The Stone-Čech compactification of Tychonoff spaces

Jordan Bell
June 27, 2014

1 Completely regular spaces and Tychonoff spaces

A topological space X is said to be completely regular if whenever F is a nonempty closed set and xXF, there is a continuous function f:X[0,1] such that f(x)=0 and f(F)={1}. A completely regular space need not be Hausdorff. For example, if X is any set with more than one point, then the trivial topology, in which the only closed sets are and X, is vacuously completely regular, but not Hausdorff. A topological space is said to be a Tychonoff space if it is completely regular and Hausdorff.

Lemma 1.

A topological space X is completely regular if and only if for any nonempty closed set F, any xXF, and any distinct a,bR, there is a continuous function f:XR such f(x)=a and f(F)={b}.

Theorem 2.

If X is a Hausdorff space and AX, then A with the subspace topology is a Hausdorff space. If {Xi:iI} is a family of Hausdorff spaces, then iIXi is Hausdorff.

Proof.

Suppose that a,b are distinct points in A. Because X is Hausdorff, there are disjoint open sets U,V in X with aU,bV. Then UA,VA are disjoint open sets in A with the subspace topology and aUA,bVA, showing that A is Hausdorff.

Suppose that x,y are distinct elements of iIXi. x and y being distinct means there is some iI such that x(i)y(i). Then x(i),y(i) are distinct points in Xi, which is Hausdorff, so there are disjoint open sets Ui,Vi in Xi with x(i)Ui,y(i)Vi. Let U=πi-1(Ui),V=πi-1(Vi), where πi is the projection map from the product to Xi. U and V are disjoint, and xU,yV, showing that iIXi is Hausdorff. ∎

We prove that subspaces and products of completely regular spaces are completely regular.11 1 Stephen Willard, General Topology, p. 95, Theorem 14.10.

Theorem 3.

If X is Hausdorff and AX, then A with the subspace topology is completely regular. If {Xi:iI} is a family of completely regular spaces, then iIXi is completely regular.

Proof.

Suppose that F is closed in A with the subspace topology and xAF. There is a closed set G in X with F=GA. Then xG, so there is a continuous function f:X[0,1] satisfying f(x)=0 and f(F)={1}. The restriction of f to A with the subspace topology is continuous, showing that A is completely regular.

Suppose that F is a closed subset of X=iIXi and that xXF. A base for the product topology consists of intersections of finitely many sets of the form πi-1(Ui) where iI and Ui is an open subset of Xi, and because XF is an open neighborhood of x, there is a finite subset J of I and open sets Uj in Xj for jJ such that

xjJπj-1(Uj)XF.

For each jJ, XjUj is closed in Xj and x(j)Uj, and because Xj is completely regular there is a continuous function fj:Xj[0,1] such that fj(x(j))=0 and fj(XjUj)={1}. Define g:X[0,1] by

g(y)=maxjJ(fjπj)(y),yX.

In general, suppose that Y is a topological space and denote by C(Y) the set of continuous functions Y. It is a fact that C(Y) is a lattice with the partial order FG when F(y)G(y) for all yY. Hence, the maximum of finitely many continuous functions is also a continuous functions, hence g:X[0,1] is continuous. Because (fjπj)(x)=0 for each jJ, g(x)=0. On the other hand, FXjJπj-1(Uj), so if yF then there is some jJ such that πj(y)XjUj and then (fjπj)(y)=1. Hence, for any yF we have g(y)=1. Thus we have proved that g:X[0,1] is a continuous function such that g(x)=0 and g(F)={1}, which shows that X is completely regular. ∎

Therefore, subspaces and products of Tychonoff spaces are Tychonoff.

If X is a normal topological space, it is immediate from Urysohn’s lemma that X is completely regular. A metrizable space is normal and Hausdorff, so a metrizable space is thus a Tychonoff space. Let X be a locally compact Hausdorff space. Either X or the one-point compactification of X is a compact Hausdorff space Y of which X is a subspace. Y being a compact Hausdorff space implies that it is normal and hence completely regular. But X is a subspace of Y and being completely regular is a hereditary property, so X is completely regular, and therefore Tychonoff. Thus, we have proved that a locally compact Hausdorff space is Tychonoff.

2 Initial topologies

Suppose that X is a set, Xi, iI, are topological spaces, and fi:XXi are functions. The initial topology on X induced by {fi:iI} is the coarsest topology on X such that each fi is continuous. A subbase for the initial topology is the collection of those sets of the form fi-1(Ui), iI and Ui open in Xi.

If fi:XXi, iI, are functions, the evaluation map is the function e:XiIXi defined by

(πie)(x)=fi(x),iI.

We say that a collection {fi:iI} of functions on X separates points if xy implies that there is some iI such that fi(x)fi(y). We remind ourselves that if X and Y are topological spaces and ϕ:XY is a function, ϕ is called an embedding when ϕ:Xϕ(X) is a homeomorphism, where ϕ(X) has the subspace topology inherited from Y. The following theorem gives conditions on when X can be embedded into the product of the codomains of the fi.22 2 Stephen Willard, General Topology, p. 56, Theorem 8.12.

Theorem 4.

Let X be a topological space, let Xi, iI, be topological spaces, and let fi:XXi be functions. The evaluation map e:XiIXi is an embedding if and only if both (i) X has the initial topology induced by the family {fi:iI} and (ii) the family {fi:iI} separates points in X.

Proof.

Write P=iIXi and let pi:e(X)Xi be the restriction of πi:XXi to e(X). A subbase for e(X) with the subspace topology inherited from P consists of those sets of the form πi-1(Ui)e(X), iI and Ui open in Xi. But πi-1(Ui)e(X)=pi-1(Ui), and the collection of sets of this form is a subbase for e(X) with the initial topology induced by the family {pi:iI}, so these topologies are equal.

Assume that e:Xe(X) is a homeomorphism. Because e is a homeomorphism and fi=πie=pie, e(X) having the initial topology induced by {pi:iI} implies that X has the initial topology induced by {fi:iI}. If x,y are distinct elements of X then there is some iI such that pi(e(x))pi(e(y)), i.e. fi(x)fi(y), showing that {fi:iI} separates points in X.

Assume that X has the initial topology induced by {fi:iI} and that the family {fi:iI} separates points in X. We shall prove that e:Xe(X) is a homeomorphism, for which it suffices to prove that e:XP is one-to-one and continuous and that e:Xe(X) is open. If x,yX are distinct then because the fi separate points, there is some iI such that fi(x)fi(y), and so e(x)e(y), showing that e is one-to-one.

For each iI, fi is continuous and fi=πie. The fact that this is true for all iI implies that e:XP is continuous. (Because the product topology is the initial topology induced by the family of projection maps, a map to a product is continuous if and only if its composition with each projection map is continuous.)

A subbase for the topology of X consists of those sets of the form V=fi-1(Ui), iI and Ui open in Xi. As fi=pie we can write this as

V=(pie)-1(Ui)=e-1(pi-1(Ui)),

which implies that e(V)=pi-1(Ui), which is open in e(X) and thus shows that e:Xe(X) is open. ∎

We say that a collection {fi:iI} of functions on a topological space X separates points from closed sets if whenever F is a closed subset of X and xXF, there is some iI such that fi(x)fi(F)¯, where fi(F)¯ is the closure of fi(F) in the codomain of f.

Theorem 5.

Assume that X is a topological space and that fi:XXi, iI, are continuous functions. This family separates points from closed sets if and only if the collection of sets of the form fi-1(Ui), iI and Ui open in Xi, is a base for the topology of X.

Proof.

Assume that the family {fi:iI} separates points from closed sets in X. Say xX and that U is an open neighborhood of x. Then F=XU is closed so there is some iI such that fi(x)fi(F)¯. Thus Ui=Xifi(F)¯ is open in Xi, hence fi-1(Ui) is open in X. On the one hand, f(xi)Ui yields xifi-1(Ui). On the other hand, if yfi-1(Ui) then fi(y)Ui, which tells us that yF and so yU, giving fi-1(Ui)U. This shows us that the collection of sets of the form fi-1(Ui), iI and Ui open in Xi, is a base for the topology of X.

Assume that the collection of sets of the form fi-1(Ui), iI and Ui open in Xi, is a base for the topology of X, and suppose that F is a closed subset of X and that xXF. Because XF is an open neighborhood of x, there is some iI and open Ui in Xi such that xfi-1(Ui)XF, so fi(x)Ui. Suppose by contradiction that there is some yF such that fi(y)Ui. This gives yfi-1(Ui)XF, which contradicts yF. Therefore Uifi(F)=, and hence XiUi is a closed set that contains fi(F), which tells us that fi(F)¯XiUi, i.e. fi(F)¯Ui=. But fi(x)Ui, so we have proved that {fi:iI} separates points from closed sets. ∎

A T1 space is a topological space in which all singletons are closed.

Theorem 6.

If X is a T1 space, Xi, iI, are topological spaces, fi:XXi are continuous functions, and {fi:iI} separates points from closed sets in X, then the evaluation map e:XiIXi is an embedding.

Proof.

By Theorem 5, there is a base for the topology of X consisting of sets of the form fi-1(Ui), iI and Ui open in Xi. Since this collection of sets is a base it is a fortiori a subbase, and the topology generated by this subbase is the initial topology for the family of functions {fi:iI}. Because X is T1, singletons are closed and therefore the fact that {fi:iI} separates points and closed sets implies that it separates points in X. Therefore we can apply Theorem 4, which tells us that the evaluation map is an embedding. ∎

3 Bounded continuous functions

For any set X, we denote by (X) the set of all bounded functions X, and we take as known that (X) is a Banach space with the supremum norm

f=supxX|f(x)|,f(X).

If X is a topological space, we denote by Cb(X) the set of bounded continuous functions X. Cb(X)(X), and it is apparent that Cb(X) is a linear subspace of (X). One proves that Cb(X) is closed in (X) (i.e., that if a sequence of bounded continuous functions converges to some bounded function, then this function is continuous), and hence with the supremum norm, Cb(X) is a Banach space.

The following result shows that the Banach space Cb(X) of bounded continuous functions X is a useful collection of functions to talk about.33 3 Stephen Willard, General Topology, p. 96, Theorem 14.12.

Theorem 7.

Let X be a topological space. X is completely regular if and only if X has the initial topology induced by Cb(X).

Proof.

Assume that X is completely regular. If F is a closed subset of X and xXF, then there is a continuous function f:X[0,1] such that f(x)=0 and f(F)={1}. Then fCb(X), and f(x)=0{1}=f(F)¯. This shows that Cb(X) separates points from closed sets in X. Applying Theorem 5, we get that X has the initial topology induced by Cb(X). (This would follow if the collection that Theorem 5 tells us is a base were merely a subbase.)

Assume that X has the initial topology induced by Cb(X). Suppose that F is a closed subset of X and that xU=XF. A subbase for the initial topology induced by Cb(X) consists of those sets of the form f-1(V) for fCb(X) and V an open ray in (because the open rays are a subbase for the topology of ), so because U is an open neighborhood of x, there is a finite subset J of Cb(X) and open rays Vf in for each fJ such that

xfJf-1(Vf)U.

If some Vj is of the form (-,af), then with g=-f we have f-1(-,af)=g-1(-af,). We therefore suppose that in fact Vf=(af,) for each fJ. For each fJ, define gf:X by

gf(x)=sup{f(x)-af,0},

which is continuous and 0, and satisfies f-1(af,)=gf-1(0,), so that

xfJgf-1(0,)U.

Define g=fJgf, which is continuous because each factor is continuous. This function satisfies g(x)=fJgf(x)>0 because this is a product of finitely many factors each of which are >0. If yg-1(0,) then yfJgf-1(0,)U, so g-1(0,)U. But g is nonnegative, so this tells us that g(XU)={0}, i.e. g(F)={0}. By Lemma 1 this suffices to show that X is completely regular. ∎

A cube is a topological space that is homeomorphic to a product of compact intervals in . Any product is homeomorphic to the same product without singleton factors, (e.g. ××{3} is homeomorphic to ×) and a product of nonsingleton compact intervals with index set I is homeomorphic to [0,1]I. We remind ourselves that to say that a topological space is homeomorphic to a subspace of a cube is equivalent to saying that the space can be embedded into the cube.

Theorem 8.

A topological space X is a Tychonoff space if and only if it is homeomorphic to a subspace of a cube.

Proof.

Suppose that I is a set and that X is homeomorphic to a subspace Y of [0,1]I. [0,1] is Tychonoff so the product [0,1]I is Tychonoff, and hence the subspace Y is Tychonoff, thus X is Tychonoff.

Suppose that X is Tychonoff. By Theorem 7, X has the initial topology induced by Cb(X). For each fCb(X), let If=[-f,f], which is a compact interval in , and f:XIf is continuous. Because X is Tychonoff, it is T1 and the functions f:XIf, fCb(X), separate points and closed sets, we can now apply Theorem 6, which tells us that the evaluation map e:XfCb(X)If is an embedding. ∎

4 Compactifications

In §1 we talked about the one-point compactification of a locally compact Hausdorff space. A compactification of a topological space X is a pair (K,h) where (i) K is a compact Hausdorff space, (ii) h:XK is an embedding, and (iii) h(X) is a dense subset of K. For example, if X is a compact Hausdorff space then (X,idX) is a compactification of X, and if X is a locally compact Hausdorff space, then the one-point compactification X*=X{}, where is some symbol that does not belong to X, together with the inclusion map XX* is a compactification.

Suppose that X is a topological space and that (K,h) is a compactification of X. Because K is a compact Hausdorff space it is normal, and then Urysohn’s lemma tells us that K is completely regular. But K is Hausdorff, so in fact K is Tychonoff. A subspace of a Tychonoff space is Tychonoff, so h(X) with the subspace topology is Tychonoff. But X and h(X) are homeomorphic, so X is Tychonoff. Thus, if a topological space has a compactification then it is Tychonoff.

In Theorem 8 we proved that any Tychonoff space can be embedded into a cube. Here review our proof of this result. Let X be a Tychonoff space, and for each fCb(X) let If=[-f,f], so that f:XIf is continuous, and the family of these functions separates points in X. The evaluation map for this family is e:XfCb(X)If defined by (πfe)(x)=f(x) for fCb(X), and Theorem 6 tells us that e:XfCb(X)If is an embedding. Because each interval If is a compact Hausdorff space (we remark that if f=0 then If={0}, which is indeed compact), the product fCb(X)If is a compact Hausdorff space, and hence any closed subset of it is compact. We define βX to be the closure of e(X) in fCb(X)If, and the Stone-Čech compactification of X is the pair (βX,e), and what we have said shows that indeed this is a compactification of X.

The Stone-Čech compactification of a Tychonoff space is useful beyond displaying that every Tychonoff space has a compactification. We prove in the following that any continuous function from a Tychonoff space to a compact Hausdorff space factors through its Stone-Čech compactification.44 4 Stephen Willard, General Topology, p. 137, Theorem 19.5.

Theorem 9.

If X is a Tychonoff space, K is a compact Hausdorff space, and ϕ:XK is continuous, then there is a unique continuous function Φ:βXK such that ϕ=Φe.

Proof.

K is Tychonoff because a compact Hausdorff space is Tychonoff, so the evaluation map eK:KgCb(K)Ig is an embedding. Write F=fCb(X)If, G=gCb(K)Ig, and let pf:FIf, qg:GIg be the projection maps.

We define H:FG for tF by (qgH)(t)=t(gϕ)=pgϕ(t). For each gG, the map qgH:FIgϕ is continuous, so H is continuous.

For xX, we have

(qgHe)(x) = (qgH)(e(x))
= pgϕ(e(x))
= (pgϕe)(x)
= (gϕ)(x)
= g(ϕ(x))
= (qgeK)(ϕ(x))
= (qgeKϕ)(x),

so

He=eKϕ. (1)

On the one hand, because K is compact and eK is continuous, eK(K) is compact and hence is a closed subset of G (G is Hausdorff so a compact subset is closed). From (1) we know H(e(X))eK(K), and thus

H(e(X))¯eK(K)¯=eK(K).

On the other hand, because βX is compact and H is continuous, H(βX) is compact and hence is a closed subset of G. As e(X) is dense in βX and H is continuous, H(e(X)) is dense in H(βX), and thus

H(e(X))¯=H(βX)¯=H(βX).

Therefore we have

H(βX)eK(K).

Let h be the restriction of H to βX, and define Φ:βXK by Φ=eK-1h, which makes sense because eK:KeK(K) is a homeomorphism and h takes values in eK(K). Φ is continuous, and for xX we have, using (1),

(Φe)(x)=(eK-1he)(x)=(eK-1He)(x)=ϕ(x),

showing that Φe=ϕ.

If Ψ:βXK is a continuous function satisfying f=Ψe, let ye(X). There is some xX such that y=e(x), and f(x)=(Ψe)(x)=Ψ(y), f(x)=(Φe)(x)=Φ(y), showing that for all ye(X), Ψ(y)=Φ(y). Since Ψ and Φ are continuous and are equal on e(X), which is a dense subset of βX, we get Ψ=Φ, which completes the proof. ∎

If X is a Tychonoff space with Stone-Čech compactification (βX,e), then because βX is a compact space, C(βX) with the supremum norm is a Banach space. We show in the following that the extension in Theorem 9 produces an isometric isomorphism Cb(X)C(βX).

Theorem 10.

If X is a Tychonoff space with Stone-Čech compactification (βX,e), then there is an isomorphism of Banach spaces Cb(X)C(βX).

Proof.

Let f,gCb(X), let α be a scalar, and let K=[-|α|f-g,|α|f+g], which is a compact set. Define ϕ=αf+g, and then Theorem 9 tells us that there is a unique continuous function F:βXK such that f=Fe, a unique continuous function G:βXK such that g=Ge, and a unique continuous function Φ:βXK such that ϕ=Φe. For ye(X) and xX such that y=e(x),

Φ(y)=ϕ(x)=αf(x)+g(x)=αF(y)+G(y).

Since Φ and αF+G are continuous functions βXK that are equal on the dense set e(X), we get Φ=αF+G. Therefore, the map that sends fCb(X) to the unique FC(βX) such that f=Fe is linear.

Let fCb(X) and let F be the unique element of C(βX) such that f=Fe. For any xX, |f(x)|=|(Fe)(x)|, so

f=supxX|f(x)|=supxX|(Fe)(x)|=supye(X)|F(y)|.

Because F is continuous and e(X) is dense in βX,

supye(X)|F(y)|=supyβX|F(y)|=F,

so f=F, showing that fF is an isometry.

For ΦC(βX), define ϕ=Φe. Φ is bounded so ϕ is also, and ϕ is a composition of continuous functions, hence ϕCb(X). Thus ϕΦ is onto, completing the proof. ∎

5 Spaces of continuous functions

If X is a topological space, we denote by C(X) the set of continuous functions X. For K a compact set in X (in particular a singleton) and fC(X), define pK(f)=supxK|f(x)|. The collection of pK for all compact subsets of K of X is a separating family of seminorms, because if f is nonzero there is some xX for which f(x)0 and then p{x}(f)>0. Hence C(X) with the topology induced by this family of seminorms is a locally convex space. (If X is σ-compact then the seminorm topology is induced by countably many of the seminorms, and then C(X) is metrizable.) However, since we usually are not given that X is compact (in which case C(X) is normable with pX) and since it is often more convenient to work with normed spaces than with locally convex spaces, we shall talk about subsets of C(X).

For X a topological space, we say that a function f:X vanishes at infinity if for each ϵ>0 there is a compact set K such that |f(x)|<ϵ whenever xXK, and we denote by C0(X) the set of all continuous functions X that vanish at infinity.

The following theorem shows first that C0(X) is contained in Cb(X), second that C0(X) is a linear space, and third that it is a closed subset of Cb(X). With the supremum norm Cb(X) is a Banach space, so this shows that C0(X) is a Banach subspace. We work through the proof in detail because it is often proved with unnecessary assumptions on the topological space X.

Theorem 11.

Suppose that X is a topological space. Then C0(X) is a closed linear subspace of Cb(X).

Proof.

If fC0(X), then there is a compact set K such that xXK implies that |f(x)|<1. On the other hand, because f is continuous, f(K) is a compact subset of the scalar field and hence is bounded, i.e., there is some M0 such that xK implies that |f(x)|M. Therefore f is bounded, showing that C0(X)Cb(X).

Let f,gC0(X) and let ϵ>0. There is a compact set K1 such that xXK1 implies that |f(x)|<ϵ2 and a compact set K2 such that xXK2 implies that |g(x)|<ϵ2. Let K=K1K2, which is a union of two compact sets hence is itself compact. If xXK, then xXK1 implying |f(x)|<ϵ2 and xXK2 implying |g(x)|<ϵ2, hence |f(x)+g(x)||f(x)|+|g(x)|<ϵ. This shows that f+gC0(X).

If fC0(X) and α is a nonzero scalar, let ϵ>0. There is a compact set K such that xXK implies that |f(x)|<ϵ|α|, and hence |(αf)(x)|=|α||f(x)|<ϵ, showing that αfC0(X). Therefore C0(X) is a linear subspace of Cb(X).

Suppose that fn is a sequence of elements of C0(X) that converges to some fCb(X). For ϵ>0, there is some nϵ such that nnϵ implies that fn-f<ϵ2, that is,

supxX|fn(x)-f(x)|<ϵ2.

For each n, let Kn be a compact set in X such that xXKn implies that |fn(x)|<ϵ2; there are such Kn because fnC0(X). If xXKnϵ, then

|f(x)||fnϵ(x)-f(x)|+|fnϵ(x)|<ϵ2+ϵ2=ϵ,

showing that fC0(X). ∎

If X is a topological space and f:X is a function, the support of f is the set

suppf={xX:f(x)0}¯.

If suppf is compact we say that f has compact support, and we denote by Cc(X) the set of all continuous functions X with compact support.

Suppose that X is a topological space and let fCc(X). For any ϵ>0, if xXsuppf then |f(x)|=0<ϵ, showing that fC0(X). Therefore

Cc(X)C0(X),

and this makes no assumptions about the topology of X.

We can prove that if X is a locally compact Hausdorff space then Cc(X) is dense in C0(X).55 5 Gerald B. Folland, Real Analysis: Modern Techniques and Their Applications, p. 132, Proposition 4.35.

Theorem 12.

If X is a locally compact Hausdorff space, then Cc(X) is a dense subset of C0(X).

Proof.

Let fC0(X), and for each n define

Cn={xX:|f(x)|1n}.

For n, because fC0(X) there is a compact set Kn such that xXKn implies that |f(x)|<1n, and hence CnKn. Because x|fn(x)| is continuous, Cn is a closed set in X, and it follows that Cn, being contained in the compact set Kn, is compact. (This does not use that X is Hausdorff.)

Let n. Because X is a locally compact Hausdorff space and Cn is compact, Urysohn’s lemma66 6 Gerald B. Folland, Real Analysis: Modern Techniques and Their Applications, p. 131, Lemma 4.32. tells us that there is a compact set Dn containing Cn and a continuous function gn:X[0,1] such that gn(Cn)={1} and gn(XDn){0}. That is, gnCc(X), 0gn1, and gn(Cn)={1}. Define fn=gnfCc(X). (A product of continuous functions is continuous, and because f is bounded and gn has compact support, gnf has compact support.) For xCn, fn(x)-f(x)=(gn(x)-1)f(x)=0, and for xXCn, |fn(x)-f(x)|=|gn(x)-1||f(x)|11n. Therefore

fn-f1n,

and hence fn is a sequence in Cc(X) that converges to f, showing that Cc(X) is dense in C0(X). ∎

If X is a Hausdorff space, then we prove that Cc(X) is a linear subspace of C0(X). When X is a locally compact Hausdorff space then combined with the above this shows that Cc(X) is a dense linear subspace of C0(X).

Lemma 13.

Suppose that X is a Hausdorff space. Then Cc(X) is a linear subspace of C0(X).

Proof.

If f,gCc(X) and α is a scalar, let K=suppfsuppg, which is a union of two compact sets hence compact. If xXK, then f(x)=0 because xsuppf and g(x)=0 because xsuppg, so (αf+g)(x)=0. Therefore {xX:(αf+g)(x)0}K and hence supp(αf+g)K¯. But as X is Hausdorff, K being compact implies that K is closed in X, so we get supp(αf+g)K. Because supp(αf+g) is closed and is contained in the compact set K, it is itself compact, so αf+gCc(X). ∎

Let X be a topological space, and for xX define δx:Cb(X) by δx(f)=f(x). For each xX, δx is linear and |δx(f)|=|f(x)|f, so δx is continuous and hence belongs to the dual space Cb(X)*. Moreover, the constant function f(x)=1 shows that δx=1. We define Δ:XCb(X)* by Δ(x)=δx. Suppose that xi is a net in X that converges to some xX. Then for every fCb(X) we have f(xi)f(x), and this means that δxi weak-* converges to δx in Cb(X)*. This shows that with Cb(X)* assigned the weak-* topology, Δ:XCb(X)* is continuous. We now characterize when Δ is an embedding.77 7 John B. Conway, A Course in Functional Analysis, second ed., p. 137, Proposition 6.1.

Theorem 14.

Suppose that X is a topological space and assign Cb(X)* the weak-* topology. Then the map Δ:XΔ(X) is a homeomorphism if and only if X is Tychonoff, where Δ(X) has the subspace topology inherited from Cb(X)*.

Proof.

Suppose that X is Tychonoff. If x,yX are distinct, then there is some fCb(X) such that f(x)=0 and f(y)=1, and then δx(f)=01=δy(f), so Δ(x)Δ(y), showing that Δ is one-to-one. To show that Δ:XΔ(X) is a homeomorphism, it suffices to prove that Δ is an open map, so let U be an open subset of X. For x0U, because XU is closed there is some fCb(X) such that f(x0)=0 and f(XU)={1}. Let

V1={μCb(X)*:μ(f)<1}.

This is an open subset of Cb(X)* as it is the inverse image of (-,1) under the map μμ(f), which is continuous Cb(X)* by definition of the weak-* topology. Then

V=V1Δ(X)={δx:f(x)<1}

is an open subset of the subspace Δ(X), and we have both δx0V and VΔ(U). This shows that for any element δx0 of Δ(U), there is some open set V in the subspace Δ(X) such that δx0VΔ(U), which tells us that Δ(U) is an open set in the subspace Δ(U), showing that Δ is an open map and therefore a homeomorphism.

Suppose that Δ:XΔ(X) is a homeomorphism. By the Banach-Alaoglu theorem we know that the closed unit ball B1 in Cb(X)* is compact. (We remind ourselves that we have assigned Cb(X)* the weak-* topology.) That is, with the subspace topology inherited from Cb(X)*, B1 is a compact space. It is Hausdorff because Cb(X)* is Hausdorff, and a compact Hausdorff space is Tychonoff. But Δ(X) is contained in the surface of B1, in particular Δ(X) is contained in B1 and hence is itself Tychonoff with the subspace topology inherited from B1, which is equal to the subspace topology inherited from Cb(X)*. Since Δ:XΔ(X) is a homeomorphism, we get that X is a Tychonoff space, completing the proof. ∎

The following result shows when the Banach space Cb(X) is separable.88 8 John B. Conway, A Course in Functional Analysis, second ed., p. 140, Theorem 6.6.

Theorem 15.

Suppose that X is a Tychonoff space. Then the Banach space Cb(X) is separable if and only if X is compact and metrizable.

Proof.

Assume that X is compact and metrizable, with a compatible metric d. For each n there are open balls Un,1,,Un,Nn of radius 1n that cover X. As X is metrizable it is normal, so there is a partition of unity subordinate to the cover {Un,k:1kNn}.99 9 John B. Conway, A Course in Functional Analysis, second ed., p. 139, Theorem 6.5. That is, there are continuous functions fn,1,,fn,Nn:X[0,1] such that k=1Nnfn,k=1 and such that xXUn,k implies that fn,k(x)=0. Then {fn,k:n,1kNn} is countable, so its span D over is also countable. We shall prove that D is dense in C(X)=Cb(X), which will show that Cb(X) is separable.

Let fC(X) and let ϵ>0. Because (X,d) is a compact metric space, f is uniformly continuous, so there is some δ>0 such that d(x,y)<δ implies that |f(x)-f(y)|<ϵ2. Let n be >2δ, and for each 1kNn let xkUn,k. For each k there is some αk such that |αk-f(xk)|<ϵ2, and we define

g=k=1Nnαkfn,kD.

Because k=1Nnfn,k=1 we have f=k=1Nnffn,k. Let xX, and then

|f(x)-g(x)|=|k=1Nn(f(x)-αk)fn,k(x)|k=1Nn|f(x)-αk|fn,k(x).

For each 1kNn, either xUn,k or xUn,k. In the first case, since x and xk are then in the same open ball of radius 1n, d(x,xk)<2n<δ, so

|f(x)-αk||f(x)-f(xk)|+|f(xk)-αk|<ϵ2+ϵ2=ϵ.

In the second case, fn,k(x)=0. Therefore,

k=1Nn|f(x)-αk|fn,k(x)k=1Nnϵfn,k(x)=ϵ,

showing that |f(x)-g(x)|ϵ. This shows that D is dense in C(X), and therefore that Cb(X)=C(X) is separable.

Suppose that Cb(X) is separable. Because X is Tychonoff, by Theorem 10 there is an isometric isomorphism between the Banach spaces Cb(X) and C(βX), where (βX,e) is the Stone-Čech compactification of X. Hence C(βX) is separable. But it is a fact that a compact Hausdorff space Y is metrizable if and only if the Banach space C(Y) is separable.1010 10 Charalambos D. Aliprantis and Kim C. Border, Infinite Dimensional Analysis: A Hitchhiker’s Guide, third ed., p. 353, Theorem 9.14. (This is proved using the Stone-Weierstrass theorem.) As βX is a compact Hausdorff space and C(βX) is separable, we thus get that βX is metrizable.

It is a fact that if Y is a Banach space and B1 is the closed unit ball in the dual space Y*, then B1 with the subspace topology inherited from Y* with the weak-* topology is metrizable if and only if Y is separable.1111 11 John B. Conway, A Course in Functional Analysis, second ed., p. 134, Theorem 5.1. Thus, the closed unit ball B1 in Cb(X)* is metrizable. Theorem 14 tells us there is an embedding Δ:XB1, and B1 being metrizable implies that Δ(X) is metrizable. As Δ:XΔ(X) is a homeomorphism, we get that X is metrizable.

Because βX is compact and metrizable, to prove that X is compact and metrizable it suffices to prove that βXe(X)=, so we suppose by contradiction that there is some τβXe(X). e(X) is dense in βX, so there is a sequence xnX, for which we take xnxm when nm, such that e(xn)τ. If xn had a subsequence xa(n) that converged to some yX, then e(xa(n))e(y) and hence e(y)=τ, a contradiction. Therefore the sequence xn has no limit points, so the sets A={xn:n odd} and B={xn:n even} are closed and disjoint. Because X is metrizable it is normal, hence by Urysohn’s lemma there is a continuous function ϕ:X[0,1] such that ϕ(a)=0 for all aA and ϕ(b)=1 for all bB. Then, by Theorem 9 there is a unique continuous Φ:X[0,1] such that ϕ=Φe. Then we have, because a subsequence of a convergent sequence has the same limit,

Φ(τ) = Φ(limne(xn))
= Φ(limne(x2n+1))
= limn(Φe)(x2n+1)
= limnϕ(x2n+1)
= 0,

and likewise

Φ(τ)=limnϕ(x2n)=1,

a contradiction. This shows that βXe(X)=, which completes the proof. ∎

6 𝐶*-algebras and the Gelfand transform

A C*-algebra is a complex Banach algebra A with a map :*AA such that

  1. 1.

    a**=a for all aA (namely, * is an involution),

  2. 2.

    (a+b)*=a*+b* and (ab)*=b*a* for all aA,

  3. 3.

    (λa)*=λ¯a* for all aA and λ,

  4. 4.

    a*a=a2 for all aA.

We do not require that a C*-algebra be unital. If a=0 then a*=0=a. Otherwise,

a2=a*aa*a

gives aa* and

a*2=a**a*=aa*aa*

gives a*a, showing that * is an isometry.

We now take Cb(X) to denote Cb(X,) rather than Cb(X,), and likewise for C(X),C0(X), and Cc(X). It is routine to verify that everything we have asserted about these spaces when the codomain is is true when the codomain is , but this is not obvious. In particular, Cb(X) is a Banach space with the supremum norm and C0(X) is a closed linear subspace, whatever the topological space X. It is then straightforward to check that with the involution f*=f¯ they are commutative C*-algebras.

A homomorphism of C*-algebras is an algebra homomorphism f:AB, where A and B are C*-algebras, such that f(a*)=f(a)* for all aA. It can be proved that f1.1212 12 José M. Gracia-Bondía, Joseph C. Várilly and Héctor Figueroa, Elements of Noncommutative Geometry, p. 29, Lemma 1.16. We define an isomorphism of C*-algebras to be an algebra isomorphism f:AB such that f(a*)=f(a)* for all aA. It follows that f1 and because f is bijective, the inverse f-1 is a C*-algebra homomorphism, giving f-11 and therefore f=1. Thus, an isomorphism of C*-algebras is an isometric isomorphism.

Suppose that A is a commutative C*-algebra, which we do not assume to be unital. A character of A is a nonzero algebra homomorphism A. We denote the set of characters of A by σ(A), which we call the Gelfand spectrum of A. We make some assertions in the following text that are proved in Folland.1313 13 Gerald B. Folland, A Course in Abstract Harmonic Analysis, p. 12, §1.3. It is a fact that for every hσ(A), h1, so σ(A) is contained in the closed unit ball of A*, where A* denotes the dual of the Banach space A. Furthermore, one can prove that σ(A){0} is a weak-* closed set in A*, and hence is weak-* compact because it is contained in the closed unit ball which we know to be weak-* compact by the Banach-Alaoglu theorem. We assign σ(A) the subspace topology inherited from A* with the weak-* topology. Depending on whether 0 is or is not an isolated point in σ(A){0}, σ(A) is a compact or a locally compact Hausdorff space; in any case σ(A) is a locally compact Hausdorff space.

The Gelfand transform is the map Γ:AC0(σ(A)) defined by Γ(a)(h)=h(a); that Γ(a) is continuous follows from σ(A) having the weak-* topology, and one proves that in fact Γ(a)C0(σ(A)).1414 14 Gerald B. Folland, A Course in Abstract Harmonic Analysis, p. 15. The Gelfand-Naimark theorem1515 15 Gerald B. Folland, A Course in Abstract Harmonic Analysis, p. 16, Theorem 1.31. states that Γ:AC0(σ(A)) is an isomorphism of C*-algebras.

It can be proved that two commutative C*-algebras are isomorphic as C*-algebras if and only if their Gelfand spectra are homeomorphic.1616 16 José M. Gracia-Bondía, Joseph C. Várilly and Héctor Figueroa, Elements of Noncommutative Geometry, p. 11, Proposition 1.5.

7 Multiplier algebras

An ideal of a C*-algebra A is a closed linear subspace I of A such that IAI and AII. An ideal I is said to be essential if IJ{0} for every nonzero ideal J of A. In particular, A is itself an essential ideal.

Suppose that A is a C*-algebra. The multiplier algebra of A, denoted M(A), is a C*-algebra containing A as an essential ideal such that if B is a C*-algebra containing A as an essential ideal then there is a unique homomorphism of C*-algebras π:BM(A) whose restriction to A is the identity. We have not shown that there is a multiplier algebra of A, but we shall now prove that this definition is a universal property: that any C*-algebra satisfying the definition is isomorphic as a C*-algebra to M(A), which allows us to talk about “the” multiplier algebra rather than “a” multiplier algebra.

Suppose that C is a C*-algebra containing A as an essential ideal such that if B is a C*-algebra containing A as an essential ideal then there is a unique C*-algebra homomorphism π:BC whose restriction to A is the identity. Hence there is a unique homomorphism of C*-algebras π1:CM(A) whose restriction to A is the identity, and there is a unique homomorphism of C*-algebras π2:M(A)C whose restriction to A is the identity. Then π2π1:CC and π1π2:M(A)M(A) are homomorphisms of C*-algebras whose restrictions to A are the identity. But the identity maps idC:CC and idM(A):M(A)M(A) are also homomorphisms of C*-algebras whose restrictions to A are the identity. Therefore, by uniqueness we get that π2π1=idC and π1π2=idM(A). Therefore π1:CM(A) is an isomorphism of C*-algebras.

One can prove that if A is unital then M(A)=A.1717 17 Paul Skoufranis, An Introduction to Multiplier Algebras, http://www.math.ucla.edu/~pskoufra/OANotes-MultiplierAlgebras.pdf, p. 4, Lemma 1.9. It can be proved that for any C*-algebra A, the multiplier algebra M(A) is unital.1818 18 Paul Skoufranis, An Introduction to Multiplier Algebras, http://www.math.ucla.edu/~pskoufra/OANotes-MultiplierAlgebras.pdf, p. 9, Corollary 2.8. For a locally compact Hausdorff space X, it can be proved that M(C0(X))=Cb(X).1919 19 Eberhard Kaniuth, A Course in Commutative Banach Algebras, p. 29, Example 1.4.13; José M. Gracia-Bondía, Joseph C. Várilly and Héctor Figueroa, Elements of Noncommutative Geometry, p. 14, Proposition 1.10. This last assertion is the reason for my interest in multiplier algebras. We have seen that if X is a locally compact Hausdorff space then Cc(X) is a dense linear subspace of C0(X), and for any topological space C0(X) is a closed linear subspace of Cb(X), but before talking about multiplier algebras we did not have a tight fit between the C*-algebras C0(X) and Cb(X).

8 Riesz representation theorem for compact Hausdorff spaces

There is a proof due to D. J. H. Garling of the Riesz representation theorem for compact Hausdorff spaces that uses the Stone-Čech compactification of discrete topological spaces. This proof is presented in Carothers’ book.2020 20 N. L. Carothers, A Short Course on Banach Space Theory, Chapter 16, pp. 156–165.