Author: Munkres J.R.  

Tags: mathematics   physics  

Year: 1974

Text
                    TOPOLOGY
A First Course
JAMES R. MUNKRES
Professor of Mathematics
Massachusetts Institute of Technology
Prentice-Hall, Inc., Englewood Cliffs, New Jersey


Contents Preface xl A Note to the Reader xv Set Theory and Logic ' 1-1 Fundamental Concepts 4 1-2 Functions 15 1-3 Relations 21 1-4 The Integers and the Real Numbers 29 1-5 Arbitrary Cartesian Products 36 1-6 Finite Sets 39 1-7 Countable and Uncountable Sets 45 *l-8 The Principle of Recursive Definition 53 1-9 Infinite Sets and the Axiom of Choice 57 1-10 Weil-Ordered Sets ' 63 *1-11 The Maximum Principle 68 •Supplementary Exercises: Well-Ordering 72 vii
viii Contents Chapter 2. Topological Spaces and Continuous Functions Topological Spaces Basis for a Topology The Order Topology The Product Topology on X x Y The Subspace Topology Closed Sets and Limit Points Continuous Functions The Product Topology The Metric Topology The Metric Topology (continued) The Quotient Topology 2-1 2-2 2-3 2-4 2-5 2-6 2-7 2-8 2-9 2-10 ♦2-11 ♦Supplementary Exercises: Topological Groups Chapter 3. Connectedness and Compactness 3-1 Connected Spaces 3-2 Connected Sets in the Real tine *3-3 Components and Path Components *3-4 Local Connectedness 3-5 Compact Spaces 3-6 Compact Sets in the Real Line 3-7 Limit Point Compactness *3-8 Local Compactness ♦Supplementary Exercises: Nets Chapter 4. Countability and Separation Axioms 4-1 The Countability Axioms 4-2 The Separation Axioms 4-3 The Urysohn Lemma 4-4 The Urysohn Metrization Theorem *4-5 Partitions of Unity ♦Supplementary Exercises: Review of Part I PARTn 227 Chapter 5. The Tychonoff Theorem 5-1 The Tychonoff Theorem 5-2 Completely Regular Spaces 5-3 The Stone-Cech Compactification 229 229 235 238
Contents ix Chapter 6. Metrization Theorems and Paracompactness 6-1 Local Finiteness 6-2 The Nagata-Smirnov Metrization Theorem (sufficiency) 6-3 The Nagata-Smirnov Theorem (necessity) 6-4 Paracompactness 6-5 The Smirnov Metrization Theorem 244 245 247 251 254 260 Chapter 7. Complete Metric Spaces and Function Spaces 7-1 Complete Metric Spaces 7-2 A Space-Filling Curve 7-3 Compactness in Metric Spaces 7-4 Pointwise and Compact Convergence 7-5 The Compact-Open Topology 7-6 Ascoli's Theorem 7-7 Baire Spaces 7-8 A Nowhere-Differentiable Function 7-9 An Introduction to Dimension Theory 262 263 271 274 280 285 289 293 297 301 Chapter 8. The Fundamental Group and Covering Spaces 8-1 8-2 8-3 8-4 8-5 8-6 8-7 8-8 8-9 8-10 8-11 8-12 8-13 8-14 Homotopy of Paths The Fundamental Group Covering Spaces The Fundamental Group of the Circle The Fundamental Group of the Punctured Plane The Fundamental Group of S" Fundamental Groups of Surfaces Essential and Inessential Maps The Fundamental Theorem of Algebra Vector Fields and Fixed Points Homotopy Type The Jordan Separation Theorem The Jordan Curve Theorem The Classification of Covering Spaces Bibliography Index 316 318 326 331 336 343 348 351 357 361 364 369 374 378 387 399 401
Preface This book is intended as a text for a one- or two-semester introduction to topology, at the senior or first-year graduate level. The subject of topology is of interest in its own right, and it also serves to lay the foundations for future study in analysis, in geometry, and in algebraic topology. There is no universal agreement among mathematicians as to what a first course in topology should include; there are many topics that are appropriate to such a course, and not all are equally relevant to these differing purposes. In the choice of material to be treated, I have tried to strike a balance among the various points of view. Prerequisites. There are no formal subject matter prerequisites for study- studying most of this book. I do not even assume the reader knows much set theory. Having said that, I must hasten to add that unless the reader has studied a bit of analysis or "rigorous calculus," he will be missing much of the motivation for the concepts introduced in the first part of the book. Things will go more smoothly if he already has had some experience with continuous functions, open and closed sets, metric spaces, and the like, although none of these is actually assumed. In Chapter 8, we do assume lamihanty with the elements of group theory. ^ Most students in a topology course have, in my experience, some knowl- ge ot the foundations of mathematics. But the amount varies a great deal JCI
xii Preface from one student to another. Therefore I begin with a fairly thorough chapter on set theory and logic. It starts at an elementary level, and works up to a level that might be described as "semi-sophisticated." It treats those topics (and only those) which will be needed later in the book. Most students will already be familiar with the material of the first few sections, but many of them will find their expertise disappearing somewhere about the middle of the chapter. How much time and effort the instructor will need to spend on this chapter will thus depend largely on the mathematical sophistication and experience of his students. Ability to do the exercises fairly readily (and correctly!) should serve as a reasonable criterion for determining whether the student's mastery of set theory is sufficient for him to begin the study of topology. How the book is organized. When this book is used for a one-semester course, some choices will have to be made concerning what material to cover. I have attempted to organize the book as flexibly as possible, so as to enable the instructor to follow his own preferences in this matter. Part I of the book, consisting of the first four chapters, deals with that body of material which in my opinion should be included in any introductory topology course worthy of the name. This may be considered the "irreducible core" of the subject, treating as it does topological spaces, connectedness, compactness (through compactness of finite products), and the counta- bility and separation axioms (through the Urysohn metrization theorem). Certain sections are marked with an asterisk; these do not form part of the basic core and may be omitted or postponed with no loss of continuity. Part II of the book consists of four chapters which are entirely indepen- independent of one another. They depend only on the material of Part I; the instruc- instructor may take them up in any order he chooses. Furthermore, if he wishes to cover only a portion of one of these later chapters, he can consult the introduction to that chapter, where there appears a diagram showing the relations of dependence among the sections of the chapter. The instructor who wishes, for instance, to conclude his course with a proof of the Jordan curve theorem can determine from this diagram which of the earlier sections of Chapter 8 are essential, and which peripheral, to his purpose. Some of the material of the later chapters depends on one or more of the asterisked sections in Part I. Each such dependence is indicated in a footnote at the beginning of the asterisked section, and again in the introduction to the chapter in question. Some of the exercises also depend on earlier aster- asterisked sections, but in such cases the dependence is obvious. Possible course outlines. Most instructors who use this text for a one- semester course will wish to cover the "core" material of Part I, along with the Tychonoff theorem (§5-1). Many will cover additional topics as well. One might, for instance, treat some of the asterisked sections of Part I. (l
Preface xiii ally do local compactness, at least.) Or he may choose one or more ^ cs from Part II. Possibilities include: the Stone-Cech compactification r«5-3)» metrization theorems (Chapter 6), the Peano curve (§7-2), one or both versions of Ascoli's theorem (§7-3 and §7-6), dimension theory (§7-9), the fundamental group and applications (§8-l-§8-10), or the Jordan curve theorem (§8-13). I have in different semesters followed each of these options. For the instructor who wishes to emphasize algebraic topology, one ssible course outline would consist of Chapters 1 to 3 followed by Chapter 8 in its entirety. Omitting Chapter 4 will cause no difficulty, provided one skips Exercise 5 of §8-12, which involves the concept of normality. Still another possible outline is the one suggested by the Committee on the Undergraduate Program in Mathematics (of the Mathematical Associa- Association of America) for a one-semester course in topology at the first-year graduate level. It would consist of Chapters 2, 3, and 4, followed by §5-1; §6-1, §6-3, §6-4; §7-1; §8-1 through §8-5, §8-8 through §8-11, and §8-14. This program assumes that the student has already had an introduction to set theory equivalent to our Chapter 1. In a two-semester course, one can reasonably expect to cover the entire book. Acknowledgements. Most of the topologists with whom I have studied, or whose books I have read, have contributed in one way or another to this book; I mention only Edwin Moise, Raymond Wilder, Gail Young, and Raoul Bott, but there are many others. For their helpful comments con- concerning this book, my thanks to Robert Mosher and John Hemperly, and to my colleagues George Whitehead and Kenneth Hoffman. My apprecia- appreciation goes to Miss Viola Wiley, who deciphered my handwriting and con- converted it into neat copy, and to the employees of Bertrick Associate Artists, Inc., who drew the illustrations. But most of all, to my students go my most heartfelt thanks. From them I learned at least as much as they did from me; without them this book would be very different. J.R.M.
1. Set Theory and Logic We adopt, as most mathematicians do, the naive point of view regarding set theory. We shall assume that what is meant by a set of objects is intuitively clear, and we shall proceed on that basis without analyzing the concept further. Such an analysis properly belongs to the foundations of mathematics and to mathematical logic, and it is not our purpose to initiate the study of those fields. Logicians have analyzed set theory in great detail, and they have formula- formulated axioms for the subject. Each of their axioms expresses a property of sets that mathematicians commonly accept, and collectively the axioms provide a foundation broad enough and strong enough that the rest of mathematics «an be built on them. It is unfortunately true that careless use of set theory, relying on intuition alone, can lead to contradictions. Indeed, one of the reasons for the axio- roatjzation of set theory was to formulate rules for dealing with sets that would ajoid these contradictions. Although we shall not deal with the axioms explic- 'ily.the rules we follow in dealing with sets derive from them. In this book, y°u will learn how to deal with sets in an "apprentice" fashion, by observing you ^ handle them and by workin« with them yourself. At some point of detr.^tu°les y°u may wish to study set theory more carefully and in greater '■» then a course in logic or foundations will be in order.
Chap, i 4 Set Theory and Logic 1-1 Fundamental Concepts Here we introduce the ideas of set theory, and establish the basic terrnU noloTv and notation. We also discuss some points of elementary logic that, ,n our experience, are apt to cause confusion. Basic Notation Commonly we shall use capital letters A, B,... to denote sets, and lower- lowercase letters a, b,... to denote the objects or elements belonging to these sets. If an object a belongs to a set A, we express this fact by the notation a e A. If a does not belong to A, we express this fact by writing a $ A. The equality symbol = is used throughout this book to mean logical identity. Thus when we write a = b, we mean that "a" and "b" are symbols for the same object. This is what one means in arithmetic, for example, when one writes £ = \. Similarly, the equation A = B states that "A" and "B" are symbols for the same set; that is, A and B consist of precisely the same ob- objects. If a and b are different objects, we write a ^ b; and if A and B are different sets, we write A ^ B. For example, if A is the set of all nonnegative real numbers, and B is the set of all positive real numbers, then A ^ B, because the number 0 belongs to A and not to B. We say that A is a subset of B if every element of A is also an element of B; and we express this fact by writing A c B. Nothing in this definition requires A to be different from B- in fact if A = B, it is true that both A <= B and B «= A. If A cz B and A is different from B, we say that A is a proper subset of B and we write B. one^fr.T f , a Set ? If the set has only a few elements, 2dmems^18 ^ ^CtS ^ thE SEt' Witin« "A is the *« consisting of elements a, b, and c." In symbols, this statement becomes A={a,b,c}, where braces are used to enclose the list of elements The usual way to specify a set, however, isTo Uke some set A of objects
Fundamental Concepts §!-' property that elements of A may or may not possess, and to form a"d Consisting of all elements of A having that property. For instance, one AC ht take the set of real numbers and form the subset B consisting of all ttI integers. In symbols, this statement becomes B = [x\x is an even integer}. the braces stand for the words "the set of," and the vertical bar stands fofthe words "such that." The equation is read,  is the set of all x such that x is an even integer." The Union of Sets and The Meaning of "or" Given two sets A and B, one can form a set from them that consists of all the elements of A together with all the elements of B. This set is called the union of A and B and is denoted by A U B. Formally, we define A u B = {x\x £ A or x e B}. But we must pause at this point and make sure exactly what we mean by the statement "x e A or x e B." In ordinary everyday English, the word "or" is ambiguous. Sometimes the statement "P or Q" means "P or Q, or both" and sometimes it means "P or Q, but not both." Usually one decides from the context which meaning is intended. For example, suppose I spoke to two students as follows: "Miss Smith, every student registered for this course has taken either a course in linear algebra or a course in analysis." "Mr. Jones, either you get a grade of at least 70 on the final exam or you will flunk this course." In the context, Miss Smith knows perfectly well that I mean "everyone has had linear algebra or analysis, or both," and Mr. Jones knows I mean "either he gets at least 70 or he flunks, but not both." Indeed, Mr. Jones would be exceedingly unhappy if both statements turned out to be true! In mathematics, one cannot tolerate such ambiguity. One has to pick just °ne meaning and stick with it, or confusion will reign. Accordingly, mathe- mathematicians have agreed that they will use the word "or" in the first sense, so «pat the statement "P or Q" always means "P or Q, or both." If one means °r Q, but not both," then one has to include the phrase "but not both" explicitly, ^te 'th th'S understandinS> the equation defining Au Bis unambiguous; it h&t a u Bis the set consisting of all elements x that belong to A or to ° °r to both.
6 Set Theory and Logic Chap, j The Intersection of Sets, The Empty Set, and The Meaning of "If... Then" Given sets A and B, another way one can form a set is to take the com- common part of A and B. This set is called the intersection of A and B and is denoted by A n B. Formally, we define A n B = {x\x e A and x e B}. But just as with the definition of A U B, there is a difficulty. The difficulty is not in the meaning of the word "and"; it is of a different sort. It arises when the sets A and B happen to have no elements in common. What meaning does the symbol A n B have in such a case? To take care of this eventuality, we make a special convention. We intro- introduce a special set which we call the empty set, denoted by 0, which we think of as "the set having no elements." Using this convention, we express the statement that A and B have no elements in common by the equation A n B = O. We also express this fact by saying that A and B are disjoint. Now some students are bothered by the notion of an "empty set." "How," they say, "can you have a set with nothing in it?" The problem is similar to that which arose many years ago when the number 0 was first introduced. The empty set is only a convention, and mathematics could very well get along without it. But it is a very convenient convention, for it saves us a good deal of awkwardness in stating theorems and in proving them. Without this convention, for instance, one would have to prove that the two sets A and B do have elements in common before one could use the notation A D B, Similarly, the notation C = {x\x e A and x has a certain property} could not be used if it happened that no element x of A had the given prop- property. It is much more convenient to agree that A n B and C equal the empty set in such cases. Since the empty set 0 is merely a convention, we must make conventions relating it to the concepts already introduced. Because 0 is thought of as "the set with no elements," it is clear we should make the convention that for each object x, the relation x e 0 does not hold. Similarly, the definitions of union and intersection show that for every set A we should have the equations A U 0 = A and A n 0 = 0. The inclusion relation is a bit more tricky. Given a set A, should we agree that 0 <= A ? Once more we must be careful about the way mathematicians use the English language. The expression 0 <= A is a shorthand way of writ-
Fundamental Concepts §1-1 "Every element that belongs to the empty set also belongs ing ^ sent ^, ^ ^ ^ ^^ formally) «For every object x, if * belongs to t0 ^ ntv set, then x also belongs to the set A." th£ "Xfc statement true or not? Some might say "yes" and others say no. Ml never settle the question by argument, only by agreement. This is a Y°U nt of the form "If P, then Q," and in everyday English the meaning stateme^ ^en,, construction is ambiguous. It always means that if P is °f then Q is true also. Sometimes that is all it means; other times it means Tmething more: that if P is false, Q must be false. Usually one decides from thTcontext which interpretation is correct. The situation is similar to the ambiguity in the use of the word "or." One n reformulate the examples involving Miss Smith and Mr. Jones to illus- illustrate the ambiguity. Suppose I said the following: "Miss Smith, if any student registered for this course has not taken a course in linear algebra, then he has taken a course in analysis." "Mr. Jones, if you get a grade below 70 on the final, you are going to flunk this course." In the context, Miss Smith understands that if a student in the course has not had linear algebra, then he has taken analysis, but if he has had linear algebra, he may or may not have taken analysis as well. And Mr. Jones knows that if he gets a grade below 70, he will flunk the course, but if he gets a grade of at least 70, he will pass. Again, mathematics cannot tolerate ambiguity, so a choice of meanings must be made. Mathematicians have agreed always to use "if. . . then" in the first sense, so that a statement of the form "If P, then Q" means that if P is true, Q is true also, but if P is false, Q may be either true or false. As an example, consider the following statement about real numbers: Ifx > 0, then x3 ^ 0. It is a statement of the form, "UP, then Q," where P is the phrase ux > 0" icalled the hypothesis of the statement) and Q is the phrase "x3 ^ 0" (called me conclusion of the statement). This is a true statement, for in every case for ™ch the hypothesis x > 0 holds, the conclusion x3 ^ 0 holds as well. Another true statement about real numbers is the following: Ifx2 <0, then x = 23; ^uT^tT0 f°r Whi°h thC hypothesis holds>the conclusion holds as well. Of esjs hold PCnS'" th'S example that there are no cases for which the hypoth- To t Statcment of this sort is sometimes said to be vacuously true. 0 c a\Trn,now t0 the emPty set and inclusion, we see that the inclusion * € 0 th" f°» CVCry SCt A' Writin8 0 c A is the same as saying, "If • en*e A," and this statement is vacuously true.
8 Set Theory and Logic Contrapositive and Converse Our discussion of the "if... then" construction leads us to consider another point of elementary logic that sometimes causes difficulty. It concerns the relation between a statement, its contrapositive, and its converse. Given a statement of the form "If P, then Q," its contrapositive is defined to be the statement "If Q is not true, then P is not true." For example, the contrapositive of the statement Ifx > 0, then x1 ^ 0, is the statement Ifx3 = 0, then it is not true that x > 0. Note that both the statement and its contrapositive are true. Similarly, the statement Ifx1 < 0, then x = 23, has as its contrapositive the statement If x^ 23, then it is not true that x2 < 0. Again, both are true statements about real numbers. These examples may make you suspect that there is some relation between a statement and its contrapositive. And indeed there is; they are two ways of saying precisely the same thing. Each is true if ind only if the other is true; they are logically equivalent. This fact is not hard to demonstrate. Let us introduce some notation first. As a shorthand for the statement "If P, then Q,'' we write P=>Q, which is read "P implies Q." The contrapositive can then be expressed in the form (not Q) =* (not P), where "not Q" stands for the phrase "Q is not true." Now the only way in which the statement "P => Q" can fail to be correct is if the hypothesis P is true and the conclusion Q is false. Otherwise it is correct. Similarly, the only way in which the statement (not Q) => (not P) can fail to be correct is if the hypothesis "not Q" is true and the conclusion not P" is false. This is the same as saying that Q is false and P is true. And ttns, in turn, is precisely the situation in which P => Q fails to be correct Thus we see that the two statements are either both correct or both incorrect; mey are logically equivalent. Therefore, we shall accept a proof of the state- statement not Q => not />» as a proof of the statement "i> => Q." Tt i«\k !* anOth°r statement that ca" be formed from the statement P => Q- U is the statement
Fundamental Concepts ■ h • called the converse of P => Q. One must be careful to distinguish W ' a statement's converse and its contrapositive. Whereas a statement betWftTcontrapositive are logically equivalent, the truth of a statement says a"tiiin at all about the truth or falsity of its converse. For example, the true nt statement > has as its converse the statement Ifx3 ^ 0, then x > 0, which is false. If it should happen that both the statement P => Q and its converse n => P are true, we express this fact by the notation P<=>Q, which is read "P holds if and only if Q holds." Negation If one wishes to form the contrapositive of the statement P => Q, one has to know how to form the statement "not P," which is called the negation of P. In many cases, this causes no difficulty; but sometimes confusion occurs with statements involving the phrases "for every" and "for at least one." These phrases are called logical quantifiers. To illustrate, suppose that X is a set, A is a subset of X, and P is a state- statement about the general element of X. Consider the following statement: (*) For every x e A, statement P holds. How does one form the negation of this statement? Let us translate the prob- problem into the language of sets. Suppose that we let B denote the set of all those elements x of X for which P holds. Then statement (*) is just the state- statement that A is a subset of B. What is its negation? Obviously, the statement that A is not a subset of B; that is, the statement that there exists at least one dement of A that does not belong to B. Translating back into ordinary lan- language, this becomes For at least one x e A, statement P does not hold. ,lfoerefore'wt0 form the negation of statement (*), one replaces the quantifier or every" by the quantifier "for at least one," and one replaces statement r ^ Us negation. e process works in reverse just as well; the negation of the statement . For at least one x e A, statement Q holds, ls the statement For every x e A, statement Q does not hold.
Chap, i 10 Set Theory and Logic The Difference of Two Sets w. «>turn now to our discussion of sets. There is one other operation on t TtaHoccasionally useful. It is the difference of two sets, denoted by J^Tand defined as the set consisting of those elements of A that are not in B. Formally, A _ b = [x\x e A and x $ B}. It is sometimes called the complement of B relative to A, or the complement of B in A. . „ . _. , Our three set operations are represented schematically in Figure 1. AVB A ~B Rules of Set Theory Given several sets, one may form new sets by applying the set-theoretic operations to them. As in algebra, one uses parentheses to indicate in what order the operations are to be performed. For example, A U (B n C) denotes the union of the two sets A and B n C, while (A U B) n C denotes the intersection of the two sets A\J B and C. The sets thus formed are quite different, as Figure 2 shows. Figure 2 *U(BnC) (AUB)nC combinations of operations lead to the same set;
Fundamental Concepts 11 Figure 3 , The equation is illustrated in Figure 3; the shaded region represents Tesetin question, as you can check mentally. This equation can be thought fas a "distributive law" for the operations n and U. Other examples of set-theoretic rules include the second "distributive law," AU(B nC) =(Aij B) n(Au C), and DeMorgarCs laws, A-{B\jC)={A- B)n{A- C), A-(BnC) = (A- B)U(A- C). We leave it to you to check these rules. One can state other rules of set theory, buttheseare the most important ones. DeMorgarfs laws are easier to remem- remember if you verbalize them as follows: The complement of the union equals the intersection of the complements. The complement of the intersection equals the union of the complements. Collections of Sets The objects belonging to a set may be of any sort. One can consider the setof all even integers, and the set of all blue-eyed people in Nebraska, and 4e set of all decks of playing cards in the world. Some of these are of limited mathematical interest, we admit! But the third example illustrates a point we Mvenotyet mentioned: namely, that the objects belonging to a set may themselves be sets. For a deck of cards is itself a set, one consisting of pieces i Pasteboard with certain standard designs printed on them. The set of all -f ^ of cards in the world is thus a set whose elements are themselves sets W Pieces of pasteboard). A e n°w have another way to form new sets from old ones. Given a set ^nskk^ °Onsider sets whose elements are subsets of A. In particular, we can symboT ^ SCt °f al1 subsets of A- This set is sometimes denoted by the latert° ^ and 's called the power set of A (for reasons to be explained ^ W° have a Set whose elements are sets, we shall often refer to it as a n of ^ts and denote it by a script letter such as a or <B. This device
12 Set Theory and Logic aP-1 ill helo us in keeping things straight in arguments where we have to consider ^Irts and sets of objects, and collections of sets of objects, all at the same f£fFor example, we might use a to denote the collection of all decks of cards in the world, letting an ordinary capital letter A denote a deck of cards and a lower-case letter a denote a single playing card. A certain amount of care with notation is needed at this point. We make a distinction between the object a, which is an element of a set A, and the one- element set {a}, which is a subset of A. To illustrate, if A is the set [a, b, c], then the statements a e A, [a}<= A, and {a} e 6>(A) are all correct, but the statements [a] e A and a <= A are not. Arbitrary Unions and Intersections We have already defined what we mean by the union and the intersection of two sets. There is no reason to limit ourselves to just two sets, for we can just as well form the union and intersection of arbitrarily many sets. Given a collection a of sets, the union of the elements of a is defined by the equation {jAea A = {x\x e A for at least one A e a}. The intersection of the elements of a is defined by the equation fW A ={x\x e A for every A e a}. There is no problem with these definitions if one of the elements of G happens to be the empty set. But it is a bit tricky to decide what (if anything) these definitions mean if we allow a to be the empty collection. Applying the definitions literally, we see that no element x satisfies the defining property for the union of the elements of a. So it is reasonable to say that Daeo. A = 0 if a is empty. On the other hand, every x satisfies (vacuously) the defining property for the intersection of the elements of a. The question is, every x in what set? If one has a given large set X that is specified at the outset of the discussion to be one's "universe of discourse," and one considers only subsets of X throughout, it is reasonable to let h*~Pty- Not a" nafowiMlclans follow this convention, however. dlffic""ty. we shall not define the intersection when a is empty. Cartesian Products theStTonoff «n°!,her WEy °f f°rmin« new sets fr°m old ones; it involves notion of an ordered pair" of objects. When you studied analytic geome-
Fundamental Concepts 13 §!-» thing you did was to convince yourself that after one has chosen try,thenrs^ axjs jn the p]ane, every point in the plane can be made to an x-axis a ^ unique ordered pair (x, y) of real numbers. (In a more sophis- po" { of geometry, the plane is more likely to be defined as the ^rordered pairs of real numbers!) h notion of ordered pair carries over to general sets. Given sets A and d fine their cartesian product A x B to be the set of all ordered pairs * AI for which a is an element of A and b is an element of B. Formally, A X B = {(a, b)\a e A and b e B}. This definition assumes that the concept of "ordered pair" is already given. It can be taken as a primitive concept, as was the notion of "set"; or it can be given a definition in terms of the set operations already introduced. One defini- definition in terms of set operations is expressed by the equation {a,b) = [{a},{a,b}}; it defines the ordered pair {a, b) as a collection of sets. If a =£ b, this definition says that (a, b) is a collection containing two sets, one of which is a one-element set and the other a two-element set. The first coordinate of the ordered pair is defined to be the element belonging to both sets, and the second coordinate is the element belonging to only one of the sets. If a = b, then (a, b) is a collec- collection containing only one set [a], since {a, b] = [a, a) = {a} in this case. Its first coordinate and second coordinate both equal the element in this single set. I think it is fair to say that most mathematicians think of an ordered pair as a primitive concept rather than thinking of it as a collection of sets! Let us make a comment on notation. It is an unfortunate fact that the notation (a, b) is firmly established in mathematics with two entirely different meanings. One meaning, as an ordered pair of objects, we have just discussed. The other meaning is the one you are familiar with from analysis; if a and b are real numbers, the symbol (a, b) is used to denote the interval consisting of aN numbers* such that a < x < b. Most of the time this conflict in notation * cause no difficulty, because the meaning will be clear from the context, enever a situation occurs where confusion is possible, we shall adopt a 1 went notation for the ordered pair (a, b), denoting it by the symbol a x b Exercises I. jj* *° d'Stributive Iaws f0r U and n, and DeMorgan's laws. D- Had° wh'^h °f the foIIowinS statements are true for all sets A, B, C, and °ut>Ie implication fails, determine whether one or the other of the pos-
Set Theory and Logic ChaP-1 sible implications holds. If an equality fails, determine whether one or the other of the possible inclusions holds. (a) A => C and B => C <^=> (A U B) => C. {h) A ^ Cot B => C<=>(A u B) ^ C. (c) A => Cand B => C<=> (A n B) => C. (d) A => Cor B => C<=^(A n B) => C. (e) A-(A-B) = B. (f) A-(B-A) = A-B. (g) An(B-C) = (AnB)-(An C). (h) AU(B-C) = (AUB)-(AU C). (i) (/)ni)uM-*) = '<. (j) ^cCand«cfl^(/(xB)c(Cxi)). (k) The converse of (j). A) The converse of (j), assuming that A and B are nonempty. (m)(A x B)U(C x £>) = (/( U C) x E u £>). (n) (/4 x B) n (C x D) = M n C) x E n /)). (o) /(x(J-C) = Mx«)-(/lx C). (p) (/(-S)x(C-fl) = (/)xC-fi x C)-/(xfl. (q) (A xB)-(Cx D) = (A - C) x E - D). 3. (a) Write the contrapositive and converse of the following statement: "If x < 0, then x1 — x > 0," and determine which (if any) of the three state- statements are true, (b) Do the same for the statement "If .v > 0, then .v2 — .v > 0." 4. Let A and B be sets of real numbers. Write the negation of each of the following statements: (a) For every a e A, it is true that a2 e B. (b) For at least one a e A, it is true that a1 e B. (c) For every a e A, it is true that a1 £ 5. (d) For at least one a £ A, it is true that a2 e B. 5. Let a be a nonempty collection of sets. Determine the truth of each of the following statements, and of their converses: (a) x e {JAea A => x e A for at least one A e a. (b) a: e IJ-iea A =>■ x e A for every /4 e ft. (c) x e H^ea ^ => x e /4 for at least one A e a. (d) x e H-iea -4 =>■ x e /f for every A e ft. 6. Write the contrapositive of each of the statements of Exercise 5. 7. Given sets A, B, and C. Express each of the following sets in terms of A, B, and C, using the symbols u, O, and —. D = [x\x e Aand(x e Bor x e C)}, E ~ [x\ (x e A and x e B) or x e C], F = [x\x e A and (x e B =»• x e C)}. 8. If a set /f has two elements, show that 6>(A) has four elements. How many ele- elements does <9(A) have if A has one element? Three elements? No elements? Why is <P(/f) called the power set of A ?
Functions 15 §1-2 te the set of real numbers. For each of the following subsets of '• Ut ^ dUermine whether it is equal to the cartesian product of two subsets ,°.wrv v)\x is an integer}. fd) f(r'y)|v is not an integer and y is an integer}. 1-2 Functions The concept of function is one you have seen many times already, so it is ., necessary to remind you how central it is to all mathematics. Tn this ction we give the precise mathematical definition, and we explore some of the associated concepts. A function is usually thought of as a rule that assigns, to each element of aset/4, an element of a set B. In calculus a function is often given by a simple formula such as f{x) = 3x2 + 2, or perhaps by a more complicated formula such as One often does not even mention the sets A and B explicitly, agreeing to take A to be the set of all real numbers for which the rule makes sense and B to be the set of all real numbers. As one goes further in mathematics, however, one needs to be more pre- precise about what a function is. Mathematicians think of functions in the way we just described, but the definition they use is more exact. First, we define the following: Definition. A rule of assignment is a subset r of the cartesian product c X D of two sets, having the property that each element of C appears as the fost coordinate of at most one ordered pair belonging to r. Thus a subset cofCx D is a rule of assignment if [(c, d)&r and (c, d') e r] ==> [d = d']. for V k °f r aS a Way of ass'gning to the element c of C the element d of D rule of assignment r, the domain of r is defined to be the subset of gOf a11 ^rSt coordinates of elements of r, and the image set of r is ^the subset of D consisting of all second coordinates of elements of domain r = {c\ there exists de D such that (c, d) e r}, image r = {d\ there exists c e C such that (c,d) & r}.
Chap. U Set Theory and Log* B «■ .• „ fh a rule of assignment r, together with a set Definition. A function/.a rule or^^ ^ ^ ^ ^ ^ ^^ ^^ ^ rroT^SnVTlhe image set of Ms also called the image set of/; and the set B is called the range of/.t ,f/is a function having domain A and range B, we express this fact by ting f:A->B, which is read '/is a function from A to B," or "/is a mapping from A into B," or simply '/maps /I into B." One sometimes visualizes/as a geometric trans- transformation physically carrying the points of A to points of B. Iff: A—>B and if a is an element of A, we denote by /(a) the unique element of B which the rule determining/assigns to a; it is called the value of/at a, or sometimes the image of a under/ Formally, if/- is the rule of the function/ then f(a) denotes the unique element of B such that (a, f(a)) e r. Using this notation, one can go back to defining functions almost as one did before, with no lack of rigor. For instance, one can write (letting R denote the real numbers) "Let/be the function whose rule is {(x, x3 + l)|.r e R] and whose range is R," or one can equally well write "Let/: A—> R be the function such that/(.v) = x3 + 1 " Example 1 Let R h negative reals. Consider'the^nc^ns1 """^ ^ let *+ denote the definedby defined by g(x)
Functions 17 i1*2 h: R —»- R+ defined by h(x) k: R + —> R+ defined by k(x) ■ n * is different from the function /, because their rules are different The fa"ction* R. .t ,s ^ restriction of/to the set Jf+. The function A is also subsets or* ^ thQugh thejr ru,es are thc same set; because the range dmTd for h is different from the range specified for/ The function * is differ- fm all of these. These functions are pictured in Figure 4. Figure 4 Let/: A — B.IT Ao is a subset of A, we denote by f(A0) the set of all images of points of Ao under the function/; this set is called the image of Ao under/ Formally, f(A0) ={b\b =/(«) for some a e Ao}. On the other hand, if Bo is a subset of B, we denote by f~l{B0) the set of all elements of A whose images under/lie in Bo; it is called the preimage of Bo under/(or the "counterimage," or the "inverse image," of Bo). Formally, rl(B0)={a\f(a)eB0}. Of course, there may be no points a of A whose images lie in Bo; in that case, /~W is empty. Some care is needed if one is to use the/and /"' notation correctly. The operation /"'.for instance, when applied to subsets of B, behaves very nicely; * Preserves inclusions, unions, intersections, and differences of sets. But the Peration/, when applied to subsets of A, does not preserve all these set 0Perat.ons. See Exercises 2 and 3. /(/-uan°ther examPle. it is not true in general that /"'(/(^o)) = ^o and o)) = Bo, as the following example shows. The relevant rules, which 'e to you to check, are the following: If/: A —> B, then f-'(f(A0)) = Ao {otA0cA, f(f-'{B0)) c Bo foi BocB.
IS Set Theory and Logic Chap, 1 6 T 5 I 4 \ 3 2 1 I I -2 1 y = f(x) J i i 1 2 Figure 5 Example 2. Consider the function f:R —> R given by f(x) = 3x2 + 2 (Figure 5). Let [a, b] denote the closed interval a < x < b. Then /-'(/([0, 1])) = /-i([2, 5]) = [-1, 1], Restricting the domain of a function and changing its range are two ways of forming a new function from an old one. Another way is to form the com- composite of two functions. Definition. Given functions /: A — B and g : B — C, we define the composite y o/of/and g as the function g of: A — C defined by the equa- Formally, * „/: ^ — C is the function whose rule .g {{a, c) | For some b e B, f(a) = A and *F) = c}. e thC COmP°site « ° /as involving a physical movement of the pomt /(«), and then to the point g(f(a)), as illustrated in Note that g of is defined only when ^ fange of/e?I/fl& ^ domain of ^ E the the g: R _ j, given by function
Functions 19 §1-2 f.£--+R given by {gofKxi- *(/(•*» = * C*2 + 2) - 5C*2 + 2). can also be formed in this case; it is the quite different = /Ex) = 3ExJ + 2. g(l(a)) = g(b) = c 9 Definition. A function /: ^ —► B is said to be injective (or one-to-one) if foreach pair of distinct points of A, their images under/are distinct. It is said to be surjecdve (or/is said to map A onto B) if every element of B is the image of some element of A under the function/. If/ is both injective and surjective, it is said to be bijective (or is called a one-to-one correspondence). More formally,/is injective if [f{a)=f{al)]=>[a = al], and/is surjective if [b e B] => [b = f{a) for at least one a e A]. Injectivity of/depends only on the rule of/; surjectivity depends on the range «/as well. You can check that the composite of two injective functions is "njective, and the composite of two surjective functions is surjective; it fol- ows that the composite of two bijective functions is bijective. Iti H b'Jective> there exists a function from B to A called the inverse of/ of \ftedby ■/"' and is defined by letting /"'(&) be that unique element a thatth"^ Wh'Ch ^ = b' Given b e B> the fact that/is surjective implies thatthCre eXhtS SUCh a" element a e A> the fact that/is injective implies i ^. * only °"e such element a. It is easy to see that if/is bijective, /"« The fw" Consider aSain *e functions /, g, h, and k of Figure 4. Its restt' '°n^ * ~* R given by ^^ = xl is neltrier injective nor surjective. tion^10" 8 t0-the n°nnegative reals is injective but not surjective. The infecf 'R~* R* obtained from / by changing the range is surjective but ve. The function k : £+ —► jR+ obtained from / by restricting the
20 Set Theory and Logic Chap, j domain and changing the range is both infective and surjective, so it has at, inverse. Its inverse is, of course, what we usually call the square-root function. A useful criterion for showing that a given function/is bijective is the following, whose proof is left to the exercises: Lemma 2.1. Letf:A~^B. If there are functions g : B -^ A and h : B —. A such that g(f(a)) = a for every a in A and f(h(b)) = b for every b in B, then f is bijective and g = h = f. Note that if/: A — > B is bijective and Bo c B, we have two meanings for the notation f~l(B0). It can be taken to denote the preimage of Bo under the function / or to denote the image of Bo under the function /-' : B —► A. These two meanings give precisely the same subset of A, however, so there is no ambiguity. Exercises 1. Let/: A—>B. Let Ao a A and Bo a B. (a) Show that /-'(/(Ao)) => /40 and that equality holds if/is injective. (b) Show that /(/~l(B0)) a Bo and that equality holds if/is surjective. 2. Let /: /4 — > B and let /4, a A and B, <= B for / = 0 and / = 1. Show that /-' preserves inclusions, unions, intersections, and differences of sets: (a) Bo cz B, =y/-'(B0) c /-'(By). (b) /"'(Bo uB,)=/-i(B0) U/-'(B,). (c) /-'(Bo n B,) =/-'(B0) n /-'(B,). (d) /"'(Bo -B,)=/-'(Bo)-/-■(£,). Show that/preserves inclusions and unions only: (e) /*o<=/<i=>/(/io) <=/(/*,)• @ /(^o U ^i) =/(^0) U/(/4,)- (g) /(^o n /<i) c= /(^0) n /(Ai); give an example where equality fails, (h) /(Ao — A,) => /(Ao) — f(A{); give an example where equality fails. 3. Show that (b), (c), (f), and (g) of Exercise 2 hold for arbitrary unions and intersections. 4. Let/: A —► B and ^: B —► C. (a) If Co c C, show that (*■ <./)-'(C0) =/-'(^-'(C0)). (b) If /and *■ are injective, show that g ° /is injective. (c) If £ o/is injective, what can you say about injectivity of/and g ? (d) If/and ^ are surjective, show that g o/is surjective. (e) If *■ o/is surjective, what can you say about surjectivity of/and g ? (f) Summarize your answers to (b)-(e) in the form of a theorem. 5. In general, let us denote the identity function for a set C by ic. That is, define ic: C —► C to be the function given by the rule ic(x) = x for all x e C. Given
Relations 21 A—*B we say that a function g : B —♦ A is a left inverse for/iff /= iA\ d we say that /;: B — /f is a right inverse for / if /° h = ;B. an sh£)W that if/has a left inverse,/is injective; and if/has a right inverse, /Ms surjective. h\ Give an example of a function that has a left inverse but no right inverse. i Give an example of a function that has a right inverse but no left inverse. A\ Can a function have more than one left inverse? More than one right inverse ? ) Show that if/has both a left inverse g and a right inverse /;, then /is bijective and g = *=/-'. Let/' /?—>./? be the function f(x) = .v3 — x. By restricting the domain and ' range of/appropriately, obtain from/a bijective function g. Draw the graphs of g and g~l- (There are several possible choices for g.) 1-3 Relations A concept that is in some ways more general than that of function is the concept of a relation. In this section we define what mathematicians mean by a relation, and we consider two types of relations that occur with great fre- frequency in mathematics: equivalence relations and simple order relations. Definition. A relation on a set A is a subset C of the cartesian product AxA. If Cisa relation on A, we use the notation xCy to mean the same thing as (x,y) G C. We read it "x is in the relation C to y." A rule of assignment /-for a function/: A —> A is also a subset of A x A. But it is a subset of a very special kind: namely, one such that each element of A appears as the first coordinate of an element of/■ exactly once. Any subset of Ax Aha relation on A. Example 1. Let P denote the set of- all people in the world, and define D <= P x P by the equation D = ((*> y) \ x is a descendant of y}. D is a relation on the set P. The statements "x is in the relation D to /' x is a descendant of y" mean precisely the same thing, namely, that (*>H 6 D. Two other relations on P are the following: B = {(x,y)\x has an ancestor who is also an ancestor of y], s ~ {(*, y) \ the parents of x are the parents of y). Ca" B the "blood relation" (pun intended), and we can call S the rel?tion ' These three relations have quite different properties. The relationship is symmetric, for instance (if x is a blood relative of y,
22 Set Theory and Logic Cha p, then y is a blood relative of x), whereas the descendant relation is not. tye shall consider these relations again shortly. Equivalence Relations and Partitions An equivalence relation on a set A is a relation C on A having the follow- following three properties: A) (Reflexivity) xCx for every x in A. B) (Symmetry) If xCy, then yCx. C) (Transitivity) If xCy and yCz, then xCz. Example 2. Among the relations defined in Example 1, the descendant relation D is neither reflexive nor symmetric, while the blood relation B is not transitive (I am not a blood relation to my wife, although my children are!) The sibling relation S is, however, an equivalence relation, as you may check. There is no reason one must use a capital letter—or indeed a letter of any sort—to denote a relation, even though it is a set. Another symbol will do just as well. One symbol that is frequently used to denote an equivalence relation is the "tilde " symbol ~. Stated in this notation, the properties of an equivalence relation become A) x ~ x for every x in A. B) If x ~ y, then y ~ x. C) If x ~ y and y ~ z, then x ~ z. There are many other symbols that have been devised to stand for particular equivalence relations; we shall meet some of them in the pages of this book. Given an equivalence relation ~ on a set A and an element x of A, we define a certain subset E of A, called the equivalence class determined by x, by the equation Note that the equivalence class E determined by x contains x, since x ~ x. Equivalence classes have the following property: Lemma 3.1. Two equivalence classes E and E' are either disjoint or equal. Proof. Let E be the equivalence class determined by x, and let E' be the equivalence class determined by x'. Suppose that E n E' is not empty; l be a point of E n E'. (See Figure 7.) We show that E = E'. Figure 7
Relations 23 • 1-3 • ■ «,e have v ~ x and y ~ x'. Symmetry allows us to conclude By definition, we^ ^ trans;tivity it follows that x ~ x'. If now w is that x ~ ■''f r Jehave w ~ * by definition; it follows from another applica- any point °'f' thflt >v „ X'. we conclude that £<=£'. o-on of transm j< sl-tuatl-on allows us to conclude that E' c £ as well, The symmetry ol LUC ° sothat^^'' □ ■ an equivalence relation on a set A, let us denote by S the collection „ ^equivalence classes determined by this relation. The preceding lem- h vs that distinct elements of S are disjoint. Furthermore, the union of lements of S equals all of A, because every element of A belongs to an 'he C ^ence class. The collection S is a particular example of what is called a partition of A: Definition. A partition of a set A is a collection of disjoint subsets of A whose union is all of A. Studying equivalence relations on a set A and studying partitions of A are really the same thing. Given any partition 3D of A, there is exactly one equiv- equivalence relation on A from which it is derived. The proof is not difficult. To show that the partition 3D comes from some equivalence relation, let us define a relation C on A by setting xCy if x and y belong to the same element of 3D. Symmetry of C is obvious; reflexivity fol- follows from the fact that the union of the elements of 3D equals all of A; transi- transitivity follows from the fact that distinct elements of SD are disjoint. It is simple to check that the collection of equivalence classes determined by C is precisely the collection 3D. To show there is only one such equivalence relation, suppose that C, and C2are two equivalence relations on A that give rise to the same collection of equivalence classes 3D. Given x e A, we show that yC,x if and only if yC2x, trom which we conclude that C, = C2. Let E, be the equivalence class deter- determined by x relative to the relation C,; let E2 be the equivalence class deter- determined by x relative to the relation C2. Then Et is an element of 3D, so that it relation C2. Then Et is an element of 3D, so that it equalTw ^ U"'qUC element D of D that contains x. Similarly, E1 must sL f \i °W by definii « oi all y such that yC^x sinc£ ^ = D = Tw ^ U"'qUC element D of D that contains x. Similarly, E1 f \i °W by definition> £1 consists of all y such that yC,x; and E oi all y h th 2 con- the saAMPLE 3' Defin° tW° points in the plane t0 ^ equivalent if they lie at hold trT- f'Stance from the or'gin- Reflexivity, symmetry, and transitivity v'aily. The collection S of equivalence classes consists of all circles at the origin, along with the set consisting of the origin alone. same^ Define two P°"nts of the plane to be equivalent if they have the straight Hi**-?' The c0llectI0n of equivalence classes is the collection of all "nes m the plane parallel to the *-axis.
Chap I 24 Set Theory and Logic Fxample 5 Let £ be the collection of all straight lines in the plane parallel m the line v = -x Then £ is a partition of the plane, since every point lies on som such line and each pair of distinct lines are disjoint. The partition £ comes from the equivalence relation on the plane that declares the po.nts (x,,yt) and (x2,yi) to be equivalent if .v, + y, = x2 + yi- Example 6 Let £' be the collection of all straight lines in the plane. Then £' is not a partition of the plane, for distinct elements of £' are not necessarily disjoint; two lines may intersect without being equal. Order Relations A relation C on a set A is called an order relation (or a simple order, or a linear order) if it has the following properties: A) (Comparability) For every x and y in A for which .v ^ y, either xCy or yCx. B) (Nonreflexivity) For no x in A does the relation .vC.v hold. C) (Transitivity) If xCy and yCi, then xC:. Note that property A) does not by itself exclude the possibility that for some pair of elements x and y of A, both the relations xCy and yCx hold (since "or" means "one or the other, or both"). But properties B) and C) combined do exclude this possibility; for if both xCy and yCx held, transitivity would imply that xCx, contradicting nonreflexivity. Example 7. Consider the relation on the real line consisting of all pairs (x, y) of real numbers such that .v < y. It is an order relation, called the "usual order relation," on the real line. A less familiar order relation on the real line is the following: Define xCy if x1 < y2, or if x1 = y1 and x < y. You can check that this is an order relation. Example 8. Consider again the relationships among people given in Exam- Example 1. The blood relation B satisfies none of the properties of an order relation, and the relation S satisfies only C). The descendant relation D does some- somewhat better, for it satisfies both B) and C); but comparability still fails. Rela- Relations that satisfy B) and C) occur often enough in mathematics to be given a. ttemlaterm(SeeT§ieyifre ^"^ sMcl par"aI order relations; we shall consider "lestfhl^T"tildC^ 7' iS thC gCneric symbo1 for an equivalence relation, the in thi notaSf ?h <> '* COmmonly used to denote an order relation. Stated th, notation, the propen.es of an order relation become 1 \][x*rth™^ 1 \][r>™ B) If x < y, then x ^ y. C) If x < y and y < z, then x < z.
Relations 25 II se the notation x < y to stand for the statement "either x < y or ^C "■ nd we shall use the notation y > x to stand for the statement * ^y '" We write x < y < z to mean "x < y and j> < z." . .- _ if X is a set and < is an order relation on X, and if a < b, we use *e notation (a, b) to denote the set [x\a < x < b); lied an open interval in X. If this set is empty, we call a the immediate 11 'd cessor of b, and we call b the immediate successor of a. Definition. Suppose that A and B are two sets with order relations <A nd <«» respectively. We say that A and B have the same order type if there is a bijective correspondence between them that preserves order; that is, if there exists a bijective function/: A —* B such that ai<Ad2 ==> f(a,) <Bf(a2). Example 9. The interval ( — 1, 1) of real numbers has the same order type as the set R of real numbers itself, for the function /: ( —1, 1) —► R given by Ax) = 1 - x1 is an order-preserving bijective correspondence, as you can check. It is pictured in Figure 8. y = x/A-x2) FigureS
Chap.j 26 Set Theory and Logic Example 10. The subset A - {0} U A. 2) of it has the Same order type „ the subset [0, 1) = W»^ * < ^ of*.ThefuncL-on/:^[0,l)definedby /@) = 0, /(jr) = jr-l for x 6 A,2) is the required order-preserving correspondence. One interesting way of defining an order relation, which will be useful to us later in dealing with some examples, is the following: Definition. Suppose that A and B are two sets with order relations <4 and <B, respectively. Define an order relation < on A x B by defining • a, X &i < a2 X b2 if a, <A a2, or if a, = a2 and 6, <B b2. It is called the dictionary order relation on A X B. Checking that this is an order relation involves looking at several separate cases; we leave it to you. The reason for the choice of terminology is fairly evident. The rule defin- defining < is the same as the rule used to order the words in the dictionary. Given two words, one compares their first letters and orders the words according to the order in which their first letters appear in the alphabet. If the first letters are the same, one compares their second letters and orders accordingly. And so on. Example 11. Consider the dictionary order on the plane R X R. In this order, the point p is less than every point lying above it on the vertical line through p, and p is less than every point to the right of this vertical line. Example 12. Consider the set [0,1) of real numbers and the set Z+ of positive integers, both in their usual orders; give Zt x [0, 1) the dictionary function S6t HaS thC ^^ °rder tyPC 3S thE SEt °f nonneSative reals; the f(nxt)=n + t~l fotUntfv? *T:t-iVe order-Preservi"g correspondence. On the other hand, for examni X '" dict'onary order has quite a different order type; i^^'SSTiifordered set has an immediate successor' irrroruieHS °f thE real numbers tha' you may have seen before need some preliminary definiUons.
Relations 27 2+ X [0,1) [0,1) X Figure 9 Suppose that /f is a set ordered by the relation <. Let Ao be a subset of A We say that the element b is the largest element of Ao if b e Ao and if x< b for every x e /40. Similarly, we say that a is the smallest element of Aja e Ao and if a< x for every x e /<„. It is easy to see that a set has at most one largest element and at most one smallest element. We say that the subset Ao of A is bounded above if there is an element b dA such that x < b for every x e Ao; the element b is called an upper bound IotAo. If the set of all upper bounds for An has a smallest element, that ele- element is called the least upper bound of Av. It is denoted by lub At); it may or may not belong to An. If it does, it is the largest element of Av. Similarly, Ao is bounded below if there is an element a of A such that a < x for every xe Ao; the element tf is called a lower bound for Aa. If the set of all lower bounds for/40 has a largest element, that element is called the greatest lower bound of Aa. It is denoted by gib Aa; it may or may not belong to Ao. K it does, it is the smallest element of Aa. Now we can define the least upper bound property. definition. An ordered set A is said to have the least upper bound property ^every nonempty subset Aa of A that is bounded above has a least upper e nal°g°usly, the set A is said to have the greatest lower bound property if boundn\vemPty SUbSEt A° °^A that is bounded below has a greatest lower pro ' .e leave 't to the exercises to show that A has the least upper bound y lf and only if it has the greatest lower bound property. orderXA"PLE J.3- Consider the set A = (-1,1) of real numbers in the usual Propert •Um8 the faCt that the real numbers have the least "PP151" bound y. it follows that this set has the least upper bound property. For, given
Set Theory and Logic Chap, j any subset of A having an upper bound in A, it follows that its least uPPer bound (in the real numbers) must be in A. For example, the subset {-l/2n|n 6 Z+] of /f, though it has no largest element, does havea least up. per bound in A, the number 0. On the other hand, the set B = (-1, 0) u @, 1) does not have the least upper bound property. The subset {-l/2n\it e Z+], for instance, is bounded above by any element of @, 1); but it has no least upper bound in B. Exercises Equivalence Relations 1. Define two points (.t0, >"o)and(A-|, >>,)of the plane to be equivalent if.y0 - xl = yt — Xi. Check that this is an equivalence relation and describe the equiva- equivalence classes. 2. Let C be a relation on a set A. If Ao <= A, define the restriction of C to Ao to be the relation C n (Ao x Ao). Show that the restriction of an equivalence relation is an equivalence relation. 3. Here is a "proof" that every relation C that is both symmetric and transitive is also reflexive: "Since C is symmetric, aCb implies bCa. Since C is transitive, aCb and bCa together imply aCa, as desired." Find the flaw in this argument. 4. Let/: A —► B be a surjective function. Let us define a relation on A by setting ao ~ cti if (a) Show that this is an equivalence relation. (b) Let A* be the set of equivalence classes. Show there is a bijective corre- correspondence of A* with B. 5. Let 51 and 5" be the following subsets of the plane: S = {(x,y)\y = x + 1 and 0 < x < 2}, •S' = {(x, y)\y - x is an integer}. (a) Show that 5" is an equivalence relation on the real line and 5" => 51. Describe the equivalence classes of 5". (b) Show that given any collection of equivalence relations on a set A, their intersection is an equivalence relation on A. (c) Describe the equivalence relation Ton the real line that is the intersection of all equivalence relations on the real line that contain 51. Describe the equivalence classes of T. Order Relations 6. Define a relation on the plane by setting (xo,yo) <(*!,>>,) if either y0 - xl < yt - x\, or y0 - xl = >>, - x] and x0 < *,. Show that ■ this js an order relation on the plane and describe it geometrically.
The Integers and the Real Numbers 29 §1-4 that the restriction of an order relation is an order relation. 7# Check that the relation defined in Example 7 is an order relation. Check that the dictionary order is an order relation. 9 i i Show that the map /: (-1, 0 —► R of Example 9 is order-preserving. "•(* |J|0W that the equation g(y) = lylU + A + 4>>2I2] defines a function (W ./{_(-1, 1) that is both a left and a right inverse for /. Show that an element in an ordered set has at most one immediate successor d at most one immediate predecessor. Show that a subset of an ordered set has at most one smallest element and at most one largest element. 2. LetZ+ denote the set of positive integers. Consider the following order relations Z (i) The dictionary order, (ii) (-vo.^o) < (•*!• y^ if either x0 — y0 < x, - yx, or .v0 — y0 = x, — yt and>>o < yt- (iii) (x0, yo) < (xi, yi) if either xo + >'o <xi + yu or x0 + y0 = .v, + yv and>-o <^i- In these order relations, which elements have immediate predecessors ? Does the set have a smallest element? Show that all three order types are different. 13. Prove the following: Theorem. If an ordered set A has the least upper bound property, then it has the greatest lower bound property. 14. If C is a relation on a set A, define a new relation D on A by letting (b, a) e D \({a,b)eC. (a) Show that C is symmetric if and only if C = D. (b) Show that if C is an order relation, D is also an order relation. (c) Prove the converse of the theorem in Exercise 13. 15. Assume that the real line has the least upper bound property, (a) Show that the sets [0, 1]=}> [0, l) = {x\0<x < 1} have the least upper bound property. ) Does [0, l] x [0,1] in the dictionary order have the least upper bound property? What about [0, 1] x [0, 1)? What about [0, 1) x [0, 1]? 1-4 The Integers and the Real Numbers fo>">dathtSr WE have been discussing what might be called the logical Now we t ourstudy of topology—the elementary concepts of set theory •""• What WC m'ght CaU the mathemati^l foundations for our and the real number system. We have already used them
Set Theory and Logic Chap j in an informal way in the examples and exercises of the preceding sections Now we wish to deal with them more formally. One way of establishing these foundations is to construct the real number system, using only the axioms of set theory—to build them with one's bare hands, so to speak. This way of approaching the subject takes a good deal of time and effort and is of greater logical than mathematical interest. A second way is simply to assume a set of axioms for the real numbers and work from these axioms. In the present section we shall sketch this ap- approach to the real numbers. Specifically, we shall give a set of axioms for the real numbers and shall indicate how the familiar properties of real numbers and the integers are derived from them. But we shall leave most of the proofs to the exercises. If you have seen all this before, our description should refresh your memory. If not, you may want to work through the exercises in detail in order to make sure of your knowledge of the mathematical founda- foundations. First we need a definition from set theory. Definition- A binary operation on a set A is a function / mapping A x A into A. When dealing with a binary operation/on a set A, we usually use a nota- notation different from the standard functional notation introduced in §1-2. Instead of denoting the value of the function/at the point (a, a') by f(a, a), we usually write the symbol for the function between the two coordinates of the point in question, writing the value of the function at (a, a') as afd. Furthermore (just as was the case with relations), it is more common to use some symbol other than a letter to denote an operation. Symbols often used are the plus symbol +, the multiplication symbols • and o, and the asterisk *; but there are many others. Assumption We assume there exists a set R, called the set of real numbers, two binary operations + and • on R, called the addition and multiplication operations, respectively, and an order relation < on R, such that the following properties hold: Algebraic Properties A) (x + y) + 2 = x + (y + 2), (x-y)-z = x-(yz) for all x, y, z in R. B) x + y=y + x, x-y = y-x for all x, y in R. C) There exists a unique element of R called zero, denoted by 0, such that x + 0 = xfora.l\x e R.
The Integers and the Real Numbers 31 §1-4 There exists a unique element of R called one, different from 0 and denoted by 1, such that x-\ = x for all * e *. For each x in R, there exists a unique j» in R such that * + y = 0. Foreach * in R different from 0, there exists a unique y in R such that E) + z) = (*..y) + (*-z) for all x, y, z e R. A Mixed Algebraic and Order Property F) Ifx>J'. then* + z> y + z. If .v > y and z > 0, then x-z > .y-z. Order Properties G) The order relation < has the least upper bound property. (8) If x < y, there exists an element z such that x < z and z < y. From properties (l)-E) follow the familiar "laws of algebra." Given x, one denotes by —x that number y such that x -|- y = 0; it is called the nega- negate of x. One defines the subtraction operation by the formula z — x = 2+ (—x). Similarly, given x ^t 0, one denotes by l/.v that number;* such that x-y=\\ it is called the reciprocal of x. One defines the quotient z/x by the formula z/x = z-(l/;c). The usual laws of signs, and the rules for adding and multiplying fractions, follow as theorems. These laws of algebra are listed in Exercise 1 at the end of the section. We often denote x-y simply by xy. When one adjoins property F) to properties (l)-E), one can prove the usual "laws of inequalities," such as the following: If x > y and z < 0, then x-z < yz. — 1 < 0 and 0 < 1. The laws of inequalities are listed in Exercise 2. We define a number x to be positive if x > 0, and to be negative if x < 0. We denote the positive reals by R+ and the nonnegative reals (for reasons to M explained later) by^+. Properties (l)-F) are familiar properties in modern algebra. Any set with fi ,°.?!nary operations satisfying (l)-E) is called by algebraists a field; if the (Uias an order relation satisfying F), it is called an ordered field. rert ^ and ^' on the other hand> are familiar properties in topo- mvolve only the order relation; any set with an order relation satisf • N8 y 'S Called by toPologists a linear continuum. h Now8- ^ and ^ 'S Called by toPologists a linear continuum. [pro ll haPPens that when one adjoins to the axioms for an ordered field the result" r~^ the axioms for a linear continuum [properties G) and (8)], ** Proved"8 contains some redundancies. Property (8), in particular, can z ■= (x + wf conse(luence of the others; given x < y, one can show that W/(l + 1) satisfies the requirements of (8). Nevertheless, we in-
Set Theory and Logic Chap. 1 elude (8) in our list of basic properties of the real numbers to emphasize the fact that it and the least upper bound property are the two crucial properties of the order relation for R. From these two properties many of the topofo- gical properties of R may be derived, as we shall see in Chapter 3. Now there is nothing in this list as it stands to tell us what an integer is. We now define the integers, using only properties (l)-F). First, we make the following definition: A subset A of the real numbers is said to be inductive if for every x in A, the number x + 1 is also in A. Definition. Let a be the collection of all inductive subsets of R that contain 1. Then the set Z+ of positive integers is defined by the equation Z+ = I lAea A- Note that the set R+ of positive real numbers contains 1 and is inductive (if x > 0, then x + 1 > 0), so that R+ belongs to a. Therefore, Z+ c R+> $o the elements of Z+ are indeed positive, as the choice of terminology suggests. Indeed, one sees readily that 1 is the smaliest element of Z+, because the set of all real numbers x for which x > 1 is inductive and contains 1. The basic properties of Z+, which follow readily from the definition, are the following: A) 1 £ Z+. B) Z+ is inductive. C) (Principle of induction). If Zo is an inductive set of positive integers that contains 1, then Zo = Z+. We define the set Z of integers to be the set consisting of the positive integers Z+, the number 0, and the negatives of the elements of Z4. One proves that the sum, difference, and product of two integers are integers, but the quo- quotient is not necessarily an integer. The set Q of quotients of integers is called the set of rational numbers. One proves also that, given the integer n, there is no integer a such that n < a <n + 1. All these are familiar facts. One property of the positive integers that may not be quite so familiar, but will be very useful, is the following: Theorem 4.1 (Well-ordering property). Every nonempty Subset ofZ+ has a smallest element. Proof. If n e Z+, we use {1,. .. ,«) to denote the set [x\x e Z+ and 1 ^ x <,«). Because there is no integer a between n and n + 1, H «+l} = [l «)U{n+l). We first prove "by induction" that, for each n e Z+> the following state- statement holds: Every nonempty subset of [I «} has a smallest element. Let Zo be the set of all positive integers n for which this statement holds.
The Integers and the Real Numbers 33 ontains 1, since if n = 1, the only nonempty subset of {1 n] 1 fil itself. Then, supposing Zo contains n, we show that it contains is the set {U^beanonempty subset of the set {1>- . .,« + 1). If C consists njr 1. So e ^ment w _|_ j^ then that element is the smallest element of C. of the smg ^ons.der thg set c n [1,. . . , «), which is nonempty. Because OthT'this set has a smallest element, which will automatically be the small- " "' of C also. Thus Zo is inductive and contains 1, so we conclude esteem^ ^ whence the statement is true for all n e Z+. ^Now we p'rove the theorem. Suppose that D is a nonempty subset of Z+. ^ an element n of D. Then the set A = £> n {1 , n} is nonempty, h M has a smallest element A:. The element k is automatically the smallest dement of 2> as well. D Everything we have done up to now has used only the axioms for an ordered field, properties (l)-F) of the real numbers. At what point do you need G), the least upper bound axiom? For one thing, you need the least upper bound axiom to prove that the setZ+ of positive integers has no upper bound in R. This is the Archimedean ordering property of the real line. To prove it, we assume thatZ+ has an upper bound and derive a contradiction. If Z¥ has an upper bound, it has a least upper bound b. There exists seZ, such that n > b — 1; for otherwise, ft- 1 would be an upper bound for Z+ smaller than b. Then n + 1 > b, contrary to the fact that b is an upper bound for Z+. The least upper bound axiom is also used to prove a number of other things about R. It is used for instance to prove the existence of a unique positive square root v^for every positive real number. This fact, in turn, can be used to demonstrate the existence of real numbers that are not rational numbers; the number *J~2 is an easy example. We use the symbol 2 to denote 1 + 1, the symbol 3 to denote 2 + 1, and so on through the standard symbols for the positive integers. It is a fact that s procedure assigns to each positive integer a unique symbol, but we never need thls fact and shall not prove it. afew°?hS °f thESe properties of the integers and real numbers, along with follow r°PertieS WC ShaU "eed later' are outlined ln the exercises that Exercises ' (i°If !hlf°ll0Wlng "laws of algebra" for R, using only axioms (l)-E): (b) 0 _ = *' then y = °' (c) 11 ~ °"lHint: Compute (x + 0).x.]
Set Theory and Logic Chap, (e) x(-y) = -(xy) = (~x)y- . (f) (-1)*= -*• (g) x(y - r) = *v - Arz- (i) If x *= 0 and X-.V = x, then >> = 1. (j) .r/.vr = l ifx*=O. (k) x/1 = x. A) a: =*= 0 and y =*= 0 => a->> =*= 0. /()if ^0 (m)(/>)(W (n) (x/yKw/z) = (xw)l(yz) if y, z * 0. (o) (a7>0 + (»'/•?) = (•« + K>0/(^) if y> z (p) x^0=*l/.v^0. (q) l/(if/z) = 2/Vf if w, z^O. (r) (xly)l(w/z) = (-vz)/(>-»') if >-, »'. ^ *= °- (s) (av)/>- = «(.v/>-) if >- ^= 0. (t) (-x)ly = -r/(->-) = -(.v/y) if >- ^ 0. 2. Prove the following "laws of inequalities" for R, using axioms (l)-F) along with the results of Exercise 1: (a) x > y and w>z=>x + w>y + z. (b) x > 0 and >>>0=>.r + >>>0 and x-y > 0. (c) x > 0 o -a- < 0. (d) a- > >-o -.V < -y- (e) a: > y and z < 0 => xz < yz. (f) x ^ 0 => x2 > 0, where x2 = x-x. (g) -1 <0< 1. (h) x>> > 0 o .r and y are both positive or both negative. (i) x> 0=> \/x> 0. (j) x > y > 0 => I/a- < l/>>. (k) a- < y => a- < (a- + >-)/2 < >-. 3. (a) Show that if a is a collection of inductive sets, then the intersection of the elements of a is an inductive set. (b) Prove the basic properties A), B), C) of Z+. 4. (a) Show that tfeZ+=><7-leZ+u {0}. [Hint: Let X = \x\x e R and ■* — 1 £ -Z+ u {0}j; show that Xis inductive and contains 1.] (b) Show that if n e Z, there is no a e Z such that n < a < n + 1. [#'>rf; Prove it first for « e Z+.] 5. (a) Prove by induction that given n e Z+> every nonempty subset of {1,.. ■.") has a largest element. (b) Explain why you cannot conclude from (a) that every nonempty subset of Z+ has a largest element. 6. Prove the following properties of Z and Z+: (a) a,b<=Z+=>a + be Z+. [Hint: Show that given a e Z+, the set X = [x\x 6 /? and a + a- e Z+) is inductive and contains 1J (b) a,b£Z+=>a-beZ+. (c) c,deZ=>c + deZ. [Hint: Prove it first for d = 1J
The Integers and the Real Numbers 35 - deZ. , uta6/?. Define inductively '" a1 = a, a"*' = tf"-tf Z (See §1-7 for a discussion of the process of inductive definition.) Show that for », m e Z+ and a. b e R, a"o» = a"*m, (a")m = a"m, ambm = (ab)m. These are called the laws of exponents. [Hint: For fixed n, prove the formulas by induction on «.] 8 Letoe /? and a^ 0. Define fl° = 1, and for // e Z., a"" = I/a". Show that the laws of exponents hold for a, b =£ 0 and //, m e Z. 9. Assume the least upper bound axiom. (a) Show that glb{l//i|« eZ4) =0. (b) Show that given a with 0 < a < 1, gib [a"\u e Z-,} = 0. [i//7i/: Let /j = A - a)la, and show that a" < 1/A + «/;).] 10. Assume the least upper bound axiom. (a) Show that every nonempty subset of Z that is bounded above has a largest element. (b) If x <£ Z, show there is exactly one n e Z such that n < x < n + 1. (c) If x — y > 1, show there is at least one « e Z such that y < n < x. (d) If y < x, show there is a rational number z such that y < z < x. 11. Assuming the least upper bound axiom, show that every positive number a has exactly one positive square root, as follows: (a) Show that if x > 0 and 0 < I, < 1, then (x + hY < x1 + hBx + 1), (x - hY ^ x2 - hBx). (b) Let x > o. Show that if x1 < a, then (x + hI < a for some h > 0; and ? > a> then <-x ~h)*> a for some h > 0. Given o > o, let B be the set of all real numbers x such that x2 < a. Show at B is bounded above and contains at least one positive number. Let b = lub B; show that />* = a tHat lf b and ^ afe P°sitive and bl = c2) then * = c- ' We Say that m is even if m/2 e Z) and m is odd otherwise, at ,f m is odd> m = 2n + j for some nsZ [Hint. Choose „ so (b)Sh0 /2<" + L] °w that if p and 9 are odd, so arep-q and/?", for any n e Z+.
Chap, i 36 Set Theory and Logic «. t if a > 0 is rational, then a = mln for some j»,»6Z( where (C> Sh°bothaL and » are even. [Hint: Let n be the smallest element of the set }] (d) Theorem. ^ b irrational. Arbitrary Cartesian Products We have already defined what we mean by the cartesian product A x B f two sets Now we introduce cartesian products of arb.tranly many sets. We consider first some special cases, and then we give the general definition. As usual let {1 . • • , m] denote the set of all positive integers a such that 1 < a<, m.'Given a set X, we define an m-tuple of elements of X to be a function x:{l,... ,m]—> X. If x is an /n-tuple, we often denote the value of x at i by the symbol x, rather than x(i). And we often denote the function x itself by the symbol (x,,...,.rj. We let Xm denote the set of all w-tuples of elements of X. Now we can define the cartesian product. Definition. Suppose that {A,, . . . , Am] is a collection of sets, indexed with the positive integers from 1 to m. Let X = A, U • • • U Am. The carte- cartesian product of this indexed collection of sets, denoted by II" i A, or A, X ••• X Am, is defined to be the set of all m-tuples (x,, . . . , xm) of elements of Zsuch that x, e A, for each ;'. Example 1. We now have two definitions for the symbol A X B. One definition is, of course, the one given earlier, under which A X B denotes the set of all ordered pairs (a, 6) such that a e A and b e B. The second definition, just given, defines/f X B as the set of all functions x:{l, 2] — AuB such that x(l) e A and xB) e B. There is an obvious bijective correspondence between these two sets, under which the ordered pair {a, 6) corresponds to the function x defined by x(l) = a and xB) = b. Since we commonly denote this function X in tuple notation" by the symbol (a, b), the notation itself suggests th correspondence. Thus for the cartesian product of two sets, the general defini tion of cartesian product reduces essentially to the earlier one. Example 2. How does the cartesian product A x B x C differ from the cartesian products A x (B x C) and (A x B) x C? Very little. There are e
Arbitrary Cartesian Products 37 correspondences between these sets, indicated as follows: (a, b, c) <—> (a, F, c)) <—> ((a, b), c). eralize the preceding definition, we proceed by analogy. Given a IT define an (o-tuple of elements of X to be a function set x' x . Z+ —> X; call such a function a sequence, or an infinite sequence, of elements of Weif is an cu-tuple, we often denote the value of x at / by x, rather than x(i), fnd Je denote x itself by the symbol (x,,x2,...) or (xn)nCZ,. We let X" denote the set of all co-tuples of elements of X. Definition. Suppose that {A, /(„,...} is a collection of sets, indexed with the positive integers. Let X be the union of the sets in this collection. The cartesian product of this indexed collection of Sets, denoted by Tliez, A, or A, X A2 X •••, is defined to be the set of all co-tuples (xu x2, . ..) of elements of X such that x, e A, for each i. Nothing in these definitions requires the sets A, to be different from one another. Indeed, they may all equal the same set X. In that case, the cartesian product At x • • • X Am is just the set Xm of all w-tuples of elements of X; and the product A, X A2 x ••• is just the set X" of all co-tuples of elements dX. Example 3. If R is the set of real numbers, then Rm denotes the Set of all m-tuples of real numbers; it is often called euclidean m-space (although Euclid would never recognize it). Analogously, R" is sometimes called "infinite- dimensional euclidean space"; it is the set of all co-tuples (*,, x2, . . .) of real numbers, that is, the set of all functions x : Z+ —> R. j °w we turn to the general definition of cartesian product, which will by " «theSe ^ sPecial cases- First we must make more precise what we mean jan indexed collection of sets," a concept tacitly assumed above. surjecf T"' Leiabea collection of sets. An indexing function for d is a a> togeth .tion-ffrom some set J, called the index set, to a. The collection with the indexing function/, is called an indexed family of sets. c** And we"!i!hn fllement a of J> we shall denote the set /(a) by the symbol A.. na" denote the indexed family itself by the symbol
38 Set Theory and Logic Chap, i which is read "the family of all A,, as a ranges over/." Sometimes we write merely {A.}, if it is clear what the index set is. Note that although an indexing function is required to be surjective, it is not required to be injective. It is entirely possible for A. and Af to be the same set of ft, even though a ^/?. One way in which indexing functions are used is to give a new notation for arbitrary unions and intersections of sets. Suppose that/:/—, a is an indexing function for a; let Aa denote /(a). Then we define U«e/ A* = ix Ifor at least °ne ct e J, x e A.}, and n«e/A"= f* ifor evefy *e J'x 6 A*}~ These are simply new notations for previously defined concepts; one Sees at once (using the surjectivity of the index function) that the first equals the union of all the elements of a and the second equals the intersection of all the elements of a. The main use of indexing functions comes when we define arbitrary cartesian products: Let / be an index set. Given a set X, we define a /-tuple of elements of X to be a function x : / — > X. If a is an element of/, we often denote the value of x at a by x, rather than x(a). And we often denote the function x itself by the symbol which is as close as we can come to a "tuple notation" for an arbitrary index set /. We denote the set of all /-tuples of elements of X by X1. Definition. Let [A.},BJ be an indexed family of sets; let X = |J«E/ A.. The cartesian product of this indexed family, denoted by II«e/ ^«. is defined to be the set of all /-tuples (x.).e, of elements of X such that x, e A. for each a e /. Said differently, this cartesian product is just the set of all functions such that x(a) e A. for each a e /. pI ,??KTly T den°te the prodoct simP'y by n^., and its general element by (*„), ,f the index set is understood all the sets A. are equal to one set X, then the cartesian product LL.,A. is just the set X' of all /-tuples of elements of X. In dealing with the we use'T St°met;mes use "tuP'e notation" for its elements, and sometimes we use funct.onal notation, depending on which is more convenient.
Finite Sets 39 §1-6 Exercises how there is a bijective correspondence of A X B with B X A. 1 (a)°show that if n > 1 there is a bijective correspondence of AxX---xAB with (/f, X ••• X /*„_,) X /(„. Let./ be an index set; write J = K U L, where K and £ are disjoint and non- nonempty- Show there is a bijective correspondence of U^jA" with (U^kA.) X(T[^lA.). Let/ be a (nonempty) index set. Consider two indexed families {A.}.e, and ' <B i CJ, Prove the following: (a)" If X "= A* for a" a e ^ then ^'eJ A* C ^-6/ ^.-t (b) The converse of (a) holds if n^'- is nonempty. (c) oi.^.)n (n«^ B'~> = n.« (^«n #«)• (d) (ii.e/ ^«)u (n»^ 5«)c n.e/ m« u sj. (e) If at least one/(« is empty, thenJI.e/-4. is empty. What can you say if each ,4. is nonempty? (See Exercise 4 of §1-9.) 4. Letm,ne^+- Let Z^ 0. (a) If m < n, find an injective map/: A'm —> X". (b) Find a bijective map g : Xm x X" —> Xm+". (c) Find an injective map h : X" —► Z™. (d) Find a bijective map & : X" x A'™ —> Z™. (e) Find a bijective map I: X" x X™ —> Z™. (f) If A a B, find an injective map m : XA —► Xs. 5. Which of the following subsets of Ra can be expressed as the cartesian product of subsets of Rl (a) {x\x, is an integer for all i}. (b) Wat, > i for all i}. (c) {x\x, is an integer for all i ^ 100}. 1-6 Finite Sets are 'nite sets and infinite sets, countable sets and uncountable sets, these are disT SdS ^at ^°U may have encountered before. Nevertheless, we shall «andSth 'n th'S Section and the next' not on'y t0 make sure y°u under- •hat ...em.thorou«hly, but also to elucidate some particular points of logic ^^'se later on. First we consider finite sets. ^ t*°» brforeSjfakln£.if We aK g'Ven a function x mapping/into \JA'a, we must change its technicality whT"^ .?)ns'dered as a function mapping/into \JA.. We shall ignore this «y when dealing with cartesian products.
Chap., 40 Set Theory and Log* Reca.1 that if, is a positive integer, we use A -} to denote the set {x|xeZ+and I <*<"); it is ca..ed a section of the positive integers. The sets {1 n] are the pr0- lotypes for what we call the finite sets. Definition. A set A is said to be finite if it is empty or if there is a bijective correspondence f:A—>{l,..-,"} for some n e Z+. In the former case, we say that A has 0 elements; in the lat- latter case, we say that A has n elements. For instance, the set {1,...,«} itself has n elements, because it is in bijective correspondence with itself under the identity function. Now note carefully: We have not yet shown that, given the finite set A, the number of elements "which A has" is uniquely determined by A. As far as we know, there might exist bijective correspondences of a given set A with two different sets {1, ...,/(} and [1, . . . , m}. The possibility may seem ridic- ridiculous, for it is like saying that it is possible for two people to count the marbles in a box and come out with two different answers, both correct. Our experience with counting in everyday life suggests that such is impossible, and in fact this is easy to prove when n is a small number such as 1,2, or 3. But a direct proof when n is 5 million would be impossibly demanding. Even empirical demonstration would be difficult for such a large value of n. One might, for instance, construct an experiment by taking a freight car full of marbles and hiring 10 different people to count them independently. If one thinks of the physical problems involved, it seems likely that the counters would not all arrive at the same answer. Of course, the conclusion one could draw is that at least one person made a mistake. But that would mean assuming the correctness of the result one was trying to demonstrate empirically. An alternative explanation could be that there do exist bijective correspondences between the given set of marbles and two different sections of the positive integers. In real life we accept the first explanation. We simply take it on faith that our experience in counting comparatively small sets of objects demonstrates a truth that holds for arbitrarily large sets as well t»V tvTT; m mathematics (« opposed to real life), one does not have to teke this statement on faith. If it is formulated in terms of the existence of it k21TT rCCS mther thSn in terms of the Phy^cal act of counting, h re do not .T?.em?tical Proof- We shall prove shortly that if n * m. ttere do not exist bijecfve functions mapping a given set A onto both the
Finite Sets 41 §1-6 are a number of other "intuitively obvious" facts about finite sets re capable of mathematical proof; we shall prove some of them in this tha' ar6and leave the rest to the exercises. Here is an easy fact to start with: Lemma 6.1- Let nbea positive integer. Let A be a set; let a0 be an element . 77ien there exists a bijective correspondence f of the set A with the set n n+ 1] if and only if there exists a bijective correspondence g of the set A'i{a0}with{l,..-,n}. Proof There are two implications to be proved. Let us first assume that there is a bijective correspondence g: A — {a0} —*{1,...,«}. We then define a function/: A —> [1, . . . , n + 1} by setting f(x) = g(x) for x e A - {a0}, f(ao)=n+ 1. One checks at once that /is bijective. To prove the converse, assume there is a bijective correspondence f:A—>[l,...,n + 1}. If/maps a0 to the number n + 1, things are especially easy; in that case, the restriction f\ A — {a0} is the desired bijective correspondence of A — [a0] with {1,..., «}. Otherwise, let f(a0) = m, and let a, be the point of A such that/(a,) =n + 1. Then at =£ a0. Define a new function h:A—>{{,...,„ + 1} by setting h{a0) =/i + 1, = f{x) for * g /( - {«„} - {«,}. (See Figure 10.) It is easy to check that h is a bijection irrtiith {ery car^thnestriction b' A~ [a°} is the desired ^j^. . '77^ A ^r . a, , . a, . 7 ^^ ^ ( i..... m, . . . , n + 1 J Figure 10
42 Set Theory and Logic From this lemma a number of useful consequences follow: Theorem 6.2. Let A be a set; suppose that th?re exists a bijection f: A -_ {1 , n}for some n e Z+. Let B be a proper subset of A. Then there exists no bijection g : B — {1, .. • , nj; but (provided B ^ 0) there does exist a bijection h : B — [1, .. • , m}/or some m < n. Proof. The case in which B = 0 is trivial, for there cannot exist a bijec- bijection of the empty set B with the nonempty set {1, ...,«}. We prove the theorem "by induction." Let Zo be the subset of Z+ con- consisting of those integers n for which the theorem holds. We shall show that Zo contains the number 1 and is inductive. From this we conclude that Zo =Z+, so the theorem is true for all positive integers n. First we show the theorem is true for n = 1. In this case A consists of a single element [a], and its only proper subset B is the empty set. Now assume that the theorem is true for n; we prove it true for /; -f- 1, Suppose that /: A —* [1, ...,« + 1] is a bijection, and B is a nonempty proper subset of A. Choose a point a0 in B and a point at in A — B. We apply the preceding lemma to conclude there is a bijection g:A-[ao}—>[l,...,n}. Now B — [a0] is a proper subset of A — [a0], for ax belongs to A — [a0] and not to B — {a0}. Because the theorem has been assumed to hold for the integer n, we conclude the following: A) There exists no bijection h : B — [a0] —> {1, . . . , /;}. B) Either B — [a0] = 0, or there exists a bijection k:B — {a0)—>{\,...,p] for some p < n. The preceding lemma, combined with A), implies that there is no bijection of B with [1,.. . , „ + 1}. This is the first half of what we wanted to prove. To prove the second half, note that if B - [a0] = 0, there is a bijection of B with the set {1}; while if B - {a0} * 0, we can apply the preceding lemma, along with B), to conclude that there is a bijection of B with f 1, ...,/>+ 1}. as desired0356' ^^ " * biJeCti°n °f B with 0. ••..*} for some m < n + 1, principle now shows that the theorem is true for all is finite, there is no bijection of A with a proper subset B 1S S Pr°per SUbset of A a«d that /: A - B is a lT i; Tis ahbUetction g: A- <>. ••..»}for some * Edicts th'e preceding theorem □' ° * ^ lh '''' "}'
Finite Sets 43 orolhry 6.4. The number of elements in a finite set A Is uniquely determined ^ Proof. Let m < n. Suppose there are bijections f:A —>[\,...,n], g:A —>[l,...,m}. Then the composite gcf-> :{1, n)—>{\ ,m} •Sa bijectionof the finite set {1, ...,«} with a proper subset of itself, contra- contradicting the corollary just proved. □ Corollary 6.5. //B is a subset of the finite set A, then B is finite. //B is a roper subset of A, then the number of elements in B is less than the number of elements in A. Corollary 6.6. Z+ is not finite. Proof. The function /: Z+ — Z+ — {1} defined by /(«) = n + 1 is a bijection of Z+ with a proper subset of itself. □ Theorem 6.7. Let B be a nonempty set; let n be a positive integer. Then the following are equivalent: A) There is a surjective function f : [1, . . . , n} —► B. B) There is an injective function g '. B —» [1, . . . , n}. C) B is finite and has at most n elements. Proof. We prove that A) => B) => C) => A); this will suffice to show the statements are equivalent. Let A = [1, . . . , n}. A)=> B). Given a surjective function f:A—*B, define g : B —* A by the equation g(b) = smallest element of f'l({b}). ttause/is surjective, the set f-\{b}) is nonempty. We know by the well- ™ «'ng property that every nonempty subset of Z+ has a unique smallest •*Srrwf IS Wdl defined Th° map 8 'S inJective' for if b * b'> then the different ^ ^~'^*'^ are disJoint> so their smallest elements must be the funrtion*' ^f/ : ^ ~* A be injective- Let C be the ima«e set S(B); then Because r -g C obtained from g by changing its range is bijective. 0. D)f'S a SUbSet °f A = U> • • •, bJ, there is a bijection h : C-* O.-.".',pj OTICp^H- The composite A o^' is a bijection of B with
Chap, i 44 Set Theory and Logic 1 m^nts where 1< m < >'■ Then there is a C) =>A). Let B have m elements, Wti bijection /,:{!,..., />.}-* * ■ ntinn of ,4 onto 5, as desired. If m < «, we extend If «, = n, then /, is a surject.on of ^ on ^ ^.^ ^ = ^^ ^ A to the desired surject.on/. 1U ■ • ■ > i m <i< "• D JZi\nhe index set Us finite, LU->A« a"d TL*£>A* are finite. To put it succinctly, this theorem states that/?,;;/, unioas and finite pro- ducts of finite sets are finite. The proof is left to the exercses. Exercises 1. (a) Make a list of all the injective maps /:{1,2,3}—> [1,2, 3,4}. Show that none is bijective. (This constitutes a direct proof that a set A having three elements does not have four elements.) (b) How many injective maps /:[!,....8}—>{!,..., 10} are there? (You can see why one would not wish to try to prove directly that there is no bijective correspondence between these sets.) 2. Show that if B is not finite and B <= A, then A is not finite. 3. Let X be the two-element set [0, 1}. Find a bijective correspondence between X" and a proper subset of itself. 4. Let A be a nonempty finite simply ordered set. (a) Show that A has a largest element. [Hint: Proceed by induction on the number of elements in A.] (b) Show that A has the order type of a section of the positive integers. 5. (a) Show that A U B is finite if and only if A and B are finite. (b) Lety ,6 0. Show that if J is finite and each set A. for a G J is finite, then U«€/ At is finite. (c) To what extent does the converse of (b) hold? 6. (a) Show that A x B is finite if A and B are finite. fa"finite 0' SH0W tHat if y iS finite and eaCh set A' is finite' then n"€/ A' (c) To what extent does the converse of (b) hold ?
Countable and Uncountable Sets 45 i. i if A is finite, then the set <P(A) of all subsets of A is finite. [Hint: 1. Sh°*J^ the ,wo element set {0, 1}. Define a bijection of <P(/f) with the cartesian product XA-] A and B be sets; let ff be the set of all functions /: A *■ ^nd ^"are finite, then ff is finite. B. Show that if 1-7 Countable and Uncountable Sets I t as the sets [1, • • • , "} are tne prototypes for the finite sets, the set fall the positive integers is the prototype for what we call the countably ■ *fi He sets. In this section we shall study such sets; we shall also construct ome sets that are neither finite nor countably infinite. This study will lead us into a discussion of what we mean by the process of "inductive definition." Definition. A set A is said to be infinite if it is not finite. It is said to be countably infinite if there is a bijective correspondence /: Z+ —> A. Example 1. The set Z of all integers is countably infinite. One checks easily that the function /: Z —> Zf defined by f2n if n > 0, = \ 1-2/; + 1 if n< 0 is a bijection. Example 2. The product Z+ x Z+ is countably infinite. If we represent the elements of the product Z+ x Z+ by the integer points in the first quadrant, then the left-hand portion of Figure 11 suggests how to "count" the points! that is, how to put them in bijective correspondence with the positive integers. Y\ • • "«• a, b5. t t • • • t \ \ Figure 11
46 Set Theory and Logic ChaP-1 A picture is not a proof, of course, but this picture suggests a proof. FiRt> we cTs^uct a bijec.ion /: Z+ x Z+ - A, where ^ .s the subset of z. x £ consTsting of pairs (,,,) for which ,£x,by the equanon f(x,y)*=(x + y- Uy)- Then we construct a bijection of A with the positive integers, defining g: A —> Z± by the formula *(.v,j>)=4(*-l).v + j'. We leave it to you to show that/and g are bijections. Another proof that Z¥ x Z± is countably infinite will be given later. Definition. A set is said to be countable if it is either finite or countably infinite. A set that is not countable is said to be uncountable. There is a theorem concerning countable sets that is similar to Theorem 6.7 about finite sets. It is the following: Theorem 7.1. Let B be a nonempty set. Then the following are equivalent: A) There is a surjective function f : Z+ —> B. B) There is an infective function g : B —> Z+. C) B is countable. Proof. We prove the implications A) => B) => C) => A). A) => B). Let/: Z+ —» B be a surjection. Define g : B —► Z+ by the equa- equation g(b) = smallest element of f'l{{b}). Because/is surjective, /"'({A}) is nonempty; thus g is well defined. The map g is injective, for if b ^ b', the sets /"'({A}) and /"'({*'}) are disjoint, so their smallest elements are different. B) => C). Let g : B —> Z+ be an injection; we wish to prove Bis counta- countable. By changing the range of g, we can obtain a bijection of B with a subset of Z+. Thus to prove our result, it suffices to show that every subset of Z+ is countable. So let C be a subset of Z+. If Cis finite, it is countable by definition. So suppose that C is infinite. We nave to prove Cis countably infinite; that is, we have to construct a bijection " • •£+ —► C. "by induction" Define h{\) to be the smallest element of C; Kn) = smallest element of [C - Ml, ., „ _ 1})]. The set C— Ml „ in • ' A • {1 „ _ W '''r " ls not emPty; for if it were empty, then Theorem 6 7) Ih^hl^^ TrJeCtive' whence C would be finite (by for aU „ Hi () Wdl defined By induction, we have denned h(n)
Countable and Uncountable Sets 47 th t h is injective is easy. Given m < n, note that h{m) belongs foshowtna ^ ^ whereas h(n), by definition, does not. Hence • *" S.7 '" " b(n) * **■ ''. t h js surjective, let c be^ny element of C; we show that c lies Xoshow ^ ^ ^.^t ^^^ that ^Zj cannot be contained in the finite jn the image becausg ;^Zj js innnite (since h is injective). Therefore, there setf1'-;'' such that /»(«) > c. Let w be the smallest element of Z+ such is an « in * Then for all / < m, we must have A(z) < c. Thus c does not that«W-he"set ^^ m_ jj) Sjnce h(m) is defined as the smallest ^'""nt'of the set C - K{U ■■■>>»- 1}), we must have h{m) < c. Putting f ™wo inequalities together, we have h(m) = c, as desired. m=>n\ Suppose that B is countable. If B is countably infinite, there is . ctjon/- z+ — B by definition, and we are through. If B is finite, there ,* abgection h : {1, ...,»}-> ^ for some n > 1. (Recall that B * 0.) We „„ extend A to a surjection/: Z+ — B by defining ja(/) for 1 < / < n, \h(\) for / > n. Xhis completes the proof of the theorem. □ Corollary 7.2. A subset of a countable s?t is countable. Corollary 7.3. The set Z+ x Z+ is countably infinite. Proof. In view of the preceding theorem, it suffices to construct an injec- injective map/: Z+ x Z+ -*Z+. We define/by the equation /(«, m) = 2-3™. It is easy to check that/is injective. For suppose that 2m = 2«. If n < p, then 3" = 2'-', contradicting the fact that 3m is odd for all m. Therefore, »-P- As a result, 3? =3'. Then if m < q, it follows that 1 = 3'"", another contradiction. Hence m=q. Q For^ 3' THe SEt ?+ °f positive rational numbers is countably infinite. we can define a surjection g; Z+ x Z+ -* Q+ by the equation g(ny m) = mjn. * f+X Z+ 'S Countable' there is a surjection /: Z+-^Z+x Z+. Then *e tomtf' re is a surjection /: Z+^Z+x Z+. Then ofcoun=P AE^ ■ ;Z+~^ ^+ is a surjection, so that Q+ is countable. And, We lw • 'S ' ite because » contains Z+. ab'y infinite " ^ Sn CXercise t0 show the set 2 of all rational numbers is count- ciPles of loricP01 utinTthC proof of Theorem 7.1 where we stretched the prin- ^ occurred at the point, in proving the implication we said that "using the induction principle" we had defined
4S Set Theory and Logic Cl>ap. 1 ,hP function h for all positive integers ,,. You may have seen arguments like the function n i v jons raised concerning their legitimacy. We ^erreaSy usldluch In lament ourselves, in the exercises of §1-4,^ WE Rufthtrfis a problem here. After all, the induction principle states only that ifZ is an inductive set of positive integers and Zo contains 1, thenZ0 = Z To use the principle to prove a theorem "by induction," one begins the proof with the statement "Let Zo be the set of all positive integers n for which the theorem is true," and then one goes ahead to prove that Zo contains 1 and that Zo is inductive, whence Zo must be all of Z+. In the preceding theorem, however, we were not really proving a theorem by induction, but defining something by induction. How then should we start the proof? Can we start by saying, "Let Zo be the set of all integers n for which the function h is defined" ? But that's silly; the symbol h has no meaning at the outset of the proof. It only takes on meaning in the course of the proof. So something more is needed. What is needed is another principle, which we call the principle of recur- recursive definition. In the proof of the preceding theorem, we wished to assert the following: Given the infinite subset C of Z+, there is a unique function h :Z+—+C satisfying the formula: . . /'(') = smallest element of C, fi(O = smallest element of [C — A({1,... , / — 1})] for all / > 1. The formula (*) is called a recursion formula for /;; it defines the function/; in terms of itself. A definition given by such a formula is called a recursive defini- definition. Now one can get into logical difficulties when one tries to define something recursively. Not all recursive formulas make sense. The recursive formula h(>) = smallest element of [C — /;({1,...,/+ 1})], for example, is self-contradictory; although /;(;) necessarily belongs to the set (U,..., i + l}), this formula says that it does not belong to the set. Another example is the classic paradox: Let the barber of Seville shave every man of Seville who does not shave himself. Who shall shave the barber? hE bSrber aPpears twice> once i" 'he phrase "the barber of of whom ?h TT ^ Clement °fthe set "men of Seville"; and this definition ' Sha" ShSVe iS S reCursive one-II also haPPenS t0 be Sdf'
Countable and Uncountable Sets 49 §1-7 ome recursive formulas do make sense, however. Specifically, one has the following principle: • • le of recursive definition. Let Abe a set. Given a formula that defines niaue element ofA, and for i > 1 defines h(i) uniquely as an element kintermsofthe values of h for positive integers less than i, this formula "Iter'mines a unique function h : Z+ — A. This is the principle we actually used in the proof of Theorem 7.1. You simply accept it on faith if you like. It may however be proved rigorously, ^ing the principle of induction. We shall formulate it more precisely in the ^ext section and indicate how it is proved. Mathematicians seldom refer to this principle specifically. They are much more likely to write a proof like our proof of Theorem 7.1 above, a proof in which they invoke the "induction principle" to define a function when what they are really using is the principle of recursive definition. We shall avoid undue pedantry in this book by following their example. Now we give some further rules for determining whether or not a set is countable. Theorem 7.4. A countable union of countable sets is countable. Proof. Let {^,},eJ be an indexed family of countable sets, where the index set/is countable. The case J = 0 is trivial; so let us assume that J ^ 0. Assume also that each set A, is nonempty, for convenience; this assumption does not change anything. Because each Aa is countable, we can choose, for each a, a surjective function /„: Z+ —> Aa. Similarly, we can choose a surjective function ?:Z+—>/. Now define h : 2+ X Z+ —> U«eJ Aa by the equation h(n, m)=fg(n){m). !pond hSt * 'S surJ'ective- Slnce z* X z+ ^ in bijective corre- 71 nCE W"h Z+> the C0Untability °f the union follows from Theorem eorent 7.5. A finite product of countable sets is countable. %<intabie Th* fet USsh°W that the Product °f two countable sets A and B is f V reSUlt 'S tr'il f A B i Oh <intabie Th P e ses A and B is functions V 7 reSUlt 'S tr'vial 'f A or B is emPty- Otherwise, choose surjective AxB defin 7f A and * : Z+ ~^ B' Then the function h : Z+ X Z+ — A X B thC eqUat'On h(" m> (/() ()) i jti t B defin 7f + + A X B ^ count blthC eqUat'On h("' m> = (/(«), g(m)) is surjective, so that
SO Set Theory andLoff'c C^P.\ r oneral we proceed by induction. Assuming that A, x ... x A W^f each X is countable, we prove the same thing for the p^'* untable d eactyj^ ^ ^ ^ .g & ^.^ correspondence ««t W^f each X is countab, p countable d eactyj^ ^ ^ ^ .g & ^.^ correspondence defined by the equation «(*„....*.) = ((*i. ••••*-')•*«>• Since the set A, X • •• X Am.t is countable by the induction assumption and A is countable by hypothesis, the product of these two sets is countable, as proved in the preceding paragraph. We conclude that At x •. • x A, h countable as well. Q It is very tempting to assert that countable products of countable sets should be countable; but this assertion is in fact not true: Theorem 7.6. Let X denote the two element set [0, 1}. T/isii the set X" is uncountable. Proof. We show that, given any function g '■ Z+ —> Xm, g is not surjective. For this purpose, let us denote g(n) as follows! g(n) = (*„„ xnl, xn3, ..., xnm,.. .), where each xtJ is either 0 or 1. Then we define a pointy = (ylt y2, ...,/„,...) of X" by letting fO if jc.b = 1, y" (If we write the numbers xnl in a rectangular array, the particular elements xm appear as the diagonal entries in this array; we choose y so that its nth coordinate differs from the diagonal entry *„„.) Now y is a point of X", and y does not lie in the image of g; given n, the point g{n) and the point y differ in at least one coordinate, namely, the «th. Thus g is not surjective. Q The cartesian product {0, l}-» is one example of an Uncountable set. Another is the following: Theorem 7.7. The set 6>(Z+) of all subsets of Z+ is uncountable. Proof One proof consists of showing that (P(Z+) is in bijecti ve correspon- correspondence with the set {0,1}». This we leave to the exercises. Anotner proof is more direct. We shall prove that if A is an arbitrary set,
Countable and Uncountable Sets SI here is no surjective function g: A - 9{A). Then, in particular, the set S{Z^) is "°t.C°U^a(p(/4) be a function. For each a e A, the image g{a) of a S°'h et of A which may or may not contain the point a itself. Let B be 'S a ubset oM consisting of all those points a such that g(a) does not contain "; B=[a\a e A - g{a)}. B may be empty, or it may be all of A, but that does not matter. We t that B is a subset of A that does not lie in the image of g. For suppose tat B = giflo) for some a° e A- We ask the 9uestion: Does a° belong t0 B or'not? By definition of B, a0 e B <=> a0 e A - g(a0) <> a0 e A - B. In either case, we have a contradiction. □ Now we have proved the existence of uncountable sets. But we have not yet mentioned the most familiar uncountable set of all—the set of real num- numbers. You have probably seen the uncountability of R demonstrated already. If one assumes that every real number can be represented uniquely by an infinite decimal (with the proviso that a representation ending in an infinite string of 9's is forbidden), then the uncountability of the reals can be proved by a variant of the diagonal procedure used in the proof of Theorem 7.6. But this proof is in some ways not very satisfying. One reason is that the in- infinite decimal representation of a real number is not at all an elementary consequence of the axioms but requires a good deal of labor to prove. Another reason is that the uncountability of R does not, in fact, depend on the infinite decimal expansion of R or indeed on any of the algebraic properties of R; it depends only on the order properties of R. We shall demonstrate the uncountability of R, using only its order properties, in a later section (§3-6). Exercises 1- Show that Q is countably infinite. 2" Show that the maps/and g of Examples 1 and 2 are bijections. • t A be a set; let X be the two-element set {0, 1}. Show that there is a bijective JTespondence between the set <P(/f) of all subsets of A and the cartesian prod- • a) A real number x is said to be algebraic (over the rationals) if it satisfies me polynomial equation of positive degree *" + a*-,*"-' + • • • + a,x + a0 = 0 coefficients "•■ Assuming that each polynomial equation has any roots, show that the set of algebraic numbers is countable.
52 Set Theory and Logic (b) A real number is said to be transcendental if it is not algebraic. Assu • the reals are uncountable, show that the transcendental numbers are "^ countable. (It is a somewhat surprising fact that only two transcend """ numbers are familiar to us: c and n. Even proving these two numb^ transcendental is highly nontrivial.) R 5. Determine, for each of the following sets, whether or not it is countable. Justif your answers. (a) The set A of all functions/: {0, 1} —* Z+. (b) The set Bn of all functions/: {1,. . ., n] — Z+. (c) The set C = U»ez. B- (d) The set D of all functions/: Z+ —> Z+. (e) The set E of all functions /: Z+ —-> {0, 1}. (f) The set F of all functions/: Z+ —> {0, 1} that are "eventually zero." [\ye say that/is eventually zero if there is a positive integer N such that f(n) = o for all /( > N.] (g) The set C of all functions /: Z+ —> Z+ that are eventually 1. (h) The set H of all functions /: Z+ —> Z+ that are eventually constant, (i) The set / of all two-element subsets of Z+. (j) The set J of all finite subsets of Z+. 6. We say that two sets A and B have the same cardinality if there is a bijection of A with B. (a) Show that if B a A and if there is an injection /: A —> B, then A and B have the same cardinality. [Hint: Define A\ — A,BX = B, and for n > 1, An = /Mn-i) and Bn = /(#„_,). (Recursive definition again!) Note that A, r> B, r> A2 => B2 => A3 => • ■ ■. Define /; : A ~> B by the rule if x G An — Bn for some n, otherwise.] (b) Theorem {SchrOeder-Bemstein theorem). If there are injections f:A-^>C and g: C —► A, then A and C have the same cardinality. 7. Show that the sets D and E of Exercise 5 have the same cardinality. 8. Let A" denote the two-element set {0, 1}; let <B be the set of all countable subsets of X". Show that X" and <B have the same cardinality. 9. (a) The recursion formula A0) = 1, (*) hB) = 2, h(n) = [h(n + I)]2 - [h(n — I)]2 for n ^ 2 is not one to which the principle of recursive definition applies. Show that nevertheless there does exist a function h:Z+—+R satisfying this formula. [Hint: Reformulate (*) so that the principle will apply and require A to be positive.] , . .. ,
The Principle of Recursive Definition S3 thftt the formula (.) of part (a) does not determine h uniquely. (W gE- SisapSe funciion satisfying (♦), tot /@ - «0 for « * 3, and function /,: Z+ - * satisfying the recursion formula Ml) = 1, h{2) = 2, A(/l) = [/,(/, + DP + [h(n - DP for n > 2. *i-# r/»e Principle of Recursive Definition Before considering the general form of the principle of recursive defini- definition let us first prove it in a specific case, the case that was used in the proof of Theorem 7.1. That should make the underlying idea of the proof much clearer when we consider the general case. So, given the infinite subset C of Z4, let us consider the following recursion formula for a function h : Z+ — > C: /,(]) = smallest element of C, ^ /((() = smallest element of [C - h([ 1, .. . , ;" — 1})] for i > 1. We shall prove that there exists a unique function // : Z+ —» C satisfying this recursion formula. The first step is to prove that there exist functions defined on sections {1,...,«} of Z+ that satisfy (*): Lemma 8.1. Given n e Z+, there exists a function f:{l,...,„}—>C that satisfies (*)for all i in its domain. Proof. The point of this lemma is that it is a statement that depends on n, and therefore it is capable of being proved by induction. Let Zo be the set of Ji«'torAvhuch the lemma holds. We show that Zo contains 1 and is inductive. «then follows that Zo =Z+. equation™ * "^ {oTn = l' since the function/: {1} - C defined by the satisfies (.). /(O = smallest element of C f0r " - l> we Prove [i true f^ »■ By ■ • U« " " >J - C satis^-B M ^ all _► c by the equations /(')= /'(.-) for/g {!,...,„_ 1}> /(») = smallest element of [C - /'({I, .,.,„_ i))].
Chap. 1 54 Set Theory and Log* c ■* fh not surjective; hence the set C - /'({I, ...,«_ 1}. Knee C is infinite./ snot ^ ^^ ^ that this definitiOn is an ac- is not empty, *™ ^ define/in terms of ft«//but in terms of thegiVen ceptable one; it Qo ■ function f. that/satisfies (*) for all i in its domain. The function/ satisfies^) for' < n - I because it equals /' there. And/satisfies (.) £ fa=Sn because, by definition, m = smallest element of [C - /'({I ,...,*•- I})] and /•({I,---."- '))=/({'. ■■■.»- '»■ O »* 5.2. SWC« »** f: {1, ■ ■ ■. "} - C ««/ g: {1, m] -.c 6*fA wfu/V (*)for all i w ^i> respective domains. Then f(i) = g(i)/or a//, ;„ 6of/i domains. Proof. Suppose not. Let i be the smallest integer for which /(i) ^ g(f). The integer f is not 1, because /(I) = smallest element of C = g(l), by (*). Hence i > I, and for ally < i, we have f(j) = ^(y). Because/andg satisfy (*), /@ = smallest element of [C — /({I,...,('— I})], g@ = smallest element of [C — g({l,. .. ,i — 1})]- Since /({I,..., i - 1}) = g([l,...,(- 1}), we have /(;") = ^@, contrary to the choice of i. Q Theorem 8.3. There exists a unique function h: Z+ —► C satisfying (*) for all i g Z+. Proof. By Lemma 8.1, there exists for each n a function that maps [I,...,«} into C and satisfies (*) for all i in its domain. Given n, Lemma 8.2 shows that this function is unique; two such functions having the same domain must be equal. Let/,: {I, ...,«}—> C denote this unique function. Now comes the crucial step. We define a function h : Z+ — C by defining its rule to be the union V of the rules of the functions /„. The rule for/, is a subset of {!,...,«} x C; therefore, U is a subset of Z+ X C. We must show that V is the rule for a function h:Z+—>C. That is, we must show that each element i of Z+ appears as the first coor- coordinate of exactly one element of V. This is easy. The integer i lies in the domain of/, if and only if „ > ,-. Therefore, the set of elements of U of which i is the first coordinate is precisely the set of all pairs of the form (,-,/.@). for f T^82 te"S US thSt m -'"■« if "' m * «■ The^fore, all
The Principle of Recursive Definition SS ft-* To show that h satisfies (.) is also easy; it is a consequence of the fol- '°*ingfaCtS: Hi) =/.(») for.<«, /„ satisfies (*) for all; in its domain, proof of uniqueness is a copy of the proof of Lemma 8.2. \J Now we formulate the general principle of recursive definition. There are new ideas involved in its proof, so we leave it as an exercise. Theorem 8.4 (Principle of recursive definition). Let A be a set; let a0 be „ element of A. Suppose p is a function that assigns, to each function f map- mapping a section of the positive integers into A, an element of A. Then there exists a unique function h : Z+ —> A such that h(l)=a0) W l,...,i - 1}) for\> 1. The formula (*) is called a recursion formula for h. It determines /((I), and it expresses the value of /; at i > I in terms of the values of /; for positive integers less than i. Example 1. Let us show that Theorem 8.3 is a special case of this theorem. Given the infinite subset C of Z+, let a0 be the smallest element of C, and define p by the equation p(f) = smallest element of [C — (image set of /)]. Because C is infinite and /is a function mapping a section of Z+ into C, the image set of/is not all of C; therefore, p is well defined. By Theorem 8.4 there exists a function h:Z+-,C such that h(\) = a0, and for i > 1, = smallest element of [C - (image set of h\{\,. . ., i - 1})] = smallest element of [C - h({\,. .., i — 1})], as desired. Example 2. Given a e R, we "defined" a", in the exercises of §1-4, by the recursion formula a1 = a, a" = a"~1-a. •"Attat 5 T-y Theorem 84 t0 define a function h:Z+ — R rigorously "nd defin iT °"' T° ap?Iy this theorem'Iet "o denote the element a of R ne P by the equation />(/) = f(m)-a, where /maps a section of 2*
56 Set Theory and Logic into R and m is the largest element of the domain of/. Then there exists a uniqile function h:Z+-* R such that MO = "o, //(/)=/>(//1 {!.- -.,''-I)) fori> 1. This means that *(!)=«, and //(/) = «i - !)•« for / > 1. If we denote *0) by a', we have a' = a, at _ ,,/- i . Oj as desired. 1. Let F,, 62,. ■ ■) be an infinite sequence of real numbers. The sum 2*-i bk is defined by induction as follows: S*-i**=*i for// = l, 2Z= ■ 6* = (S*" > **> + 6» for " > ' • Let A be the set of real numbers; choose p so that Theorem 8.4appliesto define this sum rigorously. We sometimes denote the sum £" =, bk by the symbol bt + b2 + ■ ■ ■ + bn. 2. Let Fj, b2,...) be an infinite sequence of real numbers. We define the product IK=i bk by the equations ILUi6*=6i, m=■ bk = cnr=' **) • 6n for«> 1. Use Theorem 8.4 to define this product rigorously. We sometimes denote II*=i bk by the symbol b,b2 ■ ■ ■ bn. 3. Obtain the definitions of a" and n\ for « g Z+ as special cases of Exercise 2. 4. The Fibonacci numbers of number theory are defined recursively by the formula A, = A2 = r, An = AB_i + An_2 for n > 2. Define them rigorously by use of Theorem 8.4. 5. Show that there is a unique function h:Z+-*R+ satisfying the formula MD = 3, [( ) + i] for/> j «• (a) Show that there is no function h : Z+ _» j?t satisfying the formula = 3, for/>1. why this example does not violate the principle of recursive defini-
§1-9 Infinite Sets and the Axiom of Choice 57 Consider the recursion formula «1) = 3. for/>1 Show that there exists a unique function //: Z+ —» R+ satisfying this for- formula. 7, Prove Theorem 8.4. 1-9 Infinite Sets and the Axiom of Choice We have already obtained several criteria for a set to be infinite. We know for instance that a set A is infinite if it has a countably infinite subset, or if there is a bijection of A with a proper subset of itself. It turns out that either of these properties is sufficient to characterize infinite sets. This we shall now prove. The proof will lead us into a discussion of a point of logic we have not yet mentioned—the axiom of choice. Theorem 9.1. Let Abe a set. The following statements about A are equiva- equivalent: A) There exists an injective function f: Z+ —> A. B) There exists a bijection of A with a proper subset of itself. C) A is infinite. Proof. We prove the implications A) => B) => C) => A). To prove that A) => B), we assume there is an injective function/: Z+ —<■ A. Let the image set /(Z+) be denoted by B; and let /(«) be denoted by «„. Because/is injec- " ' '" m. Define tive, an^am if n ■ by the equations g:A~>A -{a,} £(«„) = an+l for an 6 B, g(x) =x for x s A ~ B. S is indicated schematically in Figure 12; one checks easily that it A -8 Figure 12
S8 Set Theory and Logic aP-1 T, „ lmni1Cation B) => C) is just the contrapositive of Corollary 6.3, so ,t his a X ten proved. To prove that C) ~ A) we assume that A » fit, and construct "by induction" an injective function /: Z+ — A ' First, since the set A is not empty, we can choose a point a, of A; define f(\~\ to be the point so chosen. Then assuming that we have defined /(I), ...,/(«- D, we wish todefine /■(„) The set ^ - /({I. ...,»-!})« not emP?: for |f " w«e empty, the a /"• f 1 , « — 1} —' ^ would be a surjection and ^ would be finite. Hence we can choose an element of the set A - /({I,...,«- 1}) and define /(/,) to be this element. "Using the induction principle," we have defined/ for all// e Z+. It is easy to see that/is injective. For suppose that m < n. Then /(«) belongs to the set /({I, ...,«- 1}), whereas /(«), by definition, does not. Therefore, f{n) ^ f(m). D Let us try to reformulate this "induction" proof more carefully, so as to make explicit our use of the principle of recursive definition. Given the infinite set A, we attempt to define/: Z¥ — > A recursively by the formula v ' /(/) = an arbitrary element of [A - /((I, ...,/— 1})] for / > 1. But this is not an acceptable recursion formula at all! For it does not define /(/) uniquely in terms of / |{1, ...,/— 1}. In this respect this formula differs notably from the recursion formula we considered in proving Theorem 7.1. There we had an infinite subset C of Z+, and we defined h by the formula h(\) = smallest element of C, h(i) = smallest element of [C - h({\ /" — 1))] for i > 1. This formula does define h(i) uniquely in terms of h \ {1 i - 1). Another way of seeing that (*) is not an acceptable recursion formula is to note that if it were, the principle of recursive definition would imply that there is a unique function/: Z+ — A satisfying (*). But by no stretch of the imagination does (*) specify/uniquely. In fact, this "definition" of/involves infinitely many arbitrary choices. What we are saying is that the proof we have given for Theorem 9.1 is not actually a proof. Indeed, on the basis of the properties of set theory we i^S ^ ^ * " ^ P°SSible t0 Pr°Ve th'S theorem Something more is^Sd ^ " ^ P°SSible t0 Pr°Ve th'S theorem- Something ^Previously, we described certain definite allowable methods for specifying A) Defining a set by listing its elements, or by taking a given set A and
Infinite Sets and the Axiom of Choice 59 §1-9 specifying a subset B of it by giving a property that the elements of B „. ^unions or intersections of the elements of a given collection of ( its o°r taking the difference of two sets. n\ Taking the set of all subsets of a given set. ia\ Takin" cartesian products of sets. thi rule for the function/is really a set: a subset of Z+ X A. Therefore, the existence of the function /we must construct the appropriate t0 pr°VfZ X A using the allowed methods for forming sets. The methods s"bSed' ° iven simply are not adequate for this purpose. We need a new way ofasserting the existence of a set. So, we add to the list of allowed methods of forming sets the following: Axiom of choice. Given a collection Q, of disjoint nonempty sets, there exists a set C having exactly one element in common with each element of (X; that is, such that for each A e a, the set C n A contains a single element. This axiom asserts the existence of a set that can be thought of as having been obtained by choosing one element from each of the sets A in Ct. The axiom of choice certainly seems an innocent-enough assertion. And, in fact, most mathematicians today accept it as part of the set theory on which they base their mathematics. But in years past a good deal of contro- controversy raged around this particular assertion concerning set theory, for there are theorems one can prove with its aid that some mathematicians were reluctant to accept. One such is the well-ordering theorem, which we shall discuss shortly. For the present we shall simply use the choice axiom to clear up the difficulty we mentioned in the preceding proof. First we prove an easy consequence of the axiom of choice: lemma 9.2 {Existence of a choice function). Given a collection <$, of non- nonempty sets {not necessarily disjoint), there exists a function c : (B —> \JBea B ** Aat c(B) is an element of B, for each B s <B. The dffCti°n C ^ Ca"ed a °h0iCe function for the collection <B. this lemm thCnCe betWeen this lemma and the axiom of choice is that in example * **** °f the collection ® are not required to be disjoint. For Siven set.01* ^ a"°W ® t0 be the collection of all nonempty subsets of a 00 of the lemma. Given an element B of <B, we define a set B' as follows: 5,j . B' = [(B,x)\x<=B}. e collection of all ordered pairs, where the first coordinate of
Set Theory and Logic Chap. | the ordered pair is the set B, and the second coordinate is an element or* The set B' is a subset of the cartesian product <B X \JB&a B. Because B contains at least one element x, the set B' contains at least the plement (B x), so it is nonempty. Now we claim that if Bt and B2 are two different sets in <B, then the sets B' and B\ are disjoint. For the typical element of B\ is a pair of the form (B x ) and the typical element of B2 is a pair of the form (Blt x2). No two such elements can be equal, because their first coordinates are different. Now let us form the collection Q = {B'\B e <&}; it is a collection of disjoint nonempty subsets of « X (Joe® B. By the choice axiom there exists a set c having exactly one element in com- common with each element of 6. Our claim is that c is the rule for the desired choice function. In the first place, c is a subset of (B x LU® B- In the second place, c contains exactly one element from each set B'\ therefore, for each B e (&, the set c contains exactly one ordered pair (B, x) whose first coordinate is B. Thus c is indeed the rule for a function from the collection <B to the set \JBe® B. Finally, if (B, x) <= c, then x belongs to B, so that c(B) g B, as desired. □ A second proof of Theorem 9.1. Using this lemma, one can make the proof of Theorem 9.1 more precise. Given the infinite set A, we wish to construct an injective function/: Z+ —» A. Let us form the collection (B of all nonempty subsets of ,4. The lemma just proved asserts the existence of a choice function for <B; that is, a function c:<&—*\JM&tB = A such that c(B) s B for each B e (B. Let us now define a function /: Z+ — A by the recursion formula /(/) = C(A - /({l / _ i})) f0r ,- > i. Because A is infinite, the set A - f{{\ z - 1}) is nonempty; therefore, the right side of this equation makes sense. Since this formula defines f(i) uniquely m terms of /1 [1,...,/_ i}; the principle of recursive definition I appliesfWe conclude that there exists a unique function/: Z, - A satisfy- ^jng (*) for all t e Z+. Injectivity of/follows as before. Q
Infinite Sets and the Axiom of Choice 61 §1-? • mnhasized that in order to construct a proof of Theorem 9.1 Hargc2 correct, one must make specific use of a choice function, we that is l°g ca y nd admjt that in practice most mathematicians do no such D0VV Thev so on with no qualms giving proofs like our first version, proofs thl"^ olve an infinite number of arbitrary choices. They know that they are thatl" • the choice axiom; and they know that if it were necessary, they d out their proofs into a logically more satisfactory form by introducing hoice function specifically. But usually they do not bother. aeAnd either will we. You will find few further specific uses of a choice ction in this book; we shall introduce a choice function only when the roof would become confusing without it. But there will be many proofs in which we make an infinite number of arbitrary choices, and in each such case we will actually be using the choice axiom implicitly. Now we must confess that in an earlier section of this book there is a proof in which we constructed a certain function by making an infinite num- number of arbitrary choices. And we slipped that proof in without even men- mentioning the choice axiom. Our apologies for the deception. We leave it to you to ferret out which proof it was! Let us make one final comment on the choice axiom. There are two forms of this axiom. One can be called the finite axiom of choice; it asserts that given a finite collection Q, of disjoint nonempty sets, there exists a set C having exactly one element in common with each element of a. One needs this weak form of the choice axiom all the time; we have used it freely in the preceding sections with no comment. No mathematician has any qualms about the finite choice axiom; it is part of everyone's set theory. The stronger form of the axiom of choice, the one that applies to an arbitrary collection a of nonempty sets, is the one that is really properly "lied "the axiom of choice." When a mathematician writes, "This proof epends on the choice axiom," it is invariably this stronger form of the axiom that is meant. Exercises 'ln^'^u111^ /: Z+ - X"' where X is the two-element set {0,1], using the choice axiom. ^ch? a Ch°iCe funCtiOn for each of the ^'lowing collections, without noiCe axiom: ( ) T noiCe axiom: I!6*'0" * °f nonemPty subsets of Z+. ® °f n°t b & TheSM' pty subsets of Z- (d> The coll!w!0n i°f nonemPty subsets of the rational numbers Q. 3> SuPPose thaM nonemPty subsets of X", where X = fO, 1}. ■' 1S a set and if'}'**, is a given indexed family of injective functions /-•"{I n}—> A.
Set Theory and Logic Chap. j Show that A is infinite. Can you define an injective function f:Z.^A without using the choice axiom? 4 Show that the choice axiom is equivalent to the statement that for any indexed family {4»W of nonempty sets, with J^0, the cartesian product is not empty. 5 There was a theorem in §1-7 whose proof involved an infinite number of arbi- ' trary choices. Which one was it? Rewrite the proof so as to make explicit the use of the choice axiom. (Several of the earlier exercises have used the choice axiom also.) 6. (a) Use the choice axiom to show that if/: A —> B is surjective, then/has a right inverse h : B —> A. (b) Show that if /: A —> B is injective and A is not empty, then/has a left inverse. Is the axiom of choice needed ? (c) Show that if A is any set, there is no injective map /: 6>(A) —* A, where <9{A) is the set of all subsets of A. 7. Most of the famous paradoxes of naive set theory are associated in some way or other with the concept of the "set of all sets." None of the rules we have given for forming sets allows us to consider such a set. And for good reason—the concept itself is self-contradictory. For suppose that Ct denotes the "set of all sets." (a) Show that (P(Ct) <= Q,; derive a contradiction. (b) (Russell'sparadox). Let (B be the subset of ft consisting of all sets that are not elements of themselves; (B = {A | A e a and A £ A}. (Of course, there may be no set A such that A e A; if such is the case, then <B = ft.) Is (B an element of itself or not ? 8. Let A and B be two nonempty sets. If there is an injection of B into A, but no injection of A into B, we say that A has greater cardinality than B. (a) Conclude from Theorem 9.1 that every uncountable set has greater cardi- cardinality than Z+. (b) Show that if A has greater cardinality than B, and B has greater cardinality than C, then A has greater cardinality than C. (c) Find a sequence Au A2,... of infinite sets, such that for each n £ Z+, the set A^ has greater cardinality than An. (d) Find a set that for every n has cardinality greater than Aa. *9. Show that (P(Z+) and R have the same cardinality. A famous conjecture of set theory, called the continuum hypothesis, asserts that there exists no set having greater cardinality than Z+ and lesser cardinality than R. The generalized continuum hypothesis asserts that given A infinite, there is no set having greater cardinality than A and lesser cardinality than <?&)• Surprisingly enough, both of these assertions have been shown to be indepen- independent of the usual axioms for set theory. For a readable expository account, see [Sm].
§1-10 Well-Ordered Sets 63 1-10 Weil-Ordered Sets f the useful properties of the set Z+ of positive integers is the fact °h fits nonempty subsets has a smallest element. Generalizing this thate!f i^,ds to the concept of a well-ordered set. propt* v ,. • • * A set /* with an order relation < is said to be well-ordered if every nonempty subset of A has a smallest element. Example 1. Consider the set {1, 2] x Z+ in the dictionary ordering. Sche- tically <t can be represented as one infinite sequence followed by another infinite sequence: at' a2' a3> • ■ •> "i> , b3, . . . with the understanding that each element is less than every element to the right of it It is not hard to see that every nonempty subset C of this ordered set has a smallest element: If C contains any one of the elements an, we simply take the smallest element of the intersection of C with the sequence at,a2, . . .; while if Ccontains no am then it is a subset of the sequence bu bz,. ■ . and as such has a smallest element. Example 2. Consider the set Z+ x Z+ in the dictionary order. Schemat- Schematically, it can represented as an infinite sequence of infinite sequences. We assert that it is well-ordered. Let X be a nonempty subset of Z+ x Z+. Let A be the subset of Z+ consisting of all first coordinates of elements of X. Now A has a smallest element; call it aQ. Then the collection {b\aa x be X] is a nonempty subset of Z+; let b0 be its smallest element. By definition of the dictionary order, a0 x b0 is the smallest element of X. (See Figure 13.) Figure 13
Set Theory and Logic Chap. \ Example 3 The set of integers is not well-ordered in the usual order- the subset consisting of the negative integers has no smallest element. Nor is the set of real numbers in the interval 0 < x < 1 well-ordered; the subset consisting of those x for which 0 < x < 1 has no smallest element (although it has a greatest lower bound, of course). There are several ways of constructing well-ordered sets. Two of them are the following: A) If A is a well-ordered set, then any subset of A is well-ordered in the restricted order relation. B) If A and B are well-ordered sets, then A X B is well-ordered in the dictionary order. The proof of A) is trivial; the proof of B) follows the pattern given in Exam- Example 2. It follows that the set Z+ x (Z+ X Z+) is well-ordered in the dictionary order; it can be represented as an infinite sequence of infinite sequences of infinite sequences. Similarly, (Z+L is well-ordered in the dictionary order. And so on. But if you try to generalize to an infinite product of Z+ with itself, you will run into trouble. We shall examine this situation shortly. Now, given a set A without an order relation, it is natural to ask whether there exists an order relation for A that makes it into a well-ordered set. If A is finite, any bijection f.A—>{\,...,n] can be used to define an order relation on A under which A has the same order type as the ordered set {1, . . . , «). In fact, every order relation on a finite set can be obtained in this way: Theorem 10.1. Every nonempty finite ordered set has the order type of a section {1,..., n) ofZ+, so it is necessarily well-ordered. Proof. This was given as an exercise in §1-6; we prove it here. First, one shows that every finite ordered set A has a largest element: If A has one ele- element, this is trivial. Supposing it true for sets having n — 1 elements, let A have n elements and let a0 e A. Then A — {a0} has a largest element au and the larger of {aQ, a,) is the largest element of A. Second, one shows there is an order-preserving bijection of A with U>.-.,«} for some n: If A has one element, this fact is trivial. Suppose that it is true for sets having n - 1 elements. Let b be the largest element of A. By hypothesis, there is an order-preserving bijection f':A-{b}—+[l,...,n-l). Define an order-preserving bijection/: A -> [1,.. . , n] by setting /(*)-/'(*) for x *b,
Well-Ordered Sets 65 §1-10 u a finite ordered set has only one possible order type. For an infinite seJhin^ are quite different. The well-ordered sets A,...,n] X Z+, Z+ X Z+, Z+ X (Z+ X Z+) * al, countably infinite, but they all have different order types. Ill the examples we have given of well-ordered sets are ordenngs of ntable sets. It is natural to ask whether one can find a well-ordered un- C0UThe obvious uncountable set to try is the countably infinite product /=ztxztx-= (z+y of z with itself. One can generalize the dictionary order to this set in a natural way, by defining (a,,fl2, ...) <(bl,b2,. ..) if for some n> 1, a, = b; for i < n and an < bn. This is, in fact, an order relation on the set X; but unfortunately it is not a well-ordering. Consider the set A of all elements x of X of the form x = (l,..., 1,2, 1, 1,...), where exactly one coordinate of x equals 2, and the others are all equal to 1. The set A clearly has no smallest element. We have seen that the dictionary order at least does not give a well-order- well-ordering of the set (Z+)". Is there some other order relation on this set that is a well-ordering? No one has ever constructed a specific well-ordering of (Z+)". Nevertheless, there is a famous theorem that says such a well-ordering exists: Theorem (Well-ordering theorem). If A is a set, there exists an order rela- "on on A that is a well-ordering. matM the°rem was Proved by Zermelo in 1904, and it startled the mathe- ' world. There was considerable debate as to the correctness of the "nco ' \ C 'aCk °f a"y COnstructive procedure for well-ordering an arbitrary dosel th ^ 'ed many t0 be skePtIca'- When the proof was analyzed "as a co °nly.P°int at which !t was found that there might be some question aeonstr UCtl°n involving an infinite number of arbitrary choices, that is, Some" I involvin8-«« ^oice axiom. l« a le£r matlcians reJected the choice axiom as a result, and for many »ice ajdo^ qUestion about a new theorem was: Does its proof involve °m or not? A theorem was considered to be on somewhat shaky
Set Theory and Logic ChaP. I around if one had to use the choice axiom in its proof. Present-day mathe- ScTans, by and large, do not have such qualms. They accept the axiom of choice as a reasonable assumption about set theory, and they accept the well- ordering theorem along with it. In this book we are not going to prove the well-ordenng theorem. It is lone (although not exceedingly difficult) and primarily of interest to logicians. For those interested, a proof is outlined in the supplementary exercises at the end of the chapter. One seldom needs the well-ordenng theorem to prove theorems of topol- topology in any case; we shall use it to prove a theorem only twice, once in the exercises of §4-5 and again in §6-3. For us, its primary function will be to provide us with a particular counterexample that will be very useful. We shall not, in fact, need the full strength of the well-ordering theorem to construct the example we want. We shall need only the following weaker result, which we now assume: Corollary. There exists an uncountable well-ordered set. (It is of some interest to note that one can prove this weaker assertion without use of the choice axiom; see the supplementary exercises.) To construct the desired example, we first need some terminology. Definition. Let X be an ordered set. Given a, 6 X, the set S, = [x\x e Zand x < <x} is called the section of X by a. Theorem 10.2. There exists an uncountable well-ordered set, every section of which is countable. Proof. By assumption, there exists an uncountable well-ordered set X. From this fact it follows that there exists an uncountable well-ordered set Y, at least one section of which is uncountable. The set [1, 2} x X in the diction- dictionary order is one such set; its section by any element of the form 2 X X is uncountable. Consider the subset of Y consisting of those elements, a for which the section S« is uncountable; let Q be the smallest such element. Then Sa is a well-ordered set, it is uncountable, and every section of Sa is count- The well-ordered set of this theorem is called a minimal uncountable well- ordered set; the reason for the choice of terminology is fairly obvious. Its order type ,s, in fact, uniquely determined by the requirements of the theorem. (bee the supplementary exercises.) We shall denote it throughout this book Z? f ??oT USe the Symbo1 *• thro"ghout to denote the well-ordered set ^o u |S2].
Well-Ordered Sets 67 fl-lO The most useful property of the set Sa for our purposes is expressed in the following corollary: Corollary 10.3- //A is a countable subset o/SD, then A has an upper bound 10 p t Let A be a countable subset of Sa. For each a e A, the section So ntable Therefore, the union B = \JasA 5« is also countable. Since Sa l f S l b it f 5 tht is nt in ntable Therefore, the union JasA « 'S ountable the set B is not all of Sa; let .v be a point of 5n that is not in « "-Then -v is an upper bound for A. For if x<a for some a in ^, then x belongs to Sa and hence to B, contrary to choice. □ Exercises 1 Show that every well-ordered set has the least upper bound property. 2 (a) Show that in a well-ordered set, every element except the largest (if one exists) has an immediate successor. (b) Find a set in which every element has an immediate successor that is not well-ordered. 3. Both A,2} x Z+ and Z+ x {1,2} are well-ordered in the dictionary order. Do they have the same order type ? 4. (a) Let Z. denote the set of negative integers in the usual order. Show that a simply ordered set A fails to be well-ordered if and only if it contains a subset having the same order type as Z._. (b) Show that if A is simply ordered and every countable subset of A is well- ordered, then A is well-ordered. 5. Let Sa be the minimal uncountable well-ordered set. (a) Show that Sa has no largest element. (b) Show that for every a, <Q, the set {x\a <x <Q] is uncountable. (c) Let Xa be the subset of Sa consisting of all elements x such that x has no immediate predecessor. Show that XQ is uncountable. LeU be a well-ordered set. A subset Jo of / is said to be inductive if for every (■S. <= -A>) ==> « e Jo. eorem (The principle of transfinite induction). If J is a well-ordered set and ° "an lnduct™ subset ofJ, then Ja = /. ' foT^r "?' ^ Ca" a SUbSet J° of the we"-°rdered set / "semi-inductive" if tymeanl T™31' K s J°'the immediate successor of a (if any) is in y0. Show inductive °ra" examP|e that if-^o contains the smallest element of /and is semi- g Llve, Jo need not equal all of /. ^' and Az ** disJ°int sets> well-ordered by <, and <2) respectively, an order relation on Ax u Az by letting a < b either if a, b e Ax
68 Set Theory and Logic Chap. \ and a <i b, or if a, b e /*2 and a <2 6, or if a e /«, and 6 s ^ that this is a well-ordering. (b) Generalize (a) to an arbitrary family of disjoint well-ordered sets, indexed by a well-ordered set. 9. (a) Show that (Z+)" has the order type of a section of (Z+)"+>, both in the dic- dictionary order. (b) Find a well-ordered set that for every // has a section having the order type of(Z+)». 10. Theorem. Let J and C be well-ordered sets: assume that there is no surjeclive function mapping a section of J onto C. Then there exists a unique function h:J —> C satisfying the equation (*) h(.\) = smallest [C - //E,)] for each x e J, where Sx is the section ofjby x. Proof. (a) If/; and k map sections of J, or all of J, into C and satisfy (*) for all x in their respective domains, show that h(x) = A-(.v) for all .v in both domains. (b) If there exists a function h : Sx --> C satisfying (*), show that there exists a function k: 5a U (a} —► C satisfying (*). (c) If K <= J and for all (X e K there exists a function //„ : 5, —> C satisfying (*), show that there exists a function satisfying (*). (d) Show by transfinite induction that for every a e J, there exists a function h : 5. —> C satisfying (*). (e) Prove the theorem. 11. Let A and B be two sets. Assuming the well-ordering theorem, prove that either they have the same cardinality, or one has cardinality greater than the other. [Hint: If there is no surjection/: A —> B, apply the preceding exercise.] 12. Theorem (Principle of transfinite recursive definition). Let J be a well-ordered set with smallest element <X0. Let C be a set. Let ff be the set of all functions mapping sections of] into C (including the empty function mapping S^o into C). Suppose that a function p-.'S —> C is given. Then there exists a unique function h:J—*C such that for each x e J. [Hint: The proof follows the pattern outlined in Exercise 10.] *1-11 The Maximum Principle* We have already indicated that the axiom of choice leads to the deep theorem that every set can be well-ordered. The axiom of choice has another •fThis section will be assumed in Chapter 5.
The Maximum Principle 69 § 1-11 nee which is even more important in mathematics. It is the principle C°lTthe maximum principle, which we now discuss F t we make a definition. Given a set A, a relation < on A is called a . v ' a;tial order on A if it has the following two properties: nWNonreflexivity) The relation a < a never holds. il) (Transitivity) If a < b and b < c, then a<c. Thse are just the second and third of the properties of a simple order (see J1.3V the comparability property is the one that is omitted. In other words, strict partial order behaves just like a simple order except that it need not be true that for every pair of distinct points ,v and y in the set, either x -< y If-< is a strict partial order on aset^, it can easily happen that some sub- subset B of A is simply ordered by the relation; all that is needed is for every pair of elements of B to be comparable under -<. Now we can state the maximum principle. Theorem {The maximum principle). Let A be a set; let -<. be a strict partial order on A; let B be a subset of A that is simply ordered by ~<. Then there exists a maximal simply ordered subset C of A containing B. Said differently, there exists a subset C of A containing B such that C is simply ordered by -< and such that no subset D of A properly containing C is simply ordered by -<. Example I. If a is any collection of sets, the relation "is a proper subset or is a strict partial order on a. Suppose that a is the collection of all circular regions (interiors of circles) in the plane. One maximal simply ordered sub- collection of a consists of all circular regions with centers at the origin. Another maximal simply ordered subcollectton consists of all circular regions bounded by circles tangent from the right to the j>-axis at the origin. See Figure 14. Now the maximum principle guarantees not only that a has a maximal simply ordered subcollection but also that starting with any simply ordered Figure 14
Chan i Set Theory and Logic , ■ ffi of a there is a maximal simply ordered subcollection e of ""? nmimtfIPi» insists of a single circular region * i, is easy ,0 „„] a t«il simply °rdered -collections of a tha, contain the several different ^^.^ g| and g2 indicated in F.gure 14 are two such. B consists of a more complicated collection, such as the three regions [in Figure 15, it is more work to find a maximal simply ordered sub- ._ ^tainfnii'(B. We leave it to you. collection containing'(B. We leave it to you. Figure 15 Example 2. If (xo,yo) and (.v,,;',) are two points of the plane R1, define (•vo.;'o) -< Ui,;'i) if Jo = y\ and .v0 < -v,. This is a partial ordering of /?- under which two points are comparable only if they lie on the same horizontal line. The maximal simply ordered sets are the horizontal lines in R-. One can give an intuitive "proof" of the maximum principle that is rather appealing. It involves a step-by-step procedure, which one can describe in physical terms as follows. Take a box, and put into the box all the elements of B. Then examine the remaining elements of A one by one. First pick an element of A not in B. If it is comparable with everything in B, put it in the box; if not, throw it away. At the general step, you will have a collection of elements in the box and a collection of elements that you have tossed away. Take one of the remaining elements of A. If it is comparable with everything in the box, toss it in the box, too; otherwise, throw it away. Similarly contin- continue. After you have checked all the elements of A, the elements you have in the box will be comparable with one another, and thus they will form a simply ordered set. Every element not in the box will be noncomparable with at least ZT^Tl-" u b°X> f°r that was why il was toss<=d away. Hence the simply A y ^ ^Ifi v°n?h thG Ta,k P°int in the »recedinS "Proof" comes when we r-f 1 h"Vet,ChfCked a" the elements of ^." How do you know you r geuhrough checking all the elements of A ? * Should happen to be countable, it is not hard to make this intui-
The Maximum Principle 71 §1-" real proof Let us take the countably infinite case; the finite e,ements of A _ B with the positive integers, a- tinct integers correspond to distinct points of A - B. This indexing letting dls"" decjding what order to test the elements of A - B in, and how ^TnoCwhen one has tested them all. <; cifically we define a function h : Z+ — {0, 1 ], which assigns the 0 to / if we "put a, in the box," and the value 1 if we "throw a, away." To be more formal, consider the following recursion formula: /,(,) = 0 if a, is comparable with every element of the set B U {a, | 1 <j < I - 1 and /;(/) = 0}, f,(f) = 1 otherwise. Bv the principle of recursive definition, this formula determines a unique function h : Z+ — {0, 1}. It is easy to check that is a maximal simply ordered subset of A. If A — B is not countable, a variant of this procedure will work, if we as- assume the well-ordering theorem. Instead of indexing the elements of A — B with the set Z+, we index them (in a bijective fashion) with the elements of some well-ordered set J: A- B = {aJaeJ}. For this we need the well-ordering theorem, so that we know there is a bijec- bijective correspondence between A — B and a well-ordered set J. Then we can proceed as in the previous paragraph, letting the section 5a replace the section (I,...,;— 1} in the argument. Strictly speaking, you need to generalize the principle of recursive definition to well-ordered sets as well. (See Exercise 12 of the preceding section.) This discussion is far from complete. But it does indicate that the well- Ordering theorem and the maximum principle are related. It turns out that P'are in fact equivalent; either of them implies the other. Furthermore, each '"em <s equivalent to the choice axiom. would T nOt 8°inS l° pr°Ve the maximum Principle in this book; that «inni US to° far afield- For thoSe interested, a proof is outlined in the suPPlementary exercises. on a "etfi"al remark- We have defined what we mean by a strict partial order Partial ' rf W° haVC not Said what a Partial order itself is- Let ~< be a strict *• Then th °n * SCt A' Suppose that we define a ^ b if either a -< b or a = 6 rdti l rf *• Then th °n * SCt A' Suppose that we define a ^ b if either a -< b or a = sion relatj6 rdation ^ is called a Partial order on A. For example, the inclu- inclusion '°n <= on a collection of sets is a partial order, whereas proper >s a strict partial order.
72 Set Theory and Logic Chap. J Many authors prefer to deal with partial orderings rather than strict nartial orderings; the maximum principle is often expressed in these terms. Which formulation is used is simply a matter of taste and convenience. Exercises 1 If a and b are real numbers, define a < b if b - a is positive and rational. Show ' this is a strict partial order on R. What are the maximal simply ordered subsets? 2. Complete the discussion in Example 1. 3. (a) Let < be a strict partial order on the set A. Define a relation on A by letting a < b if either a < b or a = b. Show that this relation has the following properties, which are called the partial order axioms: (i) a < a for all a e A. (ii) a <, b and b <, a => a = b. (iii) a < b and b < c =* a < c. (b) Let P be a relation on A that satisfies properties (i)-(iii). Define a relation S on A by letting aSb if a/Y> and a ^ b. Show that S is a strict partial order on A. 4. Let A be a set with a strict partial order <; let x e A. Suppose that we wish to find a maximal simply ordered subset C of A that contains x. One plausible way of attempting to define C is to let C equal the set of all those elements of A that are comparable with x; C — {y\y e A and either x < y or y < x}. But this will not always work. In which of Examples 1 and 2 will this procedure succeed and in which will it not? 5. Complete the proof of the maximum principle in the case where A — discount- discountable, by showing that B u {a,\h(j) = 0] is a maximal simply ordered subset of A, *6. Given two points (x0, y0) and (*„ y,) of R\ define (Xo,yo)< (xt,yi) if x0 < xi and yo<,yi. Show that the curves y = x3 and y = 2 are maximal simply ordered subsets of R\ and the curve y = x1 is not. Find all maximal simply ordered subsets. * Supplementary Exercises: Well-Ordering aJ]h!le/OlkTn8 exerdSes> w »* you t0 Prove- w«hout using the choice axiom, !lll f,an "ncountab'e well-ordered set. We also ask you to prove the EfT 1CC aX1'°m' the ^"-oKkring theorem, and the maximum
2 Supplementary Exercises: Well-Ordering 73 l. that the proof of the principle of transfinite recursive definition does not 21 choice axiOm. (See Exercise .2 of §.-«00 • o the choice axiom, prove the following: Without us,ng he en wdl_ordered sets. lfJ has the order type of a subset Xn7 Proof. Let e0 be a fixed element of E Define A : 7 -> £ by the formula (smallest (E - h(Sj) if h(S.) =£ E, ) \ /I(a) = \eo fa) By hypothesis, there is an order-preserving map /: 7 — £. Show that we have WO < i(a) for all <X e /; conclude that A satisfies the formula = smallest (E — h(S,)) for all a. (b) Show that A(Sa) is a section of E for each a; conclude that A(/) equals E or a section of f. (c) Show that h is order preserving. (d) Show that if k : J —♦ E is any order-preserving map such that AG) equals £ or a section of E, then k satisfies (*) and hence is unique. 3. Without using the choice axiom, prove the following: Theorem. If A andB are two well-ordered sets, then either they have the same order type, or one has the order type of a section of the other. [Hint: Consider the well-ordered set constructed in Exercise 8(a) of §1-10.] Corollary. If A and B are two uncountable well-ordered sets all of whose sections are countable, then A and B have the same order type. 4. Without using the choice axiom, construct an uncountable well-ordered set, as follows. Let a be the collection of all pairs (A, <), where A is a subset of Z+ and < is a well-ordering of A. (We allow A to be empty.) Define (A, <) ~ ^.' 5'^ ""^' <^ an<* (^'> </) 'nave tne same order type. It is trivial to show this is an equivalence relation. Let [(A, <)] denote the equivalence class of (A, <); let E denote the collection of these equivalence classes. Define {.(A, <)] < [{A', <')] <f W, <) has the order type of a section of (A', <'). (a) Show that the relation « is well defined, and is a simple order on E. Note (V\ Qh thC equivalence class [@,0)] is the smallest element of E. W bhow that if a = [(A, <)] is an element of E, then (A, <) has the same order VPe as the section Sa(E) of E by a. [Hint: Define a map /: A — £ by set- to ' e ',W = KJ*(/Q, restriction of <)] for each x e A.] (I erClUde that E is well-ordered by «. gim M E iS uncountab'e- Mint: If *: IT-. Z+ is a bijection, then A Sjves rise to a well-ordering of Z+J Proves'S(wT afrgument' with z+ rePlaced by an arbitrary well-ordered set X, whose cZ r USC °f the Choice axiom) the ex»stence of a well-ordered set E * cardinality )s greater than that of AT.
Set Theory and Logic 5 Let A- be a set; let a be the collection of all pairs (A, <), where A is a subset of A" and < is a well-ordering of A. Define (A, <) < (A', <0 if (A, <) equals a section of (A', <')• (a) Show that < is a strict partial order on <*• rb Let CB be a subcollection of a that is simply ordered by <. Define B' to be the union of the sets B, for all (B, <) e CB; and define <' to be the union of the relations <, for all (B, <) e CB. Show that (B', <') is a well-ordered set. 6. Assuming Exercises I and 5, prove the following: Theorem. The maximum principle is equivalent to the well-ordering theorem. 7. Assuming Exercises 1-3 and 5, prove the following: Theorem. The choice axiom is equivalent to the well-ordering theorem. Proof. Let X be a set; let c be a fixed choice function for the nonempty sub- subsets of X. If T is a subset of X and < is a relation on T, we say that G", <) js a tower in X if < is a well-ordering of 7" and if for each .v e 7", x = c(X - 5,G")), where 5,G") is the section of 7" by .v. (a) Let G",, <:) and G\, <2) be two towers in X. Show that either these two ordered sets are the same, or one equals a section of the other. [Hint: If T, and T2 are well-ordered sets and h : 7", —> T2 is order preserving and h(T{) equals T2 or a section of 7, then h(Sx(Tt)) — Sh(x)(T2) for all x. If 7": and 7 are towers, then h(x) = ,v for all .v.] (b) Let {(Tk, <k)\k e K] be the collection of all towers in X. Let T={Jk&KTk and < = U*£«(<*)- Show that (r, <) is a tower in X Conclude that T = X.
2. Topological Spaces and Continuous Functions The concept of topological space grew out of the study of the real line andeudidean space and the study of continuous functions on these spaces. In this chapter we define what a topological space is, and we study a number of ways of constructing a topology on a set so as to make it into a topological space. We also consider some of the elementary concepts associated with topological spaces. Open and closed sets, limit points, and continuous func- functions are introduced as natural generalizations of the corresponding ideas for the real line and euclidean space. 2-1 Topological Spaces in ?edefinltlon of a topological space that is now standard was a long time other'- formulated- Var<ous mathematicians—Frechet, Hausdorff, and decades Pf°P0Sed d'fferent definitions over a period of years during the first on theS 'S CentUry> but il took q^^ a while before mathematicians settled thatWa°ne k seemed most su<table. They wanted, of course, a definition the vario^ d aS possible' so that it would include as special cases all eXamples tht l the vario p' at it would include as special cases all '"finite-dim eXamples that were useful in mathematics—euclidean space, ensional euclidean space, and function spaces among them— 75
76 Topological Spaces and Continuous Functions Chap. but they also wanted the definition to be narrow enough that the standard theorems about these familiar spaces would hold for topological spaces in general. This is always the problem when one is trying to formulate a new mathematical concept, to decide how general its definition should be. The definition finally settled on may seem a bit abstract, but as you work through the various ways of constructing topological spaces, you will get a better feeling for what the concept means. Definition. A topology on a set X is a collection 3 of subsets of X having the following properties: A) 0 and X are in 3. B) The union of the elements of any subcollection of 3 is in 3. C) The intersection of the elements of any finite subcollection of 3 is in 3. A set X for which a topology 3 has been specified is called a topological space. Properly speaking, a topological space is an ordered pair (X, 3) consisting of a set X and a topology 3 on X, but we often omit specific mention of 3 if no confusion will arise. If X is a topological space with topology 3, we say that a subset U of X is an open set of X if U belongs to the collection 3. Using this terminology, one can say that a topological space is a set X together with a collection of subsets of X, called open sets, such that c and X are both open, and such that arbitrary unions and finite intersections of open sets are open. Example 1. Let X be a three-element set, X = [a, b, c}. There are many possible topologies on X, some of which are indicated schematically in Figure 1. The diagram in the upper right-hand corner indicates the topology in which Figure I
Topological Spaces 77 §2-1 Y 0 ta b], {h}, and [b, c}. The topology in the upper left- the open sets are^^ only' x and 0, while the topology in the lower right- hand corner ^^.^ every subset of x. You can get other topologies on X by permuting ^a^p]e you can see that even a three-element set has many Fr°mt oologies. But not every collection of subsets of X is a topology on difl!,re"herofthe collections indicated in Figure 2 is a topology, for instance. Figure 2 Example 2. If Xis any set, the collection ofa//subsets of Xis a topology on r- it is called the discrete topology. The collection consisting of X and 0 only is also a topology on X; we shall call it the indiscrete topology, or the trivial topology. Example 3. Let X be a set; let 3/ be the collection of all subsets U of X such that X — U either is finite or is all of X. Then 3/ is a topology on X, called the finite complement topology. Both X and O are in 37, since X — X is finite and X — 0 is all of X. If {(/»} is an indexed collection of elements of 3f, to show that \JUX is in 37, we compute The latter set is finite because each set X — U. is finite. If U,, . .., (/„ are elements of 37, to show that f^\U, is in 37, we compute x-m.i i// = Ui"=i(*-i/i). The latter set is a finite union of finite sets and therefore finite. Example 4. Let X be a set; let 3C be the collection of all subsets U of X such that X-U either is countable or is all of X. Then 3C is a topology on X, as you can check. 3, ^finit'on. Suppose that 3 and 3' are two topologies on a given set X If => 3, we say that 3' is finer than 3; if 3' properly contains 3, we say that 3 * strictly finer than 3. sdp-,6alS0say that 3 is coarser than 3', or strictly coarser, in these two re- Wive situations. S0metWniT 'S suggested by linking of a topological space as being Elections f * a truckload ful1 of gravel—the pebbles and all unions of smallerone Pebbles being the open sets. If now we smash the pebbles into liice the graS> i Co!lection of °Pen sets has been enlarged, and the topology, ve|. is said to have been made finer by the operation.
78 Topologkal Spaces and Continuous Functions Chap.2 Two topologies on X need not be comparable, of course. In Figure , preceding, the topology in the upper right-hand corner is strictly finer than each of the three topologies in the first column and strictly coarser than each of the other topologies in the third column. But it is not comparable with any of the topologies in the second column. Other terminology is sometimes used for this concept. If 3' =, a, Some mathematicians would say that 3' is larger than 3, and 3 is smaller than 3'. This is certainly acceptable terminology, if not as vivid as the words "finer" and "coarser." Many mathematicians use the words "weaker and "stronger" in this context. Unfortunately, some of them (particularly analysts) are apt to say that 3' is stronger than 3 if 3' zd 3, while others (particularly topologists) are apt to say that 3' is weaker than 3 in the same situation! If you run across the terms "strong topology" or "weak topology" in some book, you will have to decide from the context which inclusion is meant. We shall not use these terms in this book. 2-2 Basis for a Topology For each of the examples in the preceding section, we were able to specify the topology by describing the entire collection 3 of open sets. Usually this is too difficult. In most cases one specifies instead a smaller collection of subsets of X and defines the topology in terms of that. Definition. If A" is a set, a basis for a topology on A" is a collection (B of subsets of X (called basis elements) such that A) For each x e X, there is at least one basis element B containing x. B) If .v belongs to the intersection of two basis elements 5, and B2, then there is a basis element B, containing x such that B, <= 5, n B2- Definition. If (B is a basis for a topology on X, the topology 3 generated by (B is described as follows: A subset U of X is said to be open in * (that is, to be an element of 3) if for each x e U, there is a basis element B e (B such that x e B and B a U. Note that each element of (B is open in X under this definition, so that « c 3. It is easy to check that this collection of subsets of X is a topology on X. But first let us consider some examples. Example 1. Let (B be the collection of all circular regions (interiors of circles) in the plane. Then (B satisfies both conditions for a basis. The second condition is illustrated in Figure 3. In the topology generated by (B, a subset U 01 the plane is open if every x in U lies in some circular region contained in V.
Basis for a Topology 79 B\ • X Figure 3 Figure 4 Example 2. Let (&' be the collection of all rectangular regions (interiors of rectangles)in the plane, where the rectangles have sides parallel to the coordinate axes Then (&' satisfies both conditions for a basis. The second condition is illustrated in Figure 4; in this case, the condition is trivial, because the inter- intersection of any two basis elements is itself a basis element (or empty). As we shall see later, the basis (&' generates the same topology on the plane as the basis (B given in the preceding example. Example 3. If Xis any set, the collection of ail one-point subsets of X is a basis for the discrete topology on X. Let us check now that the collection 3 generated by the basis (B is, in fact, a topology on X. If U is the empty set, it satisfies the defining condition of openness vacuously. Likewise, X is in 3, since for each x e X there is some basis element B containing x and contained in X. Now let us take an indexed family [U.],^ of elements of 3 and show that U = U.« U. ^longs to 3. Given x e U, there is an index <x such that x e c/a. Since U, ■sopen, there is a basis element B such that x e B cz Ua. Then x e B and B c V, so that U is open, by definition. Now let us take two elements U, and C/2 of 3 and show that [/, n V2 such"* t0 3' G'Ven X £ U' ° Ul' choose a basis element *i containing x ^ at B, c; t/,; choose also a basis element Bz containing x such that ment fi2 s^ond condition for a basis enables us to choose a basis ele- and B 3 C°"taininS x such that ^3 <= Bl n 52. See Figure 5. Then x e B3 FinaH ° Ul> S° Ux ° Ul belongs t0 3> by definition. of eleme ^ WtSh°W by inducti°n that any finite intersection U, n • • • O Un n-1 a^S 'S '" 3- This fact is trivial for » = I; we suppose it true for ' and Prove it for „. Now _,) n
*v and Continuous Functions Chap, 2 Figure 5 By hypothesis, U, n • • • n (/„_, belongs to 3; by the result just proved, the intersection of Ul n • • ■ n £/„-, and i/» also belongs to 3. Thus we have checked that the collection of open sets generated by a basis (B is, in fact, a topology. Another way of describing the topology generated by a basis is given in the following lemma: Lemma 2.1. Let X be a set; let (B be a basis for a topology 3 on X. Then 3 equals the collection of all unions of elements of ($>, Proof. Given a collection of elements of (B, they are also elements of 3. Because 3 is a topology, their union is in 3. Conversely, given U e 3, choose for each x e U an element Bx of <B such that jc e Bx c U. Then U = LUf B*> so £/ equals a union of elements of (B. □ When topologies are given by bases, it is useful to have a criterion in terms of the bases for determining whether one topology is finer than another. One such criterion is the following: Lemma 2.2. Let (B and <&' be bases for the topologies 3 and 3', respectively, on X. Then the following are equivalent: A) 3' is finer than 3. B) For each x e X and each basis element B e <$> containing x, there is a basis element B' e ffl' such that x £ B' c B. Some students find this condition hard to remember. "Which way does tr™ S thCy aSk- II ^ be easier t0 remember if you recall the analogy between a topological space and a truckload full of gravel. Think
Basis for a Topology 81 §2-2 c u nebbles as the basis elements of the topology; after the pebbles are t d to dust the dust particles are the basis elements of the new topology, fjnew topology is finer than the old one, and each dust particle was con- conned inside a pebble, as the criterion states. Proof of the lemma. B) => A). Given an element U of 3, we wish to show V £ 3'. Let x e U. Since (B generates 3, there is an element B e (B h that x e B aU. Condition B) tells us there exists an element B' e (&' S"ch that xe B' c B. Then x e B' c U, so U e 3', by definition. S" (i) => B). We are given x e X and B e (B, with xe.fi. Now 5 belongs to 3 by definition and 3 c 3' by condition A); therefore, B e 3'. Since 3'is generated by «', there is an element B' e «' such that x e B' cz B. □ Example 4. One can now see that the collection (B of all circular regions in the plane generates the same topology as the collection (B' of all rectangular regions; Figure 6 illustrates the proof. We shall treat this example more formally when we study metric spaces. Figure 6 We have described in two different ways how to go from a basis to the topology it generates. Sometimes we need to go in the reverse direction, from a topology to a basis generating it. Here is one way of obtaining a basis for a given topology; we shall use it frequently. Lemma 2.3. Let X be a topologkal space. Suppose that e is a collection of open sets ofX such that for each open set VofXand each x in U, there is an element Cofe such that x £ C cz U. Then Q is a basis for the topology oj X. e P™f- We must show that 6 is a basis. The first condition for a basis is asy. Given x b X, since X is itself an open set there is by hypothesis an x v . ® SUCn that x e C cz X. To check the second condition, let are °n8 to C.' n C*' where c' and C2 are elements of 6. Since C, and C2 C in^"* S° 1S C' ° C*" Therefore, there exists by hypothesis an element j m l. such that x e C3 cz C, n C2. of X Tfe den°te thC toPol°gy on x generated by 6; let 3 be the topology • me preceding lemma shows that 3' is finer than 3. Conversely, since
82 Topologkal Spaces and Continuous Functions Vbn^x each element of 6 is an element of 3, so are arbitrary unions of elements of e. Therefore, by Lemma 2.1, 3' c 3. We conclude that 3' = 3. Q There are two interesting topologies on the real line R which can be de- described in terms of bases: Definition. If ® is the collection of all open intervals in the real line (a,b) = {x\a <x<b], the topology generated by (B is called the standard topology on the real line. If (B' is the collection of all half-open intervals of the form [a, b) = {x | a < x < b], where a < b, the topology generated by (&' is called the lower limit topology on R. When R is given the lower limit topology, we denote it by R,. It is easy to see that both (B and (B' are bases; the intersection of two basis elements is either another basis element or is empty. Whenever we consider R, we shall suppose that it is given the standard topology unless we specifically state otherwise. The lower limit topology will prove very useful to us in constructing counterexamples. The relation between these two topologies is the following: Lemma 2.4. The lower limit topology 3' on R is strictly finer than the standard topology 3. Proof, Given a basis element (a, b) for 3, and a point x of (a, b), the basis element [x, b) for 3' contains x and lies in (a, b). Therefore, 3' is finer than 3. On the other hand, given a basis element [x, d) for 3', there is no open in- interval (a, b) satisfying the condition x <= (a, b) <= [x, d); therefore, 3 is not finer than 3'. fj A question may occur to you at this point. Since the topology generated by a basis (B may be described as the collection of arbitrary unions of ele- elements of (B, what happens if you start with a given collection of sets and take finite intersections of them as well as arbitrary unions? This question leads to the notion of a subbasis for a topology. Definition. A subbasis S for a topology on A' is a collection of subsets of X whose union equals X. The topology generated by the subbasis S is defined to be the collection 3 of all unions of finite intersections of elements of §• We must of course check that a is a topology. For this purpose it will suffice to show that the collection (B of all finite intersections of elements of
Basis for a Topology 83 §2-2 • hasis for then the collection 3 of all unions of elements of (B is a topol- bv Lemma 2.1. Given x e X, it belongs to an element of g and hence 08I'n element of <&; this is the first condition for a basis. To check the second condition, let Bi=s,n ■■■ nsm and B2 = S\ n • ■ ■ n 5; be two elements of®. Their intersection B1nB1 = (S1n ■■■ nSjn(S\n •■■ n SI) is also a finite intersection of elements of §,, so it belongs to (B. Subbases will prove useful to us only occasionally, so we shall not study them in detail. Exercises 1. Let X be a topological space; let A be a subset of X. Suppose that for each x e A there is an open set U containing x such that U cr A. Show that A is open in X. 2. Consider the nine topologies on the set X — {a, b, c) indicated in Example 1 of §2-1. Compare them; that is, for each pair of topologies, determine whether they are comparable, and if so which is the finer. 3. Show that the collection 3C given in Example 4 of §2-1 is a topology on the set X. Is the collection 3» = {U\ X — U is infinite or empty or all of X] a topology on XI 4. (a) If {3.} is a collection of topologies on X, show that |"K. is a topology on X. Is (JO. a topology on XI (b) Let {3a} be a collection of topologies on X. Show that there is a unique smallest topology on ^containing all the collections 3a, and a unique largest topology contained in all 3a. to If *= {a, 6, c}, let 3, = [0, X, {a}, {a, b}} and 32 = {0, X, {a}, {b, c}}. Find the smallest topology containing 3, and 32, and the largest topology contained in 3j and 32. • Show that if a is a basis for a topology on X, then the topology generated by equals the intersection of all topologies on X that contain a. Prove the same " a is a subbasis. • Consider the following collections of subsets of R: ®i={(a6)|6} ^Zil^ers^b]lx\a<x<,b}, «« ~ «, u {B - K\B e «,}, where K = {l/«|« e Z+}, 3-{(«,+oo)|ae/f}, h(+) {|}
84 Topological Spaces and Continuous Functions Chap.j (B6 = {(-oo, a)\a e R), where (—00, a) -» {x|x < a}, (B7 = {B|/f- Sis finite}. (a) Show that each ffi, is a basis for a topology on R. (b) Compare these seven topologies with one another. (c) Show that (B5 u (B6 is a subbasis that generates the same topology as ffi 7. (a) Apply Lemma 2.3 to show that the countable collection <S>\ = [(a, b)\a < b, a and b rational} is a basis that generates the standard topology on R. (b) Show that the collection (B^ = {[o; b)\a < b, a and 6 rational] is a basis that generates a topology different from the lower limit topology on R. 2-3 The Order Topology If X is a simply ordered set, there is a standard topology for X, defined using the order relation. It is called the order topology; in this section we consider it and study some of its properties. Suppose that X is a set having a simple order relation <. Given elements a and b of X such that a < b, there are four subsets of X that are called the intervals determined by a and b. They are the following: (a, b) = {x I a < x < b], (a,b) = {x\a<x<b}, [a,b) = {x\a<^x<b}, [a,b) = {x\a<x<b}. The notation used here is familar to you already in the case where Xis the real line, but these are intervals in an arbitrary ordered set. A set of the first type is called an open intervai in X, a set of the last type is called a dosed intervai in X, and sets of the second and third types are called half-open intervals. The use of the term "open" in this connection suggests that open intervals in X should turn out to be open sets when we put a topology on X. And so they will. Definition. Let AT be a set with a simple order relation assume X has more than one element. Let (B be the collection of all sets of the following types: A) All open intervals (a, b) in X. B) All intervals of the form [o0, b), where aa is the smallest element (if any) of X. C) All intervals of the form (o, ba\, where b0 is the largest element (if \ _ C tr J any) of X.
§2-3 topology- The Order Topology 85 (& is a basis for a topology on X, which is called the order ogy If X has no smallest element, there are no sets of type B), and if X has n0 largest element, there are no sets of type C). One has to check that (B satisfies the requirements for a basis. First, te that every element x of A" lies in at least one element of (B: The smallest element (if any) lies in all sets of type B), the largest element (if any) lies in e.. sets of type C), and every other element lies in a set of type A). Second, te that the intersection of any two sets of the preceding types is again a set of one of these types, or is empty. Several cases need to be checked; we leave it to you. Example 1. The standard topology on R, as defined in the preceding section, is just the order topology derived from the usual order on R. Example 2. Consider the set R x Rin the dictionary order; we shall denote the general element of R x R by jc x y, to avoid difficulty with notation. The set R x R has neither a largest nor a smallest element, so the order topology on R x R has as basis the collection of all open intervals of the form (a x b,c x d) for a < c, and for a = c and b < d. These two types of intervals are indicated in Figure 7. The subcollection consisting of only intervals of the second type is also a basis for the order topology on R x R, as you can check. aXb cXd x „ Figure 7 3' V* P0Sitive inte§ers Z+ form an o^ered set with a smallest ° Ter t°P°1°?y °" Z+ * the diSCrete topolo8y' for ev^ °ne- = , I' the one-Point set M - (« - 1. « + 1) is a basis " = 1. the one-point set {1} = [1, 2) is a basis element. I1!? Se! X=^>V x z+ jI> *e dictionary order is another S6t With a smallest element- Denoting 1 x n by a m we can represent X by
* Topological Spaces and Continuous Functions Chap. 2 The order topology on * is not the discrete topology. Most one-point sets are w,t there is an exception-the one-point set {6,}. Any open set containing T^cSS^Sl^ ab0Ut *« (by definiti°n)' and any baSiS 6lement containing 6, contains points of the a, sequence. Definition If Jf is an ordered set, and a is an element of X, there are f0Ur subset! of X that are called the rays determined by a. They are the following: (—oo, ci\ = i^n-*" ~*~~~ Qj9 [a, + oo) = {*|*>«}, (-oo,fl] = {*|*<fl}. Sets of the first two types are called open rays, and sets of the last two types are called ciosed rays. The use of the term "open" suggests that open rays in JVare open sets in the order topology. And so they are. Consider, for example, the ray (a, +00). If X has a largest element b0, then (a, + 0°) equals the basis element (a, b0]. If X has no largest element, then (a, +00) equals the union of all basis elements of the form (a, x), for x > a. In either case, (a, +00) is open. A similar argument applies to the ray ( — 00, a). The open rays, in fact, form a subbasis for the order topology on X; we leave this for you to check. 2-4 The Product Topology on X x Y If X and Y are topological spaces, there is a standard way of defining a topology on the cartesian product X X Y. We consider this topology now and study some of its properties. Definition. Let X and Y be topological spaces. The product topology on X x Y is the topology having as basis the collection (B of all sets of the form [ U X V, where l/isan open subset of X and Kis an open subset of Y. Ck ?at ® iS a basis- The first condition is trivial, since XxY r nt' The SeCond condition is almost as easy, since the aDy tW° bai d th r and th I fl X Kl) ° {Ul X Vz) = iUi n ^ X Note that the collection ffi is not a topology on^x F. The union of the
The Product Topology on X X Y 87 §2-4 two rectangles pictured in Figure 8, for instance, is not a product of two sets, so it cannot belong to (B; but it is open in X X Y. Each time we introduce a new concept, we shall try to relate it to the concepts that have been previously introduced. In the present case, we ask: What can one say if the topologies on X and Y are given by bases? The answer is as follows: Theorem 4.1. Jf(& is a basis for the topology of X, and Q is a basis for the topology of Y, then the collection 20 = {BxC|Bs(B and C e 6} is a basis for the topology ofXxY. Proof. We apply Lemma 2.3. Given an open set W of X X Y and a point x X y of W, by definition of the product topology there is a basis element V X Ksuch that xy.yeUy.VcW. Because <S and e are bases for X and y, respectively, we can choose an element B of (B such that x e B c U, and an element C of e such that y e C c V. Then xxysByCcfV. Thus the collection £ meets the criterion of Lemma 2.3, so 30 is a basis for XXY. □ Example 1. We have a standard topology on R: the order topology. n>e product of this topology with itself is called the standard topology on h? h R1' U haS as baSis the collection of a11 products of open sets of R, "t the theorem just proved tells us that the much smaller coliection of all Products (a, b) x (c, d) of open intervais in R will also serve as a basis for the pci ogy of Ri. Each such set can be pictured as the interior of a rectangie in 2 of §22 thC Standard toPoi°8y on R2 is i^t the one we considered in Exampie
Topical Spaces and Continuous Functions Chap. 2 the product topology in terms of a sub- , , ,Kvy_,Ibe defined by the equation efinition. Let «,. * * ■* Definition let *2: * X 7-> 7 be defined by the equation The maps «, and », are caUed the projections of X X 7 onto its first and second factors, respectively. We use the word "onto" because «, and *2 are surjective (unless one of Te spaces X or Y happens to be empty in wh.ch case X X Y „ empty ^of VS - J^U is Precise, the set , x , which is open in X X 7. Similarly, if K is open in Y, then «2-'(K) = X X F, which is also open in X X Y. The intersection of these two sets is the set t/ x V, as indicated in Figure 9. This fact leads to the following theorem: Theorem 4.2. The collection § = {a-\\J)| U open in X} U {nl\V) \ V open in Y} is a subbasisfor the product topology on X X Y. (V) u Figure 9
The Subspace Topology 89 of dements of S, since Therefore, V X V belongs to 5', so that 3 c 5' as well. Q 2-5 JAe Subspace Topology Definition. Let X be a topological space with topology 5. If Y is a subset of X, the collection is a topology on 7, called the subspace topology. With this topology, Y is called a subspace of X; its open sets consist of all intersections of open sets of X with Y. It is easy to see that 3y is a topology. It contains 0 and Y because 0 =Y n 0 and Y = Y n X, where 0 and Xare elements of 3. The fact that it is closed under finite in- intersections and arbitrary unions follows from the equations (£/, n r) n • • • n (£/, n 7) = (£/, n • • • n £/„) n r, U.« (tf. n 7) = (U.« tf.) n 7. Lemma 5.1. If(& is a basis for the topology ofX, then the collection (BY = {Bn Y|B G (&} ^ a basis for the subspace topology on Y. Proof Given C/ open in X and given y e £/ O 7, we can choose an ele- element 2? of ffi such that j, G B czU. Then ysBnYczUnY. It follows Lemma 2.3 that (By is a basis for the subspace topology on 7. □ ^knlT n8 W'th a SPaCe ^ and a subsPace Y> one needs t0 be careful of y Tne US6S thC term "open set-" Does one mean an element of the topology If y isran e'ement of the topology of XI We make the following definition: t0 Y) if -?^aCe °f X' WC Say that a Set U ls open in Y (°r °Pen relative • wbsrt ofy°uf t0 thC t0P°l08y of Y> this imPlies in particular that it is There is ^ that V 'S °pen in Xlt lt belon8s t0 the topology of X. in X; a sPecial situation in which every set open in 7 is also open
90 Topological Spaces and Continuous Functions Chap. 2 Lemma 5.2. Let Y be a subspace ofX. IfV is open in Y and Y is open /„ X, then U is open in X. Proof. Since U is open in Y, U = Y O V for some set V open in X. Since Kand Fare both open in X, so is Y O K □ Example i. Consider the subset Y= [0, i] of the reai iine R, in the subspace topology. The subspace topology has as basis all sets of the form (a, b) n Y, where (a, b) is an open interval in R. Such a set is of one of the fol- following types: (a, b) if a and b are in Y, [0, b) if only 6 is in Y, (a,b)n Y= (flJ] ifonlyflisin Y, Y or 0 if neither a nor 6 is in y. By definition, each of these sets is open in Y. But sets of the second and third types are not open in the larger space R. Note that sets of the first three types are the basis elements for the order topology on Y. Thus we see that in the case of ihe set Y = [0, I], its subspace topology (as a subspace of R) and its order topology are the same. Lest you think this remark utterly trivial, here is a subset of R for which the two topologies are not the same: Example 2. Let Ybe the subset [0, i) u B) of R. in the subspace topology on Y the one-point set B} is open, because it is the intersection of the open set (J, §) with Y. But in the order topology on Y, the set B} is not open. Any basis element for the order topology on Y that contains 2 is of the form (a- I .v 6 Y and a < x < 2} for some a e Y; such a set necessarily contains points of Y less than 2. The anomaly illustrated in Example 2 does not occur if Fis an interval or a ray in the ordered set X; one has the following theorem, whose proof is left to the exercises: Theorem 5.3. If X is an ordered set in the order topology and if Y is an interval or a ray in X, then the subspace topology and the order topology on Y are the same. To avoid any ambiguity, let us agree that whenever we consider a sub- subset of an ordered set X, we shall assume that it is given the subspace topology unless we specifically state otherwise. Fortunately, there is no such ambiguity involving subspaces when one is dealing with the product topology: Theorem 5.4. If A is a subspace ofX and B is a subspace ofY, then the lZtZ AXB " "
The Subspace Topology 91 §2-5 f The set UX Vis the general basis element for X X Y, where U Pr°°f' JandVisopenin Y Therefore, A/ X V) n M X S) is the general pen m tl o i X « Now £ dement for the subspace topology on A X B. Now n A and V O B are the general open sets for the subspace topologies on A and B, respectively, the set (U O A) X {V O B) is the general basis °," ent for the product topology on A X B. The conclusion we draw is that the bases for the subspace topology on A x B and for the product topology on A X B are the same. Hence the topologies are the same. G Exercises 1 Show that if Y is a subspace of X, and A is a subset of Y, then the subspace topology on A as a subspace of Y is the same as the subspace topology on A as a subspace of X. 2. If 3 and 3'are topologies on A" and 3' is strictly finer than 3, what can you say about the corresponding subspace topologies on the subset Y Of XI 3. Consider the set Y = [— 1, 1] as a subspace of R. Which Of the following sets are open in Yl Which are open in /?? A = {x\l<\x\<\). = {x\i<\x\<\}, 1}, E = {x\0 <\x\ < 1 and l/x $ Z+). 4. A map/: X —» y is said to be an open map if for every open set U of X, the set /(t/) is open in K. show that n,: X x Y—* Xanditi : X X Y—* Y are open maps. • Let A" and A" denote a single set in the topologies 5 and 5', respectively; let r and y denote a single set in the topologies 01 and 01', respectively. W Snow that if 3' => 5 and 01' => 01, then the product topology on X' X Y' 'I er than the Product topology on X x Y. (I xvt* the COnverse of (a> h°ld? Justify your answer, w What can you say if 3' => 3 and 01' => 01 and 01' ^ 01? Show that the collection of open rays in an ordered set A is a subbasis fc5r" (b) U t0P°'°gy on ,4. in ty- M a" °rdered set in the order topology. Let Y be an interval or ray then' . (~°°' "> and (°. +°°) be open rays in X. Show that if a e I7, ' the nrH f the ^ .(~°°' O) n Y and (fl' +OO) n Y is an °Pen ray of of y red *** Y> whi'e if a £ r, each of these sets is either empty or all
92 Topological Spaces and Continuous Functions Ch»p. 2 (c) Conclude that if Y is an interval or ray in X, then the order topology tnd the subspace topology on Y are the same. 7. Show that the countable collection {(a, b) x\c,d)\a <b and c < d, and a, b, c, dare rational} is a basis for R1. 8. Show that the dictionary order topology on the set R x R is the same as the product topology Rd x R, where Rd denotes the set R in the discrete topology. Compare this topology with the standard topology on R2. 9. Let R denote the reals in their usual topology; let R, denote the reals in the lower limit topology. If/, is astraight line in the plane, describe the topology/, inherits as a subspace of R, x R and as a subspace of Rt x R,. In each case it is a familiar topology. 10. Let /denote the subspace [0, 1] of R. Compare the product topology on / x /, the dictionary order topology on / x /, and the topology Id x /, where /,, denotes / in the discrete topology. 2-6 Closed Sets and Limit Points Now that we have a few examples at hand, we can introduce some of the basic concepts associated with topological spaces. In this section we treat the notions of closed set, closure of a set, and limit point. These lead naturally to consideration of a certain axiom for topological spaces called the Hausdorff axiom. Closed Sets A subset A of a topological space X is said to be closed if the set X — A is open. ~ Example 1. The subset [a, b] of R is closed because its complement ! R — [", b] = (-oo, a) U (b, +oo), is open. Similarly, [a, +oo) is closed, because its complement (-oo, a) is open. These facts justify our use of the terms "closed interval" and "closed ray." The subset [a, b) of R is neither open nor closed. Example Z In the plane Ri, the set . ' . "■''-• [xXy\x^.O and y^.0] ...... ... I - . is closed, because its complement is the union of the two sets (-oo.O) XR and /Jx(-oo.O), each of which is a product of open sets of R and is therefore open in &■
Closed Sets and Limit Points 93 S*6 Examples. In the discrete topology on the set X, every set is open; it follows that every set is closed as well. Example 4. Consider the following subset of the real line: Y = [0, 1] U B, 3), •„ thesubspace topology. In this space, the set [0, 1] is open, since it is the inter- tv of the opeTset (-J, |) of R with Y. Similarly, B, 3) is open as a subset Tv it is even open as a subset of R. Since [0, 1] and B, 3) are complements in K of each other, we conclude that both [0, 1] and B, 3) are closed as subsets oiY These examples suggest that an answer to the mathematician's riddle, "How is a set different from a door?" should be, "A door must be either open or closed, and cannot be both, while a set can be open, or closed, or both, or neither!" The collection of closed subsets of a space X has properties similar to those satisfied by the collection of open subsets of X: Theorem 6.1. Let X be a topological space. Then the following conditions hold: A) 0 andX are closed. B) Arbitrary intersections of closed sets are closed. C) Finite unions of closed sets are closed. Proof. A) 0 and Xare closed, because they are the complements of the open sets Jfand 0, respectively. B) Given a collection of closed sets {Ax}xeJ, we apply DeMorgan's law, X -n.« A. = {J^,(X -A,). Since the sets X — Ax are open by definition, the right side of this equation represents an arbitrary union of open sets, and is thus open. Therefore, I 1-4, is closed. C) Similarly, if A, is closed for i = 1 n, consider the equation aJ ^'.l" the "ght side of this equation is a finite intersection of open sets a 1S therefore open. Hence \jAt is closed. Q »PacTbTd-°-f USm8 0Pe" SetS' °ne C0Uld inst as we" sPecify a topology on a threepronT8 afC°lleCti°n °f SetS (t0 be called "closed sets") satisfying the Pkments Ofi th'S theorem-One could then define open sets as the com- Particularad f ^^ a"d proceed just as before- This procedure has no Prefer tn n V " ge °ver the one We have adopted, and most mathematicians «3 use open sets to define topologies. .
Topological Spaces and Continuous Functions Chap. 2 Now when dealing with subspaces, one needs to be careful in using the term "closed set." If Y is a subspace of X, we say that a set A is closed in y if A is a subset of Fand if A is closed in the subspace topology of y (that is, if Y — A is open in Y). We have the following theorem: Theorem 6.2. Let Y be a subspace ofX. Then a set A is closed in Y if and only if it equals the intersection of a closed set ofX with Y. Proof. Assume that A = C n Y, where C is closed in X. (See Figure 10.) Then X - C is open in X, so that (X - C) n Y is open in Y, by definition of the subspace topology. But (X - C) n r = Y - A. Hence y - a is open in Y, so that ,4 is closed in Y. Conversely, assume that A is closed in y. (See Figure 11.) Then Y — A is open in Y, so that by definition it equals Figure 10 Figure 11 the intersection of an open set U of X with Y. The set X — U is closed in X, and A = Y r\ (X — U), so that A equals the intersection of a closed set of X with 7, as desired. □ A set A that is closed in the subspace Y may or may not be closed in the larger space X. As was the case with open sets, there is a criterion for A to be closed in X; we leave the proof to you : Theorem 6.3. Let Y be a subspace ofX. If A is closed in Y and Y is closed in X, then A is closed in X. Closure and Interior of a Set Given a subset A of a topological space X, the interior of A is defined as the union of all open sets contained in A, and the closure of A is defined as the intersection of all closed sets containing A.
Closed Sets and Limit Points 95 §2-6 ■ fPrior of A is denoted by Int A or by A, and the_ closure of'A is ThVh CM or by A. Obviously A is an open set and A is a closed set; denoted by furthermore, A cz A cz A. ■ nen A = A' while if A is closed, /I = A. UAwl shall'not make much use of the interior of a set, but the closure of will be quite important. "vflien dealing w,th a topological space X and a subspace Y, one needs erase care in taking closures of sets. If A is a subset of Y, the closure 'f Tin Kand the closure of A in X will in general be different. In such a ° ation, we reserve the notation A to stand for the closure of A in X. The T'ure of /I in Yean be expressed in terms of A, as the following theorem shows: Theorem 6.4. Let Y be a subspace ofX; let A. be a subset ofY; let A denote the closure of A in X. Then the closure of A in Y equals A n Y. Proof. Let B denote the closure of A in Y. The set A is closed in X, so /n Yis closed in Y by Theorem 6.2. Since A n Y contains A, and since by definition B equals the intersection of all closed subsets of Y containing A, we must have B <= {A n Y). On the other hand, we know that B is closed in Y. Hence by Theorem 6.2, B = C n Y for some set C closed in X. Then C is a closed set of X containing/!; because A is the intersection of all such closed sets, we conclude that A c C. Then (A D Y) cz (C n Y) = B. □ The definition of the closure of a set does not give us a convenient way for actually finding the closure of specific sets, since the collection of all closed sets in X, like the collection of all open sets, is usually much too big to work with. Another way of describing the closure of a set, useful because it involves only a basis for the topology of X, is given in the following theorem, irst let us introduce some convenient terminology. We shall say that M intersects a set B if the intersection A n B is not empty. (a) 7ThZem 6'5i Let Abea S"bset °fthe t0P°l°Sical sPace x- Xe ^ and onty if every open set U containing x intersects A. e topol°sy ofX is given by a basis, then x e A if and only element B containing x intersects A. tetrafder *e !tatement in (a)-II is a statement of the form PoQ. *el°gicall °™ CaCh imPlication t0 its contrapositive, thereby obtaining following: y eqmvalent statement (not P) o (not 0. Written out, it is the X at Jd ^s^^ , ■ ■ intersect ,* * CX1StS an °pen Set ^ containing x that does not
96 Topological Spaces and Continuous Functions Chap ^ In this form our theorem is easy to prove. If x is not in A, the set U = X - £ is an open set containing x that does not intersect A, as desired. Conversely if there exists an open set U containing x which does not intersect A, then X — U is a closed set containing A. By definition of the closure A, the set X — U must contain A; therefore, x cannot be in A. Statement (b) follows readily. If every open set containing x intersects A, so does every basis element B containing x, because B is an open set. Conversely, if every basis element containing x intersects A, so does every open set U containing x, because U contains a basis element that contains x. □ Mathematicians often use some special terminology here. They shorten the statement "U is an open set containing x" to the phrase "U is a neighborhood of x." Using this terminology, one can write the first half of the preceding theorem as follows: If A is a subset of the topological space X, then x e A if and only if every neighborhood of x intersects A. Example 5. Let X be the real line R. If A = @, 1], then A = [0,1], for every neighborhood of 0 intersects A, while every point outside [0, 1] has a neighborhood disjoint from A. Similar arguments apply to the following sub- subsets of X: If B = {l/n\n e Z+}, then i = [0|u8, If C = {0} U (I, 2), then C = {0} u [1, 2]. If Q is the set of rational numbers, then Q = R. If Z+ is the set of positive integers, then Z+ = Z+. If R+ is the set of positive reals, then the closure of R+ is the set R+ u {0}. (This is the reason we introduced the notation R+ for the set R+ u {0}, back in §1-2.) Example 6. Consider the subspace Y = @, 1] of the real line R. The set A = @, £) is a subset of Y; its closure in R is the set [0, £], and its closure in Y is the set [0, £] n Y = @, £]. Some mathematicians use the term "neighborhood" differently. They say that A is a neighborhood of x if A merely contains an open set containing x. We shall not follow this practice. Limit Points There is yet another way of describing the closure of a set, a way that involves the important concept of limit point, which we consider now. UA 1S a subset of the topological space X and if x is a point of X, we say that x is a limit point (or "cluster point," or "point of accumulation") . or a it every neighborhood of x intersects A in some point other than x itself-
Closed Sets and Limit Points 97 §2-6 *•* mntlv a: is a limit point of A if it belongs to the closure of A - {*}• Said dinere y, ^ ^ ^ nQt_ fof thjs definition it does not matter, •rjiepoint-t m«v mple 7 Consider the real line R. If A = @, 1], then the point 0 is a •t oint of A and so is the point |. In fact, every point of the interval [0, 1] hm' limit point of A, but no other point of R is a limit point of A. 'S If B = {l/"l" e Z+J' then ° is the only limit p0'nt °f B' Every °ther P°'nt TR has a neighborhood that either does not intersect B at all, or it intersects Ronly in the point x itself. If C = {0} u A, 2), then the limit points of C are the points of the interval [1,2]. If Q is the set of rational numbers, every point f R is a limit point of Q. If Z+ is the set of positive integers, no point of R is a limit point of Z+. If R+ is the set of positive reals, then every point of {0} u R+ is a limit point of R+- Comparison of Examples 5 and 7 suggests a relationship between the closure of a set and the limit points of a set. That relationship is given in the following theorem: Theorem 6.6. Let A. be a subset of the topological space X; let A' be the set of all limit points of A. Then A = A U A'. Proof. If x is in A', every neighborhood of x intersects A (in a point different from x). Therefore, by Theorem 6.5, x belongs to A. Hence A' c A. Since by definition A <= A, it follows that A U A' cz A. To demonstrate the reverse inclusion, we let x be a point of /fand show that* e A u A'. If x happens to lie in A, it is trivial that x e A U A'; suppose that x does not lie in A. Since x e A, we know that every neighbor- neighborhood U of x intersects A; because x $ A, the set U must intersect A in a Point different from x. Then x e /!', so that x <= A\J A',zs desired. □ 5,7- A subset Qj- a (opoIogical e is dosed y and . .r i( Wins all its limit points. '-T1* SCt A is closed if and on'y if -4 = ^T, and the latter holds if if A' c A. Q paces Une aTdn^06 W'th °Pen a"d C'°Sed SetS and Hmit points in the real topo'ogical C °an be misleadin8 when one considers more general always has fPa°es:.ForexamPle. «n the order topology on R the interval (a, b} etal. as a llmit PoJnt. But this is not true in order topologies in gen- another example, consider the fact that in R or R\ each one-point
98 Topological Spaces and Continuous Functions set {x0} is closed. This fact is easily proved; every point different fro a neighborhood not intersecting {.v0}, so that [x0] is its own cloT *' f i f bit tplil *1* has a a neighborhood not intersecting {.v0}, so that [x0] is its own cloT *' But this fact is not true for arbitrary topological spaces. Consider on*1*' the very first examples we had of a topological space; the tl ^f three-point set [a, b, c} indicated in Figure 12. In set b is not closed, for its complement is not open. Figure 12 Topologies like this one are not really very interesting to mathematicians for they seldom occur in other branches of mathematics. And the theorems that one can prove about topological spaces are rather limited if such ex- examples are allowed. Therefore, one often imposes an additional condition that will rule out examples like this one, bringing the class of spaces under consideration closer to those to which one's geometric intuition applies. The condition was suggested by the mathematician Felix Hausdorff, so math- mathematicians have come to call it by his name. Definition. A topological space X is called a Hausdorff space if for each pair xu x2 of distinct points of X, there exist neighborhoods Ut and U2 of xt and x2, respectively, that are disjoint. Theorem 6.8. Every finite point set in a Hausdorff space X is closed. Proof. It suffices to show that every one-point set {*„} is closed. If x is a point of X different from x0, then x and x0 have disjoint neighborhoods V and V, respectively. Since U does not intersect {*„}, the point x cannot belong to the closure of the set {x0}. As a result, the closure of the set [xa] is {*„} itself, so that it is closed. □ Theorem 6.9. Let X be a Hausdorff space; let A. be a subset of X. Then the point x is a limit point of A. if and only if every neighborhood ofx contains infinitely many points of A. Proof. If every neighborhood of x intersects A in infinitely many points, it certainly intersects A in some point other than x itself, so that x is a limit point of A. Conversely, suppose that x is a limit point of A, and suppose some neig ' borhood U of x intersects A in only finitely many points. Then U also inter- intersects A - {x} in finitely many points; let {x, xm) be the points o
Closed Sets and Limit Points 99 M)ThesetX-(x1,..>xm]isanopensetofX,sincethefinite --W- lSed;then n(X_{xi,...,Xm]) intersects the set A - [*J not at all. This contra- UOV--W- }iscl0Sed;then pointsetf*,,...> Un(X_ T. Hausdorff condition is stronger than is needed to prove Theorems T H ^9 The proofs would still hold if one assumed only the follow.ng 6^er condition, which is usually called the T, axiom: Given two distinct points a and b of X, each has a ne.ghborhood not containing the other. We shall give a few exercises involving this axiom; apart from that, we shall not use the T, axiom in this book. In order to prove many of the interesting theorems of topology one needs the Hausdorff condition anyhow. Furthermore, most of the spaces that are important to mathematicians are Hausdorff spaces. The following theorem, whose proof is left to the exercises, gives some substance to the latter remark. Theorem 6.10. Every simply ordered set is a Hausdorff space in the order topology. The product of two Hausdorff spaces is a Hausdorff space. A sub- space of a Hausdorff space is a Hausdorff space. For this reason the Hausdorff condition is generally considered to be a very mild extra condition to impose on a topological space. Indeed, in a first course in topology some mathematicians go so far as to impose this condition at the outset, refusing to consider spaces that are not Hausdorff spaces. We shall not go this far, but we shall certainly assume the Hausdorff condition whenever it is needed in a proof without having any qualms about limiting seriously the range of applications of the results. The Hausdorff condition is one of a number of extra conditions one can impose on a topological space. Each time one imposes such a condition, one ran prove stronger theorems, but one limits the class of spaces to which the wems apply. Much of the research that has been done in topology since ^ Jgmnings has centered on the problem of finding conditions that will satisfy"8 en°U8h t0 enab'e °ne t0 prove interesting theorems about spaces the Jn c conditions' and yet not so strong that they limit severely Je applications of the results. The Hau'lrf" T^ * nUmber of such conditions in the next two chapters. Editions ,C°ndltiOn 3nd thC Tl axiom are but two of a ejection of •*ww Othe t0°ne another that arc called collectively the separation Pac">ess andf C ltlons incl"de the countability axioms, and various com- ^ir Cmnectedness conditions. Some of these are quite stringent . as you will see. • .
700 Topological Spaces and Continuous Functions Chap. 2 Exercises 1 Let 6 be a collection of subsets of the set X. Suppose that 0 and X are in e ' and that finite unions and arbitrary intersections of elements of e are in e' Show that the collection 5 = {#- C\C e 6} is a topology on X. 2. Show that if A is closed in Y and Y is closed in X, then /( is closed in X. 3. Show that if A is closed in X and S is closed in Y, then /( xfiis closed in 4. Show that if U is open in X and A is closed in X, then (/ - ^ is open in X, and A — U is, closed in X. 5. Let X be an ordered set in the order topology. Show that (o, b) a. [a, b]. Under what conditions does equality hold? 6. Let A, B, and /4a denote subsets of a space X. Prove the following: (a) A U B = /T_u S. (b) \JA, => U^aJ give an example where equality fails. 7. Criticize the following "proof" that [JA, a \JA^: If {A*} is a collection of sets in X and if a: e U^** then every neighborhood U of .v intersects IJ'4.- Thus (/ must intersect some /<„, so that x must belong to the closure of some A.. Therefore, x e U^- 8. Let A, B, and Ax denote subsets of a space X; let A' denote the set of limit points of A. Determine whether the following equations hold; if an equality fails, determine whether one of the inclusions aorc holds. (a) A n B = A_n B. (b) fl^a = fVa- (c) A - B = A - B. (d) (A U B)' ■= A' \J B'. (e) (A n BY = A' n B'. 9. Let /( <= X and S c Y. Show that in the space X x Y A x B = A x B. 10. Show that every order topology is Hausdorff. 11. Show that the product of two Hausdorff spaces is Hausdorff. 12. Show that a subspace of a Hausdorff space is Hausdorff. 13. Show that X is Hausdorff if and only if the diagonal A = [x x x\x e X] is closed in X x X. 14. (a) Show that the T, axiom is equivalent to the requirement that finite point sets be closed. (b) Given a set X, show that the finite complement topology 3/, defined in example 3 of §2-1, satisfies the T, axiom and is contained in every Ti topology on X. Does 3r satisfy the Hausdorff axiom ?
Continuous Functions 101 {2 ,.n topologies on R given in Exercise 6 of §2-2. „ O*** «»f ^ c,osure of the set K - Vln\n e ZJ under each of these {a) De'ermine ,opologi«-these topologies satisfy the Hausdorff axiom? The T, axiom? (b) WhichJ ^ topo|ogies on R given in Exercise 7 of §2-2. Determine the and B- nder each of these topologies. et A" = [0 1] x [0. I] in the dictionary order topology. Determine "■Siosut'of the following subsets of X: A = {(\ln) x0|« e Z+}, B = {(| -l/«) x (\)\n e Z+], C = {x x0\0<x< 1}, D = {.v x j|0 < a: < 1}, £ = {£ x j|0 <j < I}. U, if ^ c X, we define the boundary of A by the equation Bd A = /I n (X- /(). (a) Show that Int A and Bd ^ are disjoint, and A = Jnt A u Bd ^. (b) Show that BdA = 0 o A is both open and closed. (c) Show that U is open ■£> Bd U =-■ 0 — U. (d) If U is open, is it true that U = Int {0I Justify your answer. 19. Find the boundary and the interior of each of the following subsets of R2: (t)A = {xXy\y = 0} (b) B = {* x y\x > 0 and y ^ 0} (c) C = ^uS (d) /) = {x x j|a: is rational} (e) £ = {a: x j|0<a:2 + j2 < 1} (f) f = {x x j|a: 9i 0 and y < 1/^} **" *u7?towski) Consider the collection of all subsets A of the topological space X. The operations of closure A —» A and complementation /4 — X - /* are wicuons from this collection to itself. how that starting with a given set A, one can form no more than 14 distinct (b) Fd applying these two operations successively. md a subset A of R (in its usual topology) for which the maximum of 14 B obtained. 2-7 Continuous Functions T1.^ C°ntinU0US function is basic to much of mathematics. cont "S °n thC feal 'ine aPPear in the first Pages of a"y calculus "uous functions in the plane and in space follow not far
102 Topological Spaces and Continuous Functions Chap^ behind More general kinds of continuous functions arise as one goes further in mathematics. In this section, we shall formulate a definition of continuity that will include all these as special cases; and we shall study various prcJ. erties of continuous functions. Many of these properties are direct general! zations of things you learned about continuous functions in calculus and analysis. Continuity of a Function Let X and / be topological spaces. A function /: X—> Y is said to be continuous if for each open subset V of /, the set f^\V) is an open subset of A". Recall that f'l(V) is the set of all points x of X for which f(x) e K; it is empty if V does not intersect the image set f(X) off. Continuity of a function depends not only upon the function/itself, but also on the topologies specified for its domain and range. If we wish to emphasize this fact, we can say that / is continuous relative to specific to- topologies on X and /. Let us note that if the topology of the range space Y is given by a basis <B, then to prove continuity of/it suffices to show that the inverse image of every basis element is open: The arbitrary open set V of Y can be written as a union of basis elements Then r\y) = \j.e,r\Bj, so that f'l(V) is open if each set /"'E«) is open. If the topology on Y is given by a subbasis §, to prove continuity of/ it will even suffice to show that the inverse image of each subbasis element is open: The arbitrary basis element B for Y can be written as a finite inter- intersection S, n • ■ • n 5, of subbasis elements; it follows from the equation /-1(B) = /-E,)n---n/-E.) that the inverse image of every basis element is open. Example 1. Let us consider a function like those studied in analysis, a real-valued function of a real variable," f:R >R. In analysis, one defines continuity of/via the "e-S definition," a bugaboo over the years for every student of mathematics. As one would expect, the e-0 definition and ours are equivalent. To prove that our definition implies the e-0 definition, for instance, we proceed as follows- Given x0 in R, and given e > 0, the interval V - (/(*„) - e, /(*o) + ()
Continuous Functions 103 s 2-7 , of the range space R. Therefore, /-»(>0 is an open set in the fc an open set of the £« » con(ains (he point ^ ■„ contains some bas.s domainfrT> aboTvo We choose <5 to be the smaller of the two numbers el^aLd6 i Ttenif |* - jc.1 < 3,the point x must be in (a, b), so that ^l^L^haUhl^^finltil^plSourdefinition is no harder; we leave it to you We shall return to this example when we study metric spaces. pv^ple 2 In calculus one considers the property of continuity for many kinS^f functions. For«xample, one studies functions of the following types: f. r ■> & (curves in the plane) f- r —> R} (curves in space) f. ri >. r (functions /(*, y) of two real variables) f. ri —v R (functions f(x, y, z) of three real variables) j. ri —j. ri (vector fields v(<, y) in the plane) Each of them has a notion of continuity defined for it. Our general definition of continuity includes all these as special cases; this fact will be a consequence of general theorems we shall prove concerning continuous functions on product spaces and on metric spaces. Example 3. Let R denote the set of real numbers in its usual topology, and let R, denote the same set in the lower limit topology. Let f.R >R be the identity function; /(*) = x for every real number x. Then/is not a continuous function; the inverse image of the open set [a, b) of R, equals itself, which is not open in R. On the other hand, the identity function g : R, —> R "continuous, because the inverse image of (a, b) is itself, which is open in R,. In analysis, one studies several different but equivalent ways of formulat- and L lhon of c°ntinuity. Some of these generalize to arbitrary spaces, finition" 3I! C°nsldered in the theorems that follow. The familiar "e-S de- denary ,J ™ "convergent sequence definition" do not generalize to arbi- spaces; they will be treated when we study metric spaces in §2-10. topolog'Cal sPaceS' let f: X — Y. Then the U) * Is continuous. Z Tl Tel A lX ^ f(A) '" Y> '^ ^ f" sh°w thanf(?^U?e that/ is c°ntinuous. Let A be a subset of X. XZA, then fix) e f(A). Let V be a neighborhood of
104 Topological Spaces and Continuous Functions Chap. f-HV) is an open set of X containing x; it must intersect Thei KinterScts /(^) in the point /(,). Hence /W 6 6 35 midC) Let B be closed in Y and let /( = /"(«). We wish to prove that i is closed in X; we show that A <= A._ By elementary set theory, We have f(A) a B. Therefore, if x is a poin^of A, ) S B so that x e/-E) = 1- Thus / e= /( as desired. C) ==> A). Let K be an open set in Y. Let B = r — K; then 5 is closed in Y. Since C) holds, /"'(S) <s closed in X. By elementary set theory, f-i{V)= f~\Y- B) = f-\Y)- f->{B)= X - f~\B), so that /"'(P) <s open, as desired. □ Homeomorphisms Let A" and rbe topological spaces; let/: X —♦ F be a bijection. If both the function/and the inverse function /-':r—> X are continuous, then/is called a homeomorphism. The condition that /"' be continuous says that for each open set U of X, the inverse image of U under the map f~x:Y—> A' is open in Y. But the inverse image of U under the map /"' is the same as the image of Uunder the map/ See Figure 13. So another way to define a homeomorphism is to say that it is a bijective correspondence/: A'—* Y such that /(£/) is open if and only if U is open. Figure 13 This remark shows that a homeomorphism/: *— Y gives us a bijective correspondence not only between X and Y but between the collections of dSLh °; andr°f Y- As a result' a"y Pr°Pe'ty °f * ^at is entirely ex- expressed m terms of the topology of X (that is, in terms of the open sets of Y Surf71 conespondence/, the corresponding property for the space Y. Such a property of Xis called a topological proper^ of X.
Continuous Functions 105 9 r have studied in modern algebra the notion of an isomorphism Y°U Teebraic objects such as groups or rings. An isomorphism is a between Correspondence that preserves the algebraic structure involved. bijecuve ^ COncept in topology is that of homeomorphism; it is a bijective ^"ondence that preserves the topological structure involved. /"WreSD suppose that f:X—Y is an injective continuous map, where are topological spaces. Let Z be the image set f{X), considered as of Y; then the function f: X—-Z obtained by restricting the asuu^F- bijective. If/' happens to be a homeomorphism of X with Z, ""say that the map /: X—> Y is a topological imbedding, or simply an imbedding, of A'in Y. Example 4. The function f:R —> R given by f(x) = 3x + 1 is a home- homeomorphism. See Figure 14. If we define g :/?—>/? by the equation then one can check easily that f(g(y)) = y and g(f(x)) = .v for all real numbers xandy. It follows that /is bijective and that g =/"'; the continuity of /and g is a familiar result from calculus. Example 5. The function F: (—1, 1) —> R defined by is a homeomorphism. See Figure 15. We have already noted in Example 9 of §1-3 that Fh a bijective order-preserving correspondence; its inverse is the function C defined by °{y) = 1 +A +4^2)" The fact that F is a homeomorphism can be proved in two ways. One way is onote that because F is order preserving and bijective, /^carries a basis element me order topology in (-1, I) onto a basis element for the order topology f(x) = 3x figure U I Figure 15
106 Topological Spaces and Continuous Functions q^ ^ !„ R and vice versa. As a result, F is automatically a homeomorphism « -1 'n with R (both in the order topology). Since the order topology ^ (_1,' 1) and the usual (subspace) topology agree, Fisa homeomorphiSm of(-l, 1) with/?. . A second way to show Fa homeomorphism is to use the continuity Of the algebraic functions and the square-root function to show that both F and G are continuous. These are familiar facts from calculus. Example 6. A bijective function /: X—> Y can be continuous without being a homeomorphism. One such function is the identity map g:Rt-^R considered in Example 3. Another is the following: Let S1 denote the unit circle, Sl ={x X y\.xz + y2 = 1], considered as a subspace of the plane R1, and let be the map defined by /(/) = (cos 2nt, sin 2nt). The fact that / is bijective and continuous follows from familiar properties of the trigonometric functions. But the function/" is not continuous. The image under/of the open set U = \p,\) of the domain, for instance, is not open in S1, for the point p =/@) lies in no open set V of Rz such that Vn S1 c f(U). See Figure 16. HU) 0 JL Figure 16 Example 7. Consider the function ^ : [0, 1)—>ri : function / of the preceding example by expanding the mni?he P«Ung example by expanding ^bedding ' " *" eXample °f a conti™°™ injective map that is not an Constructing Continuous Functions \lter9Thg COntinUOus functions from °T another? There are a number of methods used in anal-
Continuous Functions 107 §2-7 .. h some generalize to arbitrary topological spaces and others »* °V tudy first some constructions that do hold for general topological donOtTf,rrine consideration of the others until later, spaces, deterruie 72 (Rules for constructing continuous functions). Let X, Y, and l^%Ttan?funcUon) Iff.X^Y maps all ofX into the single point Vv then fis continuous. ^ (b) (Inclusion) If A is a subspace ofX, the inclusion function j : A —> X it continuous. (c) (Composites) Iff- X — Y and g : Y —» Z are continuous, then the map f • x —-» Z is continuous. (d) (Restricting the domain) If f: X — Y is continuous, and if A is a subspace of X, then the restricted function f|A:A—»Y is continuous. (e) (Restricting or expanding the range) Let f: X —> Y be continuous. If Us a subspace of Y containing the image set f(X), then the function g-X—♦ Z obtained by restricting the range of f is continuous. If' Z is a space hating Y as a subspace, then the function h : X —> Z obtained by expanding the range of f is continuous. (f) (Local formulation of continuity) The map f: X —► Y is continuous if X can be written as the union of open sets U. such that f j Ua /.s continuous for each a. (g) (Continuity at each point) The map f:X—> Y w continuous if for eachx e X a«rf eacA neighborhood V o/f(x), /Aere « a neighborhood UofX. such that f(U) <= V. If the condition in (g) holds for a particular point x of Z, We say that/ is continuous at the point x. nan'(a) Ut /W =>70 for every * in x- Let K be °Pen in Y- The set IK J equals JTor 0, depending on whether Kcontains >;„ or not. In either v*c) it is open. K then^'^) = VnA, which is open in A by defini- subspace topology. ^ ^ 0Pe" ^ Z the" r'(U) is °Pen in ^andZ-fe-'d/)) is open ^^ set theory. m the map /•. jT v^wequals the composite of the inclusion mapy :A-*X (e) Let ?• v- ~C th of which are continuous. iJ^^be_continuous. If /(JT) <= Z <= 7, we show that the obtained from/is continuous. Let B be open in Z. Then
Topological Spaces and Continuous Functions B=Z C\ U for some open set U of Y. Because Z contains the entire set f{X), /-(i/)=r'W. by elementary set theory. Since f''(U) is open, so is £-'(#)• To show h: *—> Z is continuous if Z has Y as a subspace, note th /; is the composite of the map f:X—> Y and the inclusion map j; Y—,7 (f) By hypothesis, we can write A' as a union of open sets V. such that f\U, is continuous for each a. Let V be an open set in Y. Then because both expressions represent the set of those points x lying in U, for which /(a-) e V. Since /| l/« is continuous, this set is open in V, and hence open in X. But so that /"'(K) is also open in X. (g) Let Kbe an open set of Y\ let .v be a point of f~l{V). Then f(x) e F, so that by hypothesis there is a neighborhood Ux of x such that f{Ux) c V. Then O, <= /"'(K). It follows that /"'(K) can be written as the union of the open sets Ux> so that it is open. □ Theorem 7.3 {The pasting lemma). Let X = A U B, where A and B are closed in X. Let f: A —> Y awe/ g : B —> Y be continuous. If f(x) = g(x)/or every x e An B, /Ae/z f and g combine to give a continuous function h: X —> Y, defined by setting h(x) = f(x) ;/x e A, aw/h(x) = g(x) ;/x e B. Proof. Let C be a closed subset of Y. Now /r'(C)=/-'(C)ur'(C), by elementary set theory. Since/is continuous,/"'(C) is closed in A and therefore closed in X. Similarly, g~'(Q is closed in B and therefore closed in X. Their union h~l(C) is thus closed in X. □ This theorem also holds if A and B are open sets in X; this is just a special case of the "local formulation of continuity" rule [Theorem 7.2@]- Example 8. Let us define a function h:/{—»/{ by setting _ [x for x ^ 0> ~ L/2 for i^O. Each of the "pieces" of this definition is a continuous function, and they agree on the overlapping part of their domains, which is the one-point set {0}. Since their domains are closed in R, the function h is continuous. (See Figure 1'•) ■ One needs the "pieces" of the function to agree on the overlapping part of their domains in order to have a function at all. The equations
§2-7 Continuous Functions 109 k(x) = f <x-2 forx<0, x + 2 for x ^ 0, ce do not define a function. On the other hand, one needs some on the sets A and B to guarantee continuity. The equations jx - 2 for x < 0, \x + 2 for x S; 0, instance, do define a function / mapping R into R, and both of the pieces °r continuous. But / is not continuous; the inverse image of the open set m 3), for instance, is the nonopen set [0, 1). '{X) Figure 17 Theorem 7.4 (Maps into products). Let f: A —♦ X X Y be given by the equation f(a) = (f,(a), f2(a)). Then Us continuous if and only if the functions fi : A —y X and f2 : A —>- Y are continuous. maps/, and/2 are called the coordinate functions of/. ^lcn^f* ,r~" X and *2: X X Y^ Y be Pr°Jections onto «T»(Cnj n Ty ia °rS' resPective'y- These maps are continuous: For n. Note ^1^7/^ K> and th£Se "^ ^ °Pen lf U and V fun«i onS are composites of continuous -~u, We shDw"ttot7"T u y*™"**' suPP°se that/, and/2 are •p x y.itsinvelL 7 h baS1S element ^ X K for the topology ThetefOre( WS('XI'. that is, if and only if /,(«) 6 y and y^j 6 ^
HO Topological Spaces and Continuous Functions q^ /-■(£/x v)=f;'(U)nfV(V). Since both of the sets /r W and fV{V) are open, so is their intersec- intersection. □ There is no useful criterion for the continuity of a map f:AXB~,x whose domain is a product space. One might conjecture that/is continuous if it is continuous "in each variable separately, but this conjecture is not true. (See Exercise 13.) Example 9 In calculus, a parametrized curve in the plane is defined as a continuous map /: [a, b] -> R2. It is often expressed in the form /(,) = (v@,K'));andone frequently uses the fact that/is a continuous function of /if both a'and y are. Similarly, a vector field in the plane v(.v, y) = P(.\; y)\ + Q(x, y)'i = (P(x, y), Q(x, y)) is said to be continuous if both P and Q are continuous functions, orequiva- lently, if v is continuous as a map of R2 into R2. Both of these statements are simply special cases of the preceding theorem. One way of forming continuous functions that is used a great deal in analysis is to take sums, differences, products, or quotients of continuous real-valued functions. It is a standard theorem that if/, g : X—<■ R are con- continuous, then/+g,/— g, and f-g are continuous, and f/g is continuous if g(x) jt 0 for all a:. We shall consider this theorem in §2-10. Yet another method for constructing continuous functions that is familiar from analysis is to take the limit of an infinite sequence of functions. There is a theorem to the effect that if a sequence of continuous real-valued func- functions of a real variable converges uniformly to a limit function, then the limit function is necessarily continuous. This theorem is called the Uniform Limit Theorem. It is used, for instance, to demonstrate the continuity of the trigonometric functions, when one defines these functions rigorously using the infinite series definitions of the sine and cosine. This theorem generalizes to a theorem about maps of an arbitrary topological space X into a metric space Y. We shall prove it in §2-10. Exercises 1. Prove that for functions /:/?_/?, the e_g definition of continuity implies the open set definition. 2. Suppose that/: *_ Y is continuous. If x is a limit point of the subset A of X, is it necessarily true that /(*) is a limit point of /(/*)? 3' Ym f ^ J("^enote a single set in the two topologies 3 and 3', respectively. "it i: X — X be the identity function.
Continuous Functions 111 SZ-7 ♦hat / is continuous o 3' is finer than 3. (a>Sh0Wh! is a homeomorphism o 3' = 3. (b) Sho* *atifine a topology 3. on It by adjoining to the usual basis of open (c) If" e " ° oint set {n}. Show that (R, 3,) and (R, 32) are homeomorphic, sets the one-p but that 3i 9* ^2- rand >'o e r, show that the maps/: * y() X y0 and ^0') = *0 X y areimbeddings. nd y(^ two ordered sets in the order topology. Show that if the 5. (») ** , x_^ y is bijective and order preserving, it is a homeomorphism. w n e 2V Suppose you are given that for each real number .v 2^0, there (W l unique real number b > 0 such that 6" = .v. Denote b by ^"/ .v . Show i R R defined b g(x) ' x is continuous isl unq that the function g: R+—> R+ defined by g(x) = v' x is continuous, (c) ( 1) u [0 o) of R Define/: X R by t that the function g: R+ + y g() v Let A1 be the subspace (-co, -1) u [0, oo) of R. Define/: X --* R by the equation +1 if-v < -1, Show that/is bijective and order preserving and continuous. Is/a homeo- homeomorphism ? {.Show that the subspace (a, b) of R is homeomorphic with @, 1), and the subspace [a, b] of R is homeomorphic with [0, ]]. 7. Find a function/: R —> R that is continuous at precisely one point. 8. (a) Suppose that/: R —» R is "continuous from the right," that is, lim,^+ f{x) = f(a), for each a e /f. Show that/is continuous when considered as a function from R, to R. (b) What sort of functions/: R —, /f are continuous when considered as maps from R to R,? As maps from Rt to R,? • Let y'bean ordered set in the order topology. Let / g ; X —» K be continuous. W Show that the set [x\fM <; ^W} is ciosed in X. *'Let*:Jr—y be the function 10 ^ at * 'S continuous- [««'• Use the pasting lemma.] that/|i,k.aC0llecti0n of subsets of X- let JT= U ^.- Let/: JT-. r; suppose (a) s^ IS ^ntrnuous for each a. bcontinuous6 C°llect'°n {/4-} is finite and ^ch set A' is closed' then/ clonsedanh,e!TPle Where the collect'on {^.} is countable and each A. is (c) AnIndex S "°l Continuous' a T f SEtS {A'} is sa!d t0 ^ locaUy fin!te ff each P°!nt x eighborhood that intersects Am for only finitely many values of
112 Topological Spaces and Continuous Functions Chap. 2 a. Show that if the family {A.} is locally finite and each A. is closed, then /is continuous. 11 Let / A -* B and g : C -* D be continuous functions. Let us define a map ' fx g:A x C—> B x D by the equation ) () fx g {fxg)(a x Show that / x ,? is continuous. Show that / x ,? 12 Let /=■■ A" x Y-* Z. We say that F is continuous in each variable separately ' if for each yQ in Y, the map //: X —> Z defined by /i(a) = f (* x Jo) is con- continuous, and for each xa in X, the map k : Y —> Z defined by *(>-) = /r(Xo x ^ is continuous. Show that if^rs continuous, then F\s continuous in each variable separately. 13. Let F: R x R —► R be defined by the equation Jaj/(a-2 + y1) lfxXy^OxO. F()[ (a) Show that .Fis continuous in each variable separately. (b) Compute the function g:R-^R defined by g(x) = F(x x x). (c) Show that Fis not continuous. 14. Let A c X; let/: A —» Y be continuous; let y be Hausdorff. Show that if/ may be extended to a continuous function g : A —> Y, then g is uniquely deter- determined by/ 15. Recall that a map /: X—> Y is an open map if for every set U that is open in X, the set f(V) is open in K. Which of the statements (a)-(O of Theorem 7.2 remain true if one replaces the word "continuous" throughout by the word "open"? 2-8 The Product Topology We now return, for the remainder of the chapter, to the consideration of various methods for imposing topologies on sets. Previously, we defined a topology on the product X X Y of two topolog- topological spaces. In the present section we generalize this definition to arbitrary cartesian products. There are two ways of generalizing the definition; the one that will later prove to be the more important we shall call the product to- topology. One way to impose a topology on a product space is the following; it is a direct generalization of the way we defined a basis for the product to- Pology on X x Y: Definition. Let [X.].6J be an indexed family of topological spaces. Let us take as a basis for a topology on the product space
The Product Topology 113 §2-8 X for each <x e /. The topology generated by this basis • ollection satisfies the first condition for a basis because IT x. 's 'ts ^'element- and it satisfies the second condition because the intersection basis elements is another basis element: This topology is not the most useful one for the product space n*«> as we S £'a second way to generalize the previous definition is to generalize the subbasis formulation of the definition. Let ft? • n«e^ %* —y x# be the function assigning to each element of the product space its /?th coor- coordinate, it is called the projection mapping associated with the index 0. Definition. Let §fl denote the collection S/» = [n}\Uf)\Uf open in Xp), and let § denote the union of these collections, The topology generated by the subbasis $ is called the product topology. In this topology J[,ej X, is called a product space. How does the product topology differ from the box topology? It is easier ^answer this question if we look at the basis <B that $ generates. The collec- V01181 -°f a11 fin'tC 'ntersect;°ns of elements of $. If we intersect - °ns of elements of $. If we intersect v on8'ng t0 ^ same one of the sets $B we do not get anything because 15 ^ahTaTT °f tW° elements of S" or of finitely many such elements, elements f" & ^M °f ^"' We get something new only when we intersect ** ^sctib^111 erent sets §/»• The typical element of the basis (B can thus ft0*theind " f°Ws: LetA Abea finite set of distinct indices ex set J<Md I* Vf, be an open set in Xfi, for / = 1,..., n. Then «anP, ''^n^n-nii'W ' . ' 911 element of®. - -..
114 Topological Spaces and Continuous Functions There is another way to describe this basis element which is particuiaH useful. Note that a point x = (*.) is in B if and only if its /?,th coordinate is in £/„, its jff2th coordinate is in Vfl, and so on. There is no restriction whatever on the ath coordinate of x if a is not one of the indices /?„.. o As a result, we can write B as the product * - B = II.e, U.> where i/« denotes the entire space X, if a ^ /?,,. . ., /?„. All this is summarized in the following theorem: Theorem 8.1 {Comparison of the box and product topologies). The box topology on Y[X. has as basis all sets of the form II U«, where U, is open in X. for each a. The product topology on ]7Xa has as basis all sets of the form ]TU., where U. is open in X,for each a, and U. e^ya/j X. except for finitely many values of a,. Two things are immediately clear. First, for finite products T[!Ui ■*« the two topologies are precisely the same. Second, the box topology is in general finer than the product topology. What is not so clear is why we prefer the product topology to the box topology. The answer will appear as we continue our study of topology. We shall find that a number of important theorems about finite products will also hold for arbitrary products if we use the product topology, but not if we use the box topology. As a result, the product topology is extremely important in mathematics. The box topology is not so important; we shall use it primarily for constructing counterexamples. Therefore: Whenever we consider the product T[ %•> we shall assume it is given the product topology unless we specifically state otherwise. Some of the theorems we proved for the product X x Y hold for the product JIX. no matter which topology we use. We list them here: Theorem 8.2. Suppose the topology on each space X. is given by a basis ©«. The collection of all sets of the form IL« B., where B. £ (Ba for each a,, will serve as a basis for the box topology on n€,x a f a,, will serve as a basis for the box topogy n.€,x.. The collection of all sets of the same form, whereh. £ (B. for finitely many indices a, and B. = X. for alt the remaining indices, will serve as a basis for the product topology on ]J.6, X.. Theorem 8.3. Let A. be a subspace of X., for each a £ J. Then IIA. is a subspace of J[X. if both products are given the box topology, or if both products are given the product topology. ■ .
The Product Topology 115 §2-8 8 4 If each space X. Is a Hausdorff space, then FIX. is a Hausdorff b the box and product topologies. ofs of these theorems follow the pattern of the proofs already f fv x Y, so the details are left to you. ? r no reason has appeared for preferring the product to the box i° It is when we try to generalize our previous theorem about con- topology. .nto product spacgs (hat a difference first arises. Here is a I that does not hold if II Xm is given the box topology: Theorem 8.5. Let f: A — II«e, X* be given by the equation f(a) = (f.(a)).6J, where f.: A — X, for each a. Let fl*. have the product topology. Then the function (is continuous if and only if each function fa is continuous. Proof. Let nf be the projection of the product onto its /?th factor. The function nf is continuous, for if Uf is open in Xf, the set Kj\Uf) is a subbasis element for the product topology on Xa. Now suppose that /: A — > J[X, is continuous. The function/^ equals the composite 7^ of; being the com- composite of two continuous functions, it is continuous. Conversely, suppose that each coordinate function/, is continuous. To prove that/is continous, it suffices to prove that the inverse image under/ of each subbasis element is open in A; we remarked on this fact when we defined continuous functions. A typical subbasis element for the product topology on J[X. is a set of the form n}\Uf), where ft is some index and V, is open in Xf. Now /~'(VW) =fi\V,), because ff = nf °/ Since /, is continuous, this set is open in A, as de- ared. □ Why dOes this theorem fail if we use the box topology ? Probably the most wincing thing to do is to look at an example. See Example 2 following. oJlbUrV- CDOnsider eucl'dean n-space R". A basis for R consists of all of the fom '" henCe a baS'S f0r the toP°lo§y of R" consists of all products sin m (fl>. *i) x (a2, bi) x • • • x (a., b,). *ett,reidera/?^n"eprt!dUCt> the box and product toP°'ogies agree. Whenever fica»y state otheroise ^^^ tha' " '* SiVe" th'S topology- unless we sPeci" C°DSider jR"' the countab'y infin«e product of /{ with
116 TopicalSpaces WContinuous functions Chap.2 v n for each n Let us define a function /: R -> #» by the equation where*,-* /(/) = (,,,,,,...); „,• ,P function of /is the function /„(/) = r. Each of the coordinate the/rthcoord.natefuncuo^^^ therefOre, the function/is continuous if functions/:*--; topology. But/is not continuous if R« is given the £.'££&'. <°- —* r TV1:::' ±±) (fJ) x r TV::: TV s = (_l,l) x (±,±) x (fJ) i WP assert that /"'(B) is not open in /?. If /"'(S) were f°r then? itTuSSn* n some interval (-*.*) about the point 0. This wouU open mi?, it would£o ^ ^ ^ appiying ^ to both sides Qf ^ .^.^ meantnai/u w, ^// S'thft'/^^^^so'^raPP^ng ^'to both sides of the for all n, a contradiction. 1. Prove Theorem 8.2. 2. Prove Theorem 8.3. 3. Prove Theorem 8.4. 4. Show that (X, x • • • x *„_,) x Xn is homeomorphic with X, x • • • X Xn. 5. Let {X,} be a family of spaces; let A, <= Xr for each a. (a) Show that if A. is closed in Xr, then J\AX is closed in n*.- (b) Show that IM. = IL*~ (c) Which of (a) and (b) remain true if you use the box topology instead of the product topology? 6. A sequence (xn) of points of X is said to converge to the point x of X if for every neighborhood U of x there is an TV such that xn e U for n > N. Let (xn) be a sequence of points of the product space n«e; *«• Show that the sequence (x.) converges to x if and only if the sequence 7r«(xn) converges to 7T«(x) for each a e /. Is this true if you use the box topology instead of the product topology • 7. Let R" be the subset of Rf» consisting of all sequences that are "eventually zero," that is, all (x,, x2,...) such that x, ^ 0 for only finitely many values of i. What is the closure of R" in R" in the box and product topologies ? Justify your answer. 8. Show that the product topology is the coarsest (smallest) topology on relative to which each projection function nf is continuous. 9. Let A be a set; let {X.}xSJ be an indexed family of spaces; and let {/«}«e/ ** an indexed family of functions /« : A —> X,. (a) Show there is a unique coarsest topology 3 on A relative to which each of the functions /« is continuous. ;
The Metric Topology 117 (b) U* ^ = [f;'(Uf)\U, is open in Xf}, , . c - I IS» Show that § is a subbasis for 3. and let J> uo/»- continuous relative to 3 if and only if each (A Show that a map g ■ f °g is continuous. -> II^« >« defined by the ectuatl0n Z denote the subspace /(/I) of the product space n*- Show that the image under/of each element of 3 is an open set of Z. 2-9 The Metric Topology One of the most important and frequently used ways of imposing a topology on a set is to define the topology in terms of a metric on the set. Topologies given in this way lie at the heart of modern analysis, for example. In this section, we shall define the metric topology and shall give a number of examples. In the next section, we shall consider some of the properties that metric topologies satisfy. Definition. A metric on a set X is a function d:Xx X—>R having the following properties: A) d(x,y)>0 for all x,y & X; equality holds if and only if x = y. B) d(x,y) = d(y, x) for all x, y e X. C) (Triangle inequality) d(x, y) + d(y, z) > d(x, z), for all x, y, z e X. Given a metric d on X, the number d(x, y) is often called the distance between x and y in the metric d. Given e > 0, consider the set Ux,e) = [y\d(x,y)<e} ofan points y whose distance from x is less than e. It is called the e-ball thk hn a* *' Sometimes we omit the metric d from the notation and write fall s.mply as B(x> f)> when no confusion win afise or v °n the set X' then the collection of all f-balls ;n/, ^J , and e > 0> is a basis for a apology on X, called the metric niauced by d. ^(JScteu r a basis is trivia1' s!nce * G B(x> 0 forany e > 0. "*the basis elem SD/°nd condition for a basis> we show that if y is a point y that« contain m"' 'f)'then there is a basIs element B(y, 8) centered at nea in B(x, €). Define 8 to be the positive number e - rf(x, y).
Topological Spaces and Continuous Then B(y, <*) <= B(x< f)' for lf Z which we conclude that Functions Chap. rf(., ) SeeNo8wIocheck the second condition for a basis, let Bx and fi2 be two basis ■ . nH M ve B n B,. We have just shown that we can choose posi- rs£ " ^:»« *■ «=* - SS' '^n; u"in8' be the smaller of*, and <J2, we conclude that B{y,5) cz B, n B2. Figure 18 Using what we have just proved, we can rephrase the definition of the metric topology as follows: A set U is open in the metric topology induced by d if and only if for each y G U, there is a 5 > 0 such that Bd{y, 5) c U. Clearly this condition implies that U is open. Conversely, if U is open, it contains a basis element B = Bd(x, e) containing y, and B in turn contains a basis element Bd(y, 5) centered at y. Example 1. Given a set X, define d{x, y) = 0 if x = y. It is trivial to check that d is a metric. The topology it induces is the discrete topology; the basis element B(x, 1), for example, consists of the point x alone. Example 2. The standard metric on the real numbers R is defined by the equation d(x,y) = \x-y\. t0, Check that d is a metric- ^e topology it induces is the same as the °£By: ^ bl l ln£ ^ bSS elemetU ("' 6> f0r tlle for the metric topology; indeed,
The Metric Topology 119 §2-9 and € = (b - a)/2. And conversely, each f-ball B{x, f) the interval (x — e,x + e). a >•„„ If A- is a topological space, X is said to be metrizable if there DCfi Zicdon the set X that induces the topology of X. A metric space jfmeXable space X together with a specific metric d that gives the to- topology of x- Many of the spaces important for mathematics are metrizable, but some not Metrizability is always a highly desirable attribute for a space to are " s'for the existence of a metric gives one a valuable tool for proving fheorems about the space. It is therefore a problem of fundamental importance in topology to rind conditions on a topological space which will guarantee that it is metrizable. One of our goals in Chapter 4 will be to find such conditions; they are ex- expressed there in the famous theorem called Urysohn's melrization theorem. Further metrization theorems appear in Chapter 6. In the present section we shall content ourselves with proving merely that R" and R'" are metrizable. Although the metrizability problem is an important problem in topology, the study of metric spaces as such does not properly belong to topology as much as it does to analysis. Metrizability of a space depends only on the topology of the space in question, but properties that involve a specific metric for X in general do not. Therefore, they are not topological properties. For instance, one can make the following definition in a metric space: Definition. Let Xbe a metric space with metric d. A subset A of X is said to be bounded if there is some number M such that B)< M for every pair au a, of points of A. If A is bounded, the diameter of A is defined to be the number diam/l = PaSrmetriclhf 7 g Ppty' for Jt dePends with metric "hen ,.? * """ * For inStanCe' if X is a metric P relativetowhth, I eX'St/ a metfiC d WWch giv6S the tOpol°gy <* x> which every subset of X is bounded. It is defined as follows! *>*eqZbn'UlXbe°me'rkSpaCe Wi melrkd- Defi"ea: X X X-, R u {(,y), 5 a melric <h" Educes the topology ofX. j, called th
120 Topological Spaces and Continuous Functions Chap. 2 Proof Checking the first two conditions for a metric is trivial. Let us check the triangle inequality: Now if either d(x y) 2;1 or d{y, z)^\, then the right side of this inequality U at least 1 • since the left side is (by definition) at most 1, the inequality holds. It remains io consider the case in which d(x, y) < 1 and d(y, z)< 1. In this case, we have d(x, z) < d(x, y) + d(y, z) = d(x, y) + d(y, z). Since d\x, z) < d{x, z) by definition, the triangle inequality holds for d. The fact that rfand rf induce the same topology follows from the inclusions Bd(x, e) cz B3(x, e), Bd(x, S) c= Bd(x, e), where 5 = min {e, 1}. We merely apply the following lemma, fj Lemma 9.2. Let d and d' be two metrics on the set X; let 3 and 3' be the topologies they induce, respectively. Then 3' is finer than 3 if and only if for each x in X and each e > 0, there exists a 5 > 0 such that Bd.(x, S) cz Bd(x, e). Proof. Suppose that 3' is finer than 3. Given the basis element Bd(x, e) for 3, there is by Lemma 2.2 a basis element B' for the topology 3' such that x e B' cz Bd(x, e). Within B' we can find a ball Bd(x, 5) centered at x. Conversely, suppose the S-e condition holds. Given a basis element B for 3 containing x, we can find within B a ball Bd(x, e) centered at x. By the given condition, there is a S such that Bd(x, 5) <= Bd(x, e). Then Lemma 2.2 applies to show 3' is finer than 3. □ Now let us show that R" and R" are metrizable. Definition. Given x = (x,,... , xn) in R", we define the norm of x by the equation llx||=M+ •••+x>)">; and we define the euclidean metric d on R" by the equation </<*,y) = ||x - y|| = [(x, - Viy + ... + (Xn _ y.yy>\ We define the square metric p by the equation The proof that d is a metric requires some work; it is probably already ramiiiar to you. If not, a proof is outlined in the exercises. We shall seldom have occasion to use the euclidean metric on R\ To show that p is a metric is easier. Only the triangle inequality is non-
The Metric Topology 'IL. From the triangle inequality for R it follows that for each positive integer'' ^ _ Zt\<\x,-y,\ + \y, - z,\. by definition of p, y \<(xy) + P(y*) AS a reSU ' p(x, z) = max [| x, - z, |} < />(x, y) + p(y, z), ^if'The real line /? = /?', these two metrics coincide with the standard •"for/? In the plane R\ the basis elements under dean be pictured as drcular regions, while the basis elements under p can be pictured as square 16 Each of these metrics induces the usual topology on R": Theorem 9.3- The topologies on R" induced by the euclidean metric d and the square metric p are the same as the product topology on R". Proof. Let x = (*,,..., *„) and y = 0',, . . . , yn) be two points of R". It is simple algebra to check that p(x, y) < d(x, y) < ^/Tp(x, y). The first inequality shows that Bd(x, e) cz Bf(x, €) for all x and f, since if d(x, y) < e, then p(x, y) < e also. Similarly, the second inequality shows that Bp(x> f/V"»~) c Bd(x, e) for all x and f. It follows from the preceding lemma that the two metric topologies are the same. s that given by the B = (al,bl) X ... x (aa,bn) choose e - ^ ~ '" *' ^ t tni'n,\\'' ff}-Then B>(x' f) c 5> as you can readily check, ert iZ (Tfu™ u™ ^ Pr°dUCt t0P°l0^ y\B>(*>e), we need to L? \ dement f°r the ^toPol°gy- Given s«chthat 6ed t0 find a basis element B for the product topology b
122 Topological Spaces and Continuous Functions Chap ^ But this is trivial, for Bf(x, e) = (x, — e, x, + f) x • • • x (xn — f, xn + e) is itself a basis element for the product topology. □ The preceding theorem is a metrization theorem for R\ We seek to gen- generalize it to R". As a first try for a metric on R", it is natural to try to generalize either the euclidean metric or the square metric, by defining and p(x,y) = lub{\xl-yl\}. But these formulas simply do not make sense on all of R°>; the series need not converge, and the set may not be bounded. We can avoid the difficulty with the metric p, however, by first replacing the metric \x — y\ on R by its bounded counterpart d. This idea gives us our second try for a metric on R°>. Suppose that we take the standard bounded metric on R, d(a,b) = min [\a — b\, 1); and, given two points x = (x,, x2,, . .) and y = (y,,y2, . ..) of/?", define It is easy to check that p is a metric on Rm. Unfortunately, however, it does not induce the product topology, as we shall see. Therefore, it is of no use for proving R" metrizable. It is, however, very important in its own right; it will appear frequently in this book (and in other parts of mathematics). There is nothing special about R°> as far as this metric is concerned; one can define it on RJ for arbitrary / as follows: Definition. Given an index set /, and given points x = (x.).ey and y = CvJ.ey of RJ, let us define a metric p on R1 by the equation />(x, y) = lub {d(x., y,) | a e /}, where d is the standard bounded metric on R. The metric p is called the uniform metric on RJ, and the topology it induces is called the uniform to- topology. The relation between this topology and the product topology is the fol- following: Theorem 9.4. The uniform topology on R} is finer than the product topology! they are different ifj is infinite.
The Metric Topology 123 §2-9 „ f SuDDOse that we are given a point x = (x.).e, and a product J?basement UU. about x. Let «„..., «„ be the .nd.ces for wh.ch y°V f Then for each a,, choose f, > 0 so that ' ' Bd(x.t, e.) c t/..; ^s we can do because t/., is open in R. Let e = min {f,,..., f.}; then B,(x, e) c n^.- For if z 's a Point of R' SUch that ^(x> Z^ < f> then ^" ^ < f f°r aU "' It is easy to show that these two topologies are different if/ is infinite; we leave it to you. □ We still have not found what we want, a metric that gives the product topology on R". But we are almost there. It turns out that by modifying the uniform metric only slightly one can obtain the desired metric: Theorem 9.5. Let d(a, b) = min{\a — b|, 1} be the standard bounded metric on R. If'x and y are two points ofR°>, define Then D is a metric that induces the product topology on R". Proof. The properties of a metric are satisfied trivially except for the tnangle inequality, which is proved by noting that for all /, whence iopology requires a ■ we find an open an f-ball BD(le) that llN<e-Finaiiy> iet lh°
124 Topological Spaces and Continuous Functions _ '-nap. 2 If y is in V, this expression is less than f, so that V c BD(x, e), as de ' Conversely, consider a basis element for the product topology, where U, is open in R for i = a,,. . ., an and U — R for all other indices i. Given x e U, we find an open set V of the metr? topology such that x e V cz U. Choose an interval (x, — f„ x, + f) ;n R centered about x, and lying in U, for i = a,, . . ., an; choose each e, ^ i Then define f = min {eji\i = a,,..., a,}. We assert that x £ 5c(x, f) cz U. Let y be a point of BD(x, e)■ Then for all i, Now if ;" = a a,,, then f < eji, so that d(xit y,) < f, < 1; it follows that \x, — y,\ < €,. Therefore, y e \\U,, as desired. □ It is natural to ask whether this theorem can be generalized still further. Can one perhaps prove that RJ is metrizable for arbitrary /? The answer is "no," as we shall see in the next section. The situation is no better if we use the box topology instead of the product topology; in fact, under the box topology not even Ra is metrizable, as we shall see. Exercises 1. (a) In R", define d'(.x, y) = l*i - y\ I + • • • + \x» - y,\. Show that d' is a metric that induces the usual topology of R". Sketch the basis elements under d' when « = 2. (b) More generally, given p > 1, define d'(.x,y) = [?,U,\xl-y,\>'V"> for x, y e R". Assume that d' is a metric. Show that it induces the usual topology on R". 2. Let A' be a set; let d be a metric on X. Show that the topology on X induced by d is the coarsest topology relative to which the function d: X X X—+ & 1S continuous. • , 3. Show that R X R in the dictionary order topology is metrizable. 4. (a) Compare the box, product, and uniform topologies on R°.
The Metric Topology 125 A() (,iU ,n which topologies do the following sequences converge? Wl-A,1,1.1....), x, -A,1,1,1,...). W2 =@,2,2,2,...), x,=@.i.i.i....), w, = @,0,3,3,...), x3 = @,0,4,-$,...), y, ='A,0,0,0,...), z, =0,1,0,0....). 5. Let /f" be the subset of if" consisting of all sequences that are eventually zero. What is the closure of R~ in R" in the uniform topology? Justify your answer. 6. Let p be the uniform metric on if". Given x = (*,, Xi,...) e R" and given 0<f < 1, let t/(x,e) = (x, - f, x, + f) x • • • x (xn - f, xn + f) x • • •. (a) Show that U(x, f) is not equal to the f-ball .ff/x, f). (b) Show that t/(x, f) is not even open in the uniform topology. (c) Show that Bf(x, f) = {Ji« U{x, 5). 7. Let A" be a subset of R°> having the property that for every pair x, y of points of X, the series converges. On X we have the three topologies it inherits as a subspace of R» w the box, uniform, and product topologies. We also have the topology given by the metric (The fact that d is a metric follows from Exercise 9 ) jLT ** ln 8eneral ab°Ut h th f0Ur toPol°g''es °a ^e set (b) What ^ wu say if X is the subset to «*. ^" = fei. x2,.. .)|x, = 0 for i W What can you say if A-is the subset canyousayifA-istheHIlbertcube '
126 Topological Spaces and Continuous Functions 8. Show that the euclidean metric d on R" is a metric, as follows: If x,y g », and c e R, define x + y = (a-, + yx xn CX = (C.V|, .. . ,CXn), x • y = -Y|.yi + ••• +-v,^. (a) Show that x • (y + z) = (x • y) + (x • z). (b) Show that |x • y| < ||x|| ||y ||. [Hint: If x, y jt o, let a = |/||x|| and b = l/||y||, and use the fact that \\ax ± 6y|| > 0.] (c) Show that ||x + y|| < l|x|| + ||y||. [Hint: Compute (x + y) . (x + y) and apply (b).] (d) Verify that d is a metric. 9. Let I1 denote the subset of ^consisting of all sequences (x,, x2,...) such that 2 x? converges. (You may assume the standard facts about infinite series. In case they are not familiar to you, we shall give them in Exercise 11 of the next section.) (a) Show that if x, y e C2, then £ l-v,>v I converges. [Hint: Use (b) of Exercise 8 to show that the partial sums are bounded.] (b) Let c e R. Show that if x, y e I1, then so are x 4- y and ex. (c) Show that </(x,y) = [££.(*,-tfJ]''2 is a well-defined metric on (,l. It is called the I1-metric. 10. Show that d'(x,y) = d(x,y)/(] + d(x,y)) is a bounded metric for X if d is a metric for X. 2-10 The Metric Topology (continued) In this section we discuss the relation of the metric topology to the con- concepts we have previously introduced. Subspaces of metric spaces behave the way one would wish them to," if A is a subspace of the topological space X and d is a metric for X, then the restriction of d to A X A is a metric for the topology of A. This we leave to you to check. About order topologies there is nothing to be said; some are metrizable (for instance, Z+ and R), and others are not. See Example 3. The Hausdorff axiom is satisfied by every metric topology. If x and y are distinct points of the metric space (X, d), we let f = %d(x, y); then the triangle inequality implies that Bd(x, f) and Bd(y, f) are disjoint. The product topology we have already considered in special cases; we have proved that the products R" and R<° are metrizable. It is true in general that countable products of metrizable spaces are metrizable; the proof fol- follows a pattern similar to the proof for R", so we leave it to the exercises.
The Metric Topology {continued) 127 Consideration ,nuous functions there is a good deal to be said. AboU^C°"n"ccupy the remainder of the section. of this topic win « c^tinuous functions on metric spaces, we are about When we siuj ^ calcu|us and analysis as we shall come in this book. 85 cl°Se "iTn Slings we want to do at this point. The«are nt *o show that the familiar "e-5 definition" of continuity FifSt> ^Jgeneral metric spaces, and so does the "convergent sequence definition ° ^°wa]]t tQ consjder two additional methods for constructing SeCOn id i §27 O i h ^wa] us functions, besides those discussed in §2-7. One is the process C°D"kL sums, differences, products, and quotients of continuous real- ^Ud functions'. The other is taking limits of uniformly convergent sequences of continuous functions. Theorem 10.1. Let f: X —> Y; let X and Y be metrizable with metrics i and d respectively. Then continuity of f is equivalent to the requirement tkt given x e X and given e > 0, there exists 3 > 0 such that dx(x,y)<<* =* dY(f(x), f(y)) < f. Proof. Suppose that/is continuous. Given x and f, consider the set /-Wto.0), which is open in Xand contains the point x. It contains some <J-ball B(x, 8) centered at x. If y is in this <J-ball, then f{y) is in the f-ball centered at f(x), as desired. Conversely, suppose that the e-S condition is satisfied. Let V be open in rjweshow that/-'(K)is open in X. Let x be a point of the set/-' (V). Since /We V, there is an f-ball B(f(x), e) centered at /(*) and contained in V. flax* C°ndltlOn' there is a ^-b311 5(^. ^) centered at x such that (uik C, 5(/W> f)- Then 5(*> J) is a neighborhood of x contained in / n so that /--(K) is open, as desired. Q w we turn to the convergent sequence definition of continuity. TT ^ 3 Sequence of P°'nts of a set X ^ simply a function S> 'tlS anelementof the space *». We usually denote (*.) or (x,,Xl,...). "" Point * o'f y^Uence (*» x^ • • •) of points of X is said to converge to ■*•* wch th«t r r 6Very neighborh°°d J7 of x there exists a positive in- nat x' "es in U for all i > iV. sequence (x.) converges to x, we write ■x.
Topological Spaces and Continuous Functions Chap ^ Of course, a sequence need not converge at all. But if it does converge jt converges to only one point, provided that X is Hausdorff. For if (*„) converge to x and if y ^ x, we need merely choose disjoint neighborhoods V and y of And y, respectively, and note that since U contains the points x, for all but finitely many values ofi, the set Vcannot. It is intuitively believable from one's experience in analysis that if x ijes in the closure of a subset A of the space X, then there should exist a sequence of points of A converging to x. This is not true in general, but it is true fOr metrizable spaces. Lemma 10.2 (The sequence lemma). Let X be a topological space; let A <= X. If there is a sequence of points of A. converging to x, then x e A; the converse holds ifX is metrizable. Proof. Suppose that xn —• x, where xn 6 A. Then every neighborhood U of x contains a point of A, so x e A by Theorem 6.5. Conversely, suppose that X is metrizable and x e A. Let d be a metric for the topology of X. For each positive integer n, take the neighborhood Bd(x, 1/n) of radius 1/n of x, and choose xa to be a point of its intersection with A. We assert that the sequence xn converges to x: Any open set U containing x contains an f-ball Bd(x, f) centered at x\ if we choose N so that \/N < e, then U contains x, for all i>N. O Theorem 10.3. Let f: X —► Y; let X be metrizable. The function f is con- continuous if and only if for every convergent sequence xn —► x in X, the sequence f(xn) converges to f(x). Proof. Assume that/is continuous. Given xn —► x, we wish to show that /(*«)—»/(*)• Let Vbe a neighborhood of f(x). Then f~l(V) is a neighbor- neighborhood of x, and so there is an N such that xn £ f~'(V) for n > N. Then /(*„) e V for n > N. To prove the converse, assume that the convergent sequence condition is satisfied. Let A be a subset of X; we show that f(A) <= f(A). If x e A, then there is a sequence xn of points of A converging to x (by the preced- preceding lemma). By assumption, the sequence /(*„) converges to f(x). Since /(*«) e /(/*), the preceding lemma implies that f(x) e /(^j. (Note that metrizability of 7 is not needed.) Hence f(A) c J\A), as desired. Q Incidentally, in proving Lemma 10.2 and Theorem 10.3 we did not use the full strength of the hypothesis that the space X is metrizable. All we really needed was the countable collection Bd(x, 1/n) of balls about x. This fact leads us to make a new definition. A space X is said to have a countable basis at the point x if there is a countable collection {t/n}nez, of neighborhoods of x such that any neighborhood U of x contains at least one of the sets Un. A space X that has a countable basis at each of its points is said to satisfy the first conntability axiom.
The Metric Topology {.continued) 129 countable basis [«.} at a-, then the proof of Lemma 10.2 goes [hZcsTZ\y replaces the ball B&. Un) throughout by the set thr0US ' B. = U,nU2n"'nU,. u nrnnf of Theorem 10.3 goes through unchanged. a trizable space always satisfies the first countability axiom, but the "I true, as we shall see. Like the Hausdorff axiom, the first counta- i is a requirement that we sometimes impose on a topological space n'orderto prove theorems about the space. We shall study it in more detail in Chapter 4. Now we consider additional methods for constructing continuous func- functions. We need the following lemma: Lemma 10.4. The addition, subtraction, and multiplication operations are continuous functions from R X R 1/7/0 R; and the quotient operation is a continuous function from R X (R — {0}) into R. You have probably seen this lemma proved before; it is a standard "f-<5 argument." If not, a proof is outlined in Exercise 12 below; you should have no trouble filling in the details. Theorem 10.5. IfX is a topological space, and if f, g : X -— R are continu- continuous functions, then f + g, f - g, and f-g are continuous. //g(x) * Ofor all x, then f/g is continuous. Proof. The map h : X -> R x R denned by "(x)=f(x)x g(x) Won operation fUnCt''°n ' + 8 ^ the Composite of + :RxR—+r; jis u g is continuous. Similar arguments apply to /- g, f.g, and Finally vve r- * functions. ^ t0 the notion »f ""iform convergence for a sequence Definition. Let f ■ y ££?spa- *£Z* th emaetSSr/y°w Unctf°,ns from the set x t0 »n,formIy to the func^en^ £r_JyWe say that the : > 0, tire forall«
r vprpence depends not only on the topology of y, but Uniformity of c°nv^ following theorem about uniformly conver- also on its metric, we rui also on its gent sequences: it theorem). Let fD: X -> Y be a sequence of X£z£"' """"-v" converges uniformly to t, men Proof, Let V be open in K; let *. be a point of /-("). We wish to find ihborhood ^ of^such^t /(t/)- Proof, Let V be open in ; . a neighborhood ^ of^such^t /(t/)^ ^ ^ ^ ^ ^ K. Then^using uniform convergence, choose N so that for all n > N and all xe X, d(f.(x),Ax))<e/4. Finally, using continuity of/*, choose a neighborhood 1/ of ^0 such that/,, carries U into the f/2 ball in Y centered at /*(■*„)• We claim that/carries £/ into B(y0, e) and hence into K, as desired. For this purpose, note that if x e U, then d{f(x),Mx)) < f/4 (by choice of N), d(fN(x)JN(xJ) < ell (by choice of I/), rf(/*(-*-o),/(*o)) < f/4 (by choice of N). Adding and using the triangle inequality, we see that d(f(x), f{x0)) < e, as desired. Q Let us remark that the notion of uniform convergence is related to the definition of the uniform metric, which we gave in the preceding section. Consider, for example, the space Rx of all functions /: X —> R, in the uni- uniform metric p. It is not hard to see that a sequence of functions/,: X—> R converges uniformly to/if and only if the sequence (/„) converges to/when they are considered as elements of the metric space (Rx, p). We leave the proof to the exercises. We conclude the section with some examples of spaces that are not tnzable metnzable. Example 1 r. /„ the box Wpohgy fa ^ melfUabk ^ t°W thtth the^etor t that-thC SeqUence lemma does not "old f°r *"• Let A •" tnesubset or*- cons.sting of those points all of whose coordinates are positive: , xl3.. .)|*, > o for all /
The Metric Topology {continued) 131 E ery coordinate,,, is positive, so we can construe, a basis element B' for the Apology on R by setting B' = (—-Vn, a'h) x (—xn.xzx) x •••. ■n. «' contains the origin 0, but it contains no member of the sequence TTthe p"canno. belong to B' because its ,,th coordinate *„ does not Sng to the interva. (-*„„, -O- Hence the sequence (..) cannot converge to 0 in the box topology. Example 2. An uncountable product of R with itself is not metrizable. Let J be an uncountable index set; we show that R1 does not satisfy the sequence lemma (in the product topology). Let A be the subset of R1 consisting of all points (.v«) such that x« — 0 for finitely many values of a and a-. = I for all other values of a. Let 0 be the "origin" in R', the point each of whose coordinates is 0, We assert that 0 belongs to the closure of A. Let l\Ux be a basis element containing 0. Then C/« ^ R for only finitely many values of a, say for (X = Hi,...,a,. Let (xa) be the point of A defined by letting xa = 0 for a = a,,..., <*„ and xa = I for all other values of a; then (a-,) g A O Y[U«, as desired. But there is no sequence of points of A converging to 0. For let an be a sequence of points of A. Each point an is a point of the product space having only finitely many coordinates equal to 0. Given //, let /„ denote the subset of /consisting of those indices a for which the ath coordinate of an is zero. The union of all the sets /„ is a countable union of finite sets and therefore countable. Because J itself is uncountable, there is an index in J, say /?, that does not lie m any of the sets yn. This means that for each of the points an, its /Jth coordinate n-<(Ln 'etr" te the °Pen interval (-1> ') in /?> and let U be the open set points '" hi ThC *' U 'S a ne!8nbornood °f 0 that contains none of the »; therefore, the sequence an cannot converge to 0. RecdUhar' V"*- Wf'ordered set S" » »°> metrizable in the order topology. §1-'O, and that ? !f C mmmal unc°untable well-ordered set constructed in l°8y, n is a Hmi" °teVhe set Sa U W- No(e first that, in the order topo- §> for the set S u'fOi ^ ^ 'S the reason we introduced the notation Staining n it m J '" §10-) For if (a> Q] is any bas's element • "ust intersect Sa; otherwise we would have although c • ft ^ v Itf j» »s« ■ uncountable and Sa is countable. -. . , , *
m Topoloeical Spaces and Continuous Functions ^ ^ We assert that there is no sequence of points of Sa converging to a. FOr ■f t O is^ny s quence of points of Sa, then the set {x | „ e Z+) 1S a countabl ,f (a.) is any seque^ ^^ ^ ^ upper bound & (ha( ,,es m ^ (Coro,,aiy » of Chapter"). Then (b,Q] is a basis element containing Q that contains no Joint whatever of the sequence (.vj. Exercises 1 Let A <= Xlfrfis a metric for the topology ofA", show that rf| ,4 x yf is a metric for the subspace topology on yf. X Let A" and Y be metric spaces with metrics ds and rfr, respectively. Let r. x—> rhave the property that for every pair of points x,, a-2 of X, Show that/is an imbedding. It is called an isometric imbedding of A" in Y. 3. Show that a countable product n^ of metrizable spaces is metrizable. [Hint: Let dn be a metric for Xn that is bounded by I. Define D(x,y) = lub {tUxn,y.)ln}.] 4. Show that the spaces Sa and R, satisfy the sequence lemma. (This does not, of course, imply that they are metrizable.) 5. Theorem. Let xn —» x andyn —► y in the space R. Then Xn + yn > x + y, xn - yn > x - y, Xnyn > xy, and provided that each yn^0 and y jt 0, xjyn —y x/y. [Hint: Apply Theorem 10.3 and Lemma 10.4; recall that if xn —* x and y, ~+ y, then xn x y, —► x X y.] 6' ?^"^ f"''[0> '] ~" R by the ecluat'on fix) = x". Show that the sequence AW) conver8es for each x e [0, I], but that the sequence (/„) does not converge uniformly. f^rnf^,3 **" ^ l6t/": ' -* * ** a sec»uence °f functions. Let p be the uni- ^ !lC°n '^ SP3Ce **• Show that the «Q«»««* (/») converges uniformly f ? ^ * r metric »»»«• Let /,: X-^ K be Let xn be a sequence of points of ^con- sequence (/»} conver^ -«to /•then
12-10 The Metric Topology {continued) 133 the function rsee Figure 19.) Let/: R — R be the zero function. L Show that/„(-*•) — /(-v) for each x e R. ' cjjfnv thatX does not converge uniformly to /. (This shows that the con- converse of Theorem 10.6 does not hold; the limit function/may be continuous even though the convergence is not uniform.) Figure 19 10. Using the closed set formulation of continuity (Theorem 7.1) show that the following are closed subsets of R2: 4 ={x X y\xy= 1}, S' ={■*• Xy\x2 +y2 = i], B2 ={x X y\x2 +y2<l}. u p S6t B1 is called 'he (closed) unit ball in R2. "^Show £SJ* StandKafd facts ab0"' '"finite series: „ * X'S l~* SCqUenCe °f rCal -*« a"d - ^ - to W ** a sa>uence of real numbers; define W ■<■» *, we say that the infinite series to -d 2c,
134 Topological Spaces and Continuous Functions q^ (d) Given a sequence of functions/.: * — R, let *„(*) = 2?= ■/<(•*>• Prove the Weierstrass M-test for uniform convergence: if I A*)I < b, for all x e X and all i, W 1/ the series 2>, converges, then the sequence fo.) «««»»« "/"/"""/y W a function s. [Hint: Let r. = £,"-.+ > 6" Show that if * > "' then l*kM ~ s^\ S r.; conclude that |*(.v) - j.(a)| < »•»•] IX Prove continuity of the algebraic operations on R, as follows: Use the metric d(a, b) = \a - b\ on R and the metric p((x, y), (-Vo, yo)) = max (l-v ~ -Yo I. \y ~ y» 13 (a) Show that addition is continuous. [Hint; Given f, let 5 = e/2 and note that d(x + y, x0 + yo) < \x - -vo I + \y - (b) Show that multiplication is continuous. [Hint: Given (.v0> .Vo) and f > 0, let _ 3<5 = min[f/(|xo| + \y01 + D, V f ] and note that d(xy, xoyo) <\xo\\y - yo\ + \y<,\\x - xo\ + \x - xo\\y - yo\.] (c) Show that the operation of taking reciprocals is a continuous map from R - [0} to R. [Hint: Given ,v0 ^ 0 and f > 0, let 5 = min (| a |/2, xJf/2}. Note that (/(I/a-, I/.y0) = |-v — xo|/|.v.vo|. If \x — x0 \ < 5, then \xx0 — xl\ < |-vo|2/2, so .vx0 — xl > —xo/2 and xx0 > xl/2 > 0.] (d) Show that the subtraction and quotient operations are continuous. *2~11 The Quotient Topology1! Unlike the topologies we have already considered in this chapter, the quotient topology is not a natural generalization of something you have already studied in analysis. Nevertheless, it is easy enough to motivate. One motivation comes from geometry, where one often has occasion to use "cut-and-paste" techniques to construct such geometric objects as surfaces. The torus (surface of a doughnut), for example, can be constructed by taking a rectangle and "pasting" its edges together appropriately, as in Figure 20. And the sphere (surface of a ball) can be constructed by taking a disc and collapsing its entire boundary to a single point; see Figure 21. Formalizing these constructions involves the concept of quotient topology. tThis section will be used when we prove the Jordan curve theorem in Chapter 8.
The Quotient Topology 135 b <L bk\ Figure 20 Figure 21 Definition. Let X and Y be topological spaces; let p : X— Y be a sur- surjective map. The map p is said to be a quotient map, provided a subset U of yisopen in 7if and only if/>-'(£/) is open in X. This condition is stronger than continuity; some mathematicians call it "strong continuity." An equivalent condition is to require that a subset A of Ybe closed in Y if and only if p~l(A) is closed in X. Equivalence of the two conditions follows from the equation Another way of describing a quotient map is as follows: We say that a subset C of X is saturated (with respect to the surjective map p: X —> Y) if C contains every set p'l{{y}) that it intersects. Thus C is saturated if it ^ as p 'Q>(C)). To say that p is a quotient map is equivalent to saying P is continuous and p maps saturated open sets of X to open sets of atUrated d°Sed SelS °f X t0 ClOSed setS of y)' R n Wnds °f 9uotient maps are the open maps and the closed h a map-(: * -"* y is said to be an °Pen maP if for each °Pen /(f/) is open in Y-ll is said t0 be a closed maP if for each iS °l0Sed in Y' U follows immediately from 's a surJective continuous map that is either ei*er oi*n ' p is a ^"otient map. There are quotient maps that are °I*n nor closed. (See Exercise 2.) . ....... open or closed ei*er ' p is a
136 Topological Spaces and Continuous Functions q^ 2 Definition. If X is a space and A is a se/ and if p : X -— ,4 is a surjective map then there exists exactly one topology 3 on A relative to which p \s a quotient map; it is called the quotient topology induced by p. The topology 3 is of course defined by letting it consist of those subsets V of A such that p~ l(U) is open in X. It is easy to check that 3 is a topology. The sets 0 and A are open because p'\0) = 0 and p-'(A) = X. The other two conditions follow from the equations Example 1. Let 7T, : A" x r — X be the projection map; n, is continuous and surjective. If U x K is a basis element for A" x Y, the image set n,(U x V)= U is open in X It follows readily that 71, is an open map. In general, tt, is not a closed map; the projection nt : R x R —> R carries the closed set [x x y\xy = 1} onto the nonclosed set R - {0}, for instance. Example 2. Let A"be the subspace[0, I] u [2, 3] of R, and let Ybe the sub- space [0, 2] of R. The map p: X —>Y defined by \x for x e [0, I], P(X)=U-1 ror.ve[2,3] is readily seen to be a surjective. continuous, closed map. It is not, however, an open map; the image of the open set [0, 1] of A" is not open in Y. Example 3. Let p be the map of the real line R onto the three-point set A = [a, b, c] defined by (a if x > 0, p(x) = \b ifx<0, [c if x = 0. You can check that the quotient topology on A induced by p is the one indi- indicated in Figure 22. Figure 22 There is a special situation in which the quotient topology occurs par- particularly frequently. It is the following: Definition. Let X be a topological space, and let X* be a partition of X into disjoint subsets whose union is X. Let p : X -+ X* be the surjective map that carries each point of JTto the element of X* containing it. In the quotient topology induced by p, the space X* is called a quotient space of X.
The Quotient Topology 137 12-11 •hematicians call X* a decomposition space or an identification /.tor they think of X* as having been obtained by "identifying all space ot *■ partition class to a single point." t"eelemen describe ihe topology of X* in another way. A subset U of X* ^Son of equivalence classes, and the «,p-(C/) is just the union of isaatence classes belonging to U. Thus the typical open set of X* is the """'ion of equivalence classes whose union is an open set of X. Example 4. Let X be the rectangle [0, 1] x [0, 1]. Define a partition X* Xas follows: It consists of all the one-point sets {.v x y] where 0 < x < 1 and 0 <y < '. the following types of two-point sets {.v x 0, x x 1} where 0 < x < 1, {0 x .y, 1 x ^} where 0 < y < 1, and the four-point set [0 x 0,0 x 1, 1 x 0, 1 x 1}. Typical open sets in A" of the form ;>-'(£/) are pictured by the shaded regions in Figure 23; each is an open set of X that equals a union of equivalence classes. x X 1 0X0 u /\ \ r \ V { 1 X y XOX y X X 0 Figure 23 xue3Ch °f theSe Sets UndeTp is an °Pen set of X*> as indicated in I description of X* is J«st the mathematical way of saying what S '" P'CtUreS Wh d he edges of a rectangle t08ether t0 P(V) p{W) Figure 24
138 Topological Spaces and Continuous Functions Chap.2 Example 5. Let X be the closed unit ball {* xy\x2+yz^l] in J?2 and let X* be the partition of X consisting of all the one-point sets [x x y} for which ** + y2 < 1, along with the set S» = {x x y |*» + y« * n Typical open sets in X of the form />-«(#) are pictured by the shaded regions in Figure 25. One can show that A'* is homeomorphic with the subspace of lp called the unit 2-sphere, defined by S* = {(x, y, z)\x2 +y2 +z* = I}. It is pictured in Figure 26. p{U) ■p(V) Figure 25 Figure 26 Now we explore the relationship between this new concept, that of quo- quotient topology, and the concepts we have treated previously. It is interesting to note that many of these concepts do not behave the way one might wish. It is not hard to see that subspaces do not behave well: If p: X —* Y is a quotient map and A is a subspace of X, then the map p': A —* p(A) obtained by restricting both the domain and range of p need not be a quotient map. See Example 6 below. If it should happen, however, that A is open in Xand p is an open map, then p' is a quotient map; the same is true if both A and/? are closed. This we leave to you to check. Example 6. Consider the subspace A = [0,1] u B, 3] of R; it is a subspace of the space X of Example 2. Suppose that we restrict the map p of Example 2 to A. Then q = p\A : A — [0, 2] is continuous and surjective, but it is not a quotient map. The set for instance, is closed in the domain space A; but A, 2] is not closed in the image space Y = [0, 2]. Composites of maps behave nicely; it is trivial to check that the composite of two quotient maps is a quotient map. On the other hand, the product of two quotient maps need not be a quo-
The Quotient Topology 139 1 *'" _ s and ^ : C — /) are quotient maps, it does not fol- v x c) =/>(«) x <?(c) is a <luotient maP' See Example defined W <* * «J conditions on either the maps or the spaces in order 8 One needStrue Onesuch, a condition on the spaces, is called heal com- for this to ru • ^ exercises of §3-8.) Another, a ^erncw. wc ma is the condition that both the maps p and q be condition of^ ^ ^ .^ ^ ^ {q ^ ^ p x q is also an open map, *iUS*^"continuousfunctions on quotient spaces there is something to be vfwhen we studied product spaces, we had a criterion for determining h map/:Z— \JX. into a product space was continuous. Its coun- " ^ in the theory of quotient spaces is a criterion for determining when f. x* —- Z out of a. quotient space is continuous. One has the follow- following theorem: Theorem 11.1. Let p: X —» Y be a quotient map. Let Z be a space and let g: X — Z be a continuous map that is constant on each sei p'' ({y)), for y e Y. Thin g induces a continuous map f: Y —• Z such that f ° p — g. Y—~ f Proof. For each y e Y, the set #(/r'({>})) is a one-point set in Z (since fisconstanton /r'(M). If we let /(>■) denote this point, then we have defined a map/; Y-^Z such that for each x e X, f(p(x)) = g(x). To show that /iscontmuous, let Kbe an open set in Z. Continuity of g implies that BoPen in X. Because p is a quotient map, f'(V) must be open in Y. \J does not behave weU: even though one starts with ^Hausdorff Th ^ 'S "° reaS0" that the potent space needs to Quotient space^If ^ * S'mple c.ondition for °ne-point sets to be closed in *>$ed in y if h ^= X ~* Y is a ^"ot'ent map, then one-point sets are ^Wition Tx ? , CaCh setp"(W) is closed in JT. Therefore, if X* space a-* r n,t0 closed sets- a» one-point sets are closed in the quo- ^ There'is a", 1OnS that W'U CnSUre that X* is "ausdorff are harder * °™CWhat implicated condition called upper semicon- n impose on the partition X* to ensure that it is Hausdorff
140 Topohgical Spaces and Continuous Functions q^ (See Exercise 13 of §4-2.) There is, however, one situation in which it is ^ to see that X* is a Hausdorff space: Theorem 11.2. Let g : X — Z be a surjeetive continuous map. Let X* be the following collection of subsets ofX: X» = [g"({z])|zeZ). Give X* the quotient topology. (a) If 7. is Hausdorff, so is X*. (b) The map g induces a bijective continuous map f: X* — Z, which is a homeomorphism if and only if g is a quotient map. K X*—~Z f Proof. By the preceding theorem, g induces a continuous map/: X*-^Z; it is clear that / is bijective. If Z is Hausdorff, then given distinct points of X*, their images under / are distinct and thus possess disjoint neighbor- neighborhoods t/and V. Then f~'(U) and f~'(V) are disjoint neighborhoods of the two given points of X*. Suppose that/is a homeomorphism. Then both / and the projection p : X—* X* are quotient maps, so that g — /° p is a quotient map. Con- Conversely, suppose that g is a quotient map. Given an open set V of X*, we compute which is open in X (because p is continuous). Therefore, f(V) is open in Z, because g is a quotient map. Hence / maps open sets to open sets, so it is a homeomorphism. □ When dealing with quotient spaces, one must be careful not to rely too much on one's intuition as to what the spaces look like. Consider the follow- following example: Example 7. Let A' be the union of all straight lines in the plane of the form L, = R x {«}, for n e Z+. LetZ be the union of all straight lines in the plane passing through the origin and having positive integral slope. That is, letZ = U£i» where L', = {x x (nx)\xz R). Both are subspaces of RK Let g: X — Z be the map g(x x n)=x x (nx), . which carries each line £„ linearly onto L'n. See Figure 27. We assert that the map g is not a quotient map.
The Quotient Topology 141 Figure 27 That is, we assert that the quotient topology on Z induced by the map g is not the same as the subspace topology on Z. To prove this fact, let us take, for each n, an open interval {/„ of length \/n in the line L'm centered at the origin; and let us define U = \JUn. Then U is open in Z in the quotient topology induced by 4', since the set,?- '(U) is open in X. But Uis not open in the subspace topology on Z; in that topology, the origin is a limit point of the complement Z — U of U. To say the same thing in a different way, suppose that we form a quotient space X* from A'by collapsing the set 0 x Z+ to a single point. It is tempting to think of X* as being essentially the same as the subspace Z of the plane. But it is not. The map g induces a bijective continuous map h : X* —► Z. But since g is not a quotient map, h is not a homeomorphism. Example 8. The product of two quotient maps need not be a quotient map. Let us take the spaces X and Z of the preceding example, but now let us give Z the quotient topology induced by g. Consider the space R" (in the product topology). Let i :/?"—> R" be the identity map; being a homeomorphism, it is a quotient map. We shall prove tfiat the map f = g x i: X x R°> > Z x R°> 8 not a quotient map. Recall that X = R x Z+. •ven« 6 ^+! define ^ t0 be the following subset of ,R x z+) x /po. ^ U" = K' x «) x (xu x2,...);\ /jr. | < 1}. »K open in A' x R°>, because it is the cartesian product of the open set K» P'ane, the open set "<*• See Figure 28. {/»} in Z+, and the open sets R in all the other
Topical Spaces and Continuous Functions chap. i Figure 28 Define U to be the open set = VjnEZ. of X x R". We assert the following: (a) /"'(/(£/)) = £/. (b) /(£/) is not open in Z x R". It follows, since /"'(/(£/)) is open in X x /f» and /(£/) is not open in Z x R°>, that the map / is not a quotient map. (a) If /-'(/(£/)) is different from £/, then there must exist two points x e U and y £ £/ such that /(x) = /(y). But in the present situation, /(x) = /(y) for x ^fc y only if x = @ x m) x (x,, .y2, . ..), y = @ x n) x (x,,x2,. . .), for some point (.v,, x2,....) of R°>. And these two points both belong to U, since |0-xm| < 1 and |0-xn| < 1. (b) Suppose that /(£/) is open in Z x R°>. The set /(£/) contains 0x0, so it contains a basis element V x YlW, about 0x0. Here V is open in Z, and W, is open in R; and W, = R for all but finitely many i. Choose N so that WN = R. Since /((/) contains K x n W/, the set £/ = /"'(/(£/)) contains /-1(KxIIW()=^-"(K) x IIW|.Now^-'(K)isanopensetin ^containing the entire set 0 x Z+. In particular, g-*(V) contains some point t0 X # for r0 # 0. Choose aw > i/|^0|, and consider the point (h x N) x @ 0, aN, 0,...) •>■. ■
The Quotient Topology 143 $2-1' *» Because WH = *. *«* Point lies in rW X II"'/- Because rf * X San I, ft does not lie in U. But U contains ,-»(»0 * II^- Contra*"- Exercises Check the details of Example 3. R y. R—.R be projection on the first coordinate. 1 fut -f be the subspace @ x R) u (/? x 0) of /? x /?; let g = nl\X. Show that ? is a closed map butnot an open map. ,w Let Y be the subspace (J?+ x /?) u (/? x 0) of R x /?; let h=nt\ Y. Show that A is neither open nor closed, but it is a quotient map. [Hint: h-'(U) n(K x0) = U xOJ Letc A" ► y be surjective and continuous; let A be a subspace of X. Show that if A is open in Jf and p is an open map, then p\A is an open map; con- conclude ihaXp'.A —*P(A) is an open map. Show that if A and p are closed, p\A is closed, and so is p'. 4. Let Xbe a space; let A be a set; let p : X —> A be a surjective map. Show that the quotient topology on A induced by p is the finest (largest) topology relative to which p is continuous. 5. (a) Define an equivalence relation on the plane X as follows: x0 X y0 ~ xi x y, if .v0 + yl = x, + y\. Let X* be the collection of equivalence classes, in the quotient topology. X* is homeomorphic to a familiar space; what is it? (b) Repeat (a) for the equivalence relation *o X y0 ~ x, x y, if xl + yl = xf + JV?. «• (a) Define «■: /?* —, r by the equation ^(a- x y) = x + y*. Show that g is a quotient map. (b) Define^: Ri — ^+ by the equation ^(.v x >) = x2 + y2. Show that g is a quotient map. (c) Compare (a) and (b) with Exercise 5. 1 (R X 0) U (° X ^ Of ^2- Define * '■ R% —'Z by the «(^X)') = xxO if a;^ 0, f@x>) = 0x>. *) Show r'-611' maP? Is * continu°«s? Hausdorff '" 'he quotient topology induced by g, the space Z is not " tC- be the subspace of R> defined by . . ...
144 Topological Spaces and Continuous Functions Ox^.x Let Y be the subspace r=LUz.C. of R2, and let X be the subspace C, X Z+ of R2 x tf. Define g:X-*Y by the equation £((* X y) X «) = (x/n,y/n). Show that,? is continuous and surjective, but g is not a quotient map. 9. Let {X.} be a family of spaces and let [p.] be a family of maps pa : X. — ^ where /4 is a set. (a) Show there is a unique finest topology 3 on A relative to which each map p, is continuous. (b) Show that a map/: A —> r is continuous relative to 3 if and only if each map/° p. is continuous. * Supplementary Exercises: Topological Groups In these exercises we consider topological groups and some of their properties. The quotient topology gets its name from the special case that arises when one forms the quotient of a topological group by a subgroup. A topological group G is a group that is also a topological space, satisfying the requirements that the map of G x G into G sending x x y into x-y, and the map of G into G sending x into x'1, are continuous. 1. Show that G is a topological group if the map ofCx C into G sending x x y into x'^-y is continuous, and conversely. 2. Show that the following are topological groups: (a) (Z, +) (b) (R, +) (c) (*♦• •) (d) (S\ •), where we take S' to be the space of all complex numbers z for which 1*1 = 1. (e) The general linear group GL(«), under the operation of matrix multiplica- multiplication. (GL(«) is the set of all nonsingular n by n matrices, topologized by considering it as a subset of iJ'"'i in the obvious way.) @ (.&, +) in the product, uniform, and box topologies. 3. Let (G, •) be a topological group; let a be an element of G. Show that the maps /«,g.:G—*G defined by /«(*) = cc-x and g,(x) = x-a. are homeomorphisms of G. Conclude that G is a homogeneous space. (This means that for every pair x, y of points of G, there exists a homeomorphism of G Onto itself that carries x to y.)
Supplementary Exercises: Topological Groups 145 utaroupofff.If^e G, define .vtf-[*■*!* 6 H); this set is called 4 te,//be a subgroup^ ^ ^ ^^ the co,lection of ,eft cosets of W in G; * left c*Vt.°'"f C |f G is a topological group, give C/// the quotient topology. iHsapartili ^ ff the mapyi Qf the preceding exercise induces a homeo- (a) Show^ ^ ^^ carrying A// to (a-.v)M Conclude that G/// is a homo- peneous space. ,M show that the quotient map p:G-> G\H is open. \ Show that if H is a closed set in the topology of G, then one-point sets are closed in GjH. Show that if//is a normal subgroup of G, then G/H is a topoiogical group. f2 +) is a normal subgroup of (/?,+). The quotient R/Z is a familiar topolog- topological group; what is it? Let G be a topological group with identity element e. If A and B are subsets of G let -4-fi denote the set of all points a-b (or a e A and b e B. Let -4"' denote the set of all points a"', for a e A. (a) If 1/ is a neighborhood of e, show there is a neighborhood V of e such that K-F-' c £/. (b) If A is a closed set not containing e, show there is a neighborhood V of e such that the open sets V and are disjoint. (/« is the map defined in Exercise 3.) (c) Suppose that one-point sets are closed in G. Show that C is Hausdorff. In fact, show that C satisfies the following stronger separation axiom, which is called the regularity axiom: Given a closed set C and a point .v not in C, there exist disjoint open sets containing C and x, respectively. 7. Let H be a subgroup of G that is closed in the topology of G; let p : G —■+ G/H be the quotient map. Show that G/H satisfies the regularity axiom. [Hinf Urader ftst the case where A is a closed set in G/H not containing p(e). Let «~P W). Find a neighborhood U of e disjoint from B and let V»• Kc C/ inen consider the sets K-//and K-5.] »• Show that if H is a subgroup of G, so is the closure of H.
3. Connectedness and Compactness In the study of calculus, there are three basic theorems about continuous functions, and on these theorems the rest of calculus depends. They are the following: Intermediate value theorem. If/: [a, b] —* R is continuous and if r is a real number between f{a) and f{b), then there exists an element c e [a, b) such that f{c) = r. Maximum value theorem. If/: [a, b] —> R is continuous, then there exists an element c e [a, b] such that /(*) < /(c) for every x e [a, 6]. Uniform continuity theorem. \f f;[a,b]—* R is continuous, then given f > 0, there exists 5 > 0 such that |/(x,) — /(x2)| < f for every pair of numbers *„ x2 of [a, 6] for which \x,- x2\<d. These theorems are used in a number of places. The intermediate value theorem is used for instance in constructing inverse functions, such as^x and arcsin x; and the maximum value theorem is used for proving the mean value theorem for derivatives, upon which the two fundamental theorems of calculus depend. The uniform continuity theorem is used, among other things, for proving that every continuous function is integrable. We have spoken of these three theorems as theorems about continuous Junctions. But they can also be considered as theorems about the closed 146
Connected Spaces 147 of real numbers. The theorems depend not only on the con- [a, °l °l Qn properties of the topological space [a, b]. uity Vb":v of the space [a, b] on which the intermediate value theorem ^e propeny ty caned connectedness, and the property on which the ends is 'he P _i^ property called compactness. In this chapter, we shall er wo d^. f arbitrary topological spaces and shall prove the depen ^ property called compactness. In ths cp, other wo d^.es for arbitrary topological spaces, and shall prove the de/ine these P ^j.^ versions of these theorems. appf°Prla s e qUOted theorems are fundamental for the theory of calculus, ASt!r 'otLns of connectedness and compactness fundamental in higher so*1*. ometIy and topology—indeed, in almost any subject for which jS'ff topological space itself is relevant. 3-1 Connected Spaces The definition of connectedness for a topological space is a quite natural e One says that a space can be "separated" if it can be broken up into two "slobs"—disjoint open sets. Otherwise, one says that it is connected. From this simple idea much follows. Definition. Let X be a topological space. A separation of X is a pair U, V of disjoint nonempty open subsets of X whose union is X. The space X is said to be connected if there does not exist a separation of X. Connectedness is obviously a topological property, since it is formulated entirely in terms of the collection of open sets of X. Said differently, if X is connected, so is any space homeomorphic to X. Another way of formulating the definition of connectedness is the fol- following: A space X is connected ifand only if the only subsets ofX that are both open and closed in X are the empty set and X itself. mfA is a nonempty proper subset of X which is both open and closed in they aTe "*'- = A a"d V = X ~ A constitute a separation of X, for f and P'f1'611' dlS^°int> and nOnempty, and their union is X. Conversely, if anditi<h°!raSeparation of X>then Uis nonempty and different from X, ¥oS b°th open and closed in X. LbtlPa!e Kof a toPological space X, there is another useful way of the definition of connectedness: none>nPty sets'k VS ° subspace °/X> a separation ofY is a pair of disjoint of">e other n ^"^ U"io" " Y) neither °f which contains a limit point space Y is connected if there exists no separation ofY.
148 Connectedness and Compactness Chap. 3 Proof Suppose first that A and B form a separation of Y, Then A is both open and closed in Y. The closure of A in J is the set AnY (where ^ as 2S denotes the closure of ^ in X). Since ^_,s closed in Y, A- AnY;Ot to say the same thing, AnB=0. Since A is the union of A and its limi, points, B contains no limit points of A. A similar argument shows that A contains no limit points of B. . Conversely, suppose that /4 and 5 are disjoint nonempty sets whose union is Y neither of which contains a limit point ofthe other. Then A_ n B = 0 and'/4 n 2? = 0 ; whence we conclude that An Y = /4 and 5 n r = a. Thus both /4 and 5 are closed in Y, and since A = 7 - B and 5 = y _ A> they are open in Y as well. D Example 1. Let X denote a two-point space in the indiscrete topology. Obviously there is no separation of X, so X is connected. Example 2. Let Y denote the subspace [-1, 0) u @, 1] ofthe real line R. Each ofthe sets [ — 1, 0) and @, 1] is nonempty and open in Y (although not in R); therefore, they form a separation of Y. Alternatively, note that neither of these sets contains a limit point of the other. (They do have a limit point 0 in common, but that does not matter.) Example 3. Let X be the subspace [-1, 1] ofthe real line. The sets [-1, 0] and @, 1] are disjoint and nonempty, but they do not form separation of X, because the first set is not open in X. Alternatively, note that the first set contains a limit point, 0, ofthe second. Indeed, there exists no separation ofthe space [—1, 1]. We shall prove this fact shortly. Example 4. The rationals Q are not connected. Indeed, the only connected subspaces of Q are the one-point sets: If Y is a subspace of Q containing two points p and q, one can choose an irrational number a lying between p and q, and write Y as the union of the open sets Yn(—oo, a) and Kn(a, +oo). Example 5. Consider the following subset of the plane R2: X = {x X y\y = 0} u {x X y\x> 0 and ^ = 1/x}. Then X is not connected; indeed, the two indicated sets form a separation of X, because neither contains a limit point of the other. See Figure 1. We have given several examples of spaces that are not connected. How can one construct spaces that are connected? We shall now prove several Figure 1
Connected Spaces 149 Pi u , ,,11 how to form new connected spaces from given ones. Then theorems that tell how theorems to construct some specific I* nex« ^"^s"nteTvals in X. and balls and cubes in *■. F.rst, a ^teo sp lemma: 2 If the sets C and Dform a separation ofX, and if Y is a con- Umml tofX, then Y lies entirely within either C or D. mC'edSUf w Cand D are both open in X, the sets C n Kand D n Xare y Ttesc two sets are disjoint and their union is Y; if they were both ™'" , would constitute a separation of Y. Therefore, one of them Y must lie entirely in C or in D. □ r*<w<?m /.5. The union of a collection of connected sets that have a point jn common is connected. Proof Let {A,} be a collection of connected subsets of a space X; let p be a point of f]A.. We prove that the set Y = \JA. is connected. Suppose that Y= C U X) is a separation of r. The point p is in one of the sets C or D- suppose/) G C. Since the set A, is connected, it must lie entirely in either C or D, and it cannot lie in D because it contains the point p of C. Hence A, c C for every a; whence [JA, c C, contradicting the fact that D is non- nonempty, n Theorem 1.4. Let A be a connected subset o/X. If A c B c A, then B is also connected. Said differently: If B is formed by adjoining to the connected set A some or all of its limit points, then B is connected. Proof. Lex A be connected and let A c B c A. Suppose that B = C uD H?n]iOn °f * % LCmma L2' A- mUSt lie entirely in C or in ^ SUP" P that A c C. Then /I c C; since C and D are disjoint, B cannot intersect • ins contradicts the fact that D is a nonempty subset of B. Q ^ 'mage °fa comected sPace ™der a continuous map is Pr°°f. Let/"- x vk 10 Pfove the irnapp ~7 / cont'nuous map; let X be connected. We wish /fay restricts itf™ =/W 1S connected- Since the map obtained from «the case of a I T Space Z is also c°ntinuous, it suffices to consid- a continuous surjective map »„ g:X—>Z. suPpose that Z = j ** Open « Z. Then r >uf and "-1 m*^ ^ '^ tW° diSJ°int n°nempty S (A) and g iE) are disjoint sets whose union is X;
ISO Connectedness and Compactness Chap. 3 thev are open in X because g is continuous, and nonempty because g is SUrjec_ live. Therefore, they form a separation of X, contradicting the assumption that X is connected. D Theorem 1.6. Tlie cartesian product of connected spaces is connected. Proof We prove the theorem first for the product of two connected spaces X and Y. This proof is easy to visualize. Choose a "base point" a x b in the product X X Y. Note that the "horizontal slice" X x b is connected, being homeomorphic with X, and each "vertical slice" x x Yh connected, being homeomorphic with Y. As a result, each "T-shaped" space Tx = (X X b) U (x X Y) is connected, being the union of two connected sets that have the point x x b in common. See Figure 2. Now form the union \JxSX Tx of all these T-shaped spaces. This union is connected because it is the union of a collection of con- X Y a X b X x Figure 2 nected sets that have the point a X b in common. Since this union equals X x Y, the space X x Y is connected. The proof for any finite product of connected spaces follows by induction, using the fact (easily proved) that X, x ■■■ x Xn is homeomorphic with (X, X ■■■ XX..,) XX.. The proof for an arbitrary product now proceeds as follows. Let {X,},*i be an indexed family of connected spaces, and let Choose a "base point" b = (&.).„ for X. Given any fin.te set {a «„} of indices in J, let us define a subspace Wi. • •-,«„) otX; it consists of all points (xj.e, such that xa = b. for t ""' ^^ ^ X(a" •••'«») is homeomorphic with the product X.. X andhence is connected: For the obvious b XX. ijection between these spaces, map-
Connected Spaces 151 a - a- and J'« = b* for a" other values of a> car" .'x ••• X X., to a basis element for X(oc,, . . .,<*„). ^ ^ sUb'space r of ^ which is the union of these subspaces ; this is the subspace the union extends over all finite subsets fa, a_} of J. For the vi a ) are connected, and they all contain the base point spaces JT(Oi. - • • - ""•' * Ids'tempting to think that the proof is finished. But it is not, for Y is not f X The space /consists of all points (.vj,ey of X having the property t • =b for all but finitely many values of a. So in some sense Y is only ' * rv'small'part of X. But now the fact that we are using the product topo- for /comes in. We assert that under the product topology for X, the closure of /equals all of X. Once we prove this fact, the connectedness of Xfollows from Theorem 1.4. Let us take an arbitrary point (xj of X, and an arbitrary basis element {/= \\U, about (.v.), and prove that U intersects Y. Each set £/„ is open in I,, and U, = X, except for finitely many indices, say a = a,, . . . , «,„. Con- Construct a point O'J of X by setting _ f.x. for a = a,,..., a,, [&„ for all other values of a. Then (y,) is a point of Y, because it belongs to the space X(alt . . . , a_). Also 0'.) is a point of U, because ya = x. e £/. for a = a, a. and /. =4. e X, for all other values of a. Hence £/ intersects Y, as desired. □ ^ This theorem is not true if one uses the box topology on X. See Exercise Exercises of X ^LtTmp^att !( 1 Sh<>w that 7 n* co™ectedness in the other? " 'S C°nneCted and nonemPty. then each X. is connected. °f '•such tha< ^ - ^- - * . ^ow^t^ j 7d s"b-'^ of X; let A be a connected subset Sh0* that if v: ' 0 f°r a" a' then ^ ^ CJA.) is connected.
152 Connectedness and Compactness £. 6. A space is totally disconnected if its only connected subsets are one-point„ Show that a finite Hausdorff space is totally disconnected. sets- 7. Is it true that if X has the discrete topology, then X is totally disconnected7 Does the converse hold? 8. Is it true that if X is connected, then for every nonempty proper subset A of X, we have Bd A ^ 0 ? Does the converse hold? (Recall that Bd A L A n X - A.) 9. Let A c= X. Show that if C is a connected subset of X that intersects both a and X - A, then C intersects Bd A. 10. Is the space H, connected? Justify your answer. 11. Show that R'° is not connected in the box topology by showing that the set A of all bounded sequences is both open and closed. What happens if R<° has the uniform topology? 12. Let Y <= X; let X and Y be connected. Show that if A and B form a separation of X — Y, then Y U A and 7 U B are connected. J-2 Connected Sets in the Real Line The theorems of the preceding section show us how to construct new con- connected spaces out of given ones. But where ran we find some connected spaces to start with ? The best place to begin is the real line. We shall prove that every interval and every ray in R is connected. One application is the intermediate value theorem of calculus, suitably generalized. As another application, we show that such familiar spaces as balls and spheres in euclidean space are connected. The proof of this fact leads to a new notion, called path connectivity, which we also discuss. The fact that intervals and rays in R are connected may be familiar to you from analysis. We prove it again here, in generalized form. It turns out that this fact does not depend on the algebraic properties of R, only on its order properties. To make this clear, we shall prove the theorem for an arbitrary ordered set that has the order properties of R. Such a set is called a linear continuum. Definition. A simply ordered set L having more than one element is called a linear continuum if the following hold: A) L has the least upper bound property. B) If x < y, there exists z such that x < z < y. Theorem 2.1. If X is a linear continuum in the order topology, then L is Connected and so is every interval and ray in L. Proof. Let Y be a subset of L that equals either L or an interval or a ray in L. The set Y is "convex," in the sense that if a and b are any two points ol
Connected Sets In the Real Line 153 h then the entire interval [a, b] of points of L is contained in the * Y- A B be two disjoint nonempty sets that are open in Y. We shall Let v* AUB, thereby demonstrating that there exists no separation jhow ^at / *= oiY' pointaof/«anda point b of B; assume the notation so chosen Ch°°T (Do not make the mistake of assuming that every point of A is *at I every point of B; that need not be true.) Because Y is "convex," we less than ^^ find a int of ^ ^ tha{ belongs t0 neither A nor have [o. °J c ' D ' Consider the sets Ao = A n [a, b] and 50 = B n [a, 6]. These sets are open in [a, 6] in the subspace topology (which is the same as the order topology). Let c = lub Ao. We shall show that c belongs to neither Ao nor Bo. Case I. Suppose that c e Bo. Then c ^ a, so either c -- b or a < c < b. In either case, it follows from the fact that Bo is open in [a, b] that there is some interval of the form (d, c] contained in /?„. See Figure 3. If <■ -- b. we have a contradiction at once, for d is a smaller upper bound on A0 than c. If c < A, we note that (c, b] does not intersect Ao (because c is an upper bound on At). Then (d, b] = (d, c] U (c, b] does not intersect ,*„. Again, d is a smaller upper bound on Ao than c, con- contrary to construction. Case 2 Suppose that c e /<„. Then c ^ b, so either c = a or a < c < b Z '7Pe" 'lnJfl *L thCre mUSt be SOme interval of the form I*. 0 ... of ose a point z of L such that c < z <" » ' that c is an upper bound for Ao. fj ~ z a Figure 4
154 Connectedness and Compactness Chap. 3 Corollary 2.2. The real line R is connected and so is every interval and ray in R. As an application, we prove the intermediate value theorem of calculus, suitably generalized. Theorem 2.3 (Intermediate value theorem). Let f : X — Y be a continuous map of the connected space X into the ordered set Y in the order topology. jZld b are two points of X and if r is a pomt ofY lymg between f(a) and f(b), then there exists a point c ofX such that f(c) = r. The intermediate value theorem of calculus is the special case of this theo- theorem that occurs when we take X to be a closed interval in R and Y to be R. Proof Assume the hypotheses of the theorem. The sets A = f(X) n (-00, r) and B = f(X) n 0, +00) are disjoint, and they are nonempty because one contains f(a) and the other contains f(b). Each is open in f(X), being the intersection of an open ray in y with f(X). If there were no point c of X such that f(c) = r, then f(X) would be the union of the sets A and B. Then A and B would constitute a separation of f(X), contradicting the fact that the image of a connected space under a continuous map is connected. □ Example 1. One example of a linear continuum different from R is the space / x / in the dictionary order topology, where / = [0, 1]. We check the least upper bound property. (The second property of a linear continuum is trivial to check.) Let A be a subset of / x /; let n, : I x I —► / be projection on the first coordinate; let b = lub7T,(/0. If b e k,(A), then A intersects the subset b x / of / x /. Because b x / has the order type of /, the set A n {b x /) will have a least upper bound b X c, which will be the least upper bound of A. See Figure 5. If b £ n,{A), then b x 0 is the least upper bound of A; no element ir,(A)X0 bXc K Figure S
Connected Sets in the Real Line 1S5 , x , with b> < b can be an upper bound for A, for then b' would If Xis a well-ordered set, then X x [0, 1) is a linear continuum ofd r; this we .eave to you to check. This set can be thought o ' b ufit(i .„„ a set of the order type of @, 1) lof Xand its immediate successor, provided X has no largest element. Connectedness of intervals in H gives rise to an especially useful criterion rshowing that a space X is connected; namely, the condition that every p£rSof points of A-can be joined by apatli in X: Definition. Given points x and y of the space X, a path in X from x to y isacontinuous map/: [a, b] -> X of some closed interval in the real line into X, such that f(a) = x and f(b) = y. A space A' is said to be path connected if every pair of points of X can be joined by a path in X. It is easy to see that a path-connected space X is necessarily connected. Suppose X = A U B is a separation of ^V. Let /: [a, b] —► X be any path in I. Being the continuous image of a connected set, the set f([a, b]) is connect- connected, so that it lies entirely in either A or B. Therefore, there is no path in X joining a point of A to a point of B, contrary to the assumption that X is path connected. The converse does not hold; a connected space need not be path con- connected. See Examples 6 and 7 below. Example 3. Define the unit ball B" in R" by the equation 5» = {X;||X||<1}> where /(/) = A - /)x + /y lles "> *. For if x and y are in B" and / is in [0, 1], ^^ft'oS in TT^ ?"*?*" SP3Ce t0 te the sPace ^" " {0}> ?rdiff=rent from 0 we cfn'rl >,th'S SpaCe is path connec'ed: Given x and **« Path do« noTgo thr k 1^ y by thC straight-'ine P^h between them Zn°t on ,he lineS ° f "* °ngin- °therwise> we ™ *°™ a point « Path do« noTgo thr k 1 y traight'ine P^h between them Zan°t on ,he lineSg x ° f "* °ngin- °therwise> we ™ *°™ a point 411A «*n from z to y y> Md take the br0^n-Iine path from x to z,
156 Connectedness and Compoctness Cha 'P. 3 Example 5, Define the unit sphere S"-' in R« by the equation S"~' = [x;||x|| = 1], If n > 1, it is path connected. For the map g : R" - [0] —► S"-i defin g(x) = x/)|x|| is continuous and surjective; and it is easy to show that the tinuous image of a path connected space is path connected, COn" Example 6, The space I x I in the dictionary order topology is con but not path connected. Being a linear continuum, / x / is connected. Let p = 0 x 0 and 1 x 1, We suppose there is a path/: [a, b] —► / x /joining p and q and der* a contradiction. The image set f([a, b]) must contain every point x x vof/ y f by the intermediate value theorem. Therefore, for each x e /, the set ' Vx =/"'(* x @,1)) is a nonempty subset of [a, b]; by continuity, it is open in [a, b]. See Figure 6 Choose, for each x e /, a rational number qx belonging to Ux. Since the s t Figure 6 * X @, 1) Vx are disjoint, the map x —> qx is an infective mapping of / into Q. This con- contradicts the fact that the interval /is uncountable (which we shall prove later). Example 7, Here is another connected space that is not path connected; it is a subspace of the plane. Let K denote the set A/,, |,, e Z+], and define C = ([0, 1] x 0) u (K x [0, 1]) u @ x [0, 1]), The space C is called the comb space, for obvious reasons. See Figure 7, The pace S obtained by deleting from C the points of the vertical interval 0 x @,1) is called the deleted comb space X X 3 2 Figure 7
Connected Sets in the Real Line 157 The comb space Cis clearly path connected. The deleted comb space D dily seen to be connected, for it is the union of the path connected set A = ([0, 1] X 0) u (.K X [0, 1]) d the limit point p = 0 x 1 of A We shall prove that D is not path con- Suppose that/: [a, b] —► D is a path in D beginning at p. We assert that the setf~'({pT>is *""h open and closed ln ["• bk from this it follows by connectivity that /"'rt"l) = ["•*]• We coriclude that 'here exists no path in D joining p i^hat r'ttPj) «s closed is trivial' sinCe fP) is closed and/is continu- TTn show 1. /->({/») is open, we choose a neighborhood Cof,.n vhfch dolsno. intersect .hex-axis, as illustrated in Figure 8. Then, given an Figure 8 arbitrary point xa of /"'([/>]), we choose a basis element t/about x0 such that /(J/)c v. We assert that C/is contained in /-'([/>}). so /"'([p]) is open. Note that U is connected, being a basis element for the order topology on [a, b]. Therefore, f{U) is connected. Then f{U) cannot contain any point different from p; Given a point q = A/n) x /0 of D different from p and Vmg in V, choose r so that l/(n + 1) < r < 1/n. Then consider the following •wo disjoint open subsets of R*: (—oo, r) x .R and (r, +oo) x R. *«use/(t/) lies in Z> and does not touch the x-axis, it does not intersect the ir n theref!ore' " lies in the Union of these tw0 sets. Since f(U) is con- nwteir ! nf th and COntains the Point p of the first set, it cannot contain the point q me second set. Thus /({/) - {p\ as asserted. . Let 5 denote the following subset of the plane:
m Connect a,,* Compactness Chap, 3 c ■ ,he image of the connected set @, 1] under a continuous map, s Because S is *?«"»«"* dosure S in R> is also connected. The set $ is a is connected. T™r!" , caMed the topologist's sine curve. It is illustrated classical example n topo k curye ^ cQnnected but not path connec SSSiS^1'* p-f we gave for the deleted Mmb space- Figure 9 Exercises 1. (a) Show that no two of the spaces @, I), @, I], and [0, I] are homeomorphic, (b) Suppose that /: X —> Y and g : Y—-> X are imbeddings. Show by means of an example that X and Y need not be homeomorphic, 2. Let n be a positive integer. Given that the function f; R —► R defined by /(.v) = x" is continuous, show that for each a > 0 there is exactly one b >l 0 such that tf = a. 3. Let/: -f— X be continuous. Show that if X = [0, I], there is a point x such that f(x) = jr. The point x is called a fixed point of/. What happens if X equals [0, 1) or @, 1)? 4. Let X be an ordered set in the order topology. Show that if X is connected, then X is a linear continuum, 5. Consider the following sets in the dictionary order. Which are linear continua? (a) Z+ x [0, 1) (b) [0,1) x Z+ (c) [0,1) x [0, 1] (<J) [0,1] x [0, 1) 7' !K J! (cl If tCt f Pa'h-^ec'ed spaces necessarily path connected? v U Pa'h connected- is -4" necessarily path connected?
Components and Path Components 159 of path-connected subsets of X and if (V- * ®> ^^*H connect R' and J? are not homeomorphic if n > 1. j. show that untab|e. Show that if /* is a countable subset of *2, , Ass"* IW * ls "h connected. [Hint: How many lines are there passing 'thCn*hagiven point of *»?] thr0" r,; is an open connected subset of R\ then t/ is path connected, l0 Show that ii u £ ^ the set of points that can be joined to x0 by "SKi open and dosed in <,.] SKibo* p onnected subset of X, are Int A and Bd /( necessarily connected? 11 Jjs'theconverse hold? Justify your answers, hat the topologist's sine curve is connected but not path connected, 11 L denote the ordered set Sn X [0, 1) in the dictionary order, with its '* allest element deleted. The set L is a classical example in topology called the Theorem. The long line is path connected and locally homeoniorphic to R, but it cannot be imbedded in R. Proof. (a) Let X be an ordered set with smallest element a0. Let A = [flo] U {x\[a0, x) has the order type of [0, I)}. Show that if for each x e X there is a y > x such that [,v, y) has the order type of [0, I), then A is open in X. Show that if X satisfies the "sequence lemma," then A is closed in X. [Hint: If a is a limit point of A not in A, there is an increasing sequence of points of A converging to a.] (b) Show that L is locally homeomorphic to R; that is, each point of Z. has a neighborhood homeomorphic to an open subset of R. (c) Show that L is path connected. (d) Show that L cannot be imbedded in R, or indeed in R" for any n. [Hint: Anysubspace Y of R" has a countable basis for its topology; therefore, any collection of disjoint open sets in Y is countable.] *3-3 Components and Path Components^ ' arbltrary space X, there is a natural way to break it up into re connected (or path connected). We consider that process now. "efinUion, Given v a c !fthere'Saconn t h *" equivalence rela"on on X by setting x~y ^sarecalledTh ^^ of A'COntaining b°th x and y. The equivalence he components (or "connected components") of X. Symmetry and r fl ■ • -^ enexivity of the relation are obvious. Transitivity follows B ^"on win be assumed in Chapter 8.
160 Connectedness and Compactness Chap, j by noting that if A is a connected set containing x and y, and if B is a "ec"ed Jt containing>' and r, then /( U B is a set conta,mng x and z, which s connected because A and * have the po.nt ym common. The components of X can also be descnbed as follows: Theorem 3.1. The components of X are connected disjoint subsets of ^ whose union is X, such that each connected subset ofX intersects only one of them them. Proof. Being equivalence classes, the components of X are disjoint and their union is X. Each connected set A in X intersects only one of them. For if A intersects the components C, and C2 of X, say in points x{ and x2 respectively, then x, ~ x2 by definition; this cannot happen unless C, = c2' To show the component C is connected, choose a point x0 of C. For each point x of C, we know that x0 ~ x, so there is a connected set /(, containing x0 and x. By the result of the preceding paragraph, Ax <= C. Therefore, c = LUc 4,. Since the sets Ax are connected and have the point x0 in common, their union is connected. Q Definition. We define another equivalence relation on the space X by defining x ~ y if there is a path in X from x to y. The equivalence classes are called the path components of X. Let us show this is an equivalence relation. First we note that if there exists a path/: [a, b] —> X from x to y whose domain is the interval [a, b], then (because any two closed intervals [a, b] and [c, d] in R are homeomor- phic) there is also a path g from x to y having [c, rf] as its domain. Now the fact that x ~ x for each x in A' follows from the existence of the constant path/: [a, b] —> A'defined by the equation /(/) = x for all t. Symmetry fol- follows from the fact that if/: [0, 1] — A-is a path from x to y, then the "reverse patrT g: [0, 1] — A-defined by g(t) = /(I - /) is a path from >> to x. Finally, transitivity is proved as follows: Let/: [0, 1] — X be a path from x to .y, and let g: [I, 2] —► X be a path from j> to z. We can "paste/and g together" to get a path h : [0, 2] — X from x to z; the path h will be continuous by the pasting lemma," Theorem 7.3 of Chapter 2. One has the following theorem, whose proof is immediate: Theorem 3.2. The path components ofX are path-connected disjoint subsets oj X whose union is X, such that each path-connected subset ofX intersects only one of them. Example 1. The components of the subspace y=l-l, 0)u@, 1]
,e real line of Local Connectedness 161 R are the two sets [-1, 0) and @, 1]. These are also the path , The deleted comb space D of the previous section is a space Example 2. lheu (because it is connected) and two path corn- having * single,'° a °ace Y by adjoining to the deleted comb space D all ponents. If w>™ q{ t"he jnterva, 0 x [0, 1], we obtain a space having only thC wm'ponent but uncountably many path components. *3~4 Local Connectedness4* r nnectivity is a useful property for a space to possess. But for some pur- it is more important that the space satisfy a connectivity condition f*//'•' Roughly speaking, local connectivity means that each point has 'Arbitrarily small" neighborhoods that are connected. More precisely, one has the following definition: Definition. A space X is said to be locally connected at x if for every neiehborhood (/of x, there is a connected neighborhood Kof x contained in {/.If X is locally connected at each of its points, it is said simply to be locally connected. To phrase it differently, X is locally connected if there is a basis for X consisting of connected sets. Local connectedness and connectedness of a space are not related to one another; a space may possess one or both of these properties, or neither. Definition. A space A'is said to be locally path connected at x if for every neighborhood (/of x, there is a path-connected neighborhood V of x con- connedm U If AT is locally path connected at each of its points, then it is said 10 be locally path connected. and^liv '" EaCh 'nterVal and each ray in the real line is b0'h connected but i?.lTeCted- ThC SubSpace C—1.0) ^ @, I] of R is not connected, «»nnec.J bu nnTr6^- "^ de'e'ed COmb Space (Example 7 of §3> is "Or locally connected C°nnected- The nationals Q are neither connected fo. bTx .. ? w 'S !°fa!'y path con"ected, since each of the basis elements "nnected; each ofTf, , *\ Pa'h connected. Similarly, R» is locally path h °f "s standard basis elements is path connected.
162 Connectedness and Compactness Chap, Example 3. If we form a space Y by adjoining to the deleted „" the points of the form 0 x (l/n) for n e Z+, we obtain a L.1s locally connected a. the origin but no. locally path connected , origin. The most important facts about locally connected spaces are given in the following theorems: Theorem 4.1. A space X is locally connected if and only if for every open set U ofX, each component ofV is open in X. Proof. Suppose that A'is locally connected; let U be an open set in X; let Cbe a component of U. If x is a point of C, we can choose a connected neigh- borhood V of* such that V <= U. Since Kis connected, it must lie entirely in the component C of U. Therefore, C is open in A'. Conversely, suppose that components of open sets in A'are open. Given a point xof Xand a neighborhood i/of .v, let Cbe the component of Ucon- Ucontaining .v. Now C is connected; since it is open in X by hypothesis, X is locally connected at x- □ A similar proof holds for the following theorem: Theorem 4.2. A space X is locally path connected if and only if for every open set U ofX, each path component of\J is open in X. The relation between path components and components is given in the following theorem: Theorem 4.3. ffX is a topological space, each path component of X lies in a component ofX. IfX is locally path connected, then the components and the path components ofX are the same. Proof. Let C be a component of X; let x be a point of C; let P be the path component of X containing x. Since P is connected, P c C We wish to show that if A'is locally path connected, P = C. Suppose that P & C. Let Q denote the union of all the path components of X that are different from? and intersect C; each of them necessarily lies in C, so that C = P u Q. Because X is locally path connected, each path component of X is open in X. iherefore, P (which is a path component) and Q (which is a union of path STCut8) 3re °Pen '" X' So they institute a separation of C. This con- conthe fact that C is connected. □
Local Connectedness 163 I*4 Exercises ,he components and path components of *,? t^^e components and path components of JT (in the product C) topology)? unjform t0pology. Show ^ x and y lie in the same com- (W ConS^eof>. if and only if the sequence x -y; ., ^ in the box topology. Show that x and y lie in the same com- (C) Z, nf /f» if the sequence x - y is "eventually zero." »(d) Prove the converse of (c). , m a snace X, let us define x ~ y if there is no separation X = A U B of X So Sotat open sets such that * e /( and y e B. Show this is an equivalence relation We shall call the equivalence classes the quasi components of X. Show [hat each component of X lies in a quasicomponent of X. 4 Determine the quasicOmponents, components, and path components of the ' following subspaces of R2. [Here K denotes the set {l//i|« e Z+] and -K denotes the set {-l/»|n e Z+}.] A = (K x [0,1]) U @ x [0, 1]), C = 5 U ([0,1] x 0), D = (Kx [0,1]) u (-AT x [-1, 0]) U ([0, 1] x -K) U ([-1, 0] x K), 5. Show that if X is locally connected, then the quasicomponents of X are the same as the components of X. t- If/: X—> Yis continuous and X is locally connected, is f(X) necessarily locally connected? What if/is both continuous and open? 7. Show that / x I in the dictionary order topology is locally connected but not locally path connected. What are the path components of this space? ■ L« Xbe locally path connected. Show that every connected open set in X is P«h connected. ■ ^t * denote the rational points of the interval [0, 1] x 0 of A2. Let T denote (a) ShoTtht r'me Segments •ioinin8 the P°int P = 0 x 1 to points of X. 0» Find a subs i* f^i COnnected' but is loca"y connected only at the point p. °f 'ts points tHat 'S Path connected but is loca"y connected at none 10. ofTfterekfl'0 te c°nnect«« im kleinen at * if for every neighborhood U Sl)o* that if A-is Z SUbSCt ^ °f V that contains a neighborhood of x. wnaected. W;«- lu ^ lm kleinen at <*■<* of its points, then X is locally '• bhow that components of open sets are open.]
164 Connectedness and Compactness 11 Consider the "infinite broom" X pictured in Figure 10. Show that ' locally connected at p, but X is connected im kleinen at p. [Hint: Any neighborhood of p must contain all the points a,.] Figure 10 3-5 Compact Spaces The notion of compactness is not nearly so natural as that of connected- connectedness. From the beginnings of topology, it was clear that the closed interval [a, b] of the real line had a certain property that was crucial for proving such theorems as the maximum value theorem and the uniform continuity theorem. But for a long time it was not clear how this property should be formulated for an arbitrary topological space. It used to be thought that the crucial prop- property of [a, b] was the fact that every infinite subset of [a, b] has a limit point, and this property was the one dignified with the name of compactness. Later, mathematicians realized that this formulation does not lie at the heart of the matter, but rather that a stronger formulation, in terms of open coverings of the space, is more central. The latter formulation is what we now call com- compactness. It is not as natural or intuitive as the former; some familiarity with it is needed before its usefulness becomes apparent. Definition. A collection G of subsets of a space X is said to cover X, or to be a covering of X, if the union of the elements of G is equal to X. It is called an open covering of X if its elements are open subsets of X. Definition. A space A'is said to be compact if every open covering & of X contains a finite subcollection that also covers X. Example 1. The real line R is not compact, for the covering of R by open intervals « ={(/i,/i+ 2) |« eZ} contains no finite subcollection that covers R. Example 2. The following subspace of R is compact: X = {0] u{l/«|« e Z+}. Given an open covering e of X, there is an element U of C containing 0. The J i
Compact Spaces 165 & . ., bt finitely many of the points 1/n; choose, for each point t {/contains ail ^ ofa containing it. The collection consisting of these of *aot K*' Ha with the element U, is a finite subcollection of a that covers X- Any space X containing only finitely many points is necessarily ^"because in this case every open covering of X is finite. Ple4. The interval @, 1] is not compact; the open covering d ={(!/". 1]|" e Z+} • s no finite subcollection covering @, 1]. Nor is the interval @, 1) mcf the same argument applies. On the other hand, the interval [0, 1] is C°»cf' you are probably already familiar with this fact from analysis. In my case' we shall prove it shortly. In general, it takes some effort to decide whether a given space is compact otnot^First we shall prove some general theorems that show us how to con- construct new compact spaces out of given ones. Then in the next section we shall show certain specific spaces are compact. These spaces include all closed intervals in the real line, and all closed and bounded subsets of R". Let us first prove some facts about subspaces. If Y is a subspace of X, a collection a of subsets of X is said to cover Y if the union of its elements contains Y. Lemma 5.1. Let Y be a subspace of X. Then Y is compact if and only if mry catering ofY by sets open in X contains a finite subcollection covering Y. Proof. Suppose that Y is compact and G = {A,}.QJ is a covering of Y by *ts open in X. Then the collection {A. ny|«e/j 8»covering of y by sets open in Y; hence a finite subcollection covers ^" ° *''''' A" ° y] Conversed ^"" " ' *""*■ K * Subcollection of a that covers Y- PactLetO' = JmPkT thC g'Ven condition holds; we wish to prove 7com- '* A open in X Such thaT^ °f * ^ ^ °Pe" '" Y' ¥°T ™Ch "' ch°°SC * finite subcollectio "(A °0Vering of Y by sets °Pen in x- By hypothes: ^lection of n' .l n "" ' * ■' ^«n} covers y. Then \A! A' ) is n 01« that covers Y. Q l " '''' "J Theorems2 £ veryc'osed subset ofa compact space is compact.
166 Connectedness and Compactness Proof. Let Y be a closed subset of the compact space X. Give Chap.j :nacOVfefin. Proof. Let Y be a p p . Given a of Y by sets open in X, let us form an open covering <B of X by to a the single open set X — Y, 8 <B = fl U {X - Y}. Some finite subcollection of (B covers X. If this subcollection contains the X — Y, discard A'— 7; otherwise, leave the subcollection alone. TheresuT' ing collection is a finite subcollection of a that covers Y. Q Theorem 5.3. Every compact subset of a Hausdorff space is closed. Proof. Let Y be a compact subset of the Hausdorff space X. We shall prove that X — Y is open, so that Y is closed. Let x0 be a point of X — Y. For each point y of Y, let us choose disjoint neighborhoods U, and V, of the points x0 and y, respectively (using the Hausdorff condition). The collection [Vy\y e Y] is a covering of Y by sets open in X; therefore, finitely many of them V,,, . . . , Vyn cover Y. The open set V = Vyi U • • • U Vy, contains Y, and it is disjoint from the open set U = Uyt n • • • n £/„„ formed by taking the intersection of the corresponding neighborhoods of xt. See Figure 11. For if z is a point of V, then z G Vyi for some i, whence z $. Uyi and so z <£ U. Figure 11 Therefore, £/is a neighborhood of x0 disjoint from 7. Hence X-Y& open, as desired. □ The statement we proved in the course of the preceding proof will be useful to us later, so we repeat it here for reference purposes:
Compact Spaces 167 y is a compact subset of the Hausdorff space X and x0 is A- J :,t Aivoint open sets U and V ofX containing x0 and Y, ist disjoint open Once we prove that the interval [a, b] in R is compact, it follows Example 5. ^ any closed subse( of fa y {s cOmpact. on the other from Theorem ^ -^^ TheOrem 5.3 that the intervals (a, b] and (a, b) in R hand, '< act (Which we knew already) because they are not closed in the Hau" iotff space R- 6 One needs the Hausdorff condition in the hypothesis of ExAMP5L3 consider, for example, the topology 3/ on the real line consisting 'ftoand all complements of finite sets. The only proper subsets of R that closed in this topology are the finite sets. But every subset of R is com- ^ this topology, as you can check. Theorem 5.5- The image of a compact space under a continuous map is compact- Proof. Let/: X — Ybe continuous; let X be compact. Let G be a cover- iogoftheset /(A') by sets open in Y. The collection [f'KA)\A G G) is a collection of sets covering X; these sets are open in X because/is con- continuous. Hence finitely many of them, say . Then the sets /4, /4n cover f(X). Q One important use of the preceding theorem is as a tool for constructing "wneomorphisms: """Pact a //v l ^: X ~* Y be a bijective continuous function. If X is * h Hausdorff, then f is a homeomorphism. k y;this^in ShaU Pr°VC that 'mageS of closed sets of X under/are closed °Ve Cii • -> by Th°Ve C0ntinuity of the maP /"'• If A is closed in X, then A is """Pact Since v'-T 5" Therefore> by the theorem just proved, f{A) is is Hausdorff, /(/() is closed in Y, by Theorem 5.3. Q worem 5.7 t-j """ f. W ' ^ Pr°duct °f finitely many compact spaces is compact. ren) follow<fh°V*thatthe Product of two compact spaces is compact; oy induction for any finite product.
168 Connectedness and Compactness Step 1. Chap. 3 : that we are given spaces A'and Y, with K compact X int of X, and N is an open set of X x Y containing tt x x Y of XX K. We prove the following: s ne There is a neighborhood W o/x0 in X such that N contains the entire set W X Y. The set W X Y is often called a tube about x0 X Y. First let us cover x0 X 7 by basis elements £/ X K (for the topology of XX Y) lying in N. The space *0 X Y is compact, being homeomorphic to Y. Hence we can cover x0 X Y by finitely many such basis elements U,xV,,...,UnX Vn. (We assume that each of the basis elements U, X V, actually intersects x x Y, since otherwise that basis element would be superfluous; we could discard it from the finite collection and still have a covering of x0 x Y.) Define W=(/,n • ■ ■ n Un. The set W is open, and it contains x0 because each set U, X V, intersects x0 x Y. We assert that the sets U( X V,, which were chosen to cover the slice x0 X Y, actually cover the tube W x Y. See Figure 12. Let x x y be a point of If x 7. Consider the point ,v0 X y of the slice x0 x Y having the same ^-coordinate as this point. Now x0 x y belongs to U, X V, for some /, so "lat y e v But x e V, for every j (because x e W). Therefore, we have * X y e C/( x V,, as desired. Since all the sets U, X V, lie in N, and since they cover W X Y the tube rr x. i nes in TV also.
Compact Spaces 169 **"* ,w .heorem Let X and Y be compact spaces. Let SKP2. Now we P-ve'he theor-L ^ heslice,o x yis compact abc'nopenc-nngo ^y. ^ ^ d 0 7mav therefore be cov containing x0 X Y, by ^ "'« V^nta^e * X y about ,D X 7, where ^is p 1, the^1 ^ y°s covered by finitely many elements At, ...,Amof l0 w in J we can choose a neighborhood Wx of x such that ^rrcanb^ covered by finitely many elements of a. The col- W xfo^ ds Wx is an open covering of X; therefore by e"" ofV, there exists a finite subcollection covering X. The union of the tubes W,xY,...,WtxY •sail of X X 7; since each may be covered by finitely many elements of a, so may XX Y be covered. Q The statement proved in Step 1 of the preceding proof will be useful to us later, so we repeat it here as a lemma, for reference purposes: Lemma S.8 (The tube lemma). Consider the product space X X Y, where Y is compact. If N is an open set ofXxY containing the slice x0 X Y of X X Y, then N contains some tube W X Y about x0 X Y, where W is a neighborhood of x0 in X. Example 7. The tube lemma is certainly not true if Y is not compact. For example, let Y be the >>-axis in R2, and let N = [xxy;\x\<l/(y* + l)}. Then N is an open set containing the set 0 X R, but it contains no tube about 0 x R, It is illustrated in Figure 13. nere is an obvious question to ask at this point. Is the product of infinitely M\nrmP°C* SPaCa Comp.actl One would h°Pe that the answer is "yes," *e nam 11' ^ feSUlt 'S imP°rtant (and difficult) enough to be called by h Ptov! hT Wh° Pr°Ved k; k 'S Ca"ed the Tychonoff theorem. nected WeTrS Z t 'hat * cartesian Product of connected spaces is con- *" ^ T St f°r ^ products and derived the general case from d WeTr Z t p ^ provTn ft St f°r ^ products and derived the general case from ver there ^ ^ Cartesian products of compact spaces are compact, ecasedem "h ^ l° g° fr°m finite Products to infinite ones. The Because ona^S«anfntirely new aPProach> and the proof is a difficult ssicm itt this ch ^ and als° to avoid los!ngthe main thread °f our chapter, we have decided to postpone it until later. How-
170 Connectedness and Compactness Chap. 3 Figure 13 ever, you can study it now if you wish; the section in which it is proved (§5-1) can be studied immediately after this section without causing any disruption in continuity. There is one final criterion for a space to be compact, a criterion that is formulated in terms of closed sets rather than open sets. It does not look very natural nor very useful at first glance, but it in fact proves to be useful on a number of occasions. First we make a definition. Definition. A collection 6 of subsets of X \s said to satisfy the finite inter- intersection condition if for every finite subcollection of e, the intersection C, (~*\ • • • O Cn is nonempty. Theorem 5.9. Let Xbe a topological space. Then X is compact if and only if for every collection Q of closed sets in X satisfying the finite intersection con- condition, the intersection Hcse c of all the elements ofQ is nonempty. Proof. Given a collection a of subsets of X, let e = {X-A\A<=a) be the collection of their complements. Then the following statements hold: A) a is a collection of open sets if and only if 6 is a collection of closed sets. B) The collection a covers *if and only if the intersection (\<*Cd all the elements of e is empty. C) The finite subcollection {A,,..., A,} of a covers X if and only if the intersection of the corresponding elements C, = X - A, of <5 's empty. .
Compact Spaces ^ statement is trivial, while the second and third follow from DeMor- J lw: s law: P x (J f of the theorem now proceeds in two easy steps: taking the PIX!(of the theorem), and then the complement (of the sets)! contrapos'ti i ^ ^^ compact ;s equivalent to saying: "Given any col- ^ f open subsets of X, if a covers X, then some finite subcollection )ection a o ^ ^ statement is equivalent to its contrapositive, which is °f *f llowing- "Given any collection a of open sets, if no finite subcollection % overs X, then fl does not cover X" Letting 6 be, as above, the collection fv~A\A e a} and applying (l)-C), we see that this statement is in turn equivalent to the following: "Given any collection 6 of closed sets, if verv fi"'te intersection of elements of 6 is nonempty, then the intersection of all the elements of 6 is nonempty." This is just the condition of our theorem. □ A special case of this theorem occurs when we have a nested sequence C, d C2 3 ... Cn => Cn+, => ... of closed sets in a compact space X. If each ofthesetsCn is nonempty, then the collection & = {Cn)nez> automatical- automatically satisfies the finite intersection condition, as you can easily check. Then the intersection is nonempty. We shall use the closed set criterion for compactness in the next section toprove the uncountability of the set of real numbers, in Chapter 5 when we prove the Tychonoff theorem, and again in Chapter 7 when we prove the S^f? heOrem Atll i i p f? n y' in Pr°ving the Tychonoff theorem, we the foBowmg slight reformulation of it, whose proof we leave to * tS C°mpaCt V^only if for every collection the finite intersection condition, the intersection **+ closures is nonempty. ^ * iUta,,,,,. Exercises • does ^pactne^of ^und'f °" °? SCt *'* SUPPose that 3' = *■ What . ft) J** under the other" ™* °* """ topol^"es ™P* about com *■ ' **flOW that if v • * 3=3' - •% iSSSSTunder both
172 Connectedness and Compactness Chap. 3 2 (a) Show that every subset of R is compact in the topology 3/. (See Example 6) (b) Is [0, 1] compact as a subspace of R in the topology •> 3 = {/j I ^{ — A is countable or all of R) ? In the lower limit topology «;? 3. Show that a finite union of compact sets is compact. 4 Show that every compact subset of a metric space is bounded in that metric " and is closed. Find a metric space in which not every closed bounded subset is compact. 5. Let A and B be disjoint compact subsets of the Hausdorff space X. Show that " there exist disjoint open sets V and V containing A and B, respectively. 6. Show that if/: X —» Y is continuous, where X is compact and Y is Hausdorff, " then/is a closed map (that is,/carries closed sets to closed sets). 7. Prove Corollary 5.10. 8. Show that if Y is compact, then the projection n, : X x Y—> X is a closed map. 9. Theorem. Let f: X —> Y; let Y be compact Hausdorff. Then f is continuous if and only if the graph off, Gf = {x X /(a) | a e X), is closed in X X Y. [Hint: If Gf is closed and V is a neighborhood of /(.v0), find a tube about x0 x (Y — V) not intersecting Gf.] 10. Generalize the tube lemma as follows; Theorem. Let A and B be subsets of X and Y, respectively; let N be an open set in X x Y containing A x B. If B is compact, then there exists an open set U in X such that AxB^UxBczN. If A and B are both compact, then there exist open sets U and V in X and Y, respectively, such that 11. (a) Prove the following partial converse to the uniform limit theorem: Theorem. Letf,:X—*Rbea sequence of continuous functions, withfn(x) —> fix) for each x e X. Iff is continuous, and if the sequence f, is monotone increas- increasing, and if X is compact, then the convergence is uniform. [We say that f, is monotone increasing if /„(*) <fn+l(x) for aI, „ and x] (b) Give examples to show that this theorem fails if you delete the requirement that X be compact, or if you delete the requirement that the sequence be monotone. [Hint: See Exercise 9 of §2-10.] e xercise 9 of §2-10.] rln nJ°rCm' U' Xbea ™mP"ct Hausdorff space. Let a be a collection of closed connected subsets of X which is simply ordered by proper inclusion. Then Is connected.
Compact Sets in the Real Line 173 '** , n is a separation of Y, choose disjoint open sets V and V l*»': IfC°r and /respectively, and show that ^^ containing0 anu ' M _ (# u K)) isnot empty-] hav£ s(udied t0p0|0gica| groups: osed ^ y) = ^.^ ]f c ^ A. 3-6 Compact Sets in the Real Line ■n, theorems of the preceding section enable us to construct new com- , m, from old ones; but in order to get very far we have to find some Sc compact sets. The natural place to start is the real line. We prove ht every closed interval in R is compact. As an application, we prove the maximum value theorem of calculus, suitably generalized. Other applications include a characterization of all compact subsets of /?", and a proof of the uncountability of the set of real numbers. It turns out that in order to prove every closed interval in 7? is compact, we need only one of the order properties of the real line—the least upper bound property. We shall prove the theorem using only this hypothesis; then it will apply not only to the real line, but to well-ordered sets and other ordered sets as well. Theorem 6.1. Let X be a simply ordered set haying the least upper bound property, h the order topology, each closed interval in X is compact. Proof. Given a < b, let G be a covering of [a, b] by sets open in [a, b] in foesubspace toP°l°gy (which is the same as the order topology). We wish to prove the existence of a finite subcollection of a covering [a, b]. from?tt F'f WC pr°Ve the fol[owinS: If * « a point of [a, b] different fecovered '* " ^ V > * '" ["' b] SUCh that the interVal ^ ^ can '«i Dy at most two elements of a. Wx vlon" ''T2?316 SU°CeSSor in X> let y be this ^mediate successor. 111051 two eCn S °p^ tW° P°'ntS * and y> S0 that H can be covered by at ^«Uofa7nf'a' "° immediate successor in X, choose an ittterval of the form rimnf f Be°aUSe X * b and A is °Pen> A contains an ften * interval r?ir C)> 'Z™ ° ''" k b]' Ch°0Se a Point V in (*• &. ™ [x, y] ,s covered by the single element A of a. Y > ° °f ^ ^ ^ that the interval many elements of a. Applying Step 1 to the
174 Connectedness and Compactness Chap. 3 case x = a, we see that there exists at least one such>\ Let c be the least bound of the set C; then a < c < b. Upper Step 3. We show that c belongs to C; that is, we show that the [a, c] can be covered by finitely many elements of a. Choose an elemenT!! of a containing c; since A is open, it contains an interval of the form (rf \ for some d in [a, b]. If c is not in C, there must be a point z of C lying in th interval (d, c), because otherwise d would be a smaller upper bound on C than c. See Figure 14. Since z is in C, the interval [a, z] can be covered b finitely many, say /;, elements of G. Note that [z, c] lies in the single element A of G, whence [a, c] = [a, :] U [z, c] can be covered by n + 1 elements of G. Thus c is in C, contrary to assumption. or y a d c a c b Figure 14 Figure 15 Step 4. Finally, we show that c = b, and our theorem is proved. Suppose that c < b. Applying Step 1 to the case x = c, we conclude that there exists a point y > c of [a, b] such that the interval [c, y] can be covered by finitely many elements of G. See Figure 15. We proved in Step 3 that c is in C, so [a, c] can be covered by finitely many elements of G. Therefore, the interval [a, y] = [a, c] U [c, y] can also be covered by finitely many elements of G. This means that;' is in C, contradicting the fact that c is an upper bound on C. Q Corollary 6.2. Every closed interval in R is compact. Now we characterize compact subsets of R": Theorem 6.3. A subset A of Rn is compact if and only if it is closed and is bounded in the euclidean metric d or the square metric p. Proof. It will suffice to consider only the metric p\ the inequalities p(x, y) <, d(x, y) <> */~n~p(x, y) imply that A is bounded under d if and only if it is bounded under p- Suppose that A is compact. Then, by Theorem 5.3, it is closed. Consider the collection of open sets ,
Compact Sets in the Real Line 175 l>$ ■ ,„ of R" Some finite subco.lection covers A. It followsi that ^e union '"» ** M Therefore, for any two points x and y of A, we A c B,@> MJ° Thus A is bounded under A have^'v)r nose that /I is closed and bounded under p; suppose that /-inversely. suPr r __•_,, c , rhnncp n noint x« of A. and ave^r «nose that /I is closed and b p Conversely, suppose th ^ ^^ g pojnt ^ of ^ &»*£ 'rSSngie'neiality implie. that „(*, 0) < N -f-6 for let />(-*<>• W 'if^ = ^ + 6> then ^ is a subset of the cube [-P, P]", which is SactnBeing closed, ^ is also compact. Q often remember this theorem as stating that the collection of t sets in a metric space equals the collection of closed and bounded compTahCjs statement is clearly ridiculous as it stands, because the question as ^k-h'ch sets are bounded depends for its answer on the metric, whereas vhich sets are compact depends only on the topology of the space. Example 1. The unit sphere S"~' and the closed unit ball B" in R" are compact, since they are closed and bounded. The set A ={.v x (I/.v)|0<x< 1} is closed in R1, but it is not compact because it is not bounded. The set S = {x x (sin (i/.v))|0 <-v< 1} is bounded in R1, but it is not compact because it is not closed. Now we prove the maximum value theorem of calculus, in suitably gener- generalized form. Theorem 6.4 (Maximum and minimum value theorem). Let f : X —> Y be continuous, where Y is en ordered set in the order topology. IfX is compact, 'hen there exist points c and d in X such that f(c) < f(x) <£ f(d) for every T °fca'CuIus is the sPecial ca^ of this theo- en we take X to be a closed interval in R and Y to be R com- W^lXZT /W com ^"'and MbeLit ! geSt e'ement M a"d a Smallest el™t »• Then ^s cand dot X ' WC mUSt h3Ve m = ^c) and " = /CO for some lU h "° 'arSest e'ement, then the collection formsann K~°°, a)\a e A] f;- Si"Ce ^ is compact, some finite subcollection ^ £££££? ■>■■»-«, bd.-,, ,„ simi'arareum f u thcy cover A- r^ent shows that ^has a smallest element. D
176 Connectedness and Compactness _, Finally, we prove that the real numbers are uncountable. The thing about this proof is that it involves no algebra at all—no decimV binary expansions of real numbers or the like—just the order properties f ^ In fact, we prove the following more general result: • Theorem 6.S. Let X be a (nonempty) compact Haiisdorff space If ev point ofX is a limit point ofX, then X is uncountable. ry Proof. Step I. First we show that, given a (nonempty) open set V of V and given x e X, there exists a (nonempty) open set V contained in V such that V does not contain x. The point x may or may not be in U. But in either case, we can choose point y in U that is different from x. This is possible if x is in U because x i a limit point of X (so that U must contain a point y different from x). And 't is possible if x is not in U because [/is nonempty. Let W, and I-K, be disjoint neighborhoods of x and y, respectively; then V = U n W-, is the desired open set, whose closure does not contain x. See Figure 16. x or x Figure 16 Step 2. We show that given/: Z+ — X, the function/is not surjective. It follows that X is uncountable. Let x, = f(n). Apply Step 1 to the nonempty open set U = X to choose a nonempty open set Vl cz X such that K, does not contain .v,. In general, given KB_, open and nonempty, choose Vn to be a nonempty open set such K" c K»-' and v« d°es not contain xn. Consider the nested sequence Pi = ^ ^ ... of nonempty closed sets of X Because Zis compact, there is a point x B f\K oy ineorem 5.9. The point x cannot equal *„ for any n, since a: belongs to K and jr. does not. Q OmXfaty <J.,J. Every closed interval in R is uncountable.
Compact Sets in the Real Line 177 Exercises titxb an ordered set in which every closed interval is compact, , prove that i bound property. *" \Ta clntld Iric space having more than one point is uncountable. * ^^ric space with metric d; let A be a subset of X Recall that the '0% is defined by the equation exists. Show that if A is compact, then diam A exists and equals </(a,, «2) for some <*i, e ^> , ffc,i metric space with metric d; let A <= X 4Jf,f,e™, define «/(*,,« = gib (*«)l<£4 Show that the map carrying a:' into d(x, A) is a continuous map of X into R. (b) Is it true that for some a e /f, we have rf(x, -4) = d(x, a)l Is this true if A is closed? If A is compact ? (c) Is it true that x e Ait and only if d(x, A) = 0? (d) Define the (-neighborhood of A in X to be the union V{A, €) = LUx #/(«. f). Show that £/(/f, f) equals the set {x\d(x,A)<e}. (e) Show that if A and B are disjoint closed subsets of X and 5 is compact, then for some f, the f-neighborhoods of A and B are disjoint. (f) Does (e) hold if B is not compact? 5. Let Xbe a compact Hausdorff space. Show that if{/4n] is a countable collection ofclosedsetsin X, each of which has empty interior in X, then there is a point of Jwhich is not in any set An. [Hint: Imitate the proof of Theorem 6.5.] This is a special case of the Batre category theorem, which we shall study in Chapter 7. •6. LeMo be the closed interval [0, 1] in R. Let At be the set obtained from AD oy deleting its "middle third" (j, *). Let A2 be the set obtained from A, by delet- «* its "middle thirds" (J, |) and (J, f). In general, define An by the equation A intersection V 3 3 / C n«-z.A. (aiT* tl^Cantor set^ i» is a subspace of [0,1]. S°W 3' °is tot" disconnected. slowS. of length [li^A 1" ^ auUnion °f finitely many disJ°int closed intervals Show that Ive'rvlr, ^ ** end P°intS °f theSe intervaIs Iie in C. Conclude thl^^ °f CIS a limit point of C. "ciude that Cis uncountable.
l78 Connectedness and Compactness Chap. 3 3.7 Limit Point Compactness As indicated when we first mentioned compact sets, there are other fOr. Ulat ons of the notion of compactness that are frequently useful. In this s^ct on we introduce one of them. Weaker in general than compactness, it coincides with compactness for metrizable spaces One application is the thWd theorem of calculus quoted at the beginning of the chapter, the uniform continuity theorem, which we shall prove in suitably generalized form. Definition. A space X is said to be limit point compact if every infinite subset of* has a limit point. In some ways this property is more natural and intuitive than that of compactness. In the early days of topology, it was given the name "com- "compactness," while the open covering formulation was called "bicompactness." Later, the word "compact" was shifted to apply to the open covering defini- definition, leaving this one to search for a new name. It still has not found a name on which everyone agrees. On historical grounds, some call it "Frechet com- compactness"; others call it the "Bolzano-Weierstrass property." We have invented the term "limit point compactness." It seems as good a term as any; at least it describes what the property is about. Theorem 7.1. Compactness implies limit point compactness, but not con- conversely. Proof. Let *be a compact space. Given a subset A of X, we wish to prove that if A is infinite, then A has a limit point. We prove the contrapositive—if A has no limit point, then A must be finite. So suppose that A has no limit point. Then A contains all its limit points and is therefore closed. Being a closed subset of a compact space, it is com- compact. For each a in A, we can choose a neighborhood Ua of a such that Ua does not intersect A - {a}, since a is not a limit point of A. The set A is cov- covered by the open sets Ua; being compact, it can be covered by finitely many, say n, of them. Since each set Ua contains only one point of A, the set A itself contains n points. The example following describes a space that is limit point compact but not compact. Q *' Consider the minimal uncountable well-ordered set Sa, to °Og? Th Space * is not comPact, for it is not closed in&. ?? omPact, for it is not clos a sublet flnf J?v! P°im C°mpaCt: Ut A be an infinite subset of Sa- Ch°°Se f 1iat a sublet flnf J?v! upper bofnd i1niat.1thC°U»tabIy infi"ite- Bei"g COUntabIe> th6 Set B ^ *" is thesmanl °\ f" BlS* subset of the interv*l l<">- « of Sn, where*. * the smallest element of Sa. Since Sa has the least upper bound property, the
Limit Point Compactness > r Al is compact. By the preceding theorem, B has a limit point x .n(erval [ao, *I* aIsQ a Ijmjt poim of A Thus Sa is Iimit point compact. >n [a" \ 1 know 5n is not metrizable, for it does not satisfy the "sequence WC file 3 of §7-10). Once we prove that compactness and limit point lemma' <fc^™ equivaIent in metrizable spaces, it follows that Sa is not a<eeither even though it does satisfy the sequence lemma. For Sn is t compact but not compact. derto prove the uniform continuity theorem, we need first to prove !" lemma concerning open coverings of metric spaces. Recall that of a bounded subset A of a metric space (X, d) is the number 2)|fli,fl2 G A}. Lemma 7.2 {The Lebesgue number lemma). Let d be an open covering of the metric space (X, d). IfX is compact, there is a 8 > 0 such that for each subset o/X /laving diameter less than d, there exists an element of Q. containing it. The number 5 is called a Lebesgue number for the covering G. Proof. Step 1. Because X is compact, it is necessarily limit point compact. We shall prove that limit point compactness of X in turn implies that every infinite sequence CO in X has a convergent subsequence. That is, there is an increasing sequence n, < n2 < ■ ■ ■ <//,< of positive integers such that the sequence *„„ *„„ ..., *„„ ... converges. *fS£S£X£?Y has a conversent subsequence> we say Given the sequence (*„), let us consider the set = {x.\neZ+}. Sthat^-Tffi WC assert that ^"e is a *rtion,iet/;z !~,'S ' ,"' Y "^ ValU£S °f "• tTo P™ve this as- J« Z is the union of heI'" X'?,g fUnCti°n defined * A») = *.. Then ^S f suppose th *.. e^(*,l). Ae P°Sitive inte^r «'- • » given. Because the ball B(x, 1/0
180 Connectedness and Compactness Chap. 3 intersects A in infinitely many points, we can choose an index n, > ^ tuch a-., e B(x, I/i). Then the subsequence *.„ .*_,. • - converges to x. Step 2. Now we show that if X is sequentially compact, then every open covering a of X has a Lebesgue number <5. We shall prove the contrapositive: Suppose there is no 5 > 0 such that every set of diameter less than 5 lies in at least one element of a. Then X is not sequentially compact. So let us assume there is no such S. This means that for each <5 > 0, there exists a subset of X having diameter less than 5 which does not lie inside any element of a. In particular, for each // e Z+, we can choose a set C, having diameter less than I/// which is not contained in any element of a. Choose, for each n, a point a-, of Cn. We assert that the sequence (*„) has no convergent subsequence. Suppose that (*„) had a convergent subsequence (.%•„,), converging to x, say. Now x belongs to some element A of a, and because A is open, there is an e > 0 such that 5(.v, e) <= /I. Choose / large enough that </(*„„ -v) < f/2 and I//i, < e/2. Now Cni lies in the I/«, neighborhood of .v,,,: it follows that C., c fl(.v, e). Then Cn, <= /I, contradicting the choice of the sets Cn. □ The preceding proof is a typical proof of nonconstructive nature—it shows 5 exists without giving any hint of how to find it. Now we prove the third of the theorems of calculus stated at the begin- beginning of the chapter, in suitably generalized form. Theorem 7.3 (Uniform continuity theorem). Let f : X —» Y be a continuous map of the compact metric space (X, dx) to the metric space (Y, dY). Then f is uniformly continuous. That is, given e > 0, there exists ad>0 such that for any two points xls x2 of X, dx(x1)x2)<5 ==> dy(f(x,), f(x2))< f. «dSS Lelabet? °' takC the°Pen C0Venn8 °f Y * ba"S £()'' f/2) °f under fChn si °Pe" C°Venng of X ^ the inverse images of these balls ^0"! S t0 be a LebesS"e number for the covering a. Then if ft f/2, Then ^^^^''SKf*^ ^'"
Limit Point Compactness 181 that compactness and limit point compactness are equiva- 7d LetXbeantetrizablespace. Then the following are equivalent: Theorem /•*• i\\ X is compact. filx't'segue'>''iallyC0"'paC'- /■We have proved above that A) =* B) => C); see the proofs of '7 1 and Lemma 7.2. We have also proved that sequential com- ^"^implies that every open covering of X has a Lebesgue number. We pactness 1 ^ ^ proof by showing that sequential compactness of X implies ""jTcUiess of X So assume that X is sequentially compact. Step I- First we show that for every f > 0, there exists a finite covering f/by f-balls. And once again, we prove the contrapositive: If for some ° 0 /cannot be covered by finitely many f-balls, then X is not sequentially compact. So suppose that X cannot be covered by finitely many f-balls. Construct asequence of points xn of Afas follows: First choose x, to be any point of X. Noting that the ball B(xu e) is not all of AT (otherwise A'could be covered by a single f-ball), choose xt to be a point of X not in B(xu e). In general, given xt,. ■ ■, x,, choose xn+, to be a point not in the union B(xt,e)u ••• uB(xn,e), using the fact that these balls do not cover X. Note that by construction rf(*»+i>*() > € for i = 1, .. ., n. Therefore, the sequence (a-,) can have no convergent subsequence; in fact, any ball of radius f/2 can contain xn for at most one value of n. Step 2. Now we prove X compact. Let a be an open covering of X. ince X B sequentially compact, the covering a has a Lebesgue number 5. XhJu n V' Ch°°Se a finite covering of X by balls of radius 512. Each of an Z, t Idmmeter at m°St 2S/3'So We can choose for each of the^ balls covers ™ q contain'ng it. We thus obtain a finite subcollection of a that Exercises f omPa:: topoiogy-show that ^ >< ^is Sh°w that [0 1T» ,• ■• topologies. ' nOt llmlt point impact in either the box or the uniform
182 Connectedness and Compactness Chap. 3 3. Show that the set [0,1] is not limit point compact as a subspace of R 4. A space X is said to be countably compact if every countable open cover" X contains a finite subcollection covering X. ng °f (a) Show that countable compactness implies limit point compactness (b) Prove the converse if X is Hausdorff. [Hint: If no finite subcollecf {£/„} covers X, choose xn i £/, U • • • U £/„.] on of 5. Let X be limit point compact. (a) If/: X—> Y is continuous, is f(X) necessarily limit point compact? (b) It A is a closed subset of X, is A necessarily limit point compact? (c) If X is a subspace of Z, is X necessarily a closed set in Z? (d) Repeat (a)-(c) under the additional assumption that the spaces involved are Hausdorff. *(e) It is not in general true that the product of two limit point compact spaces is limit point compact, even if the Hausdorff condition is assumed. But the examples are fairly sophisticated. See [S-S], Example 112. 6. Let X, Y, and Z be metric spaces. (a) If/: X —► Y and g : Y—> Z are uniformly continuous, is g°f: X—>Z necessarily uniformly continuous? (b) If /: X—> Y is a homeomorphism and / is uniformly continuous, is /-' necessarily uniformly continuous? 7. Let (X, d) be a compact metric space. Let /: X —> X be a continuous map. (a) We say that / is a contraction if there is a number a < 1 such that d(f(x), f(y)) <, <*d(x, y) for all points x, y e X. Show that if/is a contraction, there is a unique point x of X such that f(x) = x. [Hint: Consider C\f"(X).] (b) We say that/is an isometry if d(f(x), f{y)) = d(x, y) for all points x,y e X. Show that if/is an isometry, then /is surjective. [Hint: If x £ f(X), let the f-neighborhood of* be disjoint from f(X). Define x, = * and, in general, *»+i = /(*„). Show that d(xm xm) > € for n ^ m.] (c) Give examples to show these theorems fail if X is not compact. *8. Show that [0, I]1" is compact in the product topology. *3-8 Local Compactness^ In this section we study the notion of local compactness, and we prove the basic theorem that any locally compact Hausdorff space can be imbedded in a certain compact Hausdorff space, called its one-point compactification. Definition. A space X is said to be locally compact at x if there is some compact subset C of X that contains a neighborhood of x. If X is locally compact at each of its points, X is said simply to be locally compact. fThis section will be assumed in §5-3, in Chapter 7, and in §8-12.
Local Compactness 183 a co mpact space is automatically locally compact. P , The real line R is locally compact. The point .v lies in some *Z») JchTn turn is contained in the compact se, [a. bl The sub- ^' of rational numbers is not locally compact, as you can check. ■> The space R" is locally compact; the point .v lies in some basis Exam -■ ^ ^ ^ whjch Jn (urn |ies in the conlpact set eIenie')tx .'. x [o../>J. The space R" is not locally compact; none of its fasis elements are contained in compact sets. For if B = (a,.b,) x ■ • • x (</„. bn) x R x ■ ■ • x i? x • • • were contained in a compact set, then its closure B = [a,, b,] x ■ • ■ x [<?„. 6,,] X /? x • ■ • would be compact, which it is not. Example 3. Every simply ordered set A" having the least upper bound prop- property is locally compact: Given a basis element for X, it is contained in a closed interval in X, which is compact. Two of the most well-behaved classes of spaces to deal with in mathe- mathematics are the metrizable spaces and the compact Hausdorff spaces. Such spaces have many useful properties, which one can use in proving theorems and making constructions and the like. Jf a given space is not of one of these types, the next best thing one can hope for is that it is a subspace of one of these spaces. Of course, a subspace of a metrizable space is itself metrizable, so one does not get any new spaces in this way. But a subspace of a compact Hausdorff space need not be compact. Thus arises the question: Under what conditions is a space homeomorphic with a subspace of a compact Hausdorff space? Wegive one answer here. We shall return to this problem in Chapter JWhen we study compactifications in general. outff"k°"' UtXbe a locally comPact Hausdorff space. Take some object formiT t den°ted by the symbo1 °° for convenience, and adjoin it to X, ODenl8, hC S5 y== XU f°°}- ToPologize Y by defining the collection of ,',,ln Y t0 be all sets of the following types: g ^. where U is an open subset of X, C h The spa yC> Whefe ° 'S a comPact subset of X. e * is called the one-point compactification of X. '■"P'/seth!aVt116^ that th'S collection is> in ^ct, a topology on Y. The that'he intersect T @> a"d the SpaCe Yis a set of tyPe &■ Checking section of two open sets is open involves three cases-
l84 connectedness and Compactness Chap. 3 £/, n Ui isof type(l). (Y — Ci) n (y w) ~ ^ ' rfn(Y— C,) = ^1 n (A"— Ci) is of type A), because G is closed in X. Similarly, one checks that the union of any col- u e C, of open sets is open: U(y _ C,) = Y - (fl^,) = r - C is of type B). (UW u or- c,)) = t/ u (r- C) = r- (c - t/), which is of type B) because C-Uisa closed subset of C and therefore com- Pa<The basic properties of the one-point compactification are given in the following theorem. Theorem 8.1. Let X be a locally compact Hausdorff space which is not compact; let Y be the one-point compactification of X. Then Y is a compact Hausdorff space; X is a subspace 0/Y; the set Y — X consists of a single point; and X = Y. Proof. First we show that A" is a subspace of Y and X = Y. Given any open set of Y, its intersection with X is open in X, since U O X = U and (K — C) n Jf = A" — C, both of which are open in X. Conversely, any set open in JT is a set of type (l)and therefore open in Y. Since A" is not compact, each open set Y — C containing the point 00 intersects X. Therefore, 00 is a limit point of X, so that X = Y. To show that Y is compact, let a be an open covering of Y. The collection a must contain an open set of type B), say Y — C, since none of the open sets of type A) contain the point 00. Take all the members of a different from Y - C and intersect them with X; they form a collection of open sets in X covering C- Because C is compact, finitely many of them cover C; the cor- corresponding finite collection of elements of a will along with the element Y-C, cover all of Y. To show that /is Hausdorff, let x and y be two points of Y. If both of them he in X, there are disjoint sets U and V open in X containing them, respectively. On the other hand, if x e Zand y = 00, we can choose a com- dki ^1 '\u COntaim'ng a neighborhood U of x. Then U and Y - C are d.sjo.nt ne.ghborhoods of x and 00, respectively, in Y. Q comPactification of the real line R is homeo- the snarl r r h<™eomorphic to the sphere S*. If & is
Local Compactness 1S5 § our definition of local compactness is not very satisfying. In some ways ^ ^ ^ satisfies a given property "locally" if every Usually onefybitrariiy small" neighborhoods having the given property. x e f-^nf local compactness has nothing to do with "arbitrarily small" "•'■"n° js some question whether we should call it local compactness^an^. formuIation of local compactness, one more truly "local" HerelSjt is equivalent to our definition when X is Hausdorff. 8 2 Let X be a Hausdorff space. Then X is locally compact at , '' An,,lv if for every neighborhood U ofx, there is a neighborhood V ofx Jch that Vis compact and V czV. Proof it is clear that this new formulation implies local compactness; the set C = V is the desired compact set containing a neighborhood of x. Conversely, suppose that Cis a compact set containing a neighborhood of x. Let £/be an arbitrary neighborhood of x. Let A = C — U; being closed in C, the set A is compact. Apply Lemma 5.4 to choose disjoint open sets W and W about x and A, respectively. See Figure 17. Let V = W n (int C); recall thatlnt Cis the union of all open sets contained in C. Then Kisa neigh- neighborhood of x. Since X\s Hausdorff, C is closed in X; therefore, V a C, so that V is compact. Because Kis contained in W, which is disjoint from IV', the set V cannot intersect A. Therefore, V a C — A, so that V cz U. □ Figure 17 A s°mewhat slmnlPr nrnnC . . ot this lemma is available if X is locally he fact that the complement of U in the . - ■ WIu kk of v a a v !r ^P*** One chooses disjoint neigh- 1Sc°mpact and ic • ~ ' resPect»vely. Then the closure V of V "u ls contained in U.
186 Connectedness and Compactness £. Corollary 8.3. Let X be locally compact Hausdorff; let Y be a subspa ofX. IfY is closed in X or open in X, then Y is locally compact. c* Proof. Suppose that Y is closed in X. Given y e Y, let C be a compact set in X containing the neighborhood C/of y in Z. Then Cr\ Yis closed in C and thus compact, and it contains the neighborhood U r\ Y of y {„ y (We have not used the Hausdorff condition here.) Suppose that Y is open in X. Given ye Y, we apply the previous lemma to choose a neighborhood V of .y in X such that K is compact and V c y Then C = K is a compact set in Y containing the neighborhood V of y [a Y □ Corollary 8.4. A space X is homeomorphic to an open subset of a compact Hausdorff space if and only ifX is locally compact Hausdorff. Proof. This follows from Theorem 8.1 and Corollary 8.3. □ Exercises 1. Show that the rationals Q are not locally compact. 2. (a) Show that if \[Xa is locally compact, then each X* is locally compact and Xx is compact for all but finitely many values of a. (Assume that each X* is nonempty.) (b) Prove the converse, assuming the Tychonoff theorem. 3. Let X be a locally compact space. If/: X—> Y is continuous, is the space f(X) necessarily locally compact? What if / is both continuous and open? Justify your answer. 4. Show that [0, Vf is not locally compact in the uniform topology. 5. Show that the conditions in Theorem 8.1 characterize the one-point compacti- fication up to homeomorphism. 6. Show that the one-point compactification of R is homeomorphic with the circle S". 7. Show that the one-point compactification of Sa is homeomorphic with Sa- 8. Show that the one-point compactification of Z+ is homeomorphic with the subspace {0} u {l/n\n e Z+} of R. *9. Show that the one-point compactification of R2 is homeomorphic with the two- sphere S2. 10. Show that if G is a locally compact topological group and H is a subgroup, then G/H is locally compact. •11. (a) Prove the following: Lemma. lfp:X-^Yisa quotient map and ifZ is a locally compact Haus- Hausdorff space, then the map
Supplementary Exercises; Nets 187 jt = p x tg : X x Z >- Y x Z iS<">Hlt- If s~'(/0 is open and contains -v x >> choose open sets I/, and K • h /compact, such that .v x ^ e I/, x K and I/, x V c *->(/<). Given *"' y c s->(/4), choose an open set l/i+, containing p~'(p(U,)) such that J^j/c «-»(/<)■ Let u = U^i; show that ji(C/ x K) is a neighborhood of'«(.* x ^ containecl in ^ (b) Prove: Theorem. Let p: A~-> B and q : C —> D be quotient maps. If B and C are locally compact Hausdorff spaces, then p x q ; A X C —> B X D is a quotient map- * Supplementary Exercises: Nets We have already seen that sequences are "adequate" to detect limit points, continuous functions, and compact sets in metric spaces. There is a generalization wtion of sequence, called a net, that will do the same thing for an arbitrary ■ical space. We give the relevant definitions here, and leave the proofs as js Recall that a relation < on a set A is called a partial order relation if the following conditions hold : A) a < a for all a. B) If a < 0 and 0 < a, then a = p. C) If a < P and P < y, then a < y. Now we make the following definition: A directed set J is a set with a partial order < such that for each pair &, P of elements of J, there exists an element y of J having the property that a < y and Hi- 1. Show that the following are directed sets: (a) Any simply ordered set, under the relation <. (b) The collection of all subsets of a set S, partially ordered by inclusion (that is, A < B if A c b). (c) A collection a of subsets of S that is closed under finite intersections, parhally ordered by reverse inclusion (that is, A < B if A => B). Hie collection of all closed subsets of a space X, partially ordered by inclu- "ion. sudfT KoUis said to ** cofinaI in J if f°r each « e J, there exists 0 e K that a ^ p. show that if J is a directed set and K is cofinal in J, then K 1Sa directed set. into X^ t0pological sPace- A ne* in X is a function / from a directed set / symbol /A6 J> WC USUally denote/(oc) by ^^ Wedenote the net/i'self by the Thenit T''^ mCKly by ^ if the index set is understood. e U«) is said to converge to the point x oT X (written x* —* x) if for
188 Connectedness and Compactness Chap. 3 each neighborhood U<Jx, there exists « e/such that Show that these definitions reduce to familiar ones when / = Z+. A. Suppose that (Xa)tS). >xir\X and Ck.)«£/-—>- y in Y. Show that (jr. X ya) *■ x x y in X x j . 5. Show that if Xis Hausdorff, a net in Xconverges to at most one point. 6. Theorem. Let A e X Then x e A if and only if there is a net ofpoints ofA converging to x. [#/>//■ To prove the implication =>, take as index set the collection of all neighborhoods of x, partially ordered by reverse inclusion.] 7. Theorem. Let f: X— K. TVn / w continuous if and only if for every con- ' vergent net (jr.) in X converging to x, say, the net (/(x.)) converges to f(x), 8. Let/: J—> Xbe a net in X; let /(a) = *a. If .STis a directed set and g : K-*J is a function such that (i) i<J=>g(O<g(J), (ii) ,?(#) is cofinal in /, then the composite function/" g : K —> X is called a subnet of (a-J. Show that if the net (*,) converges to x, so does any subnet. 9. Let (*a)«e/ be a net in X. We say that x is an accumulation point of the net (jr.) if for each neighborhood U of x, the set of those a for which *a e G is cofinal in/ Lemma. The net (*„) Aas the point x as an accumulation point If and only if some subnet of(x^) converges to x. [Hint: To prove the implication =>■, let K be the set of all pairs (a, U) where a e / and U is a neighborhood of x containing x*. Define (a, U) < @, V) if a < P and V c G. Show that JsTis a directed set and use it to define the subnet.] 10. Theorem. X Is compact if and only if every net in X has a convergent subnet. [Hint: To prove the implication =>, let Bx = {xp \ 0C < 0} and show that {£«) satisfies the finite intersection condition. To prove <=, let a be a collection of closed sets satisfying the finite intersection condition, and let (B be thecollec- tion of all finite intersections of elements of a, partially ordered by reverse inclusion.] 11. Corollary. Let G be a topological group; let A and B be subsets ofG. If A is closed In Gand B is compact, then A-Bis closed in G. 1 '• tlTst glve a Proof using sequences, assuming that G is metrizable.] exercises remain correct if condition B) is °mtt» '. Many 1
4, Countability and Separation Axioms The concepts we are going to introduce now, unlike compactness and connectedness, do not arise naturally from the study of calculus and analysis. They arise instead from a deeper study of topology itself. Such problems as imbedding a given space in a metric space or in a compact Hausdorff space are basically problems of topology rather than analysis. These particular problems have solutions that involve the countability and separation axi- axioms. We have already introduced the first countability axiom; it arose in con- connection with our study of convergent sequences in §2-10. And we have also studied one of the separation axioms—the Hausdorff axiom. In this chapter we shall introduce other, and stronger, axioms like these and explore some of their consequences. Our basic goal is to prove the Urysohn metrization theorem. It says that if a topological space X satisfies a certain countability axiom (the second) and a certain separation axiom (the regularity axiom), then A" can be imbedded in a metric space and is thus metrizable. This classi- 081 theorem may be considered the culmination of Part I of the book. Another imbedding theorem, important to geometers, appears in the hirfwi-On °f thC chapter- Given a sPaCe which is a compact manifold (the gner-d,mensional analogue of a SUrfaCe)) We show that it can be imbedded some finite-dimensional euclidean space. 189
190 Countability and Separation Axioms ^ 4-1 The Countability Axioms Recall the definition we gave in §2-10: Definition. A space X is said to have a countable basis at x if there is a countable collection (B of neighborhoods of x such that each neighborhood of x contains at least one of the elements of (B. A space that has a countable basis at each of its points is said to satisfy the first countability axiom. We have already noted that every metrizable space satisfies this axiom- see §2-10. The most useful fact concerning spaces that satisfy this axiom is the fact that in such a space, convergent sequences are adequate to detect limit points of sets and to check continuity of functions. We have noted this before; now we state it formally as a theorem: Theorem 1.1. Let X be a space satisfying the first countability axiom. (a) The point x belongs to the closure A of the subset AofX if and only if there is a sequence of points of A converging to x. (b) The function f : X — > Y is continuous if and only if for every convergent sequence (xn) in X, converging to x, say, the sequence (f(xn)) converges to f(x). The proof is a direct generalization of the proof given in §2-10 under the hypothesis of metrizability, so it will not be repeated here. Of much greater importance than the first countability axiom is the fol- following: Definition. A topological space Xis said to satisfy the second countability axiom if X has a countable basis for its topology. Obviously the second axiom implies the first: If (B is a countable basis for the topology of X, then the subset of (B consisting of those basis elements containing the point x is a countable basis at x. The second axiom is, in fact, much stronger than the first; it is so strong that not even every metric space satisfies it. Why then is the property interesting? Well, for one thing, many familiar spaces do have this property. For another, it is a crucial tool used in proving the Urysohn metrization theorem, as we shall see. Example 1. The real line R has a countable basis—the collection of all open intervals (a, 6) with rational end points. Likewise, R" has a countable basis—the collection of all products of intervals having rational end points- Even 7f» has a countable basis—the collection of all products J[,ez. V" where Un is an open interval with rational end points for finitely many values of n, and Un = R for all other values of n.
The Countability Axioms 191 1 , m.heuniformtopology,^satisfiesthefirstcountabrf.tyaxiom ***** \X But U does not sa.isfy the second. Consider the uncountable MmetT»»ble)nBu ■ nces of O's and IV If (B is a bas.s for the TZ7T^ for each xeC'an eIement m a( ^ Jf ^ and andI W^ ^ B/Jox p(.x,y) = l.sothat, £ A. We conclude well behaved with respect to the operat.ons 12 A subspace of a first-countable space is first-countable, and of first-countable spaces is first-countable. A subspace bl d tbl dt f Zvroduct of firstcouna p fi " cond countable space is second-countable, and a countable product of icmd-cmmtable spaces is second-countable. Proof. Consider the second countability axiom. If (B is a countable basis X then {B n A\B e <&} is a countable basis for the subspace A of X. If(B i'sa countable basis for the space X,, then the collection of all products Till, where U, £ <B, for finitely many values of / and U, = X, for all other values of i, is a countable basis for X[Xt. The proof for the first countability axiom is similar. □ Two consequences of the second countability axiom that will be useful to us later are given in the following theorem. First, a definition: Definition. A subset A of a space X is said to be dense in X if A = X. Theorem 1.3. Suppose that X has a countable basis. Then: (a) Every open covering ofX contains a countable subcollection covering X. (b) There exists a countable subset ofX which is dense in X. Proof. Let {£„} be a countable basis for X. (a) Leta bean open covering of X. For each positive integer n for which The Mi?- Ch°°Se a" element A- of a containing the basis element Bn. subset fTn G' °f the SCtS A" is countable. since it is indexed with a o. the poSlt1Ve integers. Furthermore, it covers X: Given a point « a hii, T C a" dement A of a containing x. Since /(is open, there ^hesetTH 5 SUCh th3t X e B" C ^ SinCe B» lies in an element of ^1' is a"rn, * ufd f°r the index n' since A" contains Bn, it contains x. (b) From ?We SubcoUectjon of « that covers X k?nnempty b3Sis dement B« choose a P°int *-• The set X e X> every basis element about x 1 he two ftive«o«ntab,°1w USted '" TheOrem 1-3 are sometimes taken as alter- 1 Notable wbcl*^.IOms- A sPac« for which every open covering contains venng ls usuaUy caUed a Lindelbf space. A space having
192 Countability and Separation Axioms Chap., a countable dense subset is often said to be separable (an unfortunate of terminology).t Weaker in general than the second countability axiom of these properties is equivalent to the second countability axiom wh '^ space is metrizable (see Exercise 7). They are less important than the < countability axiom, but you should be aware of their existence, for th sometimes useful. It is often easier, for instance, to show that a space Xh* a countable dense subset than it is to show that X has a countable basi ?r the space is metrizable (as it usually is in analysis), solving the easier proble solves the harder problem as well. em We shall not use these properties to prove any theorems, but one of the —the Lindelof condition—will be useful in dealing with some exampl Neither of them is as well behaved as one might wish under the operations of taking subspaces and cartesian products, as we shall see in the examples that follow. (See also Exercise 10.) Example 3. The space R, satisfies all the countability axioms but the second Given x e R,, the set of all basis elements of the form [a-, x + \jn) is a countable basis at x. And it is easy to see that the rational numbers are dense in Rt. To see that Rt has no countable basis, let (B be a basis for R,. Choose for each a-, an element Bx of (B such that a- e Bx cz [x, x + 1). If x =£ y, then Bx =£ By, since .v = gib Bx and y = gib By. Therefore, (B must be uncountable. To show that Rt is Lindelof requires more work. It will suffice to show that every open covering of Rt by basis elements contains a countable subcollection covering R,. (You can check this.) So let a = {[a., b.)Uj be a covering of R by basis elements for the lower limit topology. We wish to find a countable subcollection that covers R. Let C be the set considered as a subspace of R. Then C satisfies the second countability axiom. Because the collection {(o«, b*)} consists of sets open in C, it must contain a countable subcollection that covers C, consisting, say, of the elements (a,, b,) for a = a,, a2, Then the collection also covers C. We assert that the set R — C is countable. From this our result follows: Choose for each point of R - C an element of a containing it; adjoining these elements to a', one obtains a countable subcollection of a that covers all of A. So let x be a point of R — C. Necessarily x = a. for some <X e J. Choose qx to be a rational number belonging to the interval {a., ba); because this inter- tThis is a good example of how a word can be overused. We have already define what we mean by a separation of a space; and we shall discuss the separation axioms shortly. .
The Countability Axioms 193 <*■ „ ■ r so is the interval (.v.*). It follows that the function . ,s contain^ in C, * c int0 the set Q of rational numbers, so that , - * iS a" TS [For if * and y are two points of R - C and a- <y, then * „ r is countable.^ otherwlSe > would belong to (.v.*,). although y is in ■^KvlTis stained inc.] ' ' ^ 7Sf wo<*/f / of two Lindelof spaces need not be Lindelof For Example 4. " J and we s]ia|] show that the space ^2 is noti ' 'S 'is an extremely useful example in topology called the Sor- e. plane. It'has as basis all sets of the form [a, b) x [r, d) in the plane, raider the subspace C L={xx(.-x)\xe R,} o' it is easy to check that L is closed in Rf. Let us cover Rf by the open set # 11 and by all basis elements of the form [a, b) x [-a, d). Each of these basis elements intersects L in at most one point. Since L is uncountable, no countable subcollection covers Rf. See Figure 1. aX(-a) Figure 1 [a,fa) X [-a,cQ * " countable dlTZ °f "space havi"S « countable dense subset need not 1*! coorditdT/n nV*** '° ** that the Set of Points havin6 Jkt \ h BlU thesubsP^eL has the discrete topology] it cannot have a countable dense subset.
194 Countabiiity and Separation Axioms q^ Example 6. The space Sa satisfies the first countabiiity axiom, but has mn. of the other countabiiity properties. Given a in Sa, let b be the immediate successor of a. The collection of all intervals whose end points are less than or equal to b is countable, so £Q has a countable basis at a. Clearly, Sa has no countable dense subset, since any countable subset of Sa has an upper bound in Sa. Similarly, Sa is not Lindelof: Consider thecollec- tion of open sets a ={[ao,b)\b e Sa] covering So, where a0 is the smallest element of Sa. If &' is a countable sub- collection of a, the right-hand end points of the intervals in a' have an upper bound b in 5Q; then b is contained in no element of a'. Therefore, a' does not cover Sa, so Sa is not Lindelof. As a consequence, Sa cannot have a countable iis. Note that the space Sa is Lindelof (being compact), and the subspace SQ is basis not. Exercises 1. (a) A Gs set in a space A'is a set A that equals a countable intersection of open sets of X. Show that in a first-countable Hausdorff space, every one- point set is a Gs set. (b) There is a familiar space in which every one-point set is a Gs set, which nevertheless does not satisfy the first countabiiity axiom. What is it? 2. Show that every compact metrizable space X has a countable basis. [Hint: Let an be a finite covering of X by 1/n-balls.] 3. Show that if X has a countable basis, every collection of disjoint open sets in X is countable. 4. Let JThave a countable basis; let A be an uncountable subset of X. (a) Show that the subspace topology on A is not the discrete topology. (b) Show that A contains at least one of its limit points. (c) Show that A contains uncountably many of its limit points. 5. Show that if X has a countable basis {Bn}, then every basis 6 for X contains a countable basis for X. [Hint: For every pair of indices n, m for which it is possible, choose C,m e C such that Bn <= Cn m <= Bm. Let C be the collection of the sets C,,m.] 6. Show that in a Hausdorff space with a countable basis, all four varieties of compactness (compactness, limit point compactness, sequential compactness, countable compactness) are equivalent. 7. Let X be metrizable. (a) Show that if X has a countable dense subset, X has a countable basis. . - (b) Show that if X is Lindelof, X has a countable basis. [Hint: CmP Exercise 2.]
The Separation Axioms 195 i*-2 nd / x / in the dictionary order topology are not metrizable. t Sho* that *J^ ^^ are satisfied by XT in the uniform topology? ,_ vi*i<* cow* u ^ countabIe product of spaces having countable dense sub- ,„. Sho* that >fa;JShas a C0Untable dense subset. ** , if X is Lindelof and Y is compact, then XxYh Lindelof. „. Show that^se( e(/> ^ of aII continuous functions/: I-^ R, where / = [0, 1], tl in the metric ^ ^ = ^ {| /w ^ ]} ., MCe has a countable dense subset, and therefore has a countable Show that lnCo^sjder (he set ^ of aii continuous functions whose graphs are **& !f'of finitely many line segments having rational end points.] "Tow that the product space /', where / = [0, 1], has a countable dense subset. ' Qh w that the space t\ in the ^-metric, has a countable dense subset and il therefore has a countable basis. (See Exercise 9 of §2-9.) 4-2 The Separation Axioms In this section, we introduce three separation axioms and explore some of their properties. One you have already seen—the Hausdorff axiom. The others are similar but stronger. As always when we introduce new concepts, we shall examine the relationship between these axioms and the concepts introduced earlier in the book. Recall that a space A" is said to be Hausdorff if for each pair x, y of distinct points of X, there exist disjoint open sets containing x and y, respectively. Definition. Suppose that one-point sets are closed in X. Then X is said to be regular if for each pair consisting of a point x and a closed set B dis- jwnt from x, there exist disjoint open sets containing x and B, respectively. of jrth*Xh- sa'd '° be normal if for each pair A> B of disi°int closed sets . Here exist disjoint open sets containing A and B> respectively. regular (W * "^ * K&XhT SpaCe is Hausdorff. and that a normal space is Wofthed fi"^ '° mclude the condition that one-point sets be closed as ^•Atwo-poi " °f regularity and normality in order for this to be the ftedtfinitio'nsof ^T ■" 'hC indiscrete topology satisfies the other part of Fot wamples hreg. nty and normality, even though it is not Hausdorff.) '"""■".andnorm^1"8 "* reSularity axiom stronger than the Hausdorff lhe« axiomT y Str°nger than ^^ari'y. see Examples 1 to 3. U?re Called i lhe« axiomT ger than ^^ari'y. see Examples 1 to 3. V0lveUseparatw?re Called seParation axioms for the reason that they in- ^'^ have usL ih kmdS °f SetS from one another by disjoint open me word "separation" before, of course, when we studied
196 Conntability and Separation Axioms Cha, 'P. 4 Hausdorff Regular Figure 2 connected spaces. But in that case, we were trying to find disjoint open sets whose union was the entire space. The present situation is quite different, be- because the open sets need not satisfy this condition. The three separation axioms are illustrated in Figure 2. There are other ways to formulate the separation axioms. One formulation that is sometimes useful is given in the following lemma: Lemma 2.1. Let X be a topological space. Let one-point sets in X be closed. (a) X is regular if and only if given a point xofX and a neighborhood U off., there is a neighborhood V of x such that Vcl). (b) X is normal if and only if given a closed set A and an open set U con- containing A, there is an open set V containing A such that VcU. Proof, (a) Suppose that X is regular, and suppose that the point x and the neighborhood U of x are given. Let B = X — U; then B is a closed set. By hypothesis, there exist disjoint open sets V and W containing x and B, respectively. The set V is disjoint from B, since if y e B, the set W is a neigh- neighborhood of y disjoint from V. Therefore, V cz U, as desired. To prove the converse, suppose the point x and the closed set B not con- containing x are given_. Let U = X — B. By hypothesis, there is a neighborhood V of x such that VcU. The open sets V and X — V are disjoint open sets containing x and B, respectively. Thus X is regular. (b) This proof uses exactly the same argument; one just replaces the point x by the set A throughout. □ Now we seek to relate the separation axioms with the concepts previously introduced. First we study the various kinds of topologies that were intro- introduced in Chapter 2 and see what separation axioms they satisfy. And we prove two theorems which show that in the presence of compactness, or o
The Separation Axioms 197 weak separation properties become strong. This gives us of theorems to remember, but luckily most of the proofs oft 17 fa) A subspace of a Hausdorffspace is Hausdorff; a product theorem i-*- w ' ' ffofa "regular space ts regular; a product of regular spaces ^ subspace of a normal space need not be normal; a product of normal spaces need not be normal(!) , /a^ Let ;^be Hausdorff. Let .v and / be two points of the subspace fX if U and Kare disjoint neighborhoods in Jf of * and y, respectively, ° y'n yan<j Vn Yare disjoint neighborhoods of .v and y in Y. Let {X.} be a family of Hausdorff spaces. Let x = [xj and y = (;'«) be distinct points of the product space \\X«. Since x ^ y, there is some index n such that Xfj^Xf Choose disjoint open sets U and V in Xp containing v and;> respectively. Then the sets n'f\U) and nj, \V) are disjoint open sets in ][XI containing x and y, respectively. (b) Let 7 be a subspace of the regular space X. Since Y is Hausdorff, one-point sets are closed in Y. Let .v be a point of Y and let B be a closed subset of 7disjoint from x. Now B n Y =- B, where B denotes the closure of Bin X. Therefore, .v ^ B, so, using regularity of X, we can choose dis- disjoint open sets t/and V of Xcontaining .v and B, respectively. Then U n Y and Vn y are disjoint open sets in Y containing x and B, respectively. Let {X.} be a family of regular spaces; let X ~ l\X,. By (a), X is Haus- Hausdorff, so that one-point sets are closed in X. We use the preceding lemma to prove regularity of X. Let x = (.Y.) be a point of X and let U be a neigh- neighborhood of x in X. Choose a basis element \[U. about x contained in V. Ujoose, for each a, a neighborhood Ka of *. in *. such that V a U ■ if jt happens that f/a = X., choose Ka = X.. Then K = I[ Ka is a'neighbor- ^KZZ:assert that ^ »p-h foiiows ^ 8 theT}!!n}ie show that if "■ - z»for . n/ L,t ;1°V; „ l^5 an exercise in §2"8-} Supp°se that y = l §} ppe ha TtS.^ t \ ^ ^^ C°ntaininS ^ Since t /"so we can choose a P°int z-^ ' £rtt thi Thus y is in >< h at for a^ Siven inde* A ? neighborhood of y»- ^en W?'(^) is a neigh-
198 Countability and Separation Axioms O»p.A mal is rather difficult. So is the problem of finding a product of nor spaces that is not normal. There happens to be a single space that will Suffi ' for both purposes; it is given in Example 2. Q The following three theorems give three very important sets of hypotheses under which normality of a space is assured. Theorem 2.3. Every metrizable space is normal. Proof. Let X be a metrizable space with metric d. Let A and B be disjoint closed subsets of X. For each a G A, choose ea so that the ball B{a, ej does not intersect B. Similarly, for each b in B, choose eb so that the ball B{b, ffc) does not intersect A. Define U = [JaeA B(a, eJ2) and V = U<.-:« A*, ft/2). Then U and K are open sets containing A and B, respectively; we assert they are disjoint. For if r G U n V, then z e B(a, eJ2) n B(bt eb!2) for some a e A and some b ^ B. The triangle inequality applies to show that d{a, b) < (ea + f6)/2. If f, < f6, then d(a, b) < eb, so that the ball B{b, eb) contains the point a. If eb < ea, then d(a, b) < e», so that the ball B{a, eJ contains the point b. Neither situation is possible. Q Theorem 2.4. Every compact Hcuisdorff space is normal. Proof. Let A" be a compact HausdorlT space. We have already essentially proved that X is regular. For if ,v is a point of X and B is a closed set in X not containing x, then B is compact, so that Lemma 5.4 of Chapter 3 applies to show there exist disjoint open sets about .v and B, respectively. Essentially the same argument as given in that lemma can be used to show that X is normal: Given disjoint closed sets A and B in X, choose, for each point a of A, disjoint open sets Ua and Va containing a and B, respectively. (Here we use regularity of X.) The collection {£/„} covers A; because A is compact, A may be covered by finitely many sets £/„„ . . . , £/„„. Then U =Uai\J ... \j Uan and V =Vatn ■■■ C\ Va. are disjoint open sets containing A and B, respectively. □ Theorem 2.5. Every regular space with a countable basis is normal. Proof. Let X be a regular space with a countable basis <B. Let A and B be disjoint closed subsets of X. Each point x of A has a neighborhood V not intersecting B. Using regularity, choose a neighborhood V of * whose closure lies ,n U; finally, choose an element of <B containing x and contained m K. ay choosing such a basis element for each * in A, we construct a count- countable covering of A by open sets whose closures do not intersect B. Since
The Separation Axioms 199 §4-2 ■ oof A is countable, we can index it with the positive integers; this covering smflariy ^h countabIe coIIection fK») of °Pen sets coverine B> ^h that each set K, is disjoint from A. The sets U = \JU* and K= UK» are °Pen sets containing ,4 and B, nectively, but they need not be disjoint. We perform the following simple re kto construct two open sets that are disjoint. Given n, define u: = U.-\j!.iV, and K = V.-\J;^Ut. See Figure 3. Note that eachset U'n is open, being the difference of an open set U, and a closed set (J"-i K- Similarly, each set V'n is open. The collection \U',} covers A, because each x in A belongs to Un for some /;, and x belongs to "none of the sets Vt. Similarly, the collection {¥'„} covers B. Finally, the open sets U' = \J»ez. U'm and V = U,z. K Figures
200 Countability and Separation Axioms are disjoint. For if x e U' n V, then x e V, n £ for some, and fc. S nose thaty < *. « follows from the defimt.on of V, that x e l/,; and ,;,£ j<^ k it follows from the definition of Vk that x $ V,. A similar contradic- contradiction arises if y ;> £• D We shall use the following theorem in dealing with some examples: Theorem 2.6. Every well-ordered set X is normal in the order topology It is, in fact, true that every order topology is normal (see Example 39 of FS-SJ); but we shall not have occasion to use this stronger result. Proof. Let X be a well-ordered set. We assert that every interval of the form {x,y\ is open in X: If X has a largest element and y is that element, (x,y] is just a basis element about ;•. If y is not the largest element of X, then (.v,;'] equals the open set (.y, /), where >■' is the immediate successor of y. Now let A and B be disjoint closed sets in X; assume for the moment that neither ,4 nor B contains the smallest element a0 of X. For each a e A, there exists a basis element about a disjoint from B; it contains some interval of the form (.v, a]. (Here is where we use the fact that a is not the smallest element of X.) Choose, for each a e A, such an interval (.va, a] disjoint from B, Similarly, for each b e B, choose an interval (yb, b] disjoint from A. The sets U = \J.eAx.. a] and V = \Jb(=B (yb, b] are open sets containing A and B, respectively; we assert they are disjoint. For suppose that z e U n V. Then z e {xa, a] n (yb, b] for some a e A and some b e B. Assume that a < b. Then if a < yb, the two intervals are disjoint, while if a > yb, we have a e {yb, b], contrary to the fact that (yb, b] is disjoint from A. A similar contradiction occurs if b < a. Finally, assume that A and B are disjoint closed sets in X, and A contains the smallest element a0 of X. The set {«„} is both open and closed in X. By the result of the preceding paragraph, there exist disjoint open sets t/and V containing the closed sets A - {a0} and B, respectively. Then U U {a0} and Kare disjoint open sets containing A and B, respectively. Q Example 1. Here is an example showing the regularity axiom stronger than the Hausdorff axiom. Let AT be the subset of the real line R. Define a topology for R by taking as basis all sets of the following types: A) Every open interval (a, b). B) Every set (a, b) — K. It is easy to check that this is a basis for a topology on R; the intersection of
The Separation Axioms 201 another basis element or empty. The space is points of R have disjoint open intervals bout them- _. , K is cIosed in this topology, and it does not V * U nOit08 ppose «ha« thereexist disjoint open sets t/and Kcontam- °"'ainlhe/re £2'Choose a basis element containing 0 and lying .n U , respect veiy f B) since each basls element of fSekment about l/« contained in K; it must be a bas.s Coose a bass e ^^ ^ ^ that z < ]//( and r > max {^ „„ M> MoLs to both J/and K, so they are not disjoint. See Figure 4. hen z feionB5 Figure 4 Example 2. 77i<? product space Sa x 5n u not normal.'] Consider the well-ordered set Sa, in the order topology, and consider the subset Sa, in the subspace topology (which is the same as the order topology). Both spaces are normal, by Theorem 2.6. We shall show that the product space & x Sa is not normal. This example serves three purposes. First, it shows that a regular space need not be normal, for Sn x Sa is a product of regular spaces and therefore regular. Second, it shows that a subspace of a normal space need not be normal, forSn x So is a subspace of Sn x Sa, which is a compact Hausdorff space and therefore normal. Third, it shows that the product of two normal spaces need not be normal. First consider the space Sa x Sa, and its "diagonal" A = [x x x \ x e £„}. Because^is Hausdorff, A is closed in Sa x Sa: If U and Kare disjoint neigh- neighborhoods of * and y, respectively, then [/xKisa neighborhood of x x y that does not intersect A. Therefore, in the subspace Sa x Sn, the set A = A n (,sa x So) = A - [Q. x £2} B closed. Likewise, the set ?and/aredton?w "u'^" °f thl'S Pr°duCt space' See Fi§ure 5- The ets* X 4 con an T6 " 3SSUme there exist disioint °Pen sets ^and Given x % °ntainin«/ «nd B, respectively, and derive a contradiction. «** Point jj ;Tth"S'de; the^«tical slice * x 5O. We assert that there is ^^d all points ^v R ? SUf '^ * X fi "eS outside tt For if U ^ rants x x fi for ^ < p < Q> then ,he (op pojnt ^ x Q Qf {he attnbutes ,his example ,o J. Dieudonni and A. P. Morse independently.
202 Countability and Separation Axioms OX|) Figure 5 slice would be a limit point of U, which it is not because Kis an open set disjoint from U containing this top point. Choose y?(.v) to be such a point: just to be definite, let P(x) be the smallest element of Sn such that x < P(x) < Q. and x x /?(.v) lies outside U. Define a sequence of points of Sa as follows: Let x, be any point of Sa. Let x2 = /?(*,), and in general, xn+l = (}(xn). We have A'l < A < . . . , because /?(*) > x for all x. The set {*„} is countable and therefore has an upper bound in Sa; let b e Sn be its least upper bound. Because the sequence is increasing, it must converge to its least upper bound; thus xn —<■ b. But P(xn) = avm, so that /?(*„) —> A also. Then xn x /?(*„) —> b x b in the product space. See Figure 6. Now we have a contradiction, for the point b X b lies in the set A, which is contained in the open set U; and U contains none of the points xn x ft(xn). Example 3. The space R, is normal, but the space Rf is not. This example serves two purposes. It shows that a regular space Rl need not be normal, and it shows that the product of two normal spaces need not be normal. It is easy to see that R, is normal. Suppose that A and B are disjoint closed sets in R,. For each point a of A choose a basis element [a, x.) not intersecting
The Separation Axioms 203 x, X /5(x,r b x b Figure 6 B; and for each point b of B choose a basis element [b, xb) not intersecting /(. The open sets V = U«^ [". -v«) and K = U*. « [*. ■»'*) are disjoint open sets about A and 5, respectively. It is much harder to show that Rf is not normal. Let L be (he line L = {x x (-.v)|-v e /?,}, as usual. Then L is closed in Rf, and i has the discrete topology as a subspace of S}. Therefore every subset A of L, being closed in L, is closed in Rf. If /?;2 is normal, this means that for every nonempty proper subset A of L, one can choose disjoint open sets of Rf containing the closed sets A and L — A of Rf, respectively. Using a simple "cardinality argument," one can prove that this is not possi- b|e; the proof is outlined in Exercise 14 below. One specific subset of L for which it proves impossible to pick such open s is the subset A consisting of points having irrational coordinates. The Proof that this particular choice will work requires a strong tool: the special that thC BairC cate8orv 'heorem we gave as an exercise in §3-6. Assuming s^lse- we Prove that this particular set will work. s«ppose th k *e SUbSCt °f L consistin8 of Points having irrational coordinates. ^PecihJI V and V are disjoint °Pen sets in Ri containing A and L ~ A, «>at the hi'- , each point x x (-x^ of A>there is a Positive integer n such K oasis element X [-a:, -a:
204 Countabitity and Separation Axioms U. Let Km denote the set consisting of those irr which this basis element lies in a. Then U*. -S^SScSSr^ in V I. .hen since U*n con,ains •, data's aM irrational numbers in [0, 1]. Therefore, we can write [0,1] The union of the closed sets X, (for n e Z+) and the one-point s T a a. ona" number in [0, 1]). By Exercise 5 of §3-6, at leas, one mt interior in [0 1] It cannot be one of t r a a. on , 1]). By Exercise 5 of §36, at leas, one of ,£ Z must have nonempty interior in [0 1]. It cannot be one of the one-poim lets {</}. Therefore, there is an n such that the set Kn contains an open interval ("' By^efi'nition, for each c in K. the basis element [c, c + [-c, -c + [, for /?/ lies in £/. We assert that the union of these basis elements, fore e Kn contains the entire "parallelogram region" P = {.v x y\a < x < b and -.v < y < -x + l/«}, so that /" lies in U. See Figure 7. To prove this fact, let ,v0 x y0 be a point of/1. Using the fact that Kn n (a, b) Figure 7
The Separation Axioms 205 § , m «* can choose a point c of K,C\ (a, 6) such that c X (-c) is dense *(«.**« ,. ,0 and to the left of the line x = x0. It is then simple to IieSber J X^ to in «he basis element [c,c+Mn) x [-e -c + !/«)• ^ ?o££ to be a point of *. in the interval (max {a, -yo], xB). Then ^ lities ; have i c < x0 and — c < yo < —*o + i/n. imply that c < x0 < c + 1/n and -c < y0 < -c + U"-] w-../ Take any rational number q of the interval (a, b). It is clear 1!teIDoinV7>< (-</) is a Iimit point of the ParaIIeI°gram re8ion P' in ^topology of Rf- Then because P a U, this point is a limit point of U. This ntradicts the fact that Kis an open set in Rf containing? x (-</) and disjoint from V. Example 4. If J is uncountable, the product space RJ is not normal. The proof is fairly difficult; we leave it as a challenging exercise (see Exercise 15). This example serves three purposes. It shows that a regular space R1 need not be normal. It shows that a subspace of a normal space need not be normal, for R' is homeomorphic to the subspace @, iy of [0, 1]J, which (assuming the Tychonoff theorem) is compact Hausdorff and therefore normal. And it shows that an uncountable product of normal spaces need not be normal. It leaves unsettled the question as to whether a finite or a countable product of normal spaces might be normal. For that we need one of the two preceding examples. Exercises 1. Show that if jt is regular, every pair of points of X have neighborhoods whose closures are disjoint. pair of disjoim ciosed sets have -^borhoods 3- Show that a closed subspace of a normal space is normal. <• Show that every order topology is regular. ' *^/. iI^J^Olff' °r regUlar' or nomaI>then s° * X,. (Assume ^T^tTf^r^r' d tW° topO'°gies 3 and 3'- respectively; tha, ijir S" " "^^ ( c z riocaiiy compact pce . U, Af„' ^ rC8Ular Li«delbf space is normal. ^TStT that Yis Hausdo^ Show that fa ^ Pr°duct toP°lo8y? In the uniform topology?
206 Countability end Separation Axioms Chap. 4 It is not known whether R»is normal in the box topology. Mary-Ellen Rudin has shown that the answer is affirmative if one assumes the continuum hypo- hypothesis [R]. In fact, she shows it satisfies a stronger condition called paracom- pactness. 11. A space X is said to be completely normal if every subspace of X is normal. Show that X is completely normal ifand only if for every pair A, B of separated sets in X (that is, sets such that A n B = 0 and A n B = 0), there exist disjoint open sets containing them. [Hint: If X is completely normal, consider X-(An B).] 12. Which of the following spaces are completely normal ? Justify your answers. (a) A subspace of a completely normal space. (b) The product of two completely normal spaces. (c) A well-ordered set in the order topology. (d) A metrizable space. (e) A compact Hausdorff space. (f) A regular space with a countable basis. (g) The space R,. 13. Let p:X-> Y be a quotient map. Show that if p is closed and X is normal, then Y is normal. (Classically, a partition X* is said to be upper semicontinuous if the quotient map p: X —» X* is closed.) *14. Show that Rf is not normal, as follows: Suppose that for each nonempty proper subset A of L, there exist disjoint open sets UA and VA of Rf containing A and L — A, respectively. Let D be the set of points of Rf having rational coordinates. Let 9(L) and 6>{D) denote the collections of all subsets of L and D, respectively. Define 9 : ff'(L) — 9(D) as follows: 6{A) = UAn D ifA^0 and A ^ L, 0@) - 0, B(L) = D. (a) Show that 6 is injective. (b) Construct an injective map <9(D) —> L. (c) Derive a contradiction. *15. Theorem. If J is uncountable, then R1 is not normal. Proof. (This proof is due to A. H. Stone, as adapted in [S-S].) It will suffice to show that {Z+Y is not normal, since (Z+Y is closed in R'. Let X denote the space (Z+Y; use functional notation for the elements of X. (a) Define i>, cr X to be the set of those functions x : J — Z+ such that for each ; ^ 1 the set x"'@ contains at most one element. Define P2 to be the set of those x such that for each i ^ 2, the set x~'(') contains at most one element. Show that P, and Pz are disjoint, and are closed in X. (b) Assume that U and Pare disjoint open sets containing i>, and Plt respec- respectively. Show that you can choose a sequence of distinct elements of /, and a sequence of integers "o = 0 < nx < rii < • • •,
The Urysohn Lemma 207 '** f r each positive integer U the se, V, defined be.ow is contained in ^that1 t of all x such that positive integ U ,Ms1h set of all x such that ^He*^" for 1 x(a/)"tl for n,-, <;<"'■ ,,,sZ, Show you can choose a finite subsetB of/ such @ No-« ^^nefbe.ow is contained in K Here K, is the set of all that the set yb th0SexsucMha, ^^ hr ^ e B n A, x(«) = 2 for « e 5 - -4. (d) Show that for some t«n^0. A is every regular topological group normal ? 4-5 T/te Urysohn Lemma Now we come to the first deep theorem of the book, a theorem that is commonly called the "Urysohn lemma." It asserts the existence of certain ^-valued continuous functions on a normal space X. It is the crucial tool used in proving a number of important theorems. One of these, the Tietze extension theorem, we prove in this section. Another, the Urysohn metriza- tion theorem, we prove in the next section. And yet another, an imbedding theorem for manifolds, appears in the last section of the chapter. Whydowecall the Urysohn lemma a "deep" theorem? Because its proof involves a really original idea, which the previous proofs did not. Perhaps we can explain what we mean this way: By and large, one would expect tiiatifone went through this book and deleted all the proofs we have given up to now and then handed the book to a bright student who had not studied topology, that student ought to be able to go through the book and work out the proofs by himself. (It would take a good deal of time and effort, Tn81 °ne W°Uid not eXpeCt him t0 handle the trickier examples.) ■ rySOhn lemma is on a different level. It would take considerably l8inaIlt.y tha" most of us P°ssess t0 Prove this lemma unless we pious hints! ^rysohn lemma)- Let X be a normal space; let A and B be , °fX' ^ [3> b] be a Closed interVal in the real line- s a continuous map ~ , f:X—>[a,b] every x in K md f(x> = hf Where the intcrvaI i aSe follows from th^ one. proof is to construct, using normality, a certain famUy
208 Countability and Separation Axioms Chap. 4 U of open sets of X, indexed by the rational numbers. Then one uses these sets to define the continuous function/. f/*o ' Let P be the set of all rational numbers in the interval [0, i].t We shall define, for each p in P, an open set U, of X, in such a way that when- whenever p < ?, we have Thus the sets U. will be simply ordered by inclusion in the same way their subscripts are ordered by the usual ordering in the real line. Because P is countable, we can use induction to define the sets Up (or rather the principle of recursive definition). Arrange the elements of P in an infinite sequence in some way; for convenience, let us suppose that the numbers 1 and 0 are the first two elements of the sequence. Now define the sets Up as follows: First, define Ut = X — B. Second, since A is a closed set contained in the open set Uu we may by normality of X choose an open set Uo such that A c Uo and Uo <= £/,. In general, let Pn denote the set consisting of the first n rational numbers in the sequence. Suppose that Up is defined for all rational numbersp belong- belonging to the set Pn, satisfying the condition (*) P < q => U, <= Ur Let r denote the next rational number in the sequence; we wish to define U,. Consider the set Pn+t — Pn u {/•}. It is a finite subset of the interval [0, I], and, as such, it has a simple ordering derived from the usual order relation < on the real line. In a finite simply ordered set, every element (other than the smallest and the largest) has an immediate predecessor and an immediate successor. (See Theorem 10.1 of Chapter 1.) The number 0 is the smallest element, and 1 is the largest element, of the simply ordered set P,+ l, and r is neither 0 nor 1. So r has an immediate predecessor p in />„+, and an immediate successor q in Pn+l. The sets V„ and Uq are already defined, and U, c U, by the induction hypothesis. Using normality of X, we can nnd an open set U, of X such that V, <= Ur and U, <= V,. ^tt% 1 ThoIds for every pair of elements °f *.♦,.If both ele- SUh other "' W ° dS by the induction hypothesis. If ow of them is r the other „ a polnt , of ^ then dther g ^ ^ ^.^ ^ U,c:Upc Ur, anycountabIe ^se subset of [0,1] will do, providing it contains the points
The Urysohn Lemma 209 7 in which case '' UrczU,c: U,. of elements of />„, relation (*) holds. of elements of /„, re have U, defined for all p e P. To illustrate, let us suppose we started with the standard way of arranging the elements ofP in an infinite sequence: P = {l,0.i,b hhlhl f.-.J After denning Ua and U,, we would define £/i/2 so that Uo <= Ul/Z and Q 1<zUi- Then we would fit in U1/3 between Ua and UU1; and [/2/3 between ^ and Ui. And so on. At the eighth step of the proof we would have the situation pictured in Figure 8. And the ninth step would consist of choosing an open set J/2,j to fit in between U\/3 and £A/2. And so on. Figure 8 all 11 " **» 'me ifp<0, if P > L numbers p in R by defini y0U
210 Countability and Separation Axioms ^ Step 3. Given a point x of X, let us define g(x) to be the set of tho rational numbers p such that the corresponding open sets v, contain * 0(*) = {/>|* e I/,}. This set contains no number less than 0, since no x is in U, Tor p < 0. And it contains every number greater than 1, since every x is in Uf for p > , Therefore, Qix) is bounded below, and its greatest lower bound is a pOjn| ofthe interval [0,1]. Define f(x) = gib Qix) = gib [p | x e I/,}. 5/fp ¥. We show that/is the desired function. If x e A, then * e (/ for every /> > 0, so that Qix) equals the set of all nonnegative rationals' and f(x) = gib Qix) = 0. Similarly, if x e 5, then x e Ur for no p ^ 1' so that g(x) consists of all rational numbers greater than 1, and f(x) = 1' All this is easy. The only hard part is to show that/is continuous. For this purpose, we first prove the following elementary facts: A) x e V, =* f{x) < r. B) xiUr=> f{x) > r. To prove A), note that if x e 0r, then iei/, for every s > r. Therefore, Q(x) contains all rational numbers greater than 1; so that by definition we have f(x) = gib Qix) < r. To prove B), note that if x ^ {/,, then x is not in £/, for any s < r. Therefore, Qix) contains no rational numbers less than r, so that /(*) = gib Qix) > r. Now we prove continuity of/ Given a point x0 of Xand an open interval (c, rf) in R containing the point fix0), we wish to find a neighborhood U of x0 such that /(£/) c (c, rf). Choose rational numbers p and q such that c < P < /Oo) < ? < d. We assert that the open set v=vq-u, is the desired neighborhood of xB. See Figure 9. /(x0)
«** The Urysohn Lemma 211 ut x eU Necessarily x0 e U,, because xB $ U, wlsh°T ^\ >V Also, Xo t V,, because *„ e V, imp'-es by j^^ (Crf). Let * e V. Then * e tfc= U Second, we sl»w /^ ^ ^ ^ ^ ^ ^ ^ ^ and /(jf) ^ ^ by B). so*»«/5%T ' j c (r, rf), as desired. D Thtts/(-v t [/"'J • if i and B are two subsets of the topological space X, and if ^t" luous funct.cn /: *- [0. 1] such that f(A) = {0j and *"* ''rn 1 «v that A and 5 can be separated by a continuous function. f(B)~W> * The Hrvsohn lemma says that if every pair of disjoint closed sets in X ^narated by disjoint open sets, then each such pair can be separated Ta Sous function. The converse is trivial, for if/: X -> [0, 1] is the function, then /"([<),]-)) ™d /-(& 0) are disjoint open sets conta.n.ng ^ and 5, respectively. This fact leads to a question that may already have occurred to you: Why cannot the proof of the Urysohn lemma be generalized to show that in a regular space, where you can separate points from closed sets by disjoint open sets, you can also separate points from closed sets by continuous func- functions? At first glance, it seems that the proof of the Urysohn lemma should go through. You take a point a and a closed set B not containing a, and you begin the proof just as before by defining U, = X — B and choosing Uo to be an open set about a whose closure is contained in £/, (using regularity of X). But at the very next step of the proof, you run into difficulty. Suppose that/i is the next rational number in the sequence after 0 and 1. You want to find an open set Up such that UB <= Up and fjp <=[/,. For this regularity is not enough. Requiring that one be able to separate a point from a closed set by a Minuous function is, in fact, a stronger condition than requiring that one new Jarate them by disJ°int °Pen sets. We make this requirement into a *»2? °naX'Om' WhlCh WCSha11 Study further in ChaPter 5: A sPace for every ° C°mpletely ^^^ 'f °ne-point sets are closed in X, and if exisu a co^mt a °[X and every closed set B not containing a, there a{1] a cont,nuous function /: X-. [0, 1] such that f(a) = 0 and f(B) S^ ther Comment on the Urysohn lemma: Note that the 1101 assert the existence of a continuous function/ this condition is satisfied; see Exercise 5. corollary of the Urysohn lemma is the useful theorem xiauwn theorem. It states that any continuous function
Countability and Separation Axioms Chap 4 /mapping a closed subset of a normal space into the reals may be extended to a continuous map of the entire space into the reals. Although we shall not use the Tietze theorem in this book (except in the exercises), it is very important in the applications of topology. Theorem 3.2 (Tietze extension theorem). Let X be a normal space; let A be a closed subset o/X. (a) Any continuous map of A into the closed interval [a, b] of R may be extended to a continuous map of all ofX into [a, b]. (b) Any continuous map of A into the reals R may be extended to a con- continuous map of all ofX into R. Proof. The idea of the proof is to construct a sequence of continuous functions sn defined on the entire space X, such that the sequence sn Converges uniformly, and such that the restriction of sn to A approximates/more and more closely as n becomes large. Then the limit function will be continuous, and its restriction to A will equal/. Step 1. The first step is to construct a particular function g defined on all of X such that g is not too large, and such that g approximates/on the set A to a fair degree of accuracy. To be more precise, let us take the case/: A —> [—r, >•]. We assert that there exists a continuous function g : X—> R such that \g(x)\<%r for all x e X, \g(a)- f(a)\<\r for all fl e A. The function g is constructed as follows: Divide the interval [—r, r] into three equal intervals of length %r; I, = [~r, -\r], I, = [-\r, \r], /3 = [\r, r). Let B and C be the subsets B =/-(/,) and C = /-•(/,) of A. Because/is continuous, B and C are closed disjoint subsets of A. There- Therefore, they are closed in X. By the Urysohn lemma, there exists a continuous function g:X—>[-$r,}r] having the property that g(x) = -\r for each x in B, and g(x) = jr for each x in C. See Figure 10. Then \g(x)\ ^ \r for all x. We assert that for each a in A, There are three cases. If a e B, then both f(a) and g(a) belong to/,. If a e C, then f(d) and g(d) are in /,. And \f a $ B \J C, then f(a) and g(a) are in h- In each case, |g(tf) -/(a)| f
The Urysohn Lemma 213 1*3 Figure 10 fS IT theorem in the f[ A into an ^itrary <*«d i ^f f-A - [-»• »]• Here , = 1. We gt defined on all of X such that We -*.(«) I for ^ e step l-we obtain a I } ~ for x eX, ; Q). for fl e ^. • -" -*»•And so -• functions ?I(.... ^ defined on
214 Countability and Separation Axioms Chap. 4 all of X such that for a e A. Applying Step 1 to the function/- g, g., with r = (|)»( we obtain a real-valued function gntt defined on all of A'such that tor x eX, \A°) - •?.(") *■♦ 'WI ^ ®"+l for fl e ^ By induction, the functions gn are defined for all n. We now define «C*) = S.--.«^*) for all -v in X Of course, we have to know that this infinite series converges. But that follows from the comparison theorem of calculus; it converges by comparison with the geometric series Since this geometric series converges to 1 (as you can check), it follows that \g(x)\< 1, so that g maps X into [— I, 1], as desired. We assert that g{a) = /(a) for a e A. Let sn(x) = 2"-i &(*). the nth partial sum of the series. Then g(x) is by definition the limit of the infinite sequence s,(x) of partial sums. Since \A<>) - L?-i g/L«)\ = l/(fl) - ^Wl < A)" for all a in A, it follows that jn(a) —> /(a) for all a e /4. Therefore, we have /(a) = ^(o) for a e A. To show that g is continuous, we have to show that the sequence s» con- converges to g uniformly. This fact follows at once from the "Weierstrass M- test" of analysis. Without assuming this result, one can simply note that if k > n, then Holding n fixed and letting k —> 00, we see that for all igl Hence sn converges to g uniformly. Step 3. We now prove part (b) of the theorem, in which/maps A into R. We can replace R by the open interval (—1, 1), since this interval is homeomorphic to jR. So let /be a continuous map from A into (-1, 1). The half of the Tietze theorem already proved shows that we can extend/to a continuous map g : X—[—l, 1] mapping X into the closed interval. How can we find a map h carrying Zinto the open interval?
The Urysohn Lemma 215 us define a subset D of X by the equation ra]) closed subset of X. Because g{A) = /(/<), which '^Hftiset " is disjoint fr°m * By thC ^Thn oTaTd' tainedcon inuous function # :*-*>,!] such that *!>> = {0} and ** iS;n Define ^) = 1J' *(*) = #*)*(*>• discontinuous, being the product of two continuous functions. Also, ^ex'nsionof/.sincefor.in,, () #&() I «(«) = /(«) = I r »Hv /. maps all of X into the open interval (- I, 1). For if x e D, then Sjio gW =0. And if .v $ D, then |g(.v)| < 1; it follows at once that |Atv)|<l-U(.v)l<l- □ Example 1. The hypothesis that the sets involved are dosed is needed in the Urysohn lemma and the Tietze theorem. For example, the sets ^ = @,1) and B = A,2) are disjoin' subsets of the normal space R, but there is no continuous function /: R —> [0,1] that carries A into 0 and B into 1. Where could/carry the point lofinB? / lofinB? Exercises 1. Examine the proof of the Urysohn lemma, and show that for given r, P, q rational. 1 Show that 'he Tietze extension theorem implies the Urysohn lemma. • W Show that a connected normal space having more than one point is uncount- uncountable. Show that a connected regular space having more than one point is uncount- aole.T [Hint: Any countable space is Lindelof.] OOSho* \re8Ular Spa°e W'th a countable basis; let u ^ °Pen >n X. ft) Show h V eqUa'S a countable union °f dosed sets of X. for* i „ lS 3 continuous function /: Z-> [0, 1] such that f(x) > 0 5, ^ * e Uand /W = 0 for x $ V. of °Pen sets of &x°' %<A" '" XlfAis the 'ntersection of a countable collection "- e°rem(S'ro»gform of the Urysohn lemma). Let A and B be closed disjoint Exampi. W rf^j0"* Cxist a connected Hausdorff space that is countably
216 Countability and Separation Axioms ChaM subsets of the normal space X. There exists a continuous fu r such that /-'({O}) = A andf{B) = {1} if and only ifA is ac'Zr^1^ 6. Let X be metrizable. Show that the following are equivalent- (i) X is bounded under every metric that gives the topology of v (ii) Every continuous function 0 : X—* R is bounded, (iii) X is compact. [Hint: If 0 is not bounded, map X into X x R by the map X—> If (*„) is a sequence of distinct points having no convergent subsequ* X ^'' a continuous function 0 with <f)(xn) = n.] nce. find 7. Let Zbe a topological space. If Y is a subspace of Z, we say that yi a ofZ if there is a continuous map /■: ^— ysuch that r(y) = ^ for each vgy (a) Show that if Z is Hausdorff and Y is a retract of Z, then y js do^ in ,' (b) Let /4 be a two-point set in Rz. Show that A is not a retract of Jp (c) Let S1 be the unit circle in R2; show that 5' is a retract of Jjz — [0], where 0 is the origin. Can you conjecture whether or not Sl is a retract of 'jp? 8. A space Y is said to have the universal extension property if for each triple consisting of a normal space X, a closed subset A of X, and a continuous func- function/: A —► Y, there exists an extension of/to a continuous map of Jfinto Y. (a) Show that RJ has the universal extension property. (b) Show that if yis homeomorphic to a retract of RJ, then yhas the universal extension property. 9. Let Xx e= X2 <= • • • be a sequence of spaces, where X, is a closed subspace of Xn.! for each ;. Let X = \JX,. Suppose that we define a subset t/of JTtobe open in X if [/ n ^- is open in Xt for each i. The topology we obtain is calW the topology on X coherent with the spaces X,. (a) Show that X, is a subspace of X in this topology. (b) Show that if each space Xt is normal, so is the space X. [Hint: Use the lie theorem.] *10. The subspace R" of R", consisting of all sequences (,xux2,.. •) such lha xi for i> n, has a standard topology. Suppose that we give the set K 'f the topology coherent with the subspaces R". Compare this topology with the topology it inherits from the box topology on /?"■ 4-4 The Urysohn Metrization Theorem thai S'**5 Now we come to the major goal of this chapter, a thfPfe roOfweaveS us conditions under which a topological space is metrizable.T e.°uSeS fesulB together a number of strands from previous parts of the boo , ^ ^jt- on metric topologies from Chapter 2 as well as facts colfer^r The W5* ability and separation axioms proved in the present chap ^i construction used in the proof is a simple one, but very use • it several times more in this book, in various guises. ^ There are two versions of the proof, and since each na
§ 4-4 The Urysohn Metrization Theorem 217 zations that will appear subsequently, we present both of them here. The first version is the one we shall use in Chapter 5 to prove an imbedding theorem for completely regular spaces. The second version will be generalized in Chapter 6 when we prove the Nagata-Smirnov metrization theorem. Theorem 4.1 {Urysohn metrization theorem). Every regular space X with a countable basis is metrizable. Proof. We shall prove that X is metrizable by imbedding x in a metriz- metrizable space Y; that is, by showing X homeomorphic with a subspace of Y. The two versions of the proof differ in the choice of the metrizable space Y. In the first version, Y is the space R" in the product topology, a space that we have previously proved to be metrizable (Theorem 9.5 of Chapter 2). In the second version, the space Y is also R°, but this time in the topology given by the uniform metric p (see §2-9). In each case, it turns out that our proof actually imbeds X in [0, 1]". Let [Bn] be a countable basis for X. Step 1. We prove the following: There exists a countable collection of continuous functions /„ : X—► [0, 1] having the property that given a point x0 of X and given a neighborhood U of x0, there exists an index n such that /, is positive at x0 and vanishes outside U. The construction of the functions/, is simple. We apply Exercise 4 of the preceding section to choose, for each basis element Bn, a continuous function /„ : X —> [0, 1] such that f{x) > 0 for x e Bn and f(x) = 0 for x <£ Bn. The collection {/„} satisfies our requirement. Given x0 and given a neighbor- neighborhood U of x0, we can choose a basis element Bn with x0 e Bn cz U; then the function/, will be positive at x0 and vanish outside U. Another way to construct the collection {/„}, which does not rely on the exercise cited, is the following: For each pair n, m of indices for which Bn a Bm, apply the Urysohn lemma to choose a continuous function gn,m : X—> [0, I] such that ?„,„(£„) = [1} and gn.m(X - Bm) = [0}. Then the collection {>„,„,} satisfies our requirement: Given x e U, one can choose a basis element Bm such that x e Bm a U. Using regularity, one can then choose Bn so that x0 e Bn and §„ <= Bm. Then the function g,_m is defined, and^n,m is positive at -v0 and vanishes outside U. Because this collection is indexed by a subset of Z+ x Z+, it is countable; therefore it can be reindexed with the positive integers, giving us the desired indexed family {/,}. Step 2 (First version of the proof). Given the functions/„ of Step 1, take -R" in the product topology and define a map F : X — R™ by the rule F(x) = (/,(*),/,(*), • • •)• We assert that F is an imbedding. First, Fis continuous because R" has the product topology and each/;
Coumability and Separation Axioms Chap. 4 is continuous. Second, F is injective because given x ^ y, We know there' an index n such that/.fr) > 0 and/Xv) = 0; therefore, F(x) * f^) K Finally, we must prove that F is a homeomorphism of X onto itsima the subspaceZ = F(X) of #». We know that /"defines a continuousbijectfo" of X with Z, so we need only show that for each open set U in Xt the *" /•([/) is open in Z. Let z0 be a point of F(U). We shall find an open set Wof Z such that W zo F(U). Let ^0 be the point of U such that F(x0) = z0. Choose an index ^for which fN(x0) > 0 andfN(X - U) = {0}. Take the open ray @, +oo) in Rt and let K be the open set K = 7^'(@, +«.)) of /?". Let PK = K D Z; then f-K is open in Z, by definition of the subspace topology. See Figure 11. We assert that z0 e W cz F(U). First, z0 e W w = vnz Figure 11 because Second, n^z) e @, F(U). For if Si = Kjv(F(*o))=/v(*o)>O. and for some x e -^ .. vanishes op®* e W> then z = = »-(fW) =AW. an k he" ^ = ^W * in W as desired. F is an imbedding of X in *-.
The Urysohn Metrization Theorem 219 Step 3 (Second version of the proof). In this version, we imbed X in the metric space (K°, p). Actually we imbed x in the subspace [0,1]» on which p equals the metric Kx,y)=lub{|*/-.v,|}. We use the countable collection of functions/„: X—[0, 1] constructed in Step I. But now we impose the additional condition that/0) <; \jn for all x. (This condition is easy to satisfy; we can just divide each function/n by «.) Define F: X—»[0, 1]™ by the equation as before. We assert that F is now an imbedding relative to the metric p on [0, 1]™. We know from Step 2 that F is injective. Furthermore, we know that if we use the product topology on [0, 1]", the map F carries open sets of X onto open sets of Z = F(X). This statement remains true if one passes to the finer (larger) topology on [0, 1]™ induced by the metric p. The one thing left to do is to prove that F is continuous. This does not follow from the fact that each component function is continuous, for we are not using the product topology on R" now. Here is where the assumption /(.v) < 1/n comes in. Let a-0 be a point of X, and let e > 0. To prove continuity, we need to find a neighborhood U of .v0 such that -v e U=>p(F(x),F(xo))<e. First choose N large enough that l/N < f/2. Then for each n = I,..., N, use the continuity of/, to choose a neighborhood £/„ of xa such that l/Xv)-/*(*o)l<e/2 for x e U,. Let U = U, n ■ ■ ■ D UN; we show that U is the desired neigh- neighborhood of .v0. Let x e U. If /; < N, by choice of U. And if n > N, then because/, maps X into [0, 1/n]. Therefore for all x e U, p(F(x),F(xo))<e/2<e, as desired, fj In Step 2 of the preceding proof, we actually proved something stronger than the result stated there. For later use, we state it here:
220 Countability and Separation Axioms V4 Theorem 4.2 {Imbedding theorem). Let X be Hausdorff, s { f«}«£i w « collection of continuous functions f. : X — R satisfP-'">S( lha quirement that for each point x0 o/X and each neighborhood M of ^ ""'*' an index a such that f. is positive at x0 and vanishes outside U TA X°' '*Wf * tionFiX — R1 */««/ &y ' "*tf F(x) = (f.(x)).eI w fln imbedding o/X /« R1. The proof is almost a copy of Step 2 of the preceding proof; one merely replaces n by a and R° by i?J throughout. ' A collection {/„} of continuous functions satisfying the hypotheses of this theorem is said to separate points from closed sets in X. Exercises 1. Give an example showing that a Hausdorff space with a countable basis need not be metrizable. 2. Give an example showing that a space can be completely normal, and satisfy the first countability axiom, the Lindelof condition, and have a countable dense subset, and still not be metrizable. 3. Let X be a compact Hausdorff space. Is it true that if X has a countable basis, then X is metrizable? What about the converse? 4. Let A" be a locally compact Hausdorff space. 1^ (a) Is it true that if X has a countable basis, then X is metrizable? What a the converse ? uo(. (b) Let ybe the one-point compactification of X. Is it true that ifJfhasa able basis, then Y is metrizable? What about the converse? ^ 5. A space X is locally metrizable if each point x of * has a neie^borhd°o0rfr space metrizable in the subspace topology. Show that a compact H^us.(eunionof * is metrizable if it is locally metrizable. [Hint: Show that Jf isatm open subspaces, each of which has a countable basis.] (rizal* 6. Show that a regular Lindelof space is metrizable if it is l0,C^(yisessential; [Hint: A closed subspace of a Lindelof space is Lindelof.] Regula where do you use it in the proof? . 7. Show that if Y is a normal space with basis <B, then Yean be l [0, I]7, where J is some subset of CB x CB. 8. Check the details of the proof of Theorem 4.2. »-tha|se(* 9. Show that if {/.} is a family of real-valued continuous f^^ rates points from closed sets, then the topology of X is t ^^ topology relative to which all the functions/, are contm • ^for 10. Show that if X is comple tely regular, then JTcan be imbedd J.
§4-4 The Urysohn Metrization Theorem 221 11. Let Y be a normal space. Then Y is said to be an absolute retract (AR) if when ever one has an imbedding h : Y— Z of Y into a normal spaced such tha't h(Y) is closed in Z, then h(Y) is a retract of Z. Show that if Y is compact the following statements are equivalent: (i) Y is homeomorphic to a retract of [0, 1]J for some /. ' (ii) y is homeomorphic to a retract of RJ for some /. (iii) Yhas the universal extension property. (iv) Yis an absolute retract. In fact, show that (i) => (ii) => (iii) => (iv) wi(hou( assuming Y compact and (iv) => (i) if Y is compact. [Hint: Assume the Tychonoff theorem so you know that [0, \]> is normal.] 12. (a) Show the logarithmic spiral C = {0 X 0} U {e< cos t x e- sin t\ t e R] is a retract of R2. Can you define a specific retraction r: R1 —» C? (b) Show that the "knotted *-axis" K of Figure 12 is a retract of R3. Figure 12 *13. Prove the following: Theorem. Let Y be a normal space. Then Y is an absolute retract if and only if Y has the universal extension property. [Hint: If *and /are disjoint normal spaces, A is closed in X, and/: A-+Y is a continuous map, define the adjunction space Zf to be the quotient space obtained from XUXby identifying each point a of A with the point J(a) and with all the points of/-({/(a)})- Using the Tietze theorem, show that tt is normal.] *14. A subspace Y of a space Z is called a neighborhood retract of Z if Y is a retract of some open set of Z. A normal space Y is called an absolute n^rjood retract (ANR) if i, is a neighborhood retract of every normal space Z.n M it is imbedded as a closed subspace. We say that Y has the umvereaI ne.ghbor hood extension property if every continuous map /: A -> Y, where /.is a subset of the normal space X, has a continuous extension to an open set m X about A. Generalize Exercises 11 and 13 to neighborhood retracts.
222 Countability and Separation Axioms Chap. 4 *4-5 Partitions of Unity"! We have shown that every regular space with a countable basis imbedded in the "infinite-dimensional" euclidean space R> it ■ a to ask under what conditions a space Jean be imbedded in somefi dimensional euclidean space RN. One answer to this question is given bekT" we show that every compact manifold can be imbedded in R" for some V A more general answer will be obtained in Chapter 7, when we study diW sion theory. The proof uses a certain family of functions {<j>a} on X that form what is called a "partition of unity." Such families have proved to be very useful tools in analysis, geometry, and topology. We shall consider only the case of finite partitions of unity (which is all we need) and leave the general case to the exercises. First, we need some terminology. If <j>: X—► R, then the support of <f> is defined to be the closure of the set <f)~l(R — {0}). Thus if x lies outside the support of0, there is some neigh- neighborhood of .v on which 0 vanishes. Definition. Let {(/,,..., U,} be a finite indexed open covering of the space X. An indexed family of continuous functions <j>,:X—>[0, 1] for / = ),...,//, is said to be a partition of unity dominated by {U,} if.' A) (support 0,) c: U, for each ;. B) Z?=i 0.-W = ' for each x. Theorem 5.1 (Existence of finite partitions of unity). Let {Uu--->^ be a finite open covering oft lie normal space X. Then there exists a par of unity dominated by {U,}. ., Proof Step 1. First, we prove that one can "shrink" the coverm to an open covering {K,,. . . , Vn] of X such that V, cz U, for each i. We proceed by induction. First, note that the set A = X - [Uz U • ■ • U U.) is a closed subset of X. Because [U,, .. . , U.} covers X, the set in the open set [/,. Using normality, choose an open set K, such that V, cz U,. Then the collection [Vlt U2,. ■ ■ , U.) co^e In general, given open sets K,, . .. , Vk-, such that th {Vu...,Vk-t,Uk,Uk+i,...,U») covers X, let tThis section will be used when we study dimension theory in § - ■
§4-5 Partitions of Unity 223 *-*-(r.U...UFi_1)-.Wl+lU...ul0 Then A is a closed subset of X which is contained in the open set U Chn Vk to be an open set containing A such that Vk c [/4. 7^ *- Choose covers X At the nth step of the induction, our result is proved Step 2. Now we prove the theorem. Given the open covering {U, [/„} of X, choose an open covering {F, K] of x such ^ p-'- for each ;. Then choose an open covering [W.,. \y) of Y *Ju >u [ Wt = F, for each /. Using the Urysohn lemma, choosef'or each fa corZnuo* function Vt-X~>[0,l] such that Wi(W,) = {1} and Vl{X- V,) = {0}. Since Vr'(R - {0}) is con- ■ tained in V,, we have (support y/,) c K, c: ^,. Because the collection {W,} covers Z, the sum <P(x) = 2,-., ^.(x) is positive for each x. Therefore, we may define, for each j, It is easy to check that the functions <f>lt. . . , <j>n form the desired partition of unity. Q Definition. An m-manifold is a HausdorflF space X with a countable basis such that each point x of X has a neighborhood that is homeomorphic with an open subset of Rm. A 1-manifold is often called a curve, and a 2-manifold is called a surface. Manifolds form a very important class of spaces; they are much studied in differential geometry and algebraic topology. We shall prove that if Jf is a compact manifold, then Zcan be imbedded in a finite-dimensional euclidean space. The theorem holds without the assumption of compactness, but the proof is a good deal harder. Theorem 5.2. If X is a compact m-manifoid, then X can be imbedded in RN/or some positive integer N. Proof. Cover X by finitely many open sets [Uu .... U.}, each of which may be imbedded in R™. Choose imbeddings *,:#, — *™ f°r each !\BelnS compact and HausdorflF, X is normal. Let 0,,. . . , & be a partition of unity dominated by {£/,}; let A, = support <(>,. For each / = 1 »> define a function h,\ X —* Rm by the rule for* 6%
Countability and Separation Axioms Chap. 4 [Here <f>,(x) is a real number c and g,(x) is a point y — (v the product c-y denotes of course the point (cyu ... ,cy )'of T^'ru ^*' tion h, is well defined because the two definitions of/;, agree on th ■ func- tion of their domains, and ht is continuous because its restricti lnters*" open sets U, and X — A, are continuous. "* Now define i7: X—> (R X ■■ ■ X Rx Rm X ■■■ XRm) n times n times by the rule F(x) = ((*,<»,.... <*„<», *,(*),.... *,(*)). Clearly /" is continuous. To prove that F is an imbedding we need only to show that F is injective (because X is compact). Suppose that F(x) =F(y) Then <f)t(x) = 0,O) and //,(*) = K{y) for all /. Now 0,(x) > 0 for some i [since 2 0/W = !]• Therefore, 0,(^) > 0 also, so that ^,;£ £/,. Then Because 0,(x) = 0,0') > 0, we conclude that g,(x) = g,(y)- But g,::U<—>A" is injective, so that * = ^, as desired. □ In many applications of partitions of unity, such as the one just given, all one needs to know is that the sum ^2 <f>,(x) is positive for each x. In others, however, one needs the stronger condition that 2 4>i(.x) — '■ ^ee ^*'- Exercises 1. Prove that every manifold is regular and hence metrizable. Where do you the Hausdorff condition? 2. Let X be a compact Hausdorff space. Suppose that for each x 6 ^'"^j j, neighborhood U of x and a positive integer k such that £/can be imW /?*. Show that X can be imbedded in RN for some positive integer A. 3. An indexed collection {/<«} of subsets of X is said to be a point-fin*1" *' family if each x e X belongs to A. for only finitely many values ol -^ ^ tf Lemma (The shrinking lemma). Let X be a normal space: let I jJ'J^jig point-finite indexed open covering ofX. Then there exists an indexed op { V.Uj of X such that V. <= U. for each a. „ & it* (a) Prove the lemma in the case J = Z+> by induction. Where ao / fact that [£/„} is point-finite? . vWH-or*re'i *(b) Use transfinite induction to prove the lemma for an arbitrary index set J. 4. Let Xbeaspace. An indexed family{/<.}«6/of subsets of A'issaid finite indexed family if each point of A" has a neighborhood th for only finitely many values of a.
Supplementary Exercises: Review of Part I 225 lions'' {U'Uj ^ ^ indCXed °Pen C0Venne °f * A famUy of con«nuous func- 4>a : X—► [0, l], indexed by a e /, is said to be a partition of unity dominated by \U \ if (i) (support 0.) c Um for each a e J, V l a] " (ii) {support $.}.<=,, is locally finite, (iii) S«ey 0,(x) = 1 for each x. [The sum makes sense because for given x, we have fa(x) ^ 0 for only finitely many values of a.] Assuming the shrinking lemma, prove the following- Theorem. If X is a normal space and (f{Um)met is a locally finite indexed family of open sets covering X, then there exists a partition of unity dominated by [U.l 5. If [Ut]ae, is a collection of subsets of X, a collection {VP}^K is said to refine {[/,} if for each set Vf, there is at least one set U, containing it. Theorem. Let X be a normal space, and let[U.].SJbe an indexed open covering of X. If there is an open covering {Vf]fsK of X which refines {U.} and is locally finite, then there exists a partition of unity dominated by {£/„}. [Hint: Choose/: K— ► / so that Vf cr U/if). Let W, be the union of those Vf for which /(/?) = a; show that [W,],e/ is locally finite.] A Hausdorff space X is said to be paracompact if every open covering of X has a locally finite refinement that is an open covering of X. We shall study paracompact spaces in Chapter 6. 6. The Hausdorff condition is an essential part of the definition of a manifold; it is not implied by the other parts of the definition. Consider the following subset of i?2: X=(RxO)U(R+ x 1). Topologize X by taking as basis all sets of the following types: (i) (a, b) x 0 for a < b, (ii) (a, b) x 1 for 0 < a < b, (iii) ((a, 0) x 0) U ([0, ft) x 1) for a < 0 < ft. (a) Show that this is a basis. _ «■*«•*- (b) Show that each of the subspaces R x 0 and (R- x 0) U (R+ x 1) ot X is homeomorphic to R. . Y (c) Show that X has a countable basis, that finite point sets are closed in A, and that each point of X has a neighborhood homeomorphic with an open set in R. (d) Show that X is not Hausdorff. * Supplementary Exercises: Review of Part I Consider the following properties a space may satisfy: A) connected B) path connected
226 Countability and Separation Axioms C) locally connected ^1 D) locally path connected E) compact F) limit point compact G) locally compact Hausdorff (8) Hausdorff (9) regular A0) normal A1) first-countable A2) second-countable A3) Lindelof A4) has a countable dense subset A5) locally metrizable A6) metrizable 1. For each of the following spaces, determine (if you can) which of these prop- properties it satisfies. (Assume the Tychonoff theorem if you need it) (a) Sa (b) Sa (c) Sa x Sa (d) Sn x [0,1) in the dictionary order (e) / x / in the dictionary order, where / = [0,1] (f) Hi (g) Rf (h) R* in the box, uniform, and product topologies (i) R' in the product topology (j) R in the finite complement topology 2. Which of these properties does a metric space necessarily have? 3. Which of these properties does a compact Hausdorff space have? 4. Which of these properties are preserved when one passes to asubspace. closed subspace? To an open subspace? ^ 5. Which of these properties are preserved under finite products, products? Arbitrary products? 6. Which of these properties are preserved by continuous maps- ^ 7. After studying Chapters 5, 6, and 7, repeat Exercises 1-6 for the o erties: A7) completely regular A8) paracompact A9) topologically complete ^ You should be able to answer all but one of the 336 q.ue^°£i* 7. H** ercises 1-6, and all but one of the 63 questions involved in are unsolved; see the remark in Exercise 10 of § 4-2-
5. The Tychonoff Theorem We now return to a problem we left unresolved in Chapter 3. We shall prove the Tychonoff theorem, to the effect that arbitrary products of compact spaces are compact. The proof makes use of the maximum principle of set theory (see §1-11). The Tychonoff theorem is of great usefulness to analysts (less so to geo- geometers). We apply it in §5-2 to obtain an interesting characterization of completely regular spaces. Another application appears in §5-3, where we construct the Stone-Cech compactificationand explore its properties; in this section we assume §3-8, Local compactness. 5-1 The Tychonoff Theorem Like the Urysohn lemma, the Tychonoff theorem is what we call a "deep" theorem. Its proof involves not one but several original ideas; it is anything but straightforward. We shall discuss the crucial ideas of the proof in some detail before turning to the proof itself. In Chapter 3, we proved the product X X Y of two compact spaces to be compact. For that proof the open covering formulation of compactness was quite satisfactory. Given an open covering of X x Yby basis elements, 229
230 The Tychonoff Theorem Chap. we covered each slice x x Y by finitely many of them, and proceeded fr0 that to construct a finite covering of X x Y. m Try as we may, we cannot make this approach work for an infinite pr0(j uct of compact spaces. So the first idea involved in the proof of the Tychonoff theorem is to abandon open coverings and to approach the problem usin instead the closed set formulation of compactness. Actually what we use the formulation expressed in Corollary 5.10 of Chapter 3 which states that Xh compact if and only if for every collection a of subsets of X satisfying the f.i.c. (finite intersection condition), the intersection PUa^of the closures of the elements of d is nonempty. To see how this approach might work, let us consider first the simplest possible case: the product of two compact spaces X, x X2. Suppose that a is a collection of subsets of X, x X2 satisfying the f.i.c. Consider the projection map vt, : X, x X2 —* X,. The collection {iti(A)\A e a] of subsets of X, also satisfies the f.i.c, since if x e At n • • • n An, then 7t,(jc) e n^Ai) n ■ • ■ D n,(An), so the latter set is nonempty. Compactness of X, guarantees that the intersection of all the sets n^A) is nonempty. Let us choose a point xt belonging to this intersection. Similarly, let us choose a point x2 belonging to all the sets ti2(A). The obvious conclusion we would like to draw is that the point xt x x2 lies in n.<e« A, for then our theorem would be proved. But that is unfortunately not true. Consider the following example, in which X, = X2 = [0, 1] and the collection a consists of all elliptical regions bounded by ellipses that have the points p — A, j) and q = (\, |) as their foci. See Figure 1. Certainly a satisfies the f.i.c/Now let us pick a point x, in the intersection of the sets [n,(A)\A e a}. Any point of the interval Figure I
§5-1 The Tychonoff Theorem 231 r^l will do; suppose we choose x, = \. Similarly, choose a point x2 in the intersection of the sets {n1(,A)\A e a}. Any point of the interval [j, Z] will do; suppose we pick x2 = \. This proves to be an unfortunate choice, for the point *i X x2 = \ x \ does not lie in the intersection of the sets A. "Aha!" you say, "you made a bad choice. If after choosing *, = i you had chosen x2 = 7. then you would have found a point in f\A^a A" The difficulty with our tentative proof is that it gave us too much freedom in picking Xi and x2; it allowed us to make a "bad" choice instead of a "good" choice. How can we alter the proof so as to avoid this difficulty? This question leads to the second idea of the proof: Perhaps if we expand the collection a (retaining the f.i.c, of course), that expansion will restrict the choices of xt and x2 sufficiently that we will be forced to make the "right" choice. To illustrate, suppose that in the previous example we expand the collection a to the collection D consisting of all elliptical regions bounded by ellipses that have/7 = (j, ])as one focus and any point of the line segment pq as the other focus. This collection is illustrated in Figure 2. The new collec- Figure 2 ''on a still satisfies the f.i.c. But if you try to choose a point x, in 'he only possible choice for x, is ^. Similarly, the only possible choice_for xi « \. And ^ x ^ does belong to every set D, and hence to every set A. In °ther words, expanding the collection a to the collection 2) forces the proper ch°ice on us.
232 The Tychonoff Theorem Chap.j Now of course in this example we chose D carefully so that th would work. What hope can we have for choosing D correctly in * Here is the third idea of the proof: Why not simply choose ffl to be a cT^ tion that is "as large as possible"—so that no larger collection satisfied th" f.i.c.—and see whether such a D will work? It is not at all obvious that a collection D exists; to prove it, we must appeal to the maximum princil of set theory. But after we prove that D exists, we shall in fact be able t show that 2) is large enough to force the proper choices on us. ° Now let us apply these ideas: Theorem 1.1 (The Tychonoff theorem). An arbitrary product of compact spaces is compact in the product topology. Proof. Step 1. Let X be a set; let a be a collection of subsets of Xsatis- Xsatisfying the f.i.c. We show that there is a collection D of subsets of X such that A) D r> a. B) D satisfies the f.i.c. C) If S ^ D, then S does not satisfy the f.i.c. We often express conditions B) and C) by saying that D is "maximal with respect to the f.i.c." Recall the maximum principle. It states that if -< is a strict partial order relation on the set A, and if B is a subset of A that is simply ordered by -<, then there is a maximal simply ordered subset C of A containing B. We need a special case of this principle, the case where B is a one-element subset of A; such a subset is automatically simply ordered. Thus what we need is the following: If A is a set with a strict partial order and if a. is an element of A, then there is a maximal simply ordered subset C of A such that a e C. We shall apply this principle in the case where A is not a set, nor even a collection of sets, but rather a set whose elements are collections of sets. For purposes of this proof, we shall call such a set a "superset" and shall denote it by an outline letter. To summarize the notation: c is an element of X, C is a subset of X, C is a collection of subsets of X, C is a superset of collections of subsets of X. We are given the collection a of subsets of X, satisfying the f.i.c. w A denote the superset consisting of all collections of subsets of A'that sa isj the f.i.c. Then a e A. Let us use proper inclusion ? as the stn<* P\' order on A. By the special case of the maximum principle just cited, tne exists a maximal simply ordered subsuperset C of A such that a e C. Define D to be the union of the elements of C; We assert that 33 is the collection we want.
§54 The Tychonoff Theorem 233 First, note that 8cS, because 8e£. Second, we show that 33 satisfies the f.i.c Let Du ..., D. be elements of 3). Since 3) is the union of the ele- elements of C, for each i there is an element e, of £ such that D, e 6 Now C is simply ordered by proper inclusion. Therefore, the superset {e, ' e \ is also simply ordered by proper inclusion. Being finite, it has a largest'ele- largest'element; that is, there is an index k such that 6, c ek for i = 1,..., „. ln par. ticular, the sets Du ..., Dn all belong to ek. Since ek satisfies t'he'f.i.c. by hypothesis, D, O • • • O Dn is nonempty. Third, we show 33 is maximal with respect to the f.i.c. Let 3D £ S and suppose that S satisfies the f.i.c. Then we can form a new superset C by ad- adjoining the element S to C; C' = C u {s}. Now 6 c 3) for every 6 e C, and D £ S. Therefore, every two elements of C are comparable under proper inclusion, so that C is simply ordered. This contradicts the maximality of C. Step 2. Let D be a collection of subsets of X that is maximal with respect to the f.i.c. Then: (a) Any finite intersection of elements o/2D belongs to 33. Let B be the intersection of finitely many elements of 33. Define a collec- collection S by adjoining B to D, so that S = D U {B}. We show that S satisfies the f.i.c.; then maximality of D implies that S = 3D, whence B e 3D as de- desired. Take finitely many elements of S. If none of them is the set B, then their intersection is nonempty because 3D satisfies the f.i.c. If one of them is the set B, then their intersection is of the form D, n ■ ■ ■ n Dm n b. Since B equals a finite intersection of elements of 3D, this set is nonempty. (b) If A is any subset ofX that intersects every element o/3D, then A be- belongs to ID. Given A, define S = 3D U {A}. We show that S satisfies the f.i.c, from which we conclude that A belongs to 3D. Take finitely many elements of S. If none of them is the set A, their intersection is automatically nonempty. Otherwise it is of the form D, n • • • n Dn n A. Now D, n • • ■ n Dn belongs to 3D, by (a); therefore, this intersection is nonempty, by hypothesis. Step 3. Now we prove the Tychonoff theorem. Let X = Ha ej Xa) where each space X. is compact. Let a be a collection of subsets of X satis-
234 The Tychonoff Theorem Chap. 5 fying the f.i.c. We prove that the intersection CUmtA is nonempty. Compactness of AT follows. Applying Step 1, choose a collection D of subsets of X such that S> and S> is maximal with respect to the f.i.c. It will suffice to prove Z?£ intersection f)DeI) D is nonempty. te Given i6/, let n, : X -> X, be the projection map, as usual. Consider the collection ' [n.(D)\D e 33} of subsets of X,. This collection satisfies the f.i.c. because 33 does. By com- compactness of Xa, we can for each a choose a point x. of Xa such that x. e CVo *«@). Let x be the point (.va)a6J of X. We shall show that x e D for every D e £>; then our proof will be finished. First we show that if 7i^[(Uf) is any subbasis element (for the product topology on X) containing x, then n'f\Vf) intersects every element of 3); The set Uf is a neighborhood of xf in Xr Since x0 e ~nJP) by definition, 11^ intersects n^D) in some point ^(y), where y e D. Then it follows that y e n;'(Uf) r\D. It follows from (b) of Step 2 that every subbasis element containing x belongs to D. And then it follows from (a) of Step 2 that every basis element containing x belongs to 2X Since D has the f.i.c, this means that every basis element containing x intersects every element of 33; whence x e D for every D e D, as desired. □ Exercises 1. Let X be a space. Let D be a maximal collection of subsets of Xsatisfying the (a) Show that x e 5 for every D e D if and only if every neighborhood of* belongs to D. Which implication uses maximality of £>? (b) Let D E D. Show that if /I = A then ^eD. . t0 (c) Show that if X is Hausdorff, there is at most one point belongi s 2. A collection a of subsets of X satisfies the countable fnterse^tfon (c i c) if every countable intersection of elements of a is nonempty- jsfy. X is a Lindelbf space if and only if for every collection a of subsets ing the c.i.c, pUea^^ 0• 3. Consider the three statements: K,atisfyingtheci'C" (i) If Xi% a set and a is a collection of subsets of X satistying
§5 Completely Regular Spaces 235 then there is a collection 33 of subsets of X such that SD => a and 3) f. maximal with respect to the c.i.c. 00 Suppose 33 is maximal with respect to the c.i.c. Then countable intersec- intersections of elements of 33 are in 33. And if A is a subset of Xthat interacts every element of 33, then necessarily A e 33. (iii) Products of Lindelof spaces are Lindelof. (a) Prove (ii). (b) Show that (i) and (ii) together imply (iii). (c) Given a set X, show there exists a maximal collection of subsets of X satisfying the c.i.c. Why is this not adequate to prove (i)? (d) The space R, is Lindelof and the space R} is not. (See Example 4 of §4-1). Therefore, (i) does not hold. Where does the proof of (i) break down? 4. Here is another theorem whose proof uses the maximum principle. Recall that if A is a space and if x, y e A, we say that a- and y belong to the same quasi- component of A if there is no separation A = C \j D oiA into two disjoint sets open in A such that x e C and y e D. (a) Let X be a compact Hausdorff space; let x, y e X. Let ft be a collection of closed sets of A'such that for each A e 8, the points x and y lie in the same quasicomponent of A. If ft is simply ordered by proper inclusion, show that x and y lie in the same quasicomponent of Y = fWa /*• [//;>//: See Exercise 12 of §3-5.] (b) Prove the following: Theorem, If X is compact Hausdorff, the quasicomponents of X equal the components of X. [Hint: If x and y lie in the same quasicomponent of X, let ft be the collec- collection of all closed sets A in JTsuch that .v and y lie in the same quasicomponent of A. Let ft' be a maximal simply ordered subcollection and show that rw a is connected.] (c) Prove the following: Corollary. Let X be compact Hausdorff; let x e X. The intersection of all those sets A containing x which are both open and closed in X equals the com- component of X containing x. 5-2 Completely Regular Spaces We have already mentioned that the failure of the proof of the Urysohn lemma to generalize to regular spaces leads us to formulate a new separa- separation axiom, the axiom of complete regularity. In this section we study some of the properties of spaces satisfying this axiom. One jus^cat'on for oU interest in this class of spaces comes from an imbedding theorem we shal prove, which shows that the class of completely regular spaces is identical
The Tychonoff Theorem Chap. 5 with a class of spaces we have already considered, the class of all subspaces of compact Hausdorff spaces. Definition. A space X is completely regular if one-point sets are closed in Zand if for each point x0 and each closed set A not containing *„, there is a continuous function/: X—> [0, 1] such that f{x0) = 1 and f(A) = [0}. A normal space is completely regular, by the Urysohn lemma, and a completely regular space is regular, since given/, the sets /~'([0,|)) and /""'(Oz, 1]) are disjoint open sets about A and x0, respectively. As a result, this new axiom fits in between regularity and normality in the list of separa- separation axioms. Note that in the definition one could just as well require the function to map x0 to 0, and A to {1}, for g{x) = 1 — f{x) satisfies this condition. Theorem 2.1. A subspace of a completely regular space is completely regular. A product of completely regular spaces is completely regular. Proof Let X be completely regular; let Y be a subspace of X. Let x0 be a point of Y, and let A be a closed set of Y disjoint from x0. Now A = A n Y, where /denotes the closure of A in X. Therefore, x0 <£ A. SinceX is completely regular, we can choose a continuous function /: X—> [0,1] such that/(.v0) = 1 and /(/) = {0}. The restriction of/to Yis the desired continuous function on Y. Let X = JlXa be a product of completely regular spaces. Let b =(A.) be a point of Xand let A be a closed set of X disjoint from b. Choose a basis element J[U. containing b that does not intersect A; then Ua = Xa except for finitely many a, say a = a,, .. . , «,„. Given / = I, ... ,11, choose a con- continuous function /,: Xai -> [0, 1] such that f{bj = 1 and f{X- UJ = {0}. Let 0,(x) =/(*«,(x)); then <j>, maps X continuously into R and vanishes outside 7r«,'(£4,). The product /W=D,(x).^2(x) 0.(x) is the desired continuous function on X, for it equals 1 at b and vanishes outside n^«. □ Example 1. The space Sn x Sa is a space that is completely regular, since it is a subspace of the normal space Sn x Sa. But it is not normal; see Example 2 of §42 2 of §4-2. 2 of §42. A space that is regular but not completely regular is much harder to find- Most of the examples that have been constructed for this purpose are difficult, and require considerable familiarity with cardinal numbers. Fairly recently, however, John Thomas [T] has constructed a much more elementary example, which we utlin i Ei 6 which we outline in Exercise 6.
§5-2 ComPleteIy Regular Spaces 237 The separation of disjoint closed sets by continuous real-valued functions is the crucial tool used in proving the Urysohn metrization theorem- one uses it to imbed the space ATinto the product space R°. In a completely regular space, we can separate points from closed sets by continuous functions, so it is natural to ask whether one can prove a similar theorem under the hy- hypothesis of complete regularity. One can. It does not turn out to be a metriza- metrization theorem, but it is an interesting imbedding theorem: Theorem 2.2. IfX is completely regular, then X can be imbedded in [0, \]s for some J. Proof. Let {/„}«€/ denote the set of all continuous functions from X into the interval [0, 1], indexed by some index set J. Complete regularity of X guarantees that this collection separates points from closed sets in X. Therefore, by the Imbedding theorem of §4-4, the map is an imbedding of X in [0, \)J. □ Corollary 2.3. Let X he a space. The following are equivalent: A) X is completely regular. B) X is homeomofphic to a subspace of a compact Hausdorff space. C) X is homeomorphic to a subspace of a normal space. Exercises 1. Show that every locally compact Hausdorff space is completely regular. 2. Show that although R} is not normal, it is completely regular. 3. Let A" be completely regular; let A and B be disjoint closed subsets of X. Show that if A is compact, there is a continuous function /: X-*[0,1] such that /G4) {/ { *4. Show that &• in the box topology is completely regular. (This fact is an imme- immediate consequence of the following theorem, but give a direct proof.) *S. Prove the following: Theorem. Every Hausdorff topological group is completely regular. Proof. Let Ko be a neighborhood of the identity element e, in the topolo- topological group G. In general, choose V. to be a neighborhood of e such. tha y.-V.c K.-,. Consider the set of all dyadic rationals p that is. all rational numbers of the form A/2", with k and n integers. For each dyadic rational p i @,1], define an open set U(p) inductively as follows: CAD - V* a"a i = K,. Given n, if C/(*/2") is defined for 0 < W<, 1, define in
238 The Tychonoff Theorem q. for 0 < k < 2". For p <. 0, let U(p) = 0; and for p > 1, iet Show that Vn • U(k/2") <= U((k for all k and n. Proceed as in the Urysohn lemma. This exercise is adapted from [M-Z], to which the reader is referred fo further results on topological groups. ♦6. Define a set X as follows: For each even integer m, let Lm denote the line segment m x [-1,0] in the plane. For each odd integer n and each integer k 2> 2, let Cn,t denote the union of the line segments (n + 1 — l/k) x [-1 oj and (« — 1 + 1/*) x [—1. 0] and the semicircle {x xy[(x-nJ+y2 = (l-l/kJ and y^O] in the plane. Let />„.* denote the topmost point ux(l-l|t) of this semi- circle. Let A" be the union of all the sets Lm and Cn, k, along with two extra points a and b. Topologize X by taking sets of the following four types as basis ele- elements : (i) The intersection of A" with a horizontal open line segment that contains none of the points pn.k. (ii) A set formed from one of the sets Cn,t by deleting finitely many points, (iii) For each even integer m, the union of [a] and the set of points x x y of X for which x < m. (iv) For each even integer m, the union of [b] and the set of points x x y of X for which x > m. (a) Sketch X; show that these sets form a basis for a topology on X. (b) Let / be a continuous real-valued function on X. Show that for any c, the set/-•(<:) is a G6 set in X. (This is true for any space X.) Conclude that the set Sn,k consisting of those points p of Cn%k for which f(p) ¥= /(?»*)is countable. Choose d e [-1,0] so that the line y = d intersects none of the sets Sn,k. Show that for n odd, /((n - 1) x d) = \\m f(p^k) = /((« + 1) x d). Conclude that f(a) = f(b). (c) Show that X is regular but not completely regular. 5-3 The Stone-Cech Compactification We have already studied one way of compactifying a topological sPa'* X, the one-point compactification (§3-8); it is in some sense the ^nan compactification of X. The Stone-Cech compactification of X is in so sense the maximal compactification of X. We introduct it here as anintere ing application of the Tychonoff theorem. It has a number of appli^00 in modern analysis, but these lie outside the scope of this book.
§ 5-3 " The Stone-dech Compactification 239 Definition. A compactification of a space X is a compact Hausdorff space Y containing X such that X is dense in Y (that is, such that X = Y). Two compactifications Y{ and Y2 of X are said to be equivalent if there is a homeomorphism h : Y\ —♦ Y2 such that h(x) = x for every x e X. In order for Af to have a compactification, A'must be completely regular. Conversely, every completely regular space has at least one compactification. One way of obtaining a compactification of X is as follows: Let X be completely regular. Choose an imbedding h : X —> Z of X in a compact Hausdorff space Z. Let A^ denote the subspace h(X) of Z, and let Ya denote its closure in Z. Then Ko is a compact Hausdorff space and Xa = Ya; therefore, Yo is a compactification of A"o. We now construct a space K containing X such that the pair (X, Y) is homeomorphic to the pair (Xo, Yo). Let us choose a set A disjoint from X that is in bijective correspondence with the set Yo — Xo under some map k : A —» Ko — AV Define K = A" U A, and define a bijective correspon- correspondence H : Y — y0 by the rule //(x) = A(a-) for x e X, H(a) = A(fl) for a e A. Then topologize K by declaring U to be open in Y if and only if H{U) is open in Ya. The map // is automatically a homeomorphism; and the space A" is a subspace of Y because H equals the homeomorphism /; when restricted to the subset X of Y. The space Y is called the compactification of A" induced by the imbedding h. We summarize the preceding construction as follows: If h\X-^>Z is an imbedding of X in the compact Hausdorff space Z, then h induces a compactification Y of X. It has the property that the imbedding h can be extended to an imbedding H : Y —> Z. In general, there are many different ways of compactifying a given space X. Consider for instance the following compactifications of the open interval * = @,l): Example 1. Take the unit circle S1 in Rz and let h: @, 1) — S> be the map h(t) = (cos 2jw) x (sin 2nt). The compactification induced by the imbedding h is equivalent to the one- point compactification of X. Example 2. Let Y be the space [0, 1]. Then Y is a compactification of X; it is obtained by "adding one point at each end of @, 1)." Example 3. Consider the square [-1, l]2in«2and let A:@, 1) — [-1, I]2 be the map
240 The Tychonoff Theorem Chap. 5 h(x) = x x sin(l/;t). The space Yo = hJX) is the topologist's sine curve (see Example 8 f s The imbedding h gives rise to a compactification of @, 1) quite differ the other two. It is obtained by adding one point at the right-hand ^ @, 1), and an entire line segment of points at the left-hand end! °f A basic problem that occurs in studying compactifications is the foil ing: w" If Y is a compactification ofX, under what conditions can a continuous real-valued function f defined on X be extended continuously to Y1 The function / will have to be bounded if it is to be extendable, since its extension will carry the compact space Y into R and will thus be bounded But boundedness is not enough, in general. Consider the following example: Example 4. Let X = @, 1). Consider the one-point compactification of A" given in Example 1. A bounded continuous function/: @, 1) —► R is extendable to this compactification if and only if the limits lim /(.v) and lim /(.v) exist and are equal. For the ''two-point compactification" of X considered in Example 2, the function/is extendable it and only if both these limits simply exist. For the compactification of Example 3, extensions exist for a still broader class of functions. It is easy to see that /is extendable if both the above limits exist. But the function /(.v) = sin (l/.v) is also extendable to this compacli- fication: Let H be the imbedding of Kin R~ which equalsh on thesubspace*. Then the composite map is the desired extension of/. For if .v e X, then H(x) =*(■*)=* x ^"C^ so that n2(H(x)) = sin (l/.v), as desired. There is something especially interesting about this last compactification- We constructed it by choosing an imbedding A:@, 1)—>R2 whose component functions were the functions a- and sin(l/-v- ^m_ found that both the functions x and sin (l/.v) could be extended to j* ded pactification. This suggests that if we have a whole collection o\ ° „, continuous functions defined on @, 1), we might use them as c°Tb(al-n functions of an imbedding of @, 1) into R1 for some /, and there ^^ a compactification for which every function in the collection is e; J( This idea is the basic idea behind the Stone-Cech compacting is defined as follows: ruction of11 Let A" be a completely regular space. Let {/.}.« be the coiieci
§ 5-3 The Stone-dech Compactification 241 bounded continuous real-valued functions on X, indexed by some index set J. For each cc e J, choose a closed interval /, in R containing f,(X). To be definite, choose h = [glb/.W, Iub/.(A0]. Then define h : X —<■ H«e/ h by the rule Kx) = (Ux))aeJ. By the Tychonoff theorem, ]T h is compact. Because X is completely regular, the collection {/,} separates points from closed sets in X. Therefore, by the Imbedding theorem of §4-4, h is an imbedding. The compactification of X induced by h is called the Stone-Cech com- compactification of X. It is commonly denoted by fi(X). Many mathematicians avoid using the arbitrary index set /when defining the Stone-Cech compactification. Instead, they use the collection JF of all bounded continuous real-valued functions on X as its own index set, letting / = JF and letting the indexing function f-.J—> JF be the identity function. This choice of indexing is not essential for the definition, however; and we feel it leads to notational confusion. The crucial property of the Stone-Cech compactification is the following extension property: Theorem 3.1. Let X be completely regular; let fi(X) be its Stone-Cech compactification. Then every bounded continuous real-valued function on X can be uniquely extended to a continuous real-valued function on fi(X). Proof. The compactification fi(X) is induced by the imbedding h : X —> II h defined above. This means there is an imbedding H: fi(X) -^Yih that equals h when restricted to the subspace X of fi(X). Given a continuous bounded real-valued function on X, it equals ff for some fi e /. Now if *t '■ II h —<■ If is projection onto the /fth coordinate, then the composite map 7cf o H: fi(X) —■> If is the desired extension of ff. For if x e X, we have nf{H(x)) = nf{h{x)) = np((fMUj) =/,>«• Uniqueness of the extension is a consequence of the following lemma. Q Lemma 3.2. Let A c X; let f .• A — Z be a continuous map of A into the Hausdorff space Z. There is at most one extension off to a continuous fund ion g : A — Z. Proof. This lemma was given as an exercise in §2-7; we give a proof here. Suppose that g, g' : A—> X are two different extensions of/; choose x so that g(.v) 7t g'(x). Let U and U' be disjoint neighborhoods of g(x) and g'(x), respectively. Choose a neighborhood V of x so that g{V) c Uand g\V) c
The Tychonoff Theorem Chap. 5 U' Now V intersects A in some point y; then g(y) e U and g'(y) 6 „, But since jeiwe have g(.y) = /GO and g'(y) = /G>). This contradict the fact that U and V are disjoint. Q We now prove a theorem to the effect that the Stone-Cech compactig. cation is essentially unique, and is characterized by the extension property. Theorem 3.3. Let X be completely regular. Let Y, and Y2 be two compactifi. cations o/X having the extension property of Theorem 3.1. Then there is a homeomorphism $ o/Y, onto Y2 such that 0(x) = xfor each x e X. Proof. Step 1. We first prove the following fact: Suppose Y is a com- pactification of X having the extension property stated in Theorem 3.1. If Z is any compact HausdorfF space and g : X—> Z is any continuous func- function, then g can be extended to a continuous function k mapping Y into Z. To prove this fact, note that Z is completely regular, so that it can be imbedded in [0, l]y for some J. So we may as well assume that Z cz [0, If. Now consider g : X—► Z cz [0, l]y cz R>, Each component function & of the map g is a continuous bounded real-valued function on X; by hypothe- hypothesis, g, can be extended to a continuous map k« of Y into R. Define k : Y—> RJ by setting k{y) = (k,(y))«ej- The map k is continuous because R' has the product topology. We assert that k actually maps Y into the subspace Z. For g{X) is contained in Z, and k(X) ■-=■- g(X). Since Z is closed in R', it follows that k(X) cz Z. By continuity of k, k{Y) = k{X) cz k(X). Therefore, k maps Y into Z. Step 2. Now we prove the theorem. Consider the inclusion mapping Ji '■ X — Y2. It is a continuous map of ^Tinto the compact HausdorfF space Yz. Because K, has the extension property, we may, by Step I, extend/: to a continuous map/, : Y, — Yz. Similarly, we may extend the inclusion map y, : X—> Yx to a continuous map/, : Y2 — K, (because r2 has the extension property and Y, is compact Hausdorff). The composite /, «,/, : y, — y, has the property that for every x e X, one has/,(/,(*)) = x. Therefore,/, o/, is a continuous extension of the iden- identity map ix:X~> X. But the identity map of y, is also a continuous ex- extension of /,. By uniqueness of extensions (Lemma 3.2),/, o/2 must equal the ufcntity map of y,. Similarly,/, o/l must equal the identity raapof Tr lhus/, and/2 are homeomorphisms. Q
S 5'3 The Slone-dech Compactification 243 Exercises 1. Verify the statements made in Examples 1 and 4. 2. Show that the bounded continuous function g: (o,!)—»/? defined by g(x) = cos (l/.v) cannot be extended to the compactification of Example 3. Define an imbedding h : @, 1) —>■ R* such that the functions x, sin (I/*), and cos (l/x) are all extendable to the compactification induced by h. 3. Under what conditions does a metrizable space have a metrizable compactifica- compactification? 4. Let Y be an arbitrary compactification of X; let fi(X) be the Stone-Cech com- compactification. Show there is a continuous surjective closed mapg-.fi(X) ► Y that equals the identity on X. [This exercise makes precise what we mean by saying that fi(X) is the "maxi- "maximal" compactification of X. If you are familiar with quotient spaces, you will recognize that g is a quotient map. Thus every compactification of A"is equiva- equivalent to a quotient space of fi{X).] 5. (a) Show that every continuous real-valued function defined on Sn is "even- "eventually constant." [Hint: First prove that for each f, there is an element a of Sa such that |/(yff) - f{OL)\ < e for all fi > a. Then let e = 1/n for n e Z+ and consider the corresponding points an.] (b) Show that the one-point compactification of Sa and the Stone-Cech compactification are equivalent. (c) Conclude that every compactification of Sn is equivalent to the one-point compactification. 6. Let X be completely regular. Show that X is connected if and only if P(X) is connected. [Hint: If X = A U B is a separation of X, let /(.v) = 0 for x e A and /(.v) = 1 for x e B.] 7. Let X be discrete; consider fi(X). (a) Show that if A c X, then A and X - A are disjoint [closures taken in fi(X)]. (b) Show that if V is open in fi(X), then 0 is open in fi(X). (c) Show that P(X) is totally disconnected. 8. Show that 0{Z+) has cardinality at least as great as 1', where / = [0, 1]. [Him: I1 has a countable dense subset, by Exercise 13 of §4-1.] 9. Show that if fi(X) * X, then 0(X) is not metrizable. [Hint: Show that if A"is normal and y is a point of fi(X) - X, then y is not the limit of a sequence of points of X.]
6. Metrization Theorems and Paracompactness The Urysohn metrization theorem of Chapter 4 was the first step—a giant one—toward an answer to the question: When is a topological space metriz- able? It gives conditions under which a space X is metrizable: that it be regular and have a countable basis. But mathematicians are never satisfied with a theorem if there is some hope of proving a stronger one. In the present case, one can hope to strengthen the theorem by rinding conditions on X which are both necessary and sufficient for A'to be metrizable, that is, con- conditions which are equivalent to metrizability. We know that the regularity hypothesis in the Urysohn metrization theo- theorem is a necessary one, but the countable basis condition is not. So the obvi- obvious thing to do is try to replace the countable basis condition by something weaker. Finding such a condition is a delicate task. The condition has o strong enough to imply metrizability, and yet weak enough that all metnza spaces satisfy it. In a situation like this, discovering the right hypothesis more than half the battle. . y The condition that was eventually formulated, by J. Nagata an ^ Smirnov independently, involves a new notion, that of local finiteness.f say that a collection a of subsets of a space X is locally finite if every po & X has a neighborhood that intersects only finitely many eI^men tabIe'js Now one way of expressing the condition that the basis <& iscoun to say that (B can be expressed in the form 244
§ 6-1 ■■■.:■■ z,ocai Finiteness 245 where each collection <$>„ is finite. This is an awkward way of saying that (B is countable, but it suggests how to formulate a weaker version of it. The Nagata-Smirnov condition is to require that the basis <B can be expressed in the form where each collection <$>„ is locally finite. We say that such a collection (B is countably locally finite. Surprisingly enough, this condition, along with regularity, is both necessary and sufficient for metrizability of X. In §6-1 we study the notion of local finiteness. In §6-2 we show that the Nagata-Smirnov condition for X, along with regularity, implies metrizability of X; the proof is an adaptation of the proof of the Urysohn metrization theorem. In §6-3 we prove that metrizability implies the Nagata-Smirnov con- condition; this proof uses the well-ordering theorem. There is another concept in topology which involves the notion of local finiteness. It is a generalization of the concept of compactness called "para- compactness." Although of fairly recent origin, it has proved useful in many parts of mathematics. We introduce it here so that we can give another set of necessary and sufficient conditions for a space X to be metrizable. It turns out that X is metrizable if and only if it is both paracompact and locally metrizable. This we prove in §6-5. Some of the sections of this chapter are independent of one another. The dependence among them is expressed in the following diagram: . §6-1 Local finiteness. Nagata-Smirnov metrization theorem (sufficiency). §6-3 The Nagata-Smirnov theorem (necessity). §6-4 Paracompactness. §6-5 The Smirnov metrization theorem. 6-1 Local Finiteness In this section we prove some elementary properties of locally finite col- collections. Definition. Let A' be a topological space. A collection a of subsets of X is said to be locally finite if every point of X has a neighborhood that inter- intersects only finitely many elements of a. Example 1. The collection of intervals G = {(«, n + 2)| n eZ]
246 Metrization Theorems and Paracompactness is locally finite in the topological space X, as you can check. So is the coll <B={(«,2n)|neZ+}. eC<1On On the other hand, the collection C = {@, l/n)|neZ+j is not locally finite in R, nor is the collection Lemma 1.1. Let a be a locally finite collection of subsets ofX. Then: (a) Any subcollection of a is locally finite. (b) The collection <B = {A}Aea of the closures of the elements of a is locally finite. (c) Ul^A = UAea A. Proof. Statement (a) js trivial. To prove (b), note that any open set U which intersects the set A necessarily intersects A. Therefore, if U is a neigh- neighborhood of x that intersects only finitely many elements A of a, then C/can intersect at most the same number of sets of the collection (B. (It might inter- intersect fewer sets of (B, since At and /F2 can be equal even though A, and Az are different.) To prove (c), let Y denote the union of the elements of «: In general, [JA c Y; we prove the reverse inclusion, under the assumption of local finiteness. Let x e Y; let U be a neighborhood of x that intersects only finitely many elements of a, say Au . . . , Ak. We_assert that x belongs to one of the sets Alt. . . , Ak, and hence belongs to \JA. For otherwise, the set U — A, — ■ ■ ■ — Ak would be a neighborhood of x that intersects no ele- element of a and hence does not intersect Y, contrary to the assumption that x e Y. □ In the exercises of §2-7 and §4-5, we defined a concept of local finiteness for an indexed family of subsets of X. The indexed family {A.].*i is said to a locally finite indexed family if every x e A" has a neighborhood that k*1** A. for only finitely many values of a. What is the relation between Ite ^ formulations of local finiteness? It is easy to see that [A.].*/ is alocally indexed family if and only if it is locally finite as a collection of sets ana nonempty subset A of A" equals A* for at most finitely many values ot . We shall not be concerned with locally finite indexed families in tnis en v except in one or two exercises. Definition. A collection <B of subsets of X is said to be conntably finite if (B can be written as the countable union of collections «„ which is locally finite.
§ 6-2 Ue NaS"f-Smirnov Metrization Theorem (.sufficiency) 247 Most authors use the term "cr-Iocally finjte" for this concept. The<r comes from measure theory and stands for the phrase "countable union of" Note that both a countable collection and a locally finite collection are countably locally finite. Exercises 1. Check the statements in Example l. 2. Check the relation stated above between local finiteness for a collection of sets and for an indexed family of sets. 3. Find a point-finite open covering G of R that is not locally finite. [The collection G is point-finite if each point of R lies in only finitely many elements of «.] 4. Give an example of a collection of sets a that is not locally finite, such that the collection (B = [A\ A e «} is locally finite. 5. Show that if A" has a countable basis, a collection ft of subsets of A"is countably locally finite if and only if it is countable. 6. Take R" in the box topology. Find a collection of subsets of R" that is countably locally finite but is neither countable nor locally finite. 7. Many spaces have countable bases; but no Hausdorff space has a locally finite basis unless it is discrete Prove this fact. 8. Find a space that has a countably locally finite basis but does not have a count- countable basis. 6-2 The Nagata-Smirnov Metrization Theorem (sufficiency) Now we prove that regularity and the Nagata-Smirnov condition suffice to prove a space metrizable. The proof follows very closely the second proof we gave of the Urysohn metrization theorem. In this proof we constructed a map of the space X into R" that was an imbedding relative to the uniform metric p on R". So let us review the major elements of that proof. The first step of the proof was to prove that every regular space with a countable basis is normal. The second step was to construct a countable collection {/„} of real-valued functions on ^that separated points from closed sets. And the third step was to use the functions/, to imbed A'in the metric space (/?", p). Each of these steps needs to be generalized in order to prove the Nagata- Smirnov theorem. First, we prove that a regular space X with a countably locally finite basis is normal. Second, we construct a certain collection of real-valued functions {/.} on X. Third, we use these functions to imbed X in the metric space (RJ, p) for some /.
in 248 Metrization Theorems and Paracompactness Chap. Before we start, we need to recall a notion we have already intiwi the exercises, that of a Gs set. lr°auced Definition. A subset A of a space X is called a G, set in X if it equals th intersection of a countable collection of open subsets of X. Example 1. Each open subset of X is a G6 set, trivially. In a first-countabl Hausdorff space, each one-point set is a G6 set. The one-point subset {Qj Of Sa is not a G6 set, as you can check. Example 2. In a metric space X, each closed set is a Gd set. Given AcX define ) = LUx B(-v, e). If .4 is closed, you can check that a = rw. Lemma 2.1. Let X be a regular space with a basis (B that is countably locally finite. Then X is normal, and every closed set in X is a G6 set in X. Proof. Step I. Let W be open in X. We show there is a countable col- collection {£/„] of open sets of X such that W --= \JU. = [jUn. Since the basis (B for X is countably locally finite, we can write (B = \J&,, where each collection (&„ is locally finite. Let 6n be the collection of those basis elements B such that 5 ~ «„ and 5 c W. Then e, is locally finite, being a subcollection of ($>„. Define Then [/„ is an open set, and by Lemma I.I, Therefore, £/„ c W7, so that U u We assert that equality holds. Given x e IV, there is by regularity a basis element B e ffl such that x e 5 and 5 c H7. Now 5 e <&n for some". " fi g e, by definition, so that x e Un. Thus ff <= IJt/,,, as desired. Step 2. We show that every closed set C in X is a G6 set in Jf. Given ^ let fK = A' — C. By Step I, there are sets U, in X such that W - UU'- so that C equals a countable intersection of open sets of X. Step 3. Now we show X is normal. Let Cand D be disjoint closed seM» X. Applying Step I to the open set X - A we construct a countable
§ 6-2 The Nagata-Smirnov Metrizathn Theorem (sufficiency) 249 tion {£/„} of open sets such that \JUn = \JUn = X - D. Then {£/„} covers C and each set Un is disjoint from D. Similarly, we construct a countable covering [Vn] of D by open sets whose closures are disjoint from C. Now we are back in the situation which arose in the proof that a regular space with a countable basis is normal (Theorem 2.5 of Chapter 4). We can repeat that proof verbatim. Define U'n = Un- \J{., V, and Vn = V. - \J-_, &,. Then the sets U' = \J.ez.V. and V' = \J.ez.K are disjoint Open sets about C and D, respectively. Q Theorem 2.2. Let X be a regular space with a basis (B that is countably locally finite. Then X is metrizable. Proof. Step I. First we show that if W is open in X, there is a continuous function f:X—+ [0, I] such that f(x) > 0 for x e W and f(x) = 0 for x $ W. By the preceding lemma, each closed set of A' is a countable intersection of open sets of X. Passing to complements, it follows that the open set Wis a countable union of closed sets An of X. Using normality, choose for each positive integer n, a continuous function/^ : X —> [0, I] such that/,(^n) = {1} and fXX - W) = {0}. Define /(.v) = S/n(.v)/2". The series converges uniformly, by comparison with ]T] 1/2", so that / is continuous. Also, / is positive on f^and vanishes outside W. Step 2. Let (B = |J(Bn, where each collection <&„ is locally finite. For each positive integer n, and each basis element B e <&n, choose a continuous function /..»:*—►[0,I/«] such that fn<B(x) > 0 for x e B and /» = 0 for x $ B. The collection {/«,a} separates points from closed sets in X: Given a point xa and a neigh- neighborhood U of xa, there is a basis element B such that xa e B a U. Then B e <S>n for some n, so that/nfl(.v0) > 0 and/B.a vanishes outside U. Let / be the subset of Z+ X (B consisting of all pairs (n, B) such that B belongs to (Bn. Define F:X—^[0>I]/ by the equation F(x) = (/n,flW)(,,B)s/. Relative to the product topology on [0, \]J, the map Fis an imbedding, by the Imbedding theorem of §4-4. Of course, [0, l]J is not metrizable in general, so we have not yet proved the metrization theorem.
Metrization Theorems and Paracompactness Chap. 6 Step 3. Now we give [0, I]' the topology induced by the uniform • p and show that F is an imbedding relative to this topology as well Th' uniform topology is finer (larger) than the product topology. Theref relative to the uniform metric, the map F is injective and carries open s t"*' .AT onto open sets of the image space Z = F{X). We must give a separate pr* f that F is continuous. Note that on the subspace [0, 1]' of R1, the uniform metric equals the metric To prove continuity, we take a point x0 of X and a number e > 0, and find a neighborhood W of xa such that x g W=>p(F(x),F(xo))<e. Let n be fixed for the moment. Choose a neighborhood Un of xa that intersects only finitely many elements of the collection (Bn. This means that as B ranges over (£„, all but finitely many of the functions/, B are identically equal to zero on £/„. Because each function /, fl is continuous, we can now choose a neighborhood Vn of .v0 contained in £/„ on which each of the remain- remaining functions/,, jj, for B e ($>„, varies by at most e/2. Choose such a neighborhood Vn of a for each n e Z+. Then choose # so that l/N < e/2, and define W = K, n • • • D VN. We assert that W\% the desired neighborhood of .v0. Let x s IV. If n <Z N, then !/„,*(*)-/».*(*o)l<f/2 because the function/, B either vanishes identically or varies by at mostf/2 on W. \fn > N, then !/,.*(*)-/U-OI <i/«<f/2 because/,^ maps X into [0, 1/n]. Therefore, p(F(x), F(x0)) < e/2 < e, as desired. □ Exercises 1. (a) Show that in a metric space, every closed set is a G6 set. (b) Show that the irrationals are a Gd set in R. 2. A subset Wot Xh said to be an "F_ set" in Xit Wequals ac°unfiJ!f!"f^i closed sets of X Show that Wis an Fa set in ^if and only if X- " [The terminology comes from the French. The "F"'stands f°^*' which means "closed," and the "a" for "somme," which means
§6'3 The Nagata-Smirnov Theorem (necessity) 251 3. Let (X, d) be a metric space. Show that if u is open in X, then the function /to = d(x, X - V) defined in 6-3 The Nagata-Smirnov Theorem (necessity) We now prove that every metrizable space has a countably locally finite basis. This will complete the proof of the Nagata-Smirnov metrization theo- theorem. Definition. Let G be a collection of subsets of the space X. A collection (B of subsets of X is said to be a refinement of G (or is said to refine a) if for each element B of <&, there is an element A of a containing B. If the elements of® are open sets, we call (B an open refinement of G; if they are closed sets, we call (B a closed refinement. The first step in proving the necessity of the Nagata-Smirnov condition is the following lemma. It will be generalized in the next section. Lemma 3.1. Let X be a metrizable space. If G is an open covering ofX, then there is a collection D of subsets ofX such that: A) D is an open covering ofX. B) D is a refinement ofd. C) 2) is countably locally finite. Proof. We shall need to use the well-ordering theorem in order to prove this theorem. Choose a well-ordering < for the collection G. Let us denote the elements of G generically by the letters U,V,W, .. .. Choose a metric for X. Let >i be a positive integer, fixed for the moment. Given an element U of G, let us define Sn(U) to be the subset of Uobtained by "shrinking" U a distance of l/»/. More precisely, let SJLU) = {x\B(x, I/n) <= U}. (It happens that S (U) is a closed set, but that is not important for our pur- purposes.) Now we use the well-ordering < of G to pass to a still smaller set. For each U in G, define S'lU) = SlU) - VJv<o V- The situation where G consists of the three sets V<V<W* pictured in Figure 1. Just as the figure suggests, the sets we have formed are disjoint. Indeed, we assert that they are separated by a distance of at least l/». That is,
252 Metrization Theorems and Paracotnpactness Figure 1 U < V < W if Kand W are distinct elements of a, we assert that (*) x e S'n(V) and y e S'n{W) => d(x,y)>\/n. To prove this fact, assume the notation has been so chosen that V< W. Now a- e S'n(V) implies that x e 5n(K). And>> e S'n(W) implies by definition that y $ V (since V < W). Since .v e 5,(K) and >■ £ K, we must have rf(A-,j-)> 1/". The sets 5^( U) are not yet the ones we want, for we do not know that they are open sets. (In fact, they are closed.) So let us expand each of them slightly to obtain an open set £„(£/). Specifically, let £„(£/) be the I/3n neighborhood ofS'n(U); that is, In the case U < V < W, we have the situation pictured in Figure 2. As the figure suggests, the sets we have formed are disjoint, and indeed they are separated by a distance of at least I/3«. That is, if V and J-Kare distinct ele- elements of a, we assert that x e En(V) and y e En(W) => d(x,y)> I/3«- This follows at once from (*) and the triangle inequality. Note also that for each V e a, the set En(V) is contained in V. Now let us define £„ = [E.(U)\U e a}. We claim that £„ is a locally finite collection of open sets and that S, refines a. The fact that S, refines a comes from the fact that En(V) c V for « V e a. The fact that Sn is locally finite comes from the fact that for any X, the I/6n neighborhood of x can intersect at most one element of S.- Of course, the collection S, will not cover*. (Figure 2 illustrates that
§6'3 The Nagata-Smirnov Theorem (necessity) 253 £n(V) Figure 2 But we assert that the collection s = LUz.sn does cover X. Let x be a point of X. The collection a with which we began covers X; let us choose U to be ihe first element of a (in the well-ordering <) that con- contains x. Since U is open, we can choose n so that B(x, 1/n) cz U. Then, by definition, x e Sn(U). Now because U is the first element of a that contains x, the point x belongs to S'»(U). Then x also belongs to the element £„(£/) of &„, as desired. The lemma follows; S is the required collection of open sets. Q Theorem 3.2. Let X be a metrizable space. Then X has a basis that is countably locally finite. Proof. Choose a metric for X. Given m, let a- be the open covering of X by all open balls of radius 1/m; er = [B(x,l/m)\x e X}. By the preceding lemma, there is an open covering »" of Refining a"such that »- is countably locally finite. Note that each element of ST hasd.ameter at most 2/m. Let Because each collection O- is a countTbte union of locallyfinite collect ions, so is 3). We assert that 3) is a basis for X; then our theorem proved^ We prove that given, e Xznc.giver,e > CJ. ere, an element ^ ^ containing * and contained in *(*, e). Fir^n ^fV1 containing x. Then because IT covers X, we can choose an e emenl:D• ol ^.^ ^ Since Z) contains x and has diameter at most 2/m < f. ^follows from Lemma 2.3 of Chapter 2 that 3) is a basis for X. □
254 Metrication Theorems and Paracompactness Exercises 1. Let a be the following collection of subsets of R: a = {(«,« + 2)|« e Z}. Which of the following collections refines ft? (B = {(*,* + 1I* e /{}, A collection ft of subsets of X is said to be locally discrete if each point of X has a neighborhood that intersects at most one element of a. A collection (B is countably locally discrete (or "<7-locally discrete") if it equals a countable union of locally discrete collections. Prove the following: Theorem (Bing metrizalion theorem). A space X is metrizable if and only if it is regular and has a basis that is countably locally discrete. 6-4 Paracompactness The concept of paracompactness is one of the most useful generalizations of compactness that has been discovered in recent years. Particularly is it useful for applications in algebraic topology and differential geometry. We shall give just one application, a metrization theorem, which we prove in the next section. Many of the spaces that are familiar to us already are paracompact. For instance, every compact Hausdorff space is paracompact; this will be an im- immediate consequence of the definition. It is also true that every metrizable space is paracompact; this is a theorem of A. H. Stone, which we shall prove. Thus the class of paracompact spaces includes the two most important classes of spaces we have studied, the compact Hausdorff spaces and the metrizable spaces. It includes other spaces as well. (See Exercise 2.) To see how paracompactness generalizes compactness, we recall the defi- definition of compactness: A space X\s said to be compact if every open covering a of X contains a finite subcollection a' that covers X. An equivalent way of saying this is the following: A space X is compact if every open covering a of X has a finite open refinement (B that covers X. This definition is equivalent to the usual one; given such a refinement . one can choose for each element of (B an element of a containing it; in thlS way one obtains a finite subcollection of a that covers X. This new formulation of compactness is an awkward one, but it sugg«s a way to generalize:
Paracompactness 255 Definition. A space A'is paracompact if it is Hausdorff and if every open covering a of A' has a locally finite open refinement (B that covers X. Example 1. Any compact Hausdorff space is paracompact, trivially. The real line R is a paracompact space that is not compact. The fact that R is para- paracompact is a consequence of the theorem that every metrizable space is para- paracompact. But a direct proof can be given as follows: Suppose we are given an open covering ft of R. For each integer n, choose a finite number of elements of ft that cover the interval [n, n + 1] and intersect each one with the open interval (n — 1, n + 2). Let the resulting collection of open sets be denoted (Bn. Then the collection is a locally finite open refinement of ft that covers R, as you can check. Some of the properties of a paracompact space are similar to those of a compact Hausdorff space. For instance, a subspace of a paracompact space is not necessarily paracompact; but a closed subspace is paracompact. Also, a paracompact space is necessarily normal. In other ways, a paracompact space is not similar to a compact Hausdorff space; in particular, the product of two paracompact spaces need not be paracompact. One of the most useful properties of a paracompact space is the fact that given an indexed open covering {£/„} of a paracompact space X, there exists a partition of unity dominated by {£/J. A proof of this fact was outlined in the exercises of §4-5. We shall prove in this section the other properties of para- paracompact spaces we have mentioned. Theorem 4.1. Every paracompact space X is normal. Proof. The proof is somewhat similar to the proof that a compact Haus- Hausdorff space is normal. First one proves regularity. Let a be a point of X and let B be a closed set of X disjoint from a. The Hausdorff condition enables us to choose, for each b in B, an open set Ub about b whose closure is disjoint from a. Cover X by the open sets Ub, along with the open set X — B; take a locally finite open refinement e that covers X. Form the subcollection 3D of 6 consisting of every element of e that intersects B. Then 3D covers B. Furthermore, if D e 3D, then D is disjoint from a. For D intersects B, so it lies in some set Ub, whose closure is disjoint from a. Let then Vis an open set in ^containing B. Because 3D is locally finite, v = LUd d, so that V is disjoint from a. Thus regularity is proved.
256 Metrization Theorems and Paracompactness Chap. 6 To prove normality, one merely repeats the same argument renla • by the closed set A throughout and replacing the Hausdorff condit 1" regularity. □ lon ty Theorem 4.2. (a) Every closed subspace of a paracotnpact space is na compact. p^a' (b) An arbitrary subspace of a paracotnpact space need not be paracotnpact (c) A product of paracotnpact spaces need not be paracotnpact. Proof, (a) Let Y be a closed subspace of the paracompact space X- let « be a covering of Y by sets open in Y. For each A e «, choose an open set A' of X such that A' n Y = A. Cover X by the open sets A\ along with the open set X — Y. Let fflbea locally finite open refinement of this covering that covers X. The collection e = {B n Y\B e (B} is the required locally finite open refinement of a. Examples demonstrating (b) and (c) follow. □ Example 2. The spaced x Sa is compact Hausdorff, so it is paracompact. The subspace Sa x Sa is not paracompact, for it is not even normal (see §4-2). Example 3. The space Ri is paracompact; we leave the proof to the exer- exercises. But Rf is not paracompact, for it is not normal (see §4-2). Theorem 4.3 (Stone's theorem). Every metrizable space is paracompact. Proof. Let X be a metrizable space. We already know from Lemma 3.1 that every open covering of X has an open refinement that covers A'and is coimtably locally finite. It remains to prove that the latter condition implies that every open covering of X has an open refinement that covers * and is locally finite. This is a consequence of the following lemma. □ Lemma 4.4- Let X be regular. Then the following conditions on X are equivalent: Every open covering o/X has a refinement that is: A) An open covering o/X and countably locally finite. B) A covering ofX and locally finite. C) A closed covering o/X and locally finite. D) An open covering ofX and locally finite. Proof. It is trivial that D) =» A). What we need for Stone's *«^°" the converse. In order to prove the converse, we must go 'hrougtv ^ A) _» B) => C) => D) anyway, so we have for convenience listed tn ditions in the statement of the lemma. ^fi A) =» B). Let « be an open covering of X. Let (B be an open ren
Paracompactness 257 of a that covers X and is countably locally finite; let <B = (J(B, where each <&„ is locally finite. Denote the elements of <B senericallv bv u,v,w,.... y Now we apply essentially the same sort of shrinking trick we have used before to make sets from different (Bn's disjoint. Given i, let K = Uuea, U. Then for each n e Z+ and each element U of (Bn, define £„(£/) = U - U/<- K. [Note that £„(£/) is not necessarily open, nor closed.] Let 1 u e ay. Then e, is a refinement of ©„, because £„(£/) c £/ for each U e (Bn. Let 6 = |jen. We assert that e is the required locally finite refinement of Ct, covering X. Let x be a point of X. We wish to prove that x lies in an element of e, and that x has a neighborhood intersecting only finitely many elements of e. Consider the covering (B = |J (B,,; let N be the smallest integer such that x lies in an element of CBA.. Let U be an element of ($>N containing x. First, note that since .v lies in no element of (B, for / < N, the point x lies in the element SN(U) of 6. Second, note that since each collection <&„ is locally finite, we can choose for each /; = 1, . . . , N a neighborhood Wn of x that intersects only finitely many elements of (Bn. Now if W'„ intersects the element Sn{V) of Qn, it must intersect the element Kof ($>„, since Sn(V) c: K Therefore, ^intersects only finitely many elements of 6,. Furthermore, because U is in ($>N, U intersects no element of £„ for /; > N. As a result, the neighborhood w, n w» n • • • n W* n tf of .v intersects only finitely many elements of 6. B) => C). Let G be an open covering of X. Let (B be the collection of all open sets U of X such that 0 is contained in an element of a. By regularity, <B covers X. Using B), we can find a refinement e of (B that covers X and is locally finite. Let » = {C| C e e}. Then » also covers A'; it is locally finite by Lemma 1.1; and it refines a. C) =» D). Let a be an open covering of X. Using C), choose & to be a refinement of « that covers X and is locally finite. (We can take (B to be a closed refinement if we like, but that is irrelevant.) We seek to expand each element B of (B slightly to an open set, making the expansion slight enough that the resulting collection of open sets will be still be locally finite and will still refine a.
Metrization Theorems and Paracompactness Chap. 6 This step involves a new trick. The previous trick, used several times, con sisted of ordering the sets in some way and forming a new set by subtractine off all the previous ones.t That trick shrinks the sets; to expand them we need something different. We shall introduce an auxiliary locally finite closed covering e of X and use it to expand the elements of (B. For each point x of X, there is a neighborhood of x that intersects only finitely many elements of (B, The collection of all open sets that intersect only finitely many elements of (B is thus an open covering of X. Using C) again let e be a closed refinement of this covering that covers A'and is locally finite' Each element of e intersects only finitely many elements of (B. For each element B of (B, let eE) = [C\ C e e and C c X - B}, Then define E(B) = X — (Jceew C, Because C is a locally finite collection of closed sets, the union of the elements of any subcollection of e is closed, by Lemma I.I. Therefore, the set E(B) is an open set. Furthermore, E(B) =d B by definition. (See Figure 3, in which Figure 3 the elements of (B are represented as closed circular regions and line segments, and the elements of C are represented as closed square regions.) Now we may have expanded each B too much; the collection {E(B)j may not be a refinement of a. This is easily remedied. For each B e <&, choose an tSee Lemma 3,1 and the proof A) => B) just given.
Paracompactness 259 element F(B) of « containing 5. Then define V = {E(B) n F(B)\ B e <&}. The collection » is a refinement of «. Because S c (E(B) n F(S)) and (B covers X, the collection 3D also covers X. We have finally to prove that 2D is locally finite. Given a point x of X, choose a neighborhood Wof x that intersects only finitely many elements of C say C,, .. ., Q. Then W <= C, U • • • u Ct, because e is a covering of X Now if an element C of e intersects the set E(B) n F(B), it cannot lie in X— B (by definition of £E)); thus C must intersect B. Because the element C of e intersects only finitely many elements B of (B, it intersects at most the same number of elements of the collection 2D = {£(£) n F{B)}. Then since each C, intersects only finitely many elements of 2D, so does the set W. □ 1. Give an example to show that if X is paracompact, it does not follow that for every open covering & of X, there is a locally finite subcollection of G that covers A". 2. (a) Show that every regular Lindelof space is paracompact, (b) Conclude that Ri is paracompact. 3. (a) Show that the product of a paracompact space and a compact Hausdorff space is paracompact. [Him: Use the tube lemma of §3-5.] (b) Conclude that Sa is not paracompact. (c) Show that Sn x [0, 1) in the dictionary order is not paracompact. [Him: Sa is homeomorphic to the closed subspace Sa x 0.] 4. Let A" be a Hausdorff space. Suppose there exists a countable open covering {!/„} of X such that for each n, Un is compact and U, <= l/n+t. Show that A"is paracompact. [Him: See Example 1.] 5. Let A" be a regular space. Show that if there is a countable covering [V.) of A" by open sets whose closures are paracompact, then X is paracompact. 6. Is every locally compact Hausdorff space paracompact? 7. One has a "shrinking lemma" for point-finite open coverings of a normal space; see the exercises of §4-5. Here is an "expansion lemma" for locally finite indexed families in a paracompact space: Lemma. Lei [A.Uj be a locally finite indexed family of subsets of the para- paracompact space X. Then there is a locally finite indexed family [V.],ej of open sets such that A, <= U*for each a e J. 8. Determine which of the following spaces are paracompact in the dictionary order topology: (a) [0, 1] x [0, 1] (b) [0, 1) x [0, 1] *(c) [0, 1] x [0, 1)
260 Metrization Theorems and Paracompactness ' _ *9. Let Let G be a locally compact, connected, Hausdorff topological grou that G is paracompact. [Hint: Choose a neighborhood £/, of e having en closure. In general, define £/„+, = £?„• £/,. Then G = IJf?,.] "^Wct *10. Show that if G is a paracompact topological group and H is a compact k. group, then G/tf is paracompact. [Hint: Show that />: G —-► G/# is a 1 map. Then show that if (B is a locally finite closed covering of G, the collect [p(B) 15 e (B) is a locally finite closed covering of G/H.] ' '°n 6-5 The Smirnov Metrization Theorem The Nagata-Smirnov metrization theorem gives one set of necessary and sufficient conditions for metrizability of a space. In this section we give anoth- another such metrization theorem. It is a corollary of the Nagata-Smirnov theorem and was first proved by Smirnov. Definition- A space X is locally metrizable if every point x of X has a neighborhood U that is metrizable in the subspace topology. Theorem 5.1 {Smirnov metrization theorem). A space X is metrizable if and only if it is paracompact and locally metrizable. Proof. Suppose that X is metrizable. Then X is locally metrizable; it is also paracompact, by Stone's theorem. Conversely, suppose that X is paracompact and locally metrizable. We shall show that X has a basis ffi that is countably locally finite. Since Jis regular (being paracompact), it will then follow from the Nagata-Smirnov theorem that X is metrizable. The proof is an adaptation of the proof of Theorem 3.2. Cover X by open sets that are metrizable; then choose a locally finite open refinement 6 of this covering that covers X. Each element C of e is metrizable; let the function dc : C x C —» /? be a metric that gives the topology of C. Given x e C.Iet 5c(.v, e) denote the set of all points y of C such that dc(x, y) < e- Being open in C, the set Bc(x, e) will also be open in X. Given m e Z+, let <Xm be the covering of X by all open balls of radius \/m in sets C of e; am = {Bc(x, \/m)\C e e and x e C}. Let Dm be a locally finite open refinement of am that covers X. (Here we use paracompactness.) Let D = \JDm; then D is countably locally finite. We assert that D is a basis for X; our theo- theorem follows.
S fi 5 The Smimov Metrlzation Theorem 261 Let x be a point of JTand let V be a neighborhood of x We seek to fi^ an element D of j) such that x e /> e tf. Now x belongs! S finiS many elements of e, say to C,,... Ct Then r/nr;c ,, y x in the set C,, so there is an I > 0 su"h that ' * ne'ghborhood * BCi{x, e,) <=(Un c,). Choose m so that l/« < * min {f, fJ. Because the coIIection ^ covers X, there must be an element D of »" containing x. Because D™ refines ft", there must be an element Bc(y, \jm) of a™, for some C e e and some y e C, that contains D. Then x belongs to C, so that C must be one of the sets Ct,...,Ck. Say that C = C,. Then, using the triangle inequality, we have ■° c 5C,(.V, 1/m) c Sc,(a-, f() <= [/, as desired. Q Example 1. We show that the space Sa is not paracompact. One proof of this fact appeared in the preceding set of exercises; here is another. Note first that Sa is locally metrizable: Let x e Sn; choose y so that x < y < Cl. The open set Sy has a countable basis; indeed, the collection of all intervals with end points in Sy is countable. It follows from the Urysohn metrization theorem that Sy is metrizable. If Sa were paracompact, it would follow from the Smirnov theorem that Sn is metrizable. But we know that Sa is not metrizable, for it is limit point compact but not compact (see Example 1 of §3-7). Exercises 1. Compare Theorem 5.1 with Exercises 5 and 6 of §4-4. 2. Show that every connected locally compact metric space has a countable basis. [Hint: Let e be a locally finite covering of X by open sets having compact closures. Let £/, be a nonempty element of C and ingeneral let £/„+, bethe union of all elements of e that intersect fjr Show U. is compact and X- 3. A more general notion of manifold than that defined in §4-5 is the J A space X is called a paracompact-manifold if it is ^ Para^mpad pace and there is an integer m such that each point of A"has a neighborhood that is home morphic to an open subset of Rm. (a) Show that a paracompact-manifold is metrizable. H (b) Show that if X is a paracompact-manifold having only countably many components, then Jfis a manifold; and conversely.
7. Complete Metric Spaces and Function Spaces The concept of completeness for a metric space is one you may have seen already. It is basic for all aspects of analysis. Although completeness is a metric property rather than a topological one, there are a number of theorems involving complete metric spaces that are topological in character. In this chapter, we shall study the most important examples of complete metric spaces and shall prove some of these theorems. The most familiar example of a complete metric space is euclidean space in either of its usual metrics. Another example, just as important, is the set Q(X, Y) of all continuous functions mapping a space X into a space Y. This set has a metric called the uniform metric, analogous to the uniform metric defined for RJ in §2-9. If Y is a complete metric space, then 6(J, Y) is complete in the uniform metric. This we demonstrate in §7-1. As an ap- application, we construct in §7-2 the well-known Peano space-filling <^rye' One theorem of topological character concerning complete metric spaces is a theorem relating compactness of a space to completeness. We prove i in §7-3. An immediate corollary is a theorem concerning compact sets in * function space e(X, /?"); it is the classical version of a famous theorem call Ascoli's theorem. ..^ There are other useful topologies on the function space e(A", Y) oes\ the one derived from the uniform metric. We study some of them in S and §7-5. We also prove a general version of Ascoli's theorem, in S - ■ 262
§7-1 Complete Metric Spaces 263 Another theorem of topological character concerning complete metric spaces is a theorem we prove in §7-7, to the effect that every complete metric space belongs to the class of topological spaces called the Baire spaces. The defining condition for a Baire space is a bit complicated to state, but it is often useful in the applications. One application is the proof we give in §7-8 of the existence of a continuous nowhere-differentiable real-valued function. Another application arises in that branch of topology called dimension theory. In §7-9 we define a topological notion of dimension (due to Lebesgue), and prove the classic theorem that every compact metrizable space of to- topological dimension m can be imbedded in euclidean space R" of dimension ft = 2m + I. This theorem generalizes the imbedding theorem for mani- manifolds proved in §4-5. Some of the sections of this chapter are independent of one another. The dependence among them is expressed in the following diagram: §7-1 Complete metric spaces A space-filling curve §7-3 Compactness in metric spaces Pointwvse and compact convergence §7-5 j The compact-open topology §7-6 Ascoli's theorem §7-7 Baire spaces \\7-8 A nowhere-differentiable function V-9 An introduction to dimension theory Throughout the chapter, we assume §3-8, study dimension theory, we shall make use of as well as a bit of linear algebra. 7-1 Complete Metric Spaces In this section we define the notionof completeness,«ci show that* ¥ is a complete metric space, then the funct.on P»«^^ imbedded in the uniform metric. We also show that every metnc space isometrically in a complete metric space.
264 Complete Metric Spaces and Function Spaces Chap. 7 Definition. Let (X, d) be a metric space. A sequence (x ) of nnint* t is said to be a Cauchy sequence in (X, d) if it has the property th.t ■ f > 0, there is an integer N such that Y ^ 4*« xm) < e whenever n,m>N. The metric space {X, d) is said to be complete if every Cauchy sequence • A'converges. m Any convergent sequence in A'is necessarily a Cauchy sequence, of course- completeness requires that the converse hold. ' Note that a closed subset A of a complete metric space (X, d) is necessarily complete in the restricted metric. For a Cauchy sequence in A is also a Cauchy sequence in X, whence it converges in X. Because A is a closed subset of X, the limit must lie in A. Note also that if X is complete under the metric d, then X is complete under the standard bounded metric d(x, y) = min {d(x, y), 1} corresponding to d, and conversely. For a sequence (x,) is a Cauchy sequence under d if and only if it is a Cauchy sequence under d. And a sequence converges under d if and only if it converges under d. A useful criterion for a metric space to be complete is the following: Lemma I.I. A metric space X is complete if every Cauchy sequence in X has a convergent subsequence- Proof. Let (>„) be a Cauchy sequence in (X, d). We show that if (xj has a subsequence (*„,) that converges to a point x, then the sequence (xj itself converges to x. Given e > 0, first choose N large enough that d(xn, xm) < 5/2 for all n,m>N [using the fact that (x.) is a Cauchy sequence]. Then choose an integer / large enough that n,>N and <<*,.., x) < 5/2 ^d [using the fact that n, < n2 < . .. is an increasing sequence of ln.te^^u]t x_, converges to x]. Putting these facts together, we have the desired that for n > N, d(xn, x) < d(xm xj + <Kxm, x)<e. □ Theorem 1.2. Euclidean space R" /* complete in either of its usual metric , the euclidean metric d or the square metric p. Cauchy Proof. To show the metric space (Rk, p) is complete, let W^ * tf sequence in (**. /»• Then the set {*„} is a bounded subset of (* , P>
S7 Complete Metric Spaces 265 we choose N so that />(*„> *„) <, 1 for all n, m ;> iV, then the number M = max {/>(*„ 0) /?(*„_„ 0), />(*„, 0) + 1} is an upper bound for p(xm 0). Thus the points of the sequence (xn) all lie in the cube [—M, M]k. Since this cube is compact, the sequence (xn) has a convergent subsequence, by Theorem 7.4 of Chapter 3. Then (Rk, p)\s com- complete, by Lemma 1.1. To show that (i?*, d) is complete, note that a sequence is a Cauchy se- sequence relative to d if and only if it is a Cauchy sequence relative to p, and a sequence converges relative to d if and only if it converges relative to p. Q Example 1. An example of a noncomplete metric space is the space Q of rational numbers in the usual metric d(x,y) = \x—y\. For instance, the sequence 1.4, 1.41, 1.414, 1.4142,1.41421,... of finite decimals converging (in R) to ^/~T is a Cauchy sequence in Q that does not converge (in Q). Example 2. Another noncomplete space is the open interval ( — 1,1) in R, in the metric d(x, y) =■ \ x — y \. In this space the sequence (*„) defined by x, = 1 - 1/" is a Cauchy sequence that does not converge. This example shows that com- completeness is not a topological property, that is, it is not preserved by homeomor- phisms. For (-1, 1) is homeomorphic to the real line R, and R is complete in its usual metric. Example 3. There is a metric for the product topology on R" relative to which R°" is complete. Consider the metric D(\, y) = lub [d{x,, y,)/i] on R°>, where d(a, b) = min [| a - b \, 1}. We showed in Theorem 9.5 of Chapter 2 that D is a metric for the product space /f. To show R" complete under D, let (x_) be a Cauchy sequence under D. Let / be fixed for the moment. Let n,: Rm —► R be projection on the /th coordinate. Since dGt,(x), n,(y)) <, iD(x, y), the sequence Gt((xn)) is a Cauchy sequence in (R, d). Therefore, it converges, say to at. . Let a be the point (a^z. of R°>- We assert that x, — a. For a typical basis element containing a is of the form TJU,, where V, = R for 12; m. <~a°°s* N, large enough that 7t,(x.) e V, for n 2; N,. Let N = max {M, • • •. /v»J> then x, e HUt for n ^ M
266 Complete Metric Spaces and Function Spaces Chap. 7 Although both the product spaces R" and Ra have metrics relaf which they are complete, one cannot hope to prove the same result f"* ** product space RJ in general, because RJ is not even metrizableif Jisun"^ table (see §2-10). There is, however, a different topology on the set JT one given by the uniform metric. Relative to this metric, R1 is complet C we shall see. e>as We define the uniform metric for Y' in general as follows: Definition. Let (Y, d) be a metric space. Let d(a, b) = min {d(a,b\ \\ be the standard bounded metric on Y corresponding to d. Given an'index set J, define a metric on the set YJ of all functions/../—, y by the rule p(f, g) = lub W/(a), g(z))\ a e J}. The function p is called the uniform metric on Y1 corresponding to the metric d on Y; it is easy to check that it is a metric. Note that here we have used functional notation for the elements of Y' rather than "tuple" notation. We shall follow this practice thoughout the chapter. Theorem 1.3. If the space Y is complete in the metric d, then the space Y' is complete in the uniform metric p corresponding to d. Proof. Recall that if (Y, d) is complete, so is (Y, d), whererfis the bounded metric corresponding to d. Now suppose that/,,/2, ... is a sequence of points of YJ that is a Cauchy sequence relative to p. Given a in J, the fact that <?(/,(«). J"»,(«))< PiLJJ for all n, m means that the sequence/,(a),/2(a),... is a Cauchy sequence in (K, d). Hence this sequence converges, say to a point y,. Let/:/—* Ybe the function defined by /(a) = ya. We assert that the sequence (/„) converges to/in the metric p. Given e > 0, first choose N large enough that pTf.JJ < f/2 whenever n, m > N. Then, in particular, </(/„(«)> /.(«)) < f/2 .f for n, m > N and a e X Letting n and a be fixed and m become arbitrany large, we see that </(/„(«), /(«)) < e/2. This inequality holds for any a in J, provided merely that n ^ N. There p(Uf)<e/2<e for n>N, as desired. □ Now let us specialize somewhat, and consider the set Y* wheIJffect on topological space rather than merely a set. Of course, this has n
§7"J Complete Metric Spaces 267 what has gone before; the topology of X is irrelevant when considering the set of all functions /: X—► Y. But suppose that we consider the subset e(X, Y) of Yx consisting of all continuous functions/: X —> Y. It turns out that if Y is complete, this subset is also complete in the uniform metric. Theorem 1.4. Let X be a topological space and let (Y, d) be a metric space. The set C(X, Y) of continuous functions is closed in Yx under the uniform metric. Therefore, ifY is complete, e(X, Y) is complete in the uniform metric. Proof. This theorem is just the uniform limit theorem (Theorem 10.6 of Chapter 2) in a new guise. First, we show that if a sequence of elements /„ of Yx converges to the element / of Yx relative to the metric p on Yx, then it converges to /uniformly in the sense defined in §2-10, relative to the metric d on Y. Given e > 0, choose an integer N such that P(f, /,) < e for all n > N. Then for all x e X and all n > N, kfJLx), f(x)) < p{fn, /) < e. Thus (/„) converges uniformly to/ Now we show that Q(X, Y) is closed in Yx relative to the metric p. Let /be an element of Yx that is a limit point of <2(X, Y). Then there is a sequence (/J of elements of G(X, V) converging to/in the metric p. By the uniform limit theorem,/is continuous, so that/e Q(X, Y). Since Yx is complete under p, so is the closed subset e(X, Y). □ Definition. Let (V, d) be a metric space; let Xbt either a set or a topolog- topological space. Suppose that 5 is a subset of Yx having the property that for each pair/ g of elements of ff, the set {d(f(x),g(.x))\x e X] is bounded. We define a metric p on 3 by the formula />(/, g) = lub {</(/(*), g(.x))\x e X}; it is called the square metric or, more commonly, the sup (supremum) metric. How does this metric compare with the uniform metric p defined above? It is an easy exercise to show that for/ g e 3, p(f,g) = min{p(f,g), 1}- This means that on a set 3 where both p and p are defined, p is just the standard bounded metric derived from p. (This is the reason we originally introduced the notation p for the uniform metric.) It follows that p and p give the same topology on 3. And it follows that 3 is complete under p it and only if it is complete under p. Therefore, for all practical purposes, they are equivalent metrics. _ Many authors do not bother to define the metric p at all, but simply
268 Complete Metric Spaces and Function Spaces _ Chap. 7 limit themselves to spaces 3= where p is defined. We prefer to deal with rh more general case and specialize to the metric p whenever it is useful t a possible) to do so. ^ Example 4. Let (Y, d) be a metric space and let J be a finite index set Then the square metric p is defined on all of Y'. Of course, in this case all th various topologies for Y'—the box topology, the uniform topology, and the product topology—are the same. Example 5. Consider the space e(X, R) of all continuous real-valued functions on the compact space X. Each such function is bounded, so the metric is defined on this space. Indeed, the maximum value theorem tells us that max The space G(X, R) is complete under the metric p. We now prove a classical theorem, to the effect that every metric space can be imbedded isometrically in a complete metric space. (A different proof, somewhat more direct, is outlined in Exercise 8.) Although we shall not need this theorem, it is useful in other parts of mathematics. First, a lemma: Lemma 1.5. Let X be a topological space. The set (B(X, R) of all bounded functions f: X —> R is complete under the sup metric p. Proof We show that <&{X, R) is closed in Rx in the uniform metric p. It follows that <&{X, R) is complete under p. Since the metric p is defined on this set, it is also complete under p. Let / be the limit of a sequence (/„) of elements of (S(X, R). Choose N so that p{fn, f) < 1 for n > N. Then since fN belongs to &(X, R), there is an M such that \fN{x)\ < M for all x e X. Then for all x e X, so that/is bounded. □ Theorem 1.6. Let (X, d) be a metric space. There is an isometric imbedding ofX into a complete metric space. Proof Let &{X, R) be the set of all bounded functions mapping R. Let x0 be a fixed point of X. Given a e X, define 4>a : X-» R equation 4>a(x) = d(x, a) — d(x, x0). We assert that <f>a is bounded. For It follows from the inequalities
** 7"! Complete Metric Spaces 269 that we have | <f>£x) \ < d(a, x0). Define <D : A' — (&(X, R) by setting «(«) = K We show that <D is an isometric imbedding of (X, d) into the complete metric space {&{X, R), p). That is, we show that for every pair of points a, b e X, we have />(&, <j>b) = d(a, b). By definition, <)-t,lx)\lx eX] = lub{\d(x, a) - d(x, b)\ ; x e X}. By the triangle inequality, \d(x,a)~d(.x,b)\<d(a,b), from which we conclude that p(<f>a- 0*) < d(a, b). On the other hand, this inequality cannot be strict, for when x = a, \d(.x,a)~d(.x,b)\ = d(a,b). □ Definition. Let X be a metric space. If /; : X —> Y is an isometric fmbed- ding of X into a complete metric space Y, then the subspace /i(A) of Y is a complete metric space. It is called the completion of A'. The completion of X is uniquely determined up to an isometry. See Ex- Exercise 9. Exercises 1. Let A" be a metric space. (a) Suppose that for some e > 0, every e-ball in A" has compact closure. Show that X is complete. (b) Suppose that for each x e X there is an e > 0 such that the ball B(x, e) has compact closure. Show by means of an example that A"need not be complete. 2. Let (X, dx) and (Y, dr) be metric spaces; let Y be complete. Let A <= X Show that if/: A -> Y is uniformly continuous, then/can be uniquely extended to a continuous function g : A —> Y, and g is uniformly continuous. 3. Consider the subset R" of R» consisting of sequences that are eventually zero. Is R" complete in the uniform metric ?
270 Complete Metric Spaces and Function Spaces Chap. 7 4. Show that the metric space (X, d) is complete if and only if for anv sequence Ax => Az => • • • of nonempty closed sets of X such- that d An —► 0, C\n,Z, An =£ 0- 5. Let (X, d) be a complete metric space. Recall that a continuous map/- * —» y is said to be a contraction if there is a number a < 1 such that ~* </(/(*),/O)):S <*</(*, ?) for all x,;y e X. Show that if/is a contraction, then there is a unique noint x of X such that /(*) = x 6. A space X is said to be topologically complete if there exists a metric for the topology of X relative to which X is complete. (a) Show that a closed subspace of a topologically complete space is topolog. ically complete. (b) Show that a countable product of topologically complete spaces is topo- topologically complete (in the product topology). (c) Show that an open subset of a topologically complete space is topologically complete. [Hint: If U <= X and X is complete under the metric d, define (j>: U —> Ji by the equation <f)(x) = l/d(x, X - U), where d(x, A) = gib {d(x, a)\a e A}. Then define/: U —► X x Rby the equation fix) = x x <j>(x).] (d) Show that if A is a Gs set in a topologically complete space, then A is topo- topologically complete. [Hint: Let A be the intersection of the open sets V,, for n e Z+. Consider the diagonal imbedding f(a) = (a, a,...) of A into IU/.0 (e) Show that the irrationals are topologically complete. 7. Show that C2 is complete in the £2-metric (see Exercise 9 of §2-9). 8. Let (X, d)bea metric space. Show that there is an isometric imbedding h oiX into a complete metric space (y, D), as follows: Let ;? denote the set ot au Cauchy sequences x = (xu x2,...) of points of X. Define x ~ y if d(x,,y,)-^ 0. Let [x] denote the equivalence class of x; and let Y denote the set of these equivalence classes. Define a metric D on Y by the equation £([x],[y])= Jim </(*„,>-„). (a) Show that ~ is an equivalence relation, and show that D is a well metric. f ,h constan' (b) Define h:X-^Y by letting AW be the equivalence class ot the sequence (x, x,...);
§7-2 A Space-Filling Curve 271 Show that h is an isometric imbedding (c) Show that h(X) is dense in Y; indeed, given x = (x the sequence A(x.) of points of y converges to the U mtricspac^ sequence in A converges in 2T, then Z is complete (e) Show that (Y, D) is complete. ' " f v adif r 9. Theorem (Uniqueness of the completion). Let h- X — Y and h' ■ Y v be isometric imbeddings of the metric space (X d) in the comnhtp „,„,', (£*> "rfCr.D-), respectively. Then /*« * « SCSffS (h'(X), D') that equals h'h~« on the snbspace h(X). }> ] 'h 7-2 A Space-Filling Curve As an application of the completeness of the metric space Q(X, Y) in the uniform metric when Y is complete, we shall construct the famous "Peano space-filling curve." Theorem 2.1. Let I = [0, 1]. There exists a continuous map f :I—► I2 whose image fills up the entire square P. The existence of this path violates one's naive geometric intuition in much the same way as does the existence of the continuous nowhere-differentiable function (which we shall come to later). Proof. Step 1. We shall construct the map/as the limit of a sequence of continuous functions/„. First we describe a particular operation on paths which will be used to generate the sequence/,. Begin with an arbitrary closed interval [a, b] in the real line and an arbi- arbitrary square in the plane with sides parallel to the coordinate axes, and con- consider the triangular path g pictured in Figure 1. It is a continuous map of la, b] into the square. The operation we wish to describe replaces the path g by the path g' pictured in Figure 2. It is made up of four triangular paths, each half the size of g. Note that g and g' have the same initial point and the same final point. You can write the equations for g and g' if you like. Figure I
272 Complete Metric Spaces and Function Spaces Chap. 7 Figure 2 This same operation can also be applied to any triangular path conn,* ing two adjacent corners of the square. For instance, when applied to Z path h pictured in Figure 3, it gives the path h'. Figure 3 Step 2. Now we define a sequence of functions/, : /—» P. The first func- function, which we label /„ for convenience, is the triangular path pictured in Figure 1, letting a == 0 and 6 = 1. The next function/, is the function ob- obtained by applying the operation described in Step 1 to the function/0; it is pictured in Figure 2. The next function/2 is the function obtained by applying this same operation to each of the four triangular paths that make up/,. It is pictured in Figure 4. The next function/ is obtained by applying the operation to each of the 16 triangular paths that make up/2; it is pictured Figure 4
§7-2 A Space-Filling Curve 273 mlililiniiillnllli.lnilinlnil Figure 5 V /\ in Figure 5. And so on. At the general step,/, is a path made up of 4° trian- triangular paths of the type considered in Step 1, each lying in a square of edge length 1/2". The function/n+1 is obtained by applying the operation of Step I to these triangular paths, replacing each one by four smaller triangular paths. Step 3. For purposes of this proof, let d(x, y) denote the square metric onR1, cl(x, y) = max {| x, -yl\,\x1-y11}. Then we can let p denote the corresponding sup metric on S(/, P); p(f,g) = iub{d(f(t),g(t))\t e 1} Since P is closed in R2, it is complete in the square metric; then 6G, P) is complete in the metric p. We assert that the sequence of functions (/„) defined in Step 2 is a Cauchy sequence under p. To prove this fact, let us examine what happens when we pass from/, to/n+1. Each small triangular path in/n lies in a square of edge length 1/2". The operation by which we obtain /„+, replaces this triangular path by four triangular paths that lie in the same square. Therefore, in the square metric on I2, the distance between /„(/) and /„+,(/) is at most 1/2". As a result, />(/„, /n+1) < 1/2". It follows that (/„) is a Cauchy sequence, since Pi/., /.«) < 1/2" + l/2"+1 + • • • + 1/2"—1 < 2/2" for all n and m. Step 4. Because 6G, P) is complete, the sequence /„ converges to a continuous function/: I—- P. We prove that/is surjective. Let x be a point of P; we show that x belongs to /(/). First we note that, given n, the path/ comes within a distance of 1/2" of the point x. For the
274 Complete Metric Spaces and Function Spaces Chap. 7 path/, touches each of the little squares of edge length l/2» into wV >. have divided P. S ' ° whic}» we Using this fact, we shall prove that, given e > 0, the f-neiehborh a x intersects /(/). Choose N large enough that of />(/*> /) < f/2 and 1/2" < f/2. By the result of the preceding paragraph, there is a point /„ e /such «... rf(*,/v('o)) < 1/2*. Then since d{fN{t), /(/)) < e/2 for all /, it follows t£ rf(x> /Co)) < f, so the f-neighborhood of x intersects /(/). It follows that x belongs to the closure of /(/). But / is compact, so /(/) is compact and is therefore closed. Hence x belongs to /(/), as desired, n Exercises 1. Given », show there is a continuous surjective map g : I —> /". [Hint: Consider fxf:lx I— >/2 x IK] 2. Is there is continuous surjective map/: /—► K21 3. Show there is a continuous surjective map/: 7? —> R". 4. Is there a continuous surjective map/: 7? —> X if A" is the subspace R" of /P° consisting of sequences that are eventually zero? What if X— /?"? *5. (a) Let A" be a Hausdorff space. Show that if there is a continuous surjective mapping /: / —> X, then X is compact, connected, locally connected, and metrizable. (b) The converse also holds; it is a famous theorem of point set topology called the Hahu-Mcmirkiewicz theorem (see [H-Y], p. 129). Assuming this theorem, show there is a continuous surjective map /: / —> /". A Hausdorff space that is the continuous image of the closed unit interval is often called a Peano space. 7-3 Compactness in Metric Spaces We have already shown that compactness, limit point compactness, an^ sequential compactness are equivalent for metric spaces. There isstil ^.^ formulation of compactness for metric spaces, one that involves the no^ of completeness. We study it in this section. As an application, w prove a theorem characterizing those subsets of e(X, *) which arecomp in the uniform metric; it is the classical version of Ascoli's the°rel"- f J? How is compactness of a metric space X related to comPle.te"!cessarlly It follows from Lemma 1.1 that every compact metric space is
§7'3 ■■■-". Compactness in Metric Spaces 275 complete. The converse does not hold_a complete metric space need not be compact It ,s reasonable to ask what extra condition one needs to impose on a comp etc space to be assured of its compactness. Such a condition is the one called total boundedness. Definition. A metric space (X, d) is said to be totally bounded if for everv f > 0, there is a finite covering of X by f-balls. Total boundedness of a metric space clearly implies boundedness, but the converse does not hold. Example 1. Under the metric d(x, y) = \x-y\, the real line R is neither bounded nor totally bounded. Under the metric d(x,y) = min[|;r — y\ ]} the real line is bounded but not totally bounded. Example 2. Under the metric |x - y\, the subspace (-1, 1) of R is totally bounded, and so is the subspace Qn[-l,lJ. Neither of these spaces is com- complete, however. The subspace [—1, 1] is both complete and totally bounded. Theorem 3.1. A metric space (X, d) is compact if and only if it is complete and totally bounded. Proof. If A" is a compact metric space, then X is automatically complete, as noted above. The fact that X is totally bounded is a consequence of the fact that the covering of X by all open e-balls must contain a finite subcover- ing. Conversely, let X be complete and totally bounded. We shall prove that A" is sequentially compact. This will suffice. Let (*„) be a sequence of points of X. We shall construct a subsequence of (*„) that is a Cauchy sequence, so that it necessarily converges. First cover X by finitely many balls of radius I. At least one of these balls, say 51( contains x. for infinitely many values of n. Let/, be the subset ofZ+ consisting of those indices i> for which xn e B,. Next, cover X by finitely many balls of radius {. Because /, is infinite, at least one of these balls, say B2, must contain x, for infinitely many values of n in /,. Choose J2 to be the set of those indices n for which n e/, and xn e B2. In general, given an infinite set Jk of positive integers, choose Jk+l to be an infinite subset of Jk such that there is a ball Bk+l of radius l/(fc + 1) which contains *„ for all n e Jk+i. Choose n, e/,. Given nk, choose nk+l e Jk+l such that nk+l>nk; this we can do because Jk+l is an infinite set. Now for i,j >. k, the indices - », and „, both belong to Jk (because /, => Jt => ■ - ■ is a nested sequence o sets). Therefore, for all /, / >. k, the points xm and xv are contained in a ball Bk of radius l/k. It follows that the sequence (*„,) is a Cauchy sequence, as desired. □
276 Complete Metric Spaces and Function Spaces Chap. 7 Now we apply this result to the function space e(X R«) in th topology. We are going to be concerned only about the case whereV" "* pact, so we shall use the metric s COm- />(/, g) = max {d(f(x), g(x))} throughout. [Here d may denote either the euclidean metric or the so metric on R".] q are We have proved earlier that a subset of (R", d) is compact if and onlv if it is closed and bounded. One might hope that the corresponding theorem is true for (fi(X, R"), p). But it is not. One needs to assume that the subset of e(X, R") satisfies an additional condition, called equicontinuity. It is defined as follows: Definition. Let (Y, d) be a metric space. Let 5 be a subset of the function space G(X, Y). If ,v0 e X, the set 5 of functions is said to be equicontinuous at x0 if given e > 0, there is a neighborhood U of x0 such that for all x e U and all/ e 5, d(f(x\ /(*„)) < e. If the set 5 is equicontinuous at ,y0 for each x0 e X, it is said simply to be equicontinuous. Continuity of the function/at x0 means that given/and given e > 0, there exists a neighborhood U of x0 such that d(f(x), f(x0)) < e for x e U. Equicontinuity of 5 means that a single neighborhood [/can be chosen that will work for all the functions/in the collection '5. Equicontinuity has the following interesting interpretation when both X and Y are compact: Lemma 3.2. Let X be a compact space; let (Y, d) be a compact metric space. Let 5 be a subset of 6(X, Y). Then 5 is equicontinuous if and only if 'S is totally bounded in the sup metric p. Proof Suppose that SF is totally bounded under p. Given x0, we shoJ is equicontinuous at jr0. Let e > 0 be given. Choose positive numbe fl and e2 so that 2e, + e2 < e. Cover ff by finitely many open f,-e<" Each function /, is continuous; therefore, we can choose a neighbor U of x0 such that for x e U and i = 1,...,«, d{f,{x\flxo))<e,. We assert that if x e U and / e 5, then rf(/W. /W) < f'" thlS equicontinuity. ,..,„,. *„ at least one of the Let/be an arbitrary element of ff. Then/belongs to at leas above erballs, say to 5,(/,, f ,)• Then
§7 ComPa<:»>ess in Metric Spaces 277 ),/,(*))< fl> < e2, The first and third inequalities hold because/is in £,(/„ e,), and the second holds because x e U. It follows from the triangle inequality that for all x e £/, we have d{f(x), f(x0)) < e, as desired. Conversely, suppose that <F is equicontinuous. Let f > 0 be given. We wish to cover SF by finitely many open e-balls. Choose f, and e2 so that2e, + f2 < f. Using equicontinuity of <F and compactness of A", cover A" by finitely many open sets Vi,...,Uk, containing points xu . . . , xk, respectively, such that difix), fix,)) < e, for x e U, and all / e 5. Cover Y by finitely many open sets Vu...,Vm of diameter less than e2. Let J be the collection of all functions a : {I,.. ., k} —> {I,..., tn}. Given a e J, if there exists a function/of 5 such that /(>,) e K«(o for each / = I,. . . , k, choose one such function and label it/,. The collection {/„} is indexed by a subset/'of the set 7 and is thus finite. We assert that the open balls Bf(f., e), for a e J', cover SF. Let /be an element of 3\ For each / = I,. . ., k, choose an integer a@ such that f(x,) e V^»- Then the function a is in/'. We assert that/belongs to the ball £,(/,,, e). Let x be a point of A". Choose / so that x e U,. Then (/.(,),/.W)<fi- The first and third inequalities hold because x e t/<( and the second holds because fix,) and /„(*,) are in V.w. We conclude that d(f(x), f.(x)) < e. Since this inequality holds for every x e A", «) = max Thus / belongs to Bpif,, e), as asserted. □ Now we prove the classical version of Ascoli's theorem. A more general version, whose proof does not depend on this one, appears in §7-6. Theorem 3.3 (Ascoli's theorem, classical version). Let X be a<c°™Pact space; consider C(X, R") in the sup metric p. A subset 5 oj 6(A, K ; is compact if and only if it is closed, bounded, and equicontinuous. Proof. Step 1. We first show that if ff is bounded under p, then there is a compact subset Y of R" having the property that /(x) e Y tor au/ t
278 Complete Metric Spaces and Function Spaces Chap. 7 and all x e X. It then follows that 3= is contained in the subsnace Wv, ofecarU") P«*e(*,y) Choose an element /„ e 3\ Because 3= is bounded under p the • number M such that />(/„, f)<M for all/ e 3=. Because X is compact? is MX); therefore, we can choose a number N so that/0(Z) lies m the\ °w 5,@, N) in iJ». Then f{X) lies in the baU 5,@, M+ JV), for every/6 5 Choose y to be the closure of this ball; then /(x) e y for every x e Y a every/ e 3\ Step 2. Suppose that 3= is compact. Then 3= is closed and bounded under p. By Step 1, 3= is contained in some subspace e(X, Y) of Q{X, R"), where y is compact. Since $ is compact, 31 is totally bounded under p, by Theorem 3.1. Because both X and Y are compact, Lemma 3.2 applies to show that J is equicontinuous. 5. Suppose that 'S is closed and bounded and equicontinuous. Since <S is a closed subset of the complete metric space (S(X, R"), p), it is necessarily complete. Because $ is bounded, by Step 1 it is contained in some subspace S(A", Y) of Q{X, R"), where Y is compact. Then since 'S is equicontinuous, and X and Y are compact, Lemma 3.2 implies that $ is totally bounded. It follows from Theorem 3.1 that tF is compact. Q Exercises 1. Let (A", rf) be a metric space. (a) Show that if A" is totally bounded, then A" is bounded. (b) Show that X is totally bounded under d if and only if it is totally bounded under d(x, y) = min [d(x, y), 1 j. 2. Let A be a su bset of the complete metric space X. Show that A is totally bounded if and only if A is compact. 3. Consider a countable product X = TR« of metrhable spaces; use on metric Z> that gives the product topology. (See Exercise 3 of §2-10.) Show ma if each space A, is totally bounded, so is X. Conclude without using the Tyclio- noff theorem that a countable product of compact metrizable spaces is compa 4. (a) Show that any finite set of continuous functions is equicontinuous. ^ (b) Suppose that 3= is a collection of differentiate functions /: P, >J ^ whose derivatives are uniformly bounded. [This means there is such that | f'(x) \ <, M for all /in $ and all *.] Show that 3= is eq"ic°n«" (c) Let /„: [0, 1] — R be the function /„(*) = *"; let 3= be the ~"«°" "^ Show that 3= is closed and bounded but has no limit point in M.'. what point or points does 3= fail to be equicontinuous? 5. Let A" be compact; give 6(X, R") the sup metric p. We say that " e(X, R") is uniformly bounded if it is bounded under p. We say it
8 '■■"' ComPactness in Metric Spaces 279 bounded if for each x e X, the set ** = {/(*)!/ 6 3=} is a bounded subset of R". (a) Show that if 3= is equicontinuous and pointwise bounded, it is uniformly bounded. J (b) Show that if 3= is equicontinuous, so is its closure. (c) Theorem. Let X be compact; consider e(X, R») in the metric p. A subset <5 ofe(X, R") has compact closure if and only if it is equicontinuous andpointwise bounded. (d) Theorem (Arzela's theorem). Let X be compact; letfn e e(X, Rk). If the col- collection {/„} is pointwise bounded ond equicontinuous, then the sequence (fy has a uniformly convergent subsequence. 6. Let A" be locally compact HausdorfF. A subset 3= of e(X, R) is said to vanish uniformly at infinity if given e > 0, there is a compact subset C of A"such that | f(x) | < e for x e X - C and / e 3\ If 3= consists of a single function/, we say simply that / vanishes at infinity. Let G0(X, R) denote the set of continuous functions/: X—> R that vanish at infinity. The sup metric p is defined on Ct(X, R). Theorem. A subset $ of(Q0(X, R), p) is compact if and only if it is closed, bounded, equicontiituous, and vanishes uniformly at infinity. [Hint: Let Y denote the one-point compactification of X. Show that Qo(X, R) is isometric with a subspace of Q(Y, R).] *7. Let (X, d) be a metric space. If A <= X and e > 0, define Let OC be the collection of all (nonempty) closed, bounded subsets of X. If A, B e 3C, define D(A,B) = gib [e\ A c U(B,€) and B <= U(A,€)}. (a) Show that D is a metric on OC; it is called the Hausdorff metric. (b) Show that if (X, d) is complete, so is CC, D). [Hint: Let A. be a Cauchy sequence in 3C; by passing to a subsequence, assume D(A^ An+i) < 1/2". Define A to be the set of all points x that are the limits of sequences a-,,.v2, ... such that .v, e A, for each iandd(x,,xl+l) < 1/2'. ShowAn->A.} (c) Show that if (X, d) is totally bounded, so is CC, D). [Hint: Given e, choose 8 < e and let S be a finite subset of X such that the collection {Blx, S) | a- e S] covers A". Let G be the collection of all nonempty subsets of S\ show that [BB(A, e)\A e «} covers 3C.] (d) Ttorm. // A" * co/npoc/ ft rte metric d, then the spaced of all nonempty closed bounded subsets of X is compact in the Hausdorff metric V. 8. Which half of the proof of Lemma 3.2 uses compactness of Y1
280 Complete Metric Spaces and Function Spaces 7-4 Pointwise and Compact Convergence Chap. 7 There are other useful topologies on the spaces Yx and e^, Y) in add" tion to the uniform topology. We shall consider two of them here; they a'" called the topology of pointwise convergence and the topology of compact convergence. let Definition. Given a point x of the set Xand an open set U of the space Y S(x, U) = {f\feYx and /(x) e U}. The sets S(x, U) are a subbasis for topology on Yx, which is called the topology of pointwise convergence (or the point-open topology). The general basis element for this topology is a finite intersection of subbasis elements S(x, U). Thus a typical basis element about the function /consists of all functions g that are "close" to/at finitely many points. Such a neighborhood is illustrated in Figure 6; it consists of all functions g whose graphs intersect the three vertical intervals pictured. Figure 6 The topology of pointwise convergence on Yx is nothing new. It is J^ the product topology we have already studied. If we replace A"by/and o ^ the general element of / by a to make it look more familiar, hen 5(a, U) of all functions x :/ -> Y such that x(a) e U rs just the
§ 7-4 Pointwise and Compact Convergence 281 s;i(i/) of YJ, which is the standard subbasis element for the product to- topology. The reason for calling it the topology of pointwise convergence comes from the following theorem: Theorem 4.1. A sequence fn of functions converges to the function f in the topology ofpointwise convergence if and only if for each x in X, the sequence fB(x) of points of Y converges to the point f(x). Proof. This is a standard fact about the product topology; we prove it here using functional notation. Suppose that/, converges in the topology of pointwise convergence. Given x e X, and given an open set U about /(*), the set S(x, U) is a neighborhood of/ Therefore, there is an integer N such that/, e S(x, U) for all n > N. Then/,(x) e U for all n > N. Conversely, suppose/„(.*) converges to f(x) for each x. To show that/ converges to/in the topology of pointwise convergence, it suffices to show that if S(x, U) is an arbitrary subbasis element about/ then5(x, U) contains all/, for n sufficiently large. (Why?) But since/„(*) converges to f{x) and f{x) e U, there must be an integer N such that//*) e U for n > N. Then / e S(x, U) for n > jV. G Example 1. Consider the space R', where / = [0, 1]. The sequence (/„) of continuous functions given byfn(x) = x" converges in the topology ofpoint- ofpointwise convergence to the function/defined by fO forO<.v<], /(•v) forx=l. This example shows that the subspace e(/, R) of continuous functions is not closed in R> in the topology of pointwise convergence. We know that a sequence (/„) of continuous functions which converges in the uniform topology has a continuous limit, and the preceding example shows that a sequence which converges only in the topology of pointwise convergence need not. One can ask whether there is a topology intermediate between these two which nevertheless will ensure that the limit of a conver- convergent sequence of continuous functions is continuous. The answer is "yes"; assuming the (fairly mild) restriction that the space X be compactly generated, it will suffice if/n converges to /in the topology of compact convergence, which we now define. Definition. Let (Y, d) be a metric space; let Ibea topological space. Given an element / of Yx, a compact subset C in X, and a number e > 0. Iet *c(/, e) denote the set of all those elements g of Y* for which lub {d(f(x), g(x))lx e C]<e.
282 Complete Metric Spaces and Function Spaces _ Chap, t The sets Bc(f, e) form a basis for a topology on Yx. It is called the topolo of compact convergence (or sometimes the "topology of uniform convere on compact sets"). This topology differs from the preceding topology in that the general basis element containing/consists of functions that are "close" to/not iust at finitely many points, but at all points of some compact set. Note that if g e Bc(f, e), then there is a 5 > 0 such that we have Bc(g, 5) cBc(f,e); simply let S=e-\ub{d(f(x),g(x))\x e C}. It is then easy to check that these sets form a basis. The justification for the choice of terminology comes from the following theorem, whose proof is immediate. Theorem 4.2. A sequence fn .- X —> Y of functions converges to the function f in the topology of compact convergence if and only if for each compact subset C ofX, the sequence f„ | C converges uniformly to f |C. Definition. A space X is said to be compactly generated if it satisfies the following condition: A set A is open in X if and only if A n Cis open in C for each compact set C in X. This condition is equivalent to requiring that a set B be closed in X'fi and only if B n C is closed in C for each compact set C. It is a fairly mild restriction on the space; many familiar spaces are compactly generated. For instance: Lemma 4.3. IfX is locally compact, or ifX satisfies the first countability axiom, then X is compactly generated. Proof. Suppose that X is locally compact. Let A n C be open in C for every compact subset C of X. We show A is open. Given x e A, choose a neighborhood U of x which lies in a compact set C. Since A n Cis openm C by hypothesis, A n U is open in U, and hence open in X. Then A n is a neighborhood of x contained in A, so that A is open in X. Suppose that X satisfies the first countability axiom. If B n C is clos in C for each compact subset C of A', we show that B is closed in X. e x be a point of 5; we show that x e B. Since JT has a countable basis x, there is a sequence (.vj of points of B converging to x. The set C={x]U{x.\n e Z+] is compact, so that B n Cis by assumption closed in C. Since 5 O Ccon x» for every n, it contains * as well. Therefore, x e 5, as desired. U
§ 7-4 Pointwise and Compact Convergence 283 Theorem 4.4. Let Xbea compactly generated space; let (Y, d) be a metric space. Then e(X, Y) is closed in Yx in the topology of compact convergence. Proof. Let/ e Yx be a limit point of e(X, Y); we wish to show/is con- continuous. First we show that/|C is continuous for each compact set C in X. For each n, consider the neighborhood Bc(f, \/n) of/; it intersects Q(X, Y), so we can choose a function/, e Q(X, Y) lying in this neighborhood.'The sequence of functions /„ | C : C —* Y converges uniformly to the function f\ C, so that by the uniform limit theorem, f\ C is continuous. It follows that /is continuous. Let V be an open subset of Y; we show that f'l(V) is open in X. Given any subset C of X, If C is compact, this set is open in C because f\ C is continuous. Since X is compactly generated, it follows that f~'(V) is open in X. □ Corollary 4.5. Let X be a compactly generated space; let (Y, d) be a metric space. If a sequence of continuous functions fn : X —> Y converges to f in the topology of compact convergence, then f is continuous. Now we have three topologies for the function space Yx, when Y is metric. The relation between them is stated in the following theorem, whose proof is straightforward. Theorem 4.6. Let Xbea space; let (Y, d) be a metric space. For the function space Yx, one has the following inclusions of topologies: (uniform) => (compact convergence) => (pointwise convergence). JfX is compact, the first two coincide, and if X is discrete, the second two coincide. Exercises 1. Show that the sets Bc(f, e) form a basis for a topology on Yx. 2. Prove Theorem 4.2. 3. Prove Theorem 4.6. 4. Consider Yx in the topology of compact convergence. (a) Show Yx is regular. Is it normal ? (b) Show that if Xequals a countable union of open sets with compact closures, then Yx satisfies the first countability axiom. 5. Show that in general the set of bounded functions/: X-* R is not closed in Kx in the topology of compact convergence.
284 Complete Metric Spaces and Function Spaces 6. Consider the sequence of continuous functions fn: K+ . = i/nx. Chap. 7 ■ -R defined by In which of the three topologies of Theorem 4.6 does this sequence conver What about the sequence given in Exercise 9 of §2-10? 7. Consider the sequence of functions/„: (— 1,1) —> Rt defined by /.(*) = 21., fa*. (a) Show that (/„) converges in the topology of compact convergence; conclude that the limit function is continuous. (This is a standard fact about power series.) (b) Show that (/„) does not converge in the uniform topology. 8. A function /: X—> Ji is said to have compact support if/vanishes outside some compact set. Let GC(R, R) be the set of all continuous functions/: R—>R having compact support. Define a metric on this set by the rule - Si*)J (a) Show that D is a metric. The topology given by D is called the topology of mean convergence. [Hint: Define a "scalar product" by the rule then imitate the proof given in Exercise 8 of §2-9 for the euclidean metric] (b) In which of the four topologies we have for ec(X, Y) does the sequence/, pictured in Figure 7 converge? What about the sequence £„ pictured in Figure 8 ? Figure 7 Figure S (c) Show that the mean convergence topology is not comparable with the othet three.
The Compact-Open Topology 285 9. Let (Y, d) be a. metric space; let X be a space Define a t^ i as follows: Given fe e(X, Y), and given a i3 iopolo&onQ(X, Y) 5 : X-* R+ on X, let g e" * Posltlve ««hnuous funaion B(f, S)={g\ d(f(x), g(x)) < S(X) fop all x e *}. (a) Show that the sets B(f, 8) form a basis for a topology on er* Y\ w, call it the fine topology. ^ n CA> Y>- We (b) Show that the fine topology contains the uniform topology (c) Show that if X is compact, the fine and uniform topologies agree (d) Show that if X is discrete, then e(X, Y) = Y* and the fine and box topol- ogies agree. ^ 10. Here is a theorem about compactly generated spaces. We say that/- X—* Y is a proper map if for each compact set D in Y, the set /->(£) is compact in X. Theorem. Let X be a space; let Y be a compactly generated Hausdorff space. Iff'- X —» Y is continuous, injective, and proper, thenfisan imbedding andf(X) is closed in Y. 7-5 The Compact-Open Topology In defining the uniform topology and the topology of compact conver- convergence on the function space Yx, we needed to assume a metric for the space Y; both definitions involved this metric. (The topology of pointwise con- convergence on the function space Yx does not involve a metric for Y.) In general, topologists are concerned primarily with matters that depend only on the topology of a space, not on its metric. So one naturally asks whether either of these first two topologies depends only on the topology of Y, rather than on the particular metric d. There is no satisfactory answer to this question for the space Yx of all functions mapping X into Y. But for the space e(X, Y) of continuous func- functions, one can prove something; for this space, the topology of compact convergence is independent of the particular metric chosen for Y. To prove this, we define a certain topology on &{X, Y) called the compact- open topology, whose definition does not involve a metric for Y. Then we show that if Y happens to have a metric d, this topology agrees with the topology of compact convergence. Therefore, the latter topology on e(Jf, /; depends only on the space Y, not on the metric d. ..-..« The compact-open topology is very important in its own right; it turn out to have many useful properties. We shall use it ,n the next sect,on m proving a generalized version of Ascoli's theorem. It also has close^connec- close^connections with the notion of "homotopy," which is one of the fundamental concepts in algebraic topology. Definition. Let X and Y be topological space,. I f C is a compact subset of Xand {/is an open subset of Y, define
286 Complete Metric Spaces and Function Spaces Chap. 7 S(C, U) = {/|/ e e(X, Y) and f(Q c^. .. The sets S(C, U) form a subbasis for a topology on e(X, Y) calleH * compact-open topology. ' ** Note that both the compact-open topology and the point-open topolo are defined without assuming a metric on Y. It is clear from the definiti that the compact-open topology is in general finer than the point-open to! pology. Theorem S.I. Let X be a Space and let (Y, d) be a metric space. For the space S(X, Y), the compact-open topology and the topology of compact con- convergence coincide. Proof. Step I. If A is a subset of Y and e > 0, we define We call this set the "e-neighborhood" of A. If A is compact and Kisanopen set containing A, we show there is an e > 0 such that U(A, e) cz V. This was given earlier as an exercise (Exercise 4 of §3-6); we give a proof here. For each a e A, choose 5(a) > 0 so that Bd(a, S(a)) cz V. Cover A by finitely many open sets of the form Bd(ai,U(a,)),..., Bd(an, tf(a,)). Let e = min {|c5(a,)}. Then if a e A, the point a is in at least one of the sets Bd(at, {E(a,)); the triangie inequality implies that Bd(a, e) cz Bd(a,, 5@,)). Since this holds for each a in A, it follows that U(A, e) cz V, as desired. Step 2. We prove that the topology of compact convergence is finer than the compact-open topology. Let 5(C, U) be a subbasis element for the compact-open topology on e(X, Y) and let / be an element of S(C, U). Because/is continuous, f(C) is a compact subset of the open set U. By Step 1, we can choose e so the e-neighborhood of /(C) lies in U. Then BJLf, e) cz S(C, U). Step 3. We prove the compact-open topology is finer than the topology of compact convergence. Given an open set about/in the topology of com- compact convergence, it contains a basis element of the form Bc(f, t)■ We find a basis element for the compact-open topology that contains/and m 5Ea(h point x of Xhas a neighborhood Vx such that f(K) lies m an open set U of Y having diameter less than e. [For example,_choose V, so f(Vx) lies in the f/4-neighborhood of f(x). Then ftVJ lies in the ei
§ 7-5 The Compact-Open Topology 287 neighborhood of /(*), which has diameter at most 2e/3.] Cover Cby finitely many such sets V,, say for x = *,,..., xa. Let Cx = Vx n C. Then C is compact, and the basis element * CXI, uXl) n ■ • • n ^(c^, ux,) contains/and lies in Bc(f, e), as desired. □ Corollary 5.2. Let Y be a metric space. The compact convergence topology on the subspace S(X, Y) o/Yx does not depend on the metric o/Y. The fact that the definition of the compact-open topology does not in- involve any metric is just one of its useful features. Another is the fact that it satisfies the requirement of "joint continuity." Roughly speaking, this means that the expression f(x) is continuous not only in the single "variable" x, but is continuous jointly in both the "variables" x and/. More precisely, one has the following theorem : Theorem 5.3. Let X be locally compact Hausdorff; let <2(X, Y) have the compact-open topology. Then the map e:X x e(X, Y)— Y defined by the equation e(x, f) = f(x) is continuous. The map e is called the evaluation map. Proof. Given a point (x, f) of X X e(X, Y) and an open set V in Y about the image point e(x, f) = /(a), we wish to find an open set about (x, f) that e maps into V, First, using the continuity of/and the fact that Xte locally compact Hausdorff, we can choose an open set U about x having compact closure U, such that/carries U into V. Then consider the open set U X S(U, V) 'n X x e(X, Y). It is an open set containing (x, /). And if (x\ /') belongs to this set, then e{x', /') = /'(*') belongs to V, as desired. Q A consequence of this theorem is the following theorem. We shall not need to use it, but it is important in algebraic topology. Corollary 5.4. Let X be a locally compact Hausdorff space; let 6(X, Y) have the compact-open topology; let Z be an arbitrary topological space. Then a map F : X X Z —* Y is continuous if and only if the induced map F:Z —6(X,Y)
288 Complete Metric Spaces and Function Spaces is continuous, where F is defined by the rule ) = F(x, z). Chap. 7 Proof. Suppose that / is continuous. It follows that F is contin since F equals the composite 0USi X x Z -^ X x e(X, Y) —'-~ Y. Suppose that Fh continuous; we prove that F: Z — e(X, Y) is continuous. (This part of the proof does not use the local compactness of X.) To prove continuity, we take a point z0 of Z and a subbasis element S(C, U) o(e(X, Y) containing F(z0), and find a neighborhood Wofz0that is mapped by / into S(C, U). This will suffice. The statement that F(z0) lies in S(C, U) means simply that (/(zo))(x) = F(x, z0) is in U for all x e C. That is, F(C X z0) c U. Continuity of F implies that F~\U) is an open set in X X Z containing C x z0. Then /-•(£/) n(C X Z) is an open set in C X Z about the slice C X z0; the tube lemma of §3-5 shows that there is a neighborhood IV of r0 in Z such that the entire tube C x W lies in F^{U). See Figure 9. Then for z e Wand* e C, we have F(x, z) e I/. Hence f(W) cr S(C, t/), as desired. Q We discuss briefly the connections between the compact-open topology and the concept of homotopy. . m0. If/and g are continuous maps of X into Y, we say that /and ir are topic if there is a continuous map f:a-x[o, ij—y Fh such that f(.v, 0) = /(.v) and f(.v, 1) - g(x) for each a- e X TTie map called a homotopy between/and £. Roughly speaking, a homotopy is a "continuous o
§7-6 Ali's Theorem 289 I,]Q(X,Y which assigns, to each parameter valuer in [0 map from JTto Y. Assuming that *is locaHyJ^HaSS """""^ F is continuous if and only if / is continuous. TOs ml"s tW h^ "* that between two continuous maps f,g : X-* y !«!!, homotopy F Fin the function space Q(X, y) jLng the po^Tq/y) to^ Vf We sha.. return to a more detai.ed study of homotopy i^henexi 1. Let C be a subspace of A'. Show that the restriction map r: <S(X, Y) — e(C, Y) is continuous if both spaces have the point-open or compact-open topologies. 2. Show that in the compact-open topology, Q(X, Y) is Hausdorff if Y is Haus- dorff, and regular if Y is regular. [Hint: If 0 <= V, then S(C, U) <= ^(C, V).] 3. Let y be a metric space. Show that if X is compact, the uniform topology on G(X, Y) is independent of the metric chosen for Y. 4. Show that if y is locally compact Hausdorff, then composition of maps e(x, Y) x G(Y,z) — e(x,z) is continuous, provided the compact-open topology is used throughout. [Hint: ltg°feS{C, U), find V such that/(C) c Kand g(V) c U.] 5. Lete'(A", Y) denote the set 6( A", y) in some topology 3. Show that if the evalua- evaluation map e : X X e'(X, Y)-*Y is continuous, then 3 contains the compact-open topology. [Hint: Consider i : Q'(X, Y) —> e(X, Y). Note that local compactness of X is not needed.] 6. Here is an (unexpected) application of Corollary 5.4 to quotient maps: Show that if, : A-> 5 is a quotient map and * is locally compact Hausdorff then ixxp XX A-^XXB is* quotient map. Compare Exerc.se 11 of §3-8 [Hint: LeT(X x B)' denote the set * X 5 in the quotient topology.nduced^y ix x p- le w : A" X ^ - (X x 5)' be the quot.ent map. If *-^^ (AT x BY is the identity map, show that i°P = «; include that # ,s contmu ous.] /'j Theorem p^j the earlier one as a special case, we course of the proof.
290 Complete Metric Spaces and Function Spaces Chap. 7 Theorem 6.1 (Ascoli's theorem). Let X be locally compact ftw „ let (Y, A) be a metric space. Consider e(X, Y) in the compact-one, tonl* ( ) p e(X, Y) in the compact A subset S of 6(X, Y) has compact closure if and only if it is eZicon^ and the subset »nwus o/Y has compact closure for each x. Proof Step 1. We first show that if a subset <5 of e(X, Y) is equicontinn- ous, then the topologies it inherits from the topologies of pointwise conver- convergence and compact convergence on Yx are the same. In general, the compact convergence topology is finer than the pointwise convergence topology; we wish to prove the reverse inclusion holds for the subspace JF. Given/ e 3% consider the basis element 5C(/, e) for the compact convergence topology. It consists of all functions g such that Iub {d(g(x), f(x))\x e C}<e. We find a basis element B for the pointwise convergence topology such that /eBnJc Bc(f, e) n ff. Choose e, and e2 so that 2e, + e2 < f • Using equicontinuity of? and com- compactness of C, cover C by finitely many open sets £/,,..., Un of X such that on each set Ut, each of the functions in % varies by at most e,. Choose x, e U, for each /. Consider the subset B of Yx given by the equation B = {g\d(gix,), fix,)) < e2 for i = 1,..., n}; we assert that it is the desired basis element for the pointwise convergence topology. Let g be an element of B n ST. Given x e C, choose i so that x belongs to U,. Then , /(*<)) < f 2' (/(*,),/W)<e.. that d(g{x), f(x)) < e. This inequality holds for each x in C. Therefore, max {digix), fix)) \x e Cj < e, so that g e Bcif, e), as desired. Step 2. Second, we show that if a subset iF of Yx is equicontinuous, JJ its closure S in Yx relative to the topology of pointwise convergence tqU^Z. Given ,. e ^and f > 0, choose a neighborhood such that (*) dif(x), fixoy) < f/3 for all/ e ? and all x e K
§7-6 - ■ - Ascoli's Theorem 291 We assert that if g belongs to 9, then d{g(X) g(x )) < „ f „ follows that 9 is equicontinuous. «W-*W>» < e for all x e 17. It To prove this assertion, let x be a point of U Let V h» <u u Y*. open in the pointwise convergence topology cons SL of n t^ °f h of Yx such that Sy> consistlng of all elements (**) dWx),g{x))<e/3 and «*(*„), «(*„))< e/3 Because g belongs to the closure of <F in this topology, the neighborhood V of g must contain an element/ of <F. Applying the triangle inequaty^ (*)and(«),weseethatrfte(x),^0))<e>asasserted# 4 "^to Now we prove the theorem. Suppose first that iF is equicontinuous and that the set 3\, has compact closure in Y for each x e x Let g denote the closure of JF in y* relative to the pointwise convergence topology. Let S' denote the closure of JF in Yx relative to the topology of compact convergence. Then JF <= g' <= g. Because iF is equicontinuous, it follows from Step 2 that S is equicontinu- equicontinuous. Therefore, by Step I, the topologies g inherits from the topologies of pointwise convergence and compact convergence on Yx are the same. We conclude that g' = g. So far, we have used only the equicontinuity of JF. Now let Cx denote the closure in Y of ffA. Consider the product J[xEX Cx; we have by elementary set theory. The topology of pointwise convergence on Yx is the same as the product topology. In this topology, the set nQ« is compact, by the Tychonoff theorem. Since g is a closed subset of J[CX in this topology, it is compact in this topology. Therefore, S is also compact in the topology of compact convergence. Now g is equicontinuous, so each function in S is continuous. Hence g is contained in the set e(X, Y), a set on which the topology of compact convergence equals the compact-open topology. Therefore, the cl0SUIVjt SF in e(A\ Y) in the compact-open topology is the compact set S, as desirea. Step 4. Give e(X, Y) the compact-open topology; suppose that iF has compact closure <F in e(X, Y). We show that JF is equ.contmuousiand* has compact closure. The proof makes use of the fact that the evaluation map (XY)Y is continuous; here we make our first use of the local First, given X e X, We show ff, ^^Pf^ set x x § is compact in X X e(X, Y); therefore Y. But e{x X JF) contains e(x X SF) = SF,. § is equicOn- Second, we show JF equicontinuous at *„. In taci, we
292 Complete Metric Spaces and Function Spaces Chap. 7 tinuous at x0. Choose a compact set C that contains a neighborhood f It Will suffice to show the collection of functions *0> is equicontinuous, since C contains a neighborhood of x0. Consider the restriction map r:e(X,Y)-*G(C,Y) defined by r{f) =/ | C. Then (R = rE). If we give both spaces the compact- open topology, r is continuous, as you can easily check. Therefore, (R is a compact subset of 6(C, Y) in the compact-open topology. On the space 6(C, Y), the compact-open topology equals the uniform topology (because C is compact). Therefore, (R is compact in the uniform topology; by Theorem 3.1, CR is totally bounded in the metric p. We assert that there is a compact subset Yo of Ksuch that (R lies in the subspace e(C, Ya) of 6(C, Y). The theorem then follows at once from Lemma 3.2, for since C and Yo are compact, total boundedness of (R implies equicontinuity of (R. To find Yo is easy. The set C X J is compact in X x e(X, Y); let r0 denote its image under e. Because e is continuous, Yo is a compact subset of Y; it contains /(.v) for every x e C and every / e ff. It follows that <R e= e(C, y0). n An even more general version of Ascoli's theorem may be found in [K] or [Wd]. There it is not assumed that Y is a metric space, but only that it has what is called a uniform structure, which is a generalization of the notion of metric. Exercises 1. (a) Which step in the proof of Ascoli's theorem uses the local compactness of XI Which step uses the Tychonoff theorem ? (b) Show that the classical version of Ascoli's theorem is a special caseoftfl general version, and so is Exercise 5(c) of §7-3. 2. Prove the following generalized version of Arzela's theorem: ^ Theorem. Let X be locally compact Hausdorff with a countable baslS' fn:X—>Rk. If the collection {/„} is pointwisc bounded and equicontinuous'. the sequence (/„) has a subsequence that converges in the topology ofco convergence. 3. Which of the following subsets of Q(.R, R) are pointwise bounded? have compact closure relative to the topology of compact convergen (a) The collection {/.}, where /„(.*) = x + sin nx. (b) The collection {gn}, where ^,(.v) = x". (c) The collection [hj, where h,(x) = \x\w:
§ 7 ....... Baire Spaces 293 (d) The collection of all polynomials of degree at most 10, whose coefficients have absolute value less than 1. 4. Let ybe a metric space; let/,:X — Y be a sequence of continuous functions- let /: X —* Y be a function (not necessarily continuous). ' (a) Show that if {/„} is equicontinuous and/, —/in the topology of pointwise convergence, then/, -—/in the topology of compact convergence. (b) Prove the converse of (a) if X is locally compact Hausdorff. (c) Show that in either case,/is continuous. 7-7 Baire Spaces The defining condition for a Baire space is probably as "unnatural look- looking" as any condition we have yet introduced in this book. But bear with us awhile. In this section, we shall define Baire spaces and shall show that two im- important classes of spaces—the complete metric spaces and the compact HausdorfF spaces—are contained in the class of Baire spaces. Then we shall give some applications which, even if they do not make the Baire condition seem any more natural, will at least show what a useful tool it can be. In fact, it turns out to be a very useful and fairly sophisticated tool in both analysis and topology. Definition. Recall that if A is a subset of a space X, the interior of A is defined as the union of all open sets of X that are contained in A. To say that A has empty interior is to say then that A contains no open set of X other than the empty set. Example 1. The set Q of rationals has empty interior as a subset of R, but the interval [0, 1] has nonempty interior. The interval [0, 1] X 0 has empty interior as a subset of the plane R2, and so does the subset Q X R. Definition. A space Xis said to be a Baire space if the following condition holds: Given any countable collection [An] of closed sets of A" each of which has empty interior in X, their union [jAn also has empty interior in X. Example 2. The set Q of rationals is not a Baire space. For each one-point set in Q is closed and has empty interior in Q; and Q is the countable union of its one-point subsets. , r The set Z+, on the other hand, does form a Baire space. Every subset oi Z+ is open, so that there exist no subsets of Z+ having empty interior, except for the empty set. Therefore, Z+ satisfies the Baire condition vacuously. More generally, every closed subset of R, being a complete metric space, is a Baire space. Somewhat surprising is the fact that the irrationals in R also form a Baire space; see Exercise 7.
294 Complete Metric Spaces and Function Spaces Chap. 7 The terminology originally used by R. Baire for this concent i the word "category." A space X was said to be of the first ca equaled the union of a countable collection of closed sets in A'ha interiors; otherwise, it was said to be of the second category. ^ terminology, we can say the following: 8 s A space X is a Baire space if and only if every nonempty open set in X is of the second category. We shall not use the terms "first category" and "second category" in this book. The preceding definition is the "closed set definition" of a Baire space There is also a formulation involving open sets that is frequently useful. It is given in the following lemma. Lemma 7.1. X is a Baire space if and only if given any countable collection {UJ of open sets in X, each of which is dense in X, their intersection f]\J, is also dense in X. Proof Recall that a set C is dense in X if C = X. The theorem now fol- follows at once from the two remarks: A) A is closed in A'if and only if X — A is open in X. B) B has empty interior in X if and only if X — B is dense in X. Q There are a number of theorems giving conditions under which a space is a Baire space. The most important is the following: Theorem 7.2. If X is a compact Hausdorff space, or a complete metric space, then X is a Baire space. Proof. Given a countable collection {An} of closed sets in * having empty interiors, we want to show that their union \JAn also has empty interior in X. So, given the nonempty open set Uo of X, we must find a point x of C/o that does not lie in any of the sets An. Consider the first set A,. By hypothesis, A, does not contain fore, we may choose a point y of Uo that is not in A,. Regularity o^ with the fact that Ax is closed, enables us to choose a neighborhood U, y such that 0, n a, = 0, °' ^ U"' thanl- If X is metric, we also choose U, small enough that its diameter isIes* a°of In general, given the nonempty open set Un.u we choose a P" T. Un.l that is not in the closed set An, and then we choose U, to be a n s hood of this point such that
Baire Spaces 295 diam V, < \jn in the metric case. We assert that the intersection f|&. is nonempty. From this fact our theorem will follow. For if x is a point of (~}U,, then x is in Uo, because 0 c Ua. And for each n, the point x is not in A,, because £/„ is disjoint from A The proof that C\U, is nonempty splits into two parts, depending on whether X is compact Hausdorff or complete metric. If X is compact Haus- Hausdorff, we consider the nested sequence £/, => Ut =3 ... of nonempty subsets of X. The collection {£/„} satisfies the finite intersection condition; since X is compact, the intersection (~}U, must be nonempty. If X is complete metric, we apply the following lemma. □ Lemma 7.3. Let C, zd C, => • • • be a nested sequence of nonempty closed sets in the complete metric space X. If diam Cn —> 0, then p|Cn 56 0. iVoo/. We gave this as an exercise in §7-1. Here is a proof: Choose xn e Cn for each n. Because xn, xm e CN for n, m > N, and because diam CN can be made less than any given e by choosing N large enough, the sequence (*„) is a Cauchy sequence. Suppose that it converges to x. Then for given k, the subsequence xk, xk+t,. ■ ■ also converges to x. Thus x neces- necessarily belongs to Ck — Ck. Then x e Hc*> as desired. □ The theorem just proved about Baire spaces is the one most frequently applied, but it is not the strongest one that can be proved. One can prove for instance that any G6 set in a compact Hausdorff space or a complete metric space is a Baire space. These and other generalizations are given in the exercises. Example 3. The set Q of rationals is not a Gs set in the reals. Suppose that Q is the intersection of the countable collection {IVr) of open sets of R. For each q e Q, let V, be the open set R - {q}. Then the collection a = [IVn\n eZ+]U{K,|?e Q) is a countable collection of open sets of R. Each W» is dense in R, since Wn => Q. And each set V, is dense in R, being the complement of a single point. Since R is a Baire space, the intersection A of the elements of the countable collection a must be dense in R. But any point belonging to all the sets IV, is rational, and any point belonging to all the sets V, is irrational. Hence A is empty. Example 4. The second application is more amusing than profound It arises from a question that could be posed to a class in elementary calculus: Is there a function f-.R-^R that is continuous precisely at the rational points ofRl
296 Complete Metric Spaces and Function Spaces The answer is no, and the proof goes as foil ows: Let /: R _» & ^ trary function. Give the positive integer n, let U. be the union of the Jwfbi* of the following collection: elements {U\ U is open in R and diameter /(t/) < l/n}. Then [/„ is open in R. Furthermore, the set c = n.ez. i/» is precisely the set of points at which the function/is continuous, as you can readily check. Thus the set of points at which /is continuous forms a G» set in R. The set of rationals is not a Gs set in R; therefore, there can be no function /: R —> R that is continuous precisely at the rational points. Surprisingly enough, there does exist a function h that is continuous pre- precisely at the irrational points. Choose a bijective function /: Z+ —♦ Q; ]et gn = f(n). Then define h : R —» R by the rule h(qn) = I In for n e Z+, A(a-) = 0 for x irrational. We leave it to you to check that h is the desired function. Exercises 1. Let A' equal the countable union Ufi,. Show that if X is a Baire space and nonempty, at least one of the sets 5, has a nonempty interior. 2. The Baire theorem implies that R cannot be written as a countable union of closed subsets having empty interiors. Show this fails if the sets are not required to be closed. 3. Show that every open subset of a Baire space is a Baire space. 4. Show that every locally compact Hausdorff space is a Baire space. 5. Show that if every point .v of X has a neighborhood that is a Baire space, then X is a Baire space. [Hint: Use the open set formulation of the Baire condition.! 6. Show that if Y is a Gt set in X, and if X\s compact Hausdorff or complete^ memc then Y is a Baire space in the subspace topology. [Hint: Suppose that Y-[] » where IV, is open in X, and that £„ is closed in Kand has empty interior in ^ Given Uo open in X with Ua n Y ^ 0, find a sequence of open sets I/. ° with Un n y nonempty, such that £?„ c £/„_„ 0,r\5.= 0, diam C < l/n in the metric case, U. cz Wn.} 7. Show that the irrationals are a Baire space. 8. Show that every G, set in a locally compact Hausdorff space is a
§ 7-8 AN°»here'Differentiable Function 297 9. Check the statements made in Example 4 about C and h 10. Theorem {Uniform boundedness principle) Let X h? and let S be a subset ofe(X, H) such that for eachxe ** = {/(*)(/£?] is bounded. Then there is a nonempty open set U nf y „ u, , , r 9 are uniformly bounded; that is, thereTsZasftu/^^^ '" \f{X)\^Mfora,,x e Uandallfls " "dammberMs"chthat [Hint: Let AN = {x;|/(a)|^ Nfor all/e <F}j 11. Determine whether or not R, is a Baire space. 12. Show that IV is a Baire space in the box, product, and uniform topologies *13. Show that if A is any Gt set in *, there is a function/: * _> *,ha, is continuous at all points of A and discontinuous at all other points. *14. Let A" be a topological space; let Y be a complete metric space. Show that <B(X, Y) is a Baire space in the fine topology (see Exercise 9 of §7-4) [Hint- Given basis elements £(/>, £•) such that <5, <£ i and dul£d,ft and /;+i 6 5(/f, E,/3), show that 7-8 A Nowhere-Differentiable Function We prove the following result from analysis: Theorem 8.1. Let h : [0, 1] —> R be a continuous function. Given e > 0, there is a function g : [0, 1] — R with |h(x) - g(x)j < e for all x, such that g is continuous and nowhcre-differentiable. Proof. Let / = [0, 1]. Consider the space e = e(/, R) of continuous maps from / to R, in the metric />(/,£) = max {|/(*)-£(*) |}. This space is a complete metric space, and therefore is a Baire space. We shall define, for each n, a certain subset U. of e that is open in 6 and dense in e, and has the property that the functions belonging to the intersection are nowhere-differentiable. Since e is a Baire space this intersection is dense in e, by Lemma 7.1. Therefore, given /, and e, th,s intersection must contam a function g such that p{h,g) < e. The theorem Allows The tricky part is to define the set Um properly. We fin£ take^a /and consider its difference quotients. Given x e / and given 0 consider the expressions f(x and -h
298 Complete Metric Spaces and Function Spaces Chap. 7 Since h <, \, at least one of the numbers x + h and x — h belongs t so that at least one of these expressions is defined. Let Af(x, h) denoted larger of the two if both are defined; otherwise let it denote the one that' defined. If the derivative /'(*) of/at x exists, it equals the limit of these difference quotients, so that \f'ix)\ = Km &f(x,h). We seek to find a continuous function for which this limit does not exist This gives us the idea for defining the set Un. Roughly speaking, U, will consist of those functions such that Af(x, h) is large (bounded below by a number greater than n) for values of h which are small (less than 1/n). More precisely, given positive numbers a and h, with h < £, we define a set {/(a, h) <= C(/, R) by setting U(<x, h) = {/\ Af(x, h)>a, for every x e /}. Then define £/„ to be the union of all those sets U(a, h) for which a>n and /; < I/n. Note that Un => Un+l => Example 1. Let a be given. The function /(.v) = 4<X;t(l — x), whose graph is a parabola, belongs to U(a, h) for /; = \. Geometrically speaking, what this says is that for each x, at least one of the indicated secant lines of the parabola in Figure 10 has slope of absolute value at least a. The function g pictured in Figure 10 belongs to U(a, li) for any h<\; and the function k belongs to U(a, ti) for any h < \. x-l x x + 1 Figure 10 Now we prove the following facts about the set Un: A) V. is open in C. Suppose that/6 U.. Then/e U(z,h) for « > n and some h < I In. Let
fi 7-8 A Nowhere-Differentiable Function 299 We assert that if g is any function such that n(f g) < * thp , . UQt\ h), where a' = (a + n)/2. Since a' > /this Jam W ? ? ^ t0 Un. Thus the ^-neighborhood of/belongs fo l^JToPen „ e ** * T°POVe "* ^ g(x + h)- g(x) If the first difference quotient is at least a in absolute value, then the second one is in absolute value at least a - (a - «)/2 = a'. A similar computation applies if Af(x, h) equals the other difference quotient. B) Ua is dense in e. We must show that given/in e, given e > 0, and given n, we can find an element g of Un within e of/ Choose a > /;. We shall construct g as a "piecewise-linear" function, that is, a function whose graph is a broken line segment; each line segment in the graph of g will have slope at least a in absolute value. It follows at once that such a function g belongs to (/„. For let 0 = x0 < x, < .Vj < ■ ■ ■ < xk = 1 be a partition of the interval [0, 1] such that the restriction of g to each subinterval /, = [.*,_,, .vj is a linear function. Then choose h so that If* is in [0, 1], then x belongs to some subinterval /,. If* belongs to the first half of the subinterval /,, then x + h belongs to /, and (g(x + A) -g(x))/h equals the slope of the linear function g | /, Similarly, if* Wongs tc.the^sec- tc.the^second half of /„ then x - /. belongs to /, and (g(* -I,) -^f'^^ the slope of g | /, In either case, Ag(*. h) > a, so g e f/(a, A) c: ^, as des red Now given/, e, and a, we must show how to construct the desired^piece wise-linear function g. First, we use uniform continuity of / to choose partition of the interval h,,inE ,he propTO J/^-'rS.-.^,1^.** of this partition. For each i = 1, • • ■, m>'=no°sc L ns then define a piecewise-linear function gl by the equations for* e [f<-i. «J» SiW = ],•/, ..g&^l&^Cx-a,) forxe[aB/J.
300 Complete Metric Spaces and Function Spaces Chap. 7 a, ft 1 </ a2 i 9 a3 n \J - a. '3 U Figure 11 /(';-i)> we require a, to be close enough to /,■ that /, — a, \f(t.) -/('■-■)I. a Then the graph of gl will consist entirely of line segments of slope zero and line segments of slope at least a in absolute value. Furthermore, we assert that p(g,, f) < e/2: On the interval /„ both gl(x) and /(*) vary by at most e/A from /(/,_,); therefore, they are within e/2 of each other. Then p(gl, f) = max {\gx{x) - f(x)\] < e/2. The function gt is not yet the function we want. We now define a function g by replacing each horizontal line segment in the graph of gt by a "sawtooth graph that lies within e/2 of the graph of gl and has the property that each edge of the sawtooth has slope at least a in absolute value. We leave t part of the construction to you; it is not hard. The result is the desired piec wise-linear function g. See Figure 12. C) C\Un consists of nowhere-differentiable functions. Let / e \\V>- shall prove that given x in [0, 1], the limit lim Af(x, h) h—*a - j does not exist: Given n, the fact that/belongs to Un means that we can a number hn with 0 < hn < \/n such that , h.) > n.
§ 7-9 ' An Introduction to Dimension Theory 301 Figure 12 Then the sequence (/«„) converges to zero, but the sequence (Af(x, hj) does not converge. As a result, /is not differentiable at x. □ You may find this proof frustrating, in that it seems so abstract and non- constructive. Implicit in the proof, however, is a procedure for constructing a specific sequence/ of piecevvise-linear functions that converges uniform y to the nowhere-differentiable function/ And defining the funct.on/.n this way is just as constructive as the usual definition of the sine function, for instance, as the limit of an infinite series. Exercises 1. Check the stated properties of the functions /, g, and k of Example 1 2. Given „ and «. define a continuous function /: /- * such that/6 tt and 7.9 An Introduction to Dimension Theory We showed in §4-5 that if X is a bedded in R» for some positive integer N
302 Complete Metric Spaces and Function Spaces Chap. 7 theorem to arbitrary compact metrizable spaces, and we also d how large a value of N is required. ■ erniine We shall define, for an arbitrary topological space X, & notion of to gical dimension. It is the "covering dimension" originally defined K Lebesgue. We shall prove that each compact subset of Rm has topolo dimension at most m. We shall also prove that the topological dimensio^of any compact w-manifold is at most m. (His, in fact, precisely m, but this L shall not prove.) The major theorem of this section is the theorem that any compact metrizable space of topological dimension m can be imbedded in R" for N = 2m + 1. This value of N is in general the best possible. The proof is ■i application of the Baire theorem. It follows that every compact w-mani- old can be imbedded in R2m+l. It follows also that a compact metrizable space can be imbedded in RN for some N if and only if it has finite topological dimension. Much of what we shall do holds without requiring the space in question to be compact. But we shall restrict ourselves to that case whenever it is convenient to do so. Generalizations to the noncompact case are given in the exercises. Definition. A collection Q, of subsets of the space A'is said to have order m + 1 if some point of X lies in m + 1 elements of G, and no point of X lies in more than m + 1 elements of G. Now we define what we mean by the topological dimension of a space X. Recall that given a collection G of subsets of X, a collection (B is said to refine G, or to be a refinement of G, if each element B of 63 is contained in at least one element of G. Definition. A space JHs said to be finite-dimensional if there is some integer m such that for every open covering G of X, there is an open covering « o JT that refines G and has order at most m + 1. The topological dimensim" X is defined to be the smallest value of m for which this statement hold . Some basic facts about topological dimension are given in the following theorems: Theorem 9.1. If Y is a closed subset of X, and if X has finite dimension, then so does Y; and dim Y <; dim X. Proof. Let dim X = m. Let G be a covering of Y by se£ ^" ^over For each A e G, choose an open set A' of X such thai.A' Oj ' onent X by the open sets A', along with the open set X - Y. Let » »" j of this covering that is an open covering of X and has order at most m Then the collection {Bn Y\B e &}
§ 7-9 - ■ An Introduction to Dimension Theory S03 is a covering of Y by sets open in Y, it has order at most m + 1, and it refines a. □ Theorem 9.2. Let X = Y U Z, w/im? Y anrf Z ai-e c/oflv/ «?/.y /» X having finite topological dimension. Then dim X = max {dim Y, dim Z}. Proof. Let m = max {dim K, dim Z). By the preceding theorem, we know that if X is finite dimensional, dim AT;> m. It remains to prove that dim X exists and is at most m. Step 1. First we prove the following: Suppose Y is a closed subspace of X, and dim Y < m. If G is an open covering of X, then there is an open cover- covering © of X that refines a , such that no point of Y lies in more than m+1 elements of C. (If C satisfies this condition, we say that "the restriction of C to Y has order at most m +1".) To prove this fact, consider the collection {a n y | a e a}. It is an open covering of Y, so it has a refinement (B that is an open covering of Y and has order at most /w + 1. Given 5 e (B, choose an open set UB of X such that £/* n Y = B. Choose also an element AB of a such that B c= AB. Let e be the collection consisting of all the sets UB C\ AB, for B e (B, along with all the sets /4 — Y, for ^ e a. Then e is the desired open covering of X. Step 2. Now let a be an open covering of X. Let (B be an open covering of A'refining a, whose restriction to Y has order at most m + 1. Then let C be an open covering of X refining (B, whose restriction to Z has order at most m+1. We form a new covering © of A' as follows: Define/: © — <B by choosing for each C e e an element f(C) of (B such that C c /(C). Given B e (B, define 2)E) to be the union of all those elements C of C for which /(C) = 5. (Of course, D(B) is empty if B is not in the image of/.) Let © be the collec- collection of all the sets D(B), for B e (B. Now 2D refines (B, because £>E) <= 5 for each B; therefore, 3D refines a. Also, 2D covers * because C covers * and C c= 2>(/(C)) for each C e e. We show that 2D has order at most m + 1; then our theorem will be proved. Suppose x e D(B,) n ■ ■ ■ H £Ek), where the sets £>(£,) are distinct. We wish to prove that k <; m + 1. Note that the sets B,,...,Bk must be distinct because the sets D(B,) are. Because x e £>E,), we can choose for each i, a set C, £ C such that x e C, and /(C,) = 5,. The sets C, are distinct because the sets B, are. Furthermore, €[c:n ■■■ nck]c [D(B%) n ••• n D(Bk)] <= [5, n • • • n 5J.
Complete Metric Spaces and Function Spaces Chap. 7 If* happens to lie in Y, then k<m+\ because the restriction of«t v has order at most m + 1; and if x is in Z, then k^m+ blnVV restriction of e to 2 has order at most m + 1. Q Se the Corollary 9.3. Let X = Y, U • • • U Yt, where each Y, /, closed in X an* is finite-dimensional. Then X and dim X = max [dim Y,,.. ., dim Yk}. Obviously, the dimension of X (if it exists) is a topological invariant In the case where X is compact and metric, the definition can be formulated in a way that involves the metric: Lemma 9.4. Let X be a compact metric space. Then dim X < m if and only if for every e > 0 there is a finite open covering (B ofX by sets of diameter less than e such that (B has order at most m + 1. Proof. Suppose that dim X < m. Take a covering a of A' by open f/3- balls, and choose <&' to be a refinement of a that is an open covering of A' and has order at most m + 1. Let (B be a finite subcollection of (B' that covers X. Then (B has order at most m + 1, and the elements of (B have diameter less than e. To prove the reverse implication, let a be an arbitrary open covering of X. Let e > 0 be a Lebesgue number for a, and choose a finite open covering (B of X by sets of diameter less than e, such that (B has order at most m + 1. Then (B is the desired refinement of a. [^] Example I. The interval [a, b] has topological dimension 1. Given f > 0, one can choose a subdivision a - t0 < tt < ■■■ < /„ = b such that /, — /,-_ i < f/2 for each /. Then the collection <B = {[a, /,), (to, h), (/,, t}), (h, u),.... (t.-2, O, ('.-i. b]] is an open covering by sets of diameter less than f, and (B has order 2. Hence dim [a, b] < I. ,, On the other hand, given € < b - a, let (B be any open covering of [a, m lt 2 S"PP° On the other hand, given € < b a, le y p by sets of diameter less than e. We assert that (B has order at least 2. that (B has order 1; then no two elements of (B intersect. Since the elemen of (B have diameter less than e, the collection (B must contain more_ man o element; let U be one element of <B, and let V be the union of all the oin ^j Then V and V form a separation of [a, b], contrary to the fact that this mie is connected. Example 2. A (finite) linear graph C is a space that is homeomf* ^J| union of finitely many line segments, each pair of which intersect.i common end point. The end points of the line segments are caBed^ }} of the linear graph. Two examples of linear ^^^^Jr; *» The first is a diagram of the familiar ^BS^Mr-^^^ i second is called the "complete graph on five vertices. Neither oi
§7-9 An Introduction to Dimension Theory 305 Figure 13 imbedded in the plane. [This is a highly nontrivial fact to prove; it depends on the Jordan curve theorem. See the exercises of §8-13.] Any linear graph has topological dimension I. There are of course many spaces that are not linear graphs which also have dimension I (see Exercise 3). Example 3. Any compact subset C of R1 has topological dimension at most 2. To prove this fact, we construct a certain open covering fl of R2 that has order 3. We begin by defining Ci2 to be the collection of all open unit squares in R2 of the following form: G2 = R", Hi I) x (m, m -+ I)| n, m integers}. Note that the elements of G2 are disjoint. Then we define a collection Oli by taking each (open) edge e of one of these squares, e = {«} x (m, m -I- I) or e = (n, n + I) x [m], and expanding it slightly to an open set £/, of R2, being careful to ensure that if e ^ e', the sets Uc and £/„■ are disjoint. We also choose each £/, so that its diameter is at most 2. Finally, we define <20 to be the collection consisting of all open balls of radius % about the points /; x m. See Figure 14. The collection of open sets ft = G2 U a, u flo covers Rz. Each of its ele- elements has diameter at most 2. And it has order 3, since no point of R1 can lie in more than one set from each G,-. A t a, Figure 14
306 Complete Metric Spaces and Function Spaces ^. Chap. 7 Now we prove that any compact set C in R* has topological dimensio most 2. Given e > 0, consider the homeomorphism /: & -* rz defined h /(x) = ($e)x. Take the images under / of the open sets of the collection ft They will form an open covering of & by sets of diameter less than € Cho finitely many of them that cover C; this will be the desired open covering of°c Of course, not every compact subset of R* has dimension exactly 2. A. one point set has dimension 0, for instance, and a closed interval on the jc-axis has dimension 1. It is, in fact, surprisingly nontrivial to prove that some subsets of R2 do have topological dimension 2; the techniques of algebraic topology are required. See Example 6 below. Example 4. Every compact 2-manifold X has topological dimension at most 2. For X can be written as a finite union of subspaces that are homeomorphic to the closed unit ball in R2; then Corollary 9.3 applies. The 2-sphere, the torus, and the Klein bottle are typical examples of 2-manifolds (see Figure 27 of §8-10).' Example 5. A finite simplicial 2-complex (called a 2-complex for short) is a space X that is homeomorphic to a finite union of closed triangles and line segments such that any two of them intersect in at most a common vertex or (in the case of two triangles) a common edge. By Corollary 9.3, every 2-cornplex has topological dimension at most 2. Pictured in Figure 15 is a 2-complex that is not a manifold. C \ \9 Figure 15 Properly speaking, a 2-complex must contain at least one triangle; ot er- wise it would be a linear graph. Example 6. Assuming a theorem we shall prove in the next^chapter (in §8-8), we now show that any closed triangular region Tin R has i°p ^ dimension 2. It follows that every compact 2-manifold, and every 2-comp , topological dimension precisely 2. The theorem in question is the following: Let Bd T denote the union of the edges of the closed triangula. Then there is no continuous map f: T— Bd T that carries each into itself.
6 7-9 An Introduction to Dimension Theory 307 We know that T has dimension at most 2. To st « * -*'"«--«««» less than 3. For each i = l „, choose a ver(ex Qf T intersects two edges of T, let „ denote the vertex common to these edges 5 U, intersects only one edge of T, let „, be one of the vertices of this edge If U intersects no edge of T, let v, be any vertex of T. ' Now let {(/>,] be a partition of unity dominated by f [/, f/1 rq,.i> ut > Define*:r—i?* by the equation »J-^ee 54-5.) Then g is continuous. Now each point of T lies in at most two elements of a. Therefore, for each x e T, at most two of the functions fax) are nonzero. Then g(x) = v, if x lies in only one open set V,, and g(x) = /c, +(I -/)i;y for some / with 0 < / <; 1 if x lies in two open sets U, and Uj. [Here we use the fact that 20((-v) = '•] Thus g maps all of T into the union of the edges of T. We claim further that g maps each edge of T into itself. Suppose that x belongs to the edge vw of T. Then any open set U, containing x necessarily intersects this edge, so that the vertex v, must be either v or w. Likewise, if .v e Uj, we have vj = v or w. The above equations then show that,?(A-) belongs to vw. The existence of the map g violates the above-quoted theorem, Now we prove the imbedding theorem mentioned earlier, to the effect that every compact metrizable space of topological dimension m can be imbedded in Rlm+i. This theorem is another "deep" theorem; it is not at all obvious, for instance, why 2m + 1 should be the crucial dimension. We shall discuss some special cases of the theorem before turning to the proof; this discussion should illuminate why the dimension 2m + 1 occurs. First let us consider the simplest possible case, that of a finite linear graph G. In this case, the imbedding theorem states that every finite linear grapn can be imbedded in RK It is not hard to give an ad hoc argument that will work here. We shall give instead an argument that will generalize to higner dimensions, an argument that involves the notion of general position. A set A of points of R> is said to be in general position in R> *«°*"°f the points are equal, no three are collinear, and no four are cophw.«B easy to find sets of points in general position in *'. For instance, the set . consisting of all points of the curve
308 Complete Metric Spaces and Function Spaces Chap. 7 {(t, t\ t>)\t e R] . . . is in general position in R3, as you can prove. . Now, given a finite linear graph G with vertices »„ . v iet u a set {Zl,. .., zj of points in general position in R3. We define a Map/: C -> R> by letting/map ,, to z,, and letting/map the ££%$ A - />, of the line segment in G joining v, and », (if any) to the poj t tz, + A - /)z, of the line segment in R3 joining z, and z,. It is clear thatf is continuous. Also, / is injective: Let e = »,», and e' = ^, be two line segments of G. If <? and e' have no vertex in common, then f(e) and /■('e'-v are disjoint, for otherwise the points z,, z,, z*, and z, would be coplanar And if e and e' have a vertex in common, say i = k, then f(e) and /(e') can intersect only in the point z,; for otherwise z;, zp and z, would be col- linear. Now let us consider the next simplest case, that of a 2-complex. In order to generalize the preceding argument, we must first study a bit of the analytic geometry of RN. We shall need these results later anyway when we prove the general imbedding theorem. Analytic geometry in RN is really nothing more than the usual linear algebra of vector spaces translated into somewhat different language. First we make the following definition: Definition. A set {x0, . . . , xA.} of points of RN is said to be geometrically independent if E*-=o ",xf = 0 with £*.„ a, =0]=> [each a, = 0]. It is clear that any set consisting of only one point is geometrically in- independent. What does geometric independence mean if k > 0? It is simple algebra to show that the set of points {x0,. .., xA.} is geometrically indepen- independent if and only if the set of vectors {x, — x0, x, — x0,.. ., x* — x0] is linearly independent in the sense of ordinary linear algebra. This gives us something to visualize: Any two distinct points form a geometrically in- independent set. Three points form a geometrically independent set it my are not collinear. Four points in R> form a geometrically independent if they are not coplanar. And so on. It follows from these remarks that the points 0 = @, 0,.... 0), e, =A,0, ...,0), e,=@,0 1) . . are geometrically independent in R". It also follows that R" conta.ns more than N + 1 geometrically independent points.
§7"9 An Introduction to Dimension Theory 309 Definition. Let {x0 x ] be a set of nnmu nf dn *u * • independent. The plane , determined byl^i sis tl^Z^ of all points x of R" such that ° t0 be the set x = S*=o '(X,, where £*,„ ,, = i. / It is simple algebra to check that P can also be expressed as the set of all points x such that for some scalars a „ ..., ak. Thus P can be described not only as "the plane determined by the points x0,..., xk," but also as "the plane passing through the point x0 parallel to the vectors x, — x0,..., xk — x0." Consider now the homeomorphism T : R» — R» defined by the equation T(x) = x - x0. It is called a translation of RN. Expression (*) shows that this map carries the plane P onto the vector subspace V of RN having as basis the vectors x, — x0, . .. , xk — x0. For this reason, we often call P a A-plane in RN. Two facts follow at once: First, if k <N, the fe-plane P necessarily has empty interior in RN (because Vk does). And second, if y is any point of R" not lying in P, then the set [x0, . . . , xk, y} is geometrically independent. For if y ^ P, then T{yj = y — x0 is not in V. By a standard theorem of linear algebra, the vector [x, — x0,..., xk — x0, y — x0} are linearly independent, from which our result follows. Definition. A set A of points of RN is said to be in general position in RN if every subset of A containing N + 1 or fewer points is geometrically in- independent. In the case of R\ this is the same as the definition given earlier, as you can check. Lemma 9.5. Given a finite set [x,, . . . , xn] of points of R* and given S > 0, there exists a set [y,, . . . , y0} of points of RN in general position in R , such that | X[ — y, | < 5 for all i. Proof We proceed by induction. Set y, = x,. Suppose that we are given y,, •.., y, in general position in *". Consider the set of all planes in* determined by subsets of [y,,. .., y,} that contain N or fewer demente. Every such subset is geometrically independent and determines a *-plane of R" for some k<.N-\. Each of these planes has empty interior mR. Since there are only finitely many of them, their union also has empty interior in R». (Recall that *" is a Baire space.) Choose y,+ 1 to be a pom of R» within 5 of x,+, that does not lie in any of these planes. It follows at once that the set
310 Complete Metric Spaces and Function Spaces Chap. 7 ■' ' ' ' c = {yi yP,yP+i} • ■, ,. , is in general position in RN. For let D be any subset of C containing V + or fewer elements. If D does not contain yp+l, then D is geometrically dependent by the induction hypothesis. If D does contain yp+i, then I)"~ {y,+,} contains N or fewer points and y,+ I is not in the pla'ne'determined by these points, by construction. Then as noted above, D is geometricall independent. □ ^ Now we can show how to imbed any 2-complex X in R$. Given the 2- complex X with vertices «,,...,«„, let us choose a set z,,..., z, of points that are in general position in Rs. We define a continuous map/: X—» R* as follows: The general point of the triangle with vertices vt, v,y and wkcan be written in the form SV, + tVj + (\ — S — t)vk, where s, t, and I — s — t are all nonnegative. (You can convince yourself of this fact.) We let/map this point to the point sz, + tzs + (\ —s — t)ik of the triangle in Rs having z,, z;, and zk as vertices. If v,vf is a line segment that is not the edge of a triangle, we let/map it onto the line segment z,z;, as before. Because the points z, are in general position, the map/will be injective. Several cases need to be checked. For instance, if vu vt, v3 andvitv5,v6are the vertices, respectively, of two disjoint triangles a and a' of X, then f(a) and f{a') cannot intersect, for that would imply that the points z,, for i = 1, .. . , 6, are not independent. [If x0 were a point of the intersection, we would have where Y^s,- = 2]'; = 1, or where the sum of the coefficients is zero.] Similar arguments apply in the other cases. . With these special cases as motivation, let us now prove the genera imbedding theorem. Theorem 9.6 (The Imbedding theorem). Every compact metrizable space X of topological dimension m can be imbedded in R2m+1. Proof. Let N = 2m + 1. Let us denote the square metric for R" by |x-y| = max {| *,-;■,!;/= l,...,N}. Then we can use p to denote the corresponding sup metric on the spac Q(X RN)' p(f, S) = max [| f{x) - g(x)\; x e X}.
§7-9 An Introduction to Dimension Theory 311 The space e(X, R») is complete in the metric n since ™ ■ square metric. p'slnce R IS complete in the Choose a metric d for the space X; because X is Given a continuous map /: X —* R", 'let us define C°mpact> ls bounded. = lub {diam /-'({z})| z e R"}. The number A(/) measures how far f "deviate" f™m A(/) - 0, each set /-.«,), consists of at^on/Z Now, given * > 0, define U. to be the set of all those /: X-,r for which A(/) < f; it consists of all thc^KSS from being injective by less than e. We shall show that U is both open and dense in Q(X, RN). It follows that the intersection HnezMvn is dense in Q(X, RN) and is in particular nonempty. If/is an element of this intersection, then A(/) < l/n for every n. Hence A(/) = 0 and/is injective. Because A'is compact,/is an imbedding. Thus the imbedding theorem is proved. (J) Uc is open in G(X, RN). Given an element/of Uc, we wish to find some ball Bp(f, 5) about/that is contained in Ut. First choose a number b such that A(/) < b < €. Note that if f(x) = f(y) = z, then x and y belong to the set /"'({»)> so that <r/(.v, ;■) must be less than ^. It follows that if we let A be the following subset of X X X, A = {x xy\d(x,y)>b), then the function j/O) — f(y)\ is positive on A. Now A is closed in X X JT and therefore compact; hence the function \f(x)-f(y)\ has a positive minimum on A. Let <J = \ min fl/(;c) - f(y)\; x x y £ A}. We assert that this value of 5 will suffice. Suppose that g is a map such that /</,*) < * If * ^ e ^ then |/W - /(,), * 2, by definition; since ,(x) and£), .« within. /W and /(y), respectively, we must have |g(.v) - g(y)\> 0- wentc "• ,, I * W-i0') I is positive on ^. As a result, if * and y are two.points such Mt g(V=/w! then necessarily *,,)<*. We conclude that A* <:*<*. as desired. ^ B) c. f, ^ /, «X. *>). This is the hard part of the proof^Wenced to use the analytic geometry of*" discussed above. Let/ e c( , , e > 0 and given <J > 0, we wish to find a function * 6 e(X, K ) g s Ut and />(/, g) < <J- , „ i such that - Let us cover X by finitely many open sets {£/, <J A) diam i7, < e/2 in Z.
312 Complete Metric Spaces and Function Spaces Chap. 7 B) diam f(U,) < d/2 in R". C) {17, J7D} has order <,m+ 1. (A Lebesgue number argument is involved here.) Let {0,} be a partition unity dominated by {Us} (see §4-5). For each i, choose a point x, e U Th choose, for each /, a point z, e /?" such that zt is within <5/2 'of the r^"! /(*,), and such that the set P°mt is in general position in RN. Finally, define g : X-+ R" by the equation We assert that g is the desired function. First, we show that />(/, g) < E. Note that s(x) ~ f(x) = S"=i 0,(-v)z( — 2"=i &(*) /(*); here we use the fact that £0,(x) = 1. Then g(x) ~ f(x) = S0f(-v)(z< — /(x.)) + £0,-(*)(/(*i) — /(*)). Now |z,. — /(x,)| < E/2 for each ;, by choice of the points z,. And if i is an index such that 0,(.v) ?t 0, then x e £/,; because diam/((/,)< E/2, it follows that |/(x,) — /(.y)| < E/2. Since 5]0,(x) = 1, we conclude that - f(x)\ < 5. Therefore, p(g, f) < 5, as desired. Second, we show that g e U,. We shall prove that if x, y e A'and g(x) = ), then x and >■ belong to one of the open sets Ult so that necessarily d(x, y) < f/2 (since diam £/, < f/2). As a result, Afe) < f/2 < f, as desired. So suppose g(x) = g(y). Then Because the covering {t/,-} has order at most w + 1, at most m + 1 of ^ numbers 0,(x) are nonzero, and at most m + 1 of the numbers <j>t(y) are non- nonzero. Thus the sum £ [0,(x) - 0,(>')]z, has at most 2m + 2 nonzero terms. Note that the sum of the coefficients vanishes since £ [«*) - *O0J =1-1 = 0. The points z( are in general position in RN, so that any subset of them having N + 1 or fewer elements is geometrically independent. And by hypothesis N -f 1 = 2m + 2. (Aha!) Therefore, we conclude that Now <j>,(x) > 0 for some i, so that x e Ut. Since 0,0') = <t>t(x)> we haV* y e Ut also, as asserted. □ To give some content to the imbedding theorem, we need some mow examples of spaces that are finite dimensional. We prove the to h theorem:
§ 7-9 ! An Introduction to Dimension Theory 313 Theorem 9.7. Every compact subset of RN has topological dimension at mostN. Proof. The proof is a generalization of the proof given in Example 3 above for R2. Let p be the square metric on RN. Step I. We begin by breaking RN up into "unit cubes." Define <J to be the following collection of open intervals in R: 3 = {(«, n+l)[neZ), and define 3C to be the following collection of one-point sets in R: X = {{n}\n e Z}. If Mis an integer such that 0 < M < N, let QM denote the set of all products C = Ai X A2 X • ■ • x AN, where exactly M of the sets A, belong to g, and the remainder belong to 3C. If M > 0, then C is homeomorphic to the product @, \)M, and will be called an M-cube. If M = 0, then C consists of a single point, and will be called a 0-cube. Let e = Co U C, U • • • ue,v. Note that each point x of RNlies in pre- precisely one element of C, because each real number .v, lies in precisely one ele- element of § U 3C. We shall expand each clement C of C slightly to an open set U(C) of RN of diameter at most 2, in such a way that if C and C" are two different M-cubes, then U(C) and U(C) are disjoint. First, note that if C and C are two different M-cubes, then C n C" = 0. This is easily checked; let C — A, X • • • X AN and C" = A\ X • • • X A'Nl Choose an index / so that A, # ^,!- If A, and ^,- are disjoint, then so are C and C". Otherwise, ^, must be an interval (n, n + 1) and A', must equal {n} or {/; + 1). In that case, since both C and C" are A/-cubes, there must be another index j for which A, is a one-point set {A-} and A) is an interval (A / + 1). Then A, and A) are disjoint, so that C and C" are disjoint. Second, note that the collection e.v, of all M-cubes is locally finite. Indeed, if a is a point with integer coordinates, then the open set Bp(a, 1) intersects only 3N different cubes of all dimensions. Let x e C, where C is an M-cube. There is a neighborhood of x intersect- intersecting only finitely many A/-cubes C" different from C. Since C n C = 0 for each such C", we can choose f(x) so that the f(x)-neighborhood of x does not intersect any M-cube C" different from C. Also, let f(x) < 1. Then we define U(C) = Ux.c B/x, ±f(x)). It is straightforward to check that if C and C are different M-cubes, U(Q and U(C) are disjoint. S/ep 2. Given A/ with 0 < M < N, define a.w to be the collection of all sets U(C), where C e eM. The elements of dM are disjoint, and each has
314 Complete Metric Spaces and Function Spaces m Chap. 7 diameter at most 2. (For C has diameter 1, and each point of U(C) Ii« «,•«.• ie(x) < J of a point x of C.) ^ Wlthln The remainder of the proof is a copy of the proof given in Examni i for R2. □ p J Corollary 9.8. Every compact m-manifold has topological dimension at most m. Corollary 9.9. Every compact m-manifold can be imbedded in RJ«+i Corollary 9.10. Let X be a compact metrizable space. Then X can be im- imbedded in some euclidean space RN ;/ and only if X has finite topological dimension. As mentioned earlier, much of what we have proved holds without assumption of compactness. We ask you to prove the appropriate generaliza- generalizations in the exercises that follow. One thing we do not ask you to prove is the fact that the topological dimension of an m-manifold is precisely m. And for good reason; the proof requires the tools of algebraic topology, which we have not studied. Nor do we ask you to prove that A' = 1m + 1 is the smallest value of N such that every compact metrizable space of topological dimension m can be imbedded in RN. The reason is the same. Even in the case of a linear graph, where m — I, the proof is nontrivial, as we remarked in Example 2 above. For further results in dimension theory, the reader is referred to the classi- classical book of Hurewicz and Wallman [H-W]. Exercises 1. Show that the Cantor set has dimension 0. 2. Show that any connected Hausdorff space having more than one point has dimension at least 1. 3. Show that the comb space and the topologist's sine curve have dimension 4. Show that the subspace ([0, 1] x 0) u {.v x y\y = x sin (l/.v) and 0<x<,\] of Rz has dimension 1. } 5. Show that the points 0,e,,ez,e3, and A, 1, 1) are in general ^ Sketch the corresponding imbedding into R3 of the complete vertices. 6. (a) Verify that {x0 x*} is geometrically independent If and only ' vectors x, - x0,..., x* - x0 are linearly independent.
§?"9 An Introduction to Dimension Theory 315 ^^fr=rr£^p-x:rinedbylxo ^"«-t—i.** (c) Prove that a ^-dimensional subspace V of R« has empty interior in *» 7. Complete the proof that every 2-complex can be imbedded in RK 8. Show that the points of the curve x = t y => t1 z = O in R\ Generalize to *. [Hint: Use theVander'monde SjJ P°SUi°n *9. (a) Prove the following: Theorem. Let X be a locally compact Hausdorff space having a countable basis. Suppose that every point of X has a neighborhood whose closure has tope logical dimension at most m. Then X has topological dimension at most m. [Hint: Write X = \JCm where Cn is compact and CD c Int C,+1. Generalize the proof of Theorem 9.2 to show that dim X <, m.] (b) Prove: Corollary. Every m-manifold has topological dimension at most m. (c) Prove: Corollary. Every closed subset of RN has topological dimension at most N. *10. Theorem. Let X be a locally compact Hausdorff space with a countable basis, having topological dimension m. Then X can be imbedded as a closed subset of Proof. (a) Let N = 2m + 1. Let d be a metric on X. Given a compact subset C of X and given e > 0, define U,(C) = {f\f 6 e(A", **<) and A(/1C) < e]. Show U,(C) is open and dense in (Q(X, R"), p), where p is the uniform metric. [Hint: Use the Tietze extension theorem.] (b) Show that given/ 6 C(A", /?") and given 5 > 0, there is a continuous injec- tive map g : X — RN such that #/,*) < d. [Him: Write *= UC,, where each €„ is compact and C, a CB+1 for each n. Consider ntfiMOJ (c) Let /• X —♦ RN The map / is said to be proper if for each compact set O in R», the set /-(£) is compact. Let X* and (*")* be the one-point.corn- pactifications of these two spaces; extend /to a map F: X -* (K ) oy letting F(co) = co. Show that /is proper and continuous if and only it t (e) Construct a proper theorem. 11. Assume the two preceding exercises. ^ IZem. Every m-manifold can be imbedded in *~" os a closed subset. (t>) IZem. A space X can be if and only if X is locally compact and Hau finite topological dimension.
8. The Fundamental Group and Covering Spaces One of the basic problems of topology is to determine whether two given topological spaces are homeoniorphic or not. There is no method for solving this problem in general, but there do exist techniques that apply in particular cases. Showing that two spaces are homeomorphic is a matter of constructing a continuous mapping from one to the other having a continuous inverse, and constructing continuous functions is a problem that we have developed techniques to handle. Showing that two spaces are not homeomorphic is a different matter. For that one must show that a continuous function with continuous inverse does not exist. If one can find some topological property that holds for one space but not for the other, then the problem is solved—the spaces cannot be homeomorphic. The closed interval [0, I] cannot be homeomorphic to the open interval @, 1), for instance, because the first space is compact and second one is not. And the real line R cannot be homeomorphic to the ong line" L, because R has a countable basis and L does not. Norca" esa line R be homeomorphic to the plane Rz; deleting a point from R2 lea connected space remaining, and deleting a point from R does not s But the topological properties we have studied up to now do not ca / 316
Chap. 8 The Fundamental Group and Covering Spaces 317 very far in solving the problem. For instance, how does one show that the plane R2 is not homeomorphic to three-dimensional space W> As one goes down the list of topological properties—compactness, connectedness, local connectedness, metrizability, and so on—one can find no topological'prop- topological'property that distinguishes between them. As another example, consider such surfaces as the 2-sphere S2, the torus T (surface of a doughnut), and the dou- double torus T2 (surface of a two-holed doughnut). None of the topological prop- properties we have studied up to now will distinguish between them. So we must introduce new properties and new techniques. One of the most natural such properties is that of simple connectivity. You probably have studied this notion already, when you studied line integrals in the plane. Roughly speaking, one says that a space A" is simply connected if every closed curve in A'can be shrunk to a point in X. (We shall make this more precise later.) The property of simple connectivity, it turns out, will distinguish between R2 and R3; deleting a point from R3 leaves a simply connected space remaining, but deleting a point from R2 does not. It will also distinguish between S2 (which is simply connected) and the torus T(which is not). But it will not distinguish between T and 7\; neither of them is simply connected. There is an idea more general than the idea of simple connectivity, an idea that includes simple connectivity as a special case. It involves a certain group that is called the fundamental group of the space. Two spaces that are homeomorphic have fundamental groups that are isomorphic. And the con- condition of simple connectivity is just the condition that the fundamental group of A'be the trivial (one-element) group. Thus the proof that S2 and Tare not homeomorphic can be rephrased by saying that the fundamental group of S2 is trivial and the fundamental group of T\% not. The fundamental group will distinguish between more spaces than the condition of simple connectivity will. It can be used, for example, to show that Jand T2 are not homeomor- homeomorphic; it turns out that T has an abelian fundamental group and Tz does not. In this chapter we define the fundamental group and study its properties. Then we apply it to a number of problems, including the problem of showing that various spaces, such as those already mentioned, are not homeomorphic. Other applications include applications to vector fields and fixed points and antipode-preserving maps of the sphere, as well as the well-known fun- fundamental theorem of algebra, which says that every polynomial equation with real or complex coefficients has a root. Finally, there is the famous Jordan curve theorem, to the effect that every simple closed curve C in the plane separates the plane into two components, of which C is the common bound- boundary. Some of the sections of this chapter are independent of one another. The dependence among them is expressed in the following diagram:
318 The Fundamental Group and Covering Spaces Chap, g §8-1 Homotopy of paths ... §8-2 The fundamental group ' • ' §8-3 Covering spaces ' ' \\ -§8-4 The fundamental group of the circle §8-5^ The fundamental group of the punctured plane i-6 The fundamental group of S" i-7 Fundamental groups of surfaces Essential and inessential maps k SS-9 The fundamental theorem of algebra \\ §8-10 Vector fields and fixed points §8-11 Homotopy type §8-12 The Jordan separation theorem §8-13 The Jordan curve theorem §8-14 Classification of covering spaces Throughout the chapter, we assume familiarity with components and local connectivity (§3-3 and §3-4). We also assume that the reader is familiar with the elements of group theory. In §8-12, we assume familiarity with local compactness (§3-8), and in §8-13, with quotient spaces (§2-11). 8-1 Homotopy of Paths Before defining the fundamental group of a space X, we shall consider paths on JT and an equivalence relation called path homotopy between the ■ And we shall define a certain operation on the collection of these equivaien classes that makes it into what is called in algebra a groupoid. Definition. If/and/' are continuous maps of the space JTinto *espa<* Y, we say that/is homotopic to /' if there is a continuous map F: X X
6 8-1 ■* ysuch that Homotopy of Paths 319 and F{x,\)=f'{x) for each x e X. (Here / = [0, ]].) The map Fis called a homotopy between/ and / . If/is homotopic to /', we write/~ /'. J We think of a homotopy as a continuous one-parameter family of maps from X to Y. If we imagine the parameter / as representing time, then the homotopy F represents a continuous "deforming" of the map/to the map /', as / goes from 0 to 1. Now we consider the special case in which/is a path in X. Recall that if /: [0, 1] —» X is a continuous map such that /@) = x0 and /(I) =*,, we say that/is a path in Xfrom x0 to *,. We also say that x0 is the initial point, and Xi the final point, of the path/ In this chapter, we shall for convenience use the interval / = [0, 1] as the domain for all paths. If/and /' are two paths in X, there is a stronger relation between them than mere homotopy. It is defined as follows: Definition. Two paths / and /', mapping the interval / = [0, 1] into X, are said to be path homotopic if they have the same initial point x0 and the same final point .v,, and if there is a continuous map F: I X /—> A'such that F(s, 0) --= f(s) and F(s, 1) = f'{s), F@,t)=xo and F{\,t)=xu for each s e /and each / e /. We call Fa path homotopy between/and/'. See Figure 1. If/is path homotopic to /', we write/~, /'. The first condition says simply that F is a homotopy between/and /', and the second says that for each /, the path Figure 1
320 The Fundamental Croup and Covering Spaces Chap. 8 is a path from x0 to *,. Said differently, the first condition says thatF sents a continuous way of deforming the path / to the path /', and the s^"*" condition says that the end points of the path remain fixed during the defo0^ Lemma J.I. The relations ^ and c±p are equivalence relations. If/is a path, we shall denote its path-homotopy equivalence class by [/] Proof. Let us verify the properties of an equivalence relation. Given/, it is trivial that/-/; the map F(x, t) = /(*) is the required homotopy. If/is a path, /"is a path homotopy. Given/^ /', we show that /' ~/ Let F be a homotopy between/and /'. Then G(x, t) = F(x, I — r) is a homotopy between /' and/ If Fis a path homotopy, so is G. Suppose that/~ /' and /' ~ /". We show that/~ /". Let F be a homotopy between/and /', and let F' be a homotopy between /' and/". Define G : X X I —> Y by the equation (F(x, 2t) for t e [0, {], \f\x. It - I) for t e [', I]. The map G is well defined, since if / = \, F(x, 2t) = F(x, I) = f\x) = F'(x, 0) = F'{x, 2r - 1). Because G is continuous on the two closed subsets X X [0, £] and X x [{, I] of X x /, it is continuous on all of X X 7, by the pasting lemma. Thus G is the required homotopy between /and /". You can check that if F and F' are path homotopies, so is G. Figure 2 illustrates this situation. □ Example 1. Let / and g be any two maps of a space X into R2. It is easy to see that/and g are homotopic; the map G(x, t) = Figure 2
8-1 is a homotopy between them. Homotopy of Paths 321 Z222£5S£L If/and g are paths from xB to v th • can check. This situation is pictured in F^uiT" * " *** h°moto'*.« you figure 5 Figure 4 Example 2. Let X denote the punctured plane, & - [0], which we shall denote by & - 0 for short. The following paths in X, f(s) = (cos ns, sin ns), g(s) = (cos ns, 2 sin 7w) are path homotopic; the straight-line homotopy between them is an acceptable path homotopy. But the straight-line homotopy between/and the path h(s) = (cos ns, —sin ns) is not acceptable, for its image does not lie in the space X = R2 - 0. See Figure 4. Indeed, there exists no path homotopy in X between paths/and h. This result is hardly surprizing; it is intuitively clear that one cannot "deform/past the hole at 0" without introducing a discontinuity. But it takes some work to prove. We shall return to this example later. This example illustrates the fact that you must know what the range space is before you can tell whether two paths are path homotopic or not. The paths /and h would be path homotopic if they were paths in R1. Now we introduce some algebra into this geometric situation. We define a certain operation on path-homotopy classes as follows:
322 The Fundamental Croup and Covering Spaces * Chap. 8 Definition. If/is a path in * from *0 to *„ and \t g is a path in Xr to Xi, we define the composition/* g of/and g to be the path h eivenT the equations y n by h{s) = \f^ for,e[0,i], UBj-I) forje[|, I]. The function A is well-defined and continuous, by the pasting lemma- and it is a path in * from x0 to xz. We think of h as the path whose first half is the path/and whose second half is the path g. We shall show that the operation of composition on paths induces a well- defined operation on path-homotopy classes, so that we can define [/]♦[*] = [/♦*]. Furthermore, the operation ♦ on path-homotopy classes turns out to satisfy properties that look very much like the axioms for a group. They are called the groupoid properties of *. The only difference from the properties of a group is that [/] * [g] is not defined for every pair of classes, but only for those pairs [/], [g] for which /(I) = g@). Theorem 1.2. The operation * is well-defined on path-homotopy classes. It has the following properties: A) (Associativity) If[t] * ([g] * [h]) is defined, so is ([f] ♦ [g]) ♦ [h] and they are equal. B) (Right and left identities) Given x e X, let ex denote the constant path e,. : I — ► X carrying all of I to the point x. If'fis a path in Xfrom x0 to x,, then [n*[e,,] = [f] and [ej ♦ [f] = [f]. C) (Inverse) Given the path f in Xfrom x0 to x,, let Tbe the path definedby f(s) = f(l — s). It is called the reverse o/f. Then [f]*[f] = [ej and [f] * [f] = [ex,]. Proof Each of the preceding statements has an elementary geometric proof. To show the operation well-defined, let F be a path homotopy between /and /' and let G be a path homotopy between g and g'. Define (FBs, t) for s e [0, \], m') = \GBs-~ I,/) for^e & I]. Because F(I,/) = *.= G@, f) for all t, the map H is well-defined; it is con- continuous by the pasting lemma. You can check that H is the required pa homotopy between/* g and /' * g'. It is pictured in Figure 5. A) To prove associativity, we need to show that/* ^*^({*,fr*/» What exactly, do the paths /•(*♦*) and (/* *) * A look I.k£ V^ (g * A) the image of the point s traces out the .mage of/as i goes rom £ it traces out the image of g as , goes from i to |, and it traces out the .mag
§8-1 Homotopy of Paths 323 Figure 5 of h as j goes from \ to I. Under (/♦ g) ♦ h, the point traces out the same image, but at a different rate. It traces out the image of/as s goes from 0 to \, of g as s goes from \ to £, and of/; as s goes from ^ to 1. We define the homotopy F as follows: First map the unit square P onto / by the continuous map p which collapses each of the three quadrilaterals A, B, and C in Figure 6 onto its base. Follow p by the map (/♦ g) * h. The result will be the desired path homotopy F, as you can check mentally. More formally, define fDsl(t + 1)) for s e [0, (/ + l)/4], F(s, 1) = gDs -t-l) for s e [(t + I)/4, (/ + 2)/4], /i(Dj - / - 2IB - /)) for s e [(t + 2)/4, I]. You can check that 4s/0 + 1), 4s - t - I, and Dj - / - 2)/B - /) lie in the domain [0, 1] of/ g, and A when they are supposed to, so these formulas make sense, and that F is well-defined and hence continuous by the past.ng lemma. It is straightforward to show that F is the required path homotopy.
324 The Fundamental Group and Covering Spaces For instance, Chap, if* (g * A)X») = fBs) for * 6 [0, f/Bi) \(g * h){2s - 1) for s e [£, 1] or 2s - 1 e [o, \] Thus we have for*e[0,£], I) for 2s- 1 6 [O.iJorie^,^ [hBBs - I) - I) for 2s - 1 e [J, 1] or j 6 [|, i]. And this is just F(s, 1), as you can check. Similarly, ((/♦ g) * h)(s) =F(s, 0) B) We prove thatf~pf* ext. The desired path homotopy is constructed as follows: First map the unit square I2 onto / by a continuous mapp that collapses the quadrilateral A of Figure 7 onto its base, and collapses the Figure 7 triangle B to its bottom vertex. Follow p by the map / The result is the desired path homotopy G, as you can check mentally. Formally, define _ |/Bj/B - /)) for s e [0, B - 0/2], <7(*'')-'U for. e [B -0/2,U- You can check that G is well defined and continuous, and is the desired pa homotopy. The proof that ext *f~pf\s similar. C) Now we show that/*/~, en. The intuitive idea °f theProoff,2 simple. The path/*/is a path that goes from x0 to *, and then comes oa to x0 along the same route. Suppose that, for parameter value-I, «* '« ' the path which goes out from x0 partway along the route of/ as ta ^ point /(/). Then the homotopy we want is just the path «, * «,^ ceptable path homotopy, because it leaves both end points fixed at*- Figure 8.
§8-1 ' Homotopy of Paths 325 Figure 8 Formally, we define \fBt(l-s)) It is easy to check that Its and 2t{\ — s) lie in the domain of/ when they are supposed to, so the formulas make sense, that H is well-defined (and hence continuous by the pasting lemma), and that H is the required path homotopy between eXt and / * /. A similar argument could be used to show that/*/~, exr But better yet, note the following: We have shown that for any pathg, wehaveg*g~, ex, where x is the initial point of g. In particular,/*/^, ex,, where/is the reverse of/. But the reverse of/ is just/! Thus/*/— ,exi, as desired. □ Exercises 1. Show that ith,h':X-+Y are homotopic and k, k': Y^Zm homotopic, then k ° h and k' ° h' are homotopic. 2. Suppose that X is a convex set in R"; that is, for every pair x,y of points of X the line segment joining them lies in X. Show that any two paths in JThavmg the same end points are path homotopic. 3. Consider the proof of Theorem 1.2. _ (a) Check that f (,, 0) = ((/ * g) * *X*) and F@,t) =■ *. andI/tU) - ^,- (b) Check the statements made about the maps G and tf in B) and C). 4. Given spaces X and Y, let [X, n denote the set of homotopy classes of maps « Uta 11. Show that for any *,the set M^ a (b) Show that if r is path connected, the set [/, Yl has a 5. A space Xis said to be contracfible if the identity map ,x. X - X to a constant map. (a) Show that / and R are contractible.
326 The Fundamental Group and Covering Spaces Chap. 8 (b) Show that a contractible space is path connected (c) Show that if Y is contractible, then for any X, the set IX n h« ■ element. ' J1MS> s'ngle (d) Show that if X is contractible and Y is path connected then [ single element. L 5-2 77re Fundamental Group The set of path-homotopy classes of paths in a space X does not form a group under the operation *, only a groupoid. But suppose we pick out a point .v0 of X to serve as a "base point" and restrict ourselves to those paths that begin and end at x0. The set of these path-homotopy classes does form a group under *. It will be called the fundamental group of X. In this section we shall prove several properties of the fundamental group. In particular, we shall show that the group is, up to isomorphism, independent of the choice of base point (provided that X is path connected). We shall also show that the group is a topological invariant of the space X, the fact that is of crucial importance in using it to study homeomorphism problems. Definition. Let X be a space; let x0 be a point of X. A path in X that begins and ends at x0 is called a loop based at x0. The set of path homotopy classes of loops based at x0, with the operation *, is called the fundamental group of X relative to the base point .v0. It is denoted by n\{X, x0). It follows from the preceding theorem that the operation *, when restrict- restricted to this set, satisfies the axioms for a group. Given two loops/and g based at x0, the composition/* g is always defined and is a loop based at *o- As- Associativity, the existence of an identity element [<?,,], and the existence of an inverse [/] for [/] are immediate. Sometimes this group is called the first homotopy group of X, which term implies that there is a second homotopy group. There are indeed groups nn(X, x0) for all n e Zt, but we shall not study them in this book. They are part of the general subject called homotopy theory. Example 1. Let R" denote euclidean n-space. Then ni(R",x0) is group (the group consisting of the identity alone). For if /is a loop m Ji at x0, the straight-line homotopy F(s, t) = tx0 + A - t)f(s) is a path homotopy between/and the constant loop ex,. Example 2. More generally, if Xis any ™^ subset °f *" is the trivial group. The straight-line homotopy «f «^ «"? vexity of X means that for any two points x and>> of X, the
8'2 The Fundamental Group 327 between them lies in X. In particular, the unit ball Bn in R» B" = {x\x\ + ... +xi^,l], has trivial fundamental group. To find a space with a nontrivial fundamental group is more difficult; this we do in §8-4. An immediate question one asks is the extent to which the fundamental group depends on the base point. The answer is given in Corollary 2.2, which follows. Definition. Let a be a path in X from x0 to *,. We define a map a : n,(X, x0) —> n^X, *,) by the equation «(M) = [«] * [/] * W. The map a is pictured in Figure 9. It is well-defined because the operation ♦ is well-defined. If/is a loop based at ,v0, then a ♦ (/♦ a) is a loop based at xy. Hence a maps 7ii(X, xa) into ni(X, xt), as desired. Figure 9 Theorem 2.1. The map a, is a group isomorpliism. Proof. To show that a is a homomorphism, we compute = ([«] * I/] * W) * (W * [8] * = [a] This proof uses the groupoid properties of *, which we proved in Theorem 1.2. To show that & is an isomorphism, we show that if fi denotes the path a,
328 The Fundamental Group and Covering Spaces Chap. 8 which is the reverse of a, then j? is an inverse for &. We commite r element [h] of n,{X, *,), compute, for each L»] * [h] * [fi] = [a] * [A] * [a], = [a] * ([a] * [A] * fa]) * fa] = \h]. A similar computation shows that j9(<2([/])) = f/] for each f/] e s,( Corollary 2.2. If X As /wA connected and x0 <wrf x, aw /wo /,0/nw o/ X n 7ti(X, x0) As isomorphic to re,(X, x,). ' Suppose that X is a topological space. Let C be the path component of ^containing x0. It is easy to see that Sl(C, *0) = *,(*, *„), since all loops and homotopies in X that are based at x0 must lie in the subspace C. Thus 7i\(X, x0) depends only on the path component of Xcontainingx0, and gives us no information whatever about the rest of X. For this reason, it is usual to deal only with path-connected spaces when studying the fundamental group. If X is path connected, all the groups n^X, x) are isomorphic, so it is tempting to try to "identify" all these groups with one another, and to speak simply of the fundamental group of the space X, without reference to base point. The difficulty with this approach is that there is no natural way of identifying n,(.X, x0) with n,(X, x,); different paths a and fi from xa to x, may give rise to different isomorphisms between these groups. For this reason, omitting the base point can lead to error. It turns out that the isomorphism of 7ti(X, .v0) with n,(X, x,) is indepen- independent of path if and only if the fundamental group is abelian. (See Exercise 2.) This is a stringent requirement on the space X. Definition. A space X is said to be simply connected if it is a path-connected space and if n^X, x0) is the trivial (one-element) group for some xa& X, and hence for every .v0 e X. We often express the fact that n^X, x0) is the trivial group by writing n,(X, x0) = 0. Lemma 2.3. In a simply connected space X, any two paths having the same initial and final points are path homotopic. Proof. Let/and g be two paths from xa to *,. Then/* g is defined and is a loop on X based at xa. Since X is simply connected,/* £—,*,.. Appiy»>s the groupoid properties, we see that f( f * Sr\ * frl — \p * ff] = \s I. Ivy * 5) oj 1 -*• oJ loJ [(/*£) * g] = lf*(i*g)] = If* e*J = [-fl- Thus/and g are path homotopic. □ It is intuitively clear that the fundamental group is a topoloi of the space X. A convenient way to prove this fact formally i Se notion of the "homomorphism induced by a continuous map.
§ 8-2 * ,o .he poi., „ „, r we .CSS £«* * P"»,,. of ing „(*. ,.) into ^Y,ya). We define it fl^fcST ""* Definition. Let h : W ,o) _ (y> by the equation *:*'<**.) **(!/]) = [A o/]. The map A, is called the homomorphism induced by A relative to n, k point x0. y • relal've to the base homotopy between the loops h o/and /, „/'. It is a)so eaSy * che kthatl is a homomorphism. For y hat "* It follows that - 1)) forj e [|, 1]. Thus h°(f*g) equals the composition (/; o /) * (/, o ^). it follows that so that ht is a homomorphism. The induced homomorphism /i* depends not only on the map h : X—> Y but also on the choice of the base point xa. (Once xa is chosen, ya is deter- determined by /;.) So some notational difficulty will arise if we want to consider several different base points for X. If x0 and xt are two different points of A", we cannot use the same symbol /;„ to stand for two different homomorphisms, one having domain n^(X, x0) and the other having domain tzi(X, xt). Even if AT is path connected, so these groups are isomorphic, they are still not the same group. In such a case, we shall use the notation (A..),:*i(*,*o)—>^(Y,y0) for the first homomorphism and (hXl)t for the second. If there is only one base point under consideration, we shall omit mention of the base point and denote the induced homomorphism merely by ht. The induced homomorphism has two properties that are crucial in the applications. They are called the "functorial properties" of the induced homomorphism, and are given in the following theorem:
330 The Fundamental Group and Covering Spaces Chap. 8 Theorem 2.4. If h : (X, x0) -> (Y, y0) and k : (Y, y0) _> (Z 7 \ (k o h). - k. . h.. //i: (X, x0) - (X, x0) is the identity map, Lf' identity homomorphism. * / The proof is a triviality. By definition, (k ° hU[f]) = [(k ° h) ° f], (K ° **)([/]) = kt(h,([/])) = *,([* o /]) = [/t o (h of)]. Similarly, /,([/]) = [i o /] = [/]. rj Corollary 2.5. Ifh: (X, x0) — (Y, y0) ;j a homeomorphism of X h+ ij an isomorphism of m(X, Xo) wM re,(Y, y0). Proof. Let & : (Y, ya) —> (X, x0) be the inverse of h. Then lct°ht = (k o £)„ = /„, where i is the identity map of (X, x0); and h\ o &„ = (A 0*% =/„, where/ is the identity map of ( Y, y0). Since /„ and/,, are the identity homomorphisms of the groups n^X, x0) and ni(Y,ya), respectively, kt is the inverse of ht. □ Exercises 1. A subset /( of Rn is said to be star convex if for some point a0 of A, all the line segments joining a0 to other points of A lie in A. (a) Find a star convex set that is not convex. (b) Show that if A is star convex, A is simply connected. (c) Show that if A is star convex, any two paths in A having the same initial and final points are path homotopic. 2. Let .v0 and .v, be two given points of the path-connected space X. Show that 7ii(X, .v0) is abelian if and only if for every pair a and 0 of paths from x0 to a,, we have a = fi. 3. Let A c X and let r : X —> A be a retraction (see Exercise 7 of §4-3). Given a0 e A, show that r* : Tti(X, aa) —> iii(A, a0) is surjective. [Hint: Consider the inclusion/: (A, a0) —► (X, a0)] 4. Let A be a subset of R"; let h ■ (A, a0) —> (Y, y0). Show that if h is extendable to a continuous map of R" into Y, then h* is the zero homomorphism (the tnvi homomorphism that maps everything to the identity element). 5. Let h : X —> Y be continuous. Show that if X is path connected, the homoI"^'e phism induced by h is independent of base point, up to isomorphisms or groups involved. More precisely, if h(x0) = y0 and h(x,) -*.jf™ there are isomorphisms 0 and y/ such that the following diagram "commutes":
6 8-3 ■ „ * Covering Spaces 331 that is, such that V o (A,,), = (^,), 0 0. Conclude that if (A, v is the zero homomorphism (or injective, or surjective), then so is (hx,)t. 6. Let G be a topological group with operation • and identity element x0 (See the Supplementary Exercises for Chapter 2.) Let £2@, *0) denote the set of all loops in G based at x0. Iff, g e D.(G, *„), let us define a loop /® *■ by the rule (a) Show that this operation makes the set Q(G, *0) into a group. (b) Show that this operation induces a group operation ® on 71,(G, x0). (c) Show that the two group operations * and ® on 7t,(G,x0) are the same. [Hint: Compute (/* ex.) ® (ex, * g).] (d) Show that 7it(G, x0) is abelian. 8-3 Covering Spaces We have not yet computed a nontrivial fundamental group. We shall remedy that deficiency in the next section, when we compute the fundamental group of the circle S'. We need first to introduce the notion of a covering space. For us, covering spaces will serve mainly as a tool for computing fundamental groups. But they are important in their own right, particularly in the study of Riemann surfaces and complex manifolds. (See [A-S].) Definition. Let p: E—> B be a continuous surjective map. The open set U of B is said to be evenly covered by p if the inverse image p~'(U) can be written as the union of disjoint open sets V. in E such that for each a, the restriction of p to Va is a homeomorphism of Fa onto U. The collection {Fj will be called a partition of p~\U) into slices. If U is an open set that is evenly covered by p, we often picture the set P'\U) as a "stack of pancakes," each having the same size and shape as C/, floating in the air above U; the map p squashes them all down onto U. See Figure 10. Definition. Let p:E-> B be continuous and surjective. If every point bofB has a neighborhood U that is evenly covered by p, then p is called a covering map, and E is said to be a covering space of B. Note that if p : £ — B is a covering map, then for each beB the subset P-I(f>) of E necessarily has the discrete topology. For each slice V. is open
332 The Fundamental Group and Covering Spaces Chap, P 1 U) Figure 10 E and intersects the set p~' (b) in a single point; therefore this point is open in the subspace topology on p~l(b). Example J. Let A' be any space; let /: X —-> X be the identity map. Then / is a covering map (of the most trivial sort). More generally, let E be the space X x {1,.. ., n] consisting of n disjoint copies of X. The map p:E—*X given by p{xt i) — x for all / is again a (rather trivial) covering map. In this case, we can picture the entire space £asa stack of pancakes over X. To eliminate trivial coverings of the pancake-stack variety, one often requires the space E to be connected. A useful example of a covering of this sort is given in the following theorem: Theorem 3.1. The map p: R —* SJ given by the equation p(x) = (cos 27ix, sin 2nx) is a covering map. One can picture p as a function that wraps the real line R around the circle S1, and in the process maps each interval [n, n + 1] onto Sl. Proof. The fact that/? is a covering map comes from elementary propertws of the sine and cosine functions. Consider, for example, the subset V ot consisting of those points having positive first coordinate. The set p I consists of those points x for which cos 2nx is positive; that is, it is the u of the intervals
8-3 Covering Spaces 333 Figure II for all n e Z. See Figure 11. Now, restricted to any closed interval Vn, the map/p is injective, because sin 2nx is strictly monotonic on such an interval. Furthermore, p carries Vn surjectively onto U, and Fn to U, by the interme- intermediate value theorem. Sip.ce Vn is compact, p\ Vn is a homeomorphism of K, with U. In particular, p\Vn\s& homeomorphism of Vn with U. Similar arguments can be applied to the intersections of S1 with the upper and lower open half-planes, and with the open left-hand half-plane. These open sets cover Sl, and each of them is evenly covered by p. Hence p : R—» Sl is a covering map. [J] Example 2. Consider the space T= S' x S1; it is called the torus. It is a general fact that the product of two covering maps is a covering map. (See Exercise 2.) Therefore, as pictured in Figure 12, the product p x p:Rx R—►S1 x 5' is a covering of the torus by the plane R>, where p denotes the covering map of pXp Figure 12
334 The Fundamental Group and Covering Spaces ~ v-hap. g Theorem 3.1. Each of the unit squares [n, n + 1] x [m, m + 1] gets wraoned by p x p entirely around the torus. In this figure we have pictured the torus not as the product Sl x 51 ^ h is a subspace of R* and therefore hard to visualize, but as the familiar 'dough nut-shaped surface D in R3 obtained by rotating the circle C, in the^rz-plane of radius j centered at A, 0, 0) about the z-axis. It is not hard to see that S1 x S> k homeomorphic with the surface D. Let C2 be the circle of radius 1 in ihexy- plane centered at the origin. Then let us map Cx x C2 into D by defining f(a x b) to be that point into which a is carried when one rotates the circle C\ about the z-axis until its center hits the point b. See Figure 13. The map/ will be a homeomorphism of Ci x C2 with D, as you can check mentally. If you wish, you can write equations for /and check continuity, injectivity, and surjectivity directly. (Continuity of /"' will follow from compactness of C, x C.) Figure 13 Example 3. Consider the covering map p x i:Rx R*—*$' x R+> where / is the identity map of *+ and p is the map of Theorem "^ the standard homeomorphism of 5' x R. wi.h IP - 0, sendmg the composite gives us a covering R x R+ —+ R1 - 0 of the punctured plane by the open upper half-Pla-Itis pfctured in This covering map appears in ihe study of comP^J«™Wes surface corresponding to ihe complex logar.thm funct.on If p -.E^Bis, covering map, then p is with B. That is, each point e of E has a ^f as the following example shows.
§8-3 - Covering Spaces 335 I—X RXR. \ Figure 14 Example 4. The map p: R+ —> S1 given by the equation p(x) — (cos 2nxy sin 2nx) is surjective, and it is a local homeomorphism. See Figure 15. Bui it is not a covering map, for the point b0 = A, 0) has no neighborhood U that is evenly covered by p. The typical neighborhood U of b0 has an inverse image consist- consisting of small neighborhoods Vn of each integer n for n > 0, along with a small interval Vo of the form @, e). Each of the intervals Vn for n > 0 is mapped homeomorphically onto U by the map p, but the interval Vo is only imbedded in V by p. Figure 15 Example 5. The preceding example might lead you to think that the real line J? is the only connected covering space of the circle S1. This is not so. Consider, for example, the map p: Sl — Sl given in equations by PV) = *2- [Here we consider S • as the subset of the complex plane C consisting of those complex numbers z with | z | = 1.] We leave it to you to check that p is a covering map.
336 The Fundamental Group and Covering Spaces Exercises \ 1. Show that if p:E—* B is a covering map, then p is an open map. 2. Show that if p:E—* B and p': E'—* B' are covering maps, then p X p': E X E' —* B X B' is a covering map. 3. Let Y have the discrete topology. Show that if p : X x Y —> X is proiectio onto the first coordinate, then p is a covering map. n 4. Show that the map p of Example 5 is a covering map. Generalize to the ma p(z) = z". P 5. Let p : E —♦ B be a covering map; let B be connected. Show that if/r'F0) has k elements for some b0 e B, then p~{(b) has k elements for every b e B In such a case, E is called a A-fold covering of A 6. Let />: A1 —► Y and 9 : Y —► Z be covering maps. (a) Show that \{q~\z) is finite for each z e Z, then^o^; A'—>Zis a covering map. *(b) Show that the theorem fails if q~l(z) is not finite. 7. Let p: E —> B be a covering map. Assume that B is connected and locally connected. Show that if C is a component of E, then /> | C: C —> B is a covering map. 8. Write equations for the map f:C\ x C2 —► £> of Example 2 and check that it is a homeomorphism. ITte Fundamental Group of the Circle The study of covering spaces of a space X is intimately related to the study of the fundamental group of X. In this section, we establish the crucial links between the two concepts, and compute the fundamental group of the circle. Definition. Let p : E — B be a map. If/is a continuous mapping of some space JTinto B, a lifting of/is a map/: JT-+ £ such that/>»/=/ The existence of liftings when /»is a covering map is an important ^ studying covering spaces and the fundamental group. First, we slJ0* for a covering space, paths can be lifted; and then we show that path n topies can be lifted as well. First, an example:
The Fundamental Group oj the Circle 337 31" Example 1. Consider the covering p ■ R /: [0,1] -> Jt beginning at b0 = Af 0* *£ I? to the path/to = ,/2 beginning at 0 and ending at { ThlnS ^ -sin ? lifts to the path #W = _j/2 beginni^ a O^nd'e "dinlr ^ path A(j) = (cos 4ns, sin 4u) lifts to the path iti* - , ™ .ng.at ~i- The ending at 2. Intuitively, * wraps the !„«[ Iround X^8 ,* ° ?* this is reflected in the fact that the lifted path /beg ns Ttzl h ,' ^ number 2. These paths are pictured in Rgure 16 Md ^ &t the Figure 16 Lemma 4.1. Let p : E —► B be a covering map; let p(e0) = b0. Any path f: [0, 1] —* B beginning at b0 has a unique lifting to a path f in E beginning at e0- Proof. Cover B by open sets U each of which is evenly covered by p. Find a subdivision of [0, 1], say sa,. .. , sn, such that for each / the set f([s,, s,+i]) lies in such an open set U. (Here we use the Lebesgue number lemma.) We define the lifting / step by step. First, define /@) = ea. Then, supposing f(s) is defined for 0 < s < s,, we define/ on [s,, sl+i] as follows: The set f([s,, st+i]) lies in some open set U that is evenly covered by p. Let {Va} be a partition of p~'(tf) into slices; each set F. is mapped homeomorphically onto U by p. Now f(st) lies in one of these sets, say in Vo. Define f(s) for s e [s,, sl+i] by the equation /M = (pl^r'(/(*))• Because p | Vo: Fo -» {/ is a homeomorphism, / will be continuous on [■Sj, £j+]]. Continuing in this way, we define /on all of [0,1]. Continuity of/follows from the pasting lemma of §2-7; the fact that p ° f =/ is immediate from the definition of/. ». The uniqueness of/ is also proved step by step. Suppose that/is another Kfting of/ beginning at e0. Then /@) = e0 = /@). Suppose that/W -/W for all , such that 0 ^ s <: s,. Let Ko be as in the preceding paragraph; then f°r ^ [,,, ,/+J,/(,) fs defined as (^ Fo)-'(/W). What can/(,)equa^? Since /is a lifting of/, it must carry the interval [,„ *,+ 1] into the setp- (U) - U»V
338 The Fundamental Group and Covering Spaces Chap. 8 The slices V. are open and disjoint; because the set ?(\s, , i\;« it must lie entirely in one of the sets V.. Because /(I) J}(^ 1™™*' V.. 1 must carry all of W ,,+1] into the set Vo. Thus for /£tisT i "I* must equal some point j> of Fo lying in/,-•(/•(,)). But there is oriv'o.J. u point* namely, Q,| Ko)-(/W)- Hence/W = /(*) for, e [,„ J,J "* Lemma 4.2. Let p : E —► B be a covering map; let p(e0) = b Let th map F:IxI-^B be continuous, with F@, 0) = b0. There is a lifting of I to a continuous map F:I x I—>E such that F@, 0) = e0. 7/ F is a path homotopy, then F is a path homotopy The lifting F of F is, in fact, unique, but this we shall not prove. Proof. Given F we first define F@, 0) = e0. Next, we use the preceding lemma to extend F to the left-hand edge Ox/ and the bottom edge/ x 0 of 7x7. Then we extend F to all of 7 x 7 as follows: Choose subdivisions s0 < si < • ■ ■ < sm, «„<«, <■■■<«„ of 7 fine enough that each rectangle I, XJ, = [Si-uS,] X [tj-i.tj] is mapped by JF" into an open set of B that is evenly covered by/>. (Use the Lebesgue number lemma.) We define the lifting F step by step, beginning with the rectangle 7, x Ju continuing with the other rectangles I, X Jt in the "bottom row"; then with the rectangles I, X J2 in the next row, and so on. In general, given /„ andy'o, assume that F is defined on the set .4 which is the union of 0 x 7 and 7 x 0 and all the rectangles "previous" to I,, xJj. (those rectangles 7, x J, for which j < j0 and those for which; ==ja and / < i0). Assume also that F is a continuous lifting of F\ A. We defineF on7,, X /,,. Choose an open set U of B that is evenly covered by p and contains the set F(Iia x 7Jo).Let{Fa} be a partition of p"'(^) into slices; each set Fa is map- mapped "homeomorphically onto V by p. Now F is already defined on the se C = A n (/,. X ■/,„). This set is the union of the left and bottom edges the rectangle /,„ X Jh, so it is connected. Therefore, F(C) is connected, a^ must lie entirely within one of the sets K«. Suppose it lies in Vo. The situation is as pictured in Figure 17. _ . Let pQ: Vo — V denote the restriction of p to Vo. Since F is a JF"J A, we know that for x e C, so that F(x) ^p-^CFCx)). Hence we may extend F by defining
§8-4 The Fundamental Croup of the Circle 339 '■OXJ'o Figure 17 for x e Jit X Jh. The extended map will be continuous by the pasting lemma. Continuing in this way, we define F on all of/2. Now suppose that F is a path homotopy. We wish to show that F is a path homotopy. The map /"carries the entire left edge 0 X /of/2 into a single point b0 of B. Because F is a lifting of F, it carries this edge into the set/r '(Z>0). But this set has the discrete topology as a subspace of E. Since 0 x /is con- connected and F is continuous, F@ x /) is connected and thus must equal a one-point set. Similarly, /(I x /) must be a one-point set. Thus F is a path homotopy. fj The following theorem establishes the crucial link between covering spaces and the fundamental group: Theorem 4.3. Let p: E —• B be a covering map; let p(e0) = b0. Let f and g be two paths in Bfrom b0 to b,; fe/ F<»«/ I be their respective liftings to paths in E beginning at e0. If f and g are path homotopic, then f and send at the same point of E and are path homotopic. Proof. Let F: I x /-> B be the path homotopy between /and ^ Then 0, 0) = *0. Let F : / X / - £ be a lifting of f to £ such that P^, 0) - ^ the preceding lemma, / is a path homotopy, so that F(O X /; X /) is a one-point set fe,]. _ r The restriction f | / x 0 of F to the bottom edge of/ x i at <■„ that is a lifting off | / X 0. By uniqueness of 7 "■ homotopy between them. Q
340 The Fundamental Group and Covering Spaces Chap, g Now we apply this theorem to the case of 5'. Theorem 4.4. The fundamental group of the circle is infinite cyclic. Proof. Let b0 be the point A, 0) of 5'. We shall construct an isomnmi,- of the group *,(£', b0) with the group (Z, +) of integers. 1SomorPhlsm For this purpose, consider the covering map p:R—*sl given bv th equation p(x) = (cos 2nx, sin 2nx). If/ is a loop on 5' based at 6 \£ft the lifting of/to a path on R beginning at 0. The point /(I) must be a noi t of the set /T'(*>o); that is, /(I) must equal some integer n. The preceding theorem tells us that this integer depends only on the path homotopy class of/. Therefore, we can define </>:n1(S\b0)—+Z by letting 0([/]) be this integer. We assert that 0 is a group isomorphism. The map </> is surjective. Let n be a point of /r >(Z>0). Because R is path con- connected, we can choose a path /: [0, I] —► R in R from 0 to n. Define/= p o /. Then / is a loop in S1 based at Z>0, and / is its lifting to a path in R beginning at 0. By definition, 0([/]) = n. The map 0 is injective. Assume that 0([/]) = n = 0([g]); we shall prove that [/] = [g]. Let / and g be the liftings of/and g, respectively, to paths on R beginning at 0; both / and g end at n, by hypothesis. Because R is simply connected, / and g are path homotopic; \ziF be the path homotopy between them. The map F = p ° F will be a path homotopy between/and g, as you can check. The map 0 is a homomorphism. Let/and g be two loops in S' based at b0; let / and g be their liftings, respectively, to paths on R beginning at 0. Let /(I) = n and g(l) = m. Define a path h on R by the equations If Bs) forje[fl,i], "{S) \n + gBs - I) for s e [{, I]. Then h is a path on .R beginning at 0. We assert that A is a lifting of/* g. Note first that for all x, we have p(n + x) = /»(*), because the functions sine ana cosine have period 2n. Then jrtA*))-/(*) ^r, e [0,1 Thus poh=f* g, so that A is the lifting of/* g which begins at 0. definition, <f,([f*g]) is A(I), which equals n + m. Therefore, ) as desired. □ Most of the proof of the preceding theorem f connected covering spaces. The only part specal to
§ 8 ^ . The Fundamental Group oj the Circle 341 was the existence of the addition operation in R; that operation enabled us t0 show that 0 was a homomorphism. In a general covering space we have no convenient addition operation; nevertheless, we can still get a good bit of information about the fundamental group. Theorem 4.5. Let p : (E, e0) — (B, b0) be a covering map. If E is path connected, then there is a surjection IfE is simply connected, 0 is a bijection. Proof. The proof is almost a copy of the proof of Theorem 4.4; we leave the details to you. □ Definition. If £ is a simply connected space, and if p : E—► B is a covering map, then we say that £ is a universal covering space of B. If B has a universal covering space £, then £ is uniquely determined up to homeomorphism (provided B is locally path connected). See Exercise 12. The reason we call £ a universal covering space is given in the same exercise. We shall study in §8-14 conditions under which a space B possesses a universal covering space. Exercises 1. What goes wrong with the "path-lifting lemma" (Lemma 4.1) for the local homeomorphism of Example 4 of §8-3? 2. In defining the map F in the proof of Lemma 4.2, why were we so careful about the order in which we considered the small rectangles? 3. Let p : E —> B be a covering map. Let a and P be paths in B with a(l) = P@)\ let a and /? be liftings of them such that &A) = 0@). Show that a * P is a lifting of a* p. 4. Consider the covering map p : R x R* — R1 - 0 of Example 3, §8-3. Find liftings of the paths /(/) = B - /, 0), g(/) = ((l +/)cos27r/, A +/)sin: A(/)=/*g. Sketch these paths and their liftings. 5. LetneZ+. Consider the maps g, A : S1 — Sl rah<.nmie »alucl , V*". [Here we represent 5- as the set of complex numbers z of absolute^^ Compute the induced homomorphisms *♦.*• of the infinite cychc group ni(S1,b0) into itself.
342 The Fundamental Group and Covering Spaces Chap. 8 «. Show there is no retraction ,: ** _ *>, where B* is the unit ball fa * 7. Prove Theorem 4.5. , 8. Show that if B is simply connected, then any covering map p ■ E — s r v £ is path connected is a homeomorphism. * for which 9. Consider the covering map p x p: ^ x 7? — 51 x 5' of Examni,. 1 «. Consider the path Sample 2, §8-3. /« = (cos 27:/, sin 2tt/) x (cos Ant, sin 4tt/). in 5' x SK Sketch what/looks like when S> x 5' is identified with thedouah nut surface D. Find a lifting/ of / to R x R, and sketch it. Formulate a~ conjecture about the fundamental group of the torus, and prove it. 10. Prove the following generalization of Theorem 4.5: Theorem. Let p : E -^ B be a covering map; let E be path connected- let p(e0) = b0. Then (a) Pt :ndE, e0) ^nt(B, b0) is injective. (b) There is a bijection ^■■n^B^o^H—rp-Hbo), where H = p*(nt(E, e0)) and nt(B, bo)IH is the collection of right cosets of H inn,(B,b0). 11. Assume the hypotheses of the preceding theorem. If n,(B, b0) is infinite cyclic, what can you say about nt(E, eo)l What if n,(B, b0) is finite and/? is a Wold covering ? *12. (a) Lemma (A lifting lemma). Let p : E —► B be a covering map; let p(e0) = 60. Let f:(Y,y0) —► (B,b0) be a continuous map. If Y is locally path connected and simply connected, then f can be lifted uniquely to a continuous map f:Y—'E such that f(y()) = en. [Hint: Given y e Y, choose a path a in yfrom^o to>s lift/" a to a path in E beginning at ea, and define/OO to be the end point of this path.l (b) Theorem. Let B be locally path connected. Suppose p:E—>B is a simply connected covering space of B. If p': E' —> B is any path-connected covering space of B, then there is a covering map q: E—^ E' such that p -P °1- [Hint: Lift p to £".] This theorem shows why E is called a wiivewo/ covering space of B; it covers every other path-connected covering space n (c) Theorem (Uniqueness of the universal covering space). If B is localy P" connected, and ifp:E-*B andp': E' — Bare two simply connected<cov ing spaces of B, then there is a homeomorphism h:E-*E' such thatp ° P' *13. Theorem. Let G be a topologicalgroup with operation ■; suppose p. is a simply connected covering space of G. If C is locally path «£**' md there is a multiplication QonG relative to which 0 is a topologxcal group p is a homomorphism. ... . [Hint: Let e be the identity element of G; choose Imp M-
§ 8-5 ,- The Fundamental Group of the Punctured Plane 343 in G, choose paths * and 0 from S to 3 and y, respectively. Let a(s) - DE.u\\ and fi(s) = ptfm. Take the path y(s) = «(,).#,) in G, lift it to a path y * G that begins at S, and define X © j» to be the end point of y.] The group G is called the universal covering group of G. 8-5 The Fundamental Group of the Punctured Plane In this section, we shall prove that the fundamental group of the punc- punctured plane R2 — 0 is infinite cyclic. We shall need this fact in several of the applications. The proof will lead us into a short discussion of deformation retracts and their relation to the fundamental group. Theorem 5.1. Let x0 e S1. The inclusion mapping j:(S\x0)—>(R2-0,x0) induces an isomorphism of fundamental groups. Proof. Let r : R1 — 0 —► Sl be the continuous map defined by r(x) = xj\\x\\, where 11 x11 denotes the distance of x from the origin 0 in the euclidean metric. The map r can be pictured as collapsing each radial ray in Rz — 0 onto the point where the ray intersects Sl; it maps each point x of 5' to itself. We claim that r.M is an inverse Torj*. First consider the composite map OS", x0) -±-~ (R1 - 0, jt,)-^. (S1, *0); this map equals the identity map of S'. Therefore, by the functorial properties of the induced homomorphism, r.M % is the identity isomorphism of To show that jt ° r* is the identity isomorphism of n^R1 — 0, x0) with itself, let us take a loop/in Rz - 0 based at x0. Then/*(/■*[/]) is the homo- topy class of the loop g =j o r of: 1^ Rz - 0, which is given in equa- equations by 8is)=urn' We need to show that g ~p /; but this is easy. As indicated in Figure 18, we just move the point /(j) gradually along its radial ray until it hits the point 8(s). More formally, define F: I X /-> R1 - 0 by the equation ir/§ir ft is clear that F(s, t) is never equal to 0, for + Ao*o and
344 The Fundamental Group and Covering Spaces Chap. 8 t(s) Figure 18 Also, F@, t) = F{\, f) = x0 for all t. Thus F is the required path homotopy between /and g. □ There is nothing special about the circle and the plane in the preceding proof. Exactly the same proof applies to prove the following theorem: Theorem 5.2. If\0 e S", the inclusion j:(S°-\x0)—>(R°-0,x0) induces an isomorphism of fundamental groups. Of course, we have not yet computed the fundamental group of S"~l, so this does not tell us much. What made the preceding proof work? Roughly speaking, it worked because we had a natural way of deforming the path/, which lies in R2 — »> into the path r of which lies in Sl. Another way of looking at this proof is to note that we can deform the entire space R% — 0 into the circle S1 if we like, by gradually collapsing each radial line to the point where it intersec s Sl. This deformation is the one that deforms the path / into the path r ■>/. it is called a strong deformation retraction of R2 — 0 onto Sl. Analyzing the proof in this way leads to a generalization of Theorem 5.^ Although we shall not have occasion to use it in this book, it is a theo that is very useful in computing homotopy groups.
§8-5 The Fundamental Group ofthe Punctured Plane 345 Recall that if A is a subspace of X, we say that A is a retract of X if there is a continuous map r : X—<■ A such that r(a) = a for every a e A. The map r is called a retraction of JT onto ,4. Definition. Let ,4 be a subspace of X Then /< is said to be a strong defor- deformation retract of X if there is a continuous map H: X X /—> X such that H(x, 0) = x for x 6 JT, #(*, 1N/4 for x 6 X, H(a, t) = a for <? 6 A and r 6 /. The map H is called a strong deformation retraction. Said differently, the space A is a strong deformation retract of X if JTcan be deformed gradually into /4, with each point of A remaining fixed during the deformation. At the end of the deformation, we have a retraction of X onto A, mapping x into H(x, 1). Example 1. The map H:(R" - 0) x /— (R" - 0) defined by 11*11 is a strong deformation retraction of R" — 0 onto 5""'; it gradually collapses each radial line into the point where it intersects S"~l. Theorem 5.3. Let A be a strong deformation retract of X. Let a0 6 A. Then the inclusion map j : (A, a0) —> (X, a0) induces an isomorphism of fundamental groups. The proof is similar to that of Theorem 5.1; we leave it to you. This theo- theorem can be used to compute fundamental groups for a number of spaces, by reducing the spaces involved to more familiar ones. Typical examples follow. Example 2. Let B denote the z-axis in R3. Consider the space R3 - B. It has the punctured .ry-plane (Rz - 0) x 0 as a strong deformation retract. The map H defined by the equation is a strong deformation retraction; it gradually collapses each line parallel to the z-axis into the point where the line intersects the *>-plane. We conclude that the space R3 — B has an infinite cyclic fundamental group. Example 3. Consider /?* - p - q, the doubly punctured plane. We assert it has the "figure eight" space as a strong deformation retract. Rather than writing equations, we merely sketch the deformation retraction; it is the three- stage deformation indicated in Figure 19.
346 The Fundamental Group and Covering Spaces Chap. Figure 19 Example 4. Another strong deformation retract of R1 —p—q is the "theta space" O^S* u@ x[-l,l]); we leave it to you to sketch the maps involved. As a result, the figure eight (8) and the theta space F) have isomorphic fundamental groups, even though neither is a strong deformation retract of the other. Of course, we do noi know anything about the fundamental group of the figure eight as yet. But we shall. We should remark that Theorem 5.3 remains true if one replaces the condition H(a, t) = a for a e A and t e I by either of the weaker conditions H(a, t) e A for a e A and t e /, or H(a, 1) = a for a e A. But the proof in either of these cases is more difficult. (See Exercise 5 o §8-11.) In practice, the theorem one usually applies is Theorem 5.3 anyway. Exercises 1. Show that if A is a strong deformation retract of X, and Bis a strong deform tion retract of A, then B is a strong deformation retract of X
§8 undamen'alG^^punctwl(me 347 2. Show that the paths/and A in Example 2 of 58 1 a 3. Prove Theorem 5.2. ^ "» "Ot path h°™toPic. 4. Prove Theorem 5.3. * ' " 5. For each of the following spaces, the fundamental <™ • • cyclic, or isomorphic to the tm^mS^X^ ^- ***"-inflnite for each space which of the three alternatives holds "^ ^^ Determine (a) The "solid torus," Bz x S1. (b) The torus T with a point removed. (c) The cylinder Si x I. (d) The infinite cylinder Si x R. (e) R3 with the nonnegative x, y, and z axes deleted. The following subsets of Rz: (f) {Jt|||jt||> 1} (g) WII*II>1} (>) «S (j) * (k) S A) f :> u ■' u :' u [2 - (R+ (R+ (R (R+ X 0) x R) x 0) X 0) 6. Let C denote the complex plane; let /be a loop in C — o based at x0. (a) Let p : R —► S' be the standard covering map. Consider the loop h(s) = f(s)/\\ f(s)\\ in .<?'; lift it to a path h in R. Then h(\) = £@) + n for some integer n. Show that the path homotopy class of /determines n uniquely, and conversely. The integer // is called the winding number of/with respect to 0. (b) Theorem. Suppose f is a piecewise-differenthble loop in C - 0. Then the winding number off with respect to 0 equals the integral J 2n i [Hint: If * = g(s) is an arbitrary pIccewIseHdiffe^tiaHe path fa C - (^ consider the path , -> g(s)l\\Hs)\\ in 5' and let B be a lifting of it to the covering space R. Then g(s) = \\8(s)\\e™». Compute J//] equals the image of [/] under these isomorphisms. G 7- (a) Show that *3 - 0 is simply connected, t^"; Gl^" ^e up of finitely h hotopic toaloop mad P ow that * i p ^; ^ ^e up o based at .v0) first show that/is path homotopic toaloop mad P^ many straight-line segments, none of wh.ch are copia segment joining x0 and the origin.] (b) Show that S2 is simply connected.
348 The Fundamental Group and Covering Spaces Chap, g 8-6 The Fundamental Group of Sn Now we compute the fundamental group of S", showing that S" is sim 1 connected if n > 2. A proof for the case n = 2 was outlined in Exercise 7 of the preceding section. Here is a different proof, which uses a theorem w shall need later anyway. Theorem 6.1 {The special Van Kampen theorem). Let X = U U V, where U and V are open in X and V n V is path connected. Let x0 be a point of U n V. If both inclusions i:(U, x0)—>-(X,x0) and j : (V, x0)—>-(X,x0) induce zero homomorphisms of fundamental groups, then n^X, x0) = 0. Note that both /'„ and j\ are necessarily zero homomorphisms if U and Fare simply connected. This is the case we need for the following theorem. We shall need the more general result in proving the Jordan separation theo- theorem (§8-12). Proof Let/: /—> X be a loop based at x0. We wish to show that/is path homotopic to a constant loop. Step 1. By the Lebesgue number lemma, there is a subdivision 0 = aa < a{ < ■ ■ ■ < an = 1 of the interval [0, 1] such that for each /, the set /([a,.,, a,]) lies entirely in one of the open sets U or V. Among all such subdivisions, choose one for which the number n of subintervals is minimal. Then it follows that for each /, the point f{a) lies in U n V. Suppose that /(a,) <£ U, for instance. Then neither /([a,_i,aj) nor f([at, al+i\) lies entirely in U. Therefore, both of them must lie entirely in V. We can then discard a, from the subdivision, and still have a subdivision of [0, 1] for which the image of each subinterval lies either in U or in V. This contradicts minimality. Hence /(a,) must belong to U. Step 2. Consider the restriction of/to the interval [<?,_,, at]. Reparame- trize it so that it is based on the interval [0, 1], defining /((!- s)a,.t + sa,) for s e [0,1]. M) We show that/, is path homotopic to a path that lies entirely in V. Of course, if/, itself lies in U, nothing needs to be done. We denne-r, be the trivial path homotopy F,(j, t) =/,(*) of/, to itself. nnnectivjty If/ does not lie in U, then/ must lie entirely in V. Using path connect y of U O V, let us choose paths g and h in U O Ffrom the base point *,
§8-6 The Fundamental Croup ofS* 349 Figure 20 points/@) and/((I), respectively. (See Figure 20.) Consider the composition (g */■) * h; it is a loop in V based at .y0. Under the inclusion mapy, this loop goes into a loop in X based at ,v0. Since j^ is the zero homomorphism, this loop is path homotopic, in X, to the constant loop ext. Then/, is path homo- topic (in X) to the path g * h (using the groupoid properties of * proved in Theorem 1.2). Let F, be this path homotopy; it is a path homotopy (in X) between/, and a path lying in U n V. Now we reparametrize each path homotopy F, so as to get a map of [a,_u a,] x / into -V, and then we paste the pieces together to get a path homotopy F between /and a path that lies entirely in U. Specifically, define F: I x / —> X by the equation F(s, t) = ^ for s Because each path homotopy F, leaves the end points fixed, /"is well-defined; the pasting lemma shows it is continuous. Let f\s) = F(s, 1). Step 3. We have shown/is path homotopic (in X) to a loop/' that lies entirely in U. Because ■s the zero homomorphism, the loop /' must be path homotopic (in X) to.the constant loop Bjc, Thus/^,/' ~, en. as loops in X This .s what wshed to prove. □ This theorem is a special case of a famous theorem of topology called tbe Van Kampen theorem, which expresses the fundamental group oHhe space X quite generally in terms of the fundamental groups of tf and V ana homomorphisms induced by the inclusions of Un V into U and r, pro vided that U n K is path connected. (See [M].)
350 The Fundamental Group and Covering Spaces Chap, g Theorem 6.2. For n :> 2, the ^-sphere S" is simply connected. . Proof. Let p = @, ..., 0, 1) 6 R*+i be the "north pole" of *. i q = @,..., 0, -1) be the "south pole." P W ^' kt Step I. First, we show that 5" — p is homeomorphic with R" Let x = (*„.. ., jc-+i) be a point of S" different from/,. Suppose that«» take the straight line in *+• determined by x and/, and intersect it with th plane x.+ l = 0; let f(x) denote this point of intersection. The resulting man f:S° — p-^R« is called stereographic projection. More precisely, we define/by the equation Obviously/is continuous. To show that/is a homeomorphism, one checks that the map g : R" —> S" — p given by g(yi,-.., y.) = 0ylt ty2, ...,ty.,\-1), where t = 2/A + (>>,J + • • • + O'J2), is an inverse for/ Since 5" — q is homeomorphic to S" — p under the reflection map p(.\\ , ,Vn, -Vn+ ,) = (X,, . . . , *„, —*„+,), 5" —17 is also homeomorphic to R". Step 2. Now let U = 5" — p and V = S» — q. The space 5" equals the union of the open sets £/ and V. Furthermore, the sets U and V are simply connected, being homeomorphic with R". Once we show that U D Vis path connected, we can conclude from Theorem 6.1 that the space 5° is simply connected. The intersection of U and V is the set 5" — p — q- To show that this set is path connected, note first that it is homeomorphic with R" — 0 under stereographic projection. It is easy to show that R" — 0 is path connected: Any point x of R" — 0 can be joined to the point x0 = (I, 0,..., 0) by the straight-line path in R- — 0, except for points x of the form (a, 0,..., Q). where a < 0. In that case, we can take the straight-line path from x to xt - @, 1, 0, . . . , 0), followed by the straight-line path from *, to *„. (Here we use, of course, the fact that /» > 1.) □ Corollary 6.3. R" — 0 is simply connected if n > 2. Proof. By Theorem 5.2, R- - 0 and 5-' have isomorphic fundamental groups. □ Corollary 6.4. R° and R2 are not homeomorphic for n > 2. Proof Deleting a point from R" leaves a simply connected space, win e deleting a point from R2 does not. Q
§8 Fundamental Croups of Surfaces 351 This corollary has a generalization to higher dimensions- R° and R- are not homeomorphic if n ?t m. But the proof requires more tools of algebraic topology than we have developed. Exercises 1. Let X be the union of two copies of S2 having a point in common. What is the fundamental group of XI Prove that your answer is correct. [Be careful! The union of two simply connected spaces having a point in common is not neces- necessarily simply connected. See [S], p. 59.] 2. Let X be the union of two open sets U and V; let U n V be path connected; let .vo e U n V. Let / and j be the inclusion maps, as in Theorem 6.1. If you are given thaty* is the zero homomorphism, what can you say about itl 3. Criticize the following "proof" that S2 is simply connected: Let/be a loop in Sz based at ,v0. Choose a point p of S2 not lying in the image of/ Since S2 — p is homeomorphic with R2, and R2 is simply connected, the loop/is path homo- topic to the constant loop. 8-7 Fundamental Groups of Surfaces Recall that a surface is a HausdorfiT space with a countable basis, every point of which has a neighborhood that is homeomorphic with an open sub- subset of/?2. Surfaces are of interest in various parts of mathematics, including geometry, topology, and complex analysis. What we do here is to consider four familiar surfaces—the sphere S1, the projective plane P2, the torus T, and the double torus 7\—and show by comparing their fundamental groups that these surfaces are not homeomorphic. One can, in fact, use the fundamental group to classify all compact surfaces completely; but this we shall not attempt. The interested reader is referred to Chapter 4 of [M]. First we consider the torus. To compute its fundamental group, we shall need a theorem about the fundamental group of a product space. Theorem 7.1. tt,(X x Y, x0 x y0) isisomorphic with 7r,(X, x0) x 7r,(Y, y0). Proof. Recall that if A and B are groups with operation -, then the cartesian product A X B is given a group structure by using the operation (a X b)-(a' X b') = (a-a) X (*•*')• Recall also that if h : C — > A and k : C — B are group homomorphisms, then the map 0> : C— A X B defined by O(c) = h(c) X k(c) is a group homo- homomorphism. Now let p : X x K-. X and q: X x Y— Y be the projection map- mappings. If we use the base points indicated in the statement of the theorem, we
352 The Fundamental Group and Covering Spaces Chap. 8 have induced homomorphisms q*:n1(XX Y, x0 Xyo)~ We define a homomorphism *: *,(* X Y, x0 X ya) ~> „,(*, x0) x *,(}-, y0) by the equation *([/"]) = />*([/]) X <?,([/]) = [p°f]X[qof]. We shall show that 0> is an isomorphism. The map 0> issurjective. Let # : /_ * be a loop based at *„• let A • / y be a loop based at y0. We wish to show that the element [g] x [h] lieshlhe image of 0>. Define/:/—. X x Y by the equation Then/is a loop in ;r x Y based at x0 x ;■„, and *([/]) = [/»»/] X [? o/] = [?] x[A], as desired. 77;e Ae/vic/ o/O vanishes. Suppose that/: / —> X x 7is a loop in X x y based at .v0 X ;•„, and 0>([/]) = [p ° f] X [f/ o/] is the identity element. This means that p °/~p eXa and q °f~p eyc; let G and // be the respective path homotopies. Then the map F': / X / —► X x 7 defined by FO, /) = G(i, /) X H{s, t) is a path homotopy between/and the constant loop based at xa x y0. D Corollary 7.2. The fundamental group of the torus T = S1 x S1 is isomor- phic to the group Z X Z. Now we define the projective plane and compute its fundamental group. Definition. The projective plane P2 is the space obtained from S2 by iden- identifying each point x of S1 with its antipodal point —.v. Formally, define an equivalence relation on S2 by setting x ~ x and x „ (_.V); then P> is the set of equivalence classes. Up : S2 — P1 maps each point x to its equivalence class, we topologize P2 by defining Fto be open in P2 if and only if/r'(fO>s °Pen in SK . ■ rl;ve The projective plane is the fundamental object of study in projecuv goemetry, just as the euclidean plane R* is in ordinary euclidean geomet y. Topologists are primarily interested in it as an example of a surface. Theorem 7.3. The projective plane P2 is a surface, and the map p : S2 — « tf covering map.
§ " Fundamental Groups of Surfaces 353 &; hence «({/) is open in s* Since ~ * ls a homeomorphism of P'l(p(U))=Uua(U), this set also is open in S>. Therefore, by definition, rftf) is open in i>2 Now we show that p is a covering map. Given a point y of i», choose x e /r'OO- Then choose an e-neighborhood U of ^ in S2 for some f <-T using the euclidean metric d of R>. Then {/contains no pair {z afe))ofantmn' dal points of S\ since d(z, a(z)) = 2. As a result, the map ' is bijective. Being continuous and open, it is a homeomorphism. Similarly, p:a(U) — p is a homeomorphism. The set p~l(p(U)) is thus the union of the two disjoint open sets {/and a(U), each of which is mapped homeomorphically by p onto />({/). Then p(U) is a neighborhood of p(x) = y that is evenly covered by p. Hencep is a covering map. Since S2 has a countable basis {{/„}, the space ./" has a countable basis We show P2 is Hausdorff. Let yt and ^2 be two points of P1. The set P~'(yi) U /'"'O'i) consists of four points; let 2f be the minimum distance between them. Let U, be the f-neighborhood of one of the points of p~:()'i)> and let U2 be the f-neighborhood of one of the points of p-l(y2)- Then {/, u a(Ut) and U2 U ff(^2) are disjoint. It follows that />({/,) and p(U2) are disjoint neighborhoods of yx and ^2, respectively, in P2. Since 52 is a surface and every point of P2 has a neighborhood homeomor- phic with an open subset of S2, the space P2 is also a syrface. □ Corollary 7.4. s,(P2, y) « <? ^ro«/» of order 2. Proof. The projection />: 52 — P2 is a covering map. Since 52 is simply connected, we can apply Theorem 4.5, which tells us there is a bijective cor- correspondence between n,(P\ y) and the set p-'(y). Since this set is a two-ele- two-element set, 7i,(P2, y) is a group of order 2. Any group of order 2 is isomorphic to Z2, the integers mod 2, of course. U One can proceed similarly to define P; for any n e Zt, as the space obtained from S' by identifying each point x with its ant.pode -*, it is called projective n-space. The proof of Theorem 7.3 goes throu«h^^SS prove that the projection , : S- - P" is a covering map. Then because S*
354 The Fundamental Group and Covering Spaces Chap. 8 simply connected for n ^ 2, it follows that *,(i>", y) is a two-element grouD for n-^2. We leave it to you to figure out what happens when „ = i We have not yet found any space whose fundamental group is not abelian' Let us remedy that deficiency now. Lemma 7.5. The fundamental group of the figure eight is not abelian. Proof. The figure eight is the union of two circles A and B with a point x0 in common. We now describe a certain covering space E for the figure eight. The space E is the subspace of the plane consisting of the *-axis and the ^-axis, along with tiny circles tangent to these axes, one circle tangent to the ,v-axis at each nonzero integer point and one circle tangent to the j^axis at each nonzero integer point. See Figure 21. The projection map, wraps the.v-axis
§8-7 case the " ^^ Qroups °f Surfaces Z^1TZETZZ^°lz- wislKd'but the origin to the point 1 x 0. Let // t u V°ing alon8 the x-axis fm along the ,axis from the origin^£^0*** *> =0 ^ pog; then/and ^are loops in the figure eiehihl ^ / = P°^ and * = circles A and B, respectively. We asserUha /* '^ " ?'8Oin8 ar°Und *e topic, so that the fundamental group of the fii,! .**^are not P^h honio- To prove this assertio lt lif f ^1S "Ot abelian- *~ the ntal group of the fii,! To prove this assertion, let us lift eachof fh the origin. The path/., lifts to a path Z^ origin to 1 x 0, and then goes once around 1 X 0. On the other handfthe paS"! the ^-axis from the origin to 0 X tangentto the ^-axis at^ x 1. ^Z^ point,/* g and g */cannot be path homotopic. The fundamental group of the figure eight is, in fact, the group that algebraists call the "free group on two generators." But this we shall not prove. Theorem 7.6. The fundamental group of the double torus T2 is not abelian. Proof. The double torus Tz is the surface obtained by taking two copies of the torus, deleting a small open disc from each of them, and pasting the remaining pieces together along their edges. We assert that the figure eight is a retract of Tz. This fact implies that inclusion./ induces an injective homo- morphism j# : ;r,(8, x0) —> nt(T2, x0), so that 7r,(r2, x0) is not abelian. One can write equations for the retraction r:T2~+ 8, but it is simpler to indicate it in pictures, as we have done in Figure 22. First one collapses the figure 22
336 The Fundamental Group and Covering Spaces Chap, g dotted circle to a point, obtaining two tori with a point in common TV collapses each of the two tori onto their "meridianal circles™ Q Corollary 7J. The surfaces S'.P.T, and T2 are toPologically distinct. Exercises 1. Compute the fundamental groups of the "solid torus" 5' x B* and theorod » space S1 x S2. 2. Let X be the quotient space obtained from B1 by identifying each point x of Sl with its antipode -x. Show that X is homeomorphic to the projective plane/12. 3. Let p : E —► 8 be the map constructed in the proof of Lemma 7.5. Let E' be the subspace of E which is the union of the .v-axis and the .y-axis. Show that/?|£' is not a covering map. 4. Consider the covering map indicated in Figure 23. Here/? wraps A\ around A twice and wraps B\ around B twice; p maps Ao and Bo homeomorphically onto A and B, respectively. Use this covering space to show that the fundamental group of the figure eight is not abelian. Figure 23 5. Given a group G and a space X, an action of G on X is a function assigning, to each element a of G, a continuous map /!<X:A"—*X ,n such a_*grt[iheaf;ent.ty e)ement of G then ht is the identity map of X. (ii) If a = jff-y, then A,, = hf ° hr. v 4-« defines an (a) Show that denning h,:R-^Rby the equation h&) = * + " ^ action of the integers Z on. Jl. an ^ ofthe group integers mod n) on S2.
ft Q Q ' Essential and Inessential Maps 357 (d) Show that the antipodal map of 5. defines an action of the group Zl 6. Given an action ofCon X, define the orbit space X1G to be the mm.- ♦ X determined by the following equivalent re.atlonxl* fr sT some a e C. The orbit spaces under the actions defined it ^ 5 are homeomorphic to familiar spaces. What are they? 7. An action of Con Xis said to be fixed point free if the only map A that possesses a fixed point is the map he. Which of the actions defined in Exercise 5 arS point free? 8. Theorem. Let there be given an action of the finite group G on the path-con- path-connected Hausdorff space X; assume that the action is fixed point free. If X is simply connected, then tt, (X/G, xj is isomorphic to G. 9. Consider S3 as the space of all pairs of complex numbers (z,, z2) satisfying the equation | z, |* +1 z212 = 1. Given relatively prime positive integers n and ky define h: S3 —► S3 by the equation h(z,,z2) = ^e1'"", z2e2"""'»). (a) Show that the maps /;, h1 = h » h, h3,.. . can be used to define an action of Zn on Si that is fixed point free. The orbit space L(n, k) is called a lens space. (b) Show that L(n, k) is a compact 3-manifold. (c) Assume Exercise 8. Show that if L(n, k) and L(n', k') are homeomorphic, then n = n'. (In fact, L(n, k) is homeomorphic to !(«', k1) if and only if n = n' and either k = k' (mod n) or **' = 1 (mod «)• The proof is decid- decidedly nontrivial.) 8-8 Essential and Inessential Maps We have already used the fundamental group to study the problem of determining whether or not two given spaces are homeomorphic. Now we apply it to a second problem of topology, that of determining whether or not two given maps of one space into another are homotopic. ° . /. , ,,_•„*„ ^frminp whether or not a given map o given maps of one space into anoinci <uc i,u—k~ The simplest such problem is to determine whether or not a h : X-+ Y is homotopic to a constant map. We shall prove v^ - ^ L ... . . *i *h» Jn^rieed homomorphism n* ui'"" n . A —► I IS hOmOtOpiC IO a tuna""" —"K- - . j. fun_ homotopic to a constant map, then the induced homom°rp^Snm t* ta the damental groups is trivial. (The converse holds it x n ppc circle S1; see Exercise 2.) . . ,,,h-ther two arbitrary We defer the more general problem of determining whe her two maps h, k : X -^ K are homotopic to a later section (§8-11). Definition. A n,ap /, : X- K is said to be initial if* I. homotopic to a constant map. Otherwise, it is said to be essential.
The Fundamental Croup and Covering Spaces Y. Then the following are equivalent: Chap. 8 Lemma &2. Let h: S1 A) h is inessential. B) h can be extended to a continuous map g: B2 —♦ Y. Proof. A) => B). Suppose that H: Sl X I-^ Y is a homotopy between h and the constant map k which carries S1 to y0. Let n : S1 x I—> £* be the map ?r(x, 0 = A - 0*. Then ?r carries S1 X [0, 1) bijectively onto B2 — 0, and it maps S1 x 1 to the point 0. (You can check this.) Because S1 X /is compact and B2 is Hausdorff 7t is a closed map; that is, it carries closed sets of S1 X /to closed sets of B1. Now the map H: S1 x I—* Y is constant on the set S'xl. Therefore, H induces a map g : E1 —> T such that g o % = H. [It is defined by letting g(x) = Nin'^x)) if x = o and g(x) = y0 if * = o.] The map # is continu- continuous; this follows from the fact that n is a quotient map. [ Or one can prove it directly by noting that if C is closed in Y, then H~l(C) is closed in 51 x /, so that n{H~l{C)) = g~l(C) is closed in B2.] The map g is the desired extension of A; for if x e Sl, then g(x) - £(*(*, 0)) = H(x, 0) = A(x). See Figure 24. Figure 24 B, - A). Let ,:*• - y.be . co^nwj- F.S1 X /-* y by the equation F(x, /) = topy between A and a constant map. p Theorem 8.X us define , homo.
§8-8 . Essential and Inessential Maps 359 Proof. Consider first the case X = S1 We annlv th Let g: 2»» - 7 be an extension of A; let / • 'SiZ K' M T^'"8 lemma' jh L 6 & £ ^ "¥" T** g; t / SZ K M T Then goj-h. Let 6, be a point of & and kt £ =^ "¥ T induced homomorphisms yi (t>l>- Consider the By the functorial properties of the induced homomorphism h = e ; Buty is the zero homomorphism, because its range is the' trivial group Therefore, /;* is the zero homomorphism. Now we consider the general case. Let/:/-. Jf be a loop in X based atx0. Let p:I —> 51 be the standard loop 0CO = (cos 27r.y, sin ) Now/induces a map k:Sl ~> X, defined by the equation k(a) = It is easy to check that k is continuous. By hypothesis, h is homotopic to a constant; let H: X x / —► Y be the homotopy. Then /; ° A' is also homotopic to a constant; the map H'(a, t) — H(k(a), t) is the required homotopy. It follows from the preceding paragraph that (/; o A')* is the zero homomorph- homomorphism. In particular, (/; o k),.{[^]) = 0. But = [h c k o <f>] = [h o f] = //*([/]). □ As an application, we prove the theorem we used in §7-9, when we com- computed the topological dimension of a triangular region. Corollary 8.3. Let T be a closed triangular region in R2/ let Bd T denote 'he union of the edges of T. There is >io continuous map f: T -* Bd T that m"ps each edge ofT into itself. Proof. We suppose there is such a continuous map/: T— Bd T, and derive a contradiction. . , r Let g : Bd T — Bd T be the restriction of/. Because g maps each edge ot T into itself, g is homotopic to the identity map. Indeed, the map G(x, /) = /* + (!- t)g(x) Bd r x / into Bd T; for if x belongs to Bd T, then x and g(x) lie in the
The Fundamental Croup and Covering Spaces Chap, same edge of T, so that the line segment joining them lies in this edge a well. Thus G is a homotopy between g and the identity. Now we use the fact that Bd T is homeomorphic to &. The identity ma of Bd T does not induce the zero homomorphism of the fundamental grou^ so that it cannot be homotopic to a constant, by the preceding theorem Therefore, g is not homotopic to a constant. On the other hand, g is extendable to the continuous map/of Tinto Bd T, so that g is homotopic to a constant. QJ Exercises 1. Here is an easy "proof" that if h : X —> Y is inessential, then A* is the zero homomorphism. Where is the fallacy? Let H: X x I—* K be a homotopy between h and a constant map. Given a loop /in X based at x0, define a map F: / x / —> Y by the equation F(s, t) = H(f(s), /). The map F is a homotopy between h °/and a constant path, as desired. 2. Let fr.S1 —> Y. Show that if /i* is the zero homomorphism, then h is inessen- inessential. [Hint: Let <f>: I —» 51 be the standard loop; let/= /» o 0. Show there is a path homotopy Fbetween/and eyol and that F induces a map HiS1 X /—» K such that // o @ x /y) = F. See Figure 25.] ^/ s'x/ 25 3. Let ybe path connected. Show that ;tl(y,>'o)=0if and only if every map /i; s1 —* Yis inessential.
§ s'9 Tne Fundamental Theorem of Algebra 36\ 4. Recall that a space Y is said to be contractible if the identity map ir;Y—*Y is inessential. Show that if Y is contractible, then Y is simply connected. 5. A map h : S" —> Sm is said to be antipode-preserving if h(-x) = ~h(x) for every x 6 S". Theorem. If h : Sl —> Sl is antipode-preserving and continuous, then h is essential. (a) Letp : 51' --* 51' be the map/>(z) = z%, where 2 is a complex number. Show that h induces a continuous map g: S1 — Sl such that p°h =goP. (b) Show that if/is any path in Sl from a point * to its antipode —x, then /> <>/is a loop in 51' that is not path homotopic to a constant. (c) Show that both p and g induce monomorphisms (injective homomor- phisms) of the fundamental groups. (d) Conclude that h is essential. 6. Assume Exercise 5. (a) Prove the following: Theorem (Borsuk-Ulam theorem for S1). There is no continuous antipode- preserving map f: S2 —> S1. [Hint: Consider the equator in S2.] It is in general true that there is no con- continuous antipode-preserving map/: S" —» Sm if m < n. But the proof requires more tools than we now possess. (b) Prove: Theorem. Given a continuous map f: S2 —> ,/J2, there is a point x of S2 such that f(x) = /(-.*). [Hint: Consider (/(.*) - f(~x))l\\f(x) - f(-x)\\.] (c) Prove: Theorem (A "theorem of meteorology"). At any given moment in time, there exists a pair of antipodal points on the surface of the earth at which both the tem- temperature and the barometric pressure are equal. (d) Prove: Theorem. Ifg:S2—>S2 is continuous and g(x) ^ g(-x) for all x, then g is surjective. *7. Generalize Exercise 5 as follows: Assume that h : 5' —» S> is antipode-pre- antipode-preserving; let h(x0) = x,. Then A, maps a generator of nt(S\ x0) to an odd multi- multiple of a generator of nt(S', .v,). 8-9 The Fundamental Theorem of Algebra It is a basic fact about the complex numbers that every polynomial equa- equation of degree n with real or complex coefficients has n roots (if the roots are counted according to their multiplicities). You probably first were told tnis fact in high school algebra, although it is doubtful that it was proved for you at that time.
362 The Fundamental Group and Covering Spaces Chap. 8 The proof is, in fact, rather hard; the most difficult part is to prove tW every polynomial equation of positive degree has at least one root There a various ways of doing this. One can use only techniques of algebra- thi proof is long and arduous. Or one can develop the theory of analytic functions of a complex variable to the point where it becomes a trivial corollary of Liouville's theorem. Or one can prove it as a relatively easy corollary of our computation of the fundamental group of the circle; this we do now. Theorem 9.1 {The fundamental theorem of algebra). A polynomial equation x° + a^.x-1 + ■ ■ ■ + a,x + a0 = 0 of degree n > 0 with real or complex coefficients has at least one {real or com- complex) root. Proof. Step I. Consider the map /;: S' —► S' given by h{z) = z", where z is a complex number. We show that the induced homomorphism //»:*,(£', 60)—>nl{S',b0) carries a generator of this infinite cyclic group to n times itself. Let <j): /—> S' be the standard loop <f>(s) = (cos 2ns, sin 2ns) = elnls in S1. Its image under //* is the loop li(<t>(s)) ■--= (e2"'')" = e2"'" = (cos Inns, sin Inns). This loop lifts to the path s --> us in the covering space R. Therefore, the loop h o <j) corresponds to the integer n under the standard isomorphism of ni{S\ b0) with the integers, whereas <f> corresponds to the number 1. Step 2. Now we prove a special case of the theorem. Given a polynomial equation z- + an_,z"-' + ■ ■ ■ + a,z + a0 = 0, we assume that l 1 and show that the equation has a root lying in the unit ball B\ Assume that the equation has no root in B*. Then we can define a map g ■ gz _ /?: — 0 by the equation g(z) = «" + *.-i*"-' + ■ • • + ".* + fl°" Let ,. 5, _, Rl _ 0 be the restriction of g to 5'. Because/is extendable t0 the'map , of B> into * - 0, the Wf"^*^? On the other hand,/is homotopic: tc.the j.p* ■ Sj* k{z) = z". For the homotopy F. £' X / a
§ 8'9 The Fundamental Theorem of Algebra 363 is the required homotopy; F(z, t) never vanishes because = 1 -Kl«.-,l+ •■• + |«0|)>0. Furthermore, the map k is essential. For k equals the composite of the map h : S' — S' of Step 1, given by h(z) = Z", and the inclusion map j-.S1 — /?* - 0. Since A* is "multiplication by «" and;,, is an isomorphism, k* is not the zero homomorphism. Therefore, k must be essential. Since/is homotopic to k, the map/also must be essential. Thus we reach a contradiction. Step 3. Now we prove the general case. Given a polynomial equation x" + a,,-,*"-' + ••• + a,x + a0 = 0; let us choose a real number c > 0 and substitute x = cy. We obtain the equation (cy)n + on_,(n')"'' + • • • + a,(cy) + a0 = 0 or n ' an-\ ,n-l I _L ° > _i_ "o ft ^ ' "^"•J/ c"~'y c- If this equation has the root y = yB, then the original equation has the root x0 = cy0. So we need merely choose c large enough that C + c2 + ■■■ + a, c"-' + to reduce the theorem to the special case considered in Step 2. □ 1. Given a polynomial equation x" + a,.,*" + ■ • • + "ix + ao = ° with real or complex coefficients. Show that if |a,_, | + •; • +\"i\ +. Kl <j. then fl// the roots of the equation lie interior to the unit ban Jf • I™ • *U) = 1+«_,*+.■■ +«,*•- +-0*", and show that ,<*) * 0 2. Find a circle about the origin containing all the roots of thepo.ynomia. equation f 1=0.
364 The Fundamental Group and Covering Spaces Chap. 8 8-10 Vector Fields and Fixed Points In this section we apply the fundamental group to two problems «f geometry. One problem concerns the existence of vector fields taneent T given surfaces. The second deals with the "fixed-point problem": Given X does every continuous map/: X — X necessarily have a fixed point? ' We shall obtain only some of the simpler results. Deeper theorems, includ- including some of current research interest, demand much more of the machinery of algebraic topology than we have studied. Theorem 10.1 Given a nonvanishing vector field on B2, there exists a point o/S' where the vector field points directly inward and a point ofS' where it points directly outward. Proof. A vector field on B2 is an ordered pair (x, v(x)), where x is in B2 and v is a continuous map of B2 into R2. In calculus, one often uses the notation v(x) = «,(x)i + v2(x)j for the function v, where i and j are the standard unit basis vectors in R2. But we shall stick with simple functional notation. To say that a vector field is nonvanishing means that v(x) =£ 0 for every x; in such a case v actually maps B1 into R2 — 0. We show first that given v, it must point directly inward at some point of S'. Consider the map w : S1 —> R2 - 0 obtained by restricting v to S'. If there is no point x of 5' at which the vector field points directly inwand, then for no x in 5' is w(x) equal to a negative multiple of x. It follows that w is homotopic to the inclusion mapy: S1 —> R1 — 0, for the map F:S' X /—• R2 — 0 given by the equation F(x, t) = tx + A - t)w(x) is the required homotopy. It is pictured in Figure 26. Obviously, Fis con- continuous. To show that F never vanishes, note that if F(x, t) = 0, then A — t)w(x) = —tx. This equation is clearly false for t = 0 or t = 1, since x eSl and h"M£ For 0 < t < I, it says that w{x) = -«/(! - 0. so that w{x) equals a multiple of x, which is forbidden. _ Since w is homotopic to the inclusion map j : S -> K «. essential. On the other hand, W is extendable to the continuous mpj i^ - 0, so that it is inessential. Thus we arrive at a contradiction, mer v points directly inward at some point of SK result just Consider now the nonvanish.ng vector field (*, -**))■ »Y
"l0 Vector Fields and Fixed Points 365 (*,/(*)) {x,w(x)) (Y.) M) B, i B)) FigUre 26 proved, it points directly inward at some point of S'. Then v points directly outward at that point. Q We have already seen that every continuous map/: [0, 1] —► [0, 1] neces- necessarily has a fixed point (see Exercise 3 of §3-2). The same is true for the ball B1, although the proof is deeper: Theorem 10.2 {Brouwer fixed-point theorem for the disc). Iff : B2 —► Bz is continuous, then there exists a point x s B2 such that f(x) = x. Proof. We proceed by contradiction. Suppose that f(x) ^ x for every x in B1. Then defining v(x) = f(x) — x gives us a nonvanishing vector field (x, v{x)) on B1. But the vector field v cannot point directly outward at any point x of S', for that would mean f(x) — x = ax for some positive real number a, whence f{x) = A + a)x lies outside the unit ball B2. We thus arrive at a contradiction. □ One might well wonder why fixed-point theorems are of interest in mathe- mathematics. It turns out that many problems, such as problems concerning exist- existence of solutions for systems of equations, for instance, can be formulated as fixed-point problems. Here is one example, a classical theorem of Frobenius. We assume some knowledge of linear algebra at this point. Corollary 10.3. Let A be a 3 by 3 matrix of positive real numbers. Then A has a positive real eigenvalue (characteristic value). Proof Let T- R3 — R3 be the linear transformation whose matrix (rela- (relative to the standard basis for R3) is A. Let B be the intersect.on of the
366 The Fundamental Group and Covering Spaces 2-sphere S2 with the first octant {(*i, x2, x3) | x, > 0 and x2 > 0 and x3 ^ 0} of J?3. It is easy to show that B is homeomorphic to the ball B1, so that the fixed-point theorem holds for continuous maps of B into itself. ' Now if x = (*„ x2, x3) is in B, then all the components of x are non- negative and at least one is positive. Because all entries of A are positive, the vector T(x) is a vector all of whose components are positive. As a result, the map x —> T(x)/\\ T(x)\\ is a continuous map of B to itself, which therefore has a fixed point x0. Then so that T (and therefore the matrix A) has the positive real eigenvalue nn*a)ii. n Before leaving the subject of vector fields, we should mention one of the most interesting problems to which they lead: Given a surface S, does if possess a nonvanishing tangent vector fieldl Seemingly a problem of differential geometry, this is, in fact, a problem of topology, for the answer depends only on the topological type of 5". It turns out that among compacl surfaces, only the torus T and the Klein bottle K (if you know what that is) have nonzero tangent vector fields. They are illus- illustrated in Figure 27. [The torus, in fact, has two linearly independent vector fields.] The proof of this fact is outside the scope of this book. (See [G-P] for a proof.) But we can proceed one step toward a solution by showing that the sphere Sz, at least, possesses no nonvanishing vector field. We shall sketch the proof geometrically, rather than carry it out formally using equations. Figure 27
§ 8-10 Theorem 10.4. The sphere S2 possesses field. Vector Fields and Fixed Points no nonvanishing tangent vector 367 Proof. Suppose that 5» had a nonvanishing vector field (* v(xi) Con sider the north po ep = @,0, 1) of S*; we shall assume for con^en^e thai at p the vector field is parallel to the ,-axis. Take a small openTil U in £ centered at p, so small that on the ball V the vector field does not vary more than a few degrees from being parallel to the y-axis. Now consider the map/: S> -p -* R* given by "stereograph* projec- projection." (See Step 1 of the proof of Theorem 6.2.) The map/is, in fact, a homeomorphism of Sz —p with R2. More than that, it carries tangent vectors to Sz continuously into tangent vectors to R2. How ? The easiest way to see what happens is to take a given tangent vector v at a point .*, and to find a curve C in S having that vector as its velocity vector at x. The map/carries the curve C into a curve in R2, and we let it take the vector v into the velocity vector of the image curve /(C) at the point f(x). Let us denote the image of (x, v) by (y, w). You can check by direct computation that / is smooth (w is a well-defined, continuous function of (x, v)) and nonsingular (w is nonzero if v is nonzero). Now we ask the question: What happens to the nonvanishing vector field {x, v) under this map? It is carried into a nonvanishing vector field {y, w) on R2. In particular, consider the subspace S2 - U of S2, which the B Figure 28
368 The Fundamental Group and Covering Spaces Chap, g map/carries onto a ball B in R* of Urge radius about the origin WW a the image of the vector field look like? We have sketched in Figure 28 it looks like on the large circle S that is the boundary of A So far so good. But let us examine this vector field (y, w(y)) on fi mo closely, particularly its restriction to the circle S. You can see from the nictur! that the map h : S —> R* — 0 defined by carries a generator of ^,E, x0) to twice a generator of n,(R2 - 0, x0). Intui- lively, as y goes around the circle S once, the point h(y) goes around the origin twice. Now comes the contradiction. We know that h is extendable to a map of B into R1 — 0, because (y, w(y)) is a nonvanishing vector field on B. Thus A must be the zero homomorphism of fundamental groups. On the other hand, we know both fundamental groups in question are infinite cyclic groups, and that h* carries a generator of the first to twice a generator of the second. In particular, h^ is not the zero homomorphism. □ These theorems have generalizations to higher dimensions, which are discussed in Exercises 7 and 8. Exercises 1. Show that if A is a retract of B1, then every continuous map/: A —* A has a fixed point. 2. Show that if A is a nonsingular 3 by 3 matrix having nonnegative entries, then A has a positive real eigenvalue. 3. Show that the set B of Corollary 10.3 is homeomorphic to B1. *4. Try to give an algebraic proof of Corollary 10.3. (This exercise is for those stu- students who are tempted by the thought that this result is trivial! You might try the 2 by 2 case first.) 5. Show that if/: S' — S' is inessential, then/has a fixed point and/carries some point x to its antipode ~x. 6. Show that given a continuous map/: S* — S\ either/has a fixed point or/ carries some point x to its antipode -x. [Hint: Use Theorem 10.4.J 7. Suppose that you are given the fact that for each n there is «^ , . B"* ■ — S". (This result can be proved using techniques of algebraic ogy). Derive the following corollaries: (a) The identity map i: S" —<■ S" is essential. (b) The antipodal map a : S" —► S" is essential. (c) The inclusion map/: 5" — *"+1 - 0 is essential.
§8-11 . Homotopy Type 369 (d) Every nonvanishing vector field on S"+' points dir«-tiv n , a point of S-. and direct.y inward at some pZtoiT* " " ""• (e) Every continuous map /: B« —> B" has a fixed point (f) Every n by « matrix with positive real entries has a positive real (8) !4: 7"" meSSentia' th/h « 8. It is a theorem that S» has a nonvanishing tangent vector field if and only if n is odd. The "if" part is easy. Show that if,, = 2/n - 1, then v(x) = (-*2, at,, -x4, a-3 -x2m, x2m.,) is a nonvanishing tangent vector field on 5". The "only if" part is more difficult- see Exercise 7 of §8-11. 8-11 Homotopy Type In §8-8 we considered the problem of determining whether or not a given map h: X—> Y was homotopic to a constant map. Now we consider the more general problem of determining whether or not two given maps h, k : X—> Fare homotopic. We prove that a necessary condition for h and k to be homotopic is that, up to an isomorphism of the groups involved, h* and k; are equal. It turns out that this theorem has an application to the problem of com- computing fundamental groups. We have already noted, in §8-5, that two spaces have isomorphic fundamental groups if one is a strong deformation retract of the other, and we have used this fact in computing some fundamental groups. But there is a relation more general than this, called homotopy equivalence, which also implies that the spaces in question have isomorphic fundamental groups. We study that relation in this section. First, let us consider what happens when two maps are homotopic. Theorem 11.1. Let h, k : X — Y. Let h(x0) = y0 and k(x0) = y,. Ifh and k are homotopic, then there is a path a, in Y from y0 to y, such that k* - S. o h*. 7/y0 = y,, and if the base point x0 remains fixed during the homotopy, then k» = h». *,(,y,) Proof. Let H: X X /-> Y be the homotopy between A an _d *• then *<*. 0) = /,(.v) and H(x, 1) = k(x) for x 6 X. Let « :I-Y be dtf nedby the equation «(,) = H(x0, /); then a is a path in Y from ^o to y,. We assert that kt = a o A*.
370 The Fundamental Croup and Covering Spaces Chap. 8 That is, we assert that for every loop/: / —* X in X based at *0, we have **([/]) = <*(**([/]))• We must prove that [k o /] = [a] * [h o f] * [a], or, equivalently, that (*) [a] * [Arc/]*[«] = [WJ- Checking this equation requires the construction of a path homotopy c. We describe G geometrically as follows: For given parameter value /, let a, be the path that goes partway along a, from >-0 to a(f), reparametrized so as to be based on the interval [0, 1]. And Jet fi, be the loop fi£s) = H(J(s), t) / V m 'i a I _—*- w ' A / \ i t (h-f) f) ) / / in Y based at a(/). See Figure 29. Then consider the loop (a, * J3,) * a, based at y0. When / = 0, this loop equals 0™ which is path homotopic to A ^/; and when / = 1, it equals the loop (a * (k o /)) * a. ( Thus setting G(s, t) - ((at * #) * a,)(s) gives us the path homotopy rove equation (*). Formally, we define required Thus setting G(, ) to prove equation (*). l define G : I x I'—» Kas follows: for , 0 = f/(/D* - 1), 0 for s e for j e aB*(l - s)) for j e [i, 1]-
§8-11 Homotopy Type 371 You can check that these formulas make sense, that they agree when the domains overlap, and that G is the required path homotopy Finally, we note that if the homotopy H leaves the base point x fixed at the point , then the path «(,) _/,(,,,,) h the ConstanPt „,£ ^ /J. Le/ h, k: X -> Y; to h(x0) = y0 Wk(x0) = y,. Suppose that h and k are homotopic. //k* /J injective (or surjecthe, or the zero homomor. phism), then so is h*. In particular, if h is homotopic to a constant map then h* is the zero homomorphism. Now we turn to the second problem we mentioned at the beginning of this section, the problem of computing fundamental groups. We obtain a gen- general condition under which two spaces have isomorphic fundamental groups. Definition. A continuous map/: X—* Yh called a homotopy equivalence if there is a continuous map g : Y—► Xsuch that go/is homotopic to the identity map ix of X and/o g is homotopic to the identity map ir of Y. The map g is said to be a homotopy inverse for the map/ It is easy to show that given any collection C of topological spaces, the relation of homotopy equivalence is an equivalence relation on e. See Exer- Exercise 2. Two spaces that are homotopy equivalent are said to have the same homotopy type. The notion of homotopy equivalence generalizes the notion of strong deformation retraction, defined in §8-5; if A is a strong deformation retract of X, then A has the same homotopy type as X. For \et j: A — X be the inclusion mapping and let r: X-* A be the retraction mapping. Then the composite r oj equals the identity map of A, and the compositey ° r is by hypothesis homotopic to the identity map of X (and in fact each point of A remains fixed during the homotopy). Under the hypothesis that two spaces are homotopy equivalent, we can prove that they have isomorphic fundamental groups: Theorem 11.3. Le,f:X->Y be continues; let f(x0) = y0. Iff is a homo- topy equivalence, then f,:*,(X,xo)—**i(Y.yo) is an isomorphism. Proof. Lctg: Y^ X be a homotopy inverse for/ Consider the maps (X,Xo)-^(Y,yo)-^(X,Xl)-^ (Y.K). f(r \ We have the corresponding induced where x, ^g(y0) and y, =/(*i)- we nave homomorphisms:
372 The Fundamental Group and Covering Spaces Chap, g homom°rphisms ^entity map, there is In particular, (g °/)» = g* o (/J+ is an isomorphism. Similarly, because/og is homotopic to ;V, the homomorphism f/.rt = (/*,)* ° g* ls an isomorphism. The first fact implies that g* is surjective, and the second implies thatg, is injective. Therefore, g* is an isomorphism. Applying the first equation once again, we conclude that (/«)• = (*•)-'• «> so that (/,„)* is also an isomorphism. Note that although g is a homotopy inverse for/, the homomorphism £, is not an inverse for the homomorphism (/„)*. Q The relation of homotopy equivalence is clearly more general than the notion of strong deformation retraction. The theta space and the figure eight are both strong deformation retracts of the doubly punctured plane; we noted this fact in Examples 3 and 4 of §8-5. Therefore, they are homotopy equivalent to the doubly punctured plane, and hence to each other. But neither is homeomorphic to a strong deformation retract of the other; in fact, neither of them can even be imbedded in the other. (See Exercise 4.) It is a striking fact that the situation which occurs for these two spaces is the standard situation with regard to homotopy equivalences. Martin Fuchs has proved a theorem to the effect that two spaces X and Y have the same homotopy type if and only if they are homeomorphic to strong deformation retracts of a single space Z. The proof, while it uses only elementary tools, difficult [F]. Exercises 1. Check the details concerning the homotopy G constructed in the proof of Theo- 1. Thow^that given a collection 6 of spaces, the relation of homotop is an equivalence relation on C.
§ 8-1' '■.'-• Homotopy Type 373 3. Show that X has the homotopy type of a one-point space if and only if X is contractible. (See Exercise 4 of §8-8.) 4. Show that neither of the spaces 6 and 8 can be imbedded in the other. 5. Let/< be a subspace of X;letj:A—> A"be the inclusion mapping; let/: X—* A be a continuous map. Suppose that the map; °/: X—> Xis homotopic to the identity map ix : X —► X under a homotopy H: X x / —► X. (a) Show that if H(a, t) e A for every a e A, then/, and/, are isomorphisms. (b) Show that if/is a retraction, then/, and/, are isomorphisms. [In this case, H is called merely a deformation retraction rather than a strong deformation retraction.] (c) Give an example to show that/, and/, need not be isomorphisms in general. 6. We define the degree of a continuous map h : S1 —> S1 as follows: Let b0 be the point A, 0) of S'; choose a generator y for the infinite cyclic group it\(Sl, b0). (There are only two, differing in sign.) If x0 is any point of S', choose a path a in S' from b0 to a-0, and define y(.x0) = dt(y). Then y(x0) generates n^S1, .v0). The element y(x0) is independent of the choice of the path a, since the fundamental group of Sl is abelian. Now given h : S1 —► S', choose .v0 e S< and let h(x0) = x,. Consider the homomorphism //, :*,(£', *„)—>ni(S',xl). Since both groups are infinite cyclic, we have (*) /'*0'(.v0)) = d.y(X>) for some integer d. The integer d is called the degree of h and is denoted by deg ft. The degree of/; is independent of the choice of the generator y; choosing the other generator would merely change the sign of both sides of (*). (a) Show that d is independent of the choice of .v0- (b) Show that if //, k: 5' —> S1 are homotopic, they have the same degree. (c) Show that deg (/; ° k) = (deg /;)-(deg A). (d) Compute the degrees of the identity map, the Teflection p(x,, x2) = (ai, -.v2), the constant map, and the map h(z) = z", where z is a complex number. *(e) Show that if //, k : S< —♦ S1 have the same degree, they are homotopic. 7. Suppose that to every map //:£"—♦ S" we have assigned an integer, denoted by deg/; and called the degree of//, such that: (i) Homotopic maps have the same degree, (ii) deg (// o k) = (deg /;)• (deg A). (iii) The identity map has degree 1, the constant map has degree 0, and the reflection map p(x,, .. . , avu) = (-Vi, • • •. -*»+i) has deS^e ~L [One can construct such a function, using the tools of algebraic topology. Intuitively, deg // measures how many times // wraps S" about itself; the sign tells you whether ft preserves orientation or not.] Prove the following: (a) There is no retraction r : B"+l —► S". (b) If//: S» —♦ S" has degree different from (-1)"*1, then ft has a fixed point. [Hint: Show that if ft has no fixed point, then ft is homotopic to the antipodal map.]
374 The Fundamental Group and Covering Spaces Chap 8 (c) If h : S" -r* S" has degree different from 1, then h maps some point* to its antipode —x. (d) If Sn has a nonvanishing tangent vector field v, then n is odd. [Hint: If v exists, show the identity map is homotopic to the antipodal map.] (e) Every map h : Slm —► Slm has either a fixed point or a point x such that h(x) = -x. (f) Compare these results with Exercises 6, 7, and 8 of §8-10. For the use of the notion of degree in differential geometry, see [G-P]. 8-12 The Jordan Separation Theorem We now consider one of the classical theorems of mathematics, the Jordan curve theorem. It states a fact that is geometrically quite "obvious," the fact that a simple closed curve always separates the plane into two pieces. But it is quite difficult to prove by direct geometric arguments. Here we prove it as a corollary of our study of covering spaces and the fundamental group. There are actually three separation theorems involved. The first, which we call the Jordan separation theorem, states that a simple closed curve in S1 (or in the plane) separates it into at least two components. The second says that an arc does not separate S2 (or the plane). And the third, the Jordan curve theorem proper, says that a simple closed curve C in Sz (or in the plane) separates it into precisely two components, of which Cis the common boun- boundary. We shall prove the first of these theorems now. We need the following fact: Lemma 12.1. Let a and b be points o/S2; let A be a compact space; let f:A —► S2 — a — b be a continuous map. Jfz. and b lie in the same component o/S2 — f(A), then f is inessential. Intuitively, this lemma says that if the set f(A) does not separate a from b in Sz, then /can be shrunk to a point without touching a or b. Proof. Step 1. We show there is a homeomorphism /; of Sz with the one- point compactification Rz U {oo} of R2 such that h(a) = co and h(b) = 0. To construct /;, we first take a rotation h, of Sz that carries a to the north pole/?. We then take stereographic projection hz, mapping S2 — p homeomor- phically onto R2; we extend it to S2 by letting hz(p) = oo. Finally, we take a translation h3 of/?2 carrying //2(/z,F)) to the origin 0, and we extend it to R2 u {o°}by letting/;3(oo) = oo. It is easy to check that h2 and h3 arehomeo- morphisms, as is //,. Then h3 o h2 o h, is the desired homeomorphism. Step 2. In view of Step 1, the lemma reduces to the following statement: Let A be compact; let/: A —» R1 — 0 be a continuous map. If 0 lies in the unbounded component of i?2 — f{A), then/is inessential.
§8-12 The Jordan Separation Theorem This statement is easy to prove. Assume that 0 lies in the unbounded com- component of Rz — /(^)- Choose a ball B centered at the origin, of sufficiently large radius that it contains the set f(A). Choose a point p of R1 lying out- outside B. Then 0 and p lie in the same component of R* — f(A). Because R* is locally path connected, so is the open set R2 — f(A). Therefore, the components and path components of Rz — f(A) are the same. Hence we can choose a path a in R1 — /(A) from 0 to p. We define a homo- homotopy F: A X /-— R1 — 0 by the equation it is pictured in Figure 30. The homotopy Fis a homotopy between the map /and the map g defined by g(x) = /(*) — p. Note that F(x, /) ^ 0 because the path a does not intersect the set f{A). Figure 30 Now we define a homotopy G : A X /--> R2 - 0 by the equation G(x, I) = //(,) - p. It is a homotopy between the map g and a constant map. Note that G(x, t) =* 0 because //(*) lies inside the ball B and p does not. Thus we have proved that/is inessential. □ A ^efi!Ut'?"' A" ar° 'S a Space h°meomorphic to the unit interval [0, 11 A s.mple dosed curve is a space homeomorphic to the circle S'. cseparation theorem)-Let cbe asimple chsed — C is not connected.
575 The Fundamental Group and Covering Spaces Chap. 8 ■ Proof. Because S2 - C is locally path connected, its components and path components are the same. We assume that S2 - C is path connected and derive a contradiction. We are going to apply Theorem 6.1, the special case we proved of the Van Kampen theorem. It expresses the fundamental group of X = U u V'm terms of the fundamental groups of U and V, provided the intersection U r\ V is path connected. Let us first write C as the union of two arcs A v and A2 that intersect only at their end points a and b. Then let X denote the space S2 — a — b. The space AT is homeomorphic to the punctured plane R2 — 0, so its fundamental group is infinite cyclic. Let XJ be the open set S2 — A v and let V be the open set S1 — A2; then X = U U V. Now u n v = s2 - (Al u a2) = s2 ~ c, so that U D Vis path connected by assumption. Let x0 be a point of U n V. If we can prove that the inclusions i:(U,xo)—»(X,xa) and j: (V, xD) —* {X, x0) induce zero homomorphisms of the fundamental group, Theorem 6.1 will apply to show that n^X, x0) = 0. This contradicts the fact that x^X, x0) is infinite cyclic, so our assumption that S2 — C is path connected must have been mistaken. Let us prove that /* is the zero homomorphism. Given a loop/:/—* V based at x0, we show that /*([/]) is trivial. For this purpose, let <f> :I-~*Sl be the standard loop generating n^S1,^). The map/:/—* U induces a continuous map h : S1 —> U such that h o <j> =f. See Figure 31. Consider the map i ° h : Sl -* £2 - tf - 6. By hypothesis, the set /(ACS1)) = KS1) does not intersect the arc Ax joining a and b. Therefore, a Figure 31
§8-12 The Jordan Separation Theorem 377 and * lie in the same component of Sz — i(h(S<)). By the preceding lemma, the map i o h is inessential. It follows from Theorem 8.2 that (/ o h)* is the zero homomorphism of fundamental groups. But Therefore, /"*([/]) 's trivial, as desired. □ Exercises 1. By identifying S1 with R1 U {co}, prove that every simple closed curve in R1 separates R1. 2. We assumed in the proof of Theorem 12.2 that the simple closed curve C cannot equal all of S2. Why is this justified? 3. Give examples to show that a simple closed curve in the torus may or may not separate the torus. 4. (a) Lemma. Let A and B be closed connected subsets of S1 whose intersection consists of precisely two points. Then A u B separates S1. (b) Let C be a subset of S1 homeomorphic to the space which is the union of the topologist's sine curve and the broken-line path from @, I) to @, —2) to A, -2) to A, sin 1). (See Figure 32.) Show that C separates S1. Figure 32 *5. This exercise leads to a proof of the invariance of domain theorem for R2. (a) Prove the following: Lemma (Homotopy extension lemma). Let X and X x / be normal; let A be a closed subset ofX. Iff: A — R1 — 0 is a continuous map andf is inessential, then f can be extended to a continuous map g ; X —► R2 — 0. [Hint: Define a map of the subspace Y = (A x /) u (X x 1) of X x / into R1 - 0 by using the homotopy between /and a constant map. Extend to a neighborhood V of Y in X x /. Then construct a map of X x 0 into V that equals the identity on A x 0J (b) Prove: Theorem (Borsuk). Let X be a compact subset o/R1. I/O lies in a bounded
378 The Fundamental Group and Covering Spaces Chap. 8 component C of R} - X, then the inclusion map j- * — jp _ 0 /. and conversely. ° b essential; [Hint: Let B be a large ball that contains C and X. that the inclusion map of B - C into R2 - 0 can map of B into .R2 — 0.] (c) Prove: Weore/n (/< nonseparation theorem). No arc separates R2 ■ no soao hn morphic to B1 separates R2. "°meo- (d) Prove the following. Another proof is outlined in §8-13, Exercise 9 Theorem (Bronwer theorem on invariance of domain for R1). If Uisano subset ofR2 andf: U —> R2 is continuous and injective, then /({/) & Open inZ and f is an imbedding. [Hint: If/: B2 — R2 is an imbedding, note that f(S<) separates W and f(B2) does not; conclude that /(Int B2) must equal one of the components of R2-f(S').] ^ 8-13 The Jordan Curve Theorem The special case of the Van Kampen theorem, which we just applied in proving the Jordan separation theorem, tells us something about the funda- fundamental group of the space X = U U V in the case where the intersection U n Fis path connected. In the next two lemmas we examine what happens when U n Fis not connected. In the first lemma we suppose that U C\ Khas at least two components, and in the second we assume that it has at least three. These lemmas will enable us to complete the proof of the Jordan curve theorem. The proof we give is, as far as we know, a new one. It involves a construction similar to that used in complex analysis to construct Riemann surfaces. But it makes no use of normality or the Tietze theorem. Lemma 13.1. Let X be the union of two open sets U and V. Suppose that U n V can be written as the union of two disjoint open sets A and B. Let a e A and b e B; suppose that a and b can be joined by paths in U and in V. Then Proof. The proof is in many ways an imitation of the proof in §8-4 that the fundamental group of the circle is infinite cyclic. As in that proof, the crucial step is to find an appropriate covering space E for the space X. Step 1. {Construction of E). We construct E by pasting together copies of the subspaces U and V. Let us take countably many cop.es of U countably many copies of V, all disjoint, say U X {In) and V X B/t + 1)
§8-13 The Jordan Curve Theorem 379 space of Y by identifying the points x X B«) and x x B/r — 1) for x e A and by identifying the points x X B«) and x X B« + 1) for x e 5. Let ;r: Y~* E be the quotient map. Now the map />: Y —* X defined by p(x x m) —x induces a map p: £ , X; the map p is continuous because £ has the quotient topology. The map p is also surjective. We shall show that p is a covering map. The space E is illustrated in Figure 33. u (-2) ! ! u Figure 33
380 The Fundamental Group and Covering Spaces Chap. 8 First let us show that the map n is an open map. Since Y\s the n • , the disjoint open sets [U X Bn)} and {V X Bn +1)}, it" m sufficeT* * that n\iU X 2n) and *,(K x B« + ,)) are open ips. And Thfs 2 £? Take an open set in U X 2«, for example; it will be of the form W *T" Where W is open in U. Then ^X2"> tT'Gr(^ X 2n)) = [ICx 2n] U [(W n 5) x Bn + 1)] U K^ n A) X B« - 1)], which is the union of three open sets of yand hence open in Y. By definition of the quotient topology, n(W X 2n) is open in E, as desired. Now we prove that/? is a covering map; we show that the open sets Vmi V are evenly covered by p. Consider U, for example. The set p-'(U) is the union of the disjoint sets n(U X In) for n e Z. Each of these sets is open in E because n is an open map. Let nln denote the restriction of n to the open set U X 2n, mapping it onto n{U X In). It is a homeomorphism, because it is bijective, continuous, and open. Then when restricted to n{U X In), the map p is just the composite of the two homeomorphisms n(U X 2n) -^*- U X In -^-* U and is thus a homeomorphism. Therefore, p | n(U X 2n) maps this set hoineo- morphically onto U, as desired. Step 2. Now we prove that n^X, a) ^ 0. Take a path a in f/from a to b, and a path /? in F from 6 to a. We assert that the loop a * /? based at a is not path homotopic to a constant. Let us lift a * fi to a path on £ beginning at the base point e0 = n(a X 0). Define a(/) = n{a.(t) x 0), £(/) = 7r(/?@ x 1). Since a(l) = 7tF X 0) = n(b X 1) = A@), 'he path & * $ is defined. It is easily seen to be a lifting of a * /?, because /? ° a = a and p ° ^ = /?• Since the lifted path a * j5 begins at the base point e0 and ends at a point different from the base point, the loop a * /? cannot be path homotopic to the constant loop. □ Lemma 13.2. Let X be the union of two open sets UandV. f^^ U O V can be written as the union of three disjoint open sets A, A, an • Let a e A, a' e A>Wb e B; suppose that all three points can be joined by paths in U and in V. Then 7r,(X, a) is not infinite cyclic Proof. Let . and P be paths in U and ^spectively,^JJjSn from b to a, respectively. Let 5 and y be paths ,n </and^P™of tbe fl to *' and from a' to a, respectively. Wnt.ng J/ n Kas the un.o disjoint open sets
§8-13 The Jordan Curve Theorem 381 • A and A' u B> we see from the proof of the preceding lemma that both the loop a * fi and the loop S * y represent nontrivial elements of nx(Xt a). Now if 7Ci(Xt a) were infinite cyclic, each of these loops would represent a multiple of the generator of the group. Thus there would exist nonzero integers n and m such that n[a * ft] = m[S * y]. We shall prove that such integers do not exist. Apply the construction of the preceding lemma to this situation, writing U C\ V now as the union of the two disjoint open sets A u A' and B. One obtains the covering space £ of A' as before. See Figure 34. (Remember that we never assumed A and B were connected when we constructed E.) Figure 34 Now consider .he loops « . f and S . y and lift then, to paths ring at <■„. It is easy to see that S . , lifts to a loop in E wh.Ie . . More generally, every ,m,MPle of « . r »«' lift t0 a loOP "*^ nonzero multiple oTt.fi lifts to a path that begms at *. and ends no mu,tiple of , . , can „ path homo.op,c to a of Now we prove that no arc in S2 scpmai*- ^ - —-- r Exercise 5(c) of the preceding section. Here is another, if you have worked the exercise cited.
382 The Fundamental Group and Covering Spaces Chap, g Theorem 13.3 {A nonsepamtion theorem). Let A be an arc in s* S2 — A is connected. m * • not in Z). We assert that if a and 6 can be joined by paths in S2 _ n S2 - ZJ, then they can be joined by a path in S1 - D Figure is ,n ! the fact that this assertion is not entirely trivial. ' '"""rates D, fiffnrc 55 We suppose that a and b cannot be joined by a path in S1 — Z>and derive a contradiction. We apply Lemma 13.1. Let X be the space S1 — d. Let U and K be the open sets U = S1 — Z>, and K = S2 — Z>2. Then X = (/ U K, and (/ n K = 52 - Z>. By hypothesis, o and ft are points of S2 — D that cannot be joined by a path in S1 — D. Therefore, Un V \s not path connected. Let A be the path component of U n V containing a; let B be the union of the other path components of U n V. Since U n ^1S locally path connected (being open in S-), the path components of U O Kare open; hence A and B are open in X We are given that a and b can be joined by paths in U = S- - D, and K = S2 - D2. We conclude from Lemma 1J.I that b^A', a) ^ 0. , But A- is the space S2 - d, which is homeomorphic to R1 and is m simply connected. Hence it must be possible to join a and b by a pa »<•/> 2. Now, given the arc A and the points a and b in S2 -A « pose that a and ft cannot be joined by a path in 5J - A and derive tradiction. This proves our theorem. , Choose a homeomorphism A : [0, 1] — A; let A, = MO, i» «"«>
§8-13 , \ We Jordan Curve Theorem 383 B2 = A([i, 1]). We conclude from Step ] tha TandfftT" r^'.*1* paths in both S> - B, and S2 - 5, an"Ot bejOlned Continue similarly. In this way we define a sequence 1 => 7, => ]2 =) ... of closed intervals such that 7. has length ($)■ and such that for each* the points a and ft cannot be joined by a path in S2 - *(/„). Compactness of the unit interval guarantees there is a point x in f)lal since the lengths of the intervals converge to zero, there is only one such point. Then consider the space S1 - h(x). Since this space is homeomorphic to R2, the points a and ft can be joined by a path a. in S1 — h{x). Since h(x) = f~)h(In), the space S2 — h(x) equals the union of the open sets (S2 - /;(/,)) <= (S* - /;G2)) c •••. Because a(/) is compact, it must lie in the union of finitely many of these sets, and hence in some one of them, say S2 — h{lm). But then a, is a path in the set S1 — h(Im) joining a and b, contrary to hypothesis. □ Theorem 13.4 (The Jordan curve theorem). Let C be a simple closed curve in S2. Then S2 — C has precisely two components W, anclW2, of which C is the common boundary. [That is, C = W, - W, = W2 - W2.] Proof. Step J. We first prove that S2 - C has exactly two components. We apply Lemma 13.2. Write C as the union of two arcs C, and Cz having precisely two points p and q in common. Let X be the space S1 - p - q- Let U and V be the open sets U = S2-C, and V = S*-C1. Then X=UuV.AndUnV equals S2 - C, which has at least two com- components, by Theorem 12.2. , ., b Suppose that S2 - C has more than two components. Let A«AA* two of the components and let B be the union of the remaining ones.1Kcause S> - C is locally connected, each of the sets A, A', and i^^ and ft be points of ^, ^, and 5, respectively. By the precedmg *^ sets C/ = S2 - C, and V = 5> - C2 are path connected (s mce no rates 5>). There/ore, a, a', and ft may be joined by pahs ,n U mistaken.
384 The Fundamental Group and Covering Spaces ~ . Step 2. Now we show that C is the common boundary of W, and W Because S2 is locally connected, each of the components W, and W of S1 — Cis open in S2. In particular, neither contains a limit point of the other so that (K,nH'!=»',n^ = 0. Since S2 is the disjoint union of W' Wl3 and C, both the sets W, - W, and Wt - W% must be contained in C To prove the converse, we show that if x is a point of C, every neighbor- neighborhood U of x intersects the closed set w,—W,. It follows that x is jn the set W, — W,. So let U be a neighborhood of x. Because C is homeomorphic to the circle S', we can break C up into two arcs C, and C2 that intersect only in their end points, such that C, is small enough that it lies inside U See Figure 36. Figure 36 Let a and b be points of W', and H^, respectively. Because C2 does not separate S2 (by Theorem 13.3), we can find a path a, in S2 — C2 joining a and ft. The set a{l) must contain a point y of the set fV, - Wu because otherwise a(l) would be a connected set lying in the union of the disjoint open sets Wx and S* — Wx and intersecting each of them. The point y necessarily belongs to the closed curve C, since (W, - W,) a C. Because the path a does not intersect the arc C2, the point y must therefore lie in the arc C,, which in turn lies in the open set U. Thus U intersects W, — W, in the point y, as desired. □ Example 1. The second half of the Jordan curve theorem, to the effect that C is the common boundary of W, and Wlt may seem so obvious as hardly to require comment. But, in fact, it depends crucially on the fact that C is homeo- homeomorphic to S1.
§8-13 < -.:'.. •:■■:■ The Jordan Curve Theorem 385 For example, the subset C of the plane consisting of the union of the unit circle S1 and the line segment [1, 2] x 0 separates S* into two components W\ and fVz just as the circle does. But C does not equal the common boundary of W, and Wt in this case. See Figure 37. W, Figure 37 There is a fourth theorem that is often considered along with these three separation theorems. It is called the Schoeirflies theorem, and it states that if Cis a simple_closed curve in S2 and C/and Fare the components of S2 — C, then {/and V are each homeomorphic to the closed unit ball B2. We are not going to prove this theorem. It is, however, a corollary of theRiemann map- mapping theorem of complex variable theory. The separation theorems can be generalized to higher dimensions as follows: A) Any subset C of S" homeomorphic to S"~' separates S". B) No subspace A of S" homeomorphic to [0, 1] or to some ball Bm sepa- separates S". C) Any subset C of S" homeomorphic to S"'1 separates S" into two components, of which C is the common boundary. These theorems can be proved quite readily once one has studied singular homology groups in algebraic topology. (See [S], p. 198.) The Brouwer theo- theorem on invariance of domain for R" follows as a corollary. The Schoenflies theorem, however, does not generalize to higher dimen- dimensions without some restrictions on the way the space C is imbedded in S\ This is shown by the famous example of the "Alexander horned sphere," a homeomorphic image of S2 in S3, one of whose complementary domains is not simply connected! (See [H-Y], p. 176.) The separation theorems can be generalized even further than this. The definitive theorem along these lines is the famous Alexander-Pontryagin du- duality theorem, a rather deep theorem of algebraic topology, which we shall not attempt to state here. (See [S].) It implies that if C separates S" into k components, so does any subset of S" that is homeomorphic to C (or even homotopy equivalent to C). The separation theorems (l)-C) are immediate corollaries.
386 The Fundamental Group and Covering Spaces Chap. 8 Exercises 1. (a) Show that no arc in Ri separates R1. (b) Show that a simple closed curve C in R* separates R* into two component of which C is the common boundary. 'Ponents, 2. Show that under the hypotheses of Lemma 13.1, *,(*, a) contains an infinite cyclic subgroup. '■■■■rate 3. Lemma. Let A and B be closed connected subsets of S* whose Intersection consists of precisely twopoints. If neither A nor B separates S2,then S* ~(A\JB) has precisely two components. 4. Recall that a theta space is a space which is the union of three arcs a, b, and c each two of which intersect precisely in their end points. (a) Show that a theta space in S2 separates S2 into precisely three components [Hint: If C is the component of S2 - (a u b) not intersecting c, then Cue separates S2 into precisely two components, by the preceding exercise.] (b) Show that these three components have boundaries a u b,auc,andb\jc, respectively. [Hint: Show that C is one of these components; then apply symmetry.] (c) Theorem. The gas-water-electricity graph cannot be imbedded in S2. (See Figure 13 of §7-9.) [Hint: Consider the theta space obtained by joining H, and H2 to G, W, and E.I 5. Suppose that a,, a2, a3, and a4 are four distinct points of S2, and that foreach pair a,, a, of these points (i^j), we have an arc a,a, in S2 joining them. Suppose, moreover, that each pair of these arcs intersect in at most a common end point. The union G4 of these arcs is called the complete graph on four vertices. (a) Let C be the curve (aiaz) U (a2a3) u (a3ai) U (a4"i). whicn we denote by 0,0203040, for short. Let A and B be the components of S2 - C. Show that if A is disjoint from ata3, then B is disjoint froma204- [Hint: Assume Exercise 4. Note that A is one of the three components of S2 -(Cu 0^3), and these three components have boundaries C, ata2a3au and axa3a,.ax, respectively.] (b) Show that S2 — G4 has four components, whose boundaries are axa2a3ax, a,a3a4a,, a,a2a4a,, and a1a3aia1, respectively. (c) Theorem. The complete graph oil five vertices cannot be imbedded in 6 .(see Figure 13 of §7-9.) 6. (a) Assume the hypotheses and notation of Lemma 13.1. Suppose V and\V are path connected and *,(**) is infinite cyclic. Show that f**™*. connected and [« * 0 ,«««<« *,(* ")■ t^ Use Exerc.se 10 of &U (b) Consider the complete graph Gt on four vert.ces a be, d; Wl«*« a subset of S>. Show that Up is an interior point of the arc « and , h th l bda enerates «« ^Jajn a subset of S. Show that Up is a p interior point of the arc bd, then the loop abeda generates «« ^Jajn group of S* - /> - q. [Hint: Let tf = S2 - paba and V -S J^ x be an interior point of ad and y be an interior point of be,
§84 • The Classification ofCovering Spaces 387 T^ycZ}^ ^ d'fferent C°mpOnentS Of U" V- ^ *=xaby and 7. Prove the following: Theorem. Let A be a simple closed curve in &; let 0 lie in the bounded com. ponent of R* - A. Then A generates the fundamental group ofRi-o Proof. Let 5 = «>u {oo} be the one-point compactification of ** Let B be the bounded component of R* — A. (a) Construct an arc D in S containing oo that intersects A in precisely its end points a and c. (b) Construct an arc E in S containing 0 that intersects A in precisely its end points b and d, where b and d lie in different components of A - a - c. [Hint: Show that if x and y are two points of the open, connected subset U of S2, then (i) there is an arc in (/joining x and y, and (ii) there is an homeomorphism h:S* —► S2 which equals the identity outside U and maps x to y. Let A, and A2 be the components of A — a — c. Choose a point x of B; join x to oo by arcs in S — A, and S — A2; obtain arcs joining x to b and rf; then an arc from b to rf; then apply (ii) to obtain an arc containing 0.] (c) Apply Exercise 6. 8. Prove the following basic theorem of complex analysis: Theorem. Let A be a piecetvise-differentiable simple closed curve in the com- complex plane C; let z0 be a point of C not on A. Then the integral has value ±1 ifz0 lies in the bounded component ofC — A, and it has value 0 lfz0 lies in the unbounded component. [Hint: Use the preceding exercise and Exercise 6 of §8-5.] 9. A proof of the invariance of domain theorem for R2 was outlined in the exercises of §8-12. Give an independent proof based on Exercise 7 of this section. 8-14 The Classification of Covering Spaces Up to now we have used covering spaces primarily as a tool forcomputing fundamental groups. Now we are going to turn things around and use tne fundamental group as a tool for studying covering spaces. It turns out tnat one can determine all possible covering spaces of a given space B, merely oy examining the fundamental group of B. We explain this process now If H, and H> are subgroups of a group G, you may recall from^gebn. that they are said to be conjugate subgroups if H2 =«•«,-« ">r > ment /of G. Said differently, they are conjugate if the .somorpus* ofG with itself which maps x to «•*•«-' carries the group H, onte ft »^UP ^ It is easy to check that conjugacy is an equivalence relat.on on the collects
388 The Fundamental Group and Covering Spaces Chap, $ of subgroups of G. The equivalence class of the subgroup H is called * conjugacy class of H. " caned the Now suppose we have a fixed path-connected space B and a fixed bas point b0 e B.LetpiE-^Bbea. covering map, where E is path connected6 If we choose a point e0 in ;r ](&0) as a base point for E, then we have an induced homomorphism (which is, in fact, injective). The image group Ho =p*(ti1(E, e0)) depends of course on the choice of the base point e0 in E, but in a very simple way. We shall show that as e0 ranges over all the various points of/r>(&0), the image subgroup ranges over those subgroups of G that are conjugate to Ho. Thus to each path-connected covering space p : E —> B of B we can assign a certain conjugacy class of subgroups of n^B, &„), the collection of the images of the groups n,(E, e) under the homomorphisms induced byp, for e e p~'(b0). It turns out that every conjugacy class corresponds to a covering space of B under this assignment, and that this conjugacy class determines the cover- covering space uniquely, up to homeomorphism. This is the basic classification theorem for covering spaces. More precisely, we define two covering spaces p : E —► B and p': E' —♦ B to be equivalent if there is a homeomorphism h:E'—>E such that p ° h =p'. B Then the two main theorems concerning classification of covering spaces can be stated as follows: A) For every conjugacy class of subgroups of itt(B,bB), there exisib path-connected covering space p : E - B of B such that the groups />,(*,(£, e% for e in/r'(*„), form this conjugacy class B) Two path-connected covering spaces of B are equ.valent if andI o ny if they correspond to the same conjugacy class of subgroups *,(*,&,) «nder this correspondence. By means of these theorems, the topolog.cal problem of . f pace ^ By means of these theorems, the topolog.ca p .f^. possible path-connected covering spaces of a S'ven space ^ »£*£ ^ pletely to the algebraic problem of determining all conjugacy classe arbitrary path-connected space B. We shall ^ niceness" conditions on B in order to carry out the 14.3 and 14.4.)
§8-14 The Classification of Covering Spaces 389 Lemma 14.1. Let p : E —»• B be a covering map, where E and B are path connected. Let b0 G B. As e ranges over the points of p~'(b0), the group p,i(fti(E, e)) ranges precisely over a conjugacy class of subgroups ofn,(B, b0). Proof. Given e0, e, e />"'(&„), let ffo=P*(nl(E,eo)) and H, = pt(n,(E, ej). Step I. Let y be a path in E from e0 to e,; let a, be the loop p » y in B based at i0. We show that [a]*H,*[a]-1 <=«■„. Let [h] be an element of Hx; then [A] = pt[h] for some loop h in £ based at e,. Let £ = (j> * Ti) * y. Then £ is a loop based at e0, and Thus [a] * [h] * [a] e //■„, as desired. See Figure 38. Figure 38 Step 2. It follows that Ho and H, are conjugate subgroups. For let y be a path in E from e0 to eu and let a = p o j>. Then j> is a path from e, to eo> and & =p o y. Applying Step 1 twice, we have [a] • H, * [a.]-' <= Ho, [&] * Ho * [a] c:Hx. Thus Ha = [a] * H, * [a], as desired. Step 3. Now given e0 e /r'Fo)» and given a subgroup H of 7^E, fr0) conjugate to Ho, we wish to find a point e, of p-'(b0) such that the corre- corresponding subgroup Hi equals #. By hypothesis, Ho = [a] * H * [a]"' for
390 The Fundamental Group and Covering Spaces p. some loop a in B based at b0. Let y be the lifting of a to a path in £beginni at e0, and let e, = y(l). Let H, = pt(n,(E, e,)). It follows from Step 2 that Ho = [a] *Ht* [a]. Then H = Hu as desired. □ Equivalence of coverings Now we prove that two coverings are equivalent if and only if they cor- correspond to the same conjugacy class. We need the following generalized version of the "lifting lemmas" of §8-4: Lemma 14.2 (The general lifting lemma). Let p : E —► B be a covering map; let p(e0) = b0. Let f: Y —► B be a continuous map, with f(y0) = b0. Suppose Y is path connected and locally path connected. The map f can be lifted to a map f: Y —► E such that f(y0) = e0 if and only if f,(»,(Y,y,))cp»(S,(E,e,)). Furthermore, if such a lifting exists, it is unique. Proof. If the lifting / exists, then />*(*,(£, e0)). This proves the "only if" part of ihe theorem. Now we prove that if/ exists, it is unique. Given y, e Y, choose a path a in Y from y0 to y,. Take the path/o a in B and lift it to a path y in E beginning at e0. If a lifting/ of/ exists, then /(;',) must equal the end point y(\) of j>, for / o a is a lifting of/o a that begins at e0, and path liftings are unique. Finally, we prove the "if" part of the theorem. The uniqueness part of the proof gives us a clue how to proceed. Given y, e Y, choose a path a in Y from y0 to yt. Lift the path/o a to a path y in E beginning at e0, and define /(_,;,) = j»(l). See Figure 39. It is a certain amount of work to show that/is well-defined, independent of the choice of a- Once we prove that, continuity of/ is proved easily, as follows: Given a neighborhood N of f(yt), we shall find neighborhood Wofy. such that/(»O c: N. First, choose a neighborhood Uoff(y,) tHat is evenly covered by p. Let Vo be the slice of /r'(t/) that contains/(.y,). Let/>0 be the restriction of p to Vo. By passing to smaller neighborhoods if necessary, can assume that Vo c N. Now choose a neighborhood Wof yt that is pa connected and lies in/" <(U). We claim that f(W) c Vo- For, given/ & > we can choose a path /?in Wfromy, to y. Consider the path/°)?; -tis the path 5 =Po< o/o p in E, beginning at /(;>,)• Then y * 5 is defined,i a lifting of/o (a * p), and it begins at e0. By definition, f(y) equals the point of y * 3, which lies in Vo- Now we show / is well defined. Given two paths a and fi in yt, their images/o a and/o fi are paths in B. Let y and 5
§8-14 The Classification of Covering Spaces 391 Figure 39 respective liftings to paths on E beginning at <?„. We wish to show that Let f be a lifting of/° ft to a path on £ beginning at y(\). See Figure 40. Then y * e is defined and is a lifting to E of the loop (/o a) * (/o ft) in 5. The homotopy class of this loop is/*([a * /?]), which by hypothesis belongs to pJjtx{E, <?„)). Then there is a loop (j> in E based at <?„ such that figure 40
392 The Fundamental Group and Covering Spaces Chap. 8 It follows that e must end at e0. For by Theorem 4.3, if two path • which are path homotopic are lifted to paths on E beginning at th * point, they must end at the same point. Since p o 0 and (/o a) ^rT- are path homotopic, their liftings 0 and y * e must end at the same ° Then y * e ends at e0> so e ends at erf. point- Nowe is aliftingof/0/that begins at y(l) and ends ateo.Thenfisaliftin of/°/3 that begins at e» and ends at y(l). The path 8 is another such lifting B? uniqueness of path liftings, 5 = e. In particular, 5A) = },(]), as desired, n Theorem 14.3. Let B &e /?a?A connected and locally path connected Let p : E —► B and p': E' —► B be path-connected covering spaces of B- let P(e0) = P'(e'o) = b0. Then p and p' ore equivalent covering maps if and only '/P*(ti(E. e0)) fl«^ p'*(?ti(E'. e'o)) ore conjugate subgroups o/s,(B, b0). Proo/. Suppose that the coverings are equivalent. Let A :E'—*E be a homeomorphism such that p o h =p'. Let /z(e'o) = e,. Then because A is a homeomorphism, A#(s,(£", e'o)) = Ki(£, e,). Applying /»# to both sides, we have />;(*,(£•', e'o)) =/>*(*,(£>,))■ By Lemma 14.1, the latter subgroup is conjugate to pJjtx(E, e0)). Now we prove the converse. Suppose that the two subgroups are con- conjugate. In view of Lemma 14.1, we may by choosing a different base point in E' obtain the situation where the two subgroups are equal. So let us suppose this is done, and let e'o now denote the new base point. We apply the preceding lemma. Consider the diagram E' E^B where p' is a covering map. The space E is path connected; it is also locally path connected, being locally homeomorphic to B. Furthermore, />*(«,(£, eo» <=/4(«i(£'. e'o); in fact, these two groups are equal. By the preceding lemma, we can i map/?: E —► B to a continuous map h : E—* E' such that h(e0) = <"o- Reversing the roles of E and E' in this argument, we see that.the ma p' :E' —* B can be lifted to a continuous map k:E'' —» E with kro) ' To show that h and k are inverses of each other, consider the
7116 Classification of Covering Spaces 393 Note that £ o A: £ _»£■ js a lifting to £ of the the condition (* o h)(e0) = e.. Theidentity m p\J^ ^i lifting*,. By uniqueness of liftings, * „ * nJt equal iE. sintilady Example 1. Consider covering spaces of the circle B = S' Because 7it(B, b0) is abehan, two subgroups of «,(*, 60) are conjugate if and only if thev are equal. Therefore two coverings of B are equivalent if and only if thev correspond to the same subgroup of n,(B, b0). Now n,(B, b0) is isomorphic to the integers Z. What are the subgroups of Z1 It is standard theorem of modern algebra that, given a nontrivial subgroup of Z, it must be the group Gn consisting of all multiples of n, for some n e Z+. We have studied one covering space of the circle, the covering/7: R —» S*. It must correspond to the trivial subgroup of 7t,(S<, b0), because R is simply connected. We have also considered the covering p:S' —* S' defined by p(z) = z", where 2 is a complex number. In this case, the map pt carries a gen- generator of 7i,(S',b0) into n times itself. Therefore, the group pt(n,(S',*„)) corresponds to the subgroup Gn of Z under the standard isomorphism of nx{S\ b0) with Z. We conclude from the preceding theorem that every path-connected cover- covering space of S' is equivalent to one of these coverings. Existence of coverings Now we show that given a subgroup H of n,(B, b0), there exists a covering space /?:£—» B and a point e0 of /r'F0) such that pJn^E, e0)) = H. We need to assume an additional condition on the space B, called "semilocal simple connectivity." Definition. A space B is said to be semilocally simply connected if every point b of B has a neighborhood Ksuch that the homomorphism of n^V, b) into n,(B, b) induced by inclusion is trivial. If V is such a neighborhood of b and U is any neighborhoodI of b, then the neighborhood U n V of b, which lies in U, also sat.sfies ^^T Semilocal simple connectivity is weaker than true >^ ""P which would require that within each neighborhood a neighborhood V of b that is itself simply connected. Theorem 14.4. Let B be par/, connected, locally path connected locally simply connected. Let b, e B. <"">£?*% ?«£% exists a path connected covering space p. E — such that p* (*,(E, e0)) = H.
394 The Fundamental Group and Covering Spaces da. 8 Proof. Step 1. {Construction of E). The procedure for constructing v • reminiscent of the procedure used in complex analysis for construct1 'S Riemann surfaces. Let (P denote the set of all paths in B beginning at b& Define an equivalence relation on 6> by setting a ~ 0 if a and 0 end at th same point of B and the class [a • 0] belongs to the subgroup H. This is easi^ seen to be an equivalence relation. ^ Let E denote the collection of equivalence classes; denote the equivalence class of the path a by a*. Define p : E -* B by the equation />(a*) = a(l) Because B is path connected, p is surjective. We shall topologize E in such a way that/; is a covering map. We first note two facts: A) If a =:„ 0, then a* = 0*. B) If a* = 0*, then (a * 5)* = @ * d)* for any path 5 in B beginning at a(l). The first follows by noting that if a —, P, then [a * 0} is the identity element, which belongs to H. The second follows by noting that a * 5 and p * 8 end at the same point, and [(a * 5) * {0 * S)] = [(a * 5) * (<5 * fj)] = [a * 0], which belongs to H by hypothesis. Step 2. (Topo/ogizing E). One way to topologize E is to give (P the com- compact-open topology (see Chapter 7) and E the corresponding quotient topo- topology. But we can topologize E directly as follows: Let a be any element of (?, and let {/be any path-connected neighborhood of a(l). Define B(U, a) = {(a * df 15 is a path in U beginning at a(l)}. We assert that the sets B(U, a) form a basis for a topology on E. See Figure 41. Note that a* belongs to B(U, «,). First we show that if P* e B(U, a), then B(U, 0) = B{U, a). We are given that 0* = (a * 5)* for some path 5 in U. The general element of B(U, 0) is of the form @ * y)*, for some path y in U. By A) and B) above, we have @*y)* = ((a * S) * y)* = (a * E * y))*, which lies in B{ol, U) by definition. Hence B{U, 0) c B{U, a). Conversely, the general element of B(U, a) is of the form (a * e)*, where f is a path in v. Again by A) and B), (a * e)* = ((a *5)*(S* e))* = @*(S* f))#, which lies in B(U, 0). Thus B(U, a) c B(U, 0). ^ Now we show the sets B{U, a) form a basis. If 0* belongs io intersection B(U,,oc,)n B{U2, a2), we need merely choose a path-conn
§8-14 The Classification of Covering Spaces 395 Figure 41 neighborhood Kof y?(l) contained in U, n U2. The inclusion B(V,fi)c:BXU,,fi)nB(Ut,fi) follows from the definition of these sets, and the right side of the equation equals B(UU a,) n B(Ut, a2) by the result just proved. Step 3. The map p is continuous and open. It is easy to see that p is open, for the image of the basis element B(U, a) is the open subset U of B: Given x e U, we choose a path 8 in U from a(l) to x; then (a * 8)* is in B(U, a) and /?((« * 8)*) = x. To show that/? is continuous, let us take an element a* of £ and a neigh- neighborhood W of p{a*). Choose a path-connected neighborhood U of the point />(«*) = a(l) lying in W. Then B(U, a) is a neighborhood of a* that/? maps into W. Thus p is continuous at a*. Step 4. The map p is a covering map. Given bt e B, choose U to be a path-connected neighborhood of Z», that satisfies the further condition that the homomorphism nx{U, bx) —> n,(B, bt) induced by inclusion is trivial. We assert that U is evenly covered by p. First, note that /?-'({/) equals the union of the sets B(U, a), as a ranges over all paths in B from b0 to b,: Since/? maps each set B(U, a) onto U, it is clear that p~'(U) contains this union. On the other hand, if fi* is in p~l(U), then/?(l)isjn {/and we can choose a path 5 in {/frorriyff(l) to Z»,. Let a be the
396 The Fundamental Group and Covering Spaces _, Chap. 8 path fi * 5 from b0 to b,: Then fi ~p a, * §, so that p* = (a „ §* .. belongs to /?(£/, a). Thusp-'(U) is contained in the union of the sets 5A/°? Second, note that distinct sets B(U, a) are disjoint. For if B* belons t B(U, a,) n B{U, a2), then /?(</, a,) = B{U, fi) = /?«/, a2), by Step fl0 Third, we show that p defines a bijective map of B(JJ, a) with £/. it fol- follows that p\B{U, a) is a homeomorphism, being bijective and continuous and open. We already know that/? maps B(U, a) onto U. To prove injectivity suppose that />((«• <5,)#)=/>((a*<52)*), where £, and dz are paths in *7. Then 5,0) = <52A). Because the homomor- phism n,(U, b{) -* n,(B, b,) induced by inclusion is trivial, 5, * 52 is path homotopic in B to the constant loop. Then a * <5, =zp a * <52, so that we have (a * di)* — (a * 5Z)*, as desired. Step 5. The space E is path connected and the subgroup H equals Pt(n,(E, e0)) for some e0 in p~'(b0). This completes the proof of the theorem. First, we calculate specifically what the lifting of a path a in B looks like. Let a be a path in B beginning at b0. Given t e [0, 1], let a,: /—♦ B denote that "portion" of a, between the points b0 and a,(t), defined by the equation a,(j) = aits) for s e [0, 1]. Then define a : /—► E by the equation «(/) = («,)*• We assert that a is a lifting of a; note that 3, begins at the point e0 = (ej*, the equivalence class of the constant path, and ends at the point a*. It is easy to see that p ° a = a, since />(«(')) =/>((<*,)*) = <*,0) = ^C''1)- The hard part is to show that & is continuous. Once we prove that fact, our theorem follows readily. To show Eis path connected, we note that if a* is any point of E, then a, is a path in B begin- beginning at b0, and its lifting g is a path in E beginning at e0 and ending at a ■ To show that PiGi,(E, e0)) => H, take an element [a] e H. Let 8 be the lifting of a to a path on E beginning at e0 \ then & ends at a*. Now a is equiv- equivalent to the constant path eb,, because [a. * ej = [a] e H. Therefore, a* = (ej#, so that £ is a loop in £ based at e0. Since />*([«]) = [a], it follo*s that [a] e />»(*,(£, £„))• Thus /f c ^(^(f, e0))- To prove the reverse inclusion, take an element [a] e w,^, «). Lei ~ p o 8. and note that a is the unique lifting of a to £ beginning at e0. There to a ends at a*; since a is a loop in E, we have dt(l) = e0- Therefore, a* °~ ^^ so a is equivalent to e6,, and hence [a] belongs to H. Thusp*(n,(E, e"^Cm. We complete the proof by showing that £ is continuous. Take a PaI? eter value f0 and a neighborhood B(U, a,,) in £ of the point 8(/^ (aj*. We shall find a number e > 0 such that whenever \t, — to\ < e>
§8-14 The Classification of Covering Spaces 397 have (a,,)* e B(U, a,,). This will prove continuity. Choose e small enough that 11, - to| < € => <x(t,) g U. We assert that if\t1-tB\<e, then a,, is path homotopic to a,. * d for some path 5 lying in U, whence a(?,) = (a,,)* lies in B(U, a,,), as desired. Assume for convenience in notation that tB < t,. Let 6 be that part of the path a lying between x(t0) and afrO, reparametrized appropriately. Figure 42 ■"Co) Figure 42 makes it clear that a,, is path homotopic to a,, * S. Formally, let S(s) = a(d - s)tB + stt); then 5 lies in */. Define F: / X /—♦ i? by the equation F(s rt = /"(['•O ~ 0 + *i*]2j/[1 + 'D for s e [°. 0 + 0/2], *' ~ f«('o(l - Bj - 1)) + *,Bj - 1)) for je[A+ i)\l, 1]. It is the required path homotopy. □ Corollary 14.5. Let B be path connected and locally path connected. (a) If B has a universal covering space, that covering space is uniquely determined up to equivalence. (b) IfB is semilocally simply connected, then B has a universal covering space. Proof. Part (a) is an immediate consequence of Theorem 14.3; \ip -E-* ■Bis a universal covering space of B, then the groupp*(n,(E, e0)) is the trivial subgroup of 7i,(fl, b0). To prove (b), we apply the preceding theorem to con- construct a path-connected covering space /?:£■—► B such that/?*(&,(£ e0)) is the trivial subgroup of n,(B, b0). Since pt is injective (a fact we leave to you to prove), this means that E is simply connected and is thus a universal cover- covering space of J3. Q
The Fundamental Croup and Covering Spaces „ Chap, g Exercises 1. We know that n,(S' x S',box b0) is isomorphic to Z x Z, the isomorphism being induced by the projections of £' x 51' onto its two factors. (a) Consider the infinite cyclic subgroup of Z x Z generated by the element 1 X 0; find the corresponding covering space of S1 x S1. (b) Find the covering space of £' x S< corresponding to the infinite cyclic subgroup generated by 1 X 1. (c) Find the covering space corresponding to the subgroup // = {Bn, 2m)\n, m e Z}. 2. Let B be a topological space. Show that if B has a universal covering space, then B is semilocally simply connected. 3. Let An be a circle of radius l/«in the A>>-pIane tangent to the >>-axis at the origin Let X = {JAn. (a) Show that X is not semilocally simply connected. (b) Let C(X) be the subspace of R1 which is the union of all line segments joining points of X to the point @, 0, 1). (It is called the cone on X.) Show that C(X) is semilocally simply connected at the origin but not locally simply connected there. 4. Let E and B be path connected and locally path connected. Lei p : E —>B be a covering map. A homeomorphism /; : E —► E is called a deck transformation (or covering transformation) if p ° h = p. (a) Let e0 and <?, be points ofp- 'F0). Show that there exists a deck transforma- transformation h : E—► E such that h(e0) = e, if and only if pt(Kt(E, e0)) - p*Gli(E, e,)). Show that if h exists, it is unique. (b) If //„ = p*Gi,(£, <?o)) '"s normal in x,(B, bo\ then p : E—> B is said to be a regular covering map. Show that in this case, the group of deck trans- transformations of E is isomorphic to the quotient group K,(B, ba)IHo- (Com- (Compare Exercise 10 of §8-4.) (c) Up:E—> B is a universal covering space of B, what can you say about the group of deck transformations of El 5. Consider the usual covering map p: R x R—>S'xi'. Describe the group of deck transformations of R x R. 6. Determine all deck transformations of the covering space of the figure eight constructed in Lemma 7.5. *7. Generalize Exercise 13 of §8-4, concerning covering spaces of topologica groups, to the case where the covering space is path connected but not simp connected.
Index Abelian fundamental group, 330, 354, 355 Absolute neighborhood retract, 221 Absolute retract vs. universal extension properly, 221 Accumulation point of a net, 188 Action of a group on a space, 356 Addition operation. 30 Adjunction space, 221 Alexander horned sphere. 385 Algebraic numbers, 51 a, 327 independence of path, 330 is isomorphism. 327 a", definition, 35, 55 ANR, 221 Antipodal map, 353 Antipodal point, 352 Antipode-preserving map, 361 AR, 221 Arc, 375 Archimedean ordering, 33 Arzela's theorem, 279, 292 Ascoli's theorem: classical version, 277 general version, 290 Axiom of choice, 59 equivalent statements, 62, 74 finite, 61 B Baire category theorem, 294 special case, 177, 203 Baire space, 293 compact Hausdorff space, 294 complete metric space, 294 fine topology on^lX Y), 297 G5 set, 296 irrationals, 296 locally compact Hausdorff space, 296 open subset, 296 RJ, 297 Ball, unit (see Bn) Barber of Seville paradox, 48 Base point, 326 Base-point choice: effect on fi», 330 effect on n,, 328 Basis for a topology, 78,81 Bd A, 101 . 401
402 Index P(X) (see Stone-Cech compactification) Bijective function, 19 Binary operation, 30 Bing metrization theorem, 254 B", 155 compactness, 175 path connectedness, 155 simple connectedness, 327 Bolzano-Weierstrass property. 178 Borsuk theorem. 377 Borsuk-Ulam theorem forS2, 361 Boundary, 101 Bounded above, 27 Bounded below, 27 Bounded metric, standard, 119 Bounded set, 119 Box topology, 113 basis for, 114 compactness properties, 181 Hausdorff condition, 115 vs. product topology, 114 subspace, 114 Brouwer fixed-point theorem: for B", 369 for B2, 365 Brouwer invariance of domain, 378, 387 B2,133 (see also B") fi(x, e), 117 Cantor set, 177 Cardinality: comparability for two sets, 68 greater, 62 same, 52 Cartesian product: general, 36~38 of two sets, 13 Cauchy sequence, 264 Choice axiom (see Axiom of choice) Choice function, 59 Circle, unit (seeS1) CIA, 95 Classification of covering spaces, 388,392,393 Closed graph, 172 Closed interval, 84 Closed map, 135,172 Closed ray, 86 Closed refinement, 251 Closed set, 92 in subspace, 94 Closure, 94 of a cartesian product, 100,116 of a connected set, 149 Closure (cont.) in a subspace, 95 of a union, 100, 246 via basis elements, 95 via limit points, 97 via nets, 188 via sequences, 128, 190 Cluster point, 96 Coarser topology, 77 Cofinal, 187 Coherent topology, 216 Collection, 11 Comb space, 156 Compact convergence topology, 282 vs. compact-open topology, 286 convergent sequences in, 282 furst-countability, 283 independence of metric, 287 vs. pointwise convergence topology 283,290 regularity, 283 vs. uniform topology, 283 (see also Compact-open topology) Compact Hausdorff space: isBaire, 294 components, 235 G6 set is Baire, 296 imbedding in, 237 metrizability, 220 normality, 198 paracompactness, 255 quasicomponents, 235 Compactification, 238 induced by an imbedding, 239 one-point, 183 of@, 1), 239 v (see also Stone-Cech compactification) Compactly generated space, 282 Compactness, 164 of box topology, 181 of closed intervals, 174 closed set criterion for, 170 vs. completeness, 275 of continuous image, 167 of countable products, 278 in<g(X.Rn), 277, 279 \n<i(X, Y), 290 of finite products, 167 in Hausdorff metric, 279 ,,4 vs. limit point compactness, 178, !•'• via nets, 188 in order topology, 173,177 of products, 232 in R, 174 in/?", 174 vs. second-countability, 194 via sequences, 181
Index 403 Compactness (cont.) ofsubspace, 165 in .Tp 167 in uniform topology, 181 (see also Compact Hausdorff space) Compact-open topology, 286, 290 vs. compact convergence topology, 286 continuity of evaluation map, 287 Hausdorff condition, 289 regularity, 289 Compact space, 164 (see also Compactness) Compact support, 284 Comparability: of cardinalities, 68 of well-ordered sets, 73 Complement, 10 Complete graph: on five vertices, 304, 386 on four vertices, 386 Completely normal space, 206 Completely regular space, 211, 236 (see also Complete regularity) Complete metric space, 264 Completeness: vs. Baire condition, 294 of closed subspace, 264 vs. compactness, 275 of rif(X, Y) in uniform metric, 267 ofG6 set, 270 of irrationals, 270 of82, 270 of open subspace, 270 of products, 270 of/?", 264 of/?", 265 of sup metric, 267 of uniform metric, 266 Complete regularity; of locally compact Hausdorff space. 237 vs. normality, 236 of products, 236 vs. regularity, 238 of/?2, 237 of R" in box topology, 237 of subspace, 236 of topological group, 237 Completion, 269, 270 uniqueness. 271 Component. 159 in a compact Hausdorff space, 235 vs. path component, 162 vs. quasicomponent, 163, 235 Composite, 18 of continuous functions, 107 of covering maps. 336 of quotient maps, 138 Composition of paths, 322 Conclusion, 7 Conjugacy class, 388 Conjugate subgroups, 387 Connected im kleinen, 163 Connectedness, 147 of box topology, 152 of closure, 149 of continuous image, 149 in a linear continuum, 152 vs. path connectedness, 156,158 of products, 150 \nR, 154 forsubspaces, 147 of !Tf, 151 (see also Path connectedness) Connected space, 147 (see also Connectedness) Constant path, 322 Continuity: of algebraic operations in R, 129, 134 via bases, 102 after change of range, 107 closed set formulation, 103 via closures, 103 of composites. 107 of constant function. 107 e-6 formulation, 127 of fxg. 112 of inclusion, 107 local formulation, 107 of maps from quotient spaces, 139 of .maps into products, 109, 115 of metric, 124 via nets, 188 of nth root function, 111 at a point, 107 of restriction, 107 via sequences, 128, 190 via subbases, 102 of sums, differences, products, and quotients, 129 of uniform limit, 130 in each variable separately, 112 Continuous function, 102 (see also Continuity) Continuous image: of a compact space, 167 of a connected space, 149 Continuum hypothesis, 62 Contractible space, 325, 373 simple connectedness, 361 Contraction, 182,270 Contrapositive, 8 Convergent net, 187 Convergent sequence, 116, 127 (.seealso Sequences) Converges uniformly. 129
404 Index Converse, 9 Convex set, 152, 325 Coordinate functions, 109 Coset, 145 Countability, 46 of algebraic numbers, 51 of countable unions, 49 of finite products, 49 ofrationals, 47 of subsets, 47 of Z, 45 ofZ, X Z,,45,47 Countable basis, 190 (see also Second-countability) Countable basis at a point, 128. 190 (see also First-countability) Countable compactness, 182, 194 Countable dense subset, 191 in3°(/, R), 195 in/7, 195 in t2, 195 in product space, 195 vs. second-countability, 191, 194 in subspace, 193 Countable intersection condition, 234 Countable set, 46 (seealso Counlability) Countably infinite, 45 Countably locally discrete, 254 Countably locally finite, 246 Counterimage, 17 Covering, 164, 165 Covering dimension, 302 Covering map, 331 composites, 336 induces injective liomomorphism, 342 vs. local homeomorphism, 334 products, 336 R ->-5l, 332, 393 R X R —*■ torus, 333 R X/?,->fl2-0,334 5'->5', 335,336,393 S2^-P2, 352 Covering space, 331 classification, 388, 392, 393 equivalence, 388 existence, 393 ft-fold,336 vs. it, of base space, 341, 342, 388 regular, 398 of topological group, 342, 398 Covering transformation, 398 Cube, 313 Curve, 223 simple closed, 375 ^sednessin Yx, 267, 281, 283 compact subsets of, 290 10 (cont.) completeness, 267 (see also Compact convergenc Compact-open topo.ogy^ 3, 119 Deck transformation, 398 Decomposition space, 137 Deformation retraction, 373 (seealso Strong deformation retraction) Degree of a map, 373 Deleted comb space, 156 components, 161 connectedness, 157 local connectedness, 162 De Morgan's laws, 11 Dense subset, 191 Diagonal map, 270 Diagonal set, 100,201 Diameter of a set, 119, 177 Dictionary order relation, 26 Difference of two sets, 10 Dimension, topological, 302 (seealso Topological dimension) Directed set, 187 Discrete topology, 77 Disjoint sets, 6 Distance. 117 Distributive laws for u and n, 11 Domain, 15, 16 Double torus, 355 Doubly punctured plane, 345 Doughnut surface, 334 Element of a set, 4 Empty interior, 293 Empty set. 6 c-ball, 117 6-neighborhoodofaset, 17/ Equality symbol, 4 Equicontinuity, 276 ofclosure.279,290 vs. compactness, 277, 2yu vs. total boundedness, 276 Equivalence class. 22 ,J Equivalence or compact.ficat.ons 23 Equivalence of covering spaces, » . Equivalence relation, 22 Essential map, 357 Euclidean metric, 120,126 Euclidean space, 37 Evaluation map, 287
Index 405 Even integer, 35 Evenly covered, 331 Eventually zero, 52 ex (constant path), 322 Expansion lemma, 259 Exponents, laws of, 35 1/1,319 Family of sets, indexed, 37 F. i, c, 230 Field, 31 Figure eight space, 345, 373 fundamental group, 354 Final point, of a path, 3 19 Finer topology, 77 criterion for, 80 Fine topology on <<?(X Y), 285 isBaire, 297 Finite axiom of choice, 61 Finite complement topology, 77 (see also .Tf) Finite-dimensional, 302 Finite intersection condition, 170 Finiteness, 40 of finite products, 44 of finite unions, 44 of subsets, 43 Finite set, 40 First category, 294 First coordinate of ordered pair, 1 3 First-countability, 128. 190 implies adequacy of sequences, 190 implies compactly generated, 282 of metric space, 128 of product, 191 of R,, 192 of Sn, 194 of subspace, 191 First-countable space, 128, 190 (see also First-countability) First homotopy group, 326 (see also Fundamental group) Fixed point, 158 Fixed-point free action. 357 Fixed-point theorem: for B", 369 for B2, 365 for a contraction, 182,270 for a retract of B2, 368 for S", 373 for[0, 1), 158 Fourteen-set problem, 101 Frechet compactness, 178 Frobenius theorem, 365, 369 Fo set, 250 Function, 16 Functorial properties of ht, 329 Fundamental group, 326 when abelian, 330 of double torus, 355 of figure eight, 354 and homotopy equivalence, 371 nonabelian, 354, 355 of orbit space, 357 of a product, 351 0fP2, 352 of/?"'-0,344 of R2~ 0,343 of y, 350 of Sl, 340 of S2, 347, 350 of strong deformation retract, 345 of torus, 352 Fundamental theorem of algebra, 362 Gas-water-electricity graph, 304 nonimbeddability, 386 Gg set, 194,248 is fiaire, 296 closed set is a, 248, 250 f'Hc) is a, 215,238 irrationals, 250 points of continuity form a, 296, 297 rationals, 295 and strong Urysohn lemma, 215 is topologically complete, 270 Generalized continuum hypothesis, 62 General linear group, 144 General position: in RN, 309 in/?3, 307 Geometrically independent, 308 G/N, 145 paracompactness, 260 regularity, 145 gib A, 27 Graph, 172 Greater cardinality, 62 Greatest lower bound, 27 Greatest lower bound property, 27 Groupoid properties, 322 H dependence on base point, 330 dependence on homotopy class of ft, 30* functotial properties, 329
406 Index Hahn-Mazurkiewicz theorem, 274 Half-open interval, 84 Has n elements, 40 Hausdorff condition, 98 for box topology, 115 and closedness of diagonal, 100, 201 for manifold, 225 for metric space, 126 for order topology, 99 for product, 99, 115, 197 for quotient space, 139,140 vs. regularity, 195, 200 for subspace, 99, 197 vs, T| axiom, 99, 100 for topological group, 145 and uniqueness of extensions, 112, 241 (see also Compact Hausdorff space) Hausdorff metric, 279 Hausdorff space, 98 (see also Hausdorff condition) Hilbertcube. 125 Homeomorphism, 104 vs. continuous bijective map, 106. 167 Homogeneous space, 144 Homomorpliism: induced by a map. 329 (see h *) induced by a path. 327 (see a) product, 351 zero, 330 Homotopic maps, 318 Homotopy. 319 effect on/it> 369 equals path in K(X, Y). 289 straight-line, 321 Homotopy equivalence, 371 induces isomorphism of it p 371 vs. strong deformation retraction, 372 Homotopy extension lemma. 377 Homotopy inverse. 371 Homotopy type, 371 (ftXo)»,329 (see also A*) Hypothesis, 7 Identification space, 137 Identity function, 20 "If... then," 7 / X / in dictionary order: closures in, 101 connectedness, 156 linear continuum, 154 local connectedness, 163 metrizability, 195 path connectedness, 156 //, countable dense subset, 195 Image, 16, 17 ■ Image set, 15, 16 Imbedding, 105 (see also Isometric imbedding) Imbedding theorem: for a compact manifold, 223,314 for a completely regular space, 237 for a linear graph, 308 for a manifold, 315 for an m-dimensional space, 310 315 into RJ, 220 for a 2-complex, 310 Immediate predecessor, 25 Immediate successor, 25 Indexed family of sets, 37 Indexing function, 37 Index set, 37 Indiscrete topology, 77 Induced homomorphism (seeh,) Induction, principle of, 32 transfinite, 67 Inductive definition (see Recursive definition) Inductive set: in R, 32 in a well-ordered set, 67 Inessential map, 357 induces Zero homomorphism, 358, 371 Infinite broom, 164 Infinite sequence, 37 Infinite series, 133 Infinite set, 45,57 Initial point of a path, 319 lnjective function, 19 Int A, 95 Integers, 32 Interior of a set, 94 Intermediate value theorem: of calculus. 146 general. 154 Intersection, 6, 12, 38 Intersects, 95 Interval, 25, 84 Intervals in /?: compactness, 174 connectedness, 154 topological dimension, 304 lnvariance of domnii , 378, 387 Inverse function, 19 Inverse image, 17 Irrationality of v/2, 36 Irrationals: Baire, 296 G6 set, 250 topologically complete, 2/u Isometric imbedding, 132 into a complete space, 268, ^" surjectivity, 182
Index 407 Isometry, 182 Isomorphism, 105 Jordan curve theorem, 383 Jordan separation theorem, 375 /-tuple, 38 K £-fold covering, 336 fc-plane, 309 Kuratowski 14-set problem, 101 Larger topology, 78 Largest element of an ordered set, 27 Laws of algebra, 33 Laws of exponents, 35 Laws of inequalities, 34 Least upper bound, 27 Least upper bound property, 27 vs. greatest lower bound property, 29 for/?, 31 for well-ordered sets, 67 Lebesgue dimension, 302 Lebesgue number, 179 Lebesgue number lemma, 179 Left inverse, 21 Lens space, 357 Lifting of a map, 336 Lifting lemma: general, 342, 390 for path homotopies, 338 for paths, 337 Limit point, 96 Limit point compactness: vs. compactness, 178, 181,194 vs. countable compactness, 182 of products, 182 Limit point compact space, 178 Lindelof condition, 191 for products, 193, 195,235 for/?,, 192 for/?2, 193 vs. second-countability, 191,194 for Sn and Sn,194 for subspaces, 194, 220 (see also Regular Lindelof space) Lindelof space, 191 Linear continuum, 31,152 compact sets in, 173 Linear continuum (cent.) connected sets in, 152 Linear graph, 304 imbedding in R3, 308 topological dimension, 305 Linear order, 24 Local compactness, 182 implies compactly generated, 282 of products, 186 of R" andRw, 183 (see also Locally compact Hausdorff space) Local connectedness, 161 vs. connectedness imkleinen, 163 vs. local path connectedness, 162 ofR"andRw, 161 Local homeomorphism, 334 Locally compact Hausdorff space: is Baire, 296 complete regularity, 237 imbedding in RN,315 one-point compactification, 183 regularity, 205 second-countability, 261 Locally discrete collection, 254 Locally finite collection, 245 Locally finite indexed family, 111,224 vs. locally finite collection, 246 Local metrizability, 220 vs. metrizability, 220, 260 Local path connectedness, 161 Local simple connectedness, 393, 398 Logical equivalence, 8 Logical quantifiers, 9 Long line, 159 paracompactness, 259 Loop,326 Lower bound, 27 Lower limit topology, 82 (see also /?;) B2, 126 completeness, 270 countable dense subset, 195 lub A, 27 M Manifold, 223 Hausdorff condition is needed, 225 imbeds in RN, 223 imbeds in R2m*\ 314, 315 metrizability, 224 topological dimension, 306, 314, 315 Mapping, 16 Maximum principle, 69 applied, 232, 235 intuitive proof, 70 proof, 74
408 Index Maximum value theorem: of calculus, 146 general, 175 Mean convergence topology, 284 Metric, 117 Metric space, 119 Hausdorff condition, 126 82, 126 normality, 198 paracompactness, 256 R, 118 R , uniform metric, 122 R", 121 subspace, 126 YJ, uniform metric, 266 Mej-izability: of / X / in dictionary order, 195 of manifolds, 224 of products, 126,132 of R, 118 otRJ, 131 of R,, 195 of R", 121 of /?<*», 123 of Rw in box topology, 130 of Sn, 179 of Sn, 131 Metrizable space, 119 Minimal uncountable well-ordered set, 66 (see also Sn) m.tuple, 36 Multiplication operation, 30 N Nagata-Smirnov condition, 245 Nagata-Smirnov metrization theorem: necessity, 253 sufficiency, 2.49 Negation, 9 Negative number, 31 Neighborhood, 96 Neighborhood retract, 221 Nested sequence of sets, 171 Net, 187 Normality, 195 of adjunction space, 221 of closed subspace, 205 of coherent topology, 216 of compact Hausdorff space, 198 vs. complete regularity, 236 of metric space, 198 of paracompact space, 255 of product, 197,201,202, 205 vs regularity, 195, 201,202, 205 of regular Lindeldf space, 205 Normality (com.) of Rh 202 of subspace, 197,201,205 of well-ordered set 200 Normal space, 195 (see also Normality) Nonseparation theorems, 378 382 Norm, 120 ' Nowhere-differentiable function, 297 nth root function: continuity, 111 existence, 158 Number of elements in a set, 40 uniqueness, 43 o Odd integer, 35 w-tuple, 37 One, 31 One-point compactifjcation, 183 One-to-one function, 19 "Onto" function, 19 Open covering, 164 Open interval, 25, 84 Open map, 91, 135 Open ray, 86 Open refinement, 251 Open set, 76 relative to subspace, 89 Operation, binary, 30 "Or," meaning of, 5 Orbit space, 357 fundamental group, 357 (see also G/H) Order of a collection, 302 Ordered field, 31 Ordered pair, 13 Order-preserving function, 25 Order relation, 24 Order topology, 85 compact sets in, 173,177 Hausdorff condition, 99 normality, 200 regularity, 205 subbasis, 91 vs. subspace topology, 90 (see also Well-ordered set) Order type, 25 0(A),U Pasting lemma, 108 Paracompact-manifold, 261
Index 409 Paracompactness, 255 of compact Hausdoiff space, 255 of long line, 259 vs. metrizability, 256,260 vs. normality, 255 and partitions of unity, 225 of product, 256, 259 of R, 255 of regular Lindelof space, 259 of 7?/, 256, 259 of Sn, 259, 261 of subspace, 256 Paracompact space, 225, 255 Partial order, 71 axioms, 72, 187 strict, 69 Partition of a set, 23 Partition of unity, 222, 225 Path, 155 Path component, 160 vs. component, 162 Path connectedness, 155 of B", 155 vs. connectedness, 156, 158 of deleted comb space, 157 of / X /in dictionary order, 156 of long line, 159 of R"- 0,155,350 of S", 156 of topologist'ssine curve, 158 Path homotopy, 319 Path-homotopy class, 320 Path-induced homomorphism, 327 Peano curve, 271 Peano space, 274 Piecewise linear function, 299 ffj(X, xQ), 326 (Seealso Fundamental group) Plane in RN, 309 Point of accumulation, 96 Point-finite collection, 247 Point-finite indexed family, 224 Point-open topology, 280 vs. compact convergence topology, 283,290 vs. compact-open topology, 286 convergent sequences in, 281 Pojntwise bounded collection, 278 Pointwise convergence topology, 280 (.see also Point-open topology) Positive integers, 32 Positive number, 31 Power set, 11 Preimage, 17 Principle of induction, 32 Principle of recursive definition, 49, 55 Principle of tiansfinite induction, 67 Principle of ttansfinite recursive definition, Products: .of continuous maps, 112 of covering maps, 336 of open maps, 139 of quotient maps, 138,141,187,289 Product space, 86,113 basis for, 87, 114 vs. box topology, 114 closures in, 100, 116 compactness, 167, 232, 278 complete regularity, 236 connectedness, 150 convergent sequences in, 116, 281 countable dense subset, 195 equals point-open topology, 280 first-countability, 191 fundamental group of, 351 Hausdorff condition, 99,115, 197 limit point compactness, 182 local compactness, 186 Lindelof condition, 193, 195, 235 metrizability, 131, 132 normality, 197, 201, 202, 205 paracompactness, 256, 259 vs. quotient topology, 138, 141,187, 289 regularity, 197 second-countability, 191 subbasis, 88, 113 vs. subspace topology, 90, 114 vs. uniform topology, on R , 122 Projection maps, 88,113 Projective n-space, 353 Projective plane, 352 (see also P*) Proper map, 285,315 Proper subset, 4 P2,352 fundamental group, 353 surface, 352 Punctured euclidean space, 155 (see also R"-0) Punctured plane, 321 (see also R* - 0) Q, 32 Quantifiers, logical, 9 Quasicomponent, 163 vs. component, 163,235 Quotient map, 135 composites, 138 products, 138,141,187, 289 Quotient operation in R, 31 Quotient space, 136
410 Index Quotient topology, 136 Hausdorff condition, 139, 140, 206 normality, 206 vs. product topology, 138, 141,187 289 subspace, 138 ' R R, 30 compact sets in, 174 connected sets in, 154 metric for, 118 paracompactness, 255 standard topology, 82 uncountability, 176 Rt,Rt,31 Range of a function, 16 Rational numbers, 32 Ray in ordered set, 86 Real numbers, 30 (see also R) Recursion formula, 48, 55 Reciprocal, 31 Recursive definition, 48 principle, 49, 55 transfinite, 68 Refinement of a collection, 225, 251 Regular covering map, 398 Regularity. 195 vs. complete regularity, 238 aiG/H, 145 vs. Hausdorff condition, 195,200 vs. metrizability, 217, 249 vs. normality, 195, 200, 201, 205 of products, 197 of subspaces, 197 of topological groups, 145 Regular Lindelof space: metrizability, 220 normality, 205 paracompactness, 259 Regular space, 195 {see also Regularity) Relation, 21 Restriction: of a function, 16 of a relation, 28 Retract, 2l6, 345 (see also Absolute retract and Retraction) Retraction, 330 gn + l _* y, 368,373 £2-»S',342 deformation, 373 r3 onto knotted x-axis, 221 /?2 onto logarithmic spiral, 221 strong deformation, 345 Reverse of a path, 320 p--\2°< 2" Sup metric) R~, 116, 125,274 R in box topology: is Baire, 297 is topological group, 144 RJ in product topology: is Baire, 297 metrizability, 131 normality, 205 is topological group, 144 R in uniform topology, 122 is Baire, 297 completeness, 266 is topological group 144 /?/, 82 countability axioms, 192 metrizability, 195 normality, 202 paracompactness, 256, 259 vs. standard topology on R, 82 Rf. 193 complete regularity, 237 Lindelof condition, 193 normality, 202 paracompactness, 256 R", 37 basis for, 115 compact sets in. 174 completeness. 264 local compactness, 183 local path connectedness, 161 metrics for, 121 second-countability, 190 simple connectedness, 326 Rn-0, 155 fundamental group, 344 path connectedness, 155, 350 /?", 37 Ru in box topology: complete regularity, 237 components, 163 connectedness, 152 metrizability, 130 normality, 206 paracompactness, 206 (see also RJ in box topology) Rw in product topology: completeness, 265 local compactness, 183 local path connectedness, 161 metrizability, 123 second-countability, 190 (see also RJ in product topology) Rw in uniform topology: components, 163
Index 41A Ru in uniform topology (com.) second-countability, 191 (see also RJ in uniform topology) R2, standard topology, 87 R2- 0,321 covering by R X /?,, 334 fundamental group, 343 Rule of assignment, 15 Russell's paradox, 62 Saturated set, 135 Schoenflies theorem, 385 Schroeder-Bernstein theorem, 52 Second category, 294 Second coordinate of ordered pair, 13 Second-countability, 190 of compact metric space. 194 of connected locally compact metric space, 261 vs. countable dense subset, 191, 194 vs. Lindeldf condition, 191, 194 of products, 191 of/?", 190 ofRu, 190 of subspaces, 191 Second-countable space, 190 (see also Second-countability) Section: of ordered set, 66 of Z,, 40 Semilocally simply connected, 393 Separable, 192 (see also Countable dense subset) Separation by continuous functions: of closed sets, 211 of points from closed sets, 220 Separation of a space, 147 Separation theorems: * by simple closed curve, 377, 386 S", 385 S2 by AuB, 377,386 S by simple closed curve, 375, 383 Sequence lemma, 128 Sequences, 37 vs. closure, 128, 190 vs. compactness, 179, 181 vs. continuity, 128, 190 in product spaces, 116,281 sums, differences, products and quotients, 132 Sequential compactness, 179 vs. compactness, 181, 194 Set, 3,4 ^rules for specifying, 58 'Set of all sets" paradox, 62 Shrinking lemma, 224 o-locally discrete, 254 o-locally finite, 247 Simple closed curves 375 generates tt, of R* - 0, 387 separates R2, 377, 386 separatesS2,375,383 Simple order, 24 Simplicial 2-complex, 306 (see also 2-complex) Simply connected space, 328 B", 327 contractible space, 361 R", 326 Rn- 0,350 S",350 S2, 347, 351 Slice: in covering space, 3 31 in product space, 168 Smaller topology, 78 Smallest element of ordered set, 27 Smirnov metrization theorem, 260 S", 156 compactness, 175 fixed-point theorem for, 374 path connectedness, 156 simple connectedness, 350 vector fields on, 369, 374 sfi. 66 compactness properties, 178 countability axioms, 194 metrizability, 179 paracompactness, 259, 261 Stone-fiech compactification, 243 uniqueness, 73 Sn,66 metrizability, 131 sn x in- incomplete regularity, 236 normality, 201 paracompactness, 256 S1, 106 covering by R, 332 coverings by Sl, 335, 336 covering spaces classified, 393 fundamental group, 340 Sorgenfrey plane, 193 (seealso Rt) Space-filling curve, 271 Special Van Kampen theorem, 348 Sphere, unit, 138, 156 (see also S2; S") Square metric, 120, 267 (see also Sup metric) Square root of 2, irrationality, 36 Square roots, existence, 35 Star convex set, 330 Stereographic projection, 350, 367
412 Index Stone-fech compactification, 241 connectedness, 243 metrizability, 243 of Sn, 243 uniqueness, 242 of Z,, 243 Stone's theorem, 256 Straight-line homotopy, 321 Strictly coarser topology, 77 Strictly finer topology, 77 Strict partial order, 69 Strong deformation retract, 345 fundamental group, 345 Strong deformation retraction, 345 vs. homotopy equivalence, 372 of /?"-(> onto yl, 345 of/?2 -p-q onto 6 and eight, 345, 346 Stronger topology, 78 S2, 138 simple connectedness, 347, 351 vector fields on, 367 Subbasis for a topology, 82 Subnet, 188 Subsequence, 179 Subset, 4 Subspace, 89 basis for, 89 vs. box topology, 114 compactness, 165 complete regularity, 236 connectedness, 147 countable dense subset, 193 first-countability, 191 Hausdorff condition, 99, 197 Lindelbf condition, 194, 220 of metric space, 126 normality, 197, 201.205 vs. order topology, 90 paracompactness, 256 vs. product topology, 90, 114 vs. quotient topology, 138 regularity, 197 seeond-countability, 191 Subtraction operation, 31 Superset, 232 Sup metric, 267 completeness, 267, 268 vs. uniform metric, 267 Support, 222 Surface, 223 Surjective function, 19 ^,77 compactness, 167 connectedness, 151 ,Tf (com.) Tj axiom, 100 Theta space, 346, 373 separates S2, 386 Tietze extension theorem 21? Tj axiom, 99 Topological dimension, 302 of [a, b], 304 of a linear graph, 305 of a manifold, 314, 315 in a metric space, 304 of a set in/?W, 313,315 of a set in R2> 305 of a subspace, 302 of a triangular region, 306, 359 of a 2-complex, 306 of a 2-manifold, 306 of a union, 304, 315 Topological group, 144 closedness of A • B, 173, 188 complete regularity, 237 covering spaces of, 342, 398 Hausdorff condition, 145 normality, 207 paracompactness, 260 ff| is abelian, 331 regularity, 145 Topological imbedding, 105 Topologically complete space, 270 {seealso Complete metric space) Topological property, 104 Topological space, 76 Topologist's sine curve, 158 Topology, 76 generated by a basis, 78, 80 generated by a subbasis, 82 Torus, 134, 333 covering by R X R, 333 double, 355 equals doughnut surface, 334 fundamental group, 352 vector fields on, 366 Totally bounded, 275 Totally disconnected, 152 Tower, 74 Transcendental number, 52 Transfinite induction, 67 Transfinite recursive definition, 68 Translation of RN, 309 Triangle inequality, 117 Trivial group, 328 Trivial topology, 77 Tube, 168 Tube lemma, 169 generalized, 172 2-complex, 306 imbedding in Rs, 310 topological dimension, 306
2-sphere, 138 (see also S2) Tychonoff theorem, 232 countable, 278 finite, 167 u U(A, e),177 Uncountabtlity: of/?, 176 of 0\Z,)> 50 of transcendental numbers, 52 of {0,1}W, 50 Uncountable set, 46 Uncountable well-ordered set, 66 existence, 73 (see also Sq) Uniform boundedness principle, 297 Uniform continuity theorem: of calculus, 146 general, 180 Uniform convergence, 129 Weierstrass Af-test, 134 Uniform convergence on compact sets, 282 (see also Compact convergence topology) Uniform limit theorem, 130 converse fails, 133 partial converse, 172 Uniformly bounded, 278 Uniform metric, 122, 266 completeness, 266 vs. sup metric, 267 (see also Uniform topology) Uniform structure, 292 Uniform topology,-122, 266 vs. compact convergence topology, 283 compactness properties, 181, 277, 279 independence of metric, 289 vs. pointwise convergence topology, 283 vs. product topology, 122 (see also Uniform metric) Union, 5,12, 38 Universal covering group, 343 Universal covering space, 341 existence, 397 uniqueness, 342, 397 Universal extension property, 216 vs. absolute retract, 221 Universal neighborhood extension property, 221 Upper bound, 27 Utysohn lemma, 207 for completely regular spaces, 237 strong form, 215 Urysohn tnebization theorem, 217 /ndearl Vacuously true, 7 Value of a function, 16 Vanishing at infinity, 279 Van Kampen theorem, 349 special, 348 Vector field, 364 on B", 369 On B2, 364 on Klein bottle, 366 on S", 369,374 on S2, 367 on torus, 366 Vertices of a linear graph, 304 W Weaker topology, 78 Weierstrass Af-test, 134, 214 Well-ordered set, 63 AX B in dictionary order, 64 compactness in, 173 countable, 65 finite, 64 normality, 200 subsets, 64 uncountable, 66, 73 Z«,32 Zt X Z», 63 Well-ordering theorem, 65 applied, 224, 251 vs. axiom of choice, 74 vs. maximum principle, 74 Winding number, 347 as an integral, 347 of simple closed curve, 387 A, 36 A^.37 [X, H.325 Z, 32 Zermelo, 65 Zero, 30 Zero homomotphism, 330 Z.,32