MAS61015 Algebraic Topology

8. The Hausdorff property, and compactness

Video (Definition 8.1 to Proposition 8.8)

Let X be a metric space, and let a and b be points in X such that ab. By the first axiom of metric spaces, this means that d(a,b)>0. We can choose r with 0<rd(a,b)/2 and put U=OB(a,r) and V=OB(b,r). Then U and V are open, and aU and bV, and we have UV=. This is an obvious and natural construction that occurs frequently in the theory of metric spaces.

Now suppose that X is a general topological space, and we again have points a,bX with ab. We might again want to have open sets U and V with properties as above, but there is no longer any obvious way to produce them, and in fact, they need not exist in all cases. This leads us to introduce the following definition.

Definition 8.1.

Let X be a topological space. If a,bX with ab, then a Hausdorff separation for a and b is a pair of open sets U and V such that aU and bV and UV=. We say that X is a Hausdorff space if every pair of distinct points has a Hausdorff separation.

Example 8.2.

As in our previous discussion, if a and b are distinct points in a metric space X, then the open balls U=OB(a,d(a,b)/2) and V=OB(b,d(a,b)/2) give a Hausdorff separation of a and b. Thus, all metric spaces are Hausdorff spaces.

Example 8.3.

Let X be a space with the indiscrete topology, as in Example 3.15, so the only open sets are and X. Suppose also that |X|2, so we can choose two distinct points a,bX with ab. It is then clear that there can be no Hausdorff separation of a and b (because both sets U and V would have to be equal to X). Thus, X is not Hausdorff.

Example 8.4.

The Sierpiński space X is defined as follows: the underlying set is {0,1}, and the sets ,{1} and {0,1} are declared to be open, but the set {0} is not open. One can check that this is indeed a topology, and that there is no Hausdorff separation for 0 and 1, so we do not have a Hausdorff space.

Example 8.5.

Consider the line with doubled origin X/E as in Example 7.27. This contains two distinct points a=π(0,1) and b=π(0,-1). One can check that there is no Hausdorff separation for these points, so X/E is not a Hausdorff space. In particular, this illustrates the fact that a quotient of a Hausdorff space need not be Hausdorff.

Although non-Hausdorff spaces are important in various different branches of mathematics, we will mostly restrict attention to Hausdorff spaces.

Proposition 8.6.

Let X be a Hausdorff space, and let Y be a subset of X, with the subspace topology. Then Y is also Hausdorff.

Proof.

Let a and b be distinct points in Y. As X is Hausdorff, we can choose sets U,V that are open in X with aU and bV and UV=. Put U=UY and V=VY. By the definition of the subspace topology, these are open in Y. It is also clear that aU and bV and UV=UVY=, so U and V give a Hausdorff separation of a and b in Y.

Proposition 8.7.

Let Y and Z be disjoint topological spaces, and put X=YZ, and give X the coproduct topology as in Definition 7.1. Suppose that Y and Z are both Hausdorff; then X is also Hausdorff.

Proof.

Consider a pair of distinct points a,bX. There are four possible cases:

  • (1)

    Both a and b lie in Y

  • (2)

    Both a and b lie in Z

  • (3)

    a is in Y and b is in Z

  • (4)

    a is in Z and b is in Y.

In case (a), we use the fact that Y is Hausdorff. We can thus find sets U and V that are open in Y with aU and bV and UV=. From the definition of the coproduct topology we see that U and V are still open when considered as subsets of X, so they provide the required Hausdorff separation of a and b. Case (2) is essentially the same. In case (3) the pair (Y,Z) is the required Hausdorff separation, and in case (4) we use (Z,Y) instead.

Proposition 8.8.

Let Y and Z be topological spaces, and consider the space Y×Z with the product topology. Suppose that Y and Z are both Hausdorff; then Y×Z is also Hausdorff.

Proof.

Consider a pair of distinct points a=(y,z)Y×Z and b=(u,v)Y×Z. As ab, we must either have yu or zv. First suppose that yu in Y. As Y is Hausdorff, we can choose a Hausdorff separation (U,V) for y and u in Y. It is then not hard to see that the pair (U×Z,V×Z) is a Hausdorff separation for a and b in Y×Z. Similarly, if zv then we can choose a Hausdorff separation (P,Q) for z and v in Z, and we find that the pair (Y×P,Y×Q) is a Hausdorff separation for a and b in Y×Z.

We now turn to the notion of compactness. Compactness for metric spaces was covered in some detail in [MS, Section 7]. In particular, it was shown as [MS, Theorem 7.8] that a subset Xn is compact iff it is bounded and closed. For our purposes, you should think of this as a prototypical example: a space is compact if it is similar to a bounded, closed subset of n. It was also proved in [MS, Theorem 7.20] that a metric space is compact iff it has the Heine-Borel property. In the context of general topological spaces, we take the Heine-Borel property as the definition of compactness. We now recall the relevant details.

Video (Definition 8.9 to Example 8.16)

Definition 8.9.

Let X be a topological space. By an open cover of X we mean a family of open subsets (Ui)iI such that X=iIUi. Equivalently, for every point xX there must exist a set Ui in the family such that xUi. By a finite subcover we mean a subcollection Ui1,,Uir, containing only finitely many of the subsets Ui, such that we still have X=Ui1Uir. We say that X is compact if every open cover of X has a finite subcover.

Example 8.10.

Consider the space . The sets Un=(n-1,n+1) (for n) form an open cover of . However, we claim that there is no finite subcover. In other words, if we take any finite collection of these sets, say Un1,,Unr, we claim that the union U*=Un1Unr is not all of . This is clear: if p=min(n1,,nr) and q=max(n1,,nr) then U*(p-1,q+1) and so q+2U*, for example. From this it follows that is not compact. (We could also have deduced this from the results in [MS, Section 7].)

Example 8.11.

Now consider the space X=(0,1) and the subspaces Un=(1/n,1-1/n) (for n>1). These form an open cover of X but there is no finite subcover, so X is again not compact.

Definition 8.12.

Let X be a subset of n. We say that X is bounded if there is a constant R0 such that xR for all xX.

Lemma 8.13.

Let X be a space with only finitely many points; then X is compact.

Proof.

Supose that X={x1,,xn}. Let (Ui)iI be an open cover of X. For each t, the element xt must lie in some set of the cover, say xtUit. This means that XUi1Uin, so we have a finite subcover as required. ∎

Proposition 8.14.

Let X be a subset of n, with the subspace topology. Then X is compact iff it is bounded, and closed in n.

Proof.

This follows from [MS, Theorems 7.8 and 7.20]. (It is apparently identical to Theorem 7.8, but that is slightly misleading because [MS] uses a different definition of compactness; we need Theorem 7.20 as well to show that the two versions of compactness are the same.) ∎

Example 8.15.

The spaces Sn-1, Δn-1 and Bn are bounded and closed in n, so they are all compact.

Example 8.16.

Recall that On={AMn()|ATA=I}. We will show that this is closed and bounded in Mn()n2, so it is compact. We can define f:Mn()Mn() by f(A)=ATA-I. This is continuous, and {0} is closed in Mn(), and On=f-1{0}, so On is closed in Mn()n2. We have also explained that the standard metric on Mn() can be expressed in the form

d(A,B)=trace((A-B)T(A-B)).

For the identity matrix IMn() we have trace(I)=n. From this we see that d(A,0)=n for all AOn, so On is bounded.

Video (Proposition 8.17 to Corollary 8.19)

Proposition 8.17.

Let X be a compact space, and let Y be a closed subset of X, equipped with the subspace topology. Then Y is also compact.

Proof.

Let (Vi)iI be an open cover of Y. By the definition of the subspace topology, we can find sets Ui that are open in X such that Vi=UiY. As Y=iVi, we see that YiUi. Note that the set Yc=XY is also open in X (because Y is assumed to be closed). The set Yc together with the sets Ui then form an open cover of X. As X is compact, we can choose a finite subcover, which will consist of some finite collection of sets Ui1,,Uir, possibly together with Yc. These sets cover all of X, so in particular they cover Y. It is clear that Yc cannot contribute to covering Y, so we must have YUi1Uir. This means that Y=Vi1Vir, so we have a finite subcover of the original cover, as required. ∎

Proposition 8.18.

Let X be a Hausdorff space, and let Y be a subset that is compact with respect to the subspace topology. Then Y is closed in X.

Proof.

We must show that the set Yc=XY is open. Choose a point xYc. For each point yY we have xy, so we can choose a Hausdorff separation, say (Uy,Vy). This means that Uy and Vy are open in X and xUy and yVy and UyVy=. Now put Wy=VyY, which is open for the subspace topology on Y. As yWy, we see that the sets Wy cover Y. By the compactness property, we can choose a finite subcollection Wy1,,Wyr that still covers Y. This means that YVy1Vyr. Now put U#=Uy1Uyr. This is a finite intersection of open sets, so it is still open in X. Each of the sets Uyi contains x, so we have xU#. We also claim that U#Yc, or equivalently U#Y=. To see this, suppose (for a contradiction) that we have yU#Y. This means that yY, and YVy1Vyr, so we have yVyt for some t. On the other hand, we have yU#=Uy1Uir, so in particular yUyt. We now see that yUytVyt, which is impossible because UzVz= for all z. This contradiction shows that we in fact have U#Yc. We started with an arbitrary element xYc, and we produced an open set U# that contains x and is contained in Yc. By Lemma 3.33, this is enough to prove that Yc is open. ∎

Corollary 8.19.

Let X be a compact Hausdorff space, and let Y be a subset of X. Then Y is closed in X iff it is compact (in the subspace topology).

Proof.

This follows easily by combining Propositions 8.17 and 8.18. ∎

Video (Proposition 8.20 to Proposition 8.23)

Proposition 8.20.

Let X and Y be topological spaces. Suppose that X is compact, and that we have a surjective, continuous map f:XY. Then Y is also compact.

Proof.

Consider an open cover (Vi)iI of Y. Put Ui=f-1(Vi)X, so that Ui is open by the definition of continuity. If x is any point in X, then f(x)Y, and the sets Vi cover Y, so we can choose i with f(x)Vi, which means that xUi. This shows that the sets Ui form an open cover of X. As X is assumed to be compact, there must be a finite subcover, say Ui1,,Uir. We claim that the corresponding sets Vi1,,Vir cover Y. To see this, consider an arbitrary point yY. As f is assumed to be surjective, we can choose xX with f(x)=y. As the sets Ui1,,Uir cover X, we can choose t such that xUit. Here Uit was defined to be f-1(Vit), so we must have f(x)Vit, or in other words yVit as required. Thus, the list Vi1,,Vir is a finite subcover of our original cover. ∎

Corollary 8.21.

Let X be a compact space, and let E be an equivalence relation on X. Then the quotient space X/E is also compact.

Proof.

The quotient map π:XX/E is surjective and continuous, so this follows from Proposition 8.20. ∎

Example 8.22.

We can define an equivalence relation E on Sn by xEy iff y=±x. The real projective space Pn is defined to be the quotient space Sn/E. This is compact, because Sn is compact. Similarly, we can regard S2n+1 as the unit sphere in the space n+1=2n+2. We can then define another equivalence relation F on S2n+1 by declaring that xFy iff y=eiθx for some θ. The complex projective space Pn is defined to be S2n+1/F; this is again compact.

Proposition 8.23.

Let Y and Z be disjoint spaces, and take X=YZ, with the coproduct topology. If Y and Z are both compact, then so is X.

Proof.

Let (Ui)iI be an open cover of X. Then each set Ui must have the form ViWi, where Vi is an open subset of Y, and Wi is an open subset of Z. It is easy to see that the sets Vi form an open cover of Y, so there must be a finite subcover, say Yi1,,Yir. Similarly, the sets Wi form an open cover of Z, so there must be a finite subcover, say Wj1,,Wjs. We then find that the list Ui1,,Uir,Uj1,,Ujs forms a finite subcover of our original cover. ∎

Video (Theorem 8.24 to Lemma 8.26)

Theorem 8.24.

Let Y and Z be compact spaces. Then the product Y×Z is also compact (under the product topology).

We will give the proof after some preliminary definitions and results.

Definition 8.25.

Let Y and Z be as above, and let (Ui)iI be an open cover of Y×Z. We will say that a subset KY×Z is finitely covered if there is a finite subcollection Ui1,,Uir such that KUi1Uir.

We need to prove that the whole space Y×Z is finitely covered, but we will build up to this by proving that certain other sets are finitely covered first.

Lemma 8.26.

In the above context, let y be a point in Y. Then there is an open set SY such that yS and S×Z is finitely covered.

Proof.

For each zZ, choose an index iz such that (y,z)Uiz. As Uiz is open in the product topology, we can then choose a box Bz=Vz×Wz such that yVz and zWz and Vz×WzUiz. As zWz for all z, we see that the sets (Wz)zZ form an open cover of Z. As Z is compact, we can choose a finite subcover, say Z=Wz1Wzr. Now put S=Vz1Vzr. This is the intersection of a finite list of open sets, each of which contains y; so S itself is an open set containing y. We next claim that S×ZBz1Bzr. Indeed, suppose we have a point (s,z)S×Z. As Z=Wz1Wzr, we can choose k such that zWzk. Now sS=Vz1Vzr, so we also have sVzk. This proves that (s,z)Vzk×Wzk=Bzk, as required. We also have BzkUizk, so

S×ZUiz1Uizr.

This shows that S×Z is finitely covered. ∎

Proof of Theorem 8.24.

For each yY, the lemma tells us that we can choose an open set Sy containing y such that Sy×Z is finitely covered. The family (Sy)yY is then an open cover of Y. As Y is assumed to be compact we can choose a finite subcover, say Y=Sy1Syr. It follows that Y×Z=i=1r(Syi×Z). Here each of the subsets Syi×Z is finitely covered, and there are only finitely many of them, so the union is also finitely covered. This means that Y×Z is finitely covered, or in other words that our original cover of Y×Z has a finite subcover, as required. ∎

Video (Proposition 8.27 and Example 8.28)

Proposition 8.27.

Let X be a compact space, and let Y be a Hausdorff space. Let f:XY be a continuous bijection. Then the inverse map f-1:YX is also continuous, so f is a homeomorphism.

Proof.

Let g:YX be the inverse of f, so the claim is that g is continuous. By Proposition 3.23, it will be enough to check that for every closed subset FX, the preimage g-1(F) is closed in Y. Here F is a closed subset of a compact space, so it is compact by Proposition 8.17. As g is inverse to f, we have g-1(F)=f(F). We can regard f as a continuous surjective map from F to f(F) (where F and f(F) are given the respective subspace topologies). It follows by Proposition 8.20 that f(F) is also compact. Now f(F) is a compact subspace of the Hausdorff space Y, so it is closed by Proposition 8.18. In other words, the preimage g-1(F) is closed, as required. ∎

Example 8.28.

In Example 7.24 we considered a space X consisting of two disjoint discs, and the space Y=S2. We introduced an equivalence relation E on X (corresponding to the idea of gluing the boundary circles of the two discs) and defined a continuous bijection f¯:X/EY. Now X/E is a quotient of a closed bounded subset of 3, so it is compact, and Y is a metric space, so it is Hausdorff. It follows from Proposition 8.27 that f¯ is actually a homeomorphism.

Example 8.29.

In Example 7.26 we constructed a continuous bijection f¯:2/ET, where T is the torus and the equivalence relation E is given by xEy iff x-y2. Here T is a metric space and therefore Hausdorff. It is easy to see that the composite

[0,1]2inc2𝜋2/E

is surjective, and [0,1]2 is compact, so 2/E is compact by Proposition 8.20.

We conclude with some slightly different results about open covers, that are only relevant for metric spaces. Their use will become apparent later.

Video (Definition 8.30 and Proposition 8.31)

Definition 8.30.

Let X be a metric space, and let (Ui)iI be an open cover. A Lebesgue number for this cover is a number ϵ>0 with the following property: for every point xX, there is an index i such that OB(x,ϵ)Ui.

You should think of a Lebesgue number as something like the minimum size of overlaps between adjacent sets in the cover.

Proposition 8.31.

Let X be a compact metric space. Then every open cover of X has a Lebesgue number.

Proof.

Let (Ui)iI be an open cover of X. For each point xX, we can choose an index ix such that xUix. As Uix is open, we can then choose rx>0 such that OB(x,rx)Uix. Put Vx=OB(x,rx/2), so Vx is an open subset of X containing x. The sets Vx then form an open cover of X, so we can choose a finite subcover, say X=Vx1Vxn. Put ϵ=min(rx1,,rxn)/2. We claim that this is a Lebesgue number. To see this, let x be an arbitrary point in X. As the sets Vx1,,Vxn form a cover, we can choose k such that xVxk, or in other words d(x,xk)<rxk/2. If uOB(x,ϵ) then d(u,x)<ϵrxk/2 and d(x,xk)<rxk/2 so d(u,xk)<rxk. This proves that OB(x,ϵ)OB(xk,rrk)Uixk, so OB(x,ϵ) is contained in one of the sets Ui, as required. ∎