Skip to main content

Notes

The Countability Axioms

In Chapter 2, it is discussed that every metrizable space satisfies the following first countability axiom:

Definition. A space XX is said to have a countable basis at xx if there is a countable collection B\mathcal{B} of neighbourhoods of xx such that each neighbourhood of xx contains at least one of the elements of B\mathcal{B}. A space that has a countable basis at each of its points is said to satisfy the first countability axiom, or to be first-countable.

It turns out that metrizable spaces are not the only family of spaces that satisfy the first countability axiom, and this is further discussed in this chapter.

Theorem 30.1 shows that it suffices to determine limit points of sets and check continuity of functions via convergent sequences if the space is first-countable.

[Theorem 30.1] Let XX be a topological space.

(a) Let AA be a subset of XX. If there is a sequence of points of AA converging to xx, then xAx \in \overline{A}; the converse holds if XX is first-countable.

(b) Let f:XYf : X \to Y. If ff is continuous, then for every convergent sequence xnxx_n \to x in XX, the sequence f(xn)f(x_n) converges to f(x)f(x). The converse holds if XX is first-countable.

The proof of Theorem 30.1 is basically the same as the proofs of Lemma 21.2 and Theorem 21.3 presented in Chapter 2. Even though metrizability (which is stronger than first countability) is assumed there, in the proof itself actually only first countability is applied instead of the full strength of metrizability.

There is also a more important countability axiom: the second countability axiom.

Definition. If a space XX has a countable basis for its topology, then XX is said to satisfy the second-countability axiom, or to be second-countable.

info

Note that the second countability axiom is stronger than the first countability axiom: the former implies the latter. In fact, it is so strong that not even every metric space satisfies it; examples include Rω\mathbb{R}^\omega in the uniform topology (see Chapter 2, Metric Topology).

Both countability axioms are preserved under the operations of taking subspaces or countable products.

[Theorem 30.2] A subspace of a first-countable space is first-countable, and a countable product of first-countable spaces is first-countable. A subspace of a second-countable space is second-countable, and a countable product of second-countable spaces is second-countable.

Definition. A subset AA of a space XX is said to be dense in XX if A=X\overline{A}=X.

note

An example of a dense subset of a space is the rational numbers Q\mathbb{Q} as a subset of the real numbers R\mathbb{R}, because every real number is the limit point of a sequence in Q\mathbb{Q}, applying Theorem 30.1.

[Theorem 30.3] Suppose that XX has a countable basis, then:

(a) Every open covering of XX contains a countable subcollection covering XX. (This property is also called the Lindelöf condition. A space having this property is called a Lindelöf space.)

(b) There exists a countable subset of XX that is dense in XX. (A space with this property is called separable.)

info
  • The two properties above are weaker in general than the second countability axiom, but are equivalent to it when the space is metrizable (see Exercise 30.5).
  • If a space is metrizable, it follows that XX is also second-countable.
note

To show that a space XX is Lindelöf, it in fact suffices to check that every open covering of XX by its basis elements contains a countable subcollection covering XX.

Why? Suppose that XX is a space such that every open covering of XX by its basis elements contains a countable subcollection covering XX. Note that an open set is a union

warning

Do not confuse separable spaces with a separation of a space which is used to define the connectedness of a space!

Separable spaces ≠ Disconnected spaces!

info
  • The product of two Lindelöf spaces need not be Lindelöf. For example, Rl\mathbb{R}_l is Lindelöf but the product space Rl×Rl\mathbb{R}_l \times \mathbb{R}_l is not.

    The space Rl2\mathbb{R}_l^2 is also called the Sorgenfrey plane.

  • A subspace of a Lindelöf space need not be Lindelöf. For example, the ordered square Io2I_o^2 is Lindelöf but the subspace A=I×(0,1)A = I \times (0,1) is not.

The Separation Axioms

Note that the Hausdorff axiom mentioned in Chapter 2 is also a separation axiom.

Definition. Suppose that one-point sets are closed in XX, then XX is said to be regular if for each pair consisting of a point xx and a closed set BB disjoint from xx, there exists disjoint open sets containing xx and BB, respectively. The space XX is said to be normal if for each pair A,BA,B of disjoint closed sets of XX, there exists disjoint open sets containing AA and BB, respectively.

info

Note that normality is stronger than regularity, and regularity is stronger than the Hausdorff condition.

The T1T_1 axiom is necessary in the definition of regularity of normality in order for this to be the case. A two-point space in the indiscrete topology satisfies the other part of the definitions of regularity and normality despite not being Hausdorff.

[Lemma 31.1] (Equivalent formulation of regularity and normality) Let XX be a topological space. Let one-point sets in XX be closed.

(a) XX is regular if and only if given a point xx of XX and a neighbourhood UU of xx, there is a neighbourhood VV of xx such that VU\overline{V} \subseteq U.

(b) XX is normal if and only if given a closed set AA and an open set UU containing AA, there is an open VV containing AA such that VU\overline{V} \subseteq U.

[Theorem 31.2] (a) A subspace of a Hausdorff space is Hausdorff; a product of Hausdorff spaces is Hausdorff.

(b) A subspace of a regular space is regular; a product of regular spaces is regular.

(a) Let XX be Hausdorff. Let xx and yy be two (distinct) points of the subspace YY of XX. If UU and VV are neighbourhoods in XX of xx and yy, respectively, such that UV=U \cap V = \emptyset, then UYU \cap Y and VYV \cap Y are also disjoint neighbourhoods of xx and yy respectively in YY.

Let {Xα}\{X_\alpha\} be a family of Hausdorff spaces. Let x=(xα)\mathbf{x}=(x_\alpha) and y=(yα)\mathbf{y}=(y_\alpha) be distinct points of the product space Xα\prod X_\alpha, so that there exists some index β\beta such that xβyβx_\beta \neq y_\beta.

Using the Hausdorff condition, take disjoint open sets UU and VV in XβX_\beta containing xβx_\beta and yβy_\beta, respectively, then the sets πβ1(U)\pi_\beta^{-1}(U) and πβ1(V)\pi_\beta^{-1}(V) are disjoint open sets in Xα\prod X_\alpha containing x\mathbf{x} and y\mathbf{y} (projections are continuous, and inverse images preserve intersections), respectively.

(b) Let YY be a subspace of the regular space XX. Take a point yy in YY (which is also in XX) and a closed subset BB of YY that does not contain xx.

Note that BY=B\overline{B} \cap Y = B, where B\overline{B} denotes the closure of BB in XX (see Theorem 17.4 in Chapter 2). This implies that xBx \notin \overline{B}, so, since XX is regular, we can obtain disjoint open sets UU and VV of XX containing xx and B\overline{B}, respectively. It follows that UYU \cap Y and VYV \cap Y are disjoint open sets in YY containing xx and BB, respectively.

Let {Xα}\{X_\alpha\} be a family of regular spaces and X=XαX = \prod X_\alpha. By (a), XX is Hausdorff, so that one-point sets are closed in XX.

Let x=(xα)\mathbf{x}=(x_\alpha) be a point of XX and UU be a neighbourhood of x\mathbf{x} in XX. Choose a basis element Uα\prod U_\alpha about x\mathbf{x} that lies in in UU. For each α\alpha, choose a neighbourhood VαV_\alpha of xαx_\alpha in XαX_\alpha such that VαUα\overline{V_\alpha} \in U_\alpha; if it happens that Uα=XαU_\alpha = X_\alpha, choose Vα=XαV_\alpha = X_\alpha. It follows that V=VαV = \prod V_\alpha is a neighbourhood of x\mathbf{x} in XX. By Theorem 19.5, we have that V=Vα\overline{V}=\prod \overline{V_\alpha}, which implies that VUαU\overline{V} \subseteq \prod U_\alpha \subseteq U, and thus by Lemma 31.1, XX is regular.

info
  • The space RK\mathbb{R}_K is Hausdorff but not regular.
  • The space Rl\mathbb{R}_l is normal.
  • The Sorgenfrey plane Rl2\mathbb{R}_l^2 is not normal.

Normal Spaces

Recall that a space XX (with one-point sets being closed) is said to be normal if for each pair A,BA,B of disjoint closed sets of XX, there exists disjoint open sets containing AA and BB, respectively.

[Theorem 32.1] Every regular space with a countable basis is normal.

note

Recall that XX is said to be regular if for each pair consisting of a point xx and a closed set BB disjoint from xx, there exists disjoint open sets containing xx and BB, respectively.

Let XX be a regular space with a countable basis B\mathcal{B}. Let AA and BB be disjoint closed subsets of XX.

Each point xx of AA has a neighbourhood UU not intersecting BB. Using Lemma 31.1, choose a neighbourhood VV of xx such that VU\overline{V} \subseteq U. Finally, choose an element of B\mathcal{B} that contains xx and is contained in VV. Doing this for each xx in AA, we obtain a countable covering of AA by open sets whose closures do not intersect BB, which we can then index with positive integers and denote it by {Un}\{U_n\}.

We then repeat the steps for BB and obtain a countable covering {Vn}\{V_n\} for BB such that the closure of each set in the covering does not intersect AA.

The sets U=UnU=\bigcup U_n and V=VnV=\bigcup V_n are open sets containing AA and BB respectively, but they need not be disjoint. However, we can perform a trick to construct two disjoint open sets UU' and VV' that contain AA and BB respectively from them.

Define U=nZ+UnU'=\bigcup_{n \in \mathbb{Z}_+}U_n' and V=nZ+VnV'=\bigcup_{n \in \mathbb{Z}_+}V_n', where Un=Un\i=1nViU_n'=U_n \backslash \bigcup_{i=1}^{n} \overline{V_i} and Vn=Vn\i=1nUiV_n'=V_n \backslash \bigcup_{i=1}^{n} \overline{U_i}.

Note that each set UnU_n', as well as VnV_n' is open since it is the difference of an open set and a closed set. We see that UU' contains AA, because each xx in AA belongs to UnU_n for some nn, and xx belongs to none of the sets Vi\overline{V_i}. Similarly, VV' contains BB.

Finally, UU' and VV' are disjoint. If not, then there exists some xUVx \in U' \cap V', which implies that xUjVkx \in U_j' \cap V_k' for some jj and kk. If jkj \leqslant k, then by definition of UjU_j', we have that xUjx \in U_j but from the definition of VkV_k', we also have that xUjx \notin U_j, causing a contradiction. The situation is similar for jkj \geqslant k.

[Theorem 32.2] Every metrizable space is normal.

Let XX be a metrizable space with metric dd. Let AA and BB be disjoint closed subsets of XX.

For each aAa \in A, choose εa\varepsilon_a such that the ball B(a,εa)B(a,\varepsilon_a) does not intersect BB. Similarly, for each bBb \in B, choose εb\varepsilon_b such that B(b,εb)B(b,\varepsilon_b) does not intersect AA.

Define

U=aAB(a,εa/2),V=bBB(b,εb/2).U=\bigcup_{a \in A}B(a,\varepsilon_a/2), \quad V=\bigcup_{b \in B}B(b,\varepsilon_b/2).

Note that UU and VV are open sets containing AA and BB, respectively; we claim that they are disjoint.

If UVU \cap V \neq \emptyset, then there exists zz such that

zB(a,εa/2)B(b,εb/2)z \in B(a,\varepsilon_a/2) \cap B(b,\varepsilon_b/2)

for some aAa \in A and some bBb \in B. By triangle inequality, we have that d(a,b)d(a,z)+d(z,b)<(εa+εb)/2d(a,b) \leqslant d(a,z)+d(z,b) < (\varepsilon_a+\varepsilon_b)/2. If εaεb\varepsilon_a \leqslant \varepsilon_b, then d(a,b)<εbd(a,b)<\varepsilon_b, so that the ball B(b,εb)B(b,\varepsilon_b) contains the point aa. If εbεa\varepsilon_b \leqslant \varepsilon_a, then d(a,b)<εad(a,b)<\varepsilon_a, so that the ball B(a,εa)B(a,\varepsilon_a) contains the point bb. Both of the situations are impossible.

The bolded statement above is the key that motivates the definitions of UU and VV.

[Theorem 32.3] Every compact Hausdorff space is normal.

Recall that Lemma 26.4 in Chapter 3 states that: if YY is a compact subspace of the Hausdorff space XX and x0x_0 is not in YY, then there exists disjoint open sets UU and VV of XX containing x0x_0 and YY, respectively.

Let XX be a compact Hausdorff space. We have already essentially proved that XX is regular due to Lemma 26.4.

Essentially the same argument as given in Lemma 26.4 can be applied to show that XX is normal: Given disjoint closed sets AA and BB in XX, thanks to the regularity of XX, we can choose, for each point aa of AA, disjoint open sets UaU_a and VaV_a containing aa and BB, respectively.

The collection {Ua}\{U_a\} covers AA. Since AA is compact, AA may be covered by finitely many sets Ua1,,UamU_{a_1},\ldots,U_{a_m}, then

U=i=1mUaiandV=i=1mVaiU=\bigcup_{i=1}^{m} U_{a_i} \quad \text{and} \quad V=\bigcap_{i=1}^{m}V_{a_i}

are disjoint open sets containing AA and BB, respectively.

note

Why not define V=i=1mVaiV=\bigcup_{i=1}^{m}V_{a_i} ?

This is because while each VaiV_{a_i} does not contain some aiAa_i \in A, it does not necessarily mean that it does not intersect other elements in AA as well.

[Theorem 32.4] Every well-ordered set XX is normal in the order topology.

note

In fact, the stronger result is true: every order topology is normal. See here for proofs.

Let XX be a well-ordered set. We assert that every interval of the form (x,y](x,y] is open in XX: If XX has a largest element and yy is that element, (x,y](x,y] is just a basis element about yy. If yy is not the largest element of XX, then (x,y](x,y] equals the open set (x,y)(x,y'), where yy' is the immediate successor of yy.

Now, let AA and BB be disjoint closed sets in XX; assume first that neither AA nor BB contains the smallest element a0a_0 of XX. For each aAa \in A, there exists a basis element about aa disjoint from BB; it contains some interval of the form (x,a](x,a], since aa is not the smallest element of XX.

For each aAa \in A, choose such an interval (xa,a](x_a,a] disjoint from BB. Similarly, for each bBb \in B, choose an interval (yb,b](y_b,b] disjoint from AA. The sets

U=aA(xa,a]andV=bB(yb,b]U=\bigcup_{a \in A}(x_a,a] \quad \text{and} \quad V=\bigcup_{b \in B}(y_b,b]

are open sets containing AA and BB, respectively; we claim that they are disjoint.

If there exists some zUVz \in U \cap V, then z(xa.a](yb,b]z \in (x_a.a] \cap (y_b,b] for some aAa \in A and bBb \in B. Assume that a<ba < b, then if ayba \leqslant y_b, the two intervals are disjoint, while if a>yba > y_b, we have a(yb,b]a \in (y_b,b], contrary to the fact that (yb,b](y_b,b] is disjoint from AA. A similar contradiction occurs if b<ab < a.

Finally, assume that AA and BB are disjoint closed sets in XX, and AA contains the smallest element a0a_0 of XX. The set {a0}\{a_0\} is both open and closed in XX. (This is because {a0}=[a0,a)\{a_0\} = [a_0, a'), where aa' is the immediate successor of a0a_0, whilst X\{a0}=(a0,+)X \backslash \{a_0\} = (a_0,+\infty) , both of which are open.) By the result of the preceding paragraph, there exists disjoint open sets UU and VV containing the closed sets A\{a0}A \backslash \{a_0\} and BB respectively, then U{a0}U \cup \{a_0\} and VV are disjoint open sets containing AA and BB respectively.

info
  1. If JJ is uncountable, the product space RJ\mathbb{R}^J is not normal.
  2. The product space SΩ×SΩS_\Omega \times \overline{S_\Omega} is not normal. (SΩS_\Omega denotes the minimal uncountable well-ordered set whereas SΩ\overline{S_\Omega} denotes the set SΩ{Ω}S_\Omega \cup \{\Omega\}, where Ω\Omega is the first uncountable ordinal.)

The Urysohn Lemma

[Theorem 33.1] (Urysohn lemma). Let XX be a normal space; let AA and BB be disjoint closed subsets of XX. Let [a,b][a,b] be a closed interval in the real line. There then exists a continuous map

f:X[a,b]f : X \to [a,b]

such that f(x)=af(x)=a for every xx in AA, and f(x)=bf(x)=b for every xx in BB.

note

Such continuous function is also called a Urysohn function.

In fact, the converse of Urysohn lemma also holds. See here for a proof.

We need only consider the case where the interval in question is the interval [0,1][0,1]; the general case follows from that one (through a continuous bijection from [0,1][0,1] to [a,b][a,b] defined by x(ba)x+ax \mapsto (b-a)x+a).

Step 1. Using normality, we construct a certain family UpU_p of open sets of XX, indexed by the rational numbers (or just any countable dense subset of [0,1][0,1] that contains the points 00 and 11), then use these sets to define the continuous function ff.

Let PP be the set of all rational numbers in the interval [0,1][0,1]. We shall define, for each pp in PP, an open set UpU_p of XX, such that whenever p<qp<q, we have UpUq\overline{U_p} \subseteq U_q. Thus, the sets UpU_p will be simply ordered by inclusion (\subseteq) in the same way their subscripts are ordered by the usual ordering in the real line, i.e.

p<q    UpUq.(*)p < q \implies \overline{U_p} \subseteq U_q. \tag{*}

Since PP is countable, we can use induction (principle of recursive definition) to define the sets UpU_p.

Arrange the elements of PP in an infinite sequence in some way; for simplicity, suppose that the numbers 11 and 00 are the first two elements of the sequence. Define the sets UpU_p as follows: [Base case] First, let U1=X\BU_1=X \backslash B. Next, since AA is a closed set contained in the open set U1U_1, by normality of XX (Theorem 31.1), we may choose an open set U0U_0 such that

AU0andU0U1.A \subseteq U_0 \quad \text{and} \quad \overline{U_0} \subseteq U_1.

In general, let PnP_n denote the set consisting of the first nn rational numbers in the sequence. Suppose that UpU_p is defined for all rational numbers pp belonging to the set PnP_n, satisfying the condition (*) [Induction hypothesis].

Let rr denote the next rational number in the sequence; we want to define UrU_r.

[Inductive case] Consider the set Pn+1=Pn{r}P_{n+1}=P_n \cup \{r\}. It is a finite subset of the interval [0,1][0,1], thus it has a simple ordering derived from the usual order relation << on the real line. In a finite simply ordered set, every element (except the smallest and the largest) has an immediate predecessor and an immediate successor. The number 00 is the smallest element, and 11 is the largest element, of the simply ordered set Pn+1P_{n+1}, and rr is neither 00 nor 11. This implies that rr has an immediate predecessor pp and an immediate successor qq in Pn+1P_{n+1} (so that p<r<qp < r < q). The sets UpU_p and UqU_q are already defined, and UpUq\overline{U_p} \subseteq U_q by induction hypothesis. By normality of XX (Theorem 31.1), we can find an open set UrU_r of XX such that

UpUrandUrUq.\overline{U_p} \subseteq U_r \quad \text{and} \quad \overline{U_r} \subseteq U_q.

We claim that (*) now holds for every pair of elements of Pn+1P_{n+1}. If both elements lie in PnP_n, (*) holds by induction hypothesis. If one of them is rr and the other is a point of PnP_n, denoted here by ss, then either sps \leqslant p, in which case UsUpUr\overline{U_s} \subseteq \overline{U_p} \subseteq U_r, or sqs \geqslant q, in which case UrUqUs\overline{U_r} \subseteq U_q \subseteq U_s. This proves the claim.

Hence, we manage to define UpU_p for all pPp \in P by induction.

Step 2. We now extend the definition of UpU_p from all rational numbers in the interval [0,1][0,1] to all rational numbers pp in R\mathbb{R} by defining

{Up=if p<0,Uq=Xif p>1.\begin{cases} U_p = \emptyset & \text{if } p<0, \\ U_q = X & \text{if } p>1. \end{cases}

Indeed, if r<0r<0, then r<qr<q for all q[0,1]q \in [0,1], whilst Ur=Uq\overline{U_r}=\emptyset \subseteq U_q. Similarly, if r>1r>1, then r>qr>q for all q[0,1]q \in [0,1]; meanwhile, UqX=Ur\overline{U_q} \subseteq X = U_r. Therefore, it is still true that for any pair of rational numbers pp and qq, we have p<q    UpUqp<q \implies \overline{U_p} \subseteq U_q .

Step 3. Given a point xx of XX, let us define Q(x)\mathbb{Q}(x) to be the set of those rational numbers pp such that the corresponding open sets UpU_p contain xx, i.e.

Q(x)={p:xUp}\mathbb{Q}(x)=\{p : x \in U_p\}

This set contains no number less than 00, since no xx is in UpU_p for p<0p<0, and it contains every number greater than 11, since every xx is in UpU_p for p>1p > 1. Therefore, Q(x)\mathbb{Q}(x) is bounded below, and its greatest lower bound (infimum) is a point of the interval [0,1][0,1]. Define

f(x)=infQ(x)={p:xUp}.f(x)=\inf\mathbb{Q}(x)=\{ p : x \in U_p\}.

Step 4. We show that ff is the desired continuous function.

If xAx \in A, then xUpx \in U_p for every p0p \geqslant 0, so that Q(x)\mathbb{Q}(x) equals the set of all nonnegative rationals, and f(x)=infQ(x)=0f(x)=\inf\mathbb{Q}(x)=0. Similarly, if xBx \in B, then xUpx \in U_p for no p1p \leqslant 1, so that Q(x)\mathbb{Q}(x) consists of all rational numbers greater than 11, and f(x)=1f(x)=1.

Now here comes the hard part: we wish to show that ff is continuous. For this purpose, we first prove the following facts:

  1. xUr    f(x)rx \in \overline{U_r} \implies f(x) \leqslant r.
  2. xUr    f(x)r.x \notin U_r \implies f(x) \geqslant r .

To prove 1., note that if xUrx \in \overline{U_r}, then xUsx \in U_s for every s<rs<r (by definition of UpU_p in Step 1). Therefore, Q(x)\mathbb{Q}(x) contains all rational numbers greater than rr, so that by definition we have f(x)=infQ(x)rf(x)=\inf\mathbb{Q}(x) \leqslant r.

To prove 2., note that if xUrx \notin U_r, then xUsx \notin U_s for any s<rs < r. Therefore, Q(x)\mathbb{Q}(x) contains no rational numbers less than rr, so that f(x)=infQ(x)rf(x)=\inf\mathbb{Q}(x) \geqslant r.

Now we prove continuity of ff. By Theorem 18.1 (see Chapter 2), it suffices to show that given a point x0x_0 of XX and an open interval (c,d)(c,d) in R\mathbb{R} containing the point f(x0)f(x_0), we can find a neighbourhood UU of x0x_0 such that f(U)(c,d)f(U) \subseteq (c,d).

Choose rational numbers pp and qq such that c<p<f(x0)<q<dc < p < f(x_0) < q < d. (This is possible since rational numbers are dense in R\mathbb{R}.) We assert that the open set U=Uq\UpU = U_q \backslash \overline{U_p} is the desired neighbourhood of x0x_0.

First, we note that x0Ux_0 \in U, since the fact that f(x0)<qf(x_0)<q implies by the contraposition of 2. that x0Uqx_0 \in U_q, whereas the fact that f(x0)>pf(x_0)>p implies by the contraposition of 1. that x0Upx_0 \notin \overline{U_p}.

Finally, we show that f(U)(c,d)f(U) \subseteq (c,d). Let xUx \in U, then xUqUqx \in U_q \subseteq \overline{U_q}, so that by 1., f(x)qf(x) \leqslant q. Meanwhile, xUpx \notin U_p so by 2., f(x)>pf(x)>p. Hence, f(x)[p,q](c,d)f(x) \in [p,q] \subseteq (c,d), as desired.

Definition. If AA and BB are two (not necessarily closed) subsets of the topological space XX, and if there is a continuous function f:X[0,1]f : X \to [0,1] such that f(A)={0}f(A)=\{0\} and f(B)={1}f(B)=\{1\}, we say that AA and BB can be separated by a continuous function.

info

Is there a version of Urysohn lemma for regular spaces?

Not quite. Regularity is not enough to separate a point from a closed set by a continuous function, as one runs into difficulty when trying to find an open set UpU_p such that U0Up\overline{U_0} \subseteq U_p and UpU1\overline{U_p} \subseteq U_1, following the proof of Urysohn lemma given above.

In fact, being able to separate a point from a closed set by a continuous function is a stronger condition, which gives rise to a new separation axiom, as defined below.

Definition. A space XX is completely regular if one-point sets are closed in XX and if for each point x0x_0 and each closed set AA not containing x0x_0, there is a continuous function f:X[0,1]f : X \to [0,1] such that f(x0)=1f(x_0)=1 and f(A)={0}f(A)=\{0\}.

info
  • A normal space is completely regular.
  • A completely regular space is regular (to see this, consider the sets f1([0,12))f^{-1}([0,\frac{1}{2})) and f1((12,1])f^{-1}((\frac{1}{2},1])).

Complete regularity behaves nicely with regards to subspaces and products:

[Theorem 33.2] A subspace of a completely regular space is completely regular. A product of completely regular spaces is completely regular.

Let XX be completely regular; let YY be a subspace of XX. Let x0x_0 be a point of YY, and let AA be a closed set of YY not containing x0x_0. Note that A=AYA = \overline{A} \cap Y (see Theorem 17.4 in Chapter 2 or here for why), where A\overline{A} is the closure of AA in XX. This implies that x0Ax_0 \notin \overline{A}. Since XX is completely regular, we can choose a continuous function f:X[0,1]f: X \to [0,1] such that f(x0)=1f(x_0)=1 and f(A)={0}f(\overline{A})=\{0\}. The restriction of ff to YY is the desired continuous function on YY.

Let X=XαX = \prod X_\alpha be a product of completely regular spaces. Let b=(bα)\mathbf{b}=(b_\alpha) be a point of XX and AA be a closed set of XX disjoint from b\mathbf{b}. Choose a basis element Uα\prod U_\alpha containing b\mathbf{b} that does not intersect AA, then Uα=XαU_\alpha = X_\alpha except for finitely many α\alpha, say, α=α1,,αn\alpha=\alpha_1,\ldots,\alpha_n (see Theorem 19.1 in Chapter 2). Given i=1,,ni=1,\ldots,n, choose a continuous function

fi:Xαi[0,1]f_i : X_{\alpha_i} \to [0,1]

such that fi(bαi)=1f_i(b_{\alpha_i})=1 and fi(X\Uαi)={0}f_i(X \backslash U_{\alpha_i})=\{0\}. Let ϕi(x)=fi(παi(x))\phi_i(\mathbf{x})=f_i(\pi_{\alpha_i}(\mathbf{x})), then ϕi\phi_i maps XX continuously into R\mathbb{R} and vanishes outside παi1(Uαi)\pi_{\alpha_i}^{-1} (U_{\alpha_i}). The product

f(x)=ϕ1(x)ϕ2(x)ϕn(x)f(\mathbf{x})=\phi_1(\mathbf{x}) \cdot \phi_2(\mathbf{x}) \cdot \cdots \cdot \phi_n(\mathbf{x})

is the desired continuous function on XX, because it equals 11 at b\mathbf{b} and vanishes outside Uα\prod U_\alpha.

The Urysohn Metrization Theorem

[Theorem 34.1] (Urysohn metrization theorem). Every regular space XX with a countable basis is metrizable.

We shall prove that XX is metrizable by embedding XX in a metrizable space YY, i.e. by showing XX is homeomorphic with a subspace of YY.

Two versions of the proof are presented in this book. The first version defines YY as the space Rω\mathbb{R}^\omega in the product topology, which is shown in Theorem 20.5 (see Chapter 2) to be metrizable. In the second version, the space YY is also Rω\mathbb{R}^\omega, but this time in the topology induced by the uniform metric ρˉ\bar{\rho}. In each case, it turns out that the construction actually embeds XX in the subspace [0,1]ω[0,1]^\omega of Rω\mathbb{R}^\omega.

Step 1. We prove the following:

There exists a countable collection of continuous functions fn:X[0,1]f_n : X \to [0,1] having the property that given any point x0x_0 of XX and any neighbourhood UU of x0x_0, there exists an index nn such that fn(x0)>0f_n(x_0)>0 and fn(x)=0f_n(x)=0 for points xx that are outside UU.

Due to Theorem 32.1 and the Urysohn lemma, given x0x_0 and UU, there exists such a function. However, if we choose one such function for each pair (x0,U)(x_0,U), the resulting collection will not in general be countable. Our task is to cut the collection down to size. Here is one way to proceed:

Let {Bn}\{B_n\} be a countable basis for XX. For each pair n,mn,m of indices for which BnBm\overline{B_n} \subseteq B_m, apply the Urysohn lemma to choose a continuous function gn,m:X[0,1]g_{n,m} : X \to [0,1] such that gn,m(Bn)={1}g_{n,m}(\overline{B_n})=\{1\} and gn,m(X\Bm)={0}g_{n,m}(X \backslash B_m)=\{0\}. The collection {gn,m}\{g_{n,m}\} then satisfies our requirement: Given x0x_0 and a neighbourhood UU of x0x_0, one can choose a basis element BmB_m containing x0x_0 that is contained in UU. Using regularity, one can then choose BnB_n so that x0Bnx_0 \in B_n and BnBm\overline{B_n} \subseteq B_m, then n,mn,m is a pair of indices for which the function gn,mg_{n,m} is defined, and it is positive at x0x_0 and vanishes outside UU. Since the collection {gn,m}\{g_{n,m}\} is indexed with a subset of Z+×Z+\mathbb{Z}_+ \times \mathbb{Z}_+, it is countable; hence it can be reindexed with the positive integers, giving us the desired collection {fn}\{f_n\}.

Step 2 (first version). Now that we have the functions fnf_n from Step 1, take Rω\mathbb{R}^\omega in the product topology and define a map F:XRωF : X \to \mathbb{R}^\omega as such:

F(x)=(f1(x),f2(x),).F(x)=(f_1(x),f_2(x),\ldots).

We claim that FF is an embedding. First, FF is continuous because Rω\mathbb{R}^\omega has the product topology and each fnf_n is continuous (see Theorem 19.6 in Chapter 2). Next, FF is injective because given xyx \neq y, we know that there is an index nn such that fn(x)>0f_n(x)>0 and fn(y)=0f_n(y)=0 (note that the fact in Step 1 applies for every point in XX); thus, F(x)F(y)F(x) \neq F(y). Finally, we prove that FF is a homeomorphism of XX onto its image, the subspace Z=F(X)Z=F(X) of Rω\mathbb{R}^\omega. We know that FF defines a continuous bijection of XX with ZZ, so we need only show that for each open set UU in XX, the set F(U)F(U) is open in ZZ (see the 'Homeomorphisms' section of Chapter 2 notes).

Let z0z_0 be a point of F(U)F(U). We shall find an open set WW of ZZ such that z0WF(U)z_0 \in W \subseteq F(U) (see Exercise 13.1). Let x0x_0 be the point of UU such that F(x0)=z0F(x_0)=z_0. Choose an index NN for which fN(x0)>0f_N(x_0)>0 and fN(X\U)={0}f_N(X \backslash U)=\{0\}. Take the open ray (0,+)(0,+\infty) in R\mathbb{R}, and let VV be the open set V=πN1((0,+))V=\pi_N^{-1}((0,+\infty)) of Rω\mathbb{R}^\omega. Let WW = VZV \cap Z. Note that WW is open in ZZ, by definition of subspace topology. We assert that z0WF(U)z_0 \in W \subseteq F(U). First, z0Wz_0 \in W because πN(z0)=πN(F(x0))=fN(x0)>0\pi_N(z_0)=\pi_N(F(x_0))=f_N(x_0)>0. Next, WF(U)W \subseteq F(U), because if zWz \in W, then z=F(x)z = F(x) for some xXx \in X, and πN(Z)(0,+)\pi_N(Z) \in (0,+\infty). Since πN(z)=πN(F(x))=fN(x)\pi_N(z)=\pi_N(F(x))=f_N(x), and fNf_N vanishes outside UU, the point xx must be in UU, then z=F(x)z = F(x) is in F(U)F(U), as desired. Thus, FF is an embedding of XX in Rω\mathbb{R}^\omega.

Step 3 (second version). In this version, we embed XX in the metric space (Rω,ρˉ)(\mathbb{R}^\omega,\bar{\rho}). Actually, we embed XX in the subspace [0,1]ω[0,1]^\omega, on which ρˉ\bar{\rho} equals the metric ρ(x,y)=sup{xiyi}\rho(\mathbf{x},\mathbf{y})=\sup\{|x_i-y_i|\} (since the interval [0,1][0,1] is bounded).

We use the countable collection of collections fn:X[0,1]f_n : X \to [0,1] constructed in Step 1, but now we impose the additional condition that fn(x)1nf_n(x) \leqslant \frac{1}{n} for all xx. (This condition can be satisfied by just dividing each function fnf_n by nn.)

Define F:X[0,1]ωF : X \to [0,1]^\omega by the equation F(x)=(f1(x),f2(x),)F(x)=(f_1(x),f_2(x),\ldots) as in Step 2. We assert that FF is now an embedding relative to the metric ρ\rho on [0,1]ω[0,1]^\omega.

We know from Step 2 that FF is injetive. Furthermore, we know that if we use the product topology on [0,1]ω[0,1]^\omega, the map FF carries open sets of XX onto open sets of the subspace Z=F(X)Z = F(X). This statement still holds if one passes to the finer topology (see Theorem 20.4) on [0,1]ω[0,1]^\omega induced by the metric ρ\rho.

It remains to prove that FF is continuous. This does not follow from the fact that each component function is continuous, as we are not using the product topology on Rω\mathbb{R}^\omega now. Here is where the assumption that fn(x)1nf_n(x) \leqslant \frac{1}{n} comes in.(

Let x0x_0 be a point of XX, and let ε>0\varepsilon>0 (be arbitrary). To prove continuity, we need to find a neighbourhood UU of x0x_0 such that xU    ρ(F(x),F(x0))<εx \in U \implies \rho(F(x),F(x_0))<\varepsilon.

First choose NN large enough so that 1/Nε/21/N \leqslant \varepsilon/2, then for each n=1,,Nn=1,\ldots,N use the continuity of fnf_n to choose a neighbourhood UnU_n of x0x_0 such that fn(x)fn(x0)ε/2|f_n(x)-f_n(x_0)| \leqslant \varepsilon/2 for xUnx \in U_n.

Let U=U1UNU = U_1 \cap \cdots \cap U_N; we show that UU is the desired neighbourhood of x0x_0. Let xUx \in U. If nNn \leqslant N, then fn(x)fn(x0)ε/2|f_n(x)-f_n(x_0)| \leqslant \varepsilon/2 by choice of UU, and if n>Nn > N, then fn(x)fn(x0)<1/Nε/2|f_n(x)-f_n(x_0)|<1/N \leqslant \varepsilon/2 because fnf_n maps XX into [0,1/n][0,1/n] (and 1/n<1/N1/n < 1/N). Therefore, for all xUx \in U, ρ(F(x),F(x0))ε/2<ε\rho(F(x),F(x_0)) \leqslant \varepsilon/2 < \varepsilon, as desired.

[Theorem 34.2] (Embedding theorem). Let XX be a space in which one-point sets are closed. Suppose that {fα}αJ\{f_\alpha\}_{\alpha \in J} is an indexed family of continuous functions fα:XRf_\alpha : X \to \mathbb{R} satisfying the requirement that for each point x0x_0 of XX and each neighbourhood UU of x0x_0, there is an index α\alpha such that fαf_\alpha is positive at x0x_0 and vanishes outside UU, then the function F:XRJF : X \to \mathbb{R}^J defined by

F(x)=(fα(x))αJF(x)=(f_\alpha(x))_{\alpha \in J}

is an embedding of XX in RJ\mathbb{R}^J. If fαf_\alpha maps XX into [0,1][0,1] for each α\alpha, then FF embeds XX in [0,1]J[0,1]^J.

note

One needs one-point sets in XX to be closed to ensure that given xyx \neq y, there is an index α\alpha such that fα(x)fα(y)f_\alpha(x) \neq f_\alpha(y), due to the fact given in Step 1 of the proof of Theorem 34.1 given above.

The proof is almost identical to Step 2 of the proof for Theorem 34.1 above, with nn replaced by α\alpha, and Rω\mathbb{R}^\omega by RJ\mathbb{R}^J, throughout.

info

A family of continuous functions that satisfies the hypotheses of Theorem 34.2 is said to separate points from closed sets in XX.

[Theorem 34.3] (Corollary from Theorem 34.2) A space XX is completely regular if and only if it is homeomorphic to a subspace of [0,1]J[0,1]^J for some JJ.

The Tietze Extension Theorem

[Theorem 35.1] (Tietze extension theorem). Let XX be a normal space; let AA be a closed subspace of XX.

(a) Any continuous map of AA into the closed interval [a,b][a,b] of R\mathbb{R} may be extended to a continuous map of all of XX into [a,b][a,b].

(b) Any continuous map of AA into R\mathbb{R} may be extended to a continuous map of all of XX into R\mathbb{R}.

The idea of the proof is to construct a sequence of continuous functions sns_n defined on the entire space XX, where the sequence sns_n converges uniformly, and the restriction of sns_n to AA approximates ff more and more closely as nn increases. The limit function (called gg) will then be continuous (see 'Uniform limit theorem' in Chapter 2), and its restriction to AA will be equal to ff.

Step 1. We first construct a particular function gg defined on all of XX such that gg is 'not too large', and such that gg approximates ff on the set AA 'fairly accurately'. More precisely, take a continuous map f:A[r,r]f : A \to [-r,r]. We assert that there exists a continuous function g:XRg : X \to \mathbb{R} such that

g(x)13rfor all xX,(1)|g(x)| \leqslant \frac{1}{3}r \quad \text{for all } x \in X, \tag{1}g(a)f(a)23rfor all aA.(2)|g(a)-f(a)| \leqslant \frac{2}{3}r \quad \text{for all } a \in A. \tag{2}

The claim can be proven by dividing the interval [r,r][r,r] into three equal closed intervals of length 23r\frac{2}{3}r, taking the inverse images of the first and third invervals, applying the Urysohn lemma (thanks to normality) to obtain a continuous function from XX to the second interval having the desired property (1), and considering three cases (whether aa is in the first or third interval, or neither) to prove property (2).

Step 2. We now prove part (a) of the theorem. Without loss of generality, we can replace the arbitrary closed interval [a,b][a,b] of R\mathbb{R} by the interval [1,1][-1,1].

Let f:X[1,1]f : X \to [-1,1] be a continuous map, then ff satisfies the hypotheses in Step 1, with r=1r = 1, so there exists a continuous function g1:XRg_1 : X \to \mathbb{R} such that

g1(x)13for all xX,|g_1(x)| \leqslant \frac{1}{3} \quad \text{for all } x \in X,f(a)g1(a)23for all aA.|f(a)-g_1(a)| \leqslant \frac{2}{3} \quad \text{for all } a \in A.

By induction (by considering fg1gkf-g_1-\cdots-g_k and applying Step 1 repeatedly), we can define a sequence of functions gng_n such that

gn(x)13(23)n1for all xX,(1’)|g_{n}(x)| \leqslant \frac{1}{3}\left(\frac{2}{3}\right)^{n-1} \quad \text{for all } x \in X, \tag{1'}f(a)g1(a)gn(a)(23)nfor all aA.(2’)|f(a)-g_1(a)-\cdots-g_n(a)| \leqslant \left(\frac{2}{3}\right)^{n} \quad \text{for all } a \in A. \tag{2'}

Next, we define g(x)=n=1gn(x)g(x)=\sum\limits_{n=1}^{\infty}g_n(x). By comparison test with the geometric series 13n=1(23)n1\frac{1}{3}\sum\limits_{n=1}^{\infty}\left(\frac{2}{3}\right)^{n-1} (and triangle inequality), we know that gg converges.

To show that gg is continuous, we must show that the partial sum sequence sns_n converges to gg uniformly. This follows from the 'Weierstrass MM-test' in analysis.

We also see that g(a)=f(a)g(a)=f(a) for all aAa \in A, as gg is the limit of the infinite sequence sn(x)s_n(x) of partial sums, and the property (2') implies that sn(a)f(a)s_n(a) \to f(a) for all aAa \in A.

Finally, we show that gg maps XX into the interval [1,1][-1,1]. Pretending that we do not see that (1/3)(2/3)n(1/3)\sum(2/3)^n converges to 11 and all we knew was that gg mapped XX into R\mathbb{R}, we can still compose gg with the 'normalisation map' to 'scale down' values with norms exceeding 11 to within 11 to obtain the desired outcome.

Step 3. We now prove part (b) of the theorem. Take a continuous function ff that maps AA into R\mathbb{R}. We can replace R\mathbb{R} by the open interval (1,1)(-1,1) because it is homeomorphic to R\mathbb{R}.

Part (a) of the theorem gives us a continuous map g:X[1,1]g : X \to [-1,1]. From here, we define a subset DD of XX like so: D=g1({1})g1({1})D = g^{-1}(\{-1\}) \cup g^{-1}(\{1\}). Note that DD is closed in XX due to the continuity of gg, and the set AA is disjoint from DD because g(A)=f(A)(1,1)g(A)=f(A) \subseteq (-1,1). By the Urysohn lemma, there is a continuous function ϕ:X[0,1]\phi : X \to [0,1] such that ϕ(D)={0}\phi(D)=\{0\} and ϕ(A)={1}\phi(A)=\{1\}.

Now, define h(x)=ϕ(x)g(x)h(x)=\phi(x)g(x). Note that hh is the product of two continuous functions, so hh is continuous. It is also an extension of ff, since ϕ(a)=1\phi(a)=1 for all aAa \in A. Finally, hh maps all of XX into the open interval (1,1)(-1,1), because ϕ(x)=0\phi(x)=0 if xDx \in D; if xDx \notin D, then g(x)<1    h(x)1g(x)<1 |g(x)| < 1 \implies |h(x)| \leqslant 1 \cdot |g(x)| < 1.

Embeddings of Manifolds

What conditions are required so that a space XX can be embedded in some finite-dimensional Euclidean space RN\mathbb{R}^N?

Definition. An mm-manifold is a Hausdorff space XX with a countable basis such that each point xx of XX has a neighbourhood that is homeomorphic with an open subset of Rm\mathbb{R}^m.

info
  • A 1-manifold is often called a curve, and a 2-manifold is called a surface.
  • If ϕ:XR\phi : X \to \mathbb{R}, then the (closed) support of ϕ\phi is defined to be the closure of the set ϕ1(R\{0})\phi^{-1}(\mathbb{R} \backslash \{0\}), that is, if xx lies outside the support of ϕ\phi, then there is some neighbourhood UU of xx on which ϕ(u)=0\phi(u)=0 for all uUu \in U (note that this is the negation of Theorem 17.5, see Chapter 2).

Definition. Let {U1,,Un}\{U_1,\ldots,U_n\} be a finite indexed open covering of the space XX. An indexed family of continuous functions

ϕi:X[0,1]for i=1,,n,\phi_i : X \to [0,1] \quad \text{for } i =1,\ldots,n,

is said to be a partition of unity dominated by {Ui}\{U_i\} if:

  1. supp(ϕi)Ui\mathrm{supp}(\phi_i) \subseteq U_i for each ii.
  2. i=1nϕi(x)=1\sum\limits_{i=1}^{n}\phi_i(x)=1 for each xx.

[Theorem 36.1] (Existence of finite partitions of unity). Let {U1,,Un}\{U_1,\ldots,U_n\} be a finite open covering of the normal space XX, then there exists a partition of unity dominated by {Ui}\{U_i\}.

Step 1. We first prove by induction that one can 'shrink' the covering {Ui}\{U_i\} to an open covering {V1,,Vn}\{V_1,\ldots,V_n\} of XX such that ViUi\overline{V_i} \subseteq U_i for each ii.

In general, given open sets V1,,Vk1V_1, \ldots, V_{k-1} such that the collection {V1,,Vk1,Uk,Uk+1,Un}\{V_1,\ldots,V_{k-1},U_k,U_{k+1},U_n\} covers XX, let AA be the relative complement of the union of such covering (excluding UkU_k) with respect to XX, then AA is a closed subset of XX which is contained in the open set UkU_k. Choose VkV_k to be an open set containing AA such that VkUk\overline{V_k} \subseteq U_k by normality, then {V1,,Vk1,Vk,Uk+1,,Un}\{V_1,\ldots,V_{k-1},V_k,U_{k+1},\ldots,U_n\} covers XX. Repeating until the nn-th step yields the required result.

Step 2. Given the finite open covering {Ui}\{U_i\} of XX, choose a finite open covering {Vi}\{V_i\} of XX such that ViUi\overline{V_i} \subseteq U_i for each ii. Again, choose an open covering {Wi}\{W_i\} of XX such that WiVi\overline{W_i} \subseteq V_i for each ii. Using the Urysohn lemma, for each ii, take a continuous function ψi:X[0,1]\psi_i : X \to [0,1] such that ψi(Wi)={1}\psi_i(\overline{W_i})=\{1\} and ψi(X\Vi)={0}\psi_i(X \backslash V_i)=\{0\}.

Since ψi1(R\{0})\psi_i^{-1}(R \backslash \{0\}) is contained in ViV_i, we have supp(ψi)ViUi\mathrm{supp}(\psi_i) \subseteq \overline{V_i} \subseteq U_i. Since the collection {Wi}\{W_i\} covers XX, the sum Ψ(x)=i=1nψi(x)\Psi(x)=\sum_{i=1}^{n}\psi_i(x) is positive for each xx (as ψi(x)=1\psi_i(x)=1 for some ii). Thus, we may define, for each jj,

ϕj(x)=ψj(x)Ψ(x).\phi_j(x)=\frac{\psi_j(x)}{\Psi(x)}.

We see that the functions ϕi,,ϕn\phi_i,\ldots,\phi_n form the desired partition of unity, as each ϕj\phi_j maps XX to [0,1][0,1] (as ψj\psi_j also maps XX to [0,1][0,1] and Ψ\Psi is positive for each xx), we have supp(ϕj)Uj\mathrm{supp}(\phi_j) \subseteq U_j for each jj (since supp(ϕj) \mathrm{supp}(\phi_j) is determined by supp(ψj) \mathrm{supp}(\psi_j)), and j=1nϕj(x)=1\sum_{j=1}^{n}\phi_j(x)=1 for each xx (as the numerator in the definition of ϕj\phi_j adds up to Ψ(x)\Psi(x) when we sum over the indices jj).

[Theorem 36.2] If XX is a compact mm-manifold, then XX can be embedded in RN\mathbb{R}^N for some positive integer NN.

Let {U1,,Un}\{U_1,\ldots, U_n\} be finitely many open sets that cover XX, each of which may be embedded in RM\mathbb{R}^M. Take embeddings gi:UiRmg_i : U_i \to \mathbb{R}^m for each ii. Since XX is compact and Hausdorff, it is also normal. Let ϕ1,,ϕn\phi_1,\ldots,\phi_n be a partition of unity dominated by {Ui}\{U_i\} and AiA_i be supp(ϕi)\mathrm{supp}(\phi_i). Define, for each i=1,,ni=1,\ldots,n, a function hi:XRmh_i : X \to \mathbb{R}^m as follows:

hi(x)={ϕ(x)gi(x)for xUi,0=(0,,0)for xX\Ai.h_i(x)=\begin{cases} \phi(x) \cdot g_i(x) & \text{for } x \in U_i, \\ \mathbf{0}=(0,\ldots,0) & \text{for } x \in X \backslash A_i. \end{cases}

The function hih_i is well-defined as the two definitions agree on the intersection of their domains. It is also continuous as its restrictions to the open sets UiU_i and X\AiX \backslash A_i are continuous (see point 6. of Theorem 18.2 in Chapter 2).

Now define F:X(R××Rn times×Rm××Rmn times)F : X \to (\underbrace{\mathbb{R} \times \cdots \times \mathbb{R}}_{n \text{ times}} \times \underbrace{\mathbb{R}^m \times \cdots \times \mathbb{R}^m}_{n \text{ times}}) as such:

F(x)=(ϕ1(x),,ϕn(x),h1(x),,hn(x)).F(x)=(\phi_1(x),\ldots,\phi_n(x),h_1(x),\ldots,h_n(x)).

Due to Theorem 18.4, FF is continuous. Since XX is compact, to prove that FF is an embedding, it suffices to show that FF is injective (see Theorem 26.6 in Chapter 3). Suppose that F(x)=F(y)F(x)=F(y), then ϕi(x)=ϕi(y)\phi_i(x)=\phi_i(y) and hi(x)=hi(y)h_i(x)=h_i(y) for all ii. Since ϕi(x)=1\sum\phi_i(x)=1, for some ii we have that ϕi(x)>0\phi_i(x)>0, so ϕi(y)>0\phi_i(y)>0 as well, and thus x,yUix, y \in U_i, then

ϕi(x)gi(x)=hi(x)=hi(y)=ϕi(y)gi(y).\phi_i(x) \cdot g_i(x)=h_i(x)=h_i(y)=\phi_i(y) \cdot g_i(y).

Since ϕi(x)=ϕi(y)>0\phi_i(x)=\phi_i(y)>0, it follows that gi(x)=gi(y)g_i(x)=g_i(y), and gig_i is injective, so x=yx=y, as desired.