Real and Functional Analysis - Serge Lang
Real and Functional Analysis - Serge Lang
Real and Functional Analysis - Serge Lang
Editorial Board
S. Axler F.W. Gehring P.R. Halmos
Real and
Functional Analysis
Third Edition
With 37 Illustrations
, Springer
Serge Lang
Department of Mathematics
Yale University
New Haven, CT 06520
USA
Editorial Board
S. Axler F.W. Gehring P.R. Halmos
Department of Department of Department of
Mathematics Mathematics Mathematics
Michigan State University University of Michigan Santa Clara University
Bast Lansing, MI 48824 Ann Arbor, MI 48109 Santa Clara, CA 95053
USA USA USA
The previous edition was published as Real Analysis. Copyright 1983 by Addison-Wesley.
Production coordinated by Brian Howe and managed by Terry Komak; manufacturing supervised by
Vincent Scelta.
Typeset by Aseo Trade Typesetting Ltd., North Point, Hong Kong.
9 8 7 6 5 4 3 2
SPIN 10545036
Foreword
This book is meant as a text for a first year graduate course in analysis.
Any standard course in undergraduate analysis will constitute sufficient
preparation for its understanding, for instance, my Undergraduate Anal-
ysis. I assume that the reader is acquainted with notions of uniform con-
vergence and the like.
In this third edition, I have reorganized the book by covering inte-
gration before functional analysis. Such a rearrangement fits the way
courses are taught in all the places I know of. I have added a number of
examples and exercises, as well as some material about integration on the
real line (e.g. on Dirac sequence approximation and on Fourier analysis),
and some material on functional analysis (e.g. the theory of the Gelfand
transform in Chapter XVI). These upgrade previous exercises to sections
in the text.
In a sense, the subject matter covers the same topics as elementary
calculus, viz. linear algebra, differentiation and integration. This time,
however, these subjects are treated in a manner suitable for the training
of professionals, i.e. people who will use the tools in further investiga-
tions, be it in mathematics, or physics, or what have you.
In the first part, we begin with point set topology, essential for all
analysis, and we cover the most important results.
I am selective here, since this part is regarded as a tool, especially
Chapters I and II. Many results are easy, and are less essential than
those in the text. They have been given in exercises, which are designed
to acquire facility in routine techniques and to give flexibility for those
who want to cover some of them at greater length. The point set topol-
ogy simply deals with the basic notions of continuity, open and closed
sets, connectedness, compactness, and continuous functions. The chapter
vi FOREWORD
both the linear algebra and integration theory in the study of such
operators. One may view the treatment of spectral measures as providing
an example of general integration theory on locally compact spaces,
whereby a measure is obtained from a functional on the space of contin-
uous functions with compact support.
I find it appropriate to introduce students to differentiable manifolds
during this first year graduate analysis course, not only because these
objects are of interest to differential geometers or differential topologists,
but because global analysis on manifolds has come into its own, both in
its integral and differential aspects. It is therefore desirable to integrate
manifolds in analysis courses, and I have done this in the last part, which
may also be viewed as providing a good application of integration theory.
A number of examples are given in the text but many interesting
examples are also given in the exercises (for instance, explicit formulas for
approximations whose existence one knows abstractly by the Weierstrass-
Stone theorem; integral operators of various kinds; etc). The exercises
should be viewed as an integral part of the book. Note that Chapters
XIX and XX, giving the spectral measure, can be viewed as providing
an example for many notions which have been discussed previously:
operators in Hilbert space, measures, and convolutions. At the same
time, these results lead directly into the real analysis of the working
mathematician.
As usual, I have avoided as far as possible building long chains of
logical interdependence, and have made chapters as logically independent
as possible, so that courses which run rapidly through certain chapters,
omitting some material, can cover later chapters without being logically
inconvenienced.
The present book can be used for a two-semester course, omitting
some material. I hope I have given a suitable overview of the basic tools
of analysis. There might be some reason to include other topics, such as
the basic theorems concerning elliptic operators. I have omitted this
topic and some others, partly because the appendices to my SL 2 (R}
constitutes a sub-book which contains these topics, and partly because
there is no time to cover them in the basic one year course addressed to
graduate students.
The present book can also be used as a reference for basic analysis,
since it offers the reader the opportunity to select various topics without
reading the entire book. The subject matter is organized so that it makes
the topics availab1e to as wide an audience as possible.
There are many very good books in intermediate analysis, and inter-
esting research papers, which can be read immediately after the present
course. A partial list is given in the Bibliography. In fact, the determina-
tion of the material included in this Real and Functional Analysis has
been greatly motivated by the existence of these papers and books, and
by the need to provide the necessary background for them.
viii FOREWORD
Finally, I thank all those people who have made valuable comments
and corrections, especially Keith Conrad, Martin Mohlenkamp, Takesi
Yamanaka, and Stephen Chiappari, who reviewed the book for Springer-
Verlag.
PART ONE
General Topology
CHAPTER I
Sets ... .. . ..... ..... .. ......... . . .. .......... . ...... . ...... . ... .. 3
§1. Some Basic Terminology ......... .. ............................. 3
§2. Denumerable Sets .............................................. 7
§3. Zorn's Lemma ...... . .. . .. . .. . ........ . .... . . ............ .. .... 10
CHAPTER II
Topological Spaces 17
§1. Open and Closed Sets .......... ... ........................ .... . 17
§2. Connected Sets ................... . .......................... . . 27
§3. Compact Spaces ............................................... 31
§4. Separation by Continuous Functions .... .... ......... . .. .... .. ... 40
§5. Exercises .. ..... . ... . ........ .. .. ... ........ .. ............. . ... 43
CHAPTER III
Continuous Functions on Compact Sets 51
§1. The Stone-Weierstrass Theorem ...................... . ...... . ... 51
§2. Ideals of Continuous Functions .................................. 55
§3. Ascoli's Theorem . . .............. . .......... . .................. 57
§4. Exercises ........ .. .......... . ... . ..................... . ... .. .. 59
x CONTENTS
PART TWO
Banach and Hilbert Spaces 63
CHAPTER IV
Banach Spaces 65
§1. Definitions, the Dual Space, and the Hahn-Banach Theorem 65
§2. Banach Algebras . .. ....................... ... ....... .. ....... . 72
§3. The Linear Extension Theorem ................................. . 75
§4. Completion of a Normed Vector Space ... . ...................... . 76
§5. Spaces with Operators ....................... . ................ . 81
Appendix: Convex Sets .............. . ......................... . 83
1. The Krein-Milman Theorem ............ .... ........ .. ...... . 83
2. Mazur's Theorem ...................................... .. .. . 88
§6. Exercises 91
CHAPTER V
Hilbert Space 95
PART THREE
Integration 109
CHAPTER VI
The General Integral 111
§1. Measured Spaces, Measurable Maps, and Positive Measures ....... 112
§2. The Integral of Step Maps .. .... . ....... .. .. . ......... ... . . .. . . 126
§3. The L1-Completion .......... . ............................... . 128
§4. Properties of the Integral: First Part ...... . ..................... 134
§5. Properties of the Integral: Second Part ...................... . .. . 137
§6. Approximations . ............ .. .............................. . 147
§7. Extension of Positive Measures from Algebras to (I-Algebras ....... 153
§8. Product Measures and Integration on a Product Space ........... . 158
§9. The Lebesgue Integral in RP ................ ... ................ 166
§10. Exercises .......................................... . ..... . ... 172
CHAPTER VII
Duality and Representation Theorems 181
CHAPTER VIII
Some Applications of Integration 223
§1. Convolution ...... . .. .. ..... ... .. .. .. . .. .. .. ... . .. ... . . . . . .. .. 223
§2. Continuity and Differentiation Under the Integral Sign .... .. .. ... .. 225
§3. Dirac Sequences . .. . . .. . . . .... . . . .. . . . ...... . .... .. . .. . ........ 227
§4. The Schwartz Space and Fourier Transform . . . . . .. ...... . . .. . . . .. 236
§5. The Fourier Inversion Formula . . ... . . .. .. .. . . . .. .... . .. . .... . . . 241
§6. The Poisson Summation Formula .. .. .... ....... ..... ... .. . . .. .. 243
§7. An Example of Fourier Transform Not in the Schwartz Space . . .. . .. 244
§8. Exercises . . .. . .. . . . . .... .. .. . ....... .. .... ... ......... . . . . ... . 247
CHAPTER IX
Integration and Measures on Locally Compact Spaces 251
§1. Positive and Bounded Functionals on CAX) .. . .. .. . . . . . ... .. .. . . . 252
§2. Positive Functionals as Integrals ... . ... . .. .. ... . ...... . .. . . ... . . . 255
§3. Regular Positive Measures . ... .. . . .. . .. . .. . .. . . . .. . .... .. . .. .... 265
§4. Bounded Functionals as Integrals . . ..... . . .. . .. .... .. . . .. .... .... 267
§5. Localization of a Measure and of the Integral . . . ...... . . .. . ... ... . 269
§6. Product Measures on Locally Compact Spaces .. .. . . .. . .. . . . . . .. .. 272
§7. Exercises . . .... . ...... . ...... . . . . .. . .. .. . . . . .. . . . ... . . . . . . .... 274
CHAPTER X
Riemann-Stieltjes Integral and Measure 278
§1. Functions of Bounded Variation and the Stieltjes Integral . ... .. .... 278
§2. Applications to Fourier Analysis . . . .. .. . .. . . . . . . . ... . .... .. .. . .. . 287
§3. Exercises . .. .. ... .. . .. .. ... . .. . ... .. . ..... . .. . . .. ... .... ..... . 294
CHAPTER XI
Distributions 295
§1. Definition and Examples . .. . .... . .. ...... . ... . . ......... . .. . .. . 295
§2. Support and Localization . .. ..... ... . ... ..... .... . . .. . . . ... .... . 299
§3. Derivation of Distributions . . . . .. . . ... . .... . . . ... ... . . ......... . 303
§4. Distributions with Discrete Support 304
CHAPTER XII
Integration on Locally Compact Groups 308
§1. Topological Groups .. . ... . . ... . . ...... . ... . .. .... .. . .. . .. .. . . . . 308
§2. The Haar Integral, Uniqueness .. . .... . .. . . .. .... . . .... ....... . . . 313
§3. Existence of the Haar Integral . . .... . .. . ............ . .. . . .... .. . . 319
§4. Measures on Factor Groups and Homogeneous Spaces .. .. ..... ... . 322
§5. Exercises . ... . ... .. .. .. . . . .. .. .. .. ........ . . .. . . . .. . ... .. . . .. . 326
PART FOUR
Calculus . .. . . ..... .. . . ... . . ... . ...... . . . . .. . . . .. . .. .. . . . .. . .... . 329
xii CONTENTS
CHAPTER XIII
Differential Calculus 331
§l. Integration in One Variable . . . .. . .... .. .. . . . ... . . . .. . .... . . . .. 331
§2. The Derivative as a Linear Map .. . . . .. .. . ... . .. . .. .. ... . . .. . . . 333
§3. Properties of the Derivative . . ..... . ... . . . .. . ..... . ..... . . . . .. . 335
§4. Mean Value Theorem . . . . .. . .. . .. .. . . ... . . . . . . .. .. .. ... .. . . . . 340
§5. The Second Derivative . . ...... .. . .. .. . . .. .... . .. . . ... . ... .. . . . 343
§6. Higher Derivatives and Taylor's Formula .... . .. . .... . . . .. . . .. . . 346
§7. Partial Derivatives . .. . . . . . . . .... . . .. .... . ... . . . .. ..... . .. . . . . 351
§8. Differentiating Under the Integral Sign ... . . .. . . .. .. .. .. ... . . .. .. 355
§9. Differentiation of Sequences ... . . . . .. . . . .. . .. ...... . .. ..... ... . 356
§10. Exercises . .. .. . . .. .. . . . . .. . . ... .. . . . . . .. . . .. . . . ... . . . . .. . . ... 357
CHAPTER XIV
Inverse Mappings and Differential Equations 360
§l. The Inverse Mapping Theorem . . . ... . . . .. . ......... .. .. . . .. . .. . 360
§2. The Implicit Mapping Theorem . ... . .. ...... . . . . . . .... .. .. . ... . . 364
§3. Existence Theorem for Differential Equations . . . .. . ... .. . .. . .. . ... 365
§4. Local Dependence on Initial Conditions . ... .. ..... .. ... . .. . .. ... 371
§5. Global Smoothness of the Flow . . ... .. . . . .. . .... .. . .. . .. . ....... 376
§6. Exercises . ...... . ... . .... . .. . .. . ... . . .. . ... ........ . . . .. .. .. . . 379
PART FIVE
Functional Analysis 385
CHAPTER XV
The Open Mapping Theorem, Factor Spaces, and Duality 387
§l. The Open Mapping Theorem .. .. . ... .. . .. . . . ... . . .. . . . . . .. .. . .. 387
§2. Orthogonality . . . . ... ... ..... .. ... .. .. .. . .. .. .. .. .. . ...... . . .. 391
§3. Applications of the Open Mapping Theorem .. .. . ... . . . . . . . . . .. .. 395
CHAPTER XVI
The Spectrum 400
§1. The Gelfand-Mazur Theorem .. . .. . . . . . . . .. .. .. .... . . .. . . ..... . 400
§2. The Gelfand Transform .. ... .. . . .. .. ... . .... .. . . ... .. . . ... . . . .. 407
§3. C*-Algebras . . . .. ....... . .. . . .. .. . . .. ...... .... . ... . . . . .. . . . .. 409
§4. Exercises . . .. . ..... .. . . .. . ..... . ..... . .. . . . .. . ... .. ......... . . 412
CHAPTER XVII
Compact and Fredholm Operators 415
§1. Compact Operators ... . ..... .. . . . . .. .... . .. . . . . ... .. . . .. . .. .. . 415
§2. Fredholm Operators and the Index ...... ... . . . . . . . . . ... . . . .. . ... .. . 417
§3. Spectral Theorem for Compact Operators . . .. . . . . .... . ... . .. . ... . 426
§4. Application to Integral Equations .... .. .. ... . .. .. ..... . . . .. . . ... 432
§5. Exercises .. .. ... . ... . . . . . . .. . . . . .. .. . . .. . . . ... . . .... . . .. . ... . . 433
CONTENTS xiii
CHAPTER XVIII
Spectral Theorem for Bounded Hermitian Operators 438
§l. Hermitian and Unitary Operators ............................... 438
§2. Positive Hermitian Operators . . ............ .. ................... 439
§3. The Spectral Theorem for Compact Hermitian Operators .......... 442
§4. The Spectral Theorem for Hermitian Operators ................... 444
§5. Orthogonal Projections .... . ... . ... ...... .. .................... 449
§6. Schur's Lemma .... . .. ... ......... . ........................... 452
§7. Polar Decomposition of Endomorphisms ........................ . 453
§8. The Morse-Palais Lemma . ... . . .. . .. .. ..... ...... ..... ........ 455
§9. Exercises ..................................................... 458
CHAPTER XIX
Further Spectral Theorems 464
§l. Projection Functions of Operators ........................ ..... . 464
§2. Self-Adjoint Operators . ... ... ........ . . . .... . ........ ... .... .. . 469
§3. Example: The Laplace Operator in the Plane .. . .................. 476
CHAPTER XX
Spectral Measures 480
§l. Definition of the Spectral Measure ...... ..... ... ... . ............ 480
§2. Uniqueness of the Spectral Measure:
the Titchmarsh-Kodaira Formula ... .... . . .......... ... .. . . . .... 485
§3. Unbounded Functions of Operators ............................. 488
§4. Spectral Families of Projections ... ... ... ... .. ... . .... .. ......... 490
§5. The Spectral Integral as Stieltjes Integral .. .. . .. .. ..... .. . ... . .... 491
§6. Exercises . .................................................... 492
PART SIX
Global Analysis 495
CHAPTER XXI
Local Integration of DiHerential Forms 497
§l. Sets of Measure 0 .. . ............ ... ......... . ..... . . . . .. ...... 497
§2. Change of Variables Formula . ......... .. ........ ... ......... ... 498
§3. Differential Forms ... . ................... . ...... .. ............. 507
§4. Inverse Image of a Form ... .... ... .. . . ..... ... ....... ... . . .. ... 512
§5. Appendix ......................... . .......................... 516
CHAPTER XXII
Manifolds ....................................................... 523
§l. Atlases, Charts, Morphisms ... .. ..... ... .......... ...... .... . ... 523
§2. Submanifolds ........... .. ....... .. ...... . . ..... ... . . . ........ 527
§3. Tangent Spaces ............ .. . . .... . . .. . . .... . ................ 533
§4. Partitions of Unity . . ... ....... ... .. ... .. ... .. ... .. .. ... ... .... 536
§5. Manifolds with Boundary ... ... .. . . . . . ... . .. . . ... . ... . .. ....... 539
§6. Vector Fields and Global Differential Equations .................. 543
XIV CONTENTS
CHAPTER XXIII
Integration and Measures on Manifolds 547
§l. Differential Forms on Manifolds ................ . ......... ... .... 547
§2. Orientation ................................................... 551
§3. The Measure Associated with a Differential Form . ................ 553
§4. Stokes' Theorem for a Rectangular Simplex . . .. . .. . ............... 555
§5. Stokes' Theorem on a Manifold .. . .............................. 558
§6. Stokes' Theorem with Singularities .. . ........................ . . .. 561
General Topology
CHAPTER
Sets
We assume that the reader understands the meaning of the word "set",
and in this chapter, summarize briefly the basic properties of sets and
operations between sets. We denote the empty set by 0. A subset S' of
S is said to be proper if S' =1= S. We write S' c S or S => S' to denote the
fact that S' is a subset of S.
Let S, T be sets. A mapping or map f : T --+ S is an association which
to each element x E T associates an element of S, denoted by f(x), and
called the value of f at x, or the image of x under f. If T' is a subset of
T, we denote by f(T') the subset of S consisting of all elements f(x) for
x E T'. The association of f(x) to x is denoted by the special arrow
X 1--+ f(x).
We usually reserve the word function for a mapping whose values are in
the real or complex numbers. The characteristic function of a subset S' of
S is the function X such that X(x) = 1 if XES' and X(x) = 0 if x ¢ S'. We
often write Xs' for this function.
Let X, Y be sets. A map f : X --+ Y is said to be injective if for all x,
x ' E X with x =1= x' we have f(x) =1= f(x'). We say that f is surjective if
f(X) = Y, i.e. if the image of f is all of Y. We say that f is bijective if it
is both injective and surjective. As usual, one should index a map f by
its set of arrival and set of departure to have absolutely correct notation,
but this is too clumsy, and the context is supposed to make it clear what
these sets are. For instance, let R denote the real numbers, and R' the
4 SETS [I, §l]
ff : R-+R'
flS
and stands for the map f : Z+ -+ R such that f(i) = Xi . As before, note
that a sequence can have all its elements equal to each other, that is
{l , l , l, ... }
which to each i E I associates a set Si' The sets Si mayor may not have
elements in common, and it is conceivable that they may all be equal.
As before, we write the family {SJieI '
We can define the intersection and union of families of sets, just as for
the intersection and union of a finite number of sets. Thus, if {SJieI is a
family of sets, we define the intersection of this family to be the set
consisting of all elements x which lie in all Si' We define the union
USi
ieI
00
nSi
i=l
nSi
iel
(X u Y) x Z = (X x Z) u (Y x Z).
(X U Y) x Z c (X x Z) U (Y x Z) .
(X n Y) x Z = (X x Z) n (Y x Z).
~x(Y V Z) = ~x Y n ~xZ,
In particular,
Thus the operation 1-1 commutes with all set theoretic operations.
We also use the word sequence for mappings of the natural numbers
into a set, thus allowing our sequences to start from 0 instead of 1. If we
8 SETS [I, §2]
need to specify whether a sequence starts with the O-th term or the first
term, we write
or
Proof. For each YES, there exists an element Xy E D such that f(xy) =
Y because f is surjective. The association y ~ Xy is an injective mapping
of S into D, because if y, Z E Sand Xy = x z , then
D2 : {X2l ' X 22 , X 23 , . • ·· }
f(i,j) = xij
ORO 1. We have x ~ x.
ORO 2. If x ~ y and y ~ z then x ~ z.
ORO 3. If x ~ y and y ~ x then x = y.
b ~ f(b) ~ b,
a < f(a) < j2(a) < ... < f"(a) < ....
f(x) ~ f(c).
f(x) ~ c ~ f(c).
f(x) = f(c) E E
by what has already been proved, and so f(x) ~ b. This proves that
bEE, that E is admissible, and thus proves Lemma 3.3.
x ~ f(x) ~ y,
Proof. Suppose that A does not have a maximal element. Then for
each x E A there exists an element Yx E A such that x < Yx' Let f: A ~ A
be the map such that f(x) = Yx for all x E A. Then A, f satisfy the hypoth-
eses of Theorem 3.1 and applying Theorem 3.1 yields a contradiction.
The only difference between Corollary 3.4 and Zorn's lemma is that in
Corollary 3.4, we assume that a non-empty totally ordered subset has a
least upper bound, rather than an upper bound. It is, however, a simple
16 SETS [I, §3]
Z= UX
iE I
i.
Topological Spaces
Similarly, one defines the notion of normed vector space bver the
complex numbers. The axioms are the same, except that we then take
the number e to be complex in NVS 2.
By an open ball B in E centered at a point v, and of radius r > 0, we
mean the set of all x E E such that Ix - vi < r. We denote such a ball by
Br(v). We define a set U to be open in E if for each point r E U there
exists an open ball B centered at x and contained in U. Again it is easy
to verify that this defines a topology, also called the ordinary topology of
the normed vector space. It is but an exercise to verify that an open ball
is indeed an open set of this topology.
Let {x.} be a sequence in a normed vector space E. This sequence is
said to be Cauchy if given c (always assumed > 0) there exists N such
that for all m, n ~ N we have
Ix - x.1 < c.
sup meaning least upper bound. It can be easily shown that the set of
bounded maps B(S, F) of S into F is a vector space, and that II II is a
norm on this space, called the sup norm.
L
For fEE define
balls just as we did in the case of normed vector spaces, and also define
a topology in a metric space just as we did for a normed vector space.
Every open set is then a union of open balls. This topology is said to be
determined by the metric.
In a normed vector space, we can define the distance between elements
x, y to be d(x, y) = Ix - YI. It is immediately verified that this is a metric
on the space. Conversely, the reader will see in Exercise 5 how a metric
space can be embedded naturally in a normed vector space, in a manner
preserving the metric, so that the "generality" of metric spaces is illusory.
For convenience, we also make here the following definition: If A, Bare
subsets of a normed vector space, we define their distance to be
Example. The reader will verify easily that two norms I 11 and I 12 on
a vector space E give rise to the same topology if and only if they satisfy
the following condition: There exist C 1 , C2 > 0 such that for all x E E we
have
Ix - vi ~ r.
We define the sphere centered at v, of radius r, to be the set of points x
such that
Ix - vi = r.
Warning. In some books, what we call a ball is called a sphere. This
is not good terminology, and the terminology used here is now essen-
tially universally adopted.
[II, §1] OPEN AND CLOSED SETS 21
The first condition is clear, and the other two come from the fact that
the complement of the union of subsets is equal to the intersection of
their complements, and that the complement of the intersection of subsets
is equal to the union of their complements.
Conversely, given a collection fF of subsets of a set X (not yet a
topological space), we say that it defines a topology on X by means of
22 TOPOLOGICAL SPACES [II, §1]
If f!8 satisfies these two conditions, then there exists a unique topology
whose open sets are the unions of sets in f!4. Indeed, such a topology is
uniquely determined, and it exists because we can define a set to be open
if it is a union of sets in f!4. The axioms for open sets are trivially
verified.
Example. The open balls in a normed vector space form a base for
the ordinary topology of that space.
of the sets f- 1 (W), where W is open in Y and f ranges over fF. Then
we leave to the reader the verification of the following facts:
2. This topology is the coarsest topology (the one with the fewest
open sets) such that every map f E fF is continuous.
is seen to be open.
As usual, we observe that a continuous image of an open set is not
necessarily open.
A continuous map f: X ~ Y which admits a continuous inverse map
g : Y ~ X is called a homeomorphism, or topological isomorphism. It is
clear that a composite of homeomorphisms is also a homeomorphism.
As usual, we observe that a continuous bijective map need not be a
homeomorphism. In fact, later in this course, we meet many examples
of. vector spaces with two different norms on them such that the identity
map is continuous but not bicontinuous.
Let {X;}ieI be a family of topological spaces and let
X = TI Xi
ieI
26 TOPOLOGICAL SPACES [II, §1]
Such sets have arbitrary open sets at a finite number of components, and
the full space at all other components.
The product topology is the unique topology with the fewest open sets
in X which makes each projection map
{/;: S ~ li};el
IIxll = maxlx;l
[II, §2] CONNECTED SETS 27
if x = (Xl' ... ,xn ) is given in terms of its coordinates. Then the topology
determined by this norm is clearly the same as the product topology.
Remark. A map f : X --+ Y which maps open sets onto open sets is
said to be open. A map which maps closed sets onto closed sets is said
to be closed. A continuous map need not be either. For instance, the
graph of the tangent is closed in the plane, but the projection map on
the x-axis maps it on an open interval:
Figure 2.1
The map which folds the plane over the real axis maps the open plane
on the closed half plane. If f : X --+ Y is continuous and bijective, then a
necessary and sufficient condition that f be a homeomorphism is that f
be open. This is simply a rephrasing of the continuity of the inverse
mapping f-l .
f: X --+ {p, q}
Proof Clear.
and (J(r(br) = y.
Of course, if such a path exists, then it is easy to define just one continu-
ous map
(J(: [a, b] --+ X
from some interval [a, b] into X such that (J((a) = x and (J((b) = y. One
can even take the interval [a, b] to be [0, 1].
Proof We give the proof for a closed interval J = [a, b] and leave the
other cases (open, half-open, infinite intervals) as exercises. Suppose that
we can write J = A u B where A, B are closed, disjoint, and non-empty.
Say that a E A. Let c be the greatest lower bound of B. Then c lies in
30 TOPOLOGICAL SPACES [II, §2]
is connected.
Let
z = {a.'1' ... , a·'n' (b') '4' . )
'1.,...11. ·· ·.'n
so that the coordinates of z are the same as those of a for i l , ..• ,in and
the same as those of b for the other indices. Then z E V and fez) = p.
[II, §3] COMPACT SPACES 31
g( x.'1 ) = (x '1'
. a·12"'"
a·In' (b')'4' .)
""""'1 •...• In •
Let X be a set and {S~}~EA a family of subsets. We say that this family
is a covering of X if its union is equal to X. If X is a topological space,
and {U~}~EA is a covering, we say it is an open covering if each U~ is
open. If {S~}~EA is a covering of X, we define a subcovering to be a
covering {Sp}PEB where B is a subset of A. In particular, a finite sub-
covering of {S~} is a covering {S~" ... ,S~J.
Let X be a topological space. We shall say that X is compact if any
open covering of X has a finite subcovering. As usual, we can express a
dual condition relative to closed sets. Let {F~laEA be a family of subsets
of X. We say that this family has the finite intersection property if any
finite intersection
is not empty.
is not empty.
32 TOPOLOGICAL SPACES [II, §3]
F.ell n···nF.an
Since V is disjoint from S, it follows that already Val' ... ,Va" cover S,
thus proving our assertion.
The converse of the preceding assertion is almost true but not quite.
A topological space X is said to be Hausdorff if given points x, Y E X
and x =1= y there exist disjoint open sets V , V such that x E V and Y E V
If X is Hausdorff, then each point of X is obviously closed.
Then the intersection Uy, n ... n UYn is open, contains x, and is contained
in the complement of S, thus proving what we want.
A topological space X is said to be normal if it is Hausdorff, and if
given two disjoint closed sets A, B in X there exist disjoint open sets U,
V such that A c U and B c V.
Proof. The proof is similar to the previous one, and involves merely
one further application of the same principle. Using the same trick as in
this previous proof, we know that for each x E A there exist disjoint open
sets Ux , w" such that x E Ux and Be w" w".. (One would take the finite
union of the open sets v", ...
... ,v'n to obtain w" in the analogous situa-
tion.) The family of open sets {Ux } X E A covers A, and there exists a finite
subcovering
UXl U"'uUXn .
t
ball C l of radius such that C l n S is not covered by a finite number of
Ui' We let Xl be a point of Cl n S. Suppose that we have obtained a
sequence of closed balls
such that Cn has radius 1/2n , with a point Xn E Cn n S, and such that
Cn n S is not covered by a finite number of Ui ' Since S itself can be
covered by a finite number of closed balls of radius 1/(2 n +1), it follows
that Cn n S can also be so covered, and hence there exists a closed ball
Cn + l of radius 1/(2n+1) and such that Cn+1 n S cannot be covered by a
finite number of Ui' We let Xn+1 be a point of Cn + l n S. This constructs
our sequence as desired. We see that {xn} is a Cauchy sequence in S,
which co verges to a point X in S. But x lies in some Uia which contains
Cn for all sufficiently large n, a contradiction which proves our theorem.
Proof. The image f(A) is compact, so closed and bounded. The least
upper bound of f(A) lies in f(A), thus proving our assertion.
We recall the proof briefly. Given £, for each x E A we let r(x) > 0 be
such that if Iy - xl < r(x), then If(y) - f(x) I < £. We can cover A by
open balls Bi of radius
bi = r(xJ/2,
is compact.
be the projection on the O(-th factor. For each 0(, the family of closed sets
{tra(F) }, FE !IF*,
element x" in each set n,,(F) for all FE ff'*. Let x = (x,,). We contend
that x belongs to all sets F E ff'* . This will prove our theorem.
To prove our contention, we observe that the intersection of a finite
number of sets in ff'* also lies in ff'* because of the maximality of ff'*.
Let V be an open set of X containing x, of the form
with each V". open in X". Then V" contains x". for all i, and therefore
V", contains ~ point of n,.;(F) for all FE ff'*. Hen'ce
belongs to ff'*, and hence the finite intersection of these sets for
i = 1, .. . ,n
also belongs to ff'* . But this finite intersection is nothing else but our
set V, and hence V intersects each F in ff'*, so a fortiori each FE ff'.
Hence x lies in the closure of each FE ff', whence x E F for all FE ff', as
was to be shown.
Proof Let I I be the sup norm, and I I any other norm. It will
suffice to prove that these two norms are equivalent. If e l , ... ,en are the
usual unit vectors of Rn, then for x = Xl e l + ... + xne n we get
[II, §3] COMPACT SPACES 39
with C = n · max IeJ This proves one of the desired inequalities, and also
shows that the other norm is continuous, because
Drawing a closed ball around x intersecting F, and using the fact that
the intersection of F and this ball is compact, we conclude that there is
some Z E F such that d(x, F) = Ix - zl, and we have x - z :f. 0 since F is
closed in E. Then there is some Xi such that
IIxx -- z I 1
zl- Xi < 2"
40 TOPOLOGICAL SPACES [II, §4]
d(x, F) = Ix - zl
°
B be disjoint closed subsets. Then there exists a continuous function f
on X with values in the interval [0, 1] such that f(A) = and f(B) = 1.
[II, §4] SEPARATION BY CONTINUOUS FUNCTIONS 41
Inductively, for each integer k with 0 ~ k ~ 2n, we find V k /2 " such that if
r < s, then Vr C Dr C Vs . We then define the function f by
f(x) = 1 if x E B,
f(x) = inf of all r such that x E Vr if x ~ B.
f- 1 [0, a) = U Vr
r<a
because f(x) < a if and only if x lies in some Ur with r < a. Similarly, we
have f(x) > b if and only if x ~ Dr for some r > b, so that
f- 1 (b, 1] = U l6'D
r>b
r •
V= v:
Xl
U"'uV:Xn .
Proof Assume first that f has values in [0, 1]. If A, B are disjoint
closed subsets of X, we denote by gA,B a function with values in [0, 1]
such that g(A) = 0 and g(B) = 1. Such a function exists by Theorem 4.2.
We then define
We have
and
[II, §5] EXERCISES 43
°
that by Urysohn's lemma there exists a continuous function h on X with
values in [0, 1] such that h is 1 on A and on B. Then hf* has values
in the open interval (-1, 1), as desired. This concludes the proof of
Theorem 4.4.
Remark 2. The theorem also holds in the complex case dealing sepa-
rately with the real and imaginary parts. The extra condition on the
restriction of the values can then be formulated analogously by requiring
that
11/*11 ~ 11111·
Indeed, suppose that we have extended 1 to a bounded continuous com-
plex valued function g. Let b = 11111. Let h be the function such that
h(z) = z if Izl ~ b, and h(z) = bz/ lzl if Izl > b. Then h is continuous,
IIhll ~ b, and hog fulfills our requirement.
Show that this is a norm on [1, and that [I is complete under this norm.
(b) Let f3 = {bn } be a fixed sequence in [I. Show that the set of all 0( E [I
such that lanl ~ Ibnl is compact. Show that the unit sphere in [I is not
compact.
4. Let 0( be a real number, 0 < 0( ~ 1. A real valued function f on [0, 1] is said
to satisfy a Holder condition of order 0( if there is a constant C such that for
all x, y we have
If(x) - f(y)1 ~ Clx - yl'·
(a) Show that the set of functions satisfying such a HOlder condition is a
vector space, and that II 11« is a norm on this space.
(b) Show that the set of functions f with Ilfll« ~ 1 is a compact subset of
C([O, 1]).
5. Metric spaces. (a) Let X be a metric space with distance function d. Define
d'(x, y) = min {l, d(x, y)}. Show that d' is a distance function, and that the
notion of convergence and limit with respect to d' is the same as with
respect to d.
(b) As in normed vector spaces, one can define Cauchy sequences, i.e. se-
quences {x n } such that given e, there exists N such that for all m, n ~ N
we have d(xn' x m ) < e. A metric space is called complete if every Cauchy
sequence converges. Show that if a metric space X as in part (a) is
complete with respect to d, then it is complete with respect to d'.
(c) For each x E X define the function fx on X by
x ....... d(x, A)
and
(d) Show that a sequence {f.} converges uniformly on compact sets if and
only if it converges in the above metric.
(e) Let K be a compact set and e > O. Given f, let V(f, K, e) be the set of
all maps 9 such that 111- gilK < e. Show that V(f, K, e) is open in the
topology defined by the metric. Show that the family of all such open
[II, §5] EXERCISES 47
sets for all choices of jj,, K, e is base for the topology. This proves that
the topology does not depend on the choice of exhaustive sequence {KJ.
(f) If E is complete, i.e.
i.e. aa Banach space, show that M(X, E) is complete in
the metric defined above.
(g) If X is locally compact, show that the space of continuous maps C(X, E)
is closed in M(X, E) for the metric.
to. Let U be the open unit disc in the plane. Show that there is an exhaustive
sequence of compact subsets of U.
U.
11. Let U be a connected open set in the plane (or in Euclidean space Rk). Show
that there is an exhaustive sequence of compact subsets of U.
12. Let U be an open subset of a normed vector space. Show that U is con-
nected if and only if U is arcwise connected.
13. The diagonal ~ in a product X x X is the set of all points (x, x).
(a) Show that a space X is Hausdorff if and only if the diagonal is closed in
X x X.
(b) Show that a product of Hausdorff spaces is Hausdorff.
Separable Spaces
15. A topological space having a countable base for its open sets is called separa-
ble. Show that a separable space has a countable dense subset.
ble.
16. (a) If X is a metric space and has a countable dense subset, then X is
separable.
(b) A compact metric space is separable.
17. (a) Every open covering of a separable space has a countable subcovering.
(b) A disjoint collection of open sets in a separable space is countable.
(c) A base for the open sets of a separable space contains a countable base.
19. Let X be separable. Show that the following conditions are equivalent:
(a) X is compact.
(b) Every sequence {xn} in X has at least one point of accumulation, that is
X is sequentially compact.
(c) Every decreasing sequence {An} of non-empty closed sets has a nonempty
intersection.
20. Prove that a normal separable space X is metrizable (Urysohn metrization
theorem). [Hint: Let {Un} be a countable base for the topology. Let (Un" Um)
be an enumeration of all pairs of elements in this base such that Vn, cUm,.
For each i let I; be a continuous function satisfying 0 ~ I; ~ 1 and such that
I; is 0 on Vn , and 1 on the complement of Um ,. Let
00 1
d(x, y) = i~ 2i 11;(x) - l;(y)I·]
Show that d is a metric and that the identity mapping is continuous with
respect to the given topology on X and the topology obtained from the
metric. You will use the fact that given x E X and some open set Um in the
base containing x, there exists another set Un in the base such that
f x I z = fz: X x Z --+ Y x Z
23. Let f: X ---> X' and g: X' ---> X" be continuous maps. Prove:
(a) If f and 9 are proper, so is 9 0 f.
(b) If 9 0 f is proper and f is surjective, then 9 is proper.
(c) If 9 0 f is proper and 9 is injective, then f is proper.
(d) If 9 0 f is proper and X'is Hausdorff, then f is proper.
24. Let X be a topological space, {p} a set consisting of one element p. The map
f: X ---> {p} is proper if and only if X is compact. [Hint: Assume that f is
proper. To show that X is compact, let {S.} be a family of non-empty closed
sets having the finite intersection property. Let Y = Xu {p}, where p is dis-
joint from X. Define a base for a topology of Y by letting a set be in this
base if it is of type S. u {p}, or if it is an arbitrary subset of X. Show this is
a base. The projection n: X x Y ---> Y is a closed map. Let D be the subset of
X x Y consisting of all pairs (x, x) with x E X. Then n(D) is closed and
therefore contains p. Hence there exists x E X such that (x, p) E D, whence
give an open U in X containing x, and any S., the set U x (S. u {p}) inter-
sects D, whence U intersects S., and x lies in n S•. ]
25. Let f: X ---> Y be a continuous map. Show that the following properties are
equivalent:
(a) f is proper.
(b) f is closed and for each y E Y the set f-1(y) is compact.
26. Let f: X ---> Y be proper. If B is a compact subset of Y, then f-1(B) is
compact.
27. (The marriage problem so baptized by Hermann Weyl.) Let B be a set of boys,
and assume that each boy b knows a finite set of girls Gb • The problem is to
marry each boy to a girl of his acquaintance, injectively. A necessary condi-
tion is that each set of n boys know collectively at least n girls. Prove that
this condition is sufficient. [Hint: First assume that B is finite, and use
induction. Let n > 1. If for all 1 ;;;; k < n each set of k boys knows > k girls,
marry off one boy and refer the others to the induction hypothesis. If for
some k with 1 ;;;; k < n there exists a subset of k boys knowing exactly k girls,
marry them by induction. The remaining n - k boys satisfy the induction
hypothesis with respect to the remaining girls (obvious!) and thus the case of
finite B is settled. For the infinite case, which is really the relevant problem
here, take the Cartesian product TI Gb over all b E B, each Gb being finite,
discrete, and use Tychonoff's theorem. For this elegant proof, cf. Halmos
and Vaughn, Amer. J. Math. January 1950, pp. 214-215.]
28. The Cantor set. Let K be the subset of [0, 1] consisting of all numbers
having a trecimal expansion
co a
L -3:'
n=l
where bm = a m /2. Show that f is well defined. Show that f is surjective and
continuous. One can then extend f to a continuous map of the interval onto
the square. This is called a Peano curve. Note that the interval has dimen-
sion 1 whereas its image under the continuous map f has dimension 2. This
caused quite a sensation at the end of the nineteenth century when it was
discovered by Peano.
30. The semi parallelogram law (Bruhat-Tits). Let X be a complete metric space.
We say that X satisfies the semi parallelogram law, or is seminegative, if
given two points Xl' X2 E X there is a point z such that for all X E X we have
Prove that under this law, d(z, xd = d(XI, x 2 )/2, and z is uniquely deter-
mined. We call z the midpoint of Xl, X2.
31. (Serre, after Bruhat-Tits) Let X be a seminegative complete metric space. Let
S be a bounded subset of X. Show that there exists a unique closed ball
B,(xd of minimal radius containing S. [Use the semiparallelogram law both
for uniqueness and existence. For existence, show that if {B,.(xn )} is a se-
quence of closed balls containing S with lim rn = r (the inf of all
ali radii of
closed balls containing S), then {xn} is Cauchy.] The center of that closed
ball is called the circumcenter of S.
32. (Bruhat-Tits fixed point theorem) Let X be a complete seminegative metric
space. Let G be a group of isometries of X, i.e. bijective maps f: X --+ X
which preserve distance. Denote the action of G by (g, x) I-> g . x. Suppose G
has a bounded orbit (i.e. there is a point X such that the set S of all elements
g . X, g E G, is bounded). Then G has a fixed point (the circumcenter) of the
orbit.
For the above exercises, cf. Bruhat-Tits, Groupes Reductifs sur un Corps
Local I, Pub. IRES 41 (1972) pp. 5-251; and K. Brown, Buildings, Springer
Verlag, 1989, Chapter VI, Theorem 2 of §5.
CHAPTER III
Continuous Functions
on Compact Sets
Let E be a normed vector space (over the real or the complex numbers).
We can define the notion of Cauchy sequence in E as we did for real
sequences, and also the notion of convergent sequence (having a limit). If
every Cauchy sequence converges, then E is said to be complete, and is
also called a Banach space. A closed subspace of a Banach space is
complete, hence it is also a Banach space.
Proof. We give the proof in three steps. First, we prove that given Xl'
X2 E S and Xl #- X2 ' and given real numbers 0(, p, there exists hE A such
that h(x l ) = 0( and h(X2) = p. By hypothesis, there exists qJ E A such that
qJ(X I) #- qJ(X2)' Let
f + g If - gl
max(J, g) = - 2- + - 2 - '
min(f ) = f + g _ If - gl
,g 2 2·
-c ~f(x) ~ c
for all XES. The absolute value function can be uniformly approximated
by ordinary polynomials on the interval [-c, c] by Exercises 6, 7, or 8,
which are very simple ad hoc proof. Given 6, let P be a polynomial such
that
Ip(t)-ltll<6
for - c ~ t ~ c. Then
then
i.e.
p(J(x») = anf(x)n + ... + ao·
This concludes the proof of the Stone-Weierstrass theorem.
J(X) = f(x).
Proof. Let A R be the set of all functions in A which are real valued.
We contend that AR is an algebra over R which satisfies the hypotheses
of the preceding theorem. It is obviously an algebra over R. If Xl "# x 2
are points of S, there exists f E A such that f(x l ) = 0 and f(x 2) = 1.
(The proof of the first step of Lemma 1.2 shows this.) Let g = f + 1.
Then g(x l ) = 0 and g(X2) = 2, and g is real valued, so AR separates
points. It obviously contains the real constants, and so the real S-W
theorem applies to it. Given a complex continuous function <p on S, we
write <p = u + iv, where u, v are real valued. Then u, v are continuous,
and u, v can be approximated uniformly by elements of A R, say f, g EAR
such that lIu - !II < e and ttv - gil < e. Then f + ig approximates
u + iv = <p, thereby concluding the proof.
The second theorem of this chapter deals with ideals of continuous func-
tions. Let S be a topological space, and R a ring of continuous functions
(real valued) pn S. An ideal J of R is a subset of R satisfying the
following properties: The zero function 0 is in J. If f, g E J, then f + g
and -fare in J, and if hER, then hf E J. The reader should really have
56 CONTINUOUS FUNCTIONS ON COMPACT SETS [III, §2]
9 = g2YI + .. . +g2Ym .
Then 9 is in J, is continuous, is nowhere 0 on S, and ~ O. Since 9 has a
minimum on S, there is a number a > 0 such that g(x) ~ a for all XES.
The function
ng
1 + ng
f~
1 + ng
and so fng/(l + ng) lies within 2e of f. Thus we have shown that flies
in the closure of J . Since J is assumed closed, we conclude that f lies in
J, thereby proving our theorem.
In the examples of Chapter XVIII, §4, we shall deal with compact subsets
of function spaces, and we need a criterion for compactness, which is
theorem. It is also used in other places in analysis,
provided by Ascoli's theorem.
for instance in a proof of the Riemann mapping theorem in complex
analysis. Therefore, we give a proof here in the general discussion of
compact spaces.
Let X be a subset of a metric space, and let F be a Banach space. Let
<1> be a subset of the space of continuous maps C(X, F). We shall say
that <1> is (or its elements are) equicontinuous at a point Xo E X if given e,
there exists D such that whenever x E X and d(x, x o ) < DD,, then
ASC 2. For each x E X, the set <1>(x) consisting of all values f(x) for
f E <1> is relatively compact.
58 CONTINUOUS FUNCTIONS ON COMPACT SETS [III, §3]
Proof Assume that <I> satisfies the two conditions. We shall prove
that <I> is relatively compact. For this it is sufficient to show that <I> can
be covered by a finite number of balls of prescribed radius (Corollary 3.9
of Chapter II). Let r > O. By equicontinuity, for each x E X we select an
open neighborhood V(x) such that if y E V(x), then If(y) - f(x)1 < r for
all f E <1>. Then a finite number V(x 1)' ... , V(x n ) cover X. Each set
<I> (X d, .. . <I>(X
, n)
is relatively compact, and hence so is their union
Y = <I>(X 1)
1) u ... U
U <I>(Xn ))..
Let B(a 1 ), ... ,B(am) be open balls of radius r centered at points a 1 , ... ,am
which cover Y. Then f(x d, ... ,J(xn ) lie in these balls. In fact, for each
i = 1, ... ,n we have
where u: {1, ... ,n} ~ {1, ... ,m} is some mapping. For each such map u
let <1>" be the set of f E
E <I> such that for all i, we have
Then the finite number of <1>" cover <1>. It suffices now to prove that each
<1>" has diameter < 4r. But if f, 9 E <1>" and x E X, then x lies in some
V(x;), and then:
If(x) - g(x)1 ~ If(x) - f(x;)1 + If(x;) - a,,;! + la"i - g(xi)1 + Ig(x;) - g(x)1
< 4r.
This proves our implication, and the part of Ascoli's theorem which
is used in the applications. The converse is trivial and left to the reader.
Corollary 3.2. Let X be a compact subset of a metric space, and let <I>
be a subset of the space of continuous functions on X with sup norm.
Then <I> is relatively compact if and only if <I> is equicontinuous and
bounded (for the sup norm, of course).
Remark. Since cI>cl> has a metric defined by the sup norm, as a rela-
tively compact set it has the property that any sequence has a convergent
subsequence, converging in its closure. Sometimes one deals with a lo-
cally compact set X which is a denumerable union of compact sets. In
that case, one obtains the following version of Ascoli's theorem.
Proof. We can find a sequence {V;} of open sets such that Vi C V;+1'
such that Vi is compact, and such that the union of the V; is X. For
each i, by the previous version of Ascoli's theorem, there exists a sub-
sequence which converges uniformly on Vi . The diagonal sequence with
respect to all i converges uniformly on every compact set. This proves
the corollary.
o. Let S be a subset of a normed vector space (or a metric space), and let {f.}
be a sequence of continuous maps of S into a Banach space F. Assume that
{f.} is a Cauchy sequence (for the sup norm). Show that {In} converges to a
continuous function f (for the sup norm). Show that BC(S, F) is closed in
B(S, F).
1. Let X be a compact set and let R be the ring of continuous (real valued)
functions on X. Let J, J' be closed ideals of R. Show that J c: J' if and only
if Z(J) => Z(J').
2. Let S be a closed subset of X. Let J be the set of all fER such that f
vanishes on S. Show that J is a closed ideal. Assume that X is Hausdorff.
Establish a ring-isomorphism between the factor ring RjJ and the ring of
continuous functions on S. (We assume that you have had the notion of a
factor ring in an algebra course.)
3. Let X be a compact space and let J be an ideal of C(X). If the set of zeros
of J is empty, show that J = C(X). (This result is valid in both the real and
the complex case.)
60 CONTINUOUS FUNCTIONS ON COMPACT SETS [III, §4]
r.
O~vt-Pn(t)~
20 r. '
2 + nvt
whence 0 ~ 0 - J>,,(t) ~ 2/n.
8. Look at Example 1 of Chapter VIII, §3 to see another explicit way of
proving Weierstrass' approximation theorem for a continuous function on a
finite closed interval. Do Exercise 1 of that chapter.
9. Let X be a compact set in a normed vector space, and let {in} be a sequence
of continuous functions converging pointwise to a continuous function i, and
such that {J.} is a monotone increasing sequence. Show that the convergence
is uniform (Dini's theorem ; cf. Chapter IX, §1).
10. Let X be a compact metric space (whence separable). Show that the Banach
space C(X, R) or C(X, C) of continuous functions on X is separable.
[Hint : Let {xn } be a countable dense set in X and let gn be the function on
X given by
(f ® g)(x, y) = f(x)g(y).
The two chapters of this part are absolutely basic for everything else that
follows, and introduce the most useful of all the spaces encountered in
analysis, namely Banach and Hilbert spaces. The reader who wishes to
study integration theory as soon as possible may continue these chapters
with Chapter VI, which will make essential use of the basic properties of
these spaces, especially the completion of a normed vector space and the
linear extension theorem. Indeed, the integral of the absolute value of a
function defines a seminorm on a suitable space of functions, whose com-
pletion will be the main object of study of the chapters on integration.
On the other hand, readers may look directly at the functional anal-
ysis, as a continuation of the linear theory of Banach and Hilbert spaces.
At some point, of course, these come together when we study the spectral
theorems and the existence of spectral measures.
As in the algebraic theory of vector spaces, we shall consider continu-
ous linear maps L: E -+ F of a normed vector space into another. The
kernel and image of L are defined as in the algebraic theory, namely the
kernel is the set of elements x E E such that L(x) = O. The image is
simply L(E). Both Ker L and 1m L are subspaces, of E and F respec-
tively. However, now that we have the norm, we note that the kernel is
a closed subspace (being the inverse image of the closed set {O}). Warn-
ing: the image if not necessarily closed. For conditions under which the
image is closed, see Chapter XV.
For the integration theory, we do not need such considerations of
subspace and factor space. However, we shall consider the dual space in
the context of integration, showing that various spaces of functions are
dual to each other. Thus we deal at somewhat greater length with the
dual space in this chapter. An application of the duality theory 10 the
context of Banach algebras will be given in Chapter XVI.
CHAPTER IV
Banach Spaces
Let E be a Banach space, i.e. a complete normed vector space. One can
deal with series L Xn in Banach spaces just as with series of numbers, or
of functions, and the most frequent test for convergence (in fact absolute
convergence) is the standard one:
the set of all x E E such that Ixl = 1. Then a bound for A is immediately
seen to be the same thing as a bound for the values of A on Sl' The
least upper bound of all values IA(x)l, for x E Sl' is called the norm of A,
and the map
is a norm on L(E, F). It is immediately seen that IAI is the greatest lower
bound of all numbers C > 0 such that
Let E, F, G be normed vector spaces, let u E L(E, F), and let v E L(F, G).
Then v 0 u is in L(E, G) and we have
Iv 0 ul ~ Ivllul.
and
cp A. (x, y) = A(X)(Y)·
Then CPA. is obviously bilinear, and we have
Then
68 BANACH SPACES [IV, §1]
so that by definition,
Hence
L(E, F) x E-F
from the space of repeated continuous linear maps to the space of con-
tinuous multilinear maps exactly as in the bilinear case. If F is complete,
then all these spaces are also complete.
We now consider a specially important space of linear maps.
The normed vector space L(E, R) [or L(E, C) in the complex case] is
called the dual space of E, and is denoted by E'. Elements of E' are
called functionals on E. Functionals can be used as substitutes for coor-
dinates. Indeed, suppose that E = R\ and let Ai be the i-th coordinate
function, that is
Then it is easily verified that {A1"" ,An} is a basis for the dual space of
Rk. Furthermore, the values of A1 , ••• ,An on an element X E Rk character-
ize this element. Although we do not have such convenient bases in the
infinite dimensional case, we still have such a characterization of elements
of E in terms of the values of functionals. This is based on the following
theorem.
[IV, §1] DEFINITIONS, THE DUAL SPACE 69
IA(Y) + al ~ Iy + vi
for all Y E F, or equivalently that for all Y E F,
Corollary 1.2. Let E be a normed vector space, and vEE, v "# O. Then
there exists a functional A on E such that A(V) "# O.
IA(X)I ~ IAlixl
shows that Ifxl ~ Ixi. We leave to the reader the opposite inequality
Ixi ~ Ifxl, which concludes the proof that we have an isometric em-
bedding of E in E".
[IV, §1] DEFINITIONS, THE DUAL SPACE 71
Theorem 1.4 (Alaoglu's Theorem). Let E be a Banach space, and let E'l
be the unit ball in the dual space E'. Then E'l is compact for the weak
topology.
K= n Kx
XEE
Ixl;§; 1
be the Cartesian product of all closed discs of radius 1, taken over all
x E E satisfying Ixl;£ 1. We give K the product topology, so that by
Tychonoff's theorem, K is compact. We map E'l into K by the map
Immediately from the definition, one sees that the map f is injective, and
thus gives an embedding of E'l into the product space. Furthermore, also
from the definition of the weak topology defined in Chapter II, §1, we
observe that the weak topology determined by the family ff is the same
as the weak topology determined by the family ffl of functionals fx with
x EEl (the closed unit ball in E), because any x E E, x # 0 is a scalar
multiple of a unit vector. More precisely, we also have an imbedding
f: E' c... nC
XEE
x given by AH n A{X),
XEE
E' ~ n ex
l l
XEE
E'1 ~ n Kx
xEE t
72 BANACH SPACES [IV, §2]
1 fTC
f * g(x) = 2n -TC f(t)g(x - t) dt.
tive, assocIatIve Banach algebra. Note that E does not have a unit
element. In this direction, see Chapter III, §1.
Theorem 2.1. Let A be a Banach algebra with unit element e. Then the
set of invertible elements is open in A. If v E A and Ivl < 1, then e + v
is invertible.
Proof. Let Ivl < 1. Then the series e + v + v2 + ... converges (abso-
lutely) and since
Then
Thus v has a right inverse and a left inverse, say v l , v2 , such that
and
A.:F~G
Hence F is a subspace of E.
The uniqueness of X is clear from continuity. We show its existence.
76 BANACH SPACES [IV, §4]
and Y = lim Yn
X(X + y) = lim A(Xn + Yn) = lim(Axn + AYn) = lim AXn + lim AYn
= Xx + Xy.
Similarly, X(cx) = cX(x). Hence X is linear, and since for x E F we have
x = lim x, it follows that Xx = AX if x E F. Thus X is an extension of A.
Finally, we have
it follows that
the method of Cauchy sequences. For another method, cf. Exercise 25.
We define a completion of E to be a pair (E, <p) consisting of a Banach
space E and a continuous linear map
which is injective, such that <p(E) is dense in E, and such that <p preserves
the norm, i.e. l<pxl = Ixl for all x E E. We shall now prove that such a pair
is essentially uniquely determined. In fact, if (F, t/I) is another completion,
then there exists a unique invertible element A E L(E, F) such that the
following diagram is commutative, in other words t/I = A 0 <po
~\ /~
E
is continuous and linear (it even preserves the norm) and consequently,
by the linear extension theorem, it has a unique continuous linear exten-
sion of E into F, which we denote by A. Similarly, the continuous linear
map
IAxl = Ixl
for all x E E. This again follows by continuity.
We shall now give two proofs of the existence of a completion. So let
E be a normed vector space and let E' be its dual. As we saw in
Proposition 1.3, we have a natural norm-preserving injection E -+ E".
But E" is complete because E" = L(E', F) with complete F (F = scalars).
So the completion of E is simply the closure E in E". (Do Exercise 15.)
Next we give another proof, based on the same construction as the
78 BANACH SPACES [IV, §4]
real numbers from the rational numbers. This construction will be used
in the integration theory. See the examples after the construction.
The Cauchy sequences of elements of E form a vector space, which we
denote by S, As usual, we have the notion of null sequences, that is
sequences {xn} in E such that given e, there exists N such that for all
n > N we have IXnl < e. The null sequences form a subspace. We define
e
two Cauchy sequences = {xn} and '1 = {Yn} to be equivalent if there
exists a null sequence IX = {an} such that e= '1 + IX (in other words
Xn = Yn + an for all n). This is an equivalence relation, and we denote
e
the equivalence class of by~. Then the equivalence classes of Cauchy
sequences form a vector space in a natural way, and we have (for C E R):
and
Then we define
n-->oo
be the map such that <p(x) is the class of the Cauchy sequence {x, x, .. . }.
Then it is clear that <p is linear, and preserves norms. Furthermore, one
sees at once that if ~ is the class of a Cauchy sequence e,
and x = {xn},
then
because <p(E) is dense in if. The sequence {xn} is then Cauchy (in E).
[IV, §4] COMPLETION OF A NORMED VECTOR SPACE 79
Indeed, we have
for n sufficiently large. This proves that E is complete, and concludes the
proof for the existence of a completion of E.
One can also take the vector space of continuous functions on R, van-
ishing outside some bounded interval, and define the U-norm similarly.
Then this space is not complete, and its completion is called L 1. It then
becomes a problem to identify elements of L 1 with certain functions, and
this is what we shall do.
Ixi ~O
j: E -+ E,
the only difference being that j has a kernel, which the reader will verify
to be precisely Eo. In fact, we have a norm on the factor space E/ Eo if
we define the norm of a coset Ix + Eol to be Ixl (independent of the coset
representative x since we have
Ix + yl = Ixl
for all y E Eo). Thus we can say that if E has a seminorm, the comple-
tion E is simply the completion of E/ Eo as discussed in this section.
A vector space E with a seminorm I I can be called a seminormed
space. We can define Cauchy sequences using the same definition as in
the normed case. We shall say that E is complete if every Cauchy
sequence in E converges-in other words, if given a Cauchy sequence
{x n } in E, there exists x E E such that given e, there exists N such that
for all n ~ N we have
IX n - xl < e.
Of course, the element x to which our sequence {x n } converges is not
uniquely determined, only up to an element of Eo . However, examples of
this situation arise in practice, in integration theory. One must then
distinguish between a complete seminormed space, and the completion of
E/ Eo mentioned above.
°
valued) on R, vanishing outside a compact set (i.e. infinitely differentiable
functions 1 such that I(t) = if t is outside some bounded interval). We
define the HO-norm on E by
f:
where
°
Proof If x E E and Bx = 0, then ABx = BAx = for all A E S, so the
kernel of B is S-invariant. Similarly, also from the relation ABx = BAx,
we see that the image of B is S-invariant.
where
dk = L crbs·
r+s=k
But
(cp)(A) = cp(A).
All these rules are useful when considering the evaluation of polynomials
on operators. In algebraic terminology, they express the fact that the
map
P 1--+ p(A)
where AI' ... ,An are elements of S, and the coefficients are numbers.
Indeed, if F is A- and B-invariant, then it is also (A + B)-invariant and
AB-invariant.
If an operator B commutes with all elements of S, then it is clear that
B also commutes with all elements in the ring of operators generated
by S, because if B commutes with Al and A 2 , then B commutes with
Al + A2 and also with Al A 2 . Furthermore, if F is a closed subspace
and is S-invariant, then it is also S-invariant, where S is the closure of
S. Indeed, if {Bn} is a sequence of operators in S converging to some
operator B, and if x E F, then the sequence {Bnx} is Cauchy, and hence
converges to Bx which lies in F.
In Chapters XVII and XVIII we study a pair (E, A) consisting of a
space E and an operator A, and analyze this pair, describing its structure
completely in important cases. The idea is to apply in the present con-
text an all-pervasive point of view in mathematics, which is to decompose
an object into a direct sum of simpler objects. In the present context, let
us make some general definitions.
[IV, App.] THE KREIN-MILMAN THEOREM 83
FxG-+E
given by
(y, z)~ y +z
is a toplinear isomorphism, then we say that E is the direct sum of the
subspaces F and G. Observe that our requirements involve both an
algebraic and a topological condition. It follows from our conditions
that
E=F+G and Fn G = {O}.
It will be proved later that, in fact, these two conditions are sufficient; in
other words, if they are satisfied, then the map
not only has an algebraic inverse, but this inverse is continuous (corol-
lary of the open mapping theorem). When E is a direct sum of F and G,
we write
E=F$G.
Although we shall not use the theorem of this section later in the book
(except for some exercises), it is worthwhile giving it since it is used
at the beginning of more advanced and specialized courses, in a wide
variety of contexts. The exposition follows that of Artin (cf. Collected
Works).
Throughout this section, we let E be a vector space over the reals (not
normed). We let E* be a vector space of linear maps of E into R (not
necessarily the space of all such linear maps), and assume that E* separates
E, that is given x E E, x =f; 0 there exists A. E E* such that A.(x) =f; O. We
84 BANACH SPACES [IV, App.]
give E the topology having the smallest amount of open sets making all
A E E* continuous. A base for this topology is therefore given by the
following sets : We take x E E, and ..1. 1 , • • • ,An E E*, and e > 0. We let B
be the set of all y E E such that
(1 - t)x + ty, ° ~ t ~ 1,
joining x to y is contained in S.
Lemma 1.1. Let Xl' .•. ,X.ES. Any convex set containing Xl ' .. . ,x.
also contains all linear combinations
°
with ~ ti ~ 1 for all i, and t 1 + ...
such linear combinations is convex.
+ tn = 1. Conversely, the set of all
The following properties of convex sets also follow at once from the
definitions.
Let A E E*, A '# 0, and let Ho be the kernel of A (i.e. the set of all x E E
such that A(x) = 0). Then H.o is a closed subspace, and if vEE is such
that ..1.(v) '# 0, then
E = Ho + Rv.
[IV, App.] THE KREIN-MILMAN THEOREM 85
If AI' A2 are non-zero functionals with the same kernel H, then there
exists c E R, c#-O such that Al = cA. 2 • Indeed, one sees at once that
c = Al (v)/ A2(V).
Let A#-O be an element of E*, and let c E R. By the hyperplane He we
mean the set of all x E E such that A(X) = c. In other words, He = A-I (c).
If Ho is the kernel of A, then He consists of all elements y + Yo with
y E Ho and Yo any fixed element of E such that A(Yo) = c.
The set of x E E such that A(X) ~ c will be called a closed half space
determined by the hyperplane, and so will the set of all x such that
A(X) ~ c. Similarly, we have the open half spaces, determined by the
inequalities A(X) > c and A(X) < c respectively.
If S is a closed subset of E and Xo a point, we say that a hyperplane
H separates Sand Xo if S is contained in one of the closed half spaces
determined by H, and Xo is not contained in this half space.
X·N~c,
86 BANACH SPACES [IV, App.]
given by
The image of S is a convex set <p(S) in Rn, which does not intersect the
neighborhood of <p(xo) determined by the inequality
Its closure does not contain <p(x o). By our result in the finite dimen-
sional case, there exists a non-zero vector
such that <p(S) lies in the closed half spaces determined by N and a
suitable constant c. We let
Remark. All that we need in the sequel is that, the assumptions being
as in the theorem, there exists a functional A E E* such that A(Xo) is not
contained in A(S).
x = tY1 + (1 - t)Y2
is not empty, and clearly is again in fi'. Hence by Zorn's lemma, there
exists a minimal element So in fi'. We contend that So consists of one
point. (This will prove our theorem.) Since elements of E* separate
points, it will suffice to prove that fo{ each A E E*, the set A(So) consists
of one point. But A(So) is convex and compact, whence a closed bounded
interval. Let c be a right end point of this interval. Then the set
A-l(C) n So is non-empty, convex, compact. We contend that it lies in fi'.
Let x be an element in r1(c) n So, and suppose that we can write
x = tY1 + (1 - t)Y2
with Y1' Y2 E Sand 0 < t < 1. Since So E fi', we get Y1' Y2 E So. Applying
A, we find that
Hence Y1' Y2 also lie in A-l(C), and this shows that A-l(C) n So is in fi'.
Since we took So minimal, we conclude that So is contained in A-l(C),
thereby proving our theorem.
x = tYl + (1 - t)Y2
with 0 < t < 1, then A.(x) = c = tA(Yl) + (1 - t)A(Y2), and hence A(Yd =
A(Y2) = c, so that YI' Y2 E A-I(C) Il S. From this we conclude that
YI = Y2, and hence that x is also an extreme point of S itself.
Let c be the right end point of the interval A(K). By Corollary 1.4, the
set A-I (c) Il K contains an extreme point of K, contradicting the fact that
A(S) < c, and proving our theorem.
o o
e-a
(1 - r)x - ra = h E H,
whence
r 1
x = -1-- r a + -1-- r h.
IxYI
Ilxll =sup- .
y"O Iyl
FnG = {O}.
I~I = L Ix.1
then this is a norm, and E is complete.
14. Let E be a Banach space, and P, Q two operators on E such that P + Q = I,
and PQ = QP = o. Show that
E = Ker P + Ker Q,
and that Ker P = 1m Q. Show that Ker P n Ker Q = {O}, and that Ker P
and 1m P are closed subspaces.
15. Let E be a Banach space and let F be a vector subspace. Let F be the
closure of F. Prove that F is a subspace, and is complete.
16. Let A be a subset of a Banach space. By c(A) we denote the convex closure
of A, i.e. the intersection of all convex sets containing A. We let c(A) denote
the closure of c(A). Then c(A) is convex. Prove: If K is compact, then c(K)
is also compact. [Hint: Show c(K) is totally bounded as follows. First find a
finite number of points Xl' ... ,X. such that K is contained in the union of
the balls of radius e around these points. Let C be the convex closure of
the set {Xl' ... ,x.}. Show that C is compact, expressing C as a continuous
image of a compact set. Let Yl' ... ,Ym be points of C such that C is con-
tained in the union of balls of radius e around these points. Then get the
desired result.]
[IV, §6] EXERCISES 93
17. Let F be the complete normed vector space of continuous periodic functions
on [ -n, n] of period 2n, with the sup norm. Let E be the vector space of all
real sequences IX = {an} such that L lanl converges. Define
IIXI = L'"
n=1
lanl·
18. Let K be a continuous function of two variables defined for (x, y) in the
square [a, b] x [a, b]. Assume that IIKII ~ C for some constant C> 0, where
II II is the sup norm. Let E be the Banach space of continuous functions on
r
[a, b], and let T: E -+ E be the linear map such that
Show that T is bounded and II TIl < C(b - a). For more on T, see Chapter
XIV, Exercise 5.
19. Let A be a commutative Banach algebra with unit element e, over the reals,
and define the exponential and logarithm maps by
u2
exp u = 1 + u + - + .. .
2!
and
(u - e)2 (u - e)3
log u = (u - e) - - - + - - - ...
2 3
Show that exp converges absolutely for an u E A, and that log converges
absolutely for all u with lu - el < 1. Show that the exp and log give inverse
continuous mappings from a neighborhood of 0 onto a neighborhood of e in
A . Show that they satisfy the usual function equations
20. Let X be a compact Hausdorff space and let C(X) be the Banach space of
real continuous functions on X. If A. is a functional on C(X) (sup norm) such
that A.(1) = IA.I, show that A. is positive, in the sense that if f E C(X), f ~ 0,
then A.(f) ~ o.
CHAPTER V
Hilbert Space
Essentially all of this chapter goes through over the real or the complex
numbers with no change. Since the theory over the complex does intro-
duce the extra conjugation, we use the complex language, and point out
explicitly in one or two instances those results which are valid only over
the complex.
Let E, F be vector spaces over C and let L: E -+ F be a map. We say
that L is antilinear, or semi-linear, if L is R-linear, and L(ax) = iXL(x) for
all x E E and a E C.
Let E be a vector space over the complex numbers. A sesquilinear
form or scalar product on E is a map
ExE-+C
denoted by
(x, Y)f-+ (x, y)
all x E E, we say that the form is positive. We say the form is positive
definite if it is positive, and <x, x> > 0 if x "# O. We shall assume through-
out that our form <, > is positive, but not necessarily definite. We ob-
serve that a sesquilinear form is always R-bilinear.
We define v to be perpendicular or orthogonal to w if <v, w> = O. Let
S be a subset of E. The set of elements VEE such that <v, w> = 0 for
all w E S is a subspace of E. This is easily seen and will be left as an
exercise. We denote this set by SJ.. Let Eo consist of all elements vEE
such that v E E1, that is <v, w> = 0 for all wEE. Then Eo is a subspace,
which will be called the null space of the hermitian product.
Theorem 1.1. If wEE is such that <w, w> = 0, then WE Eo, that is
<w, v> = 0 for all VEE.
Proof. Let t be real, and consider
If Re<v, w> "# 0 then we take t very large of opposite sign to Re<v, w>.
Then <v, v> + 2t Re<v, w> is negative, a contradiction. Hence
Re<v, w> = O.
This is true for all VEE. Hence Re<iv, w> = 0 for all vEE, whence
Im<v, w> = O. Hence <v, w> = 0, as was to be shown.
But
Hence
Proof. The first assertion follows from Theorem 1.1. The second is
left to the reader. The third is proved with the Schwarz inequality. It
suffices to prove that
To do this, we have
(v - cw, w) = 0,
(v, w)
c= - - .
(w,w)
98 HILBERT SPACE [V, §1]
The proofs come immediately from expanding out the norm according to
the definitions.
Let {V;} i E I be a family of elements of E such that Ivd i= 0 for all i.
For each finite subfamily, we can take the space generated by this sub-
family, i.e. linear combinations
with complex coefficients ci . The union of all such spaces is called the
space generated by the family {V;}iEI . Let us denote this space by F. We
say that the family {v;} is total in E if the closure of F is equal to all
of E.
As a matter of notation, we shall omit the double indices and write
VI' ... ,vn instead of Vi" ... ,Vin •
We say that the family {Vi} is an orthogonal family if its elements are
mutually perpendicular, that is <Vi' Vj ) = 0 if i i= j, and if in addition
Iv;! i= 0 for all i. We say that it is an orthonormal family if it is ortho-
gonal and if Iv;! = 1 for all i. One can always obtain an orthonormal
family from an orthogonal family by dividing each vector by its norm.
A total orthonormal family is called a Hilbert basis, or also ~n ortho-
normal basis. (Warning: It is not necessarily a "basis" in the sense of
[V, §1] HERMITIAN FORMS 99
abstract algebra, i.e. not every element of the space is a linear combina-
tion of a finite number of elements in a Hilbert basis.)
a=lx-Yol·
100 HILBERT SPACE [V, §1]
because of the definition of a. This shows that {Yn} is Cauchy, and thus
converges to some vector Yo' The lemma follows by continuity.
Yo
Figure 5.1
We let ex = t<z, Y), with t real "# O. We can then cancel t and get a
contradiction for small t, if <y, z) "# O. This proves the theorem.
Corollary 1.7. Let E be a Hilbert space, E"# {O}. Then there exists a
total orthogonal basis for E.
x=y+Z
Px = Y
Xi E F;.)
Let Pi be the orthogonal projection on F;. Then Xi = p;x, and for any
choice of elements Yi E F; we have
102 HILBERT SPACE [V, §1]
Proof Since
n
X - L p;x
i=l
is orthogonal to F1 , ' " ,Fn we can use exactly the same argument as in
Theorem 1.4, and the Pythagoras theorem to show the last inequality,
writing
2
Ix- t Yil2 = Ix - t p; x + It (p;x _ YJI2.
1=1 1=1
1
1=1
X= LXi
i=l
n=l
n n
V = V -
k=l
L: akvk
L akvk + k=l
and apply Pythagoras' theorem.
Conversely, we can define directly a set [2 consisting of all sequences
{an} such that L: lan l2 converges. If ex = {an} and P= {bn} are two se-
quences in this space, then using the Schwarz inequality, on finite partial
sums, one sees that
<ex, P) = L: anbn·
Again from the above convergence, we conclude that [2 is in fact a vector
space, because
is linear and preserves the norm. In this way, we get a map from our
space of periodic functions into [2, which is injective and preserves the
norm. It extends therefore uniquely to the completion.
In general, if two Hilbert spaces have total orthonormal families with
104 HILBERT SPACE [V, §2]
the same cardinality, then any bijection between these families extends to
a unique norm-preserving linear map of one space to the other.
Theorem 2.1. For every y in the Hilbert space E, the map Ay such that
Ay(X) = <x, y) is a functional. The association
Proof. The Schwarz inequality shows that IAyl ~ Iyl, and evaluating Ay
at y shows that IAyl = Iyl, so we get a norm-preserving semi-linear map
of E into E', semi-linear because of the hermitian nature of the scalar
product, namely for complex oc,
There remains to show that every functional comes from some y E E. Let
A be a functional, and let F be its kernel (the closed subspace of all x
such that A(X) = 0). If F =1= E, there exists Z E E, Z =1= 0 such that Z is
perpendicular to F (by Theorem 1.6). We contend that some scalar
multiple of Z achieves our purpose, say ocz. A necessary condition on oc is
that
<z,ocz) = A(Z)
or in other words, ii = A(Z}/<Z, z). This is also sufficient. Indeed, for any
x E E, we can write
and
XI---+CP(x, y)
Hence IAxl ~ ICPAllxi. This proves that IAI ~ ICPAI, whence our theorem
follows.
IA*I = IAI,
Proof. The first four properties are immediate from the definitions.
For instance,
and conversely,
q(x) = q>(x, x)
<Ax, y) + <Ay, x) = 0,
i<Ax, y) - i<Ay, x) = O.
V, §3. EXERCISES
For the first two exercises, recall that a sequence {xn} in a Hilbert space H
converges weakly to 0 if for all Y E H we have lim<x n , y) = o.
1. Let {v n } (n = 1,2, . . . ) be a denumerable Hilbert basis for the Hilbert space H.
Show that the sequence {v n } converges weakly to 0, and hence that the unit
sphere is not closed in the unit ball for the weak topology.
2. Suppose the Hilbert space H has a countable basis. Let x E H be such that
Ix I ;;i; 1. Show that there exists a sequence {un} in H with IUn I = 1 for all h
such that {un} converges weakly to x.
108 HILBERT SPACE [V, §3]
3. Let X be a closed convex subset of a Hilbert space. Show that there exists a
point in X which is at smallest distance from the origin.
4. Let E be a Hilbert space, and let {xn} be an orthonormal basis. Let {cn} be a
sequence of positive numbers such that L c;
converges. Let C be the subset of
E consisting of all sums L
anxn where lanl ;;; Cn. Show that C is compact.
S. Show that a Hilbert space is separable (has a countable base for the topology)
if and only if it has a countable orthonormal basis.
6. Let A be an operator on a Hilbert space. Show that
r
eigenfunction for K, with respect to a real number r, if
r
We take E with the L 2 -norm of the hermitian product given by
<f,g) = fg·
Prove that if fl' ... ,f. are in E, mutually orthogonal, and of L 2- norm equal to
1, and if they are eigenfunctions with respect to the same number r, then n is
bounded by a number depending only on K and r. [Hint: Apply Bessel's
inequality.]
8. Let E be a pre Hilbert space.
(a) If E is complex, then Im<x, y) = Re<x, iy).
(b) Let x, y E E. If E is real, then
If E is complex, then
(c) Let F be a normed vector space such that the parallelogram law holds for
its norm. Define <x, y) by the formula in (b). Show that this is a positive
definite scalar product.
PART THREE
Integ rati on
This part deals with integration in multiple contexts. We start with the
integral on arbitrary measured spaces, setting the basic framework in a
context which makes its structure particularly clear. The main idea is
that one starts the theory of the integral by defining the integral on a
natural space of simple functions where one sees immediately what the
integral means. The space of step functions is the one which covers all
cases, from the most general to the most special. As we shall also see, if
one wants integration on the reals, or in euclidean space, then the space
generated by characteristic functions of intervals or cubes, or the Coo
functions with compact support, also form a natural starting space for
integration.
It turns out that for the basic framework of integration, all one needs
for the space of values is linearity and completeness, so a Banach space.
I think it obscures matters to assume (as is often done) that values are
first taken in the real numbers, and to make abusive use of the ordering
properties of the reals and of positivity in setting up the integral. Fur-
ther comments on this will be made in Chapter VI, especially the intro-
ductory comments.
However, doing general Banach valued integration on measured spaces
does not mean that one eventually slights special properties of complex
valued integration over the real numbers. This entire part will mix gen-
eral considerations with particular situations and examples, especially on
euclidean space and the real line. Readers can see how having the gen-
eral machinery of integration on measured spaces, or locally compact
spaces, is used to make easier the formulation of more concrete results.
For instance, in Chapter VIII, we give specific results on approximations
on R or Rn with Dirac sequences and families. In Chapter IX, two
110 INTEGRA TION [PART THREE]
00
U An
n=I
A - B = A nCCxB
00
fields, vector spaces, etc.) it has become more or less standard practice
to assume that a ring has a unit element for multiplication, while an
"algebra" is merely an additive group with a bilinear law of composition.
Our definitions have therefore been made to fit these conventions, in the
analogous situation of algebras of subsets. Here, of course, the "unit
element" is the whole space.
Example 1.
Let X be a topological space, and let !/ be the collection of all open
sets. The (J-algebra generated by these open sets is called the algebra of
Borel sets. An element of this algebra is called Borel measurable. In
particular, every denumerable intersection of open sets and every de-
numerable union of closed sets is Borel measurable.
Example 2.
Let (X,.A) be a measurable space. Let f: X -+ Y be a mapping of X
into some set Y. Let % be the collection of subsets S of Y such that
f-1(S) is measurable in X. Then % is a (J-algebra. The proof for this is
immediate from basic properties of inverse images of sets. We call % the
direct image of .A under f, and could denote it by f*(.A). (Cf. Exercise
1.)
Example 3.
Let X be a measurable space, and let Y be a subset. If.A is the
collection of measurable sets of X, we let .Ay consist of all subsets AnY,
where A E.A. Then it is clear that .Ay is a (J-algebra, which is said to be
induced by .A on Y. Then (Y, .Ay) is a measurable space.
Measurable Maps
From now on, our maps will have values in a topological space, with
the Borel sets as measurable sets.
We note at once that taking complements, we could have defined
measurability by the condition that the inverse image of a closed set is
measurable. Furthermore, we see that the inverse image of a countable
union of closed sets, and the inverse image of a countable intersection of
open sets is measurable because if {Un} is a sequence of open sets, then
for n = 1, 2, .. . .
We shall now give a large" number of criteria for mappings and sets to
be measurable, and we shall see that limit operations preserve measur-
ability, and algebraic operations likewise, under extremely mild hypo-
theses on the image space Y. These hypotheses will always be satisfied in
practice, and trivially so in the case when we deal with maps into the
real or complex numbers, or into Euclidean n-space.
For this last assertion, we note that the product is composed of the
map (f, g) and the product C x C -+ C, which is continuous.
[VI, §l] MEASURED SPACES, MEASURABLE MAPS 117
m=l k=m
00
1
On the other hand, let A be a closed set. Suppose that x lies in every
union
for all positive integers m. Then for arbitrarily large k, we see that Jk(X)
lies in A, and hence by assumption the limit J(x) lies in A because A is
closed. Hence we obtain the reverse inclusion
n u Jk- (A)
00
m=l k=m
00
1 c J-l(A).
Let V be a fixed open set. For each positive integer n let An be the
closed set of all y E Y such that d(y, t;6'V) ~ lin, and let v" be the open
set of all y E Y such that d(y, t;6'V) > lin. Then
and
00 00
V = U An = n=l
n=l
U v".
Thus we have the inclusions
U J- U n U Jk-
00 00
and
This proves that the equality holds, and shows that J-l(V) is measurable.
118 THE GENERAL INTEGRAL [VI, §1]
This last result is really the main thing we were after. We need it
immediately in the next section to know that if f is a limit of measurable
real valued functions, then for every real a, the set
for k = 1, . .. ,N + 1
so that each Ak is measurable, the sets Ak (k = 1, .. . ,N + 1) are disjoint,
and their union is X . On each Ak we define a constant map t/ln by
if k = 1, ... ,N.
[VI, §1] MEASURED SPACES, MEASURABLE MAPS 119
and
Then the sequence {t/ln} converges pointwise to f, and each t/ln is a simple
function. This proves that measurability implies the other condition. The
converse is already known from M7, and thus our characterization of
measurable maps is proved.
The construction of the case we just discussed yields a useful addi-
tional property in the positive case:
M9. Let f: X --+ R ;;; o be a positive real valued measurable map. Then
f is a pointwise limit of an increasing sequence of simple maps.
Proof. The functions t/ln defined above are all ~ f, and we let
Positive Measures
oo·a=a·oo=O if a = 0,
oo·a = a · 00 = 00 if 0< a ~ 00 ,
oo+a=a+oo=oo if °
~ a ~ 00.
A =
n=1
then
as was to be shown.
Proposition 1.2. A map Jl:.# ...... [0, 00] is a measure if and only if
Jl(0) = 0, Jl is finitely additive, and if {An} is an increasing sequence of
measurable sets whose union is A, then
00
A = nAn'
n=l
then
The set of step maps St(Jl, E) is a vector space. If f is a step map, then
so is If I· If f: X -+ E is a step map and g: X -+ C is a step function,
then gf (also written fg) is a step map.
We shall not use the rest of this section until the corollaries of the
dominated convergence theorem in §5.
We shall define the integral on certain maps which are limits of step
maps. The present discussion is devoted to such limits. We define a map
to be J.L-measurable if it is a pointwise limit of a sequence of step maps
almost everywhere. In other words, if there exists a set Z of measure 0
and a sequence of step maps {cp.} such that {cp.(x)} converges to f(x) for
all x ~ Z. Let f: X -+ Y be J.L-measurable, and let A c X and BeY be
measurable subsets with f(A) c B. Then the induced map f : A -+ B is
J.L-measurable. Instead of MI, we have: if f : X -+ E is j.L-measurable, and
g: E -+ F is continuous, then go f is J.L-measurable.
Proof. All statements are clear, except possibly the last, for which we
give 'the argument: If {cp.} is a sequence of step functions converging
pointwise to f, then we let t/I.(x) = ll cp.(x) if CP.(x) #- 0 and t/I.(x) = 0 if
CP.(x) = O. Then t/I. is step, and the sequence {t/I.} converges pointwise to
1If.
set which IS a countable union of finite sets. Thus we now have two
necessary conditions for a measurable map to be jl-measurable, namely
countability conditions on its domain and range. It turns out that these
are sufficient.
Proof. We have already proved that (ii) implies (i), using our preced-
ing remarks, and M7. Conversely, assume (i). We may assume that X is
a disjoint union of subsets X k (k = 1,2, ... ) of finite measure. If we can
prove that the restriction flX k of f to each X k is jl-measurable, then for
each k there is a sequence {cp?l} (j = 1,2, .. . ) of step maps on X k which
converges almost everywhere to flX k • We define CPn by the following
values:
Then each CPn is a step map, and the sequence {CPn} converges almost
everywhere to f. This reduces the proof that f is jl-measurable to the
case when X has finite measure.
Suppose therefore that X has finite measure. We may also assume
that the image of f contains a countable dense set {Vk} (k = 1,2, ... ).
For each positive integer n, let B1/n(vd be the open ball of radius l/n
centered at vk • The union of these balls for all k = 1, 2, ... covers the
image of f, whence the union of the inverse images under f covers X
itself. If we take k large, it follows that the finite union of inverse images
differs from X by a set y" such that jl( y,,) < 1/2n. We let
We simply define the map CPn inductively to have the value Vi on the
inverse image of B 1/ n (V 1 ), the value V 2 on the inverse image of
and so forth. We let t/ln be equal to CPn on X - Zn and give t/ln the value
o on Zn . Then t/ln is a step map, and the sequence {t/In} converges
pointwise to J, except possibly on the set Z equal to the intersection of
all Zn, which has measure O. This proves what we wanted.
Remark 4. Let vi{ be the a-algebra of all subsets of the set X. Let
f: X ~ E be an arbitrary map into· a Banach space. Then f is measur-
able, and J1.-measurable if J1. is such that J1.(Y) = 0 for all subsets Y of X.
This shows that it is reasonable to exclude the behavior on a set of
measure 0 in our definition of J1.-measurability.
Proof This is clear by using (i) of Mll, and the following facts: A
denumerable union of sets of measure 0 has measure O. A denumerable
union of sets having countable dense subsets has a countable dense
subset. [If {Dk} is a sequence of denumerable sets in a metric space, then
'"
'" Dk => Dn U Dk => n=l
U Dn,
-",--
so that
For the rest of this chapter, we let (X, vIl, JI) be a measured space, i.e.
vIl is a a-algebra in X, and JI is a positive measure on vIl. We let E be a
Banach space. At first reading, the reader may assume that all maps fare
complex or real valued, that is E = C or R. No proof or notation would
be made shorter by this assumption.
Ix f dfl = ;~ fl(A;)f(AJ
If {BJ (j = 1, ... ,s) is another partition of A, then f is step with respect
to the partition {A; (\ Bj } and we have
s
L fl(A; (\ Bj)f(A;) =
j=l
fl(A;)f(AJ
Summing over i shows that our integral does not depend on the partition
of A. If f is step with respect to a partition of a set A and a set B, then
it is also step with respect to a partition of A u B, and we see that our
integral is therefore well defined.
If A is an arbitrary measurable subset and f is a step map on X,
recall that fA is the map such that fA(X) = f(x) if x E A and fA(X) = 0 if
x ~ A. Then fA is a step map both on A and on X, and we define
Ix f instead of Ix f dfl,
[VI, §2] THE INTEGRAL OF STEP MAPS 127
and even omit the X if the total space X is fixed, so that we also write
f f instead of Ix f.
If we integrate over a subset of X, then we shall always specify this
subset, however. We now have trivial properties of the integral.
First, the integral is obviously a linear map
f: St(,u, E) --+ E
(1)
J f=Jf+ff.
AuB A B
This is clear from the linearity, and the fact that fAuB = fA + fB'
Over the rea Is, the integral is an increasing function of its variables.
This means: If E = Rand f ~ g, then
(2)
Furthermore, if f ~ °
and A c B, then
(3)
(4)
128 THE GENERAL INTEGRAL [VI, §3]
we take a partition of a set of finite measure such that both f and g are
step maps with respect to this partition, and then we estimate using the
triangle inequality. This semi norm will be called the U-seminorm.
Note. The results of this section are at the level of a first course in
calculus. We don't take limits, and our results depend only on the
presence of an algebra (not necessarily a u-algebra) and a map Jl of this
algebra into the reals ~ 0 which is additive, i.e.
Proof. For each integer k there exists Nk such that if m, n ~ Nk, then
if m ~ n.
L (gk+l(X) -
00
gl(X) + gk(X))
k;1
Since gn and gn+1 are step mappings, it follows that Y" has finite measure.
On Y", we have the inequality
1
2n ~ Ign+1 - gnl
whence
Hence
Let
Zn = Y" u Yn+l u ....
Then
L
00
is absolutely and uniformly convergent, for x ~ Zn. This proves the state-
ment concerning the uniform convergence. If we let Z be the intersection
of all Zn, then Z has measure 0, and if x ~ Z, then x ~ Zn for some n,
whence our series converges for this x. This proves the lemma.
and tlfnl
converge to 0.
Given E, there exists N such that if m, n ~ N we have
f A-Z
Ifni < E.
L ~L
If" I If" - fNI + L IfNI
desired bound,
lim II fn II 1
Theorem 3.4. The space 'p 1 is complete, under the seminorm II 111'
Proof. Let Un} be a Cauchy sequence in 'pl. For each n there exists
an element gn E St(J.l) such that
which gives a 3e-proof of the fact that {gn} is a Cauchy sequence. For a
subsequence of n, we know by Lemma 3.1 that {gn} converges almost
everywhere to a function I in 'pl. For this subsequence, we then have
and this is < 2e for n sufficiently large in the subsequence. Hence the
subsequence is L 1-convergent to f. It follows that the sequence {In} itself
is L 1-convergent to I, and concludes the proof.
Cauchy sequence {f,,} approximating f Lemma 3.2 shows that this map
is well defined, and it is obviously linear. The definition of the seminorm
on 21 means that in this notation, we have
IIfll1 = Ily(f)111'
Ixf= Ix y(f).
This relation is true for step maps f, and consequently holds for the
extension of our continuous linear map to the completion. Therefore, it
holds also for elements of 21 by Lemma 3.3 and the definition of the
seminorm II 111 on 21. The preceding relation also shows that the inte-
gral has norm ~ 1, as a linear map.
is linear.
We observe that if f, 9 are in 'p 1(/1) then If I, Igl are in 'p 1(/1, R), and
consequently if E = R, then
The expression for the sup also shows that if {f,,}, {gn} are sequences
in 'p 1(/1, R) which are L l-convergent to functions f, 9 respectively, then
sup(fn, gn) is L 1-convergent to sup(f, g).
If f is a real function, then we can write
For any measurable set A and any f E 'p 1(/1) the map fA is also in 'pl.
(Recall that fA is the same as f on A, and zero outside A.) Proof: If {<Pn}
is a sequence of step maps approximating f, then {<PnA} converges almost
everywhere to fA' and is Cauchy because
Hence {<PnA} approximates fA' From the linearity of the integral, we thus
obtain :
(1)
JAuB
f=Jf+Jf.
A B
(2)
Furthermore, if f ~ °
and A c B are measurable, then
(3)
(4)
where II II is the sup norm. (We recall that 0· 00 = 0.) This is immediate,
taking an approximating sequence {<Pn} of step maps to J, using continu-
ity for the first inequality, and (2) for the second. When IIfII or Jl(A) is
infinite, the inequality is clear, and when both are finite, we use (2).
The next properties are general properties, immediate from the conti-
nuity of the integral. We make the Banach space explicit here.
by
fl-+ A. 0 f ,
and we have
f = g + ih
where g, h are real, we note that a sequence of complex step functions
approximates f if and only if its real part approximates g and its imagi-
nary part approximates h (with our definition of approximation, that is
U-Cauchy , and convergence almost everywhere). Thus
We first generalize the basic and crucial Lemma 3.1 to arbitrary maps in
21 . This will be formulated as Theorem 5.2. We need a minor lemma
to use in the proof, which was automatically satisfied when we dealt with
step maps. We define a measurable set to be a-finite if it is a countable
union of sets of finite measure.
138 THE GENERAL INTEGRAL [VI, §5]
Lemma 5.1. Let f E 'p 1(J.l) be measurable. Let c > O. Let Sc be the set
of all x E X such that If(x)1 ~ c. Then Sc has finite measure. Further-
more, f vanishes outside a (J-finite set.
This proves that Sc has finite measure. Taking the values c = 11k for
k = 1, 2, ... shows that f vanishes outside a (J-finite set. Actually we can
see this even more easily, since each <Pn vanishes outside a set of finite
measure, and f is the limit almost everywhere of {<Pn}, whence f vanishes
outside a countable union of sets of finite measure.
whence
[VI, §5] PROPERTIES OF THE INTEGRAL: SECOND PART 139
are bounded. Then {f,,} is Cauchy, and is both L1 and almost every-
where convergent to some function f E 21.
140 THE GENERAL INTEGRAL [VI, §5]
(J. = sup
k
f X
fk·
and f
inf fn ~ inf f
fn·
lim inf f.
k-oo n~k
exists, we call it the lim inC of the sequence {f.} . It is clear that if Un}
converges pointwise, then its lim inf exists and is equal to the limit.
Actually, in the next corollary,
will exist for almost all x, and the resulting function, which we may
[VI, §5] PROPERTIES OF THE INTEGRAL : SECOND PART 141
exists (so is a real number ~ 0). Then lim inf fn(x) exists for almost all
x, the function lim inf fn is in 2 1 , and we have
for j = 1, ... ,m
f f. ~ ff" ~
inf
• ~k
inf
.~k
lim inf
k-+oo .~k
ff.
Let hk = inf f •. Then {h k} is an increasing sequence for k = 1, 2, ... , and
.~k
we can apply the monotone convergence theorem to hk. The limit lim hk
k-+oo
is precisely lim inf f., and Fatou's lemma drops out as desired.
Note. Fatou's lemma is used most often in the simple case when {f.}
is pointwise convergent almost everywhere, and when the L l- seminorms
11f,,111 are bounded, thus ensuring that the pointwise limf" is in 21.
We now refer for the first time since the definition of f£1 to the
notion of ,u-measurability. The point is that we want to give criteria for
the limit of a sequence of maps to be in f£1, and ,u-measurability is the
natural hypothesis here. We refer the reader to Mll and emphasize the
countability implications arising from a map being in f£\ and hence
,u-measurable (by definition).
n~ Ix Ifni dJ1.
converges. Then the series
co
f(x) = L: fn(x)
n=l
f x
f dJ1. = f f fn dJ1..
n=l X
144 THE GENERAL INTEGRAL [VI, §5]
The sets Bn are increasing, and without loss of generality we may assume
[VI, §5] PROPERTIES OF THE INTEGRAL: SECOND PART 145
as the average of f over A. The theorem will assert that if the average of
f over all such A lies in some closed set S, then in fact the values f(x)
must lie in S for almost all x. We call this the averaging theorem.
Corollary 5.17. Let f E 2 1 (/1). For each step function g the map fg is
in 21 (/1), and if
for all sets A of finite measure, then If(x)1 ~ b for almost all x.
The next corollary is included for later applications. The reader inter-
ested only in the case of complex or real functions may omit it.
Proof. The proof is really just like that of Corollary 5.17. First we
may assume that the image of f is contained in a separable Hilbert
subspace. Let e be a unit vector. For any measurable set A of finite
measure, the step map eXA having value e in A and 0 outside A is
bounded measurable. Let us denote by fe the Fourier coefficient of f
along e so that fe is a function. We have
This being true for all A it follows that f e is equal to 0 almost every-
where. Since there is a countable Hilbert basis in our Hilbert space, it
follows that f is 0 almost everywhere.
Corollary 5.20. Let E be a Hilbert space and f E !i'1(j1., E). For each
unit vector e E E, let fe be the component of f along e. Let b ~ O.
Assume that for each unit vector e and each set of finite measure A we
have
If we let CP1 = inf(cp, 1), then IXy - CP11 ~ IXy - cpl, and hence
For what follows, we also observe that St(dA , R) is closed under the
operations of sup and info
Then ~ is a a-algebra in A.
B
IlxY n - CPnlll < 2n '
Let
00
y=UY,,·
n=l
Then
We take n so large that the first term on the right is < B. The second
term on the right is estimated by
n
L
k=l
IlxYk - CPklll < B.
Since Y is the union of all sets Y n An' we can find some n such that
It follows that
II XY - t
k=l
CPk II 1 ~ II XY - t
k=l
XYnA k t
I 1 + II k=l XYnAk - t
k=l
CPk II·
1
< 26.
This proves our special case.
[VI, §6] APPROXIMA TIONS 151
The general case is now obvious: a step map f ;/: 0 with respect to all
sets of finite measure is a finite linear combination
with Vj E E, Vj ;/: 0 for all j, and such that the sets lj have finite measure.
By definition, the space of these maps is U-dense in fel(p., E). For each
XY j we can find a step function CPj with respect to d such that
for all real step functions cP with respect to d. Let Y be a set of finite
measure. By Theorem 6.3 and Lemma 3.1, we can find a sequence of
step functions {cp.} with respect to d which converges almost everywhere
to XY and is also L1-convergent to Xy. Taking inf(cp., 1) and sup(cp., 0)
if necessary, we may assume without loss of generality that 0 ~ CPo ~ 1.
Then
tf=O
for all sets of finite measure Y. By the a-finiteness, every measurable set
is a countable union of sets of finite measure. Since fXy = 0 almost
everywhere by Theorem 5.15, we conclude that f = 0 almost everywhere,
thus proving our corollary.
for all COO functions cP vanishing outside some compact set, then f is equal
to 0 almost everywhere. We shall state this result formally later in R".
fJ, : d~[O,oo]
such that fJ,(0) = 0, and such that fJ, is countably additive on d. This
means that if {An} is a sequence of disjoint elements of d, and if their
union U An is also in d, then
L
00
the inf being taken over all sequences {An} in d whose union contains
Y. If X can be expressed as a countable union of sets of finite measure
in d, then there exists a unique extension of jJ. to a positive measure on
.A.
Proof. The proof will proceed in two steps and needs the notion of an
outer measure.
OM 1. We have jJ.(0) = O.
OM 2. If A, BE % and A c B, then jJ.(A) ~ jJ.(B).
OM 3. If {An} is a sequence of elements of %, then
L jJ.(An),
00
jJ.*(Y) = inf
n=l
the inf being taken over all sequences {An} of elements of d whose
union contains Y. Then jJ.* is an outer measure which extends jJ..
Proof. We first show that if A Ed, then jJ.*(A) = jJ.(A), in other words
jJ.* extends jJ.. Since
A=Au0u0u'"
Jl(Y);;;;; ;;;;;
n.j j=l
Proof. Since we deal only with Jl, we omit the prefix Jl-. We first
prove that !/' is an algebra. It obviously contains the empty set, and if
A is measurable, it is clear that ~A is measurable (the definition of
measurable is symmetric in A and ~A). Let A, B be measurable. We
show that A n B is measurable. Let Z be any subset of X. Since B is
measurable, we get
But this is seen by using the fact that A is measurable, and writing
Thus A n B is measurable.
Next we observe that if A l ' ... ,An are disjoint measurable sets, and Z
is arbitrary, then
n
Jl(Zn(A1u' '' uA n »)= L Jl(ZnAd·
k=l
156 THE GENERAL INTEGRAL [VI, §7]
n
~ L p.(Z 11 A k ) + p.(Z 11 CCA)
k=l
L
<Xl
To prove the existence part of the theorem, all we need to show now
is that the sets of our original algebra d are measurable. Let A E d and
let Z be any subset of X . The inequality
Corollary 7.4. Let (X, vIt, p.) be a measured space, and let .91 be a
subalgebra of vIt consisting of sets of finite measure, generating .1(, and
such that X is u-finite with respect to d. A subset Z of X has p.*-
measure 0 if and only if given e, there exists a sequence {An} in .91
whose union covers Z and such that
L p.(An) < e.
00
n=l
and
If P, Q E.9I x fJl these show that both P (') Q and P - Q E.9I x fJl. Since
P u Q = (P - Q) u Q
Proof Since (.91 x fJl) c (.91" x fJl") c (.91" ® fJl") it follows that
and we must prove the reverse inclusion. For each BE fJl consider the
a-algebra in X x B generated by all sets A x B with A E.9I. It is con-
tained in (.91 x fJl)", which therefore contains .91" x {B} for all B E fJl.
Now for any A E .91", it follows that {A} x fJl" is contained in (.91 x fJl)".
Thus finally,
.91" x fJl" c (.91 X fJl)",
Proof. Let Y' be the collection of subsets Q E vii ®.AI such that
Qx E.AI for all x. Then Y' contains all rectangles A x B with A E vii
and B E.AI. It will suffice to prove that Y' is a u-algebra. The point is
that the operation Qf-+ Qx commutes with all the operations of set
theory. Indeed, X x Y E Y'. If Q E Y', then ~Q E Y' because (~Q)x =
~(Qx)' If Q, P are in Y', then
For the rest of this section we let (X, vii, J1.) and (Y,.AI, v) be u-finite
measured spaces. Let .91 and f!4 be the algebras of sets of finite mea-
sure in vii and .AI respectively.
f = VXAxB
and
tfx dv.
160 THE GENERAL INTEGRAL [VI, §8]
If J is a step map with respect to .91 x f!J, we conclude that the map
XH t Jx dv
Ix df.l(x) t Jx dv,
Proof For each positive integer n, let Sn be the set of all x such that
v(ZJ ~ lin. Let S = USn. It will suffice to prove that S is contained in
a set of measure O. Given e, let {Rd be a sequence of rectangles whose
union contains Z and such that
00 e
k~ (f.l x v)(Rd < n2n'
1
-n ~ k=l
L V(Rk.J·
00
1
-f.l(T,.) ~
n
Loofx v(Rk.x) df.l = L (f.l x V)(Rk) < -2"
k=l
00
k=l n
e
'
162 THE GENERAL INTEGRAL [VI, §8]
This shows that Jl(T,,) < e/2n, whence S is contained in a set of measure 0,
thereby proving our lemma. The converse will follow from Corollary 8.5.
Theorem 8.4 (Fubini's Theorem, Part 1). Let f E ,Pl(Jl ® v). Then for
almost all x, the map fx is in 'pl(V), the map given by
x~ IfxdV
for almost all x (and defined arbitrarily for other x) is in ,Pl(Jl); and we
have
f Xx Y
f d(Jl ® v) = ff
x y
fx dv dJl(x).
is a map of X into St(fJI). Indeed, <Pn.x is a step map with respect to fJI,
[VI, §8] PRODUCT MEASURES AND INTEGRATION 163
shows that II>n is step with respect to d. We view the space St(~) as
having the Ll-seminorm. We contend that {lI>n} is a Cauchy sequence.
This is easily seen, because
(Of course, the U-seminorms taken on the right and left of the preceding
equation refer to different spaces.)
By the fundamental Lemma 3.1, we may assume without loss of gener-
ality (using a subsequence if necessary) that there exists a set T of mea-
°
sure in X such that for x ¢ T the sequence {lI>n(x)} is Cauchy. [Lemma
3.1 and its proof are valid for values in a Banach space. For our
purposes, we note that the proof of this lemma applies as well in the
seminormed case to yield a pointwise Cauchy sequence for almost all x.
Alternatively, we may also take the natural map of St(Pl) into U(v) and
apply the lemma with respect to the Banach space U(v).] This means
that for each x ¢ T, the sequence
for all x ¢ S u T.
Finally, we note that the map
is a step map with respect to d. [It is in fact the composite map of II>n
164 THE GENERAL INTEGRAL [VI, §8]
'P(X) = t fx dv.
fx fy
fP•. xdv dJl(x) = fxxY
fP. d(Jl ® v),
We can apply the corollary of Fubini's theorem (by linearity), and the
monotone convergence theorem once more to conclude that the sequence
given by
decreases to O.
k = 1, 2, .. . ; X E I, x ¢ U,
L J.(x) dx ~ CB + 1(l)B,
which proves our theorem.
with B. = A1 U .. . u A.
For the rest oj this section, we let /1 denote Lebesgue measure. One
customarily writes y1(RP) instead oj y1(/1) in this case.
fRP
fqJ dJi. =0
a b
The function
g: x f--+ f:oo h(t) dt = g(x)
[VI, §9] THE LEBESGUE INTEGRAL IN RP 169
a b
/ \
a b
(ii) We have
In fact, if
fRP
(ljI - <p) dJ.l < 8.
[a l - 8, bl + 8] x ... x Cap - 8, bp + 8]
Observe that in deriving this lemma, we are dealing with the simplest
case of Riemann integration. The lemma is at the level of elementary
calculus.
The result of Theorem 9.3 really concerns the values of the map I on
bounded sets of RP. In many applications, it is not convenient to restrict
oneself to elements of ,21, and one needs a formulation which allows us
to deal with maps locally. Thus we say that a map I: RP -+ E is locally
integrable if for each compact set K in RP the map IK (equal to I on K
and 0 outside K) is in ,21 (Jl).
Corollary 9.5. Let I be a locally integrable map on RP such that lor all
qJ E C~(RP, C) we have
Proof This is really what Theorem 9.3 proved, since all we have to
consider is IA for every bounded rectangle A.
f
we have
fRP
ra! dJl =
RP
I dJl.
[VI, §9] THE LEBESGUE INTEGRAL IN RP 171
Proof. By Theorem 6.3 we can find a sequence {q>n} of step maps with
respect to finite unions of rectangles which converges both L 1 and almost
everywhere to f. If q> is a step map as above, it is clear that its integral
is the same as the integral of a translation !aq>, because if R is a rectan-
gle, then
Thus if Y has finite measure, given e there exists an open set U and a
compact set K such that
~ J.l(K) + e.
This proves our assertion when Y is bounded. The general case follows
at once by considering the intersections of Y with a sequence {Rn} of
rectangles such that Rn c Rn+l for all n, and such that the union of the
Rn is the entire euclidean space.
172 THE GENERAL INTEGRAL [VI, §1O]
I: X -+ y and g: Y -+ Z
lim sup a.
we mean the least upper bound of all points of accumulations of the sequence
{a.} . We allow -00 and +00 as points of accumulation, taking the obvious
ordering in [ -00, 00] where
Ix 1. dJ-l
exist and are bounded (so in particular each In is 0 outside a set of finite
measure), define the integral of I to be their least upper bound, and if
unbounded, define the integral of I to be 00. Show that this is well
defined, i.e. independent of the sequence {I.} increasing to f. Formulate
and prove the monotone convergence theorem in this context. Nate: In-
stead of redoing integration theory, you can quote results from the text to
shorten the procedure.
(d) For each measurable A and positive measurable map I: X -+ [0,00]
define
lim L 1. dJ-l = 0,
If (X, .II, /1) and (Y, .;V, v) are measured spaces, show that
8. (a) Direct image of a measure. Let (X,.II) and (Y, .;V) be measurable spaces.
Let I : X -+ Y be a map such that for each BE .;V we have I-I (B) E.II.
Let /1 be a positive measure on .II, and let /1. = 1./1 be defined on .;V
by /1.(B) = /1(J-1(B»). Show that /1. is a measure. Show that if g is in
2 1(/1.), then go I is in 2 1(/1), and that
is measurable (with respect to vIt). [Hint: Show that the set of Q in vIt ® %
having the above property is a monotone family containing the rectangles.]
12. Show that if c, is the (Lebesgue) measure of the closed n-ball in R' of radius
1, centered at the origin, then
and therefore
C, = C.-I
fn/2
-n/2
cos' t dt,
, rG + 1)
C = ----
13. Let T be a metric space and let f be a map on X x T such that for each
t E T the partial map
J,: x f-+ f(x, t)
is in Ie I. Assume that for each x the map t f-+ f(x, t) is continuous. Finally
assume that there is some g E Ie I (Il, R) such that Jf(x, t)J ~ Jg(x)J for all x.
Show that the function <II given by
L
tive is given by
D<II(t) = Dd(x, t) dll(X).
f * g(x) = f Rp
f(t)g(x - t) dJl(t).
(a) Show that f*gE2'1(Jl) and that Ilf*gI11:::::; Ilf11111g111' We call f*g the
convolution of f and g.
(b) Show that convolution is commutative, associative, bilinear, and that
2'1(Jl) is therefore a Banach algebra. Does there exist a unit element in
this algebra?
16. Let M be the set of all finite positive Borel measures on RP. For each Jl E M
define IJlI = Jl(RP). For Jl, v E M, and any Borel subset A of RP define
(g) After you have read about complex measures in the next chapter, show
that all the previous properties apply as well to such measures, and that
these measures therefore form a Banach algebra under convolution.
17. Let X = [-n, n], and let J1 be Lebesgue measure. Let f E 2'1(J1, C). Show
that one can define the Fourier coefficients of f in the usual way, by
Cn =
1
-
fn .
f(x)e- InX dx.
2n -n
lim
t-c.o
f R
f(x)e- itx dx = O.
[VI, §lOJ EXERCISES 177
!/J H f R
dt
!/J(t)fiI'
for !/J a step function with respect to intervals not containing 0, defines a
positive Borel measure on R*. We denote this measure by f.1.*. Show that
a function I is in ..'f'1(f.1.*) if and only if I(x)/Ixl is in ..'f'1(f.1.), where f.1. is
Lebesgue measure, and that in this case,
f R*
I df.1.* = f R-{O}
I(x)lxl- 1 dx.
(b) Show f.1.* is invariant under multiplicative translations, and so is the inte-
gral on R* with respect to f.1.*. (Multiplicative translations are of type
x H ax for a =1= 0.)
20. Not all sets are measurable. Consider the reals modulo the rational numbers,
and in each coset x + Q, x real, select an element y such that 0 ~ y < 1.
Show that the set consisting of all such elements cannot be Lebesgue measur-
able. [Hint: Use the countable additivity to show that this set cannot be
measurable.]
21. Let X be a measured space with finite measure f.1.(X). Let IE ..'f'1(f.1.). Com-
pute the limit
X=nXn
be the product space. Let At be the a-algebra generated by all sets of the
form
f.1.(A) = nf.1.n(An)·
(b) Let In E ..'f'l(f.1.n) and assume that In is the characteristic function of Xn for
almost all n. Show that the product function 1= ®In is in ..'f'1(f.1.), and
178 THE GENERAL INTEGRAL [VI, §10]
that
r
all n we have for all B ~ a. Let
l.(x) = F" .
Assume :
(a) The sequence {F,,} converges pointwise to a function F, uniformly on each
finite interval [a, B].
(b) The sequence {I.} converges uniformly on [a, (0).
(c) The improper integrals
f F"OO
a
= lim
8- 00
f8 F"
a
and fa
OO F = lim
8-00
f8 F
a
exist.
Then
lim f oo F" = f oo F.
"-CO a a
for all B.
dist(B, C) > 0,
we have
by considering the sums L Il(Z II I-I B k) for k even and k odd, and applying
the hypothesis (* ).]
25. Let X be a metric space and :F a family of subsets of X whose union covers
X. Let
q>: :F -+ R U { 00 }
IlAA) = inf
~, FE~
L q>(F),
(b) If all elements of ~ are Borel sets, prove that for any subset A of X we
have
Jl(A) = inf Jl(B),
where VIft is the volume of the m-dimensional ball in Rift, and let ~ be the
family of all open sets of Rift. The associated Caratheodory measure is called
the m-dimensional Hausdorff measure .?fm .
CHAPTER VII
Consider first complex valued functions. We let 22(fJ.) be the set of all
functions f on X that are limits almost everywhere of a sequence of step
functions (i.e. fJ.-measurable), and such that Ifl2 lies in 21. Thus
the value of <f, g) at x being given by the scalar product <f(x), g(x»
in E.
The reader interested only in the complex numbers can take E = C
and the product to be <f, g) = fg, where the bar denotes complex conju-
gation. Not a single proof, however, will be made shorter or simpler.
(f,g)r-. Ix <f,g)dfJ.
182 DUALITY AND REPRESENTATION THEOREMS [VII, §1]
Proof We apply the theorem and the Schwarz inequality to the pair
If I and Ix (the constant 1 on X).
Then y" has finite measure, and the proof of Lemma 3.1 in the preceding
chapter goes through as before. We have J-L(y") ~ 1/2n, and we let
Zn = Y" U y"+1 U . .. .
2 1
Ih+1 (x) - h(x)1 < 2k
so that the series
is the U -seminorm of Ifn - fml2 . We fix m and take the limit as n -+ 00.
We can apply Fatou's lemma, and conclude that If - fml 2 is in 21,
184 DUALITY AND REPRESENTATION THEOREMS [VII, §1]
so that for large m we see that Ilf - fml12 is small, i.e. the sequence Um}
is L 2 -convergent to f. This proves our theorem.
Proof. Obvious.
Proof. The proof is essentially the same as in the !f1 case. For each
positive integer k let
gk = sup Ifn - fml·
m. n ~k
Then l/In is a step map for each n, and the sequence {l/In} converges
pointwise to f Furthermore, Il/Inl ~ 21fl for all n. The theorem shows
that {l/In} is L 2-convergent to f Any element of 22 is equivalent to one
for which you can find a sequence {qJn} as above. Hence our corollary is
proved.
where I II is the sup norm, and the inf is taken over all bounded
j,L-measurable maps g equal to f almost everywhere. Alternatively, for
each c ~ 0 let Sc be the set of all x such that If(x) I ~ c. We could have
defined Ilfll", by the condition:
Theorem 2.1.
(i) The space ,:e00(11) is complete. If Un} is an L oo-Cauchy sequence in
,:e00(11), then there exists a set Z of measure 0 such that the conver-
gence of {f,, } is uniform on the complement of Z.
(ii) If E is finite dimensional, then the simple maps are dense in
L 00(11, E).
(iii) If I1(X) is finite, then given I: and f E ,:e00(11), there exists a step
map cp and a set Z with I1(Z) < I: such that
for some n, or some pair m, n. Then Z has measure 0, and the conver-
gence of the sequence is uniform on the complement of Z. We let f have
value 0 in Z and be the uniform limit of the sequence {f,,} on the
complement of Z. Then f E ,:e00(11), and clearly is the L 00 limit of {f,,}.
Now assume that E is finite dimensional, or say equal to the complex
numbers for concreteness. Let f E ,:e00(11, C). After replacing f by an
equivalent function, we may assume that f is measurable and bounded.
Say the values of f are contained in a square. We cut up the square into
small I:-squares which are disjoint, and take their inverse images in X.
These give a partition of X and we can define a simple function with
respect to this partition by giving the function anyone of its values in a
given square. To get our small squares, let, say, e 1 , e2 be the standard
unit vectors in C = R2, and let S be the square
-N ~ n ~ N and -N~m~N,
then our small squares Sm.n cover the image of f as desired. The argu-
ment also works in any finite dimensional space, taking unit vectors
e 1 , .. . ,ep in RP.
[VII, §2] DUALITY BETWEEN L 1 (Ji) AND Loo(Ji) 187
Finally, for the third part of the theorem, if Ji(X) is finite, then any
element of 2 OO (Ji) is in 21 (Ji), and our assertion follows from the fact
that elements of 21 are L I-limits of step maps, together with the funda-
mental lemma of integration, or Theorem 5.2 of Chapter VI.
given by
(f, g)H Ix fg dJi = [f, g]11
We investigate this situation when we deal with the spaces L 1 (f.l) and
L OO(f.l).
Theorem 2.2. Let f.l be a-finite. The kernels on the right and left of the
bilinear map
The maps g f--+ Ag and ff--+ AI for g E YOO(f.l) and f E y 1(f.l) induce norm-
preserving linear maps of L OO(f.l) and L 1 (f.l), respectively, into the other's
dual space. In the case of L OO(f.l), the map g f--+ Ag is a norm-preserving
isomorphism between L OO(f.l) and the dual space of U (f.l), i.e. the map is
surjective.
and shows that IAgl ~ Ilglioo. For the reverse, let b = IAgl. For each
subset A of finite measure, we have
for all step maps f For any measurable set A of finite measure, we then
190 DUALITY AND REPRESENTATION THEOREMS [VII, §2]
obtain
for all step maps f, this same relation must hold true for all f E 2 1 (J-l)
because the step maps are dense in 21. This proves our last assertion
when X has finite measure.
The general case when J-l is a-finite follows easily. We write X as a
disjoint union of sets of finite measure X k (k = 1,2, ... ). Let f E 21 (J-l),
and let fk = fXk be the same as f on X k and 0 outside X k • Then the
series
We let
L
00
9= gk
k=1
We now repeat the statement of Theorem 2.2 for the Hilbert case.
U(J-l, E) x L OO(J-l, E) -+ C
The kernels on both sides are O. The map g 1-+ Ag induces a norm-
preserving linear map of L OO(J1., E) onto the dual of U (J1., E) (so the map
is surjective), and the map fl-+ AJ induces a norm-preserving antilinear
map of L 1 (J1., E) into the dual of L OO (J1., E) (not necessarily surjective).
Proof. Exactly the same, except that when for instance we considered
in the proof of Theorem 2.2, we now have to write <eXA, g) for some
unit vector e E E, and apply Corollary 5.20 instead of Corollary 5.18 of
the averaging theorem.
v(A) = L f dJ1..
X=AuB
of X into a disjoint union of measurable sets such that Vi is concentrated
in A and V2 is concentrated in B.
Let ft'OO(vIt, C) denote the space of bounded measurable functions on
X. We make no reference to any measure here at all, and we take the
sup norm on this space. Let V be a finite positive measure on vIt. This
means that v(X) < 00. Then V gives rise to a functional on ft'OO(vIt, C) by
the map
Theorem 2.4 (Radon- Nikodym and Lebesgue). Assume that fJ. is (f-
finite, and let V be a finite positive measure on vIt. Then there exists a
unique decomposition
with Jl' J2 E ft'1(fJ.), and vs, v; singular with respect to fJ., then
~ 11q>1I2111xI12'
where Ix is the function equal to 1 on X. Hence the map
(uniquely determined up to equivalence) such that for all step functions q>
t t
we have
q> dv = q>h d(Jl + v).
(1)
194 DUALITY AND REPRESENTATION THEOREMS [VII, §2]
Let Y be the set of all x E X such that 0 ~ h(x) < 1, and let Z be the set
of all x E X such that h(x) = 1. First let g be the characteristic function
of Z. From (1) we see that ft(Z) = O. Let g be arbitrary (bounded
measurable) and iterate (1). By induction we obtain
(2)
t L
that
gh" dv -+ g dv as n -+ 00.
Let
h
f= -
1- h
Corollary 2.5. Assume that ji. is a-finite. Let E be a Hilbert space and
let v be a positive measure on Jt, absolutely continuous with respect to ji..
Let hE ,;el(V, E). Then there exists f E ,;el(ji., E), uniquely determined
up to equivalence, such that h dv = f dji.. In other words, if a functional
on ,;eOO(ji., E) can be represented by a finite E-valued measure, then it
can be representeq by a map fin ,;el(ji., E).
°
Proof. Let l/Ihl denote the function equal to at a point x such that
h(x) = 0, and equal to l /lh(x)1 if h(x) #- 0. Then h/ lhl is ji.-measurable
and bounded, and Ihl is in ,;el(V, R). Then Ihl dv is a positive measure
on Jt, which is absolutely continuous with respect to ji.. By the positive
Radon-Nikodym theorem, we conclude that Ihl dv = k dji., where k is
positive and in ,;el(ji.). Then
h
f= - k
Ihl
is in ,;el(ji., E), being the product of a bounded ji.-measurable map and an
element in ,;el(ji.). It is then clear that this f satisfies our requirements.
We shall see in the next section that, in fact, we can start from the
"measure" point of view to arrive at our functionals.
of disjoint measurable sets whose union is A. (We don't use the word
partition, which was used for a finite decomposition of a set of finite
measure with respect to a positive measure.) A map
v:..# -. E
L v(An)·
00
v(A) =
n=1
the sup being taken over all decompositions {An} of A. We shall prove
that Ivl is a positive measure, and that if E = C (or is finite dimensional),
then Ivl is in fact real valued, i.e. finite.
Ivl(A) ~ Ivl(B).
°
Proof. Let {An} be a decomposition of A E.It. Let bn be a real
number ~ such that bn ~ Ivl(An). Let {Anj } be a decomposition of An
such that
In
Ivl(An) ~ Ivl(A).
~ I
hn
Iv(An n B)I ~ Ln Ivl(An)·
This is true for all decompositions {Bj } of A, whence we get the reverse
inequality
Ivl(A) ~ I Ivl(An),
n
Proof The general case reduces at once to the real case (compo-
nentwise). We deal with the real case as in Saks [Sa]. Suppose that
Ivl(X) = 00. We first observe that there exist measurable subsets of X
whose measures have arbitrarily large absolute values. This is seen as
follows. We take a decomposition {Xn} of X such that
198 DUALITY AND REPRESENTATION THEOREMS [VII, §3]
is large. We combine all those terms with indices n such that v(Xn) have
the same sign. For either + or -, the corresponding sum will be large.
We take a finite number of such n, but sufficiently many so that the sum
of the corresponding Xn is a subset B with Iv(B)1 large. All we need here
is the finite additivity of v.
Now we construct a decreasing sequence of subsets of X having
measures whose absolute values tend to infinity. Let X = AI. By what
we have just seen, there exists a subset B c: Al such that
Iv(B)1 ~ IV(Adl + 2.
If Ivl(B) = 00, we let A2 = B. If Ivl(B) is finite, then
Ivl(AI - B) = 00
and we let A2 = Al - B. Then
Let v(A) be the sequence whose n-th term is vn(A). It is clear that the
total variation of v is infinite and that v is countably additive on the
positive line X consisting of all real numbers ~ o.
[VII, §3] COMPLEX AND VECTORIAL MEASURES 199
v: vIt -+ E
such that Ivl(X) is finite, i.e. such that Ivl is a real valued positive
measure. [Recall that if A c B, then Ivl(A) ~ Ivl(B).] For simplicity, we
also call a vectorial measure a measure, and when we have to make a
distinction with the objects discussed in Chapter VI or the preceding
sections, we emphasize this and say positive measure for the former
object. Another way of making the distinction is to say (even more
correctly) an E-valued measure for our map v: vIt -+ E.
It is clear that E-valued measures form a vector space denoted by
Mi(vIt, E), or simply Mi. For such a measure, we define
IIvil = Ivl(X).
IIIlJII ~ IIfil 1·
We shall prove that these two conditions are equivalent. It is clear that
AC 2 implies AC 1. Conversely, assume AC 1. If AC 2 is false, for
each positive integer n there exists a set y" such that Jl(y") < 1/2n , but
Iv(y")1 > t:. Then Ivl(y") > t:. Let
because y" c Zn. Hence there is some measurable subset Z' of Z such
that v(Z'}"# 0, contradictin'g AC 1 because Jl(Z'} = O. This proves the
equivalence between our two conditions.
Ix If I dJl - t: ~ L If I dJl.
By Jl-continuity, there exists (j such that if Z is a set with Jl(Z) < (j, then
Write
n
cP = L ViXA i '
i=1
=L cP + L I>Il(AJ
i Ai i Il(A)
Proof For each measurable set A, we can apply Theorem 3.3 with
respect to A and get
n
cP = L
i=1
ViXA i
dv = h dlvl.
qJ E St(1 vi, C)
t
we have
the set of all x E X such that Ih(x)1 < r. Let {An} be a decomposition of
Sr. Then
In
IV(An)1 =I
n
If XA
X
n h divil ~I n
rlvl(An) = rlvl(Sr)'
and
dv = h dlvl.
Then h takes on only the values 1 and - 1. Let A be the set of points
where h takes the value 1, and let B be the set where h takes the value
- 1. Let v;; and vi: now be defined by the formulas
and
It is then clear that vi: and vi: are mutually singular, and it is immedi-
I
204 DUALITY AND REPRESENTATION THEOREMS [VII, §4J
ately verified that v;; = v+ , vi; = V- . We leave the uniqueness and the
proof of the last properties as an exercise.
Proof. Identical with that of Theorem 3.5, except that we must insert
unit vectors e and write eXAn or eXA in the appropriate place.
dlvl = k dfl
with some positive k in !l'l(fl, R), whence by the theorem, on step maps
we get (cf. Exercise 15)
dv = h dlvl = hk dfl,
as was to be shown.
A: St(Jl, E) --. C
Bn = A 1 U ... u An·
As for its total variation, let en be the unit vector in the direction of
v(An), and consider the series
00 00
..1.(g) = L ..1.(ekXA.) = L IV(Ak)1 =
k=l k=l
1..1.(g)l,
1..1.(g)1 ~ r(A).
Remark. The part of the proof showing that v is a measure does not
depend in an essential way on the assumption that E is a Hilbert space,
208 DUALITY AND REPRESENTATION THEOREMS [VII, §4]
and goes through with very minor modifications in the arbitrary Banach
case. The definition of Jl-continuity of a functional A. applies in this case,
and one can characterize such functionals as measures in the following
manner:
Assume that Jl(X) is finite. Let E be a Banach space and E' its dual
space. There exists a unique norm-preserving linear map
from the space of Jl-continuous E' -valued measures into the dual space
of L OO(Jl, E), denoted by v ~ dv, whose image is the space of Jl-
continuous functional on L OO(Jl, E), and such that on step maps VXA
(v E E and A measurable) we have
is relatively compact.
R'. Given A measurable with 0 < Jl(A) < 00, there is some B c A and a
compact subset K of E not containing 0 such that Jl(B) > 0 and
m(Y) is contained in the cone generated by K for all Y c B.
Note. The cone generated by K is the set of all positive finite linear
combinations of elements of K. Condition R' may be expressed by say-
ing that m has compact direction locally somewhere. Condition R' is
obviously satisfied in the finite dimensional case. A discussion of the
literature and applications will also be found in Rieffel's paper. For an
example when the measure m cannot be written as JlJ' even though it is
Jl-continuous, cf. Exercise 21.
[VII, §5] THE LP SPACES, 1< P< 00 209
1 1
- + - = 1,
p q
a 1/ P b 1/ q ~ ~ + ~,
p q
There are several easy proofs for this. Either take the log of both sides
and use the convexity of the log, or proceed as follows. If t ~ 1, then
Theorem 5.1. Let 1 < p < 00. Then 2P(p,) is a vector space. If we
define
First this shows that Ifllgl is in !l'1 (corollary of the dominated conver-
gence theorem), and second it yields the last inequality stated in the
theorem, after we integrate over X. To show that II lip is a seminorm,
write
Integrating and using Holder's inequality yields the fact that II lip is a
seminorm, and concludes the proof of the theorem.
Proof. As before.
Corollary 5.4. The step maps are dense in ff P, and U(j,t) is the comple-
tion of the step maps in the U-seminorm.
Proof. As before.
Theorem 5.5. Assume that j,t is a-finite. For f E ffP(j,t) and 9 E ffq(j,t),
we let
and define Ag by Ag(f) = <J; g)w Then the map 9 1--+ Ag is norm-
preserving isomorphism of Lq(j,t) onto the dual space of U(j,t).
Proof. We consider first as usual the case when j,t(X) is finite. Our
map 9 1--+ Ag is certainly an injective linear map, and we have
co
Ah = L A(hk)·
k=l
For each k we have Ahk = <hk> fk)/l. If we let f = LA, it follows that
A = f dJl on ffP(Jl). This concludes the proof of the U-duality theorem.
Remark. The proof follows the classical pattern (see Rudin [Ru 1] or
Loomis [Lo]), granted the L 2 and (L \ L CO)-duality theorem. For the
general case when E is a Banach space, and one wants U(Jl, E') to be
dual to U(Jl, E) for 1 ~ p < 00, cf. Dinculeanu [Din], §13, Corollary 1 of
Theorem 8, where this is proved under some countability assumption.
and
Let (Xn' .An' Jln) be a sequence of measured spaces such that Jln(Xn) = 1
for almost all n (meaning for all but a finite number of n). Let.A be
the a-algebra in the product space
unique measure J-l on (X, vii) such that Jor all such sets A we have
The above theorem has a simple intuitive content, but some applica-
tions require a stronger version, as follows.
Let
n
Hn(x) = L hk(x)
k=l
be the partial sum. Assume that L IIhkll~ converges. Then the limit
Z = {x E X, max Hff(x)
1 ~k ~n
~ e}.
Then
n
ejl(Z) ~ L
k=l
Ilhnll~·
Proof Let
Y,. = {x E X such that Hff(x) ~ e and H/(x) < e for all i < k}.
In other words, Y,. is the set of points x such that Hff(x) is the first
partial sum at least equal to e. Then the sets Y,. are disjoint, and we get
the inequality
Write
The last term is negative, and we shall leave it out when we integrate.
On the other hand, the middle term gives
216 DUALITY AND REPRESENTATION THEOREMS [VII, §6]
L
k=l
Ilhkll~ < 00.
Define
for large n, so that w,; has measure ~ 1/2n - 1 . Then the partial sums
L hk{x) converge for x not in w,;. Hence if we let W be the intersection
W= nw,;,
then these partial sums converge for x not in W, and W has measure
zero, thereby proving the lemma.
f fn df.1.n = O.
Let {bn} be a sequence of positive real numbers monotonically increasing
to infinity. If
"L... b211fnl12
1 2 < 00,
n
[VII, §7] EXERCISES 217
Proof Let hn = /,, /bn and apply the lemma to the partial sums
The lemma says that these partial sums converge for almost all x. It is a
trivial fact (proved by summation by parts) that if ak is a convergentL
seq uence, then
n
L
k=l
akbk = o(bn )·
is a constant cn . Define
and the triangle inequality is satisfied]. The only difference from a metric is
that we may have d(A, B) = 0 and yet A # B. In this way, d becomes a
topological space, and II-continuity corresponds to the topological notion.
2. Radon-Nikodym derivative. Let II, v be positive measures, and let m be a
complex measure. Suppose that dm = f dll, where f E ,!l'1(1l) and dll = g dv
where g E ,!l'1(V, R). Prove that fg E ,!l'1(V), and that dm = fg dv. If we use the
notation dm/dll = f and dll/dv = g, then we have the old formalism
dm dll dm
dll dv = dv
Assume that Il(X) < co. Show that F is finite dimensional and that
dim F ~ CIl(X).
and also
Show that [00 consists of all complex sequences r:x = {a.} with norm
(b) Show that [1 is the dual of the subspace Co of [00 conslstmg of all
sequences r:x = {a.} such that a. -+ 0 as n -+ 00. Show that the dual of [I
is [00 quoting any theorem from the text.
(c) Show that Co and [1 are separable, but that [00 is not separable.
The space H •. (For applications to PDE, cf. SL 2 (R), Appendix 4.)
6. Let s be an integer. On the integers Z define
Show that Hs = L2(Z, /-Is), and in particular is complete for the norm
associated with this scalar product.
(b) Show that the finite sequences! = {an} such that a. = 0 for all but a
finite number of n form a dense subspace of Hs .
7. For each function! E Coo(T), where T = R/Z is the circle, or if you wish, for
each COO function on R, periodic of period 1, associate the Fourier series
(a) Integrating by parts, show that the coefficients satisfy the inequality
la.l« rnr1
for each positive integer k. The symbol « means that the left-hand side
is less than some constant times the right-hand side for Inl-+ 00.
(b) Prove that Coo(T) c L2(Z, /-Is) for all s E Z, and that Coo(T) is dense in this
space L2. [Look at the finite Fourier series
L a.e2ninx.]
i.i;>N
8. Let r < s. Prove that the unit ball in Hs is relatively compact in Hr, in other
words that this unit ball is totally bounded in Hr.
220 DUALITY AND REPRESENTATION THEOREMS [VII, §7]
9. Let {In} be a sequence in 22(X, II) such that 11!.112 --> 0 as n --> 00. Prove that
St(lvl, C) --> E,
and that this map is U(lvl)-continuous. This linear map can therefore be
extended linearly by continuity to 2 1(lvl, C), thus allowing you to define
Sf dv, for f E 21(1vl, C).
15. Let E be a Banach space and E' its dual. In the bilinear map
given by
show that IAtl = IIflll and IAgl = Ilglloo,just as in the Hilbert case. [Hint: Use
step maps, and for a constant map, use the Hahn-Banach theorem to see
that given vEE, there exists v' E E' such that Iv'l = Ivl and (v, v') = Ivl.]
16. Assume that II(X) is finite. Let E be a Banach space and E' its dual space.
The definition of a II-continuous functional A on L 00(11, E) is as in the text.
Show that such a functional can be written in the form A = dv for some
[VII, §7] EXERCISES 221
E' -valued measure v, in the sense that on a map VX,c (v E E and A measur-
able) we have
l(vX,c) = <v, v(A) .
17. Prove the statement included in the remark following Corollary 4.4 of Theo-
rem 4.1, concerning that part of the dual of L'XJ(Il, E) represented by a mea-
sure in E'.
18. Let f be a Il-measurable map of X into a Banach space E. Given a measur-
able set A with Il(A) finite, and e, show that there exists Z c A such that
Il(Z) < e and f(A - Z) is relatively compact (or equivalently, totally bounded).
We may say that f is locally almost compact valued.
19. The essential image. Let E be a Banach space. Let f be a measurable map
and let A be a measurable set. The essential image of f on A is defined to be
the set of all vEE such that for every r > 0 the measure of the set
A nr'(Br(v)
= closure of U ei,c.(f).
<Xl
ei,c(f)
"=1
belongs to H.
In view of the result on convex sets in §2 of the Appendix to Chapter 4, it
follows that the above "average" in fact lies in the closure of the convex set
generated by the image f(X), i.e. the smallest closed convex set containing
f(X) .
21. Let E = L'(Il, C) where X = [0, 1] and Il is Lebesgue measure on the algebra
of Borel sets. For each Borel set A let
(a) Show that Iml is Lebesgue measure itself. (b) Show that m is an E-valued
measure which cannot be written as J-lj . [Hint: View dm as a functional
on step functions, say real valued, so that for any step function cp and
measurable set A we have
22. Let E be a Banach space. Let P denote the set of all partitions, i.e. collec-
tions 1t consisting of a finite number of disjoint measurable sets of finite
measure. We let 1tl ~ 1t if every element of 1t is, up to a set of measure 0, the
union of elements of 1t 1. For each 1t E P and IE !l'1(J-l, E) we define
Ix = T,J = L [J-lj(A)/J-l(A)]XA
A.x
(b) Prove the same thing replacing 1 by p for 1 < p < 00.
CHAPTER VIII
Some Applications of
Integration
Suppose first we deal with functions f, g on the real line. We shall study
their convolution, defined by the integral
Ln f(x)g(y - x) dx = f
f(x)g(y - x) dx.
even Coo, and the resulting convolution is also continuous or Coo. To see
this one must be able to take a limit or differentiate under the integral
sign, and the next section gives basic conditions under which this is
legitimate. We shall see several examples after the main approximation
theorem is proved in Theorem 3.1.
We now come to the basic tests for absolute convergence of the
convolution integral.
Theorem 1.1. Let f, g E 21 (Rn). Then for almost all y E Rn the function
x ~ f(x)g(y - x)
f * g(y) = ff(x)g(y - x) dx
The integrals converge absolutely, and we have the trivial estimate from
[VIII, §2] CONTINUITY AND DIFFERENTIATION 225
Lemma 2.1. Let X be a measured space with positive measure J1.. Let
U be an open subset of Rn. Let f be function on X x U. Assume :
(i) For each y E U the function Xf-+ f(x, y) is in g>1(J1.).
(ii) For each x E X and Yo E U, we have
XH f(x, y),
Lemma 2.2. Let X be a measured space with positive measure J1.. Let
U be an open subset of Rn. Let f be a function on X x U. Assume:
(i) For each y E U the function XH f(x, y) is in 21 (J1.).
(ii) For each y E U, each partial derivative Djf(x, y) (taken with respect
to the j-th y-variable) is in 21 (J1.).
(iii) There exists a function f1 E 2 1(J1.) such that for all y E U,
Using the mean value theorem and (iii), together with the dominated
convergence theorem, we conclude that the right-hand side has a limit,
equal to
[VIII, §3] DIRAC SEQUENCES 227
[As in the previous proof, we have to use the device of taking a sequence
{hd to apply the dominated convergence theorem in its standard form.]
f
have
f * cp(y) = f(x)cp(y - x) dx.
Lemmas 2.1 and 2.2 show that f * cp is Coo, and allow us to differentiate
repeatedly under the integral sign.
f
~
The third condition shows that for large k, the volume under CPk is concen-
trated near the origin. Thus in one variable, the sequence looks like this:
228 SOME APPLICA nONS OF INTEGRATION [VIII, §3]
DIR 3s. Each q>k has compact support, and given b, the support of q>k is
contained in the ball of radius b, centered at the origin, for all
k sufficient large.
f q>(x) dx = 1.
and by DIR 2,
f
Hence
We then write
The integral over Iyl < 1> is then bounded by B. For the other integral
with Iyl ~ 1>, we use DIR 3 to conclude that this integral is bounded by
211fliooB for k sufficiently large. This concludes the proof.
for Ixl ~ 1,
so that
f1
-1
<Pk(X) dx = 1.
230 SOME APPLICATIONS OF INTEGRATION [VIII, §3]
We define CPk(X) = °
if x is outside the interval [-1, 1]. It
exercise to show that {cpd is a Dirac sequence. We have
IS an easy
and therefore CPk * f is a polynomial. By Theorem 3.1 the sequence {Cf>k *J}
converges to f uniformly on [0, 1], thus proving Weierstrass' theorem.
Sn,J(x) =
n
L Ck eikx where Ck = -1 f" .
f(t)e- lkt dt.
k=-n 2n: _"
1 n-l m 1
Kn(x) = - L
2n:n m=O
L e
k= -m
ikx = -(Do
n
+ ... + Dn-d·
[VIII, §3] DIRAC SEQUENCES 231
1 sin 2 nx/2
Kn{x) = -2 . 2 x/2'
nn sm
Then Pr{O) satisfies the three conditions DIR 1, DIR 2, DIR 3 where k is
replaced by rand r --+ 1 instead of k --+ 00. In other words :
f~It Pr{O) dO = 1.
DIR 3. Given B and 15, there exists ro , 0 < ro < 1, such that if ro < r < 1,
then
f- fit
6
-It
Pr +
6
Pr < B.
P{O) = ~ 1 - r2
r 2n 1 - 2r cos 0 + r 2 '
Theorem 3.1 concerning Dirac sequences applies to the family {Pr }, again
letting r --+ 1 instead of k --+ 00. In other words, let f be a bounded
232 SOME APPLICATIONS OF INTEGRATION [VIII, §3]
°
function on the circle, that is f«(}) is periodic, as usual. We want to find
a function on the disc, that is a function u(r, (}) with ~ r ~ 1, satisfying
the Laplace equation du = 0, where d is the Laplace operator, given in
polar coordinates by
Example 4 (The Heat Equation for the Laplace Operator). For t > 0,
and a real variable x, define
1
K t (X ) -- K( x, t ) -_ (4nt)1/2e -x 2 /4t
.
°
Then {K t } is a Dirac family, replacing k by t in the definition of Dirac
sequences, and letting t -4 instead of k -400. We define the heat opera-
tor to be
on functions of two variables (x, t). You can easily verify that HK = 0.
Thus K satisfies the heat equation. On R", we can define K t and H
[VIII, §3] DIRAC SEQUENCES 233
similarly, by
1
K t (X ) -- K( x, t ) -_ (41tt)"/2e -x 2 /4t
.
Here x E R" is an n-tuple, and x 2 is the dot product of x with itself. The
heat operator is then written
H=d--
o
ot'
et + e- t et _ e- t
cosh t = 2 ' sinh t = 2 '
cosh t 1
coth t = -'-h- ' cscht= ~h .
sm t sm t
A - (
t -
coth t
-csch t
-csch
coth
t).
t
(b) Let
f.
operation
F * f(x, t) = F(x, y, t)f(y) dy.
Proof. The proof is identical to the proof of Theorem 3.1 except for
the following final modification. At the very last estimate, for k large, the
support of lPk is contained in the ball of radius b, whence the integral
expressing lPk * f(y) - f(y) is concentrated on that ball, and is obviously
estimated bye, thus proving our theorem.
Proof We have
CfJk * f(x) =F 0,
Corollary 3.4. Let f E 2 P(R n ) for 1 ~ p < 00. Then {CfJk * J} is U con-
vergent to f
Since
for k large. The last of the three terms in our estimate above is < e,
thus concluding the proof.
236 SOME APPLICATIONS OF INTEGRATION [VIII, §4]
xERn,
is bounded for Ixl sufficiently large. Here as in the rest of this chapter,
Ixl is the euclidean norm of x. Equivalently, the preceding condition can
be formulated by saying that for every polynomial P (in n variables) the
function Pf is bounded, or that the function
{
e-i/(X- a)(b-Xl
if a < x < b,
f(x) = 0
otherwise.
so that
(MPf)(x) = Xfl ... x:nf(x).
In what follows, we shall take the integral of certain functions over Rn,
and we use the following notation:
f
f(X) dx = fan
f(x) dx = fOCi ...
-00-00
foo f(x 1 , · · · ,xn ) dX 1 ••• dx n •
C
If(x) I ~ (1 + xi)"'(1 + x;)'
and we can view the integral as a repeated integral, the order of
integration being arbitrary. The justification is at the level of elementary
calculus. Furthermore, we differentiate under the integral sign, using the
formula
a: f K(x, y) dx = f a: K(x, y) dx
j j
f
f(x - y) dx = ff(x) dx, f f( -x) dx = f f(x) dx.
The general change of variables formula, of which these are but elemen-
tary cases, will be proved in detail in Chapter XXI, §2.
Finally, for normalization purposes, we shall write formally
This makes some formulas come out more symmetrically at the end.
We now define the Fourier transform of a function f E S by
Remember that xy = x· y.
Since
we see that we can differentiate under the integral sign, and that
By induction, we get
and integrate by parts with respect to the j-th variable first. We let
u = f(x) and
Then v = ie- ixy and the term uv between -00 and +00 gives zero contri-
[VIII, §4] SCHWARTZ SPACE AND FOURIER TRANSFORM 239
The expression for MP! in terms of the Fourier transform of DPf, which
is in S, shows that M p! is bounded, so that ! tends rapidly to zero
at infinity. Similarly, one sees that MP Dq! is bounded, because we let
g = Dqf, g E S, and
f * g(x) = f
f(t)g(x - t) d 1 t.
This integral is obviously absolutely convergent, and the reader will ver-
ify at once that the map
Furthermore,
= fff(t)g(x-t)e-iXYdltdlx,
We change variables, letting U = x - t, d1u = d1x and see that our last
integral is equal to
Dh(y) = - yh(y).
g(x) = h(ax).
Then
g(y) = ~h(y/a).
an
This is proved trivially, changing the variable in the integral defining the
Fourier transform.
= f f(t)g(t + y) d1t.
242 SOME APPLICA nONS OF INTEGRATION [VIII, §5]
g(u) = ~h(u/a),
a
f
and hence
f
f(x)e-'XYh(ax) d1x =
A . 1t -+
f(t) ann (t y)
a- d1t
f
= f(au - y)h(u) d 1u
(t
U= - - ,
+ y)
a
Theorem 5.2. For every f E S there exists a function <p E S such that
f = <p. If f, 9 E S, then
Proof First, it is clear that applying the roof operation four times to
a function f gives back f itself. Thus f = <p, where <p = F' '''' . Now to
prove the formula, write f = <p and 9 =.[1. Then j = <p- and 9 = ljI- by
Theorem 5.1. Furthermore, using Theorem 4.2, we find
as was to be shown.
We observe that the first step of the proof in Theorem 5.1 yields
f j(x)g(x) dx = f f(x)g(x) dx
[VIII, §6] THE POISSON SUMMATION FORMULA 243
by letting y =
definitions
°on both sides. Furthermore, we have directly from the
<I, g) = <i. g)
and hence
Proof We have
A function g on R" will be called periodic if g(x + k) = g(x) for all k E Z".
We let T" = R"/Z" be the n-torus. Let g be a periodic Coo function. We
define its k-th Fourier coefficient for k E Z" by
Ck = f
T"
g(x)e- 21tikx dx.
The integral on T" is by definition the n-fold integral with the variables
(Xl' ... ,X") ranging from °to 1. Integrating by parts d times for any
integer d > 0, and using the fact that the partial derivatives of g are
bounded, we conclude at once that there is some number C = C(d, g)
such that for all k E Z" we have Ickl ~ Clllkll d, where IIkll is the sup norm.
Hence the Fourier series
g(x) = L cke21tikx
kE Z"
converges to g uniformly.
244 SOME APPLICATIONS OF INTEGRATION [VIII, §7]
f
this section by
J(y) = f(x)e-21tiXY dx
a"
L:
me zn
f(m) = L
me Z"
J(m).
Proof Let
g(x) = L f(x + k).
ke Z"
L Cm = g(O) = L f(k).
me Z" ke Z"
L: f
T" ke Z" T"
f
ke Z" T"
= f(X)e-21timx dx = J(m).
a"
Then cp has compact support, and is certainly in 'pl . Its Fourier trans-
form is therefore given by the integral
<j?(y) = r
JIxl ~ 1
e-21tix'Y dx = r
JIxl ~ 1
e21tix .y dx.
Thus
for t --+ 00 .
(The sign « means that the left-hand side is bounded in absolute value
by a constant times the right-hand side, namely O(t- 1/2).)
I o
1t e it cos 9 dO = fl e itu
-1 ~
du .
I I
o
e itu
~
1
du = O(t-l/2).
246 SOME APPLICATIONS OF INTEGRA nON [VIII, §7J
f 1
o
e itu
~
1 g(u) du,
where g(u) = I/ J1+u is Coo over the interval. Integrating by parts (cf.
also Lemma 7.3), we see that the desired integral satisfies the bound :
f b
eltu
. 1
du = O(t- 1/2 ).
a ~
Proof Let v = 1 - u, and then tv = r. Then the integral is estimated
by the absolute value of
t- 1/2fB e ir _ 1_ dr
A Jr '
where 0 ~ A ~ B. But writing e ir = cos r + i sin r, and noting that IIJr
is monotone decreasing, we see that the integral on the right-hand side is
uniformly bounded independently of A, B. This proves the lemma, and
also concludes the proof of the proposition for n = O.
r
interval [a, bJ. Then the Fourier transform satisfies the estimate
Theorem 7.4. Let qJ be the characteristic function of the unit disc in the
plane. Then
Lt
r
[setting ur = t, r du = dt] = cos(ts)(l - (t/ rftl/2 dt dr
Estimating this last integral as in Lemmas 7.2 and 7.3 concludes the
proof.
L
00
(a) Show that 0 satisfies the conditions of a Dirac family, except for the
positivity condition. Note that 0, is periodic in the variable x.
(b) For f continuous periodic, show that 0, *f converges to f uniformly as
t -+ O.
(c) Show that 0 satisfies the heat equation, and so does 0, * f(x), as a function
of (x, t). The heat equation is normalized here in the form
8. Let f E g>'(R) n g>2(R) and suppose the function xf"(x) is in g>'(R). Prove
that there is a C' function g such that g = f almost everywhere, and give a
formula for g.
[VIIi, §8] EXERCISES 249
to. Prove that there is a function h in the Schwartz space such that h A has
compact support and h(O) # O. Show that for such a function, hA = h for all
A sufficiently large (notation as in Exercise 9).
11. Let f E 2' 1 (R). Prove that!" is uniformly continuous on R.
12. The lattice point problem. Let N(R) be the number of lattice points (that is,
elements of Z2) in the closed disc of radius R in the plane. A famous
conjecture asserts that
for every e > O. It is known that the error term cannot be O(Rl/2(log R)k) for
any positive integer k (result of Hardy and Landau). Prove the following
best-known result of Sierpinski-Van der Corput-Vinogradov-Hua:
Sketch of Proof. Let cP be the characteristic function of the unit disc, and put
CPR(X) = cp(x/R).
Let IjJ be a COO function with compact support, positive, and such that
Let
fR'
ljJ(x) dx = 1.
Then N.} is a Dirac family for e -+ 0, and we can apply the Poisson summa-
tion formula to the convolution CPR >I< 1jJ. to get
LZ2 CPR
me
>I< 1jJ.(m) = L q;R(m)~.(m).
me Zl
= nR2 + L nR2q;(Rm)~(em).
m .. O
Splitting off the term with m = 0 on the right-hand side, we find (using
Theorem 7.4):
Therefore we find
IAfI ~ Cllfil
for all f E Cc(X). Thus A is bounded if and only if A is continuous for the
norm topology.
A linear map A of CAX) into the complex numbers is said to be
positive if we have Af ~ 0 whenever f is real and ~ o.
A(bg) ± Af ~ 0
and IAfl ~ bA(g). Thus Ag is our desired bound.
[IX, §1] POSITIVE AND BOUNDED FUNCTIONALS ON CAX) 253
Hence
Ag = A{inf(f1' g)) + A{g - inf(f1' g))
~ A+f1 + .A.+f2·
Taking the sup on the left implies that .A.+(f1 + f2) ~ A+f1 + .A.+f2' thus
proving that .A. + is additive.
We extend the definition of .A.+ to all elements of Cc(X, R) by expres-
sing an arbitrary f as a difference
sup cP =f
<pee!>
Assume that if cP, 1/1 E <1>, then sup(cp, 1/1) E <1>. Given e, there exists cp E <I>
such that Ilf - cpll < e. If A is a positive functional on Cc(X), then
Af = sup Acp.
<pee!>
then Ilf - <p11 < 8. This proves the first part of the theorem. The last
assertion follows by the continuity of A. (Lemma 1.1).
We shall prove the converse (Riesz' theorem), and first obtain a positive
measure from a positive functional.
If f is a function on X, we define its support to be the closure of the
set of all x such that f(x) # O. Thus the support is a closed set. We
denote it by supp(f).
We use the following notation as in Rudin [Ru 1], which we more or
less follow for the proof of Theorem 2.3. If V is open, we write
f -< V
to mean that f is real, f E CAX), 0 ~ f ~ 1, and supp(f) c V. Similarly,
if K is compact we write
K -<f
Remark 1. From (ii) and the remarks before Theorem 2.3, we see at
once that for any compact K we have
Note that the property that Jl{V) = sup Jl{K) for compact K c V is satis-
fied by open V. This is convenient because even in pathological situa-
tions, we are able to define the measure of Theorem 2.3 on Borel sets
rather than on a more restricted algebra (e.g. that generated by the
compact sets, as is sometimes done in the literature). Observe that if we
know that property (iv) is satisfied by all sets A of finite measure, then it
follows at once for any a-finite A. Indeed, let {An} be a disjoint sequence
of sets of finite measure, and let Kn be a compact subset of An such that
Jl{A n - Kn) < e/ 2n. Then K 1 U ... u Kn is compact, and Jl{K 1 u ... u Kn)
tends to the measure of U
An as n -+ 00, whether this measure is finite or
not.
Lemma 2.4. Let,.1. be a positive functional on Cc{X). For each open set
V, define
Jl{V) = sup ,.1.g for g -< v.
258 LOCALL Y COMPACT SPACES [IX, §2]
(1)
Let h -< VI u V2 . Let <D be the family of all functions SUp(gl' g2) with
gi -< V; for i =:= 1, 2. Then <D is closed under the sup operation (on a finite
number of elements), and we have
sup g = Xv, vv 2 •
ge<!>
gi-<V;, i=I,2.
Taking the sup over all h on the left yields our inequality (1).
Now let {An} be a sequence of subsets of X, and A = U An· Let v,.
be open, An C v,., and
Remark. The special role played by compact and open sets in con-
structing the algebra of measurable sets and the measure on it will stem
from the following property:
The right inequality is obvious. As for the left one, let W be the set of x
such that f(x) > 1 - 1>. Then K c W Let 9 -< W be such that
Jl(W) ~ Ag + 1>.
We have
(1 - I»g ~ f whence (1 - I»Ag ~ Af
Hence
Af
Jl(K) ~ Jl(W) ~ Ag + I> ~ -1 -- I> + 1>.
Our remark shows the main idea of what follows. We recover charac-
teristic functions of certain sets by squeezing them between compact and
open sets, and comparing them with functions f E Cc(X) on which the
given functional is defined. The K and V allow us to use the old
technique of lower and upper sums respectively. We first have to recover
the measure itself, however, and we proceed to do this. For convenience
of notation, the outer measure Jl described in Lemma 2.4 will be called
the outer measured determined by A.
Then d is an algebra containing all compact sets and all open sets of
260 LOCALLY COMPACT SPACES [IX, §2]
We may assume that /1(V) > O. To cover the case when /1(V) = 00, we let
r be a real number such that 0 < r < /1 (V). There exists f such that
r < Af ~ /1(V).
and therefore r ~ /1(K). This proves that /1(V) = sup /1(K) for K compact
c V. In particular, if /1(V) is finite, then V E d.
Before proving that .91 is an algebra, we find it convenient to have the
finite additivity. Actually, it is no more troublesome to prove the count-
able additivity. First we prove that if K 1, K 2 are disjoint and compact,
then
Then
/1(K i ) + /1(K 2 ) ~ /1(W n Vd + /1(W n V2 )
~ Ag i + Ag2 + 26
= A(gi + g2) + 26
L J.I.(An) = J.I.(A).
00
n=l
This gives the countable additivity and also proves that A E d if J.I.(A) is
finite.
We can now prove that d is an algebra. Clearly the empty set is in
d. If A 1 , A2 Ed, we can find compact sets K 1 , K2 and open sets V1 ,
V2 such that
Ki C Ai C V; (i = 1, 2)
J.I.(V; - KJ < s.
Since
we get
The measure of Theorem 2.3 (or Theorem 2.6) will be called the
associated measure of A, or the measure determined by A. In applications,
one needs it mainly on the Borel sets (or the completion of the Borel
sets).
Partitions of Unity. Let K be compact and let {U1, ... ,Un} be an open
covering of K . There exist functions;; (i = 1, ... ,n) such that;; --< Ui
and such that
n
L J;(x) = 1, all x E K.
i=l
Let
f1 = gl'
f2 = g2(1 - gl),
The functions {/;} are said to form a partition of unity over K, subor-
dinate to the covering {U 1 , ... ,Un}.
<P = L CiXA;
i=1
and also that f ~ C i on V;. [For instance, cut an interval containing the
image of f into eI2-subintervals, say half closed to make them disjoint,
and let c; be the right end point of each subinterval. Let Ai be the
inverse image in K of the i-th subinterval. Let Ci = c; + e12. For each i
let Jt; be open => Ai such that f ~ Ci on Jt;, and shrink Jt; to an open
V; => Ai satisfying (*).]
Let {hI' ... ,h n } be a partition of unity over K subordinate to
{VI' ... ,y"}. Then fh i has support in V;, and fh i ~ cih i . Furthermore,
K « inf(l, L h;), so that
L (c
n
~ j + c)J.l(l';) - cJ.l(K)
j=1
This proves our inequality since the integral of f over K is the same as
the integral of f over X, and concludes the proof of our theorem.
Proof Theorems 2.3 and 2.7 show that the map J.ll-+ dJ.l is surjective.
Let J.ll' J.l2 be positive measures satisfying conditions (ii), (iii), and (iv) of
Theorem 2.3 and assume that dJ.ll = dJ.l2. To show that J.ll = J.l2' it
suffices to prove that the two measures coincide on compact sets, because
then (iv) shows that they coincide on open sets, and (ii) shows that they
are equal on Borel sets. Let K be compact, and let V be an open set
containing K such that
This proves one inequality, and the other follows by symmetry. Thus we
get a bijection between Mo and the set of positive functionals on Cc(X).
This bijection is obviously additive. This proves our corollary.
Proof. The step functions are dense in U(/l), and it thus suffices to
prove that for any set A, of finite measure given e we can find some
f E Cc(X) such that
We take K compact, V open such that K cAe V, and /l(V) < /l(K) + e.
Let K -< f -< V. Then
and
Corollary 3.2. If f E 2 1 (/l) and Jf ({J d/l = 0 for all ({J E Cc(X), then
f = 0 almost everywhere.
and
mf-+dm
is then a linear map of Mo into the dual space of Cc(X) (sup norm),
because we have the inequality
Idml ~ Ilmll,
or written out explicitly,
In fact:
Theorem 4.2 (Riesz Theorem, Part 3). The map m f-+ dm is a norm-
preserving isomorphism between the space of regular complex Borel mea-
sures on X and the dual space of Cc(X) (with sup norm topology).
is regular and represents A., i.e. we have A. = dm, thus proving that our
map is surjective.
[IX, §5] LOCALIZATION OF A MEASURE AND THE INTEGRAL 269
To show that the map is injective, we have to prove that its kernel is
O. Suppose that dm = O. Let Jl = Iml and dm = h dJl with Ihl = 1. Then
<f, h\ = 0 for all ! E Cc(X). But Cc(X) is U-dense in !l'l(Jl, C) by Theo-
rem 3.1. We have the inequality
for all ! E !l'l(Jl, C). It follows that <cp, h)1l = 0 for all step functions cp,
whence h is equal to 0 almost everywhere. Since Ihl = 1 we must have
Jl(X) = 0, thus proving m = O.
Finally, write again dm = h dJl with Jl = Iml and Ihl = 1. Let A = dm.
We have to show that Jl(X) ~ IAI. By Lusin's theorem, §3, we can find a
function g E Cc(X) such that g = Ii except on a set Z of measure < e, and
such that Igl ~ 1 on Z. Consequently
~ Jl(X) - 2e.
This proves the desired inequality, and concludes the proof of the
theorem.
which does not depend on the choice of K. This function JlA.' defined on
Borel-measurable subsets A of compact sets, will also be called a mea-
sure, and more specifically the measure associated with A. For instance,
suppose Jl is a positive Borel measure on X , and! is a measurable
function, bounded on each compact set, then A = ! dJl defines such a
functional, which has such an associated measure. The measure JlA. could
also be called the direct limit of the measures JlK' taken over all compact
sets K.
We contend that this sum is independent of the choice of a(i), and also
of the choice of partition of unity. Once this is proved, it is then obvious
(see Exercise 10) that A is a functional which satisfies our requirements.
We now prove this independence. First note that if ~'(i) is another one
of the open sets ~ in which the support of hJ is contained, then hJ has
support in the intersection ~(i) II ~'(i» and our assumption concerning
our functionals Aa shows that the corresponding term in the sum does
not depend on the choice of index a(i). Next, let {gk} be another parti-
tion of unity over K subordinated to some covering of K by a finite
number of the open sets ~. Then for each i,
whence
(The sum is taken over all i, but is in fact finite for any given x in view
of PU 3.) As a matter of notation, we often write that {(Jt;, l/J;)} or sim-
ply {l/JJ is a partition of unity if it satisfies the previous four conditions.
In the proof of the next theorem, we use the facts (trivially proved)
that if a space X has a countable base, then any open covering has a
countable subcovering, and any base contains a countable base.
Theorem 5.3. Let X be locally compact Hausdorff, and assume that the
topology of X has a countable base. Then X admits continuous parti-
tions of unity, subordinated to a given open covering d/I.
Proof. Let V 1 , V 2 , ••• , ••• be a base for the open sets, such that each
Vi is compact. We construct first inductively a sequence A 1, A 2 , • • • of
compact sets whose union is X and such that Ai is contained in the
interior of Ai+!. We let A1 = V1. If we have constructed Ai inductively,
then we let j be the smallest integer such that Ai is contained in
Let Int abbreviate interior. For each point x of Ai+l - Int(Ai) we can
find a pair (Wx, Vx ) of open sets containing x such that Wx c Wx c Vx ,
such that Vx is contained in Int(A i+ 2 ) - Ai-I, and such that Vx is con-
tained in one of the open sets of the given covering U. There is a finite
number of pairs such that already the open sets Wx cover the compact
set A i + 1 - Int(AJ Taking all such finite collections of pairs for i = 1, 2,
... , we obtain a countable collection of pairs {(~, v,.)} such that the
{v,.} form a locally finite covering of X, the {~} is also an open cover-
ing, and ~ c Vk . Let hk be such that ~ -< hk -< v,. (see the beginning of
§2 for the notation -<.) Let
co
h= Lh
k=1
k•
Let
in the sense that the sum is absolutely convergent, and is equal to the
integral on the right.
Proof. Let
n
fn = L hJ
i=1
Proof. Let f
E Cc(X) and let K be the compact support of f. Let
ex E A be 1 onK . Let U be an open set containing the support of ex,
and having compact closure V. The restrictions to V of elements of A
form an algebra, which clearly satisfies the hypotheses of the Stone-
Weierstrass theorem. Therefore the restriction fl V can be uniformly
approximated by elements of A IV. Denote by II Ilv the sup norm over
V. If we can approximate f by an element PEA over V, say
Ilf - Pllv < t:,
then
lIexf - exPllv < t:llexll,
274 LOCALLY COMPACT SPACES [IX, §7]
(x, y) H <p(x)t/!(y),
and call it the product function. The set of finite sums of product
functions is an algebra, which we shall call the algebra generated by the
product functions.
Theorem 6.3. Let X, Y be locally compact Hausdorff spaces and let Jl,
v be positive a-regular Borel measures on X and Y, respectively. Assume
that X, Yare a-finite with respect to these measures. Then all functions
in Cc(X x Y) are in ,21(Jl ® v), and there exists a unique a-regular
Borel measure on X x Y which restricts to Jl ® v on 8l(X) ® 81(Y).
Proof. Lemma 6.1 shows that functions in Cc(X x Y) are (Jl ® v)-
measurable, and combined with Fubini's theorem shows that these func-
tions are in ,21(Jl ® v). The map
fHf f d(Jl® v)
XxY
5. Let J.l, v be regular Borel measures on R". Define the convolution J.l * v by
where (J: R" x R" -> R" is the sum. Show that J.l * v is regular.
6. Assume that X is (J-compact. Let J.l be a regular Borel measure on X. If A is
measurable, show that there exists a closed set B c A and an open set V:::J A
such that J.l(V - B) < e.
7. Assume that every open set in X is (J-compact. If v is a positive Borel
measure which is finite on compact sets, show that v is regular. [Hint: Show
that v = J.l if J.l is the regular measure associated with dv as in the text. Do it
first for open sets.]
8. (a) Let M denote the Banach space of complex regular Borel measures on R".
If m, m' are in M, show that for f E C,(R") the integral
[Note: This obvious extension of the text, and of Theorem 5.4 in particular,
is useful when dealing with manifolds. Cf. for instance Chapter XXIII, §3, §5,
and §6.]
10. Verify in detail the "obvious" fact in the proof of Theorem 5.l that A is a
functional, in particular that for each compact set K there is a number AK
such that for any f E C,(X) with support in K we have IAfl ~ AK Ilfll.
11. Let J.l be a regular positive measure on R. (a) Show that the functions of type
e- Xg(x) (where g is a polynomial) are dense in :;el(R+, J.l). (b) Show that the
276 LOCALLY COMPACT SPACES [IX, §7]
for all real t, show that f(x) = 0 for J-l-almost all x. [Hint : By a Fourier
series argument, show that
t
i.e.
f(x)x" dJ-l(x) = 0
for all n ~ 0, then f(x) = 0 for J-l-almost all x . Note : Actually, (b) implies (a).
[Hint for (a) : Show that the integral in Exercise 12 is analytic in t for t at a
distance ;:;; c/q from the real line, and 0 near the origin. You can also use
the exercises at the end of Chapter III.]
Examples. Taken dJ-l(x) = e- x2 dx. We get the completeness of the Her-
mite polynomials. For the Laguerre polynomials, one takes dJ-l(x) = h(x) dx,
where h(x) = 0 if x < 0 and h(x) = e- if x ~ o. And similarly for the other
X
n
!~~ 1 i~
" f(xJ = f f dJ-l.
Let N(A, n) be the number of indices i;:;; n such that Xi E A. Prove that
. N(A, n)
lIm - - =J-l(A).
p.- oo n
[IX, §7] EXERCISES 277
for all n E Z.
Prove that J1 = 0.
CHAPTER X
Riemann-Stieltjes Integral
and Measure
Let us start with a finite interval [a, b] on the real line. To each
partition
P = [a = Xo, Xl' ... ,X. = b]
we associate its size,
Let
f: [a, b] --+ E
.-1
Vp(f) = L If(x k +1) - f(Xk) I·
k=O
[X, §1] FUNCTIONS OF BOUNDED VARIATION 279
where the sup (least upper bound if it exists, otherwise 00) is taken over
all partitions. If V(f) is finite, then f is called of bounded variation, and
f is bounded.
Examples. If f is real valued, increasing, and bounded on [a, b], then
f is obviously of bounded variation, in fact bounded by f(b) - f(a).
If f is differentiable on [a, b] and f' is bounded, then f is of bounded
variation (mean value theorem). This is so in particular if f is of class
CI .
Proof For (i), we note that if x < y, then we can always refine a
partItIon of [a, y] to include the number x. Furthermore, if p i is a
partition refining P, then
Then (i) follows at once. For (ii), we again use the fact that a partition of
[a, y] can be refined to contain x. Finally, suppose that f is continuous.
By (ii), the continuity from the right of J.f amounts to proving that
lim V(f, x, y) = o.
Suppose that the limit is not O. Then there exists a number fJ > 0 such
that
V(f, x, t) > fJ for all x < t ~ y.
If(x 1 ) - f(x)1 by
Now we repeat this procedure with Y replaced by Y1' and find Y2 with
x < Y2 < Y1 such that
Since V(f, Y., y) ~ V(f, x, Y), this gives a contradiction, concluding the
proof.
and 2h = Vf - f + f(a).
If f is continuous, so are 9 and h by Theorem 1.1 (iii). In any case,
°
g(a) = h(a) = and the two formulas of the theorem are valid. There
remains only to prove that g, h are increasing. Let a ~ x ~ y ~ b. Then
by additivity of Proposition 7.1(ii),
Denote by a(P) the size of the partition P. We say that the limit
lim S(P, c)
a(P)-O
exists if there exists LEG such that given e there exists {) such that
whenever a(P) < {) then jS(P, c) - Lj < e. If
lim S(P, c, f, g)
a(P)-O
ffd9.
When g(x) = x, then the integral is just the Riemann integral, and we call
the function Riemann integrable.
282 RIEMANN -STIEL TJES INTEGRAL AND MEASURE [X, §lJ
r r
as a function of g, that is
f d(ag) = a f dg.
r
Proposition 1.3. Assume f E RS(g), and 9 of bounded variation. Then
I f dg I ~ IlfII V(g),
where IlfII is the sup norm of f on [a, b].
r r
case, the formula for integration by parts holds, namely
f dg + 9 df = f(b)g(b) - f(a)g(a).
where
n-I
S(Q, g, f) = g(a) [f(co) - f(a)] +L g(xd [f(cd - f(ck-d]
k=1
with the intermediate points a, x I' ... ,Xn - I , b. When the size of P
approaches 0, so does the size of Q, and by hypothesis, the sum S(Q, g, f)
approaches the integral J~ g df, thereby completing the proof of the
proposition.
Proof. Given e let D be such that if Ix - yl < D then If(x) - f(y)1 < e.
Let P, P' be partitions of size < D. To estimate IS(P, c) - S(P'. c')I, we
may assume without loss of generality that P' is a refinement of P. Thus
it suffices to prove two estimates: if P' = P but we change the choice of
intermediate points c to c', then the difference of the sums is small; and
if P' is obtained from P by inserting one more point in the partition,
then again the difference of the sums is small. As to the first step, letting
P = P', we have
which gives us the desired estimate. Secondly, suppose that P' is ob-
tained from P by inserting one point, say xi with Xj ~ xi ~ x j +!. Then
the size of P' is still < D. By the first step, to get the desired estimate for
IS(P, c) - S(P', c')1 we may assume without loss of generality that for
i 1= j we have x; = Xi' that Cj = xj, and xi is also selected as the interme-
diate point for the two intervals [Xj' xj] and [xi, xj + l ] of the partition
P'. Then
S(P, c) - S(P', c') = 0,
r r
tiable on [a, b] with Riemann-integrable g'. Then f E RS(g), and
f dg = f(x)g'(x) dx,
where the integral on the left is the RS-integral, and the integral on the
right is the usual Riemann integral.
r
we obtain a bounded functional
ff--+ f dg.
t r fdg = fdg,
the right side being independent of the choice of [a, b]. Suppose that
there exists a number B > 0 such that
Indeed, suppose f(x) ~ c > 0 for infinitely many x. Then f(x) ~ cl2 (say)
for all x sufficiently large, otherwise f would not be in Ll(R) n BV(R).
Under these circumstances, in a situation when integration by parts is
valid for finite intervals as in Proposition 1.4, the extra terms
vanish.
Proposition 1.8.
(a) Let f be of bounded variation. Then the set of points of discontinu-
ity of f is countable.
r
(b) Assume f continuous at a and b with a ~ b. Then
Proof We leave part (a) as an easy exercise (see Exercise 4), done by
a routine estimate. As for part (b), we note that the constant function 1
is continuous on [a, b], and all Riemann-Stieltjes sums give the same
value f(b) - f(a), as desired.
r r
creasing. Then there exists c E [a, b] such that
r
the Riemann sum
•
S = L f(xi)g(Xi)(Xi -
i=l
Xi-I) approximates f(x)g(x) dx .
•
Bi = L
k=i
g(Xk)(Xk - xk-d,
n
S = f(xo)BI + L Bi(J(Xi) -
i=l
f(X i- I »)·
XH J: g(t) dt
lim
A-+oo
f""
-00
f(x)e iAX dx = O.
We may consider next the variation when we must take an end point
into consideration.
We now introduce the condition of bounded variation. According
to Zygmund [Zy], Dirichlet was the first who proved Fourier series
288 RIEMANN -STIEL TJES INTEGRAL AND MEASURE [X, §2J
e- itx
-/(x)-.- Joo +:-1 foo e- itx dl(x).
It -00 It -00
lim
.4-'00
foo-00
l(x)e iAX dx = O.
Proof. Let 9 E C~(R) be such that III - gill < e. Then the integral can
be written and estimated in the form
The first integral is bounded in absolute value bye. The second integral
is < e for large A by Proposition 2.1. This proves the proposition.
sense that
f" "(x) = lim fA (x).
A-+oo
fA(X) =
1 fA "
M:
itx 1
f (t)e dt = -
foo f(y)
sin A(x - y)
dy.
y2n -A n - 00 x-y
Proof. We have
The next theorem gives conditions for f /\/\ = f-. It involves an appli-
cation of the Bonnet mean value theorem. The theorem is in Titchmarsh
[Ti], who attributes the result to Prasad, Pringsheim, and Hobson.
1
reduced to proving
n . 00
sin Ay
-2 f (0) = lim f(y) - -dy,
A-+ oo 0 Y
290 RIEMANN -STIEL TJES INTEGRAL AND MEASURE [X, §2]
.1
hm
A--+ oo
00
b
f(y) - Ay dy = 0
sin -
Y
AYd _ IA sin u d
Io -sinY
b
- y-
oU
-- u-+-
7t
2
as A -+ 00 .
Ib
o
f(y) sin Ay dy =
Yo"
I" Y
Ib
+ f(y) sin Ay dy.
I"o
f(y) sin Ay dy
y
= f(J) I"
c
sin Ay dy
Y
= f(J) fA" sin u duo
Ac U
Since If(J)1 < B, and since the integral over [Ac, AJ] is bounded, we have
concluded the proof of the theorem.
.
CI(X) = - fcc -cos-t dt and
.
SI(X) = - fcc -sin t dt.
x t x t
gy
cos Ay
() = - -
y
and G(y) = - f y
CC cos At
- - dt = cl(Ay)
t
.
r r
for c5 < a < 00, we get:
We now estimate each one of the three terms on the right side.
Since f is bounded by hypothesis, and G(a) --+ 0 as a --+ 00, we see that
f(a)G(a) --+ 0 as a --+ 00 , so the first term approaches 0 as a --+ 00.
292 RIEMANN-STIELTJES INTEGRAL AND MEASURE [X, §2]
We claim that f(b)G(b) --+ 0 as b --+ O. To see this, we split the inte-
.( )_II
gral:
CIX--
cos-tdt -
- cos-tdt
- fro
x t i t
= f x cos t - 1d
t+ fX -dt -cll
.( )
I t i t
f: G(y) df(y)
r r
exists. There remains only to prove that the tail end
lim
a-+oo
fa G(y) df(y)
0
exists.
This concludes the proof of existence for the improper integral of Theo-
rem 2.6.
Estimate of the integral. We shall now prove the stated estimate. We
let A. = 1/(1 + e). We decompose the integral:
fro
o
f(y) cos A Y dy =
y
f
0
I/A'
+ fro
I /A'
f(y) cos A Y dy
Y
= f o
lM'
+ [f(y) ci(AY)]lIA' -
fro
I/A'
ci(Ay) df(y)
[X, §2] APPLICATIONS TO FOURIER ANALYSIS 293
= f o
IlA' f(y)
-cos(Ay) dy - f(l / A).) ci(A 1-).)
y
- fro l / AA
ci(Ay) df(y).
We shall estimate each one of the three terms on the right side of the
equality.
For the first term, the integral is taken near zero where
If(y)/y I = O(y'-l)
1 1
« -e -A t / ). for A --+ 00,
). . 1-), 1 2 2
f(l / A ) cl(A )« Ad A1-). ~ A 2t /(1+,)
If b
l / AA
ci(Ay) df(y) I ~ A ;-). VR(f)
for every b, and hence for the integral to 00 . This concludes the proof of
the theorem.
X, §3. EXERCISES
S(x) = f x sin t
-dt.
o t
Show that for A > 0:
If'"
-00
I(y) sin A~ - Y) dyl
x Y
~ 211SI1 VR(f),
or better,
lim fb In dg = fb I dg.
11-00 a a
Assume that g.(x) converges to g(x) for some bounded function g and all
x E [a, bJ. Show that
4. Let I be of bounded variation on [a, bJ. For each x E [a, b] define the jump
Distributions
°
sion as above determines the coefficients a p uniquely. Indeed, suppose
that D = 0. To prove that a p = it suffices to prove that for any a E Rn
we have ap(a) = 0. For a given p we consider a function given locally
near a by
Then
Dqf(a) = {op! ~f p :/; q,
If p=q,
T: C;"(U) -+ C
such that, for every compact set K contained in V, there exists a con-
stant AK and an integer m for which
all qJ E C;"(K).
Let {<pj} be a sequence in C;o(U), such that all <Pj have support in a
compact set K, and such that for every p, {DP<pj} converges to 0 uni-
formly on K. Then T<pj -+ O.
am = supi Tfl·
the sup being taken for those f E C;o(K) such that 1tm (f) ~ 1. It will
suffice to show that for some m, we have am "# 00. Suppose that am = 00
for all m. Choose fm E C;o(K) such that 1tm(fm) ~ 1, but ITfml ~ m. Let
gm = fm /m. Then
and if k ~ m, then
so
But
then
if qJ E C;'(K).
Furthermore, the map fl--+ T.r induces an injective linear map of U(U)
into the space of distributions on U, because we know from Corollary
9.5 of Chapter VI that if T.r = I'g for two locally integrable functions f, g,
then f is equal to 9 almost everywhere. Thus from now on, we can
interpret locally integrable functions as distributions.
(TD)(ep) = T(Dep).
In particular, if T can be represented by a locally integrable function J,
then
This yields what we want, except for the fact that the indices may not be
j = 1, . . . ,m. But it is trivial to adjust this as desired. All we have to do
is to find for each k and index j(k) such that Bk is contained in Vj , and
then for each j = 1, ... ,m take the sum of those CX k such that j(k) = j, to
obtain qJj.
To get the function t/lk as in the preceding proof, we combine a
function whose graph is indicated below with the square of the euclidean
norm to get a Coo function which is 1 on a ball, and 0 outside another
ball of slightly bigger radius.
Proof Let <p E C:'(U), and let K be the support of <po For each a E K
we can find an open set Ua such that T is zero on Ua . We can cover K
with a finite number of such open sets, say UI , ... ,Um • Let {<pJ be a C" )
partition of unity over K with j = 1, ... ,m, such that supp <Pj is contained
in Uj • Then
m
and T<p = 2:
j;1
T(<pj<p) = 0,
Corollary 2.2. Two distributions which are locally equal everywhere are
equal.
Proof Let K = supp <p, and Q = supp T. There exists an open neigh-
borhood V of K which does not intersect Q and is contained in U . Let
a E C:'(V) be such that a = 1 on K and the support of a is contained in
V Then <p = a<p and
T(<p) = T(a<p).
°
Proof Since <p - IjJ is equal to on an open neighborhood of K, it
follows that the supports of <p - IjJ and T are disjoint, whence we can
apply Corollary 2.3 to conclude the proof.
Tf = T(af).
function f such that f(x) = x (say in one variable x). If T has compact
support, then we can speak of the value T(x) = Tf(x) using the definition
we just made.
Using partitions of unity over a whole open set, one can prove the
following result, left as an exercise.
Let {V;} be an open covering of an open set V in Rn. For each i, let 1i
be a distribution on Vi' and assume that for each pair i, j the restric-
tions of 1i and 1j to Vi n ~ are equal. Then there exists a unique
distribution T on V which is equal to 1i on each Vi'
But
DP(lXjCP) = L ",jqDqcp
Iql;:i!lpl
all cP E C:'(K),
[XI, §3] DERIVA TION OF DISTRIBUTIONS 303
whence
k
IT(cp)1 ~ L Aj Bj 1tm (cp), all cp E C~(K).
j=l
Proof. For the existence, we may restrict ourselves to the case when
D = aDP for some a E C<Xl(U). Then we have
This proves the existence. As for uniqueness, suppose that D* and D' are
differential operators such that
f (D'f)cp f
= (D*f)cp
f ((D* - D')f)cp = °
for all cp, so that (D* - D')f = 0, whence D* = D'.
We shall call D* the adjoint of D. The map D 1-+ D* is an anti-
automorphism of the ring of differential operators; anti because
304 DISTRIBUTIONS [XI, §4]
DT = To D* = TD*
on C:'(U). In particular, if D = D; is the i-th partial derivative, then
D;* = -D; and
D~ = 15,
where D = Dl is the derivative in one variable.
Let
1
g(x, y) = 2n log r
T = L cpDPJ.
Ipl ;> m
In fact, Cp = (-1)lpIT(x P)/pL
(1)
and it will suffice to prove that this value tends to 0 as r -+ 00. Since m
is the order of T, there exists a constant A such that
Ipi ~m.
The support of xqa,h lies in the disc of radius 1/r. The usual formula for
the derivative of a product yields
and hence Dka, is bounded by rlkl times a bound for the derivatives of a
itself, up to order m. In the circle of radius 1/r we have
But Ikl ~ m - Ijl < Iql - UI. This proves that DP(xqa,h) tends to 0 as
r -+ 00, and concludes the proof of our theorem.
Integration on Locally
Compact Groups
This chapter is independent of the others, but is interesting for its own
sake. It gives examples of integration in a different setting from eu-
clidean space, for instance integration on a group of matrices. For an
application of integration and some functional analysis to compact
groups, see Exercise 11.
which define a group law and the inverse mapping in the group, such
that these maps are continuous. After this section, i.e. from §2 to the end
of the chapter, we assume always in addition that G is Hausdorff.
(5) The Galois group of the algebraic numbers over the rational num-
bers, with the Krull topology.
(6) The additive group of p-adic numbers.
(If you don't know these last two examples, don't panic; forget about
them. They won't be used in this book.)
Proof. Just the same as when S is in a metric space. For each XES
we can find an open neighborhood Ux of e such that if y E xUx , then
V= v. n· ·· n
Xl
V.Xn .
Let x, YES and suppose that X E yv. We have y E Xi v,,; for some i, and
hence
so that
If(x) - f(y)1 ;£ If(x) - f(x;) I + If(x i ) - f(y)1
< 26,
Proof. This is proved purely formally using the fact that translations
and the inverse map are homeomorphisms. Namely, if h E H, then
hH=HcR
[XII, §1] TOPOLOGICAL GROUPS 311
n : G -+ G/ H,
which to each x E G associates the left coset xH. We give G/H the
topology having the minimum amount of open sets making n continuous.
Thus a subset W in G/H is defined to be open if and only if n-l(W) is
open in G. We have the following characterization of open sets in G/H:
for a, bEG, and this formula is true with the present definition. Func-
tions form a contravariant system, i.e. if T: X -+ Y is a map of sets, then
it induces a map in the reverse direction
We can apply this when T = fa is the translation, thus forcing the inverse
a -1 when applying translation to functions.
The original proof for the existence of Haar measure due to Haar
provides the standard model for all known proofs. We shall prove the
existence of the functional in §2, following Weil's exposition [W]. Here,
we discuss the relation between the measure and the functional; we prove
uniqueness and give examples.
First we prove a lemma which shows that a locally compact group
has a certain a-finiteness built into it. We recall that a set is called
a-compact if it is a countable union of compact sets.
xK c K n+1 c H,
G= U xiH
i€I
for i in some indexing set I, and H is open, closed, and a-compact. Let
f.1 be a Haar measure. By the remarks following Theorem 2.3, Chapter
14, it follows that the measure on each coset xiH is regular. If A is an
arbitrary measurable set, then we can write
Theorem 2.2. If f.1 is a Haar measure, then for any f E ,g>1(f.1) and any
L L
a E G we have
Theorem 2.3. Let J.l and v be Baar measures on G. Let dJ.l and dv be
the functionals on CA G) associated with J.l and v. Then there exists a
number c > 0 such that dv = c· dJ.l.
We shall first give a simpler proof when the Haar measure is also
right invariant (which applies for instance when G is commutative). Let
hE Cc(G) be a positive function such that
L h dJ.l = 1.
= c LfdJ.l
where
This proves our theorem in the present case. It also gives us an explicit
determination of the constant c involved in the statement of theorem.
The proof of uniqueness when the Haar measure is not also right
invariant is slightly more involved, and runs as follows. For each non-
zero positive function f E Cc(G), we consider the ratio of the integrals
(taken over G):
r(f) =
f-f-.
f dJL
f dv
fh dv = 1.
XH f(xy) - f(yx)
[XII, §2] THE HAAR INTEGRAL, UNIQUENESS 317
has support in the set (supp J)K- 1 u K-1(supp f), which is compact, and
whose ji-measure is bounded by a fixed number CJ depending only on f,
as K shrinks to the origin. Since h is positive, we get the estimate (using
the fact that f h dv = 1):
with an obvious notation concerning the use of the limit symbol. The
left-hand side is independent of f This proves what we wanted, and
concludes the proof of the uniqueness of Haar measure in general.
Corollary 2.4. The map ji 1-+ dji is a bijection between the set of H aar
consequently ji(K) = °
pact set K can be covered by a finite number of translates of V, and
for all compact K. If f E Cc(G), f #- 0, then
flllfil -< W for some open W with compact closure. It follows that
° ~ Af ~ ji(W) Ilfll = 0, contradicting the non-triviality of A. This proves
that ji(V) > 0, and hence that ji is a Haar measure. Thus the map
ji 1-+ dji from Haar measures to Haar functionals is surjective. The map
is injective by the Corollary of Theorem 2.7, Chapter IX. The last state-
ment is now clear.
!~ f a)
-a)
dx
!(X)jXj.
Thus we let J1* be the measure such that dJ1*(x) = dx/ixi. This is easily
seen to be invariant under multiplicative translations. Namely, suppose
that a < O. We compute
But the limits of integration 00 and -00 get reversed, and we conclude
at once that our integral is equal to
f a)
-a)
du
!(u)fUI'
°
f, gEL + and if we assume that g is not identically 0, then there exist
numbers Ci > and elements Si E G (i = 1, . . . ,n) such that for all x we
have
n
f(x) ~ L Cig(Si X ).
i=l
For instance, we let V be an open set and m > such that g ~ m > on
V. We can take all Ci = sup f lm and cover the support K of f by
° °
translates Sl V, ... 'Sn V. We define
(f : g)
to be the inf of all sums L Ci for all choices of {c;}, {s;} satisfying the
above inequality. If g = 0, we define (f: g) to be 00. The symbol (f: g)
satisfies the following properties. The first expresses an in variance under
translation, where fa(x) = f(a- 1 x).
The first four properties are obvious. For (5), we note that if
then
f(x) ~ L cjdjg(tAx).
j,j
A (f) = (f : g) .
9 (h o : g)
Then we have
(7)
For each fixed g, the map Ag will give an approximation of the Haar
functional, which will be obtained below as a limit in a suitable sense.
We note that Ag is left invariant, and satisfies
c ~O.
and h2 = f2 1f.
and
[XII, §3] EXISTENCE OF THE HAAR INTEGRAL 321
But by (7), we know that Aill + 12) and Ag(h) are bounded from above
by numbers depending only on 11' 12 (and h, which itself depends only
on 11' 12). Hence for small <5 we conclude the proof of the lemma.
The additivity of A. shows that this is well defined, i.e. independent of the
choice of fl' f2' and it is immediately verified that A. is then linear on L.
Furthermore, from the properties of Ag , we also conclude that A is left
invariant, and that for any f E L + we have
Ul--+ L
f(uy) dy
ff-+ ff
GIH H
f(uy) dy du
is a Haar functional on G.
h(x) = I'(n(x»)
gH(n(x»)
h(x) =0
Then h is continuous on G, and is constant on cosets of H. Let f = gh.
Then it is clear that fH = 1', thus proving our first assertion, and the
theorem.
),(ra!) = "'(a)),(f).
Theorem 4.2. The map J.l.1-+ dJ.l. is a bijection between the set of (1-
regular positive relatively invariant measures on S and the set of positive
relatively invariant functionals on Cc(S).
In the case of the coset space, we then have the analog of Theorem
4.1.
ff
the map
fl-+ f(uy)",-1(uy) dy du
GI H H
is a Haar functional on G.
ff
GI H H
f(auy)",-l(uy) dy du = "'(a) ffGIH H
f(auy)",-l(auy) dy duo
and
() = angle of 13.
iSU(2)
f 1
dJi. = -22
TC
f1
-1
f~
- J1-u1
f21< F(u
0
1, u2, B) dB dU2 dU 1 •
ff--> i i'"
l
o 0
f(re 21ti8 ) -dr d8.
r
For each fixed a, show that there exists a number ~(a) such that for all
[XII, §5] EXERCISES 327
where dx is Haar measure, and that L\(x -I) dx is right Haar measure.
8. If G is compact and t/!: G --> R + is a continuous homomorphism into R + ,
show that t/! is trivial, i.e. t/!(G) = 1. In this case, Haar measure is also right
invariant.
9. Compute the modular function for the group G of all affine maps x I--> ax + b
with a E R* and bE R. In fact, show that L\(a, b) = a. In this case the right
Haar measure is not equal to the left Haar measure. Show that the right
Haar measure is the Cartesian product measure on R* x R.
to. Let G be a locally compact group with Haar measure, let L\ be the modular
function as in Exercise 7. Let MI be the set of regular complex measures on
G.
(a) Just as in Exercise 8 of Chapter IX, prove that M' is a Banach space, if
we define the convolution m * m' to be the measure associated with the
functional
for f E CAG).
(b) For mE M', define mV to be the direct image of m under the mapping
Xl-->X-' of G. Show that Ilmv II = Ilmli.
(c) If J1. is a right measure, prove that J1.v = L\J1..
11. Let G be a compact abelian group. By a character of G we mean a con-
tinuous homomorphism t/!: G --> C* into the multiplicative group of non-zero
complex numbers.
(a) Show that the values of t/! lie on the unit circle.
(b) If G = R"/Z" is an n-torus, show that the characters separate points.
Assume this for the general case.
(c) Let 0": G --> G be a topological and algebraic automorphism of G, or an
automorphism for short. Show that 0" preserves Haar measure, and in-
duces a norm-preserving linear map T: L 2(J1.) --> L 2(J1.) by fl--> f 00".
(d) If t/! is a non-trivial character on G, show that Jt/! dJ1. = O. If t/! is trivial,
that is t/!(G) = 1, then Jt/! dJ1. = 1, assuming that J1.(G) = 1, which we do.
Prove that the characters generate an algebra which is dense for the sup
norm in the algebra of continuous functions. Prove that the characters
form a Hilbert basis for L 2 .
328 INTEGRATION ON LOCALLY COMPACT GROUPS [XII, §5]
LX) h(r)r 3 dr = 1.
Calculus
and elements Vi' ""V n E E such that if ai-l < t < ai' then f(t) = Vi' We
then say that f is stt;p with respect to P. The notion of a refinement of a
partition is the usual one, and if J, g are two step maps of [a, bJ into E,
then there exists a partition P such that both J, g are step with respect
to P. From this we see that the step maps form a subspace of the space
of all bounded maps, and we deal with the sup norm on this space.
We define the integral of a step map f with respect to a partition P by
n
Jp(f) = l: (ai -
i=l
ai-dv i ,
(1)
If a ~ c<d ~ b, we define
Then formula (1) actually holds for any three points a, b, c in any order,
lying in an interval on which f is in the closure of the space of step
maps.
Since a continuous map is uniformly continuous on a compact set,
one concludes that the continuous maps of [a, b] into E lie in the closure
of the space of step maps, so that the integral is defined over continuous
maps.
If E = E1 X ••• x En is a product of Banach spaces, and
ff?;'O
as one sees first for step maps, and then by continuity for uniform limits
of step maps.
For convenience, the closure of the space of step maps will be called
the space of regulated maps. Thus a map is called regulated if it is a
uniform limit of step maps.
[XIII, §2] THE DERIVATIVE AS A LINEAR MAP 333
additive term f(x), of course, with an error term described by the limiting
properties of r/J or cp described above.
It is clear that if f is differentiable at x, then it is continuous at x.
We contend that if the continuous linear map A exists satisfying (*),
then it is uniquely determined by f and x. To prove this, let AI' A2 be
continuous linear maps having property (*). Let vEE. Let t have real
values> 0 and so small that x + tv lies in U. Let h = tv. We have
Take the limit as t ~ O. The limit of the right side is equal to O. Hence
A1 (V) - A2(V) = 0 and Al(V) = A2(V). This is true for every vEE, whence
Al = ,{2, as was to be shown.
In view of the uniqueness of the continuous linear map A, we call it
the derivative of f at x and denote it by f'(x) or Df(x). Thus f'(x) is a
continuous linear map, and we can write
Ah
Df = f' : V -+ L(E, F)
from V into the space of continuous linear maps L(E, F), and thus to
each x E V, we have associated the linear map f'(x) E L(E, F). If f' is
continuous, we say that f is of class C 1 • Since f' maps V into the
Banach space L(E, F), we can define inductively f to be of class CP if all
derivatives Dkf exist and are continuous for 1 ~ k ~ p.
If f : [a, b] -+ F is a map of a real variable, then its derivative
f'(t): R -+ F
(cf)'(x) = cf'(x).
Al + A2 = (f + g),(x),
as was to be shown. The statement with the constant is equally clear.
VH (f'(x)v)g(x) + f(x}(g'(x)v).
Note that f'(x): E ~ FI is a linear map of E into FI , and when applied to
VEE yields an element of Fl. Furthermore, g(x) lies in F2 , and so we
can take the product
(f'(x)v)g(x) E G.
Similarly for f(x) (g'(x)v). In practice we omit the extra set of parenthe-
ses, and write simply
f'(x)vg(x).
Ivwl ~ Ivllwl
We have:
The map
hf-'--> f(x)g '(x)h + f'(x)hg(x)
is the linear map of E into G, which is supposed to be the desired
derivative. It remains to be shown that each of the other three terms
appearing on the right are of the desired type, namely o(h). This is
immediate. For instance,
Proof We have
from which we see that Ik(h)11/I1 (k(h)) = o(h). We argue similarly for the
other term.
Corollary 3.2. Let f: U --+ F be a differentiable map, and let A.: F --+ G
be a continuous linear map. Then for each x E U,
(A 0 f)'(x) = A(j'(X)),
(A 0 f)'(x)v = A(j'(X)V).
Proof This follows from Theorem 3.1 and the chain rule. Of course,
one can also give a direct proof, considering
Proof Suppose that f(t) #- f(a) for some t E [a, b]. By the Hahn-
Banach theorem, let A be a functional such that A(j(t)) #- A(j(a)). The
map A0 f is differentiable, and its derivative is equal to O. Hence A0 f is
constant on [a, b], contradiction.
340 DIFFERENTIAL CALCULUS [XIII, §4]
cp(c + h) - cp(c) = f C
C
+h
f
f
and
C +h
cp(c + h) - cp(c) - hf(c) = C (f - f(c)).
The mean value theorem essentially relates the values of a map at two
different points by means of the intermediate values of the map on the
line segment between these two points. In vector spaces, we give an inte-
gral form for it.
We shall be integrating curves in the space of continuous linear maps
L(E, F).
We shall also deal with the association
L(E, F) x E ..... F
given by
(A, Y) 1-+ A(Y)
r
hand, we can integrate the curve a, and
a(t) dt
r r
interval J
r
where the dot on the right means the application ofthe linear map
a(t) dt
to the vector y.
AH A(Y) = AY
is a continuous linear map of L(E, F) into F. Hence our lemma follows
from the last property of the integral proved in §l.
Proof. Let g(t) = f(x + ty). Then g'(t) = f'(x + ty)y. By the funda-
mental theorem of calculus we find that
But g(1) = f(x + y) and g(O) = f(x). Our theorem is proved, taking into
account the lemma which allows us to pull the y out of the integral.
342 DIFFERENTIAL CALCULUS [XIII, §4]
(Note. The sup of the norms of the derivative exists because the seg-
ment is compact and the map t t-+ If' (x + ty)1 is continuous.)
Proof. We can either apply Corollary 4.3 to the map g such that
g(x) = f(x) - f' (xo)x, or argue directly with the integral:
and find
We shall call Theorem 4.2 or either one of its two corollaries the
mean value theorem in vector spaces. In practice, the integral form of the
remainder is always preferable and should be used as a conditioned
[XIII, §5] THE SECOND DERIVATIVE 343
reflex. One big advantage it has over the others is that the integral, as a
function of y, is just as smooth as f', and this is important in some
applications. In others, one needs only an intermediate value estimate,
and then Corollary 4.3, or especially Corollary 4.4, may suffice.
Df = f' : U -+ L(E, F)
We have seen in Chapter IV, §1 that we can identify L(E, L(E, F)) with
L(E, E; F), which we denote by L2(E, F), i.e. the space of continuous
bilinear maps of E into F .
we have
This proves the first assertion, and also the second, since each term on
the right is linear in both (Xl ' X2) = x and h = (hI ' h2)' We know that
the derivative of a linear map is constant, and the derivative of a con-
stant map is 0, so the rest is obvious.
A.: E x E -+ F
344 DIFFERENTIAL CALCULUS [XIII, §5]
A(V, W) = A(W, v)
is said to be symmetric if
for any permutation (J of the indices 1, ... ,no In this section we look at
the symmetric bilinear case in connection with the second derivative.
We see that we may view a second derivative D2f(x) as a continuous
bilinear map. Our next theorem will be that this map is symmetric. We
need a lemma.
and that
IA(V, w)1 ~ II/J(v, w)llvllwl·
Then A = 0.
Proof. This is like the argument which gave us the uniqueness of the
derivative. Take v, WEE arbitrary, and let s be a positive real number
sufficiently small so that I/J(sv, sw) is defined. Then
f Il
Then
+ Il Il t/I(sv, tw)v' w ds dt
= D2f(x)(v, w) + cp(v, w)
where cp(v, w) is the second integral on the right, and satisfies the
estimate
Icp(v, w)1 ~ sup It/I(sv, tw)llvllwl·
S.I
where
ICPl(V, w)1 ~ sup It/ll(SV, tw)llvllwl·
s. 1
346 DIFFERENTIAL CALCULUS [XIII, §6]
But then
D 2f(x)(w, v) - D 2f(x)(v, w) = cp(v, w) - CPl(V, W).
as was to be shown.
For an application of the second derivative, cf. the Morse-Palais
lemma in Chapter XIII. It describes the behavior of a function in a
neighborhood of a critical point in a manner used for instance in the
calculus of variations.
Thus DPf(x) is an element of L{E, L(E, ... ,L(E, F) ... ») which we denote
by U(E, F). We say that f is of class CP on U or is a CP map if Dkf(x)
exists for each x E U, and if
and
If J. E U(E, F) we write
Lemma 6.1. Let V2' ... ,vp be fixed elements of E. Assume that f is p
times differentiable on U. Let
and
(**)
From (*) and (**) we conclude that DPf(x) is symmetric because any
permutation of (1, ... ,p) can be expressed as a composition of the permu-
tations considered in (*) or (**). This proves the theorem.
pose these
Ex···xE~F!.G
(I) ~ A0 (I),
and
so
If one wishes to omit the x from the notation in Theorem 6.3, then
one must write
Occasionally, one omits the lower * and writes simply DP(2 0 f) = 20 DPf
Df(x)y Dr1f(x)y(P-l)
f(x + y) = f(x) +- 1-! - + ... + (p _ I)! + Rp
where
Rp = f 1 (1-t)P-l
o (p-l)!
DPf(x + ty)y(P) dt.
We consider the map t f--+ Df(x + ty)y of the interval into F , and the
usual product
R x F --+ F,
This gives the next term, and then we proceed by induction, letting
(1 - t)rl
u = DPf(x + ty)y(P) and dv = (p _ I)! dt
at the p-th state. Integration by parts yields the next term of Taylor's
formula, plus the next remainder term.
350 DIFFERENTIAL CALCULUS [XIII, §6]
The other proof can be given by using the Hahn-Banach theorem and
applying a continuous linear function to the formula. This reduces the
proof to the ordinary case of functions of one variable, that is with
values in R. Of course, in that case, we also proceed by induction, so
there is really not much to choose from between the two proofs.
The remainder term Rp can also be written in the form
R =
P
I i
0
(1 - W- 1
(p - I)!
DPf(x + ty) dt . y(P)
.
The mapping
yH I 0
l(I-t)P-l
(p _ I)! DPf(x + ty) dt
lim I/!(ty) =0
y~O
Ii (1
o (p-l)!
-
t)p-l
DPf(x)y(P) dt + Ii
0
(1
(p-l)!
- W- I/!(ty)y(P) dt.
1
We integrate the first integral to obtain the desired p-th term, and esti-
[XIII, §7] PARTIAL DERIVATIVES 351
0~~~1 1I/I(ty)lIyIP
e
Jo
(1 - t)P-1
(p _ I)! dt,
where we can again perform the integration to get the estimate for the
error term O(y).
Proof We proved this for p = 1 in §3, and the general case follows by
induction.
Proof We have
V ~ L(F'G))
11 x -L(E, G)
V ~ L(E,F)
DJ(x) = A.: E i -+ F
then
n
Df(x)v = L DJ(x)v i •
i=l
satisfies
lim rjJ(hl' th 2) = o.
h-O
= o(h).
where each Aij: Ej ...... F; is itself a linear map. We thus take matrices
whose components are not numbers any more but are themselves linear
maps. This is done as follows.
354 DIFFERENTIAL CALCULUS [XIII, §7]
and
Then for any x E U, the linear map Df(x) is represented by the matrix
Proof This follows by applying Theorem 7.1 to each one of the maps
f1 and f2, and using the definitions of the preceding discussion.
Observe that except for the fact that we deal with linear maps, all that
precedes is treated just like the standard way for functions on open sets
of n-space, where the derivatives follow exactly the same formalism with
respect to the partial derivatives.
f
ous. Let
A= f Dd(t, x) dt.
We investigate
g: U -+ L(E, F).
(1)
Fix n > N. Again by the mean value theorem, there exists <5 such that if
Ix - xol <<5 we have
(2)
Finally, use the fact that Ilf: - gil < t:. We conclude from (1) and (2) that
is a C 1 (or even C OO) map on the disc of radius r centered at the origin in A.
3. Let E, F be Banach spaces, and Lis(E, F) the set of toplinear isomorphisms
between E and F. Show that the map Uf-+ u- 1 from Lis(E, F) to Lis(F, E) is
differentiable, and find its derivative (as in the case of Banach algebras).
4. Let A be a Banach algebra with unit e. Show that one can define a square
root function in a neighborhood of e, in such a manner that it is of class C 1
(or even C OO).
S. Let Z be a compact topological space, E a Banach space, and F = CO(Z, E)
the Banach space of continuous maps of Z into E, with the sup norm. Let U
be open in E, and let V be the subset of F consisting of all maps f: Z ..... U
which map Z into U, so V = CO(Z, U). Let g: U ..... G be a map of U into a
Banach space G.
(a) If 9 is continuous, show that the map
(b) If g is of class C 1, show that the above map is of class C\ and find a
formula for its derivative.
(c) If g is of class CP, show that the above map is of class CPo
6. Let J = [a, bJ be a closed interval, and let U be open in a Banach space E.
Let g: U -+ G be a C I map. Let C°(J, U) be the set of continuous maps of J
into U. Show that C°(J, U) is open in C°(J, E), and that the map
11. This exercise is a starting point for the calculus of variations. Let E be a
Banach space and U an open subset of R x E x E. Let
H: U-+R
f g(t)u(t) dt =0
uta) = u(b) = O.
Show that g = O.
(d) Let CI(J, E, 0) be the subset of curves oc in CI(J, E) such that
oc(a) = oc(b) = O.
Both the inverse mapping theorem and the existence theorem for differen-
tial equations will be based on a basic and simple lemma in complete
metric spaces.
and
and for n sufficiently large, TT n x approaches z and also Tz. This proves
the shrinking lemma.
g: V~V
such that g o f and fog are the identity maps on V and V respectively.
We say that f is a local CP-isomorphism at a point x in V, or is locally
CP-invertible at x, if there exists an open set V 1 contained in V and
containing x such that the restriction of f to V 1 is CP-invertible on V 1 •
It is clear that the composite of two CP-isomorphisms is again a
CP-isomorphism, and that the composite of two locally CP-invertible
maps is also locally CP-invertible. In other words, if f is locally CP-
invertible at x, if f(x) is contained in some open set V, and if g: V ~ G is
locally CP-invertible at f(x), then go f is locally CP-invertible at x.
The inverse mapping theorem provides a criterion for a map to be
locally CP-invertible, in terms of its derivative.
Ig'(x)1 ~ t·
From the mean value theorem we see that Ig(x)1 ~ tlxl, and hence that g
maps the closed ball Br(O) into Br/2(0). We contend that given y E Br/2(0),
there exists a unique element x E Br(O) such that f(x) = y. We prove this
by considering the map
We shall now see that this inverse is differentiable on the open ball
Br/2 (0). Indeed, fix y I E Br/2(0) and let y I = f(x I) with x I
E Br(O). Let
y E Br/2(0), and let y = f(x) with x E Br(O). Then:
If'(z) - f'(x)1 ~ s
where
1«5(xl' x2)1 ~ IXl - x 21sup II'(z) - 1'(0)1
~ slx l - x21·
Hence gy has a unique fixed point x E B,(O) which is such that f(x) = y.
This proves the lemma.
Dd(a, b): F -+ G
qJ: U x V -+ E x G
given by
qJ(x, y) = (x, f(x, y)).
DqJ(a b) = ( I E
, Dd(a, b)
0)
Dd(a, b)
= (IE
Dd(a, b)
0)
IG
0)
and is the matrix
( IE
- Dd(a, b) IG .
[XIV, §3] EXISTENCE THEOREM FOR DIFFERENTIAL EQUATIONS 365
Since ({J is invertible near (a, b) it follows that there is a unique point
(x, y) near (a, b) such that ({J(x, y) = (x, 0). Let Uo be a small ball on
which g is defined. If go is also defined on Uo , then the above argument
shows that g and go coincide on some smaller neighborhood of a. Let
x E Uo and let v = x-a. Consider the set of those numbers t with
o ~ t ~ 1 such that g(a + tv) = go(a + tv). This set is not empty. Let s
be its least upper bound. By continuity, we have g(a + sv) = go(a + sv).
If s < 1, we can apply the existence and that part of the uniqueness just
proved to show that g and go are in fact equal in a neighborhood of
a + sv. Hence s = 1, and our uniqueness statement is proved, as well as
the theorem.
a: J ~ U
We visualize this as saying that the velocity (tangent) vector of the curve
a at a point is equal to the vector associated to that point by the vector
field. We observe that an integral curve can also be viewed as a solution
of the integral equation
If(x) - f(y)1 ~ K Ix - yl
where Jb is the open interval - b < t < b, and Ba(x o ) is the open ball of
radius a centered at Xo '
of the closed interval into the closed ball of center Xo and radius 2a, such
that 0((0) = x. We view M as a subset of the space of continuous maps
of Ib into E, with the sup norm. Then M is complete. For each 0( in M
we define the curve SO( by
We can now apply the shrinking lemma to conclude the proof of our
theorem.
and
be two integral curves for f with the same initial condition Xo' Then a l
and a 2 are equal on J l n J2 •
Proof. Let Q be the set of numbers b such that a l (t) = a 2 (t) for
o ~ t < b. Then Q contains some number b > 0 by the local uniqueness
theorem. If Q is not bounded from above, the equality of a l (t) and a 2 (t)
for all t > 0 follows at once. If Q is bounded from above, let b be its
least upper bound. We must show that b is the right end point of
J l n J2 • Suppose that this is not the case. Define curves /31 and /32 near
o by
and
Then /31 and /32 are integral curves of f with the initial conditions a l (b)
and a 2(b) respectively. The values /3l(t) and /32(t) are equal for small
negative t because b is the least upper bound of Q. By continuity it
follows that al(b) = a2 (b), and finally we see from the local uniqueness
theorem that
of Q. We can argue the same way towards the left end points, and thus
prove our theorem.
For each x E U, let J(x) be the union of all open intervals containing °
on which integral curves for j are defined, with initial condition equal to
x. Then Theorem 3.3 allows us to define the integral curve uniquely on
all of J(x).
°
Remark. The choice of as the initial time value is made for conve-
nience. From Theorem 3.3 one obtains at once (making a time transla-
tion) the analogous statement for an integral curve defined on any open
interval; in other words, if J 1 , J 2 do not necessarily contain 0, and to is a
point in J 1 1\ J 2 such that (Xl (to) = (X2(tO), and also we have the differential
equations
and
then (Xl and (X2 are equal on J 1 1\ J 2 • One can also repeat the proof of
Theorem 3.3 in this case.
In practice, one meets vector fields which may be time dependent, and
also depend on parameters. We discuss these to show that their study
reduces to the study of the standard case.
j:J x U ~E
!:JxU~RxE
by
!(t, x) = (l,J(t, x))
Dependence on Parameters
g: J x V x U-.E
G: J x V x U -. F x E
by
G(t, z, y) = (0, g(t, z, y»
which gives the flow of our original vector field g depending on the
[XIV, §4] LOCAL DEPENDENCE ON INITIAL CONDITIONS 371
We shall now see that the map x ~ IXx in fact depends differentiably on x.
The proof, which depends on a very simple application of the implicit
mapping theorem in Banach spaces, was found independently by Pugh
and Robbin.
Let V be open in E and let f : V -+ E be a CP map (which we call a
vector field). Let b > 0 and let Ib be the closed interval of radius b
centered at O. Let
T(x, a) = x + L f 0 a - a.
Here we omit the dummy variable of integration, and x stands for the
constant curve with value x. If we evaluate the curve T(x, a) at t, then
by definition we have
Lemma 4.1. The map T is of class CP, and its second partial derivative
is given by the formula
D2 T(x, a) = L Df 0 a - I
372 INVERSE MAPPINGS AND DIFFERENTIAL EQUATIONS [XIV, §4]
~ L If 0 (a + h) - f 0 a - (Df 0 a)hl·
L
we have
T(x, a) =0
is precisely an integral curve for the vector field, with initial condition
equal to x. Thus we are in a situation where we want to apply the
implicit mapping theorem.
It Df(a(u))h(u) du I~ bClllhll.
is of class CPo
Proof. We take a so small and then b so small that the local flow
exists and is uniquely determined by Theorem 3.1. We then take b
smaller and a smaller so as to satisfy the hypotheses of Lemma 4.2. We
can then apply the implicit mapping theorem to conclude that the map
x f-+ (Xx is of class C p. Of course, we have to consider the flow (X and still
must show that (X itself is of class CPo It will suffice to prove that Dl (X
and D2 (X are of class c r l , by Theorem 7.1 of Chapter XIII. We first
consider the case p = 1.
We could derive the continuity of (X from Corollary 3.2 but we can
also get it as an immediate consequence of the continuity of the map
374 INVERSE MAPPINGS AND DIFFERENTIAL EQUATIONS [XIV, §4]
Then
Icp(x) - cp(y)11 wi
so that
ID21X(t, x) - D21X(t, y)1 ~ Icp(x) - cp(y)l·
I(cp(y)w)(t) - (cp(y)w)(s) I
Therefore
= W+ f
0
l 1
I [f(IX(U, x + AW)) - /(IX(U, x))] duo
On the other hand, we have already seen in the proof of Theorem 4.3
that
IAI
(cp(x)w)(t) + Tiwi tjJ(AW)(t)
= W+ f 0
l 1
I [f{IX(U, x + AW)) - /(IX(U, x))] du
= W + Lf G(u, A, v) dv du,
where
with
Letting A -+ 0, we have
By (5) we have
~ bC1Icp(x)I'lwl'lt - sl,
I
We have
We can differentiate under the integral sign with respect to the parameter
x and thus obtain
(X : !l(f) ~ U
defined on all of !l(f), letting (X(t, x) = (Xx(t) be the integral curve on J(x)
having x as initial condition. We call this the flow determined by f, and
we call !l(f) its domain of definition.
J(tox) = J(x) - to
and
are integral curves of the same vector field, with the same initial condi-
tion tox at t = O. Hence they have the same domain of definition J(tox).
Hence tl lies in J(tox) if and only if tl + to lies in J(x). This proves
the first assertion. The second assertion comes from the uniqueness of
the integral curve having given initial condition, whence the theorem
follows.
denote by 13:
13(0, x) = x,
defined for some open interval Ja = (-a, a) and open ball Ba(x l ) of ra-
dius a centered at Xl' Let (j be so small that whenever b - (j < t < b we
have
and
Dl qJ(t, x) = D l f3(t - t l , a(tl' x)
= f(f3(t - t l , a(tl' x)))
= f(qJ(t, x).
Hence both qJx and ax are integral curves for f with the same value at t l .
They coincide on any interval on which they are defined by Theorem 3.3.
If we take (j very small compared to a, say (j < a/4, we see that qJ is an
extension of a to an open set containing (tl' x o), and also containing
(b, x o). Furthermore, qJ is of class CP, thus contradicting the fact that b
is strictly smaller than the end point of J(x o ). Similarly, one proves the
analogous statement on the other side, and we therefore see that 1)(f) is
open in R x V and that a is of class CP on 1)(f), as was to be shown.
J(x o ). The flow f3 at lX(b, x o) has a fixed local domain of definition, and
we simply take t close enough to b so that f3 gives an extension of IX, as
described in the above proof.
Of course, if f is of class Coo, then we have shown that IX is of class CP
for each positive integer p, and therefore the flow is also of class Coo.
. f(2 n x)
g(x) = hm - n-
n-oo 2
exists.]
2. Generalize Exercise 1 to the bilinear case. In other words, let f : E x F -> G
be a map and assume that there is a constant C such that
for all X, Xl' Xl E E and Y, Yl, Yl E F. Show that there exists a unique
bilinear map g : E x F -> G such that f - 9 is bounded for the sup norm.
3. Prove the following statement. Let Br be the closed ball of radius reentered
at 0 in E. Let f: Br -> E be a map such that:
(a) If(x) - f(y)1 ;;i; b Ix - yl with 0 < b < 1.
(b) I f(O) I ;;i; r(l - b).
Show that there exists a unique point X E Br such that f(x) = x.
4. With notation as in Exercise 3, let 9 be another map of Br into E and let
c > 0 be such that Ig(x) - f(x) I ;;i; c for all x. Assume that 9 has a fixed point
Xl' and let Xl be the fixed point of f. Show that IXl - xli ;;i; c/(1 - b).
5. Let K be a continuous function of two variables, defined for (x, y) in the
square a;;i; X ;;i; b and a;;i; Y ;;i; b. Assume that IIKII;;i; C for some constant
C > O. Let f be a continuous function on [a, b] and let r be a real number
satisfying the inequality
1
Irl < C(b - a)"
Show that there is one and only one function 9 continuous on [a, b] such
380 INVERSE MAPPINGS AND DIFFERENTIAL EQUATIONS [XIV, §6]
that
6. Newton's method. This method serves the same purpose as the shrinking
lemma but sometimes is more efficient and converges more rapidly. It is used
to find zeros of mappings.
Let B, be a ball of radius r centered at a point Xo E E. Let f: B, -> E be a
C 2 mapping, and assume that r
is bounded by some number C G 1 on B,.
Assume that f'(x) is invertible for all x E B, and that 1f'(xt11 ~ C for all
x E B,. Show that there exists a number b depending only on C such that if
If(xo)1 ~ b then the sequence defined by
ex(x + 1) = ex(x) + 1
such that
f(ex(x)) = ex(nx).
[Hint: Follow Tate's proof. Show that f is continuous, strictly increasing,
[XIV, §6] EXERCISES 381
and let g be its inverse function. You want to solve IX(X) = g(lX(nx)). Let
M be the set of all continuous functions which are increasing (not neces-
sarily strictly) and satisfying IX(X + 1) = IX(X) + 1. On M, define the norm
(TIX)(x) = g(lX(nx)).
one says that the map x t--+ nx is conjugate to f Interpreting this on the
circle, one gets the statement originally due to Shub that a differentiable
function on the circle, with positive derivative, is conjugate to the n-th
power for some n.
(b) Show that the differentiability condition can be replaced by the weaker
condition: There exist numbers '1' '2 with 1 < '1 <'2 such that for all
x ~ 0 we have
A=G ~)
has this property.
Next we introduce the C I norm. If f is a C I map, such that both f and
I' are bounded, we define the C I norm to be
IIflll = max(lIfll, 111'11),
where I II is the usual sup norm. In this case, we also say that f is
C I-bounded.
382 INVERSE MAPPINGS AND DIFFERENTIAL EQUATIONS [XIV, §6]
that is Tp+ = (I - S)p+ where IISII < 1. So find an inverse for T on p+.
Analogously, show that Tp- = (1- SOl )p- where II So II < 1, so that SoT =
So - I is invertible on p-. Hence T can be inverted componentwise, as it
were.
To prove the theorem, write f = A + g where g is C I-small. We want to
solve for h = I + P with p E M, satisfying f 0 h = h 0 A. Show that this is
equivalent to solving
Tp = -A-log 0 h,
or equivalently,
p = - r- I (A -log 0 (I + p)).
This is then a fixed point condition for the map R: M ..... M given by
Note. With only a bounded amount of extra work, one can show that the
map h itself is CO-invertible, and so f = h 0 A 0 h- l •
12. (a) Let f be a C l vector field on an open set U in E. If f(xo) = 0 for some
Xo E U, if IX: J -> U is an integral curve for f, and there exists some to E J
such that lX(t o) = Xo, show that lX(t) = Xo for all t E J. (A point Xo such
that f(xo) = 0 is called a critical point of the vector field.)
(b) Let f be a C l vector field on an open set U of E. Let IX: J -> U be an
integral curve for f Assume that all numbers t > 0 are contained in J,
and that there is a point P in U such that
lim lX(t) = P.
Prove that f(P) = o. (Exercises 12(a) and 12(b) have many applications,
notably when f = grad g for some function g. In this case we see that P
is a critical point of the function g.)
13. Let U be open in the (real) Hilbert space E and let g: U -> R be a C 2
function. Then g': U -> L(E, R) is a C l map into the dual space, and we
know that E is self dual. Thus there is a C l map f: U -> E such that
g'(X)v = (v,f(x»
I/I'(t) ~ -cl/l(t).
Dq>(x)v = f(q>(x)).
In the language of charts (Chapter XXI) this expresses the fact that if a
vector field is not zero at a point, then after a change of charts, this
vector field can be made to be constant in a neighborhood of that point.
15. Let J be an open interval (a, b) and let U be open in E. Let f : J x U -+ E
be a continuous map which is Lipschitz on U uniformly for every compact
subinterval of J . Let IX be an integral curve of f, defined on a maximal open
subinterval (a o , bo) of J. Assume:
(a) There exists e > 0 such that the closure 1X(bo - e, bo}} is contained in U.
(b) There exists c> 0 such that If(t, lX(t)) 1 ~ C for all t in (b o - e, bolo
Then bo = b.
16. Linear differential equations. Let J be an open interval containing 0, and let
V be open in a Banach space E. Let L be a Banach space. Let A : J x V -+ L
be a continuous map, and let L x E -+ E be a continuous bilinear map.
Let Wo E E. Then there exists a unique map k J x V -+ E, which for each
x E V is a solution of the differential equation
This map A is continuous. [Hint: Use Exercise 15. We see that in the linear
case, the integral curve is defined over the whole interval J.]
17. Let U be open in a Banach space E and let f: U -+ E be a C 1 vector field.
Assume that f is bounded. Let IX be an integral curve for f, and let J be its
maximal interval of definition. Suppose that J does not contain all positive
real numbers, and let b be its right end point. Show that
lim lX(t)
,-b
exists, and that it is a boundary point of U. Cf. [La 1] and [La 2] to see
Exercises 12-17 worked out.
PART FIVE
Functional Analysis
made, i.e. either by describing a basis for the space on which the effect
of the operator is obvious, or by giving a structure theorem for the
algebra generated by the operator. These two ways permeate functional
analysis.
CHAPTER XV
Proof Suppose that this is not the case. We find Xl in the comple-
ment of Sl (which cannot be the whole space) and some closed ball
Br,(xd centered at Xl of radius r 1 > 0, contained in this complement. By
assumption, there is some X z in Br , (xd contained in the complement of
Sz and some closed ball Br2 (x z ) contained in Br , (Xl)' and which lies in
the complement of Sz. We continue inductively using a sequence r 1 , rz ,
... such that rn > 0 and rn -> o. We thus obtain a sequence of closed
balls
such that Br (xn) is disjoint from Sl u··· uSn. We then select X n+1 and
Brn +,(xn+1) c BrJx n ) disjoint from Sn+1. Then the sequence {xn} is a
Cauchy sequence, converging to a point X, and X lies in every Br (xn) for
all n. Hence X does not lie in Sn for any n, contradicting the hypothesis
that the union of all Sn is equal to X. This proves Baire's theorem.
Inductively, there exist Xl' ... ,Xn E E such that IXnl < k(jn-l r and
Hence <p(Bkr ) contains the ball Cr(l-cS) of radius r(1 - (j). This is true
for every (j > 0 whence our assertion follows that <p(Bkr ) contains Cr.
Now to conclude the proof of Theorem 1.3, let U be an open set in E,
and let x E U. Let B be an open ball centered at the origin in E
such that x + B c U. Then <p(x) + <p(B) is contained in <p(U). But <p(B)
contains an open ball centered at the origin in F. This proves that <p(U)
is open, and concludes the proof of the open mapping theorem.
FxG-+E
Ix + FI = inf Ix + YI·
YEF
Then ElF is complete under this norm, i.e. is also a Banach space. To see
this, let
cP: E -+ ElF
Icp(x)1 ~ Ixl·
Let {¢n} be a Cauchy sequence in El F. Taking a subsequence if neces-
sary, we may assume without loss of generality that
Indeed, suppose that we have found Xl' ... ,Xn satisfying these conditions.
Since I~n+l - ~nl < 2Ll' we can find y such that
1
and Iyl < 2n+1
EjF --+ G
induced by q>, namely the map such that x + F t-+ q>(x) = q>(x + F). This
map is in fact continuous, because there exists C > 0 such that for all
x E E we have
Iq>(X) I ~ Clx + FI
Since q>(x) = q>(x + y) for all Y E F it follows that
Iq>(X) I ~ Clx + FI
whence the continuity of EjF --+ G. Consequently, by Corollary 1.4 of
the open mapping theorem, if q> is surjective, it follows that the map
EjF --+ G is a toplinear isomorphism.
Let E be a vector space and F a subspace. If EjF has finite dimen-
sion, then we say that F has finite codimension, and we call dim EjF its
codimension.
E --+ EjF.
Proof We can find in the usual way (as in Corollary 1.6) a finite
dimensional subspace F of G such that G = cp(E) + F. Of course, so far,
this is an algebraic direct sum, not yet topological. Factoring out the
kernel of cp, we may assume without loss of generality that cp is injective.
We compose cp with the natural map G ~ GIF. Then the composite
We could now deal with either the real or complex case. We deal with
the latter, since it is useful to get used to the complex conjugation which
occurs, and introduces only a change of notation.
Let E be a Banach space over the complex numbers. We let E* be
the space of anti-linear maps cp: E ~ C, i.e. continuous maps which are
R-linear and satisfy
q5(x) = cp(x)
392 OPEN MAPPING, FACTOR SPACES, DUALITY [XV, §2]
u: E ---+ F
u*: F* ---+ E*
such that
CPHCPOU,
E x E* ---+ C
by
(x, cp)H<X, cp) = cp(x).
This map is continuous sesquilinear, and we shall see that it behaves very
much like the scalar product of Hilbert space for the basic formalism of
duality.
First the remark that the map
UHU*
Ex E* -+ C
E* -+F*
(4) F.l.l = F.
u: E -+ G
394 OPEN MAPPING, FACTOR SPACES, DUALITY [XV, §2]
namely
in a natural way.
The reader acquainted with the language of exact sequences will see
that our results can be expressed as follows. If
o+- F* +- E* +- G* +- 0
is also exact.
[XV, §3] APPLICA TrONS OF THE OPEN MAPPING THEOREM 395
The results of this section will not be used at all throughout the rest of
this book, and are included only for the sake of completeness. The first
two give criteria for a linear map to be continuous.
As usual, if q>: E --+ F is a map, we define the graph of q> to be the set
of all points (x, q>(x)) in E x F. If q> is linear, then the graph of q> is
obviously a subspace of E x F.
Theorem 3.1 (Closed Graph Theorem). Let q>: E --+ F be a linear map
from one Banach space into another, and assume that the graph is
closed. Then q> is continuous.
U T;(B)
ie I
is bounded.
Proof For each positive integer n let en be the set of all x E E such
that I T;x I ~ n for all i E I. Since each T; is continuous, it follows that en
is closed. By assumption, we have
00
E = U en·
n=l
whence
I7;(x) I ~ 17;(x + xo)1 + I7;(xo) I
~ 2m.
Tx = lim T"x
n-+oo
lim T"x = 0
x-+o
uniformly in n.
Then
[XV, §3] APPLICA nONS OF THE OPEN MAPPING THEOREM 397
Then
There exists X3 E E such that AX3 = Y3 and IX31 ~ rZ. Continuing induc-
tively, we find x. such that AX. = y. and
Then
Y1 = <PX1 + . .. +<PX. + Y.+1 ·
If we let
co
X = LX.,
.=1
1 .
t hen Ixi ~ -1-' and <PX = Y1' thus provmg our theorem.
-r
Xl + X2 E (1 + r)B,
and by (*) we find
Let Y3 = Axl - f(x l + X2)· There exists X3 with IX31 ~ IY31 ~ r2lxll, such
that Ax3 = Y3. Then
We have
so that we get
We let
co
X = Lx
n=1
n,
1
and we see that f(x) = Y1. Furthermore, x E - - B thus proving our
1- r '
theorem.
CHAPTER XVI
The Spectrum
f(z) = (v - zefl.
1 1
L - -k '
n
S(n) = -
n k=l v - IX r
Il (t -
n
t n - rn = IXkr)
k=l
shows that
402 THE SPECTRUM [XVI, §1]
1
S(n) = I.
V - r(rl vt
Then there exists an interval on the unit circle near ~, and there exists
8> 0 such that for all roots of unity ( lying in this interval, we have
_ 1 I<M-8.
1v- (r
S(n)= - 1[1
n
LI - -k + Ln -1
-k]'
v-ex r v-ex r
the first sum LI being taken over those roots of unity ex k lying in our
interval, and the second sum being taken over the others. Each term in
the second sum has norm ~ M because M is a maximum. Hence we
obtain the estimate
1
IS(n)1 ~ -n [ILII + ILnl]
bn
~ M - - 8.
- n
This contradicts the fact that the limit of IS(n)1 is equal to M, and proves
our theorem.
Proof. Assume first that K contains C. Then the theorem implies that
K = C. If K does not contain C, in other words does not contain a
square root of -1, we let E = K(j) where / = -1. (One can give a
formal definition of the field E as one defines the complex numbers from
ordered pairs of real numbers. Thus we let E consist of pairs (x, y) with
x, y E K, and define multiplication in E as if (x, y) = x + yj. This makes
E into a field.) We can define a norm on E by putting
~ Izllz'l.
We shall see later in this chapter that under fairly general conditions,
a Banach algebra is isomorphic to the algebra of continuous functions on
a compact set. This set is obtained in a natural way, namely it is the
maximal ideal space of A.
Finally, we remark that the fact that the spectrum is not empty can
also be proved by quoting an elementary theorem about analytic func-
tions of a complex variable, namely that a bounded analytic function is
constant. The proof runs as follows. Suppose that we have an element v
in our algebra such that (v - zer 1 exists for all complex z. Then cer-
tainly the map
is not constant, and hence there exists a functional A. on the algebra such
that the map
Z~A.[(v - ze)-l] = J(z)
404 THE SPECTRUM [XVI, §l]
00
Thus we get the same integral formula for the coefficients of the power
series for J at 0 as in the complex valued cases.
We leave it to the reader to verify that if A is a Banach algebra with
unit e, and x E A, then the function
J(z) = (x - zet 1
lim Ix nl1/ n
n--+ co
exists, and is equal to inf Ixnll /n.
n--+ co n n--+ co
The reason for the name will become clear from Theorem 1.7 below.
We are now ready to consider the power series
L
00
hx(z) = x"z"
"=0
with x"z" E A for all complex numbers z. We are interested in the do-
main of convergence of the series.
Lemma 1.6. Suppose A is a Banach algebra with unit. Then the radius
of convergence of the series hx is l/p(x).
In Theorem 1.1 we saw that the spectrum a(x) is compact and con-
tained in the disc [z[ ~ [xl. We now prove a more precise inequality.
Then [(Z-l X )"[ ~ (r/[zlt for large n, so the series L (Z-l X )" converges and
shows that z rf; a(x). Conversely, let s = sup[z[ for z E a(x), so s is the
smallest radius for a closed disc centered at the origin and containing the
spectrum a(x). Suppose WE C and [w[ > s. Then w rf; a(x), so x - we is
invertible. Thus the function
h(z) = (e - ZX)-l
whence p(x) ~ S.
Proposition 2.3. The association M 1-----+ A.M is a bijection between the set
of maximal ideals and the set of characters of A. For each character A.,
we have IA.I ~ 1.
Proof. The first statement has been proved. As to the second, let us
first prove that if Ixl ~ 1 then IA.(x) I ~ 1. If 1A.(x)1 > 1, then 1A.(xn)1 ~ 00
as n ~ 00, but Ixnl ~ 1, which contradicts the continuity of A.. Thus
1A.(x)1 ~ 1. Then for any x =F 0, let y = x/lxl so Iyl = 1. Then 1A.(y)1 ~ 1,
so IA.I ~ 1, thus proving the proposition.
In light of the fact that A. = A.M for some M, we see that we may
rephrase Proposition 2.3 as follows. If x E A and c E C is such that
x == c mod M, and if Ixl ~ 1, then lei ~ 1.
The Banach algebra A being a Banach space has its dual A'. But it
also has the set of characters, which we denote by A, and A c A'l' where
A'l is the unit ball in A'. In Chapter IV, §1 we defined the weak
topology on A'l by embedding A'l in a product space
f : A'l c:... TI
Ixl~ 1
Kx c TI
xeA
ex·
Let .it be the set of maximal ideals of A. In light of the bijection of .it
with A, we have a natural embedding (the restriction of f)
g: A or .it c:... TI Kx by
Ixl~ I
y(xy) = y(x)y(y),
gx+y = gx + gy and
A -+ C(vII, C) ~ C(,4, C)
If A has a unit element e, then the reader will verify at once that e* = e.
If further x is invertible, then (x -1)* = (X*)-I.
By a C*-algebra, we mean a Banach algebra with an involution as
above, satisfying the additional condition
for all x E A.
Ixl = Ix*l·
Examples. The standard example of a commutative C*-algebra is that
of the algebra of continuous functions on a compact Hausdorff space,
with the sup norm. The involution is given by the complex conjugate,
f* = 1. Theorem 3.3 below shows that under certain conditions, there is
no other.
Let H be a Hilbert space, and let A = End(H) be the algebra of
bounded linear maps of H into itself (operators). Then A is a C*-algebra,
the star operation being the adjoint. In Chapter XVIII, we shall consider
the commutative subalgebra generated by one hermitian operator and
reprove the basic theorem independently of the result in this section.
For another example of an involution (which does not satisfy the
condition of a C*-algebra), see Exercise 9.
and we simply use the definition of the spectral radius to conclude the
proof.
But the element y = x*x satisfies y = y*. By Theorem 1.7 and Proposi-
412 THE SPECTRUM [XVI, §4]
1. Let A be a complex Banach algebra with unit element, and let u E A. Let
a(u) be the spectrum of u. Let p be a polynomial with complex coefficients.
Show that the spectrum of p(u) is equal to p(a(u)), i.e. to the set of all
numbers p(IX), where IX lies in the spectrum of u. [Hint : For one inclusion
write
p(t) - p(lX) = (t - lX)q(t)
3. Let A be a complex Banach algebra with unit element e and let u E A. Show
that the map
so that t/I can be viewed as a function on the unit circle. Let A be the set of
[XVI, §4] EXERCISES 413
00
1= L cnl/l n
-00
L Icnl <
- 00
00 .
Prove that A is a normed algebra, with unit (1,0), and containing an isomor-
phic image of A as the subset of elements (0, x) with x E A . Show that A is
an ideal in A . Warning: If A happened to have a unit element, then this unit
is not the same one as the unit in A.
6. Suppose A is a C*-algebra but without our assuming that A has a unit
element. Prove that there exists on A a norm extending the norm on A
which makes A into a C*-algebra with unit. (Warning : It is not the norm of
Exercise 5.) [Hint: Observe that for x E A, we have
IxYI
Ixl =SUp -
yeA lyl
m*=mV =rnv.
Show that this star operation is an involution, and that Ilmil = Ilm*ll.
10. Let A be a C*-algebra with unit, and let x E A be such that x* = X-I. Show
that the spectrum of x lies on the unit circle.
11. Let A be a Banach algebra with unit e. Let v E A. Show that there exists a
unique COO mapping K: R --+ A such that:
d
H 1. We have dt K(t) = K(t)v.
H 2. K(O) = e.
Show that the image of K lies in the multiplicative group of invertible ele-
ments of A.
CHAPTER XVII
u: E -+ F
r
and c ~ y ~ d. If f is continuous on [a, b], we define
It will be shown later that S is compact. Thus our theory applies to the
study of this type of integral equation.
for some i. This implies that lu(x) - Yil < r, and hence that u(B) is
covered by a finite number of balls of radius r, as was to be shown.
is also compact.
Proof The first relation follows from the fact that a continuous image
of a compact set is compact. The second is obvious. The third comes
from the definitions.
Proof One can give a direct simple proof, but the reader will note
that our assertion is an immediate consequence of the Ascoli theorem.
We shall make no use of Theorem 1.3 in this book, and thus we leave
the details to the reader.
T: E --+ F
is said to be Fredholm if:
(i) Ker T is finite dimensional.
(ii) 1m T is closed and finite codimensional.
418 COMPACT AND FREDHOLM OPERATORS [XVII, §2]
T: E --+ E
by
E = Ker(l - u) E8 G.
It will suffice to prove that TG = TE is closed, and for this it will suffice
to prove that the inverse map
converges. Since
Iz - Yol ~ (inf Iz -
yeF
YI) (1 + 1».
We let
z - Yo
x = .
Iz - Yol
Then for Y E F we have
To apply the lemma, suppose that TE does not have finite codimen-
sion. We can find a sequence of closed subspaces
TE = Ho C HI C . •. c Hn C ...
because - TXn - Xk + TXk lies in Hn- I . This shows that the sequence
{uxn} cannot have a convergent subsequence, and contradicts the com-
pactness of u, thus proving Theorem 2.1.
In the language of linear algebra, the factor space F ITE is also called the
cokernel of T, and thus
Gx H ~ SG EB H =F
given by
(x, Y)I-+SX +Y
is a toplinear isomorphism. We know that the set of toplinear isomor-
phisms of one Banach space into another is open in the space of all
[XVII, §2] FREDHOLM OPERATORS AND THE INDEX 421
G x H -> TG EEl H =F
given by
(x, Y}H Tx +Y
is therefore also a toplinear isomorphism. Hence the kernel of T is finite
dimensional, since G n Ker T = {O}, and in fact dim Ker T is at most
equal to the codimension of G in E. The image of T has finite codimen-
sion (at most equal to the dimension of H), and is consequently closed,
by Corollary 1.8 of Chapter XV. This proves that T is Fredholm, and
proves our first assertion.
Now concerning the index, we observe that G EEl Ker T is a direct sum
of two closed subspaces, and there is some finite dimensional subspace M
such that
E = G EEl Ker T EEl M.
tHind(l- tu}.
ind(l - u} = O.
422 COMPACT AND FREDHOLM OPERATORS [XVII, §2]
1m T:::> 1m TS
Show that Tl == T2 mod K(F, E), and that Tl or T2 is thus an inverse for
T modulo compact operators.
Proof Clear.
ind(T + u) = ind T.
Proof The same proof works as for the corollary of Theorem 2.3,
namely we connect T + u with T by the segment.
T+ tu, O~t~1.
The next theorem will not be used later in a significant way and thus
its proof can be omitted if the reader is allergic to formal algebra.
be Fredholm. Then
f: V ~ f(V)
424 COMPACT AND FREDHOLM OPERATORS [XVII, §2]
be a linear map, with image f(V), which we also write fV for simplicity
of notation. If the factor space V/ W is finite dimensional, we denote by
(V : W) the dimension of the factor space V/ W We denote by lj- the
kernel of f in V, and by »j the kernel of fin W, that is W (\ lj- .
V -+ fV -+ fV/fW
(2) Wclj-+WcV
If we now use (1) and (3), we get the relation stated in the lemma, as was
to be shown.
We return to the proof of Theorem 2.8.
For simplicity of notation, we use our notation Es for the kernel of S
in E. We write down the definitions of the index for T, Sand TS:
TSE c TF c G.
Hence
we obtain
(8)
If we now substitute the values of (5), (6), (7), (8) into the expression for
sum
Then
SE = S(W (£l U) = SW EB SUo
ux = ax.
(I - u)" = / - Ul
ind(u - al)" = 0,
in other words,
(u - a;IYWj = O.
aj 0 0
o aj 0
o 0 0
o 0 0 aj
for all n ~ r.
428 COMPACT AND FREDHOLM OPERATORS [XVII, §3]
Proof It suffices to prove that Ker(l - u)' = Ker(l - u)", for all
n ~ some r. Suppose that this is not the case. Then we have a strictly
ascending chain of subs paces
x. E Ker(l - u)·
and each of the spaces occurring in this direct sum is a closed invariant
subspace of u. If P =F rx is another non-zero eigenvalue of u, and s is its
exponent, then
Ker(u - PI)' c: Im(u - rxI)'.
Ker Tn 1m T = {O}.
Ker T = Ker T2
[XVII, §3] SPECTRAL THEOREM FOR COMPACT OPERATORS 429
x=y+z
o= Sx = Sy + Sz,
and since Sy E Ker T, Sz E 1m T, it follows from the uniqueness of the
decomposition that Sy = O. But S, T are obtained as relatively prime
polynomials in u, and hence there exist polynomials in u, namely P and
Q, such that
PS + QT = I.
(We recall the proof below.) Applying this to y shows that Iy = 0 so that
y = 0 and hence x = Z E 1m T, thus proving our theorem.
Now to recall the proof of the existence of P, Q, let A =u- !Xl and
B = u - pl. There exist constants a, b such that
aA + bB = I.
We take n sufficient!y large, and raise both sides to the n-th power. We
obtain
L" ciaA)j(bB)"-j = I.
j=O
Theorem 3.4. Assume that there are irifinitely many eigenvalues. Then
the eigenvalues =I 0 of u form a denumerable set, and if we order them
as !Xl' !X2' • .• such that
then
430 COMPACT AND FREDHOLM OPERATORS [XVII, §3]
Proof. Given c > 0 we first show that there is only a finite number of
eigenvalues a such that lal ~ c. If this is not true, then we can find a
sequence of eigenvectors {w n} belonging to distinct eigenvalues {an} such
that Iwnl = 1 and lanl ~ c > 0 for all n. The vectors W l , ... 'Wn are lin-
early independent, for otherwise, if n ~ 2 and
n
Fn = L Ker(u -
i=l
aJ)\
and the sum is direct since each Ker(u - ai)'; is finite dimensional. Then
we get an ascending sequence of subspaces
Fl cF2 c···cF"c···.
[XVII, §3] SPECTRAL THEOREM FOR COMPACT OPERATORS 431
Hn = 1m nn
i=l
(u - a;!)",
°
whence x lies in E. Inductively, suppose that the kernel of (u - rxl)k
is contained in E, and suppose that (u - rxI)k+1X = for some x E if.
Then (u - rxl)x = y lies in the kernel of (u - rxl)k and hence lies in E.
Therefore
r
a square [a, b] x [a, b]. Then we obtain an operator SK such that
We shall consider this operator with respect to two norms on the space
E of continuous functions on [a, b].
r
and for f E <I> we have
r
Case 2. We take E with the L 2 -norm, arising from the hermitian
product
The spectral theorem applies therefore in the present case, and so does
Theorem 3.5, which showed that the finite dimensional spaces corre-
sponding to the eigenvalues # 0 actually had bases with elements in E
rather than in the L 2-completion of E.
If we take K to be hermitian, for instance real valued and such that
K(x, y) = K(y, x), then we have a Fourier expansion of any function in E
as an L 2-convergent series. One can then start playing the same game as
in the ordinary theory of Fourier series, and ask for uniform or pointwise
convergence. We leave this to look up for readers who have a more
direct interest in integral equations.
Several more examples of compact integral operators will be given in
the exercises.
1. Let E be the space of COO functions of one variable, periodic of period 2n.
Let D be the derivative. Denote by Eo the space E together with the norm
arising from the hermitian product
e" fg,
<f, g>o = Jo
and for each positive integer p denote by Ep the same space but with the
product
434 COMPACT AND FREDHOLM OPERATORS [XVII, §5]
Show that S is continuous, and that IS"II/" --+ 0 as n --+ 00. [Hint: Show
that I(S"J)(x) I ~ IIJllx"ln! by induction. You will need some inequality
like n! ~ nne -".]
(b) Show that 0 is the only element in the spectrum of S, and that S is
compact.
*
(c) For each IX 0, given a continuous function gEE, show that there exists
a continuous function J E E such that
SJ - IXJ = g.
Express J explicitly as an integral involving g and the exponential
function.
3. Let J = [0, 1J and let E be the vector space of all C I paths IX: J --+ R". Let
II II be the sup norm and I I the euclidean norm on R". Given two paths IX, p
f
define
the product <IX(t), P(t) > being the dot product. Its associated norm is called
the HO-norm on E, and will be denoted by II 110' Define
(a) Show that this is a positive definite scalar product, and that its norm,
which we call the HI-norm and denote by II III' is equivalent to the norm
arising from the scalar product
and
(c) Let HO(J, Rn) be the completion of E with respect to the HO-norm, and let
Hl(J, Rn) be its completion with respect to the HI-norm. Show that the
identity mapping on E induces injective continuous linear maps
and
[Hint : Use the preceding exercise. If the conclusion is false, pick a sequence
vn E H(n) of unit vectors such that IAvnl > e.]
8. Let H, be the Hilbert space defined in Exercise 6 of Chapter VII. Following
Exercise 8 of that chapter, if r < s, prove that the inclusion Hs c... H, is a
compact linear map.
The next three exercises give examples of compact operators which are called
Hilbert-Schmidt. For a systematic treatment of such operators in general Hilbert
spaces, see Exercises 17 and 18 of Chapter XVIII.
9. Hilbert-Schmidt operators in L2. Let (X, vii, dx) and (Y,.AI, dy) be measured
spaces. Assume that L2(X) and L2(y) have countable Hilbert bases.
(a) Show that if {(!1i} and {I/Ii} are Hilbert bases for L 2(X) and L 2( Y), respec-
tively, then {(!1i ® I/Ii } is a Hilbert basis for L 2(dx ® dy).
436 COMPACT AND FREDHOLM OPERATORS [XVII, §5]
from
L
given by
is compact. [Hint: Prove first that it is bounded, with bound IIKI12 '
Using partial sums for the Fourier expansion of K, show that SK can be
approximated by operators with finite dimensional images. Cf. Theorem 2
of SL2(R), Chapter I, §3.]
Let H be a Hilbert space with countable Hilbert basis. An operator A
on H is said to be Hilbert-Schmidt if there exists a Hilbert basis {q>;}
such that
For a more subtle result along these lines, cf. SL2(R), Theorem 6 of
Chapter XII, §3.]
10. Assume that (X, At, dx) = (Y, %, dy) in the preceding exercise, and that X has
finite measure. Let
K = L cm .• q>m ® CPo
m ••
be the Fourier series for K. Let Pm •• be the integral operator defined by the
function q>m ® CP.·
(a) Show that Pm •• q>. = q>m and Pm •• q>j = 0 if h #- n.
(b) Assume that the coefficients c m,. tend to 0 sufficiently rapidly. Show that
K is in Ll and that
f x
K(x, x) dx = L c•.•.
•
(c) Again assume that the coefficients cm•• tend to 0 sufficiently rapidly. Show
that
L <SKq>., q>.> = L c•.• ,
• •
Under suitable convergence conditions, this gives an integral expression
for the "trace" of SK'
[XVII, §5] EXERCISES 437
SKf(x) = f f(y)K(x, y) dy
I o
I K(x, x) dx = I
•
c•.• .
[Hint : Let {ep.} be the Hilbert basis given by ep.(x) = e2ni•x. Let
1
B= I cm.•(1 + n2 )Pm .• and C=" - - P. .
71 + / }
m.'
T: H -+ L2(X)
(A- 1 )* = (A*)-l.
The case when A = A* is the main one studied in this chapter. For a
complex Hilbert space, the following properties are equivalent, concerning
an operator A :
We have A = A*.
The form CPA: (x, Y)f--+ <Ax, y) is hermitian.
The numbers <Ax, x) are real for all x E E.
The equivalence between the first two is left to the reader. As to the
third, suppose that A = A*. Then
UN 1. A* = A-I .
UN 2. IAxl = Ixl for all x E E.
UN 3. (Ax, Ay) = (x, y) for all x, y E E.
UN 4. IAxl = 1 for every unit vector x E E.
Hence
I(Ax, y) + (Ay, x)1 ~ c(lxl 2 + lyI2).
440 BOUNDED HERMITIAN OPERA TORS [XVIII, §2]
Figure 18.1
The right-hand side remains unchanged, and for suitable e, the left-hand
side becomes
I<Ax, y)1 + I<Ay, x)l·
(In other words, we are lining up two complex numbers by rotating one
by e and the other by -e.) Next we replace x by tx and y by y/t for t
real and t > O. Then the left-hand side remains unchanged, while the
right-hand side becomes
The point at which g'(t) = 0 is the unique minimum, and at this point to
we find that
g(t o) = Ixllyl·
for all x, or equivalently, the sup of all values I<Ax, x)1 taken for x on
the unit sphere in E.
-cl ~ A ~ cl.
then we have
IJ. ~ A ~ {3,
and from Theorem 2.2,
The next two sections are devoted to generalizing to Hilbert space the
spectral theorem in the finite dimensional case. These two sections are
logically independent of each other. In the finite dimensional case, the
spectral theorem for hermitian operators asserts that there exists a basis
consisting of eigenvectors. We recall that an eigenvector for an operator
A is a vector w =1= 0 such that there exists a number c for which Aw = cwo
We then call c an eigenvalue, and say that w, c belong to each other. In
the next section, we describe a special type of hermitian operator for
which the generalization to Hilbert space has the same statement as in
the finite dimensional case. Afterwards, we give a theorem which holds
in the general case, and can be used as a substitute for the "basis"
statement in many applications. Some of these applications are described
in subsequent sections.
442 BOUNDED HERMITIAN OPERA TORS [XVIII, §3]
exI ~ A ~ {3I.
t - y = (t - a) + (a - y)
(t - a)(f3 - t)Q(t)2
pH p(A)
Proof The first assertion comes from the remarks preceding our
lemma. The second follows at once by considering q - p. Finally, if we
let
q(t) = Ilpll ± p(t)
then q ~ 0 on [ex, P] and hence q(A) ~ 0, whence the last assertion fol-
lows from Theorem 2.2.
If(A)1 ~ Ilfll,
the sup norm being taken on [ex, Pl If Pn -? f and qn -? g, then Pnqn -? fg·
Hence we obtain (fg)(A) = f(A)g(A) for any continuous functions, f, g.
In other words, our map is also a ring-homomorphism.
The kernel of our map fl--+ f(A) is a closed ideal in the ring of
continuous functions on [ex, Pl We forget for a moment our definition
of the spectrum given in Chapter XVI, §1, and here define the spectrum
a(A) to be the closed set of zeros of this ideal. We use Theorem 2.1 of
Chapter III.
If f is any continuous function on a(A), we extend f to a continuous
[XVIII, §4] THE SPECTRAL THEOREM 447
function on [a, fJ] having the same sup norm, say Jl' and define
J(A) = Jl (A).
If g is another extension of J to [a, fJ], then g - Jl vanishes on a(A),
and hence g(A) = Jl (A). Hence J(A) is well defined, independently of the
particular extension of J to [a, fJ]. We denote by I IIA the sup norm
with respect to a(A); thus
IJ(A)I ~ IIJIIA·
°
algebra oj continuous Junctions on a(A) onto the Banach algebra R[A].
A continuous Junction J is ~ on a(A) if and only if J(A) ~ o.
Proof. We had derived the norm inequality previously from the posi-
°
tivity statement. We do this again in the opposite direction. Thus we
assume first that J(A) ~ 0 and prove that J is ~ on the spectrum of A.
Assume that this is not the case. Then J is negative at some point c of
the spectrum. Let g be a continuous function whose graph is as follows:
c (3
Figure 18.2
°
Thus g is ~ 0, and has a positive peak at c. Then Jg is ~ and Jg is
negative at the point c of the spectrum. Hence - Jg ~ 0, and hence
- J(A)g(A) ~ O. But J(A) ~ 0 and g(A) ~ 0, so that by Theorem 4.3 we
also have J(A)g(A) ~ O. This implies that J(A)g(A) = 0, which is impos-
sible since Jg does not vanish on the spectrum. We conclude that J ~ °
on a(A), and in view of our previous result this proves the positivity
statement of the theorem.
448 BOUNDED HERMITIAN OPERA TORS [XVIII, §4]
IIfIIA ~ If(A)I,
Then g(t) =f 0 on a(A), and hence h(t) = l /g(t) is its inverse. Then
h(A)(A - Zl) would be an inverse for A - zI, a contradiction. This
proves that z is real.
Let ~ be real and not in the spectrum a(A). Then t - ~ is invertible
on a(A), and hence so is A - O.
Suppose that ~ is in the spectrum a(A). Let g be the continuous
function whose graph is as follows.
Figure 18.3
That is,
l / lt -
g(t) = {N ~I if It - ~I ~ l iN,
if It - ~I ~ liN.
[XVIII, §5] ORTHOGONAL PROJECTIONS 449
B(A - 0) = (A - eI)B = I.
But g(t) has a large sup on the spectrum if we take N large, and hence
Ig(A)1 is equally large, a contradiction. Theorem 4.4 is proved.
The main idea to use the positivity to get the spectral theorem is due
to F. Riesz. However, most treatments go from the positivity statement
to an integral representation of A which we give in Chapter XX. Von
Neumann always emphasized that it is much more efficient to prove at
once the statement of Theorem 4.4, which suffices for many applications,
and can be obtained quite simply from the positivity statement. In fact,
the arguments used to derive Theorem 4.4 from the positivity statement
are taken from a seminar of Von Neumann around 1950.
d
HI. ds e- sA = _Ae- sA .
lim e-sAv = v.
s-o
p2 = P, PQ = QP = 0 , P +Q= I.
E = Ker P + 1m P.
Proof This proof is independent of the spectral theorem, and uses
only basic definitions, together with Corollary 1.8 of Chapter V. Let
F = Ker P. If x E F, we have
x = I x = Px + Qx = Qx
so that x is in the image of Q. Since PQ = QP = 0, it follows that the
image of Q is in the kernel of P, whence Ker P = 1m Q. We obviously
have
Q2 = (I - p)2 = I - P = Q
°
we conclude that Qx = so F.l c Ker Q. The converse inclusion follows
from these same equalities, and our theorem is proved.
Proof. We use the spectral theorem. Let g be the function such that
g(t) = 1 if t ~ 0 and g(t) = 0 if t < O. Since A is invertible, it follows that
o is not in the spectrum of A. Hence g is continuous on the spectrum,
and g2 = g on the spectrum. Hence g(A) = P satisfies p 2 = P. Let F =
1m P. Then P is an orthogonal projection on F by Theorem 5.1. Since
A commutes with g(A), and since tg(t) ~ 0 on the spectrum of A, it
follows that AP = PAis a positive operator. Furthermore, A maps F
into itself, and since A -1 exists on E, and also maps F into itself, let A +
be the restriction of A to F. Then A + is positive, invertible on F, whence
positive definite (because the spectrum is closed, and 0 is not in the
spectrum). Similarly, let h(t) = 1 - g(t) and Q = h(A). Then th(t) ~ 0 on
the spectrum of A, and by similar arguments, letting A - be the restric-
tion of - A to F1-, we conclude that A - is positive definite on F1-. This
proves what we wanted.
Proof. It will suffice to prove that there is only one element in the
spectrum of A. Suppose that there are two, C 1 1= c 2 • There exist continu-
ous functions f, g on the spectrum such that neither is 0 on the spec-
trum, but fg is 0 on the spectrum. For instance, we can take for f, g the
functions whose graphs are indicated on the next figure.
Figure 18.4
p: G -+ Laut(E)
Among other things, this section shows some ways how Theorem 4.3 is
used. Namely, for any operator T on a Hilbert space E, the operator
T*T is hermitian positive, and so has a square root by Theorem 4.3. We
start with a special case of the polar decomposition which is of interest
for its own sake.
and since the image of T is dense, this implies that u is orthogonal to all
of E, whence u = O. Hence we get Tx = Ty, thus defining U uniquely by
our given formula. Then we find
Verify that U satisfies similar relationships. This shows that our two
representations are "isomorphic".
Next we deal with the general polar decomposition. Let H be a
Hilbert space. Let U: H -+ E be a bounded linear map into some other
Hilbert space E. We say that U is a partial isometry if there exists an
orthogonal decomposition
A= UP,
Proof. We shall give the essential steps of the proof, and leave certain
routine details as Exercise 15.
First note that A *A is symmetric positive and has a unique symmetric
positive square root, denoted by
Let U be an open set in some (real) Hilbert space E, and let f be a Cp+2
function on U, with p ~ 1. We say that Xo is a critical point for f if
Df(x o ) = O. We wish to investigate the behavior of f at a critical point.
After translations, we can assume that Xo = 0 and that f(x o ) = O. We
observe that the second derivative D2f(0) is a continuous bilinear form
on E. Let A = D2 f(O), and for each x E E let Ax be the functional
y 1---+ A(x, y). If the map x 1---+ Ax is a toplinear isomorphism of E with its
dual space E', then we say that A is non-singular, and we say that the
critical point is non-degenerate.
We recall that a local CP-isomorphism cp at 0 is a CP-invertible map
defined on an open set containing O.
that
f(x) = <Acp(x), cp(x».
Ll Ll
g(x) = D2 f(stx)t ds dt.
f(x) = <A(x)x, x)
where A: V ---. Sym(E) is a CP map of V into the space of symmetric
operators on E. A straightforward computation shows that
If we let B(x) = A(O)-l A (x), then B(x) is close to the identity I for small
x. The square root function has a power series expansion near 1, which
is a uniform limit of polynomials, and is COO on a neighborhood of I (cf.
Exercise 2 of Chapter XIII), and we can therefore take the square root of
[XVIII, §8] THE MORSE-PALAIS LEMMA 457
We contend that this C(x) does what we want. Indeed, since both A(O)
and A(x) (or A(xfl) are self adjoint, we find that
B(x)* = A(x)A(O)-I,
whence
B(x)* A(O) = A(O)B(x).
But C(x) is a power series in I - B(x), and C(x)* is the same power
series in 1- B(x)*. The preceding relation holds if we replace B(x) by
any power of B(x) (by induction), hence it holds if we replace B(x) by
any polynomial in I - B(x), and hence finally, it holds if we replace B(x)
by C(x), and thus
vanatlOns, EPa 1J and [Sm 1]. For instance, one considers a space of
paths (of various smoothness) (1: [a, bJ -4 E where E is a Hilbert space.
r
One then defines a function on these paths, essentially related to the
length
and one investigates the critical points of this function, especially its
minimum values. These turn out to be the solutions of the variational
problem, by definition of what one means by a variational problem.
Even if E is finite dimensional, so a euclidean space, the space of paths
is infinite dimensional, so that we need an infinite dimensional theory to
deal with this question.
a(A *) = a(A).
6. Prove the statement made in the text that if A is compact, hermitian, then
(Ax, x) takes on a maximum or minimum on the unit sphere.
7. Let E be the space of real valued continuous functions on [0, 1] and let
M: E -+ E be the linear map given by
(Mf)(x) = xf(x).
(f,g) = ffg,
[XVIII, §9] EXERCISES 459
and we let E2 denote the completion of E with respect to this norm. Show
that M is self-adjoint, and that for any real a, the operator M - aI is not
invertible on E 2 , for otherwise, it would be invertible on E. Show that a is
not an eigenvalue of M. Note: M is obviously injective on E, but you will
have to prove that, for instance, it is injective on E z , so deal with LZ-Cauchy
sequences in E.
Show that the spectrum of T is the unit disc and that T has no eigenvalue.
R: G --> Aut(E)
t R(x)PR(xfl dx = cI.
In fact, show that the operator on the left is a positive operator, commut-
ing with all R(a), a E G.
>,
(b) Considering (cv l , VI show that c > o.
460 BOUNDED HERMITIAN OPERA TORS [XVIII, §9]
(c) For any x E G, {R(xf1vJ is an orthonormal basis {wJ. Prove that for
any n,
n
I
i=1
<Pwi , Wi) ~ 1.
11. Let A be a hermitian operator on the Hilbert space E, and assume that the
spectrum of A is the union of two disjoint closed sets S, T. Show that E
admits a direct sum decomposition into two closed subspaces Es and ET
which are A-invariant, and such that, if we let As and AT be the restriction of
A to Es and ET respectively, then the spectrum of As is S and the spectrum
of AT is T. (Cf. Exercise 2 of Chapter XVI.)
13. Show that an operator A on a Hilbert space is hermitian positive if and only
if there exists an operator B such that A = B* B.
15. Give the details of the proofs in statements (a), (b), (c), (d) for Theorem 8.1.
The next exercise gives a complement and refinement of Theorem 8.1 when we
deal with an automorphism of the Hilbert space.
A2
exp(A) = I + A + - + . ..
2!
(A - 1)2
log A = (A - I) - + ...
2
Hilb(E) x Pos(E)
(H, P)HHP.
Hilbert-Schmidt Operators
(a) Prove that the same convergence holds for any other Hilbert basis {Vj}.
For Hilbert-Schmidt operators A and B, define their scalar product
HS 1. IIA*lIz = IIAllz.
HS 2. XA and AX are Hilbert-Schmidt, and
The first sum shows that the trace is independent of the choice of B, C.
Show :
TR 1.
[XVIII, §9] EXERCISES 463
tr(AX) = tr(XA).
TR 3. A is of trace class if and only if PA is of trace class, and
and
to prove the assertion in this case. In general write the polar decomposition
A = UP, so T A = (TU)P and you can apply the first part of the proof.]
Note: For complete proofs, cf. [La 3], Appendix of Chapter VII, and [Sh].
CHAPTER XIX
We need to extend the notion f(A) to functions f which are not continu-
ous, to include at least characteristic functions of intervals. We follow
Riesz- Nagy more or less. We let H be a Hilbert space.
Then Av is anti linear, and I<Anv, w)1 ~ Clvllwl for some C and all v,
w E H . Hence there exists an operator A such that
Proof Say gn(t) clecreases also to f(t). Given k, for large n we have
This is true for all e. Letting e -+ 0 and using symmetry, we have proved
our lemma.
466 FURTHER SPECTRAL THEOREMS [XIX, §1]
ft-+ f(A)
(fg)(A) = f(A)g(A)
for f, 9 in this vector space.
The most important functions to which we apply this extension are
characteristic functions like the function t/tJt) whose graph is drawn in
Figure 19.1. It is a limit of the functions hn(t) drawn in Figure 19.2.
Figure 19.1
c c+ l
n
Figure 19.2
c c
Figure 19.3
Hence
A - cI = -gAA)
Theorem 1.5. The family {P,} is strongly continuous from the right.
because
Let h.(t) be the function whose graph is shown in Figure 19.4. We have
and
468 FUR THER SPECTRAL THEOREMS [XIX, §1]
C c+€ c+6+€
Figure 19.4
lim (~ - ~-.) = Qc
.-0
is the projection on the c-eigenspace of A.
fb(A) = A - bI = (c - b)I
ut-+(Au, v)
on the space DA , of such vectors v. We call the pair (A*, DA ,) the adjoint
of A.
Let J: H x H --. H x H be the operator such that J(x, y) = (- y, x).
Then J2 = -I. We note that the graph GA , of A* is given by the
formula
If A is closed, then A ** = A.
is also Cauchy, hence {(A - I)u n} is Cauchy, and {un} is Cauchy, say
converging to u. But
Then
<Au, v) = <u, AU 1 + AU 1 - AV),
whence
«A + II)u, v) = «A + II)u, u 1 ),
This proves that v = u 1 , as was to be shown.
Remark. In the literature, you will find that the dimension of the
cokernel of (A + 'u)DA is called a defect index. We are concerned here
with a situation in which the defect indices are O.
Multiplying this on the left by R(z) yields the resolvent formula of the
theorem, whose proof is concluded.
[XIX, §2] SELF-ADJOINT OPERATORS 473
We write
R(i) = (A - if)-i = C + iB
where B, C are bounded hermitian. From the resolvent equation between
R(i) and R( - i) we conclude that B, C commute. We may call B the
imaginary part of (A - i/rl, symbolically
B = Im(A - il)-i .
IfuEHn , then
<un' A*v) = <Aun, v),
<un, (A*v}n) = <Aun, vn),
(A - iIr t = C + iB
as above.
1 1
n+1 n
Figure 19.5
[XIX, §2] SELF-ADJOINT OPERATORS 475
Q. = (}.(B) = PI t. - PI t(.H)·
Then
/i(t) T/(t)
(a) (b)
Figure 19.6
°
Then 1 - e = I] and I](B) = because the spectral family for B is continu-
ous at 0, in view of Lemma 2.5 (kernel B = 0) and Lorch's theorem
(Theorem 1.6).
Let s.(t) be the function whose graph is shown in Figure 19.7. Then
\ {I/I if _I_<I"!
'-- s,,(I) = n+1 n
o otherwise
I I
n+1 n
Figure 19.7
Let Hn = QnH. Then H is the orthogonal direct sum of the spaces Hn,
the restriction of A to Hn is a bounded operator An, and
Proof Since tsn(t) = 0n(t), we get Bsn(B) = On(B) = Qn. Then by Lemma
2.5
then
L =-.1
f r div F dx dy JBdr
JU
=
U
F· n ds,
of of
w(x, y) = ox f dy - oyf dx.
478 FUR THER SPECTRAL THEOREMS [XIX, §3]
Then
iPf + (af)2] 2f f
dw = [ ax2f ax dx 1\ dy - [aay2f + (aay )2] dy 1\ dx
fJur dw = r
JBdU
w.
where (r, e) are polar coordinates. (No fancy integration is needed here,
but in the language of integration, we could say that log r is locally L 1.)
Now we may view G as defining an integral operator
f
by the formula
1
-2
n
f
&2
(log \wl)f(w + z') dw.
Standard theorems from advanced calculus justify the fact that you can
differentiate under the integral sign because f is assumed to be smooth
[XIX, §3] EXAMPLE: THE LAPLACE OPERA TOR IN THE PLANE 479
with compact support, thus showing that SGf is COO. (Cf. Lemma 2.2 of
Chapter VIII.) If D is the open disc of radius 1, then it is immediate that
SG maps Ccoo(D) into BCoo(D), namely that SGf is bounded.
We now have the fundamental formula:
L 3.
For the proof, let r = Iz - z/l, and view f in terms of the polar coordi-
nates (r, e). Apply the Stokes-Green formula to the region U(e) outside a
circle of radius e, so that the boundary is the circle S(e) of radius e with
reversed orientation. We have ds = de if we parametrize the circle of
radius e by
°e~ ~ 2ne.
f (f ~
S(e)
Og
un
- g~
Of)
un
ds = 10
2 1[' 1
f(e, e)-2 de -
ne
10
2 1[' 1
-2
n
of
(log e)~ de.
ur
As e --> 0, the first term goes to f(O, 0) and the second term goes to 0.
Hence the desired formula follows.
Spectral Measures
<
fl-+ f(A) v, v)
<q;(A)v, v) = t q; dll v ·
thus giving the desired bound for the measure Ilv, and concluding the
proof.
The measure liv is called the spectral measure associated with A and v.
By polarization, for v, w E H we see the existence and uniqueness of a
complex measure liv,w such that
<q;(A)v, w) = t q; dllv,w
II Ilv,wII ~ Ivllwl·
The measure Ilv,w is also called the spectral measure associated with A, v,
w. Applying the defining formula for Ilv,w to a real valued function q;, we
see immediately that
<f(A)v, w) = f f dJ1.v.w·
We obtain
= f qJn d/L"'(A)V,W'
which converges to
f f d/L"'(A)V,w = <f(A)iP(A)v, w)
if we use the second expression on the right. This takes care of one
factor. We take care of the other by using a sequence {iPn} converging to
g in the same manner as above. This proves SPEC 1, and also proves
the equivalent formula
the series
L <f(A}vn , vn >
n
The formalism of the five SPEC properties extends at once to the case of
an unbounded operator A. For example, in the case of SPEC 4, note
that
where g(t} = tf(t}.
It follows that
L f(A}v n E DA ,
for all v E H.
d
ds K(s) = - AK(s).
The proof is an exercise, which can be carried out by using the case for
bounded operators, and the direct sum decomposition of Theorem 2.6
for unbounded operators. I took the statement from Faltings [Fa 1],
Lemma 3.4. Readers can find another idea for the existence proof in this
reference, which also proves the uniqueness.
Observe that for each positive s, the function fP.(t) = e- st is bounded
on the spectra of the bounded operators An on the components Hn, and
sufficiently uniformly so that there is no difficulty in handling the exis-
tence part of the proof by considering its effect on infinite sums L Vn • As
in Proposition 1.3 the existence proof by this method is not invariant,
but the uniqueness show that the end result fPs(A) is independent of the
direct sum decomposition of H. One may write
t CPt (x) dx = 1.
f-" -00
+ foo
"
CPt < (j'
Suppose given an association fr-. f(A) satisfying the five spectral prop-
erties. For each v, WE H there is a unique measure J-tv.w such that
(f(A)v, w) = f f dJ-tv.w ·
Let z be complex and not real. The function f(t) such that
1
f(t)=-
t-z
(A - zI)f(A) = 1.
This means that the resolvent has the integral expression
«A - zl) -1 v, w) = f R
- 1
t- z
dJ-tv.w(t).
f R t~O m
1.
t/I(A) dJ-tv(A) = lim -2 f R
([R(A + ie) - R(A - ie)]v, V)t/I(A) dA.
[XX, §2] UNIQUENESS OF THE SPECTRAL MEASURE 487
Lemma 2.3. Let j.J. be a positive regular measure on R such that j.J.(R) is
finite. Then for", E Cc(R) we have
Furthermore, if A1 < A2 are real and such that the set {A1' A2} has j.J.-
measure 0, then
1 fA2
lim -
fco (A)2
e
2 dj.J.(t) dA =
fA2
dj.J.(A) .
.... 0 1t Al -co t- +e Al
Proof. First observe that the family of functions
1 2ie
t - A- ie - t - A + ie = (t - A)2 + e2'
488 SPECTRAL MEASURES [XX, §3]
(iv) If f is bounded, then f(A) has the previous meaning, and if f(t) = t,
then f(A) = A.
Proof. Observe first that the integral in (ii) exists by the Schwarz
inequality. To prove the theorem, let
strongly.
Let Bn = I.(A), so that Bn is bounded, and operates on Hn through the
projection on Hn because InXn = Xnin = In' whence
Let I(A) = B be the self-adjoint operator whose domain is the usual one,
consisting of v = L
Vn with Vn E Hn and L
IBnvnl2 < 00. Then
= f~ 1.2 d}lv
= fI d}lv'
This proves everything except the final assertion that if I(t) = t, then
I(A) = A. But this follows from the fact that In(A) is equal to A re-
490 SPECTRAL MEASURES [XX, §4]
stricted to Hn = EnH, since fn(A) and Xn(A) have the usual meaning, as in
§1. This concludes the proof of the theorem.
The first condition means that if Ha and Hb are the subspaces on which
Pa and Pb project H, then Ha c Hb and Pa projects Hb on Ha. The second
means that for each vector v E H we have
"'b(t) = 1 if t ~ b
"'b(t) = 0 if t> b.
Figure 20.1
Then let t --+ -00 to get the first limit. For t --+ 00 consider v - p,v.
Finally, we want to prove continuity from the right, i.e. for v E H we
want to show
lim (P'H - p,)v = o.
0-0
We look at
f f
cp dh = cp dJl.v,
is approximated by a sum
Instead of starting with a self-adjoint operator as in the text, one may start with
a spectral family, develop the functional calculus, and get back (unbounded)
operators as follows.
[XX, §6] EXERCISES 493
the limit being taken in the same sense as in the text, for the size of the
partition tending to O. Deduce the existence and uniqueness of a measure Ji.v
such that
2. Conclude that there exists a unique bounded operator J(P) for each J E BM(R)
L
such that
3. Show that the map J~ J(P) is a linear map from BM(R) into the space of
operators, satisfying the five properties SPEC 1 through SPEC 5, except that
A is replaced by P.
for all t. If b is a real number and I/Ib is the function whose graph is drawn
below, then
•b
if t ~ b
otherwise
b
Proof. From the assumption it follows that if <p E Cc(R), then <p(P.) = <p(Q,)
for all t. Let {go} be a sequence decreasing to I/Ib as shown.
~g"(t)
1
b b+-
n
Then for a fine partition,
Since
Global Analysis
One of the most attractive things that can be done with analysis is to
mix it up with the global topology of geometric structures. For instance,
whereas the local existence theorem for differential equations yields inte-
gral curves in an open set of say euclidean space, one may wish to see
what happens if a differential equation is given on the sphere. In this
case, the integral curves wind around the sphere and one investigates their
behavior as time goes to infinity. Similarly, one can work on toruses, or
arbitrarily complicated similar structures, which have one thing in com-
mon: locally, they look like euclidean space, but globally they turn and
twist. The relations between the analytic properties, and the algebraic-
topological invariants associated with the topological structure, constitute
one of the central parts of mathematics. Our task here is but to lay
down the most basic definitions to prepare readers for further readings,
and to give them the flavor of global results, as distinguished from local
ones in open sets of euclidean space.
We should add, however, that even on open sets of euclidean space,
i.e. locally, we may be interested in certain objects and properties which
are invariant under CP changes of coordinate systems, i.e. under CP iso-
morphisms. The language of manifolds provides the natural language for
such properties. Thus we begin with the change of variables formula,
which gives an example how the integral changes under C 1 isomor-
phisms. The change is of such a nature that we can associate with it an
integral on manifolds. This is done in the last chapter, which includes
the basic theorem of Stokes.
We don't do too much with differential equations besides defining the
basic notions on manifolds. Readers can refer to [La 2] for further
foundations. Smale's survey [Sm 2] is an excellent starting point for the
496 GLOBAL ANALYSIS [PART SIX]
Local Integration of
Differential Forms
We recall that a set has measure 0 in Rn if and only if, given e, there
exists a covering of the set by a sequence of rectangles {R j } such that
L f.1(Rj ) < e. We denote by Rj the closed rectangles, and we may always
assume that the interiors RJ = Int(R) cover the set, at the cost of in-
creasing the lengths of the sides of our rectangles very slightly (an e/2n
argument). We shall prove here some criteria for a set to have measure
O. We leave it to the reader to verify that instead of rectangles, we could
have used cubes in our characterization of a set of a measure 0 (a cube
being a rectangle all of whose sides have the same length).
We recall that a map f satisfies a Lipschitz condition on a set A if
there exists a number C such that
for all x, YEA. Any C 1 map f satisfies locally at each point a Lipschitz
condition, because its derivative is bounded in a neighborhood of each
point, and we can then use the mean value estimate,
Note. All three lemmas may be viewed as stating that certain parame-
trized sets have measure O. Lemma 1.3 shows that parametrizing a set by
strictly lower dimensional spaces always yields an image having measure
O. The other two lemmas deal with a map from one space into another
of the same dimension. Observe that Lemma 1.3 would be false if f is
only assumed to be continuous (Peano curves).
We first deal with the simplest of cases. We consider vectors Vi' ... ,Vn in
Rn and we define the block B spanned by these vectors to be the set of
points
[XXI, §2] CHANGE OF VARIABLES FORMULA 499
with 0 1. We say that the block is degenerate (in Rn) if the vectors
~ ti ~
are linearly dependent. Otherwise, we say that the block is
VI' ... ,V n
non-degenerate, or is a proper block in Rn.
taking the + if Det(v l , ... ,Vn) > 0 and the - if Det(v l , ... ,Vn) < O. The
determinant is viewed as the determinant of the matrix whose column
vectors are V I ' ... , Vn , in that order.
We recall the following characterization of determinants. Suppose that
we have a product
which to each n-tuple of vectors associates a number such that the prod-
uct is multilinear, alternating, and such that
if e l , ... ,en are the unit vectors. Then this product is necessarily the
determinant, i.e. it is uniquely determined. "Alternating" means that if
Vi = Vj for some i =I j then
The uniqueness is easily proved, and we recall this short proof. We can
write
500 LOCAL INTEGRATION OF DIFFERENTIAL FORMS [XXI, §2]
The sum is taken over all maps 0-: {1, .. . ,n} ..... {1, . .. ,n}, but because of
the alternating property, whenever 0- is not a permutation the term corre-
sponding to 0- is equal to O. Hence the sum may be taken only over all
permutations. Since
and
for CE R,
V, V 2 , ..• 'Vn then B(kv, V2, • . • ,vn ) is the union of the two sets
Vol(kv, V2, ... ,vn ) = Vol{(k - l)v, + Vol(v, V2, ..• ,vn )
V2 , ... ,vn )
as was to be shown.
Now let
v = vdk
for a positive integer k. Then applying what we have just proved shows
that
we conclude that
502 LOCAL INTEGRATION OF DIFFERENTIAL FORMS [XXI, §2]
Letting c, c' approach r as a limit, we conclude that for any real number
r ~ 0 we have
B(v, V 2 , . .. ,vn )
by -v so that these two blocks have the same volume. This proves the
first assertion.
As for the second, we look at the geometry of the situation, which is
made clear by the following picture in case v = V1, W = V2.
VOIO(C1V 1 + .. . +CnV n ' V2, .•• ,vn ) = VoIO(C1V 1 + ... + Cn-1Vn- 1' V 2 , ... ,vn )
= VoIO(C1V 1, V 2 , • • •,vn )
From this the linearity follows at once, and the theorem is proved.
[XXI, §2] CHANGE OF VARIABLES FORMULA 503
Corollary 2.2. Let S be the unit cube spanned by the unit vectors in Rn.
Let A.: Rn -+ Rn be a linear map. Then
Proof If Vl , ... ,Vn are the images of e l , . . . ,en under A, then A(S) is the
block spanned by Vl , . . .,Vn • If we represent A by the matrix A = (a i ),
then
and hence Det(v l , .. . ,vn ) = Det(A) = Det(A). This proves the corollary.
The next theorem extends Corollary 2.3 to the more general case
where the linear map A is replaced by an arbitrary Cl-invertible map.
The proof then consists of replacing the C 1 map by its derivative and
estimating the error thus introduced. For this purpose, we define the
Jacobian determinant
where Jj(x) is the Jacobian matrix, and f'(x) is the derivative of the map
f: U -+Rn.
on the cube S. (We have made suitable translations which don't affect
volumes.) We have
A-1 0 f(x} = X +r l 0 q>(x},
so that A-1 0 f is nearly the identity map. For some constant C, we have
for XES:
IA -1 0 q>(x} I ~ Ceo
From the lemma after the proof of the inverse mapping theorem, we
conclude that A-1 0 f(S} contains a cube of radius
(1 - Ce)(radius S)
and trivial estimates show that A-1 0 f(S} is contained in a cube of radius
(1 + Ce)(radius S).
with some fixed constant C1. Summing over j and estimating ILlfl, we
see that our theorem follows at once, in case R is a cube.
We may assume g real. The sup and inf of g on f(Sj) differ only by e if b
is taken sufficiently small. Using the theorem, applied to each Sj' and
replacing g by its minimum mj and maximum Mj on Sj we see that the
corollary follows at once.
the result for rectangles and continuous functions (corollary of the pre-
ceding theorem):
so the Cauchy nature of the sequence {gn is clear from that of {gd. It
follows that the restriction of (g 0 f) IAJI to R is the L I-limit of {gn, and
is in ;tJI(R). It also follows that the formula of the theorem holds for R,
that is
when A = R.
The theorem is now seen to hold for any measurable subset A of R,
°
since f(A) is measurable, and since a function g in ;tJI(j(A)) can be
extended to a function in ;tJI(j(R)) by giving it the value outside f(A).
From this it follows that the theorem holds if A is a finite union of
rectangles contained in U. We can find a sequence of rectangles {Rm}
contained in U whose union is equal to U, because U is separable.
Taking the usual stepwise complementation, we can find a disjoint se-
quence of measurable sets
whose union is U, and such that our theorem holds if A = Am. Let
and
fJ(A)
gdfl=i (gof)IAJldfl.
A
Proof. Let Uo be the interior of A. The sets f(A) and f(Uo) differ
only by a set of measure 0, namely f(8A). Also the sets A, Uo differ only
[XXI, §3] DIFFERENTIAL FORMS 507
Note. Since step maps are dense in yl(X, E) for a Banach space E,
the preceding proof generalizes at once to the case of Banach valued
maps.
We recall first two simple results from linear (or rather multilinear) alge-
bra. We use the notation E(r) = E x Ex··· x E, r times.
Theorem B. For each pair of positive integers (r, s), there exists a
unique product (bilinear map)
The proofs for these two statements will be briefly summarized in the
appendix to this chapter.
Let E* be the dual space, E* = L(E, R). (We prefer here to use E*
rather than E', first because we shall use the prime for the derivative, and
second because we want a certain notational consistency as in §4.) If
E = R n and Ai' ... ,An are the coordinate functions, then each Ai is an
element of the dual space, and in fact {Ai"" ,An} is a basis of this dual
space.
Let U be an open set in Rn. By a differential form of degree r on U
(or an r-form) we mean a map
w: U _1\' E*
from U into the r-th alternating product of E*. We say that the form is
of class CP if the map is of class CPo (We view 1\' E* as a normed vector
space, using any norm. It does not matter which, since all norms on a
finite dimensional vector space are equivalent.)
Since {Ai"" ,An} is a basis of E*, we can express each differential
form in terms of its coordinate functions with respect to the basis
where ft.i) = h, " ' i r is a function on U. Each such function has the same
order of differentiability as W. We call the preceding expression the
standard form of W. We say that a form is decomposable if it can be
written as just one term f(X)A i, 1\ ... 1\ Air' Every differential form is a
sum of decomposable ones.
We agree to the convention that functions are differential forms of
degree O.
It is clear that the differential forms of given degree r form a vector
space, denoted by n'(U).
Let E = Rn. Let f be a function on U. For each x E U the derivative
f'(X): Rn _ R
[XXI, §3] DIFFERENTIAL FORMS 509
/' : V ~ E*
,of of
f (x)h = ~hl + ... + :;-hn •
UX I UXn
of of
df(x) = ~ dX 1 + ... +:;- dx n·
UX I UXn
Let wand IjJ be forms of degrees rand s respectively, on the open set
U. For each x E V we can then take the alternating product w(x) /\ ljJ(x)
and we define the alternating product w /\ IjJ by
f /\ W = fw
W /\ ftjJ = fw /\ tjJ.
We shall now define the exterior derivative dw for any differential form
w. We have already done it for functions. We shall do it in general first
in terms of coordinates, and then show that there is a characterization
independent of these coordinates. If
we define
dw = I dJ;i) /\ dAi l /\ ••• /\ dA ir •
(i)
af ag
= - dy /\ dx + - dx /\ dy
ay ax
= (a f _ ag ) dy /\ dx
ay ax
have
d(fw) = df 1\ W +f dw.
of
df(x) = L -OX
n
j=l
dX
j
j
and
Using the fact that the partials commute, and the fact that for any two
positive integers r, s we have dX,1\ dx. = -dx. 1\ dx" we see that the
preceding double sum is equal to O. A similar argument shows that the
theorem is true for I-forms of type g(x) dXi where 9 is a function, and
thus for all I-forms by linearity. We proceed by induction. It suffices to
prove the formula in general for decomposable forms. Let w be decom-
posable of degree r, and write
where deg tf; = 1. Using the formula of Theorem 3.1 twice, and the fact
that ddtf; = 0 and dd'1 = 0 by induction, we see at once that ddw = 0, as
was to be shown.
Jl.0A:E~R
which we visualize as
f'(x): E --+ F
(f*w)(x) = f'(x)*w(j(x)).
Example 1. Let Yl' ... ,Ym be the coordinates on V, and let J.lj be the
j-th coordinate function, j = 1, . .. ,m, so that Yj = J.lj(Yl, ... ,Ym)' Let
f: U --+ V
514 LOCAL INTEGRATION OF DIFFERENTIAL FORMS [XXI, §4]
dx dy
f*w(t) = g(x(t), y(t») dt dt + h(x(t), y(t») dt dt
if we write f(t) = (x(t), y(t»). Let G = (g, h) be the vector field whose
components are 9 and h. Then we can write
(g 0 f)*(w) = f*(g*(w)) .
f*(dw) = df*w.
Proof. We first prove this last relation. From the definitions, we have
dg(y) = g'(y), whence by the chain rule,
and this last term is nothing else but d(g 0 f)(x), whence the last relation
follows. For a form of degree 1, say
Using the fact that ddfl = 0, together with Theorem 3.1, we get
We shall give brief reviews of the proofs of the algebraic theorems which
have been quoted in this chapter.
We first discuss "formal linear combinations". Let S be a set. We
wish to define what we mean by expressions
[XXI, §5] APPENDIX 517
where {cJ are numbers, and {sJ are distinct elements of S. What do we
wish such a "sum" to be like? Well, we wish it to be entirely determined
by the "coefficients" ci , and each "coefficient" Ci should be associated
with the element Si of the set S. But an association is nothing but a
function. This suggests to us how to define "sums" as above.
For each S E S and each number C we define the symbol
CS
are linearly independent. To prove this, suppose c l ' ... ,Cn are numbers
such that
and any given element has a unique such expression, because of the linear
independence of Sl' ... 'Sn' This justifies our terminology.
E(r) = Ex' .. x E,
518 LOCAL INTEGRATION OF DIFFERENTIAL FORMS [XXI, §5]
Then
where the sum is taken over all maps 11: {I, ... ,r} -+ {l, ... ,n}. In this
sum, all terms will be 0 whenever 11 is not an injective mapping, that is
whenever there is some pair i, j with i # j such that l1(i) = 11(j), because
of the alternating property of f. From now on, we consider only injec-
tive maps 11. Then {11(1), ... ,11(r)} is simply a permutation of some r-tuple
(il' ... ,ir ) with i l < .. , < ir •
Dets(A) = L
(1eP(S)
6s(l1)a l .(1(l)··· ar.(1(r)'
where 6s(l1) is the sign of 11, depending only on 11. In terms of this
notation, we can write our expression for f(u l , ... ,ur) in the form
where Vs denotes (ViI' ... ,V;) if il < ... < ir are the elements of the set S.
The first sum over S is taken over all subsets of 1, ... ,n having precisely
r elements.
V · A . . . A V·
II lr'
is a basis of /\r E.
Proof. For each subset S of {1, . . . ,n} consIstmg of precisely r ele-
ments, we select a letter ts . As explained at the beginning of the section,
these letters ts form a basis of a vector space whose dimension is equal
to the binomial coefficient (~). It is the space of formal linear combina-
tions of these letters. Instead of t s , we could also write to) = til'" ir with
i l < .. . < ir • Let {VI' ... ,vn } be a basis of E and let U 1 , ... ,Ur be elements
of E. Let A = (aij) be the matrix of numbers such that
Define
u1 A .•• A Ur = L Dets{A)t S '
s
Proof For each r-tuple (u i , ... ,ur) consider the map of E(s) into
I\r+s E given by
To obtain the desired product I\r E x 1\' E ~ I\r+, E, we simply take the
association
fJ. 0 A.: E ~ R
which we visualize as
A. *: I\r F* -+ I\r E*
given by
Property (i) now follows by linearity and the fact that decomposable
elements III /\ ... /\ Ilr generate I\r F*. Property (ii) comes from the de-
finition. This proves Theorem C.
CHAPTER XXII
Manifolds
We call (Xl"" ,Xn) the local coordinates of x in the chart (V, cp). The
notation here is already somewhat concise, but useful. If readers feel the
need for it, they may extend this notation as follows. Denote a point
of X by P. Then in a chart cp: V --+ Rn at P, we have coordinates
(Xl (P), ... ,xn(P)) for the point cp(P), P E V, and we abbreviate this n-tuple
by x(P). In most cases, it is a useful abbreviation to do away with the
extra letter.
Let V be a subset of X and let cp: V --+ cpV be a bijection of V onto
an open subset of E. We say that the pair (V, cp) is compatible with the
atlas {(Vi' CPJ} if each map CPiCP-l (defined on a suitable intersection as in
AT 3) is a CP -isomorphism. Two atlases are said to be compatible if each
chart of one is compatible with the other atlas. The relation of compati-
bility between atlases is immediately verified to be an equivalence rela-
tion. An equivalence class of CP-atlases on X is said to define a structure
of CP-manifold on X. The number n being fixed, we say that X is then
an n-dimensional manifold.
[XXII, §1] ATLASES, CHARTS, MORPHISMS 525
If (V, cp) and (V, t/J) are two charts of a manifold, then we shall call the
map cp 0 t/J-l (defined on t/J(V (\ V), whenever V (\ V is not empty) a
transition map.
So far we have not assumed that X has a topology. In many cases, a
topology is first given, and then to make the atlases topologically com-
patible with this topology, one can require the additional condition that
the maps cp of the charts be homeomorphisms. However, it is also useful
not to do this and deal with the more general situation when X is
merely a set, and we shall in fact have an important application later
when we deal with the tangent bundle.
We shall now see how to define a topology on X by means of the
atlases. Let {( Vi' cp;)} be an atlas. A subset V of X is defined to be open
if and only if the intersection V (\ Vi with each open set of the atlas is
such that CPi(V (\ V;) is open, in E of course. It is a trivial exercise to
verify that this defines a topology. Furthermore, if {(l-j, t/Jj)} is an equiva-
lent atlas, then the two topologies coincide. We leave the formal verifica-
tion to the reader. We note merely that the basic reason is that if a
point x lies in charts Vi and l-j, then there is a subset W containing x
such that CPi Wand t/Jj Ware open. Since a topology is really determined
locally (i.e. an open set is a union of open neighborhoods of its points)
one sees at once that a set is open relative to one atlas if and only if it is
open relative to the other.
Let X be a manifold, and V an open subset of X. Then it is possible,
in the obvious way, to induce a manifold structure on V , by taking as
atlases the intersections
(Vi (\ V, CP;l(Vi (\ V»).
Example 4. Let sn be the n-sphere in Rn+l, i.e. the set of all points
(x l' . . . ,Xn+l) such that
xi + ... + X;+l = 1.
526 MANIFOLDS [XXII, §1]
The sphere is the set of points x such that f(x) = 1. For any point
a E sn, a = (a 1 , ... ,an+d, some coordinate is not equal to 0, say a 1 • Then
DJ(a) =f. 0,
It is an exercise to verify that the collection of all similar pairs (((JV, ((J-l)
is a Ceo atlas for sn. Actually, we shall obtain some theorems below
which will prove this, and give general criteria showing that certain
subsets of euclidean space are manifolds.
is an atlas for the product, and the product of compatible atlases gives
rise to compatible atlases, so that we do get a well-defined product
manifold.
[XXII, §2] SUBMANIFOLDS 527
A manifold may arise like the torus, not embedded in any particular
euclidean space, or it may be given as a subset of some euclidean space
like the sphere. We now study this second possibility.
Let X be a topological space and Y a subspace. We say that Y is
locally closed in X if every point Y E Y has an open neighborhood U in
X such that Y 11 U is closed in U. We leave it to the reader to verify
that a locally closed subset of X is the intersection of an open set and a
528 MANIFOLDS [XXII, §2]
closed set in X. For instance any open subset of X is locally closed, and
any open interval is locally closed in the plane.
Let X be a manifold and Y a subset. We shall say that Y is a
submanifold if, roughly speaking at each point Y E Y there exists a chart
such that in this chart, the points of Y correspond to a factor in a
product space. We now make this condition precise as follows. For each
Y E Y there exists a chart (V, cp) at y such that cp gives an isomorphism
All of this explains what we said about Y being locally at each point a
factor in a product.
We observe that if Y is a submanifold of X, then Y is locally closed in
X. We must also justify our terminology by showing that Y is a mani-
fold in its own right. Indeed, if (V, cp) is a chart at y as in our definition,
then cp induces a bijection
and
is also a CP map.
the first map being an injection of V1 as a factor, and the third map a
projection on the first factor.
530 MANIFOLDS [XXII, §2]
0 0 0 0
0 0 0 0
0 0 0 0
DJ(a) DJ(a)
given by
qJ(x, yz} = (f(x), 0) + (0, yz)·
Then
at f(x o) in f(Ud which maps f(U 1 ) into a factor in the product Fl x F2.
Note that if we write E = Rq and F = Rn, then the subspace Fl of F is
not necessarily equal to Rq in its usual embedding in Rn as the space of
the first q coordinates. The subspace Fl can be quite arbitrary. How-
ever, we can find a complementary subspace F2, and then a basis of Fl
and of F2 in such a way that if we take coordinates with respect to this
basis, then the coordinates of g(J(U1 )) are precisely the coordinates
(Xl' ... ,Xq , 0, ... ,0). In our geometric terminology, we can say that f(Ud
is a submanifold of F.
The next result deals with the dual situation, where instead of an
injection we deal with a projection. If we have a map
then we shall say that this map is a projection (on the first factor) if this
map can be expressed as a composite
is a local CP isomorphism at a.
( 11
Dtf(a)
0)
Dzf(a)
and
as the unique linear map having the following property: If (U, cp) is a
chart at x and (V, "') is a chart at J(x) such that J(U) c V, and v is a
tangent vector at x represented by v in the chart (U, cp), then
Txf(v)
Tf(x)(Y) --4 F
j: Y --+ X
n: T(X) -+ X
which maps each tangent space Yx(X) on the point x of X. We call n the
natural projection. Let (U, qJ) be a chart of X, with qJU is open in E. We
then obtain a map
defined by
by
536 MANIFOLDS [XXII, §4]
for x E V n V and vEE. Since the derivative D(1jI a q> -1) is of class C r1
and is a linear isomorphism at x, we conclude that our family of maps
{!rp}, for (V, q» ranging over all charts of X, is an atlas for T(X), and
therefore that T(X) is a C r1 manifold, as we predicted it would be.
We call each chart (n-lV, !rp) a trivializing chart of T(X), over the open
set V. Locally, we see that each such trivializing chart for T(X) gives an
isomorphism of the tangent bundle over V with a product q>V x E.
Let f: X -+ Y be a CP morphism, p ~ 1. We can then define a tangent
map
Tf: T(X) -+ T(Y)
n- l (V)
- t.
q>V x E
Tlj j
n-l(V)
~ IjIV x F.
(The sum is taken over all i, but is in fact finite for any given point x in
view of PU 3.)
As a matter of notation, we often write that {(Vi' !/Ii)} is a partition of
unity if it satisfies the previous four conditions.
Proof Let Vi' V 2 , ••• be a basis for the open sets of X such that each
D; is compact. We can find such a basis since X is locally compact. We
construct inductively a sequence Ai, A 2 , ... of compact sets whose union
is X, such that Ai is contained in the interior of Ai+l' We start with
Al = V1 · Suppose that we have constructed Ai, ... ,Ai' Let j be the
smallest integer such that Ai is contained in Vi u··· u~. We let Ai+1 be
the closed and compact set
by a finite number of sets Wx " . .. ,Wx ",. Let fJli denote the family
{ Vx , ' • •• , VXrn } ' and let fJI be the union of all fJl i for all i = 1, 2, . . . . Then
fJI is an open covering of X, is locally finite, and is subordinated to our
given covering il/I. It also satisfies the other requirements of the theorem.
The sum is finite at each point, and we let Yk = t/Jd t/J. Then {(~, yd} is
the desired partition of unity.
We now recall the argument giving the function t/Jk . If 0 ~ a < b, then
the function defined by
-1
(t--- a--:-)-:-:(b-----:-
exp -:- t)
in the open interval a < t < band 0 outside the interval determines a
bell-shaped C OO function from R to R. Its integral from -00 to t divided
by the area under the bell yields a function which lies strictly between 0
and 1 in the interval a < t < b, is equal to 0 for t ~ a and is equal to 1
for t ~ b.
We can therefore find a real valued function of a real variable, say
l1(t), such that l1(t) = 1 for 1tl < 1 and l1(t) = 0 for 1tl ~ 1 + D with small
D, and such that 0 ~ 11 ~ 1. Then 11(lx1)2 = t/J(x) gives us a function
which is equal to 1 on the ball of radius 1 and 0 outside the ball of
radius 1 + D. (We denote by lithe euclidean norm.) This function can
then be transported to the manifold by any given chart whose image is
the ball of radius 3.
o~ f ~ 1, f = 1 on A, f = 0 on B.
ing consisting of CIJ A and {Ux } X EA' Let J be the set of those indices j
such that supp IXj C UxUl for some x(j) E A. Let
f= L IX
jEJ
j •
For any x E X there is only a finite number of functions IXj such that
IXj(X) =1= 0, so our sum expressing f is actually a finite sum. If x E A and
IX; has support in ClJA, then IX;(X) = O. Hence for each x E A we have
f(x) = 1. If x E B, then IXj(X) = 0 for each j E J so f(x) = O. For any
x E X we have 0 ~ f(x) ~ 1 because of the definition of a partition of
unity and the fact that we take our sum for f only over a subset of the
indices {j}. This proves our corollary.
This is easily proved. Indeed, we see at once that the kernels of A and
J1. must be equal. Suppose that A =1= O. Let Xo be such that A(Xo) > O.
Then J1.(x o) > 0 also. The functional
A - CJ1.
where c = A(Xo)/J1.(xo) vanishes on the kernel of A (or J1.), and also on Xo.
Therefore it is the 0 functional and c satisfies our requirement.
If we exclude the two end circles, then what is left is just an ordinary
manifold. However, if we include the two end circles, then we have an
object which, at each point of one of the end circles, does not look like
some open set in 2-space, but rather looks like a point at the boundary
of a half plane. In other words, we have a parametrization as indicated
by the following picture:
-
We shall formulate the definitions and lemmas which allow us to give a
formal development for such parametrizations.
Let E, F be euclidean spaces, and let E1 and Fit be two half spaces in
E and F, respectively. Let V, V be open subsets of these half spaces
respectively. We shall say that a mapping
f: V -+ V
o(t)
fl.f'(O)v ~ -t-fl.(wl ) .
Proof. For each x E U, we conclude from the chain rule that f'(x) is
invertible. Our first assertion then follows from Lemma 5.2. We also see
that no interior point of U maps on a boundary point of V and con-
versely. Thus f induces a bijection of au and av, and a bijection of
Int(U) on Int(V). Since these interiors are open in their respective spaces,
it follows that f induces an isomorphism between them. As for the
boundary, it is a submanifold of the full space, and locally, our definition
of the derivative, together with the product structure, shows that the
restriction of f to au must be an isomorphism on av.
We see that Lemma 5.3 gives us the invariance of the boundary under
cP maps (p ~ 1), first for open subsets of half spaces, but then also
immediately for the boundary of a manifold since the property reduces at
once to such subsets, under charts.
We can then describe local coordinates at a point in a manifold with
boundary as follows. If the point is not a boundary point, then a neigh-
borhood of this point is described by coordinates (Xl'''' ,X n ) in some
open set of R n, which we may even take to contain 0, and such that 0
corresponds to the given point.
If the point is a boundary point, then an open neighborhood can be
described by coordinates (Xl' .,. ,xn ) satisfying
We can define a chart for part of the cylinder in terms of the three
e,
coordinates (r, z) satisfying the inequalities :
r = c,
~: X -+ T(X)
such that ~(x) lies in the tangent space Tx(X) for each x E X, or in other
words, such that n 0 ~ = id x . Thus a vector field assigns a tangent vector
to each point.
When we identify the tangent bundle of an open set U in E with the
product U x E relative to a chart (U, cp), then we see that a vector field
corresponds to a map
U-+UxE
such that
~(x) = (x, f(x))
a': J -+ T(X),
which is such that a'(t) lies in 7;.(t)(X). We shall also write da/dt instead
of a'(t), following standard notation, consistent with previous notation
when we studied vector fields on open sets of vector spaces.
We say that a is an integral curve for the vector field ~ if we have
a'(t) = ~(ex(t))
[XXII, §6] VECTOR FIELDS 545
for all t E J. If J contains 0 and a(O) = xo, we say that Xo is the initial
condition of a. The theorems on differential equations proved in Chapter
XIV, §3, §4, §5 can now be formulated on manifolds.
e
is an integral curve for with initial condition x. When we select a chart
at a point x of X, then we see that this definition of flow coincides with
the definition we gave for open sets in euclidean spaces for the local
representation of our vector field. As in Chapter XIV, we abbreviate
a(t, x) by tx.
J(tox) = J(x) - to
Corollary 6.4. Let :Dt(~) be the set of points x of X such that (t, x) lies
in :D(~). Then :Dt(~) is open for each t E R, and at is a CP-isomorphism
of :Dt(~) onto an open subset of X. In fact, at(:D t) = :D- t and a;l = a_to
denoted by 1\ T*(X).
By a differential form on X (of degree r) we shall mean a map
w: X ~ 1\' T*(X)
such that for each x the value w(x) lies in 1\' T,,* . (We shall add differ-
entiability conditions in a moment.) The set of differential forms is a
vector space denoted by n'(X).
If f : X ~ Y is a CP map of manifolds, then we obtain an induced map
f* : n'(Y) ~ n'(X),
548 INTEGRATION AND MEASURES ON MANIFOLDS [XXIII, §1]
just as in the case of subsets of euclidean spaces, and arising from the
induced linear map at each point,
cp: U ~ cpU c Rn
which to each vector vERn associates the class of (U, cp, v). If A. is a
functional on T", then A. 0 atp is a functional on Rn. Let w be a I-form on
U, and Wx the value of w at x. Then Wx can be pulled back to R", to
obtain the form
and we say that the expression on the right is the local expression of w
determined by the chart, or corresponding to the chart cpo We shall also
commit the abuse of notation, writing g(i)(X) instead of g(i)(X l , ... ,x").
We abbreviate
for simplicity.
If (V, "') is another chart such that Un V is not empty (so that our
two charts (U, cp) and (V, "') may be viewed as charts at a common
point), then we obtain a representation of w determined by",. If w is a
1-form, then W x is a functional on Tx , and the pull backs of Wx to R" by
atp and a", respectively can be visualized in the following diagram:
Rn
Rn
[XXIII, §1] DIFFERENTIAL FORMS ON MANIFOLDS 549
The vertical map on the left is simply the derivative (qJ 0 1/1-1)'(I/IX), i.e.
the derivative at I/Ix of the transition map qJ 0 1/1-1 giving the change of
charts. In terms of local coordinates, the change in the local representa-
tion of Wx is given in terms of partial derivatives, which are of class CP-1.
Similarly, for any r-form the change in the local representation is given
by certain subdeterminants of the Jacobian matrix of (qJ 0 1/1-1)'(I/IX), and
is again of class C p - 1 • The most important case is that of an n-form, and
we can then write w locally as
wi = g(x) dX 1 A ... A dx n •
defined on the space of r-forms (of class 0, into the space of forms of
class Cq-1), satisfying the properties that if deg w = r, then
If g: Y -+ Z is a CP map, then
(g 0 f)* = f* 0 g*.
can form the product aw, which is the form whose value at x is a(x}w(x}.
If a has compact support, then aw has compact support. Later, we shall
study the integration of forms, and reduce this to a local problem by
means of partitions of unity, in which we multiply a form by functions.
If X is a manifold and Y a submanifold, then any differential form on
X induces a form on Y. We can view this as a very special case of the
inverse image of a form, under the embedding (injection) map
id: Y -+ X.
where the roof over dXj means dXj is omitted. We should denote this
induced form by Wy, although occasionally we omit the subscript Y. We
shall use such an induced form especially when Y is the boundary of a
manifold X.
Then (x z , .. . ,xn ) are the local coordinates for a chart of the boundary,
namely the restriction of q> to ax (') U, and the picture is as follows.
------Xl
We may say that we have considered a chart q> such that the manifold
lies to the left of its boundary. If readers think of a domain in R Z,
having a smooth curve for its boundary, as on the following picture, they
will see that our choice of chart corresponds to what is usually visualized
as "counterclockwise" orientation.
The collection of all pairs (U (') ax, q>1(U (') ax»), chosen according to
the criteria described above, is obviously an atlas for the boundary ax,
and we contend that it is an oriented atlas.
and
f(x) where f = (f1' ... ,fn) is the transition mapping. Since we deal with
oriented charts for X, we know that .1f (x) > 0 for all x. Since f maps
boundary into boundary, we have
T. .
(0, X 2 , •• • ,xn ) is equal to
D 1f 1 (0, X2"
) _ I1m
.. ,Xn -
' f1 (h, X2 ' . • . , X n )
h '
h--O
taking the limit with h < 0 since by prescription, points of X have coor-
dinates with Xl < O. Furthermore, for the same reason we have
Consequently
Dd1(0, X 2 , • • •,xn ) > O.
From this it follows that .1~n-1)(X2' ... ,xn ) > 0, thus proving our assertion
that the atlas we have defined for ax is oriented.
w(x) = f(x) dX l 1\ dX n
is the local representation of w in this chart, then for any 9 E Cc(X) with
support in U, we have
(1) Ag = f"'u
g",(x) If(x) I dx,
w(x) = f(x) dX l 1\ . .. 1\ dX n
is the local representation of w in this chart, then for any 9 E Cc(X) with
support in U, we have
Ag = f g",(x)f(x) dx,
"'u
where g", represents 9 in the chart, and dx is Lebesgue measure.
[XXIII, §4] STOKES' THEOREM FOR A RECTANGULAR SIMPLEX 555
Ix gw = Ag.
f w=~f
x 'x
Or:iW,
all but a finite number of terms in this sum being equal to O. As usual,
it is immediately verified that this sum is in fact independent of the
choice of partition of unity, and in fact, we could just as well use only a
partition of unity over the support of w. Alternatively, if Or: is a function
in CAX) which is equal to 1 on the support of w, then we could also
define
Let
pieces
R? = [ai' bl ] x ... x {ai} x ... x [an, bn],
Rt = [al,b l ] x ... x {b;} x ... x [an,bnJ.
If
/'-.
W(XI' ... ,xn) = f(x l , ... ,xn) dX 1 1\ .. . 1\ dXj 1\ .. . 1\ dX n
is an (n - 1)-form, and the roof over anything means that this thing is to
be omitted, then we define
f cO R
= L (_1)i [f
n
i=l R?
- f] R[
.
fR dw = faOR
w.
b
2
a2
-R? D -R~
R}
clear that
fR? w = 0= f Ri
w,
so that
(The (_l)j-l comes from interchanging dXj with dx 1 , ••• ,dxj _1 • All other
terms disappear by the alternation rule.)
Integrating dw over R, we may use repeated integration and integrate
ojj OXj with respect to Xj first. Then the fundamental theorem of calculus
for one variable yields
J. L
dw = a* dw.
f f
0"
w =
oOR
a*w,
558 INTEGRATION AND MEASURES ON MANIFOLDS [XXIII, §5]
f f
x
dOJ =
oX
OJ.
and this sum has only a finite number of non-zero terms since the
support of OJ is compact. Using the additivity of the operation d, and
that of the integral, we find :
Suppose that OC i has compact support in some open set J.'i of X and that
we can prove
in other words we can prove Stokes' theorem locally in J.'i. We can write
and similarly
which yields Stokes' theorem on the whole manifold. Thus our argument
with partitions of unity reduces Stokes' theorem to the local case, namely
it suffices to prove that for each point of X there exists an open neigh-
borhood V such that if w has compact support in V, then Stokes' theo-
rem holds with X replaced by V. We now do this.
If the point is not a boundary point, we take an oriented chart (U, <p)
at the point, containing an open neighborhood V of the point, satisfying
the following conditions: <pU is an open ball, and <p V is the interior of a
rectangle, whose closure is contained in <pU. If w has compact support
in V, then its local representation in <p U has compact support in <p V.
Applying Stokes' theorem for rectangles as proved in the preceding sec-
tion, we find that the two integrals occurring in Stokes' formula are
equal to 0 in this case (the integral over an empty boundary being equal
to 0 by convention).
Now suppose that we deal with a boundary point. We take an
oriented chart (U, <p) at the point, having the following properties. First,
<pU is described by the following inequalities in terms of local coordi-
nates (x l ' ... ,Xn ):
Next, the given point has coordinates (1,0, ... ,0), and that part of U on
the boundary of X, namely Un ax, is given in terms of these coordi-
nates by the equation Xl = 1. We then let V consist of those points
whose local coordinates satisfy
-If------f
In the sum giving the integral over the boundary of a rectangle as in the
previous section, only one term will give a non-zero contribution, corre-
560 INTEGRATION AND MEASURES ON MANIFOLDS [XXIII, §5]
sponding to i = 1, which is
r dw = Jvr ncJx w,
Jv
which proves Stokes' theorem locally in this case, and concludes the
proof of Theorem 5.1.
lim Jik(Uk) = o.
k- oo
f f
X
dw =
oX
w.
[XXIII, §6] STOKES' THEOREM WITH SINGULARITIES 561
lim
k-+co
f
eX
gk W = f
eX
W.
Similarly,
lim
k-+co
f
X
gk dw = f
X
dw.
The vertex and the circle surrounding the base disc prevent the cone
from being a submanifold of R3. However, if we delete the vertex and
this circle, what remains is a submanifold with boundary embedded in
R3. The boundary consists of the conical shell, and of the base disc
(without its surrounding circle). Another example is given by polyhedra,
as on the following figure.
Xn + l = ... = XN = 0 and
and the points of the frontier of X which lie in the chart are those with
coordinates satisfying
The set of all regular frontier points of X will be denoted by ax, and
will be called the boundary of X. We may say that X u ax is a sub-
manifold of RN, possibly with boundary.
A point of the frontier of X which is not regular will be called
singular. It is clear that the set of singular points is closed in RN. We
[XXIII, §6] STOKES' THEOREM WITH SINGULARITIES 563
From our first condition, we see that gkOJ vanishes on an open neighbor-
hood of S. Since gk = 1 on the complement of Uk' we have dg k = 0 on
this complement, and therefore our second condition implies that the
measures induced on X near the singular frontier by Idgk /\ OJI (for k =
1, 2, . . .), are concentrated on shrinking neighborhoods and tend to 0 as
k ..... 00.
f
x
dw = fax w.
Proof. Let U, {Ud, and {gd satisfy conditions NEG 1 and NEG 2.
Then gkW is 0 on an open neighborhood of S, and since w is assumed to
have compact support, one verifies immediately that
We have
lim
k~co
f
X
gk dw = fX
dw.
Proof Let U, {Ud, {gd and V, {l'k}, {hd be triples associated with S
and T, respectively, as in conditions NEG 1 and NEG 2 (with V re-
placing U and h replacing g when T replaces S). Let
W=Uuv, and
Then the open sets {l¥,.} form a fundamental sequence of open neighbor-
hoods of S u T in W, and NEG 1 is trivially satisfied. As for NEG 2, we
have
Lemma 6.2. Let S be a compact subset of Rn. Let Uk be the open set
of points x such that d(x, S) < 2jk. There exists a Coo function gk on R n
which is equal to 0 in some open neighborhood of S, equal to 1 outside
Uk' 0 ~ gk < 1, and such that all partial derivatives of gk are bounded
by C1 k, where C1 is a constant depending only on n.
cp(x) = 0 if 0 ~ IIxll ~ I j 2,
cp(x) = 1 if 1~ Ilxll.
566 INTEGRA TION AND MEASURES ON MANIFOLDS [XXIII, §6]
We use I I for the sup norm in Rn. The graph of <p looks like this:
-1 -t t
For each positive integer k, let <Pk(X) = <p(kx). Then each partial deriva-
tive D;<Pk satisfies the bound
This function has the same shape as <Pk but is translated to the point
ll 2k. Consider the product
For this 1 we have d(l12k, S) < 11k, so that this 1 occurs in the product,
and
<Pk(X - ll 2k) = o.
Therefore gk is equal to 0 in an open neighborhood of S. If on the other
hand we have d(x, S) > 21k and if 1 occurs in the product, that is
d(//2k, S) ~ 11k,
then
d(x, ll 2k) > 11k
f f
x
dw =
eX
w.
beginning of the section. Criterion 2 is also the natural one when dealing
with manifolds defined by algebraic inequalities. By using the resolution
of singularities due to Hironaka one can parametrize a compact set of
algebraic singularities as in Criterion 2.
Finally, we note that the condition that ill have compact support in an
open neighborhood of X is a very mild condition. If for instance X is a
bounded open subset of R", then X is compact. If ill is any form on
some open set containing X, then we can find another form 1'/ which is
equal to ill on some open neighborhood of X and which has compact
support. The integrals of 1'/ entering into Stokes' formula will be the
same as those of ill. To find 1'/, we simply multiply ill with a suitable COO
function which is 1 in a neighborhood of X and vanishes a little further
away. Thus Theorem 6.3 provides a reasonably useful version of Stokes'
theorem which can be applied easily to all the cases likely to arise
naturally.
Bibliography
[A-B] M . ATIYAH and R. BOTT, "The Lefschetz fixed point theorem for ellip-
tic complexes," Annals of Math ., 86 (1967) pp. 374-407.
[A- Si] M . ATIYAH and I. SINGER, "The index of elliptic operators," Annals of
Math ., 87 (1968) pp. 484- 530, 546-604.
[A- Se] M . ATIYAH and G. SEGAL, "The index of elliptic operators," Annals of
Math ., 87 (1968) pp. 531 - 545.
[Ab-R] R. ABRAHAM and JOEL ROBBIN, Transversal Mappings and Flows,
Benjamin, New York, 1967.
[Ba] K. BARNER, Einfuhrung in die Analytische Zahlentheorie, 1990.
[BGV] N . BERLINE, E. GETZLER, and M . VERGNE, Heat Kernels and Dirac
Operators, Grundlehren der Math. Wiss. 298, Springer-Verlag, New
York, 1992. (This book contains an extensive useful bibliography.)
[Bo] N. BOURBAKI, General Topology, Addison-Wesley, Reading, Mass.,
1968.
[Di] J . DIEUDONNE, Foundations of Modern Analysis, Academic Press, New
York, 1960.
[Din] N. DINCULEANU, Vector Measures, Veb Deutscher Verlag, Berlin, 1966.
[Dix] J. DIXMIER, Les C*-Algebres et Leurs Representations, Gauthier Villars,
Paris, 1964.
[Du] N . DUNFORD and J. SCHWARTZ, Linear Operators, Interscience, New
York, 1958.
[Fa] L. F ADDEEV, Expansion in Eigenfunctions of the Laplace Operator on
the Fundamental Domain of a Discrete Group on the Lobacevskii
Plane, AMS Trans!. Trudy (1967) pp. 357-386.
[Fal] G . FAL TINGS, Lectures on the Arithmetic Riemann- Roch Theorem,
Annals of Math. Studies, 127, Princeton University Press, Princeton,
NJ, 1992.
[Fo 1] B. FOLLAND, Real Analysis, Wiley- Interscience, New York, 1984.
570 BIBLIOGRAPHY
A space 51
Barner's theorem 291
Absolutely continuous 191, 199 Base 23
Adherent 22 Bessel function 245
Adjoint 106, 392, 438, 469 inequality 102
of differential operator 303 Bijective 3
Alaoglu's theorem 71 Bilinear 67, 91
Algebra 51, 72, 113 Block 498
of functions 51 Bonnet mean value 286
of subsets 113 Borel set or measurable 114
Algebra automorphism 61 Bound of linear map 65
Almost all 122 Boundary
Almost compact support 560 of manifold 541
Almost everywhere 122 point 22
Alternating 507 Bounded 18
product 507, 509 functional 253
Anosov theorem 381 linear map 65, 252
Antifunctional 391 measure 199
Antilinear 95, 391 variation 279, 284
Approximate 129 Bourbaki's theorem 13
Approximation Bruhat-Tits 50
Dirac 228
LI 147
Stone-Weierstrass 52, 61, 62
Arcwise connected 29
c
Ascoli's theorem 57 Cc-functional 253
Atlas 523 CP-invertible 361
Automorphism 67 CP-isomorphism 361
Averaging theorem 145, 221 C·-algebra 410
Calculus of variation 358
B Cantor set 49
Carathi:odory
Baire's theorem 387 criterion 179
Banach algebra 73 measure 179
isomorphism 67 Carried (measure) 179, 192
576 INDEX
Fourier I
coefficients 98, 176
inversion 241, 289 Ideal 55
series 230 Ideal topology 21
transform 238, 288 Image 3
Fredholm operator 417 Implicit mapping theorem 364, 532
Frontier 562 Index 420
Fubini's theorem 162 Induced topology 23
Function 4 Inductively ordered 12
Functionals 68, 104 Initial condition 366, 371
Fundamental lemma of integration Injective 3
111,129 Integrable 132
Integral
curve 544
G equation 432
Gelfand-Mazur theorem 402 general theory 129
Gelfand-Naimark theorem 411 in one variable 331
Gelfand transform 407, 409 Lebesgue 166
Generated (-algebra) 114 mean value theorem 286
Global flow 377, 545 of step maps 126
Gradient 383 Integral curve 365, 544
Integral operators 213,432, 478
Integration by parts 282
H Interior 22, 497
HP spaces 80 Invariant subspace 81,442, 450
Hs spaces 219 Inverse image 6
Haar functional and measure 313, 324 of differential form 513
Hahn-Banach theorem 70 Inverse mapping theorem 361
Hahn decomposition 203 Invertible 66, 74
Hahn's theorem 153 Irreducible representation 459
Half space 85, 539 Isometry 67
Harmonic functions 231 Isomorphism 66, 361, 527
Hausdorff
measure 180 J
space 32
Heat Jacobian 503
equation 232, 234, 248
operator 232, 235, 248, 449, 485 K
Hermite polynomials 276
Hermitian Karamata theorem 277
form 95 Kernel 87
operator 107, 438 Kolmogoroff inequality 215
Hilbert Krein-Milman theorem 88
basis 98
Nullstellensatz 57 L
space 99
Hilbertian operator 439 Ll 129
Hilbert-Schmidt operators 435, 461 Ll seminorm 19, 128
Holder condition 44 L2 181
Holder inequality 210 L2 bound 108, 218, 220, 437
Homeomorphism 25 L2 norm 97, 182
Homogeneous space 323 U 209
Hyperbolic 381 Laguerre polynomials 276
Hyperplane 539 Landau approximation 229
578 INDEX