The uniform metric on product spaces

Jordan Bell
April 3, 2014

1 Metric topology

If (X,d) is a metric space, aX, and r>0, then the open ball with center a and radius r is

Brd(a)={xX:d(x,a)<r}.

The set of all open balls is a basis for the metric topology induced by d.

If (X,d) is a metric space, define

d¯(a,b)=d(a,b)1,a,bX,

where xy=min{x,y}. It is straightforward to check that d¯ is a metric on X, and one proves that d and d¯ induce the same metric topologies.11 1 James Munkres, Topology, second ed., p. 121, Theorem 20.1. The diameter of a subset S of a metric space (X,d) is

diam(S,d)=supa,bSd(a,b).

The subset S is said to be bounded if its diameter is finite. The metric space (X,d) might be unbounded, but the diameter of the metric space (X,d¯) is

diam(X,d¯)=supa,bXd¯(a,b)=diam(X,d)1,

and thus the metric space (X,d¯) is bounded.

2 Product topology

If J is a set and Xj are topological spaces for each jJ, let X=jJXj and let πj:XXj be the projection maps. A basis for the product topology on X are those sets of the form jJ0πj-1(Uj), where J0 is a finite subset of J and Uj is an open subset of Xj, jJ0. Equivalently, the product topology is the initial topology for the projection maps πj:XXj, jJ, i.e. the coarsest topology on X such that each projection map is continuous. Each of the projection maps is open.22 2 John L. Kelley, General Topology, p. 90, Theorem 2. The following theorem characterizes convergent nets in the product topology.33 3 John L. Kelley, General Topology, p. 91, Theorem 4.

Theorem 1.

Let J be a set and for each jJ let Xj be a topological space. If X=jJXj has the product topology and (xα)αI is a net in X, then xαx if and only if πj(xα)πj(x) for each jJ.

Proof.

Let (xα)αI be a net that converges to xX. Because each projection map is continuous, if jJ then πj(xα)πj(x). On the other hand, suppose that (xα)αI is a net, that xX, and that πj(xα)πj(x) for each jJ. Let 𝒪j be the set of open neighborhoods of πj(x)Xj. For jJ and U𝒪j, because πj(xα)πj(x) we have that πj(xα) is eventually in U. It follows that if jJ and U𝒪j then xα is eventually in πj-1(U). Therefore, if J0 is a finite subset of J and Uj𝒪j for each jJ0, then xα is eventually in jJ0πj-1(Uj). This means that the net (xα)αI is eventually in every basic open neighborhood of x, which implies that xαx. ∎

The following theorem states that if J is a countable set and (X,d) is a metric space, then the product topology on XJ is metrizable.44 4 James Munkres, Topology, second ed., p. 125, Theorem 20.5.

Theorem 2.

If J is a countable set and (X,d) is a metric space, then

ρ(x,y)=supjJd¯(xj,yj)j=supjJd(xj,yj)1j

is a metric on XJ that induces the product topology.

A topological space is first-countable if every point has a countable local basis; a local basis at a point p is a set of open sets each of which contains p such that each open set containing p contains an element of . It is a fact that a metrizable topological space is first-countable. In the following theorem we prove that the product topology on an uncountable product of Hausdorff spaces each of which has at least two points is not first-countable.55 5 cf. John L. Kelley, General Topology, p. 92, Theorem 6. From this it follows that if (X,d) is a metric space with at least two points and J is an uncountable set, then the product topology on XJ is not metrizable.

Theorem 3.

If J is an uncountable set and for each jJ we have that Xj is a Hausdorff space with at least two points, then the product topology on jJXj is not first-countable.

Proof.

Write X=jJXj, and suppose that xX and that Un,n, are open subsets of X containing x. Since Un is an open subset of X containing x, there is a basic open set Bn satisfying xBnUn: by saying that Bn is a basic open set we mean that there is a finite subset Fn of J and open subsets Un,j of Xj, jFn, such that

Bn=jFnπj-1(Un,j).

Let F=nFn, and because J is uncountable there is some kJF; this is the only place in the proof at which we use that J is uncountable. As Xk has at least two points and x(k)Xk, there is some aXk with x(k)a. Since Xk is a Hausdorff space, there are disjoint open subsets N1,N2 of Xk with x(k)N1 and aN2. Define

Vj={N1j=kXjjk

and let V=jJVj. We have xV. But for each n, there is some ynBn with yn(k)=aN2, hence yn(k)N1 and so ynV. Thus none of the sets Bn is contained in V, and hence none of the sets Un is contained in V. Therefore {Un:n} is not a local basis at x, and as this was an arbitrary countable set of open sets containing x, there is no countable local basis at x, showing that X is not first-countable. (In fact, we have proved there is no countable local basis at any point in X; not to be first-countable merely requires that there be at least one point at which there is no countable local basis.) ∎

3 Uniform metric

If J is a set and (X,d) is a metric space, we define the uniform metric on XJ by

dJ(x,y)=supjJd¯(xj,yj)=supjJd(xj,yj)1.

It is apparent that dJ(x,y)=0 if and only if x=y and that dJ(x,y)=dJ(y,x). If x,y,zX then,

dJ(x,z) = supjJd¯(xj,zj)
supjJd¯(xj,yj)+d¯(yj,zj)
supjJd¯(xj,yj)+supjJd¯(yj,zj)
= dJ(x,y)+dJ(y,z),

showing that dJ satisfies the triangle inequality and thus that it is indeed a metric on XJ. The uniform topology on XJ is the metric topology induced by the uniform metric.

If (X,d) is a metric space, then X is a topological space with the metric topology, and thus XJ=jJX is a topological space with the product topology. The following theorem shows that the uniform topology on XJ is finer than the product topology on XJ.66 6 James Munkres, Topology, second ed., p. 124, Theorem 20.4.

Theorem 4.

If J is a set and (X,d) is a metric space, then the uniform topology on XJ is finer than the product topology on XJ.

Proof.

If xXJ, let U=jJUj be a basic open set in the product topology with xU. Thus, there is a finite subset J0 of J such that if jJJ0 then Uj=X. If jJ0, then because Uj is an open subset of (X,d) with the metric topology and xjUj, there is some 0<ϵj<1 such that Bϵjd(xj)Uj. Let ϵ=minjJ0ϵj. If dJ(x,y)<ϵ then d(xj,yj)<ϵ for all jJ and hence d(xj,yj)<ϵj for all jJ0, which implies that yjBϵjd(xj)Uj for all jJ0. If jJJ0 then Uj=X and of course yjUj. Therefore, if yBϵdJ(x) then yU, i.e. BϵdJ(x)U. It follows that the uniform topology on XJ is finer than the product topology on XJ. ∎

The following theorem shows that if we take the product of a complete metric space with itself, then the uniform metric on this product space is complete.77 7 James Munkres, Topology, second ed., p. 267, Theorem 43.5.

Theorem 5.

If J is a set and (X,d) is a complete metric space, then XJ with the uniform metric is a complete metric space.

Proof.

It is straightforward to check that (X,d) being a complete metric space implies that (X,d¯) is a complete metric space. Let fn be a Cauchy sequence in (XJ,dJ): if ϵ>0 then there is some N such that n,mN implies that

dJ(fn,fm)<ϵ.

Thus, if ϵ>0, then there is some N such that n,mN and jJ implies that d¯(fn(j),fm(j))dJ(fn,fm)<ϵ. Thus, if jJ then fn(j) is a Cauchy sequence in (X,d¯), which therefore converges to some f(j)X, and thus fXJ. If n,mN and jJ, then

d¯(fn(j),f(j)) d¯(fn(j),fm(j))+d¯(fm(j),f(j))
dJ(fn,fm)+d¯(fm(j),f(j))
<ϵ+d¯(fm(j),f(j)).

As the left-hand side does not depend on m and d¯(fm(j),f(j))0, we get that if nN and jJ then

d¯(fn(j),f(j))ϵ.

Therefore, if nN then

dJ(fn,f)ϵ.

This means that fn converges to f in the uniform metric, showing that (XJ,dJ) is a complete metric space. ∎

4 Bounded functions and continuous functions

If J is a set and (X,d) is a metric space, a function f:JX is said to be bounded if its image is a bounded subset of X, i.e. f(J) has a finite diameter. Let B(J,X) be the set of bounded functions J(X,d); B(J,X) is a subset of XJ. Since the diameter of (X,d¯) is 1, any function J(X,d¯) is bounded, but there might be unbounded functions J(X,d). We prove in the following theorem that B(J,X) is a closed subset of XJ with the uniform topology.88 8 James Munkres, Topology, second ed., p. 267, Theorem 43.6.

Theorem 6.

If J is a set and (X,d) is a metric space, then B(J,X) is a closed subset of XJ with the uniform topology.

Proof.

If fnB(J,Y) and fn converges to fXJ in the uniform topology, then there is some N such that dJ(fN,f)<12. Thus, for all jJ we have d¯(fN(j),f(j))<12, which implies that

d(fN(j),f(j))=d¯(fN(j),f(j))<12.

If i,jJ, then

d(f(i),f(j)) d(f(i),fN(i))+d(fN(i),fN(j))+d(fN(j),f(j))
12+diam(fN(J),d)+12.

fNB(J,X) means that diam(fN(J),d)<, and it follows that diam(f(J),d)diam(fN(J),d)+1<, showing that fB(J,X). Therefore if a sequence of elements in B(J,X) converges to an element of XJ, that limit is contained in B(J,X). This implies that B(J,X) is a closed subset of XJ in the uniform topology, as in a metrizable space the closure of a set is the set of limits of sequences of points in the set. ∎

If J is a set and Y is a complete metric space, we have shown in Theorem 5 that YJ is a complete metric space with the uniform metric. If X and Y are topological spaces, we denote by C(X,Y) the set of continuous functions XY. C(X,Y) is a subset of YX, and we show in the following theorem that if Y is a metric space then C(X,Y) is a closed subset of YX in the uniform topology.99 9 James Munkres, Topology, second ed., p. 267, Theorem 43.6. Thus, if Y is a complete metric space then C(X,Y) is a closed subset of the complete metric space YX, and is therefore itself a complete metric space with the uniform metric.

Theorem 7.

If X is a topological space and let (Y,d) is a metric space, then C(X,Y) is a closed subset of YX with the uniform topology.

Proof.

Suppose that fnC(X,Y) and fnfYX in the uniform topology. Thus, if ϵ>0 then there is some N such that nN implies that dJ(fn,f)<ϵ, and so if nN and xX then

d¯(fn(x),f(x))dJ(fn,f)<ϵ.

This means that the sequence fn converges uniformly in X to f in the uniform metric, and as each fn is continuous this implies that f is continuous.1010 10 See James Munkres, Topology, second ed., p. 132, Theorem 21.6. We have shown that if fnC(X,Y) and fnfYX in the uniform topology then fC(X,Y), and therefore C(X,Y) is a closed subset of YX in the uniform topology. ∎

5 Topology of compact convergence

Let X be a topological space and (Y,d) be a metric space. If fYX, C is a compact subset of X, and ϵ>0, we denote by BC(f,ϵ) the set of those gYX such that

sup{d(f(x),g(x)):xC}<ϵ.

A basis for the topology of compact convergence on YX are those sets of the form BC(f,ϵ), fYX, C a compact subset of X, and ϵ>0. It can be proved that the uniform topology on YX is finer than the topology of compact convergence on YX, and that the topology of compact convergence on YX is finer than the product topology on YX.1111 11 James Munkres, Topology, second ed., p. 285, Theorem 46.7. Indeed, we have already shown in Theorem 4 that the uniform topology on YX is finer than the product topology on YX. The significance of the topology of compact convergence on YX is that a sequence of functions fn:XY converges in the topology of compact convergence to a function f:XY if and only if for each compact subset C of X the sequence of functions fn|C:CY converges uniformly in C to the function f|C:CY.