<<

1

Lecture 17: October 27 Complete spaces. Our topic this week is the C(X, Y ) of continuous functions between two topological spaces X and Y . Understanding the space of all continuous functions can be very useful, for example to find interesting examples of continuous functions. This topic actually belongs more to analysis than to , but since it involves several results from as well, it seems like a good way to finish our discussion of general topology. Before we get to the space of continuous functions, however, we have to return for a moment to the subject of metric spaces and introduce the important concept of completeness. Let (X, d) be a . We saw earlier that the metric topology on X is first countable – the open balls with radius in Q form a neighborhood at every point – and that many properties of can therefore be detected by looking at : a A X is closed iffit is sequentially closed; a subset A X is compact iffit is sequentially⊆ compact; etc. ⊆ In analysis, an important problem is to decide whether a given in a metric space converges. Typically, one does not know what the is – more often than not, the whole point of trying to show that the sequence converges is so that one can be sure that there is a limit. The notion of a “”, which you have probably already seen in your analysis course, was introduced to deal with this problem: proving that a sequence converges without knowing what the limit is.

Definition 17.1. A sequence of points xn X is called a Cauchy sequence if, for every ε>0, one can find an N such∈ that d(x ,x ) <ε for all m, n N. n m ≥ Intuitively, a Cauchy sequence is one where the points xn huddle closer and closer together as n gets large. Of course, not every Cauchy sequence actually converges: in the metric space Q, any sequence that converges to an irrational in R is a Cauchy sequence without limit in Q. Definition 17.2. A metric space (X, d) is called complete if every Cauchy sequence in X has a limit in X.

So R is complete, but Q is not; this is the reason why analysts prefer to work with real instead of rational numbers. There is a general construction for “completing” a metric space, similar to the way is which R can be obtained from Q. The result is that, given an arbitrary metric space (X, d), there is a complete metric space (X,ˆ dˆ) that contains X has a dense subset. Roughly speaking, the idea is to define an on the of Cauchy sequences in X:two Cauchy sequences x and y are equivalent if d(x ,y ) 0 as n . The points n n n n → →∞ of Xˆ are the equivalence classes of Cauchy sequences; there is a natural metric dˆ on Xˆ, and after some checking, one finds that (X,ˆ dˆ) is complete. If you have not seen this construction in a course on analysis, have a look at Theorem 43.7 or at Exercise 43.9 in Munkres’ book. Example 17.3. Completeness of a metric space is a property of the metric, not of the topology: R and (0, 1) are homeomorphic as topological spaces, but whereas R is complete, (0, 1) is not. 2

Example 17.4. Every compact metric space is complete. To see why, suppose that x X is a Cauchy sequence. Being a metric space, X is also sequentially compact, n ∈ and so the sequence has a convergent subsequence xn(k) with

x =limxn(k). k →∞ Now that we have a potential limit, we can easily show that the entire sequence xn converges to x.Fixε>0; then since xn is a Cauchy sequence, we can find N such that d(x ,x ) < ε for every n, m N. Using the triangle inequality, we get n m 2 ≥ ε ε d(x ,x) d(x ,x )+d(x ,x) < + = ε n ≤ n n(k) n(k) 2 2 for n N, by choosing k large enough so that n(k) N and d(x ,x) < ε . ≥ ≥ n(k) 2 The converse of this last result is not true: in the Euclidean metric, R is com- plete but not compact. Let us try to figure out what extraadditional condition (besides completeness) is needed to ensure that a metric space is compact. From the definition of compactness in terms of open coverings, we already know that every compact metric space has to be bounded; in fact, the following stronger form of boundedness holds. Definition 17.5. A metric space (X, d) is called totally bounded if, for every r>0, it is possible to cover X by finitely many open balls of radius r. Example 17.6. Let X be an infinite set, with the metric d(x, y)=1ifx = y, and d(x, y)=0ifx = y.ThenX is bounded, but not totally bounded. ￿ Theorem 17.7. A metric space is compact if and only if it is complete and totally bounded. Proof. We have already seen that every compact metric space is complete and totally bounded. Let us prove that the converse holds. Suppose that (X, d)isa complete metric space that is totally bounded. In a metric space, compactness is equivalent to sequential compactness, and so it suffices to show that every sequence in X has a convergent subsequence. Since X is complete, it suffices moreover to find a subsequence that is Cauchy. Let x X be a sequence of points. By assumption, we can cover X by finitely n ∈ many open balls of radius 1; evidently, at least one of these balls must contain xn for infinitely many n N. Call this ball U1, and let ∈ J1 = n N xn U1 , ∈ ∈ which is an infinite set. Similarly,￿ we can￿ cover X ￿by finitely many open balls of 1 ￿ radius 2 ;sincethesetJ1 is infinite, at least one of these balls must contain xn for infinitely many n J . Call this ball U , and let ∈ 1 2 J = n J x U , 2 ∈ 1 n ∈ 2 which is again infinite. Continuing￿ in this￿ way, we obtain￿ a nested sequence J1 ￿ ⊇ J2 J3 of infinite subsets of N, such that xn U1 Uk whenever n Jk. ⊇ ⊇··· ∈ ∩···∩ ∈ Now we can choose a suitable subsequence xn(k) of our original sequence. Pick any element n(1) J1; then pick an element n(2) J2 with n(2) >n(1), which exists because J is∈ infinite; then pick an element n(3)∈ J with n(3) >n(2); and 2 ∈ 3 so on. In this way, we obtain a subsequence xn(k) of the original sequence, with the property that x U U n(k) ∈ 1 ∩···∩ k 3

for every k =1, 2,.... This sequence is obviously Cauchy: whenever k ￿, both 1 ≤ xn(k) and xn(￿) belong to the open ball Uk of radius k , and so 2 d(x ,x ) . n(k) n(￿) ≤ k This subsequence converges because X is complete; the conclusion is that X is sequentially compact, and therefore compact. ￿ Function spaces. Given two nonempty sets X and Y , we can consider the space Fun(X, Y ) of all functions f : X Y . Since a function is uniquely determined by its values f(x) for x X, it is clear→ that ∈ Fun(X, Y )=Y X = Y x X ￿∈ is simply the Cartesian product, indexed by X, of several copies of Y : a function f : X Y corresponds to the element f(x) of the product. → x X Now suppose that X and Y are topological spaces;∈ we are interested in the subset ￿ ￿ C(X, Y )= f : X Y f is continous Fun(X, Y ) → ⊆ of all continuous functions.￿ There are￿ several ways￿ of making Fun(X, Y ) and C(X, Y ) into topological spaces; each￿ is useful in certain situations. The simplest way is to use the on Y X . The product topology is given by a basis; let us see what the usual basic open sets look like in terms of functions. Take finitely many points x1,...,xn X, and finitely many open sets U1,...,Un Y ; the corresponding basic open set∈ is ⊆ B(x ,...,x ,U ,...,U )= f : X Y f(x ) U for every i =1,...,n . 1 n 1 n → i ⊆ i Example 17.8. When does a sequence￿ of functions￿ fn : X Y converge to another￿ ￿ function f : X Y in this topology? For every x X →and every neighborhood U of f(x), the→ basic open set B(x, U) is a neighborhood∈ of f, and therefore has to contain all but finitely many of the f . In other words, f (x) U for all but n n ∈ finitely many n, which means that the sequence fn(x) Y converges to f(x) Y . So convergence in this topology is the same as pointwise∈ convergence. ∈ Definition 17.9. The product topology on Fun(X, Y )=Y X is called the topology of pointwise convergence. The disadvantage of this topology is that the subspace C(X, Y ) is not closed, because the pointwise limit of continuous functions may fail to be continuous. Example 17.10. Let X =[0, 1] and Y = R, and consider the sequence of functions n n fn :[0, 1] R, fn(x)=x .Sincex 0 for x<1, the sequence converges pointwise→ to the function → 0ifx<1, f :[0, 1] R,f(x)= → 1ifx = 1, which is no longer continuous. Another choice of topology on Fun(X, Y ) is related to the notion of uniform ® convergence. Here we assume that (Y,d) is a metric space. Given two functions f,g Fun(X, Y ), we define their uniform distance to be ∈ ρ(f,g)=supd f(x),g(x) x X ∈ ￿ ￿ 4

if the right-hand side is 1, and ρ(f,g) = 1 otherwise. (This slightly odd definition ≤ is to make sure that ρ takes values in R.) Intuitively, ρ(f,g) is a measurement of the “distance” between the graphs of f and g. You should convince yourself that ρ is in fact a metric on the space Fun(X, Y ); for obvious reasons, it is called the uniform metric. Definition 17.11. The metric topology on Fun(X, Y )definedbyρ is called the uniform topology. The basic open sets in this metric are balls of some radius: if 0 <ε<1, then B (f )= f : X Y d f(x),f (x) <εfor all x X . ε 0 → 0 ∈ So a neighborhood of￿ a given function￿ ￿ f0 consists￿ of all functions whose￿ graphs lie ￿ in a strip of width 2ε around the graph of f0.

Example 17.12. When does a sequence of functions fn : X Y converge to another function f : X Y in the uniform topology? For every 0→<ε<1, the open ball → Bε(f) is a neighborhood of f, and therefore has to contain all but finitely many of the fn. In other words, there should exist some N such that

d fn(x),f(x) <ε for every n N and every x X; but this is saying exactly that f f uniformly. ≥ ∈ ￿ ￿ n → Lemma 17.13. In the uniform topology, C(X, Y ) is a closed subset of Fun(X, Y ). Proof. The uniform topology is a metric topology, and so it is enough to show that C(X, Y ) is sequentially closed. Suppose a sequence of continuous functions fn C(X, Y ) converges to a function f Fun(X, Y ). Then fn f uniformly, and we∈ know from Lemma 15.3 that f is also∈ continuous. Thus f → C(X, Y ), and so ∈ C(X, Y ) is closed. ￿ In analysis, people are most interested in the case when (Y,d) is a complete metric space (such as Rn). Let me end today’s class by showing that the space of (continuous) functions inherits this completeness property. Proposition 17.14. If (Y,d) is complete, then both Fun(X, Y ) and C(X, Y ) are complete with respect to the uniform metric.

Proof. We first show that Fun(X, Y ) is complete with respect to ρ. Let fn Fun(X, Y ) be an arbitrary Cauchy sequence. For every ε>0, we can therefore find∈ some N such that

(17.15) d fm(x),fn(x) <ε for every x X and every m, n N. If we fix a point x X, this implies that the ∈ ≥￿ ￿ ∈ sequence fn(x) Y is a Cauchy sequence; but Y is complete, and so it converges to some limit that∈ we shall denote by f(x) Y . In this way, we obtain a function f : X Y . For given ε>0, we can now let∈m go to infinity in (17.15) to obtain → d f(x),f (x) ε n ≤ for every x X and every n N. Therefore ρ(f,f ) ε whenever n N,which ∈ ≥ ￿ ￿ n ≤ ≥ means that f is the limit of the Cauchy sequence fn. Completeness of C(X, Y ) follows from the fact that it is closed: any Cauchy sequence in C(X, Y ) has a limit in Fun(X, Y ), because Fun(X, Y ) is complete; the limit actually belongs to C(X, Y ), because C(X, Y ) is sequentially closed. ￿