Dk) = C ′j ⊂

j∈Z+

C ′j = W . Hence A ⊂ W .

Similarly, one can easily show that B ⊂ Z.

To conclude the proof, it remains to show that W and Z are disjoint.

Let us suppose that x ∈ W⋂

Z, then x ∈ C ′j

D′k for some j and k. This implies that x ∈ C ′

j

and x ∈ D′k for some j and k.

But x ∈ C ′j =⇒ x ∈ Cj and x ∈ D′

k =⇒ x /∈ Cl(Cj) ⊃ Cj. We have a contradiction.

Thus W⋂

Z = ∅. Finally, W and Z are two disjoint open sets which contain A and B respec-tively. X is then T4.

We also have the criterion of metrizability of topological spaces with a weaker separation axiomtogether with some additional conditions.

Lemma 3.1.14. A compact Hausdorff space is normal, i.e.

if X is T2 and compact, then X is T4.

Proof. First of all, we show that

if X is T2 and compact, then X is T3.

Page 21: Metrizability of Topological Spaces

Section 3.1. Urysohn’s Metrization Theorem Page 18

Let us suppose that X is T2 and compact, A a closed subset of X and x /∈ A. X being T2, thenfor each a ∈ A, there are disjoint open sets Ua and Va containing x and a respectively. Then{Va}a∈A is an open cover of A, which is also compact (being closed in X). Thus there is a finitesubcover of {Va}a∈A such that

A ⊆k⋃

i=1

Vai.

Thenk⋂

i=1

Uaiis an open set containing x which is disjoint from

k⋃

i=1

Vai. Thus X is T3.

We next prove thatif X is T3 and compact, then X is T4.

Let us suppose now X is T3 and compact, A and B two disjoint closed subsets of X.Let x /∈A,then X being T3, there are disjoint open sets U and V such that x ∈ U and A ⊂ V . Butfor each x /∈ A, the family {Ux}x∈X\A such that Ux

V = ∅ is an open cover of B which iscompact. Thus there is a finite subcover of {Ux}x∈X\A such that

B ⊆p⋃

k=1

Uxk.

p⋃

k=1

Uxkand V are two open disjoint subsets of X containing B and A respectively.

Thus X is T4 and this concludes the proof.

Hence, by combining the above Lemma with the Theorem 3.1.1, we have the following result forthe metrizability of Hausdorff spaces,

Theorem 3.1.15. If X is a second countable compact Hausdorff space, then X is metrizable.

Before giving the next result, lets us define another notion which is similar to compactness, butmore closely related to second countability.

Definition 3.1.16. A topological space X is said to be Lindelof if every open cover of X has acountable subcover. A topological space X is said to be hereditarily3 Lindelof if every subspaceof X is also Lindelof.

Then we have the following lemma.

Lemma 3.1.17. A topological space X which is normal and hereditarily Lindelof is perfectlynormal i.e.

If X is T4 and hereditarily Lindelof, then X is T6.

3A property of a space is hereditary if each of its subspaces possesses this property.For example being second

countable is a hereditary property.

Page 22: Metrizability of Topological Spaces

Section 3.1. Urysohn’s Metrization Theorem Page 19

Proof. Let A be a closed subset of a normal hereditarily Lindelof space X. For each x ∈ X \A,let us choose two open disjoint sets Ux and Vx containing x and A respectively. Then the family{Vx}x∈X\A is an open cover of X \A. But X \A is also Lindelof , being a subpace of X whichis hereditarily Lindelof. Thus there is a countable subcover of {Vx}x∈X\A such that

X \ A =

∞⋃

i=1

Vxi.

Then

A = X \(

∞⋃

i=1

Vxi

)

=∞⋂

i=1

Uxi,

Where the Uxi’s are open. Thus A is Gδ i.e. X is T6.

From the above lemma, we notice that: We used the weaker T3 separation axiom, instead ofT4, together with the hereditary Lindelof condition to prove that all closed sets are Gδ. But letus recall that T6 requires T4 as part of the definition. However Tychonoff’s Lemma, alluded toabove actually shows that

If X is T3 and Lindelof, then X is T4,

so that indeed T4 can be replaced by T3 in the above lemma.

Lemma 3.1.18. Any second countable space X is hereditarily Lindelof.

Proof. Let X be a second countable space. Since any subspace of a second countable space issecond countable, it suffices to show that X is Lindelof.

Let {Un}n∈Z+be a countable base for X and let {Vi}i∈I be any open cover of X. Let us next

consider the subcollection{

Unj

}

j∈Z+of the base consisting of those Un’s which are contained

in some open set Vi, in the given open cover. Then for each such Unj, let us pick a particular

Vi(nj) containing it (i.e. Unj). Then

X =⋃

i∈I

Vi =

∞⋃

j=1

Unj⊆

∞⋃

j=1

Vi(nj).

It follows that{

Vi(nj)

}

j∈Z+is a countable subcover of X, then X is Lindelof.

Let us now prove the Urysohn Metrization Theorem :

Proof. Urysohn Metrization Theorem.

Let {Un}n∈Z+be a countable base for the topology on X. Then by the above lemmas we have

shown that

If X is T4 and second countable =⇒ X is T4 and hereditarily Lindelof

=⇒ X is T6.

Page 23: Metrizability of Topological Spaces

Section 3.1. Urysohn’s Metrization Theorem Page 20

Thus the closed sets X \ Un are Gδ and we can find continuous functions (see Lemma 3.1.9)

fn : X =⇒ [0, 1] ,

such thatf−1

n ({0}) = X \ Un.

Let us consider the function

f :X −→∞∏

n=1

[0, 1]

x 7−→ (f1(x), f2(x), f3(x), . . .) .

It is clear that f is continuous since each of its coordinate functions fn is continuous.

f is bijective. Indeed, let x, y ∈ X such that x 6= y. Then we can find a basic open set Un suchthat x ∈ Un and y /∈ Un. Then fn(x) > 0 while fn(y) = 0. Thus f(x) 6= f(y), since they differin their nth coordinates.

In order to show that the restriction of f

f : X −→ f(X)

is a homeomorphism, it suffices to show that for any open set V ⊂ X, the direct image f(V ) isopen in f(X). Since direct images take unions to unions, it suffices to prove this when V is oneof the basic open sets Um. Let us consider the projection

pm :∞∏

n=1

[0, 1] −→ [0, 1]

into the mth coordinate. Then we have

Um = f−1m ((0, 1]) = (pm ◦ f)−1 ((0, 1]) = f−1

(

p−1m ((0, 1])

)

.

Applying f to this equation, we have

f(Um) = f(

f−1(

p−1m ((0, 1])

))

= f(X)⋂

p−1m ((0, 1]) ,

which is an open set in f(X).

Thus X is homeomorphic to the subset f(X) of the Hilbert cube∞∏

n=1

[0, 1] which is metrizable.

It follows that X is also metrizable and the topology on X arises from the metric d given by

d (x, y) = d (f (x) , f (y)) , for x,y ∈ X,

where d is the distance application on the Hilbert cube. This conclude the proof of Urysohn’sMetrization Theorem.

Page 24: Metrizability of Topological Spaces

Section 3.1. Urysohn’s Metrization Theorem Page 21

The Urysohn’s Metrization Theorem gives only sufficient conditions for the metrizability of topo-logical spaces, but does not give necessary conditions. The next metrization theorem, provedindependently by Nagata and Smirnov in 1950, gives a complete answer to the metrization tho-erem. But before giving the Nagata-Smirnov Metrization Theorem, let us first prove a theoremdue to A.H. Stone, and after define the generalized Hilbert space, which will be used in the proofof Nagata-Smirnov Metrization Theorem.

Theorem 3.1.19. [3]. The Stone’s Theorem

Every open cover of a metrizable space has an open refinement which is both locally finite andσ-discrete i.e. every metrizable space is paracompact and has a σ-discrete refinement.

Proof. Let X be a metrizable space and {Us}s∈S be an open cover of X. Let us take a metricd on X and a well-ordering relation < on the set S. Let us define families Vi = {Vs,i}s∈S

ofsubsets of X by letting

Vs,i =⋃

B(c,1

2i),

where the union is taken over all points c ∈ X satisfying the following conditions:

∗ s is the smallest element of S such that c ∈ Us , (1)

∗ c /∈ Vt,j for j < i and t ∈ S, (2)

∗ B(c, 32i ) ⊂ Us. (3)

It follows from the definition that the sets Vs,i are open, and (3) implies that Vs,i ⊂ Us.

Let x be a point of X; let us take the smallest s ∈ S such that x ∈ Us and a natural numberi such tat B(x, 3

2i ) ⊂ Us. Clearly, we either have x ∈ Vt,j for a j < i and t ∈ S or x ∈ Vs,i.

Hence, the union V =∞⋃

i=1

Vi is an open refinement of the cover {Us}s∈S.

We shall prove that for every i

if x1 ∈ Vs1,i , x2 ∈ Vs2,i and s1 6= s2, then d(x1, x2) >1

2i(4),

and this will show that the families Vi are discrete, because every 12i -ball meets at most one

member of Vi.

Let us assume that s1 < s2. By definition of Vs1,i and Vs2,i there exist points c1 and c2 satisfying(1)− (3) such that xk ∈ B(ck,

12i ) ⊂ Vsk,i for k = 1, 2. From (3) it follows that B(c1,

32i ) ⊂ Us1

and from (1) we see that c2 /∈ Us1, so that d(c1, c2) ≥ 3

2i . Hence,

d(c1, c2) ≤ d(c1, x1) + d(x1, x2) + d(x2, c2) i.e.

d(x1, x2) ≥ d(c1, c2) − d(c1, x1) − d(c2, x2) >1

2i,

which proves (4).

Page 25: Metrizability of Topological Spaces

Section 3.2. Nagata-Smirnov Metrization Theorem. Page 22

To conclude the proof of the theorem, it suffices to show that for every t ∈ S and any pair k, jof natural numbers

if B(x,1

2k) ⊂ Vt,j , then B(x,

1

2i+k)⋂

Vs,i = ∅ for i ≥ j + k and t ∈ S, (5)

because for every x ∈ X there exist j, k and t such that B(x, 12k ) ⊂ Vt,j and thus the ball

B(x, 12j+k ) meets at most j + k − 1 members of V and this will show that the families Vi are

locally finite.

It follows from (2) that the points c in the definition of Vs,i do not belong to Vt,j wheneveri ≥ j + k; since B(x, 1

2k ) ⊂ Vt,j, we have d(x, c) ≥ 12k for any such c. The inequalities

j + k ≥ k + 1 and i ≥ k + 1 imply that B(x, 12j+k )

B(c, 12i ) = ∅, and this yields (5).

Let us now define the generalized Hilbert space4

Definition 3.1.20. Let τ be an infinite cardinal number. The generalized Hilbert space of weightτ , Hτ , is described as follows:

Let A be an index set of cardinal τ . Then Hτ consists of all functions x : A −→ R such that

a) x(a) = xa 6= 0 for at most countably many a ∈ A,

b)∑

a∈A

x2a converges.

Note that the sum (b) makes sense, since it is really a countable sum. The distance function inHτ is defined, just as it was in the Hilbert space H, by

d(x, y) =√

(xa − ya)2. [14]

3.2 Nagata-Smirnov Metrization Theorem.

Theorem 3.2.1. [14] , [10]. Let X be a topological space. Then the following two conditionsare equivalent:

(a) X is metrizable.

(b) X is T3 and has a σ-locally finite base.

Proof. Let X be a metrizable space and for each n ∈ N, let Cn ={

B(x, 12n )|x ∈ X

}

the opencover of X by 1

2n -balls about x. Then, according to the Stone’s Theorem, X is paracompact.

Let Vn be a locally finite open refinement of Cn. One can easily verify that the family D =∞⋃

k=1

Vk

is a σ-locally finite base for X. Since every metric space is T3, then necessity is proved.

4The real Hilbert space H is the collection of all real sequences x = (x1, x2, . . .) such that∑

x2

k< ∞, with

the metric d(x, y) =√∑

(xk − yk)2.

Page 26: Metrizability of Topological Spaces

Section 3.2. Nagata-Smirnov Metrization Theorem. Page 23

Conversely, let us consider a T3-space which has a σ-locally finite base B =∞⋃

n=1

Bn, where the

Bn’s are locally finite. It is apparent that X is paracompact, since every open cover of X hasa σ-locally finite refinement consisting of basis elements, and hence X is normal. (see TheoremA.0.24 in the Appendix.)

Next we show that X is perfectly normal. Let G be open in X. By regularity, for each x ∈ G,there is a basis element Bx such that Cl(Bx) ⊂ G. Let

Bn =⋃

{Cl(Bx)| Bx ∈ Bn} .

Then Bn is the union of a locally finite collection of closed sets and hence is closed, and G =∞⋃

n=1

Bn. Thus every open set in X is Fσ, so X is perfectly normal.

Now each basis element Bnα has the property that for some continuous functionfnα : X −→ [0, 1],

Bnα = {x ∈ X| fnα(x) 6= 0} , by perfect normality.

Let τ be the cardinal number of the base B, and let Hτ be the generalized Hilbert space ofweight τ ; we can use the pairs n, α as the index set A in the definition of Hτ . Let us defineF : X −→ Hτ by giving coordinate functions Fnα(x) = [F (x)]nα as follows:

Fnα(x) =1

(√2)n

fnα(x)√

1 +∑

β

f 2nβ(x)

.

The denominator here makes sense because for any x in X, x ∈ Bnα for only finitely manyBnα ∈ Bn, so that fnα(x) 6= 0 for only finitely many α, if n is fixed. This also shows thatFnα(x) 6= 0 for only countably many pairs n, α. Since

α

F 2nα(x) <

1

2n,

we find that∑

n,α

F 2nα(x) <

n

1

2n= 1,

so that F (x) is indeed an element of Hτ . We claim F is a homeomorphism of X with a subsetof Hτ .

First, if x 6= y in X, then for some Bnα ∈ B, x ∈ Bnα and y /∈ Bnα. Then fnα(x) 6= 0 andfnα(y) = 0, from which it follows that Fnα(x) 6= Fnα(y), and hence F (x) 6= F (y).Thus F is bijective.

Let us now prove that F is continuous. First, note that each Fnα is continuous as a map of Xinto R. Now let x0 ∈ X and ε > 0 be given. Choose N so large such that

∞∑

n=N+1

1

2n<

ε2

4.

Page 27: Metrizability of Topological Spaces

Section 3.2. Nagata-Smirnov Metrization Theorem. Page 24

Now let U be a neighborhood of x0 meeting only finitely many Bnα for n ≤ N ; say, U meetsBn1α1

, Bn2α2, . . . , Bnkαk

. Let V ⊂ U be a neighborhood of x0 such that for x ∈ V ,

| Fniαi(x) − Fniαi

(x0) |<ε√2k

,

for i = 1, . . . , k. Now for x ∈ V and any pair n, α other than ni, αi for i = 1, . . . , k, we haveFnα(x) = Fnα(x0) = 0. Hence, for x ∈ V ,

n≤N

α

| Fnα(x) − Fnα(x0) |2=k∑

i=1

| Fniαi(x) − Fniαi

(x0) |2<ε2

2.

But we also have

n>N

α

| Fnα(x) − Fnα(x0) |2 ≤∑

n>N

α

(

F 2nα (x) + F 2

nα (x0))

<∑

n>N

(

1

2n+

1

2n

)

= 2∑

n>N

1

2n<

ε2

2

by choice of N . It follows that, for x ∈ V ,

n,α

| Fnα(x) − Fnα(x0) |2< ε2.

Hence, for x ∈ V , d(F (x), F (x0)) =√

n,α | Fnα(x) − Fnα(x0) |2 < ε, proving the continuity

of F .

Finally, we show that F is closed, i.e. the direct image by F of a closed set A is closed. IfA is closed in X, we assert F (A) = Cl(F (A)). Indeed, let us suppose F (x) /∈ F (A) i.e.x /∈ A. Then for some nα, x ∈ Bnα and Bnα

A = ∅. Hence fnα(x) 6= 0 and fnα(A) = 0.It follows that Fnα(x) 6= 0 and Fnα(A) = 0 and then, obviously, d(F (x), F (A)) > 0 so thatF (x) /∈ Cl(F (A)). Thus Cl(F (A)) ⊂ F (A). Furthermore, we always have F (A) ⊂ Cl(F (A)).So F (A) = Cl(F (A)) i.e. F (A) is closed.

The homeomorphism F imbeds X into a subspace of the generalized Hilbert Space Hτ which ismetrizable, then X is metrizable.

Remark 3.2.2. Any finite collection of sets is evidently locally finite. Thus any countablecollection of sets is σ-locally finite. Therefore a second countable space has a σ-locally finitebase, and thus the Urysohn Metrization Theorem is subsumed by the Nagata-Smirnov MetrizationTheorem.

There is a variant of the Nagata-Smirnov Metrization Theorem, proved independently by R.H.Bing in 1951:

Page 28: Metrizability of Topological Spaces

Section 3.3. Bing’s Metrization Theorem Page 25

3.3 Bing’s Metrization Theorem

Theorem 3.3.1. The following two conditions on a topological space are equivalent:

(a) X is metrizable.

(b) X is T3 and has a σ-discrete base.

Proof. Necessity follows from the Stone theorem. If Bi is an open σ-discrete refinement of the

open cover of X consisting of all 1i-balls about x ∈ X, then B =

∞⋃

i=1

Bi is a σ-discrete base for

X. (Recall that every metric space is T3.)

Sufficiency follows from the fact that a σ-discrete family is σ-locally finite and the Nagata-SmirnovMetrization Theorem implies that X is metrizable.

We shall now introduce some notions related to the notion of a cover, which will be used to provethe next metrization theorems.

Definition 3.3.2. A normal sequence in a space X is a sequence U1, U2, . . . of open covers ofX such that Un+1 star-refines Un, for n = 1, 2, . . ..

It will be called compatible normal sequence in X if and only if {St (x, U ) |n = 1, 2, . . .} is aneighborhood base at x, for each x ∈ X.

Any open cover of X which is U1 in some normal sequence in X will be called a normal cover.(Thus, every cover in a normal sequence is a normal cover.)

3.4 Pseudometrization Theorem.

All the previous metrization theorems are satisfied only for topological spaces which are at leastregular. But what about T0-spaces? The following theorem answers that question:

Theorem 3.4.1. [14]. A topological space X is pseudometrizable5 if and only if it has a compat-ible normal sequence. (Hence, a T0-space is metrizable if and only if it has a compatible normalsequence).

Proof. If X is pseudometrizable, its topology is generated by the pseudometric ρ. Let us defineUn =

{

Bρ(x, 13n )|x ∈ X

}

. Then the sequence U1, U2, . . . is a compatible normal sequence inX. The sets St(x, Un) form a neighborhood base at x, for each x in X. It is also clear that

St

(

Bρ(x,1

3n), Un

)

⊂ Bρ(x,1

3n−1),

5A pseudometric is a real-valued function ρ on X satisfying ρ(x, y) ≥ 0, ρ(x, x) = 0, ρ(x, z) ≤ ρ(x, y)+ρ(y, z)for every x, y, z ∈ X . A pseudometric is not required to satisfy ρ(x, y) > 0 for x 6= y like a metric.

Page 29: Metrizability of Topological Spaces

Section 3.4. Pseudometrization Theorem. Page 26

so that · · · < U ∗3 < U ∗

2 < U1 i.e. Un+1 star-refines Un, for n = 1, 2, . . ..

Conversely, Suppose we have a compatible normal sequence (Un) for X. Define t on X × X asfollows:

t(x, y) = 0 if y ∈ St(x, Un) for all n,

t(x, y) = 1 if y /∈ St(x, U1),

t(x, y) =1

2if y ∈ St(x, U1) , y /∈ St(x, U2),

t(x, y) =1

2nif y ∈ St(x, Un) , y /∈ St(x, Un+1).

Now for x, y ∈ X, let S (x, y) be all finite sequences s = (x1, . . . , xn) of points of X such thatx1 = x, xn = y or x1 = y, xn = x. Define

ρ(x, y) = inf

{

n∑

i=2

t(xi−1, xi)| {x1, . . . , xn} ∈ S (x, y)

}

.

Then ρ is a pseudometric on X. Indeed, for all x, y ∈ X, ρ(x, y) ≥ 0 and ρ(x, y) = 0 ify ∈ St(x, Un) for all n, and in particular when x = y, we have ρ(x, x) = 0. Furthermore, for allx, y ∈ X, ρ(x, y) = ρ(y, x) by the definition of S (x, y) and t. Moreover, ρ satisfies the triangleinequality, because for x, y ∈ X, then given any ε > 0, there exist x1 = x and xn = y such that

n∑

i=2

t(xi−1, xi) ≤ε

2+ ρ(x, y), this is by the definition of the infinimum.

Using again the same argument for y, z ∈ X, there exist x1 = y and xn = z such thatn∑

i=2

t(xi−1, xi) ≤ ε2

+ ρ(y, z). Finally,

ρ(x, z) = inf

{

n∑

i=2

t(xi−1, xi)| {x1, . . . , xn} ∈ S (x, y)

}

≤n∑

i=2

t(xi−1, xi) +

n∑

i=2

t(xi−1, xi)

≤ ε

2+ ρ(x, y) +

ε

2+ ρ(y, z).

Then, by taking the limit when ε goes to zero, we have ρ(x, z) ≤ ρ(x, y) + ρ(y, z) .

It remains to show that ρ is compatible with the topology on X.

Let Vn be the cover of X by balls Bρ(x, 12n ). It will suffice to show that, for any n,

a) Un < Vn−1,

b) Vn < Un−1,

since it will then be clear that the topologies generated by the two sequences are the same.

Page 30: Metrizability of Topological Spaces

Section 3.4. Pseudometrization Theorem. Page 27

a) Suppose U ∈ Un. Pick x ∈ U . If y ∈ U , then y ∈ St(x, Un) so t(x, y) ≤ 12n and hence

ρ(x, y) ≤ 12n < 1

2n−1 . Thus y ∈ Bρ(x, 12n−1 ), so U ⊂ Bρ(x, 1

2n−1 ) and thus Un < Vn−1.

b) To show that Vn < Un−1, it is enough to prove that whenever ρ(x, y) < 12n , then x and y lie

together in some element of Un, since then

Bρ(x,1

2n) ⊂ St(x, Un) ⊂ U,

for some U ∈ Un−1.

Hence, suppose ρ(x, y) < 12n . Then

infs∈S (x,y)

k∑

i=2

t(xi−1, xi) <1

2n,

and consequently, for some sequence {x1, . . . , xn} from S (x, y),

k∑

i=2

t(xi−1, xi) <1

2n.

We proceed now by induction on the length k of this sequence. If k = 2, then t(x, y) < 12n so

that y ∈ St(x, Um), y /∈ St(x, Um+1) for some m > n. Hence, in particular, y ∈ St(x, Un+1),from which it follows that x, y ∈ U for some U ∈ Un+1 in fact, so that certainly x, y lie togetherin some U ′ ∈ Un. (Recall, then, that if t(x, y) < 1

2n , we have x and y together in some elementof Un+1; we will use this again.)

Suppose the result is true for sequences of length < k, and supposek∑

i=2

t(xi−1, xi) < 12n . Let j

be the last number, 2 ≤ j ≤ k, such that

j∑

i=2

t(xi−1, xi) <1

2n+1.

Thenj+1∑

i=2

t(xi−1, xi) ≥1

2n+1

so thatk∑

i=j+2

t(xi−1, xi) <1

2n+1.

Now by the inductive hypothesis x1, xj lie in some U1 ∈ Un+1 while the argument above shows,since t(xj , xj+1) < 1

2n , that xj , xj+1 lie in some U2 ∈ Un+1, and finally, using the inductivehypothesis again, xj+1, xk lie in some U3 ∈ Un+1. Then x1 and xk lie in St(U2, Un+1) ⊂ U forsome U ∈ Un. This establishes our claim, by induction. X is then pseudometrizable.

We use the above theorem to prove the following metrization theorem:

Page 31: Metrizability of Topological Spaces

Section 3.5. Neighborhood Metrization Theorem Page 28

3.5 Neighborhood Metrization Theorem

Others criteria for a T0-space to be metrizable are given by the following theorem:

Theorem 3.5.1. [14]. A T0-space X is metrizable if and only if each x ∈ X possesses acountable neighborhood base {Uxn|n ∈ N} with the following properties:

a) y ∈ Uxn =⇒ Uyn ⊂ Uxn−1.

b) y /∈ Uxn−1 =⇒ Uyn

Uxn = ∅.

Proof. Let X be a T0 metrizable space, and let d be the distance defined on X.

Necessity is verified, since the properties (a) and (b) are obviously satisfied if we consider Uxn asthe disk of radius 1

2n about x. Indeed

a) If y ∈ Uxn =⇒ d(x, y) < 12n < 1

2n−1 . This means y ∈ Uxn−1 i.e. Uxn ⊂ Uxn−1.

b) If y /∈ Uxn−1 =⇒ d(x, y) ≥ 12n−1 > 1

2n i.e. x /∈ Uyn and y /∈ Uxn i.e. Uyn

Uxn = ∅.Conversely, let Un = {Uxn|x ∈ X} be a neighborhood base for each x ∈ X satisfying properties(a) and (b). We claim that St(Uxn, Un) ⊂ Uxn−2, for any n > 2.

Indeed, suppose Uzn

Uxn 6= 0. Then by property (b), z ∈ Uxn−1. Hence, by property (a),Uzn ⊂ Uxn−2, and thus St(Uxn, Un) ⊂ Uxn−2 as asserted. It now follows that Un star-refinesUn−2 for any n > 2, so that U1, U3, . . . is a normal sequence. It also follows that St(x, Un) ⊂Uxn−2 for any n > 2, so that U1, U3, . . . is compatible with X. Thus, by the above theorem(Theorem 3.4.1), X is metrizable.

The next metrization theorem due to A.H. Frink is a direct consequence of the previuos Neigh-borhood Metrization Theorem.

3.6 Frink’s Metrization Theorem

Theorem 3.6.1. A T1-space X is metrizable if and only if there is a neighborhood base{Uxn|n ∈ N} at each x ∈ X such that:

i) · · · ⊂ Ux2 ⊂ Ux1.

ii) for each n ∈ N, there is some m > n such that Uxm

Uym 6= ∅ =⇒ Uxm ⊂ Uyn.

Proof. Necessity. Let X a metrizable space. Then one can take Uxn ={

B(x, 12n )|x ∈ X

}

andUxn is a neighborhood base at x ∈ X satisfying properties (i) and (ii) of the theorem.

Sufficiency. The property (ii) follows directly from properties (b) and (a) of the NeighborhoodMetrization Theorem which give the metrizability of X. Indeed if Uxm

Uym 6= ∅, then by (b)

Page 32: Metrizability of Topological Spaces

Section 3.7. Alexandroff-Urysohn’s Metrization Theorem Page 29

we have x ∈ Uym−1 and by (a) Uxm ⊂ Uym−2 ⊂ Uyn for some n such that m > n (because ofthe property (i)).

We introduce now an idea which is obviously related to the notion of a compatible normal sequencewhich will be used in the next metrization theorem.

Definition 3.6.2. A development for a space X is a sequence U1, U2, . . . of open covers of Xsuch that Un refines Un−1, and, at each x ∈ X, {St(x, Un)|n = 1, 2, . . .} is a neighborhoodbase.

A space having a development is called developable. For example, a regular space having adevelopment is called Moore space.

The following theorem on metrizability of developable spaces is due to Alexandroff and Urysohnin 1923. ( See [14].)

3.7 Alexandroff-Urysohn’s Metrization Theorem

Theorem 3.7.1. A T0-space X is metrizable if and only if it has a development U1, U2, . . . withthe additional property that whenever U, V ∈ Un and U

V 6= ∅, then U⋃

V ⊂ W for someW ∈ Un−1.

Proof. Necessity. If X is a metrizable space, then one can take Un as the collection of 14n -balls

i.e. Un ={

B(x, 14n )|x ∈ X

}

.

To prove sufficiency, we use the Neighborhood Metrization Theorem above. Let U1, U2, . . .be a development of X with the required property. Then, for each n > 1, we find that ifU ∈ Un and x ∈ U , then St(U, Un) ⊂ St(x, Un−1). Now for n = 1, 2, . . . and x ∈ X,define Uxn = St(x, Un). Then we need only to verify properties a) and b) of the NeighborhoodMetrization Theorem.

a) If y ∈ Uxn, then for some V ∈ Un, x ∈ V and y ∈ V . But then

Uyn = St(y, Un) ⊂ St(V, Un) ⊂ St(x, Un−1) = Uxn−1 i.e. Uyn ⊂ Uxn−1.

b) If Uyn

Uxn 6= ∅, then for some U, V ∈ Un, U⋂

V 6= ∅. But then U⋃

V ⊂ W for someW ∈ Un−1, and hence y ∈ St(x, Un−1) = Uxn−1. Thus if y /∈ Uxn−1, then Uyn

Uxn = ∅.Thus, X is metrizable.

Still in order to extend the class of metrizable spaces, we need another new terminology: thecollectionwise normality, which together with the notion of development, will be used to establishfurther topological characterizations of the class of metrizable spaces. Collectionwise normalityis another extension of normality, weaker than paracompactness.

Page 33: Metrizability of Topological Spaces

Section 3.8. Bing’s Metrization Theorem for Moore Spaces Page 30

Definition 3.7.2. A topological space X is said to be collectionwise normal if X is a T1-spaceand for every discrete family {Fs}s∈S of closed subsets of X, there exists a discrete family {Vs}s∈S

of open subsets of X such that Fs ⊂ Vs for every s ∈ S.

It is clear that every collectionwise normal space is normal.

Theorem 3.7.3. A T1-space X is collectionwise normal if and only if for every discrete family{Fs}s∈S of closed subsets of X, there exists a discrete family {Us}s∈S of open subsets of X suchthat Fs ⊂ Us for every s ∈ S and Us

Us′ = ∅ whenever s 6= s′.

Proof. Necessity follows directly from the definition of collectionwise normality. To prove suf-ficiency, it suffices to show that any T1-space X satisfying the given condition in the theoremis collectionwise normal. Let us suppose that these conditions are satisfied, then it is clear thatX is normal, so that for a discrete family {Fs}s∈S of closed subsets of X, there is a family ofpairwise disjoint open sets {Us}s∈S. Let us put A =

s∈S

Fs and B = X \ ⋃s∈S

Us. Then the

closed sets A and B are disjoint. Indeed, since we have Fs ⊂ Us, then⋃

s∈S

Fs ⊂ ⋃

s∈S

Us which

means that (X \ ⋃s∈S

Us) ⊂ (X \ ⋃s∈S

Fs). This last inclusion implies (X \ ⋃s∈S

Us)⋂

(⋃

s∈S

Fs) = ∅.By the normality of X, we can find disjoint open sets U and V such that A =

s∈S

Fs ⊂ U and

B = (X \ ⋃s∈S

Us) ⊂ V . One can check that the family {Vs}s∈S where Vs = Us

U is discrete

and contains Fs for every s ∈ S.

Remark 3.7.4. A collectionwise normal space can be defined as a space such that any discretefamily of closed sets can be separated by a discrete family of open sets.

We can now prove the metrisation theorem for Moore spaces due to Bring in 1951:

3.8 Bing’s Metrization Theorem for Moore Spaces

Theorem 3.8.1. [3]. A topological space is metrizable if and only if it is collectionwise normaland has a development; or in other words, a topological space is metrizable if and only if it is acollectionwise normal Moore space.

Proof. Necessity. Let X be a metrizable space. Then, according to the Stone theorem, X isparacompact. The collectionwise normality of X is given by the following theorem:

Theorem 3.8.2. Every paracompact space is collectionwise normal.

Proof. Let {Fs}s∈S be a discrete family of closed subsets of a paracompact space X. For everyx ∈ X, let us choose a neighborhood Ux of the point x whose closure meets at most one setFs. Then X being paracompact, the open cover U = {Ux}x∈X of X has a locally finite open

refinement W . For every s ∈ S, let Vs = X \ ⋃{

W : W ∈ W and W⋂

Fs = ∅}

. Clearly,

Page 34: Metrizability of Topological Spaces

Section 3.8. Bing’s Metrization Theorem for Moore Spaces Page 31

we have Fs ⊂ Vs. Indeed, suppose y /∈ Vs then y ∈ ⋃{

W : W ∈ W and W⋂

Fs = ∅}

which

means that there is some W ∈ W suth that y ∈ W and W⋂

Fs = ∅ i.e. y /∈ Fs. To concludethe proof, it suffices to show that every W ∈ W meets at most one element of the family{Vs}s∈S. This, however, follows from the fact that W meets at most one set Fs. This establishesthe collectionwise normality of X.

Furthermore, the family Bn ={

B(x, 12n )|x ∈ X

}

is a development of X.

Sufficiency. Let us consider a collectionwise normal space X having a development B1, B2, . . .We first prove that that X is paracompact.

Let us consider an open cover {Us}s∈S of X and take a well-ordering relation < on the set S.

Let Fs,i = X \[

St(X \ Us, Bi)⋃

(⋃

s′<s

Us′)

]

. For all s ∈ S and i = 1, 2, . . ., Fs,i are closed

because complement of open sets. Furthermore these closed sets Fs,i form a cover of the spaceX. Indeed, taking for any x ∈ X the smallest element sx ∈ S such that x ∈ Usx

, and a naturalnumber ix such that6 St(x, Bix) ⊂ Usx

we have x ∈ Fsx,ix . Moreover, the family Fi = {Fs,i}s∈S

is discrete, since for a fixed i, the neighborhood Usx

St(x, Bi) of the point x meets only onemember of Fi, namely the set Fsx,i. By the collectionwise normality of X, there exists a discretefamily {Us,i}s∈S

of open sets such that Fs,i ⊂ Us,i. Since Fs,i ⊂ Us, we can choose Us,i suchthat Us,i ⊂ Us for s ∈ S and i = 1, 2, . . . and we will have Fs,i ⊂ Us,i ⊂ Us. Hence {Us,i}s∈S

,i = 1, 2, . . ., is a σ-discrete and obviously σ-locally finite open refinement of {Us}s∈S. Then Xis paracompact.

Now, for i = 1, 2, . . ., let Di be a σ-locally finite open refinement of the cover Bi (which is the

element of the development of X), then the family D =∞⋃

i=1

Di is a base for the space X and the

metrizability of X follows from the Nagata-Smirnov Metrization Theorem.

6Since {St(x, Bi)|i = 1, 2, . . .} is a neighborhood base at each x ∈ X .

Page 35: Metrizability of Topological Spaces

4. Conclusion

The theory of metrizability of topological spaces is distinguished by its long history and thediversity of methods that have been used in its study. In this work, we have studied someimportant criteria for metrizability of topological spaces. The first metrization theorem wasgiven in 1923 by Alexandroff and Urysohn (Theorem 3.7.1). Two years later, Urysohn gave aproof of metrizability of second countable normal space (Theorem 3.1.1). Unfortunately, UrysohnMetrization Theorem does not give a total answer because it only gives sufficient condition formetrizability. Topologists wanted to get general criteria which are both necessary and sufficient toestablish the metrizability of topological spaces, like in Alexandroff-Urysohn Metrization Theoremcase. The search of such conditions was long and was not satisfactorily concluded until the early1950’s when Nagata (in 1950) and Smirnov (in 1951) independently provided such conditions(Theorem 3.2.1). R.H. Bing provided the similar result in 1951 independently of works of Nagataand Smirnov (Theorem 3.3.1). Theorem 3.8.1 of metrizability of Moore spaces is also due toBing. We owe the Uniform Metrization Theorem 3.4.1 to Alexandroff and Urysohn, and theNeighborhood Metrization Theorem 3.5.1 is a slight modification of results of Nagata; whileTheorem 3.6.1, which is a consequence of Theorem 3.5.1, is due to A.H. Frink.

We have studied only a few well-known results on metrizability of topological spaces and thetheory of metrizability continues to grow up and to be an active area of research. The normalMoore space conjecture for example which states that every normal Moore space is metrizabledepends on the choice of set theory. For instance, F.B. Jones showed, assuming the continuumhypothesis, that every separable normal Moore is metrizable. Furthermore, Health and Silver(See [14], P. 309) states that it is consistent with the axioms of set theory through the axiom ofchoice to assume the existence of a non-metrizable separable normal Moore space.

32

Page 36: Metrizability of Topological Spaces

Appendix A. The Separation Axioms

The definition of a topological space is very general; not many interesting theorems can be provedabout all topological spaces. In this part, we study a certain classes of topological spaces in whichwe discuss axioms of separation which concern the ways of separating points and/or closed setsin topological spaces (see [3] and [5]). There are various degrees of separation and the firstseparation axiom is the following.

T0-spaces

Definition A.0.3. A topological space (X, τ) is said to be T0 if given any two distinct points xand y of X, there exists a neighborhood of at least one point which does not contain the otherpoint. T0-spaces are also called Kolmogorov spaces.

Example A.0.4. Let (X, τ) be a topological space, where X = {0, 1}.

• If τ = {∅, X} then X is not T0, since there is no way of separating 0 from 1 or 1 from 0 i.e.there is no neighborhood of 0 which does not contain 1 and vice versa.

• But if τ = {∅, {1} , X}, then X becomes a T0-space, since there is a neighborhood {1} of 1which does not contain 0.

The following proposition gives another characterization of a T0-space.

Proposition A.0.5. A space (X, τ) is T0 if given any two distinct points x and y of X, eitherx /∈ Cl {y} or y /∈ Cl {y}, where Cl {x} = {x} is the closure of {x}.

Proof. Suppose X is a T0-space, x and y any two distinct points of X, then there is an openset U which contains x, but not y. U being open in X, X \ U is a closed subset of X whichcontains y, but not x. Since {y} ⊂ X \ U and X \ U is closed, then Cl {y} ⊂ X \ U1; hencex /∈ Cl {y} (because X \ U does not contain x). Interchanging the role of x and y and usingthe same argument, we show that y /∈ Cl {x}.

Corollary A.0.6. In a T0-space (X, τ), every neighborhood of x or y is a neighborhood of bothx and y if Cl {x} = Cl {y}.

Proof. If x and y are distinct points of a T0-space (X, τ) such that Cl {x} = Cl {y} then, sincewe always have x ∈ Cl {x} , y ∈ Cl {y} and considering the fact that Cl {x} = Cl {y}, we havex ∈ Cl {y} and y ∈ Cl {x}. Suppose U is a neighborhood of x which does not contain y, thenX \ U is a closed set which contains y, but not x and Cl {y} ⊂ X \ U ; hence x /∈ Cl {y}, acontradiction. This means that our assumption was wrong; there is no neighborhood of x whichdoes not contain y. Similarly, there is no neighborhood of y which does not contain x.

1because Cl {y} is the smallest closed set containing {y}

33

Page 37: Metrizability of Topological Spaces

Page 34

The following definition gives a slightly stronger separation property than the one given by T0-spaces.

T1-spaces

Definition A.0.7. A topological space (X, τ) is said to be T1 if given any two distinct points xand y of X, there exists a neighborhood of each of them which does not contain the other point.T1-spaces are also called Frechet spaces.

Proposition A.0.8. A topological space (X, τ) is T1-space if and only if each one-point subsetof X is closed.

Proof. Let (X, τ) be a topological space and x ∈ X.Then the statement {x} is closed in X isequivalent to the following statements:

∗ its complement X \ {x} is open in X,

∗ each point y 6= x has a neighborhood (X \ {x}) which does not contain x,

∗ X is a T1-space.

As consequence of the above proposition, we have the following result.

Corollary A.0.9. If (X, τ) is T1-space and x ∈ X, the {x} = Cl {x}.Example A.0.10. Every metric space is T1, because every one-point subset of a metric space isclosed.

Indeed, let (X, d) be any metric space and x ∈ X. Suppose y ∈ X \ {x}, then there is aε- neighborhood of y which is disjoint from {x} (otherwise y would be in {x}). Thus that ε-neighborhood of y lies completely in X \{x}2. The point y being arbitrary chosen in X \{x}, wesay that X \ {x} is a neighborhood of each of its points and then is open. Hence {x} is closed.

Remark A.0.11. Every T1-space is a T0-space, but not the inverse.

There is a stronger separation property than the ones given by either T0 or T1-spaces due toHausdorff.

T2-spaces

Definition A.0.12. A topological space (X, τ) is said to be T2 if given any two distinct pointsx and y of X, there exist two open sets U and V such that x ∈ U , x ∈ V and U

V = ∅.T2-spaces are also called Hausdorff spaces.

2since there ε > 0 such that B(y, ε)⋂ {x} = ∅ and B(y, ε) ⊂ X \ {x}, Cf Definition 2.1.7

Page 38: Metrizability of Topological Spaces

Page 35

Example A.0.13.

• The set R of real numbers with the usual topology is a T2-space.

• Suppose that the space (X, τ) is homeomorphic to the space (Y, τ ′) and that X is T2, then Yalso is T2.

Remark A.0.14. Every T2-space is a T1-space.

We now give some separation axioms which separate either a point from a closed subset or twoclosed subsets of a topological space.

T3-spaces

Definition A.0.15. A topological space (X, τ) is said to be T3 if it is T1 and given any closedsubset A of X and any point x of X which is not in A, there exist two disjoint open sets U andV such that x ∈ U and A ⊂ V . T3-spaces are sometimes called Regular spaces.

Remark A.0.16. Every T3-space is a T2-space.

Proposition A.0.17. A topological space (X, τ) is T3 if and only if given any x ∈ X and anyneighborhood U of x, there exists a neighborhood V of x such that Cl(V ) ⊂ U , where Cl(V )denotes the closure of V .

Proof. Let us suppose that (X, τ) is T3 and U an open neighborhood of x in X. Then thecomplement X \ U of X is a closed subset of X which does not contain x. Then according tothe definition of T3-space, there exist disjoint open sets W and V such that X \ U ⊂ W andx ∈ V . Since X \U ⊂ W , we have X \W ⊂ X \(X \U) = U . Furtheremore, since W

V = ∅,we have V ⊂ X \ W ⊂ U . So X \ W is a closed set of X which contains V . Cl(V ) being thesmallest closed set containing V , we finaly have

V ⊂ Cl(V ) ⊂ X \ W ⊂ U.

Suppose now given any neighborhood U of x ∈ X, there exists a neighborhood V of x suchthat Cl(V ) ⊂ U . Let x ∈ X and let F be any closed subset of X which does not contain x.Then X \ F is an open subset of X containing x i.e. a neighborhood of x; hence there exists aneighborhood V of x such that Cl(V ) ⊂ X \ F . This implies F ⊂ X \ Cl(V ), i.e. X \ Cl(V )is an open set which contains F , and V is an open set which contains x. Since we always haveV ⊂ Cl(V ), then

(X \ Cl(V ))⋂

V = ∅.We finally have: F ⊂ X \ Cl(V ), x ∈ V and (X \ Cl(V ))

V = ∅ which is similar to say thatX is T3.

Example A.0.18.

Page 39: Metrizability of Topological Spaces

Page 36

• Rn is a regular space.

• The circle C is a regular space (as subspace of Rn).

• The torus T is regular, since T = C × C ⊂ R3.

One of the most important of the separation axioms is given by the following definition.

T4-spaces

Definition A.0.19. A topological space (X, τ) is said to be T4 if it is T1 and given any twodisjoint closed subsets A and B of X, there exist two disjoint open sets U and V such thatA ⊂ U and B ⊂ V . T4-spaces are sometimes called Normal spaces.

Remark A.0.20. Every T4-space is a T3-space.

Example A.0.21.

• Each topological space carrying the discrete topology is T4, but every topological space withthe trivial topology is not normal, since no one-point subset of X is closed.

• Every metric space is normal.

We now give another characterization of T4-space which is very important.

Proposition A.0.22. A topological space (X, τ) is normal if and only if given any closed subsetF of X and any open set U of X such that F ⊂ U , there exists an open set V such that

F ⊂ V ⊂ Cl(V ) ⊂ U.

Proof. Suppose (X, τ) is normal, F and U two subsets of X such that F ⊂ U , F closed andU open. Then X \ U is a closed set of X and (X \ U)

F = ∅. Hence there are two disjointopen sets W and V such that X \ U ⊂ W and F ⊂ V . This implies F ⊂ V ⊂ Cl(V ) ⊂ U .

Suppose now given any closed set F and any open set U which contains F , there exists an openset V such that F ⊂ V ⊂ Cl(V ) ⊂ U . Let A and B be any two disjoint closed subsets ofX. Then X \ A is an open set which contains B. By hypothesis, then there is an open set Vsuch that B ⊂ V ⊂ Cl(V ) ⊂ X \ A. Hence, A ⊂ X \ Cl(V ) i.e. X \ Cl(V ) is an open setwhich contains A and is disjoint from V . We finally have B ⊂ V , V open, A ⊂ X \ Cl(V ) and(X \ Cl(V ))

V = ∅. Therefore X is normal.

The following statement gives us the condition for a subset of a normal space to be normal.3

3for k = 0, 1, 2, 3, every subspace of a Tk-space is Tk. But for T4-spaces, it is not true. However, each closed

subset of a T4-space is T4.

Page 40: Metrizability of Topological Spaces

Page 37

Proposition A.0.23. If Y is a closed subset of a normal space X, then Y is normal.

Proof. Y is T1 because subspace of a normal space X which is T1. Since Y is closed, then if F1

and F2 are disjoint closed subsets of Y , they are also disjoint closed subsets of X since Y ⊂ X.There exist then two disjoint open sets U and V such that F1 ⊂ U and F2 ⊂ V .

Then F1 ⊂ U⋂

Y , F2 ⊂ V⋂

Y where U⋂

Y and V⋂

Y are two disjoint open subsets of Y .Thus Y is normal.

Theorem A.0.24. Every paracompact space is nornal.

Proof. We first establish regularity. Let us suppose A is closed set in a paracompact space Xand x /∈ A. For each y ∈ A. We find an open set Vy containing y such that x /∈ Cl (Vy). Thenthe sets Vy, y ∈ A, together with the set X \ A, form an open cover of X. Let W be an openlocally finite refinement and let V =

⋃ {W ∈ W | W⋂

A 6= ∅}. Then V is open, contains A,and Cl(V ) = {Cl(W )| W

A 6= ∅}. But each such set W is contained in some Vy, and henceCl(W ) is contained in Cl(Vy) and thus does not contain x. Hence x /∈ Cl(V ). Thus x and Aare separated by open sets in X.

Now let us suppose A and B are disjoint closed sets in X. For each y ∈ A, by regularity, we findopen set Vy such that y ∈ Vy and Cl (Vy)

B = ∅. Then proceeding exactly as above, we canproduce an open set V such that A ⊂ V and Cl (V )

B = ∅. Thus X is normal.

T5-spaces

Definition A.0.25. A topological space (X, τ) is said to be T5 if all subspaces of X are normal.

T5-spaces are sometimes called Completely normal spaces.

Remark A.0.26. Every T5-space is a T4-space.

T6-spaces

Definition A.0.27. A topological space (X, τ) is said to be T6 if it is normal and all closedsubsets of X are Gδ. T6-spaces are also called Perfectly normal spaces.

Definition A.0.28. A subset B of a topological space is said to be Gδ if B is the intersectionof countably many open sets i.e.

B is Gδ ⇐⇒ B =

∞⋂

j=1

Oj where all the Oj are open.

Remark A.0.29. Every T6-space is a T5-space.

Page 41: Metrizability of Topological Spaces

Page 38

Proposition A.0.30. A normal space X is perfectly normal if and only if every open subset Aof X is Fσ i.e. A is the union of countably many closed sets. In other words,

A is Fσ ⇐⇒ A =

∞⋃

i=1

Ci where all the Ci are closed.

Remark A.0.31. It is clear that Fσ sets are complement of Gδ sets. Then by taking complementfrom Definition A.0.28, we prove Proposition A.0.30.

Finally, the separation axioms form the following hierarchy of conditions :

T6 =⇒ T5 =⇒ T4 =⇒ T3 =⇒ T2 =⇒ T1 =⇒ T0.

Let us notice that any metric space satisfies all these conditions, from T0 to T6-spaces.

Indeed, let (X, d) be any metric space and A any closed subset of X. Let Gn denotes the familyof open sets defined by:

Gn =

{

x : d(x, A) <1

n

}

=⋃

x∈A

B

(

x,1

n

)

.

It is clear that∞⋂

n=1

Gn =∞⋂

n=1

x∈A

B

(

x,1

n

)

= A.

To prove that, let a ∈∞⋂

n=1

x∈A

B

(

x,1

n

)

=⇒ ∃xn ∈ A such that a ∈ B(xn,1

n) for all n,

=⇒ xn −→ a when n −→ ∞,

=⇒ a ∈ A = A ,

and this means that∞⋂

n=1

x∈A

B(

x, 1n

)

⊂ A. The inverse inclusion is obvious.

We conclude that A =∞⋂

n=1

(

x∈A

B(

x, 1n

)

)

i.e. A is Gδ.

Furthermore, for any disjoint closed subsets A and B of X , f(x) = d(x,A)d(x,A)+d(x,B)

is a continuous

real valued function on X such that 0 ≤ f(x) ≤ 1 ,and f(a) = 0 for all a ∈ A and f(b) = 1for all b ∈ B i.e. f(A) ⊂ {0} and f(B) ⊂ {1}. We have A ⊂ f−1

(

[0, 12))

= U andB ⊂ f−1

(

(12, 1])

= V , where U and V are two disjoint open subsets of X. This means that themetric space X is T6 and Gδ, i.e. X is perfectly normal.

Page 42: Metrizability of Topological Spaces

Acknowledgments

I would like to acknowledge my supervisor, Prof. Hans-Peter Kunzi, who helped me through myessay, for his valuable remarks, supports and help that were essential in this work.

My acknowledgment also goes to all AIMS 2006 lecturers, without forgetting all the tutors namelyAnahita New, Dr. Sam Webster, Dr. Laure Gouba, Paul Razafimandimby, Christian Rivasseau,Henry Amuasi, Eman and Jean-Marie Mirebeau, for their help and guidance during all academicyear.

I wish to express my deep gratitude to Prof. Fritz Hahne, AIMS director, and Prof. Neil Turok,the founder of AIMS, for the opportunity they gave me to become one of the students of thisgreat institute.

I also want to thank all AIMS staff, particularly Igsaan Kamalie, Emmanuel Kongolo, Jan Groe-newald, Andy Rabagliati, Mirjam Miske, Lynne Teixeira and Theressa Phokolo Popiel, for theirsupport and assistance.

I am also thankful to my former lecturers from the Exact Sciences Department at the EcoleNormale Superieure (ENS) of the Marien Ngouabi University for training and helping me todevelop my abilities in mathematics.

I acknowledge all the AIMS 2006 students for the special time we have spent together that madelife easy and fun at AIMS.

Lastly, I dedicate this essay to my family for their unconditional and perpetual love, best wishesand encouragement.

39

Page 43: Metrizability of Topological Spaces

Bibliography

[1] George Mc Carty. Topology, An Introduction with Application to Topological Groups. DoverPublications, Inc. New-York, 1988.

[2] Helen F. Cullen. Introduction to General Topology. D.C. Heath and Company, Boston,1967.

[3] Ryszard Engelking. General Topology, Revised and completed edition. Sigma Series in PureMathematics, Vol.6. Heldermann Verlag Berlin, 1989.

[4] Zbigniew Fiedorowicz. Lectures notes on Urysohn’s Metrization Theorem. Department ofMathematics, Ohio State University, Columbus, OH 43210 , USA, 2002.

[5] Michael C. Gemignani. Elementary Topology, second edition. Dover Publications, Inc.New-York, 1990.

[6] Sze-Tsen Hu. Introduction to General Topology. Holden-day Series in Mathematics, LosAngeles, 1969.

[7] Bruce Ikenaga. Lectures notes on the Separation Axioms. Department of Mathematics,Millersville University, P.O. Box 1002, Millersville, PA 17551-0302, USA, 1999.

[8] John L. Kelley. General Topology. D.Van Nostrand Company, Inc. Priceton, New Jersey,1963.

[9] Maynard J. Mansfield. Introduction to Topology. D.Van Nostrand Company, Inc. Priceton,New Jersey, 1963.

[10] J. Nagata. Modern General Topology, Second edition revised. Elsevier Science PublishersB.V., Amsterdam, Netherlands, 1985.

[11] George F. Simmons. Introduction to Topology and Modern Analysis. Mc Graw-Hill BookCompany, Inc. USA, 1963.

[12] I.M. Singer and J.A. Thorpe. Lecture Notes on Elementary Topology and Geometry.Springer-Verlag, New-York, 1967.

[13] W.A. Sutherland. Introduction to Metric and Topological Spaces. Oxford Science Publica-tions, Oxford, 2003.

[14] Stephen Willard. General Topology. Dover Publications, Inc. New-York, 2004.

40