Skip to main content

Notes

Complete Metric Spaces

Definition. Let (X,d)(X,d) be a metric space. A sequence (xn)(x_n) of points of XX is said to be a Cauchy sequence in (X,d)(X,d) if it has the property that for any ε>0\varepsilon>0, there exists an integer NN such that

d(xn,xm)<εwhenever n,mN.d(x_n,x_m)<\varepsilon \quad \text{whenever } n,m \geqslant N.

Definition. The metric space (X,d)(X,d) is said to be complete if every Cauchy sequence in XX converges (in the space XX itself).

info
  • A closed subset AA of a complete metric space (X,d)(X,d) is necessarily complete in the restricted metric (inherited from the original space), as a Cauchy sequence in AA is also Cauchy in XX, so it converges in XX, and the limit must lie in AA because AA is closed (see Theorem 17.6 in Chapter 2).
  • If XX is complete under the metric dd, then XX is also complete under the standard bounded metric dˉ(x,y)=min{d(x,y),1}\bar{d}(x,y)=\min\{d(x,y),1\} with respect to dd, and conversely, because a sequence is Cauchy (resp. converges) under dd if and only if it is Cauchy (resp. converges) under dd (since dˉ(x,y)d(x,y)\bar{d}(x,y) \leqslant d(x,y) for any x,yXx,y \in X and the equality holds when d(x,y)1d(x,y) \leqslant 1, see also Exercise 43.3).

[Lemma 43.1] A metric space XX is complete if every Cauchy sequence in XX has a convergent subsequence.

info

In fact, it also follows from this theorem and Theorem 28.2 (see Chapter 3) that every compact metric space is complete. However, the converse does not hold: a complete metric space need not be compact (e.g. R\mathbb{R} in its standard topology).

Let (xn)(x_n) be a Cauchy sequence in (X,d)(X,d). It follows that given ε>0\varepsilon > 0, there exists a positive integer NN' such that d(xn,xm)<ε/2d(x_n, x_m)< \varepsilon/2 whenever n,mNn,m \geqslant N'.

Let (xnk)(x_{n_k}) be a convergent subsequence of (xn)(x_n) with the limit xx, then for any ε>0\varepsilon>0, there exists a positive integer KK such that d(xnk,x)<ε/2d(x_{n_k},x)<\varepsilon/2 whenever kKk \geqslant K.

Take N=max{N,nK}N= \max\{N',n_K\}. Using the fact that nkkn_k \geqslant k for any subindex kk, it follows that whenever n,nkNn,n_k \geqslant N, we have

d(xn,x)d(xn,xnk)+d(xnk,x)<ε/2+ε/2=ε.d(x_n,x) \leqslant d(x_n,x_{n_k})+d(x_{n_k},x) < \varepsilon/2+\varepsilon/2=\varepsilon.

[Theorem 43.2] The Euclidean space Rk\mathbb{R}^k is complete in the Euclidean metric dd or the square metric ρ\rho.

To show that the metric space (Rk,d)(\mathbb{R}^k,d) is complete, let (xn)(x_n) be a Cauchy sequence in (Rk,d)(\mathbb{R}^k,d), then the set {xn}\{x_n\} is a bounded subset of (Rk,d)(\mathbb{R}^k,d), because if we choose NN such that

d(xn,xm)1 for all n,mN,d(x_n,x_m) \leqslant 1 \text{ for all } n,m \geqslant N,

then the number

M=max{d(x1,0),,d(xN1,0),d(xN,0)+1}M =\max\{d(x_1,\mathbf{0}),\ldots,d(x_{N-1},\mathbf{0}),d(x_N,\mathbf{0})+1\}

is an upper bound for d(xn,0)d(x_n,\mathbf{0}), by triangle inequality (d(xn,0)d(xn,xm)+d(xm,0) d(x_n,\mathbf{0}) \leqslant d(x_n,x_m)+d(x_m,\mathbf{0})). This implies that the points of the sequence (xn)(x_n) all lie in the kk-ball {xRk:d(x,0)M+1}\{ x \in \mathbb{R}^k : d(x,\mathbf{0}) \leqslant M+1\}. Since this sphere is compact (see Theorem 27.3 in Chapter 3), the sequence (xn)(x_n) has a convergent subsequence, by Theorem 28.2 (in Chapter 3). Thus, (Rk,d)(\mathbb{R}^k,d) is complete.

Note that a sequence is Cauchy (resp. converges) relative to ρ\rho if and only if it is Cauchy (resp. converges) relative to dd, hence (Rk,ρ)(\mathbb{R}^k,\rho) is also complete.

[Lemma 43.3] Let XX be the product space X=XαX = \prod X_\alpha and xn\mathbf{x}_n be a sequence of points of XX, then xnx\mathbf{x}_n \to \mathbf{x} if and only if πα(xn)πα(x)\pi_\alpha(\mathbf{x}_n) \to \pi_\alpha(\mathbf{x}) for each α\alpha.

Since the projection mapping πα:XXα\pi_\alpha : X \to X_\alpha is continuous, it preserves convergent sequences (see Theorem 21.3 in Chapter 2); the 'only if' (    \implies direction) part of the lemma follows.

To prove the converse (    \impliedby direction), suppose that πα(xn)πα(x)\pi_\alpha(\mathbf{x}_n) \to \pi_\alpha(\mathbf{x}) for each αJ\alpha \in J. Let U=UαU = \prod U_\alpha be a basis element for XX that contains x\mathbf{x}. For each α\alpha such that UαU_\alpha does not equal the entire space XαX_\alpha (see Theorem 19.1 in Chapter 2), choose a positive integer NαN_\alpha so that πα(xn)Uα\pi_\alpha(\mathbf{x}_n) \in U_\alpha whenever nNαn \geqslant N_\alpha. Let NN be the largest of such numbers NαN_\alpha, then for all nNn \geqslant N, we have that xnU\mathbf{x}_n \in U.

[Theorem 43.4] There is a metric for the product space Rω\mathbb{R}^\omega relative to which Rω\mathbb{R}^\omega is complete.

note

Another way of phrasing this theorem is:

Rω\mathbb{R}^\omega is completely metrizable.

See Exercise 43.6.

Let dˉ(a,b)=min{ab,1}\bar{d}(a,b)=\min\{|a-b|,1\} be the standard bounded metric on R\mathbb{R}. Let DD be the metric on Rω\mathbb{R}^\omega defined by D(x,y)=sup{dˉ(xi,yi)/i}D(\mathbf{x},\mathbf{y})=\sup\{\bar{d}(x_i,y_i)/i\}, then DD induces the product topology on Rω\mathbb{R}^\omega (see Theorem 20.5 in Chapter 2).

We verify that Rω\mathbb{R}^\omega is complete under DD. Let xn\mathbf{x}_n be a Cauchy sequence in (Rω,D)(\mathbb{R}^\omega, D). Since by definition

dˉ(πi(x),πi(y))iD(x,y),\bar{d}(\pi_i(\mathbf{x}),\pi_i(\mathbf{y})) \leqslant i D(\mathbf{x},\mathbf{y}),

we see that for fixed ii the sequence πi(xn)\pi_i(\mathbf{x}_n) is a Cauchy sequence in R\mathbb{R} (consider NN such that D(xn,xm)<ε/iD(\mathbf{x}_n,\mathbf{x}_m)<\varepsilon/i whenever n,mNn,m \geqslant N, given ε>0\varepsilon>0), so it converges, say to aia_i (see Theorem 43.2 above). The sequence xn\mathbf{x}_n then converges to the point a=(a1,a2,)\mathbf{a}=(a_1,a_2,\ldots) in Rω\mathbb{R}^\omega.

Definition. Let (Y,d)(Y,d) be a metric space; let dˉ(a,b)=min{d(a,b),1}\bar{d}(a,b)=\min\{d(a,b),1\} be the standard bounded metric on YY derived from dd. If x=(xα)αJ\mathbf{x}=(x_\alpha)_{\alpha \in J} and y=(yα)αJ\mathbf{y}=(y_\alpha)_{\alpha \in J} are points of the Cartesian product YJY^J, let

ρˉ(x,y)=sup{dˉ(xα,yα):αJ}.\bar{\rho}(\mathbf{x},\mathbf{y})=\sup\{\bar{d}(x_\alpha,y_\alpha) : \alpha \in J\}.

One can check that ρˉ\bar{\rho} is a metric; it is called the uniform metric on YJY^J corresponding to the metric dd on YY.

note

This is the generalisation of the definition of uniform metric given in Chapter 2.

info

Since the elements of YJY^J are in fact functions from JJ to YY, one could also use functional notation for them, which will be used from hereon.

The functional notation for the uniform metric is as such: If f,g:JYf,g : J \to Y, then

ρˉ(f,g)=sup{dˉ(f(α),g(α)):αJ}.\bar{\rho}(f,g)=\sup\{\bar{d}(f(\alpha),g(\alpha)) : \alpha \in J\}.

[Theorem 43.5] If the space YY is complete in the metric dd, then the space YJY^J is complete in the uniform metric ρˉ\bar{\rho} corresponding to dd.

[Theorem 43.6] Let XX be a topological space and let (Y,d)(Y,d) be a metric space. The set C(X,Y)\mathcal{C}(X,Y) of continuous functions is closed in YXY^X under the uniform metric, and so is the set B(X,Y)\mathcal{B}(X,Y) of bounded functions. Therefore, if YY is complete, these spaces are complete in the uniform metric.

note

A function ff is said to be bounded if its image f(X)f(X) is a bounded subset of the metric space (Y,d)(Y,d).

To show the first part of this theorem, we first show that if a sequence of elements fnf_n of YXY^X converges to the element ff of YXY^X relative to the metric ρˉ\bar{\rho} on YXY^X, then it converges to ff uniformly in the sense defined in §21, relative to the metric dˉ\bar{d} on YY. Let ε>0\varepsilon>0 be arbitrary, then due to the assumed convergence, we obtain an integer NN such that

ρˉ(f,fn)<ε for all nN.\bar{\rho}(f,f_n)<\varepsilon \text{ for all } n \geqslant N.

It follows that for all xXx \in X and all nNn \geqslant N, we have

dˉ(fn(x),f(x))ρˉ(fn,f)<ε.\bar{d}(f_n(x),f(x)) \leqslant \bar{\rho}(f_n,f)<\varepsilon.

Thus, (fn)(f_n) converges uniformly to ff.

Now we show that C(X,Y)\mathcal{C}(X,Y) is closed in YXY^X relative to the metric ρˉ\bar{\rho}. Let ff be an element of YXY^X that is a limit point of C(X,Y)\mathcal{C}(X,Y), then there is a sequence (fn)(f_n) of elements of C(X,Y)\mathcal{C}(X,Y) converging to ff in the metric ρˉ\bar{\rho} (see Lemma 21.2 in Chapter 2). By the uniform limit theorem, ff is continuous, so that fC(X,Y)f \in \mathcal{C}(X,Y). This implies that the closure of C(X,Y)\mathcal{C}(X,Y) is itself, and thus C(X,Y)\mathcal{C}(X,Y) is closed in YXY^X.

Finally, we show that B(X,Y)\mathcal{B}(X,Y) is closed in YXY^X. If ff is a limit point of B(X,Y)\mathcal{B}(X,Y), then there is a sequence of elements fnf_n of B(X,Y)\mathcal{B}(X,Y) converging to ff. Choose a sufficiently large NN such that ρˉ(fN,f)<1/2\bar{\rho}(f_N,f)<1/2, then for xXx \in X, we have dˉ(fN(x),f(x))<1/2\bar{d}(f_N(x),f(x))<1/2, which implies that d(fN(x),f(x))<1/2d(f_N(x),f(x))<1/2. It follows that if MM is the diameter of the set fN(x)f_N(x), then f(X)f(X) has diameter at most M+1M+1. Thus, fB(X,Y)f \in \mathcal{B}(X,Y).

If YY is complete, we conclude that C(X,Y)\mathcal{C}(X,Y) and B(X,Y)\mathcal{B}(X,Y) are complete in the metric ρˉ\bar{\rho}.

Definition. If (Y,d)(Y,d) is a metric space, one can define another metric on the set B(X,Y)\mathcal{B}(X,Y) of bounded functions from XX to YY by the equation

ρ(f,g)=sup{d(f(x),g(x)):xX}.\rho(f,g)=\sup\{d(f(x),g(x)) : x \in X\}.

Since ff and gg are bounded, the set f(X)g(X)f(X) \cup g(X) is bounded (so that the said supremum exists) if both f(X)f(X) and g(X)g*(X) are bounded, it follows that ρ\rho is well-defined. The metric ρ\rho here is called the sup metric.

info
  • There is a simple relation between the sup metric and the uniform metric, which is that if f,gB(X,Y)f,g \in \mathcal{B}(X,Y), then ρˉ(f,g)=min{ρ(f,g),1}\bar{\rho}(f,g)=\min\{\rho(f,g),1\}.

    This is because if ρ(f,g)>1\rho(f,g)>1, then there exists x0Xx_0 \in X such that d(f(x0),g(x0))>1d(f(x_0),g(x_0))>1, so by definition, dˉ(f(x0),g(x0))=1    ρˉ(f,g)=1\bar{d}(f(x_0),g(x_0)) = 1 \implies \bar{\rho}(f,g)=1. On the other hand, if ρ(f,g)1\rho(f,g) \leqslant 1, then dˉ(f(x),g(x))=d(f(x),g(x))1\bar{d}(f(x),g(x)) = d(f(x),g(x)) \leqslant 1 for all xx, so that ρˉ(f,g)=ρ(f,g)\bar{\rho}(f,g)=\rho(f,g).

    Thus, on B(X,Y)\mathcal{B}(X,Y), the metric ρˉ\bar{\rho} is in fact the standard bounded metric derived from the uniform metric ρ\rho.

  • If XX is a compact space, then every continuous function f:XYf : X \to Y is bounded, so the sup metric is well-defined on C(X,Y)\mathcal{C}(X,Y).

  • If YY is complete under dd, then C(X,Y)\mathcal{C}(X,Y) is complete under the corresponding uniform metric ρˉ\bar{\rho} (see Theorem 43.5), so it is also complete under the sup metric ρ\rho (see the first 'Info' block of this section).

[Theorem 43.7] Let (X,d)(X,d) be a metric space. There is an isometric embedding of XX into a complete metric space.

note

An alternative proof can be found in Exercise 43.9.

Definition. Let XX be a metric space. If h:XYh : X \to Y is an isometric embedding of XX into a complete metric space YY, then the subspace h(X)\overline{h(X)} of YY. is a complete metric space. It is called the completion of XX.

note

The completion of XX is uniquely determined up to an isometry. See Exercise 43.10.

A Space-Filling Curve

[Theorem 44.1] (Peano space-filling curve) Let I=[0,1]I=[0,1]. There exists a continuous map f:II2f : I \to I^2 whose image fills up the entire square I2I^2.

Step 1. We construct the map ff as the (uniform) limit of a sequence of a continuous functions fnf_n, through a series of triangular paths.

Step 2. We define a sequence of functions fn:II2f_n : I \to I^2 through recursively defined triangular paths.

Step 3. We claim that the sequence of functions (fn)(f_n) defined in Step 2 is a Cauchy sequence under the sup metric on C(I,I2)\mathcal{C}(I,I^2), which we denote by ρ\rho.

Step 4. Since C(I,I2)\mathcal{C}(I,I^2) is complete, the sequence fnf_n converges to a continuous function f:II2f : I \to I^2. We prove that ff is surjective.

Compactness in Metric Spaces

Definition. A metric space (X,d)(X,d) is said to be totally bounded if for every ε>0\varepsilon>0, there is a finite covering of XX by ε\varepsilon-balls.

[Theorem 45.1] A metric space (X,d)(X,d) is compact if and only if it is complete and totally bounded.

Definition. Let (Y,d)(Y,d) be a metric space. Let F\mathcal{F} be a subset of the function space C(X,Y)\mathcal{C}(X,Y). If x0Xx_0 \in X, the set F\mathcal{F} of functions is said to be equicontinuous at x0x_0 if given ε>0\varepsilon > 0, there is a neighbourhood UU of x0x_0 such that for all xUx \in U and all fFf \in \mathcal{F}, d(f(x),f(x0))<εd(f(x),f(x_0))<\varepsilon.

If the set F\mathcal{F} is equicontinuous at each point in XX, it is said simply to be equicontinuous.

note

Equicontinuity is stronger than continuity, but is a separate concept to uniform continuity.

While continuity of the function ff at a point x0x_0 implies the existence of a neighbourhood UU of x0x_0 such that d(f(x),f(x0))<εd(f(x),f(x_0))<\varepsilon for all xUx \in U with a given ε>0\varepsilon>0, which may change depending on the choice of x0x_0 and ε\varepsilon, equicontinuity of F\mathcal{F} ensures the existence of a single neighbourhood that works for all functions ff in the collection F\mathcal{F}.

In other words, equicontinuity gives rise to the existence of such neighbourhood that does not depend on the choice of functions ff amongst a family of functions, which is not to be confused with uniform continuity, for which the existence of such neighbourhood does not depend on the choice of points in the domain XX.

warning

Note that equicontinuity depends on the specific metric dd rather than just on the topology of YY.

[Lemma 45.2] Let XX be a space and (Y,d)(Y,d) be a metric space. If the subset F\mathcal{F} of C(X,Y)\mathcal{C}(X,Y) is totally bounded under the uniform metric correspond to dd, then F\mathcal{F} is equicontinuous under dd.

[Lemma 45.3] Let XX be a space and (Y,d)(Y,d) a metric space; assume XX and YY are compact. If the subset F\mathcal{F} of C(X,Y)\mathcal{C}(X,Y) is equicontinuous under dd, then F\mathcal{F} is totally bounded under the uniform and sup metrics corresponding to dd.

Definition. If (Y,d)(Y,d) is a metric space, a subset F\mathcal{F} of C(X,Y)\mathcal{C}(X,Y) is said to be pointwise bounded under dd if for each xXx \in X (fixed), the subset

Fx={f(x):fF}\mathcal{F}_x = \{f(x) : f \in \mathcal{F}\}

of YY is bounded under dd.

[Theorem 45.4] (Ascoli's theorem, classical version). Let XX be a compact space; let (Rn,d)(\mathbb{R}^n,d) denote Euclidean space in either the square metric or the Euclidean metric; give C(X,Rn)\mathcal{C}(X,\mathbb{R}^n) the corresponding uniform topology. A subspace F\mathcal{F} of C(X,Rn)\mathcal{C}(X,\mathbb{R}^n) has compact closure if and only if F\mathcal{F} is equicontinuous and pointwise bounded under dd.

[Corollary 45.5] Let XX be compact and dd denote either the square metric or the Euclidean metric on Rn\mathbb{R}^n. Give C(X,Rn)\mathcal{C}(X,\mathbb{R}^n) the corresponding uniform topology. A subspace F\mathcal{F} of C(X,Rn)\mathcal{C}(X,\mathbb{R}^n) is compact if and only if it is closed, bounded under the sup metric ρ\rho, and equicontinuous under dd.

Suppose that F\mathcal{F} is compact, then it must be closed and bounded (see Theorem 27.3 in Chapter 3); since the closure of F\mathcal{F} here is also compact (see Theorem 26.3 in Chapter 3 and recall that every metric space is Hausdorff), Theorem 45.4 implies that it is also equicontinuous.

Conversely, if F\mathcal{F} is closed, it is equal to its closure G\mathcal{G}; if it is bounded under ρ\rho, it is also pointwise bounded under dd; it is also assumed that F\mathcal{F} is equicontinuous, so Theorem 45.4 implies that it is compact.

Pointwise and Compact Convergence

Definition. Given a point xx of the set XX and an open set UU of the space YY, let

S(x,U)={f:fYX and f(x)U}.S(x,U) = \{f : f \in Y^X \text{ and } f(x) \in U\}.

The sets S(x,U)S(x,U) are a subbasis for topology on YXY^X, which is called the topology of pointwise convergence (or the point-open topology).

info

The general basis element for the topology of pointwise convergence is a finite intersection of subbasis elements S(x,U)S(x,U), so a typical basis element about the function ff consists of all functions gg that are 'close' to ff at finitely many points.

In fact, the topology of pointwise convergence on YXY^X is just the product topology. The set S(α,U)S(\alpha,U) of all functions x:JY\mathbf{x} : J \to Y such that x(α)U\mathbf{x}(\alpha) \in U is just the subset πα1(U)\pi_\alpha^{-1}(U) of YJY^J, which is the standard subbasis element for the product topology.

[Theorem 46.1] A sequence fnf_n of functions converges to the function ff in the topology of pointwise convergence if and only if for each xx in XX, the sequence fn(x)f_n(x) of points of YY converges to the point f(x)f(x).

This result is just a rephrasing of Lemma 43.3 in function space notation.

info

A sequence of continuous functions that converges only in the topology of pointwise convergence need not have a continuous limit. See here for (counter)examples.

Is there a topology intermediate between uniform topology and topology of pointwise convergence that preserves continuity in limits? Yes, by imposing a restriction that the space XX is compactly generated, we can define the topology of compact convergence.

Definition. Let (Y,d)(Y,d) be a metric space and XX be a topological space. Given an element ff of YXY^X, a compact subspace CC of XX and a number ε>0\varepsilon>0, let BC(f,ε)B_C(f,\varepsilon) denote the set of all the elements gg of YXY^X for which

sup{d(f(x),g(x)):xC}<ε.\sup\{d(f(x),g(x)) : x \in C\}<\varepsilon.

The sets BC(f,ε)B_C(f,\varepsilon) form a basis for a topoogy on YXY^X, which is called the topology of compact convergence (or 'the topology of uniform convergence on compact sets').

info
  • To show that BC(f,ε)B_C(f,\varepsilon) satisfies the condition for a basis, the key step is to note that if gBC(f,ε)g \in B_C(f,\varepsilon), then if we take

    δ:=εsup{d(f(x),g(x)):xC},\delta := \varepsilon - \sup\{d(f(x),g(x)) : x \in C\},

    we then have BC(g,δ)BC(f,ε)B_C(g,\delta) \subseteq B_C(f,\varepsilon).

  • The topology of compact convergence is 'stronger' than the topology of pointwise convergence in the sense that the general basis element containing ff are functions that are 'close' to fff not just at finitely many points, but at all points of some compact set.

[Theorem 46.2] A sequence fn:XYf_n : X \to Y of functions converges to the function ff in the topology of compact convergence if and only if for each compact subspace CC of XX, the sequence fnCf_n|_{C} converges uniformly to fCf|_{C}.

Let fn:XYf_n : X \to Y be a sequence of functions in the function space YXY^X under the topology of compact convergence.

Suppose that for each compact subspace of CC of XX, the sequence fnCf_n|_{C} converges uniformly to fCf|_{C}, then given ε>0\varepsilon>0, there exists an integer NN, such that

d(fnC(x),fC(x))<ε whenever nN, for all xC.d(f_n|_{C}(x),f|_{C}(x))<\varepsilon \text{ whenever } n \geqslant N, \text{ for all } x \in C.

This implies that whenever nNn \geqslant N, the set {d(fn(x),f(x)):xC}\{d(f_n(x),f(x)) : x \in C\} is bounded above by ε\varepsilon, so we have that

sup{d(fn(x),f(x)):xC}<ε.\sup\{d(f_n(x),f(x)) : x \in C\}<\varepsilon.

Note that each neighbourhood of ff in YXY^X under the topology of compact convergence must contain some basis element of the form BC(f,ε)B_C(f,\varepsilon). Note as well that if we range over all such neighbourhoods, we will run through all ε>0\varepsilon>0; conversely, if we range over all ε>0\varepsilon>0, we will run through all the basis elements in the form of BC(f,ε)B_C(f,\varepsilon), which generates all neighbourhoods of ff.

Therefore, the above implies that for each such neighbourhood UU, there exists an integer NN such that fnUf_n \in U for all nNn \geqslant N. Therefore, the sequence fn:XYf_n : X \to Y of functions converges to the function ff in the topology of compact convergence.

The steps above are reversible, so the converse also holds.

Definition. A space XX is said to be compactly generated (or is called a k-space), if it satisfies the following condition:

A set AA is open (equivalently, closed) in XX if ACA \cap C is open (equiv. closed) in CC for each compact subspace CC of XX.

[Lemma 46.3] If XX is locally compact, or if XX satisfies the first countability axiom, then XX is compactly generated.

Locally compact

Suppose that XX is locally compact, then for every point xx in XX, there exists some compact subspace CC of XX that contains a neighbourhood of xx.

Let ACA \cap C be open in CC for every compact subspace CC of XX. We want to show that AA is open in XX, so that XX is compactly generated. Given xAx \in A, choose a neighbourhood UU of xx that lies in a compact subspace CC of XX. Note that by hypothesis, ACA \cap C is open in CC, and thus AUA \cap U is open in UU, implying that AUA \cap U is open in XX as well (see Lemma 16.2 in Chapter 2). It follows that AUA \cap U is a neighbourhood of xx contained in AA, so AA is open in XX (since the choice of xAx \in A here is arbitrary).

First-countable

Suppose that XX satisfies the first countability axiom. If BCB \cap C is closed in CC for every compact subspace CC of XX, we want to show that BB is closed in XX. Let xx be a point of B\overline{B}; we show that xBx \in B (so that B=BB = \overline{B} and thus BB is closed in XX). Since XX has a countable basis at xx, there exists a sequence (xn)(x_n) of points of BB converging to xx (see Theorem 30.1 in Chapter 4). The subspace

C={x}{xn:nZ+}C = \{x\} \cup \{x_n : n \in \mathbb{Z}_+\}

is compact (why?), so that BCB \cap C is by assumption closed in CC. Since BCB \cap C contains xnx_n for every nn, it contains xx as well. Therefore, xBx \in B, as desired.

[Lemma 46.4] If XX is compactly generated, then a function f:XYf : X \to Y is continuous if for each compact subspace CC of XX, the restricted function fCf|_{C} is continuous.

We show that if VV is an open subset of YY, then f1(V)f^{-1}(V) is also open in XX. For any subspace CC of XX, we have

f1(V)C=(fC)1(V).f^{-1}(V) \cap C = (f|_{C})^{-1}(V).

If CC is compact, then by the continuity of fCf|_{C}, the set (fC)1(V)(f|_{C})^{-1}(V) is open in CC. Since XX is compactly generated, it follows that f1(V)f^{-1}(V) is open in XX.

[Theorem 46.5] Let XX be a compactly generated space and (Y,d)(Y,d) be a metric space, then C(X,Y)\mathcal{C}(X,Y) is clsoed in YXY^X in the topology of compact convergence.

Let fYXf \in Y^X be a limit point of C(X,Y)\mathcal{C}(X,Y); we wish to show that fC(X,Y)f \in \mathcal{C}(X,Y), i.e. ff is continuous.

It suffices to show that fCf|_{C} is continuous for each compact subspace CC of XX (see Theorem 46.2 above). For each nn, consider the neighbourhood BC(f,1/n)B_C(f,1/n) of ff; it intersects C(X,Y)\mathcal{C}(X,Y) (see Theorem 20.1 in Chapter 2), so we can choose a function fnC(X,Y)f_n \in \mathcal{C}(X,Y) lying in this neighbourhood. The sequence of functions fnC:CYf_n|_{C} : C \to Y converges uniformly to the function fCf|_{C} (as the definition of BC(f,1/n)B_C(f,1/n) implies that the choice of NN does not depend on the points xXx \in X), so by the uniform limit theorem, fCf|_{C} is continuous.

[Corollary 46.6] Let XX be a compactly generated space and (Y,d)(Y,d) be a metric space. If a sequence of continuous functions fn:XYf_n : X \to Y converges to ff in the topology of compact convergence, then ff is continuous.

Theorem 46.5 implies that the closure of C(X,Y)\mathcal{C}(X,Y) is itself, so all limit points will also be in C(X,Y)\mathcal{C}(X,Y), and this implies the desired result.

[Theorem 46.7] (Relationship between uniform, compact convergence, and pointwise convergence) Let XX be a space and (Y,d)(Y,d) be a metric space. For the function space YXY^X, one has the following inclusion of topologies:

(uniform)(compact convergence)(pointwise convergence).(\text{uniform}) \supseteq (\text{compact convergence}) \supseteq (\text{pointwise convergence}).

If XX is compact, then (uniform) = (compact convergence), and if XX is discrete, then (compact convergence) = (pointwise convergence).

A proof of this is in Exercise 46.2.

Definition. Let XX and YY be topological spaces. If CC is a compact subspace of XX and UU is an open subset of YY, define

S(C,U)={f:fC(X,Y) and f(C)U}.S(C,U)=\{ f : f \in \mathcal{C}(X,Y) \text{ and } f(C) \subseteq U\}.

The sets S(C,U)S(C,U) form a subbasis for a topology on C(X,Y)\mathcal{C}(X,Y) that is called the compact-open topology.

info

From the definition, we see that the compact-open topology is finer than the pointwise convergence topology (since S(C,U)S(x,U)S(C,U) \subseteq S(x,U) for any open subset of YY and compact subspace of XX and note that C(X,Y)YX\mathcal{C}(X,Y) \subseteq Y^X).

The compact-open topology can actually be defined on the entire function space YXY^X, but it is not the focus of this section.

[Theorem 46.8] Let XX be a space and let (Y,d)(Y,d) be a metric space. On the set C(X,Y)\mathcal{C}(X,Y), the compact-open topology and the topology of compact convergence coincide.

[Corollary 46.9] (from Theorem 46.8) Let YY be a metric space. The compact convergence topology on C(X,Y)\mathcal{C}(X,Y) does not depend on the metric of YY. Therefore, if XX is compact, the uniform topology on C(X,Y)\mathcal{C}(X,Y) does not depend on the metric of YY.

The first part of this corollary comes from Theorem 46.8 and the definition of compact-open topology, which does not depend on the choice of said metric. The second part of this corollary comes from the first part and Theorem 46.7.

[Theorem 46.10] Let XX be a locally compact Hausdorff space and C(X,Y)\mathcal{C}(X,Y) have the compact-open topology, then the evaluation map, given by

e:X×C(X,Y)Ye : X \times \mathcal{C}(X,Y) \to Y

which is defined by the equation

e(x,f)=f(x)e(x,f)=f(x)

is continuous.

Let (x,f)(x,f) be a point of X×C(X,Y)X \times \mathcal{C}(X,Y) and VV be an open set in YY that contains the image point e(x,f)=f(x)e(x,f)=f(x). We want to find an open set UU containing (x,f)(x,f) such that e(U)Ve(U) \subseteq V.

First, due to the continuity of ff and the fact that XX is locally compact Hausdorff, we can choose an open set UU about xx having compact closure U\overline{U}, such that f(U)Vf(\overline{U}) \subseteq V (see Theorem 29.2 in Chapter 3). Next, consider the open set U×S(U,V)U \times S(\overline{U},V) in X×C(X,Y)X \times \mathcal{C}(X,Y). It is an open set containing (x,f)(x,f). If (x,f)(x',f') belongs to this set, then e(x,f)=f(x)e(x',f')=f'(x') belongs to VV, as desired.

Definition. Given a function f:X×ZYf : X \times Z \to Y, there is a corresponding function F:ZC(X,Y)F : Z \to \mathcal{C}(X,Y), defined by the equation

(F(z))(x)=f(x,z).(F(z))(x)=f(x,z).

Conversely, given F:ZC(X,Y)F : Z \to \mathcal{C}(X,Y), this equation defines a corresponding function f:X×ZYf : X \times Z \to Y. We say that FF is the map of ZZ into C(X,Y)\mathcal{C}(X,Y) that is induced by ff.

[Theorem 46.11] Let XX and YY be spaces; give C(X,Y)\mathcal{C}(X,Y) the compact-open topology. If f:X×ZYf : X \times Z \to Y is continuous, then so is the induced function F:ZC(X,Y)F : Z \to \mathcal{C}(X,Y). The converse holds if XX is locally compact Hausdorff.

Ascoli's Theorem (see Arzelà–Ascoli theorem)

[Theorem 47.1] (Ascoli's theorem, general version). Let XX be a space and (Y,d)(Y,d) be a metric space. Give C(X,Y)\mathcal{C}(X,Y) the topology of compact convergence; let F\mathcal{F} be a subset of C(X,Y)\mathcal{C}(X,Y).

(a) If F\mathcal{F} is equicontinuous under dd and the set

Fa={f(a):fF}\mathcal{F}_a = \{ f(a) : f \in \mathcal{F} \}

has compact closure for each aXa \in X, then F\mathcal{F} is contained in a compact subspace of C(X,Y)\mathcal{C}(X,Y).

(b) The converse holds if XX is locally compact Hausdorff.

Proof of (a). Throughout, we give YXY^X the product topology. Recall that this is the same as the topology of pointwise convergence, so YXY^X is a Hausdorff space. The space C(X,Y)\mathcal{C}(X,Y), which has the topology of compact convergence, is not a subspace of YXY^X. Let G\mathcal{G} be the closure of F\mathcal{F} in YXY^X.

Step 1. We show that G\mathcal{G} is a compact closure of YXY^X.

Step 2. We show that each function belonging to G\mathcal{G} is continuous, and that G\mathcal{G} itself is equicontinuous under dd.

Step 3. We show that the product topology on YXY^X and the compact convergence topology on C(X,Y)\mathcal{C}(X,Y) coincide on the subset G\mathcal{G}.

Step 4. The set G\mathcal{G} contains F\mathcal{F} and is contained in C(X,Y)\mathcal{C}(X,Y). We know from Step 1 that it is compact as a subspace of YXY^X in the product topology. By Step 3, it is also compact as a subspace of C(X,Y)\mathcal{C}(X,Y) in the compact convergence topology.

Proof of (b). Let H\mathcal{H} be a compact subspace of C(X,Y)\mathcal{C}(X,Y) that contains F\mathcal{F}. We show that H\mathcal{H} is equicontinuous and that for each aXa \in X, the set Ha\mathcal{H}_a is compact, so that F\mathcal{F} is equicontinuous (since FH)\mathcal{F} \subseteq \mathcal{H}) and that Fa\mathcal{F}_a lies in the compact subspace Ha\mathcal{H}_a of YY, implying that Fa\overline{\mathcal{F}_a} is compact.