The Stone-Čech compactification of Tychonoff spaces
1 Completely regular spaces and Tychonoff spaces
A topological space is said to be completely regular if whenever is a nonempty closed set and , there is a continuous function such that and . A completely regular space need not be Hausdorff. For example, if is any set with more than one point, then the trivial topology, in which the only closed sets are and , 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 is completely regular if and only if for any nonempty closed set , any , and any distinct , there is a continuous function such and .
Theorem 2.
If is a Hausdorff space and , then with the subspace topology is a Hausdorff space. If is a family of Hausdorff spaces, then is Hausdorff.
Proof.
Suppose that are distinct points in . Because is Hausdorff, there are disjoint open sets in with . Then are disjoint open sets in with the subspace topology and , showing that is Hausdorff.
Suppose that are distinct elements of . and being distinct means there is some such that . Then are distinct points in , which is Hausdorff, so there are disjoint open sets in with . Let , where is the projection map from the product to . and are disjoint, and , showing that 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 is Hausdorff and , then with the subspace topology is completely regular. If is a family of completely regular spaces, then is completely regular.
Proof.
Suppose that is closed in with the subspace topology and . There is a closed set in with . Then , so there is a continuous function satisfying and . The restriction of to with the subspace topology is continuous, showing that is completely regular.
Suppose that is a closed subset of and that . A base for the product topology consists of intersections of finitely many sets of the form where and is an open subset of , and because is an open neighborhood of , there is a finite subset of and open sets in for such that
For each , is closed in and , and because is completely regular there is a continuous function such that and . Define by
In general, suppose that is a topological space and denote by the set of continuous functions . It is a fact that is a lattice with the partial order when for all . Hence, the maximum of finitely many continuous functions is also a continuous functions, hence is continuous. Because for each , . On the other hand, , so if then there is some such that and then . Hence, for any we have . Thus we have proved that is a continuous function such that and , which shows that is completely regular. ∎
Therefore, subspaces and products of Tychonoff spaces are Tychonoff.
If is a normal topological space, it is immediate from Urysohn’s lemma that is completely regular. A metrizable space is normal and Hausdorff, so a metrizable space is thus a Tychonoff space. Let be a locally compact Hausdorff space. Either or the one-point compactification of is a compact Hausdorff space of which is a subspace. being a compact Hausdorff space implies that it is normal and hence completely regular. But is a subspace of and being completely regular is a hereditary property, so is completely regular, and therefore Tychonoff. Thus, we have proved that a locally compact Hausdorff space is Tychonoff.
2 Initial topologies
Suppose that is a set, , , are topological spaces, and are functions. The initial topology on induced by is the coarsest topology on such that each is continuous. A subbase for the initial topology is the collection of those sets of the form , and open in .
If , , are functions, the evaluation map is the function defined by
We say that a collection of functions on separates points if implies that there is some such that . We remind ourselves that if and are topological spaces and is a function, is called an embedding when is a homeomorphism, where has the subspace topology inherited from . The following theorem gives conditions on when can be embedded into the product of the codomains of the .22 2 Stephen Willard, General Topology, p. 56, Theorem 8.12.
Theorem 4.
Let be a topological space, let , , be topological spaces, and let be functions. The evaluation map is an embedding if and only if both (i) has the initial topology induced by the family and (ii) the family separates points in .
Proof.
Write and let be the restriction of to . A subbase for with the subspace topology inherited from consists of those sets of the form , and open in . But , and the collection of sets of this form is a subbase for with the initial topology induced by the family , so these topologies are equal.
Assume that is a homeomorphism. Because is a homeomorphism and , having the initial topology induced by implies that has the initial topology induced by . If are distinct elements of then there is some such that , i.e. , showing that separates points in .
Assume that has the initial topology induced by and that the family separates points in . We shall prove that is a homeomorphism, for which it suffices to prove that is one-to-one and continuous and that is open. If are distinct then because the separate points, there is some such that , and so , showing that is one-to-one.
For each , is continuous and . The fact that this is true for all implies that 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 consists of those sets of the form , and open in . As we can write this as
which implies that , which is open in and thus shows that is open. ∎
We say that a collection of functions on a topological space separates points from closed sets if whenever is a closed subset of and , there is some such that , where is the closure of in the codomain of .
Theorem 5.
Assume that is a topological space and that , , are continuous functions. This family separates points from closed sets if and only if the collection of sets of the form , and open in , is a base for the topology of .
Proof.
Assume that the family separates points from closed sets in . Say and that is an open neighborhood of . Then is closed so there is some such that . Thus is open in , hence is open in . On the one hand, yields . On the other hand, if then , which tells us that and so , giving . This shows us that the collection of sets of the form , and open in , is a base for the topology of .
Assume that the collection of sets of the form , and open in , is a base for the topology of , and suppose that is a closed subset of and that . Because is an open neighborhood of , there is some and open in such that , so . Suppose by contradiction that there is some such that . This gives , which contradicts . Therefore , and hence is a closed set that contains , which tells us that , i.e. . But , so we have proved that separates points from closed sets. ∎
A space is a topological space in which all singletons are closed.
Theorem 6.
If is a space, , , are topological spaces, are continuous functions, and separates points from closed sets in , then the evaluation map is an embedding.
Proof.
By Theorem 5, there is a base for the topology of consisting of sets of the form , and open in . 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 . Because is , singletons are closed and therefore the fact that separates points and closed sets implies that it separates points in . Therefore we can apply Theorem 4, which tells us that the evaluation map is an embedding. ∎
3 Bounded continuous functions
For any set , we denote by the set of all bounded functions , and we take as known that is a Banach space with the supremum norm
If is a topological space, we denote by the set of bounded continuous functions . , and it is apparent that is a linear subspace of . One proves that is closed in (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, is a Banach space.
The following result shows that the Banach space of bounded continuous functions is a useful collection of functions to talk about.33 3 Stephen Willard, General Topology, p. 96, Theorem 14.12.
Theorem 7.
Let be a topological space. is completely regular if and only if has the initial topology induced by .
Proof.
Assume that is completely regular. If is a closed subset of and , then there is a continuous function such that and . Then , and . This shows that separates points from closed sets in . Applying Theorem 5, we get that has the initial topology induced by . (This would follow if the collection that Theorem 5 tells us is a base were merely a subbase.)
Assume that has the initial topology induced by . Suppose that is a closed subset of and that . A subbase for the initial topology induced by consists of those sets of the form for and an open ray in (because the open rays are a subbase for the topology of ), so because is an open neighborhood of , there is a finite subset of and open rays in for each such that
If some is of the form , then with we have . We therefore suppose that in fact for each . For each , define by
which is continuous and , and satisfies , so that
Define , which is continuous because each factor is continuous. This function satisfies because this is a product of finitely many factors each of which are . If then , so . But is nonnegative, so this tells us that , i.e. . By Lemma 1 this suffices to show that 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. is homeomorphic to ) and a product of nonsingleton compact intervals with index set is homeomorphic to . 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 is a Tychonoff space if and only if it is homeomorphic to a subspace of a cube.
Proof.
Suppose that is a set and that is homeomorphic to a subspace of . is Tychonoff so the product is Tychonoff, and hence the subspace is Tychonoff, thus is Tychonoff.
Suppose that is Tychonoff. By Theorem 7, has the initial topology induced by . For each , let , which is a compact interval in , and is continuous. Because is Tychonoff, it is and the functions , , separate points and closed sets, we can now apply Theorem 6, which tells us that the evaluation map 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 is a pair where (i) is a compact Hausdorff space, (ii) is an embedding, and (iii) is a dense subset of . For example, if is a compact Hausdorff space then is a compactification of , and if is a locally compact Hausdorff space, then the one-point compactification , where is some symbol that does not belong to , together with the inclusion map is a compactification.
Suppose that is a topological space and that is a compactification of . Because is a compact Hausdorff space it is normal, and then Urysohn’s lemma tells us that is completely regular. But is Hausdorff, so in fact is Tychonoff. A subspace of a Tychonoff space is Tychonoff, so with the subspace topology is Tychonoff. But and are homeomorphic, so 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 be a Tychonoff space, and for each let , so that is continuous, and the family of these functions separates points in . The evaluation map for this family is defined by for , and Theorem 6 tells us that is an embedding. Because each interval is a compact Hausdorff space (we remark that if then , which is indeed compact), the product is a compact Hausdorff space, and hence any closed subset of it is compact. We define to be the closure of in , and the Stone-Čech compactification of is the pair , and what we have said shows that indeed this is a compactification of .
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 is a Tychonoff space, is a compact Hausdorff space, and is continuous, then there is a unique continuous function such that .
Proof.
is Tychonoff because a compact Hausdorff space is Tychonoff, so the evaluation map is an embedding. Write , , and let , be the projection maps.
We define for by . For each , the map is continuous, so is continuous.
For , we have
so
(1) |
On the one hand, because is compact and is continuous, is compact and hence is a closed subset of ( is Hausdorff so a compact subset is closed). From (1) we know , and thus
On the other hand, because is compact and is continuous, is compact and hence is a closed subset of . As is dense in and is continuous, is dense in , and thus
Therefore we have
Let be the restriction of to , and define by , which makes sense because is a homeomorphism and takes values in . is continuous, and for we have, using (1),
showing that .
If is a continuous function satisfying , let . There is some such that , and , , showing that for all , . Since and are continuous and are equal on , which is a dense subset of , we get , which completes the proof. ∎
If is a Tychonoff space with Stone-Čech compactification , then because is a compact space, with the supremum norm is a Banach space. We show in the following that the extension in Theorem 9 produces an isometric isomorphism .
Theorem 10.
If is a Tychonoff space with Stone-Čech compactification , then there is an isomorphism of Banach spaces .
Proof.
Let , let be a scalar, and let , which is a compact set. Define , and then Theorem 9 tells us that there is a unique continuous function such that , a unique continuous function such that , and a unique continuous function such that . For and such that ,
Since and are continuous functions that are equal on the dense set , we get . Therefore, the map that sends to the unique such that is linear.
Let and let be the unique element of such that . For any , , so
Because is continuous and is dense in ,
so , showing that is an isometry.
For , define . is bounded so is also, and is a composition of continuous functions, hence . Thus is onto, completing the proof. ∎
5 Spaces of continuous functions
If is a topological space, we denote by the set of continuous functions . For a compact set in (in particular a singleton) and , define . The collection of for all compact subsets of of is a separating family of seminorms, because if is nonzero there is some for which and then . Hence with the topology induced by this family of seminorms is a locally convex space. (If is -compact then the seminorm topology is induced by countably many of the seminorms, and then is metrizable.) However, since we usually are not given that is compact (in which case is normable with ) and since it is often more convenient to work with normed spaces than with locally convex spaces, we shall talk about subsets of .
For a topological space, we say that a function vanishes at infinity if for each there is a compact set such that whenever , and we denote by the set of all continuous functions that vanish at infinity.
The following theorem shows first that is contained in , second that is a linear space, and third that it is a closed subset of . With the supremum norm is a Banach space, so this shows that is a Banach subspace. We work through the proof in detail because it is often proved with unnecessary assumptions on the topological space .
Theorem 11.
Suppose that is a topological space. Then is a closed linear subspace of .
Proof.
If , then there is a compact set such that implies that . On the other hand, because is continuous, is a compact subset of the scalar field and hence is bounded, i.e., there is some such that implies that . Therefore is bounded, showing that .
Let and let . There is a compact set such that implies that and a compact set such that implies that . Let , which is a union of two compact sets hence is itself compact. If , then implying and implying , hence . This shows that .
If and is a nonzero scalar, let . There is a compact set such that implies that , and hence , showing that . Therefore is a linear subspace of .
Suppose that is a sequence of elements of that converges to some . For , there is some such that implies that , that is,
For each , let be a compact set in such that implies that ; there are such because . If , then
showing that . ∎
If is a topological space and is a function, the support of is the set
If is compact we say that has compact support, and we denote by the set of all continuous functions with compact support.
Suppose that is a topological space and let . For any , if then , showing that . Therefore
and this makes no assumptions about the topology of .
We can prove that if is a locally compact Hausdorff space then is dense in .55 5 Gerald B. Folland, Real Analysis: Modern Techniques and Their Applications, p. 132, Proposition 4.35.
Theorem 12.
If is a locally compact Hausdorff space, then is a dense subset of .
Proof.
Let , and for each define
For , because there is a compact set such that implies that , and hence . Because is continuous, is a closed set in , and it follows that , being contained in the compact set , is compact. (This does not use that is Hausdorff.)
Let . Because is a locally compact Hausdorff space and 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 containing and a continuous function such that and . That is, , , and . Define . (A product of continuous functions is continuous, and because is bounded and has compact support, has compact support.) For , , and for , . Therefore
and hence is a sequence in that converges to , showing that is dense in . ∎
If is a Hausdorff space, then we prove that is a linear subspace of . When is a locally compact Hausdorff space then combined with the above this shows that is a dense linear subspace of .
Lemma 13.
Suppose that is a Hausdorff space. Then is a linear subspace of .
Proof.
If and is a scalar, let , which is a union of two compact sets hence compact. If , then because and because , so . Therefore and hence . But as is Hausdorff, being compact implies that is closed in , so we get . Because is closed and is contained in the compact set , it is itself compact, so . ∎
Let be a topological space, and for define by . For each , is linear and , so is continuous and hence belongs to the dual space . Moreover, the constant function shows that . We define by . Suppose that is a net in that converges to some . Then for every we have , and this means that weak-* converges to in . This shows that with assigned the weak-* topology, is continuous. We now characterize when is an embedding.77 7 John B. Conway, A Course in Functional Analysis, second ed., p. 137, Proposition 6.1.
Theorem 14.
Suppose that is a topological space and assign the weak-* topology. Then the map is a homeomorphism if and only if is Tychonoff, where has the subspace topology inherited from .
Proof.
Suppose that is Tychonoff. If are distinct, then there is some such that and , and then , so , showing that is one-to-one. To show that is a homeomorphism, it suffices to prove that is an open map, so let be an open subset of . For , because is closed there is some such that and . Let
This is an open subset of as it is the inverse image of under the map , which is continuous by definition of the weak-* topology. Then
is an open subset of the subspace , and we have both and . This shows that for any element of , there is some open set in the subspace such that , which tells us that is an open set in the subspace , showing that is an open map and therefore a homeomorphism.
Suppose that is a homeomorphism. By the Banach-Alaoglu theorem we know that the closed unit ball in is compact. (We remind ourselves that we have assigned the weak-* topology.) That is, with the subspace topology inherited from , is a compact space. It is Hausdorff because is Hausdorff, and a compact Hausdorff space is Tychonoff. But is contained in the surface of , in particular is contained in and hence is itself Tychonoff with the subspace topology inherited from , which is equal to the subspace topology inherited from . Since is a homeomorphism, we get that is a Tychonoff space, completing the proof. ∎
The following result shows when the Banach space is separable.88 8 John B. Conway, A Course in Functional Analysis, second ed., p. 140, Theorem 6.6.
Theorem 15.
Suppose that is a Tychonoff space. Then the Banach space is separable if and only if is compact and metrizable.
Proof.
Assume that is compact and metrizable, with a compatible metric . For each there are open balls of radius that cover . As is metrizable it is normal, so there is a partition of unity subordinate to the cover .99 9 John B. Conway, A Course in Functional Analysis, second ed., p. 139, Theorem 6.5. That is, there are continuous functions such that and such that implies that . Then is countable, so its span over is also countable. We shall prove that is dense in , which will show that is separable.
Let and let . Because is a compact metric space, is uniformly continuous, so there is some such that implies that . Let be , and for each let . For each there is some such that , and we define
Because we have . Let , and then
For each , either or . In the first case, since and are then in the same open ball of radius , , so
In the second case, . Therefore,
showing that . This shows that is dense in , and therefore that is separable.
Suppose that is separable. Because is Tychonoff, by Theorem 10 there is an isometric isomorphism between the Banach spaces and , where is the Stone-Čech compactification of . Hence is separable. But it is a fact that a compact Hausdorff space is metrizable if and only if the Banach space 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 is a compact Hausdorff space and is separable, we thus get that is metrizable.
It is a fact that if is a Banach space and is the closed unit ball in the dual space , then with the subspace topology inherited from with the weak-* topology is metrizable if and only if is separable.1111 11 John B. Conway, A Course in Functional Analysis, second ed., p. 134, Theorem 5.1. Thus, the closed unit ball in is metrizable. Theorem 14 tells us there is an embedding , and being metrizable implies that is metrizable. As is a homeomorphism, we get that is metrizable.
Because is compact and metrizable, to prove that is compact and metrizable it suffices to prove that , so we suppose by contradiction that there is some . is dense in , so there is a sequence , for which we take when , such that . If had a subsequence that converged to some , then and hence , a contradiction. Therefore the sequence has no limit points, so the sets and are closed and disjoint. Because is metrizable it is normal, hence by Urysohn’s lemma there is a continuous function such that for all and for all . Then, by Theorem 9 there is a unique continuous such that . Then we have, because a subsequence of a convergent sequence has the same limit,
and likewise
a contradiction. This shows that , which completes the proof. ∎
6 𝐶*-algebras and the Gelfand transform
A -algebra is a complex Banach algebra with a map such that
-
1.
for all (namely, is an involution),
-
2.
and for all ,
-
3.
for all and ,
-
4.
for all .
We do not require that a -algebra be unital. If then . Otherwise,
gives and
gives , showing that is an isometry.
We now take to denote rather than , and likewise for , and . 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, is a Banach space with the supremum norm and is a closed linear subspace, whatever the topological space . It is then straightforward to check that with the involution they are commutative -algebras.
A homomorphism of -algebras is an algebra homomorphism , where and are -algebras, such that for all . It can be proved that .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 -algebras to be an algebra isomorphism such that for all . It follows that and because is bijective, the inverse is a -algebra homomorphism, giving and therefore . Thus, an isomorphism of -algebras is an isometric isomorphism.
Suppose that is a commutative -algebra, which we do not assume to be unital. A character of is a nonzero algebra homomorphism . We denote the set of characters of by , which we call the Gelfand spectrum of . 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 , , so is contained in the closed unit ball of , where denotes the dual of the Banach space . Furthermore, one can prove that is a weak-* closed set in , 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 the subspace topology inherited from with the weak-* topology. Depending on whether is or is not an isolated point in , is a compact or a locally compact Hausdorff space; in any case is a locally compact Hausdorff space.
The Gelfand transform is the map defined by ; that is continuous follows from having the weak-* topology, and one proves that in fact .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 is an isomorphism of -algebras.
It can be proved that two commutative -algebras are isomorphic as -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 -algebra is a closed linear subspace of such that and . An ideal is said to be essential if for every nonzero ideal of . In particular, is itself an essential ideal.
Suppose that is a -algebra. The multiplier algebra of , denoted , is a -algebra containing as an essential ideal such that if is a -algebra containing as an essential ideal then there is a unique homomorphism of -algebras whose restriction to is the identity. We have not shown that there is a multiplier algebra of , but we shall now prove that this definition is a universal property: that any -algebra satisfying the definition is isomorphic as a -algebra to , which allows us to talk about “the” multiplier algebra rather than “a” multiplier algebra.
Suppose that is a -algebra containing as an essential ideal such that if is a -algebra containing as an essential ideal then there is a unique -algebra homomorphism whose restriction to is the identity. Hence there is a unique homomorphism of -algebras whose restriction to is the identity, and there is a unique homomorphism of -algebras whose restriction to is the identity. Then and are homomorphisms of -algebras whose restrictions to are the identity. But the identity maps and are also homomorphisms of -algebras whose restrictions to are the identity. Therefore, by uniqueness we get that and . Therefore is an isomorphism of -algebras.
One can prove that if is unital then .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 -algebra , the multiplier algebra 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 , it can be proved that .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 is a locally compact Hausdorff space then is a dense linear subspace of , and for any topological space is a closed linear subspace of , but before talking about multiplier algebras we did not have a tight fit between the -algebras and .
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.