Vous êtes sur la page 1sur 15

MATH 104INTRODUCTION TO ANALYSIS SPRING 2012 NOTES ON METRIC SPACES

RUMEN ZAREV

Contents 1. Motivation 1.1. Solution 2. Metric spaces 2.1. Distance functions on Rn 2.2. Distance functions on the circle 2.3. Uniform metric on the space of functions 3. Convergence in metric spaces 4. Topological spaces 5. Closed and compact sets 5.1. Equivalence of metrics 5.2. More topological notions in metric spaces 6. Non-topological aspects of metric spaces 1 2 2 3 4 6 7 8 10 11 12 13

1. Motivation When talking about notions such as convergence, continuity, etc. the key notion is that of points (numbers) being close to each other, or arbitrarily close. For example when working with the extended real numbers we have encountered the following situations: We say that the real numbers x and y are close to each other if |x y | is a small positive number. We say that we can nd xn arbitrarily close to y if we can nd xn such that |xn y | < for any > 0. We say that the real number x is close to +, if x is a large positive number. We say that we can nd xn arbitrarily close to + if we can nd xn > M for any M R. We can use this language to phrase some familiar notions, like the limit of a sequence of real numbers: lim xn = x if for large n, xn is arbitrarily close to x. This would work both for the case of x being a real number, and for the cases of x being . Similarly, for a function f : R R we can dene limxa f (x) = L as follows: for x close to a, f (x) becomes arbitrarily close to L.
1

RUMEN ZAREV

Unfortunately, as we have seen, if we want to use a denition like this in practice we have to consider multiple cases. Ignoring left and right limits, there are nine distinct cases, depending on whether a is real, , or +, and similarly, whether L is real, , or +. This means that really we have nine dierent denitions of limit of a function. Not only that, but if we want to prove or use any theorem about limits, we have to consider all these cases separatelywe would have to prove nine dierent theorems, even when they are all very similar. What if we want to work with convergence, limits, or continuity (and even derivatives, and integrals) in a setting other than the real numbers? What if we are considering multidimensional quantities, e.g. points in a plane, or in space? What about sequences of functions? We have already seen two notions of convergence of sequences of functionspointwise and uniform. Which one works better? Why? 1.1. Solution. As is standard in mathematics, we can solve all of these problems together, by developing a general abstract theory. That is, take the intuitive notion we are usingcloseness, and make it precise. If we can come up with a rigorous denition that captures this, then we can work in full generality. Every theorem we prove in the general case translates into many dierent results in the specic cases. In fact, we will consider two separate constructions: Topological spaces: In a topological space the fundamental notion is that of an open set. It captures the qualitative part of closeness. It makes the statement arbitrarily close precise. Metric spaces: In a metric space the fundamental notion is that of distance. It captures the quantitative part of closeness. In addition to arbitrarily close, we can make also makes statements like x is closer to y than z precise. Of the two, metric spaces are more concrete, and more similar to the real numbers. Thus they are easier to understand, and we will deal with them rst. 2. Metric spaces A metric space consists of some objects (e.g. numbers, points) and a notion of distance between pairs of objects. We can make that precise as follows. Denition 1. A metric space consists of a pair of (X, d), where X is a set, called the space, and d : X X R is a function, called the metric or distance function, assigning a real number to pairs of elements of X . This distance function satises the following conditions: (M1 ) d(x, y ) = d(y, x) for all x, y X . (M2 ) d(x, x) = 0 for all x X , and d(x, y ) > 0 whenever x = y . (M3 ) d(x, y ) + d(y, z ) d(x, z ) for all x, y, z X .

MATH 104METRIC SPACES

All three of these conditions make sense, if we think of d as distance. (M1 ) means that d is symmetricthe distance from x to y is the same as that from y to x. (M2 ) means that the distance from x to itself is always 0, while the distance between any two distinct points is strictly positive. Finally, (M3 ) is the triangle inequality going from x to y to z , cannot be shorter that going from x to z directly. For a real world example of a metric space consider the following. Suppose that X is the set of cities in a country, connected by a highway system, e.g. the United Sates. Let d(x, y ) be thew shortest distance (in miles), along any route, between the cities x and y , e.g. we can have d(Berkeley, New York) = 2897, d(Berkeley, Los Angeles) = 2777, and d(Berkeley, Los Angeles) = 374. As we see, the triangle inequality is satised for these three cities. For a more mathematical example, notice that we have already seen the properties (M1 )(M3 ). They are satised by the absolute value function. This gives us our rst mathematical example. Example 2. We can turn the set of real numbers into a metric space (R, dstd ) by using the standard distance function dstd (x, y ) = |x y |. Indeed, this is the usual notion of distance between two real numbers, and coincides with the distance between the points on the real line representing the numbers. For instance d(1, 4) = 3, and d(5, 6) = 11. 2.1. Distance functions on Rn . Another familiar example of a distance function is that between points in the plane. If we put a coordinate system, and using the Pythagorean theorem, the usual notion of a distance between the points (x1 , x2 ) and (y1 , y2 ) is (x1 y1 )2 + (x2 y2 )2 . In general, for any natural number n, we can turn the ndimensional space Rn into a metric space. Example 3. For any n 1, there is a metric space (Rn , d2 ), where the distance between two points x = (x1 , . . . , xn ) and y = (y1 , . . . , yn ) is d2 (x, y) = (x1 y1 )2 + + (xn yn )2 .

It is easy to see that this function satises (M1 ) and (M2 ). It is a little harder to see it satises the triangle inequality (M3 ). There are two other important metrics that can be put on Rn . One of them is called the taxi-cab metric. To understand it imagine the following situationwe are considering the points in the plane R2 as addresses in a city. To get from one address to another we cannot travel along a straight line as there are buildings in the way. Instead, we have to travel along a grid-like street systemwe can only move in a NorthSouth or an East West direction. The distance between two points is then the sum of the NS distance and the EW distance between them. More formally we have the following.

RUMEN ZAREV

Example 4. For any n 1, there is a metric space (Rn , d1 ), where the distance between the points x = (x1 , . . . , xn ) and y = (y1 , . . . , yn ) is d1 (x, y) = |x1 y1 | + + |xn yn |. The third metric we would like to consider is called supremum or uniform metric. In this case the distance between two points is the maximum dierence in any coordinate. Example 5. For any n 1, there is a metric space (Rn , d ), where the distance between the points x = (x1 , . . . , xn ) and y = (y1 , . . . , yn ) is d (x, y) = max{|x1 y1 |, . . . , |xn yn |}. Notice that we have used very similar notation for two of these metrics d1 and d2 . This is due to the fact that they are special cases of a more general construction, called the pmetric. Example 6. For any n 1 and real number p 1, there is a metric space (Rn , dp ), where the distance between the points x = (x1 , . . . , xn ) and y = (y1 , . . . , yn ) is dp (x, y) =
p

|x1 y1 |p + + |xn yn |p .

We can try do that for any positive number p. However, unless p is at least 1, the condition (M3 ) fails. The reason the supremum metric is denoted by d is that it is, in a sense, the limit of dp as p goes to innity. More precisely, for any xed pair x, y Rn , we have d (x, y) = limp+ dp (x, y). Exercise 7. For a xed natural number n check that dp for p [1, +) and d are indeed metrics on Rn , that is they satisfy the three conditions (M1 )(M3 ). Hint: for the triangle inequality (which is the hardest to show) rst try the case y = (0, . . . , 0). The general case is very similar. Notice for R1 = R, i.e. the real numbers, all of these metrics coincide (and are equal to the standard metric dstd ). To get a grasp of how these distance functions behave, it helps to visualize a sphere, i.e. the set of points equidistant to some xed point. For instance, in R2 we can visualize the sphere of radius 1 around (0, 0) according to dierent metrics. For the Euclidean metric d2 we get the usual circle. However, for other metrics we get dierent picturessquares for d1 and d , and something between a square and a circle for say d3 . This can be seen in Fig. 1. The points (0, 1), (0, 1), (1, 0), and (1, 0) are included in all four cases, but the rest of the points dier. 2.2. Distance functions on the circle. Consider the standard circle S 1 = {(x, y ) : x2 + y 2 = 1}. Instead of by Euclidean coordinates, we can refer to points on the circle by their angle with the xaxis. That is, we consider points [0, 2 ), where corresponds to the point in R2 with coordinates (x(), y ()) = (cos , sin ) (see Fig. 2). We can put a distance function on the circle by measuring the Euclidean distance in R2 , that is the distance along a straight line.

MATH 104METRIC SPACES

Figure 1. Points in the sphere {x R2 : d(x, (0, 0)) = 1} for dierent metrics d. From left to right we have d1 , d3 , d2 , and d .

Figure 2. The correspondence between [0, 2 ) and a point (x, y ) on the circle S 1 .

Example 8. There is a metric space ([0, 2 ), d2 ), where d2 (, ) = (cos cos )2 + (sin sin )2 = 2 sin . 2

There are two ways to check that d2 is a metric on [0, 2 ). One is to do it directly, using the formula. The other is by using the fact that we know it is a metric on the bigger space R2 , which contains S 1 . This is an example of the subspace construction. Denition 9. Suppose (X, dX ) is a metric space, and Y is any subset of Y . We can turn into a metric space by putting on it the subspace metric dY : Y Y R, which is just the restriction dY = (dX ) |Y Y . Theorem 10. For any metric space (X, dX ) and any Y X , and for dY as in Def. 9, (Y, dY ) is a metric space. Proof. The proof consists of checking that dY satises (M1 )(M3 ). This is straightforward from the fact that dX satises them. This metric on [0, 2 ) is somewhat hard to work with as it involves trigonometric functions. Instead, we can measure distances along the circle, instead of along a straight line (see Fig. 3). This distance has a much easier formula.

RUMEN ZAREV

Example 11. We can turn [0, 2 ) into a metric space ([0, 2 ), darc ), using the arc metric darc : [0, 2 ) [0, 2 ) R, dened by darc (, ) = min{| |, 2 | |}.

Figure 3. The two metrics on the circle.

2.3. Uniform metric on the space of functions. The next example we can consider is the uniform metric on the space of functions. This is a generalization of the uniform metric d on the nite dimensional spaces Rn , so we will use the same notation. Consider the following two sets: C = {f : [0, 1] R : f is continuous}, B = {f : R R : f is bounded}. C is the set of all continuous functions on the closed interval [0, 1], and B is the set of all bounded functions on R. (Note that any function f C is also bounded, by the Extreme Value Theorem). Example 12. We can turn C and B into metric spaces (called function spaces ) (C , d ) and (B , d ), where we use d to denote the (two distinct) supremum or uniform metrics dened as follows: d (f, g ) = sup |f (x) g (x)|
x[0,1]

f, g C , f, g B .

d (f, g ) = sup |f (x) g (x)|


x R

Although we abuse notation by referring to both by the same symbol, it should not cause any confusion. Notice that d (f, g ) measures the largest dierence between f and g . We are tempted to dene it for all functions R R, but we run into a problem if f or g is unbounded. For example, if f (x) = x and g (x) = x, than the formula gives supxR |2x| = +. Distances, however, have to be real. We will come back to this example in a bit.

MATH 104METRIC SPACES

3. Convergence in metric spaces After all the examples we introduced, we will nally generalize some of our denitions about R to metric spaces. This is straightforward, by replacing all mentions of absolute values by distances. Denition 13. Suppose (X, d) is a metric space, xn is a sequence of points in X , and x is another point in X . We say that xn converges to x, or that x is the limit of xn , denoted, limd xn = x, if the following statement holds: for any real > 0, there exists some N , such that n > N implies d(xn , x) < . We say that a sequence xn in X is a convergent sequence if it has a limit x X. If d is unambiguous, we can write lim xn as usual, instead of limd xn . Notice that for (R, dstd ) this agrees with usual denition. Exercise 14. For some examples of convergent sequences, consider the se1 n1 n quence x(n) = (xn in R2 . Show (using Def. 13) that 1 , x2 ) = n , n lim x(n) = lim x(n) = lim x(n) = (0, 1).
d1 d2 d

On the other hand, the sequence n = 2 [ 2n ] on the circle, where [x] denotes the oor of x, converges for neither d2 nor darc . It keeps going around and around the circle without slowing down. Example 15. Notice that convergence for the metric d in the function spaces C and B from Example 12 is equivalent to uniform convergence. This justies calling it the uniform metric. On the other hand, pointwise convergence of functions cannot be expressed as convergence in a metric space. (Unless the domain is nite. In that cases the function space is actually Rn ). Somewhat implicit in this discussion is the following result. Theorem 16. If a limit in a metric space exists, it is unique, just as with R. Proof. This is a straightforward modication of the usual proof. Suppose ) y = z are both limits for xn in (X, d). Let = d(y,z 3 . This is positive by property (M2 ). By the convergences, there are N1 and N2 , such that n > N1 implies d(xn , y ) < , and n > N2 implies d(xn , z ) < . We can pick a large enough n for which both are satised. Then d(y, xn ) + d(xn , z ) = d(xn , y ) + d(xn , z ) < 2 < 3 = d(y, z ). This contradicts the triangle inequality, i.e. property (M3 ). Notice that in these example, the existence (and value) of the limit did not depend on the metric. This is not a coincidence, and corresponds to the fact that all our metrics for R2 are equivalent. The two metrics on the circle are also equivalent. Before getting to the notion of equivalent metrics, we need to understand the denition of a limit in two alternative ways. First, we need another denition.

RUMEN ZAREV

Denition 17. Let (X, d) be a metric space. If x X is a point in the space, and > 0 is arbitrary, we can dene the open ball of radius around x, or the ball around x, to be the set B (x) = {y X : d(x, y ) < }. In (R, dstd ), this is the interval (x, x+). In (R2 , d2 ) it is a disk (without its boundary) around x, and in (R3 , d2 ) it is a round three-dimensional ball (without the boundary) around x, of radius . Hence the name. For d1 , d2 , d3 , and d on R2 , the balls of radius 1 around (0, 0) are the interiors of the curves in Fig. 1. Theorem 18. Let xn be a sequence in the metric space (X, d), and x X . The following three statements are equivalent: (1) limd xn = x, (2) For any > 0, there is an N , such that n > N implies xn B (x), (3) limn d(xn , x) = 0 (as a usual limit of real numbers). Proof. The equivalence of (1) and (2) is immediate from the denitions. The equivalence with (3) is a basic consequence of the properties of (usual) convergence for real numbers. 4. Topological spaces The statement (2) in Thm. 18 hints at a way of looking at metric spaces in a way less sensitive to the numerical value of distances. The idea is to replace B (x) by any set of points close to x, or neighborhood of x. Similar to the notion of distance, we can reduce the notion of a neighborhood to an abstract denition. This is somewhat farther from the intuition, though. Denition 19. A topological space is a pair (X, T ), where X is a set, called the space, and T is some collection of subsets of X , each of which we call an open subset of X , such that the following conditions are satised: (T1 ) , X T , (T2 ) For any index set (nite or innite) I , if Ui T for all i I , then iI Ui T . (T3 ) For any natural number n, if U1 , . . . , Un T , then U1 . . . Un T . A set V X is called closed if its complement X \ V is open. In a more understandable language, (T1 ) says that the empty set, and the space itself are open, (T2 ) says that arbitrary unions of open sets are open, and (T3 ) says that nite intersections of open sets are open. Note that by basic set algebra, and properties of complements, we see that closed sets satisfy some properties analogous to those for open sets (except unions and intersections are switched). Theorem 20. (T1 ) and X are closed. (T2 ) Any intersection (nite or innite) of closed sets is closed.

MATH 104METRIC SPACES

(T3 ) Any nite union of closed sets is closed. Example 21. For example, the usual notion of open sets in R (i.e. unions of open sets) makes R into a topological space. More generally, any metric space can become a topological space. Theorem 22. Let (X, d) be a metric space. Let T consist of all sets U X with the following property: if x U then there is some real > 0, such that B (x) U . Then (X, T ) is a topological space. (We call this the metric topology on X ). Proof. We will prove that T satises (T3 ) as this is the most interesting condition. Suppose U1 , . . . , Un T , and U = n i=1 Ui . Let x U . We need to show U contains a ball around x. Since U is an intersection, x Ui for all i = 1, . . . , n. But Ui T , so Ui contains some ball Bi (x) for i > 0. Let = min{1 , . . . , n } > 0. Then B (x) Bi (x) Ui for all i, so B (x) U . Note that this would fail for an innite intersection. Instead of a minimum we would need to take an inmum of i , but that could be 0 rather than a positive number. In other words, a set in X is open if around each of its points it contains an entire open ball. This means that any open set can be written as a (possibly innite) union of open balls. By property (T2 ), all such unions are open. This can be taken as an alternative denitiona set is open in the metric topology if it is the union of open balls. We can now translate Def. 13 into a statement about open sets without a reference to a metric. This becomes a denition for topological spaces. Denition 23. Let (X, T ) be a topological space, xn a sequence in X , and x a point in x. We say that xn converges to x, or limT xn = x if for any open set U T which contains x, there is some N such that n > N implies xn U . We can easily see that for the metric topology this is equivalent to the previous denition. Theorem 24. Let (X, d) be a metric space, and (X, T ) be its metric topology. Then limd xn = x if and only if limT xn = x. Proof. The backward direction (topological convergence implies metric space convergence) is easy. Just apply Def. 23 to the open set B (x). In the other direction, given any open set U around x, it contains some ball b (x) U for > 0. Now apply Def. 13 to this . Note that while the topological limit agrees with the metric limit for the metric topology, it is not as well behaved for more general topological spaces. For example, the limit is not necessarily unique. Exercise 25. Show that (N, T ) is a topological space, where N is the natural numbers, and T consists of all sets U with nite complements, or is empty.

10

RUMEN ZAREV

That is, U T if it contains all but nitely many of the natural numbers (or none at all). This is called the conite topology. Show that limT n = m for all m N. I.e. the sequence 1, 2, 3, . . . has all natural numbers as limits under this topology. The next example shows how we can unify nite and innite limits of sequences in R into one. Example 26. Consider (R, T ), where R = R {, +} is the set of extended real numbers, and T consists of sets U satisfying the following conditions: (1) If x U and x is real, then there is some > 0, such that (x , x + ) U . (2) If U then there is some M R, such that (, M ) U . (3) If + U then there is some M R, such that (M, +) U . This pair is a topological space, where the notion of convergence is equivalent to our usual three notions of limits. In a sense the intervals [, M ) and (M, +] play the roles of open balls around and +, respectively. This is in fact a metric space topology. However, to realize it while having d(, +) be nite, we need some not-so-pretty formulas. One option is to set the metric as follows: For x, y R, set d(x, y ) = | tan1 x tan1 y |. For x R, set d(x, ) = tan1 x + 2. For x R, set d(x, +) = tan1 x + 2. Finally, set d(, +) = . The metric topology for d is precisely T . Exercise 27. Prove the assertions in Example 26. 5. Closed and compact sets There are couple of notions that we have encountered before, for sets in R, which can be generalized to metric spaces. One is the notion of closed sets. In fact, we have already given the definition of a closed set. However, we should notice that one of the most important properties of closed sets in R generalizes to this case. Theorem 28. Let (X, T ) be a topological space, let V X be closed, and let xn be a sequence of points in V . Suppose x = limT xn (recall that x might not be unique). Then x V . (In other words, the limit(s) of any sequence in V are also in V ). Proof. The complement X \ V is open, i.e. in T . By the denition of convergence, there is some N such that n > N implies xn X \ V . However, xn V for all n. This is a contradiction. Another important notion is that of compact spaces. The general topological denition is somewhat convoluted.

MATH 104METRIC SPACES

11

Denition 29. Suppose V is a topological space, or a subset of a topological space. It is called compact if the following condition holds: if V is contained in the union aI Ua of some open sets (where the index set I could be arbitrary), then it is also contained in a nite union of some of these sets: V Ua1 . . . Uan for some a1 , . . . , an I . For metric spaces this is equivalent to something much nicer (and more familiar): Theorem 30. Let (X, d) be a metric space, and V be any subset of X . Then V is compact (with respect to the metric topology) if and only if any sequence of points in V has a subsequence which converges to a point in V . Proof. We will omit the proof. Note that if V is the whole space, then V is compact if and only if any sequence in V has a convergent subsequence. Another important theorem (for which we omit the proof) describes all compact sets in Rn (with the metric topology for any dp , p [1, +]). Theorem 31. A set V Rn is compact if and only if it is closed and bounded (i.e. there is an M > 0, such that d(x, 0) < M for all x V ). For example, the circle S 1 with either metric d2 , or darc , is compact. 5.1. Equivalence of metrics. We can get back to the question of why all the metrics dp for p [1, +] give the same notion of convergence on Rn . This is due to the fact that they they are equivalent metrics. Denition 32. Let d and d be two functions X X R, such that (X, d) and (X, d ) are both metric spaces. We say that they are equivalent metrics if the following two conditions hold: (1) For all x X and > 0, there is some > 0, such that d(x, y ) < implies d (x, y ) < . (2) For all x X and > 0, there is some > 0, such that d (x, y ) < implies d(x, y ) < . Note that the two conditions are symmetric to each other. Roughly, (1) means that points which are close when measured by d are also close when measured by d , and (2) is the reverse statement. More precisely we have the following implications. Theorem 33. If (X, d) and (X, d ) are equivalent metrics, then the following statements hold: (1) Any open ball with respect to d around x contains an open ball around x with respect to d, and vice versa. (2) Any open set with respect to d is open with respect to d, and vice versa. (3) d and d dene the same metric topology on X .

12

RUMEN ZAREV

Proof. (1) is a restatement of the denition. (2) follows quickly, and (3) is a restatement of (2). As an immediate corollary we see that limd xn = x if and only if limd xn = x. Exercise 34. Show that the metrics d1 , d2 , and d on R2 are equivalent to each other. Hint: prove that for any real a, b R, we have 2 max{|a|, |b|} |a| + |b| a2 + b2 max{|a|, |b|}.

For a bigger challenge, prove that for any xed n, all the metrics dp for p [1, ] on Rn are equivalent to each other. Exercise 35. Prove that the metrics d2 and darc on [0, 2 ) (from Examples 8 and 11) are equivalent. Hint: use the fact that for any x [0, ], we have x 2 sin x x. Finally, there is another notion of convergence in Rn that seems more familiar from calculuscoordinate-wise convergence. It is also equivalent to metric convergence. Theorem 36. Suppose x(m) = (x1 , . . . , xn ) is a sequence of points in Rm , and x = (x1 , . . . , xn ) is another point in Rn . Then limdp x(m) = x (for p [1, ])) if and only if lim xi = xi for all i = 1, . . . , n. In other words, convergence in the metric sense is equivalent to convergence of all individual coordinates. Proof. In light of Exercise 34, it is enough to prove for one p. We will do it for d . The forward direction is easy. If limd x(m) = x, then d (x(m) , x) con(m) (m) verges to 0. However |xi xi | d (x(m) , x) and lim xi = xi by the squeeze theorem, for all i. In the backward direction, suppose the individual coordinates converge. Then for any > 0, there are M1 , . . . , Mn , such that m > Mi implies (m) |xi xi | < . Take M = max{M1 , . . . , Mn }. Then for m > M all of the inequalities hold, so d (x(m) , x) < as well. 5.2. More topological notions in metric spaces. In addition to convergence and limits of sequences that are a few other notions for metric spaces that are really topological notions. For the denitions, just substitute distance for absolute value. Denition 37. Suppose (X, dX ) and (Y, dY ) are metric spaces, and f : X Y is a function. We say that limxa f (x) = L (where a X and L Y ) if for all > 0 there is some > 0 such that if dX (x, a) < and x = a, then dY (f (x), L) < . For topological spaces this generalizes to the following.
(m) (m) (m)

MATH 104METRIC SPACES

13

Denition 38. Suppose (X, TX ) and (Y, TY ) are topological spaces, and f : X Y is a function. We say that limxa f (x) = L (where a X and L Y ) if for all U TY containing L, there is some V TX containing a, such that f (V ) U . It is easy to see that for metric spaces the two denitions are equivalent, so existence and value of limits are the same for equivalent metrics. Similarly, we can dene continuous maps between metric spaces, and generalize to topological spaces. Denition 39. For (X, dX ) and (Y, dY ) metric spaces, a function f : X Y is continuous if for any x X and any > 0, there is some > 0, such that dX (x, y ) < implies dY (f (x), f (y )) < . For (X, TX ) and (Y, TY ) topological spaces, a function f : X Y is continuous if for any open set U TY , its preimage is also open, i.e. f 1 (U ) = {x X : f (x) U } TX . Again, for metric spaces the denitions are equivalent, so continuity is the same for equivalent metrics. As an example, it makes sense to talk about continuous maps Rn Rm without specifying the metric used (any choice of p [1, ] for Rn , and any (possibly dierent) choice for Rm give the same result). Thus we can pick and choose the metric for convenience. Recall that in class we mentioned that the Extreme Value Theorem holds for any compact set. In fact it holds for any compact topological space. Theorem 40. Suppose X is a compact metric (or topological) space. If f : X R is continuous (using the standard metric topology on R), then the image f (X ) R has a maximum and a minimum. Proof. For metric spaces, the proof is the same as for a nite closed interval in R. As a sketch, if we assume f (X ) is unbounded, then there is a sequence xn for which f (xn ) diverge to + or . By compactness, there is a convergent subsequence, and by continuity we get a point where f is innite. If f has a nite supremum, we can pick a sequence of xn for which f (xn ) approaches the supremum. Again, we can pass to a convergent subsequence to get a point x where f (x) is equal to the supremum. For topological spaces the proof is somewhat dierent, but we will omit it. 6. Non-topological aspects of metric spaces As we saw in the previous section, many of the notions about metric spaces are in fact topological, and not that sensitive to the metric. However, some notions are not topological. This is basically anything which compares distances between dierent pairs of points, as this cannot be done in an arbitrary topological space. One such notion is that of Cauchy sequences.

14

RUMEN ZAREV

Denition 41. A sequence of points xn in a metric space (X, d) is called Cauchy if for all > 0 there is some N , such that n, m > N implies d(xn , xm ) < . Example 42. A sequence of functions fn in one of the function spaces (C , dinf ) or (B , dinf ) is Cauchy if and only if it is uniformly Cauchy as dened previously. To really see that being Cauchy is not a topological property, consider the following example. Example 43. Let dexp : R R R be the exponential metric dexp (x, y ) = |ex ey |. Then (R, dexp ) is a metric space. The metric can be interpreted as follows: as we move toward points get closer and closer together. In fact the whole interval (, 0) has nite length 1. Recall the standard metric dstd (x, y ) = |x y |. One can check that (R, dstd ) and (R, dexp ) are equivalent metric spaces. Thus any topological notions about them should be the same. Let xn = n. This sequence is clearly not Cauchy for the standard metric (it is not Cauchy in the usual . sense). However, it is Cauchy for dexp . Indeed, for > 0, let N = log 2 n m n m For any n, m > N , we get dexp (xn , xm ) = |e e | < e + e < 2eN = . So xn is Cauchy for one metric, but not for another, equivalent, metric. On the other hand, convergence is topological. xn does not converge for dstd , so it doesnt converge for dexp either. This means that we have an example of a sequence in the metric space (R, dexp ) which is Cauchy but does not converge. We see that unlike in R, in metric spaces Cauchy and convergent are not equivalent. One direction is still true. Any convergent sequence is Cauchy. However the other direction might fail as in the above example. Denition 44. A metric space is called complete if all Cauchy sequences converge. For subspaces of (R, dstd ) (i.e. subsets of R with the usual notion of convergence), this completeness is equivalent to completeness in the sense of least upper bounds. Example 45. The two functional spaces B and C from Example 12, are complete with respect to d . The completeness of B is due to the fact that a uniformly Cauchy sequences is pointwise Cauchy so it pointwise converges. As we have seen it also uniformly converges to the same limit. The limit must be bounded, so it still in B . Similarly, a uniformly Cauchy sequence in C has a uniform limit. This limit must also be continuous, so it is also in C . The nal example of something that can be dened for metric spaces, but not for topological spaces is uniformly continuous functions.

MATH 104METRIC SPACES

15

Denition 46. Suppose (X, dX ) and (Y, dY ) are metric spaces. The function f : X Y is uniformly continuous if for all > 0 there is a > 0 such that dX (x, y ) < implies dY (f (x), f (y )) < . As in the usual denition of uniform continuity, depends only on , not on x. Thus we can not phrase the condition in term of open sets only, to make it topological.

Vous aimerez peut-être aussi