0 ratings0% found this document useful (0 votes) 2K views384 pages(Simmons) Introduction To Topology and Modern Analysis
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content,
claim it here.
Available Formats
Download as PDF or read online on Scribd
Introduction to
TOPOLOGY AND
MODERN ANALYSIS
GEORGE F. SIMMONS
Associate Professor of Mathematics
Colorado College
©
ROBERT E. KRIEGER PUBLISHING COMPANY
MALABAR, FLORIDAFor Virgie May Hatcher
and Elizabeth B. Blossom
TO EACH OF WHOM
| OWE MORE
THAN | CAN POSSIBLY EXPRESS
Original Edition 1963
Reprint Edition 1983
Printed and Published by
ROBERT E. KRIEGER PUBLISHING COMPANY, INC.
KRIEGER DRIVE
MALABAR, FLORIDA 32950
Copyright © 1963 by McGraw-Hill, Inc.
Reprinted by arrangement
All rights reserved. No part of this book may be reproduced in any
Sorm or by any electronic or mechanical means including information
storage and retrieval systems without permission in writing from the
publisher.
Printed in the United States of America
Library of Congress Cataloging in Publication Data
Simmons, George Finlay, 1925-
Introduction to topology and modern analysis.
Reprint. Originally published: New York: McGraw-
Hill, 1963 (International series in pure and applied
mathematics)
Bibliography: p.
Includes index.
1, Topology. , 2. Mathematical analysis.
1. Title. IL. Title: Topology and modern analysis.
IIL. Series: International series in pure and
QAGIL.S49 1983 34 82-1484,
ISBN 0-89874—551-9
098765Preface
For some time now, topology has been firmly established as one of
the basic disciplines of pure mathematics. Its ideas and methods have
transformed large parts of geometry and analysis almost beyond recogni-
tion. It has also greatly stimulated the growth of abstract algebra. As
things stand today, much of modern pure mathematics must remain a
closed book to the person who does not acquire a working knowledge of at
least the elements of topology.
There are many domains in the broad field of topology, of which the
following are only a few: the homology and cohomology theory of com-
plexes, and of more general spaces as well; dimension theory; the theory
of differentiable and Riemannian manifolds and of Lie groups; the theory
of continuous curves; the theory of Banach and Hilbert spaces and their
operators, and of Banach algebras; and abstract harmonic analysis on
locally compact groups. Each of these subjects starts from roughly the
same body of fundamental knowledge and develops its own methods of
dealing with its own characteristic problems. The purpose of Part 1 of
this book is to make available to the student this “hard core” of funda-
mental topology; specifically, to make it available in a form which is
general enough to meet the needs of modern mathematics, and yet is
unburdened by excess baggage best left in the research journals.
A topological space can be thought of as a set from which has been
swept away all structure irrelevant to the continuity of functions defined
on it. Part 1 therefore begins with an informal (but quite extensive)
treatment of sets and functions. Some writers deal with the theory of
metric spaces as if it were merely a fragment of the general theory of
topological spaces. This practice is no doubt logically correct, but it
seems to me to violate the natural relation between these topics, in which
metric spaces motivate the more general theory. Metric spaces are
therefore discussed rather fully in Chapter 2, and topological spaces are
introduced in Chapter 3. The remaining four chapters in Part 1 are
concerned with various kinds of topological spaces of special importance
in applications and with the continuous functions carried by them.
It goes without saying that one aspect of this type of mathematics
is its logical precision. Too many writers, however, are content with
this, and make little effort to help the reader maintain his orientation in
vilviii Preface
the midst of mazes of detail. One of the main features of this book is the
attention given to motivating the ideas under discussion. On every
possible occasion I have tried to make clear the intuitive meaning of what
is taking place, and diagrams are provided, whenever it seems feasible,
to help the reader develop skill in using his imagination to visualize
abstract ideas. Also, each chapter begins with a brief introduction
which describes its main theme in general terms. Courses in topology
are being taught more and more widely on the undergraduate level in
our colleges and universities, and I hope that these features, which tend
to soften the austere framework of definitions, theorems, and proofs,
will make this book readable and easy to use as a text.
Historically speaking, topology has followed two principal lines of
development. In homology theory, dimension theory, and the study of
manifolds, the basic motivation appears to have come from geometry.
In these fields, topological spaces are looked upon as generalized geometric
configurations, and the emphasis is placed on the structure of the spaces
themselves. In the other direction, the main stimulus has been analysis.
Continuous functions are the chief objects of interest here, and topological
spaces are regarded primarily as carriers of such functions and as domains
over which they can be integrated. These ideas lead naturally mto the
theory of Banach and Hilbert spaces and Banach algebras, the modern
theory of integration, and abstract harmonic analysis on locally compact
groups.
In Part 1 of this book, I have attempted an even balance between
these two points of view. This part is suitable for a basic semester course,
and most of the topics treated are indispensable for further study in
almost any direction. If the instructor wishes to devote a second
semester to some of the extensions and applications of the theory, many
possibilities are open. If he prefers applications in modern analysis, he
can continue with Part 2 of this book, supplemented, perhaps, with a
brief treatment of measure and integration aimed at the general form of
the Riesz representation theorem. Or if his tastes incline him toward
the geometric aspects of topology, he can switch over to one of the many
excellent books which deal with these matters.
The instructor who intends to continue with Part 2 must face a
question which only he can answer. Do his students know enough about
algebra? This question is forced to the surface by the fact that Chapters
9 to 11 are as much about algebra as they are about topology and analy-
sis. If his students know little or nothing about modern algebra, then a
careful and detailed treatment of Chapter 8 should make it possible to
proceed without difficulty. And if they know a good deal, then a quick
survey of Chapter 8 should suffice. It is my own opinion that education
in abstract mathematics ought to begin on the junior level with a course
in modern algebra, and that topology should be offered only to studentsPreface ix
who have acquired some familiarity, through such a course, with abstract
methods.
Part 3 is intended for individual study by exceptionally well-qualified
students with a reasonable knowledge of complex analysis. Its principal
purpose is to unify Parts 1 and 2 into a single body of thought, along the
lines mapped out in the last section of Chapter 11.
Taken as a whole, the present work stands at the threshold of the
more advanced books by Rickart [34], Loomis (27], and Naimark [32];
and much of its subject matter can be found (in one form or another and
with innumerable applications to analysis) in the encyclopedic treatises
of Dunford and Schwartz (8] and Hille and Phillips [20].!_ This book is
intended to be elementary, in the sense of being accessible to well-trained
undergraduates, while those just mentioned are not, Its prerequisites
are almost negligible. Several facts about determinants are used without
proof in Chapter 11, and Chapter 12 leans heavily on Liouville’s theorem
and the Laurent expansion from complex analysis. With these excep-
tions, the book is essentially self-contained.
It seems to me that a worthwhile distinction can be drawn between
two types of pure mathematics. The first—which unfortunately is
somewhat out of style at present—centers attention on particular func-
tions and theorems which are rich in meaning and history, like the gamma
function and the prime number theorem, or on juicy individual facts, like
Euler’s wonderful formula
14M+wt--- = 9/6.
The second is concerned primarily with form and structure. The present.
book belongs to this camp; for its dominant theme can be expressed in
just two words, continuity and linearity, and its purpose is to illuminate
the meanings of these words and their relations to each other. Mathe-
matics of this kind hardly ever yields great and memorable results like
the prime number theorem and Euler’s formula, On the contrary, its
theorems are generally small parts of a much larger whole and derive
their main significance from the place they occupy in that whole. In
my opinion, if a body of mathematics like this is to justify itself, it must
possess aesthetic qualities akin to those of a good piece of architecture.
It should have a solid foundation, its walls and beams should be firmly
and truly placed, each part should bear a meaningful relation to every
other part, and its towers and pinnacles should exalt the mind. It is my
hope that this book can contribute to a wider appreciation of these mathe-
matical values,
George F, Simmons
1 The numbers in brackets refer to works listed in the Bibliography.A Note to the Reader
‘Two matters call for special comment: the problems and the proofs.
The majority of the problems are corollaries and extensions of
theorems proved in the text, and are freely drawn upon at all later stages
of the book. In general, they serve as a bridge between ideas just treated
and developments yet to come, and the reader is strongly urged to master
them as he goes along.
In the earlier chapters, proofs are given in considerable detail, in an
effort to smooth the way for the beginner. As our subject unfolds
through the successive chapters and the reader acquires experience in
following abstract mathematical arguments, the proofs become briefer
and minor details are more and more left for the reader to fill in for
himself. The serious student. will train himself to look for gaps in proofs,
and should regard them as tacit invitations to do a little thinking on his
own. Phrases like “it is easy to see,” “‘one can easily show,” “evidently,”
“clearly,” and so on, are always to be taken as warning signals which
indicate the presence of gaps, and they should put the reader on his
guard,
It is @ basic principle in the study of mathematics, and one too
seldom emphasized, that a proof is not really understood until the
stage is reached at which one can grasp it as a whole and see it as a single
idea, In achieving this end, much more is necessary than merely follow-
ing the individual steps in the reasoning. This is only the beginning.
A proof should be chewed, swallowed, and digested, and this process of
assimilation should not be abandoned until it yields a full comprehension
of the overall pattern of thought.Contents
Preface vii
A Note to the Reader xi
PART ONE: TOPOLOGY
Chapter One SETS AND FUNCTIONS 3
Sets and set inclusion 3
The algebra of sets 7
Functions 14
Products of sets 21
Partitions and equivalence relations 25
Countable sets 31
. Uncountable sets 36
Partially ordered sets and lattices 43
exes eee
Chapter Two METRIC SPACES 49
9. The definition and some examples 51
10. Open sets 59
11, Closed sets 65
12, Convergence, completeness, and Baire’s theorem 70
13. Continuous mappings 75
14, Spaces of continuous functions 80
15, Euclidean and unitary spaces 85
Chapter Three TOPOLOGICAL SPACES 91
16. The definition and some examples 92
17. Elementary concepts 95
18. Open bases and open subbases 99
19. Weak topologies 104
20, The function algebras €(X,R) and e(X,C) 106
Chapter Four COMPACTNESS 110
21, Compact spaces 111
22, Products of spaces 115
xlxiv Contents
23, Tychonofi’s theorem and locally compact spaces 118
24, Compactness for metric spaces 120
25. Ascoli’s theorem 124
Chapter Five SEPARATION
26. Tr-spaces and Hausdorff spaces 130
27, Completely regular spaces and normal spaces 132
28. Urysohn’s lemma and the Tietze extension theorem 135
29. The Urysohn imbedding theorem 137
30. The Stone-Cech compactification 139
Chapter Sit CONNECTEDNESS
31. Connected spaces 143
32. The components of a space 146
33, Totally disconnected spaces 149
34, Locally connected spaces 150
Chapter Seven APPROXIMATION
35. The Weierstrass approximation theorem 153
36. The Stone-Weierstrass theorems 157
37. Locally compact Hausdorff spaces 162
38. The extended Stone-Weierstrass theorems 165
PART TWO:
Chapter Hight ALGEBRAIC SYSTEMS
39, Groups 172
40. Rings 181
41. The structure of rings 184
42, Linear spaces 191
43. The dimension of a linear space 196
44, Linear transformations 203
45, Algebras 208
Chapter Nine BANACH SPACES
46, The definition and some examples 212
47, Continuous linear transformations 219
48, The Hahn-Banach theorem 224
49, The natural imbedding of Nin N** 231
50. The open mapping theorem 235
51. The conjugate of an operator 239
Chapter Ten HILBERT SPACES
52. The definition and some simple properties 244
53. Orthogonal complements 249
54, Orthonormal sets 251
129
142
153
OPERATORS
171
211
243Contents xv
55. The conjugate space H* 260
56. The adjoint of an operator 262
57. Self-adjoint operators 266
58, Normal and unitary operators 269
59. Projections 273
Chapter Eleven FINITE-DIMENSIONAL SPECTRAL THEORY 278
60. Matrices 280
61. Determinants and the spectrum of an operator 287
62. The spectral theorem 290
63. A survey of the situation 295
PART THREE: ALGEBRAS OF OPERATORS
Chapter Twelve GENERAL PRELIMINARIES ON BANACH
ALGEBRAS 301
64. The definition and some examples 302
65. Regular and singular elements 305
66. Topological divisors of zero 307
67. The spectrum 308
68, The formula for the spectral radiue 312
69. The radical and semi-simplicity 313
Chapter Thirteen THE STRUCTURE OF COMMUTATIVE
BANACH ALGEBRAS 318
70. The Gelfand mapping 318
71. Applications of the formula r(z) = lim lz" 323
72, Involutions in Banach algebras 324
73. The Gelfand-Neumark theorem 325
Chapter Fourteen SOME SPECIAL COMMUTATIVE BANACH
ALGEBRAS 327
74, Ideals in €(X) and the Banach-Stone theorem 327
75. The Stone-Cech compactification (continued) 330
76. Commutative C*-algebras 332
APPENDICES
ONE Fixed point theorems and some applications to analysis 337
TWO Continuous curves and the Hahn-Mazurkiewics theorem 341
THREE Boolean algebras, Boolean rings, and Stone’s theorem 344
Bibliography 355
Index of Symbols 359
Subject Index 363PART ONE
CopologyCHAPTER ONE
Sets and Functions
It is sometimes said that mathematics is the study of sets and func-
tions. Naturally, this oversimplifies matters; but it does come as close
to the truth as an aphorism can.
The study of sets and functions leads two ways. One path goes
down, into the abysses of logic, philosophy, and the foundations of
mathematics. The other goes up, onto the highlands of mathematics
itself, where these concepts are indispensable in almost all of pure mathe-
matics as it is today. Needless to say, we follow the latter course. We
regard sets and functions as tools of thought, and our purpose in this
chapter is to develop these tools to the point where they are sufficiently
powerful to serve our needs through the rest of this book.
As the reader proceeds, he will come to understand that the words
set and function are not as simple as they may seem. In a sense, they
are simple; but they are potent words, and the quality of simplicity they
possess is that which lies on the far side of complexity. They are like
seeds, which are primitive in appearance but have the capacity for vast
and intricate development.
1, SETS AND SET INCLUSION
We adopt a naive point of view in our discussion of sets and assume
that the concepts of an element and of a set of elements are intuitively
clear. By an element we mean an object or entity of some sort, as, for
example, a positive integer, a point on the real line (= a real number),
34 Topology
or a point in the complex plane (= a complex number). A set is a collec-
tion or aggregate of such elements, considered together or as a whole.
Some examples are furnished by the set of all even positive integers, the
set of all rational points on the real line, and the set; of all points in the
complex plane whose distance from the origin is 1 (= the unit circle in the
plane). We reserve the word class to refer to a set of sets. We might
speak, for instance, of the class of all circles in a plane (thinking of each
circle as a set of points). It will be useful in the work we do if we carry
this hierarchy one step further and use the term family for a set of classes.
One more remark: the words element, set, class, and family are not
intended to be rigidly fixed in their usage; we use them fluidly, to express
varying attitudes toward the mathematical objects and systems we
study. It is entirely reasonable, for instance, to think of a circle not as
a set of points, but as a single entity in itself, in which case we might
justifiably speak of the set of all circles in a plane.
There are two standard notations available for designating a par-
ticular set. Whenever it is feasible to do so, we can list its elements
between braces. Thus {1, 2, 3} signifies the set consisting of the first
three positive integers, {1, 1, —1, —1} is the set of the four fourth roots
of unity, and {+1, +3, +5, . . .} is the set of all odd integers. This
manner of specifying a set, by listing its elements, is unworkable in many
circumstances. We are then obliged to fall back on the second method,
which is to use a property or attribute that characterizes the elements of
the set in question. If P denotes a certain property of elements, then
{z:P} stands for the set of all elements x for which the property P is
meaningful and true. For example, the expression
{x:a is real and irrational},
which we read the set of all x such that x is real and irrational, denotes the
set of all real numbers which cannot be written as the quotient of two
integers. The set under discussion contains all those elements (and no
others) which possess the stated property. The three sets of numbers
described at the beginning of this paragraph can be written either way:
{1, 2, 3} = {n:n is an integer and 0 < n < 4},
{1, 4, -1, —7} is a complex number and z‘ = 1},
and {+1, £3, +5, ...} = {n:nis an odd integer}.
We often shorten our notation. For instance, the last. two sets mentioned
might perfectly well be written {z:z4 = 1} and {n:nis odd}. Our pur-
pose is to be clear and to avoid misunderstandings, and if this can be
achieved with less notation, so much the better. In the same vein we canSets and Functions 5
write
the unit circle = {z:|z| = 1},
the closed unit dise = {z:|z| < 1},
and the open unit disc = {z:|z| < 1}.
We use a special system of notation for designating intervals of various
kinds on the real line. If @ and 6} are real numbers such that a < 6,
then the following symbols on the left are defined to be the indicated sets
on the right:
[a,b] = {za Sx <5},
(a,b] = {z:a A and BD A.
It will often be convenient to have a symbol for logical implication,
and = is the symbol we use. If p and q are statements, then p> q@
means that p implies q, or that if p is true, then q is also true. Similarly,
« is our symbol for two-way implication or logical equivalence. It
means that the statement on each side implies the statement on the other,
and is usually read #f and only if, or is equivalent to.
The main properties of set inclusion are obvious. They are the
following:
(1) ACA for every A;
(2) ACBand BCAA =B;
(3) ACBandBOCSACC.
It is quite important to observe that (1) and (2) can be combined into the
single statement that A = B@ACB and BCA. This remark
contains a useful principle of proof, namely, that the only way to show
that two sets are equal, apart from merely inspecting them, is to show
that each is a subset of the other.
Problems
1. Perhaps the most famous of the logical difficulties referred to in the
text is Russell’s parador. To explain what this is, we begin by
observing that a set can easily have elements which are themselves
sets, e.g., {1, {2,3}, 4}. This raises the possibility that a set might
well contain itself as one of its elements. We call such a set an
abnormal set, and any set which does not contain itself as an element
we call a normal set. Most sets are normal, and if we suspect that.
abnormal sets are in some way undesirable, we might try to confine
our attention to the set NV of all normal sets. Someone is now sure
to ask, Is N itself normal or abnormal? It is evidently one or the
other, and it cannot be both. Show that if N is normal, then it
must be abnormal. Show also that if N is abnormal, then it must
be normal. We see in this way that each of our two alternatives is
self-contradictory, and it seems to be the assumption that N exists
as a set which has brought us to this impasse. For further discussion
of these matters, we refer the interested reader to Wilder [42, p. 55]Sets and Functions = 7
or Fraenkel and Bar-Hillel [10, p. 6]. Russell’s own account of the
discovery of his paradox can be found in Russell [36, p. 75].
The symbol we have used for set inclusion is similar to that used for
the familiar order relation on the real line: if z and y are real numbers,
2 2 <2.
It also has an important additional property:
(4) for any z andy, either z < yor y < z.
Property (4’) says that any two real numbers are comparable with
respect to the relation in question, and it leads us to call the order
relation on the real line a total (or linear) order relation. Show by an
example that this property is not possessed by set inclusion. It is
for this reason that set inclusion is called a partial order relation.
(a) Let U be the single-element set {1}. There are two subsets,
the empty set § and {1} itself. If A and B are arbitrary subsets
of U, there are four possible relations of the form A € B.
Count the number of true relations among these.
(b) Let U be the set {1,2}. There are four subsets. List them.
If A and B are arbitrary subsets of U, there are 16 possible
relations of the form A C B. Count the number of true ones.
(c) Let U be the set {1, 2, 3}. There are 8 subsets. What are
they? There are 64 possible relations of the form A C B.
Count the number of true ones.
(d) Let U be the set {1, 2,..., m} for an arbitrary positive
integer n. How many subsets are there? How many possible
relations of the form A C B are there? Can you make an
informed guess as to how many of these are true?
2. THE ALGEBRA OF SETS
In this section we consider several useful ways in which sets can be
combined with one another, and we develop the chief properties of these
operations of combination.
As we emphasized above, all the sets we mention in this section are
assumed to be subsets of our universal set U. U is the frame of reference,
or the universe, for our present discussions. In our later work the frame
of reference in a particular context will naturally depend on what ideas
we happen to be considering. If we find ourselves studying sets of real8 Topology
numbers, then U is the set R of all real numbers. If we wish to study
sets of complex numbers, then we take U to be the set C of all complex
numbers. We sometimes want to narrow the frame of reference and to
consider (for instance) only subsets of the closed unit interval [0,1], or
of the closed unit dise {z:|z| < 1}, and in these cases we choose U
accordingly. Generally speaking, the universal set U is at our disposal,
and we are free to select it to fit the needs of the moment. For the
présent, however, U is to be regarded as a fixed but arbitrary set. This
generality allows us to apply the ideas we develop below to any situation
which arises in our later work.
Fig. 1. Set inclusion. Fig. 2. The union of A and B.
It is extremely helpful to the imagination to have a geometric
picture available in terms of which we can visualize sets and operations
on sets. A convenient way to accomplish this is to represent U by a
rectangular area in a plane, and the elements which make up U by the
points of this area. Sets can then be pictured by areas within this
rectangle, and diagrams can be drawn which illustrate operations on sets
and relations between them. For instance, if A and B are sets, then Fig.
1 represents the circumstance that A is a subset of B (we think of each
set as consisting of all points within the corresponding closed curve).
Diagrammatic thought of this kind is admittedly loose and imprecise;
nevertheless, the reader will find it invaluable. Nomathematics, however
abstract it may appear, is ever carried on without the help of mental
images of some kind, and these are often nebulous, personal, and diffi-
cult to describe.
The first operation we discuss in the algebra of sets is that of forming
unions. The union of two sets A and B, written A U B, is defined
to be the set of all elements which are in either A or B (including
those which are in both). A U B is formed by lumping together the
elements of A and those of B and regarding them as constituting a single
set. In Fig. 2, 4 UB is indicated by the shaded area. The aboveSets and Functions 9
definition can also be expressed symbolically:
AVB = {z:z¢A or zé B}.
The operation of forming unions is commutative and associative:
AUB=BUA and AU(BUC)=(AUB)UC.
It has the following additional properties:
AVA=A,AUG=A,andAUU=U,
We also note that
AGCBoAUB=B,
so set inclusion can be expressed in terms of this operation.
Our next operation is that of forming intersections. The inter-
section of two sets A and B, written A ( B, is the set of all elements which
are in both A and B. In symbols,
AB = {a:a¢ A and xe B}.
A(\B is the common part of the
sets A and B. In Fig. 3, AM Bis
represented by the shaded area. If
AQ Bis non-empty, we express this
by saying that A intersects B. If,
on the other hand, it happens that A
and B have no common part, or
equivalently that A B = G, then Fig. 3. The intersection of A and B.
we say that A does not intersect B, or
that A and B are disjoint; and a class of sets in which all pairs of distinct
sets are disjoint is called a disjoint class of sets. The operation of form-
ing intersections is also commutative and associative:
ANB=BONA and AN(BNC)=(ANB)NC.
It has the further properties that
ANA=A,ANG6=§$,and ANU = A;
and since
ACBeANB=A,
we see that set inclusion can also be expressed in terms of forming
intersections.
We have now defined two of the fundamental operations on sets, and
we have seen how each is related to set inclusion. The next obvious step
is to see how they are related to one another. The facts here are given by10 Topology
the distributive laws:
AN (BUC) = (ANB)U(ANO)
and AU(BNC) = (AUB) N(AUC).
‘These properties depend only on simple logic applied to the meanings of
ttf
ily
/
Fig. 4. A U(BNC) = (A UB) N(A UC).
the symbols involved. For instance, the first of the two distributive
laws says that an element is in A and is in B or C precisely when it is in
A and Bor isin A and C. We can convince ourselves intuitively of the
validity of these laws by drawing pictures. The second distributive
law is illustrated in Fig. 4, where A U (B 1 C) is formed on the left by
shading and (A U B)(\ (A UC) on the right by cross-shading. A
moment’s Stare ai baal a
< grams ought to convi th di
that one obtains the same set in each
case.
Fig. 5. The complement of A. those which make up JU, it goes with-
out saying—but it ought to be said
The last of our major operations
on sets is the formation of comple-
—that A’ consists of all those elements in U which are not in A.
Symbolically,
ments. The complement of a set A,
A’ = {xing A}.
Figure 5 (in which A’ is shaded) illustrates this operation. The operation
of forming complements has the following obvious properties:
(A'’ = 4, 8 = U, U' =
AUVA'=U,and AN A' =.Sets and Functions 11
Further, it is related to set inclusion by
ACBeBCA’
and to the formation of unions and intersections by
(AUB) =A’ OB and (AM BY = A’UB’. (1)
The first equation of (1) says that an element is not in either of two sets
precisely when it is outside of both, and the second says that it is not in
both precisely when it is outside of one or the other.
The operations of forming unions and intersections are primarily
binary operations; that is, each is a process which applies to a pair of sets
and yields a third. We have emphasized this by our use of parentheses
to indicate the order in which the operations are to be performed, as in
(A, U As) U As, where the parentheses direct us first to unite A; and
A, and then to unite the result of this with A3. Associativity makes it
possible to dispense with parentheses in an expression like this and to
write Ai\U A; U A;, where we understand that these sets are to be
united in any order and that the order in which the operations are
performed is irrelevant. Similar remarks apply to Ai A: As.
Furthermore, if {Ai, 42, ..., An} is any finite class of sets, then we
can form
AL\UA21U+++UAn and Ai NAI1N +++ OAs
in much the same way without any ambiguity of meaning whatever.
In order to shorten the notation, we let J = {1, 2,... , n} be the set
of subscripts which index the sets under consideration. J is called the
index set. ‘We then compress the symbols for the union and intersection
just mentioned to Usar Ai and Mr Ai. As long as it is quite clear what
the index set is, we can write this union and intersection even more
briefly, in the form U;A; and (;A,. For the sake of both brevity and
clarity, these sets are often written UZ, A; and Mf, Ai
These extensions of our ideas and notations don’t reach nearly far
enough. It is often necessary to form unions and intersections of large
(really large!) classes of sets. Let {A;} be an entirely arbitrary class
of sets indexed by a set J of subscripts. Then
User Ay = {u:2 ¢ A; for at least one ¢ ¢ 7}
and Over Ag = {ata € A; for every i€ I}
define their union and intersection. As above, we usually abbreviate these
notations to U:A; and (,A,; and if the class {A,} consists of a sequence
of sets, that is, if {As} = (Ai, As, As, ...}, then their union and
intersection are often written in the form UZ, A;and (2, Ai Observe
that we did not require the class {A,} to be non-empty. If it does12 Topology
happen that this class is empty, then the above definitions give (remem-
bering that all sets are subsets of U) U.A; = § and M4; = U. The
second of these facts amounts to the following statement: if we require
of an element that it belong to each set. in a given class, and if there are no
sets present in the class, then every element satisfies this requirement.
If we had not made the agreement that the only elements under consider-
ation are those in U, we would not have been able to assign a meaning to
the intersection of an empty class of sets. A moment’s consideration
makes it clear that Eqs. (1) are valid for arbitrary unions and intersections:
(UA)! = OA! and (NiAy’ = UsAY. (2)
It is instructive to verify these equations for the case in which the class
{ A.} is empty.
We conclude our treatment of the general theory of sets with a
brief discussion of certain special classes of sets which are of consider-
able importance in topology, logic, and measure theory. We usually
denote classes of sets by capital letters in boldface.
First, some general remarks which will be useful both now and
later, especially in connection with topological spaces. We shall often
have occasion to speak of finite unions and finite intersections, by which
we mean unions and intersections of finite classes of sets, and by a
finite class of sets we always mean one which is empty or consists of n
sets for some positive integer n. If we say that a class A of sets is closed
under the formation of finite unions, we mean that A contains the
union of each of its finite subclasses; and since the empty subclass
qualifies as a finite subclass of A, we see that its union, the empty set,
is necessarily an element of A. In the same way, a class of sets which is
closed under the formation of finite intersections necessarily contains
the universal set.
Now for the special classes of sets mentioned above. For the
remainder of this section we specifically assume that the universal set
U is non-empty. A Boolean algebra of sets is a non-empty class A of
subsets of U which has the following properties:
(@) AandBeAS>AUBEA;
(2) AandBeA>AMBeEA;
(3) AcAsd4’ed.
Since A is assumed to be non-empty, it must contain at least one set A.
Property (3) shows that A’ is in A along with A, and since A A’ = 6
and A U A’ = JU, (1) and (2) guarantee that A contains the empty set
and the universal set. Since the class consisting only of the empty set
and the universal set is clearly a Boolean algebra of sets, these two
distinct sets are the only ones which every Boolean algebra of sets mustSets and Functions = 13
contain. It is equally clear that the class of all subsets of U is also a
Boolean algebra of sets. There are many other less trivial kinds, and
their applications are manifold in fields of study as diverse as statistics
and electronics.
Let A be a Boolean algebra of sets. It is obvious that if
{A1, 42, ... , 4a} is a non-empty finite subclass of A, then
AiVUA2U+++UAn, and Ai MA2N +++ OAs
are both sets in A; and since A contains the empty set and the universal
set, it is easy to see that A is a class of sets which is closed under the
formation of finite unions, finite intersections, and complements. We
now go in the other direction, and let A be a class of sets which is closed
under the formation of finite unions, finite intersections, and comple-
ments. By these assumptions, A automatically contains the empty set
and the universal set, so it is non-empty and is easily seen to be a Boolean
algebra of sets. We conclude from these remarks that Boolean algebras
of sets can be described alternatively as classes of sets which are closed
under the formation of finite unions, finite intersections, and comple-
ments. It should be emphasized once again that when discussing
Boolean algebras of sets we always assume that the universal set is non-
empty.
One final comment. We speak of Boolean algebras of sets because
there are other kinds of Boolean algebras than those which consist of
sets, and we wish to preserve the distinction. We explore this topic
further in our Appendix on Boolean algebras.
Problems
1. If {Ai} and {B;} are two classes of sets such that {Aj C {B;},
show that U.A. C U;B; and NB; © MAG.
2. The difference between two sets A and B, denoted by A — B, is the set
of all elements in A and not in B; thus A — B = AM B’. Show
the following:
A-~B=4A-~(AMB)=(AUB)-B;
(A - B)-C=A-(BUC);
A-(B-C)
(AUB)-C=(A-C)UB-O);
A-(BUC)=(A-B)N(A-0C).
3. The symmetric difference of two sets A and B, denoted by AAB,
is defined by A AB = (A — B) U (B — A); it is thus the union of14. Topology
their differences in opposite orders. Show the following:
AA(BAC) = (AAB)AC;
AAG = A;ADA =;
AAB=BAA;
AM (BAC) = (ANB)A(ANC).
4. A ring of sets is a non-empty class A of sets such that if A and B are
in A, then A AB and A‘) B arealsoin A. Show that A must also
contain the empty set, AU B, and A — B. Show that if a non-
empty class of sets contains the union and difference of any pair of its
sets, then it is a ring of sets. Show that a Boolean algebra of sets is
a ring of sets.
5. Show that the class of all finite subsets (including the empty set) of
an infinite set is a ring of sets but is not a Boolean algebra of sets.
6. Show that the class of all finite unions of closed-open intervals on the
real line is a ring of sets but is not 2 Boolean algebra of sets.
7. Assuming that the universal set U is non-empty, show that Boolean
algebras of sets can be described as rings of sets which contain U.
3. FUNCTIONS
Many kinds of functions occur in topology, in a great variety of
situations. In our work we shall need the full power of the general con-
cept of a function, and since its modern meaning is much broader and
deeper than its elementary meaning, we discuss this concept in con-
siderable detail and develop its main abstract properties.
Let us begin with a brief inspection of some simple examples. Con-
sider the elementary function
yor
of the real variable x. What do we have in mind when we call this a
function and say that y is a function of z? In a nutshell, we are drawing
attention to the fact that each real number z has linked to it a specific
real number y, which can be calculated according to the rule (or law of
correspondence) given by the formula. We have here a process which,
applied to any real number z, does something to it (squares it) to produce
another real number y (the square of x). Similarly,
y=o?—3c and y= (2? +1)
are two other simple functions of the real variable z, and each is given
by a rule in the form of an algebraic expression which specifies the exact,
manner in which the value of y depends on the value of =.Sets and Functions 15
The rules for the functions we have just mentioned are expressed by
formulas. In general, this is possible only for functions of a very simple
kind or for those which are sufficiently important to deserve special
symbols of their own. Consider, for instance, the function of the real
variable z defined as follows: for each real number 2, write z as an infinite
decimal (using the scheme of decimal expansion in which infinite chains
of 9’s are avoided—in which, for example, 4 is represented by .25000 . . .
rather than by .24999 . . .); then let y be the fifty-ninth digit after
the decimal point. There is of course no standard formula for this,
but nevertheless it is a perfectly respectable function whose rule is given
by a verbal description. On the other hand, the function y = sin x of
the real variable z is so important that its rule, though fully as compli-
cated as the one just defined, is assigned the special symbol sin. When
discussing functions in general, we want to allow for all sorts of rules and
to talk about them all at once, so we usually employ noncommittal
notations like y = f(x), y = g(x), and so on.
Each of the functions mentioned above is defined for all real numbers
a. The example y = 1/z shows that this restriction is much too severe,
for this function is defined only for non-zero values of z. Similarly,
y = log z is defined only for positive values of z, and y = sin! z only for
values of x which lie in the interval [—1,1]. Whatever our conception
of a function may be, it should certainly be broad enough to include
examples like these, which are defined only for some values of the real
variable z.
In real analysis the notion of function is introduced in the following
way. Let X be any non-empty set of real numbers. We say that a
function y = f(z) is defined on X if the rule f associates a definite real
number y with each real number z in X. The specific nature of the
tule f is totally irrelevant to the concept of a function. The set X is
called the domain of the given function, and the set Y of all the values it
Assumes is called its range. If we speak of complex numbers here
instead of real numbers, we have the notion of function as it is used in
complex analysis.
This point of view toward functions is actually a bit more general
than is needed for the aims of analysis, but it isn’t nearly general
enough for our purposes. The sets X and Y above were taken to be
sets of numbers. If we now remove even this restriction and allow X
and Y to be completely arbitrary non-empty sets, then we arrive at the
most inclusive concept of a function. By way of illustration, suppose
that X is the set of all squares in a plane and that Y is the set of all
circles in the same plane. We can define a function y = f(x) by requiring
that the rule f associate with each square z that circle y which is inscribed
in it. In general, there is no need at all for either X or Y to be a set of16 Topology
numbers. All that is really necessary for a function is two non-empty
sets X and Y and a rule f which is meaningful and unambiguous in
assigning to each element z in X a specific element y in Y.
With these preliminary descriptive remarks, we now turn to the
rather abstract but very precise ideas they are intended to motivate.
A function consists of three objects: two non-empty sets X and Y
(which may be equal, but need not be) and a rule f which assigns to each
element x in X a single fully determined element y in Y. The y which
corresponds in this way to a given z is usually written f(x), and is called
the image of x under the rule f, or the value of f at the element x. This
f
fee eee aa
Fig. 6. A way of visualizing mappings.
notation is supposed to be suggestive of the idea that the rule f takes the
element z and does something to it to produce the element y = f(z).
The rule f is often called a mapping, or transformation, or operator, to
amplify this concept of it. We then think of f as mapping 2’s to y’s, or
transforming 2's into y’s, or operating on z’s to produce y’s. The sct X
is called the domain of the function, and the set of all f(z)’s for all z’s
in X is called its range. A function whose range consists of just one
element is called a constant function.
We often denote by f:X — Y the function with rule f, domain X,
and range contained in Y. This notation is useful because the essential
parts of the function are displayed in a manner which emphasizes that it
is a composite object, the central thing being the rule or mapping f.
Figure 6 gives a convenient way of picturing this function. On the
left, X and Y are different sets, and on the right, they are equal—in which
case we usually refer to f as a mapping of X into itself. If it is clear
from the context what the sets X and Y are, or if there is no real need to
specify them explicitly, it is common practice to identify the function
f:X— Y with the rule f, and to speak of f alone as if it were the function
under consideration (without mentioning the sets X and Y).
It sometimes happens that two perfectly definite sets X and Y are
under discussion and that a mapping of X into Y arises which has nu
natural symbol attached to it. If there is no necessity to invent aSets and Functions 7
symbol for this mapping, and if it is quite clear what the mapping is, it is
often convenient to designate it by x—» y. Accordingly, the function
y = 2 mentioned at the beginning of this section can be written as
2— x? or x— y (where y is understood to be the square of x).
A function f is called an extension of a function g (and g is called a
restriction of f) if the domain of f contains the domain of g and f(z) = g(z)
for each z in the domain of g.
Most of mathematical analysis, both classical and modern, deals
with functions whose values are real numbers or complex numbers.
f
emer esa
Pepe es
Fig. 7. The inverse of a mapping.
This is also true of those parts of topology which are concerned with
the foundations of analysis. If the range of a function consists of real
numbers, we call it a real function; similarly, a complex function is one
whose range consists of complex numbers. Obviously, every real function
is also complex. We lay very heavy emphasis on real and complex func-
tions throughout our work.
As a matter of usage, we generally prefer to reserve the term function
for real or complex functions and to speak of mappings when dealing
with functions whose values are not necessarily numbers.
Consider a mapping f:X — Y. When we call f a mapping of X
into Y, we mean to suggest by this that the elements f(z)—as x varies
over all the elements of X—need not fill up Y; but if it definitely does
happen that the range of f equals Y, or if we specifically want to assume
this, then we call f a mapping of X onto Y. If two different elements
in X always have different images under f, then we call f a one-to-one
mapping of X into Y. If f:X— Y is both onto and one-to-one, then
we can define its inverse mapping f-!: ¥ + X as follows: for each y in Y,
we find that unique element x in X such that f(z) = y (z exists and is
unique since f is onto and one-to-one); we then define z to be f-!(y).
The equation z = f-'(y) is the result of solving y = f(x) for x in just the
same way as x = log y is the result of solving y = e* for x. Figure 7
illustrates the concept of the inverse of a mapping.18 Topology
If f is a one-to-one mapping of X onto Y, it will sometimes be con-
venient to subordinate the conception of f as a mapping sending 2’s
over to y’s and to emphasize its role as a link between z’s and y’s. Each
z has linked to it (or has corresponding to it) precisely one y = f(x);
and, turning the situation around, each y has linked to it (or has corre-
sponding to it) exactly one z = f-'(y). When we focus our attention
on this aspect of a mapping which is one-to-one onto, we usually call it
a one-to-one correspondence. Thus f is a one-to-one correspondence
between X and Y, and f-1 is a one-to-one correspondence between Y
and X.
Now consider an arbitrary mapping f:X— Y. The mapping f,
which sends each element of X over to an element of Y, induces the
following two important set mappings. If A is a subset of X, then its
image f(A) is the subset of Y defined by
S(A) = {f(@):2€ A},
and our first set mapping is that which sends each A over to its corre-
sponding f(A). Similarly, if B is a subset of Y, then its inverse image
f-1(B) is the subset of X defined by
S-(B) = {x:f(z) € Bh,
and the second set mapping pulls each B back to its corresponding
J-\(B). It is often essential for us to know how these set mappings
behave with respect to set inclusion and operations on sets. We develop
most of their significant features in the following two paragraphs.
The main properties of the first set mapping are:
fM=9 SMOCY;
Arf Ar=f(Ai) € f(A);
HUA) = UIAd;
SOA) S Of (Ad
The reader should convince himself of the truth of these statements.
For instance, to prove (1) we would have to prove first that f(U;A,) isa
subset of Uif(Ai), and second that U,f(A,) is a subset of f(UiA).
A proof of the first of these set inclusions might run as follows: an element
in f(U;A,) is the image of some element in Ai, therefore it is the image
of an element in some Ay, therefore it is in some f(A,), and so finally it is in
Usf(As). The irregularities and gaps which the reader will notice in the
above statements are essential features of this set mapping. For exam-
ple, the image of an intersection need not equal the intersection of the
images, because two disjoint sets can easily have images which are not
disjoint. Furthermore, without special assumptions (see Problem 6)
nothing can be said about the relation between f(A)’ and f(A’).
q)Sets and Functions 19
The second set mapping is much better behaved. Its properties
are satisfyingly complete, and can be stated as follows:
f7G) =9; fr) =X;
B,C Br f(Bi) S f"(B2);
SUB) = Uif-(B); (2)
PMOB) = Of-(B)5 (3)
SB) = f(By. (4)
Again, the reader should verify each of these statements for himself.
f g
Bere EEE eae ec eer eer
(Pane.
f(z) (x),
Fig. 8. Multiplication of mappings.
We discuss one more concept in this section, that of the multiplication
(or composition) of mappings. If y = f(z) = 2* +1 and
z= gy) = sin y,
then these two functions can be put together to form a single function
defined by z = (gf)(x) = g(f(z)) = gz? + 1) = sin (2? +1). One of
the most important tools of calculus (the chain rule) explains how to dif-
ferentiate functions of this kind. This manner of multiplying functions
together is of basic importance for us as well, and we formulate it in
general as follows. Suppose that f:X — Y and g:Y— Z are any two
mappings. We define the product of these mappings, denoted by
of:X — Z, by (9f)(z) = g(f(x)). In words: an element z in X is taken
by f to the element f(z) in Y, and then g mapsf(zx) to g(f(x))in Z. Figure
8 isa picture of this process. We observe that the two mappings involved
here are not entirely arbitrary, for the set Y which contains the range of
the first equals the domain of the second. More generally, the product
of two mappings is meaningful whenever the range of the first is con-
tained in the domain of the second. We have regarded f as the first
mapping and g as the second, and in forming their product gf, their
symbols have gotten turned around. This is a rather unpleasant
phenomenon, for which we blame the occasional perversity of mathe-
matical symbols. Perhaps it will help the reader to keep this straight20 Topology
in his mind if he will remember to read the product gf from right to left:
first apply f, then g.
Problems
1, Two mappings f:X — Y and g:X — Y are said to be equal (and we
write this f = g) if f(x) = g(x) for every x in X. Let f, 9, andh
be any three mappings of a non-empty set X into itself, and show
that multiplication of mappings is associative in the sense that
S(gh) = (fg)h.
2. Let X be a non-empty set. The identity mapping ix on X is the
mapping of X onto itself defined by ix(x) = x for every x. Thus tx
sends each element of X to itself; that is, it leaves fixed each element
of X. Show that fix = txf = f for any mapping f of X into itself.
If f is one-to-one onto, so that its inverse f-' exists, show that
Sf = f-f = ix. Show further that f- is the only mapping of X
into itself which has this property; that is, show that if g is a mapping
of X into itself such that fg = gf =ix, then g =f! (hint:
9 = gix = g(ff") = Off = xf! = ft, or
9 = ixg = (F'fg = F(fo) = fix =F).
3. Let X and Y be non-empty sets and f a mapping of X into Y. Show
the following:
(a) f is one-to-one = there exists a mapping g of Y into X such
that gf = tx;
(6) f is onto + there exists a mapping h of Y into X such that
fh = ty.
4, Let X be a non-empty set and f a mapping of X into itself. Show
that f is one-to-one onto = there exists a mapping g of X into itself
such that fg = gf = ix. If there exists a mapping g with this
property, then there is only one such mapping. Why?
5. Let X be a non-empty set, and let f and g be one-to-one mappings
of X onto itself. Show that fg is also a one-to-one mapping of X
onto itself and that (fg)-! = g-f-1.
6. Let X and Y be non-empty sets and f a mapping of X into Y. If
A and B are, respectively, subsets of X and Y, show the following:
(a) ff-(B) C B, and ff-\(B) = B is true for all B + f is onto;
(b) A Cf-¥(A), and A = f-¥f(A) is true for all A = fis one-to-one;
(ce) f(A1O Az) = f(A) A f(A2) is true for all A, and A; @f is
one-to-one;
(@) f(A)’ C f(A’) is true for all A =f is onto;
(e) if f is onto—so that f(A)’ C f(A’) is true for all A—then
(A) = f(A’) is true for all A =f is also one-to-one.Sets and Functions 21
4, PRODUCTS OF SETS
We shall often have occasion to weld together the sets of a given
class into a single new set called their product (or their Cartesian product).
The ancestor of this concept is the coordinate plane of analytic geometry,
that is, a plane equipped with the usual rectangular coordinate system.
We give a brief description of this fundamental idea with a view to
paving the way for our discussion of products of sets in general.
First, a few preliminary comments about the real line. We have
already used this term several times without any explanation, and of
course what we mean by it is an ordinary geometric straight line (see
Fig. 9) whose points have been identified with—or coordinatized by—the
Ya Ye v2 er
o 1 2 3
Fig. 9. The real line.
set R of all real numbers. We use the letter R to denote the real line
as well as the set of all real numbers, and we often speak of real numbers
as if they were points on the real line, and of points on the real line as if
they were real numbers. Let, no one be deceived into thinking that the
real line is a simple thing, for its structure is exceedingly intricate. Our
present view of it, however, is as naive and uncomplicated as the picture
of it given in Fig. 9. Generally speaking, we assume that the reader is
familiar with the simpler properties of the real line—those relating to
inequalities (see Problem 1-2) and the basic algebraic operations of
addition, subtraction, multiplication, and division. One of the most
significant facts about the real number system is perhaps less well
known. This is the so-called least upper bound property, which asserts
that every non-empty set of real numbers which has an upper bound has
a least upper bound. It is an easy consequence of this that every non-
empty set of real numbers which has a lower bound has a greatest lower
bound. All these matters can be developed rigorously on the basis of a
small number of axioms, and detailed treatments can often be found in
books on elementary abstract elgebra.
To construct the coordinate plane, we now proceed as follows. We
take two identical replicas of the real line, which we call the z axis
and the y axis, and paste them on a plane at right angles to one another
in such a way that they cross at the zero point on each. The usual
picture is given in Fig. 10. Now let P be a point in the plane. We
project P perpendicularly onto points P, and P, on the axes. If z andy
are the coordinates of P; and P, on their respective axes, this process22 Topology
leads us from the point P to the uniquely determined ordered pair (z,y)
of real numbers, where x and y are called the x coordinate and y coordi-
nate of P. We can reverse the process, and, starting with the ordered
pair of real numbers, we can recapture the point. This is the manner in
which we establish the familiar one-to-one correspondence between
points P in the plane and ordered pairs (z,y) of real numbers. In fact,
we think of a point in the plane (which is a geometric object) and its
corresponding ordered pair of real numbers (which is an algebraic
object) as being—to all intents and purposes—identical with one another.
The essence of analytic geometry lies
y axis in the possibility of exploiting this
identification by using algebraic
tools in geometric arguments and
giving geometric interpretations to
ay Palay) algebraic calculations.
The conventional attitude to-
ward the coordinate plane in ana-
lytic geometry is that the geometry
is the focus of interest and the alge-
bra of ordered pairs is only a con-
xexis venient tool. Here we reverse this
point of view. For us, the coordinate
Fig. 10. The coordinate plane. plane is defined to be the set of all
ordered pairs (x,y) of real numbers.
We can satisfy our desire for visual images by using Fig. 10 as a picture
of this set and by calling such an ordered pair a point, but this geo-
metric language is more a convenience than a necessity.
Our notation for the coordinate plane is R XR, or R*. This
symbolism reflects the idea that the coordinate plane is the result of
“multiplying together” two replicas of the real line R.
It is perhaps necessary to comment on one possible source of mis-
understanding. When we speak of R? as a plane, we do so only to
establish an intuitive bond with the reader’s previous experience in ana-
lytic geometry. Our present attitude is that R? is a pure set and has no
structure whatever, because no structure has yet been assigned to it.
We remarked earlier (with deliberate vagueness) that a space is a set
to which has been added some kind of algebraic or geometric structure,
In Sec. 15 we shall convert the set 2? into the space of analytic geometry
by defining the distance between any two points (x1,y1) and (#2,y2) to be
VG ~ ay FH)
This notion of distance endows the set R? with a certain “spatial” char-
acter, which we shall recognize by calling the resulting space the Huclidean
plane instead of the coordinate plane.
Bey
1
|
'
!
I
1
It
'Sets and Functions = 23.
We assume that the reader is fully acquainted with the way in
which the set C of all complex numbers can be identified (as a set)
with the coordinate plane R®. If 2 is a complex number, and if z has
the standard form x + iy where x and y are real numbers, then we
identify z with the ordered pair (z,y), and thus with an element of R*.
The complex numbers, however, are much more than merely a set.
They constitute a number system, with operations of addition, multi-
plication, conjugation, etc. When the coordinate plane R? is thought
of as consisting of complex numbers and is enriched by the algebraic
structure it acquires in this way,
it is called the complex plane. The y,|
letter C is used to denote either
the set of all complex numbers or
the complex plane. We shall make
a space out of the complex plane
in Sec. 9.
Suppose now that Xi and X_
are any two non-empty sets. By
analogy with our above discussion,
their product X, X Xz is defined to aL x
be the set of allordered pairs (xi,22), Fig. 11. A way of visualizing X1 X Xz.
where z, is in X; and 22 is in Xo.
In spite of the arbitrary nature of X, and X:, their product can be repre-
sented by a picture (see Fig. 11) which is loosely similar to the usual
picture of the coordinate plane. The term product is applied to this set,
and it is thought of as the result of “multiplying together” Xi and X», for
the following reason: if X, and X; are finite sets with m and n elements,
then (clearly) X: X X2 has mn elements. If f:X:— X» is a mapping
with domain X; and range in X2, its graph is that subset of X; X Xe
which consists of all ordered pairs of the form (z1,f(z:)). We observe
that this is an appropriate generalization of the concept of the graph of
a function as it occurs in elementary mathematics.
This definition of the product of two sets extends easily to the case
of n sets for any positive integer n. If X1, X2,..., X, are non-empty
sets, then their product Xi X X2 X +--+ X Xn is the set of all ordered
n-tuples (21, Z2, . . . , Zn), where x; is in X; for each subscript 7. If the
X¥s are all replicas of a single set X, that is, if
Xoo =X,=4K,
then their product is usually denoted by the symbol X*.
These ideas specialize directly to yield the important sets R° and
C*, R'is just R, the real line, and R? is the coordinate plane. R*—the
set of all ordered triples of real numbers—is the set which underlies
solid analytic geometry, and we assume that the reader is familiar with
Re24 Topology
the manner in which this set arises, through the introduction of a rec-
tangular coordinate system into ordinary three-dimensional space. We
can draw pictures here just as in the case of the coordinate plane, and we
can use geometric language as much as we please, but it must be under-
stood that the mathematics of this set is the mathematics of ordered
triples of real numbers and that the pictures are merely an aid to the
intuition. Once we fully grasp this point of view, there is no difficulty
whatever in advancing at once to the study of the set R* of all ordered
n-tuples (21, 22, . . . , Zn) of real numbers for any positive integer n.
It is quite true that when n is greater than 3 it is no longer possible to
draw the same kinds of intuitively rich pictures, but at worst this is
merely an inconvenience. We can (and do) continue to use suggestive
geometric language, so all is not lost. The set C” is defined similarly:
it is the set of all ordered n-tuples (21, 22, . . . , 2n) of complex numbers.
Each of the sets R” and C* plays a prominent part in our later work.
We emphasized above that for the present the coordinate plane is to
be considered as merely a set, and not a space. Similar remarks apply
to R" and C*. In due course (in Sec. 15) we shall impart form and
content to each of these sets by suitable definitions. We shall convert
them into the Euclidean and unitary n-spaces which underlie and motivate
so many developments in modern pure mathematics, and we shall
explore some aspects of their algebraic and topological structure to the
very last pages of this book. But as of now—and this is the point we
insist on—neither one of these sets has any structure at all.
As the réader doubtless suspects, it is not enough that we consider
only products of finite classes of sets. The needs of topology compel
us to extend these ideas to arbitrary classes of sets.
We defined the product X: X X2X - ++ X X, to be the set of
all ordered n-tuples (x1, t2,..., %») such that 2; is in X; for each
subscript 7. To see how to extend this definition, we reformulate it as
follows. We have an index set J, consisting of the integers from 1 to n,
and corresponding to each index (or subscript) i we have a non-empty
set X; The n-tuple (x1, zz, . . . , 2a) is simply a function (call it z)
defined on the index set J, with the restriction that its value x(i) = xis
an element of the set X; for each iin J. Our point of view here is that
the function x is completely determined by, and is essentially equivalent
to, the array (21, 2, . . . , tn) of its values.
The way is now open for the definition of products in their full
generality. Let {X,} be a non-empty class of non-empty sets, indexed by
the elements i of an index set J. The sets X; need not be different
from one another; indeed, it may happen that they are all identical
replicas of a single set, distinguished only by different indices. The
product of the sets X;,, written Pir X;, is defined to be the set of all
functions z defined on I such that x(i) is an element of the set X; forSets and Functions 25
each index i. We call X; the tth coordinate set. When there can be no
misunderstanding about the index set, the symbol P;,; X; isoften abbrevi-
ated to P:X; The definition we have just given requires that each
coordinate set be non-empty before the product car be formed. It will
be useful if we extend this definition slightly by agreeing that if any of
the X;’s are empty, then P;X; is also empty.
This approach to the idea of the product of a class of sets, by means of
functions defined on the index set, is useful mainly in giving the definition.
In practice, it is much more convenient to use the subscript notation
a; instead of the function notation z(i). We then interpret the product
P.X, as made up of elements z, each of which is specified by the exhibited
array {x;} of its values in the respective coordinate sets X;. We call
a; the ith coordinate of the element x = {z;}.
The mapping p; of the product P;X; onto its ith coordinate set Xi
which is defined by p:(x) = x;—that is, the mapping whose value
at an arbitrary element of the product is the ith coordinate of that element
—is called the projection onto the ith coordinate set. The projection
p; selects the ith coordinate of each element in its domain. There is
clearly one projection for each element of the index set I, and the set of
all projections plays an important role in the general theory of topological
spaces.
Problems
1. The graph of a mapping f:X — Y is a subset of the product X x Y.
What properties characterize the graphs of mappings among all
subsets of X X Y¥?
2. Let X and Y be non-empty sets. If A; and A are subsets of X, and
B, and B; subsets of Y, show the following:
(Ai X By) 1 (Ai X Br) = (Ar 0 Aa) X (Br Ba);
(Ai X Bi) — (Aa X Bs) = (Air — Az) X (Br — Bz)
U (ALO Az) X (Bi — Ba)
U (Ai — A2) X (BLO B:).
3. Let X and Y be non-empty sets, and let A and B be rings of subsets
of X and Y, respectively. Show that the class of all finite unions of
sets of the form A X B with A € A and B ¢ Bisa ring of subsets of
xXxXY.
5. PARTITIONS AND EQUIVALENCE RELATIONS
In the first part of this section we consider a non-empty set X, and
we study decompositions of X into non-empty subsets which fill it out26 Topology
and have no elements in common with one another. We give special
attention to the tools (equivalence relations) which are normally used to
generate such decompositions.
A partition of X is a disjoint class {X,} of non-empty subsets of X
whose union is the full set X itself. The X,’s are called the partition sets.
Expressed somewhat differently, a partition of X is the result of splitting
it, or subdividing it, into non-empty subsets in such a way that each
element of X belongs to one and only one of the given subsets.
If X is the set {1, 2, 3, 4, 5}, then {1, 3, 5}, {2,4} and {1, 2, 3}, {4, 5}
are two different partitions of X. If X is the set R of all real numbers,
then we can partition X into the set of all rationals and the set of all
irrationals, or into the infinitely many closed-open intervals of the form
{n, n + 1) where n is an integer. If X is the set of all points in the
coordinate plane, then we can partition X in such a way that each
partition set, consists of all points with the same x coordinate (vertical
lines), or so that each partition set consists of all points with the same
y coordinate (horizontal lines).
Other partitions of each of these sets will readily occur to the reader.
In general, there are many different ways in which any given set can
be partitioned. These manufactured examples are admittedly rather
uninspiring and serve only to make our ideas more concrete. Later in
this section we consider some others which are more germane to our
present purposes.
A binary relation in the set X is a mathematical symbol or verbal
phrase, which we denote by R in this paragraph, such that for each
ordered pair (z,y) of elements of X the statement z Ry is meaningful,
in the sense that it can be classified definitely as true or false. For such
a binary relation, z R y symbolizes the assertion that x is related by R to
y, and x Wy the negation of this, namely, the assertion that x is not
related by R to y Many examples of binary relations can be given,
some familiar and others less so, some mathematical and others not.
For instance, if X is the set of all integers and R is interpreted to mean
“Gs less than,” which of course is usually denoted by the symbol <, then
we clearly have 4 <7 and 5¢ 2. We have been speaking of binary
relations, which are so named because they apply only to ordered pairs
of elements, rather than to ordered triples, etc. In our work we drop the
qualifying adjective and speak simply of a relation in X, since we shall
have occasion to consider only relations of this kind.!
We now assume that a partition of our non-empty set X is given,
1 Some writers prefer to regard a relation R in X asa subset Rof X X X. From
this point of view, z Ry and x Hy are simply equivalent ways of writing (z,y)¢ R
and (z,y)¢R. This definition has the advantage of being more tangible than ours,
and the disadvantage that few people really think of a relation in this way.Sets and Functions 27
and we associate with this partition a relation in X. This relation is
defined in the following way: we say that « is equivalent to y and write
this z ~ y (the symbol ~ is pronounced “wiggle”), if z and y belong to
the same partition set. It is obvious that the relation ~ has the follow-
ing properties:
(1) 2 ~ a for every x (reflexivity);
(2) t~y=sy~ a (symmetry);
(3) s~yandy~z=2 ~z (transitivity).
This particular relation in X arose in a special way, in connection with a
given partition of X, and its properties are immediate consequences of its
definition. Any relation whatever in X which possesses these three
properties is called an equivalence relation in X.
We have just seen that each partition of X has associated with
it a natural equivalence relation in X. We now reverse the situation
and show that a given equivalence relation in X determines a natural
partition of X.
Let ~ be an equivalence relation in X; that is, assume that it is
reflexive, symmetric, and transitive in the sense described above. If x
is an element of X, the subset of X defined by [z] = {y:y ~ 2} is called
the equivalence set of x. The equivalence set of z is thus the set of all
elements which are equivalent to x. We show that the class of all
distinct equivalence sets forms a partition of X. By reflexivity, x [z]
for each element z in X, so each equivalence set is non-empty and
their union is X. It remains to be shown that any two equivalence sets
[x1] and [z2] are either disjoint or identical. We prove this by showing
that if [z:] and [x2] are not disjoint, then they must be identical. Sup-
pose that [21] and [zs] are not disjoint; that is, suppose that they have
a common element z. Since z belongs to both equivalence sets, 2 ~ 21
and z~ 22, and by symmetry, %;~ 2. Let y be any element of [x],
so that y~a:. Since y~ a: and x ~z, transitivity shows that
y~z. By another application of transitivity, y ~~ z and z~ 2 imply
that y ~ 22, so that yisin [z.]. Since y was chosen arbitrarily in [x1], we
see by this that [z,] C [x]. The same reasoning shows that [x2] C [x], and
from this we conclude (sce the last paragraph of Sec. 1) that [z:] = [zz].
The above discussion demonstrates that there is no real distinction
(other than a difference in language) between partitions of a set and
equivalence relations in the set. If we start with a partition, we get an
equivalence relation by regarding elements as equivalent if they belong
to the same partition set, and if we start with an equivalence relation,
we get a partition by grouping together into subsets all elements which
are equivalent to one another. We have here a single mathematical
idea, which we have been considering from two different points of view,
and the approach we choose in any particular application depends entirely28 = Topology
on our own convenience. In practice, it is almost invariably the case that.
we use equivalence relations (which are usually easy to define) to obtain
partitions (which are sometimes difficult to describe fully).
We now turn to several of the more important simple examples of
equivalence relations.
Let I be the set of all integers. If a and b are elements of this set,
we write a = b (and say that a equals b) if a and b are the same integer.
Thus 2 + 3 = 5 means that the expressions on the left and right are
simply different ways of writing the same integer. It is apparent that =
used in this sense is an equivalence relation in the set J:
(1) a = a for every a;
(2) b=b=a;
(3) a=bandb=c>a=c.
Clearly, each equivalence set consists of precisely one integer.
Another familiar example is the relation of equality commonly used
for fractions. We remind the reader that, strictly speaking, a fraction
is merely a symbol of the form a/b, where a and b are integers and b is
not zero. The fractions 24 and 46 are obviously not identical, but
nevertheless we consider them to be equal. In general, we say that
two fractions a/b and c/d are equal, written a/b = c/d, if ad and be are
equal as integers in the usual sense (see the above paragraph). We leave
it to the reader to show that this is an equivalence relation in the set of
all fractions. An equivalence set of fractions is what we call a rational
number. Everyday usage ignores the distinction between fractions and
rational numbers, but it is important to recognize that from the strict
point of view it is the rational numbers (and not the fractions) which
form part of the real number system.
Our final example has a deeper significance, for it provides us with
the basic tool for our work of the next two sections.
For the remainder of this section we consider a relation between
pairs of non-empty sets, and each set mentioned (whether we say so
explicitly or not) is assumed to be non-empty. If X and Y are two
sets, we say that X is numerically equivalent to Y if there exists a one-to-
one correspondence between X and Y, i.e., if there exists a one-to-one
mapping of X onto Y. This relation is reflexive, since the identity
mapping tx: X — X is one-to-one onto; it is symmetric, since if f:X > Y
is one-to-one onto, then its inverse mapping f-': Y > X is also one-to-one
onto; and it is transitive, since if f:X — Y and g:Y — Z are one-to-one
onto, then gf: X — Z is also one-to-one onto. Numerical equivalence has
all the properties of an equivalence relation, and if we consider it as an
equivalence relation in the class of all non-empty subsets of some universal
set U, it groups together into equivalence sets all those subsets of U
which have the same number of elements. After we state and prove theSets and Functions 29
following very useful but rather technical theorem, we shall continue in
Secs. 6 and 7 with an exploration of the implications of these ideas.
The theorem we have in mind—the Schroeder-Bernstein theorem—is
the following: if X and Y are two sets each of which is numerically equivalent
to a subset of the other, then all of X is numerically equivalent to all of Y.
There are several proofs of this classic theorem, some of which are quite
difficult. The very elegant proof we give is essentially due to Birkhoff
and MacLane.
Now for the proof. We assume that f:X— Y is a one-to-one
mapping of X into Y, and that g: Y > X is a one-to-one mapping of Y
into X. Our task is to produce a mapping F': X — Y which is one-to-one
onto. We may assume that neither f nor g is onto, since if f is, we can
define F to be f, and if g is, we can define F to be g-'. Since both f and g
are one-to-one, it is permissible to use the mappings f-' and g™! as long
as we clearly understand that f—! is defined only on f(X) and g™ only on
g(Y). We obtain the mapping F by splitting both X and Y into subsets
which we characterize in terms of the ancestry of their elements. Let z
be an element of X. We apply g~! to it (if we can) to get the element
g(x) in Y. If g-*(z) exists, we call it the first ancestor of z. The ele-
ment z itself we call the zeroth ancestor of z. We now apply f-! to
g7\(z) if we can, and if (f-'g-*)(z) exists, we call it the second ancestor
of x. We now apply g™ to (f-\g-1)(z) if we can, and if (g~'f-"g—)(z)
exists, we call it the third ancestor of x. As we continue this process of
tracing back the ancestry of x, it becomes apparent that there are three
possibilities. (1) z has infinitely many ancestors. We denote by X;
the subset of X which consists of all elements with infinitely many
ancestors. (2) 2 has an even number of ancestors; this means that z
has a last ancestor (that is, one which itself has no first ancestor) in X.
We denote by X, the subset of X consisting of all elements with an even
number of ancestors. (3) z has an odd number of ancestors; this
means that x has a last ancestor in Y. We denote by X, the subset of X
which consists of all elements with an odd number of ancestors. The
three sets X;, X., X, form a disjoint class whose union is X. We decom-
pose Y in just the same way into three subsets Y;, Y., Y.. It is easy to
see that f maps X; onto Y; and X, onto Y,, and that g-! maps X, onto
Y,; and we complete the proof by defining F in the following piecemeal
manner:
f(z) ifceX:UX,
g(x) ifee X.
We attempt to illustrate these ideas in Fig. 12. Here we present two
replicas of the situation: on the left, X and Y are represented by the
vertical lines, and f and g by the lines slanting down to the right and
F(z) =30
Topology
left; and on the right, we schematically trace the ancestry of three ele-
ments in X, of which 2; has no first ancestor, z2 has a first and second
ancestor, and 2; has a first, second, and third ancestor.
x Y
al) (X)
Fig. 12, The proof of the Schroeder-Bernstein theorem.
The Schroeder-Bernstein theorem has great theoretical and practical
significance. Its main value for us lies in its role as a tool by means of
which we can prove numerical equivalence with a minimum of effort for
many specific sets. We put it to work in Sec. 7.
Problems
1
Let f: X — Y be an arbitrary mapping. Define a relation in X as
follows: 21 -~ 22 means that f(z:) = f(z2). Show that this is an
equivalence relation and describe the equivalence sets.
In the set F of all real numbers, let s ~~ y mean that x — y is an
integer. Show that this is an equivalence relation and describe the
equivalence sets.
Let I be the set of all integers, and let m be a fixed positive integer.
Two integers a and b are said to be congruent modulo m—symbolized
by a = b (mod m)—if a — b is exactly divisible by m, i.e., if a — bis
an integral multiple of m. Show that this is an equivalence relation,
describe the equivalence sets, and state the number of distinct
equivalence sets.
Decide which ones of the three properties of reflexivity, symmetry,
and transitivity are true for each of the following relations in the setSets and Functions 31
of all positive integers: m y ~ 2;
a~yand y~z>2x~z; therefore x ~z for every z. In view
of Problem 5f, this cannot be a valid proof. What is the flaw in the
reasoning?
7. Let X be a non-empty set. A relation ~ in X is called circular if
‘a~yandy ~2=2 ~ 2, and triangular ifz ~~ yandz ~z>y ~z.
Prove that a relation in X is an equivalence relation + it is reflexive
and circular + it is reflexive and triangular.
a
. COUNTABLE SETS
‘The subject of this section and the next—injintte cardinal numbers—
lies at the very foundation of modern mathematics. It is a vital instru-
ment in the day-to-day work of many mathematicians, and we shall
make extensive use of it ourselves. This theory, which was created by
the German mathematician Cantor, also has great aesthetic appeal, for
it begins with ideas of extreme simplicity and develops through natural
stages into an elaborate and beautiful structure of thought. In the
course of our discussion we shall answer questions which no one before
Cantor’s time thought to ask, and we shall ask a question which no one
can answer to this day.
Without further ado, we can say that cardinal numbers are those used
in counting, such as the positive integers (or natural numbers) 1, 2,
3, . . . familiar to us all, But there is much more to the story than this.
The act of counting is undoubtedly one of the oldest of human
activities. Men probably learned to count in a crude way at about the
same time as they began to develop articulate speech. The earliest men
who lived in communities and domesticated animals must have found
it necessary to record the number of goats in the village herd by means of
a pile of stones or some similar device. If the herd was counted in each
night by removing one stone from the pile for each goat accounted for,
then stones left over would have indicated strays, and herdsmen would
have gone out to search for them. Names for numbers and symbols for32 Topology
them, like our 1, 2, 3, . . . , would have been superfluous. The simple
and yet profound idea of a one-to-one correspondence between the
stones and the goats would have fully met the needs of the situation.
In a manner of speaking, we ourselves use the infinite set
N = {1,2,3,...}
of all positive integers as a “pile of stones.” We carry this set around
with us as part of our intellectual equipment. Whenever we want to count
a set, say, a stack of dollar bills, we start through the set N and tally off
one bill against each positive integer as we come to it. The last number
we reach, corresponding to the last bill, is what we call the number of
bills in the stack. If this last number happens to be 10, then “10” is
our symbol for the number of bills in the stack, as it also is for the number
of our fingers, and for the number of our toes, and for the number of
elements in any set which can be put into one-to-one correspondence
with the finite set {1, 2, , 10}. Our procedure is slightly more
sophisticated than that of the primitive savage. We have the symbols
1, 2, 8, . . . for the numbers which arise in counting; we can record
them for future use, and communicate them to other people, and manipu-
late them by the operations of arithmetic. But the underlying idea,
that of the one-to-one correspondence, remains the same for us as it
probably was for him.
The positive integers are adequate for the purpose of counting any
non-empty finite set, and since outside of mathematics all sets appear to
be of this kind, they suffice for all non-mathematical counting. But in
the world of mathematics we are obliged to consider many infinite sets,
such as the set of all positive integers itself, the sct of all integers, the
set of all rational numbers, the set of all real numbers, the set of all
points in a plane, and so on. It is often important to be able to count
such sets, and it was Cantor’s idea to do this, and to develop a theory of
infinite cardinal numbers, by means of one-to-one correspondences.
In comparing the sizes of two sets, the basic concept is that of
numerical equivalence as defined in the previous section. We recall
that two non-empty sets X and Y are said to be numerically equivalent if
there exists a one-to-one mapping of one onto the other, or—and this
amounts to the same thing—if there can be found a one-to-one corre-
spondence between them. To say that two non-empty finite sets are
numerically equivalent is of course to say that they have the same number
of elements in the ordinary sense. If we count one of them, we simply
establish a one-to-one correspondence between its elements and a set of
positive integers of the form {1, 2, . . . , n}, and we then say that n is
the number of elements possessed by both, or the cardinal number of both.
The positive integers are the finite cardinal numbers. We encounterSets and Functions 33
many surprises as we follow Cantor and consider numerical equivalence
for infinite sets.
The set N = {1, 2, 3, .. .} of all positive integers is obviously
“larger” than the set {2, 4, 6, . . .} of all even positive integers, for it
contains this set as a proper subset. It appears on the surface that NV has
“more” elements. But it is very important to avoid jumping to con-
clusions when dealing with infinite sets, and we must remember that our
criterion in these matters is whether there exists a one-to-one corre-
spondence between the sets (not whether one set is or is not a proper
subset of the other). Asa matter of fact, the pairing
NaS pSE Het at eects
2,4,6,...,2n,...
serves to establish a one-to-one correspondence between these sets, in
which each positive integer in the upper row is matched with the even
positive integer (its double) directly below it, and these two sets must
therefore be regarded as having the same number of elements. This is a
very remarkable circumstance, for it seems to contradict our intuition
and yet is based only on solid common sense. We shall see below, in
Problems 6 and 7-4, that every infinite set is numerically equivalent to a
proper subset of itself. Since this property is clearly not possessed by
any finite set, some writers even use it as the definition of an infinite set.
In much the same way as above, we can show that N is numerically
equivalent to the set of all even integers:
OH g ratte He ttge
0,2, —2, 4, —4,6, -6,...
Here our device is to start with 0 and follow each even positive integer
as we come to it by its negative. Similarly, N is numerically equivalent
to the set of all integers:
1,2, 3,4, 5,6, 7...
0,1, -1,2, -2,3, -3,...
It is of considerable historical interest to note that Galileo observed in the
early seventeenth century that there are precisely as many perfect
squares (1, 4, 9, 16, 25, etc.) among the positive integers as there are
positive integers altogether. This is clear from the pairing
1, 2, 3,4, 5...
1, 2, 34, 43,53...
It struck him as very strange that this should be true, considering how34 ~— Topology
sparsely strewn the squares are among all the positive integers. But.
the time appears not to have been ripe for the exploration of this phenom-
enon, or perhaps he had other things on his mind; in any case, he did not
follow up his idea.
These examples should make it clear that all that is really necessary
in showing that an infinite set X is numerically equivalent to N is that we
be able to list the elements of X, with a first, a second, a third, and so on,
in such a way that it is completely exhausted by this counting off of its
elements. It is for this reason that any infinite set which is numerically
equivalent to N is said to be countably infinite. We say that a set is
countable if it is non-empty and finite (in which case it can obviously be
counted) or if it is countably infinite.
One of Cantor’s earliest discoveries in his study of infinite sets was
that the set of all positive rational numbers (which is very large: it
contains N and a great many other numbers besides) is actually countable.
We cannot list the positive rational numbers in order of size, as we can
the positive integers, beginning with the smallest, then the next smallest,
and so on, for there is no smallest, and between any two there are infi-
nitely many others. We must find some other way of counting them,
and following Cantor, we arrange them not in order of size, but according
to the size of the sum of the numerator and denominator. We begin
with all positive rationals whose numerator and denominator add up to 2:
there is only one, }{ = 1. Next we list (with increasing numerators) all
those for which this sum is 3:14, 2{ = 2. Next, all those for which this
sum is 4:14, 34 = 1, 3 = 3. Next, all those for which this sum is
5:4, %, 34, = 4. Next, all those for which this sum is 6:14, 34 = 34,
3 = 1, 4% = 2,54 = 5. And soon. If we now list all these together
from the beginning, omitting those already listed when we come to them,
we get-a sequence
1, 34, 2, 18, 3, 14, 3%, 94, 4, 18,5,»
which contains each positive rational number once and only once.
Figure 13 gives a schematic representation of this manner of listing the
positive rationals. In this figure the first row contains all positive
rationals with numerator 1, the second all with numerator 2, etc.; and
the first column contains all with denominator 1, the second all with
denominator 2, and so on. Our listing amounts to traversing this array
of numbers as the arrows indicate, where of course all those numbers
already encountered are left out.
It’s high time that we christened the infinite cardinal number we’ ve
been discussing, and for this purpose we use the first letter of the Hebrew
alphabet (N, pronounced “aleph”) with 0 as a subscript. We say
that N> is the number of elements in any countably infinite set. OurSets and Functions 35
vomplete list of cardinal numbers so far is
1,2,3,...,%o
We expand this list in the next section.
Suppose now that m and n are two cardinal numbers (finite or
infinite). The statement that m is less than n (written m < 7) is defined
to mean the following: if X and Y are sets with m and n elements, then
2iadiaidk oo a
1 2 3 4 5
eee Heer HCeeeHeeeeH-Heee eerie a
1 2 3 4 6
ee ee eee eee eee eee aa--
1 2 3 4 5
Eee Oe ee ee ee pees
Ths tis
gE Ee tee ree ree eee et ae
1 2 3 4 5
Fig. 13. A listing of the positive rationals.
(1) there exists a one-to-one mapping of X into Y, and (2) there does not
exist a one-to-one mapping of X onto Y. Using this concept, it is easy to
relate our cardinal numbers to one another by means of
1<2<3< +++ and less than
c? No one knows the answer to this question. Cantor himself thought
that there is no such number, or in other words, that c is the next infinite
cardinal number greater than Xo, and his guess has come to be known as
Cantor’s continuum hypothesis. The continuum hypothesis can also be
expressed by the assertion that every uncountable set of real numbers
has ¢ as its cardinal number.'
There is another question which arises naturally at this stage, and
this one we are fortunately able to answer. Are there any infinite
cardinal numbers greater than c? Yes, there are; for example, the
cardinal number of the class of all subsets of R. This answer depends on
the following fact: if X is any non-empty set, then the cardinal number of
X is less than the cardinal number of the class of all subsets of X.
We prove this statement as follows. In accordance with the defini-
tion given in the last paragraph of the previous section, we must show
1 For further information about the continuum hypothesis, see Wilder (42, p. 125]
and Gadel [12].«0 Topology
(1) that there exists a one-to-one mapping of X into the class of all its
subsets, and (2) that there does not exist such a mapping of X onto this
class. To prove (1), we have only to point to the mapping z— {z},
which makes correspond to each element z that set {x} which consists of
the element x alone. We prove (2) indirectly. Let us assume that
there does exist a one-to-one mapping f of X onto the class of all its
subsets. We now deduce a contradiction from the assumed existence of
such a mapping. Let A be the subset of X defined by A = {a:2¢f(z)}.
Since our mapping f is onto, there must exist an element ain X such that
S(@ = A. Where is the element a? Jf ais in A, then by the definition
of A we have a ¢ f(a), and since f(a) = A,ag¢A. Thisisa contradiction,
soacannot belong to A. Butif ais notin A, then again by the definition
of A we have aéf(a) or ae A, which is another contradiction. The
situation is impossible, so our assumption that such a mapping exists
must be false.
This result guarantees that given any cardinal number, there always
exists a greater one. If we start with a set X, = {1} containing one
element, then there are two subsets, the empty set @ and the set {1}
itself. If Xz: = {1,2} is a set containing two elements, then there are
four subsets: 6, {1}, {2}, {1,2}. If Xs = {1, 2, 3} is a set containing
three elements, then there are eight subsets: @, {1}, {2}, {3}, {1,2}, {1,3},
{2,3}, {1, 2,3}. In general, if X, is a set with n elements, where 7 is any
finite cardinal number, then X, has 2" subsets. If we now take n to be
any infinite cardinal number, the above facts suggest that we define 2" to
be the number of subsets of any set with n elements. If n is the first
infinite cardinal number, namely, No, then it can be shown that
ok = 6,
The simplest proof of this fact depends on the ideas developed in the
following paragraph.
Consider the closed-open unit interval [0,1) and a real number x in
this set. Our concern is with the meaning of the decimal, binary, and
ternary expansions of x. For the sake of clarity, let us take z to be }4.
How do we arrive at the decimal expansion of }4? First, we split (0,1)
into the 10 closed-open intervals
(0,40), B40,3£0), - » - (Mos),
and we use the 10 digits 0, 1, . . . , 9 to number them in order. Our
number }4 belongs to exactly one of these intervals, namely, to [3{0,340).
We have labeled this interval with the digit 2, so 2 is the first digit after
the decimal point in the decimal expansion of 44:
w= 2...Sets and Functions = 41
Next, we split the interval [2{0,3{0) into the 10 closed-open intervals
(%{o,?4o0), [?4400,72f00), - - - » (?%00,340),
and we use the 10 digits to number these in order. Our number }4
belongs to [25490,28{00), which is labeled with the digit 5, so 5 is the
second number after the decimal point in the decimal expansion of 34:
Wa 2...
If we continue this process exactly as we started it, we can obtain the
decimal expansion of }4 to as many places as we wish. As a matter of
fact, if we do continue, we get 0 at each stage from this point on:
4 = .25000...
The reader should notice that there is no ambiguity in this system as we
have explained it: contrary to customary usage, .24999 . . . is not to
be regarded as another decimal expansion of 14 which is “equivalent” to
25000... . In this system, each real number z in (0,1) has one and
only one decimal expansion which cannot end in an infinite chain of 9’s.
There is nothing magical about the role of the number 10 in the above
discussion. If at each stage we split our closed-open interval into two
equal closed-open intervals, and if we use the two digits 0 and 1 to
number them, we obtain the binary expansion of any real number z in
(0,1). The binary expansion of 14 is easily seen to be .01000....
The ternary expansion of z is found similarly: at each stage we split our
closed-open interval into three equal closed-open intervals, and we use
the three digits 0, 1, and 2 to number them. A moment’s thought should
convince the reader that the ternary expansion of 14 is .020202
Just as (in our system) the decimal expansion of a number in (0,1) cannot
end in an infinite chain of 9’s, so also its binary expansion cannot end in
an infinite chain of 1’s, and its ternary expansion cannot end in an
infinite chain of 2's.
We now use this machinery to give a proof of the fact that
ae = 6,
Consider the two sets N = {1, 2,3, .. .} and J = (0,1), the first with
cardinal number No and the second with cardinal number c. If N denotes
the class of all subsets of N, then by definition N has cardinal number
2** Our proof amounts to showing that there exists a one-to-one corre-
spondence between N and J. We begin by establishing a one-to-one
mapping f of Ninto J. If A is a subset of N, then f(A) is that real num-
ber z in J whose decimal expansion z = .didads . . . is defined by the
condition that d, is 3 or 5 according asnisorisnotin A. Any other two
digits can be used here, as long as neither of them is 9. Next, we con-42 Topology
struct a one-to-one mapping g of J intoN. If z isa real number in I, and
if = .bibebs . . . is its binary expansion (so that each b, is either 0 or 1),
then g(x) is that subset A of N defined by A = {n:b, = 1}. We con-
clude the proof with an appeal to the Schroeder-Bernstein theorem, which
guarantees that under these conditions N and J are numerically equivalent
to one another.
If we follow up the hint contained in the fact that 2** = c, and
successively form 2¢, 2%, and so on, we get a chain of cardinal numbers
1<2 x. A non-empty
set P in which there is defined a partial order relation is called a partially44 Topology
ordered set. It is clear that any non-empty subset of a partially ordered
set is a partially ordered set in its own right.
Partially ordered sets are abundant in all branches of mathematics.
Some are simple and easy to grasp, while others are complex and rather
inaccessible. We give four examples which are quite different in nature
but possess in common the virtues of being both important and easily
described.
Example 1. Let P be the set of all positive integers, and let m z= y = 2, that is, if no element other than z itself is
greater than or equal toz. A maximal element in P is thus an element of
P which is not less than or equal to any other element of P. Examples
1, 2, and 4 have no maximal elements. Example 3 has a single maximal
element: the set U itself.
Let A be a non-empty subset of a partially ordered set P. An
element z in P is called a lower bound of A if z < a for each a€ A; anda
lower bound of A is called a greatest lower bound of A if it is greater than orSets and Functions 45
equal to every lower bound of A. Similarly, an element y in P is said
to be an upper bound of A if a < y for every ae A; and a least upper
bound of A is an upper bound of A which is less than or equal to every
upper bound of A. In general, A may have many lower bounds and
many upper bounds, but it is easy to prove (see Problem 1) that a
greatest lower bound (or least upper bound) is unique if it exists. It is
therefore legitimate to speak of the greatest lower bound and the least
upper bound if they exist.
We illustrate these concepts in some of the partially ordered sets
mentioned above.
In Example 1, let the subset A consist of the integers 4 and 6. An
upper bound of {4,6} is any positive integer divisible by both 4 and 6.
12, 24, 36, and so on, are all upper bounds of {4,6}. 12 is clearly its
least upper bound, for it is less than or equal to (i.e., it divides) every
upper bound. The greatest lower bound of any pair of integers in this
example is their greatest common divisor, and their least upper bound is
their least common multiple—both of which are familiar notions from
elementary arithmetic.
We now consider Example 2, the real line with its natural order
relation. The reader will doubtless recall from his study of calculus that
3 is an upper bound of the set {(1 + 1/n)":n = 1, 2, 3, . . .J} and that
its least upper bound is the fundamental constant e = 2.7182....
As we have stated before, it is a basic property of the real line that every
non-empty subset of it which has a lower bound (or upper bound) has a
greatest lower bound (or least upper bound). ‘There are several items of
standard notation and terminology which must be mentioned in connec-
tion with this example. Let A be any non-empty set of real numbers.
If A has a lower bound, then its greatest lower bound is usually called
its infimum and denoted by inf A. Correspondingly, if A has an upper
bound, then its least upper bound is called its supremum and written
sup A. If A happens to be finite, then inf A and sup A both exist and
belong to A. In this case, they are often called the minimum and
maximum of A and are denoted by min A and max A. If A consists of
two real numbers a; and a», then min A is the smaller of a; and a2, and
max A is the larger.
Finally, consider Example 3, and let A be any non-empty class of
subsets of U. A lower bound of A is any subset of U which is contained
in every set in A, and the greatest lower bound of A is the intersection
of all its sets. Similarly, the least upper bound of A is the union of
all its sets.
One of our main aims in this section is to state Zorn’s lemma, an
exceedingly powerful tool of proof which is almost indispensable in
many parts of modern pure mathematics. Zorn’s lemma asserts that4 Topology
if P is a partially ordered set in which every chain has an upper bound, then
P possesses a mazimal element. It is not possible to prove this in the
usual sense of the word. However, it can be shown that Zorn’s lemma is
logically equivalent to the aziom of choice, which states: given any
non-empty class of disjoint non-empty sets, a set can be formed which
contains precisely one element taken from each set in the given class.
The axiom of choice may strike the reader as being intuitively obvious,
and in fact, either this axiom itself or some other principle equivalent to
Fig. 16. The geometric meaning of f Ag and f V9.
it is usually postulated in the logic with which we operate. We therefore
assume Zorn’s lemma as an axiom of logic. Any reader who is inter-
ested in these matters is urged to explore them further in the literature.!
A lattice is a partially ordered set L in which each pair of elements has
a greatest lower bound and a least upper bound. If z and y are two
elements in L, we denote their greatest lower bound and least upper
bound by zany and zvy. These notations are analogous to (and are
intended to suggest) the notations for the intersection and union of two
sets. We pursue this analogy even further, and call ray and xv y the
meet and join of x and y. It is tempting to assume that all properties of
intersections and unions in the algebra of sets carry over to lattices, but
this is not a valid assumption. Some properties do carry over (see
Problem 5), but others, for instance the distributive laws, are false in
some lattices.
It is easy to see that all four of our examples are lattices. In
Example 1, m an is the greatest common divisor of m and n, and m v nis
their least common multiple; and in Example 3, AaB = AB and
AvB=AUB. In Example 2, if z and y are any two real numbers,
then xay is min {z,y} and vy is max {x,y}. In Example 4, fag is
48ee, for example, Wilder [42, pp. 129-132], Halmos [16, secs. 15-16], Birkho
[4, p. 42], Sierpinski (37, chap. 6], or Fraenkel and Bar-Hillel [10, p. 44].Sets and Functions 47
the real function defined on X by (fag)(z) = min {f(z),g(z)}, and
fvg is that defined by (fv g)(z) = max {f(x),g(z)}. Figure 16 illus-
trates the geometric meaning of fag and fvg for two real functions
f and g defined on the closed unit interval [0,1].
Let L be a lattice. A sublattice of L is a non-empty subset L, of L
with the property that if z and y are in L,, then xay and xv y are also
in L,. If L is the lattice of all real functions defined on the closed unit
interval, and if L, is the set of all continuous functions in L, then L, is
easily seen to be a sublattice of L.
If a lattice has the additional property that every non-empty subset
has a greatest lower bound and a least upper bound, then it is called a
complete lattice. Example 3 is the only complete lattice in our list.
There are many distinct types of lattices, and the theory of these
systems has a wide variety of interesting and significant applications
(see Birkhoff [4]). We discuss some of these types in our Appendix on
Boolean algebras.
Problems
1, Let A be a non-empty subset of a partially ordered set P. Show
that A has at most one greatest lower bound and at most one least
upper bound.
2. Consider the set {1, 2,3, 4,5}. What elements are maximal if it is
ordered as Example 1? If it is ordered as Example 2?
Under what circumstances is Example 4 @ chain?
Give an example of a partially ordered set which is not a lattice.
Let L be a lattice. If x, y, and z are elements of L, verify the
following: raz =2,2V2=2,0AY=YAL,LZVY=YVA,
Pre
aa(yaz) = (ray) rz,
av (yvz) = (evy)v2, (@ay)va=a,(avyrar=a.
6. Let A be a class of subsets of some non-empty universal set U. We
say that, A has the finite intersection property if every finite subclass of
A has non-empty intersection. Use Zorn’s lemma to prove that
if A has the finite intersection property, then it is contained in some
maximal class B with this property (to say that B is a maximal class
with this property is to say that any class which properly contains B
fails to have this property). (Hint: consider the family of all
classes which contain A and have the finite intersection property,
order this family by class inclusion, and show that any chain in the
family has an upper bound in the family.)
7. Prove that if X and Y are any two non-empty sets, then there exists
@ one-to-one mapping of one into the other. (Hint: choose an48
10.
Topology
element z in X and an element y in Y, and establish the obvious
one-to-one correspondence between the two single-element sets
{x} and {y}; define an extension to be a pair of subsets A of X and
B of Y such that {x} C A and {y} C B, together with a one-to-one
correspondence between them under which z and y correspond with
one another; order the set of all extensions in the natural way; and
apply Zorn’s lemma.)
Let m and n be any two cardinal numbers (finite or infinite). The
statement that m is less than or equal to n (written m < n) is defined
to mean the following: if X and Y are sets with m and n elements,
then there exists a one-to-one mapping of X into Y. Prove that
any non-empty set of cardinal numbers forms a chain when it is
ordered in this way. The fact that for any two cardinal numbers
one is less than or equal to the other is usually called the compara-
bility theorem for cardinal numbers.
Let X and Y be non-empty sets, and show that the cardinal number
of X is less than or equal to the cardinal number of Y + there exists
@ mapping of Y onto X.
Let {X;} be any infinite class of countable sets indexed by the ele-
ments 7 of an index set 7, and show that the cardinal number of
UX; is less than or equal to the cardinal number of J. (Hint: if I
is only countably infinite, this follows from Problem 6-2, and if J is
uncountable, Zorn’s lemma can be applied to represent it as the
union of a disjoint class of countably infinite subsets.)CHAPTER TWO
Metric Spaces
Classical analysis can be described as that part of mathematics
which begins with calculus and, in essentially the same spirit, develops
similar subject matter much further in many directions. It is a great
nation in the world of mathematics, with many provinces, a few of which
are ordinary and partial differential equations, infinite series (especially
power series and Fourier series), and analytic functions of a complex
variable. Each of these has experienced enormous growth over a long
history, and each is rich enough in content to merit a lifetime of study.
In the course of its development, classical analysis became so complex
and varied that even an expert could find his way around in it only with
difficulty. Under these circumstances, some mathematicians became
interested in trying to uncover the fundamental principles on which all
analysis rests. This movement had associated with it many of the great
names in mathematics of the last century: Riemann, Weierstrass, Cantor,
Lebesgue, Hilbert, Riesz, and others. It played a large part in the rise to
prominence of topology, modern algebra, and the theory of measure and
integration; and when these new ideas began to percolate back through
classical analysis, the brew which resulted was modern analysis.
As modern analysis developed in the hands of its creators, many a
major theorem was given a simpler proof in a more general setting, in an
effort to lay bare its inner meaning. Much thought was devoted to
analyzing the texture of the real and complex number systems, which are
the context of analysis. It was hoped—and these hopes were well founded
—that analysis could be clarified and simplified, and that stripping away
"950 Topology
superfluous underbrush would give new emphasis to what really mattered
from the point of view of the underlying theory.!
Analysis is primarily concerned with limit processes and con-
tinuity, so it is not surprising that mathematicians thinking along these
lines soon found themselves studying (and generalizing) two elementary
concepts: that of a convergent sequence of real or complex numbers, and
that of a continuous function of a real or complex variable.
We remind the reader of the definitions. First, a sequence
of real numbers is said to be convergent if there exists a real number x
(called the limit of the sequence) such that, given e > 0, a positive
integer no can be found with the property that
{an} = {21,22 te -
n> m= |t.— 2] 0 there exists 5 > 0 such
that
win X and |z — | < = |f(z) — fla) <6,
and f is said to be continuous if it is continuous at each point of X.
When X is an interval, this definition gives precise expression to the
intuitive requirement that f have a graph without breaks or gaps. The
corresponding definitions for sequences of complex numbers and complex
functions of a complex variable are word for word the same.
Our purpose in giving these definitions in detail here is a simple one.
We wish to point out explicitly that each is dependent for its meaning
on the concept of the absolute value of the difference between two real or
complex numbers. We wish to observe also that this absolute value is
the distance between the numbers when they are regarded as points on the
real line or in the complex plane.
In many branches of mathematics—in geometry as well as analysis—
it has been found extremely convenient to have available a notion of
distance which is applicable to the elements of abstract sets. A metric
space (as we define it below) is nothing more than a non-empty set
1 We illustrate these points in Appendix 1, where one of the basic existence
theorems in the theory of differential equations is given a brief and uncluttered proof
which depends only on the ideas of this chapter.Metric Spaces 51
equipped with a concept of distance which is suitable for the treatment
of convergent sequences in the set and continuous functions defined on
the set. Our purpose in this chapter is to develop in a systematic manner
the main elementary facts about metric spaces. These facts are impor-
tant for their own sake, and also for the sake of the motivation they
provide for our later work on topological spaces.
9, THE DEFINITION AND SOME EXAMPLES
Let X be a non-empty set. A metric on X is a real function d of
ordered pairs of elements of X which satisfies the following three
conditions:
(1) d(y) 2 0, and d(@z,y) = Oa = y;
(2) d(zy) = d(y,2) (symmetry);
(3) d(x,y) < d(z,z) + d(z,y) (the triangle inequality).
The function d assigns to each pair (2,y) of elements of X a non-negative
real number d(z,y), which by symmetry does not depend on the order
of the elements; d(z,y) is called the distance between z and y. A metric
space consists of two objects: a non-empty set X and a metric d on X.
The elements of X are called the points of the metric space (X,d). When-
ever it can be done without causing confusion, we denote the metric
space (X,d) by the symbol X which is used for the underlying set of
points. One should always keep in mind, however, that a metric space
is not merely a non-empty set: it is a non-empty set together with a
metric. It often happens that several different metrics can be defined
on a single given non-empty set, and in this case distinct metrics make the
set into distinct metric spaces.
There are many different kinds of metric spaces, some of which
play very significant roles in geometry and analysis. Our first example
is rather trivial, but it is often useful in showing that certain statements
we might wish to make are not true. It also shows that every non-
empty set can be regarded as a metric space.
Example t. Let X be an arbitrary non-empty set, and define d by
0 ife=y,
aay) = 1 ife xy.
The reader can easily see for himself that this definition yields a metric
on X.
Our next two examples are the fundamental number systems of
mathematics.52 Topology
Example 2. Consider the real line R and the real function |z| defined
on R. Three elementary properties of this absolute value function are
important for our purposes:
@) |o| > 0, and |2| = 02 = 0;
Gi) |=21 = Ia\;
ii) le + yl < lel + byl.
We now define a metric on R by
(zy) = |x — 9}.
This is called the usual metric on R, and the real line, as a metric space, is
always understood to have this as its metric. The fact that d actually
is a metric follows from the three properties stated above. This is a
piece of reasoning which occurs frequently in our work, so we give the
details. By (i), d(z,y) = |z — y| is a non-negative real number which
equals02—y=Oerxr=y. By (ii),
d(zy) = |z — yl = |-& — 2) = ly — 2] = dy,2).
And by (ii),
day) = |e -yl = |@-2+e@-Misle-a+le-yl
= d(zz) + d(z,y).
Example 3. Consider the complex plane C. We mentioned C briefly in
Sec. 4, and we described the sense in which it can be identified as a set with
the coordinate plane R?. We nowgive a somewhat fuller discussion. If z
is a complex number, and if z = a + ib where a and 6 are real numbers,
then a and b are called the real part and the imaginary part of z and are
denoted by R(z) and I(z). Two complex numbers are said to be equal
if their real and imaginary parts are equal:
a+i=c+id@a=candb=d.
We add (or subtract) two complex numbers by adding (or subtracting)
their real and imaginary parts, and we multiply them by multiplying
them out as in elementary algebra and replacing 7? by —1 wherever it
appears:
(a+) + (C+ id) = @ tc) +i +d),
and (a + ib)(c + id) = ac + iad + tbe + bd
= (ace — bd) + i(ad + be).Metric Spaces 53
Division is carried out in accordance with
ati _ @+ tb)(c— td) _ (ac + bd) + tbe ~ ad)
e+id (+ td)(c — id) oe + d?
= wtbd , be ~ ad
“ea eta
where c* + d? is required to be non-zero. If z = a + tb is a complex
number, then its negative —z and its conjugate 2 are defined by
~2= (-a) + i(-2)
and = a+ %(—b), which are usually written more informally as
—z= —a-— ibandz =a -— ib. It is easy to see that
ene,
2
Re = 242 and I(2)
The real line F is usually regarded as part of the complex plane:
R= {z:I(z) = 0} = {2:2 =z}.
Simple calculations show directly that
atasatm, a=", and F=2,
The origin, or zero, is the complex number 0 = 0 + 40. The ordinary
distance from z = a + 4b to the origin is defined by
le] = (ot +O,
le| is called the absolute value of z, and it is easy to see that
and el? = 22.
The usual metric on C is defined by
(21,22) = |zr — aa.
Exactly as in Example 2, the fact that this is a metric is a consequence
of the following properties of the real function |z|:
@) le] > 0, and [ze] = Oz = 0;
Gi) |-2| = lal;
Gii) ler + 2a] < las] + [eal
Properties (i) and (ii) are obvious. Since (—1)z, property (ii)
is also a special case of the fact that
lerzel = lal leal,54 Topology
which we prove by means of
levee]? = eveszaza = zatazaza = |eil*leel* = (leil [2el)*.
If we use the fact that |R(z)| < |e| for any z, property (iii) follows directly
from
la t+ al? = (@+a)G@iFa)=@Gtalit+e) _
eta + ante + arte + taza
= leal® + leal® + (exe + 2122)
= lel? + les? + 2220)
S fal? + leak? + laseal_
= |ail? + leal? + 2lail lel
= lea? + les|® + 2lzi] feal
= (lel + leal)®.
Whenever the complex plane C is mentioned as a metric space, its
metric is always assumed to be the usual metric defined above.
The remaining examples to be given in this section fit a common
pattern, which we have tried to exhibit in our discussion of Examples 2
and 3. We now point out several major features of this pattern, so that
the reader can see clearly how it applies in the slightly more complicated
examples that follow.
I, The elements of each space can be added and subtracted in a
natural way, and every element has a negative. Each space
contains a special element, denoted by 0 and called the origin,
or zero element.
II. In each space there is defined a notion of the distance from an
arbitrary element to the origin, that is, a notion of the “size”
of an arbitrary element. The size of an element z is a real
number denoted below by ||z|| and called its norm. Our use
of the double vertical bars is intended to emphasize that the
norm is a generalization of the absolute value functions in
Examples 2 and 3, in the sense that it satisfies the following
three conditions: (i) |lz|| 20, and |x| =O02=0; (ii)
all = [ells Gii) lle + yll < Mell + Hull
III. Finally, each metric arises as the norm of the difference between
two elements: d(z,y) = ||z — yl]. As in Example 2, the fact
that this is a metric follows from the properties of the norm
listed in II. This metric is called the metric induced by the
norm.
The knowledgeable reader will see at once that we are describing here
(though incompletely and imprecisely) the concept of a normed linear
space. Most of the metric spaces of major importance in analysis are
of this type.Metric Spaces 55
Example 4. Let f be a real function defined on the closed unit.interval
[0,1]. We say that f is a bounded function if there is a real number K
such that |f(x)| < K for every z€[0,1]. This concept is familiar to
the reader from elementary analysis, as is that of the continuity of f as
defined in the introduction to this chapter. The underlying set of points
in this example is the set of all bounded continuous real functions defined
on the closed unit interval. Actually, the boundedness of such a function
is a consequence of its other properties, but at this stage we assume it
explicitly. If f and g are two such functions, we add and subtract
them, and form negatives, pointwise:
f+ 9)(@) = f@) + (2);
GF - @) = f@) ~— 9@);
(-A@) = —f@).
The origin (denoted by 0) is the constant function which is identically
zero;
O(z) =0
for all x ¢ [0,1]. We define the norm of a function f by
Il = fo Wel as,
and the induced metric by
a(f.0) = If — oll = fo Ise2) — o(@)l de.
The integral involved in this definition is the Riemann integral of ele-
mentary calculus. Properties (i) and (ii) of the norm are easy to prove,
and (iii) follows from
toll = fy Wie) + orl de < f (seal + Iota) ax
= fp elas + ft lole)| de
= lal + loll
Example 5. The set of points in the preceding example—that is, the
set of all bounded continuous real functions defined on the closed unit
interval—has another metric which is far more important for our pur-
poses. It is defined by means of
Ill = sup {|f(@)]:2 € [0,1]},
which we usually write more briefly as
fll = sup [f@)1,
and d(fg) = If — all = sup [f(z) — 9).56 Topology
Properties (i) and (ii) of the norm are obvious, and in Problem 5 we ask
the reader to prove (iii) in a slightly more general form. This example
is typical of a large class of metric spaces which will play a major role
in all our work throughout the rest of this book. We denote this space
by e(0,1).
So much for the present for specific examples. We now turn to
several fundamental principles relating to metric spaces in general.
Let X be a metric space with metric d. Let Y be an arbitrary non-
empty subset of X. If the function d is considered to be defined only
for points in Y, then (Y,d) is evidently itself a metric space. Y, with
d restricted in this way, is called a subspace of X. This technique of
forming subspaces of a given metric space enables us to obtain an infinity
of further examples from the handful described above. For instance,
the closed unit interval [0,1] is a subspace of the real line, as is the set
consisting of all the rational points; and the unit circle, the closed unit
disc, and the open unit disc are subspaces of the complex plane. Also,
the real line itself is a subspace of the complex plane.
It is desirable at this stage to introduce the extended real number
system, by which we mean the ordinary real number system FR with
the symbols
-o and +o
adjoined. An extended real number is thus a real number or one of these
symbols. We say (by definition) that
-o <4;
also, if z is any real number, then
—ao itisa
union of open spheres.
Proor. We assume first that G is open, and we show that it is a union
of open spheres. If G is empty, it is the union of the empty class of
open spheres. If G@ is non-empty, then since it is open, each of its
points is the center of an open sphere contained in it, and it is the union
of all the open spheres contained in it.
We now assume that G is the union of a class S of open spheres. We
must show that G is open. If S$ is empty, then G is also empty, and by
Theorem A, G is open. Suppose that S is non-empty. G is also non-
empty. Let xz bea point inG. Since G is the union of the open spheres
in S, x belongs to an open sphere S,(zo) in S. By Theorem B, z is the
center of an open sphere S,,(z) C 8,(z0). Since S,(0) € G, S,,(z) G Gand
we have an open sphere centered on z and contained in G. G is there-
fore open.
The fundamental properties of the open sets in a metric space are
those stated in
Theorem D. Let X be a metric space. Then (1) any union of open sets
in X is open; and (2) any finite intersection of open sets in X is open.
proor. To prove (1), let {G:} be an arbitrary class of open sets in X.
We must show that G = U,G; is open. If {G;} is empty, then @ is62 Topology
empty, and by Theorem A, Gis open. Suppose that {G;} is non-empty.
By Theorem C, each G; (being an open set) is a union of open spheres; G
is the union of all the open spheres which arise in this way; and by another
application of Theorem C, @ is open.
To prove (2), let {Gs} be @ finite class of open sets in X. We must
show that G = OG; is open. If {G;} is empty, then @ = X; and by
Theorem A, G is open. Suppose that {G,} is non-empty and that
{Gs} = (Gi, Gs, . . . , Gn} for some positive integer n. If @ happens
to be empty, then it is open by Theorem A, so we may assume that G is
non-empty. Let z bea point in G. Since z is in each Gi, and each G; is
open, for each ¢ there is a positive real number r; such that S,,(z) € Gi.
Let r be the smallest number in the set. {r:, 72, . . . , 7a}. This number
ris a positive real number such that S,(z) € S,,(z) for each ¢, so S,(z) ©
G; for each ¢, and therefore S.(z) © G. Since S,(z) is an open sphere
centered on a and contained in G, G@ is open.
The above theorem says that the class of all open sets in a metric
space is closed under the formation of arbitrary unions and finite inter-
sections. The reader should clearly understand that Theorem A is an
immediate consequence of this statement, since the empty set is the
union of the empty class of open sets and the full space is its intersection.
The limitation to finite intersections in this theorem is essential. To see
this, it suffices to consider the following sequence of open intervals on the
real line:
(-1,D, (—34,4), (-14,%), « - -
The intersection of these open sets is the set {0} consisting of the single
point 0, and this set is not open.
In an arbitrary metric space, the structure of the open sets can be
very complicated indeed. Theorem C contains the best information
available in the general case: each open set is a union of open spheres.
In the case of the real line, however, a description can be given of the
open sets which is fairly explicit and reasonably satisfying to the intuition.
Theorem E. Every non-empty open set on the real line is the union of a
countable disjoint class of open intervals.
proor. Let G be a non-empty open subset of the real line. Let 2 bea
point of G. Since G is open, z is the center of a bounded open interval
contained in G. Define J, to be the union of all the open intervals
which contain z and are contained in G. The following three facts are
easily proved: J, is an open interval (by Theorem D and Problem 9-6)
which contains z and is contained inG; Jz contains every open interval
which contains z and is contained in G; and if y is another point in Js,Metric Spaces 63
then I, = Iy. We next observe that if z and y are any two distinct
points of G, then J, and J, are either disjoint or identical; for if they
have a common point z, then J. = I, and I, = I,, 80 Jz = Iy. Consider
the class I of all distinct sets of the form J. for points z in G. This is a
disjoint class of open intervals, and G is obviously its union. It remains
to be proved that | is countable. Let G, be the set of rational points in
G. G, is clearly non-empty. We define a mapping f of G, onto | as
follows: for each r in G,, let f(r) be that unique interval in | which con-
tains r. G, is countable by Problem 6-7, and the fact that | is countable
follows from Problem 6-8.
A firm grasp of the ideas involved in the theory of metric spaces
depends on one’s capacity to ‘“‘see” these spaces with the mind’s eye.
The complex plane is perhaps the
best metric space to use as a model
from which to absorb this necessary
intuitive understanding. When we
consider an unspecified set A of com-
plex numbers, we usually imagine
it as a region bounded by a curve,
as in Fig. 19. We think of the point
a, which is completely surrounded
by points of A, as being “inside”
the set A, or in its “interior,” while
zz ison the “boundary” of A. More pig. 19. A sot A of complex numbers
precisely, z: is the center of some with interior point z; and boundary
open sphere contained in A, and point zs.
each open sphere centered on zz in-
tersects both A and its complement A’. We formulate these ideas for a
general metric space in the next paragraph and at the end of the next
section.
Let X be an arbitrary metric space, and let A be a subset of X. A
point in A is called an interior point of A if it is the center of some open
sphere contained in A; and the interior of A, denoted by Int(A), is the set
of all its interior points. Symbolically,
Int(A) = {z:2€ A and S,(z) C A for some r}.
The basic properties of interiors are the following:
(1) Int(A) is an open subset of A which contains every open subset
of A (this is often expressed by saying that the interior of A is
the largest open subset of A);
(2) A is open = A = Int(A);
(3) Int(A) equals the union of all open subsets of A.64
Topology
The proofs of these facts are quite easy, and we ask the reader to fill in
the details as an exercise (see Problem 8).
Problems
1
10.
Let X be a metric space, and show that any two distinct points of X
can be separated by open spheres in the following sense: if x and y
are distinct points in X, then there exists a disjoint pair of open
spheres each of which is centered on one of the points.
Let X be a metric space. If {2} is a subset of X consisting of a
single point, show that its complement {z}’ is open. More gen-
erally, show that A’ is open if A is any finite subset of X.
Let X be a metric space and S,(z) the open sphere in X with center
zand radius r, Let A be a subset of X with diameter less than r
which intersects S,(z). Prove that A C S:,(z).
Let X be a metric space. Show that every subset of X is open
each subset of X which consists of a single point is open.
Let X be a metric space with metric d, and let d, be the metric
defined in Problem 9-1. Show that the two metric spaces (X,d)
and (X,d,) have precisely the same open sets. (Hint: show that
they have the same open spheres with one exception. What is this
exception?)
If X =X. X X: X ++: X X, is the product in Problem 9-4, and
if d and d are the metrics on X defined in that problem, show that
the two metric spaces (X,d) and (X,d) have precisely the same
open sets. Observe that in this case the spaces do not have the
same open spheres.
Let Y be a subspace of a metric space X, and let A be a subset of the
metric space Y. Show that A is open as a subset of Y © it is the
intersection with Y of a set which is open in X.
Prove the statements made in the text about interiors.
Describe the interior of each of the following subsets of the real
line: the set of all integers; the set of all rationals; the set of all
irrationals; (0,1); [0,1]; (0,1) U {1,2}. Do the same for each of the
following subsets of the complex plane: {z:|z| < 1}; {z:lz] < 1};
{z:I(z2) = 0}; {2:R@) is rational}.
Let A and B be two subsets of a metric space X, and prove the
following:
(a) Int(A) U Int(B) C Int(A U B);
(6) Int(A) M Int(B) = Int(A A B).
Give an example of two subsets A and B of the real line such that
Int(A) U Int(B) # Int(A U B).Metric Spaces 65
11. CLOSED SETS
Let X be a metric space with metric d. If A is a subset of X, a
point z in X is called a limit point of A if each open sphere centered on x
contains at least one point of A different from x. The essential idea here
is that the points of A different from z get “arbitrarily close” to 2, or
“pile up” at x.
The subset {1, 4, }4, . . .} of the real line has 0 as a limit point;
in fact, 0 is its only limit point. The closed-open interval 0,1) has 0 as a
limit point which is in the set and 1 as a limit point which is not in the
set; further, every real number z such that 0 < x < 1 is also a limit
point of this set. The set of all integral points on the real line has no
limit points at all, whereas every real number is a limit point of the set of
all rationals. In Example 9-1, every open sphere of radius less than 1
consists only of its center, so no subset of this space has any limit points.
A subset F of the metric space X is called a closed set if it contains
each of its limit points. In rough terms, a set is closed if its points do not
get arbitrarily close to any point outside of it. Among the subsets of the
real line mentioned in the preceding paragraph, only the set of integral
points is closed. In Example 9-1, every subset is closed.
Theorem A. In any metric space X, the empty set B and the full space X
are closed sets.
PRooF. The empty set has no limit points, so it contains them all and is
therefore closed. Since the full space X contains all points, it auto-
matically contains its own limit points and thus is closed.
The following theorem characterizes closed sets in terms of open sets.
We already know a good deal about open sets, so this characterization
provides us with a useful tool for establishing properties of closed sets.
Theorem B. Let X be a metric space. A subset F of X is closed + tts
complement F’ is open.
proor. Assume first that F is closed. We show that F’ is open. If
F’ is empty, it is open by Theorem 10-A, so we may suppose that F” is
non-empty. Let xbeapointin F’. Since F is closed and zis not in F, 2
is not a limit point of F. Since x is not in F and is not a limit point of F,
there exists an open sphere S,(z) which is disjoint from F. S,(x) is an
open sphere centered on z and contained in F”’, and since x was taken to
be any point of F’, F’ is open.
We now assume that ¥” is open and show that F is closed. The only
way F can fail to be closed is to have a limit point in F’. This cannot66 Topology
happen, for since F’ is open, each of its points is the center of an open
sphere disjoint from F, and no such point can be a limit point of F.
If zo is a point in our metric space X, and r is a non-negative real
number, the closed sphere S,{2o] with center x and radius r is the subset
of X defined by
S[zo] = {x:d(z,ao) r, 7: = d(z,20) — r is a positive real number. We take
r, as the radius of an open sphere S,,(z) centered on z, and we show that.
S,[x0]’ is open by showing that S,,(x) C S,[{xo]’. Let y be a point in
S,,(z), so that d(y,x) <1; On the basis of this and the fact that
A(xo,2) < d(xo,y) + d(y,x), we see that
d(y,a0) 2 d(x,r0) — dYy,x) > d(a,a0) — 1 = (a0) — (dear) — r] = 7,
so that y is in S,[x9]’.
The main general facts about closed sets are those given in our next
theorem.
Theorem D. Let X be a metric space. Then (1) any intersection of closed
sets in X is closed; and (2) any finite union of closed sets in X is closed.
prooF. By virtue of Eqs. 2-(2) and Theorem B above, this theorem is
an immediate consequence of Theorem 10-D. We prove (1) as follows.
If {F;} is an arbitrary class of closed subsets of X and F = Fi, then
by Theorem B, F is closed if F’ is open; but F’ = U;F'/ isopen by Theorem
10-D, since by Theorem B each F;’ is open. The second statement is
proved similarly.
In Theorem E of the previous section, we gave an explicit charac-
terization of the open sets on the real line. We now consider the struc-
ture of its closed sets. Among the simplest closed sets on the real line
are the closed intervals (which are the closed spheres) and finite unionsMetric Spaces 67
of closed intervals. Finite sets are included among these, since a set
consisting of a single point is a closed interval with equal end-points.
What is the character of the most general closed set on the real line?
Since closed sets are the complements of open sets, Theorem 10-E gives
a complete answer to this question: the most general proper closed subset.
of the real line is obtained by remov-
ing a countable disjoint class of open
intervals. This process sounds inno-
cent enough, but in fact it leads to
some rather curious and complicated
examples. One of these examples
is of particular importance. It was
studied by Cantor and is usually —.----——--—--- 2-0
called the Cantor set.
To construct the Cantor set, we
proceed as follows (see Fig. 20). es-es---ss-0.—--————ee-se---ee-s0
First, denote the closed unit interval Fig. 20. The Cantor set.
[0,1] by Fi. Next, delete from Fi
the open interval (34,3) which is its middle third, and denote the
remaining closed set by F2. Clearly,
Fr = (0,)4] U [34,1].
Next, delete from F, the open intervals (14,36) and (74,36), which are
the middle thirds of its two pieces, and denote the remaining closed set
by Fy. It is easy to see that
Fs = (0,36) U [96,24] U [36,76] U [86,1].
If we continue this process, at each stage deleting the open middle third
of each closed interval remaining from the previous stage, we obtain a
sequence of closed sets F’,, each of which contains all its successors. The
Cantor set F is defined by
F=ORi Fa,
and it is closed by Theorem D. F consists of those points in the closed
unit interval [0,1] which “ultimately remain” after the removal of all the
open intervals (14,24), (14,36), (14,86), - .. . What points do remain?
F clearly contains the end-points of the closed intervals which make up
each set Fn:
0, 1, 24, 36, 38, 3S, 26, 36, - - -
Does F contain any other points? We leave it to the reader to verify
that }4 isin F and is not an end-point. Actually, F contains a multitude
of points other than the above end-points, for the set of these end-points68 = Topology
is clearly countable, while the cardinal number of F itself is c, the cardinal
number of the continuum. To prove this, it suffices to exhibit a one-to-
one mapping f of [0,1) into F. We construct such a mapping as follows.
Let x be a point in (0,1), and let z = .bibeb; . . . be its binary expansion
(see Sec. 7). Eachb,iseitherOorl. Lett, = 2bn, and regard -titsts . . .
as the ternary expansion of a real number f(z) in (0,1). The reader will
easily convince himself that f(z) is in the Cantor set F: since ¢; is 0 or 2,
S(@) is not in (34,34); since t is 0 or 2, f(z) is not in (34,36) or [74,36); ete.
Also, it is easy to see that the mapping f: [0,1) > F is one-to-one.
According to this, F contains exactly as many points as the entire closed
unit interval [0,1]. It is interesting to compare this conclusion with the
fact that the sum of the lengths of all the open intervals removed is
precisely 1, since
M+R tS to a1
Tt is also interesting to observe (by doing a little arithmetic) that Fes is
the union of 16,777,216 disjoint closed intervals of the same length which
are rather irregularly distributed along [0,1]. These facts may suffice to
indicate that the Cantor set is a very intricate mathematical object and
is just the sort of thing mathematicians delight in. We shall encounter
this set again from time to time, for its properties illustrate several
phenomena discussed in later sections.
We conclude this section by defining two additional concepts which
are often useful.
Let X be an arbitrary metric space, and let A be a subset of X. The
closure of A, denoted by A, is the union of A and the set of all its limit
points. Intuitively, A is A itself together with all other points in X
which are arbitrarily close to A. As an example, if A is the open unit
disc {z:|z| <1} in the complex plane, then A is the closed unit disc
{z:|z| <1}. The main facts about closures are the following:
(1) A is a closed superset of A which is contained in every closed
superset of A (we express this by saying that A is the smallest
closed superset of A);
(2) A is closed + A = A;
(3) A equals the intersection of all closed supersets of A.
It is a routine exercise to prove these statements, and we leave this task
to the reader (in Problem 6).
Our second concept relates to the discussion of Fig. 19 given at the
end of the previous section. Again, let X be a metric space and A a
subset of X. A point in X is called a boundary point of A if each open
sphere centered on the point intersects both A and A’, and the boundary
of A is the set of all its boundary points. This concept possesses the
following properties;Metric Spaces 69
(1) the boundary of A equals A 1 A’;
(2) the boundary of A is a closed set;
(8) A is closed + it contains its boundary.
We ask the reader to give the proofs in Problem 11.
Problems
1,
NO
Let X be a metric space, and extend Problem 10-1 by proving the
following statements:
(a) any point and disjoint closed set in X can be separated by open
sets, in the sense that if z is a point and F a closed set which
does not contain z, then there exists a disjoint pair of open
sets G, and G, such that x¢G, and F C G,;
(b) any disjoint pair of closed sets in X can be separated by open
sets, in the sense that if F; and F; are disjoint closed sets, then
there exists a disjoint pair of open sets G; and G; such that
F, CG, and F, CG.
Let X be a metric space, and let A be a subset of X. If zis a limit
point of A, show that each open sphere centered on z contains an
infinite number of distinct points of A. Use this result to show that
a finite subset of X is closed.
Show that a subset of a metric space is bounded © it is non-empty
and is contained in some closed sphere.
Give an example of an infinite class of closed sets whose union is not,
closed. Give an example of a set which (a) is both open and closed;
(6) is neither open nor closed; (c) contains a point which is not a
limit point of the set; and (d) contains no point which is not a limit
point of the set.
Describe the interior of the Cantor set.
Prove the statements made in the text about closures.
Let X be a metric space and A a subset of X. Prove the following
facts:
(a) A’ = Int(A);
(0) A = {x:d(a,A) = 0}.
Describe the closure of each of the following subsets of the real line:
the integers; the rationals; the Cantor set; (0,+ ©); (—1,0) U (0,1).
Do the same for each of the following subsets of the complex plane:
{z:|z| is rational}; {z:1/R(z) is an integer}; {z:|z| <1 and
I(z) <0}.
Let X be a metric space, let z be a point of X, and let r be a positive
real number. One is inclined to believe that the closure of S,(z)
must equal S,{z]. Give an example to show that this is not neces-
sarily true. (Hint: see Example 9-1.)70 Topology
10. Let X be a metric space, and let G be an open set. in X. Prove that
G is disjoint from a set A & G is disjoint from A.
11, Prove the facts about boundaries stated in the text.
12. Describe the boundary of each of the following subsets of the real
line: the integers; the rationals; [0,1]; (0,1). Do the same for each
of the following subsets of the complex plane: {z:|z| < 1}; {z:|z| <
1}; {e:Z(z) > O}.
13. Let X be a metric space and A a subset of X. A is said to be dense
(or everywhere dense) if A = X. Prove that A is dense = the only
closed superset of A is X < the only open set disjoint from A is
$+ A intersects every non-empty open set + A intersects every
open sphere.
12, CONVERGENCE, COMPLETENESS, AND BAIRE’S THEOREM
As we emphasized in the introduction to this chapter, one of our
main aims in considering metric spaces is to study convergent sequences
in @ context more general than that of classical analysis. The fruits of
this study are many, and among them is the added insight gained into
ordinary convergence as it is used in analysis.
Let X be a metric space with metric d, and let
{aa} = {ay B20) Tay so}
be a sequence of points in X. We say that {z,} is convergent if there
exists a point x in X such that either
(1) for each «> 0, there exists a positive integer no such that
n > 1 = d(xn,2) < €; or equivalently,
(2) for each open sphere S,(z) centered on 2, there exists a positive
integer mo» such that 2, is in S.(x) for all n > mo.
The reader should observe that the first condition is a direct generaliza-
tion of convergence for sequences of numbers as defined in the introduc-
tion, and that the second can be thought of as saying that each open
sphere centered on z contains all points of the sequence from some place
on. If we rely on our knowledge of what is meant by a convergent
sequence of real numbers, the statement that {z,} is convergent can
equally well be defined as follows: there exists a point z in X such that
d(z,,2) > 0. We usually symbolize this by writing
In 2,
and we express it verbally by saying that xz, approaches x, or that ta
converges tox. It is easily seen from condition (2) and Problem 10-1 that
the point x in this discussion is unique, that is, that 2, > y with y #Metric Spaces 7”
is impossible. The point z is called the limit of the sequence {z,}, and
we sometimes write tz, — 2 in the form
lim tq = 2.
The statements az and limz,=2
mean exactly the same thing, namely, that {z,} is a convergent sequence
with limit z.
Every convergent sequence {z,} has the following property: for each
«> 0, there exists a positive integer m» such that m,n > 1p => d(tm,tn) <
« For if ,— 2, then there exists a positive integer mo such that n >
1 = d(tn,z) < ¢/2, and from this we see that
mn > 9 => d(tmtn) < d(m,z) + d(a,tn) < ¢/2 + ¢/2 = 6
A sequence with this property is called a Cauchy sequence, and we have
just shown that every convergent sequence is a Cauchy sequence.
Loosely speaking, this amounts to the statement that if the terms of a
sequence approach a limit, then they get close to one another. It is of
basic importance to understand that the converse of this need not be true,
that is, that a Cauchy sequence is not necessarily convergent. As an
example, consider the subspace X = (0,1] of the real line. The sequence
defined by z, = 1/n is easily seen to be a Cauchy sequence in this space,
but it is not convergent, since the point 0 (which it wants to converge to)
is not a point of the space. The difficulty which arises in this example
stems from the fact that the notion of a convergent sequence is not
intrinsic to the sequence itself, but also depends on the structure of the
space in which it lies. A convergent sequence is not convergent “on its
own”; it must converge to some point in the space. Some writers
emphasize the distinction between convergent sequences and Cauchy
sequences by calling the latter ‘intrinsically convergent” sequences.
A complete metric space is a metric space in which every Cauchy
sequence is convergent. In rough terms, a metric space is complete if
every sequence in it which tries to converge is successful, in the sense that,
it finds a point in the space to converge to. The space (0,1] mentioned
above is not complete, but it evidently can be made so by adjoining the
point 0 to it to form the slightly larger space [0,1]. As a matter of fact,
any metric space, if it isn’t already complete, can be made so by suitably
adjoining additional points. We outline this process in a problem at the
end of Sec. 14.
It is a fundamental fact of elementary analysis that the real line is a
complete metric space. The complex plane is also complete, as we see
from the following argument. Let {z,}, where 2, = Gn + tba, be a
Cauchy sequence of complex numbers. Then {a,} and {b,} are them-72 Topology
selves Cauchy sequences of real numbers, since
[dm — Gn] < [2m — 2nl
and lbm — Bal < zm — 2a].
By the completeness of the real line, there exist real numbers a and b such
that a, a and b,—b. If we now put 2 = a + ib, then we see that
2x — 2 by means of
len — 2] = [(@n + thn) — (a + 10)|
= |(a, — a) + i(b, — b)|
S lan — al + |b, — |
and the fact that both final terms on the right approach 0. The com-
pleteness of the complex plane thus depends directly on the completeness
of the realline. The metric space defined in Example 9-1 is also complete;
for in this space a Cauchy sequence must be constant (i.e., it must consist,
of a single point repeated) from some place on, and it converges with that
point as its limit.
The first three of the five metric spaces given as examples in Sec. 9
are therefore complete. What about the last two?
We ask the reader to show in Problem 5 that Example 9-4 is not.
complete. The problem of completing this space leads to the modern
theory of Lebesgue integration, and it would carry us too far afield to
pursue this matter to its natural conclusion.
On the other hand, the space €[0,1] defined in Example 9-5 is com-
plete. We prove this in a more general form in Sec. 14. The complete-
ness of this space, and of others similar to it, is one of the major focal
points of topology and modern analysis.
The terms limit and limit point are often a source of confusion for
people not thoroughly accustomed to them. On the real line, for
instance, the constant sequence {1, 1,..., 1, .. .} is convergent
with limit 1; but the set of points of this sequence is the set consisting of
the single element 1, and by Problem 11-2, the point 1 is not a limit point
of this set. The essence of the matter is that a sequence of points in a set
is not a subset of the set: it is a function defined on the positive integers
with values in the set, and is usually specified by listing its values, as in
{an} = {a1 t ... 2m... -}, where z, is the value of the function at
the integer n. A sequence may have a limit, but cannot have a limit
point; and the set of points of a sequence may have a limit point, but
cannot have a limit. The following theorem relates these concepts to
one another and is a useful tool for some of our later work.
Theorem A. If a convergent sequence in a metric space has infinitely many
distinct points, then tts limit is a limit point of the set of points of the sequence.Metric Spaces 73
proor. Let X be a metric space, and let {z,} be a convergent sequence
in X with limit z. We assume that « is not a limit point of the set of
points of the sequence, and we show that it follows from this that the
sequence has only finitely many distinct points. Our assumption
implies that there exists an open sphere S,(x) centered on z which con-
tains no point of the sequence different fromz. However, since z is the
limit of the sequence, all x,’s from some place on must lie in S,(x), hence
must coincide with z. From this we see that there are only finitely many
distinct points in the sequence.
Our next theorem guarantees the completeness of many metric
spaces which arise as subspaces of complete metric spaces.
Theorem B. Let X be a complete metric space, and let ¥ be a subspace of
X. Then Y is complete = tt ts closed.
PROOF. We assume first that Y is complete as a subspace of X, and we
show that it is closed. Let y be a limit point of Y. For each positive
integer n, Sij.(y) contains a point y, in Y. It is clear that {y,} con-
verges to y in X and is a Cauchy sequence in Y, and since Y is complete,
yisin Y. Y is therefore closed.
We now assume that Y is closed, and we show that it is complete.
Let {y,} be a Cauchy sequence in Y. It is also a Cauchy sequence in X,
and since X is complete, {y.} converges to a point zin X. We show that.
awisin Y. If {yn} has only finitely many distinct points, then z is that
point infinitely repeated and is thusin Y. On the other hand, if {y,} has
infinitely many distinct points, then, by Theorem A, z is a limit point of
the set of points of the sequence; it is therefore also a limit point of Y, and
since Y is closed, z is in Y.
A sequence {A,} of subsets of a metric space is called a decreasing
sequence if
4,.2D4:24;2°°°.-
The following theorem gives conditions under which the intersection of
such a sequence is non-empty.
Theorem C (Cantor's Intersection Theorem). Let X be a complete metric
space, and let {F,,} be a decreasing sequence of non-empty closed subsets of X
such that d(F,) > 0. Then F = (\%_, F, contains exactly one point.
pRooF. It is first of all evident from the assumption d(F,) — 0 that F
cannot contain more than one point, so it suffices to show that F is non-
empty. Let 2, be a point in F,. Since d(F,) 0, {z,} is a Cauchy
sequence. Since X is complete, {z,} has a limit x. We show that z isin
F, and for this it suffices to show that z is in F,, for a fixed but arbitrary
no If {an} has only finitely many distinct points, then z is that point74 Topology
infinitely repeated, and is therefore in F,,. If {za} has infinitely many
distinct points, then z is a limit point of the set of points of the sequence,
it is a limit point of the subset {z,: 7 > no} of the set of points of the
sequence, it is a limit point of F,,, and thus (since F,, is closed) it is in Fy,.
A subset A of a metric space is said to be nowhere dense if its closure
has empty interior. It is easy to see that A is nowhere dense <= A does
not contain any non-empty open set < each non-empty open set has a
non-empty open subset disjoint from A + each non-empty open set has a
non-empty open subset disjoint from A + each non-empty open set con-
tains an open sphere disjoint from A. Ifa nowhere dense set is thought of
as a set which doesn’t cover very much of the space, then our next
theorem says that a complete metric space cannot be covered by any
sequence of such sets.
Theorem D. If {Aq} is a sequence of nowhere dense sets in a complete metric
space X, then there exists a point in X which is not in any of the A,’s.
proor. For the duration of this proof, we abandon our usual notations
for open spheres and closed spheres. Since X is open and A, is nowhere
dense, there is an open sphere S, of radius less than 1 which is disjoint
from A;. Let F, be the concentric closed sphere whose radius is one-half
that of §;, and consider its interior. Since A is nowhere dense, Int(F;)
contains an open sphere S; of radius less than }4 which is disjoint from A:.
Let Fz be the concentric closed sphere whose radius is one-half that of S2,
and consider its interior. Since Az is nowhere dense, Int(F2) contains an
open sphere S; of radius less than 14 which is disjoint from A;. Let F3 be
the concentric closed sphere whose radius is one-half that of S;. Con-
tinuing in this way, we get a decreasing sequence {F,} of non-empty
closed subsets of X such that d(F,) > 0. Since X is complete, Theorem
C guarantees that there exists a point z in X which is in all the F,’s.
This point is clearly in all the S,’s, and therefore (since S, is disjoint from
A,) it is not in any of the A,’s.
For our purposes, the following equivalent form of Theorem D is
often more convenient.
Theorem E. If a complete metric space is the union of a sequence of its
subsets, then the closure of at least one set in the sequence must have non-
empty interior.
Theorems D and E are really one theorem expressed in two different
ways. We refer to both (or either) as Baire’s theorem. This theorem is
admittedly rather technical in nature, and the reader can hardly be
expected to appreciate its significance at the present stage of our work.