Sunteți pe pagina 1din 286

Fundamenfrals

of
Fu
Ar
ncionel
nalysis
Kutateladze
Functional Analysis
br
S. S. Kut at el adze
Jol , nl , v l n5111u1c o. f . Vt t t hr nt t t t i , t .
Stberi on IJranch o;[ the Ru,rsi on Acadent.y o! Sci ence.t
,\' ot o s i b r r.s k, R us s i a
Fundamentals of
KLUWER ACADEMI C PURLI SHE, RS
DORI ) RLCI l f i ROS' I ON / L( ) Nl ) ON
A C.I.P. Catalogue record for this book is available from the Library of Congress
ISBN 0-7923-3898-7
Published by Kluwer Academic Publishers,
P.O. Box 17, 3300 AA Dordrecht, The Netherlands.
Kluwer Academic Publishers incorporates
the publishing programmes of
D. Reidel, Martinus Nijhoff, Dr W. Junk and MTP Press.
Sold and disributed in the U.S.A. and Canada
by Kluwer Academic Publishers,
101 Philip Drive, Norwell, MA 02061, U.S.A.
In all other countries, sold and distributed
by Kluwer Academic Publishers Group,
P.O. Box 322,nn AH Dordrecht, The Netherlands.
Translated from0cnoan
@yun4raoHalrnolo aaaJlr4sa. lial; Z, Ioltornenuoe.
Sobolev Institute of Mathematics. Novosibink.
@ 1995 S. S. Kutateladze
Printed on acid-free paper
All Rights Reserved
@ 1996 Kluwer Academic Publishers
No pert of the material protected by this copyright notice may be reproduced or
utilized in any form or by any means, electronic or mechanical,
including photocopying, recording or by any information storage and
retrieval system, without written permission from the copyright owner.
Printed in the Netherlands
Pr,
Pr,
Prr
Ch
Chr
Chr
Che
Contents
Preface to the English Translation v
Preface to the First Russian Edition vi
Chapter 1. An Excursus into Set Theory 1
1.1. Correspondences . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2. Ordered Sets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.3. Filters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
Chapter 2. Vector Spaces 10
2.1. Spaces and Subspaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.2. Linear Operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.3. Equations in Operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
vi Contents
Chapter 3. Convex Analysis 21
3.1. Sets in Vector Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
3.2. Ordered Vector Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.3. Extension of Positive Functionals and Operators . . . . . . . . . . . . . 26
3.4. Convex Functions and Sublinear Functionals . . . . . . . . . . . . . . . . . 27
3.5. The HahnBanach Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
3.6. The KrenMilman Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
3.7. The Balanced HahnBanach Theorem . . . . . . . . . . . . . . . . . . . . . . . . 35
3.8. The Minkowski Functional and Separation . . . . . . . . . . . . . . . . . . . 37
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
Chapter 4. An Excursus into Metric Spaces 42
4.1. The Uniformity and Topology of a Metric Space . . . . . . . . . . . . . . 42
4.2. Continuity and Uniform Continuity . . . . . . . . . . . . . . . . . . . . . . . . . . 45
4.3. Semicontinuity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
4.4. Compactness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
4.5. Completeness Completeness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
4.6. Compactness and Completeness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
4.7. Baire Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
4.8. The Jordan Curve Theorem and Rough Drafts . . . . . . . . . . . . . . . 57
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
Chapter 5. Multinormed and Banach Spaces 60
5.1. Seminorms and Multinorms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
5.2. The Uniformity and Topology of a Multinormed Space . . . . . . . 64
5.3. Comparison Between Topologies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
5.4. Metrizable and Normable Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
5.5. Banach Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
5.6. The Algebra of Bounded Operators . . . . . . . . . . . . . . . . . . . . . . . . . . 77
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
Contents vii
Chapter 6. Hilbert Spaces 85
6.1. Hermitian Forms and Inner Products . . . . . . . . . . . . . . . . . . . . . . . . 85
6.2. Orthoprojections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
6.3. A Hilbert Basis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
6.4. The Adjoint of an Operator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
6.5. Hermitian Operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
6.6. Compact Hermitian Operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
Chapter 7. Principles of Banach Spaces 105
7.1. Banachs Fundamental Principle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
7.2. Boundedness Principles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
7.3. The Ideal Correspondence Principle . . . . . . . . . . . . . . . . . . . . . . . . . . 110
7.4. Open Mapping and Closed Graph Theorems . . . . . . . . . . . . . . . . . 113
7.5. The Automatic Continuity Principle . . . . . . . . . . . . . . . . . . . . . . . . . 117
7.6. Prime Principles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
Chapter 8. Operators in Banach Spaces 126
8.1. Holomorphic Functions and Contour Integrals . . . . . . . . . . . . . . . . 126
8.2. The Holomorphic Functional Calculus . . . . . . . . . . . . . . . . . . . . . . . . 132
8.3. The Approximation Property . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
8.4. The RieszSchauder Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140
8.5. Fredholm Operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151
Chapter 9. An Excursus into General Topology 153
9.1. Pretopologies and Topologies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153
9.2. Continuity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
9.3. Types of Topological Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158
9.4. Compactness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162
9.5. Uniform and Multimetric Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166
9.6. Covers, and Partitions of Unity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175
viii Contents
Chapter 10. Duality and Its Applications 177
10.1. Vector Topologies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177
10.2. Locally Convex Topologies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 179
10.3. Duality Between Vector Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181
10.4. Topologies Compatible with Duality . . . . . . . . . . . . . . . . . . . . . . . . 184
10.5. Polars . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185
10.6. Weakly Compact Convex Sets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187
10.8. Reexive Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189
10.8. The Space C(Q, R) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189
10.9. Radon Measures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 195
10.10. The Spaces D() and D

() . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 202
10.11. The Fourier Transform of a Distribution . . . . . . . . . . . . . . . . . . . 209
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220
Chapter 11. Banach Algebras 221
11.1. The Canonical Operator Representation . . . . . . . . . . . . . . . . . . . . 221
11.2. The Spectrum of an Element of an Algebra . . . . . . . . . . . . . . . . . 223
11.3. The Holomorphic Functional Calculus in Algebras . . . . . . . . . . . 224
11.4. Ideals of Commutative Algebras . . . . . . . . . . . . . . . . . . . . . . . . . . . . 226
11.5. Ideals of the Algebra C(Q, C) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 227
11.6. The Gelfand Transform . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 228
11.7. The Spectrum of an Element of a C

-Algebra . . . . . . . . . . . . . . . 232
11.8. The Commutative GelfandNamark Theorem . . . . . . . . . . . . . . . 234
11.9. Operator -Representations of a C

-Algebra . . . . . . . . . . . . . . . . 237
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243
References 237
Notation Index 255
Subject Index 259
Preface
to the English Translation
This is a concise guide to basic sections of modern functional analysis. Included
are such topics as the principles of Banach and Hilbert spaces, the theory of multi-
normed and uniform spaces, the RieszDunford holomorphic functional calculus,
the Fredholm index theory, convex analysis and duality theory for locally convex
spaces.
With standard provisos the presentation is self-sucient, whereas exposing
about a hundred of celebrated named theorems furnished with complete proofs
and culminating in the GelfandNamarkSegal construction for C

-algebras.
The rst Russian edition was printed by the Siberian Division of the Nauka
Publishers in 1983. Since then the monograph has served as the standard textbook
on functional analysis at the University of Novosibirsk.
This volume is translated from the second Russian edition printed by the
Sobolev Institute of Mathematics of the Siberian Division of the Russian Acad-
emy of Sciences in 1995. It incorporates new sections on Radon measures, the
Schwartz spaces of distributions, and a supplementary list of theoretical exercises
and problems for drill.
This edition was typeset using /
/
5-T
E
X, the American Mathematical Soci-
etys T
E
X system.
To clear my conscience completely, I also confess that:= stands for the denor,
the assignment operator, marks the beginning of a (possibly empty) proof, and
signies the end of the proof.
S. Kutateladze
Preface
to the First Russian Edition
As the title implies, this book treats of functional analysis. At the turn of the centu-
ry the term functional analysis was coined by J. Hadamard who is famous among
mathematicians for the formula of the radius of convergence of a power series. The
term functional analysis was universally accepted then as related to the calculus
of variations, standing for a new direction of analysis which was intensively devel-
oped by V. Volterra, C. Arzel`a, S. Pincherle, P. Levy, and other representatives
of the French and Italian mathematical schools. J. Hadamards contribution to
the recent discipline should not be reduced to the invention of the word function-
al (or more precisely to the transformation of the adjective into a proper noun).
J. Hadamard was fully aware of the relevance of the rising subject. Working hard,
he constantly advertised problems, ideas, and methods just evolved. In particular,
to one of his students, M. Frechet, he suggested the problem of inventing something
that is now generally acclaimed as the theory of metric spaces. In this connection
it is worthy to indicate that neighborhoods pertinent to functional analysis in the
sense of Hadamard and Volterra served as precursors to Hausdors famous research
heralded the birth of general topology.
For the sequel it is essential to emphasize that one of the most attractive,
dicult, and important sections of classical analysis, the calculus of variations,
became the rst source of functional analysis.
The second source of functional analysis was provided by the study directed to
creating some algebraic theory for functional equations or, stated strictly, to sim-
plifying and formalizing the manipulations of equations in functions and, in par-
ticular, linear integral equations. Ascending to H. Abel and J. Liouville, the theory
of the latter was considerably expanded by works of I. Fredholm, K. Neumann,
F. Noether, A. Poincare, et al. The eorts of these mathematicians fertilized soil
for D. Hilberts celebrated research into quadratic forms in innitely many variables.
His ideas, developed further by F. Riesz, E. Schmidt, et al., were the immediate
predecessors of the axiomatic presentation of Hilbert space theory which was un-
Preface xi
dertaken and implemented by J. von Neumann and M. Stone. The resulting section
of mathematics has vigorously inuenced theoretical physics, rst of all, quantum
mechanics. In this regard it is instructive as well as entertaining to mention that
both terms, quantum and functional, originated in the same year, 1900.
The third major source of functional analysis encompassed Minkowskis geo-
metric ideas. His invention, the apparatus for the nite-dimensional geometry
of convex bodies, prepared the bulk of spatial notions ensuring the modern devel-
opment of analysis. Elaborated by E. Helly, H. Hahn, C. Caratheodory, I. Radon,
et al., the idea of convexity has eventually shaped the fundamentals of the theory
of locally convex spaces. In turn, the latter has facilitated the spread of distribu-
tions and weak derivatives which were recognized by S. L. Sobolev as drastically
changing all tools of mathematical physics. In the postwar years the geometric no-
tion of convexity has conquered a new sphere of application for mathematics, viz.,
social sciences and especially economics. An exceptional role in this process was
performed by linear programming discovered by L. V. Kantorovich.
The above synopsis of the strands of functional analysis is schematic, incom-
plete, and approximate (for instance, it casts aside the line of D. Bernoullis super-
position principle, the line of set functions and integration theory, the line of op-
erational calculus, the line of nite dierences and fractional derivation, the line
of general analysis, and many others). These demerits notwithstanding, the three
sources listed above reect the main, and most principal, regularity: function-
al analysis has synthesized and promoted ideas, concepts, and methods from the
classical sections of mathematics: algebra, geometry, and analysis. Therefore, al-
though functional analysis verbatim means analysis of functions and functionals,
even a supercial glance at its history gives grounds to claim that functional analysis
is algebra, geometry, and analysis of functions and functionals.
A more viable and penetrating explanation for the notion of functional analysis
is given by the Soviet Encyclopedic Dictionary: Functional analysis is one of the
principal branches of modern mathematics. It resulted from mutual interaction,
unication, and generalization of the ideas and methods stemming from all parts
of classical mathematical analysis. It is characterized by the use of concepts per-
taining to various abstract spaces such as vector spaces, Hilbert spaces, etc. It nds
diverse applications in modern physics, especially in quantum mechanics.
The S. Banach treatise Theorie des Operationes Lineares, printed half a cen-
tury ago, inaugurated functional analysis as an essential activity in mathematics.
Its inuence on the development of mathematics is seminal: Ubiquitous and om-
nipresent, Banachs ideas, propounded in the book, captivate the realm of modern
mathematics.
Outstanding contribution toward progress in functional analysis belongs to
the renowned Soviet scientists: I. M. Gelfand, L. V. Kantorovich, M. V. Keldysh,
xii Preface
A. N. Kolmogorov, M. G. Kren, L. A. Lyusternik, and S. L. Sobolev. The char-
acteristic feature of the Soviet school is that its research on functional analysis is
always conducted in connection with profound applied problems. The research has
expanded the scope of functional analysis which becomes the prevailing language
of the applications of mathematics.
The next fact is demonstrative: In 1948 considered as paradoxical was even the
title of Kantorovichs insightful article Functional Analysis and Applied Mathemat-
ics which provided a basis for numerical mathematics of today. Whereas in 1974
S. L. Sobolev stated that to conceive the theory of calculations without Banach
spaces is as impossible as without electronic computers.
The exponential accumulation of knowledge within functional analysis is now
observed alongside a sharp rise in demand for the tools and concepts of the disci-
pline. A thus resulting conspicuous gap widens permanently between the state-of
the art of analysis as such and as reected in the literature accessible to the reading
community. To alter this ominous trend is the purpose of the present book.
Preface
to the Second Russian Edition
More than a decade the monograph serves as a reference book for the compulsory
and optional courses in functional analysis which are delivered at Novosibirsk State
University. The span of time proves that the principles of compiling the book are
legitimate. The present edition is enlarged with sections addressing fundamentals
of distribution theory. Theoretical exercises are supplemented and the list of refer-
ences is updated. Also, the inaccuracies are improved that were mostly indicated
by my colleagues.
I seize the opportunity to express my gratitude to all those who helped me
in preparation of the book. My pleasant debt is to acknowledge the nancial support
of the Sobolev Institute of Mathematics of the Siberian Division of the Russian
Academy of Sciences, the Russian Foundation for Basic Research, the International
Science Foundation and the American Mathematical Society during the compilation
of the second edition.
S. Kutateladze
March, 1995
Chapter 1
An Excursus into Set Theory
1.1. Correspondences
1.1.1. Definition. Let A and B be sets and let F be a subset of the product
A B := (a, b) : a A, b B. Then F is a correspondence with the set
of departure A and the set of arrival B or just a correspondence from A (in)to B.
1.1.2. Definition. For a correspondence F AB the set
domF := D(F):= a A : ( b B) (a, b) F
is the domain (of denition) of F and the set
imF := R(F):= b B : ( a A) (a, b) F
is the codomain of F, or the range of F, or the image of F.
1.1.3. Examples.
(1) If F is a correspondence from A into B then
F
1
:= (b, a) B A : (a, b) F
is a correspondence from B into A which is called inverse to F or the inverse of F.
It is obvious that F is the inverse of F
1
.
(2) A relation F on A is by denition a subset of A
2
, i.e. a correspon-
dence from A to A (in words: F acts in A).
(3) Let F AB. Then F is a single-valued correspondence if for all
a A the containments (a, b
1
) F and (a, b
2
) F imply b
1
= b
2
. In particular,
if U A and I
U
:= (a, a) A
2
: a U, then I
U
is a single-valued correspon-
dence acting in A and called the identity relation (over U on A). A single-valued
2 Chapter 1
correspondence F A B with domF = A is a mapping of A into B or a map-
ping from A (in)to B. The terms function and map are also in current usage.
A mapping F AB is denoted by F : A B. Observe that here domF always
coincides with A whereas imF may dier from B. The identity relation I
U
on A is
a mapping if and only if A = U and in this case I
U
is called the identity mapping
in U or the diagonal of U
2
. The set I
U
may be treated as a subset of U A.
The resulting mapping is usually denoted by : U A and is called the identical
embedding of U into A. It is said that F is a correspondence from A onto B
if imF = B. Finally, a correspondence F X Y is one-to-one whenever the
correspondence F
1
B A is single-valued.
(4) Occasionally the term family is used instead of mapping. Na-
mely, a mapping F : A B is a family in B (indexed in A by F), also denoted
by (b
a
)
aA
or a b
a
(a A) or even (b
a
). Specically, (a, b) F if and only
if b = b
a
. In the sequel, a subset U of A is often treated as indexed in itself
by the identical embedding of U into A. It is worth recalling that in set theory a
is an element or a member of A whenever a A. In this connection a family in B
is also called a family of elements of B or a family of members of B. By way of
expressiveness a family or a set of numbers is often addressed as numeric. Also,
common abusage practices the identication of a family and its range. This sin is
very enticing.
(5) Let F A B be a correspondence and U A. The restriction
of F to U, denoted by F[
U
, is the set F (U B) U B. The set F(U):= imF[
U
is the image of U under F.
If a and b are elements of A and B then F(a) = b is usually written instead
of F(a) = b. Often the parentheses in the symbol F(a) are omitted or replaced
with other symbols. For a subset U of B the image F
1
(U) of U under F
1
is
the inverse image of U or the preimage of U under F. So, inverse images are just
images of inverses.
(6) Given a correspondence F AB, assume that A is the product
of A
1
and A
2
, i.e. A = A
1
A
2
. Fixing a
1
in A
1
and a
2
in A
2
, consider the sets
F(a
1
, ):= (a
2
, b) A
2
B: ((a
1
, a
2
), b) F;
F( , a
2
):= (a
1
, b) A
1
B: ((a
1
, a
2
), b) F.
These are the partial correspondences of F. In this regard F itself is often sym-
bolized as F( , ) and referred to as a correspondence in two arguments. This
benecial agreement is eective in similar events.
1.1.4. Definition. The composite correspondence or composition of corre-
spondences F AB and G C D is the set
G F := (a, d) AD : ( b) (a, b) F & (b, d) G.
An Excursus into Set Theory 3
The correspondence G F is considered as acting from A into D.
1.1.5. Remark. The scope of the concept of composition does not diminish
if it is assumed in 1.1.4 from the very beginning that B = C.
1.1.6. Let F be a correspondence. Then F F
1
I
imF
. Moreover, the
equality F F
1
= I
imF
holds if and only if F[
domF
is a mapping.
1.1.7. Let F A B and G B C. Further assume U A. Then the
correspondence G F AC satises the equality G F(U) = G(F(U)).
1.1.8. Let F AB, G BC, and H CD. Then the correspondences
H (G F) AD and (H G) F AD coincide.
1.1.9. Remark. By virtue of 1.1.8, the symbol H G F and the like are
dened soundly.
1.1.10. Let F, G, and H be correspondences. Then
H G F =
_
(b,c)G
F
1
(b) H(c).
(a, d) H G F ( (b, c) G) (c, d) H & (a, b) F
( (b, c) G) a F
1
(b) & d H(c)
1.1.11. Remark. The claim of 1.1.10, together with the calculation intended
as its proof, is blatantly illegitimate from a formalistic point of view as based on
ambiguous or imprecise information (in particular, on Denition 1.1.1!). Experience
justies treating such criticism as petty. In the sequel, analogous convenient (and,
in fact, inevitable) violations of formal purity are mercilessly exercised with no
circumlocution.
1.1.12. Let G and F be correspondences. Then
G F =
_
bimF
F
1
(b) G(b).
Insert H:= G, G:= I
imF
and F := F into 1.1.10.
1.2. Ordered Sets
1.2.1. Definition. Let be a relation on a set X, i.e. X
2
. Reexivity
for means the inclusion I
X
; transitivity, the inclusion ; antisymmetry,
the inclusion
1
I
X
; and, nally, symmetry, the equality =
1
.
4 Chapter 1
1.2.2. Definition. A preorder is a reexive and transitive relation. A sym-
metric preorder is an equivalence. An order (partial order, ordering, etc.) is an an-
tisymmetric preorder. For a set X, the pair (X, ), with an order on X, is
an ordered set or rarely a poset. The notation x

y is used instead of y (x).


The terminology and notation are often simplied and even abused in common
parlance: The underlying set X itself is called an ordered set, a partially ordered
set, or even a poset. It is said that x is less than y, or y is greater than x, or
x y, or y x, or x is in relation to y, or x and y belong to , etc.
Analogous agreements apply customarily to a preordered set, i.e. to a set furnished
with a preorder. The convention is very propitious and extends often to an arbitrary
relation. However, an equivalence is usually denoted by the signs like .
1.2.3. Examples.
(1) The identity relation; each subset X
0
of a set X bearing a relation
is endowed with the (induced) relation
0
:= X
0
X
0
.
(2) If is an order (preorder) in X, then
1
is also an order (preorder)
which is called reverse to .
(3) Let f : X Y be a mapping and let be a relation on Y . Consider
the relation f
1
f appearing on X. By 1.1.10,
f
1
f =
_
(y
1
,y
2
)
f
1
(y
1
) f
1
(y
2
).
Hence it follows that (x
1
, x
2
) f
1
f (f(x
1
), f(x
2
)) . Thus, if
is a preorder then f
1
f itself is a preorder called the preimage or inverse
image of under f. It is clear that the inverse image of an equivalence is also
an equivalence. Whereas the preimage of an order is not always antisymmetric.
In particular, this relates to the equivalence f
1
f (= f
1
I
Y
f).
(4) Let X be an arbitrary set and let be an equivalence on X. Dene
a mapping : X P(X) by (x) := (x). Recall that P(X) stands for the
powerset of X comprising all subsets of X and also denoted by 2
X
. Let X :=
X/:= im be the quotient set or factor set of X by or modulo . A member
of X/ is usually referred to as a coset or equivalence class. The mapping is
the coset mapping (canonical projection, quotient mapping, etc.). Note that is
treated as acting onto X. Observe that
=
1
=
_
xX

1
(x)
1
(x).
Now let f : X Y be a mapping. Then f admits factorization through X; i.e.,
there is a mapping f : X Y (called a quotient of f by ) such that f = f
if and only if f
1
f.
An Excursus into Set Theory 5
(5) Let (X, ) and (Y, ) be two preordered sets. A mapping f : X
Y is increasing or isotone (i.e., x

y f(x)

f(y)) whenever f
1
f.
That f decreases or is antitone means f
1

1
f.
1.2.4. Definition. Let (X, ) be an ordered set and let U X. An element x
of X is an upper bound of U (write x U) if U
1
(x). In particular, x .
An element x of X is a lower bound of U (write x U) if x is an upper bound of U
in the reverse order
1
. In particular, x .
1.2.5. Remark. Throughout this book liberties are taken with introducing
concepts which arise from those stated by reversal, i.e. by transition from a (pre)ord-
er to the reverse (pre)order. Note also that the denitions of upper and lower bounds
make sense in a preordered set.
1.2.6. Definition. An element x of U is greatest or last if x U and x U.
When existent, such element is unique and is thus often referred to as the greatest
element of U. A least or rst element is dened by reversal.
1.2.7. Let U be a subset of an ordered set (X, ) and let

(U) be the
collection of all upper bounds of U. Suppose that a member x of X is the greatest
element of U. Then, rst, x is the least element of

(U); second, (x) U =


x.
1.2.8. Remark. The claim of 1.2.7 gives rise to two generalizations of the
concept of greatest element.
1.2.9. Definition. Let X be a (preordered) set and let U X. A supremum
of U in X is a least upper bound of U, i.e. a least element of the set of all upper
bounds of U. This element is denoted by sup
X
U or in short sup U. Certainly,
in a poset an existing supremum of U is unique and so it is in fact the supremum
of U. An inmum, a greatest lower bound inf U or inf
X
U is dened by reversal.
1.2.10. Definition. Let U be a subset of an ordered set (X, ). A member x
of X is a maximal element of U if (x) U = x. A minimal element is again
dened by reversal.
1.2.11. Remark. It is important to make clear the common properties and
distinctions of the concepts of greatest element, maximal element, and supremum.
In particular, it is worth demonstrating that a typical set has no greatest element
while possibly possessing a maximal element.
1.2.12. Definition. A lattice is an ordered set with the following property:
each pair (x
1
, x
2
) of elements of X has a least upper bound, x
1
x
2
:= supx
1
, x
2
,
the join of x
1
and x
2
, and a greatest lower bound, x
1
x
2
:= infx
1
, x
2
, the meet
of x
1
and x
2
.
6 Chapter 1
1.2.13. Definition. A lattice X is complete if each subset of X has a supre-
mum and an inmum in X.
1.2.14. An ordered set X is a complete lattice if and only if each subset of X
has a least upper bound.
1.2.15. Definition. An ordered set (X, ) is ltered upward provided that
X
2
=
1
. A downward-ltered set is dened by reversal. A nonempty poset is
a directed set or simply a direction if it is ltered upward.
1.2.16. Definition. Let X be a set. A net or a (generalized) sequence in X
is a mapping of a direction into X. A mapping of the set N of natural numbers,
N = 1, 2, 3 . . . , furnished with the conventional order, is a (countable) sequence.
1.2.17. A lattice X is complete if and only if each upward-ltered subset of X
has a least upper bound.
1.2.18. Remark. The claim of 1.2.17 implies that for calculating a supremum
of each subset it suces to nd suprema of pairs and increasing nets.
1.2.19. Definition. Let (X, ) be an ordered set. It is said that X is ordered
linearly whenever X
2
=
1
. A nonempty linearly-ordered subset of X is
a chain in X. A nonempty ordered set is called inductive whenever its every chain
is bounded above (i.e., has an upper bound).
1.2.20. KuratowskiZorn Lemma. Each inductive set contains a maximal
element.
1.2.21. Remark. The KuratowskiZorn Lemma is equivalent to the axiom
of choice which is accepted in set theory.
1.3. Filters
1.3.1. Definition. Let X be a set and let B, a nonempty subset of P(X),
consist of nonempty elements. Such B is said to be a lterbase (on X) if B is ltered
downward. Recall that P(X) is ordered by inclusion. It means that a greater subset
includes a smaller subset by denition; this order is always presumed in P(X).
1.3.2. A subset B of P(X) is a lterbase if and only if
1 B ,= and , B;
(2) B
1
, B
2
B ( B B) B B
1
B
2
.
1.3.3. Definition. A subset F of P(X) is a lter (on X) if there is a lter-
base B such that F is the set of all supersets of B; i.e.,
F = l B:= C P(X) : ( B B) B C.
In this case B is said to be a base for the lter F (so each lterbase is a lter base).
An Excursus into Set Theory 7
1.3.4. A subset F in P(X) is a lter if and only if
(1) F ,= and , F;
(2) (A F & A B X) B F;
(3) A
1
, A
2
F A
1
A
2
F.
1.3.5. Examples.
(1) Let F X Y be a correspondence and let B be a downward-
ltered subset of P(X). Put F(B) := F(B) : B B. It is easy to see that
F(B) is ltered downward. The notation is alleviated by putting F(B):= l F(B).
If F is a lter on X and B domF ,= for all B F, then F(F) is a lter
(on Y ) called the image of F under F. In particular, if F is a mapping then the
image of a lter on X is a lter on Y .
(2) Let (X, ) be a direction. Clearly, B := (x) : x X is
a lterbase. For a net F : X Y , the lter l F(B) is the tail lter of F. Let
(X, ) and F : X Y be another direction and another net in Y . If the tail lter
of F includes the tail lter of F then F is a subnet (in a broad sense) of the net
F. If there is a subnet (in a broad sense) G : X X of the identity net ((x)
xX
in the direction (X, )) such that F = F G, then F is a subnet of F (sometimes
F is addressed as a Moore subnet or a strict subnet of F). Every subnet is a subnet
in a broad sense. It is customary to speak of a net having or lacking a subnet with
some property.
1.3.6. Definition. Let F(X) be the collection of all lters on X. Take
F
1
, F
2
F(X). Say that F
1
is ner than F
2
or F
2
renes F
1
(in other words,
F
1
is coarser than F
2
or F
1
coarsens F
2
) whenever F
1
F
2
.
1.3.7. The set F(X) with the relation to be ner is a poset.
1.3.8. Let N be a direction in F(X). Then N has a supremum F
0
:=
sup N . Moreover, F
0
= F : F N .
To prove this, it is necessary to show that F
0
is a lter. Since N is not
empty it is clear that F
0
,= and / F
0
. If A F
0
and B A then, choosing
F in N for which A F, conclude that B F F
0
. Given A
1
, A
2
F
0
,
nd an element F of N satisfying A
1
, A
2
F, which is possible because N is
a direction. By 1.3.4, A
1
A
2
F F
0
.
1.3.9. Definition. An ultralter is a maximal element of the ordered set
F(X) of all lters on X.
1.3.10. Each lter is coarser than an ultralter.
By 1.3.8, the set of lters ner than a given lter is inductive. Recalling the
KuratowskiZorn Lemma completes the proof.
8 Chapter 1
1.3.11. A lter F is an ultralter if and only if for all A X either A F
or X A F.
: Suppose that A , F and B := X A , F. Note that A ,= and
B ,= . Put F
1
:= C P(X) : A C F. Then A , F , F
1
and
B F
1
F
1
,= . The checking of 1.3.4 (2) and 1.3.4 (3) is similar. Hence, F
1
is an ultralter. By denition, F
1
F. In addition, F is an ultralter and so
F
1
= F. Observe that B , F and B F, a contradiction.
: Take F
1
F(X) and let F
1
F. If A F
1
and A , F then X A F
by hypothesis. Hence, XA F
1
; i.e., = A(XA) F
1
, which is impossible.
1.3.12. If f is a mapping from X into Y and F is an ultralter on X then
f(F) is an ultralter on Y .
1.3.13. Let X := X
F
0
:= F F(X) : F F
0
for F
0
F(X). Then X
is a complete lattice.
It is obvious that F
0
is the greatest element of X and X is the least
element of X. Therefore, the empty set has a supremum and an inmum in X:
in fact, sup = inf X = X and inf = sup X = F
0
. By 1.2.17 and 1.3.8,
it suces to show that the join F
1
F
2
is available for all F
1
, F
2
X. Consider
F := A
1
A
2
: A
1
F
1
, A
2
F
2
. Clearly, F F
0
while F F
1
and
F F
2
. Thus to verify the equality F = F
1
F
2
, it is necessary to establish
that F is a lter.
Plainly, F ,= and , F. It is also immediate that (B
1
, B
2
F
B
1
B
2
F). Moreover, if C A
1
A
2
where A
1
F
1
and A
2
F
2
, then
C = A
1
A
2
C = (A
1
C) (A
2
C). Since A
1
C F
1
and A
2
C F
2
,
conclude that C F. Appealing to 1.3.4, complete the proof.
Exercises
1.1. Give examples of sets and nonsets as well as set-theoretic properties and non-set-
theoretic properties.
1.2. Is it possible for the interval [0, 1] to be a member of the interval [0, 1]? For the
interval [0, 2]?
1.3. Find compositions of the simplest correspondences and relations: squares, disks and
circles with coincident or distinct centers in R
M
R
N
for all feasible values of M and N.
1.4. Given correspondences R, S, and T, demonstrate that
(R S)
1
= R
1
S
1
; (R S)
1
= R
1
S
1
;
(R S) T = (R T) (S T); R (S T) = (R S) (R T);
(R S) T (R T) (S T); R (S T) (R S) (R T).
1.5. Assume X X X. Prove that X = .
1.6. Find conditions for the equations XA = B and AX= B to be solvable for X in cor-
respondences or in functions.
An Excursus into Set Theory 9
1.7. Find the number of equivalences on a nite set.
1.8. Is the intersection of equivalences also an equivalence? And the union of equivalences?
1.9. Find conditions for commutativity of equivalences (with respect to composition).
1.10. How many orders and preorders are there on two-element and three-element sets? List
all of them. What can you say about the number of preorders on a nite set?
1.11. Let F be an increasing idempotent mapping of a set X into itself. Assume that F
dominates the identity mapping: F I
X
. Such an F is an abstract closure operator or, briey,
an (upper) envelope. Study xed points of a closure operator. (Recall that an element x is a xed
point of F if F(x) = x.)
1.12. Let X and Y be ordered sets and M(X, Y ), the set of increasing mappings from X
to Y with the natural order (specify the latter). Prove that
(1) (M(X, Y ) is a lattice) (Y is a lattice);
(2) (M(X, Y ) is a complete lattice) (Y is a complete lattice).
1.13. Given ordered sets X, Y , and Z, demonstrate that
(1) M(X, Y Z) is isomorphic with M(X, Y ) M(Y, Z);
(2) M(X Y, Z) is isomorphic with M(X, M(Y, Z)).
1.14. How many lters are there on a nite set?
1.15. How do the least upper and greatest lower bounds of a set of lters look like?
1.16. Let f be a mapping from X onto Y . Prove that each ultralter on Y is the image
of some ultralter on X under f.
1.17. Prove that an ultralter rening the intersection of two lters is ner than either
of them.
1.18. Prove that each lter is the intersection of all ultralters ner than it.
1.19. Let A be an ultralter on N containing conite subsets (a conite subset is a subset
with nite complement). Given x, y s := R
N
, put x
A
y := (A A ) x[
A
= y[
A
. Denote
R:= R
N
/
A
. For t R the notation t symbolizes the coset with the constant sequence t dened
as t(n) := t (n N). Prove that R \ t : t R ,= . Furnish R with algebraic and order
structures. How are the properties of R and R related to each other?
Chapter 2
Vector Spaces
2.1. Spaces and Subspaces
2.1.1. Remark. In algebra, in particular, modules over rings are studied.
A module X over a ring A is dened by an abelian group (X, +) and a repre-
sentation of the ring A in the endomorphism ring of X which is considered as left
multiplication : A X X by elements of A. Moreover, a natural agreement
is presumed between addition and multiplication. With this in mind, the following
phrase is interpreted: A module X over a ring A is described by the quadruple
(X, A, +, ). Note also that A is referred to as the ground ring of X.
2.1.2. Definition. A basic eld is the eld R of real numbers or the eld C of
complex numbers. The symbol F stands for a basic eld. Observe that R is treated
as embedded into C in a standard (and well-known) fashion so that the operation
Re of taking the real part of a number sends C onto the real axis, R.
2.1.3. Definition. Let F be a eld. A module X over F is a vector space
(over F). An element of the ground eld F is a scalar in X and an element of X
is a vector in X or a point in X. So, X is a vector space with scalar eld F.
The operation + : X X X is addition in X and : F X X is scalar
multiplication in X. We refer to X as a real vector space in case F = R and as
a complex vector space, in case F = C. A more complete nomenclature consists
of (X, F, +, ), (X, R, +, ), and (X, C, +, ). Neglecting these subtleties,
allow X to stand for every vector space associated with the underlying set X.
2.1.4. Examples.
(1) A eld F is a vector space over F.
(2) Let (X, F, +, ) be a vector space. Consider (X, F, +,

), where

: (, x)

x for F and x X, the symbol

standing for the conventional


complex conjugate of . The so-dened vector space is the twin of X denoted by X

.
If F := R then the space X and the twin of X, the space X

, coincide.
Vector Spaces 11
(3) A vector space (X
0
, F, +, ) is called a subspace of a vector space
(X, F, +, ), if X
0
is a subgroup of X and scalar multiplication in X
0
is the re-
striction of that in X to F X
0
. Such a set X
0
is a linear set in X, whereas X
is referred to as an ambient space (for X
0
). It is convenient although not perfectly
puristic to treat X
0
itself as a subspace of X. Observe the important particularity
of terminology: a linear set is a subset of a vector space (whose obsolete title is
a linear space). To call a subset of whatever space X a set in X is a mathematical
idiom of long standing. The same applies to calling a member of X a point in X.
Furthermore, the neutral element of X, the zero vector of X, or simply zero of X,
is considered as a subspace of X and is denoted by 0. Since 0 is not explicitly
related to X, all vector spaces including the basic elds may seem to have a point
in common, zero.
(4) Take (X

, a family of vector spaces over F, and let X :=

be the product of the underlying sets, i.e. the collection of mappings


x :

such that x

:= x() X

as (of course, here is not


empty). Endow X with the coordinatewise or pointwise operations of addition and
scalar multiplication:
(x
1
+x
2
)():= x
1
() +x
2
() (x
1
, x
2
X, );
( x)():= x() (x X, F, )
(below, as a rule, we write x and sometimes x rather than x). The so-
constructed vector space X over F is the product of (X

. If := 1, 2, . . . , N
then X
1
X
2
. . . X
N
:= X. In the case X

= X for all , the designation


X

:= X is used. Given := 1, 2, . . . , N, put X


N
:= X.
(5) Let (X

be a family of vector spaces over F. Consider their


direct sum X
0
:=

. By denition, X
0
is the subset of X :=

which comprises all x


0
such that x
0
(
0
) 0 for a nite subset
0
(routinely
speaking,
0
is dependent on x
0
). It is easily seen that X
0
is a linear set in X.
The vector space associated with X
0
presents a subspace of the product of (X

and is the direct sum of (X

.
(6) Given a subspace (X, F, +, ) of a vector space (X
0
, F, +, ),
introduce

X
0
:= (x
1
, x
2
) X
2
: x
1
x
2
X
0
.
Then
X
0
is an equivalence on X. Denote X := X/
X
0
and let : X X be
the coset mapping. Dene operations on X by letting
x
1
+x
2
:= (
1
(x
1
) +
1
(x
2
)) (x
1
, x
2
X);
x:= (
1
(x)) (x X, F).
12 Chapter 2
Here, as usual, for subsets S
1
and S
2
of X, a subset of F, and a scalar , a member
of F, it is assumed that
S
1
+S
2
:= +S
1
S
2
;
S
1
:= ( S
1
); S
1
:= S
1
.
Thus X is furnished with the structure of a vector space. This space, denoted
by X/X
0
, is the quotient (space) of X by X
0
or the factor space of X modulo X
0
.
2.1.5. Let X be a vector space and let Lat(X) stand for the collection of all
subspaces of X. Ordered by inclusion, Lat(X) presents a complete lattice.
It is clear that inf Lat(X) = 0 and sup Lat(X) = X. Further, the intersection
of a nonempty set of subspaces is also a subspace. By 1.2.17, the proof is complete.
2.1.6. Remark. With X
1
, X
2
Lat(X), the equality X
1
X
2
= X
1
+ X
2
holds. It is evident that inf E = X
0
: X
0
E for a nonempty subset E
of Lat(X). Provided that E is ltered upward, sup E = X
0
: X
0
E .
2.1.7. Definition. Subspaces X
1
and X
2
of a vector space X split X into
(algebraic) direct sum decomposition (in symbols, X = X
1
X
2
), if X
1
X
2
= 0
and X
1
X
2
= X. In this case X
2
is an (algebraic) complement of X
1
to X, and
X
1
is an (algebraic) complement of X
2
to X. It is also said that X
1
and X
2
are
(algebraically) complementary to one another.
2.1.8. Each subspace of a vector space has an algebraic complement.
Take a subspace X
1
of X. Put
E := X
0
Lat(X) : X
0
X
1
= 0.
Obviously, 0 E . Given a chain E i0 in E , from 2.1.6 infer that X
1
sup E i0 = 0,
i.e. sup E i0 E . Thus E is inductive and, by 1.2.20, E has a maximal element,
say, X
2
. If x X (X
1
+X
2
) then
(X
2
+x : F) X
1
= 0.
Indeed, if x
2
+x = x
1
with x
1
X
1
, x
2
X
2
and F, then x X
1
+X
2
and
so = 0. Hence, x
1
= x
2
= 0 as X
1
X
2
= 0. Therefore, X
2
+x : F = X
2
because X
2
is maximal. It follows that x = 0. At the same time, it is clear that
x ,= 0. Finally, X
1
X
2
= X
1
+X
2
= X.
Vector Spaces 13
2.2. Linear Operators
2.2.1. Definition. Let X and Y be vector spaces over F. A correspondence
T X Y is linear, if T is a linear set in X Y . A linear operator on X (or
simply an operator, with linearity apparent from the context) is a mapping of X
and a linear correspondence simultaneously. If need be, we distinguish such T from
a linear single-valued correspondence S with domS ,= X and say that T is given
on X or T is dened everywhere or even T is a total operator, whereas S is referred
to as not-everywhere-dened or partially-dened operator or even a partial operator.
In the case X = Y , a linear operator from X to Y is also called an operator in X
or an endomorphism of X.
2.2.2. A correspondence T X Y is a linear operator from X to Y if and
only if domT = X and
T(
1
x
1
+
2
x
2
) =
1
Tx
1
+
2
Tx
2
(
1
,
2
F; x
1
, x
2
X).
2.2.3. The set L(X, Y ) of all linear operators carrying X into Y constitutes
a vector space, a subspace of Y
X
.
2.2.4. Definition. A member of L(X, F) is a linear functional on X, and the
space X
#
:= L(X, F) is the (algebraic) dual of X. A linear functional on X

is
a -linear or conjugate-linear functional on X. If the nature of F needs specifying,
then we speak of real linear functionals, complex duals, etc. Evidently, when F = R
the term -linear functional is used rarely, if ever.
2.2.5. Definition. A linear operator T, a member of L(X, Y ), is an (alge-
braic) isomorphism (of X and Y , or between X and Y ) if the correspondence T
1
is a linear operator, a member of L(Y, X).
2.2.6. Definition. Vector spaces X and Y are (algebraically) isomorphic,
in symbols X Y , provided that there is an isomorphism between X and Y .
2.2.7. Vector spaces X and Y are isomorphic if and only if there are operators
T L(X, Y ) and S L(Y, X) such that S T = I
X
and T S = I
Y
(in this
event S = T
1
and T = S
1
).
2.2.8. Remark. Given vector spaces X, Y , and Z, take T L(X, Y ) and
S L(Y, Z). The correspondence S T is undoubtedly a member of L(X, Z).
For simplicity every composite operator S T is denoted by juxtaposition ST.
Observe also that the taking of composition (S, T) ST is usually treated as the
mapping : L(Y, Z) L(X, Y ) L(X, Z). In particular, if E L(Y, Z)
and T L(X, Y ) then we let E T := (E T). One of the reasons behind the
convention is that juxtaposition in the endomorphism space L(X) := L(X, X)
of X which comprises all endomorphisms of X transforms L(X) into a ring (and
even into an algebra, the endomorphism algebra of X, cf. 5.6.2).
14 Chapter 2
2.2.9. Examples.
(1) If T is a linear correspondence then T
1
is also a linear correspon-
dence.
(2) If X
1
is a subspace of a vector space X and X
2
is an algebraic
complement of X
1
then X
2
is isomorphic with X/X
1
. Indeed, if : X X/X
1
is the coset mapping then its restriction to X
2
, i.e. the operator x
2
(x
2
) with
x
2
X
2
, implements a desired isomorphism.
(3) Consider X :=

, the product of vector spaces (X

.
Take a coordinate projection, i.e. a mapping Pr

: X X

dened by Pr

x:= x

.
Clearly, Pr

is a linear operator, Pr

L(X, X

). Such an operator is often


treated as an endomorphism of X, a member of L(X), on implying a natural
isomorphism between X

and X

, where X

:=

with X

:= 0 if ,= and
X

:= X

.
(4) Let X := X
1
X
2
. Since +
1
is an isomorphism between X and
X
1
X
2
, we may dene P
1
, P
2
L(X) as P
1
:= P
X
1
]]X
2
:= Pr
1
(+
1
) and
P
2
:= P
X
2
]]X
1
:= Pr
2
(+
1
). The operator P
1
is the projection of X onto X
1
along
X
2
and P
2
is the complementary projection to P
1
or the complement of P
1
(in
symbols, P
2
= P
d
1
). In turn, P
1
is complementary to P
2
, and P
2
projects X onto
X
2
along X
1
. Observe also that P
1
+P
2
= I
X
. Moreover, P
2
1
:= P
1
P
1
= P
1
, and so
a projection is an idempotent operator. Conversely, every idempotent P belonging
to L(X) projects X onto P(X) along P
1
(0).
For T L(X) and P a projection, the equality PTP = TP holds if and only
if T(X
0
) X
0
with X
0
= imP (read: X
0
is invariant under T).
The equality TP
X
1
]]X
2
= P
X
1
]]X
2
T holds whenever both X
1
and X
2
are invari-
ant under T, and in this case the direct sum decomposition X = X
1
X
2
reduces T.
The restriction of T to X
1
is acknowledged as an element T
1
of L(X
1
) which is
called the part of T in X
1
. If T
2
L(X
2
) is the part of T in X
2
, then T is
expressible in matrix form
T
_
T
1
0
0 T
2
_
.
Namely, an element x of X
1
X
2
is regarded as a column vector with compo-
nents x
1
and x
2
, where x
1
= P
X
1
]]X
2
x and x
2
= P
X
1
]]X
2
x; matrices are multiplied
according to the usual rule, rows by columns. The product of T and the column
vector x, i.e. the vector with components T
1
x
1
and T
2
x
2
, is certainly considered
as Tx (in this case, we also write Tx
1
and Tx
2
). In other words, T is identied
with the mapping from X
1
X
2
to X
1
X
2
acting as
_
x
1
x
2
_

_
T
1
0
0 T
2
__
x
1
x
2
_
.
Vector Spaces 15
In a similar way we can introduce matrix presentation for general operators con-
tained in L(X
1
X
2
, Y
1
Y
2
).
(5) A nite subset E of X is linearly independent (in X) provided that

eE

e
e = 0, with
e
F (e E ), implies
e
= 0 for all e E . An arbitrary
subset E of X is linearly independent if every nite subset of E is linearly inde-
pendent. A Hamel basis (or an algebraic basis) for X is a linearly independent set
in X maximal by inclusion. Each linearly independent set is contained in a Hamel
basis. All Hamel bases have the same cardinality called the dimension of X and
denoted by dimX. Every vector space is isomorphic to the direct sum of a family
(F)

with of cardinality dimX. Suppose that X


1
is a subspace of X. The codi-
mension of X
1
is the dimension of X/X
1
, with codimX
1
standing for the former.
If X = X
1
X
2
then codimX
1
= dimX
2
and dimX = dimX
1
+ codimX
1
.
2.3. Equations in Operators
2.3.1. Definition. Given T L(X, Y ), dene the kernel of T as ker T :=
T
1
(0), the cokernel of T as coker T := Y/ imT, and the coimage of T as coimT :=
X/ ker T. Agree that an operator T is an monomorphism whenever ker T = 0.
An operator T is an epimorphism in the case of the equality imT = Y .
2.3.2. An operator is an isomorphism if and only if it is a monomorphism and
an epimorphism simultaneously.
2.3.3. Remark. Below use is made of the concept of commutative diagram.
So the phrase, The following diagram commutes,
V W
X Y

encodes the containments


1
L(X, Y ),
2
L(Y, W),
3
L(X, W),

4
L(V, Y ) and
5
L(V, W), as well as the equalities
2

1
=
3
and

5
=
2

4
.
2.3.4. Definition. A diagram X
T
Y
S
Z is an exact sequence (at the
term Y ), if ker S = imT. A sequence . . . X
k1
X
k
X
k+1
. . . is exact
at X
k
, if for all k the subsequence X
k1
X
k
X
k+1
(symbols of operators are
omitted), and is exact if it is exact at every term (except the rst and the last,
if any).
16 Chapter 2
2.3.5. Examples.
(1) An exact sequence X
T
Y
S
Z is semi-exact, i.e. ST = 0. The
converse is not true.
(2) A sequence 0 X
T
Y is exact if and only if T is a monomor-
phism. (Throughout the book 0 X certainly denotes the sole element of L(0, X),
zero (cf. 2.1.4 (3)).)
(3) A sequence X
T
Y 0 is exact if and only if T is an epi-
morphism. (Plainly, the symbol Y 0 again stands for zero, the single element
of L(Y, 0).)
(4) An operator T from X to Y , a member of L(X, Y ), is an isomor-
phism if and only if 0 X
T
Y 0 is exact.
(5) Suppose that X
0
is a subspace of X. Let : X
0
X stand for
the identical embedding of X
0
into X. Consider the quotient space X/X
0
and let
: X X/X
0
be the corresponding coset mapping. Then the sequence
0 X
0

X

X/X
0
0
is exact. (On similar occasions the letters and are usually omitted.) In a sense,
this sequence is unique. Namely, consider a so-called short sequence
0 X
T
Y
S
Z 0
and assume that it is exact. Putting Y
0
:= imT, arrange the following commutative
diagram
0
0
Y
0
X
Y
Y
Y/Y
0
Z
0
0


where , , and are some isomorphisms. In other words, a short exact sequence
actually presents a subspace and the corresponding quotient space.
(6) Every T in L(X, Y ) generates the exact sequence
0 ker T X
T
Y coker T 0
which is called the canonical exact sequence for T.
2.3.6. Definition. Given T
0
L(X
0
, Y ), with X
0
a subspace of X, call
an operator T from X to Y an extension of T
0
(onto X, in symbols, T T
0
),
provided that T
0
= T, where : X
0
X is the identical embedding of X
0
into X.
Vector Spaces 17
2.3.7. Let X and Y be vector spaces and let X
0
be a subspace of X. Then
each T
0
in L(X
0
, Y ) has an extension T in L(X, Y ).
Putting T := T
0
P
X
0
, where P
X
0
is a projection onto X
0
, settles the claim.
2.3.8. Theorem. Let X, Y , and Z be vector spaces. Take A L(X, Y ) and
B L(X, Z). The diagram
Z
X Y
A
X
B


is commutative for some X in L(Y, Z) if and only if ker A ker B.
: It is evident that ker A ker B in case B = XA.
: Set X := BA
1
. Clearly, X A(x) = B(A
1
A)x = B(x+ker A) = Bx.
Show that X
0
:= X[
imA
is a linear operator. It suces to check that X is single-
valued. Suppose that y imA and z
1
, z
2
X(y). Then z
1
= Bx
1
, z
2
= Bx
2
and
Ax
1
= Ax
2
= y. By hypothesis B(x
1
x
2
) = 0; therefore, z
1
= z
2
. Applying 2.3.7,
nd an extension X of X
0
to Y .
2.3.9. Remark. Provided that the operator A is an epimorphism, there is
a unique solution X in 2.3.8. It is the right place to emphasize that the phrase,
There is a unique X, implies that X is available as well as unique.
2.3.10. Every linear operator T admits a unique factorization through its
coimage; i.e., there is a unique quotient T of T by the equivalence
codimT
.
Immediate from 2.3.8 and 2.3.9.
2.3.11. Remark. The operator T is sometimes called the monoquotient of T
and treated as acting onto imT. In this connection, observe that T is expressible
as the composition T = T, with an epimorphism, T an isomorphism, and
a monomorphism; i.e., the following diagram commutes:
coimT
X
imT
Y

T
T

2.3.12. Let X be a vector space and let f


0
, f
1
, . . . , f
N
belong to X
#
. The
functional f
0
is a linear combination of f
1
, . . . , f
N
(i.e., f
0
=

N
j=1

j
f
j
with
j
in
F) if and only if ker f
0

N
j=1
ker f
j
.
18 Chapter 2
Dene the linear operator (f
1
, . . . , f
N
) : X F
N
by (f
1
, . . . , f
N
)x :=
(f
1
(x), . . . , f
N
(x)). Obviously, ker(f
1
, . . . , f
N
) =
N
j=1
ker f
j
. Now apply 2.3.8 to
the problem
F
X F
N
(f
1
,... ,f
N
)
f
0


on recalling what F
N#
is.
2.3.13. Theorem. Let X, Y , and Z be vector spaces. Take A L(Y, X)
and B L(Z, X). The diagram
Z
X Y
A
X
B


is commutative for some X in L(Z, Y ) if and only if imA imB.
: imB = B(Z) = A(X(Z)) A(Y ) = imA.
: Let Y
0
be an algebraic complement of ker A to Y and A
0
:= A[
Y
0
. Then A
0
is an isomorphism between Y
0
and imA. The operator X := A
1
0
B is obviously
a sought solution, with the identical embedding of Y
0
into Y .
2.3.14. Remark. Provided that the operator A is a monomorphism, there is
a unique operator X in 2.3.13.
2.3.15. Remark. Theorems 2.3.8 and 2.3.13 are in formal duality. One
results from the other by reversing arrows, interchanging kernels and images,
and passing to reverse inclusion.
2.3.16. Snowake Lemma. Let S L(Y, Z) and T L(X, Y ). There
are unique operators
1
, . . . ,
6
such that the following diagram commutes:
Vector Spaces 19
0 0
ker ST ker S
0 ker T X Y coker T 0
Z
coker S coker ST
0 0

1

3
T
ST S

6

4

Moreover, the sequence


0 ker T

1
ker ST

2
ker S

3
coker T

4
coker ST

5
coker S 0
is exact.
Exercises
2.1. Give examples of vector spaces and nonvector spaces. Which constructions lead to
vector spaces?
2.2. Study vector spaces over the two-element eld Z
2
.
2.3. Describe vector spaces with a countable Hamel basis.
2.4. Prove that there is a discontinuous solution f : R R to the function equation
f(x +y) = f(x) +f(y) (x, y R).
How to visualize such an f graphically?
2.5. Prove that the algebraic dual to the direct sum of (X

) is presentable as the product


of the algebraic duals of (X

).
2.6. Let X X
0
X
00
. Prove that the spaces X/X
00
and (X/X
0
)/(X
00
/X
0
) are isomor-
phic.
2.7. Dene the double sharp mapping by the rule
x
##
: x
#
x[ x
#
) (x X, x
#
X
#
).
Show that this mapping embeds a vector space X into the second dual X
##
.
20 Chapter 2
2.8. Prove that nite-dimensional spaces and only such spaces are algebraically reexive;
i.e.,
##
(X) = X
##
dimX < +.
2.9. Are there any analogs for a Hamel basis in general modules?
2.10. When does a sum of projections present a projection itself?
2.11. Let T be an endomorphism of some vector space which satises the conditions T
n1
,=
0 and T
n
= 0 for a natural n. Prove that the operators T
0
, T, . . . , T
n1
are linearly independent.
2.12. Describe the structure of a linear operator dened on the direct sum of spaces and
acting into the product of spaces.
2.13. Find conditions for unique solvability of the following equations in operators XA = B
and AX= B (here the operator X is unknown).
2.14. What is the structure of the spaces of bilinear operators?
2.15. Characterize the real vector space that results from neglecting multiplication by imag-
inary scalars in a complex vector space (cf. 3.7.1).
2.16. Given a family of linearly independent vectors (x
e
)
eE
, nd a family of functionals
(x
#
e
)
eE
satisfying the following conditions:
x
e
[ x
#
e
) = 1 (e E);
x
e
[ x
#
e
) = 0 (e, e

E, e ,= e

).
2.17. Given a family of linearly independent functionals (x
#
e
)
eE
, nd a family of vectors
(x
e
)
eE
satisfying the following conditions:
x
e
[ x
#
e
) = 1 (e E);
x
e
[ x
#
e
) = 0 (e, e

E, e ,= e

).
2.18. Find compatibility conditions for simultaneous linear equations and linear inequalities
in real vector spaces.
2.19. Consider the commutative diagram
W X
T
Y Z

W X
T
Y Z
with exact rows and such that is an epimorphism, and is a monomorphism. Prove that
ker = T(ker ) and T
1
(im) = im.
Chapter 3
Convex Analysis
3.1. Sets in Vector Spaces
3.1.1. Definition. Let be a subset of F
2
. A subset U of a vector space is
a -set in this space (in symbols, U ()) if (
1
,
2
)
1
U +
2
U U.
3.1.2. Examples.
(1) Every set is in (). (Hence, () is not a set.)
(2) If := F
2
then a nonempty -set is precisely a linear set in a vector
space.
(3) If := R
2
then a nonempty -set in a vector space X is a real
subspace of X.
(4) By denition, a cone is a nonempty -set with := R
2
+
. In other
words, a nonempty set K is a cone if and only if K +K K and K K for all
R
+
. A nonempty R
2
+
0-set is sometimes referred to as a nonpointed cone; and
a nonempty R
+
0-set, as a nonconvex cone. (From now on we use the convenient
notation R
+
:= t R : t 0.)
(5) A nonempty -set, for := (
1
,
2
) F
2
:
1
+
2
= 1, is
an ane variety or a at. If X
0
is a subspace of X and x X then x+X
0
:= x+X
0
is an ane variety in X. Conversely, if L is an ane variety in X and x L then
L x:= L +x is a linear set in X.
(6) If := (
1
,
2
) F
2
: [
1
[ + [
2
[ 1 then a nonempty -set is
absolutely convex.
(7) If := (, 0) F
2
: [[ 1 then a nonempty -set is balanced.
(In the case F := R the term star-shaped is occasionally employed; the word
symmetric can also be found.)
(8) A set is convex if it is a -set for := (
1
,
2
) R
2
:
1
0,
2

0,
1
+
2
= 1.
22 Chapter 3
(9) A conical segment or conical slice is by denition a nonempty -set
with := (
1
,
2
) R
2
+
:
1
+
2
1. A set is a conical segment if and only
if it is a convex set containing zero.
(10) Given F
2
, observe that X () for every vector space X
over F. Note also that in 3.1.2 (1)3.1.2 (9) the set is itself a -set.
3.1.3. Let X be a vector space and let E be a family of -sets in X. Then
U : U imE (). Provided that imE is ltered upward (by inclusion),
U : U imE ().
3.1.4. Remark. The claim of 3.1.3 means in particular that the collection
of -sets of a vector space, ordered by inclusion, presents a complete lattice.
3.1.5. Let X and Y be vector spaces and let U and V be -sets, with U X
and V Y . Then U V ().
If U or V is nonempty then U V = , and there is nothing to prove.
Now take u
1
, u
2
U, v
1
, v
2
V , and (
1
,
2
) . Find
1
u
1
+
2
u
2
U and

1
v
2
+
2
v
1
V . Hence, (
1
u
1
+
2
u
2
,
1
v
1
+
2
v
2
) U V .
3.1.6. Definition. Let X and Y be vector spaces and F
2
. A correspon-
dence T X Y is a -correspondence provided that T ().
3.1.7. Remark. When a -set bears a specic attribute, the latter is pre-
served for naming a -correspondence. With this in mind, we speak about linear
and convex correspondences, ane mappings, etc. The next particularity of the
nomenclature is worth memorizing: a convex function of one variable is not a con-
vex correspondence, save trivial cases (cf. 3.4.2).
3.1.8. Let T X Y be a
1
-correspondence and let U X be a
2
-set.
If
2

1
then T(U) (
2
).
Take y
1
, y
2
T(U). Then (x
1
, y
1
) T and (x
2
, y
2
) T with some
x
1
, x
2
U. Given (
1
,
2
)
2
, observe that
1
(x
1
, y
1
) +
2
(x
2
, y
2
) T,
because by hypothesis (
1
,
2
)
1
. Finally, infer that
1
y
1
+
2
y
2
T(U).
3.1.9. The composition of -correspondences is also a -correspondence.
Let F X V and G W Y , with F, G (). Note that
(x
1
, y
1
) G F ( v
1
) (x
1
, v
1
) F & (v
1
, y
1
) G;
(x
2
, y
2
) G F ( v
2
) (x
2
, v
2
) F & (v
2
, y
2
) G.
To complete the proof, multiply the rst row by
1
; the second, by
2
, where
(
1
,
2
) ; and sum the results.
Convex Analysis 23
3.1.10. If subsets U and V of a vector space are -sets for some F
2
then
U +V () for all , F.
The claim is immediate from 3.1.5, 3.1.8, and 3.1.9.
3.1.11. Definition. Let X be a vector space and let U be a subset of X. For
F
2
the -hull of U is the set
H

(U):= V X : V (), V U.
3.1.12. The following statements are valid:
(1) H

(U) ();
(2) H

(U) is the least -set including U;


(3) U
1
U
2
H

(U
1
) H

(U
2
);
(4) U () U = H

(U);
(5) H

(H

(U)) = H

(U).
3.1.13. The Motzkin formula holds:
H

(U) = H

(U
0
) : U
0
is a nite subset of U.
Denote the right side of the Motzkin formula by V . Since U
0
U; applying
3.1.12 (3), deduce that H

(U
0
) H

(U); and, hence, H

(U) V . By 3.1.12 (2),


it is necessary (and, surely, sucient) to verify that V (). The last follows from
3.1.3 and the inclusion H

(U
0
) H

(U
1
) H

(U
0
U
1
).
3.1.14. Remark. The Motzkin formula reduces the problem of describing
arbitrary -hulls to calculating -hulls of nite sets. Observe also that -hulls for
concrete s have special (but natural) designations. For instance, if := (
1
,
2
)
R
2
+
:
1
+
2
= 1, then the term convex hull is used and co(U) stands for H

(U).
If U ,= , the notation for H
F
2(U) is lin(U); moreover, it is natural and convenient
to put lin():= 0. The subspace lin(U) is called the linear span of U. The concepts
of ane hull, conical hull, etc. are introduced in a similar fashion. Note also that
the convex hull of a nite set of points comprises their convex combinations; i.e.,
co(x
1
, . . . , x
N
) =
_
N

k=1

k
x
k
:
k
0,
1
+. . . +
N
= 1
_
.
3.2. Ordered Vector Spaces
3.2.1. Definition. Let (X, R, +, ) be a vector space. A preorder on X
is compatible with vector structure if is a cone in X
2
; in this case X is an ordered
vector space. (It is more precise to call (X, R, +, , ) a preordered vector space,
reserving the term ordered vector space for the case in which is an order.)
24 Chapter 3
3.2.2. If X is an ordered vector space and is the corresponding preorder
then (0) is a cone and (x) = x +(0) for all x X.
By 3.1.3, (0) is a cone. The equality (x, y) = (x, x) + (0, y x) yields
the equivalence (x, y) y x (0).
3.2.3. Let K be a cone in a vector space X. Denote
:= (x, y) X
2
: y x K.
Then is a preorder compatible with vector structure and K coincides with the
cone (0) of positive elements of X. The relation is an order if and only
if K (K) = 0.
It is clear that 0 K I
X
and K+K K . Furthermore,

1
= (x, y) X
2
: x y K. Therefore,
1
I
X
K (K) = 0.
To show that is a cone, take (x
1
, y
1
), (x
2
, y
2
) and
1
,
2
R
+
. Find

1
y
1
+
2
y
2
(
1
x
1
+
2
x
2
) =
1
(y
1
x
1
) +
2
(y
2
x
2
)
1
K +
2
K K.
3.2.4. Definition. A cone K is an ordering cone or a salient cone provided
that K (K) = 0.
3.2.5. Remark. By virtue of 3.2.2 and 3.2.3, assigning a preorder to X is
equivalent to distinguishing some cone of positive elements in it which is the positive
cone of X. The structure of an ordered vector space arises from selecting an ordering
cone. Keeping this in mind, we customarily call a pair (X, X
+
) with positive cone
X
+
, as well as X itself, a (pre)ordered vector space.
3.2.6. Examples.
(1) The space of real-valued functions R

with the positive cone R

+
:=
(R
+
)

constituted by the functions assuming only positive values has the natural
order.
(2) Let X be an ordered vector space with positive cone X
+
. If X
0
is a subspace of X then the order induced in X
0
is dened by the cone X
0
X
+
.
In this way X
0
is considered as an ordered vector space, a subspace of X.
(3) If X and Y are preordered vector spaces then T L(X, Y ) is a pos-
itive operator (in symbols, T 0) whenever T(X
+
) Y
+
. The set L
+
(X, Y ) of all
positive operators is a cone. The linear span of L
+
(X, Y ) is denoted by L
r
(X, Y ),
and a member of L
r
(X, Y ) is called a regular operator.
3.2.7. Definition. An ordered vector space is a vector lattice or a Riesz space
if the ordered set of its vectors presents a lattice.
3.2.8. Definition. A vector lattice X is called a Kantorovich space or briey
a K-space if X is boundedly order complete or Dedekind complete, which means
that each nonempty bounded above subset of X has a least upper bound.
Convex Analysis 25
3.2.9. Each nonempty bounded below subset of a Kantorovich space has a
greatest lower bound.
Let U be bounded below: x U for some x. So, x U. Applying 3.2.8,
nd sup(U). Obviously, sup(U) = inf U.
3.2.10. If U and V are nonempty bounded above subsets of a K-space then
sup(U +V ) = sup U + sup V.
If U or V is a singleton then the equality is plain. The general case follows
from the associativity of least upper bounds. Namely,
sup(U +V ) = supsup(u +V ) : u U
= supu + sup V : u U = sup V + supu : u U
= sup V + sup U.
3.2.11. Remark. The derivation of 3.2.10 is valid for an arbitrary ordered
vector space provided that the given sets have least upper bounds. The equality
sup U = sup U for R
+
is comprehended by analogy.
3.2.12. Definition. For an element x of a vector lattice, the positive part
of x is x
+
:= x 0, the negative part of x is x

:= (x)
+
, and the modulus of x is
[x[ := x (x).
3.2.13. If x and y are elements of a vector lattice then
x +y = x y +x y.
x +y x y = x +y + (x) (y) = y x
3.2.14. x = x
+
x

; [x[ = x
+
+x

.
The rst equality follows from 3.2.13 on letting y := 0. Furthermore, [x[ =
x (x) = x + (2x) 0 = x + 2x
+
= (x
+
x

) + 2x
+
= x
+
+x

.
3.2.15. Interval Addition Lemma. If x and y are positive elements of a vec-
tor lattice X then
[0, x +y] = [0, x] + [0, y].
(As usual, [u, v] := (u)
1
(v) is the (order) interval with endpoints u and v.)
The inclusion [0, x]+[0, y] [0, x+y] is obvious. Assume that 0 z x+y
and put z
1
:= z x. It is easy that z
1
[0, x]. Now if z
2
:= z z
1
then z
2
0 and
z
2
= z z x = z + (z) (x) = 0 (z x) 0 (x +y x) = 0 y = y.
3.2.16. Remark. The conclusion of the Interval Addition Lemma is often
referred to as the Riesz Decomposition Property.
3.2.17. RieszKantorovich Theorem. Let X be a vector space and let Y
be a K-space. The space of regular operators L
r
(X, Y ), ordered by the cone
of positive operators L
+
(X, Y ), is a K-space.
26 Chapter 3
3.3. Extension of Positive Functionals and Operators
3.3.1. Counterexamples.
(1) Let X be B([0, 1], R), the space of the bounded real-valued func-
tions given on [0, 1]; and let X
0
be C([0, 1], R), the subspace of X comprising
all continuous functions. Put Y := X
0
and equip X
0
, X, and Y with the natural
order (cf. 3.2.6 (1) and 3.2.6 (2)). Consider the problem of extending the identical
embedding T
0
: X
0
Y to a positive operator T in L
+
(X, Y ). If the problem had
a solution T, then each nonempty bounded set E in X
0
would have a least upper
bound sup
X
0
E calculated in X
0
. Namely, sup
X
0
E = T sup
X
E , where sup
X
E is
the least upper bound of E in X; whereas, undoubtedly, Y fails to be a K-space.
(2) Denote by s:= R
N
the sequence space and furnish s with the nat-
ural order. Let c be the subspace of s comprising all convergent sequences, the
convergent sequence space. Demonstrate that the positive functional f
0
: c R
dened by f
0
(x) := limx(n) has no positive extension to s. Indeed, assume that
f s
#
, f 0 and f f
0
. Put x
0
(n):= n and x
k
(n):= k n for k, n N. Plainly,
f
0
(x
k
) = k; moreover, f(x
0
) f(x
k
) 0, since x
0
x
k
0, a contradiction.
3.3.2. Definition. A subspace X
0
of an ordered vector space X with positive
cone X
+
is massive (in X) if X
0
+X
+
= X. The terms coinitial or minorizing
are also in common parlance.
3.3.3. A subspace X
0
is massive in X if and only if for all x X there are
elements x
0
and x
0
in X
0
such that x
0
x x
0
.
3.3.4. Kantorovich Theorem. If X is an ordered vector space, X
0
is massive
in X, and Y is a K-space; then each positive operator T
0
, a member of L
+
(X
0
, Y ),
has a positive extension T, a member of L
+
(X, Y ).
Step I. First, let X := X
0
X
1
, where X
1
is a one-dimensional subspace,
X
1
:= x : R. Since X
0
is massive and T
0
is positive, the set U := T
0
x
0
:
x
0
X
0
, x
0
x is nonempty and bounded below. Consequently, there is some y
such that y := inf U. Assign Tx:= T
0
x
0
+y : x = x
0
+x, x
0
X
0
, R. It is
clear that T is a single-valued correspondence. Further, T T
0
and domT = X.
So, only the positivity of T needs checking. If x = x
0
+ x and x 0, then
the case of equal to 0 is trivial. If > 0 then x x
0
/. This implies that
T
0
x
0
/ y, i.e., Tx Y
+
. In a similar way for < 0 observe that x x
0
/.
Thus, y T
0
x
0
/ and, nally, Tx = T
0
x
0
+y Y
+
.
Step II. Now let E be the collection of single-valued correspondences S
X Y such that S T
0
and S(X
+
) Y
+
. By 3.1.3, E is inductive in order
by inclusion and so, by the KuratowskiZorn Lemma, E has a maximal element T.
If x X domT, apply the result of Step I with X := domT X
1
, X
0
:=
domT, T
0
:= T and X
1
:= x : R to obtain an extension of T. But T is
maximal, a contradiction. Thus, T is a sought operator.
Convex Analysis 27
3.3.5. Remark. When Y := R, Theorem 3.3.4 is sometimes referred to as
the KrenRutman Theorem.
3.3.6. Definition. A positive element x is discrete whenever [0, x] = [0, 1]x.
3.3.7. If there is a discrete functional on (X, X
+
) then X = X
+
X
+
.
Let T be such functional and := X
+
X
+
. Take f X
#
. It suces to
check that ker f = . Evidently, T + f [0, T]; i.e., T + f = T for some
[0, 1]. If T[ = 0 then 2T [0, T]. Hence, T = 0 and f = 0. Now if T(x
0
) ,= 0
for some x
0
, then = 1 and f = 0.
3.3.8. Discrete KrenRutman Let X be a massive subspace of an ordered
vector space X and let T
0
be a discrete functional on X
0
. Then there is a discrete
functional T on X extending T
0
.
Adjust the proof of 3.3.4.
Step I. Observe that the exhibited functional T is discrete. For, if T

[0, T]
then there is some [0, 1] such that T

(x
0
) = T(x
0
) for all x
0
X
0
and so
(T T

)(x
0
) = (1 )T(x
0
). Estimating, nd
T

(x) infT

(x
0
) : x
0
x, x
0
X
0
= T(x);
(T T

)(x) inf(T T

)(x
0
) : x
0
x, x
0
X
0
= (1 )T(x).
Therefore, T

= T and [0, T] [0, 1]T. The reverse inclusion is always true.


Thus, T is discrete.
Step II. Let E be the same as in the proof of 3.3.4. Consider E
d
, the set
comprising all S in E such that the restriction S[
domS
is a discrete functional
on domS. Show that E
d
is inductive. To this end, take a chain E
0
in E
d
and put
S := S
0
: S
0
E
0
; obviously, S E . Verifying the discreteness of S will
complete the proof.
Suppose that 0 S

(x
0
) S(x
0
) for all x
0
(domS)
+
and S

(domS)
#
.
If S(x
0
) = 0 for all x
0
then S

= 0S, as was needed. In the case S(x


0
) ,= 0 for
some x
0
(domS)
+
choose S
0
E
0
such that S
0
(x
0
) = S(x
0
). Since S
0
is discrete,
deduce that S

(x

) = S(x

) for all x

domS
0
. Furthermore, = S

(x
0
)/S(x
0
);
i.e., does not depend on the choice of S
0
. Since E
0
is a chain, infer that S

= S.
3.4. Convex Functions and Sublinear Functionals
3.4.1. Definition. The semi-extended real line R

is the set R

with some
greatest element + adjoined formally. The following agreements are eective:
(+):= + ( R
+
) and ++x:= x + (+):= + (x R

).
28 Chapter 3
3.4.2. Definition. Let f : X R

be a mapping (also called an extended


function). The epigraph of f is the set
epi f := (x, t) X R : t f(x).
The eective domain of denition or simply the domain of f is the set
domf := x X : f(x) < +.
3.4.3. Remark. Inconsistency in overusing the symbol domf is illusory. Na-
mely, the eective domain of denition of f : X R

coincides with the domain


of denition of the single-valued correspondence f X R. We thus continue to
write f : X R, omitting the dot in the symbol R

whenever domf = X.
3.4.4. Definition. If X is a real vector space then a mapping f : X R

is
a convex function provided that the epigraph epi f is convex.
3.4.5. A function f : X R

is convex if and only if the Jensen inequality


holds:
f(
1
x
1
+
2
x
2
)
1
f(x
1
) +
2
f(x
2
)
for all
1
,
2
0,
1
+
2
= 1 and x
1
, x
2
X.
: Take
1
,
2
0,
1
+
2
= 1. If either of x
1
and x
2
fails to
belong to domf then the Jensen inequality is evident. Let x
1
, x
2
domf.
Then (x
1
, f(x
1
)) epi f and (x
2
, f(x
2
)) epi f. Appealing to 3.1.2 (8), nd

1
(x
1
, f(x
1
)) +
2
(x
2
, f(x
2
)) epi f.
: Take (x
1
, t
1
) epi f and (x
2
, t
2
) epi f, i.e. t
1
f(x
1
) and t
2
f(x
2
)
(if domf = then f(x) = + (x X) and epi f = ). Applying the Jensen
inequality, observe the containment (
1
x
1
+
2
x
2
,
1
t
1
+
2
t
2
) epi f for all

1
,
2
0,
1
+
2
= 1.
3.4.6. Definition. A mapping p : X R

is a sublinear functional provided


that epi p is a cone.
3.4.7. If domp ,= 0 then the following statements are equivalent:
(1) p is a sublinear functional;
(2) p is a convex function that is positively homogeneous:
p(x) = p(x) for all 0 and x domp;
(3) if
1
,
2
R
+
and x
1
, x
2
X, then p(
1
x
1
+
2
x
2
)

1
p(x
1
) +
2
p(x
2
);
(4) p is a positively homogeneous functional that is subadditive:
p(x
1
+x
2
) p(x
1
) +p(x
2
) (x
1
, x
2
X).
Convex Analysis 29
3.4.8. Examples.
(1) A linear functional is sublinear; an ane functional is a convex
function.
(2) Let U be a convex set in X. Dene the indicator function (U) :
X R

of U as the mapping
(U)(x):=
_
0, if x U
+, if x , U.
It is clear that (U) is a convex function. If U is a cone then (U) is a sublinear
function. If U is an ane set then (U) is an ane function.
(3) The sum of nitely many convex functions and the supremum (or
upper envelope) of a family of convex functions (calculated pointwise, i.e. in (R

)
X
)
are convex functionals. Sublinear functionals have analogous properties.
(4) The composition of a convex function with an ane operator, i.e.
an everywhere-dened single-valued ane correspondence, is a convex function.
The composition of a sublinear functional with a linear operator is sublinear.
3.4.9. Definition. If U and V are subsets of a vector space X then U absorbs
V if V nU for some n N. A set U is absorbing (in X) if it absorbs every point
of X; i.e., X =
nN
nU.
3.4.10. Let T X Y be a linear correspondence with imT = Y . If U is
absorbing (in X) then T(U) is absorbing (in Y ).
Y = T(X) = T(
nN
nU) =
nN
T(nU) =
nN
nT(U)
3.4.11. Definition. Let U be a subset of a vector space X. An element x
belongs to the core of U (or x is an algebraically interior point of U) if U x is
absorbing in X.
3.4.12. If f : X R

is a convex function and x core domf then for all


h X there is a limit
f

(x)(h):= lim
0
f(x +h) f(x)

= inf
>0
f(x +h) f(x)

.
Moreover, the mapping f

(x) : h f

(x)h is a sublinear functional f

(x) : X R,
the directional derivative of f at x.
Set () := f(x + h). By 3.4.8 (4), : R R

is a convex function and


0 core dom. The mapping (() (0))/ ( > 0) is increasing and
bounded above; i.e.,

(0)(1) is available. By denition, f

(x)(h) =

(0)(1).
30 Chapter 3
Given > 0 and h H, successively nd
f

(x)(h) = inf
f(x +h) f(x)

= inf
f(x +h) f(x)

= f

(x)(h).
Moreover, using the above result, for h
1
, h
2
X infer that
f

(x)(h
1
+h
2
) = 2 lim
0
f
_
x +
1
/
2
(h
1
+h
2
)
_
f(x)

= 2 lim
0
f
_
1
/
2
(x +h
1
) +
1
/
2
(x +h
2
)
_
f(x)

lim
0
f(x +h
1
) f(x)

+ lim
0
f(x +h
2
) f(x)

= f

(x)(h
1
) +f

(x)(h
2
).
Appealing to 3.4.7 ends the proof.
3.5. The HahnBanach Theorem
3.5.1. Definition. Let X be a real vector space and let f : X R

be
a convex function. The subdierential of f at a point x in domf is the set

x
(f):= l X
#
: ( y X) l(y) l(x) f(y) f(x).
3.5.2. Examples.
(1) Let p : X R

be a sublinear functional. Put (p):=


0
(p). Then
(p) = l X
#
: ( x X) l(x) p(x);

x
(p) = l (p) : l(x) = p(x).
(2) If l X
#
then (l) =
x
(l) = l.
(3) Let X
0
be a subspace of X. Then
((X
0
)) = l X
#
: ker l X
0
.
(4) If f : X R

is a convex function then

x
(f) = (f

(x))
whenever x core domf.
Convex Analysis 31
3.5.3. HahnBanach Theorem. Let T L(X, Y ) be a linear operator and
let f : Y R

be a convex function. If x X and Tx core domf then

x
(f T) =
Tx
(f) T.
By 3.4.10 it follows that x core domf. From 3.5.2 (4) obtain
x
(f T) =
((f T)

(x)). Moreover,
(f T)

(x)(h) = lim
0
(f T)(x +h) (f T)(x)

= lim
0
f(Tx +Th) f(Tx)

= f

(Tx)(Th)
for h X. Put p := f

(Tx). By 3.4.12, p is a sublinear functional. Whence, on


appealing again to 3.5.2 (4), infer that
(p) = (f

(Tx)) =
Tx
(f);
(p T) = ((f T)

(x)) =
x
(f T).
Thereby, the only claim left unproven is the equality
(p T) = (p) T.
If l (p) T (i.e., l = l
1
T, where l
1
(p)) then l
1
(y) p(y) for all y Y .
In particular, l(x) l
1
(Tx) p(Tx) = p T(x) for all x X and so l (p T).
This argument yields the inclusion (p) T (p T).
Now take l (p T). If Tx = 0 then l(x) p(Tx) = p(0) = 0; i.e., l(x) 0.
The same holds for x. Therefore, l(x) = 0; in other words, ker l ker T. Hence,
by 2.3.8, l = l
1
T for some l
1
Y
#
. Putting Y
0
:= T(X), observe that the
functional l
1
belongs to (p ), with the identical embedding of Y
0
into Y .
With (p ) (p) proven, observe that l
1
= l
2
for some l
2
(p), which
consequently yields l = l
1
T = l
1
T = l
2
T = l
2
T; i.e., l (p) T.
Thus, to complete the proof of the HahnBanach Theorem, showing that (p
) (p) is in order.
Take an element l
0
in (p) and consider the functional T
0
: (y
0
, t) tl
0
(y
0
)
given on the subspace Y
0
:= Y
0
R of Y := Y R. Equip Y with the cone
Y
+
:= epi p. Note that, rst, Y
0
is massive since
(y, t) = (0, t p(y)) + (y, p(y)) (y Y, t R).
Second, by 3.4.2, t p(y
0
) for (y
0
, t) Y
0
Y
+
, and so T
0
(y
0
, t) = t l
0
(y
0
) 0;
i.e., T
0
is a positive functional on Y
0
. By 3.3.4, there is a positive functional T
dened on Y which is an extension of T
0
. Put l(y) := T(y, 0) for y Y . It is
clear that l = l
0
. Furthermore, T(0, t) = T
0
(0, t) = t. Hence, 0 T(y, p(y)) =
p(y) l(y), i.e., l (p).
32 Chapter 3
3.5.4. Remark. The claim of Theorem 3.5.3 is also referred to as the formula
for a linear change-of-variable under the subdierential sign or the HahnBanach
Theorem in subdierential form. Observe that the inclusion (p ) (p) is
often referred to as the HahnBanach Theorem in analytical form or the Dominated
Extension Theorem and verbalized as follows: A linear functional given on a sub-
space of a vector space and dominated by a sublinear functional on this subspace
has an extension also dominated by the same sublinear functional.
3.5.5. Corollary. If X
0
is a subspace of a vector space X and p : X R is
a sublinear functional then the (asymmetric) HahnBanach formula holds:
(p +(X
0
)) = (p) +((X
0
)).
It is obvious that the left side includes the right side. To prove the reverse
inclusion, take l (p + (X
0
)). Then l (p ), where is the identical
embedding of X
0
into X. By 3.5.3, l (p) , i.e., l = l
1
for some
l
1
(p). Put l
2
:= l l
1
. By denition, nd l
2
= (l l
1
) = l l
1
= 0,
i.e., ker l
2
X
0
. As was mentioned in 3.5.2 (3), this means l
2
((X
0
)).
3.5.6. Corollary. If f : X R

is a linear functional and x core domf


then
x
(f) ,= .
Let p := f

(x) and let : 0 X be the identical embedding of 0 to X.


Plainly, 0 (p ), i.e., (p ) ,= . By the HahnBanach Theorem, (p) ,=
(otherwise, = (p) = (p )). To complete the proof, apply 3.5.2 (4).
3.5.7. Corollary. Let f
1
, f
2
: X R

be convex functions. Assume further


that x core domf
1
core domf
2
. Then

x
(f
1
+f
2
) =
x
(f
1
) +
x
(f
2
).
Let p
1
:= f

1
(x) and p
2
:= f

2
(x). Given x
1
, x
2
X, dene p(x
1
, x
2
) :=
p
1
(x
1
) +p
2
(x
2
) and (x
1
):= (x
1
, x
1
). Using 3.5.2 (4) and 3.5.3, infer that

x
(f
1
+f
2
) = (p
1
+p
2
) = (p )
= (p) = (p
1
) +(p
2
) =
x
(f
1
) +
x
(f
2
).
3.5.8. Remark. Corollary 3.5.6 is sometimes called the Nonempty Subdier-
ential Theorem. On the one hand, it is straightforward from the Kuratowski
Zorn Lemma. On the other hand, with Corollary 3.5.6 available, demonstrate
(p T) = (p) T as follows: Dene p
T
(y) := infp(y + Tx) l(x) : x X,
where l (p) and the notation of 3.5.3 is employed. Check that p
T
is sublinear
and every l
1
in (p
T
) satises the equality l = l
1
T. Thus, the Nonempty Subdif-
ferential Theorem and the HahnBanach Theorem in subdierential form constitute
a precious (rather than vicious) circle.
Convex Analysis 33
3.6. The KrenMilman Theorem
3.6.1. Definition. Let X be a real vector space. Putting
seg(x
1
, x
2
):=
1
x
1
+
2
x
2
:
1
,
2
> 0,
1
+
2
= 1,
dene the correspondence seg X
2
X that assigns to each pair of points the open
segment joining them. If U is a convex set in X and seg
U
is the restriction of seg
to U
2
, then a convex subset V of U is extreme in U if seg
1
U
(V ) V
2
. An extreme
subset of U is sometimes called a face in U. A member x of U is an extreme point
in U if x is an extreme subset of U. The set of extreme points of U is usually
denoted by ext U.
3.6.2. A convex subset V is extreme in U if and only if for all u
1
, u
2
U and

1
,
2
> 0,
1
+
2
= 1, the containment
1
u
1
+
2
u
2
V implies that u
1
V
and u
2
V .
3.6.3. Examples.
(1) Let p : X R

be a sublinear functional and let a point x of X


belong to domp. Then
x
(p) is an extreme subset of (p).
For, if
1
l
1
+
2
l
2

x
(p) and l
1
, l
2
(p), where
1
,
2
> 0,
1
+
2
= 1;
then 0 = p(x) (
1
l
1
(x) +
2
l
2
(x)) =
1
(p(x) l
1
(x)) +
2
(p(x) l
2
(x)) 0.
Moreover, p(x)l
1
(x) 0 and p(x)l
2
(x) 0. Hence, l
1

x
(p) and l
2

x
(p).
(2) Let W be an extreme set in V and let V be in turn an extreme set
in U. Then W is an extreme set in U.
(3) If X is an ordered vector space then a positive element x of X is
discrete if and only if the cone x : R
+
is an extreme subset of X
+
.
: If 0 y x then x =
1
/
2
(2y) +
1
/
2
(2(x y)). Therefore, by 3.6.2,
2y = x and 2(x y) = x for some , R
+
. Thus, 2x = ( + )x. The case
x = 0 is trivial. Now if x ,= 0 then

/
2
[0, 1] and, consequently, [0, x] [0, 1]x.
The reverse inclusion is evident.
: Let [0, x] = [0, 1]x, and suppose that x =
1
y
1
+
2
y
2
for 0;
1
,
2
>
0,
1
+
2
= 1 and y
1
, y
2
X
+
. If = 0 then
1
y
1
[0, x] and
2
y
2
[0, x];
hence, y
1
is a positive multiple of x; the same is valid for y
2
. In the case > 0
observe that (

1
/

)y
1
= tx for some t [0, 1]. Finally, (

2
/

)y
2
= (1 t)x.
(4) Let U be a convex set. A convex subset V of U is a cap of U,
if U V is convex. A point x in U is extreme if and only if x is a cap of U.
3.6.4. Extreme and Discrete Lemma. Let p : X R be a sublinear
functional and l (p). Assign:= X R,
+
:= epi , and T
l
: (x, t) t l(x) (x
X, t R). Then the functional l is an extreme point of (p) if and only if T
l
is
a discrete functional.
34 Chapter 3
: Consider T

#
such that T

[0, T
l
]. Put
t
1
:= T

(0, 1), l
1
(x):= T

(x, 0);
t
2
:= (T
l
T

)(0, 1), l
2
(x):= (T
l
T

)(x, 0).
It is clear that t
1
0, t
2
0, t
1
+t
2
= 1; l
1
(t
1
p), l
2
(t
2
p), and l
1
+l
2
= l.
If t
1
= 0 then l
1
= 0; i.e., T

= 0 and T

[0, 1]T
l
. Now if t
2
= 0 then t
1
= 1; i.e.,
T

= T
l
, and so T

[0, 1]T
l
. Assume that t
1
, t
2
> 0. In this case
1
/
t
1
l
1
(p)
and
1
/
t
2
l
2
(p); moreover, l = t
1
_
1
/
t
1
l
1
_
+ t
2
_
1
/
t
2
l
2
_
. Since, by hypothesis,
l ext (p), from 3.6.2 it follows that l
1
= t
1
l; i.e., T

= t
1
T
l
.
: Let l =
1
l
1
+
2
l
2
, where l
1
, l
2
(p) and
1
,
2
> 0,
1
+
2
= 1. The
functionals T

:=
1
T
l
1
and T

:=
2
T
l
2
are positive; moreover, T

[0, T
l
], since
T

+T

= T
l
. Therefore, there is some [0, 1] such that T

= T
l
. At the point
(0, 1), nd
1
= . Hence, l
1
= l. By analogy, l
2
= l.
3.6.5. KrenMilman Let T L(X, Y ) be a linear operator and let f :
Y R

be a convex function. If x X and Tx core domf then


ext
x
(f T) (ext
Tx
(f)) T.
Start arguing as in the proof of the HahnBanach Theorem: Put p:= f

(Tx)
and observe that the only claim left to checking is the inclusion ext (p T)
(ext (p)) T. Take l ext (p T). Since ext (p T) (p T) (p) T,
there is some f in (p) such that l = f T. Denote by f
0
the restriction of f to
Y
0
:= imT and notice that f
0
ext (p ), with the identical imbedding of Y
0
into Y .
Now in Y:= Y R consider Y
+
:= epi p and introduce the space Y
0
:= Y
0
R.
Note that Y
+
Y
0
= epi(p ). Applying 3.6.4 with X := Y
0
, l := f
0
, and
p:= p , observe that T
f
0
is a discrete functional on Y
0
; moreover, Y
0
is massive
in Y (cf. the proof of 3.5.3). By 3.3.8, nd a discrete extension S Y
#
of T
f
0
.
Evidently, S = T
g
, where g(y) := S(y, 0) for y Y . Appealing again to 3.6.4,
infer that g ext (p). By construction, l(x) = f(Tx) = f
0
(Tx) = g(Tx) for all
x X. Finally, l (ext (p)) T.
3.6.6. Corollary. If p : X R is a sublinear functional then for every x in X
there is an extreme functional l, a member of ext (p), such that l(x) = p(x).
From 3.6.5 it is easy that ext (p) ,= for every p (cf. 3.5.6). Using
this and 3.4.12, choose l in ext
x
(p

(x)). Applying 3.5.2 (2) and 3.5.2 (4), obtain


l ext
x
(p). By 3.6.3 (1),
x
(p) is extreme in (p). Finally, 3.6.3 (2) implies that
l is an extreme point of (p).
Convex Analysis 35
3.7. The Balanced HahnBanach Theorem
3.7.1. Definition. Let (X, F, +, ) be a vector space over a basic eld F.
The vector space (X, R, +, [
RX
) is called the real carrier or realication
of (X, F, +, ) and is denoted by X
R
.
3.7.2. Definition. Given a linear functional f on a vector space X, dene
the mapping Ref : x Re f(x) (x X). The real part map or realier is the
mapping Re: (X
#
)
R
(X
R
)
#
.
3.7.3. The real part map Re is a linear operator and, moreover, an isomor-
phism of (X
#
)
R
onto (X
R
)
#
.
Only the case F := C needs verifying, because Re is the identity mapping
when F:= R.
Undoubtedly, Re is a linear operator. Check that Re is a monomorphism and
an epimorphism simultaneously (cf. 2.3.2).
If Ref = 0 then 0 = Re f(ix) = Re(if(x)) = Re(i(Re f(x) + i Imf(x))) =
Imf(x). Hence, f = 0 and so Re is a monomorphism.
Now take g (X
R
)
#
and put f(x):= g(x)ig(ix). Evidently, f L(X
R
, C
R
)
and Re f(x) = g(x) for x X. It is sucient to show that f(ix) = if(x), because
this equality implies f X
#
. Straightforward calculation shows f(ix) = g(ix) +
ig(x) = i(g(x) ig(ix)) = if(x), which enables us to conclude that Re is also
an epimorphism.
3.7.4. Definition. The lear trap map or complexier is the inverse Re
1
:
(X
R
)
#
(X
#
)
R
of the real part map.
3.7.5. Remark. By 3.7.3, for the complex scalar eld
Re
1
g : x g(x) ig(ix) (g (X
R
)
#
, x X).
In the case of the reals, F:= R, the complexier Re
1
is the identity operator.
3.7.6. Definition. Let (X, F, +, ) be a vector space over F. A seminorm
on X is a function p : X R

such that domp ,= and


p(
1
x
1
+
2
x
2
) [
1
[p(x
1
) +[
2
[p(x
2
)
for x
1
, x
2
X and
1
,
2
F.
3.7.7. Remark. A seminorm is a sublinear functional (on the real carrier
of the space in question).
36 Chapter 3
3.7.8. Definition. If p : X R

is a seminorm then the balanced subdier-


ential of p is the set
[[(p):= l X
#
: [l(x)[ p(x) for all x X.
3.7.9. Balanced Subdierential Lemma. Let p : X R

be a seminorm.
Then
[[(p) = Re
1
((p)); Re([[(p)) = (p)
for the subdierentials [[(p) and (p) of p.
If F:= R then the equality [[(p) = (p) is easy. Furthermore, in this case
Re is the identity operator.
Let F:= C. If l [[(p) then (Rel)(x) = Re l(x) [l(x)[ p(x) for all x X;
i.e., Re([[(p)) (p). Take g (p) and put f := Re
1
g. If f(x) = 0 then
[f(x)[ p(x); for f(x) ,= 0 set := [f(x)[/f(x). Thereby [f(x)[ = f(x) = f(x) =
Re f(x) = g(x) p(x) = [[p(x) = p(x), since [[ = 1. Finally, f [[(p).
3.7.10. Let X be a vector space, let p : X R be a seminorm, and let X
0
be
a subspace of X. The asymmetric balanced HahnBanach formula holds:
[[(p +(X
0
)) = [[(p) +[[((X
0
)).
From 3.7.9, 3.5.5, and the results of Section 3.1 obtain
[[(p +(X
0
)) = Re
1
((p +(X
0
))) = Re
1
((p) +((X
0
)))
= Re
1
((p)) +Re
1
(((X
0
))) = [[(p) +[[((X
0
)).
3.7.11. If X and Y are vector spaces, T L(X, Y ) is a linear operator, and
p : Y R is a seminorm; then p T is also a seminorm and, moreover,
[[(p T) = [[(p) T.
Applying 2.3.8 and 3.7.10, successively infer that
[[(p T) = [[(p +(imT)) T = ([[(p) +[[((imT))) T
= [[(p) T +[[((imT)) T = [[(p) T.
3.7.12. Remark. If T is the identical embedding of a subspace and the ground
eld is C then 3.7.11 is referred to as the SukhomlinovBohnenblustSobczyk The-
orem.
3.7.13. Balanced HahnBanach Let X be a vector space. Assume further
that p : X R is a seminorm and X
0
is a subspace of X. If l
0
is a linear functional
given on X
0
such that [l
0
(x
0
)[ p(x
0
) for x
0
X
0
, then there is a linear functional l
on X such that l(x
0
) = l
0
(x
0
) whenever x
0
X
0
and [l(x)[ p(x) for all x X.
Convex Analysis 37
3.8. The Minkowski Functional and Separation
3.8.1. Definition. Let R stand for the extended real axis or extended reals
(i.e., R denotes R

with the least element adjoined formally). If X is an arbi-


trary set and f : X R is a mapping; then, given t R, put
f t:= x X : f(x) t;
f = t:= f
1
(t);
f < t:= f t f = t.
Every set of the form f t, f = t, and f < t is a level set or Lebesgue set
of f.
3.8.2. Function Recovery Lemma. Let T R and let t U
t
(t T) be
a family of subsets of X. There is a function f : X R such that
f < t U
t
f t (t T)
if and only if the mapping t U
t
increases.
: Suppose that T contains at least two elements s and t. (Otherwise there
is nothing left to proof.) If s < t then
U
s
f s f < t U
t
.
: Dene a mapping f : X R by setting f(x) := inft T : x U
t
. If
f < t is empty for t T then all is clear. If x f < t then f(x) < +, and
so there is some s T such that x U
s
and s < t. Thus, f < t U
s
U
t
.
Continuing, for x U
t
nd f(x) t; i.e., U
t
f t.
3.8.3. Function Comparison Lemma. Let f, g : X R be functions
dened by (U
t
)
tT
and (V
t
)
tT
as follows:
f < t U
t
f t;
g < t V
t
g t (t T).
If T is dense in R (i.e., ( r, t R, r < t) ( s T) (r < s < t)) then the inequality
f g (in R
X
; i.e., f(x) g(x) for x X) holds if and only if
(t
1
, t
2
T & t
1
< t
2
) V
t
1
U
t
2
.
: This is immediate from the inclusions
V
t
1
g t
1
f t
1
f < t
2
U
t
2
.
: Assume that g(x) ,= + (if not, f(x) g(x) for obvious reasons). Given
t R such that g(x) < t < +, choose t
1
, t
2
T so as to satisfy the conditions
g(x) < t
1
< t
2
< t. Now
x g < t
1
V
t
1
U
t
2
f t
2
f < t.
Thus, f(x) < t. Since t is arbitrary, obtain f(x) g(x).
38 Chapter 3
3.8.4. Corollary. If T is dense in R and the mapping t U
t
(t T) increases
then there is a unique function f : X R such that
f < t U
t
f t (t T).
Moreover, the level sets of f may be presented as follows:
f < t = U
s
: s < t, s T;
f t = U
r
: t < r, r T (t R).
It is immediate from 3.8.2 and 3.8.3 that f exists and is unique. If s < t
and s T then U
s
f s f < t. If now f(x) < t then, since T is dense,
there is an element s in T such that f(x) < s < t. Therefore, f f < s U
s
,
which proves the formula for f < t.
Suppose that r > t, r T. Then f t f < r U
r
. In turn, if x U
r
for r T, r > t; then f(x) r for all r > t. Hence, f(x) t.
3.8.5. Let X be a vector space and let S be a conical segment in X. Given
t R, put U
t
:= if t < 0 and U
t
:= tS if t 0. The mapping t U
t
(t R)
increases.
If 0 t
1
< t
2
and x t
1
S then x (
t
1
/
t
2
) t
2
S. Hence, x t
2
S.
3.8.6. Definition. The Minkowski functional or the gauge function or simply
the gauge of a conical segment S is a functional p
S
: X R such that
p
S
< t tS p
S
t (t R
+
)
and p < 0 = . (Such a functional exists and is unique by 3.8.2, 3.8.4, and 3.8.5.)
In other words,
p
S
(x) = inft > 0 : x tS (x X).
3.8.7. Gauge Theorem. The Minkowski functional of a conical segment is
a sublinear functional assuming positive values. Conversely, if p is a sublinear
functional assuming positive values then the sets p < 1 and p 1 are conical
segments. Moreover, p is the Minkowski functional of each conical segment S such
that p < 1 S p 1.
Consider a conical segment S and its Minkowski functional p
S
. Let x X.
The inequality p
S
(x) 0 is evident. Take > 0. Then
p
S
(x) = inft > 0 : x tS = inf
_
t > 0 : x
t
/

S
_
= inf > 0 : x S, > 0
= inf > 0 : x S = p
S
(x).
Convex Analysis 39
To start checking that p
S
is subadditive, take x
1
, x
2
X. Noting the inclusion
t
1
S +t
2
S (t
1
+t
2
)S for t
1
, t
2
> 0, in view of the identity
t
1
x
1
+t
2
x
2
= (t
1
+t
2
)
_
t
1
t
1
+t
2
x
1
+
t
2
t
1
+t
2
x
2
_
,
successively infer that
p
S
(x
1
+x
2
) = inft > 0 : x
1
+x
2
tS
inft : t = t
1
+t
2
; t
1
, t
2
> 0, x
1
t
1
S, x
2
t
2
S
= inft
1
> 0 : x
1
t
1
S + inft
2
> 0 : x
2
t
2
S = p
S
(x
1
) +p
S
(x
2
).
Let p : X R

be an arbitrary sublinear functional with positive values and


let p < 1 S p 1. Given t R
+
, put V
t
:= p < t and U
t
:= tS; given
t < 0, put V
t
:= U
t
:= . Plainly,
p
S
< t U
t
p
S
t; p < t V
t
p t
for t R. If 0 t
1
< t
2
then V
t
1
= p < t
1
= t
1
p < 1 t
1
S = U
t
1
U
t
2
.
Moreover, U
t
1
t
1
p 1 p t
1
p < t
2
V
t
2
. Therefore, by 3.8.3 and
3.8.4, p = p
S
.
3.8.8. Remark. A conical segment S in X is absorbing in X if and only
if domp
S
= X. Also, p
S
is a seminorm whenever S is absolutely convex. Conversely,
for every seminorm p the sets p < 1 and p 1 are absolutely convex.
3.8.9. Definition. A subspace H of a vector space X is a hypersubspace
in X if X/H is isomorphic to the ground eld of X. An element of X/H is called
a hyperplane in X parallel to H. By a hyperplane in X we mean an ane variety
parallel to some hypersubspace of X. An ane variety H is a hyperplane if Hh is
a hypersubspace for some (and, hence, for every) h in H. If necessary, a hyperplane
in the real carrier of X is referred to as a real hyperplane in X.
3.8.10. Hyperplanes in X are exactly level sets of nonzero elements of X
#
.
3.8.11. Separation Theorem. Let X be a vector space. Assume further
that U is a nonempty convex set in X and L is an ane variety in X. If LU =
then there is a hyperplane H in X such that H L and H core U = .
Without loss of generality, it may be supposed that core U ,= (otherwise
there is nothing left unproven) and, moreover, 0 core U. Take x L and put X
0
:=
L x. Consider the quotient space X/X
0
and the corresponding coset mapping
: X X/X
0
. Applying 3.1.8 and 3.4.10, observe that (U) is an absorbing
40 Chapter 3
conical segment. Hence, by 3.8.7 and 3.8.8, the domain of the Minkowski functional
p:= p
(U)
is the quotient X/X
0
; moreover,
(core U) core (U) p < 1 (U).
In particular, this entails the inequality p((x)) 1, since (x) , (U).
Using 3.5.6, nd a functional f in
x
(p ); now the HahnBanach Theorem
implies
f
x
(p ) =
(x)
(p) .
Put H:= f = p (x). It is clear that H is a real hyperplane and, undoubtedly,
H L. Appealing to 3.5.2 (1), conclude that Hcore U = . Now let f := Re
1
f
and H:= f = f(x). There is no denying that L H H. Thus, the hyperplane
H provides us with what was required.
3.8.12. Remark. Under the hypotheses of the Separation Theorem, it may be
assumed that core U L = . Note also that Theorem 3.8.11 is often referred to as
the HahnBanach Theorem in geometric form or as the MinkowskiAscoliMazur
Theorem.
3.8.13. Definition. Let U and V be sets in X. A real hyperplane H in X
separates U and V if these sets lie in the dierent halfspaces dened by H; i.e.,
if there is a presentation H = f = t, where f (X
R
)
#
and t R, such that
V f t and U f t:= f t.
3.8.14. Eidelheit Separation Theorem. If U and V are convex sets such
that core V ,= and U core V = , then there is a hyperplane separating U
and V and disjoint from core V .
Exercises
3.1. Show that a hyperplane is precisely an ane set maximal by inclusion and other than
the whole space.
3.2. Prove that every ane set is an intersection of hyperplanes.
3.3. Prove that the complement of a hyperplane to a real vector space consists of two convex
sets each of which coincides with its own core. The sets are named open halfspaces. The union
of an open halfspace with the corresponding hyperplane is called a closed halfspace. Find out how
a halfspace can be prescribed.
3.4. Find possible presentations of the elements of the convex hull of a nite set. What use
can be made of nite dimensions?
3.5. Given sets S
1
and S
2
, let S:=

01
S
1
(1)S
2
. Prove that S is convex whenever
so are S
1
and S
2
.
3.6. Calculate the Minkowski functional of a halfspace or a cone and of the convex hull
of the union or intersection of conical segments.
Convex Analysis 41
3.7. Let S := p + q 1, where p and q are the Minkowski functionals of the conical
segments S
p
and S
q
. Express S via S
p
and S
q
.
3.8. Describe sublinear functionals with domain R
N
.
3.9. Calculate the subdierential of the upper envelope of a nite set of linear or sublinear
functionals.
3.10. Let p and q be sublinear functionals in general position, i.e. such that dompdomq =
domq domp. Prove the symmetric HahnBanach formula (cf. 3.5.7): (p +q) = (p) +(q).
3.11. Let p, q : X R be total (= everywhere-dened) sublinear functionals on X. Then
the equality holds: (p q) = co((p) (q)).
3.12. Find the Minkowski functional of a ball in a Hilbert space whose center of symmetry
is not necessarily the zero of the space.
3.13. A symmetric square 2 2-matrix is called positive, provided that its eigenvalues are
positive. Does the resulting order in the space of such matrices agree with vector structure? Does
it dene the structure of a Kantorovich space?
3.14. Each ordered vector space admits a nonzero positive linear functional, doesnt it?
3.15. What are the means for transforming R
N
into a Kantorovich space?
3.16. Under what conditions does the claim of the HahnBanach Theorem in analytical
form hold for a partial (= not-everywhere-dened) sublinear functional?
3.17. Find the extreme points of the subdierential of the conventional norm on l

.
3.18. Find possible generalizations of the HahnBanach Theorem for a mapping acting into
a Kantorovich space.
3.19. Given a set C in a space X, dene the Hormander transform H(C) of C as
H(C) = (x, t) X R : x tC.
Study the properties of the Hormander transform on the collection of all convex sets.
Chapter 4
An Excursus into Metric Spaces
4.1. The Uniformity and Topology of a Metric Space
4.1.1. Definition. A mapping d : X
2
R
+
is a metric on X if
(1) d(x, y) = 0 x = y;
(2) d(x, y) = d(y, x) (x, y X);
(3) d(x, y) d(x, z) +d(z, y) (x, y, z X).
The real d(x, y) is usually referred to as the distance between x and y. The pair
(X, d) is a metric space. In this situation, it is convenient to take the liberty
of calling the underlying set X a metric space. An element of a metric space X is
also called a point of X.
4.1.2. A mapping d : X
2
R
+
is a metric if and only if
(1) d 0 = I
X
;
(2) d t = d t
1
(t R
+
);
(3) d t
1
d t
2
d t
1
+t
2
(t
1
, t
2
R
+
).
Items 4.1.2 (1)4.1.2 (3) reformulate 4.1.1 (1)4.1.1 (3).
4.1.3. Definition. Let (X, d) be a metric space and take R
+
0, a strictly
positive real. The relation B

:= B
d,
:= d is the closed cylinder of size .
The set

:=

B
d,
:= d < is the open cylinder of size . The image B

(x)
of a point x under the relation B

is called the closed ball with center x and


radius . By analogy, the set

(x) is the open ball with center x and radius .


4.1.4. In a nonempty metric space open cylinders as well as closed cylinders
form bases for the same lter.
An Excursus into Metric Spaces 43
4.1.5. Definition. The lter on X
2
with the lterbase of all cylinders of
a nonempty metric space (X, d) is the metric uniformity on X. It is denoted
by U
X
, U
d
, or even U , if the space under consideration is implied. Given X:= ,
put U
X
:= . An element of U
X
is an entourage on X.
4.1.6. If U is a metric uniformity then
(1) U l I
X
;
(2) U U U
1
U ;
(3) ( U U ) ( V U ) V V U;
(4) U : U U = I
X
.
4.1.7. Remark. Property 4.1.6 (4) reecting 4.1.1 (1) is often expressed as
follows: U is a Hausdor or separated uniformity.
4.1.8. Given a space X with uniformity U
X
, put
(x):= U(x) : U U .
Then (x) is a lter for every x in X. Moreover,
(1) (x) l x;
(2) ( U (x)) ( V (x) & V U) ( y V ) V (y).
4.1.9. Definition. The mapping : x (x) is the metric topology on X.
An element of (x) is a neighborhood of x or about x. More complete designations
for the topology are also in use:
X
, (U ), etc.
4.1.10. Remark. All closed balls centered at x form a base for the neighbor-
hood lter of x. The same is true of open balls. Note also that there are disjoint
(= nonintersecting) neighborhoods of dierent points in X. This property, ciphered
in 4.1.6 (4), reads:
X
is a Hausdor or separated topology.
4.1.11. Definition. A subset G of X is an open set in X whenever G is
a neighborhood of its every point (i.e., G Op() (( x G) G (x))).
A subset F of X is a closed set in X whenever its complement to X is open
(in symbols, F Cl() (X F Op())).
4.1.12. The union of a family of open sets and the intersection of a nite
family of open sets are open. The intersection of a family of closed sets and the
union of a nite family of closed sets are closed.
4.1.13. Definition. Given a subset U of X, put
int U :=

U := G Op(
X
) : G U;
cl U := U := F Cl(
X
) : F U.
44 Chapter 4
The set int U is the interior of U and its elements are interior points of U. The
set cl U is the closure of U and its elements are adherent to U. The exterior of U
is the interior of X U; the elements of the former are exterior to U. A boundary
point of U is by agreement a point of X neither interior nor exterior to U. The
collection of all boundary points of U is called the boundary of U or the frontier
of U and denoted by fr U or U.
4.1.14. A set U is a neighborhood about x if and only if x is an interior point
of U.
An Excursus into Metric Spaces 45
4.1.15. Remark. In connection with 4.1.14, the set Op(
X
) is also referred
to as the topology of the space X under study, since
X
is uniquely determined
from Op(
X
). The same relates to Cl(
X
), the collection of all closed sets in X.
4.1.16. Definition. A lterbase B on X converges to x in X or x is a limit
of B (in symbols, B x) if l B is ner than the neighborhood lter of x; i.e.,
l B (x).
4.1.17. Definition. A net or (generalized) sequence (x

converges to x
(in symbols, x

x) if the tail lter of (x

) converges to x. Other familiar terms


and designations are freely employed. For instance, x = lim

and x is a limit
of (x

) as ranges over .
4.1.18. Remark. A limit of a lter, as well as a limit of a net, is unique
in a metric space X. This is another way of expressing the fact that the topology
of X is separated.
4.1.19. For a nonempty set U and a point x the following statements are
equivalent:
(1) x is an adherent point of U;
(2) there is a lter F such that F x and U F;
(3) there is a sequence (x

whose entries are in U and which con-


verges to x.
(1) (2): Since x is not exterior to U, the join F:= (x)l U is available
of the pair of the lters (x) and l U.
(2) (3): Let F x and U F. Direct F by reverse inclusion. Take
x
V
V U for V F. It is clear that x
V
x.
(3) (1): Let V be a closed set. Take a sequence (x

in V such that
x

x. In this case it suces to show that x V , which is happily evident.


Indeed, were x in X V we would nd at least one such that x

X V .
4.1.20. Remark. It may be assumed that F has a countable base in 4.1.19
(2), and := N in 4.1.19 (3). This property is sometimes formulated as follows:
In metric spaces the rst axiom of countability is fullled.
4.2. Continuity and Uniform Continuity
4.2.1. If f : X Y and
X
and
Y
are topologies on X and Y then the
following conditions are equivalent:
46 Chapter 4
(1) G Op(
Y
) f
1
(G) Op(
X
);
(2) F Cl(
Y
) f
1
(F) Cl(
X
);
(3) f(
X
(x))
Y
(f(x)) for all x X;
(4) (x X & F x) (f(F) f(x)) for a lter F;
(5) f(x

) f(x) for every point x and every sequence (x

) convergent
to x.
The equivalence (1) (2) follows from 4.1.11. It remains to demonstrate
that (1) (3) (4) (5) (2).
(1) (3): If V
Y
(f(x)) then W := int V Op(
Y
) and f(x) W. Hence,
f
1
(W) Op(
X
) and x f
1
(W). In other words, f
1
(W)
X
(x) (see 4.1.14).
Moreover, f
1
(V ) f
1
(W) and, consequently, f
1
(V )
X
(x). Finally, V
f(f
1
(V )).
(3) (4): Given F x, by Denition 4.1.16 l F
X
(x). From the
hypothesis infer that f(F) f(
X
(x))
Y
(f(x)). Appealing to 4.1.16, again
reveal the sought instance of convergence, f(F) f(x).
(4) (5): The image of the tail lter of (x

under f is coarser than the


tail lter of (f(x

))

.
(5) (2): Take a closed set F in Y . If F = then f
1
(F) is also empty
and, hence, closed. Assume F nonempty and let x be an adherent point of f
1
(F).
Consider a sequence (x

converging to x and consisting of points in f


1
(F)
(the claim of 4.1.19 yields existence). Then f(x

) F and f(x

) f(x). Another
citation of 4.1.19 guarantees that f(x) F and, consequently, x f
1
(F).
4.2.2. Definition. A mapping satisfying one (and hence all) of the equivalent
conditions 4.2.1 (1)4.2.1 (5) is continuous. If 4.2.1 (5) holds at a xed point x then
f is said to be continuous at x. Thus, f is continuous on X whenever f is continuous
at every point of X.
4.2.3. Every composition of continuous mappings is continuous.
Apply 4.2.1 (5) thrice.
4.2.4. Let f : X Y and let U
X
and U
Y
be uniformities on X and Y . The
following statements are equivalent:
(1) ( V U
Y
) ( U U
X
) (( x, y)(x, y) U (f(x), f(y)) V );
(2) ( V U
Y
) f
1
V f U
X
;
(3) f

(U
X
) U
Y
, with f

: X
2
Y dened as f

: (x, y)
(f(x), f(y));
(4) ( V U
Y
) f
1
(V ) U
X
; i.e., f
1
(U
Y
) U
X
.
An Excursus into Metric Spaces 47
By 1.1.10, given U X
2
and V Y
2
, observe that
f
1
V f =
_
(v
1
,v
2
)V
f
1
(v
1
) f
1
(v
2
)
= (x, y) X
2
: (f(x), f(y)) V = f
1
(V );
f U f
1
=
_
(u
1
,u
2
)U
f(u
1
) f(u
2
)
= (f(u
1
), f(u
2
)) : (u
1
, u
2
) U = f

(U).
4.2.5. Definition. A mapping f : X Y satisfying one (and hence all)
of the equivalent conditions 4.2.4 (1)4.2.4 (4) is called uniformly continuous (the
term uniform continuous is also in common parlance).
4.2.6. Every composition of uniformly continuous mappings is uniformly con-
tinuous.
Consider f : X Y , g : Y Z and h:= g f : X Z. Plainly,
h

(x, y) = (h(x), h(y)) = (g(f(x)), g(f(y))) = g

(f(x), f(y)) = g

(x, y)
for all x, y X. Hence, by 4.2.4 (3) h

(U
X
) = g

(f

(U
X
)) g

(U
Y
) U
Z
,
meaning that h is uniformly continuous.
4.2.7. Every uniformly continuous mapping is continuous.
4.2.8. Definition. Let E be a set of mappings from X into Y and let U
X
and U
Y
be the uniformities in X and Y . The set E is equicontinuous if
( V U
Y
)

fE
f
1
V f U
X
.
4.2.9. Every equicontinuous set consists of uniformly continuous mappings.
Every nite set of uniformly continuous mappings is equicontinuous.
4.3. Semicontinuity
4.3.1. Let (X
1
, d
1
) and (X
2
, d
2
) be metric spaces, and := X X. Given
x:= (x
1
, x
2
) and y := (y
1
, y
2
), put
d(x, y):= d
1
(x
1
, y
1
) +d
2
(x
2
, y
2
).
Then d is a metric on . Moreover, for every x:= (x
1
, x
2
) in the presentation holds:
(x) = lU
1
U
2
: U
1

X
1
(x
1
), U
2

X
2
(x
2
).
4.3.2. Definition. The topology is called the product of
X
1
and
X
2
or the
product topology of X
1
X
2
. This topology on X
1
X
2
is denoted by
X
1

X
2
.
48 Chapter 4
4.3.3. Definition. A function f : X R

is lower semicontinuous if its


epigraph epi f is closed in the product topology of X R.
4.3.4. Examples.
(1) A continuous real-valued function f : X R is lower semicontinu-
ous.
(2) If f

: X R

is lower semicontinuous for all , then the


upper envelope f(x) := supf

(x) : (x X) is also lower semicontinuous.


A simple reason behind this is the equality epi f =

epi f

.
4.3.5. A function f : X R

is lower semicontinuous if and only if


x X f(x) = lim
yx
inf f(y).
Here, as usual,
lim
yx
inf f(y):= lim
yx
f(y):= sup
U(x)
inf f(U)
is the lower limit of f at x (with respect to (x)).
: If x , domf then (x, t) , epi f for all t R. Hence, there is a neighbor-
hood U
t
of x such that inf f(U
t
) > t. This implies that lim
yx
inf f(y) = + =
f(x). Suppose that x domf. Then inf f(V ) > for some neighborhood V
of x. Choose > 0 and for an arbitrary U in (x) included in V nd x
U
in
U so that inf f(U) f(x
U
) . By construction, x
U
domf and, moreover,
x
U
x (by implication, the set of neighborhoods of x is endowed with the con-
ventional order by inclusion, cf. 1.3.1). Put t
U
:= inf f(U) + . It is clear that
t
U
t := lim
yx
inf f(y) + . Since (x
U
, t
U
) epi f, from the closure property
of epi f obtain (x, t) epi f. Thus,
lim
yx
inf f(y) + f(x) lim
yx
inf f(y).
: If (x, t) , epi f then
t < lim
yx
inf f(y) = sup inf
U(x)
f(U).
Therefore, inf f(U) > t for some neighborhood U of x. It follows that the comple-
ment of epi f to X R is open.
4.3.6. Remark. The property, stated in 4.3.5, may be accepted as an initial
denition of lower semicontinuity at a point.
4.3.7. A function f : X R is continuous if and only if both f and f are
lower semicontinuous.
An Excursus into Metric Spaces 49
4.3.8. A function f : X R

is lower semicontinuous if and only if for every


t R the level set f t is closed.
: If x , f t then t < f(x). By 4.3.5, t < inf f(U) in a suitable
neighborhood U about x. In other words, the complement of f t to X is open.
: Given lim
yx
inf f(y) t < f(x) for some x X and t R, choose > 0
such that t + < f(x). Repeating the argument of 4.3.5, given U (x) take
a point x
U
in U f inf f(U) +. Undoubtedly x
U
f t + and x
U
x,
a contradiction.
4.4. Compactness
4.4.1. Definition. A subset C of X is called a compact set whenever for
every subset E of Op(
X
) with the property C G : G E there is a nite
subset E
0
in E such that still C G : G E
0
.
4.4.2. Remark. Denition 4.4.1 is verbalized as follows: A set is compact
if its every open cover has a nite subcover. The terms covering and subcov-
ering are also in current usage.
4.4.3. Every closed subset of a compact set is also compact. Every compact
set is closed.
4.4.4. Remark. With regard to 4.4.3, it stands to reason to use the term
relatively compact set for a set whose closure is compact.
4.4.5. Weierstrass Theorem. The image of a compact set under a continu-
ous mapping is compact.
The inverse images of sets in an open cover of the image compose an open
cover of the original set.
4.4.6. Each lower semicontinuous real-valued function, dened on a nonempty
compact set, assumes its least value; i.e., the image of a compact domain has a least
element.
Suppose that f : X R

and X is compact. Let t


0
:= inf f(X). In the
case t
0
= + there is nothing left to proof. If t
0
< + then put T := t R :
t > t
0
. The set U
t
:= f t with t T is nonempty and closed. Check that
U
t
: t T is not empty (then every element x of the intersection meets the
claim: f(x) = inf f(X)). Suppose the contrary. Then the sets X U
t
: t T
compose an open cover of X. Rening a nite subcover, deduce U
t
: t T
0
= .
However, this equality is false, since U
t
1
U
t
2
= U
t
1
t
2
for t
1
, t
2
T.
4.4.7. Bourbaki Criterion. A space is compact if and only if every ultralter
on it converges (cf. 9.4.4).
50 Chapter 4
4.4.8. The product of compact sets is compact.
It suces to apply the Bourbaki Criterion twice.
4.4.9. Cantor Theorem. Every continuous mapping on a compact set is
uniformly continuous.
4.5. Completeness Completeness
4.5.1. If B is a lterbase on X then B
2
: B B is a lterbase (and a base
for the lter B

) on X
2
.
(B
1
B
1
) (B
2
B
2
) (B
1
B
2
) (B
1
B
2
)
4.5.2. Definition. Let U
X
be a uniformity on X. A lter F is called
a Cauchy lter or even Cauchy (the latter might seem preposterous) if F

U
X
.
A net in X is a Cauchy net or a fundamental net if its tail lter is a Cauchy lter.
The term fundamental sequence is treated in a similar fashion.
4.5.3. Remark. If V is an entourage on X and U is a subset of X then U is
V -small whenever U
2
V . In particular, U is B

-small (or simply -small) if and


only if the diameter of U, diamU := sup(U
2
), is less than or equal to . In this
connection, the denition of a Cauchy lter is expressed as follows: A lter is
a Cauchy lter if and only if it contains arbitrarily small sets.
4.5.4. In a metric space the following conditions are equivalent:
(1) every Cauchy lter converges;
(2) each Cauchy net has a limit;
(3) every fundamental sequence converges.
The implications (1) (2) (3) are obvious; therefore, we are left with
establishing (3) (1).
Given a Cauchy lter F, let U
n
in F be a B
1/n
-small set. Put V
n
:= U
1
. . .U
n
and take x
n
V
n
. Observe that V
1
V
2
. . . and diamV
n

1
/
n
. Hence, (x
n
)
is a fundamental sequence. By hypothesis it has a limit, x := limx
n
. Check that
F x. To this end, choose n
0
N such that d(x
m
, x)
1
/
2n
as m n
0
. Then
for all n N deduce that d(x
p
, y) diamV
p

1
/
2n
and d(x
p
, x)
1
/
2n
whenever
p:= n
0
2n and y V
p
. It follows that y V
p
d(x, y)
1
/
n
; i.e., V
p
B
1/n
(x).
In conclusion, F (x).
4.5.5. Definition. A metric space satisfying one (and hence all) of the equiv-
alent conditions 4.5.4 (1)4.5.4 (3) is called complete.
4.5.6. Cantor Criterion. A metric space X is complete if and only if ev-
ery nonempty downward-ltered family of nonempty closed subsets of X whose
diameters tend to zero has a point of intersection.
An Excursus into Metric Spaces 51
: If B is such family then by Denition 1.3.1 B is a lterbase. By hy-
pothesis, B is a base for a Cauchy lter. Therefore, there is a limit: B x. The
point x meets the claim.
: Let F be a Cauchy lter. Put B:= cl V : V F. The diameters of the
sets in B tend to zero. Hence, there is a point x such that x cl V for all V F.
Plainly, F x. Indeed, let V be an

/
2
-small member of F and y V . Some
y

in V is such that d(x, y

/
2
. Therefore, d(x, y) d(x, y

) + d(y

, y) .
Consequently, V B

(x) and so B

(x) F.
4.5.7. Nested Ball Theorem. A metric space is complete if and only if every
nested (= decreasing by inclusion) sequence of balls whose radii tend to zero has
a unique point of intersection.
4.5.8. The image of a Cauchy lter under a uniformly continuous mapping is
a Cauchy lter.
Let a mapping f act from a space X with uniformity U
X
into a space Y with
uniformity U
Y
and let F be a Cauchy lter on X. If V U
Y
then, by Denition
4.2.5, f
1
V f U
X
(cf. 4.2.4 (2)). Since F is a Cauchy lter, U
2
f
1
V f
for some U F. It turns out that f(U) is V -small. Indeed,
f(U)
2
=
_
(u
1
,u
2
)U
2
f(u
1
) f(u
2
)
= f U
2
f
1
f (f
1
V f) f
1
= (f f
1
) V (f f
1
) V,
because, by 1.1.6, f f
1
= I
imf
I
Y
.
4.5.9. The product of complete spaces is complete.
The claim is immediate from 4.5.8 and 4.5.4.
4.5.10. Let X
0
be dense in X (i.e., cl X
0
= X). Assume further that f
0
:
X
0
Y is a uniformly continuous mapping from X
0
into some complete space Y .
Then there is a unique uniformly continuous mapping f : X Y extending f
0
;
i.e., f[
X
0
= f
0
.
For x X, the lter F
x
:= U X
0
: U
X
(x) is a Cauchy lter on X
0
.
Therefore, 4.5.8 implies that f
0
(F
X
) is a Cauchy lter on Y . By the completeness
of Y , there is a limit y Y ; that is, f
0
(F
x
) y. Moreover, this limit is unique
(cf. 4.1.18). Dene f(x):= y. Checking uniform continuity for f readily completes
the proof.
4.5.11. Definition. An isometry or isometric embedding or isometric map-
ping of X into

X is a mapping f : (X, d) (

X,

d ) such that d =

d f

.
A mapping f is an isometry onto

X if f is an isometry of X into

X and, moreover,
imf =

X. The expressions, an isometry of X and

X or an isometry between X
and

X and the like, are also in common parlance.
52 Chapter 4
4.5.12. Hausdor Completion Theorem. If (X, d) is a metric space then
there are a complete metric space (

X,

d ) and an isometry : (X, d) (

X,

d )
onto a dense subspace of (

X,

d ). The space (

X,

d ) is unique to within isometry
in the following sense: The diagram
(

X
1
,

d
1
)
(X, d)
(

X,

d )

commutes for some isometry : (

X,

d ) (

X
1
,

d
1
), where
1
: (X, d) (

X
1
,

d
1
)
is an isometry of X onto a dense subspace of a complete space (

X
1
,

d
1
).
Uniqueness up to isometry follows from 4.5.10. For, if
0
:=
1

1
then

0
is an isometry of the dense subspace (X) of

X onto the dense subspace
1
(X)
of

X
1
. Let be the unique extension of
0
to the whole of

X. It is sucient
to show that acts onto

X
1
. Take x
1
in

X
1
. This element is the limit of some
sequence (
1
(x
n
)), with x
n
X. Clearly, the sequence (x
n
) is fundamental. Thus,
((x
n
)) is a fundamental sequence in

X. Let x := lim(x
n
), x

X. Proceed as
follows: ( x) = lim
0
((x
n
)) = lim
1

1
((x
n
)) = lim
1
(x
n
) = x
1
.
We now sketch out the proof that

X exists. Consider the set of all fundamental
sequences in X. Dene some equivalence relation in X by putting x
1
x
2

d(x
1
(n), x
2
(n)) 0. Assign

X:= / and

d((x
1
), (x
2
)) := limd(x
1
(n), x
2
(n)),
where :

X is the coset mapping. An isometry : (X, d) (

X,

d ) is
immediate: (x):= (n x (n N)).
4.5.13. Definition. The space (

X,

d ) introduced in Theorem 4.5.12, as well
as each space isomorphic to it, is called a completion of (X, d).
4.5.14. Definition. A set X
0
in a metric space (X, d) is said to be complete
if the metric space (X
0
, d[
X
2
0
), a subspace of (X, d), is complete.
4.5.15. Every closed subset of a complete space is complete. Every complete
set is closed.
4.5.16. If X
0
is a subspace of a complete metric space X then a completion
of X
0
is isometric to the closure of X
0
in X.
Let

X:= cl X
0
and let : X
0


X be the identical embedding. It is evident
that is an isometry onto a dense subspace. Moreover, by 4.5.15

X is complete.
Appealing to 4.5.12 ends the proof.
4.6. Compactness and Completeness
4.6.1. A compact space is complete.
An Excursus into Metric Spaces 53
4.6.2. Definition. Let U be a subset of X and V U
X
. A set E in X is
a V -net for U if U V (E).
4.6.3. Definition. A subset U of X is a totally bounded set in X if for every V
in U
X
there is a nite V -net for U.
4.6.4. If for every V in U a set U in X has a totally bounded V -net then U
is totally bounded.
Let V U
X
and W U
X
be such that WW V . Take a totally bounded
W-net F for U; i.e., U W(F). Since F is totally bounded, there is a nite W-net
E for F; that is, F W(E). Finally,
U W(F) W(W(E)) = W W(E) V (E);
i.e., E is a nite V -net for U.
4.6.5. A subset U of X is totally bounded if and only if for every V in U
X
there is a family U
1
, . . . , U
n
of subsets of U such that U = U
1
. . . U
n
and each
of the sets U
1
, . . . , U
n
is V -small.
4.6.6. Remark. The claim of 4.6.5 is verbalized as follows: A set is totally
bounded if and only if it has nite covers consisting of arbitrarily small sets.
4.6.7. Hausdor Criterion. A set is compact if and only if it is complete
and totally bounded.
4.6.8. Let C(Q, F) be the space of continuous functions with domain a com-
pact set Q and range a subset of F. Furnish this space with the Chebyshev metric
d(f, g):= sup
xQ
d
F
(f(x), g(x)):= sup
xQ
[f(x) g(x)[ (f, g C(Q, F));
and, given U
F
, put
U

:=
_
(f, g) C(Q, F)
2
: g f
1

_
.
Then U
d
= l U

: U
F
.
4.6.9. The space C(Q, F) is complete.
4.6.10. AscoliArzel`a Theorem. A subset E of C(Q, F) is relatively com-
pact if and only if E is equicontinuous and the set g(Q) : g E is totally
bounded in F.
54 Chapter 4
: It is beyond a doubt that g(Q) : g E is totally bounded. To
show equicontinuity for E take U
F
and choose a symmetric entourage

such
that

. By the Hausdor Criterion, there is a nite U

-net E

for E .
Consider the entourage U U
Q
dened as
U :=

fE

f
1

f
(cf. 4.2.9). Given g E and f E

such that g f
1

, observe that

=
1
(g f
1
)
1
= (f
1
)
1
g
1
= f g
1
.
Moreover, the composition rules for correspondences and 4.6.8 imply
g

(U) = g U g
1
g (f
1

f) g
1
(g f
1
)

(f g
1
)

.
Since g is arbitrary, the resulting inclusion guarantees that E is equicontinuous.
: By 4.5.15, 4.6.7, 4.6.8, and 4.6.9, it is sucient to construct a nite U

-net
for E given U
F
. Choose

U
F
such that

and nd an open
symmetric entourage U U
Q
from the condition
U

gE
g
1

g
(by the equicontinuity property of E , such an U is available). The family U(x) :
x Q clearly forms an open cover of Q. By the compactness of Q, rene a nite
subcover U(x
0
) : x
0
Q
0
. In particular, from 1.1.10 derive
I
Q

_
x
0
Q
0
U(x
0
) U(x
0
)
=
_
(x
0
,x
0
)I
Q
0
U
1
(x
0
) U(x
0
) = U I
Q
0
U.
The set g[
Q
0
: g E is totally bounded in F
Q
0
. Consequently, there is a nite

-net for this set. Speaking more precisely, there is a nite subset E

of E with the
following property:
g I
Q
0
f
1

An Excursus into Metric Spaces 55


for every g in E and some f in E

. Using the estimates, successively infer that


g f
1
= g I
Q
f
1
g (U I
Q
0
U) f
1
g (g
1

g) I
Q
0
(f
1

f) f
1
= (g g
1
)

(g I
Q
0
f
1
)

(f f
1
)
= I
img

(g I
Q
0
f
1
)

I
imf

.
Thus, by 4.6.8, E

is a nite U

-net for E .
4.6.11. Remark. It is an enlightening exercise to translate the proof of the
AscoliArzel`a Theorem into the - language. The necessary vocabulary is in hand:
and U

stand for ,

is

/
3
, and is U. It is also rewarding and instructive
to nd generalizations of the AscoliArzel`a Theorem for mappings acting into more
abstract spaces.
4.7. Baire Spaces
4.7.1. Definition. A set U is said to be nowhere dense or rare whenever its
closure lacks interior points; i.e., int cl U = . A set U is meager or a set of rst
category if U is included into a countable union of rare sets; i.e., U
nN
U
n
with
int cl U
n
= . A nonmeager set (which is not of rst category by common parlance)
is also referred to as a set of second category.
4.7.2. Definition. A space is a Baire space if its every nonempty open set is
nonmeager.
4.7.3. The following statements are equivalent:
(1) X is a Baire space;
(2) every countable union of closed rare sets in X lacks interior points;
(3) the intersection of a countable family of open everywhere dense sets
(i.e., dense in X) is everywhere dense;
(4) the complement of each meager set to X is everywhere dense.
(1) (2): If U :=
nN
U
n
, U
n
= cl U
n
, and int U
n
= then U is meager.
Observe that int U U and int U is open; hence, int U as a meager set is necessarily
empty, for X is a Baire space.
(2) (3): Let U :=
nN
G
n
, where G
n
s are open and cl G
n
= X. Then
XU = X
nN
G
n
=
nN
(XG
n
). Moreover, XG
n
is closed and int(XG
n
) =
(since cl G
n
= X). Therefore, int(X U) = , which implies that the exterior
of U is empty; i.e., U is everywhere dense.
56 Chapter 4
(3) (4): Let U be a meager set in X; i.e., U
nN
U
n
and int cl U
n
= .
It may be assumed that U
n
= cl U
n
. Then G
n
:= X U
n
is open and everywhere
dense. By hypothesis,
nN
G
n
= X
nN
U
n
is everywhere dense. Moreover, the
set is included into X U, and so X U is everywhere dense.
(4) (1): If U is nonempty open set in X then XU is not everywhere dense.
Consequently, U is nonmeager.
4.7.4. Remark. In connection with 4.7.3 (4), observe that the complement
of a meager set is (sometimes) termed a residual or comeager set. A residual set
in a Baire space is nonmeager.
4.7.5. Osgood Theorem. Let X be a Baire space and let (f

: X R)

be a family of lower semicontinuous functions such that supf

(x) : < +
for all x X. Then each nonempty open set G in X includes a nonempty open
subset G
0
on which (f

is uniformly bounded above; i.e., sup


xG
0
supf

(x) :
+.
4.7.6. Baire Category Theorem. A complete metric space is a Baire space.
Let G be a nonempty open set and x
0
G. Suppose by way of contradiction
that G is meager; i.e., G
nN
U
n
, where int U
n
= and U
n
= cl U
n
. Take

0
> 0 satisfying B

0
(x
0
) G. It is obvious that U
1
is not included into B

0
/2
(x
0
).
Consequently, there is an element x
1
in B

0
/2
(x
0
) U
1
. Since U
1
is closed, nd
1
satisfying 0 <
1


0
/
2
and B

1
(x
1
) U
1
= . Check that B

1
(x
1
) B

0
(x
0
).
For, if d(x
1
, y
1
)
1
, then d(y
1
, x
0
) d(y
1
, x
1
) +d(x
1
, x
0
)
1
+

0
/
2
, because
d(x
1
, x
0
)

0
/
2
. The ball B

1
/2
(x
1
) does not lie in U
2
entirely. It is thus possible
to nd x
2
B

1
/2
(x
1
) U
2
and 0 <
2


1
/
2
satisfying B

2
(x
2
) U
2
= . It is
easy that again B

2
(x
2
) B

1
(x
1
). Proceeding by induction, obtain the sequence
of nested balls B

0
(x
0
) B

1
(x
1
) B

2
(x
2
) . . . ; moreover,
n+1


n
/
2
and
B

n
(x
n
) U
n
= . By the Nested Ball Theorem, the balls have a common point,
x := limx
n
. Further, x ,=
nN
U
n
; and, hence, x , G. On the other hand,
x B

0
(x
0
) G, a contradiction.
4.7.7. Remark. The Baire Category Theorem is often used as a pure exis-
tence theorem. As a classical example, consider the existence problem for contin-
uous nowhere dierentiable functions.
Given f : [0, 1] R and x [0, 1), put
D
+
f(x):= lim
h0
inf
f(x +h) f(x)
h
;
D
+
f(x):= lim
h0
sup
f(x +h) f(x)
h
.
An Excursus into Metric Spaces 57
The elements D
+
f(x) and D
+
f(x) of the extended axis R are the lower right and
upper right Dini derivatives of f at x. Further, let D stand for the set of functions
f in C([0, 1], R) such that for some x [0, 1) the elements D
+
f(x) and D
+
f(x)
belong to R; i.e., they are nite. Then D is a meager set. Hence, the functions
lacking derivatives at every point of (0, 1) are everywhere dense in C([0, 1], R).
However, the explicit examples of such functions are not at all easy to nd and
grasp. Below a few of the most popular are listed:
van der Waerdens function:

n=0
4
n
x
4
n
,
with x := (x [x]) (1 + [x] x), the distance from x to the whole number
nearest to x;
Riemanns function:
+

n=0
1
n
2
sin (n
2
x);
and, nally, the historically rst
Weierstrasss function:

n=0
b
n
cos (a
n
x),
with a an odd positive integer, 0 < b < 1 and ab > 1 +
3
/
2
.
4.8. The Jordan Curve Theorem and Rough Drafts
4.8.1. Remark. In topology, in particular, many signicant and curious facts
of the metric space R
2
are scrutinized. Here we recall those of them which are of
use in the sequel and whose role is known from complex analysis.
4.8.2. Definition. A (Jordan) arc is a homeomorphic image of a (nonde-
generate) interval of the real axis. Recall that a homeomorphism or a topological
mapping is by denition a one-to-one continuous mapping whose inverse is also
continuous. A (simple Jordan) loop is a homeomorphic image of a circle. Concepts
like smooth loop are understood naturally.
4.8.3. Jordan Curve Theorem. If is a simple loop in R
2
then there are
open sets G
1
and G
2
such that
G
1
G
2
= R
2
; = G
1
= G
2
.
4.8.4. Remark. Either G
1
or G
2
in 4.8.3 is bounded. Moreover, each of the
two sets is connected; i.e., it cannot be presented as the union of two nonempty
disjoint open sets. In this regard the Jordan Curve Theorem is often read as
follows: A simple loop divides the plane into two domains and serves as their
mutual boundary.
58 Chapter 4
4.8.5. Definition. Let D
1
, . . . , D
n
and D be closed disks (= closed balls)
in the plane which satisfy D
1
, . . . , D
n
int D and D
m
D
k
= as m ,= k. The
set
D
n
_
k=1
int D
k
is a holey disk or, more formally, a perforated disk. A subset of the plane which is
dieomorphic (= smoothly homeomorphic) to a holey disk is called a connected
elementary compactum. The union of a nite family of pairwise disjoint connected
elementary compacta is an elementary compactum.
4.8.6. Remark. The boundary F of an elementary compactum F comprises
nitely many disjoint smooth loops. Furthermore, the embedding of F into the
(oriented) plane R
2
induces on F the structure of an (oriented) manifold with (ori-
ented) boundary F. Observe also that, by 4.8.3, it makes sense to specify the
positive orientation of a smooth loop. This is done by indicating the orientation
induced on the boundary of the compact set surrounded by the loop.
4.8.7. If K is a compact subset of the plane and G is a nonempty open set
that includes K then there is a nonempty elementary compactum F such that
K int F F G.
4.8.8. Definition. Every set F appearing in 4.8.7 is referred to as a rough
draft for the pair (K, G) or an oriented envelope of K in G.
Exercises
4.1. Give examples of metric spaces. Find methods for producing new metric spaces.
4.2. Which lter on X
2
coincides with some metric uniformity on X?
4.3. Let S be the space of measurable functions on [0, 1] endowed with the metric
d(f, g):=
1
_
0
[f(t) g(t)[
1 +[f(t) g(t)[
dt (f, g S)
with some natural identication implied (specify it!). Find the meaning of convergence in the
space.
4.4. Given , N
N
, put
d(, ):= 1/ min k N :
k
,=
k
.
Check that d is a metric and the space N
N
is homeomorphic with the set of irrational numbers.
An Excursus into Metric Spaces 59
4.5. Is it possible to metrize pointwise convergence in the sequence space? In the function
space F
[0,1]
?
4.6. How to introduce a reasonable metric into the countable product of metric spaces? Into
an arbitrary product of metric spaces?
4.7. Describe the function classes distinguishable by erroneous denitions of continuity and
uniform continuity.
4.8. Given nonempty compact subsets A and B of the spaces R
N
, dene
d(A, B):=
_
sup
xA
inf
yB
[x y[
_

_
sup
yB
inf
xA
[x y[
_
.
Show that d is a metric. The metric is called the Hausdor metric. What is the meaning
of convergence in this metric?
4.9. Prove that nonempty compact convex subsets of a compact convex set in R
N
constitute
a compact set with respect to the Hausdor metric. How does this claim relate to the AscoliArzel`a
Theorem?
4.10. Prove that each lower semicontinuous function on R
N
is the upper envelope of some
family of continuous functions.
4.11. Explicate the interplay between continuous and closed mappings (in the product topol-
ogy) of metric spaces.
4.12. Find out when a continuous mapping of a metric space into a complete metric spaces
is extendible onto a completion of the initial space.
4.13. Describe compact sets in the product of metric spaces.
4.14. Let (Y, d) be a complete metric spaces. A mapping F : Y Y is called expanding
whenever d(F(x), F(y)) d(x, y) for some > 1 and all x, y Y . Assume that an expanding
mapping F : Y Y acts onto Y . Prove that F is one-to-one and possesses a sole xed point.
4.15. Prove that no compact set can be mapped isometrically onto a proper part of it.
4.16. Show normality of an arbitrary metric space (see 9.3.11).
4.17. Under what conditions a countable subset of a complete metric spaces is nonmeager?
4.18. Is it possible to characterize uniform continuity in terms of convergent sequences?
4.19. In which metric spaces does each continuous real-valued function attain the supremum
and inmum of its range? When is it bounded?
Chapter 5
Multinormed and Banach Spaces
5.1. Seminorms and Multinorms
5.1.1. Let X be a vector space over a basic eld F and let p : X R

be
a seminorm. Then
(1) domp is a subspace of X;
(2) p(x) 0 for all x X;
(3) the kernel ker p:= p = 0 is a subspace in X;
(4) the sets

B
p
:= p < 1 and B
p
:= p 1 are absolutely convex;
moreover, p is the Minkowski functional of every set B such that

B
p
B B
p
;
(5) X = domp if and only if

B
p
is absorbing.
If x
1
, x
2
domp and
1
,
2
F then by 3.7.6
p(
1
x
1
+
2
x
2
) [
1
[p(x
1
) +[
2
[p(x
2
) < ++ (+) = +.
Hence, (1) holds. Suppose to the contrary that (2) is false; i.e., p(x) < 0 for some
x X. Observe that 0 p(x) + p(x) < p(x) = p(x) < 0, a contradiction. The
claim of (3) is immediate from (2) and the subadditivity of p. The validity of (4)
and (5) has been examined in part (cf. 3.8.8). What was left unproven follows from
the Gauge Theorem.
5.1.2. If p, q : X R

are two seminorms then the inequality p q (in (R

)
X
)
holds if and only if B
p
B
q
.
: Evidently, q 1 p 1.
: In view of 5.1.1 (4), observe that p = p
B
p
and q = p
B
q
. Take t
1
, t
2
R
such that t
1
< t
2
. If t
1
< 0 then q t
1
= , and so q t
1
p t
2
.
If t
1
0 then t
1
B
q
t
1
B
p
t
2
B
p
. Thus, by 3.8.3, p q.
Multinormed and Banach Spaces 61
5.1.3. Let X and Y be vector spaces, let T XY be a linear correspondence,
and let p : Y R

be a seminorm. If p
T
(x) := inf p T(x) for x X then
p
T
: X R

is a seminorm and the set B


T
:= T
1
(B
p
) is absolutely convex.
In addition, p
T
= p
B
T
.
Given x
1
, x
2
X and
1
,
2
F, infer that
p
T
(
1
x
1
+
2
x
2
) = inf p(T(
1
x
1
+
2
x
2
))
inf p(
1
T(x
1
) +
2
T(x
2
)) inf([
1
[p(T(x
1
)) +[
2
[p(T(x
2
)))
= [
1
[p
T
(x
1
) +[
2
[p
T
(x
2
);
i.e., p
T
is a seminorm.
The absolute convexity of B
T
is a consequence of 5.1.1 (4) and 3.1.8. If x B
T
then (x, y) T for some y B
p
. Hence, p
T
(x) p(y) 1; that is, B
T
B
p
T
.
If in turn x

B
p
T
then p
T
(x) = infp(y) : (x, y) T < 1. Thus, there is some
y T(x) such that p(y) < 1. Therefore, x T
1
(

B
p
) T
1
(B
p
) = B
T
. Finally,

B
p
T
B
T
B
p
T
. Referring to 5.1.1 (4), conclude that p
B
T
= p
T
.
5.1.4. Definition. The seminorm p
T
, constructed in 5.1.3, is the inverse
image or preimage of p under T.
5.1.5. Definition. Let p : X R be a seminorm (by 3.4.3, this implies that
domp = X). A pair (X, p) is referred to as a seminormed space. It is convenient
to take the liberty of calling X itself a seminormed space.
5.1.6. Definition. A multinorm on X is a nonempty set (a subset of R
X
)
of everywhere-dened seminorms. Such a multinorm is denoted by M
X
or simply
by M, if the underlying vector space is clear from the context. A pair (X, M
X
),
as well as X itself, is called a multinormed space.
5.1.7. A set M in (R

)
X
is a multinorm if and only if (X, p) is a seminormed
space for every p M.
5.1.8. Definition. A multinorm M
X
is a Hausdor or separated multinorm
whenever for all x X, x ,= 0, there is a seminorm p M
X
such that p(x) ,= 0.
In this case X is called a Hausdor or separated multinormed space.
5.1.9. Definition. A norm is a Hausdor multinorm presenting a singleton.
The sole element of a norm on a vector space X is also referred to as the norm
on X and is denoted by | | or (rarely) by | |
X
or even | [ X| if it is necessary
to indicate the space X. A pair (X, | |) is called a normed space; as a rule, the
same term applies to X.
62 Chapter 5
5.1.10. Examples.
(1) A seminormed space (X, p) can be treated as a multinormed space
(X, p). The same relates to a normed space.
(2) If M is the set of all (everywhere-dened) seminorms on a space X
then M is a Hausdor multinorm called the nest multinorm on X.
(3) Let (Y, N) be a multinormed space, and let T XY be a linear
correspondence such that domT = X. By 3.4.10 and 5.1.1 (5), for every p in N the
seminorm p
T
is dened everywhere, and hence M:= p
T
: p N is a multinorm
on X. The multinorm N is called the inverse image or preimage of N under the
correspondence T and is (sometimes) denoted by N
T
. Given T L(X, Y ), set
M := p T : p N and use the natural notation N T := M. Observe
in particular the case in which X is a subspace Y
0
of Y and T is the identical
embedding : Y
0
Y . At this juncture Y
0
is treated as a multinormed space with
multinorm N . Moreover, the abuse of the phrase N is a multinorm on Y
0
is
very convenient.
(4) Each basic eld F is endowed, as is well known, with the standard
norm [ [ : F R, the modulus of a scalar. Consider a vector space X and
f X
#
. Since f : X F, it is possible to dene the inverse image of the norm
on F as p
f
(x) := [f(x)[ (x X). If is some subspace of X
#
then the multinorm
(X, ):= : is the weak multinorm on X induced by .
(5) Let (X, p) be a seminormed space. Assume further that X
0
is a sub-
space of X and : X X/X
0
is the coset mapping. The linear correspondence

1
is dened on the whole of X/X
0
. Hence, the seminorm p

1 appears, called the


quotient seminorm of p by X
0
and denoted by p
X/X
0
. The space (X/X
0
, p
X/X
0
)
is called the quotient space of (X, p) by X
0
. The denition of quotient space for
an arbitrary multinormed space requires some subtlety and is introduced in 5.3.11.
(6) Let X be a vector space and let M (R

)
X
be a set of seminorms
on X. In this case M can be treated as a multinorm on the space X
0
:= domp :
p M. More precisely, thinking of the multinormed space (X
0
, p

: p M),
where is the identical embedding of X
0
into X, we say: M is a multinorm, or
Consider the (multinormed) space specied by M. The next example is typical:
The family of seminorms
_
p
,
(f):= sup
xR
N
[x

f(x)[ : and are multi-indices


_
species the (multinormed) space of innitely dierentiable functions on R
N
de-
creasing rapidly at innity (such functions are often called tempered, cf. 10.11.6).
Multinormed and Banach Spaces 63
(7) Let (X, | |) and (Y, | |) be normed spaces (over the same ground
eld F). Given T L(X, Y ), consider the operator norm of T, i.e. the quantity
|T|:= sup|Tx| : x X, |x| 1 = sup
xX
|Tx|
|x|
.
(From now on, in analogous situations we presume
0
/
0
:= 0.)
It is easily seen that | | : L(X, Y ) R

is a seminorm. Indeed, putting


B
X
:= | |
X
1 for T
1
, T
2
L(X, Y ) and
1
,
2
F, deduce that
|
1
T
1
+
2
T
2
| = sup | |

1
T
1
+
2
T
2
(B
X
)
= sup | |((
1
T
1
+
2
T
2
)(B
X
)) sup |
1
T
1
(B
X
) +
2
T
2
(B
X
)|
[
1
[ sup | |
T
1
(B
X
) +[
2
[ sup | |
T
2
(B
X
) = [
1
[ |T
1
| +[
2
[ |T
2
|.
The subspace B(X, Y ), the eective domain of denition of the above semi-
norm, is the space of bounded operators; and an element of B(X, Y ) is a bounded
operator. Observe that a shorter term operator customarily implies a bounded
operator. It is clear that B(X, Y ) is a normed space (under the operator norm).
Note also that T in L(X, Y ) is bounded if and only if T maintains the normative
inequality; i.e., there is a strictly positive number K such that
|Tx|
Y
K|x|
X
(x X).
Moreover, |T| is the greatest lower bound of the set of Ks appearing in the nor-
mative inequality.
(8) Let X be a vector space over F and let || be a norm on X. Assume
further that X

:= B(X, F) is the (normed) dual of X, i.e. the space of bounded


functionals fs with the dual norm
|f| = sup[f(x)[ : |x| 1 = sup
xX
[f(x)[
|x|
.
Consider X

:= (X

:= B(X

, F), the second dual of X. Given x X and


f X

, put x

:= (x) : f f(x). Undoubtedly, (x) (X

)
#
= L(X

, F).
In addition,
|x

| = |(x)| = sup[(x)(f)[ : |f[ X

1
= sup[f(x)[ : (x X) [f(x)[ |x|
X
= sup[f(x)[ : f [[(| |
X
) = |x|
X
.
The nal equality follows for instance from Theorem 3.6.5 and Lemma 3.7.9.
Thus, (x) X

for every x in X. It is plain that the operator : X X

,
64 Chapter 5
dened as : x (x), is linear and bounded; moreover, is a monomorphism and
|x| = |x| for all x X. The operator is referred to as the canonical embedding
of X into the second dual or more suggestively the double prime mapping. As a rule,
it is convenient to treat x and x

:= x as the same element; i.e., to consider X as


a subspace of X

. A normed space X is reexive if X and X

coincide (under the


indicated embedding!). Reexive spaces possess many advantages. Evidently, not
all spaces are reexive. Unfortunately, such is C([0, 1], F) which is irreexive (the
term nonreexive is also is common parlance).
5.1.11. Remark. The constructions, carried out in 5.1.10 (8), show some
symmetry or duality between X and X

. In this regard, the notation (x, f) :=


x [ f := f(x) symbolizes the action of x X on f X

(or the action of f on x).


To achieve and ensure the utmost conformity, it is a common practice to denote
elements of X

by symbols like x

; for instance, x [ x

= (x, x

) = x

(x).
5.2. The Uniformity and Topology of a Multinormed
Space
5.2.1. Let (X, p) be a seminormed space. Given x
1
, x
2
X, put d
p
(x
1
, x
2
)
:= p(x
1
x
2
). Then
(1) d
p
(X
2
) R
+
and d 0 I
X
;
(2) d
p
t = d
p
t
1
and d
p
t = td
p
1 (t R
+
0);
(3) d
p
t
1
d
p
t
2
d
p
t
1
+t
2
(t
1
, t
2
R
+
);
(4) d
p
t
1
d
p
t
2
d
p
t
1
t
2
(t
1
, t
2
R
+
);
(5) p is a norm d
p
is a metric.
5.2.2. Definition. The uniformity of a seminormed space (X, p) is the lter
U
p
:= l d
p
t : t R
+
0.
5.2.3. If U
p
is the uniformity of a seminormed space (X, p) then
(1) U
p
l I
X
;
(2) U U
p
U
1
U
p
;
(3) ( U U
p
) ( V U
p
) V V U.
5.2.4. Definition. Let (X, M) be a multinormed space. The lter U :=
supU
p
: p M is called the uniformity of X (the other designations are U
M
, U
X
,
etc.). (By virtue of 5.2.3 (1) and 1.3.13, the denition is sound.)
5.2.5. If (X, M) is a multinormed space and U is the uniformity of X then
(1) U l I
X
;
(2) U U U
1
U ;
(3) ( U U ) ( V U ) V V U.
Multinormed and Banach Spaces 65
Examine (3). Given U U , by 1.2.18 and 1.3.8 there are seminorms
p
1
, . . . , p
n
M such that U = U
p
1
,... ,p
n
]
= U
p
1
. . . U
p
n
. Using 1.3.13, nd
sets U
k
U
p
k
satisfying U U
1
. . . U
n
. Applying 5.2.3 (3), choose V
k
U
p
k
so
as to have V
k
V
k
U
k
. It is clear that
(V
1
. . . V
n
) (V
1
. . . V
n
) V
1
V
1
. . . V
n
V
n
U
1
. . . U
n
.
Moreover, V
1
. . . V
n
U
p
1
. . . U
p
n
U .
5.2.6. A multinorm M on X is separated if and only if so is the uniformity
U
M
; i.e., V : V U
M
= I
X
.
: Let (x, y) , I
X
; i.e., x ,= y. Then p(x y) > 0 for some seminorm p
in M. Hence, (x, y) , d
p

1
/
2
p(x y). But the last set is included in U
p
,
and thus in U
M
. Consequently, X
2
I
X
X
2
V : V U
M
. Furthermore,
I
X
V : V U
M
.
: If p(x) = 0 for all p M then (x, 0) V for every V in U
M
. Hence,
(x, 0) I
X
by hypothesis. Therefore, x = 0.
5.2.7. Given a space X with uniformity U
X
, dene
(x):= U(x) : U U
X
(x X).
Then (x) is a lter for every x X. Moreover,
(1) (x) l x;
(2) ( U (x)) ( V (x) & V U) ( y V ) V (y).
All is evident (cf. 4.1.8).
5.2.8. Definition. The mapping : x (x) is called the topology of a multi-
normed space (X, M); a member of (x) is a neighborhood about x. The designation
for the topology can be more detailed:
X
,
M
, (U
M
), etc.
5.2.9. The following presentation holds:

X
(x) = sup
p
(x) : p M
X

for all x X.
5.2.10. If X is a multinormed space then
U (x) U x
X
(0)
for all x X.
By 5.2.9 and 1.3.13, it suces to consider a seminormed space (X, p). In this
case for every > 0 the equality holds: d
p
(x) = B
p
+x, where B
p
:= p 1.
Indeed, if p(y x) then y = (
1
(y x)) + x and
1
(y x) B
p
. In turn,
if y B
p
+x, then p(y x) = inft > 0 : y x tB
p
.
66 Chapter 5
5.2.11. Remark. The proof of 5.2.10 demonstrates that in a seminormed
space (X, p) a key role is performed by the ball with radius 1 and centered at zero.
The ball bears the name of the unit ball of X and is denoted by B
p
, B
X
, etc.
5.2.12. A multinorm M
X
is separated if and only if so is the topology
X
;
i.e., given distinct x
1
and x
2
in X, there are neighborhoods U
1
in
X
(x
1
) and U
2
in
X
(x
2
) such that U
1
U
2
= .
: Let x
1
,= x
2
and let := p(x
1
x
2
) > 0 for p M
X
. Put U
1
:= x
1
+

/
3
B
p
and U
2
:= x
2
+

/
3
B
p
. By 5.2.10, U
k

X
(x
k
). Verify that U
1
U
2
= . Indeed,
if y U
1
U
2
then p(x
1
y)

/
3
and p(x
2
y)

/
3
. Therefore, p(x
1
x
2
)
2
/
3
< = p(x
1
x
2
), which is impossible.
: If (x
1
, x
2
) V : V U
X
then x
2
V (x
1
) : V U
X
. Thus,
x
1
= x
2
and, consequently, M
X
is separated by 5.2.6.
5.2.13. Remark. The presence of a uniformity and the corresponding topol-
ogy in a multinormed space readily justies using uniform and topological concepts
such as uniform continuity, completeness, continuity, openness, closure, etc.
5.2.14. Let (X, p) be a seminormed space and let X
0
be a subspace of X.
The quotient space (X/X
0
, p
X/X
0
) is separated if and only if X
0
is closed.
: If x , X
0
then (x) ,= 0 where, as usual, : X X/X
0
is the coset
mapping. By hypothesis, 0 ,= := p
X/X
0
((x)) = p

1((x)) = infp(x + x
0
) :
x
0
X
0
. Hence, the ball x +

/
2
B
p
does not meet X
0
and x is an exterior point
of X
0
. Thus, X
0
is closed.
: Suppose that x is a nonzero point of X/X
0
and x = (x) for some x in X.
If p
X/X
0
(x) = 0 then 0 = infp(x x
0
) : x
0
X
0
. In other words, there is
a sequence (x
n
) in X
0
converging to x. Consequently, by 4.1.19, x X
0
and x = 0,
a contradiction.
5.2.15. The closure of a -set is a -set.
Given U (), suppose that U ,= (otherwise there is nothing worthy
of proving). By 4.1.9, for x, y cl U there are nets (x

) and (y

) in U such that
x

x and y

y. If (, ) then x

+y

U. Appealing to 4.1.19 again,


infer x +y = lim(x

+y

) cl U.
5.3. Comparison Between Topologies
5.3.1. Definition. If Mand N are two multinorms on a vector space then M
is said to be ner or stronger than N (in symbols, M N) if U
M
U
N
. If M N
and N M simultaneously, then M and N are said to be equivalent (in symbols,
M N).
Multinormed and Banach Spaces 67
5.3.2. For multinorms Mand N on a vector space X the following statements
are equivalent:
(1) M N;
(2) the inclusion
M
(x)
N
(x) holds for all x X;
(3)
M
(0)
N
(0);
(4) ( q N) ( p
1
, . . . , p
n
M) (
1
, . . . ,
n
R
+
0)
B
q

1
B
p
1
. . .
n
B
p
n
;
(5) ( q N) ( p
1
, . . . , p
n
M) ( t > 0) q t(p
1
. . . p
n
) (with
respect to the order of the K-space R
X
).
The implications (1) (2) (3) (4) are evident.
(4) (5): Applying the Gauge Theorem (cf. 5.1.2), infer that
q p
B
p
1
/
1
. . . p
B
p
n
/
n
=
_
1
/

1
p
1
_
. . .
_
1
/

n
p
n
_

_
1
/

1
_
. . .
_
1
/

n
_
p
1
. . . p
n
.
(5) (1): It is sucient to check that M q for a seminorm q in N.
If V U
q
then V d
q
for some > 0. By hypothesis
d
q
d
p
1

/
t
. . . d
p
n

/
t

with suitable p
1
, . . . , p
n
Mand t > 0. The right side of this inclusion is an element
of U
p
1
. . . U
p
n
= U
p
1
,... ,p
n
]
U
M
. Hence, V is also a member of U
M
.
5.3.3. Definition. Let p, q : X R be two seminorms on X. Say that p is
ner or stronger than q and write p q whenever p q. The equivalence of
seminorms p q is understood in a routine fashion.
5.3.4. p q ( t > 0) q tp ( t 0) B
q
tB
p
;
p q ( t
1
, t
2
> 0) t
2
p q t
1
p ( t
1
, t
2
> 0) t
1
B
p
B
q
t
2
B
p
.
Everything follows from 5.3.2 and 5.1.2.
5.3.5. Riesz Theorem. If p, q : F
N
R are seminorms on the nite-dimen-
sional space F
N
then p q ker p ker q.
5.3.6. Corollary. All norms in nite dimensions are equivalent.
5.3.7. Let (X, M) and (Y, N) be multinormed spaces, and let T be a linear
operator, a member of L(X, Y ). The following statements are equivalent:
(1) N T M;
(2) T

(U
X
) U
Y
and T
1
(U
Y
) U
X
;
(3) x X T(
X
(x))
Y
(Tx);
(4) T(
X
(0))
Y
(0) and
X
(0) T
1
(
Y
(0));
(5) ( q N) ( p
1
, . . . , p
n
M) q T p
1
. . . p
n
.
68 Chapter 5
5.3.8. Let (X, | |
X
) and (Y, | |
Y
) be normed spaces and let T be a linear
operator, a member of L(X, Y ). The following statements are equivalent:
(1) T is bounded (that is, T B(X, Y ));
(2) | |
X
| |
Y
T;
(3) T is uniformly continuous;
(4) T is continuous;
(5) T is continuous at zero.
Each of the claims is a particular case of 5.3.7.
5.3.9. Remark. The message of 5.3.7 shows that it is sometimes convenient
to substitute for Ma multinorm equivalent to Mbut ltered upward (with respect
to the relation or ). For example, we may take the multinorm M:= sup M
0
:
M
0
is a nonempty nite subset of M. Observe that unltered multinorms should
be treated with due precaution.
5.3.10. Counterexample. Let X := F

; and let X
0
comprise all constant
functions; i.e., X
0
:= F1, where 1 : 1 ( ). Set M:= p

: ,
with p

(x) := [x()[ (x F

). It is clear that M is a multinorm on X. Now let


: X X/X
0
stand for the coset mapping. Undoubtedly, M

1 consists of the
sole element, zero. At the same time M

1 is separated.
5.3.11. Definition. Let (X, M) be a multinormed space and let X
0
be a sub-
space of X. The multinorm M

1, with : X X/X
0
the coset mapping,
is referred to as the quotient multinorm and is denoted by M
X/X
0
. The space
(X/X
0
, M
X/X
0
) is called the quotient space of X by X
0
.
5.3.12. The quotient space X/X
0
is separated if and only if X
0
is closed.
5.4. Metrizable and Normable Spaces
5.4.1. Definition. A multinormed space (X, M) is metrizable if there is
a metric d on X such that U
M
= U
d
. Say that X is normable if there is a norm
on X equivalent to the initial multinorm M. Say that X is countably normable
if there is a countable multinorm on X equivalent to the initial.
5.4.2. Metrizability Crtiterion. A multinormed space is metrizable if and
only if it is countably normable and separated.
: Let U
M
= U
d
. Passing if necessary to the multinorm M, assume that
for every n in N it is possible to nd a seminorm p
n
in M and t
n
> 0 such that
d
1
/
n
d
p
n
t
n
. Put N:= p
n
: n N. Clearly, M N. If V U
M
then V d
1
/
n
for some n N by the denition of metric uniformity. Hence,
by construction, V U
p
n
U
M
, i.e., M N. Thus, M N. The uniformity U
d
Multinormed and Banach Spaces 69
is separated, as indicated in 4.1.7. Applying 5.2.6, observe that U
M
and U
N
are
both separated.
: Passing if necessary to an equivalent multinorm, suppose that X, the space
in question, is countably normed and separated; that is, M:= p
n
: n N and
M is a Hausdor multinorm on X. Given x
1
, x
2
X, dene
d(x
1
, x
2
):=

k=1
1
2
k
p
k
(x
1
x
2
)
1 +p
k
(x
1
x
2
)
(the series on the right side of the above formula is dominated by the convergent
series

k=1
1
/
2
k, and so d is dened soundly).
Check that d is a metric. It suces to validate the triangle inequality. For
a start, put (t) := t(1 + t)
1
(t R
+
). It is evident that

(t) = (1 + t)
2
> 0.
Therefore, increases. Furthermore, is subadditive:
(t
1
+t
2
) = (t
1
+t
2
)(1 +t
1
+t
2
)
1
= t
1
(1 +t
1
+t
2
)
1
+t
2
(1 +t
1
+t
2
)
1
t
1
(1 +t
1
)
1
+t
2
(1 +t
2
)
1
= (t
1
) +(t
2
).
Thus, given x, y, z X, infer that
d(x, y) =

k=1
1
2
k
(p
k
(x y))

k=1
1
2
k
(p
k
(x z) +p
k
(z y))

k=1
1
2
k
((p
k
(x z)) +(p
k
(z y))) = d(x, z) +d(z, y).
It remains to established that U
d
and U
M
coincide.
First, show that U
d
U
M
. Take a cylinder, say, d ; and let (x, y)
d
p
1
t . . . d
p
n
t. Since is an increasing function, deduce that
d(x, y) =
n

k=1
1
2
k
p
k
(x y)
1 +p
k
(x y)
+

k=n+1
1
2
k
p
k
(x y)
1 +p
k
(x y)

t
1 +t
n

k=1
1
2
k
+

k=n+1
1
2
k

t
1 +t
+
1
2
n
.
Since t(1 +t)
1
+ 2
n
tends to zero as n and t 0, for appropriate t and n
observe that (x, y) d . Hence, d U
M
, which is required.
70 Chapter 5
Now establish that U
M
U
d
. To demonstrate the inclusion, given p
n
M
and t > 0, nd > 0 such that d
p
n
t d . For this purpose, take
:=
1
2
n
t
1 +t
,
which suces, since from the relations
1
2
n
p
n
(x y)
1 +p
n
(x y)
d(x, y) =
1
2
n
t
1 +t
holding for all x, y X it follows that p
n
(x y) t.
5.4.3. Definition. A subset V of a multinormed space (X, M) is a bounded
set in X if sup p(V ) < +for all p Mwhich means that the set p(V ) is bounded
above in R for every seminorm p in M.
5.4.4. For a set V in (X, M) the following statements are equivalent:
(1) V is bounded;
(2) for every sequence (x
n
)
nN
in V and every sequence (
n
)
nN
in F
such that
n
0, the sequence (
n
x
n
) vanishes:
n
x
n
0
(i.e., p(
n
x
n
) 0 for each seminorm p in M);
(3) every neighborhood of zero absorbs V .
(1) (2): p(
n
x
n
) [
n
[p(x
n
) [
n
[ sup p(V ) 0.
(2) (3): Let U
X
(0) and suppose that U fails to absorb V . From Denition
3.4.9, it follows that ( n N) ( x
n
V ) x
n
, nU. Thus,
1
/
n
x
n
, U for all n N;
i.e., (
1
/
n
x
n
) does not converge to zero.
(3) (1): Given p M, nd n N satisfying V nB
p
. Obviously, sup p(V )
sup p(nB
p
) = n < +.
5.4.5. Kolmogorov Normability Criterion. A multinormed space is normable
if and only if it is separated and has a bounded neighborhood about zero.
: It is clear.
: Let V be a bounded neighborhood of zero. Without loss of generality,
it may be assumed that V = B
p
for some seminorm p in the given multinorm M.
Undoubtedly, p M. Now if U
M
(0) then nU V for some n N. Hence,
U
p
(0). Using Theorem 5.3.2, observe that p M. Thus, p M; and, therefore,
p is also separated by 5.2.12. This means that p is a norm.
5.4.6. Remark. Incidentally, 5.4.5 shows that the presence of a bounded
neighborhood of zero in a multinormed space X amounts to the seminormabil-
ity of X.
Multinormed and Banach Spaces 71
5.5. Banach Spaces
5.5.1. Definition. A Banach space is a complete normed space.
5.5.2. Remark. The concept of Frechet space, complete metrizable multi-
normed space, serves as a natural abstraction of Banach space. It may be shown
that the class of Frechet spaces is the least among those containing all Banach
spaces and closed under the taking of countable products.
5.5.3. A normed space X is a Banach space if and only if every norm conver-
gent (= absolutely convergent) series in X converges.
: Let

n=1
|x
n
| < + for some (countable) sequence (x
n
). Then the
sequence of partial sums s
n
:= x
1
+ +x
n
is fundamental because
|s
m
s
k
| = |
m

n=k+1
x
n
|
m

n=k+1
|x
n
| 0
for m > k.
: Given a fundamental sequence (x
n
), choose an increasing sequence (n
k
)
kN
such that |x
n
x
m
| 2
k
as n, m n
k
. Then the series x
n
1
+(x
n
2
x
n
1
)+(x
n
3

x
n
2
) + converges in norm to some x; i.e., x
n
k
x. Observe simultaneously that
x
n
x.
5.5.4. If X is a Banach space and X
0
is a closed subspace of X then the
quotient space X/X
0
is also a Banach space.
Let : X := X/X be the coset mapping. Undoubtedly, for every x
there is some x
1
(x) such that 2|x| |x| |x|. Hence, given

n=1
x
n
,
a norm convergent series in , it is possible to choose x
n

1
(x
n
) so that the
norm series

n=1
|x
n
| be convergent. According to 5.5.3, the sum x:=

n=1
x
n
is available. If x:= (x) then
|x
n

k=1
x
k
| |x
n

k=1
x
k
| 0.
Appealing to 5.5.3 again, conclude that is a Banach space.
5.5.5. Remark. The claim of 5.5.3 may be naturally translated to seminormed
spaces. In particular, if (X, p) is a complete seminormed space then the quotient
space X/ ker p is a Banach space.
5.5.6. Theorem. If X and Y are normed spaces and X ,= 0 then B(X, Y ),
the space of bounded operators, is a Banach space if and only if so is Y .
72 Chapter 5
: Consider a Cauchy sequence (T
n
) in B(X, Y ). By the normative inequal-
ity, |T
m
x T
k
x| |T
m
T
k
| |x| 0 for all x X; i.e., (T
n
x) is fundamental
in Y . Thus, there is a limit Tx := limT
n
x. Plainly, the so-dened operator T is
linear. By virtue of the estimate [|T
m
| |T
k
|[ |T
m
T
k
| the sequence (|T
n
|) is
fundamental in R and, hence, bounded; that is, sup
n
|T
n
| < +. Therefore, pass-
ing to the limit in |T
n
x| sup
n
|T
n
| |x|, obtain |T| < +. It remains to check
that |T
n
T| 0. Given > 0, choose a number n
0
such that |T
m
T
n
|

/
2
as m, n n
0
. Further, for x B
X
nd m n
0
satisfying |T
m
x Tx|

/
2
.
Then |T
n
x Tx| |T
n
x T
m
x| +|T
m
x Tx| |T
n
T
m
| +|T
m
x Tx|
as n n
0
. In other words, |T
n
T| = sup|T
n
x Tx| : x B
X
for all
suciently large n.
: Let (y
n
) be a Cauchy sequence in Y . By hypothesis, there is a norm-one
element x in X; i.e., x has norm 1: |x| = 1. Applying 3.5.6 and 3.5.2 (1), nd
an element x

[[(| |) satisfying (x, x

) = |x| = 1. Obviously, the rank-


one operator (with range of dimension 1) T
n
:= x

y
n
: x (x, x

)y
n
belongs
to B(X, Y ), since |T
n
| = |x

| |y
n
|. Hence, |T
m
T
k
| = |x

(y
m
y
k
)|=
|x

| |y
m
y
k
| = |y
m
y
k
|, i.e., (T
n
) is fundamental in B(X, Y ). Assign T :=
limT
n
. Then |Tx T
n
x| = |Tx y
n
| |T T
n
| |x| 0. In other words, Tx is
the limit of (y
n
) in Y .
5.5.7. Corollary. The dual of a normed space (furnished with the dual norm)
is a Banach space.
5.5.8. Corollary. Let X be a normed space; and let : X X

, the double
prime mapping, be the canonical embedding of X into the second dual X

. Then
the closure cl (X) is a completion of X.
By virtue of 5.5.7, X

is a Banach space. By 5.1.10 (8), is an isometry


from X into X

. Appealing to 4.5.16 ends the proof.


5.5.9. Examples.
(1) Abstract examples: a basic eld, a closed subspace of a Banach
space, the product of Banach spaces, and 5.5.45.5.8.
(2) Let E be a nonempty set. Given x F
E
, put |x|

:= sup [x(E )[.


The space l

(E ):= l

(E , F):= dom| |

is called the space of bounded functions


on E . The designations B(E ) and B(E , F) are also used. For E := N, it is
customary to put m:= l

:= l

(E ).
(3) Let a set E be innite, i.e. not nite, and let F stand for a lter
on E . By denition, x c(E , F) (x l

(E ) and x(F) is a Cauchy lter on F).


In the case E := N and F is the nite complement lter (comprising all conite sets
each of which is the complement of a nite subset) of N, the notation c:= c(E , F)
is employed, and c is called the space of convergent sequences. In c(E , F) the
Multinormed and Banach Spaces 73
subspace c
0
(E , F) := x c(E , F) : x(F) 0 is distinguished. If F is
the nite complement lter then the shorter notation c
0
(E ) is used and we speak
of the space of functions vanishing at innity. Given E := N, write c
0
:= c
0
(E ).
The space c
0
is referred to as the space of vanishing sequences. It is worth keeping
in mind that each of these spaces without further specication is endowed with the
norm taken from the corresponding space l

(E , F).
(4) Let S:= (E , X,
_
) be a system with integration. This means that
X is a vector sublattice of R
E
, with the lattice operations in X coincident with
those in R
E
, and
_
: X R is a (pre)integral; i.e.,
_
X
#
+
and
_
x
n
0 whenever
x
n
X and x
n
(e) 0 for e E . Moreover, let f F
E
be a measurable mapping
(with respect to S) (as usual, we may speak of almost everywhere nite and almost
everywhere dened measurable functions).
Denote N
p
(f) := (
_
[f[
p
)
1
/
p
for p 1, where
_
is the corresponding Lebesgue
extension of the initial integral
_
. (The traditional liberty is taken of using the
same symbol for the original and its successor.)
An element of domN
1
is an integrable or summable function. The integrability
of f F
E
is equivalent to the integrability of its real part Re f and imaginary part
Imf, both members of R
E
. For the sake of completeness, recall the denition
N(g):= inf
_
sup
_
x
n
: (x
n
) X, x
n
x
n+1
, ( e E ) [g(e)[ = lim
n
x
n
(e)
_
for an arbitrary g in F
E
. If F = R then domN
1
obviously presents the closure of X
in the normed space (domN, N).
The Holder inequality is valid:
N
1
(fg) N
p
(f)N
p
(g),
with p

the conjugate exponent of p, i.e.


1
/
p
+
1
/
p
= 1 for p > 1.
This is a consequence of the Young inequality, xy
x
p
/
p

y
p

/
p
for all
x, y R
+
, applied to [f[/N
p
(f) and [g[/N
p
(g) when N
p
(f) and N
p
(g) are both
nonzero. If N
p
(f)N
p
(g) = 0 then the Holder inequality is beyond a doubt.
The set L
p
:= domN
p
is a vector space.
[f +g[
p
([f[ +[g[)
p
2
p
([f[ [g[)
p
= 2
p
([f[
p
[g[
p
) 2
p
([f[
p
+[g[
p
)
The function N
p
is a seminorm, satisfying the Minkowski inequality:
N
p
(f +g) N
p
(f) +N
p
(g).
74 Chapter 5
For p = 1, this is trivial. For p > 1 the Minkowski inequality follows from
the presentation
N
p
(f) = supN
1
(fg)/N
p
(g) : 0 < N
p
(g) < + (f L
p
)
whose right side is the upper envelope of a family of seminorms. To prove the above
presentation, using the Holder inequality, note that g := [f[
p
/
p

lies in L
q
when
N
p
(f) > 0; furthermore, N
p
(f) = N
1
(fg)/N
p
(g). Indeed, N
1
(fg) =
_
[f[
p
/
p

+1
=
N
p
(f)
p
, because
p
/
p
+1 = p
_
1
1
/
p
_
+1 = p. Continue arguing to nd N
p
(g)
p

=
_
[g[
p

=
_
[f[
p
= N
p
(f)
p
, and so N
p
(g) = N
p
(f)
p
/
p

. Finally,
N
1
(fg)/N
p
(g) = N
p
(f)
p
/N
p
(f)
p
/
p

= N
p
(f)
p
p
/
p

= N
p
(f)
p(1
1
/
p
)
= N
p
(f).
The quotient space L
p
/ ker N
p
, together with the corresponding quotient norm
| |
p
, is called the space of p-summable functions or L
p
space with more complete
designations L
p
(S), L
p
(E , X,
_
), etc.
Finally, if a system with integration S arises from inspection of measurable step
functions on a measure space (, A, ) then it is customary to write L
p
(, A, ),
L
p
(, ) and even L
p
(), with the unspecied parameters clear from the context.
RieszFisher Completeness Theorem. Each L
p
space is a Banach space.
We sketch out the proof. Consider t :=

k=1
N
p
(f
k
), where f
k
L
p
. Put

n
:=

n
k=1
f
k
and s
n
:=

n
k=1
[f
k
[. It is seen that (s
n
) has positive entries and
increases. The same is true of (s
p
n
). Furthermore,
_
s
p
n
t
p
< +. Hence, by the
Levy Monotone Convergence Theorem, for almost all e E there exists a limit
g(e) := lims
p
n
(e), with the resulting function g a member of L
1
. Putting h(e) :=
g
1
/
p
(e), observe that h L
p
and s
n
(e) h(e) for almost all e E . The inequalities
[
n
[ s
n
h imply that for almost all e E the series

k=1
f
k
(e) converges. For
the sum f
0
(e) the estimate holds: [f
0
(e)[ h(e). Hence, it may be assumed that
f
0
L
p
. Appealing to the Lebesgue Dominated Convergence Theorem, conclude
that N
p
(
n
f
0
) =
__
[
n
f
0
[
p
_
1
/
p
0. Thus, in the seminormed space under
consideration each seminorm convergent series converges. To complete the proof,
apply 5.5.35.5.5.
If S is the system of conventional summation on E ; i.e., X :=

eE
R is the
direct sum of suitably many copies of the ground eld R and
_
x:=

eE
x(e), then
L
p
comprises all p-summable families. This space is denoted by l
p
(E ). Further,
|x|
p
:=
_
eE
[x(e)[
p
_
1
/
p
. In the case E := N the notation l
p
is used and l
p
is
referred to as the space of p-summable sequences.
(5) Dene L

as follows: Let X be an ordered vector space and let


e X
+
be a positive element. The seminorm p
e
associated with e is the Minkowski
functional of the order interval [e, e], i.e.,
p
e
(x):= inft > 0 : te x te.
Multinormed and Banach Spaces 75
The eective domain of denition of p
e
is the space of bounded elements (with
respect to e); the element e itself is referred to as the strong order-unit in X
e
.
An element of ker p
e
is said to be nonarchimedean (with respect to e). The quotient
space X
e
/ ker p
e
furnished with the corresponding quotient seminorm is called the
normed space of bounded elements (generated by e in X). For example, C(Q, R),
the space of continuous real-valued functions on a nonempty compact set Q, presents
the normed space of bounded elements with respect to 1 := 1
Q
: q 1 (q Q)
(in itself). In R
E
the same element 1 generates the space l

(E ).
Given a system with integration S:=
_
E , X,
_ _
, assume that 1 is measurable
and consider the space of functions acting from E into F and satisfying
N

(f):= inft > 0 : [f[ t1 < +,


where means less almost everywhere than. This space is called the space
of essentially bounded functions and is labelled with L

. To denote the quotient


space L

/ ker N

and its norm the symbols L

and | |

are in use.
It is in common parlance to call the elements of L

(like the elements of L

)
essentially bounded functions. The space L

presents a Banach space.


The space L

, as well as the spaces C(Q, F), l


p
(E ), c
0
(E ), c, l
p
, and L
p
(p
1), also bears the unifying title classical Banach space. Nowadays a Lindenstrauss
space which is a space whose dual is isometric to L
1
(with respect to some system
with integration) is also regarded as classical. It can be shown that a Banach
space X is classical if and only if the dual X

is isomorphic to one of the L


p
spaces
with p 1.
(6) Consider a system with integration S :=
_
E , X,
_ _
and let p 1.
Suppose that for every e in E there is a Banach space (Y
e
, | |
Y
e
). Given an arbi-
trary element f in

eE
Y
e
, dene [[[f[[[ : e |f(e)|
Y
e
. Put N
p
(f) := infN
p
(g) :
g L
p
, g [[[f[[[. It is clear that domN
p
is a vector space equipped with the
seminorm N
p
. The sum of the family in the sense of L
p
or simply the p-sum
of (Y
e
)
eE
(with respect to the system with integration S) is the quotient space
domN
p
/ ker N
p
under the corresponding (quotient) norm [[[[[[
p
.
The p-sum of a family of Banach spaces is a Banach space.
If

k=1
N
p
(f
k
) < + then the sequence (s
n
:=

n
k=1
[[[f
k
[[[) tends to some
almost everywhere nite positive function g and N
p
(g) < +. It follows that for
almost all e E the sequence (s
n
(e)) (i.e., the series

k=1
|f
k
(e)|
Y
e
) converges.
By the completeness of Y
e
, the series

k=1
f
k
(e) converges to some sum f
0
(e) in Y
e
with |f
0
(e)|
Y
e
g(e) for almost every e E . Therefore, it may be assumed that
f
0
domN
p
. Finally, N
p
(

n
k=1
f
k
f
0
)

k=n+1
N
p
(f
k
) 0.
In the case when E := N with conventional summation, for the sum Y of a se-
quence of Banach spaces (Y
n
)
nN
(in the sense of L
p
) the following notation is often
employed:
Y:= (Y
1
Y
2
)
p
,
76 Chapter 5
with p the type of summation. An element y in Y presents a sequence (y
n
)
nN
such
that y
n
Y
n
and
[[[y[[[
p
:=
_

k=1
|y
n
|
p
Y
n
_
1
/
p
< +.
In the case Y
e
:= X for all e E , where X is some Banach space over F, put F
p
:=
domN
p
and F
p
:= F
p
/ ker N
p
. An element of the so-constructed space is a vector
eld or a X-valued function on E (having a p-summable norm). Undoubtedly, F
p
is a Banach space. At the same time, if the initial system with integration contains
a nonmeasurable set then extraordinary elements are plentiful in F
p
(in particular,
for the usual Lebesgue system with integration F
p
,= L
p
). In this connection the
functions with nite range, assuming each value on a measurable set, are selected
in F
p
. Such a function, as well as the corresponding coset in F
p
, is a simple, nite-
valued or step function. The closure in F
p
of the set of simple functions is denoted
by L
p
(or more completely L
p
(X), L
p
(S, X), L
p
(, A, ), L
p
(, ), etc.) and is
the space of X-valued p-summable functions. Evidently, L
p
(X) is a Banach space.
It is in order to illustrate one of the advantages of these spaces for p = 1. First,
notice that a simple function f can be written as a nite combination of character-
istic functions:
f =

ximf

f
1
(x)
x,
where f
1
(x) is a measurable set as x imf, with
E
(e) = 1 for e E and

E
(e) = 0 otherwise. Moreover,
_
[[[f[[[ =
_

ximf
|
f
1
(x)
x|
=
_

ximf

f
1
(x)
|x| =

ximf
|x|
_

f
1
(x)
< +.
Next, associate with each simple function f some element in X by the rule
_
f :=

ximf
_

f
1
(x)
x.
Straightforward calculation shows that the integral
_
dened on the subspace of sim-
ple functions is linear. Furthermore, it is bounded because
|
_
f| = |

ximf
_

f
1
(x)
x|

ximf
_

f
1
(x)
|x|
=
_

ximf
|x|
f
1
(x)
=
_
[[[f[[[.
Multinormed and Banach Spaces 77
By virtue of 4.5.10 and 5.3.8, the operator
_
has a unique extension to an element
of B(L
1
(X), X). This element is denoted by the same symbol,
_
(or
_
E
, etc.), and
is referred to as the Bochner integral.
(7) In the case of conventional summation, the usage of the scalar the-
ory is preserved for the Bochner integral. Namely, the common parlance favours
the term sum of a family rather than integral of a summable function, and the
symbols pertinent to summation are perfectly welcome. What is more important,
innite dimensions bring about signicant complications.
Let (x
n
) be a family of elements of a Banach space. Its summability (in the
sense of the Bochner integral) means the summability of the numeric family (|x
n
|),
i.e. the norm convergence of (x
n
) as a series. Consequently, (x
n
) has at most
countably many nonzero elements and may thus be treated as a (countable) se-
quence. Moreover,

n=1
|x
n
| < +; i.e. the series x
1
+ x
2
+ converges in
norm. By 5.5.3, for the series sum x =

n=1
x
n
observe that x = lim

, where
s

:=

n
x
n
is a partial sum and ranges over the direction of all nite sub-
sets of N. In this case, the resulting x is sometimes called the unordered sum of
(x
n
), whereas the sequence (x
n
) is called unconditionally or unorderly summable
to x (in symbols, x =

nN
x
n
). Using these terms, observe that summability in
norm implies unconditional summability (to the same sum). If dimX < + then
the converse holds which is the Riemann Theorem on Series. The general case is
explained by the following deep and profound assertion:
DvoretzkyRogers Theorem. In an arbitrary innite-dimensional Banach
space X, for every sequence of positive numbers (t
n
) such that

n=1
t
2
n
< +
there is an unconditionally summable sequence of elements (x
n
) with |x
n
| = t
n
for
all n N.
In this regard for a family of elements (x
e
)
eE
of an arbitrary multinormed
space (X, M) the following terminology is accepted: Say that (x
e
)
eE
is summable
or unconditionally summable (to a sum x) and write x :=

eE
x
e
whenever x is
the limit in (X, M) of the corresponding net of partial sums s

, with a nite
subset of E ; i.e., s

x in (X, M). If for every p there is a sum

eE
p(x
e
)
then the family (x
e
)
eE
is said to be multinorm summable (or, what is more exact,
fundamentally summable, or even absolutely fundamental).
In conclusion, consider a Banach space Y and T B(X, Y). The operator T
can be uniquely extended to an operator from L
1
(X) into L
1
(Y) by putting Tf :
e Tf(e) (e E ) for an arbitrary simple X-valued function f. Then, given
f L
1
(X), observe that Tf L
1
(Y) and
_
E
Tf = T
_
E
f. This fact is verbalized
as follows: The Bochner integral commutes with every bounded operator.
5.6. The Algebra of Bounded Operators
5.6.1. Let X, Y , and Z be normed spaces. If T L(X, Y ) and S L(Y, Z)
78 Chapter 5
are linear operators then |ST| |S| |T|; i.e., the operator norm is submultiplica-
tive.
Given x X and using the normative inequality twice, infer that
|STx| |S| |Tx| |S| |T| |x|.
5.6.2. Remark. In algebra, in particular, (associative) algebras are studied.
An algebra (over F) is a vector space A (over F) together with some associative
multiplication : (a, b) ab (a, b A). This multiplication must be distributive
with respect to addition (i.e., (A, +, ) is an (associative) ring) and, moreover, the
operation must agree with scalar multiplication in the following sense: (ab) =
(a)b = a(b) for all a, b A and F. A displayed notation for an algebra is
(A, F, +, , ). However, just like on the other occasions it is customary to use
the term algebra simply for A.
5.6.3. Definition. A normed algebra (over a ground eld) is an associative
algebra (over this eld) together with a submultiplicative norm. A Banach algebra
is a complete normed algebra.
5.6.4. Let B(X) := B(X, X) be the space of bounded endomorphisms of
a normed space X. The space B(X) with the operator norm and composition
as multiplication presents a normed algebra. If X ,= 0 then B(X) has a neutral
element (with respect to multiplication), the identity operator I
X
; i.e., B(X) is
an algebra with unity. Moreover, |I
X
| = 1. The algebra B(X) is a Banach algebra
if and only if X is a Banach space.
If X = 0 then there is nothing left to proof. Given X ,= 0, apply 5.5.6.
5.6.5. Remark. It is usual to refer to B(X) as the algebra of bounded opera-
tors in X or even as the (bounded) endomorphism algebra of X. In connection with
5.6.4, given F, it is convenient to retain the same symbol for I
X
. (In partic-
ular, 1 = I
0
= 0!) For X ,= 0 this procedure may be thought as identication of F
with FI
X
.
5.6.6. Definition. If X is a normed space and T B(X) then the spectral
radius of T is the number r(T) := inf
_
|T
n
|
1
/
n
: n N
_
. (The rationale of this
term will transpire later (cf. 8.1.12).)
5.6.7. The norm of T is greater than the spectral radius of T.
Indeed, by 5.6.1, the inequality |T
n
| |T|
n
is valid.
5.6.8. The Gelfand formula holds:
r(T) = lim
n
_
|T
n
|.
Multinormed and Banach Spaces 79
Take > 0 and let s N satisfy |T
s
|
1
/
s
r(T)+. Given n N with n s,
observe the presentation n = k(n)s+l(n) with k(n), l(n) N and 0 l(n) s1.
Hence,
|T
n
| = |T
k(n)s
T
l(n)
| |T
s
|
k(n)
|T
l(n)
|

_
1 |T| . . . |T
s1
|
_
|T
s
|
k(n)
= M|T
s
|
k(n)
.
Consequently,
r(T) |T
n
|
1
/
n
M
1
/
n
|T
s
|
k(n)
/
n
M
1
/
n
(r(T) +)
k(n)s
/
n
= M
1
/
n
(r(T) +)
(nl(n))
/
n
.
Since M
1
/
n
1 and
(nl(n))
/
n
1, nd r(T) limsup|T
n
|
1
/
n
r(T) + .
The inequality liminf |T
n
|
1
/
n
r(T) is evident. Recall that is arbitrary, thus
completing the proof.
5.6.9. Neumann Series Expansion Theorem. With X a Banach space
and T B(X), the following statements are equivalent:
(1) the Neumann series 1+T +T
2
+ converges in the operator norm
of B(X);
(2) |T
k
| < 1 for some k and N;
(3) r(T) < 1.
If one of the conditions (1)(3) holds then

k=0
T
k
= (1 T)
1
.
(1) (2): With the Neumann series convergent, the general term (T
k
) tends
to zero.
(2) (3): This is evident.
(3) (1): According to 5.6.8, given a suitable > 0 and a suciently large
k N, observe that r(T) |T
k
|
1
/
k
r(T) + < 1. In other words, some tail
of the series

k=0
|T
k
| is dominated by a convergent series. The completeness
of B(X) and 5.5.3 imply that

k=0
T
k
converges in B(X).
Now let S:=

k=0
T
k
and S
n
:=

n
k=0
T
k
. Then
S(1 T) = limS
n
(1 T) = lim(1 +T + +T
n
) (1 T) = lim(1 T
n+1
) = 1;
(1 T)S = lim(1 T)S
n
= lim(1 T)(1 +T + +T
n
) = lim
_
1 T
n+1
_
= 1,
because T
n
0. Thus, by 2.2.7 S = (1 T)
1
.
80 Chapter 5
5.6.10. Corollary. If |T| < 1 then (1 T) is invertible (= has a bounded
inverse); i.e., the inverse correspondence (1 T)
1
is a bounded linear operator.
Moreover, |(1 T)
1
| (1 |T|)
1
.
The Neumann series converges and
|(1 T)
1
|

k=0
|T
k
|

k=0
|T|
k
= (1 |T|)
1
.
5.6.11. Corollary. If |1 T| < 1 then T is invertible and
|1 T
1
|
|1 T|
1 |1 T|
.
By Theorem 5.6.9,
1 +

k=1
(1 T)
k
=

k=0
(1 T)
k
= (1 (1 T))
1
= T
1
.
Hence,
|T
1
1| = |

k=1
(1 T)
k
|

k=1
|(1 T)
k
|

k=1
|1 T|
k
.
5.6.12. Banach Inversion Stability Theorem. Let X and Y be Banach
spaces. The set of invertible operators Inv(X, Y ) is open. Moreover, the inversion
T T
1
acting from Inv(X, Y ) to Inv(Y, X) is continuous.
Take S, T B(X, Y ) such that T
1
B(Y, X) and |T
1
| |S T|
1
/
2
.
Consider the operator T
1
S in B(X). Observe that
|1 T
1
S| = |T
1
T T
1
S| |T
1
| |T S|
1
/
2
< 1.
Hence, by 5.6.11, (T
1
S)
1
belongs to B(X). Put R := (T
1
S)
1
T
1
. Clearly,
R B(Y, X) and, moreover, R = S
1
(T
1
)
1
T
1
= S
1
. Further,
|S
1
| |T
1
| |S
1
T
1
|
= |S
1
(T S)T
1
| |S
1
| |T S| |T
1
|
1
/
2
|S
1
|.
This implies |S
1
| 2|T
1
|, yielding the inequalities
|S
1
T
1
| |S
1
| |T S| |T
1
| 2|T
1
|
2
|T S|.
5.6.13. Definition. If X is a Banach space over F and T B(X) then a scalar
F is a regular or resolvent value of T whenever ( T)
1
B(X). In this case
put R(T, ):= (T)
1
and say that R(T, ) is the resolvent of T at . The set
of the resolvent values of T is denoted by res(T) and called the resolvent set of T.
The mapping R(T, ) from res(T) into B(X) is naturally called the resolvent
of T. The set Fres(T) is referred to as the spectrum of T and is denoted by Sp(T)
or (T). A member of Sp(T) is said to be a spectral value of T (which is enigmatic
for the time being).
Multinormed and Banach Spaces 81
5.6.14. Remark. If X = 0 then the spectrum of the only operator T = 0
in B(X) is the empty set. In this regard, in spectral analysis it is silently presumed
that X ,= 0. In the case X ,= 0 for F:= R the spectra of some operators can also
be void, whereas for F:= C it is impossible (cf. 8.1.11).
5.6.15. The set res(T) is open. If
0
res(T) then
R(T, ) =

k=0
(1)
k
(
0
)
k
R(T,
0
)
k+1
in some neighborhood of
0
. If [[ > |T| then res(T) and the expansion
R(T, ) =
1

k=0
T
k

k
holds. Moreover, |R(T, )| 0 as [[ +.
Since |(T)(
0
T)| = [
0
[, the openness property of res(T) follows
from 5.6.12. Proceed along the lines
T = (
0
) + (
0
T) = (
0
T)R(T,
0
)(
0
) + (
0
T)
= (
0
T)((
0
)R(T,
0
) + 1) = (
0
T)(1 ((1)(
0
)R(T,
0
))).
In a suitable neighborhood of
0
, from 5.6.9 derive
R(T, ) = ( T)
1
= (1 ((1)(
0
)R(T,
0
)))
1
(
0
T)
1
=

k=0
(1)
k
(
0
)
k
R(T,
0
)
k+1
.
According to 5.6.9, for [[ > |T| there is an operator
_
1
T
/

_
1
presenting
the sum of the Neumann series; i.e.,
R(T, ) =
1

_
1
T
/

_
1
=
1

k=0
T
k

k
.
It is clear that
|R(T, )|
1
[[

1
1
|T|
/
]]
.
5.6.16. The spectrum of every operator is compact.
82 Chapter 5
5.6.17. Remark. It is worth keeping in mind that the inequality [[ > r(T)
is a necessary and sucient condition for the convergence of the Laurent series,
R(T, ) =

k=0
T
k
/
k+1
, which expands the resolvent of T at innity.
5.6.18. An operator S commutes with an operator T if and only if S commutes
with the resolvent of T.
: ST = TS S( T) = S ST = S TS = ( T)S
R(T, )S( T) = S R(T, )S = S R(T, ) ( res(T)).
: SR(T,
0
) = R(T,
0
)S S = R(T,
0
)S(
0
T) (
0
T)S =
S(
0
T) TS = ST.
5.6.19. If , res(T) then the rst resolvent equation, the Hilbert identity,
holds:
R(T, ) R(T, ) = ( )R(T, )R(T, ).
Multiplying the equality = ( T) ( T), rst, by R(T, )
from the right and, second, by R(T, ) from the left, successively infer the sought
identity.
5.6.20. If , res(T) then R(T, )R(T, ) = R(T, )R(T, ).
5.6.21. For res(T) the equality holds:
d
k
d
k
R(T, ) = (1)
k
k! R(T, )
k+1
.
5.6.22. Composition Spectrum Theorem. The spectra Sp(ST) and
Sp(TS) may dier only by zero.
It suces to establish that 1 , Sp(ST) 1 , Sp(TS). Indeed, from
, Sp(ST) and ,= 0 it will follow that
1 ,
1
/

Sp(ST) 1 , Sp
_
1
/

ST
_
1 , Sp
_
1
/

TS
_
, Sp(TS).
Therefore, consider the case 1 , Sp(ST). The formal Neumann series expan-
sions
(1 ST)
1
1 +ST + (ST)(ST) + (ST)(ST)(ST) + ,
T(1 ST)
1
S TS +TSTS +TSTSTS + (1 TS)
1
1
lead to conjecturing that the presentation is valid:
(1 TS)
1
= 1 +T(1 ST)
1
S
Multinormed and Banach Spaces 83
(which in turn means 1 , Sp(TS)). Straightforward calculation demonstrates the
above presentation, thus completing the entire proof:
(1 +T(1 ST)
1
S)(1 TS) = 1 +T(1 ST)
1
S TS +T(1 ST)
1
(ST)S
= 1 +T(1 ST)
1
S TS +T(1 ST)
1
(1 ST 1)S
= 1 +T(1 ST)
1
S TS +TS T(1 ST)
1
S = 1;
(1 TS)(1 +T(1 ST)
1
S) = 1 TS +T(1 ST)
1
S +T(ST)(1 ST)
1
S
= 1 TS +T(1 ST)
1
S +T(1 ST 1)(1 ST)
1
S
= 1 TS +T(1 ST)
1
S +TS T(1 ST)
1
S = 1.
Exercises
5.1. Prove that a normed space is nite-dimensional if and only if every linear functional
on the space is bounded.
5.2. Demonstrate that it is possible to introduce a norm into each vector space.
5.3. Show that a vector space X is nite-dimensional if and only if all norms on X are
equivalent to each other.
5.4. Demonstrate that all separated multinorms introduce the same topology in a nite-
dimensional space.
5.5. Each norm on R
N
is appropriate for norming the product of nitely many normed
spaces, isnt it?
5.6. Find conditions for continuity of an operator acting between multinormed spaces and
having nite-dimensional range.
5.7. Describe the operator norms in the space of square matrices. When are such norms
comparable?
5.8. Calculate the distance between hyperplanes in a normed space.
5.9. Find the general form of a continuous linear functional on a classical Banach space.
5.10. Study the question of reexivity for classical Banach spaces.
5.11. Find the mutual disposition of the spaces l
p
and l
p
as well as L
p
and L
p
. When is
the complement of one element of every pair is dense in the other?
5.12. Find the spectrum and resolvent of the Volterra operator (the taking of a primitive),
a projection, and a rank-one operator.
5.13. Construct an operator whose spectrum is a prescribed nonempty compact set in C.
5.14. Prove that the identity operator (in a nonzero space) is never the commutator of any
pair of operators.
5.15. Is it possible to dene some reasonable spectrum for an operator in a multinormed
space?
5.16. Does every Banach space over F admit an isometric embedding into the space C(Q, F),
with Q a compact space?
84 Chapter 5
5.17. Find out when L
p
(X)

= L
p
(X

), with X a Banach space.


5.18. Let (X
n
) be a sequence of normed spaces and let
X
0
=
_
x

nN
X
n
: |x
n
| 0
_
be their c
0
-sum (with the norm |x| = sup |x
n
| : n N induced from the l

-sum). Prove that


X
0
is separable if and only if so is each of the spaces X
n
.
5.19. Prove that the space C
(p)
[0, 1] presents the sum of a nite-dimensional subspace and
a space isomorphic to C[0, 1].
Chapter 6
Hilbert Spaces
6.1. Hermitian Forms and Inner Products
1. Definition. Let H be a vector space over a basic eld F. A mapping
f : H
2
F is a hermitian form on H provided that
(1) the mapping f( , y) : x f(x, y) belongs to H
#
for every y in Y ;
(2) f(x, y) = f(y, x)

for all x, y H, where

is the natural
involution in F; that is, the taking of the complex conjugate of a complex number.
2. Remark. It is easy to see that, for a hermitian form f and each x in H the
mapping f(x, ) : y (x, y) lies in H
#

, where H

is the twin of H (see 2.1.4 (2)).


Consequently, in case F := R every hermitian form is bilinear, i.e., linear in each
argument; and in case F := C, sesquilinear, i.e., linear in the rst argument and
-linear in the second.
3. Every hermitian form f satises the polarization identity:
f(x +y, x +y) f(x y, x y) = 4 Re f(x, y) (x, y H).
f(x +y, x +y) = f(x, x) +f(x, y) +f(y, x) +f(y, y)

f(x y, x y) = f(x, x) f(x, y) f(y, x) +f(y, y)


2(f(x, y) +f(y, x))
4. Definition. A hermitian formf is positive or positive semidenite provided
that f(x, x) 0 for all x H. In this event, write (x, y):= x [ y:= f(x, y) (x, y
H). A positive hermitian form is usually referred to as a semi-inner product on
H. A semi-inner product on H is an inner product or a (positive denite) scalar
product whenever (x, x) = 0 x = 0 with x H.
5. The CauchyBunyakovskiSchwarz inequality holds:
[(x, y)[
2
(x, x)(y, y) (x, y H).
86 Chapter 6
If (x, x) = (y, y) = 0 then 0 (x+ty, x+ty) = t(x, y)

+t

(x, y). Letting


t := (x, y), nd that 2[(x, y)[
2
0; i.e., in this case the claim is established.
If, for deniteness, (y, y) ,= 0; then in view of the estimate
0 (x +ty, x +ty) = (x, x) + 2t Re(x, y) +t
2
(y, y) (t R)
conclude that Re(x, y)
2
(x, x)(y, y).
If (x, y) = 0 then nothing is left to proof. If (x, y) ,= 0 then let :=
[(x, y)[ (x, y)
1
and x:= x. Now [[ = 1 and, furthermore,
(x, x) = (x, x) =

(x, x) = [[
2
(x, x) = (x, x);
[(x, y)[ = (x, y) = (x, y) = (x, y) = Re(x, y).
Consequently, [(x, y)[
2
= Re(x, y)
2
(x, x)(y, y).
6. If ( , ) is a semi-inner product on H then the mapping || : x (x, x)
1
/
2
is a seminorm on H.
It suces to prove the triangle inequality. Applying the CauchyBunyakov-
skiSchwarz inequality, observe that
|x +y|
2
= (x, x) + (y, y) + 2 Re(x, y)
(x, x) + (y, y) + 2|x| |y| = (|x| +|y|)
2
.
7. Definition. A space H endowed with a semi-inner product ( , ) and the
associate seminorm | | is a pre-Hilbert space. A pre-Hilbert space H is a Hilbert
space provided that the seminormed space (H, | |) is a Banach space.
8. In a pre-Hilbert space H, the Parallelogram Law is eective
|x +y|
2
+|x y|
2
= 2(|x|
2
+|y|
2
) (x, y H)
which reads: the sum of the squares of the lengths of the diagonals equals the sum
of the squares of the lengths of all sides.
|x +y|
2
= (x +y, x +y) = |x|
2
+ 2 Re(x, y) +|y|
2
;
|x y|
2
= (x y, x y) = |x|
2
2 Re(x, y) +|y|
2

9. Von NeumannJordan Theorem. If a seminormed space (H, | |) obeys


the Parallelogram Law then H is a pre-Hilbert space; i.e., there is a unique semi-
inner product ( , ) on H such that |x| = (x, x)
1
/
2
for all x H.
Hilbert Spaces 87
Considering the real carrier H
R
of H and x, y H
R
, put
(x, y)
R
:=
1
/
4
_
|x +y|
2
|x y|
2
_
.
Given the mapping ( , y)
R
, from the Parallelogram Law successively derive
(x
1
, y)
R
+ (x
2
, y)
R
=
1
/
4
_
|x
1
+y|
2
|x
1
y|
2
+|x
2
+y|
2
|x
2
y|
2
_
=
1
/
4
__
|x
1
+y|
2
+|x
2
+y|
2
_

_
|x
1
y|
2
+|x
2
y|
2
__
=
1
/
4
_
1
/
2
(|(x
1
+y) + (x
2
+y)|
2
+|x
1
x
2
|
2
_

1
/
2
_
|(x
1
y) + (x
2
y)|
2
+|x
1
x
2
|
2
)
_
=
1
/
4
_
1
/
2
|x
1
+x
2
+ 2y|
2

1
/
2
|x
1
+x
2
2y|
2
_
=
1
/
2
_
_
_
_
(x
1
+x
2
)
/
2
+y
_
_
_
2

_
_
_
(x
1
x
2
)
/
2
y
_
_
_
2
_
= 2
_
(x
1
+x
2
)
/
2
, y
_
R
.
In particular, (x
2
, y)
R
= 0 in case x
2
:= 0, i.e.
1
/
2
(x
1
, y)
R
=
_
1
/
2
x
1
, y
_
R
. Analo-
gously, given x
1
:= 2x
1
and x
2
:= 2x
2
, infer that
(x
1
+x
2
, y)
R
= (x
1
, y)
R
+ (x
2
, y)
R
.
Since the mapping ( , y)
R
is continuous for obvious reasons, conclude that
( , y)
R
(H
R
)
#
. Put
(x, y):= (Re
1
( , y)
R
)(x),
where Re
1
is the complexier (see 3.7.5).
In case F:= R it is clear that (x, y) = (x, y)
R
= (y, x) and (x, x) = |x|
2
;
i.e., nothing is to be proven. On the other hand, if F:= C then
(x, y) = (x, y)
R
i(ix, y)
R
.
This entails sesquilinearity:
(y, x) = (y, x)
R
i(iy, x)
R
= (x, y)
R
i(x, iy)
R
= (x, y)
R
+i(ix, y)
R
= (x, y)

,
since
(x, iy)
R
=
1
/
4
_
|x +iy|
2
|x iy|
2
_
=
1
/
4
_
[i[ |y ix|
2
[ i[ |ix +y|
2
_
= (ix, y)
R
.
88 Chapter 6
Furthermore,
(x, x) = (x, x)
R
i(ix, x)
R
= |x
2
|
i
/
4
_
|ix +x|
2
|ix x|
2
_
= |x|
2
_
1
i
/
4
_
[1 +i[
2
[1 i[
2
__
= |x|
2
.
The claim of uniqueness follows from 6.1.3.
10. Examples.
(1) A Hilbert space is exemplied by the L
2
space (over some system
with integration), the inner product introduced as follows (f, g):=
_
fg

for f, g
L
2
. In particular, (x, y):=

eE
x
e
y

e
for x, y l
2
(E ).
(2) Assume that H is a pre-Hilbert space and ( , ) : H
2
F is
a semi-inner product on H. It is clear that the real carrier H
R
with the semi-
inner product ( , )
R
: (x, y) Re(x, y) presents a pre-Hilbert space with the
norm of an element of H independent of whether it is calculated in H or in H
R
.
The pre-Hilbert space (H
R
, ( , )
R
) is the realication or decomplexication of
(H, ( , )). In turn, if the real carrier of a seminormed space is a pre-Hilbert space
then the process of complexication leads to some natural pre-Hilbert structure
of the original space.
(3) Assume that H is a pre-Hilbert space and H

is the twin vector


space of H. Given x, y H

, put (x, y)

:= (x, y)

. Clearly, ( , )

is a semi-
inner product on H

. The resulting pre-Hilbert space is the twin of H, with the


denotation H

preserved.
(4) Let H be a pre-Hilbert space and let H
0
:= ker | | be the kernel
of the seminorm | | on H. Using the CauchyBunyakovskiSchwarz inequality,
Theorem 2.3.8 and 6.1.10 (3), observe that there is a natural inner product on
the quotient space H/H
0
: If x
1
:= (x
1
) and x
2
:= (x
2
), with x
1
, x
2
H and
: H H/H
0
the coset mapping, then (x
1
, x
2
) := (x
1
, x
2
). Moreover, the pre-
Hilbert space H/H
0
may be considered as the quotient space of the seminormed
space (H, | |) by the kernel of the seminorm | |. Thus, H/H
0
is a Hausdor space
referred to as the Hausdor pre-Hilbert space associated with H. Completing the
normed space H/H
0
, obtain a Hilbert space (for instance, by the von Neumann
Jordan Theorem). The so-constructed Hilbert space is called associated with the
original pre-Hilbert space.
(5) Assume that (H
e
)
eE
is a family of Hilbert spaces and H is the
2-sum of the family; i.e., h H if and only if h:= (h
e
)
eE
, where h
e
H
e
for e E
and
|h|:=
_

eE
|h
e
|
2
_
1
/
2
< +.
Hilbert Spaces 89
By 5.5.9 (6), H is a Banach space. Given f, g H, on successively applying the
Parallelogram Law, deduce that
1
/
2
_
|f +g|
2
+|f g|
2
_
=
1
/
2
_

eE
|f
e
+g
e
|
2
+ +

eE
|f
e
g
e
|
2
_
=

eE
1
/
2
_
|f
e
+g
e
|
2
+|f
e
g
e
|
2
_
=

eE
_
|f
e
|
2
+|g
e
|
2
_
= |f|
2
+|g|
2
.
Consequently, H is a Hilbert space by the von NeumannJordan Theorem. The
space H, the Hilbert sum of the family (H
e
)
eE
, is denoted by
eE
H
e
. With
E := N, it is customary to use the symbol H
1
H
2
. . . for H.
(6) Let H be a Hilbert space and let S be a system with integration.
The space L
2
(S, H) comprising all H-valued square-integrable functions is also
a Hilbert space.
6.2. Orthoprojections
6.2.1. Let U

be a convex subset of the spherical layer (r +)B


H
rB
H
with
r, > 0 in a Hilbert space H. Then the diameter of U

vanishes as tends to 0.
Given x, y U

, on considering that
1
/
2
(x + y) U

and applying the


Parallelogram Law, for r infer that
|x y|
2
= 2
_
|x|
2
+|y|
2
_
4
_
_
_
(x+y)
/
2
_
_
_
2
4(r +)
2
4r
2
= 8r + 4
2
12r.
6.2.2. Levy Projection Theorem. Let U be a nonempty closed convex set
in a Hilbert space H and x H U. Then there is a unique element u
0
of U such
that
|x u
0
| = inf|x u| : u U.
Put U

:= u U : |xu| inf |U x|+. By 6.2.1, the family (U

)
>0
constitutes a base for a Cauchy lter in U.
6.2.3. Definition. The element u
0
appearing in 6.2.2 is the best approxima-
tion to x in U or the projection of x to U.
6.2.4. Let H
0
be a closed subspace of a Hilbert space H and x H H
0
.
An element x
0
of H
0
is the projection of x to H
0
if and only if (x x
0
, h
0
) = 0 for
every h
0
in H
0
.
90 Chapter 6
It suces to consider the real carrier (H
0
)
R
of H
0
. The convex function
f(h
0
) := (h
0
x, h
0
x) is dened on (H
0
)
R
. Further, x
0
in H
0
serves as the
projection of x to H
0
if and only if 0
x
0
(f). In view of 3.5.2 (4) this containment
means that (x x
0
, h
0
) = 0 for every h
0
in H
0
, because f

(x
0
) = 2(x
0
x, ).
6.2.5. Definition. Elements x and y of H are orthogonal, in symbols x y,
if (x, y) = 0. By U

we denote the subset of H that comprises all elements


orthogonal to every point of a given subset U; i.e.,
U

:= y H : ( x U) x y.
The set U

is the orthogonal complement or orthocomplement of U (to H).


6.2.6. Let H
0
be a closed subspace of a Hilbert space H. The orthogonal
complement of H
0
, the set H

0
, is a closed subspace and H = H
0
H

0
.
The closure property of H

0
in H is evident. It is also clear that H
0
H

0
=
H
0
H

0
= 0. We are left with showing only that H
0
H

0
= H
0
+H

0
= H. Take
an element h of H H
0
. In virtue of 6.2.2 the projection h
0
of H to H
0
is available
and, by 6.2.4, h h
0
H

0
. Finally, h = h
0
+ (h h
0
) H
0
+H

0
.
6.2.7. Definition. The projection onto a (closed) subspace H
0
along H

0
is
the orthoprojection onto H
0
, denoted by P
H
0
.
6.2.8. Pythagoras Lemma. x y |x +y|
2
= |x|
2
+|y|
2
.
6.2.9. Corollary. The norm of an orthoprojection is at most one: (H ,= 0 &
H
0
,= 0) |P
H
0
| = 1.
6.2.10. Orthoprojection Theorem. For an operator P in L(H) such that
P
2
= P, the following statements are equivalent:
(1) P is the orthoprojection onto H
0
:= imP;
(2) |h| 1 |Ph| 1;
(3) (Px, P
d
y) = 0 for all x, y H, with P
d
the complement of P, i.e.
P
d
= I
H
P;
(4) (Px, y) = (x, Py) for x, y H.
(1) (2): This is observed in 6.2.9.
(2) (3): Let H
1
:= ker P = imP
d
. Take x H

1
. Since x = Px + P
d
x and
x P
d
x; therefore, |x|
2
|Px|
2
= (xP
d
x, xP
d
x) = (x, x)2 Re(x, P
d
x)+
(P
d
x, P
d
x) = |x|
2
+|P
d
x|
2
. Whence P
d
x = 0; i.e., x imP. Considering 6.2.6,
from H
1
= ker P and H

1
imP deduce the equalities H

1
= imP = H
0
. Thus
(Px, P
d
y) = 0 for all x, y H, since Px H
0
and P
d
y H
1
.
(3) (4): (Px, y) = (Px, Py +P
d
y) = (Px, Py) = (Px, Py) +(P
d
x, Py) =
(x, Py).
Hilbert Spaces 91
(4) (1): Show rst that H
0
is a closed subspace. Let h
0
:= limh
n
with
h
n
H
0
, i.e. Ph
n
= h
n
. For every x in H, from the continuity property of the
functionals ( , x) and ( , Px) successively derive
(h
0
, x) = lim(h
n
, x) = lim(Ph
n
, x) = lim(h
n
, Px) = (Ph
0
, x).
Whence (h
0
Ph
0
, h
0
Ph
0
) = 0; i.e., h
0
imP.
Given x H and h
0
H
0
, now infer that (x Px, h
0
) = (x Px, Ph
0
) =
(P(x Px), h
0
) = (Px P
2
x, h
0
) = (Px Px, h
0
) = 0. Therefore, from 6.2.4
obtain Px = P
H
0
x.
6.2.11. Let P
1
, and P
2
be orthoprojections with P
1
P
2
= 0. Then P
2
P
1
= 0.
P
1
P
2
= 0 imP
2
ker P
1
imP
1
= (ker P
1
)

(imP
2
)

= ker P
2

P
2
P
1
= 0
6.2.12. Definition. Orthoprojections P
1
and P
2
are orthogonal, in symbols,
P
1
P
2
or P
2
P
1
, provided that P
1
P
2
= 0.
6.2.13. Theorem. Let P
1
, . . . , P
n
be orthoprojections. The operator P :=
P
1
+. . . +P
n
is an orthoprojection if and only if P
l
P
m
for l ,= m.
: First, given an orthoprojection P
0
, observe that |P
0
x|
2
= (P
0
x, P
0
x) =
(P
2
0
x, x) = (P
0
x, x) by Theorem 6.2.10. Consequently,
|P
l
x|
2
+|P
m
x|
2

k=1
|P
k
x|
2
=
n

k=1
(P
k
x, x) = (Px, x) = |Px|
2
|x|
2
for x H and l ,= m.
In particular, putting x:= P
l
x, observe that
|P
l
x|
2
+|P
m
P
l
x|
2
|P
l
x|
2
|P
m
P
l
| = 0.
: Straightforward calculation shows that P is an idempotent operator. Indeed,
P
2
=
_
n

k=1
P
k
_
2
=
n

l=1
n

m=1
P
l
P
m
=
n

k=1
P
2
k
= P.
Furthermore, in virtue of 6.2.10 (4), (P
k
x, y) = (x, P
k
y) and so (Px, y) = (x, Py).
It suces to appeal to 6.2.10 (4) once again.
6.2.14. Remark. Theorem 6.2.13 is usually referred to as the pairwise orthog-
onality criterion for nitely many orthoprojections.
92 Chapter 6
6.3. A Hilbert Basis
6.3.1. Definition. A family (x
e
)
eE
of elements of a Hilbert space H is or-
thogonal, if e
1
,= e
2
x
e
1
x
e
2
. By specication, a subset E of a Hilbert space H
is orthogonal if so is the family (e)
eE
.
6.3.2. Pythagoras Theorem. An orthogonal family (x
e
)
eE
of elements of
a Hilbert space is (unconditionally) summable if and only if the numeric family
(|x
e
|
2
)
eE
is summable. Moreover,
_
_
_

eE
x
e
_
_
_
2
=

eE
|x
e
|
2
.
Let s

:=

eE
x
e
, where is a nite subset of E . By 6.2.8, |s

|
2
=

e
|x
e
|
2
. Given a nite subset

of E which includes , thus observe that


|s

|
2
= |s

\
|
2
=

\
|x
e
|
2
.
In other words, the fundamentalness of (s

) amounts to the fundamentalness of the


net of partial sums of the family (|x
e
|
2
)
eE
. On using 4.5.4, complete the proof.
6.3.3. Orthoprojection Summation Theorem. Let (P
e
)
eE
be a family
of pairwise orthogonal orthoprojections in a Hilbert space H. Then for every x
in H the family (P
e
x)
eE
is (unconditionally) summable. Moreover, the operator
Px:=

eE
P
e
x is the orthoprojection onto the subspace
H :=
_

eE
x
e
: x
e
H
e
:= imP
e
,

eE
|x
e
|
2
< +
_
.
Given a nite subset of E , put s

:=

e
P
e
. By Theorem 6.2.13, s

is
an orthoprojection. Hence, in view of 6.2.8, |s

x|
2
=

e
|P
e
x|
2
|x|
2
for every x in H. Consequently, the family (|P
e
x|
2
)
eE
is summable (the net
of partial sums is increasing and bounded). By the Pythagoras Theorem, there
is a sum Px :=

eE
P
e
x; i.e., Px = lim

x. Whence P
2
x = lim

Px =
lim

lim

x = lim

lim

x = lim

lim

x = lim

x = Px. Finally,
|Px| = | lim

x| = lim

|s

x| |x| and, moreover, P


2
= P. Appealing to
6.2.10, conclude that P is the orthoprojection onto imP.
If x imP, i.e., Px = x; then x =

eE
P
e
x and by the Pythagoras Theorem

eE
|P
e
x|
2
= |x|
2
= |Px|
2
< +. Since P
e
x H
e
(e E ); therefore, x H .
If x
e
H
e
and

eE
|x
e
|
2
< + then for x :=

eE
x
e
(existence follows from
the same Pythagoras Theorem) observe that x =

eE
x
e
=

eE
P
e
x
e
= Px;
i.e., x imP. Thus, imP = H .
Hilbert Spaces 93
6.3.4. Remark. This theorem may be treated as asserting that the space H
and the Hilbert sum of the family (H
e
)
eE
are isomorphic. The identication is
clearly accomplished by the Bochner integral presenting the process of summation
in this case.
6.3.5. Remark. Let h in H be a normalized or unit or norm-one element;
i.e., |h| = 1. Assume further that H
0
:= Fh is a one-dimensional subspace of H
spanned over h
0
. For every element x of H and every scalar , a member of F,
observe that
(x (x, h)h, h) =

((x, h) (x, h))(h, h) = 0.


Therefore, by 6.2.4, P
H
0
= ( , h) h. To denote this orthoprojection, it is conve-
nient to use the symbol h. Thus, h : x (x, h)h (x H).
6.3.6. Definition. A family of elements of a Hilbert space is called orthonor-
mal (or orthonormalized) if, rst, the family is orthogonal and, second, the norm of
each member of the family equals one. Orthonormal sets are dened by specica-
tion.
6.3.7. For every orthonormal subset E of H and every element x of H, the
family (ex)
eE
is (unconditionally) summable. Moreover, the Bessel inequality
holds:
|x|
2

eE
[(x, e)[
2
.
It suces to refer to the Orthoprojection Summation Theorem, for
|x|
2

_
_
_
_
_

eE
ex
_
_
_
_
_
2
=
_
_
_
_
_

eE
(x, e)e
_
_
_
_
_
2
=

eE
|(x, e)e|
2
.
6.3.8. Definition. An orthonormal set E in a Hilbert space H is a Hilbert
basis (for H) if x =

eE
ex for every x in H. An orthonormal family of elements
of a Hilbert space is a Hilbert basis if the range of the family is a Hilbert basis.
6.3.9. An orthonormal set E is a Hilbert basis for H if and only if lin(E ), the
linear span of E , is dense in H.
6.3.10. Definition. A subset E of a Hilbert space is said to meet the Steklov
condition if E

= 0.
6.3.11. Steklov Theorem. An orthonormal set E is a Hilbert basis if and
only if E meets the Steklov condition.
: Let h E

. Then h =

eE
eh =

eE
(h, e)e =

eE
0 = 0.
: For x H, in virtue of 6.3.3 and 6.2.4, x

eE
ex E

.
94 Chapter 6
6.3.12. Theorem. Each Hilbert space has a Hilbert basis.
By the KuratowskiZorn Lemma, each Hilbert space H has an orthonormal
set E maximal by inclusion. If there were some h in H H
0
, with H
0
:= cl lin(E );
then the element h
1
:= hP
H
0
h would be orthogonal to every element in E . Thus,
for H ,= 0 we would have E |h
1
|
1
h
1
= E . A contradiction. In case H = 0
there is nothing left to proof.
6.3.13. Remark. It is possible to show that two Hilbert bases for a Hilbert
space H have the same cardinality. This cardinality is the Hilbert dimension of H.
6.3.14. Remark. Let (x
n
)
nN
be a countable sequence of linearly independent
elements of a Hilbert space H. Put x
0
:= 0, e
0
:= 0 and
y
n
:= x
n

n1

k=0
e
k
x
n
, e
n
:=
y
n
|y
n
|
(n N).
Evidently, (y
n
, e
k
) = 0 for 0 k n 1 (for instance, by 6.2.13). Also, y
n
,= 0,
since H is innite-dimensional. Say that the orthonormal sequence (e
n
)
nN
results
from the sequence (x
n
)
nN
by the GramSchmidt orthogonalization process. Using
the process, it is easy to prove that a Hilbert space has a countable Hilbert basis
if and only if the space has a countable dense subset; i.e., whenever the space is
separable.
6.3.15. Definition. Let E be a Hilbert basis for a space H and x H. The
numeric family x:= ( x
e
)
eE
in F
E
, given by the identity x
e
:= (x, e), is the Fourier
coecient family of x with respect to E or the Fourier transform of X (relative
to E ).
6.3.16. RieszFisher Isomorphism Theorem. Let E be a Hilbert basis
for H. The Fourier transformF : x x (relative to E ) is an isometric isomorphism
of H onto l
2
(E ). The inverse Fourier transform, the Fourier summation F
1
:
l
2
(E ) H, acts by the rule F
1
(x) :=

eE
x
e
e for x := (x
e
)
eE
l
2
(E ).
Moreover, for all x, y H the Parseval identity holds:
(x, y) =

eE
x
e
y

e
.
By the Pythagoras Theorem, the Fourier transform acts in l
2
(E ). By The-
orem 6.3.3, is an epimorphism. By the Steklov Theorem, is a monomorphism.
It is beyond a doubt that F
1
x = x for x H and

F
1
(x) = x for x l
2
(E ).
The equality
|x|
2
=

eE
| x
e
|
2
= | x|
2
2
(x H)
Hilbert Spaces 95
follows from the Pythagoras Theorem. At the same time
(x, y) =
_

eE
x
e
e,

eE
y
e
e
_
=

e,e

E
x
e
y

e
(e, e

) =

eE
x
e
y

e
.
6.3.17. Remark. The Parseval identity shows that the Fourier transform pre-
serves inner products. Therefore, the Fourier transform is a unitary operator or
a Hilbert-space isomorphism; i.e., an isomorphism preserving inner products. This
is why the RieszFisher Theorem is sometimes referred to as the theorem on Hilbert
isomorphy between Hilbert spaces (of the same Hilbert dimension).
6.4. The Adjoint of an Operator
6.4.1. Riesz Prime Theorem. Let H be a Hilbert space. Given x H, put
x

:= ( , x). Then the prime mapping x x

presents an isometric isomorphism


of H

onto H

.
It is clear that x = 0 x

= 0. If x ,= 0 then
|x

|
H
= sup
|y|1
[(y, x)[ sup
|y|1
|y| |x| |x|;
|x

|
H
= sup
|y|1
[(y, x)[ [(x/|x|, x)[ = |x|.
Therefore, x x

is an isometry of H

into H

. Check that this mapping is


an epimorphism.
Let l H

and H
0
:= ker l ,= H (if there no such l then nothing is to be
proven). Choose a norm-one element e in H

0
and put gradl := l(e)

e. If x H
0
then
(gradl)

(x) = (x, gradl) = (x, l(e)

e) = l(e)

(x, e) = 0.
Consequently, for some in F and all x H in virtue of 2.3.12 (gradl)

(x) = l(x).
In particular, letting x:= e, nd
(gradl)

(e) = (e, grad l) = l(e)(e, e) = l(e);


i.e., = 1.
6.4.2. Remark. From the Riesz Prime Theorem it follows that the dual
space H

possesses a natural structure of a Hilbert space and the prime mapping


x x

implements a Hilbert space isomorphism from H

onto H

. The inverse
mapping now coincides with the gradient mapping l grad l constructed in the
proof of the theorem. Implying this, the claim of 6.4.1 is referred to as the theorem
on the general form of a linear functional in a Hilbert space.
96 Chapter 6
6.4.3. Each Hilbert space is reexive.
Let : H H

be the double prime mapping; i.e. the canonical embedding


of H into the second dual H

which is determined by the rule x

(l) = (x)(l) = l(x),


where x H and l H

(see 5.1.10 (8)). Check that is an epimorphism. Let


f H

. Consider the mapping y f(y

) for y H. It is clear that this mapping is


a linear functional over H

and so by the Riesz Prime Theorem there is an element


x H = H

such that (y, x)

= (x, y) = f(y

) for every y in H. Observe that


(x)(y

) = y

(x) = (x, y) = f(y

) for all y H. Since by the Riesz Prime Theorem


y y

is a mapping onto H

, conclude that (x) = f.


6.4.4. Let H
1
and H
2
be Hilbert spaces and T B(H
1
, H
2
). Then there is
a unique mapping T

: H
2
H
1
such that
(Tx, y) = (x, T

y)
for all x H
1
and y H
2
. Moreover, T

B(H
2
, H
1
) and |T

| = |T|.
Let y H
2
. The mapping x (Tx, y) is the composition y

T; i.e.,
it presents a continuous linear functional over H
1
. By the Riesz Prime Theorem
there is precisely one element x of H
1
for which x

= y

T. Put T

y := x.
It is clear that T

L(H
2
, H
1
). Furthermore, using the CauchyBunyakovski
Schwarz inequality and the normative inequality, infer that
[(T

y, T

y)[ = [(TT

y, y)[ |TT

y| |y| |T| |T

y| |y|.
Hence, |T

y| |T| |y| for all y H


2
; i.e., |T

| |T|. At the same time


T = T

:= (T

; i.e., |T| = |T

| |T

|.
6.4.5. Definition. For T B(H
1
, H
2
), the operator T

, the member of
B(H
2
, H
1
) constructed in 6.4.4, is the adjoint of T. The terms like hermitian-
conjugate and Hilbert-space adjoint are also in current usage.
6.4.6. Let H
1
and H
2
be Hilbert spaces. Assume further that S, T B(H
1
, H
2
)
and F. Then
(1) T

= T;
(2) (S +T)

= S

+T

;
(3) (T)

;
(4) |T

T| = |T|
2
.
(1)(3) are obvious properties. If |x| 1 then
|Tx|
2
= (Tx, Tx) = [(Tx, Tx)[ = [(T

Tx, x)[ |T

Tx| |x| |T

T|.
Furthermore, using the submultiplicativity of the operator norm and 6.4.4, infer
|T

T| |T

| |T| = |T|
2
, which proves (4).
Hilbert Spaces 97
6.4.7. Let H
1
, H
2
, and H
3
be three Hilbert spaces. Assume further that
T B(H
1
, H
2
) and S B(H
2
, H
3
). Then (ST)

= T

.
(STx, z) = (Tx, S

z) = (x, T

z) (x H
1
, z H
3
)
6.4.8. Definition. Consider an elementary diagram H
1
T
H
2
. The dia-
gram H
1
T

H
2
is the adjoint of the initial elementary diagram. Given an ar-
bitrary diagram composed of bounded linear mappings between Hilbert spaces,
assume that each elementary subdiagram is replaced with its adjoint. Then the re-
sulting diagram is the adjoint or, for suggestiveness, the diagram star of the initial
diagram.
6.4.9. Diagram Star Principle. A diagram is commutative if and only if so
is its adjoint diagram.
Follows from 6.4.7 and 6.4.6 (1).
6.4.10. Corollary. An operator T is invertible if and only if T

is invertible.
Moreover, T
1
= T
1
.
6.4.11. Corollary. If T B(H) then Sp(T)

Sp(T

).
6.4.12. Sequence Star Principle (cf. 7.6.13). A sequence
. . . H
k1
T
k
H
k
T
k+1
H
k+1
. . .
is exact if and only if so is the sequence star
. . . H
k1
T

k
H
k
T

k+1
H
k+1
. . . .
6.4.13. Definition. An involutive algebra or -algebra A (over a ground
eld F) is an algebra with an involution , i.e. with a mapping a a

from A
to A such that
(1) a

= a (a A);
(2) (a +b)

= a

+b

(a, b A);
(3) (a)

( F, a A);
(4) (ab)

= b

(a, b A).
A Banach algebra A with involution satisfying |a

a| = |a|
2
for all a A is
a C

-algebra.
6.4.14. The endomorphism space B(H) of a Hilbert space H is a C

-algebra
(with the composition of operators as multiplication and the taking of the adjoint
of an endomorphism as involution).
98 Chapter 6
6.5. Hermitian Operators
6.5.1. Definition. Let H be a Hilbert space over a ground eld F. An ele-
ment T of B(H) is a hermitian operator or selfadjoint operator in H provided that
T = T

.
6.5.2. Rayleigh Theorem. For a hermitian operator T the equality holds:
|T| = sup
|x|1
[(Tx, x)[.
Put t := sup[(Tx, x)[ : |x| 1. It is clear that [(Tx, x)[ |Tx| |x|
|T| provided |x| 1. Thus, t |T|.
Since T = T

; therefore, (Tx, y) = (x, Ty) = (Ty, x)

= (y, Tx)

; i.e.,
(x, y) (Tx, y) is a hermitian form. Consequently, in virtue of 6.1.3 and 6.1.8
4 Re(Tx, y) = (T(x +y), x +y) (T(x y), x y)
t(|x +y|
2
+|x y|
2
) = 2t(|x|
2
+|y|
2
).
If Tx = 0 then it is plain that |Tx| t. Assume Tx ,= 0. Given |x| 1 and
putting y := |Tx|
1
Tx, infer that
|Tx| = |Tx|
_
Tx
|Tx|
,
Tx
|Tx|
_
= (Tx, y) = Re(Tx, y)
1
/
2
t
_
|x|
2
+
_
_
Tx
/
|Tx|
_
_
2
_
t;
i.e., |T| = sup|Tx| : |x| 1 t.
6.5.3. Remark. As mentioned in the proof of 6.5.2, each hermitian operator T
in a Hilbert space H generates the hermitian form f
T
(x, y):= (Tx, y). Conversely,
let f be a hermitian form, with the functional f( , y) continuous for every y
in H. Then by the Riesz Prime Theorem there is an element Ty of H such that
f( , y) = (Ty)

. Evidently, T L(H) and (x, Ty) = f(x, y) = f(y, x)

=
(y, Tx)

= (Tx, y). It is possible to show that in this case T B(H) and T = T

.
In addition, f = f
T
. Therefore, the condition T B(H) in Denition 6.5.1 can be
replaced with the condition T L(H) (the HellingerToeplitz Theorem).
6.5.4. Weyl Criterion. A scalar belongs to the spectrum of a hermitian
operator T if and only if
inf
|x|=1
|x Tx| = 0.
: Put t := inf|x Tx| : x H, |x| = 1 > 0. Demonstrate that
/ Sp(T). Given an x in H, observe that |x Tx| t|x|. Thus, rst, ( T)
Hilbert Spaces 99
is a monomorphism; second, H
0
:= im( T) is a closed subspace (because |(
T)x
m
( T)x
k
| t|x
m
x
k
|; i.e., the inverse image of a Cauchy sequence
is a Cauchy sequence); and, third, which is nal, ( T)
1
B(H) whenever
H = H
0
(in such situation |R(T, )| t
1
). Suppose to the contrary that
H ,= H
0
. Then there is some y in H

0
satisfying |y| = 1. For all x H, note that
0 = (x Tx, y) = (x,

y Ty); i.e.,

y = Ty. Further,

= (Ty, y)/(y, y)
and the hermiticity of T guarantees

R. Whence

= and y ker( T).


We arrive at a contradiction: 1 = |y| = |0| = 0.
: If / Sp(T) then the resolvent of T at , the member R(T, ) of B(H),
is available. Hence, inf|x Tx| : |x| = 1 |R(T, )|
1
.
6.5.5. Spectral Endpoint Theorem. Let T be a hermitian operator in
a Hilbert space. Put
m
T
:= inf
|x|=1
(Tx, x), M
T
:= sup
|x|=1
(Tx, x).
Then Sp(T) [m
T
, M
T
] and m
T
, M
T
Sp(T).
Considering that the operator T Re is hermitian in the space H under
study, from the identity
|x Tx|
2
= [ Im[
2
|x|
2
+|Tx Re x|
2
infer the inclusion Sp(T) R by 6.5.4. Given a norm-one element x of H and
invoking the CauchyBunyakovskiSchwarz inequality, in case < m
T
deduce
that
|x Tx| = |x Tx| |x| [(x Tx, x)[
= [ (Tx, x)[ = (Tx, x) m
T
> 0.
On appealing to 6.5.4, nd res(T). In case > M
T
, similarly infer that
|x Tx| [(x Tx, x)[ = [ (Tx, x)[ = (Tx, x) M
T
> 0.
Once again res(T). Finally, Sp(T) [m
T
, M
T
].
Since (Tx, x) R for x H; therefore, in virtue of 6.5.2
|T| = sup[(Tx, x)[ : |x| 1
= sup(Tx, x) ((Tx, x)) : |x| 1 = M
T
(m
T
).
Assume rst that := |T| = M
T
. If |x| = 1 then
|x Tx|
2
=
2
2(Tx, x) +|Tx|
2
2|T|
2
2|T|(Tx, x).
100 Chapter 6
In other words, the next estimate holds:
inf
|x|=1
|x Tx|
2
2|T| inf
|x|=1
(|T| (Tx, x)) = 0.
Using 6.5.4, conclude that Sp(T).
Now consider the operator S:= Tm
T
. It is clear that M
S
= M
T
m
T
0 and
m
S
= m
T
m
T
= 0. Therefore, |S| = M
S
and in view of the above M
S
Sp(S).
Whence it follows that M
T
belongs to Sp(T), since T = S+m
T
and M
T
= M
S
+m
T
.
It suces to observe that m
T
= M
T
and Sp(T) = Sp(T).
6.5.6. Corollary. The norm of a hermitian operator equals the radius of its
spectrum (and the spectral radius).
6.5.7. Corollary. A hermitian operator is zero if and only if its spectrum
consists of zero.
6.6. Compact Hermitian Operators
6.6.1. Definition. Let X and Y be Banach spaces. An operator T, a member
of L(X, Y ), is called compact (in symbols, T K (X, Y )) if the image T(B
X
) of
the unit ball B
X
of X is relatively compact in Y .
6.6.2. Remark. Detailed study of compact operators in Banach spaces is the
purpose of the RieszSchauder theory to be exposed in Chapter 8.
6.6.3. Let T be a compact hermitian operator. If 0 ,= Sp(T) then is
an eigenvalue of T; i.e., ker( T) ,= 0.
By the Weyl Criterion, x
n
Tx
n
0 for some sequence (x
n
) such that
|x
n
| = 1. Without loss of generality, assume that the sequence (Tx
n
) converges to
y := limTx
n
. Then from the identity x
n
= (x
n
Tx
n
)+Tx
n
obtain that there is
a limit (x
n
) and y = limx
n
. Consequently, Ty = T(limx
n
) = limTx
n
= y.
Since |y| = [[, conclude that y is an eigenvector of T, i.e. y ker( T).
6.6.4. Let
1
and
2
be distinct eigenvalues of a hermitian operator T. Assume
further that x
1
and x
2
are eigenvectors with eigenvalues
1
and
2
(i.e., x
s

ker(
s
T), s:= 1, 2). Then x
1
and x
2
are orthogonal.
(x
1
, x
2
) =
1

1
(Tx
1
, x
2
) =
1

1
(x
1
, Tx
2
) =

2

1
(x
1
, x
2
)
6.6.5. For whatever strictly positive , there are only nitely many eigenvalues
of a compact hermitian operator beyond the interval [, ].
Let (
n
)
nN
be a sequence of pairwise distinct eigenvalues of T satisfying
[
n
[ > . Further, let x
n
be an eigenvector corresponding to
n
and such that
|x
n
| = 1. By virtue of 6.6.4 (x
k
, x
m
) = 0 for m ,= k. Consequently,
|Tx
m
Tx
k
|
2
= |Tx
m
|
2
+|Tx
k
|
2
=
2
m
+
2
k
2
2
;
i.e., the sequence (Tx
n
)
nN
is relatively compact. We arrive at a contradiction to
the compactness property of T.
Hilbert Spaces 101
6.6.6. Spectral Decomposition Lemma. Let T be a compact hermitian
operator in a Hilbert space H and 0 ,= Sp(T). Put H

:= ker(T). Then H

is nite-dimensional and the decomposition H = H

reduces T. Moreover,
the matrix presentation holds
T
_
0
0 T

_
,
where the operator T

, the part of T in H

, is hermitian and compact, with


Sp(T

) = Sp(T) .
The subspace H

is nite-dimensional in view of the compactness of T.


Furthermore, H

is invariant under T. Consequently, the orthogonal complement


H

of H

is an invariant subspace of T

(coincident with T), since h H


( x H

) x h ( x H

) 0 = (h, Tx) = (T

h, x) T

h H

.
The part of T in H

is clearly . The part T

of T in H

is undoubtedly
compact and hermitian. Obviously, for ,= , the operator
T
_
0
0 T

_
is invertible if and only if so is T

. It is also clear that is not an eigenvalue


of T

.
6.6.7. HilbertSchmidt Theorem. Let H be a Hilbert space and let T be
a compact hermitian operator in H. Assume further that P

is the orthoprojection
onto ker( T) for Sp(T). Then
T =

Sp(T)
P

.
Using 6.5.6 and 6.6.6 as many times as need be, for every nite subset
of Sp(T) obtain the equality
_
_
_
_
_
T

_
_
_
_
_
= sup[[ : (Sp(T) 0) .
It suces to refer to 6.6.5.
6.6.8. Remark. The HilbertSchmidt Theorem provides essentially new in-
formation, as compared with the case of nite dimensions, only if the operator T
has innite-rank, that is, the dimension of its range is innite or, which is the same,
H

0
is an innite-dimensional space, where H
0
:= ker T. In fact, if the operator T
102 Chapter 6
has nite rank (i.e., its range is nite-dimensional) then, since the subspace H

0
is
isomorphic with the range of T, observe that
T =
n

k=1

k
e
k
=
n

k=1

k
e

k
e
k
,
where
1
, . . . ,
n
are nonzero points of Sp(T) counted with multiplicity, and e
1
, . . . , e
n

is a properly-chosen orthonormal basis for H

0
.
The HilbertSchmidt Theorem shows that, to within substitution of series for
sum, an innite-rank compact hermitian operator looks like a nite-rank operator.
Indeed, for ,= , where and are nonzero points of Sp(T), the eigenspaces
H

and H

are nite-dimensional and orthogonal. Moreover, the Hilbert sum

Sp(T)\0
H

equals H

0
= cl imT, because H
0
= (imT)

. Successively se-
lecting a basis for each nite-dimensional space H

by enumerating the eigen-


values in decreasing order of magnitudes with multiplicity counted; i.e., putting

1
:=
2
:= . . . :=
dimH

1
:=
1
,
dimH

1
+1
:= . . . :=
dimH

1
+dimH

2
:=
2
, etc.,
obtain the decomposition H = H
0
H

1
H

2
. . . and the presentation
T =

k=1

k
e
k
=

k=1

k
e

k
e
k
,
where the series is summed in operator norm.
6.6.9. Theorem. Let T in K (H
1
, H
2
) be an innite-rank compact operator
from a Hilbert space H
1
to a Hilbert space H
2
. There are orthonormal families
(e
k
)
kN
in H
1
, (f
k
)
kN
in H
2
, and a numeric family (
k
)
kN
in R
+
0,
k
0, such
that the following presentation holds:
T =

k=1

k
e

k
f
k
.
Put S := T

T. It is clear that S B(H


1
) and S is compact. Furthermore,
(Sx, x) = (T

Tx, x) = (Tx, Tx) = |Tx|


2
. Consequently, in virtue of 6.4.6, S
is hermitian and H
0
:= ker S = ker T. Observe also that Sp(S) R
+
by Theorem
6.5.5.
Let (e
k
)
kN
be an orthonormal basis for H

0
comprising all eigenvalues of S and
let (
k
)
kN
be a corresponding decreasing sequence of strictly positive eigenvalues

k
> 0, k N (cf. 6.6.8). Then every element x H
1
expands into the Fourier
series
x P
H
0
x =

k=1
(x, e
k
)e
k
.
Hilbert Spaces 103
Therefore, considering that TP
H
0
= 0 and assigning
k
:=

k
and f
k
:=
1
k
Te
k
,
nd
Tx =

k=1
(x, e
k
)Te
k
=

k=1
(x, e
k
)

k
Te
k
=

k=1

k
(x, e
k
)f
k
.
The family (f
k
)
kN
is orthonormal because
(f
n
, f
m
) =
_
Te
n

n
,
Te
m

m
_
=
1

m
(Te
n
, Te
m
)
=
1

m
(T

Te
n
, e
m
) =
1

m
(Se
n
, e
m
)
=
1

n
,
m
(
n
e
n
, e
m
) =

n

m
(e
n
, e
m
).
Successively using the Pythagoras Theorem and the Bessel inequality, argue as
follows:
_
_
_
_
_
_
T
n

k=1

k
e

k
f
k
_
x
_
_
_
_
_
2
=
_
_
_
_
_

k=n+1

k
(x, e
k
)f
k
_
_
_
_
_
2
=

k=n+1

2
k
[(x, e
k
)[
2

n+1

k=n+1
[(x, e
k
)[
2

n+1
|x|
2
.
Since
k
0, nally deduce that
_
_
_
_
_
T
n

k=1

k
e

k
f
k
_
_
_
_
_

n+1
0.
6.6.10. Remark. Theorem 6.6.9 means in particular that a compact operator
(and only a compact operator) is an adherent point of the set of nite-rank oper-
ators. This fact is also expressed as follows: Every Hilbert space possesses the
approximation property.
Exercises
6.1. Describe the extreme points of the unit ball of a Hilbert space.
6.2. Find out which classical Banach spaces are Hilbert spaces and which are not.
6.3. Is a quotient space of a Hilbert space also a Hilbert space?
6.4. Is it true that each Banach space may be embedded into a Hilbert space?
6.5. Is it possible that the (bounded) endomorphism space of a Hilbert space presents
a Hilbert space?
104 Chapter 6
6.6. Describe the second (= repeated) orthogonal complement of a set in a Hilbert space.
6.7. Prove that no Hilbert basis for an innite-dimensional Hilbert space is a Hamel basis.
6.8. Find the best approximation to a polynomial of degree n +1 by polynomials of degree
at most n in the L
2
space on an interval.
6.9. Prove that x y if and only if |x +y|
2
= |x|
2
+|y|
2
and |x +iy|
2
= |x|
2
+|y|
2
.
6.10. Given a bounded operator T, prove that
(ker T)

= cl imT

, (imT)

= ker T

.
6.11. Reveal the interplay between hermitian forms and hermitian operators (cf. 6.5.3).
6.12. Find the adjoint of a shift operator, a multiplier, and a nite-rank operator.
6.13. Prove that an operator between Hilbert spaces is compact if and only if so is its
adjoint.
6.14. Assume that an operator T is an isometry. Is T

an isometry too?
6.15. A partial isometry is an operator isometric on the orthogonal complement of its kernel.
What is the structure of the adjoint of a partial isometry?
6.16. What are the extreme points of the unit ball of the endomorphism space of a Hilbert
space?
6.17. Prove that the weak topology of a separable Hilbert space becomes metrizable if re-
stricted onto the unit ball.
6.18. Show that an idempotent operator P in a Hilbert space is an orthoprojection if and
only if P commutes with P

.
6.19. Let (a
kl
)
k,lN
be an innite matrix such that a
kl
0 for all k and l. Assume further
that there are also p
k
and , > 0 satisfying

k=1
a
kl
p
k
p
l
;

l=1
a
kl
p
l
p
k
(k, l N).
Then there is some T in B(l
2
) such that (e
k
, e
l
) = a
kl
and |T| =
_
, where e
k
is the
characteristic function of k, a member of N.
Chapter 7
Principles of Banach Spaces
7.1. Banachs Fundamental Principle
7.1.1. Lemma. Let U be a convex set with nonempty interior in a multi-
normed space: int U ,= . Then
(1) 0 < 1 cl U + (1 ) int U int U;
(2) core U = int U;
(3) cl U = cl int U;
(4) int cl U = int U.
(1): For u
0
int U, the set int U u
0
is an open neighborhood of zero
in virtue of 5.2.10. Whence, given 0 < 1, obtain
cl U = cl U U + (1 )(int U u
0
)
= U + (1 ) int U (1 )u
0
U + (1 )U (1 )u
0
U (1 )u
0
.
Thus, (1 )u
0
+cl U U and so U includes (1 ) int U +cl U. The last set
is open as presenting the sum of cl U and (1 ) int U, an open set.
(2): Undoubtedly, int U core U. If u
0
int U and u core U then u =
u
0
+(1 )u
1
for some u
1
in U and 0 < < 1. Since u
1
cl U, from (1) deduce
that u int U.
(3): Clearly, cl int U cl U for int U U. If, in turn, u cl U; then, choosing
u
0
in the set int U and putting u

:= u
0
+ (1 )u, infer that u

u as 0
and u

int U when 0 < < 1. Thus, by construction u cl int U.


(4): From the inclusions int U U cl U it follows that int U int cl U.
If now u int cl U then, in virtue of (2), u core cl U. Consequently, taking u
0
in the set int U again, nd u
1
cl U and 0 < < 1 satisfying u = u
0
+(1 )u
1
.
Using (1), nally infer that u int U.
106 Chapter 7
7.1.2. Remark. In the case of nite dimensions, the condition int U ,= may
be omitted in 7.1.1 (2) and 7.1.1 (4). In the opposite case, the presence of an interior
point is an essential requirement, as shown by numerous examples. For instance,
take U := B
c
0
X, where c
0
is the space of vanishing sequences and X is the
subspace of terminating sequences in c
0
, the direct sum of countably many copies
of a basic eld. Evidently, core U = and at the same time cl U = B
c
0
.
7.1.3. Definition. A subset U of a (multi)normed space X is an ideally con-
vex set in X, if U is closed under the taking of countable convex combinations.
More precisely, U is ideally convex if for all sequences (
n
)
nN
and (u
n
)
nN
, with

n
R
+
,

n=1

n
= 1 and u
n
U such that the series

n=1

n
u
n
converges
in X to some u, the containment holds: u U.
7.1.4. Examples.
(1) Translation (by a vector) preserves ideal convexity.
(2) Every closed convex set is ideally convex.
(3) Every open convex set is ideally convex.
Take an open and convex U. If U = then nothing is left to proof. If U ,=
then by 7.1.4 (1) it may be assumed that 0 U and, consequently, U = p
U
< 1,
where p
U
is the Minkowski functional of U. Let (u
n
)
nN
and (
n
)
nN
be sequences
in U and in R
+
such that

n=1

n
= 1, with the element u:=

n=1

n
u
n
failing
to lie in U. By virtue of 7.1.4 (2), u belongs to cl U = p
U
1 and so p
U
(u) = 1.
On the other hand, it is clear that p
U
(u)

n=1

n
p
U
(u
n
) 1 =

n=1

n
(cf. 7.2.1). Thus, 0 =

n=1
(
n

n
p
U
(u
n
)) =

n=1

n
(1 p
U
(u
n
)). Whence

n
= 0 for all n N. We arrive at a contradiction.
(4) The intersection of a family of ideally convex sets is ideally con-
vex.
(5) Every convex subset of a nite-dimensional space is ideally con-
vex.
7.1.5. Banachs Fundamental Principle. In a Banach space, each ideally
convex set with absorbing closure is a neighborhood of zero.
Let U be such set in a Banach space X. By hypothesis X =
nN
ncl U. By
the Baire Category Theorem, X is nonmeager and so there is some n in N such that
int ncl U ,= . Therefore, int cl U =
1
/
n
int ncl U ,= . By hypothesis 0 core cl U.
Consequently, from 7.1.1 it follows that 0 int cl U. In other words, there is a > 0
such that cl U B
X
. Consequently,
> 0 cl
1
/

B
X
.
Using the above implication, show that U

/
2
B
X
.
Principles of Banach Spaces 107
Let x
0


/
2
B
X
. Putting := 2, choose y
1

1
/

U from the condition |y


1

x
0
|
1
/
2
. Thus obtain an element u
1
of U such that
_
_
1
/
2
u
1
x
0
_
_

1
/
2
=
1
/
4
.
Now putting x
0
:=
1
/
2
u
1
+ x
0
and := 4 and applying the argument of the
preceding paragraph, nd an element u
2
of U satisfying
_
_
1
/
4
u
2
+
1
/
2
u
1
x
0
_
_

1
/
2
=
1
/
8
. Proceeding by induction, construct a sequence (u
n
)
nN
of the points
of U which possesses the property that the series

n=1
1
/
2
nu
n
converges to x
0
.
Since

n=1
1
/
2
n = 1 and the set U is ideally convex, deduce x
0
U.
7.1.6. For every ideally convex set U in a Banach space the following four
sets coincide: the core of U, the interior of U, the core of the closure of U and the
interior of the closure of U.
It is clear that int U core U core cl U. If u core cl U then cl(U u)
equal to cl U u is an absorbing set. An ideally convex set translates into an ideally
convex set (cf. 7.1.4 (1)). Consequently, Uu is a neighborhood of zero by Banachs
Fundamental Principle. By virtue of 5.2.10, u belongs to int U. Thus, int U =
core U = core cl U. Using 7.1.1, conclude that int cl U = int U.
7.1.7. The core and the interior of a closed convex set in a Banach space
coincide.
A closed convex set is ideally convex.
7.1.8. Remark. Inspection of the proof of 7.1.5 shows that the condition for
the ambient space to be a Banach space in 7.1.7 is not utilized to a full extent.
There are examples of incomplete normed spaces in which the core and interior of
each closed convex set coincide. A space with this property is called barreled. The
concept of barreledness is seen to make sense also in multinormed spaces. Barreled
multinormed spaces are plentiful. In particular, such are all Frechet spaces.
7.1.9. Counterexample. Each innite-dimensional Banach space contains
absolutely convex, absorbing and not ideally convex sets.
Using, for instance, a Hamel basis, take a discontinuous linear functional f.
Then the set [f[ 1 is what was sought.
7.2. Boundedness Principles
7.2.1. Let p : X R be a sublinear functional on a normed space (X, | |).
The following conditions are equivalent:
(1) p is uniformly continuous;
(2) p is continuous;
(3) p is continuous at zero;
(4) p 1 is a neighborhood of zero;
(5) |p|:= sup[p(x)[ : |x| 1 < +; i.e., p is bounded.
108 Chapter 7
The implications (1) (2) (3) (4) are immediate.
(4) (5): There is some t > 0 such that t
1
B
X
p 1. Given |x| 1,
thus nd p(x) t. In addition, the inequality p(x) p(x) implies that p(x)
t for x B
X
. Finally, |p| t < +.
(5) (1): From the subadditivity of p, given x, y X, observe that
p(x) p(y) p(x y); p(y) p(x) p(y x).
Whence [p(x) p(y)[ p(x y) p(y x) |p| |x y|.
7.2.2. Gelfand Theorem. Every lower semicontinuous sublinear functional
with domain a Banach space is continuous.
Let p be such functional. Then the set p 1 is closed (cf. 4.3.8). Since
domp is the whole space; therefore, by 3.8.8, p 1 is an absorbing set. By
Banachs Fundamental Principle p 1 is a neighborhood of zero. Application
to 7.2.1 completes the proof.
7.2.3. Remark. The Gelfand Theorem is stated amply as follows: If X is
a Banach space then each of the equivalent conditions 7.2.1 (1)7.2.1 (5) amounts
to the statement: p is lower semicontinuous on X. Observe immediately that
the requirement domp = X may be slightly relaxed by assuming domp to be
a nonmeager linear set and withdrawing the condition for X to be a Banach space.
7.2.4. Equicontinuity Principle. Suppose that X is a Banach space and Y
is a (semi)normed space. For every nonempty set E of continuous linear operators
from X to Y the following statements are equivalent:
(1) E is pointwise bounded; i.e., for all x X the set Tx : T E is
bounded in Y ;
(2) E is equicontinuous.
(1) (2): Put q(x) := supp(Tx) : T E , with p the (semi)norm of Y .
Evidently, q is a lower semicontinuous sublinear functional and so by the Gelfand
Theorem |q| < +; i.e., p(T(x y)) |q| |x y| for all T E . Consequently,
T
1
(d
p
) d
||
/|q| for every T in E , where > 0 is taken arbitrarily.
This means the equicontinuity property of E .
(2) (1): Straightforward.
7.2.5. Uniform Boundedness Pronciple. Let X be a Banach space and
let Y be a normed space. For every nonempty family (T

of bounded operators
the following statements are equivalent:
(1) x X sup

|T

x| < +;
(2) sup

|T

| < +.
It suces to observe that 7.2.5 (2) is another expression for 7.2.4 (2).
Principles of Banach Spaces 109
7.2.6. Let X be a Banach space and let U be a subset of X

. Then the
following statements are equivalent:
(1) U is bounded in X

;
(2) for every x in X the numeric set x [ x

: x

U is bounded
in F.
This is a particular case of 7.2.5.
7.2.7. Let X be a normed space and let U be a subset of X. Then the following
statements are equivalent:
(1) U is bounded in X;
(2) for every x

in X

the numeric set x [ x

: x U is bounded
in F.
Only (2) (1) needs examining. Observe that X

is a Banach space
(cf. 5.5.7) and X is isometrically embedded into X

by the double prime map-


ping (cf. 5.1.10 (8)). So, the claim follows from 7.2.6.
7.2.8. Remark. The message of 7.2.7 (2) may be reformulated as U is bound-
ed in the space (X, (X, X

)) or, in view of 5.1.10 (4), as U is weakly bounded.


The duality between 7.2.6 and 7.2.7 is perfectly revealed in 10.4.6.
7.2.9. BanachSteinhaus Theorem. Let X and Y be Banach spaces. As-
sume further that (T
n
)
nN
, T
n
B(X, Y ), is a sequence of bounded operators.
Put E:= x X : limT
n
x. The following conditions are equivalent:
(1) E = X;
(2) sup
nN
|T
n
| < + and E is dense in X.
Under either (and, hence, both) of the conditions (1) and (2) the mapping T
0
:
X Y , dened as T
0
x:= limT
n
x, presents a bounded linear operator and |T
0
|
liminf |T
n
|.
If E = X then, of course, cl E = X. In addition, for every x in X the se-
quence (T
n
x)
nN
is bounded in Y (for, it converges). Consequently, by the Uniform
Boundedness Principle sup
nN
|T
n
| < + and (1) (2) is proven.
If (2) holds and x X then, given x E and m, k N, infer that
|T
m
x T
k
x| = |T
m
x T
m
x +T
m
x T
k
x +T
k
x T
k
x|
|T
m
x T
m
x| +|T
m
x T
k
x| +|T
k
x T
k
x|
|T
m
| |x x| +|T
m
x T
k
x| +|T
k
| |x x|
2 sup
nN
|T
n
| |x x| +|T
m
x T
k
x|.
Take > 0 and choose, rst, x E such that 2 sup
n
|T
n
| |x x|

/
2
, and,
second, n N such that |T
m
x T
k
x|

/
2
for m, k n. By virtue of what was
110 Chapter 7
proven |T
m
x T
k
x| ; i.e., (T
n
x)
nN
is a Cauchy sequence in Y . Since Y is
a Banach space, conclude that x E. Thus, (2) (1) is proven.
To complete the proof it suces to observe that
|T
0
x| = lim|T
n
x| liminf |T
n
| |x|
for all x X, because every norm is a continuous function.
7.2.10. Remark. Under the hypotheses of the BanachSteinhaus Theorem,
the validity of either of the equivalent items 7.2.9 (1) and 7.2.9 (2) implies that
(T
n
) converges to T
0
compactly on X (= uniformly on every compact subset of X).
In other words,
sup
xQ
|T
n
x T
0
x| 0
for every (nonempty) compact set Q in X.
Indeed, it follows from the Gelfand Theorem that the sublinear functional
p
n
(x):= sup|T
m
xT
0
x| : m n is continuous. Moreover, p
n
(x) p
n+1
(x) and
p
n
(x) 0 for all x X. Consequently, the claim follows from the Dini Theorem:
Each decreasing sequence of continuous real functions which converges pointwise
to a continuous function on a compact set converges uniformly.
7.2.11. Singularity Fixation Principle. Let X be a Banach space and let
Y be a normed space. If (T
n
)
nN
is a sequence of operators, T
n
B(X, Y )
and sup
n
|T
n
| = + then there is a point x of X satisfying sup
n
|T
n
x| = +.
The set of such points xing a singularity is residual.
The rst part of the assertion is contained in the Uniform Boundedness
Principle. The second requires referring to 7.2.3 and 4.7.4.
7.2.12. Singularity Condensation Principle. Let X be a Banach space
and let Y be a normed space. If (T
n,m
)
n,mN
is a family of operators, T
n
B(X, Y ),
such that sup
n
|T
n,m
| = +for every m N then there is a point x of X satisfying
sup
n
|T
n,m
x| = + for all m N.
7.3. The Ideal Correspondence Principle
7.3.1. Let X and Y be vector spaces. A correspondence F XY is convex
if and only if for x
1
, x
2
X and
1
,
2
R
+
such that
1
+
2
= 1, the inclusion
holds:
F(
1
x
1
+
2
x
2
)
1
F(x
1
) +
2
F(x
2
).
: If (x
1
, y
1
), (x
2
, y
2
) F and
1
,
2
0,
1
+
2
= 1, then
1
y+
2
y
2

F(
1
x
1
+
2
x
2
) since y
1
F(x
1
) and y
2
F(x
2
).
: If either x
1
or x
2
fails to enter in domF then there is nothing to prove. If
y
1
F(x
1
) and y
2
F(x
2
) with x
1
, x
2
domF then
1
(x
1
, y
1
) +
2
(x
2
, y
2
) F
for
1
,
2
0,
1
+
2
= 1 (cf. 3.1.2 (8)).
Principles of Banach Spaces 111
7.3.2. Remark. Let X and Y be Banach spaces. It is clear that there are
many ways for furnishing the space X Y with a norm so that the norm topology
be coincident with the product of the topologies
X
and
Y
. For instance, it is
possible to put |(x, y)|:= |x|
X
+|y|
Y
; i.e., to dene the norm on X Y as that
of the 1-sum of X and Y . Observe immediately that the concept of ideally convex
set has a linear topological character: the class of objects distinguished in a space
is independent of the way of introducing the space topology; in particular, the class
remains invariant under passage to an equivalent (multi)norm. In this connection
the next denition is sound.
7.3.3. Definition. A correspondence F X Y , with X and Y Banach
spaces, is called ideally convex or, briey, ideal if F is an ideally convex set.
7.3.4. Ideal Correspondence Lemma. The image of a bounded ideally
convex set under an ideal correspondence is an ideally convex set.
Let F X Y be an ideal correspondence and let U be a bounded ideally
convex set in X. If U domF = then F(U) = and nothing is left unproven.
Let now (y
n
)
nN
F(U); i.e., y
n
F(x
n
), where x
n
U and n N. Let, nally,
(
n
) be a sequence of positive numbers such that

n=1

n
= 1 and, moreover,
there is a sum of the series y :=

n=1

n
y
n
in Y . It is beyond a doubt that

n=1
|
n
x
n
| =

n=1

n
|x
n
|

n=1

n
sup |U| = sup |U| < +
in view of the boundedness property of U. Since X is complete, from 5.5.3 it
follows that X contains the element x :=

n=1

n
x
n
. Consequently, (x, y) =

n=1

n
(x
n
, y
n
) in the space X Y . Successively using the ideal convexity of F
and U, infer that (x, y) F and x U. Thus, y F(U).
7.3.5. Ideal Correspondence Principle. Let X and Y be Banach spaces.
Assume further that F X Y is an ideal correspondence and (x, y) F.
A correspondence F carries each neighborhood of x onto a neighborhood about y
if and only if y core F(X).
: This is obvious.
: On account of 7.1.4 it may be assumed that x = 0 and y = 0. Since each
neighborhood of zero U includes B
X
for some > 0, it suces to settle the case
U := B
X
. Since U is bounded; by 7.3.4, F(U) is ideally convex. To complete the
proof, it suces to show that F(U) is an absorbing set and to cite 7.1.6.
Take an arbitrary element y of Y . Since by hypothesis 0 core F(X), there is
a real in R
+
such that y F(X). In other words, y F(X) for some x in X.
112 Chapter 7
If |x| 1 then there is nothing to prove. If |x| > 1 then := |x|
1
< 1. Whence,
using 7.3.1, infer that
y = (1 )0 +y (1 )F(0) +F(x)
F((1 )0 +x) = F(x) F(B
X
) = F(U).
Here use was made of the fact that |x| = 1; i.e., x B
X
.
7.3.6. Remark. The property of F, described in 7.3.5, is referred to as the
openness of F at (x, y).
7.3.7. Remark. The Ideal Correspondence Principle is formally weaker that
Banachs Fundamental Principle. Nevertheless, the gap is tiny and can be easily
lled in. Namely, the conclusion of 7.3.5 remains valid if we suppose that y
core cl F(X), on additionally requiring ideal convexity from F(X). The requirement
is not too stringent and certainly valid provided that the domain of F is bounded
in virtue of 7.3.4. As a result of this slight modication, 7.1.5 becomes a particular
case of 7.3.5. In this connection the claim of 7.3.5 is often referred to as Banachs
Fundamental Principle for a Correspondence.
7.3.8. Definition. Let X and Y be Banach spaces and let F X Y be
a correspondence. Then F is called closed if F is a closed set in X Y .
7.3.9. Remark. For obvious reasons, a closed correspondence is often referred
to as a closed-graph correspondence.
7.3.10. A correspondence F is closed if and only if for all sequences (x
n
) in X
and (y
n
) in Y such that x
n
domF, y
n
F(x
n
) and x
n
x, y
n
y, it follows
that x domF and y F(x).
7.3.11. Assume that X and Y are Banach spaces and F X Y is a closed
convex correspondence. Further, let (x, y) F and y core imF. Then F carries
each neighborhood of x onto a neighborhood about y.
A closed convex set is ideally convex and so all follows from 7.3.5.
7.3.12. Definition. A correspondence F XY is called open if the image
of each open set in X is an open set in Y .
7.3.13. Open Correspondence Principle. Let X and Y be Banach spaces
and let F X Y be an ideal correspondence, with imF an open set. Then F is
an open correspondence.
Let U be an open set in X. If y F(U) then there is some x in U such
that (x, y) F. It is clear that y core imF. By the criterion of 7.3.5 F(U) is
a neighborhood of y because U is a neighborhood of x. This means that F(U) is
an open set.
Principles of Banach Spaces 113
7.4. Open Mapping and Closed Graph Theorems
7.4.1. Definition. A member T of L(X, Y ) is a homomorphism, if T
B(X, Y ) and T is an open correspondence.
7.4.2. Assume that X is a Banach space, Y is a normed space and T is a ho-
momorphism from X to Y . Then imT = Y and Y is a Banach space.
It is obvious that imT = Y . Presuming T to be a monomorphism, observe
that T
1
L(Y, X). Since T is open, T
1
belongs to B(Y, X), which ensures the
completeness of Y (the inverse image of a Cauchy sequence in a subset is a Cauchy
sequence in the inverse image of the subset). In the general case, consider the
coimage coimT := X/ ker T endowed with the quotient norm. In virtue of 5.5.4,
coimT is a Banach space. In addition, by 2.3.11 there is a unique quotient T of T
by coimT, the monoquotient of T. Taking account of the denition of quotient norm
and 5.1.3, conclude that T is a homomorphism. Furthermore, T is a monomorphism
by denition. It remains to observe that imT = imT = Y.
7.4.3. Remark. As regards the monoquotient T : coimT Y of T, it may
be asserted that |T| = |T|.
7.4.4. Banach Homomorphism Theorem. Every bounded epimorphism
from one Banach space onto the other is a homomorphism.
Let T B(X, Y ) and imT = Y . On applying the Open Correspondence
Principle to T, complete the proof.
7.4.5. Banach Isomorphism Theorem. Let X and Y be Banach spaces
and T B(X, Y ). If T is an isomorphism of the vector spaces X and Y , i.e.
ker T = 0 and imT = Y ; then T
1
B(Y, X).
A particular case of 7.4.4.
7.4.6. Remark. The Banach Homomorphism Theorem is often referred to as
the Open Mapping Theorem for understandable reasons. Theorem 7.4.5 is briey
formulated as follows: A continuous (algebraic) isomorphism of Banach spaces
is a topological isomorphism. It is also worth observing that the theorem is
sometimes referred to as the Well-Posedness Principle and verbalized as follows:
If an equation Tx = y, with T B(X, Y ) and X and Y Banach spaces, is unique-
ly solvable given an arbitrary right side; then the solution x depends continuously
on the right side y.
7.4.7. Banach Closed Graph Theorem. Let X and Y be Banach spaces
and let T in L(X, Y ) be a closed linear operator. Then T is continuous.
The correspondence T
1
is ideal and T
1
(Y ) = X.
114 Chapter 7
7.4.8. Corollary. Suppose that X and Y are Banach spaces and T is a linear
operator from X to Y . The following conditions are equivalent:
(1) T B(X, Y );
(2) for every sequence (x
n
)
nN
in X, together with some x in X and y
in Y satisfying x
n
x and Tx
n
y, it happens that y = Tx.
(2): This is a reformulation of the closure property of T.
7.4.9. Definition. A subspace X
1
of a Banach space X is complemented
(rarely, a topologically complemented), if X
1
is closed and, moreover, there is a closed
subspace X
2
such that X = X
1
X
2
(i.e., X
1
X
2
= 0 and X
1
X
2
= X). These
X
1
and X
2
are called complementary to one another.
7.4.10. Complementation Principle. For a subspace X
1
of some Banach
space X one of the following conditions amounts to the other:
(1) X
1
is complemented;
(2) X
1
is the range of a bounded projection; i.e., there is a member P
of B(X) such that P
2
= P and imP = X
1
.
(1) (2): Let P be the projection of X onto X
1
along X
2
(cf. 2.2.9 (4)).
Let (x
n
)
nN
be a sequence in X with x
n
x and Px
n
y. It is clear that
Px
n
X
1
for n N. Since X
1
is closed, by 4.1.19 y X
1
. Similarly, the condition
(x
n
Px
n
X
2
for n N) implies that x y X
2
. Consequently, P(x y) = 0.
Furthermore, y = Py; i.e., y = Px. It remains to refer to 7.4.8.
(2) (1): It needs showing only that X
1
equal to imP is closed. Take a se-
quence (x
n
)
nN
in X
1
such that x
n
x in X. Then Px
n
Px in view of the
boundedness of P. Obtain Px
n
= x
n
, because x
n
imP and P is an idempotent.
Finally, x = Px, i.e. x X
1
, what was required.
7.4.11. Examples.
(1) Every nite-dimensional subspace is complemented.
(2) The space c
0
is not complemented in l

.
It turns out more convenient to work with X:= l

(Q) and Y := c
0
(Q), where
Q is the set of rational numbers. Given t R, choose a sequence (t
n
) of pairwise
distinct rational numbers other than t and such that t
n
t. Let Q
t
:= t
n
: n N.
Observe that Q
t
Q
t
is a nite set if t

,= t

.
Let
t
be the coset containing the characteristic function of Q
t
in the quotient
space X/Y and V :=
t
: t R. Since
t
,=
t
for t

,= t

, the set V is
uncountable. Take f (X/Y )

and put V
f
:= v V : f(v) ,= 0. It is evident
that V
f
=
nN
V
f
(n), where V
f
(n) := v V : [f(v)[
1
/
n
. Given m N and
pairwise distinct v
1
, . . . , v
m
in V
f
(n), put
k
:= [f(v
k
)[/f(v
k
) and x:=

n
k=1

k
v
k
to nd |x| 1 and |f| [f(x)[ = [

m
k=1

k
f(v
k
)[ = [

m
k=1
[f(v
k
)[[
m
/
n
.
Principles of Banach Spaces 115
Hence, V
f
(n) is a nite set. Consequently, V
f
is countable. Whence it follows
that for every countable subset F of (X/Y )

there is an element v of V satisfying


( f F) f(v) = 0. At the same time the countable set of the coordinate projections

q
: x x(q) (q Q) is total over l

(Q); i.e., ( q Q)
q
(x) = 0 x = 0 for
x l

(Q). It remains to compare the above observations.


(3) Every closed subspace of a Hilbert space is complemented in virtue
of 6.2.6. Conversely, if, in an arbitrary Banach space X with dimX 3, each closed
subspace is the range of some projection P with |P| 1; then X is isometrically
isomorphic to a Hilbert space (this is the Kakutani Theorem). The next fact is
much deeper:
LindenstraussTzafririTheorem. A Banach space having every closed sub-
space complemented is topologically isomorphic to a Hilbert space.
7.4.12. Sard Theorem. Suppose that X, Y, and Z are Banach spaces. Take
A B(X, Y ) and B B(Y, Z). Suppose further that imA is a complemented
subspace in Y. The diagram
Z
X Y
A
X
B


is commutative for some X in B(Y, Z) if and only if ker A ker B.
Only needs examining. Moreover, in the case imA = Y the sole member
X
0
of L(Y, Z) such that X
0
A = B is continuous. Indeed, X
1
0
(U) = A(B
1
(U))
for every open set U in Z. The set B
1
(U) is open in virtue of the boundedness
property of B, and A(B
1
(U)) is open by the Banach Homomorphism Theorem.
In the general case, construct X
0
B(imA, Z) and take as X the operator X
0
P,
where P is some continuous projection of Y onto imA. (Such a projection is
available by the Complementation Principle.)
7.4.13. Phillips Theorem. Suppose that X, Y, and Z are Banach spaces.
Take A B(Y, X) and B B(Z, X). Suppose further that ker A is a comple-
mented subspace of Y . The diagram
Z
X Y
A
X
B


is commutative for some X in B(Z, Y ) if and only if imA imB.
116 Chapter 7
Once again only needs examining. Using the denition of complemented
subspace, express Y as the direct sum of ker A and Y
0
, where Y
0
is a closed subspace.
By 5.5.9 (1), Y
0
is a Banach space. Consider the part A
0
of A in Y
0
. Undoubtedly,
imA
0
= imA imB. Consequently, by 2.3.13 and 2.3.14 the equation A
0
X
0
= B
has a unique solution X
0
:= A
1
0
. It suces to prove that the operator X
0
, treated
as a member of L(Z, Y
0
), is bounded.
The operator X
0
is closed. Indeed (cf. 7.4.8), if z
n
z and A
1
0
Bz
n
y
then Bz
n
Bz, since B is bounded. In addition, by the continuity of A
0
, the
correspondence A
1
0
X Y
0
is closed; and so 7.3.10 yields the equality y =
A
1
0
Bz.
7.4.14. Remark. We use neither the completeness of Z in proving the Sard
Theorem nor the completeness of X in proving the Phillips Theorem.
7.4.15. Remark. The Sard Theorem and the Phillips Theorem are in formal
duality; i.e., one results from the other by reversing arrows and inclusions and
substituting ranges for kernels (cf. 2.3.15).
7.4.16. Two Norm Principle. Let a vector space be complete in each of two
comparable norms. Then the norms are equivalent.
For deniteness, assume that | |
2
| |
1
in X. Consider the diagram
(X, | |
1
)
(X, | |
1
) (X, | |
2
)
I
X
X
I
X


By the Phillips Theorem some continuous operator X makes the diagram commu-
tative. Such an operator is unique: it is I
X
.
7.4.17. Graph Norm Principle. Let X and Y be Banach spaces and let T
in L(X, Y ) be a closed operator. Given x X, dene the graph norm of x as
|x|
gr T
:= |x|
X
+|Tx|
Y
. Then | |
gr T
| |
X
.
Observe that the space (X, | |
gr T
) is complete. Further, | |
gr T
| |
X
.
It remains to refer to the Two Norm Principle.
7.4.18. Definition. A normed space X is a Banach range, if X is the range
of some bounded operator given on some Banach space.
7.4.19. Kato Criterion. Let X be a Banach space and X = X
1
X
2
, where
X
1
, X
2
Lat(X). The subspaces X
1
and X
2
are closed if and only if each of them
is a Banach range.
Principles of Banach Spaces 117
: A corollary to the Complementation Principle.
: Let Z be a some Banach range, i.e. Z = T(Y ) for some Banach space Y
and T B(Y, Z). Passing, if need be, to the monoquotient of T, we may assume
that T is an isomorphism. Put |z|
0
:= |T
1
z|
Y
. It is clear that (Z, | |
0
) is
a Banach space and |z| = |TT
1
z| |T||T
1
z| = |T||z|
0
; i.e., | |
0
| |
Z
.
Applying this construction to X
1
and X
2
, obtain Banach spaces (X
1
, | |
1
) and
(X
2
, | |
2
). Now | |
k
| |
X
on X
k
for k:= 1, 2.
Given x
1
X
1
and x
2
X
2
, put |x
1
+ x
2
|
0
:= |x
1
|
1
+ |x
2
|
2
. Thereby we
introduce in X some norm | | that is stronger than the initial norm | |
X
. By con-
struction (X, | |
0
) is a Banach space. It remains to refer to 7.4.16.
7.5. The Automatic Continuity Principle
7.5.1. Lemma. Let f : X R

be a convex function on a (multi)normed


space X. The following statements are equivalent:
(1) U := int domf ,= and f[
U
is a continuous function;
(2) there is a nonempty open set V such that sup f(V ) < +.
(1) (2): This is obvious.
(2) (1): It is clear that U ,= . Using 7.1.1, observe that each point u of U
has a neighborhood W in which f is bounded above, i.e. t := sup f(W) < +.
Without loss of generality, it may be assumed that u := 0, f(u) := 0 and W is
an absolutely convex set. From the convexity of f, for every R
+
such that
1 and for an arbitrary v in W, obtain
f(v) = f(v + (1 )0) f(v) + (1 )f(0) = f(v);
f(v) +f(v) f(v) +f((v))
= 2(
1
/
2
f
_
v) +
1
/
2
f(v)
_
2f(0) = 0.
Therefore, [f(W)[ t, which implies that f is continuous at zero.
7.5.2. Corollary. If x int domf and f is continuous at x then the sub-
dierential
x
(f) contains only continuous functionals.
If l
x
(f) then ( x X) l(x) l(x) + f(x) f(x) and so l is bounded
above on some neighborhood about x. Consequently, l is continuous at x by 7.5.1.
From 5.3.7 derive that l is continuous.
7.5.3. Corollary. Every convex function on a nite-dimensional space is con-
tinuous on the interior of its domain.
7.5.4. Definition. A function f : X R

is called ideally convex if epi f is


an ideal correspondence.
118 Chapter 7
7.5.5. Automatic Continuity Principle Every ideally convex function on a Ba-
nach space is continuous on the core of its domain.
Let f be such function. If core domf = then there is nothing to prove.
If x core domf then put t := f(x) and F := (epi f)
1
RX. Applying the Ideal
Correspondence Principle, nd a > 0 from the condition F(t + B
R
) x + B
X
.
Whence, in particular, infer the estimate f(x + B
X
) t + 1. In virtue of 7.5.1,
f is continuous on int domf. Since x int domf; therefore, by Lemma 7.1.1,
core domf = int domf.
7.5.6. Remark. Using 7.3.6, it is possible to prove that an ideally convex
function f, dened in a Banach space on a subset with nonempty core, is locally
Lipschitz on int domf. In other words, given x
0
int domf, there are a positive
number L and a neighborhood U about x
0
such that |f(x) f(x
0
)| L|x x
0
|
whenever x U.
7.5.7. Corollary. Let f : X R

be an ideally convex function on a Banach


space X and x core domf. Then the directional derivative f

(x) is a continuous
sublinear functional and
x
(f) X

.
Apply the Automatic Continuity Principle twice.
7.5.8. Remark. In view of 7.5.7, in studying a Banach space X, only contin-
uous functionals on X are usually admitted into the subdierential of a function
f : X R

at a point x; i.e., we agree to dene

x
(f):=
x
(f) X

.
Proceed likewise in (multi)normed spaces. If a need is felt to distinguish the old
(wider) subdierential, a subset of X
#
, from the new (narrower) subdierential,
a subset of X

; then the rst is called algebraic, whereas the second is called topo-
logical. With this in mind, we refer to the facts indicated in 7.5.2 and 7.5.7 as to the
Coincidence Principle for algebraic and topological subdierentials. Observe nally
that if f := p is a seminorm on X then, for similar reasons, it is customary to put
[[(p):= [[(p) X

.
7.5.9. Ideal HahnBanach Theorem. Let f : Y R

be an ideally convex
function on a Banach space Y . Further, let X be a normed space and T B(X, Y ).
If a point x in X is such that Tx core domf then

x
(f T) =
Tx
(f) T.
The right side of the sought formula is included into its left side for obvious
reasons. If l in X

belongs to
x
(f T), then by the HahnBanach Theorem there
is an element l
1
of the algebraic subdierential of f at Tx for which l = l
1
T.
It suces to observe that, in virtue of 7.5.7, l
1
is an element of Y

and so it is
a member of the topological subdierential
Tx
(f).
Principles of Banach Spaces 119
7.5.10. Balanced HahnBanach Theorem. Suppose that X and Y are
normed spaces. Given T B(X, Y ), let p : Y R be a continuous seminorm.
Then
[[(p T) = [[(p) T.
If l [[(p T) then l = l
1
T for some l
1
in the algebraic balanced
subdierential of p (cf. 3.7.11). From 7.5.2 it follows that l
1
is continuous. Thus,
[[(p T) [[(p) T. The reverse inclusion raises no doubts.
7.5.11. Continuous Extension Principle. Let X
0
be a subspace of X and
let l
0
be a continuous linear functional on X
0
. Then there is a continuous linear
functional l on X extending l
0
and such that |l| = |l
0
|.
Take p := |l
0
| | |, and consider the identical embedding : X
0
X.
On account of 7.5.10, l
0
[[(p ) = [[(p) = |l
0
| [[(| |) . It suces to
observe that [[(| |
X
) = B
X
.
7.5.12. Topological Separation Theorem. Let U be a convex set with
nonempty interior in a space X. If L is an ane variety in X and L int U =
then there is a closed hyperplane H in X such that H L and H int U = .
7.5.13. Remark. When applying Theorem 7.5.12, it is useful to bear in mind
that a closed hyperplane is precisely a level set of a nonzero continuous linear
functional.
7.5.14. Corollary. Let X
0
be a subspace of X. Then
cl X
0
= ker f : f X

, ker f X
0
.
It is clear that (f X

& ker f X
0
) ker f cl X
0
. If x
0
/ cl X
0
then there is an open convex neighborhood about x
0
disjoint from cl X
0
. In virtue
of 7.5.12 and 7.5.13 there is a functional f
0
, a member of (X
R
)

, such that ker f


0

cl X
0
and f
0
(x
0
) = 1. From the properties of the complexier infer that the func-
tional Re
1
f
0
vanishes on X
0
and diers from zero at the point x
0
. It is also beyond
a doubt that the functional is continuous.
7.6. Prime Principles
7.6.1. Let X and Y be (multi)normed vector spaces (over the same ground
eld F). Assume further that X

and Y

are the duals of X and Y respectively.


Take a continuous linear operator T from X to Y . Then y

T X

for y

and the mapping y

T is a linear operator.
7.6.2. Definition. The operator T

: Y

, constructed in 7.6.1, is the


dual or transpose of T : X Y .
120 Chapter 7
7.6.3. Theorem. The prime mapping T T

implements a linear isometry


of the space B(X, Y ) into the space B(Y

, X

).
The prime mapping is clearly a linear operator from B(X, Y ) to L(Y

, X

).
Furthermore, since |y| = sup[l(y) : l [[(| |); therefore,
|T

| = sup|T

| : |y

| 1
= sup[y

(Tx)[ : |y

| 1, |x| 1 = sup|Tx| : |x| 1 = |T|,


what was required.
7.6.4. Examples.
(1) Let X and Y be Hilbert spaces. Take T B(X, Y ). Observe rst
that, in a plain sense, T B(X, Y ) T B(X

, Y

). Denote the prime mapping


of X by ()

X
: X

, i.e., x x

:= ( , x); and denote the prime mapping of Y


by ()

Y
: Y

, i.e., y y

:= ( , y).
The adjoint of T, the member T

of B(Y, X), and the dual of T, the mem-


ber T

of B(Y

, X

), are related by the commutative diagram:


X

()

X
()

Y
X

Indeed, it is necessary to show the equality T

= (T

y)

for y Y . Given
x X, by denition observe that
T

(x) = y

(Tx) = (Tx, y) = (x, T

y) = (T

y)

(x).
Since x is arbitrary, the proof is complete.
(2) Let : X
0
X be the identical embedding of X
0
into X. Then

: X

0
. Moreover,

(x

)(x
0
) = x

(x
0
) for all x
0
X
0
and x

and

is
an epimorphism; i.e., X

0
0 is an exact sequence.
7.6.5. Definition. Let an elementary diagram X
T
Y be given. The dia-
gram Y

is referred to as resulting from setting primes or as the diagram


prime of the original diagram or as the dual diagram. If primes are set in every
elementary subdiagram in an arbitrary diagram composed of bounded linear op-
erators in Banach spaces, then the so-obtained diagram is referred to as dual or
resulting from setting primes in the original diagram. The term diagram prime
is used for suggestiveness.
Principles of Banach Spaces 121
7.6.6. Double Prime Lemma. Let X

be the diagram that results


from setting primes in the diagram X
T
Y twice. Then the following diagram
commutes:
X
T
Y

Here

: X X

and

: Y Y

are the respective double prime mappings; i.e.


the canonical embeddings of X into X

and of Y into Y

(cf. 5.1.10 (8)).


Let x X. We have to prove that T

= (Tx)

. Take y

. Then
T

(y

) = x

(T

) = T

(x) = y

(Tx) = (Tx)

(y

).
Since y

is arbitrary, the proof is complete.


7.6.7. Diagram Prime Principle. A diagram is commutative if and only
if so is its diagram prime.
It suces to convince oneself that the triangles
X
T
Y
R S
Z
X

are commutative or not simultaneously. Since R = ST R

= (ST)

= T

;
therefore, the commutativity of the triangle on the left entails the commutativity
of the triangle on the right. If the latter commutes then by what was already
proven R

= S

. Using 7.6.6, argue as follows: (Rx)

= R

= S

=
S

(T

) = S

(Tx)

= (STx)

for all x X. Consequently, R = ST.


7.6.8. Definition. Let X
0
be a subspace of X and let X
0
be a subspace
of X

. Put
X

0
:= f X

: ker f X
0
= [[((X
0
));

X
0
:= x X : f X
0
f(x) = 0 = ker f : f X
0
.
The subspace X

0
is the (direct) polar of X
0
, and the subspace

X
0
is the reverse
polar of X
0
. A less exact term annihilator is also in use.
7.6.9. Definition. Let X and Y be Banach spaces. An arbitrary element T
of B(X, Y ) is a normally solvable operator, if imT is a closed subspace of Y . The
natural term closed range operator is also in common parlance.
122 Chapter 7
7.6.10. An operator T, a member of B(X, Y ), is normally solvable if and only
if T is a homomorphism, when regarded as acting from X to imT.
: The Banach Homomorphism Theorem.
: Refer to 7.4.2.
7.6.11. Polar Lemma. Let T B(X, Y ). Then
(1) (imT)

= ker(T

);
(2) if T is normally solvable then
imT =

ker(T

), (ker T)

= im(T

).
(1): y

ker(T

) T

= 0
( x X) T

(x) = 0 ( x X) y

(Tx) = 0 y

(imT)

.
(2): The equality cl imT =

ker(T

) follows from 7.5.13. Furthermore, by hy-


pothesis imT is closed.
If x

= T

and Tx = 0 then x

(x) = T

(x) = y

(Tx) = 0, which means


that x

(ker T)

. Consequently, im(T

) (ker T)

. Now take x

(ker T)

.
Considering the operator T acting onto imT, apply the Sard Theorem to the left
side of the diagram
X
T
imT Y
x

0
y

F
As a result of this, obtain y

0
in (imT)

such that y

0
T = x

. By the Continuous
Extension Principle there is an element y

of Y

satisfying y

0
. Thus, x

= T

,
i.e. x

im(T

).
7.6.12. Hausdor Theorem. Let X and Y be Banach spaces. Assume fur-
ther that T B(X, Y ). Then T is normally solvable if and only if T

is normally
solvable.
: In virtue of 7.6.11 (2), im(T

) = (ker T)

. Evidently, the subspace


(ker T)

is closed.
: To begin with, suppose cl imT = Y . It is clear that 0 = Y

= (cl im T)

= (imT)

= ker(T

) in virtue of 7.6.11. By the Banach Isomorphism Theorem


there is some S B(im(T

), Y

) such that ST

= I
Y
. The case r := |S| = 0 is
trivial. Therefore, it may be assumed that |T

|
1
/
r
|y

| for all y

.
Show now that cl T(B
X
)
1
/
2r
B
Y
. With this at hand, from the ideal con-
vexity of T(B
X
) it is possible to infer that T(B
X
)
1
/
4r
B
Y
. The last inclusion
implies that T is a homomorphism.
Let y / cl T(B
X
). Then y does not belong to some open convex set including
T(B
X
). Passing, if need be, to the real carriers of X and Y , assume that F:= R.
Principles of Banach Spaces 123
Applying the Topological Separation Theorem, nd some nonzero y

in Y

satisfying
|y

| |y| y

(y) sup
|x|1
y

(Tx) = |T

|
1
/
r
|y

|.
Whence |y|
1
/
r
>
1
/
2r
. Therefore, the sought inclusion is established and the
operator T is normally solvable under the above supposition.
Finally, address the general case. Put Y
0
:= cl imT and let : Y
0
Y be
the identical embedding. Then T = T, where T : X Y
0
is the operator acting
by the rule Tx = Tx for x X. In addition, im(T

) = im(T

) = T

(im(

)) =
T

(Y

0
), because

(Y

) = Y

0
(cf. 7.6.4 (2)). Thus, T

is a normally solvable operator.


By what was proven, T is normally solvable. It remains to observe that imT =
imT.
7.6.13. Sequence Prime Principle. A sequence
. . . X
k1
T
k
X
k
T
k+1
X
k+1
. . .
is exact if and only if so is the sequence prime
. . . X

k1
T

k
X

k
T

k+1
X

k+1
. . . .
: Since imT
k+1
= ker T
k+2
; therefore, T
k+1
is normally solvable. Using
the Polar Lemma, conclude that
ker(T

k
) = (imT
k
)

= (ker T
k+1
)

= im(T

k+1
).
: By the Hausdor Theorem T
k+1
is normally solvable. Once again appealing
to 7.6.11 (2), infer that
(imT
k
)

= ker(T

k
) = im(T

k+1
) = (ker T
k+1
)

.
Since T
k
is normally solvable by Theorem 7.6.12; therefore, imT
k
is a closed sub-
space. Using 7.5.14, observe that
imT
k
=

((imT
k
)

) =

((ker T
k+1
)

) = ker T
k+1
.
Here account was taken of the fact that ker T
k+1
itself is a closed subspace.
7.6.14. Corollary. For a normally solvable operator T, the following isomor-
phisms hold: (ker T)

coker(T

) and (coker T)

ker(T

).
By virtue of 2.3.5 (6) the sequence
0 ker T X
T
Y coker T 0
is exact. From 7.6.13 obtain that the sequence
0 (coker T)

(ker T)

0
is exact.
7.6.15. Corollary. T is an isomorphism T

is an isomorphism.
7.6.16. Corollary. Sp(T) = Sp(T

).
124 Chapter 7
Exercises
7.1. Find out which linear operators are ideal.
7.2. Establish that a separately continuous bilinear form on a Banach space is jointly con-
tinuous.
7.3. Is a family of lower semicontinuous sublinear functionals on a Banach space uniformly
bounded on the unit ball?
7.4. Let X and Y be Banach spaces and T : X Y . Prove that |Tx|
Y
t|x|
X
for some
strictly positive t and all x X if and only if ker T = 0 and imT is a complete set.
7.5. Find conditions for normal solvability of the operator of multiplication by a function
in the space of continuous functions on a compact set.
7.6. Let T be a bounded epimorphism of a Banach space X onto l
1
(E). Show that ker T is
complemented.
7.7. Establish that a uniformly closed subspace of C([a, b]) composed of continuously dif-
ferentiable functions (i.e., elements of C
(1)
([a, b])) is nite-dimensional.
7.8. Let X and Y be dierent Banach spaces, with X continuously embedded into Y . Es-
tablish that X is a meager subset of Y .
7.9. Let X
1
and X
2
be nonzero closed subspaces of a Banach space and X
1
X
2
= 0. Prove
that the sum X
1
+X
2
is closed if and only if the next quantity
inf|x
1
x
2
|/|x
1
| : x
1
,= 0, x
1
X
1
, x
2
X
2

is strictly positive.
7.10. Let (a
mn
) be a double countable sequence such that there is a sequence (x
(m)
) of
elements of l
1
for which all the series

n=1
a
mn
x
(m)
n
fail to converge in norm. Prove that there
is a sequence x in l
1
such that the series

n=1
a
mn
x
n
fail to converge in norm for all m N.
7.11. Let T be an endomorphism of a Hilbert space H which satises Tx[ y) = x [ Ty) for
all x, y H. Establish that T is bounded.
7.12. Let a closed cone X
+
in a Banach space X be reproducing: X = X
+
X
+
. Prove
that there is a constant t > 0 such that for all x X and each presentation x = x
1
x
2
with
x
1
, x
2
X
+
, the estimates hold: |x
1
| t|x| and |x
2
| t|x|.
7.13. Let lower semicontinuous sublinear functionals p and q on a Banach space X be such
that the cones domp and domq are closed and the subspace domp domq = domq domp is
complemented in X. Prove that
(p +q) = (p) +(q)
in the case of topological subdierentials (cf. Exercise 3.10).
7.14. Let p be a continuous sublinear functional dened on a normed space X and let T be
a continuous endomorphism of X. Assume further that the dual T

of T takes the subdierential


(p) into itself. Establish that (p) contains a xed point of T

.
7.15. Given a function f : X R

on a (multi)normed space X, put


f

(x

):= supx [ x

) f(x) : x domf (x

);
f

(x):= supx [ x

) f

(x

) : x

dom(f

) (x X).
Find conditions for f to satisfy f = f

.
Principles of Banach Spaces 125
7.16. Establish that l

is complemented in each ambient Banach space.


7.17. A Banach space X is called primary, if each of its innite-dimensional complemented
subspaces is isomorphic to X. Verify that c
0
and l
p
(1 p +) are primary.
7.18. Let X and Y be Banach spaces. Take an operator T, a member of B(X, Y ), such
that imT is nonmeager. Prove that T is normally solvable.
7.19. Let X
0
be a closed subspace of a normed space X. Assume further that X
0
and X/X
0
are Banach spaces. Show that X itself is a Banach space.
Chapter 8
Operators in Banach Spaces
8.1. Holomorphic Functions and Contour Integrals
8.1.1. Definition. Let X be a Banach space. A subset of the ball B
X

in the dual space X

is called norming (for X) if |x| = sup[l(x)[ : l


for all x X. If each subset U of X satises sup |U| < + on condition that
sup[l(u)[ : u U < + for all l , then is a fully norming set.
8.1.2. Examples.
(1) The ball B
X
is a fully norming set in virtue of 5.1.10 (8) and 7.2.7.
(2) If
0
is a (fully) norming set and
0

1
B
X
then
1
itself is
a (fully) norming set.
(3) The set ext B
X
of the extreme points of B
X
is norming in virtue
of the KrenMilman Theorem in subdierential form and the obvious equality
B
X
= [[(| |
X
) which has already been used many times. However, ext B
X

can fail to be fully norming (in particular, the possibility is realized in the space
C([0, 1], R)).
(4) Let X and Y be Banach spaces (over the same ground eld F) and
let
Y
be a norming set for Y . Put

B
:=
(y

, x)
: y


Y
, x B
X
,
where
(y

, x)
(T):= y

(Tx) for y

Y, x X and T B(X, Y ). It is clear that


[
(y

, x)
(T)[ = [y

(Tx)[ |y

| |Tx| |y

| |T| |x|;
i.e.,
(y

, x)
B(X, Y )

. Furthermore, given T B(X, Y ), infer that


|T| = sup|Tx| : |x| 1 = sup[y

(Tx)[ : y


Y
, |x| 1
= sup[
(y

, x)
(T)[ :
(y

, x)

B
.
Operators in Banach Spaces 127
Therefore,
B
is a norming set for B(X, Y ). If
Y
is a fully norming set then
B
is also a fully norming set. Indeed, if U is such that the numeric set [y

(Tx)[ :
T U is bounded in R for all x B
X
and y


Y
, then by hypothesis the
set Tx : T U is bounded in Y for every x in X. By virtue of the Uniform
Boundedness Principle it means that sup |U| < +.
8.1.3. DunfordHille Theorem. Let X be a complex Banach space and let
be a fully norming set for X. Assume further that f : D X is an X-va-
lued function with domain D an open (in C
R
R
2
) subset of C. The following
statements are equivalent:
(1) for every z
0
in D there is a limit
lim
zz
0
f(z) f(z
0
)
z z
0
;
(2) for all z
0
D and l there is a limit
lim
zz
0
l f(z) l f(z
0
)
z z
0
;
i.e., the function l f : D C is holomorphic for l .
(1) (2): This is obvious.
(2) (1): For the sake of simplicity assume that z
0
= 0 and f(z
0
) = 0. Con-
sider the disk of radius 2 centered at zero and included in D; i.e., 2D D, where
D:= B
C
:= z C : [z[ 1 is the unit disk. As is customary in complex analysis,
treat the disk D as an (oriented) compact manifold with boundary T, where T is
the (properly oriented) unit circle T:= z C : [z[ = 1 in the complex plane C
R
.
Take z
1
, z
2
D 0 and the holomorphic function l f (the functional l lies in ).
Specifying the Cauchy Integral Formula, observe that
l f(z
k
)
z
k
=
1
2i
_
2T
l f(z)
z(z z
k
)
dz (k:= 1, 2).
If now z
1
,= z
2
then, using the condition [z z
k
[ (k := 1, 2) for z 2T and
the continuity property of the function l f on D, nd

l
_
1
z
1
z
2
_
f(z
1
)
z
1

f(z
2
)
z
2
__

1
z
1
z
2

1
2i
_
2T
l f(z)
_
1
z(z z
1
)

1
z(z z
2
)
_
dz

=
1
2

_
2T
l f(z)
1
z(z z
1
)(z z
2
)
dz

M sup
z2T
[l f(z)[ < +
128 Chapter 8
for a suitable M > 0. Since is a fully norming set, conclude that
sup
z
1
,=z
2
;z
1
,z
2
,=0
]z
1
],]z
2
]
1
[z
1
z
2
[
_
_
_
_
f(z
1
)
z
1

f(z
2
)
z
2
_
_
_
_
< +.
This nal inequality ensures that the sought limit exists.
8.1.4. Definition. A mapping f : D X satisfying 8.1.3 (1) (or, which is
the same, 8.1.3 (2) for some fully norming set ) is called holomorphic.
8.1.5. Remark. A meticulous terminology is used sometimes. Namely, if f
satises 8.1.3 (1) then f is called a strongly holomorphic function. If f satises
8.1.3 (2) with := B
X
then f is called weakly holomorphic. Under the hypotheses
of 8.1.3 (2) and 8.1.2 (4), i.e. for f : D B(X, Y ),
Y
:= B
Y
and the correspond-
ing :=
B
, the expression, f is weakly operator holomorphic, is employed. With
regard to this terminology, the DunfordHille Theorem is often referred to as the
Holomorphy Theorem and verbalized as follows: A weakly holomorphic function
is strongly holomorphic.
8.1.6. Remark. It is convenient in the sequel to use the integrals of the
simplest smooth X-valued forms f(z)dz over the simplest oriented manifolds, the
boundaries of elementary planar compacta (cf. 4.8.5) which are composed of nitely
many disjoint simple loops. An obvious meaning is ascribed to the integrals: Name-
ly, given a loop , choose an appropriate (smooth) parametrization : T (with
orientation accounted for) and put
_

f(z)dz :=
_
T
f d,
with the integral treated for instance as a suitable Bochner integral (cf. 5.5.9 (6)).
The soundness of the denition is beyond a doubt, since the needed Bochner integral
exists independently of the choice of the parametrization .
8.1.7. CauchyWiener Integral Theorem. Let D be an nonempty open
subset of the complex plane and let f : D X be a holomorphic X-valued function,
with X a Banach space. Assume further that F is a rough draft for the pair (, D).
Then
_
F
f(z)dz = 0.
Moreover,
f(z
0
) =
1
2i
_
F
f(z)
z z
0
dz
for z
0
int F.
Operators in Banach Spaces 129
By virtue of 8.1.3 the Bochner integrals exist. The sought equalities follow
readily from the validity of their scalar versions basing on the Cauchy Integral
Formula, the message of 8.1.2 (1) and the fact that the Bochner integral commutes
with every bounded functional as mentioned in 5.5.9 (6).
8.1.8. Remark. The CauchyWiener Integral Theorem enables us to infer
analogs of the theorems of classical complex analysis for X-valued holomorphic
functions on following the familiar patterns.
8.1.9. Taylor Series Expansion Theorem. Let f : D X be a holomor-
phic X-valued function, with X a Banach space, and take z
0
D. In every open
disk U := z C : [z z
0
[ < such that cl U lies in D, the Taylor series expansion
holds (in a compactly convergent power series, cf. 7.2.10):
f(z) =

n=0
c
n
(z z
0
)
n
,
where the coecients c
n
, members of X, are calculated by the formulas:
c
n
=
1
2i
_
U
f(z)
(z z
0
)
n+1
dz =
1
n!
d
n
f
dz
n
(z
0
).
The proof results from a standard argument: Expand the kernel u
(u z)
1
of the formula
f(z) =
1
2i
_
U

f(u)
u z
du (z cl U)
in the powers of z z
0
; i.e.,
1
u z
=
1
(u z
0
)
_
1
zz
0
uz
0
_ =

n=0
(z z
0
)
n
(u z
0
)
n+1
.
The last series converges uniformly in u U

. (Here U

= U + qD for some
q > 0 such that cl U

D.) Taking it into account that sup |f(U

)| < +
and integrating, arrive at the sought presentation of f(z) for z cl U. Applying
this to U

and using 8.1.7, observe that the power series under study converges
in norm at every point of U

. This yields uniform convergence on every compact


subset of U

, and so on U.
130 Chapter 8
8.1.10. Liouville Theorem. If an X-valued function f : C X, with X
a Banach space, is holomorphic and sup |f(C)| < + then f is a constant map-
ping.
Considering the disk D with > 0 and taking note of 8.1.9, infer that
|c
n
| sup
zT
|f(z)|
n
sup |f(C)|
n
for all n N and > 0. Therefore, c
n
= 0 for n N.
8.1.11. Each bounded endomorphism of a nonzero complex Banach space has
a nonempty spectrum.
Let T be such an endomorphism. If Sp(T) = then the resolvent R(T, )
is holomorphic on the entire complex plane C, for instance, by 5.6.21. Further-
more, by 5.6.15, |R(T, )| 0 as [[ +. By virtue of 8.1.10 conclude that
R(T, ) = 0. At the same time, using 5.6.15, observe that R(T, )( T) = 1 for
[[ > |T|. A contradiction.
8.1.12. The BeurlingGelfand formula holds:
r(T) = sup[[ : Sp(T)
for all T B(X), with X a complex Banach space; i.e., the spectral radius of
an operator T coincides with the radius of the spectrum of T.
It is an easy matter that the spectral radius r(T) is greater than the radius
of the spectrum of T. So, there is nothing to prove if r(T) = 0. Assume now that
r(T) > 0. Take C so that [[ > sup[[ : Sp(T). Then the disk of radius
[[
1
lies entirely in the domain of the holomorphic function (cf. 5.6.15)
f(z):=
_
R(T, z
1
), for z ,= 0 and z
1
res(T)
0, for z = 0.
Using 8.1.9 and 5.6.17, conclude that [[
1
< r(T)
1
. Consequently, [[ > r(T).
8.1.13. Let K be a nonempty compact subset of C. Denote by H(K) the set of
all functions holomorphic in a neighborhood of K (i.e., f H(K) f : domf C
is a holomorphic function with domf K). Given f
1
, f
2
H(K), let the notation
f
1
f
2
mean that it is possible to nd an open subset D of domf
1
domf
2
satisfying K D and f
1
[
D
= f
2
[
D
. Then is an equivalence in H(K).
8.1.14. Definition. Under the hypotheses of 8.1.13, put H (K):= H(K)/.
The element f in H (K) containing a function f in H(K) is the germ of f on K.
Operators in Banach Spaces 131
8.1.15. Let f, g H (K). Take f
1
, f
2
f and g
1
, g
2
g. Put
x domf
1
domg
1

1
(x):= f
1
(x) +g
1
(x),
x domf
2
domg
2

2
(x):= f
2
(x) +g
2
(x).
Then
1
,
2
H(K) and
1
=
2
.
Choose open sets D
1
and D
2
such that K D
1
domf
1
domf
2
and
K D
2
domg
1
domg
2
, with f
1
[
D
1
= f
2
[
D
1
and g
1
[
D
2
= g
2
[
D
2
. Observe now
that
1
and
2
agree on D
1
D
2
.
8.1.16. Definition. The coset, introduced in 8.1.15, is the sum of the germs
f
1
and f
2
. It is denoted by f
1
+ f
2
. The product of germs and multiplication of
a germ by a complex number are introduced by analogy.
8.1.17. The set H (K) with operations dened in 8.1.16 is an algebra.
8.1.18. Definition. The algebra H (K) is the algebra of germs of holomor-
phic functions on a compact set K.
8.1.19. Let K be a compact subset of C, and let R : C K X be an
X-valued holomorphic function, with X a Banach space. Further, take f H (K)
and f
1
, f
2
f. If F
1
is a rough draft for the pair (K, domf
1
) and F
2
is a rough
draft for the pair (K, domf
2
) then
_
F
1
f
1
(z)R(z)dz =
_
F
2
f
2
(z)R(z)dz.
Let K D int F
1
int F
2
, with D open and f
1
[
D
= f
2
[
D
. Choose a rough
draft F for the pair (K, D). Since f
1
R is holomorphic on domf
1
K and f
2
R is
holomorphic on domf
2
K, infer the equalities
_
F
f
1
(z)R(z)dz =
_
F
1
f
1
(z)R(z)dz,
_
F
f
2
(z)R(z)dz =
_
F
2
f
2
(z)R(z)dz
(from the nontrivial fact of the validity of their scalar analogs). Since f
1
and f
2
agree on D, the proof is complete.
132 Chapter 8
8.1.20. Definition. Under the hypotheses of 8.1.19, given an element h
in H (K), dene the contour integral of h with kernel R as the element
_
h(z)R(z)dz :=
_
F
f(z)R(z)dz,
where h = f and F is a rough draft for the pair (K, domf).
8.1.21. Remark. The notation h(z) in 8.1.20 is far from being ad hoc. It is
well justied by the fact that w:= f
1
(z) = f
2
(z) for every point z in K and every
two members f
1
and f
2
of a germ h. In this connection the element w is said to
be the value of h at z, which is expressed in writing as h(z) = w. It is also worth
noting that the function R in 8.1.2 may be assumed to be given only in U K,
where int U K.
8.2. The Holomorphic Functional Calculus
8.2.1. Definition. Let X be a (nonzero) complex Banach space and let T be
a bounded endomorphism of X; i.e., T B(X). For h H (Sp(T)), the contour
integral with kernel the resolvent R(T, ) of T is denoted by
R
T
h:=
1
2i
_
h(z)R(T, z)dz
and called the RieszDunford integral (of the germ h). If f is a function holomorphic
in a neighborhood about Sp(T) then put f(T) := R
T
f := R
T
f. We also use more
suggestive designations like
f(T) =
1
2i
_
f(z)
z T
dz.
8.2.2. Remark. In algebra, in particular, various representations of mathe-
matical objects are under research. It is convenient to use the primary notions of
representation theory for the most algebraic objects, namely, algebras. Recall
the simplest of them.
Let A
1
and A
2
be two algebras (over the same eld). A morphism from A
1
to A
2
or a representation of A
1
in A
2
(rarely, over A
2
) is a multiplicative linear
operator R, i.e. a member R of L(A
1
, A
2
) such that R(ab) = R(a)R(b) for all
a, b A
1
. The expression, T represents A
1
in A
2
, is also in common parlance.
A representation R is called faithful if ker R = 0. The presence of a faithful repre-
sentation R : A
1
A
2
makes it possible to treat A
1
as a subalgebra of A
2
.
Operators in Banach Spaces 133
If A
2
is a (sub)algebra of the endomorphism algebra L(X) of some vector
space X (over the same eld), then a morphism of A
1
in A
2
is referred to as
a (linear) representation of A
1
on X or as an operator representation of A
1
. The
space X is then called the representation space for the algebra A
1
.
Given a representation R, suppose that the representation space X for A has
a subspace X
1
invariant under all operators R(a), a A. Then the representation
R
1
: A L(X
1
) arises naturally, acting by the rule R
1
(a)x
1
= R(a)x
1
for x
1
X
1
and a A and called a subrepresentation of R (induced in X
1
). If X = X
1
X
2
and the decomposition reduces each operator R(a) for a A, then it is said that
the representation R reduces to the direct sum of subrepresentations R
1
and R
2
(induced in X
1
and X
2
). We mention in passing the signicance of studying arbi-
trary irreducible representations, each of which has only trivial subrepresentations
by denition.
8.2.3. GelfandDunford Theorem. Let T be a bounded endomorphism
of a Banach space X. The RieszDunford integral R
T
represents the algebra of
germs of holomorphic functions on the spectrum of T on the space X. Moreover,
if f(z) =

n=0
c
n
z
n
(in a neighborhood about Sp(T)) then f(T) =

n=0
c
n
T
n
(summation is understood in the operator norm of B(X)).
There is no doubt that R
T
is a linear operator. Check the multiplicativity
property of R
T
. To this end, take f
1
, f
2
H (Sp(T)) and choose rough drafts F
1
and F
2
such that Sp(T) int F
1
F
1
int F
2
F
2
D, with the functions f
1
in
f
1
and f
2
in f
2
holomorphic on D.
Using the obvious properties of the Bochner integral, the Cauchy Integral For-
134 Chapter 8
mula and the Hilbert identity, successively infer the chain of equalities
R
T
f
1
R
T
f
2
= f
1
(T)f
2
(T) =
1
2i
1
2i
_
F
1
f
1
(z
1
)
z
1
T
dz
1

_
F
2
f
2
(z
2
)
z
2
T
dz
2
=
1
2i
1
2i
_
F
2
_
_
_
F
1
f
1
(z
1
)R(T, z
1
)dz
1
_
_
f
2
(z
2
)R(T, z
2
)dz
2
=
1
2i
1
2i
_
F
1
_
F
2
f
1
(z
1
)f
2
(z
2
)R(T, z
1
)R(T, z
2
)dz
2
dz
1
=
1
2i
1
2i
_
F
1
_
F
2
f
1
(z
1
)f
2
(z
2
)
R(T, z
1
) R(T, z
2
)
z
2
z
1
dz
2
dz
1
=
1
2i
_
F
1
f
1
(z
1
)
_
_
1
2i
_
F
2
f
2
(z
2
)
z
2
z
1
dz
2
_
_
R(T, z
1
)dz
1

1
2i
_
F
2
f
2
(z
2
)
_
_
1
2i
_
F
1
f
1
(z
1
)
z
2
z
1
dz
1
_
_
R(T, z
2
)dz
2
=
1
2i
_
F
1
f
1
(z
1
)f
2
(z
1
)R(T, z
1
)dz
1
0 = f
1
f
2
(T) = R
T
(f
1
f
2
).
Choose a circle := T that lies in res(T) as well as in the (open) disk of
convergence of the series f(z) =

n=0
c
n
z
n
. From 5.6.16 and 5.5.9 (6) derive
f(T) =
1
2i
_

f(z)

n=0
z
n1
T
n
dz
=
1
2i

n=0
_

f(z)z
n1
T
n
dz =

n=0
_
_
1
2i
_

f(z)
z
n+1
dz
_
_
T
n
=

n=0
c
n
T
n
in virtue of 8.1.9.
8.2.4. Remark. Theorem 8.2.3 is often referred to as the Principal Theorem
of the holomorphic functional calculus.
8.2.5. Spectral Mapping Theorem. For every function f holomorphic in a neigh-
borhood about the spectrum of an operator T in B(X), the equality holds:
f(Sp(T)) = Sp(f(T)).
Operators in Banach Spaces 135
Assume rst that Sp(f(T)) and f
1
() Sp(T) = . Given z
(C f
1
()) domf, put g(z) := ( f(z))
1
. Then g is a holomorphic function
on a neighborhood of Sp(T), satisfying g( f) = ( f)g = 1
C
. Using 8.2.3,
observe that res(f(T)), a contradiction. Consequently, f
1
() Sp(T) ,= ;
i.e., Sp(f(T)) f(Sp(T)).
Now take Sp(T). Put
,= z g(z):=
f() f(z)
z
; g():= f

().
Clearly, g is a holomorphic function (the singularity is removed). From 8.2.3
obtain
g(T)( T) = ( T)g(T) = f() f(T).
Consequently, if f() res(f(T)) then the operator R(f(T), f())g(T) is inverse
to T. In other words, res(T), which is a contradiction. Thus,
f() C res(f(T)) = Sp(f(T));
i.e., f(Sp(T)) Sp(f(T)).
8.2.6. Let K be a nonempty compact subset of C and let g : domg C be
a holomorphic function with domg K. Given f H(g(K)), put

g(f) := f g.
Then

g is a representation of the algebra H (g(K)) in the algebra H (K).
8.2.7. Dunford Theorem. For every function g : domg C holomorphic
in the neighborhood domg about the spectrum Sp(T) of an endomorphism T of
a Banach space X, the following diagram of representations commutes:
B(X)
H (Sp(T)) H (Sp(g(T)))
R
g(T)

g
R
T

Let f H (g(Sp(T))) with f : D C such that f f and D g(Sp(T)) =


Sp(g(T)). Let F
1
be a rough draft for the pair (Sp(g(T)), D) and let F
2
be a rough
draft for the pair (Sp(T), g
1
(int F
1
)). It is clear that now g(F
2
) int F
1
and,
moreover, the function z
2
(z
1
g(z
2
))
1
is dened and holomorphic on int F
2
for z
1
F
1
. Therefore, by 8.2.3
R(g(T), z
1
) =
1
2i
_
F
2
R(T, z
2
)
z
1
g(z
2
)
dz
2
(z
1
F
1
).
136 Chapter 8
From this equality, successively derive
R
g(T)
f =
1
2i
_
F
1
f(z
1
)
z
1
g(T)
dz
1
=
1
2i
1
2i
_
F
1
f(z
1
)
_
_
_
F
2
R(T, z
2
)
z
1
g(z
2
)
dz
2
_
_
dz
1
=
1
2i
1
2i
_
F
2
_
_
_
F
1
f(z
1
)
z
1
g(z
2
)
dz
1
_
_
R(T, z
2
)dz
2
.
Since g(z
2
) int F
1
for z
2
F
2
by construction; therefore, the Cauchy Integral
Formula yields the equality
f(g(z
2
)) =
1
2i
_
F
1
f(z
1
)
z
1
g(z
2
)
dz
1
.
Consequently,
R
g(T)
f =
1
2i
_
F
2
f(g(z
2
))R(T, z
2
)dz
2
= R
T

g(f).
8.2.8. Remark. The Dunford Theorem is often referred to as the Composite
Function Theorem and written down symbolically as f g(T) = f(g(T)).
8.2.9. Definition. A subset of Sp(T) is a clopen or isolated part or rarely
an exclave of Sp(T), if and its complement

:= Sp(T) are closed.


8.2.10. Let be a clopen part of Sp(T) and let

be some function that


equals 1 in an open neighborhood of and 0 in an open neighborhood of

. Further,
assign
P

:=

(T):=
1
2i
_

(z)
z T
dz.
Then P

is a projection in X and the (closed) subspace X

:= imP

is invariant
under T.
Since
2

, it follows from 8.2.3 that

(T)
2
=

(T). Furthermore,
T = R
T
I
C
, where I
C
: z z. Hence TP

= P

T (because I
C

I
C
).
Consequently, in virtue of 2.2.9 (4), X

is invariant under T.
8.2.11. Definition. The projection P

is the Riesz projection or the Riesz


idempotent corresponding to .
Operators in Banach Spaces 137
8.2.12. Spectral Decomposition Theorem. Let be a clopen part of the
spectrum of an operator T in B(X). Then X splits into the direct sum decompo-
sition of the invariant subspaces X = X

which reduces T to matrix form


T
_
T

0
0 T

_
,
with the part T

of T in X

and the part T

of T in X

satisfying
Sp(T

) = , Sp(T

) =

.
Since

=
Sp(T)
= 1
C
, in view of 8.2.3 and 8.2.10 it suces to es-
tablish the claim about the spectrum of T

.
From 8.2.5 and 8.2.3 obtain
0 =

I
C
(Sp(T)) = Sp(

I
C
(T)) = Sp(R
T
(

I
C
))
= Sp(R
T

R
T
I
C
) = Sp(P

T).
Moreover, in matrix form
P

T
_
T

0
0 0
_
.
Let be a nonzero complex number. Then
P

T
_
T

0
0
_
;
i.e., the operator P

T is not invertible if and only if the same is true of the


operator T

. Thus,
Sp(T

) 0 Sp(P

T) 0 = ( 0) 0 .
Suppose that 0 Sp(T

) and 0 / . Choose disjoint open sets D

and D

so
that D

, 0 / D

and

, and put
z D

h(z):=
1
z
;
z D

h(z):= 0.
By 8.2.3, h(T)T = Th(T) = P

. Moreover, since h

h, the decomposition
X = X

reduces h(T) and h(T)

= T

h(T)

= 1 for the part h(T)

of h(T)
in X

. So, T

is invertible; i.e., 0 / Sp(T

). We thus arrive at a contradiction which


implies that 0 . In other words, Sp(T

) .
Observe now that res(T) = res(T

) res(T

). Consequently, by the above


Sp(T) = C res(T) = C (res(T

) res(T

))
= (C res(T

)) (C res(T

)) = Sp(T

) Sp(T

= Sp(T).
Considering that

= , complete the proof.


138 Chapter 8
8.2.13. RieszDunford Integral Decomposition Let be
a clopen part of Sp(T) for an endomorphism T of a Banach space X. The direct
sum decomposition X = X

reduces the representation R


T
of the algebra
H (Sp(T)) in X to the direct sum of the representations R

and R

. Moreover,
the following diagrams of representations commute:
B(X

)
H (Sp(T)) H ()
R
T

B(X

)
H (Sp(T)) H (

)
R
T

Here

(f) :=

f and

(f) :=

f for f H(Sp(T)) are the representations


induced by restricting f onto and

.
8.3. The Approximation Property
8.3.1. Let X and Y be Banach spaces. For K L(X, Y ) the following
statements are equivalent:
(1) the operator K is compact: K K (X, Y );
(2) there are a neighborhood of zero U in X and a compact subset V
of Y such that K(U) V ;
(3) the image under K of every bounded set in X is relatively compact
in Y ;
(4) the image under K of every bounded set in X is totally bounded
in Y ;
(5) for each sequence (x
n
)
nN
of points of the unit ball B
X
, the se-
quence (Kx
n
)
nN
has a Cauchy subsequence.
8.3.2. Theorem. Let X and Y be Banach spaces over a basic eld F. Then
(1) K (X, Y ) is a closed subspace of B(X, Y );
(2) for all Banach spaces W and Z, it holds that
B(Y, Z) K (X, Y ) B(W, X) K (W, Z);
i.e., if S B(W, X), T B(Y, Z) and K K (X, Y ) then TKS K (W, Z);
(3) I
F
K (F):= K (F, F).
Operators in Banach Spaces 139
That K (X, Y ) is a subspace of B(X, Y ) follows from 8.3.1. If K
n

K (X, Y ) and K
n
K; then, given > 0, for n suciently large observe that
|Kx K
n
x| |K K
n
| |x| whenever x B
X
. Therefore, K
n
(B
X
) serves
as an -net (= B

-net) for K(B


X
). It remains to refer to 4.6.4 and thus complete
the proof of the closure property of K (X, Y ). The other claims are evident.
8.3.3. Remark. Theorem 8.3.2 is often verbalized as follows: The class of
all compact operators is an operator ideal. Behind this lies a conspicuous analogy
with the fact that K (X) := K (X, X) presents a closed bilateral (two-sided) ideal
in the (bounded) endomorphism algebra B(X); i.e., K (X) B(X) K (X) and
B(X) K (X) K (X).
8.3.4. Calkin Theorem. The ideals 0, K (l
2
), and B(l
2
) exhaust the list of
closed bilateral ideals in the endomorphism algebra B(l
2
) of the Hilbert space l
2
.
8.3.5. Remark. In view of 8.3.4 it is clear that a distinguishable role in op-
erator theory should be performed by the algebra B(X)/K (X) called the Calkin
algebra (on X). The performance is partly delivered in 8.5.
8.3.6. Definition. An operator T, a member of L(X, Y ), is called a nite-
rank operator provided that T B(X, Y ) and imT is a nite-dimensional subspace
of Y . In notation: T F(X, Y ). A hasty term nite-dimensional operator would
abuse consistency since T as a subspace of X Y is usually innite-dimensional.
8.3.7. The linear span of the set of (bounded) rank-one operators comprises
all nite-rank operators:
T F(X, Y )
( x

1
, . . . , x

n
X

) ( y
1
, . . . , y
n
Y ) T =
n

k=1
x

k
y
k
.
8.3.8. Definition. Let Q be a (nonempty) compact set in X. Given T
B(X, Y ), put
|T|
Q
:= sup |T(Q)|.
The collection of all seminorms B(X, Y ) of type | |
Q
is the Arens multinorm
in B(X, Y ), denoted by
B(X,Y )
. The corresponding topology is the topology of
uniform convergence on compact sets or the compact-open topology (cf. 7.2.10).
8.3.9. Grothendieck Theorem. Let X be a Banach space. The following
conditions are equivalent:
(1) for every > 0 and every compact set Q in X there is a nite-rank
endomorphism T of X, a member of F(X) := F(X, X), such that |Tx x|
for all x Q;
140 Chapter 8
(2) for every Banach space W, the subspace F(W, X) is dense in the
space B(W, X) with respect to the Arens multinorm
B(W,X)
;
(3) for every Banach space Y , the subspace F(X, Y ) is dense in the
space B(X, Y ) with respect to the Arens multinorm
B(X,Y )
.
It is clear that (2) (1) and (3) (1). Therefore, we are to show only
that (1) (2) and (1) (3).
(1) (2): If T B(W, X) and Q is a nonempty compact set in W then,
in view of the Weierstrass Theorem, T(Q) is a nonempty compact set in X. So,
for > 0, by hypothesis there is a member T
0
of F(X) such that |T
0
I
X
|
T(Q)
=
|T
0
T T|
Q
. Undoubtedly, T
0
T F(W, X).
(1) (3): Let T B(X, Y ). If T = 0 then there is nothing to be proven.
Let T ,= 0, > 0 and Q be a nonempty compact set in X. By hypothesis there
is a member T
0
of F(X) such that |T
0
I
X
|
Q
|T|
1
. Then |TT
0
T|
Q

|T| |T
0
I
X
|
Q
. Furthermore, TT
0
F(X, Y ).
8.3.10. Definition. A Banach space satisfying one (and hence all) of the
equivalent conditions 8.3.9 (1)8.3.9 (3) is said to possess the approximation prop-
erty.
8.3.11. Grothendieck Criterion. A Banach space X possesses the approx-
imation property if and only if, for every Banach space W, the equality holds:
cl F(W, X) = K (W, X), with the closure taken in operator norm.
8.3.12. Remark. For a long time there was an unwavering (and yet unprov-
able) belief that every Banach space possesses the approximation property. There-
fore, P. Enos rather sophisticated example of a Banach space lacking the approx-
imation property was acclaimed as sensational in the late seventies of the current
century. Now similar counterexamples are in plenty:
8.3.13. Szankowski Counerexample. The space B(l
2
) lacks the approxi-
mation property.
8.3.14. DavisFigielSzankowski Couterexample. The spaces c
0
and l
p
with p ,= 2 have closed subspaces lacking the approximation property.
8.4. The RieszSchauder Theory
8.4.1. -Perpendicular Lemma. Let X
0
be a closed subspace of a Banach
space X and X ,= X
0
. Given an > 0, there is an -perpendicular to X
0
in X; i.e.,
an element x

in X such that |x

| = 1 and d(x

, X
0
):= inf d
||
(x

X
0
) 1.
Take 1 > and x XX
0
. It is clear that d:= d(x, X
0
) > 0. Find x

in the
subspace X
0
satisfying |xx

| d/(1), which is possible because d/(1) > d.


Operators in Banach Spaces 141
Put x

:= (x x

)|x x

|
1
. Then |x

| = 1. Finally, for x
0
X
0
, observe that
|x
0
x

| =
_
_
_
_
x
0

x x

|x x

|
_
_
_
_
=
1
|x

x|
|(|x x

|x
0
+x

) x|
d(x, X
0
)
|x

x|
1 .
8.4.2. Riesz Criterion. Let X be a Banach space. The identity operator
in X is compact if and only if X is nite-dimensional.
Only the implication needs proving. If X fails to be nite-dimensional,
then select a sequence of nite-dimensional subspaces X
1
X
2
. . . in X such
that X
n+1
,= X
n
for all n N. In virtue of 8.4.1 there is a sequence (x
n
) satisfying
x
n+1
X
n+1
, |x
n+1
| = 1 and d(x
n+1
, X
n
)
1
/
2
, namely, some sequence of
1
/
2
-
perpendiculars to X
n
in X
n+1
. It is clear that d(x
m
, x
k
)
1
/
2
for m ,= k. In other
words, the sequence (x
n
) lacks Cauchy subsequences. Consequently, by 8.3.1 the
operator I
X
is not compact.
8.4.3. Let T K (X, Y ), with X and Y Banach spaces. The operator T is
normally solvable if and only if T has nite rank.
Only the implication needs examining.
Let Y
0
:= imT be a closed subspace in Y . By the Banach Homomorphism
Theorem, the image T(B
X
) of the unit ball of X is a neighborhood of zero in Y
0
.
Furthermore, in virtue of the compactness property of T, the set T(B
X
) is relatively
compact in Y
0
. It remains to apply 8.4.2 to Y
0
.
8.4.4. Let X be a Banach space and K K (X). Then the operator 1 K is
normally solvable.
Put T := 1 K and X
1
:= ker T. Undoubtedly, X
1
is nite-dimensional
by 8.4.2. In accordance with 7.4.11 (1) a nite-dimensional subspace is comple-
mented. Denote a topological complement of X
1
to X by X
2
. Considering that X
2
is a Banach space and T(X) = T(X
2
), it suces to verify that |Tx| t|x| for
some t > 0 and all x X
2
. In the opposite case, there is a sequence (x
n
) such that
|x
n
| = 1, x
n
X
2
and Tx
n
0. Using the compactness property of K, we may
assume that (Kx
n
) converges. Put y := limKx
n
. Then the sequence (x
n
) converges
to y, because y = lim(Tx
n
+ Kx
n
) = limx
n
. Moreover, Ty = limTx
n
= 0; i.e.,
y X
1
. It is beyond a doubt that y X
2
. Thus, y X
1
X
2
; i.e., y = 0. We
arrive at a contradiction: |y| = lim|x
n
| = 1.
8.4.5. For whatever strictly positive , there are only nitely many eigenvalues
of a compact operator beyond the disk centered at zero and having radius .
142 Chapter 8
Suppose by way of contradiction that there is a sequence (
n
)
nN
of pair-
wise distinct eigenvalues of a compact operator K, with [
n
[ for all n N and
> 0. Suppose further that x
n
satisfying 0 ,= x
n
ker(
n
K) is an eigenvector
with eigenvalue
n
. Establish rst that the set x
n
: n N is linearly inde-
pendent. To this end, assume the set x
1
, . . . , x
n
linearly independent. In case
x
n+1
=

n
k=1

k
x
k
, we would have 0 = (
n+1
K)x
n+1
=

n
k=1

k
(
n+1

k
)x
k
.
Consequently,
k
= 0 for k:= 1, . . . , n. Whence the false equality x
n+1
= 0 would
ensue.
Put X
n
:= lin(x
1
, . . . , x
n
). By denition X
1
X
2
. . . ; moreover, as was
proven, X
n+1
,= X
n
for n N. By virtue of 8.4.1 there is a sequence (x
n
) such that
x
n+1
X
n+1
, |x
n+1
| = 1 and d(x
n+1
, X
n
)
1
/
2
. For m > k, straightforward
calculation shows that z := (
m+1
K)x
m+1
X
m
and z +Kx
k
X
m
+X
k
X
m
.
Consequently,
|Kx
m+1
Kx
k
| = |
m+1
x
m+1
+Kx
m+1
+
m+1
x
m+1
Kx
k
|
= |
m+1
x
m+1
(z +Kx
k
)| [
m+1
[d(x
m+1
, X
m
)

/
2
.
In other words, the sequence (Kx
n
) has no Cauchy subsequences.
8.4.6. Schauder Theorem. Let X and Y be Banach spaces (over the same
ground eld F). Then
K K (X, Y ) K

K (Y

, X

).
: Observe rst of all that the restriction mapping x

[
B
X
implements
an isometry of X

into l

(B
X
). Therefore, to check that K

(B
Y
) is relatively
compact we are to show the same for the set V := K

[
B
X
: y

B
Y
. Since
K

[
B
X
(x) = y

K[
B
X
(x) = y

(Kx) for x B
X
and y

B
Y
, consider the
compact set Q := cl K(B
X
) and the mapping

K : C(Q, F) l

(B
X
) dened
by the rule

Kg : x g(Kx). Undoubtedly, the operator

K is bounded and, hence,


continuous. Now put S := y

[
Q
: y

B
Y
. It is clear that S is simultaneously
an equicontinuous and bounded subset of C(Q, F). Consequently, by the Ascoli
Arzel`a Theorem, S is relatively compact. From the Weierstrass Theorem derive that

K(S) is a relatively compact set too. It remains to observe that

Ky

[
Q
= K

[
B
X
for y

B
Y
; i.e.,

K(S) = V .
: If K

K (Y

, X

) then, as is proven, K

K (X

, Y

). By the Double
Prime Lemma, K

[
X
= K. Whence it follows that the operator K is compact.
8.4.7. Every nonzero point of the spectrum of a compact operator is isolated
(i.e., such point constitutes a clopen part of the spectrum).
Operators in Banach Spaces 143
Taking note of 8.4.4 and the Sequence Prime Principle, observe that each
nonzero point of the spectrum of a compact operator K is either an eigenvalue of K
or an eigenvalue of the dual of K. Using 8.4.5 and 8.4.6, conclude that, for each
strictly positive , there are only nitely many points of Sp(K) beyond the disk
centered at zero and having radius .
8.4.8. RieszSchauder Theorem. The spectrum of a compact operator K
in an innite-dimensional space contains zero. Each nonzero point of the spectrum
of K is isolated and presents an eigenvalue of K with the corresponding eigenspace
nite-dimensional.
Considering K, a compact endomorphism of a Banach space X, we must
only demonstrate the implication
0 ,= Sp(K) ker( K) ,= 0.
First, settle the case F:= C. Note that is a clopen part of Sp(K). Putting
g(z) :=
1
/
z
in some neighborhood about and g(z) := 0 for z in a suitable
neighborhood about

, observe that
]
= gI
C
. Thus, by 8.2.3 and 8.2.10,
P
]
= g(K)K. By virtue of 8.3.2 (2), P
]
K (X). From 8.4.3 it follows that
imP
]
is a nite-dimensional space. It remains to invoke the Spectral Decompo-
sition Theorem.
In the case of the reals, F := R, implement the process of complexication.
Namely, furnish the space X
2
with multiplication by an element of C which is
introduced by the rule i(x, y) := (y, x). The resulting complex vector space
is denoted by X iX. Dene the operator K(x, y) := (Kx, Ky) in the space
X iX. Equipping X iX with an appropriate norm (cf. 7.3.2), observe that
the operator K is compact and Sp(K). Consequently, is an eigenvalue of K
by what was proven. Whence it follows that is an eigenvalue of K.
8.4.9. Theorem. Let X be a complex Banach space. Given T B(X),
assume further that f : C C is a holomorphic function vanishing only at zero
and such that f(T) K (X). Then every nonzero point of the spectrum of T is
isolated and the Riesz projection P
]
is compact.
Suppose the contrary; i.e., nd a sequence (
n
)
nN
of distinct points of Sp(T)
such that
n
,= 0 (in particular, X is innite-dimensional). Then f(
n
) f()
and f() ,= 0 by hypothesis. By the Spectral Mapping Theorem, Sp(f(T)) =
f(Sp(T)). Thus, by 8.4.8, f(
n
) = f() for all suciently large n. Whence it
follows that f(z) = f() for all z C and so f(T) = f(). By the Riesz Criterion
in this case X is nite-dimensional. We come to a contradiction meaning that
is an isolated point of Sp(T). Letting g(z) := f(z)
1
in some neighborhood about
disjoint from zero, infer that gf =
]
. Consequently, by the GelfandDunford
Theorem, P
]
= g(T)f(T); i.e., in virtue of 8.3.2 (2) the Riesz projection P
]
is
compact.
144 Chapter 8
8.4.10. Remark. Theorem 8.4.9 is sometimes referred to as the Generalized
RieszSchauder Theorem.
8.5. Fredholm Operators
8.5.1. Definition. Let X and Y be Banach spaces (over the same ground
eld F). An operator T, a member of B(X, Y ), is a Fredholm operator (in symbols,
T Fr(X, Y )) if ker T := T
1
(0) and coker T := Y/ imT are nite-dimensional;
i.e., if the following quantities, called the nullity and the deciency of T, are nite:
(T):= nul T := dim ker T; (T):= def T := dim coker T.
The integer ind T := (T)(T), a member of Z, is the index or, fully, the Fredholm
index of T.
8.5.2. Remark. In the Russian literature, a Fredholm operator is usually
called a Noether operator, whereas the term Fredholm operator is applied only
to an index-zero Fredholm operator.
8.5.3. Every Fredholm operator is normally solvable.
Immediate from the Kato Criterion.
8.5.4. For T B(X, Y ), the equivalence holds:
T Fr(X, Y ) T

Fr(Y

, X

).
Moreover, ind T = ind T

.
By virtue of 2.3.5 (6), 8.5.3, 5.5.4 and the Sequence Prime Principle, the
next pairs of sequences are exact simultaneously:
0 ker T X
T
Y coker T 0;
0 (ker T)

(coker T)

0;
0 ker(T

) Y

coker(T

) 0;
0 (ker(T

))

Y
T
X (coker(T

))

0.
Moreover, (T) = (T

) and (T) = (T

) (cf. 7.6.14).
8.5.5. An operator T is an index-zero Fredholm operator if and only if so is
the dual of T.
This is a particular case of 8.5.4.
Operators in Banach Spaces 145
8.5.6. Fredholm Alternative. For an index-zero Fredholm operator T
either of the following mutually exclusive events takes place:
(1) The homogeneous equation Tx = 0 has a sole solution, zero. The
homogeneous conjugate equation T

= 0 has a sole solution, zero. The equation


Tx = y is solvable and has a unique solution given an arbitrary right side. The
conjugate equation T

= x

is solvable and has a unique solution given an arbitrary


right side.
(2) The homogeneous equation Tx = 0 has a nonzero solution. The ho-
mogeneous conjugate equation T

= 0 has a nonzero solution. The homogeneous


equation Tx = 0 has nitely many linearly independent solutions x
1
, . . . , x
n
. The
homogeneous conjugate equation T

= 0 has nitely many linearly independent


solutions y

1
, . . . , y

n
.
The equation Tx = y is solvable if and only if y

1
(y) = . . . = y

n
(y) = 0.
Moreover, the general solution x is the sum of a particular solution x
0
and the
general solution of the homogeneous equation; i.e., it has the form
x = x
0
+
n

k=1

k
x
k
(
k
F).
The conjugate equation T

= x

is solvable if and only if x

(x
1
) = . . . =
x

(x
n
) = 0. Moreover, the general solution y

is the sum of a particular solution y

0
and the general solution of the homogeneous equation; i.e., it has the form
y

= y

0
+
n

k=1

k
y

k
(
k
F).
This is a reformulation of 8.5.5 with account taken of the Polar Lemma.
8.5.7. Examples.
(1) If T is invertible then T is an index-zero Fredholm operator.
(2) Let T L(F
n
, F
m
). Let rank T := dim imT be the rank of T.
Then (T) = nrank T and (T) = mrank T. Consequently, T Fr(F
n
, F
m
)
and ind T = n m.
(3) Let T B(X) and X = X
1
X
2
. Assume that this direct sum
decomposition of X reduces T to matrix form
T
_
T
1
0
0 T
2
_
.
Undoubtedly, T is a Fredholm operator if and only if its parts are Fredholm
operators. Moreover, (T) = (T
1
)+(T
2
) and (T) = (T
1
)+(T
2
); i.e., ind T =
ind T
1
+ ind T
2
.
146 Chapter 8
8.5.8. Fredholm Theorem. Let K K (X). Then 1 K is an index-zero
Fredholm operator.
First, settle the case F := C. If 1 / Sp(K) then 1 K is invertible and
ind (1 K) = 0. If 1 Sp(K) then in virtue of the RieszSchauder Theorem
and the Spectral Decomposition Theorem there is a decomposition X = X
1
X
2
such that X
1
is nite-dimensional and 1 / Sp(K
2
), with K
2
the part of K in X
2
.
Furthermore,
1 K
_
1 K
1
0
0 1 K
2
_
.
By 8.5.7 (2), ind (1K
1
) = 0 and, by 8.5.7 (3), ind (1K) = ind (1K
1
)+ind (1
K
2
) = 0.
In the case of the reals, F := R, proceed by way of complexication as in
the proof of 8.4.8. Namely, consider the operator K(x, y) := (Kx, Ky) in the
space X iX. By above, ind (1 K) = 0. Considering the dierence between R
and C, observe that (1 K) = (1 K) and (1 K) = (1 K). Finally,
ind (1 K) = 0.
8.5.9. Definition. Let T B(X, Y ). An operator L, a member of B(Y, X),
is a left approximate inverse of T if LT 1 K (X). An operator R, a member of
B(Y, X), is a right approximate inverse of T if TR 1 K (Y ). An operator S,
a member of B(Y, X), is an approximate inverse of T if S is simultaneously a left
and right approximate inverse of T. If an operator T has an approximate inverse S
then T is called approximately invertible. The terms regularizer and parametrix
are all current in this context with regard to S.
8.5.10. Let L and R be a left approximate inverse and a right approximate
inverse of T, respectively. Then L R K (Y, X).
LT = 1 +K
X
(K
X
K (X)) LTR = R+K
X
R;
TR = 1 +K
Y
(K
Y
K (Y )) LTR = L +LK
Y

8.5.11. If L is a left approximate inverse of T and K K (Y, X) then L+K
is also a left approximate inverse of T.
(L +K)T 1 = (LT 1) +KT K (X)
8.5.12. An operator is approximately invertible if and only if it has a right
approximate inverse and a left approximate inverse.
Only the implication needs examining. Let L and R be a left approximate
inverse and a right approximate inverse of T, respectively. By 8.5.10, K:= LR
K (Y, X). Consequently, by 8.5.11, R = L K is a left approximate inverse of T.
Thus, R is an approximate inverse of T.
Operators in Banach Spaces 147
8.5.13. Remark. The above shows that, in the case X = Y , an operator S
is an approximate inverse of T if and only if (S)(T) = (T)(S) = 1, where
: B(X) B(X)/K (X) is the coset mapping to the Calkin algebra. In other
words, a left approximate inverse is the inverse image of a left inverse in the Calkin
algebra, etc.
8.5.14. Noether Criterion. An operator is a Fredholm operator if and only
if it is approximately invertible.
: Let T Fr(X, Y ). Using the Complementation Principle, consider the
decompositions X = ker T X
1
and Y = imT Y
1
and the respective nite-rank
projections P which carries X onto ker T along X
1
and Q which carries Y onto
Y
1
along imT. It is clear that the restriction T
1
:= T[
X
1
is an invertible operator
T
1
: X
1
imT. Put S:= T
1
1
(1Q). The operator S may be viewed as a member
of B(Y, X). Moreover, it is beyond a doubt that ST +P = 1 and TS +Q = 1.
: Let S be an approximate inverse of T; i.e., ST = 1 + K
X
and TS =
1 + K
Y
for appropriate compact operators K
X
and K
Y
. Consequently, ker T
ker(1+K
X
); i.e., ker T is nite-dimensional since so is ker(1+K
X
) in virtue of 8.5.8.
Furthermore, imT im(1 + K
Y
); and so the range of T is of nite codimension
because 1 +K
Y
is an index-zero Fredholm operator.
8.5.15. Corollary. If T Fr(X, Y ) and S is an approximate inverse of T
then S Fr(Y, X).
8.5.16. Corollary. The product of Fredholm operators is itself a Fredholm
operator.
The composition of approximate inverses (taken in due succession) is an ap-
proximate inverse to the composition of the originals.
8.5.17. Consider an exact sequence
0 X
1
X
2
. . . X
n1
X
n
0
of nite-dimensional vector spaces. Then the Euler identity holds:
n

k=1
(1)
k
dimX
k
= 0.
For n = 1 the exactness of the sequence 0 X
1
0 means that X
1
= 0,
and for n = 2 the exactness of 0 X
1
X
2
0 amounts to isomorphy between
X
1
and X
2
(cf. 2.3.5 (4)). Therefore, the Euler identity is beyond a doubt for
n:= 1, 2.
148 Chapter 8
Suppose now that for m n 1, where n > 2, the desired identity is already
established. The exact sequence
0 X
1
X
2
. . . X
n2
T
n2
X
n1
T
n1
X
n
0
reduces to the exact sequence
0 X
1
X
2
. . . X
n2
T
n2
ker T
n1
0.
By hypothesis,
n2

k=1
(1)
k
dimX
k
+ (1)
n1
dimker T
n1
= 0.
Furthermore, since T
n1
is an epimorphism,
dimX
n1
= dimker T
n1
+ dimX
n
.
Finally,
0 =
n2

k=1
(1)
k
dimX
k
+ (1)
n1
(dimX
n1
dimX
n
)
=
n

k=1
(1)
k
dimX
k
.
8.5.18. Atkinson Theorem. The index of the product of Fredholm operators
equals the sum of the indices of the factors.
Let T Fr(X, Y ) and S Fr(Y, Z). By virtue of 8.5.16, ST
Fr(X, Z). Using the Snowake Lemma, obtain the exact sequence of nite-
dimensional spaces
0 ker T ker ST ker S coker T coker ST coker S 0.
Applying 8.5.17, infer that
(T) (ST) +(S) (T) +(ST) (S) = 0;
whence ind (ST) = ind S + ind T.
Operators in Banach Spaces 149
8.5.19. Corollary. Let T be a Fredholm operator and let S be an approximate
inverse of T. Then ind T = ind S.
ind (ST) = ind (1 + K) for some compact operator K. By Theorem 8.5.8,
1 +K is an index-zero Fredholm operator.
8.5.20. Compact Index Stability Theorem. The property of being
a Fredholm operator and the index of a Fredholm operator are preserved under com-
pact perturbations: if T Fr(X, Y ) and K K (X, Y ) then T +K Fr(X, Y )
and ind (T +K) = ind T.
Let S be an approximate inverse to T; i.e.,
ST = 1 +K
X
; TS = 1 +K
Y
with some K
X
K (X) and K
Y
K (Y ) (that S exists is ensured by 8.5.14). It is
clear that
S(T +K) = ST +SK = 1 +K
X
+SK 1 +K (X);
(T +K)S = TS +KS = 1 +K
Y
+KS 1 +K (Y );
i.e., S is an approximate inverse of T +K. By virtue of 8.5.14, T +K Fr(X, Y ).
Finally, from 8.5.19 infer the equalities ind (T+K) = ind S and ind T = ind S.
8.5.21. Bounded Index Stability Theorem. The property of being
a Fredholm operator and the index of a Fredholm operator are preserved under
suciently small bounded perturbations: the set Fr(X, Y ) is open in the space of
bounded operators, and the index of a Fredholm operator ind : Fr(X, Y ) Z is
a continuous function.
Let T Fr(X, Y ). By 8.5.14 there are operators S B(Y, X), K
X

K (X) and K
Y
K (Y ) such that
ST = 1 +K
X
; TS = 1 +K
Y
.
If S = 0 then the spaces X and Y are nite-dimensional by the Riesz Criterion,
i.e., nothing is left to proof: it suces to refer to 8.5.7 (2). If S ,= 0 then for all
V B(X, Y ) with |V | < 1/|S|, from the inequality of 5.6.1 it follows: |SV | < 1
and |V S| < 1. Consequently, in virtue of 5.6.10 the operators 1 +SV and 1 +V S
are invertible in B(X) and in B(Y ), respectively.
Observe that
(1 +SV )
1
S(T +V ) = (1 +SV )
1
(1 +K
X
+SV )
= 1 + (1 +SV )
1
K
X
1 +K (X);
150 Chapter 8
i.e., (1 + SV )
1
S is a left approximate inverse of T + V . By analogy, show that
S(1 +V S)
1
is a right approximate inverse of T +V . Indeed,
(T +V )S(1 +V S)
1
= (1 +K
Y
+V S)(1 +V S)
1
= 1 +K
Y
(1 +V S)
1
1 +K (Y ).
By 8.5.12, T +V is approximately invertible. In virtue of 8.5.14, T +V Fr(X, Y ).
This proves the openness property of Fr(X, Y ). When left and right approximate
inverses of a Fredholm operator W exist, each of them is an approximate inverse
to W (cf. 8.5.12), Therefore, from 8.5.19 and 8.5.18 obtain
ind (T +V ) = ind ((1 +SV )
1
S)
= ind (1 +SV )
1
ind S = ind S = ind T
(because (1 + SV )
1
is a Fredholm operator by 8.5.7 (1)). This means that the
Fredholm index is continuous.
8.5.22. Nikol

ski Criterion. An operator is an index-zero Fredholm opera-


tor if and only if it is the sum of an invertible operator and a compact operator.
: Let T Fr(X, Y ) and ind T = 0. Consider the direct sum decom-
positions X = X
1
ker T and Y = imT Y
1
. It is beyond a doubt that the
operator T
1
, the restriction of T to X
1
, implements an isomorphism between X
1
and imT. Furthermore, in virtue of 8.5.5, dimY
1
= (T) = (T); i.e., there is
a natural isomorphism Id : ker T Y
1
. Therefore, T admits the matrix presenta-
tion
T
_
T
1
0
0 0
_
=
_
T
1
0
0 Id
_
+
_
0 0
0 Id
_
.
: If T := S + K with K K (X, Y ) and S
1
B(Y, X) then, by 8.5.20
and 8.5.7 (1), ind T = ind (S +K) = ind S = 0.
8.5.23. Remark. Let Inv(X, Y ) stand as before for the set of all invertible
operators from X to Y (this set is open by Theorem 5.6.12). Denote by F(X, Z)
the set of all Fredholm operators acting from X to Y and having index zero. The
Nikol

ski Criterion may now be written down as


F(X, Y ) = Inv(X, Y ) +K (X, Y ).
As is seen from the proof of 8.5.22, it may also be asserted that
F(X, Y ) = Inv(X, Y ) +F(X, Y ),
where, as usual, F(X, Y ) is the subspace of B(X, Y ) comprising all nite-rank
operators.
Operators in Banach Spaces 151
Exercises
8.1. Study the RieszDunford integral in nite dimensions.
8.2. Describe the kernel of the RieszDunford integral.
8.3. Given n N, let f
n
be a function holomorphic in a neighborhood U about the spectrum
of an operator T. Prove that the uniform convergence of (f
n
) to zero on U follows from the
convergence of (f
n
(T)) to zero in operator norm.
8.4. Let be an isolated part of the spectrum of an operator T. Assume that the part

:= Sp(T) \ is separated from by some open disk with center a and radius r so that
z C : [z a[ < r. Considering the Riesz projection P

, prove that
P

= lim
n
(1 z
n
(T a)
n
)
1
;
x im(P

) limsup
n
|(a T)
n
x|
1
/
n
< r.
8.5. Find conditions for a projection to be a compact operator.
8.6. Prove that every closed subspace lying in the range of a compact operator in a Banach
space is nite-dimensional.
8.7. Prove that a linear operator carries each closed linear subspace onto a closed set if and
only if the operator is normally solvable and its kernel is nite-dimensional or nite-codimensional
(the latter means that the kernel has a nite-dimensional algebraic complement).
8.8. Let 1 p < r < +. Prove that every bounded operator from l
r
to l
p
or from c
0
to l
p
is compact.
8.9. Let H be a separable Hilbert space. Given an operator T in B(H) and a Hilbert basis
(e
n
) for H, dene the HilbertSchmidt norm as
|T|
2
:=
_

n=1
|Te
n
|
2
_1
/
2
.
(Examine soundness!) An operator with nite HilbertSchmidt norm is a HilbertSchmidt oper-
ator. Demonstrate that an operator T is a HilbertSchmidt operator if and only if T is compact
and

n=1

2
n
< +, where (
n
) ranges over the eigenvalues of some operator (T

T)
1
/
2
(dene
the latter!).
8.10. Let T be an endomorphism. Then
im(T
0
) im(T
1
) im(T
2
) . . . .
If there is a number n satisfying im(T
n
) = im(T
n+1
) then say that T has nite descent. The
least number n with which stabilization begins is the descent of T, denoted by d(T). By analogy,
considering the kernels
ker(T
0
) ker(T
1
) ker(T
2
) . . . ,
introduce the concept of ascent and the denotation a(T). Demonstrate that, for an operator T
with nite descent and nite ascent, the two quantities, a(T) and d(T), coincide.
152 Chapter 8
8.11. An operator T is a RieszSchauder operator, if T is a Fredholm operator and has
nite descent and nite ascent. Prove that an operator T is a RieszSchauder operator if and only
if T is of the form T = U + V , where U is invertible and V is of nite rank (or compact) and
commutes with U.
8.12. Let T be a bounded endomorphism of a Banach space X which has nite descent and
nite ascent, r := a(T) = d(T). Prove that the subspaces im(T
r
) and ker(T
r
) are closed, the
decomposition X = ker(T
r
) im(T
r
) reduces T, and the restriction of T to im(T
r
) is invertible.
8.13. Let T be a normally solvable operator. If either of the next quantities is nite
(T):= dim ker T, (T):= dim coker T,
then T is called semi-Fredholm. Put
+
(X):= T B(X) : imT Cl(X), (T) < +;

(X):= T B(X) : imT Cl(X), (T) < +.


Prove that
T
+
(X) T


(X

);
T

(X) T

+
(X

).
8.14. Let T be a bounded endomorphism. Prove that T belongs to
+
(X) if and only if for
every bounded but not totally bounded set U, the image T(U) is not a totally bounded set in X.
8.15. A bounded endomorphism T in a Banach space is a Riesz operator, if for every nonzero
complex the operator ( T) is a Fredholm operator. Prove that T is a Riesz operator if and
only if for all C, ,= 0, the following conditions are fullled:
(1) the operator ( T) has nite descent and nite ascent;
(2) the kernel of ( T)
k
is nite-dimensional for every k N;
(3) the range of ( T)
k
has nite deciency for k N,
and, moreover, all nonzero points of the spectrum of T are eigenvalues, with zero serving as the
only admissible limit point (that is, for whatever strictly positive , there are only nitely many
points of Sp(T) beyond the disk centered at zero with radius ).
8.16. Establish the isometric isomorphisms: (X/Y )

and X

/Y

for Banach
spaces X and Y such that Y is embedded into X.
8.17. Prove that for a normal operator T in a Hilbert space and a holomorphic function f,
a member of H(Sp(T)), the operator f(T) is normal. (An operator is normal if it commutes with
its adjoint, cf. 11.7.1.)
8.18. Show that a continuous endomorphism T of a Hilbert space is a Riesz operator if and
only if T is the sum of a compact operator and a quasinilpotent operator. (Quasinilpotency of
an operator means triviality of its spectral radius.)
8.19. Given two Fredholm operators S and T, members of Fr(X, Y ), with ind S = ind T,
demonstrate that there is a Jordan arc joining S and T within Fr(X, Y ).
Chapter 9
An Excursus into General Topology
9.1. Pretopologies and Topologies
9.1.1. Definition. Let X be a set. A mapping : X P(P(X)) is a pre-
topology on X if
(1) x X (x) is a lter on X;
(2) x X (x) lx.
A member of (x) is a (pre)neighborhood about x or of x. The pair (X, ), as well
as the set X itself, is called a pretopological space.
9.1.2. Definition. Let T (X) be the collection of all pretopologies on X.
If
1
,
2
T (X) then
1
is said to be stronger than
2
or ner than
2
(in symbols,

1

2
) provided that x X
1
(x)
2
(x). Of course,
2
is weaker or coarser
than
1
.
9.1.3. The set T (X) with the relation to be stronger presents a complete
lattice.
If X = then T (X) = and there is nothing to be proven. If X ,=
then refer to 1.3.13.
9.1.4. Definition. A subset G of X is an open set in X, if G is a (pre)neigh-
borhood of its every point (in symbols, G Op() ( x G) (G (x))).
A subset F of X is a closed set in X if the complement of F to X is open; that is,
F Cl() X F Op().
9.1.5. The union of a family of open sets and the intersection of a nite family
of open sets are open. The intersection of a family of closed sets and the union
of a nite family of closed sets are closed.
9.1.6. Let (X, ) be a pretopological space. Given x X, put
U t()(x) ( V Op()) x V & U V.
The mapping t() : x t()(x) is a pretopology on X.
154 Chapter 9
9.1.7. Definition. A pretopology on X is a topology if = t(). The pair
(X, ), as well as the underlying set X itself, is then called a topological space. The
set of all topologies on X is denoted by the symbol T(X).
9.1.8. Examples.
(1) A metric topology.
(2) The topology of a multinormed space.
(3) Let

:= inf T (X). It is clear that

(x) = X for x X.
Consequently, Op(

) = , X and so

= t(

); i.e.,

is a topology. This
topology is called trivial, or antidiscrete, or even indiscrete.
(4) Let

:= sup T (X). It is clear that

(x) = l x for x X.
Consequently, Op(

) = 2
X
and so

= t(

); i.e.,

is a topology. This topology


is called discrete.
(5) Let Op be a collection of subsets in X which is stable under the
taking of the union of each of its subfamilies and the intersection of each of its nite
subfamilies. Then there is a unique topology on X such that Op() = Op.
Put (x):= l V Op : x V for x X (in case X = there is nothing
to prove). Observe that (x) ,= since the intersection of the empty family equals
X (cf. inf = +). From the construction derive that t() = and Op Op().
If G Op() then G = V : V Op, V G and so G Op by hypothesis.
The claim of uniqueness raises no doubts.
9.1.9. Let the mapping t : T (X) T (X) act by the rule t : t(). Then
(1) imt = T(X); i.e., T (X) t() T(X);
(2)
1

2
t(
1
) t(
2
) (
1
,
2
T (X));
(3) t t = t;
(4) T (X) t() ;
(5) Op() = Op(t()) ( T (X)).
The inclusion Op() Op(t()) holds because it is easier to be open in .
The reverse inclusion Op() Op(t()) follows from the denition of t(). The
equality Op() = Op(t()) makes everything evident.
9.1.10. A pretopology on X is a topology if and only if
( U (x))( V (x) & V U)( y) (y V V (y))
for x X.
Straightforward from 9.1.9 (5).
An Excursus into General Topology 155
9.1.11. Let
1
,
2
T(X). The following statements are equivalent:
(1)
1

2
;
(2) Op(
1
) Op(
2
);
(3) Cl(
1
) Cl(
2
).
9.1.12. Remark. As follows from 9.1.8 (5) and 9.1.11, the topology of a space
is uniquely determined from the collection of its open sets. Therefore, the set Op(X)
itself is legitimately called the topology of the space X. In particular, the collection
of open sets of a pretopological space (X, ) makes X into the topological space
(X, t()) with the same open sets in stock. Therefore, given a pretopology , the
topology t() is usually called the topology associated with .
9.1.13. Theorem. The set T(X) of all topologies on X with the relation
to be stronger presents a complete lattice. Moreover, for every subset E of T(X)
the equality holds:
sup
T(X)
E = sup
T (X)
E .
Evidently, t(sup
T (X)
E ) sup
T (X)
t(E ) sup
T (X)
E t(sup
T (X)
E ).
Thus, := sup
T (X)
E belongs to T(X). It is clear that E . Furthermore,
if
0
E and
0
T(X) then
0
and so = sup
T(X)
E . It remains to refer
to 1.2.14.
9.1.14. Remark. The explicit formula for the greatest lower bound of E is
more involved:
inf
T(X)
E = t(inf
T (X)
E ).
However, the matter becomes simpler when the topologies are given by means of
their collections of open sets in accordance with 9.1.12. Namely,
U Op(inf
T(X)
E ) ( E ) U Op().
In other words,
Op(inf
T(X)
E ) =

E
Op().
In this connection it is in common parlance to speak of the intersection of the set E
of topologies (rather than of the greatest lower bound of E ).
9.2. Continuity
9.2.1. Remark. The presence of a topology on a set obviously makes it pos-
sible to deal with such things as the interior and closure of a subset, convergence
of lters and nets, etc. We have already made use of this circumstance while in-
troducing multinormed spaces. Observe for the sake of completeness that in every
topological space the following analogs of 4.1.19 and 4.2.1 are valid:
156 Chapter 9
9.2.2. Birkho Theorem. For a nonempty subset U and a point x of a topo-
logical space the following statements are equivalent:
(1) x is an adherent point of U;
(2) there is some lter containing U and converging to x;
(3) there is a net of elements of U which converges to x.
9.2.3. For a mapping f between topological spaces the following conditions
are equivalent:
(1) the inverse image under f of an open set is open;
(2) the inverse image under f of a closed set is closed;
(3) the image under f of the neighborhood lter of an arbitrary point
x is coarser than the neighborhood lter of f(x);
(4) for all x, the mapping f transforms each lter convergent to x into
a lter convergent to f(x);
(5) for all x, the mapping f sends a net that converges to x to a net
that converges to f(x).
9.2.4. Definition. A mapping acting between topological spaces X and Y
and satisfying one (and hence all) of the equivalent conditions 9.2.3 (1)9.2.3 (5)
is called continuous. A continuous one-to-one mapping f from X onto Y whose
inverse f
1
acts continuously from Y to X is a homeomorphism or a topological
mapping or a topological isomorphism between X and Y .
9.2.5. Remark. If f : (X,
X
) (Y,
Y
) meets 9.2.3 (5) at some point x in X
then it is customary to say that f is continuous at x (cf. 4.2.2). Observe that the
dierence is immaterial between the denitions of the continuity property at a point
of X and the general continuity property (on X). Indeed, if we let
x
(x) :=
X
(x)
and
x
(x) := l x for x X, x ,= x, then the continuity property of f at x (with
respect to the topology
X
in X) amounts to that of f : (X,
x
) (Y,
Y
) (at every
point of the space X with topology
x
).
9.2.6. Let
1
,
2
T(X). Then
1

2
if and only if I
X
: (X,
1
) (X,
2
)
is continuous.
9.2.7. Let f : (X, ) (Y, ) be a continuous mapping and let
1
T(X)
and
1
T(Y ) be such that
1
and
1
. Then f : (X,
1
) (Y,
1
) is
continuous.
By hypothesis the following diagram commutes:
(X, )
f
(Y, )
I
X
I
Y
(X,
1
)
f
(Y,
1
)
An Excursus into General Topology 157
It suces to observe that every composition of continuous mappings is continu-
ous.
9.2.8. Inverse Image Topology Theorem. Let f : X (Y, ). Put
T
0
:= T(X) : f : (X, ) (Y, ) is continuous.
Then the topology f
1
():= inf T
0
belongs to T
0
.
From 9.2.3 (1) it follows that
T
0
(x X f
1
((f(x))) (x)).
Let (x) := f
1
((f(x))). Undoubtedly, t() = . Furthermore, f((x)) =
f(f
1
((f(x)))) (f(x)); i.e., T
0
by 9.2.3 (3). Thus, f
1
() = .
9.2.9. Definition. The topology f
1
() is the inverse image of under
a mapping f or simply the inverse image topology under f.
9.2.10. Remark. Theorem 9.2.8 is often verbalized as follows: The inverse
image topology under a mapping is the weakest topology on the set of departure
in which the mapping is continuous. Moreover, it is easy for instance from 9.1.14
that the open sets of the inverse image topology are precisely the inverse images
of open sets. In particular, (x

x in f
1
()) (f(x

) f(x) in ); likewise,
(F x in f
1
()) (f(F) f(x) in ) for a lter F.
9.2.11. Image Topology Let f : (X, ) Y . Put

0
:= T(Y ) : f : (X, ) (Y, ) is continuous.
Then the topology f():= sup
0
belongs to
0
.
Appealing to 9.1.13, observe that
f()(y) = (sup
T(Y )

0
)(y) = (sup
T (Y )

0
)(y) = sup(y) :
0

for y Y . By virtue of 9.2.3 (3),



0
(x X f((x)) (f(x))).
Comparing the formulas, infer that f()
0
.
9.2.12. Definition. The topology f() is the image of under a mapping f
or simply the image topology under f.
158 Chapter 9
9.2.13. Remark. Theorem 9.2.11 is often verbalized as follows: The image
of a topology under a mapping is the strongest topology on the set of arrival in which
the mapping is continuous.
9.2.14. Theorem. Let (f

: X (Y

))

be a family of mappings.
Further, put := sup

f
1

). Then is the weakest (= least) topology on X


making all the mappings f

( ) continuous.
Using 9.2.8, note that
(f

: (X, ) (Y

) is continuous) f
1

).
9.2.15. Theorem. Let (f

: (X

) Y )

be a family of mappings.
Further, assign := inf

). Then is the strongest (= greatest) topology


on Y making all the mappings f

( ) continuous.
Appealing to 9.2.11, conclude that
(f

: (X

) (Y, ) is continuous) f

).
9.2.16. Remark. The messages of 9.2.14 and 9.2.15 are often referred to as
the theorems on topologizing by a family of mappings.
9.2.17. Examples.
(1) Let (X, ) be a topological space and let X
0
be a subset of X.
Denote the identical embedding of X
0
into X by : X
0
X. Put
0
:=
1
().
The topology
0
is the induced topology (by in X
0
), or the relative or subspace
topology; and the space (X
0
,
0
) is a subspace of (X, ).
(2) Let (X

be a family of topological spaces and let X :=

be the product of (X

. Put := sup

Pr
1

), where Pr

: X X

is the coordinate projection (onto X

); i.e., Pr

x = x

( ). The topology
is the product topology or the product of the topologies (

, or the Tychono
topology of X. The space (X, ) is the Tychono product of the topological spaces
under study. In particular, if X

:= [0, 1] for all then X:= [0, 1]

(with the
Tychono topology) is a Tychono cube. When := N, the term Hilbert cube is
applied.
9.3. Types of Topological Spaces
9.3.1. For a topological space the following conditions are equivalent:
(1) every singleton of the space is closed;
(2) the intersection of all neighborhoods of each point in the space
consists solely of the point;
An Excursus into General Topology 159
(3) each one of any two points in the space has a neighborhood disjoint
from the other.
To prove, it suces to observe that
y clx ( V (y)) x V x V : V (y),
where x and y are points of a space with topology .
9.3.2. Definition. A topological space satisfying one (and hence all) of the
equivalent conditions 9.3.1 (1)9.3.1 (3), is called a separated space or a T
1
-space.
The topology of a T
1
-space is called a separated topology or (rarely) a T
1
-topology.
9.3.3. Remark. By way of expressiveness, one often says: A T
1
-space is
a space with closed points.
9.3.4. For a topological space the following conditions are equivalent:
(1) each lter has at most one limit;
(2) the intersection of all closed neighborhoods of a point in the space
consists of the sole point;
(3) each one of any two points of the space has a neighborhood disjoint
from some neighborhood of the other point.
(1) (2): If y
U(x)
cl U then U V ,= for all V (y), provided
that U (x). Therefore, the join F := (x) (y) is available. Clearly, F x
and F y. By hypothesis, x = y.
(2) (3): Let x, y X, x ,= y (if such points are absent then either X = or
X is a singleton and nothing is left unproven). There is an neighborhood U in (x)
such that U = cl U and y / U. Consequently, the complement V of U to X is open.
Furthermore, U V = .
(3) (1): Let F be a lter on X. If F x and F y then F (x)
and F (y). Thus, U V ,= for U (x) and V (y), which means that
x = y.
9.3.5. Definition. A topological space satisfying one (and hence all) of the
equivalent conditions 9.3.4 (1)9.3.4 (3) is a Hausdor space or a T
2
-space. A nat-
ural meaning is ascribed to the term Hausdor topology.
9.3.6. Remark. By way of expressiveness, one often says: A T
2
-space is
a space with unique limits.
9.3.7. Definition. Let U and V be subsets of a topological space. It is said
that V is a neighborhood of U or about U, provided int V U. If U is nonempty
then all neighborhoods of U constitute some lter that is the neighborhood lter
of U.
160 Chapter 9
9.3.8. For a topological space the following conditions are equivalent:
(1) the intersection of all closed neighborhoods of an arbitrary closed
set consists only of the members of the set;
(2) the neighborhood lter of each point has a base of closed sets;
(3) if F is a closed set and x is a point not in F then there are disjoint
neighborhoods of F and x, respectively.
(1) (2): If x X and U (x) then V := X int U is closed and x / V .
By hypothesis there is a set F in Cl() such that x / F and int F V . Put
G:= X F. Clearly, G (x). Moreover, G X int F = cl(X int F) X V
int U U. Consequently, cl G U.
(2) (3): If x X and F Cl() with x / F then X F (x). Thus, there
is a closed neighborhood U in (x) lying in X F. Thus, X U is a neighborhood
of F disjoint from U.
(3) (1): If F Cl() and int G F y cl G, then U G ,= for every
U in (y) and every neighborhood G of F. This means that y F.
9.3.9. Definition. A T
3
-space is a topological space satisfying one (and hence
all) of the equivalent conditions 9.3.8 (1)9.3.8 (3). A separated T
3
-space is called
regular.
9.3.10. Urysohn Little Lemma. For a topological space the following con-
ditions are equivalent:
(1) the neighborhood lter of each nonempty closed set has a base of
closed sets;
(2) if two closed sets are disjoint then they have disjoint neighborhoods.
(1) (2): Let F
1
and F
2
be closed sets in some space X with F
1
F
2
= .
Put G := X F
1
. Obviously, G is open and G F
2
. If F
2
= , then there is
nothing to be proven. It may be assumed consequently that F
2
,= . Then there is
a closed set V
2
such that G V
2
int V
2
F
2
. Put V
1
:= X V
2
. It is clear that
V
1
is open, and V
1
V
2
= . Moreover, V
1
X G = X (X F
1
) = F
1
.
(2) (1): Let F = cl F, G = int G and G F. Put F
1
:= X G. Then
F
1
= cl F
1
and so there are open sets U and U
1
satisfying U U
1
= , with F U
and F
1
U
1
. Finally, cl U X U
1
X F
1
= G.
9.3.11. Definition. A T
4
-space is a topological space meeting one (and hence
both) of the equivalent conditions 9.3.10 (1) and 9.3.10 (2). A separated T
4
-space
is called normal.
9.3.12. Continuous Function Recovery Lemma. Let a subset T be dense
in R and let t U
t
(t T) be a family of subsets of a topological space X. There
is a unique continuous function f : X R such that
f < t U
t
f t (t T)
An Excursus into General Topology 161
if and only if
(t, s T & t < s) cl U
t
int U
s
.
: Take t < s. Since f t is closed and f < s is open, the inclusions
hold:
cl U
t
f t f < s int U
s
.
: Since U
t
cl U
t
int U
s
U
s
for t < s, the family t U
t
(t T) increases
by inclusion. Therefore, f exists by 3.8.2 and is unique by 3.8.4. Consider the
families t V
t
:= cl U
t
and t W
t
:= int U
t
. These families increase by inclusion.
Consequently, on applying 3.8.2 once again, nd functions g, h : X R satisfying
g < t V
t
g t, h < t W
t
h t
for all t T. If t, s T and t < s, then in view of 3.8.3
W
t
= int U
t
U
t
U
s
f h;
V
t
= cl U
t
int U
s
= W
s
h g;
U
t
U
s
cl U
s
= V
s
g f.
So, f = g = h. Taking account of 3.8.4 and 9.1.5 and given t R, nd
f < t = h < t = W
s
: s < t, s T Op(
X
);
f t = g t = V
s
: t < s, s T Cl(
X
).
These inclusions readily provide continuity for f.
9.3.13. Urysohn Great Lemma. Let X be a T
4
-space. Assume further that
F is a closed set in X and G is a neighborhood of F. Then there is a continuous
function f : X [0, 1] such that f(x) = 0 for x F and f(x) = 1 for x / G.
Put U
t
:= for t < 0 and U
t
:= X for t > 1. Consider the set T of the
dyadic-rational points of the interval [0, 1]; i.e., T :=
nN
T
n
with T
n
:= k2
n+1
:
k:= 0, 1, . . . , 2
n1
. It suces to dene U
t
for all t in T so that the family t U
t
(t T := T (R [0, 1])) satisfy the criterion of 9.3.12. This is done by way
of induction.
If t T
1
, i.e., t 0, 1; then put U
0
:= F and U
1
:= G. Assume now
that, for t T
n
and n 1, some set U
t
has already been constructed, satisfying
cl U
t
int U
s
whenever t, s T
n
and t < s. Take t T
n+1
and nd the two points
t
l
and t
r
in T
n
nearest to t:
t
l
:= sups T
n
: s t;
t
r
:= infs T
n
: t s.
162 Chapter 9
If t = t
l
or t = t
r
, then U
t
exists by the induction hypothesis. If t ,= t
l
and
t ,= t
r
, then t
l
< t < t
r
and again by the induction hypothesis cl U
t
l
int U
t
r
.
By virtue of 9.3.11 there is a closed set U
t
such that
cl U
t
l
int U
t
U
t
= cl U
t
int U
t
r
.
It remains to show that the resulting family satises the criterion of 9.3.12.
To this end, take t, s T
n+1
with t < s. If t
r
= s
l
, then for s > s
l
by con-
struction
cl U
t
cl U
t
r
= cl U
s
l
int U
s
.
For t < t
r
= s
l
similarly deduce the following:
cl U
t
int U
t
r
= inf U
s
l
int U
s
.
If t
r
< s
l
then, on using the induction hypothesis, infer that
cl U
t
cl U
t
r
int U
s
l
int U
s
,
what was required.
9.3.14. Urysohn Theorem. A topological space X is a T
4
-space if and only
if to every pair of disjoint closed sets F
1
and F
2
in X there corresponds a continuous
function f : X [0, 1] such that f(x) = 0 for x F
1
and f(x) = 1 for x F
2
.
: It suces to apply 9.3.13 with F := F
1
and G:= X F
2
.
: If F
1
F
2
= and F
1
and F
2
are closed sets, then for a corresponding
function f the sets G
1
:= f <
1
/
2
and G
2
:= f >
1
/
2
are open and disjoint.
Moreover, G
1
F
1
and G
2
F
2
.
9.3.15. Definition. A topological space X is a T
3
1
/
2
-space, if to a closed set F
in X and a point x not in F there corresponds a continuous function f : X [0, 1]
such that f(x) = 1 and y F f(y) = 0. A separated T
3
1
/
2
-space is a Tychono
space or a completely regular space.
9.3.16. Every normal space is a Tychono space.
Straightforward from 9.3.1 and 9.3.14.
9.4. Compactness
9.4.1. Let B be a lterbase on a topological space and let
cl B:= cl B : B B
be the set of adherent points of B (also called the adherence of B). Then
An Excursus into General Topology 163
(1) cl B = cl l B;
(2) B x x cl B;
(3) (B is an ultralter and x cl B) B x.
Only (3) needs demonstrating, since (1) and (2) are evident. Given U (x)
and B B, observe that U B ,= . In other words, the join F := (x)
B is available. It is clear that F x. Furthermore, F = B, because B is
an ultralter.
9.4.2. Definition. A subset C of a topological space X is a compact set in X
if each open cover of C has a nite subcover (cf. 4.4.1).
9.4.3. Theorem. Let X be a topological space and let C be a subset of X.
The following statements are equivalent:
(1) C is compact;
(2) if a lterbase B lacks adherent points in C then there is a member
B of B such that B C = ;
(3) each lterbase containing C has an adherent point in C;
(4) each ultralter containing C has a limit in C.
(1) (2): Since cl B C = ; therefore, C X cl B. Thus, C
X cl B : B B = X cl B : B B. Consequently, there is a nite
subset B
0
of B such that C X cl B
0
: B
0
B
0
= X cl B
0
: B
0
B
0
.
Let B B satisfy B B
0
: B
0
B
0
cl B
0
: B
0
B
0
. Then
C B C (cl B
0
: B
0
B
0
) = .
(2) (3): In case C = , there is nothing to be proven. If C ,= , then for
B B by hypothesis B C ,= because C B. Thus, cl B C ,= .
(3) (4): It suces to appeal to 9.4.1.
(4) (1): Assume that C ,= (otherwise, nothing is left to proof).
Suppose that C is not compact. Then there is a set E of open sets such that
C G : G E and at the same time, for every nite subset E
0
of E , the
inclusion C G : G E
0
fails. Put
B:=
_

GE
0
X G : E
0
is a nite subset of E
_
.
It is clear that B is a lterbase. Furthermore,
cl B = cl B : B B = X G : G E
= X G : G E X C.
Now choose an ultralter F that is coarser than B, which is guaranteed by 1.3.10.
By supposition each member of B meets C. We may thus assume that C F
(adjusting the choice of F, if need be). Then F x for some x in C and so, by 9.4.1
(2), cl F C ,= . At the same time cl F cl B. We arrive at a contradiction.
164 Chapter 9
9.4.4. Remark. The equivalence (1) (4) in Theorem 9.4.3 is called the
Bourbaki Criterion and verbalized for X = C as follows: A space is compact if and
only if every ultralter on it converges (cf. 4.4.7). An ultranet is a net whose
tail lter is an ultralter. The Bourbaki Criterion can be expressed as follows:
Compactness amounts to convergence of ultranets. Many convenient tests for
compactness are formulated in the language of nets. For instance: A space X is
compact if and only if each net in X has a convergent subnet.
9.4.5. Weierstrass Theorem. The image of a compact set under a continu-
ous mapping is compact (cf. 4.4.5).
9.4.6. Let X
0
be a subspace of a topological space X and let C be a subset
of X
0
. Then C is compact in X
0
if and only if C is compact in X.
: Immediate from 9.4.5 and 9.2.17 (1).
: Let B be a lterbase on X
0
. Further, let V := cl
X
0
B stand for the
adherence of B relative to X
0
. Suppose that V C = . Since B is also a lterbase
on X, it makes sense to speak of the adherence W := cl
X
B of B relative to X.
It is clear that V = W X
0
and, consequently, W C = . Since C is compact
in X, by 9.4.3 there is some B in B such that B C = . Using 9.4.3 once again,
infer that C is compact in X
0
.
9.4.7. Remark. The claim of 9.4.6 is often expressed as follows: Compact-
ness is an absolute concept. It means that for C to be or not to be compact
depends on the topology induced in C rather than on the ambient space inducing
the topology. For that reason, it is customary to conne study to compact spaces,
i.e. to sets compact in themselves. A topology on a set C, making C into a com-
pact space, is usually called a compact topology on C. Also, such C is referred to as
compact with respect to .
9.4.8. Tychono Theorem. The Tychono product of compact spaces is
compact.
Let X :=

be the product of such spaces. If at least one of the


spaces X

is nonempty then X = and nothing is left to proof. Let X ,= and


let F be an ultralter on X. By 1.3.12, given and considering the coordinate
projection Pr

: X X

, observe that Pr

(F) is an ultralter on X

. Consequently,
in virtue of 9.4.3 there is some x

in X

such that Pr

(F) x

. Let x : x

.
It is clear that F x (cf. 9.2.10). Appealing to 9.4.3 once more, infer that X is
compact.
9.4.9. Every closed subset of a compact space is compact.
Let X be compact and C Cl(X). Assume further that F is an ultralter
on X and C F. By Theorem 9.4.3, F has a limit x in X: that is, F x.
By the Birkho Theorem, x cl C = C. Using 9.4.3 again, conclude that C is
compact.
An Excursus into General Topology 165
9.4.10. Every compact subset of a Hausdor space is closed.
Let C be compact in a Hausdor space X. If C = then there is nothing
to prove. Let C ,= and x cl C. By virtue of 9.2.2 there is a lter F
0
on X such
that C F
0
and F
0
x. Let F be an ultralter ner than F
0
. Then F x
and C F. By 9.4.3, F has a limit in C. By 9.3.4 every limit in X is unique.
Consequently, x C.
9.4.11. Let f : (X, ) (Y, ) be a continuous one-to-one mapping with
f(X) = Y . If is a compact topology and is a Hausdor topology, then f is
a homomorphism.
It suces to establish that f
1
is continuous. To this end, we are to demon-
strate that F Cl() f(F) Cl(). Take F Cl(). Then F is compact
by 9.4.9. Successively applying 9.4.5 and 9.4.10, infer that f(F) is closed.
9.4.12. Let
1
and
2
be two topologies on a set X. If (X,
1
) is a compact
space and (X,
2
) is a Hausdor space with
1

2
, then
1
=
2
.
9.4.13. Remark. The message of 9.4.12 is customarily verbalized as follows:
A compact topology is minimal among Hausdor topologies.
9.4.14. Theorem. Every Hausdor compact space is normal.
Let X be the space under study and let B be some lterbase on X. Assume
further that U is a neighborhood of cl B. It is clear that X int U is compact
(cf. 9.4.9) and cl B(X int U) = . By Theorem 9.4.3 there is a member B of B
such that B(Xint U) = ; i.e., B U. Putting, if need be, B:= cl B : B B,
we may assert that cl B U.
To begin with, take x X and put B:= (x). By virtue of 9.3.4, cl B = x
and, consequently, the lter (x) has a base of closed sets. Thus, X is regular.
Now take nonempty closed subset F of X. Take as B the neighborhood lter
of F. By 9.3.8, cl B = F, and, as is already established, B has a base of closed
sets. In accordance with 9.3.9, X is a normal space.
9.4.15. Corollary. Each Hausdor compact space is (to within a homeomor-
phism) a closed subset of a Tychono cube.
The compactness property of a closed subset of a Tychono cube follows
from 9.4.8 and 9.4.9. Moreover, every cube is a Hausdor space and so such is each
of its subspaces.
Now take some Hausdor compact space X. Let Q be the collection of all
continuous function from X to [0, 1]. Dene the mapping : X [0, 1]
Q
as
(x)(f) := f(x) where x X and f Q. From 9.4.14 and 9.3.14 infer that
carries X onto (X) in a one-to-one fashion. Furthermore, is continuous.
Application to 9.4.11 completes the proof.
166 Chapter 9
9.4.16. Remark. Corollary 9.4.15 presents a part of a more general assertion.
Namely, a Tychono space is (to within a homeomorphism) a subspace of a Ty-
chono cube.
9.4.17. Remark. Sometimes a Hausdor compact space is also called a com-
pactum (cf. 4.5 and 4.6).
9.4.18. Dieudonn`e Lemma. Let F be a closed set and let G
1
, . . . , G
n
be
open sets in a normal topological space, with F G
1
. . . G
n
. There are closed
sets F
1
, . . . , F
n
such that F = F
1
. . . F
n
and F
k
G
k
(k:= 1, . . . , n).
It suces to settle the case n := 2. For k := 1, 2 the set U
k
:= F G
k
is closed. Moreover, U
1
U
2
= . By 9.3.10 there are open V
1
and V
2
such that
U
1
V
1
, U
2
V
2
and V
1
V
2
= . Put F
k
:= F V
k
. It is clear that F
k
is closed and
F
k
F U
k
= F (F G
k
) G
k
for k:= 1, 2. Finally, F
1
F
2
= F (V
1
V
2
) = F.
9.4.19. Remark. From 9.3.14 we deduce that under the hypotheses of 9.4.18
there are continuous functions h
1
, . . . , h
n
: X [0, 1] such that h
k
[
G

k
= 0 and

n
k=1
h
k
(x) = 1 for every point x in some neighborhood about F. (As usual,
G

k
:= X G
k
.)
9.4.20. Definition. A topology is called locally compact if each point pos-
sesses a compact neighborhood. A locally compact space is a set furnished with
a Hausdor locally compact topology.
9.4.21. A topological space is locally compact if and only if it is homeomor-
phic with a punctured compactum (= a compactum with a deleted point), i.e. the
complement of a singleton to a compactum.
: In virtue of the Weierstrass Theorem it suces to observe that each point
of a punctured compactum possesses a closed neighborhood (since every compactum
is regular). It remains to make use of 9.4.9 and 9.4.6.
: Put the initial space X in X

:= X, adjoining to X a point taken


elsewhere. Take the complements to X

of compact subsets of X as a base for the


neighborhood lter about . A neighborhood of a point x of X in X

is declared
to be a superset of a neighborhood of x in X. If A is an ultralter in X

and K
is a compactum in X then A converges to a point in K provided that K A.
If A contains the complement of each compactum K in X to X, then A converges
to .
9.4.22. Remark. If a locally compact space X is not compact in its own right
then X

of 9.4.20 is the one-point or Alexandro compactication of X.


9.5. Uniform and Multimetric Spaces
9.5.1. Definition. Let X be a nonempty set and let U
X
be a lter on X
2
.
The lter U
X
is a uniformity on X if
An Excursus into General Topology 167
(1) U
X
l I
X
;
(2) U U
X
U
1
U
X
;
(3) ( U U
X
)( V U
X
) V V U.
The uniformity of the empty set is by denition . The pair (X, U
X
), as well
as the underlying set X, is called a uniform space.
9.5.2. Given a uniform space (X, U
X
), put
x X (x):= U(x) : U U
X
.
The mapping : x (x) is a topology on X.
Clearly, is a pretopology. If W (x) then W = U(x) for some U
in U
X
. Choose a member V of U
X
so that V V U. If y V (x) then V (y)
V (V (x)) = V V (x) U(x) W. In other words, the set W is a neighborhood
about y for every y in V (x). Therefore, V (x) lies in int W. Consequently, int W is
a neighborhood about x. It remains to refer to 9.1.6.
9.5.3. Definition. The topology appearing in 9.5.2 is the topology of the uni-
form space (X, U
X
) under consideration or the uniform topology on X. It is also
denoted by (U
X
),
X
, etc.
9.5.4. Definition. A topological space (X, ) is called uniformizable pro-
vided that there is a uniformity U on X such that coincides with the uniform
topology (U ).
9.5.5. Examples.
(1) A metric space (with its metric topology) is uniformizable (with its
metric uniformity).
(2) A multinormed space (with its topology) is uniformizable (with its
uniformity).
(3) Let f : X (Y, U
Y
) and f
1
(U
Y
) := f
1
(U
Y
), where as usual
f

(x
1
, x
2
):= (f(x
1
), f(x
2
)) for (x
1
, x
2
) X
2
. Evidently, f
1
(U
Y
) is a uniformity
on X. Moreover,
(f
1
(U
Y
)) = f
1
((U
Y
)).
The uniformity f
1
(U
Y
) is the inverse image of U
Y
under f. Therefore, the inverse
image of a uniform topology is uniformizable.
(4) Let (X

, U

be a family of uniform spaces. Assume further


that X:=

is the product of the family. Put U


X
:= sup

Pr
1

(U

). The
uniformity U
X
is the Tychono uniformity. It is beyond a doubt that the uniform
topology (U
X
) is the Tychono topology of the product of (X

, (U

))

.
168 Chapter 9
(5) Each Hausdor compact space is uniformizable in a unique fashion.
By virtue of 9.4.15 such space X may be treated as a subspace of a Ty-
chono cube. From 9.5.5 (3) and 9.5.5 (4) it follows that X is uniformizable. Since
each entourage of a uniform space includes a closed entourage; therefore, the com-
pactness property of the diagonal I
X
of X
2
implies that every neighborhood of I
X
belongs to U
X
. On the other hand, each entourage is always a neighborhood of the
diagonal.
(6) Assume that X and Y are nonempty sets, U
Y
is a uniformity on Y
and B is an upward-ltered subset of P(X). Given B B and U
Y
, put
U
B,
:= (f, g) Y
X
Y
X
: g I
B
f
1
.
Then U := l U
B,
: B B, U
Y
is a uniformity on Y
X
. This uniformity
has a cumbersome (but exact) title, the uniformity of uniform convergence on the
members of B. Such is, for instance, the uniformity of the Arens multinorm
(cf. 8.3.8). If B is the collection of all nite subsets of X, then U coincides with the
Tychono uniformity on Y
X
. This uniformity is called weak, and the corresponding
uniform topology is called the topology of pointwise convergence or, rarely, that
of simple convergence. If B is a singleton X, then the uniformity U is called
strong and the corresponding topology (U ) in Y
X
is the topology of uniform
convergence on X.
9.5.6. Remark. It is clear that, in a uniform (or uniformizable) space, the
concepts make sense such as uniform continuity, total boundedness, completeness,
etc. It is beyond a doubt that in such space the analogs of 4.2.44.2.9, 4.5.8, 4.5.9,
and 4.6.14.6.7 are preserved. It is a rewarding practice to ponder over a possibility
of completing a uniform space, to validate a uniform version of the Hausdor Cri-
terion, to inspect the proof of the AscoliArzela Theorem in an abstract uniform
setting, etc.
9.5.7. Definition. Let X be a set and put R

+
:= x R

: x 0. A map-
ping d : X
2
R

+
is called a semimetric or a pseudometric on X, provided that
(1) d(x, x) = 0 (x X);
(2) d(x, y) = d(y, x) (x, y X);
(3) d(x, y) d(x, z) +d(z, y) (x, y, z X).
A pair (X, d) is a semimetric space.
9.5.8. Given a semimetric space (X, d), let U
d
:= l d : > 0. Then
U
d
is a uniformity.
9.5.9. Definition. Let M be a (nonempty) set of semimetrics on X. Then
the pair (X, M) is a multimetric space with multimetric M. The multimetric
uniformity on X is dened as U
M
:= supU
d
: d M.
An Excursus into General Topology 169
9.5.10. Definition. A uniform space is called multimetrizable, if its uniformi-
ty coincides with some multimetric uniformity. A multimetrizable topological space
is dened by analogy.
9.5.11. Assume that X, Y , and Z are sets, T is a dense subset of R, and
(U
t
)
tT
and (V
t
)
tT
are increasing families of subsets of X Z and Z Y , re-
spectively. Then there are unique functions f : X Z R, g : Z Y R and
h : X Y R such that
f < t U
t
f t, g < t V
t
g t,
h < t V
t
U
t
h t (t T).
Moreover, the presentation holds:
h(x, y) = inff(x, z) g(z, y) : z Z.
The sought functions exist by 3.8.2. The claim of uniqueness is straightfor-
ward from 3.8.4. The presentation of h via f and g raises no doubts.
9.5.12. Definition. Let f : X Z R and g : Z Y R. The function h,
given by 9.5.11, is called the -convolution (read: vel-convolution) of f and g and
is denoted by
f

g(x, y):= inff(x, z) g(z, y) : z Z.


By analogy, the +-convolution of f and g is dened by the rule
f
+
g(x, y):= inff(x, z) +g(z, y) : z Z.
9.5.13. Definition. A mapping f : X
2
R

+
is a K-ultrametric with
K R, K 1, if
(1) f(x, x) = 0 (x X);
(2) f(x, y) = f(y, x) (x, y X);
(3)
1
/
K
f(x, u) f(x, y) f(y, z) f(z, u) (x, y, z, u X).
9.5.14. Remark. Condition 9.5.13 (3) is often referred to as the (strong)
ultrametric inequality. In virtue of 9.5.12 this inequality may be rewritten as
K
1
f f

f.
9.5.15. 2-Ultrametric Lemma. To every 2-ultrametric f : X
2
R

+
there
corresponds a semimetric d such that
1
/
2
f d f.
170 Chapter 9
Let f
1
:= f and f
n+1
:= f
n

+
f (n N). Then
f
n+1
(x, y) f
n
(x, y) +f(y, y) = f
n
(x, y) (x, y X).
Thus, (f
n
) is a decreasing sequence. Put
d(x, y):= limf
n
(x, y) = inf
nN
f
n
(x, y).
Since
d(x, y) f
2n
(x, y) = f
n

+
f
n
(x, y) f
n
(x, z) +f
n
(z, y),
for n N, it follows that d(x, y) d(x, z) + d(z, y). The validity of 9.5.7 (1)
and 9.5.7 (2) is immediate.
We are left with proving that
1
/
2
f d. To this end, show that f
n

1
/
2
f for
n N. Proceed by way of induction.
The desired inequalities are obvious when n:= 1, 2. Assume now that f f
1

. . . f
n

1
/
2
f and at the same time f
n+1
(x, y) <
1
/
2
f(x, y) for some (x, y)
in X
2
and n 2.
For suitable z
1
, . . . , z
n
in X by construction
t := f(x, z
1
) +f(z
1
, z
2
) +. . . +f(z
n1
, z
n
) +f(z
n
, y)
<
1
/
2
f(x, y).
If f(x, z
1
)
t
/
2
then
t
/
2
f(z
1
, z
2
) + . . . + f(z
n
, y)
1
/
2
f(z
1
, y). It follows
that t f(x, z
1
) and t f(z
1
, y). On account of 9.5.13 (3),
1
/
2
f(x, y)
f(x, z
1
) f(z
1
, y) t. Whence we come to
1
/
2
f(x, y) > t
1
/
2
f(x, y), which is
false.
Thus, f(x, z
1
) <
t
/
2
. Find m N, m < n, satisfying
f(x, z
1
) +. . . +f(z
m1
, z
m
) <
t
/
2
;
f(x, z
1
) +. . . +f(z
m
, z
m+1
)
t
/
2
.
This is possible, since assuming m = n would entail the false inequality f(z
n
, y)
t
/
2
. (We would have
t
/
2
f(x, z
1
) + . . . + f(z
n1
, z
n
)
1
/
2
f(x, z
n
) and so
1
/
2
f(x, y) > t f(x, z
n
) f(z
n
, y)
1
/
2
f(x, y).)
We obtain the inequality
f(z
m+1
, z
m+2
) +. . . +f(z
n1
, z
n
) +f(z
n
, y) <
t
/
2
.
An Excursus into General Topology 171
Using the induction hypothesis, conclude that
f(x, z
m
) 2(f(x, z
1
) +. . . +f(z
m1
, z
m
)) t;
f(z
m
, z
m+1
) t;
f(z
m+1
, y) 2(f(z
m+1
, z
m+2
) +. . . +f(z
n
, y)) t.
Consequently, by the denition of 2-ultrametric
1
/
2
f(x, y) f(x, z
m
) f(z
m
, z
m+1
) f(z
m+1
, y) t
<
1
/
2
f(x, y).
We arrive at a contradiction, completing the proof.
9.5.16. Theorem. Every uniform space is multimetrizable.
Let (X, U
X
) be a uniform space. Take V U
X
. Put V
1
:= V V
1
. If now
V
n
U
X
then nd a symmetric entourage V = V
1
, a member of U
X
, satisfying
V V V V
n
. Dene V
n+1
:= V . Since by construction
V
n
V
n+1
V
n+1
V
n+1
V
n+1
I
X
I
X
V
n+1
;
therefore, (V
n
)
nN
is a decreasing family.
Given t R, dene a set U
t
by the rule
U
t
:=
_

_
, t < 0
I
X
, t = 0
V
infnN: t2
n
]
, 0 < t < 1
V
1
, t = 1
X
2
, t > 1.
By denition, the family t U
t
(t R) increases. Consider a unique function
f : X
2
R satisfying the next conditions (cf. 3.8.2 and 3.8.4)
f < t U
t
f t (t R).
If W
t
:= U
2t
for t R then
U
s
U
s
U
s
W
t
for s < t. Consequently, in virtue of 3.8.3 and 9.2.1 the mapping f is a 2-ultrametric.
Using 9.5.15, nd a semimetric d
V
such that
1
/
2
f d
V
f. Clearly, U
d
V
=
l V
n
: n N. It is also beyond a doubt that U
M
= U
X
for the multimetric
M:= d
V
: V U
X
.
9.5.17. Corollary. A topological space is uniformizable if and only if it is
a T
3
1
/
2
-space.
9.5.18. Corollary. A Tychono space is the same as a separated multimetric
space.
172 Chapter 9
9.6. Covers, and Partitions of Unity
9.6.1. Definition. Let E and F be covers of a subset of U in X; i.e., E , F
P(X) and U (E ) (F). It is said that F coarsens E or E renes F, if each
member of E is included in some member of F; i.e., ( E E ) ( F F) E F.
It is also said that E is a renement of F. Observe that if F is a subcover of E
(i.e., F E ) then E renes F.
9.6.2. Definition. A cover E of a set X is called locally nite (with respect
to a topology on X), if each point in X possesses a neighborhood (in the sense
of ) meeting only nite many members of E . In the case of the discrete topology on
X, such cover is called point nite. If X is regarded as furnished with a prescribed
topology then, speaking of a locally nite cover of X, we imply the topology .
9.6.3. Lefschetz Lemma. Let E be a point nite open cover of a normal
space X. Then there is an open cover G
E
: E E such that cl G
E
E for all
E E .
Let the set S comprise the mappings s : E Op(X) such that s(E ) = X
and for E E either s(E) = E or cl s(E) E. Given functions s
1
and s
2
, put
s
1
s
2
:= ( E E ) (s
1
(E) ,= E s
2
(E) = s
1
(E)). It is evident that (S, ) is
an ordered set and I
E
S. Show that S is inductive.
Given a chain S
0
in S, for all E E put s
0
(E):= s(E) : s S
0
. If s
0
(E) =
E then s(E) = E for all s S
0
. If the case s
0
(E) ,= E observe that s
0
(E) =
s(E) : s(E) ,= E, s S
0
.
Since the order of S
0
is linear, infer that s
0
(E) = s(E) for s S
0
with s(E) ,=
E. Hence, s
0
(E ) Op(X) and s
0
S
0
. It remains to verify that s
0
is a cover of X
(and so s
0
S). By the hypothesis of point niteness, given x X, there are some
E
1
, . . . , E
n
in E such that x E
1
. . . E
n
and x / E for the other members E
of E . If s
0
(E
k
) = E
k
for some k, then there is nothing to prove, for x s
0
(E ).
In the case when s
0
(E
k
) ,= E
k
for every k, there are s
1
, . . . , s
n
S
0
meeting the
conditions s
k
(E
k
) ,= E
k
(k := 1, 2 . . . , n). Since S
0
is a chain, it may be assumed
that s
n
s
1
, . . . , s
n1
. Moreover, x s
n
(E) E for an appropriate E in E .
It is clear that E E
1
, . . . , E
n
(because x / E for the other members E of E ).
Since s
0
(E) = s
n
(E), it follows that x s
0
(E).
By the KuratowskiZorn Lemma there is a maximal element s in S. Take
E E . If F := X s(E E), then F is closed and s(E) is a neighborhood of F.
For a suitable G in Op(X) by 9.3.10 F G cl G s(E). Put s(E) := G and
s(E) := s(E) for E ,= E (E E ). It is clear that s S. If s(E) = E, then s s
and so s = s. Moreover, s(E) cl G s(E) = E; i.e., cl s(E) E. If s(E) ,= E,
then cl s(E) E by denition. Thus, s is a sought cover.
9.6.4. Definition. Let f be a numeric or scalar-valued function on a topo-
logical space X, i.e. f : X F. The set supp(f) := clx X : f(x) ,= 0
An Excursus into General Topology 173
is the support of f. If supp(f) is a compact set then f is a compactly-supported
function or a function of compact support. The designation spt (f) := supp(f) is
used sometimes.
9.6.5. Let (f
e
)
eE
be a family of numeric functions on X and let E := supp(f
e
) :
e E be the family of their supports. If E is a point nite cover of X then the
family (f
e
)
eE
is summable pointwise. If in addition E is locally nite and every
member of (f
e
)
eE
is continuous, then the sum

eE
f
e
is also continuous.
It suces to observe that in a suitable neighborhood about a point in X
only nitely many members of the family (f
e
)
eE
are distinct from zero.
9.6.6. Definition. It is said that a family of functions (f : X [0, 1])
fF
is
a partition of unity on a subset U of X, if the supports of the members of the family
composes a point nite cover of X, and

fF
f(x) = 1 for all x U. The empty
family of functions in this context is treated as summable to unity at each point.
The term continuous partition of unity and the like are understood naturally.
9.6.7. Definition. Let E be a cover of a subset U of a topological space
and let F be a continuous partition of unity on U. If the family of supports
supp(f) : f F renes E then F is a partition of unity subordinate to E .
A possibility of nding such an F for E is also verbalized as follows: E admits
a partition of unity.
9.6.8. Each locally nite open cover of a normal space admits a partition
of unity.
By the Lefschetz Lemma, such cover U

: has an open renement


V

: satisfying the condition cl V

for all . By the Urysohn


Theorem, there is a continuous function g

: X [0, 1] such that g

(x) = 1 for
x V

and g

(x) = 0 for x XU

. Consequently, supp(g

) U

. In virtue of 9.6.5
the family (g

is summable pointwise to a continuous function g. Moreover,


g(x) > 0 for all x X by construction. Put f

:= g

/g ( ). The family (f

is what we need.
9.6.9. Definition. A topological space X is called paracompact, if each cover
of X has a locally nite open renement.
9.6.10. Remark. The theory of paracompactness contains deep and surpris-
ing facts.
9.6.11. Theorem. Every metric space is paracompact.
9.6.12. Theorem. A Hausdor topological space is paracompact if and only
if its every open cover admits a partition of unity.
174 Chapter 9
9.6.13. Remark. The metric space R
N
possesses a number of additional struc-
tures providing a stock of well-behaved, smooth (= innitely dierentiable) func-
tions (cf. 4.8.1).
9.6.14. Definition. A mollier or a mollifying kernel on R
N
is a real-valued
smooth function a having unit (Lebesgue) integral and such that a(x) > 0 for
[x[ < 1 and a(x) = 0 for [x[ 1. In this event, supp(a) = x R
N
: [x[ 1 is
the (unit Euclidean) ball B:= B
R
N.
9.6.15. Definition. A delta-like sequence is a family of real-valued (smooth)
functions (b

)
>0
such that, rst, lim
0
(sup [ supp(b

)[) = 0 and, second, the equal-


ity holds
_
R
N
b

(x) dx = 1 for all > 0. The terms -sequence and -like se-
quence are also in current usage. Such a sequence is often assumed countable
without further specication.
9.6.16. Example. The function a(x):= t exp(([x[
2
1)
1
) is taken as a most
popular mollier when extended by zero beyond the open ball int B, with the con-
stant t determined from the condition
_
R
N
a(x) dx = 1. Each mollier generates
the delta-like sequence a

(x):=
N
a(x/) (x R
N
).
9.6.17. Definition. Let f L
1,loc
(R
N
); i.e., let f be a locally integrable
function, that is, a function whose restriction to each compact subset of R
N
is
integrable. For a compactly-supported integrable function g the convolution f g
is dened as
f g(x):=
_
R
N
f(x y)g(y) dy (x R
N
).
9.6.18. Remark. The role of a mollifying kernel and the corresponding delta-
like sequence (a

)
>0
becomes clear from inspecting the aftermath of applying the
smoothing process f (fa

)
>0
to a function f belonging to L
1,loc
(R
N
) (cf. 10.10.7
(5)).
9.6.19. The following statements are valid:
(1) to every compact set K in the space R
N
and every neighborhood U
of K there corresponds a truncator (= a bump function) :=
K,U
, i.e. a smooth
mapping : R
N
[0, 1] such that K int = 1 and supp() U;
(2) assume that U
1
, . . . , U
n
Op(R
N
) and U
1
. . . U
n
is a neigh-
borhood of a compact set K; there are smooth functions
1
, . . . ,
n
: R
N
[0, 1]
such that supp(
k
) U
k
and

n
k=1

k
(x) = 1 for x in some neighborhood of K.
(1) Put := d(K, R
N
U) := inf[x y[ : x K, y / U. It is clear that
> 0. Given > 0, denote the characteristic function of K + B by

. Take
a delta-like sequence (b

)
>0
of positive functions and put :=

. When
and + with := sup [ supp(b

)[, observe that is a sought function.


An Excursus into General Topology 175
(2) By the Diedonn`e Lemma there are closed sets F
k
, with F
k
U
k
, composing
a cover of K. Put K
k
:= F
k
K and choose some truncators
k
:=
K
k
,U
k
. The
functions
k
/

n
k=1

k
(k := 1, . . . , n), dened on

n
k=1

k
> 0, meet the claim
after extension by zero onto

n
k=1

k
= 0 and multiplication by a truncator
corresponding to an appropriate neighborhood of K.
9.6.20. Countable Partition Theorem. Let E be a family of open sets
in R
N
and := E . There is a countable partition of unity which is composed
of smooth compactly-supported functions on R
N
and subordinate to the cover E
of .
Rene from E a countable locally nite cover A of with compact sets so
that the family (:= int )
A
be also an open cover of . Choose an open cover
(V

)
A
of from the condition cl V

for A. In virtue of 9.6.19 (1) there


are truncators

:=
cl V

,
. Putting

(x) :=

(x)/

(x) for x and

(x):= 0 for x R
N
, arrive to a sought partition.
9.6.21. Remark. It is worth observing that the so-constructed partition of uni-
ty (

)
A
possesses the property that to each compact subset K of there corre-
spond a nite subset A
0
of A and a neighborhood U of K such that

A
0

(x) =
1 for all x U (cf. 9.3.17 and 9.6.19 (2)).
Exercises
9.1. Give examples of pretopological and topological spaces and constructions leading to them.
9.2. Is it possible to introduce a topology by indicating convergent lters or sequences?
9.3. Establish relations between topologies and preorders on a nite set.
9.4. Describe topological spaces in which the union of every family of closed sets is closed.
What are the continuous mappings between such spaces?
9.5. Let (f

: X (Y

))

be a family of mappings. A topology on X is called


admissible (in the case under study), if for every topological space (Z, ) and every mapping
g : Z X the following statement holds: g : (Z, ) (X, ) is continuous if and only if so
is each mapping f

g ( ). Demonstrate that the weakest topology on X making every


f

( ) continuous is the strongest admissible topology (in the case under study).
9.6. Let (f

: (X

) Y )

be a family of mappings. A topology on Y is called


admissible (in the case under study), if for every topological space (Z, ) and every mapping
g : Y Z the following statement is true: g : (Y, ) (Z, ) is continuous if and only if each
mapping gf

( ) is continuous. Demonstrate that the strongest topology on Y making every


f

( ) continuous is the weakest admissible topology (in the case under study).
9.7. Prove that in the Tychono product of topological spaces, the closure of the product
of subsets of the factors is the product of closures:
cl
_

_
=

cl A

.
9.8. Show that a Tychono product is a Hausdor space if and only if so is every factor.
9.9. Establish compactness criteria for subsets of classical Banach spaces.
9.10. A Hausdor space X is called H-closed, if X is closed in every ambient Hausdor
space. Prove that a regular H-closed space is compact.
176 Chapter 9
9.11. Study possibilities of compactifying a topological space.
9.12. Prove that the Tychono product of uncountably many real axes fails to be a normal
space.
9.13. Show that each continuous function on the product of compact spaces depends on at
most countably many coordinates in an evident sense (specify it!).
9.14. Let A be a compact subset and let B be a closed subset of a uniform space, with
A B = . Prove that V (A) V (B) = for some entourage V .
9.15. Prove that a completion (in an appropriate sense) of the product of uniform spaces is
uniformly homeomorphic (specify!) to the product of completions of the factors.
9.16. A subset A of a separated uniform space is called precompact if a completion of A is
compact. Prove that a set is precompact if and only if it is totally bounded.
9.17. Which topological spaces are metrizable?
9.18. Given a uniformizable space, describe the strongest uniformity among those inducing
the initial topology.
9.19. Verify that the product of a paracompact space and a compact space is paracompact.
Is paracompactness preserved under general products?
Chapter 10
Duality and Its Applications
10.1. Vector Topologies
10.1.1. Definition. Let (X, F, +, ) be a vector space over a basic eld F.
A topology on X is a topology compatible with vector structure or, briey, a vector
topology, if the following mappings are continuous:
+ : (X X, ) (X, ),
: (F X,
F
) (X, ).
The space (X, ) is then referred to as a topological vector space.
10.1.2. Let
X
be a vector topology. The mappings
x x +x
0
, x x (x
0
X, F 0)
are topological isomorphisms in (X,
X
).
10.1.3. Remark. It is beyond a doubt that a vector topology on a space X
possesses the next linearity property:
(x +y) = (x) +(y) (, F 0; x, y X),
where in accordance with the general agreements (cf. 1.3.5 (1))
U
x+y
(x) +(y)
( U
x
(x) & U
y
(y)) U
x
+U
y
U
x+y
.
In this regard a vector topology is often called a linear topology and a topological
vector space, a linear topological space. This terminology should be used on the
understanding that a topology may possess the linearity property while failing
to be linear. For instance, such is the discrete topology of a nonzero vector space.
178 Chapter 10
10.1.4. Theorem. Let X be a vector space and let N be a lter on X.
There is a vector topology on X such that N = (0) if and only if the following
conditions are fullled:
(1) N +N = N ;
(2) N consists of absorbing sets;
(3) N has a base of balanced sets.
Moreover, (x) = x +N for all x X.
: Let be a vector topology and N = (0). From 10.1.2 infer that
(x) = x + N for x X. It is also clear that (1) reformulates the continuity
property of addition at zero (of the space X
2
). Condition (2) may be rewritten
as
F
(0)x N for every x in X, which is the continuity property of the mapping
x at zero (of the space R) for every x in X. Condition (3) with account taken
of (2) may in turn be rendered in the form
F
(0)N = N , which is the continuity
property of scalar multiplication at zero (of the space F X).
: Let N be a lter satisfying (1)(3). It is evident that N l 0. Put
(x):= x+N . Then is a pretopology. From the denition of and (1) it follows
that is a topology, with every translation continuous and addition continuous
at zero in X
2
. Thus, addition is continuous at every point of X
2
. The validity
of (2) and (3) means that the mapping (, x) x is jointly continuous at zero
and continuous at zero in the rst argument given the second argument. By virtue
of the identity
x
0
x
0
=
0
(x x
0
) + (
0
)x
0
+ (
0
)(x x
0
),
we are left with examining the continuity property of scalar multiplication at zero
in the second argument given the rst argument. In other words, it is necessary
to show that N N for F. With this in mind, nd n N such that [[ n.
Let V in N and W in N be such that W is balanced and W
1
+ . . . + W
n
V ,
where W
k
:= W. Then W = n
_

/
n
W
_
nW W
1
+. . . +W
n
V .
10.1.5. Theorem. The set VT(X) of all vector topologies on X presents
a complete lattice. Moreover,
sup
VT(X)
E = sup
T(X)
E
for every subset E of VT(X).
Let := sup
T(X)
E . Since for E each translation by a vector is a topolog-
ical isomorphism in (X, ); therefore, this mapping is a topological isomorphism in
(X, ). Using 9.1.13, observe that the lter (0) meets conditions 10.1.4 (1)10.1.4
(3), since these conditions are fullled for every lter (0) with E . It remains
to refer to 1.2.14.
Duality and Its Applications 179
10.1.6. Theorem. The inverse image of a vector topology under a linear
operator is a vector topology.
Take T L(X, Y ) and VT(Y ). Put := T
1
(). If x

x and
y

y in (X, ) then by 9.2.8 Tx

Tx and Ty

Ty. So T(x

+y

) T(x+y).
This means in virtue of 9.2.10 that x

+y

x+y in (X, ). Thus, (x) = x+(0)


for all x X and, moreover, (0) + (0) = (0). Successively applying 3.4.10
and 3.1.8 to the linear correspondence T
1
, observe that the lter (0) = T
1
((0))
consists of absorbing sets and has a base of balanced sets. The reason is as follows:
by 10.1.4 the lter (0) possesses these two properties. Once again using 10.1.4,
conclude that VT(X).
10.1.7. The product of vector topologies is a vector topology.
Immediate from 10.1.5 and 10.1.6.
10.1.8. Definition. Let A and B be subsets of a vector space. It is said that
A is B-stable if A+B A.
10.1.9. To every vector topology on X there corresponds a unique uniformity
U

having a base of I
X
-stable sets and such that = (U

).
Given U (0), put V
U
:= (x, y) X
2
: y x U. Observe the obvious
properties:
I
X
V
U
; V
U
+I
X
= V
U
; (V
U
)
1
= V
U
;
V
U
1
U
2
V
U
1
V
U
2
; V
U
1
V
U
2
V
U
1
+U
2
for all U, U
1
, U
2
(0). Using 10.1.4, infer that U

:= l V
U
: U (0) is
a uniformity and = (U

). It is also beyond a doubt that U

has a base of
I
X
-stable sets.
If now U is another uniformity such that (U ) = , and W is some I
X
-stable
entourage in U ; then W = V
W(0)
. Whence the sought uniqueness follows.
10.1.10. Definition. Let (X, ) be a topological vector space. The unifor-
mity U

, constructed in 10.1.9, is the uniformity of X.


10.1.11. Remark. Considering a topological vector space, we assume it to be
furnished with the corresponding uniformity without further specication.
10.2. Locally Convex Topologies
10.2.1. Definition. A vector topology is locally convex if the neighborhood
lter of each point has a base of convex sets.
180 Chapter 10
10.2.2. Theorem. Let X be a vector space and let N be a lter on X. There
is a locally convex topology on X such that N = (0) if and only if
(1)
1
/
2
N = N ;
(2) N has a base of absorbing absolutely convex sets.
: By virtue of 10.1.2 the mapping x 2x is a topological isomorphism.
This means that
1
/
2
N = N . Now take U N . By hypothesis there is a convex
set V in N such that V U. Applying 10.1.4, nd a balanced set W satisfying
W V . Using the Motzkin formula and 3.1.14, show that the convex hull co(W)
is absolutely convex. Moreover, W co(W) V U.
: An absolutely convex set is balanced. Consequently, N satises 10.1.4 (2)
and 10.1.4 (3). If V N and W is a convex set, W N and W V ; then
1
/
2
W N . Furthermore,
1
/
2
W +
1
/
2
W W V because of the convexity
property of W. This means that N +N = N . It remains to refer to 10.1.4.
10.2.3. Corollary. The set LCT(X) of all locally convex topologies on X is
a complete lattice. Moreover,
sup
LCT(X)
E = sup
T(X)
E
for every subset E of LCT(X).
10.2.4. Corollary. The inverse image of a locally convex topology under a lin-
ear operator is a locally convex topology.
10.2.5. Corollary. The product of locally convex topologies is a locally con-
vex topology.
10.2.6. The topology of a multinormed space is locally convex.
10.2.7. Definition. Let be a locally convex topology on X. The set of
all everywhere-dened continuous seminorms on X is called the mirror (rarely, the
spectrum) of and is denoted by M

. The multinormed space (X, M

) is called
associated with (X, ).
10.2.8. Theorem. Each locally convex topology coincides with the topology
of the associated multinormed space.
Let be a locally convex topology on X and let := (M

) be the topology
of the associated space (X, M

). Take V (0). By 10.2.2 there is an absolutely


convex neighborhood B of zero, B (0), such that B V . In virtue of 3.8.7
p
B
< 1 B p
B
1.
It is obvious that p
B
is continuous (cf. 7.5.1); i.e., p
B
M

, and so p
B
< 1 (0).
Consequently, V (0). Using 5.2.10, infer that (x) = x+(0) x+(0) = (x);
i.e., . Furthermore, by denition.
Duality and Its Applications 181
10.2.9. Definition. A vector space, endowed with a separated locally convex
topology, is a locally convex space.
10.2.10. Remark. Theorem 10.2.8 in slightly trimmed form is often verbal-
ized as follows: The concept of locally convex space and the concept of separated
multinormed space have the same scope. For that reason, the terminology connect-
ed with the associated multinormed space is lavishly applied to studying a locally
convex space (cf. 5.2.13).
10.2.11. Definition. Let be a locally convex topology on X. The symbol
(X, )

(or, in short, X

) denotes the subspace of X


#
that comprises all continuous
linear functionals. The space (X, )

is the dual (or -dual) of (X, ).


10.2.12. (X, )

= [[(p) : p M

.
10.2.13. Prime Theorem. The prime mapping (X, )

from LCT(X)
to Lat(X
#
) preserves suprema; i.e.,
(X, sup E )

= sup(X, )

: E
for every subset E of LCT(X).
If E = then sup E is the trivial topology

of X and, consequently,
(X,

= 0 = inf Lat(X
#
) = sup
Lat(X
#
)
. By virtue of 9.2.7 the prime mapping
increases. Given a nonempty E , from 2.1.5 infer that
(X, sup E )

sup(X, )

: E .
If f (X, sup E )

, then in view of 10.2.12 and 9.1.13 there are topologies

1
, . . . ,
n
E such that f (X,
1
. . .
n
)

. Using 10.2.12 and 5.3.7, nd


p
1
M

1
, . . . , p
n
M

n
satisfying f [[(p
1
. . . p
n
). Recalling 3.5.7 and 3.7.9,
observe that [[(p
1
+. . . +p
n
) = [[(p
1
) +. . . +[[(p
n
). Finally,
f (X,
1
)

+. . . + (X,
n
)

= (X,
1
)

. . . (X,
n
)

.
10.3. Duality Between Vector Spaces
10.3.1. Definition. Let X and Y be vector spaces over the same ground eld
F. Also, Assume given a bilinear form (or, as it is called sometimes, a bracketing)
[ acting from XY to F, i.e. a mapping linear in each of its arguments. Given
x X and y Y , put
x [ : y x [ y, [ : X F
Y
, X[ Y
#
;
[ y : x x [ y, [ : Y F
X
, [ Y X
#
.
The mappings [ and [ are the bra-mapping and the ket-mapping of the initial
bilinear form. By analogy, a member of X[ is a bra-functional on X and a member
of [ Y is a ket-functional on Y .
182 Chapter 10
10.3.2. The bra-mapping and the ket-mapping are linear operators.
10.3.3. Definition. A bracketing of some vector spaces X and Y is a pairing,
if its bra-mapping and ket-mapping are monomorphisms. In this case we say that
X and Y are (set) in duality, or present a duality pair, or that Y is the pair-dual
of X, etc. This is written down as X Y . Each of the bra-mapping and the ket-
mapping is then referred to as dualization. For suggestiveness, the corresponding
pairing of some spaces in duality is also called their duality bracket.
10.3.4. Examples.
(1) Let X Y with duality bracket [ . Given (y, x) Y X,
put y [ x := x [ y. It is immediate that the new bracketing is a pairing of Y
and X. Moreover, the pairs of dualizations with respect to the old (direct) and
new (reverse) duality brackets are the same. It thus stands to reason to draw
no distinction between the two duality brackets unless in case of an emergency
(cf. 10.3.3). For instance, Y is the pair-dual of X in the direct duality bracket
if and only if X is the pair-dual of Y in the reverse duality bracket. Therefore,
the hair-splitting is neglected and a unied term pair-dual is applied to each
of the spaces in duality, with duality treated as a whole abstract phenomenon.
Observe immediately that the mapping x [ y
R
:= Rex [ y sets in duality the real
carriers X
R
and Y
R
. By way of taking liberties, the previous notation is sometimes
reserved for the arising duality X
R
Y
R
; i.e., it is assumed that x [ y := x [ y
R
,
on considering x and y as members of the real carriers.
(2) Let H be a Hilbert space. The inner product on H sets H and H

in duality. The prime mapping is then coincident with the ket-mapping.


(3) Let (X, ) be a locally convex space and let X

be the dual of X.
The natural evaluation mapping (x, x

) x

(x) sets X and X

in duality.
(4) Let X be a vector space and let X
#
:= L(X, F) be the (algebraic)
dual of X. It is clear that the evaluation mapping (x, x
#
) x
#
(x) sets the spaces
in duality.
10.3.5. Definition. Let X Y . The inverse image in X of the Tychono
topology on F
Y
under the bra-mapping, further denoted by (X, Y ), is the bra-
topology or the weak topology on X induced by Y . The bra-topology (Y, X)
of Y X is the ket-topology of X Y or the weak topology on Y induced by X.
10.3.6. The bra-topology is the weakest topology making every ket-functional
continuous. The ket-topology is the weakest topology making every bra-functional
continuous.
x

x (in (X, Y )) x

[ x [ (in F
Y
) ( y Y ) x

[ (y)
x [ (y) ( y Y ) x

[ y x [ y ( y Y ) [ y(x

) [ y(x) ( y Y )
x

x (in [ y
1
(
F
))
Duality and Its Applications 183
10.3.7. Remark. The notation (X, Y ) agrees perfectly with that of the
weak multinorm in 5.1.10 (4). Namely, (X, Y ) is the topology of the multinorm
[ [ y[ : y Y . Likewise, (Y, X) is the topology of the multinorm [x [ [ :
x X.
10.3.8. The spaces (X, (X, Y )) and (Y, (Y, X)) are locally convex.
Immediate from 10.2.4 and 10.2.5.
10.3.9. Dualization Theorem. Each dualization is an isomorphism between
the pair-dual and the weak dual of the pertinent member of a duality pair.
Consider a duality pair X Y . We are to prove exactness for the sequences
0 X
]
(Y, (Y, X))

0; 0 Y
])
(X, (X, Y ))

0.
Since the ket-mapping of X Y is the bra-mapping of Y X, it suces to show
that the rst sequence is exact. The bra-mapping is a monomorphism by denition.
Furthermore, from 10.2.13 and 10.3.6 it follows that
(Y, (Y, X))

= (Y, supx [
1
(
F
) : x X)

= sup(Y, x [
1
(
F
))

: x X = lin((Y, f
1
(
F
))

: f X[) = X[,
since in view of 5.3.7 and 2.3.12 (Y, f
1
(
F
))

= f : F (f Y
#
).
10.3.10. Remark. Theorem 10.3.9 is often referred to as the theorem on the
general form of a weakly continuous functional. Here a useful convention reveals
itself: apply the base form weak when using objects and properties that are
related to weak topologies. Observe immediately that, in virtue of 10.3.9, Example
10.3.4 (3) actually lists all possible duality brackets. That is why in what follows
we act in accordance with 5.1.11, continuing the habitual use of the designation
(x, y) := x [ y, since it leads to no misunderstanding. For the same reason, given
a vector space X, we draw no distinction between the pair-dual of a space X and the
weak dual of X. In other words, considering a duality pair X Y , we sometimes
identify X with (Y, (Y, X))

and Y with (X, (X, Y ))

, which justies writing


X

= Y and Y

= X.
10.3.11. Remark. A somewhat obsolete convention relates to X X

with
X a normed space. The ket-topology (X

, X) is customarily called the weak

topology (read: weak-star topology) in X

, which reects the concurrent notation


X

for X

. The term weak

proliferates in a routine fashion elsewhere.


184 Chapter 10
10.4. Topologies Compatible with Duality
10.4.1. Definition. Take a duality pair X Y and let be a locally convex
topology on X. It is said that is compatible with duality (between X and Y
by pairing X Y ), provided that (X, )

= [ Y . A locally convex topology


on Y is compatible with duality (by pairing X Y ), if is compatible with duality
(by pairing Y X); i.e., if the equality holds: (Y, )

= X[. A unied concise


term compatible topology is also current in each of the above cases.
10.4.2. Weak topologies are compatible.
Follows from 10.3.9.
10.4.3. Let (X, Y ) stand for the least upper bound of the set of all locally
convex topologies on X compatible with duality (between X and Y ). Then the
topology (X, Y ) is also compatible.
Denote the set of all compatible topologies on X by E . Theorem 10.2.13
readily yields the equalities
(X, (X, Y ))

= sup(X, )

: E = sup[ Y : E = [ Y ,
because E is nonempty by 10.4.2.
10.4.4. Definition. The topology (X, Y ), constructed in 10.4.3 (i.e., the
nest locally convex topology on X compatible with duality by pairing X Y ), is
the Mackey topology (on X induced by X Y ).
10.4.5. MackeyArens Theorem. A locally convex topology on X is com-
patible with duality between X and Y if and only if
(X, Y ) (X, Y ).
By 10.2.13 the prime mapping (X, )

preserves suprema and, in par-


ticular, increases. Therefore, given in the interval of topologies, from 10.4.2
and 10.4.3 obtain
[ Y = (X, (X, Y )) (X, )

(X, (X, Y ))

= [ Y .
The remaining claim is obvious.
10.4.6. Mackey Theorem. All compatible topologies have the same
bounded sets in stock.
Duality and Its Applications 185
A stronger topology has fewer bounded sets. So, to prove the theorem
it suces to show that if some set U is weakly bounded in X (= is bounded in the
bra-topology) then U is bounded in the Mackey topology.
Take a seminorm p from the mirror of the Mackey topology and demonstrate
that p(U) is bounded in R. Put X
0
:= X/ ker p and p
0
:= p
X/ ker p
. By 5.2.14, p
0
is
clearly a norm. Let : X X
0
be the coset mapping. It is beyond a doubt that
(U) is weakly bounded in (X
0
, p
0
). From 7.2.7 it follows that (U) is bounded
in the norm p
0
. Since p
0
= p; therefore, U is bounded in (X, p).
10.4.7. Corollary. Let X be a normed space. Then the Mackey topology
(X, X

) coincides with the initial norm topology on X.


It suces to refer to the Kolmogorov Normability Criterion implying that the
space X with the topology (X, X

) ner than the original topology is normable.


Appealing to 5.3.4 completes the proof.
10.4.8. Strict Separation Theorem. Let (X, ) be a locally convex space.
Assume further that K and V are nonempty convex subsets of X with K compact,
V closed and K V = . Then there is a functional f, a member of (X, )

, such
that
sup Re f(K) < inf Re f(V ).
A locally convex space is obviously a regular space. Since K is compact,
it thus follows that, for an appropriate convex neighborhood of zero, say W, the
set U := K +W does not meet V (it suces to consider the lterbases comprising
all subsets of the form K +W and V +W, with W a closed neighborhood of zero).
By 3.1.10, U is convex. Furthermore, K int U = core U. By the Eidelheit
Separation Theorem, there is a functional l, a member of (X
R
)
#
, such that the
hyperplane l = 1 in X
R
separates V from U and does not meet the core of U.
Obviously, l is bounded above on W and so l (X
R
, )

by 7.5.1. If f := Re
1
l
then, of view of 3.7.5, f (X, )

. It is clear that f is a sought functional.


10.4.9. Mazur Theorem. All compatible topologies have the same closed
convex sets in stock.
A stronger topology has more closed sets. So, to prove the theorem it suces
in view of 10.4.5 to show that if U is a convex set closed in the Mackey topology
then U is weakly closed. The last claim is beyond a doubt since by Theorem 10.4.8
U is the intersection of weakly closed sets of the form Re f t, with f a (weakly)
continuous linear functional and t R.
10.5. Polars
10.5.1. Definition. Let X and Y be sets and let F XY be a correspon-
186 Chapter 10
dence. Given a subset U of X and a subset V of Y, put
(U):=
F
(U):= y Y : F
1
(y) U;

1
(V ):=
1
F
(V ):= x X : F(u) V .
The set (U) is the (direct) polar of U (under F), and the set
1
(V ) is the (reverse)
polar of V (under F).
10.5.2. The following statements are valid:
(1) (u):= (u) = F(u) and (U) =
uU
(u);
(2) (

) =

(U

);
(3)
1
F
(V ) =
F
1(V );
(4) U
1
U
2
(U
1
) (U
2
);
(5) U V F (V (U) & U
1
(V );
(6) U
1
((U)).
10.5.3. Akilov Criterion. A subset U of X is the polar of some subset of Y
if and only if given x X U there is an element y in Y such that
U
1
(y), x /
1
(y).
: If U =
1
(V ) then U =
vV

1
(v) by 10.5.2 (1).
: The inclusion U
1
(y) means that y (U). Thus, by hypothesis
U =
y(U)

1
(y) =
1
((U)).
10.5.4. Corollary. The set
1
((U)) is the (inclusion) least polar greater
than U.
10.5.5. Definition. The set
1
F
(
F
(U)) is the bipolar of a subset U (under
the correspondence F).
10.5.6. Examples.
(1) Let (X, ) be an ordered set and let U be a subset of X. Then

(U) is the collection of all upper bounds of U (cf. 1.2.7).


(2) Let (H, ( , )
H
) be a Hilbert space and F := (x, y) H
2
:
(x, y)
H
= 0. Then (U) =
1
(U) = U

for every subset U of H. The bipolar


of U in this case coincides with the closed linear span of U, that is, the closure of
the linear span of U.
(3) Let X be a normed space and let X

be the dual of X. Consider


F := (x, x

) : x

(x) = 0. Then (X
0
) = X

0
and
1
(X
0
) =

X
0
for a subspace
X
0
of X and a subspace X
0
of X

(cf. 7.6.8). Moreover,


1
((X
0
)) = cl X
0
by 7.5.14.
Duality and Its Applications 187
10.5.7. Definition. Let X Y . Put
pol := (x, y) X Y : Rex [ y 1;
abs pol := (x, y) X Y : [x [ y[ 1.
To refer to direct or inverse polars under pol , we use the unied term polar (with
respect to X Y ) and the unied designations (U) and (V ). In the case of the
correspondence abs pol , we speak of absolute polars (with respect to X Y ) and
write U

and V

(for U X and V Y ).
10.5.8. Bipolar Theorem. The bipolar
2
(U) := ((U)) is the (inclusion)
least weakly closed conical segment greater than U.
Straightforward from 10.4.8 and the Akilov Criterion.
10.5.9. Absolute Bipolar Theorem. The absolute bipolar U

:= (U

is
the (inclusion) least weakly closed absolutely convex set greater than U.
It suces, rst, to observe that the polar of a balanced set U coincides with
the absolute polar of U and, second, to apply 10.5.8.
10.6. Weakly Compact Convex Sets
10.6.1. Let X be a locally convex real vector space and let p : X R be
a continuous sublinear functional on X. Then the (topological) subdierential (p)
is compact in the topology (X

, X).
Put Q:=

xX
[p(x), p(x)] and endow Q with the Tychono topology.
Evidently (p) Q, with the Tychono topology on Q and (X

, X) inducing the
same topology on (p). It is beyond a doubt that the set (p) is closed in Q by the
continuity of p. Taking note of the Tychono Theorem and 9.4.9, conclude that
(p) is a (X

, X)-compact set.
10.6.2. The balanced subdierential of each continuous seminorm is weakly
compact.
10.6.3. Theorem. Let X be a real vector space. A subset U of X
#
is the
subdierential of a (unique total) sublinear functional s
U
: X R if and only if U
is nonempty, convex and (X
#
, X)-compact.
: Let U = (s
U
) for some s
U
. The uniqueness of s
U
is ensured by 3.6.6.
In view of 10.2.12 it is easy that the mirror of the Mackey topology (X, X
#
) is
the strongest multinorm on X (cf. 5.1.10 (2)). Whence we infer that the functional
s
U
is continuous with respect to (X, X
#
). In virtue of 10.6.1 the set U is compact
in (X
#
, X). The convexity and nonemptiness of U are obvious.
: Put s
U
(x):= supl(x) : l U. Undoubtedly, s
U
is a sublinear functional
and doms
U
= X. By denition, U (s
U
). If l (s
U
) and l / U, then by the
Strict Separation Theorem and the Dualization Theorem s
U
(x) < l(x) for some x
in X. This is a contradiction.
188 Chapter 10
10.6.4. Definition. The sublinear functional s
U
, constructed in 10.6.3, is the
supporting function of U. The term support function is also in current usage.
10.6.5. KrenMilman Theorem. Each compact convex set in a locally
convex space is the closed convex hull (= the closure of the convex hull) of the set
of its extreme points.
Let U be such subset of a space X. It may be assumed that the space
X is real and U ,= . By virtue of 9.4.12, U is compact with respect to the
topology (X, X

). Since (X, X

) is induced in X by the topology (X

#
, X

)
on X

#
; therefore, U = (s
U
). Here (cf. 10.6.3) s
U
: X

R acts by the rule


s
U
(x

) := sup x

(U). By the KrenMilman Theorem in subdierential form, the


set ext U of the extreme points of U is not empty. The closure of the convex hull
of ext U is a subdierential by Theorem 10.6.3. Moreover, this set has s
U
as its
supporting function, thus coinciding with U (cf. 3.6.6).
10.6.6. Let X Y and let S be a conical segment in X. Assume further that
p
S
is the Minkowski functional of S. The polar (S) is the inverse image of the
(algebraic) subdierential (p
S
) under the ket-mapping; i.e.,
(S) = [ (p
S
)
1
R
.
If S is absolutely convex, then the absolute polar S

is the inverse image of the


(algebraic) balanced subdierential [[(p
S
) under the ket-mapping; i.e.,
S

= [ [[(p
S
)
1
.
If y Y
R
and y [ (p
S
)
1
R
then [ y
R
belongs to (p
S
). Hence, Rex [ y =
x [ y
R
= [ y
R
(x) p
S
(x) 1 for x S, because S p
S
1 by the Gauge
Theorem. Consequently, y (S).
If, in turn, y (S) then [ y
R
belongs to (p
S
). Indeed, 1 > p
S
(
1
x) for
all x in X
R
and > p
S
(x); i.e.,
1
x p
S
< 1 S. Whence
1
x [ y
R
=
Re
1
x [ y =
1
Rex [ y 1. Finally, observe that [ y
R
(x) . Since
is arbitrary, this inequality means that [ y
R
(x) p
S
(x). In other words, y
[ (p
s
)
1
R
, which implies that (S) = [ (p
S
)
1
R
. The remaining claim follows
from the properties of the complexier (cf. 3.7.3 and 3.7.9).
10.6.7. AlaogluBourbaki Theorem. The polar of a neighborhood of zero
of each compatible topology is a weakly compact convex set.
Let U be a neighborhood of zero in a space X and let (U) be the polar of U
(with respect to X X

). Since U p 1 for some continuous seminorm p,


by 10.5.2 (4), (U) (p 1) = (B
p
) = B

p
. Using 10.6.6 and recalling that
p is the Minkowski functional of B
p
, obtain the inclusion (U) [[(p). By virtue
of 10.6.2 the topological balanced subdierential [[(p) is (X

, X)-compact. By def-
inition (U) is weakly closed. To infer the (X

, X)-compactness of (U), it re-


mains to appeal to 9.4.9. The convexity property of (U) is beyond a doubt.
Duality and Its Applications 189
10.7. Reexive Spaces
10.7.1. Kakutani Criterion. A normed space is reexive if and only if its
unit ball is weakly compact.
: Let X be reexive, i.e.

(X) = X

. In other words, the image of X


under the double prime mapping coincides with X

. Since the ball B


X
is the
polar of the ball B
X
with respect to X

; therefore, B
X
is a (X

, X

)-
compact set by the AlaogluBourbaki Theorem. It remains to observe that B
X
is
(the image under the double prime mapping of) B
X
, and (X, X

) is (the inverse
image under the double prime mapping of) (X

, X

).
: Consider the duality pair X

. By denition, the ball B


X
presents
the bipolar of B
X
(more precisely, the bipolar of (B
X
)

). Using the Absolute


Bipolar Theorem and observing that the weak topology (X, X

) is induced in X
by the topology (X

, X

), conclude that B
X
= B
X
(because of the obvious
convexity and closure properties of B
X
, the latter following from compactness since
X is separated). Thus, X is reexive.
10.7.2. Corollary. A space X is reexive if and only if every bounded closed
convex set in X is weakly compact.
10.7.3. Corollary. Every closed subspace of a reexive space is reexive.
By the Mazur Theorem, such subspace and, hence, the unit ball of it are
weakly closed. It thus suces to apply the Kakutani Criterion twice.
10.7.4. Pettis Theorem. A Banach space and its dual are (or are not) re-
exive simultaneously.
If X is reexive then (X

, X) coincides with (X

, X

). Therefore, by the
AlaogluBourbaki Theorem, B
X
is (X

, X

)-compact. Consequently, X

is re-
exive. If, in turn, X

is reexive, then so is X

by what was proven. However, X,


as a Banach space, is a closed subspace of X

. Thus, X is reexive by 10.7.3.


10.7.5. James Theorem. A Banach space is reexive if and only if each
continuous (real) linear functional attains its supremum on the unit ball of the
space.
10.8. The Space C(Q, R)
10.8.1. Remark. Throughout Section 10.8 let Q stand for a nonempty Haus-
dor compact space, denoting by C(Q, R) the set of continuous real-valued func-
tions on Q. Unless specied otherwise, C(Q, R) is furnished with the natural point-
wise algebraic operations and order and equipped with the sup-norm | |:= | |

related to the Chebyshev metric (cf. 4.6.8). Keeping this in mind, we treat the
statements like C(Q, R) is a vector lattice, C(Q, R) is a Banach algebra, etc.
Other structures, if ever introduced in C(Q, R), are specied deliberately.
190 Chapter 10
10.8.2. Definition. A subset L of C(Q, R) is a sublattice (in C(Q, R)) if
f
1
f
2
L and f
1
f
2
L for f
1
, f
2
L, where, as usual,
f
1
f
2
(q):= f
1
(q) f
2
(q),
f
1
f
2
(q):= f
1
(q) f
2
(q) (q Q).
10.8.3. Remark. Observe that to be a sublattice in C(Q, R) means more
than to be a lattice with respect to the order induced from C(Q, R).
10.8.4. Examples.
(1) , C(Q, R), and the closure of a sublattice.
(2) The intersection of each family of sublattices is also a sublattice.
(3) Let L be a sublattice and let Q
0
be a subset of Q. Put
L
Q
0
:= f C(Q, R) : ( g L) g(q) = f(q) (q Q
0
).
Then L
Q
0
is a sublattice. Moreover, L L
Q
0
.
(4) Let Q
0
be a compact subset of Q. Given a sublattice L in C(Q, R),
put
L

Q
0
:=
_
f

Q
0
: f L
_
.
Therefore,
L
Q
0
=
_
f C(Q, R) : f

Q
0
L

Q
0
_
.
It is clear that L

Q
0
is a sublattice of C(Q
0
, R). Furthermore, if L is a vec-
tor sublattice of C(Q, R) (i.e., a vector subspace and simultaneously a sublattice
of C(Q, R)); then L

Q
0
is a vector sublattice of C(Q
0
, R) (certainly, if Q
0
,= ).
(5) Let Q:= 1, 2. Then C(Q, R) R
2
. Each nonzero vector sublat-
tice of R
2
is given as
(x
1
, x
2
) R
2
:
1
x
1
=
2
x
2

for some
1
,
2
R
+
.
(6) Let L be a vector sublattice of C(Q, R). For q Q, the alternative
is oered: either L
q]
= C(Q, R) or L
q]
= f C(Q, R) : f(q) = 0. If q
1
and
q
2
are distinct points of Q and L

q
1
,q
2
]
,= 0, then by 10.8.4 (5) there are some
1
,

2
R
+
such that
L
q
1
,q
2
]
= f C(Q, R) :
1
f(q
1
) =
2
f(q
2
).
Moreover, if L contains a constant function other than zero (i.e. a nonzero multiple
of the constantly-one function 1) then as
1
and
2
in the above presentation
of L
q
1
,q
2
]
the unity, 1, may be taken.
Duality and Its Applications 191
10.8.5. Let L be a sublattice of C(Q, R). A function f in C(Q, R) belongs
to the closure of L if and only if for all > 0 and (x, y) Q
2
there is a function
f := f
x,y,
in L satisfying the conditions
f(x) f(x) < , f(y) f(y) > .
: This is obvious.
: Basing it on 3.2.10 and 3.2.11, assume that f = 0. Take > 0. Fix x Q
and consider the function g
y
:= f
x,y,
L. Let V
y
:= q Q : g
y
(q) > . Then
V
y
is an open set and y V
y
. By a standard compactness argument, there are
y
1
, . . . , y
n
Q such that Q = V
y
1
. . . V
y
n
. Put f
x
:= g
y
1
. . . g
y
n
. It is
clear that f
x
L. Furthermore, f
x
(x) < and f
x
(y) > for all y Q. Now let
U
x
:= q Q : f
x
(q) < . Then U
x
is open and x U
x
. Using the compactness
of Q once again, nd x
1
, . . . , x
m
Q such that Q = U
x
1
. . . U
x
m
. Finally, put
l := f
x
1
. . . f
x
m
. It is beyond a doubt that l L and |l| < .
10.8.6. Remark. The message of 10.8.5 is often referred to as the Generalized
Dini Theorem. (cf. 7.2.10).
10.8.7. Kakutani Lemma. Every sublattice L of C(Q, R) is expressible as
cl L =

(q
1
,q
2
)Q
2
cl
_
L
q
1
,q
2
]
_
.
The inclusion of cl L into cl
_
L
q
1
,q
2
]
_
for all (q
1
, q
2
) Q
2
raises no doubts.
If f cl
_
L
q
1
,q
2
]
_
for all such q
1
and q
2
, then by 10.8.5, f cl L.
10.8.8. Corollary. Every vector sublattice L of C(Q, R) is expressible as
cl L =

(q
1
,q
2
)Q
2
L
q
1
,q
2
]
.
Observe that every set of the form L
q
1
,q
2
]
is closed.
10.8.9. Definition. A subset U of F
Q
separates the points of Q, if for all q
1
,
q
2
Q such that q
1
,= q
2
there is a function u U assuming dierent values at q
1
and q
2
, i.e. u(q
1
) ,= u(q
2
).
10.8.10. Stone Theorem. If a vector sublattice of C(Q, R) contains constant
functions and separates the points of Q, then it is dense in C(Q, R).
Given such sublattice L, observe that
L
q
1
,q
2
]
= C(Q, R)
q
1
,q
2
]
for every pair (q
1
, q
2
) in Q
2
(cf. 10.8.4 (6)). It remains to appeal to 10.8.8.
192 Chapter 10
10.8.11. Let C(Q, R)

. Put
N ():= f C(Q, R) : [0, [f[] ker .
Then there is a unique closed subset supp() of Q such that
f N () f

supp()
= 0.
By the Interval Addition Lemma
[0, [f[] + [0, [g[] = [0, [f[ +[g[].
Thus, f, g N () [f[ + [g[ N (). Since N () is an order ideal; i.e.,
(f N () & 0 [g[ [f[ g N ()), conclude that N () is a linear set.
Moreover, N () is closed. Indeed, assuming f
n
0, f
n
f and f
n
N (), for
g [0, f] nd g f
n
g and g f
n
[0, f
n
]. Whence it follows that (g) = 0;
i.e., f N ().
Since N () is an order ideal, from 10.8.8 deduce that
N () =

qQ
N ()
q]
.
Dene the set supp() as
q supp() N ()
q]
,= C(Q, R) (f N () f(q) = 0).
It is beyond a doubt that supp() is closed. Moreover, the equalities hold:
N () =

qsupp()
N ()
q]
=
_
f C(Q, R) : f

supp()
= 0
_
.
The claim of uniqueness follows from the normality of Q (cf. 9.4.14) and the Urysohn
Theorem.
10.8.12. Definition. The set supp() under discussion in 10.8.11 is the sup-
port of (cf. 10.9.4 (5)).
10.8.13. Remark. If is positive then
N () = f C(Q, R) : ([f[) = 0.
Consequently, when (fg) = 0 for all g C(Q, R), observe that f

supp()
= 0.
By analogy supp() = N () = C(Q, R) = 0. Therefore, it is quite
convenient to work with the support of a positive functional.
Duality and Its Applications 193
Let F be a closed subset of Q. It is said that F supports or carries or that
X F lacks or is void of if ([f[) = 0 for every continuous function f with
supp(f) QF. The support supp() of carries ; moreover, supp() is included
in every closed subset of Q supporting . In other words, the support of is the
complement of the greatest open set void of (cf. 10.10.5 (6)).
It stands to reason to observe that in virtue of 3.2.14 and 3.2.15 to every
bounded functional there correspond some positive (and hence bounded) func-
tionals
+
,

, and [[ dened as

+
(f) = sup [0, f];

(f) = inf [0, f]; [[ =


+
+

,
given f C(Q, R)
+
.
Moreover, C(Q, R)

is a Kantorovich space (cf. 3.2.16).


10.8.14. The supports of and [[ coincide.
By denition N () = N ([[).
10.8.15. Considering a C(Q, R) with 0 a 1, dene a : f (af) for
f C(Q, R) and C(Q, R)

. Then [a[ = a[[.


Given f C(Q, R)
+
, infer that
(a)
+
(f) = sup(ag) : 0 g f sup [0, af]
=
+
(af) = a
+
(f).
Furthermore,

+
= (a + (1 a))
+
(a)
+
+ ((1 a))
+
a
+
+ (1 a)
+
=
+
.
Consequently, (a)
+
= a
+
, whence the claim follows.
10.8.16. De Branges Lemma. Let A be a subalgebra of C(Q, R) containing
constant functions. Take ext(A

B
C(Q, R)
). Then the restriction of each
member of A to the support of is a constant function.
If = 0 then supp() = , and there is nothing to be proven. If ,= 0
then, certainly, || = 1. Take a A. Since the subalgebra A contains constant
functions, it suces to settle the case in which 0 a 1 and
q supp() 0 < a(q) < 1.
Put
1
:= a and
2
:= (1 a). It is clear that
1
+
2
= and the functionals

1
and
2
are both nonzero. Moreover,
|| |
1
| +|
2
|
194 Chapter 10
= sup
|f|1
(af) + sup
|g|1
((1 a)g) = sup
|f|1,|g|1
(af + (1 a)g) ||,
because it is obvious that
aB
C(Q, R)
+ (1 a)B
C(Q, R)
B
C(Q, R)
.
Thus, || = |
1
| +|
2
|. Since
= |
1
|

1
|
1
|
+|
2
|

2
|
2
|
,
and
1
,
2
A

, conclude that
1
= |
1
|. By 10.8.15, a[[ = [a[ = [
1
[ =
|
1
| [[. Consequently, [[((a |
1
|1)g) = 0 for all g C(Q, R). Using 10.8.13
and 10.8.14, infer that the function a is constant on the support of .
10.8.17. StoneWeierstrass Theorem. Let A be a subalgebra of the alge-
bra C(Q, R). Suppose that A contains constant functions and separates the points
of Q. Then A is dense in C(Q, R).
Proceeding by way of contradiction, assume the contrary. By the Absolute
Bipolar Theorem, the subspace A

(coincident with A

) of C(Q, R)

is nonzero.
Using the AlaogluBourbaki Theorem, observe that A

B
C(Q,R)
is a nonempty
absolutely convex weakly compact set. Thus, by the KrenMilman Theorem, the
set has an extreme point, say, .
Undoubtedly, is a nonzero functional. By the de Branges Lemma the support
of fails to contain two distinct points, since A separates the points of Q. The
support of is not a singleton, since annihilates constant functions. Thus, supp()
is empty. But then is zero (cf. 10.8.13). We arrive at a contradiction completing
the proof.
10.8.18. Corollary. The closure of a subalgebra of C(Q, R) is a vector sub-
lattice of C(Q, R).
Using the StoneWeierstrass Theorem, nd a polynomial p
n
satisfying
[p
n
(t) [t[ [
1
/
2n
for all t [1, 1]. Then [p
n
(0)[
1
/
2n
. Therefore, the polynomial
p
n
(t):= p
n
(t) p
n
(0)
maintains the inequality [p
n
(t) [t[ [
1
/
n
when 1 t 1. By construction, p
n
lacks the constant term. Now, if a function a lies in a subalgebra A of C(Q, R)
and |a| 1, then
[p
n
(a(q)) [a(q)[ [
1
/
n
(q Q).
Moreover, the function q p
n
(a(q)) is clearly a member of A.
Duality and Its Applications 195
10.8.19. Remark. Corollary 10.8.18 (together with 10.8.8) completely de-
scribes all closed subalgebras of C(Q, R). In turn, as the proof prompts, the
claim of 10.8.18 is immediate on providing some sequence of polynomials which
converges uniformly to the function t [t[ on the interval [1, 1]. It takes no
pains to demonstrate 10.8.17, with 10.8.18 available.
10.8.20. TietzeUrysohn Theorem. Let Q
0
be a compact subset of a com-
pact set Q and f
0
C(Q
0
, R). Then there is a function f in C(Q, R) such that
f

Q
0
= f
0
.
Suppose that Q
0
,= (otherwise, there is nothing to prove). Consider the
identical embedding : Q
0
Q and the bounded linear operator

: C(Q, R)
C(Q
0
, R) acting by the rule

f := f . We have to show that

is an epimorphism.
It is beyond a doubt that im

is a subalgebra of C(Q
0
, R) separating the points
of Q
0
and containing constant functions. In virtue of 10.8.17 it is thus sucient
(and, clearly, necessary) to examine the closure of im

.
Consider the monoquotient of the operator

and the coset mapping :
C(Q, R) coim

. Given f C(Q, R), put


g := (f sup [f(Q
0
)[1) (sup [f(Q
0
)[1).
By denition f

Q
0
= g

Q
0
; i.e., f := (f) = (g). Consequently, |g| |f|.
Furthermore,
|f| = inf
_
|h|
C(Q,R)
:

(h f) = 0
_
= inf
_
|h|
C(Q,R)
: h

Q
0
= f

Q
0
_
inf
_
|h

Q
0
|
C(Q,R)
: h

Q
0
= f

Q
0
_
= sup [f(Q
0
)[ = |g| |f|.
Therefore,
|f| = |

g| = |

g|
C(Q
0
,R)
= |g |
C(Q
0
,R)
= sup [g(Q
0
)[ = |g| = |f|;
i.e., is an isometry. Successively applying 5.5.4 and 4.5.15, infer rst that coim

is
a Banach space and second that im is closed in C(Q
0
, R). It suces to observe
that im

= im.
10.9. Radon Measures
10.9.1. Definition. Let be a locally compact topological space. Put K()
:= K(, F) := f C(, F) : supp(f) is compact. If Q is compact in then
let K(Q) := K

(Q) := f K() : supp(f) Q. The space K(Q) is furnished


with the norm | |

. Given E Op (), put K(E) := K(Q) : Q E. (The


notation Q E for a subset E of means that Q is compact and Q lies in the
interior of E relative to .)
196 Chapter 10
10.9.2. The following statements are valid:
(1) if Q and f C(Q, F) then
f

Q
= 0 ( g K(Q)) g

Q
= f;
moreover, K(Q) is a Banach space;
(2) let Q, Q
1
, and Q
2
be compact sets and Q Q
1
Q
2
; the linear span
in the space C(Q, F) of the restrictions to Q of the functions like u
1
u
2
(q
1
, q
2
):=
u
1
u
2
(q
1
, q
2
):= u
1
(q
1
)u
2
(q
2
) with u
s
K(Q
s
) is dense in C(Q, F);
(3) if is compact then K() = C(, F); if fails to be compact and
is embedded naturally into C(

, F), with

:= the Alexandro com-


pactication of , then the space K() is dense in the hyperplane f C(

, F) :
f() = 0;
(4) the mapping E Op () K(E) Lat (K()) preserves supre-
ma;
(5) for E

, E

Op () the following sequence is exact:


0 K(E

(E

,E

)
K(E

) K(E

(E

,E

)
K(E

) 0,
with
(E

,E

)
f := (f, f), and
(E

,E

)
(f, g):= f +g.
(1): The boundary Q of Q is at the same time the boundary of the exterior
of Q, the set int( Q).
(2): The set under study is a subalgebra. Apply 9.3.13 and 10.8.17 (cf. 11.8.2).
(3): It may be assumed that F = R. The claim will then follow from 10.8.8
since K() is an order ideal separating the points of

(cf. 10.8.11).
(4): Clearly, K(sup ) = K() = 0. If E Op () and E is ltered upward
then, for f K(E ), observe that supp(f) E for some E E by the compactness
of supp(f). Whence K(E ) = K(E) : E E . To conclude, take E
1
, . . . , E
n

Op () and f K(E
1
. . . E
n
). In accordance with 9.4.18 there are some

k
K(E
k
) such that

n
k=1

k
= 1. Moreover, f =

n
k=1

k
f and supp(f
k
)
E
k
(k:= 1, . . . , n).
(5): This is straightforward from (4).
10.9.3. Definition. A functional , a member of K(, F)
#
, is a measure
(or, amply, a Radon F-measure) on , in symbols, M() := M(, F); if

K(Q)
K(Q)

whenever Q . The following notation is current:


_

f d:=
_
f d:=
_
f(x) d(x):= (f) (f K()).
The scalar (f) is the integral of f with respect to . In this connection is also
called an integral.
Duality and Its Applications 197
10.9.4. Examples.
(1) For q the Dirac measure f f(q) (f K()) presents
a Radon measure. It is usually denoted by the symbol
q
and called the delta-
function at q.
Suppose also that is furnished with group structure so that the taking
of an inverse q q
1
and the group operation (s, t) st
are continuous; i.e., is a locally compact group. By we denote
e
where e is the
unity of (recall the concurrent terms: identity, unit, and neutral element,
all meaning the same: es = se = s for all s ). The nomenclature pertinent
to addition is routinely involved in abelian (commutative) groups.
Given a , acknowledge the operation of (left or right) translation or shift
by a in K() (in fact, every function is shifted in F):
(
a
f)(q):=
a
f(q):= f(a
1
q), (
a
f)(q):= f
a
(q):= f(qa
1
)
with f K() and q . Clearly,
a
,
a
L(K()).
A propitious circumstance of paramount importance is the presence of a non-
trivial measure on , a member of M(, R), which is invariant under left (or right)
translations. All (left)invariant Radon measures are proportional. Each nonzero
(left)invariant positive Radon measure is a (left) Haar measure (rarely Haar in-
tegral). In the case of right translations, the term (right) Haar measure is in
common parlance. In the abelian case, we speak only of a Haar measure and even
of Haar measure, neglecting the necessity of scaling. The familiar Lebesgue measure
on R
N
is Haar measure on the abelian group R
N
. That is why the conventional no-
tation of Lebesgue integration is retained in the case of an abstract Haar measure.
In particular, the left-invariance condition is written down as
_

f(a
1
x) dx =
_

f(x) dx (f K(), a ).
(2) Let M() := (K(), | |

. A member of M() is a bounded


Radon measure. It is clear that a bounded measure belongs to the space C(

, F)

(cf. 10.9.2 (2)).


(3) Given M(), put

(f) = (f

, where f

(q) := f(q)

for
q and f K(). The measure

is the conjugate of . Distinction be-


tween

and is perceptible only if F = C. In case =

, we speak of a real
C-measure. It is clear that =
1
+i
2
, where
1
and
2
are uniquely-determined
real C-measures. In turn, each real C-measure is generates by two R-measures (al-
so called real measures), members of M(, R), because K(, C) coincides with
the complexication K(, R) iK(, R) of K(, R). The real R-measures obvi-
ously constitute a Kantorovich space. Moreover, the integral with respect to such
198 Chapter 10
measure serves as (pre)integral. So, an opportunity is oered to deal with the corre-
sponding Lebesgue extension and spaces of scalar-valued or vector-valued functions
(cf. 5.5.9 (4) and 5.5.9 (5)). We seize the opportunity without much ado and spec-
ication.
To every Radon measure we assign the positive measure [[ that is dened
for f K(, R), f 0, by the rule
[[(f):= sup[(g)[ : g K(, F), [g[ f.
Observe that, in current usage, the word measure often means a positive measure,
whereas a general measure is referred to as a signed measure or a charge.
If and are measures and [[ [[ = 0 then and are called disjoint or
independent (of one another). A measure is absolutely continuous with respect
to on condition that is independent of every measure independent of . Such
a measure may be given as = f, where f L
1,loc
() and the measure f
having density f with respect to acts by the rule (f)(g) := (fg) (g K())
(this is the RadonNikodym Theorem).
(4) Given

Op () and M(), consider the restriction

:=

K(

)
of to K(

). The restriction operator

from M() to M(

),
viewed as depending on a subset of , meets the agreement condition: if

and M() then

= (

. This situation is verbalized as follows:


The mapping M : E Op () M(E) and the restriction operator (referred
to jointly as the functor M) denes a presheaf (of vector spaces over ). It stands
to reason to convince oneself that the (values of the) restriction operator need not
be an epimorphism.
(5) Let E Op () and M(). It is said that E lacks or is
void of or that E supports or carries if
E
= 0. By 10.9.2 (4) there is
a least closed set supp() supporting , the support of . It may be shown that
supp() = supp([[). The above denition agrees with 10.8.12. The Dirac measure

q
is a unique Radon measure supported at q to within scaling.
(6) Let
k
be a locally compact space and
k
M(
k
) (k := 1, 2).
There is a unique measure on the product
1

2
such that
_

2
u
1
(x)u
2
(y) d(x, y) =
_

1
u
1
(x) d
1
(x)
_

2
u
2
(y) d
2
(y)
with u
k
K(
k
). The next designations are popular:
1

2
:=
1

2
:= .
Using 10.9.2 (4), infer that for f K(
1

2
) the value
1

2
(f) may be
calculated by repeated integration (this is the Fubini Theorem).
Duality and Its Applications 199
(7) Let G be a locally compact group, and , M(G). For f K(G)
the function

f(s, t) := f(st) is continuous and [( )(

f)[ || || |f|

. This
denes the Radon measure (f):= ( )(

f) (f K(G)), the convolution of
and . Using vector integrals, obtain the presentations:
=
_
GG

s

t
d(s)d(t) =
_
G

s
d(s) =
_
G

t
d(t).
The space M(G) of bounded measures with convolution as multiplication, ex-
emplies a Banach convolution algebra.
This algebra is commutative if and only if G is an abelian group. In that event
the space L
1
(G), constructed with respect to Haar measure m, also possesses a nat-
ural structure of a convolution algebra (namely, that of a subalgebra of M(G)). The
algebra (L
1
(G), ) is the group algebra of G. Thus, for f, g L
1
(G), the deni-
tions of convolution for functions and measures agree with one another (cf. 9.6.17):
(f g)dm = fdm gdm. By analogy, the convolution of M(G) and f L
1
(G)
is dened as ( f)dm:= (fdm); i.e., as the density of the convolution of
and fdm with respect to Haar measure dm. In particular,
f g =
_
G
(
x
g)f(x) dm(x) =
_
G

x
(g)f(x) dm(x).
Wendel Theorem. Let T B(L
1
(G)). Then the following statements are
equivalent:
(i) there is a measure M(G) such that Tf = f for f L
1
(G);
(ii) T commutes with translations: T
a
=
a
T for a G, where
a
is
a unique bounded extension to L
1
(G) of translation by the element
a in K(G);
(iii) T(f g) = (Tf) g for f, g L
1
(G);
(iv) T(f ) = (Tf) for M(G) and f L
1
(G).
10.9.5. Definition. The spaces K() and M() are set in duality (induced
by the duality bracket K() K()
#
). In this case, the space M() is fur-
nished with the topology (M(), K()), usually called vague. The space K()
is furnished with the Mackey topology
K()
:= (K(), M()) (thereby, in par-
ticular, (K(),
K()
)

= M()). The space of bounded measures M() is usually


considered with the dual norm:
||:= sup[(f)[ : |f|

1, f K() ( M()).
10.9.6. The topology
K()
is strongest among all locally convex topologies
making the identical embedding of K(Q) to K() continuous for every Q with
Q (i.e.,
K()
is the so-called inductive limit topology (cf. 9.2.15)).
200 Chapter 10
If is the inductive limit topology and (K(), )

then by denition
M(), because
K(Q)
is continuous for Q . In turn, for M()
the set V
Q
:= f K(Q) : [(f)[ 1 is a neighborhood of zero in K(Q). From
the denition of , infer that V
Q
: Q = f K() : [(f)[ 1 is
a neighborhood of zero in . Thus, (K(), )

and is compatible with


duality. Therefore,
K()
.
On the other hand, if p is a seminorm in the mirror of the Mackey topology
then p is the supporting function of a subdierential lying in M(). Consequently,
the restriction q := p
K(Q)
of p to K(Q) is lower semicontinuous anyway. By the
Gelfand Theorem (in view of the barreledness of K(Q)) the seminorm q is contin-
uous. Consequently, the identical embedding
K(Q)
: K(Q) (K(),
K()
) is
continuous and
K()
by the denition of inductive limit topology.
10.9.7. A subset A of K(R
N
) is bounded (in the inductive limit topology),
if sup |A|

< + and, moreover, the supports of the members of A lie in a com-


mon compact set.
Suppose to the contrary that, for some Q such that Q R
N
, the inclusion
A K(Q) fails. In other words, for n N there are some q
n
R
N
and a
n
A
satisfying a
n
(q
n
) ,= 0 and [q
n
[ > n. Put B:= n[a
n
(q
n
)[
1

q
n
: n N and observe
that this set of Radon measures is vaguely bounded and so the seminorm p(f) :=
sup[[([f[) : B is continuous. Moreover, p(a
n
) n[a
n
(q
n
)[
1

q
n
([a
n
[) = n,
which contradicts the boundedness property of A.
10.9.8. Remark. Let (f
n
) K(R
N
). The notation f
n

K
0 symbolizes the
proposition ( Q R
N
)( n) supp(f
n
) Q & |f
n
|

0. From 10.9.7 it is
immediate that K(R
N
)
#
is a Radon measure if (f
n
) 0 whenever f
n

K
0.
Observe also that the same holds for every locally compact countable at innity
or -compact, i.e. for presenting the union of countably many compact sets.
10.9.9. Remark. There is a sequence (p
n
) of real-valued positive polynomials
on R such that the measures p
n
dx converge vaguely to as n +. Considering
products of the measures, arrive at some polynomials P
n
on the R
N
, with (P
n
dx)
converging vaguely to (here, as usual, dx := dx
1
. . . dx
N
is the Lebesgue
measure on R
N
). Now, take f K(R
N
) such that f is of class C
(m)
in some
neighborhood of a compact set Q (i.e., f has continuous derivatives up to order m
on Q). Arranging the convolutions (f P
n
), obtain a sequence of polynomials which
approximates not only f but also the derivatives of f up to order m uniformly on Q.
The possibility of such reqularization is referred to as the Generalized Weierstrass
Theorem in R
N
(cf. 10.10.2 (4)).
10.9.10. Measure Localization Pronciple. Let E be an open cover of
and let (
E
)
EE
be a family of Radon measures,
E
M(E). Assume further
Duality and Its Applications 201
that, for every pair (E

, E

) of the members of E , the restrictions of


E
and
E

to E

coincide. Then there is a unique measure on whose restriction to E


equals
E
for all E E .
Using 10.9.2 (5), arrange the sequence

,E

]
E

,E

E, E

,=E

K(E

EE
K(E)

K() 0,
where is generated by summation of the coordinate identical embeddings
(E

,E

)
and is the standard addition. Every direct sum is always assumed to be furnished
with the inductive limit topology (cf. 10.9.6).
Examine the exactness of the sequence. Since K() =
Q
K(Q) and in view
of 10.9.2 (4), it suces to settle the case of a nite cover by checking exactness at the
second term. Thus, proceeding by induction, suppose that for every n-element cover
E
1
, . . . , E
n
(n 2) the following sequence is exact:
K
n

k=1
K(E
k
)

n
K(E
1
. . . E
n
) 0,
where
n
is the restriction of to K
n
, the symbol
n
stands for addition and
K
n
:=

k<l
k,l1,... ,n]
K(E
k
E
l
).
By hypothesis, im
n
= ker
n
. If
n+1
(

f, f
n+1
) = 0 where

f := (f
1
, . . . , f
n
), then

f = f
n+1
and f
n+1
K((E
1
. . . E
n
) E
n+1
). Since
n
is an epimorphism
by 10.9.2 (5), there are some
k
K(E
k
E
n+1
) such that
n
= f
n+1
for
:= (
1
, . . . ,
n
). Whence (

f ) ker
n
, and by hypothesis there is a member
of K
n
such that
n
=

f . Clearly,
K
n+1
= K
n

k=1
K(E
k
E
n+1
)
(to within isomorphism), := (,
1
, . . . ,
n
) K
n+1
, and
n+1
= (

f, f
n+1
).
Passing to the diagram prime (cf. 7.6.13), arrive at the exact sequence
0 M()

EE
M(E)

,E

]
E

,E

E,E

,=E

M(E

).
The proof is complete.
202 Chapter 10
10.9.11. Remark. In topology, a presheaf recoverable from its patches or local
data in the above manner is a sheaf. In this regard, the claim of 10.9.10 is verbalized
as follows: the presheaf M() of Radon measures is a sheaf or, putting it more
categorically, the functor M is a sheaf (of vector spaces over , cf. 10.9.4 (4)).
10.10. The Spaces D() and D

()
10.10.1. Definition. A compactly-supported smooth function f : R
N
F is
a test function; in symbols, f D(R
N
) := D(R
N
, F). Given Q R
N
and
Op (R
N
), designate D(Q) :=
_
f D(R
N
) : supp(f) Q
_
and D() := D(Q) :
Q .
10.10.2. The following statements are valid:
(1) D(Q) = 0 int Q = ;
(2) given Q R
N
, put
|f|
n,Q
:=

]]n
|

f|
C(Q)
:=

(Z
+
)
N

1
+...+
N
n
sup [(

1
. . .

n
f)(Q)[
for a function f smooth in a neighborhood of Q (as usual, Z
+
:= N 0); the
multinorm M
Q
:= | |
n,Q
: n N makes D(Q) into a Frechet space;
(3) the space of smooth functions C

() := E () on in Op (R
N
)
with the multinorm M

:= | |
n,Q
: n N, Q is a Frechet space; moreover,
D() is dense in C

();
(4) if Q
1
R
N
, Q
2
R
M
and Q Q
1
Q
2
; then the linear span
in D(Q) of the restrictions to Q of the functions like f
1
f
2
(q
1
, q
2
):= f
1
f
2
(q
1
, q
2
):=
f
1
(q
1
)f
2
(q
2
), with q
k
Q
k
and f
k
D(Q
k
), is dense in D(Q);
(5) the mapping E Op () D(E) Lat (D()) preserves suprema:
D(E

) = D(E

) D(E

), D(E

) = D(E

) +D(E

);
D(E ) = L(D(E) : E E ) (E Op ()).
Moreover, the next sequence is exact (cf. 10.9.2 (5)):
0 D(E

(E

,E

)
D(E

) D(E

(E

,E

)
D(E

) 0.
(1) and (2) are obvious.
(3): Choose a sequence (Q
m
)
mN
such that Q
m
, Q
m
Q
m+1
and

mN
Q
m
= . The multinorm | |
n,Q
m
: n N, m N is countable and
Duality and Its Applications 203
equivalent to M

. A reference to 5.4.2 yields the metrizability of C

(). The
completeness of C

() raises no doubts.
To show that D() is dense in C

(), consider the truncator set Tr():=


D() : 0 1. Make Tr() a direction on letting
1

2
supp(
1
)
int
2
= 1. It is clear that, for f C

(), the net (f)


Tr()
approximates f
as is needed.
(4): Take q

R
N
and q

R
M
. Let a(q

, q

) := a

(q

)a

(q

), where a

and a

are molliers on R
N
and R
M
, respectively. Given f D(Q), n N, and
> 0, choose from the condition |f f a

|
n,Q

/
2
. Using the equicontinuity
property of the family F:=

f(q)
q
(a

) : [[ n, q Q
1
Q
2
, nd nite sets

and

, with

Q
1
and

Q
2
, so that the integral of each function in F
be approximated to within
1
/
2
(N +1)
n
by a Riemann sum of it using the points
of

. This yields a function f, a member of the linear span of D(Q


1
)D(Q
2
),
which was sought; i.e., |f f|
n,Q
.
(5): Check this as in 10.9.2 (4), on replacing 9.4.18 with 9.6.19 (2).
10.10.3. Remark. The Generalized Weierstrass Theorem may be applied
to the demonstration of 10.10.2 (4), when combined with due truncation providing
that the constructed approximation has compact support.
10.10.4. Definition. A functional u, a member of D(, F)
#
, is a distribution
or a generalized function whenever u[
D(Q)
D

(Q):= D(Q)

for all Q . This is


expressed in writing as u D

():= D

(, F). Sometimes a reference is appended


to the nature of the ground eld F.
The usual designations are as follows: u, f := f [ u := u(f). Often we use
the most suggestive and ubiquitous symbol
_
f(x)u(x) dx:= u(f) (f D()).
10.10.5. Examples.
(1) Let g L
1,loc
(R
N
) be a locally integrable function. The mapping
u
g
(f):=
_
f(x)g(x) dx (f D())
determines a distribution. A distribution of this type is regular. To denote a regular
distribution u
g
, a more convenient symbol g is also employed. In this connection,
we write D() D

() and u
g
= [ g.
(2) Every Radon measure is a distribution. Each positive distribution u
(i.e., such that f 0 u(f) 0) is determined by a positive measure.
204 Chapter 10
(3) A distribution u is said to has order at most m, if to every Q such
that Q R
N
there corresponds a number t
Q
satisfying
[u(f)[ t
Q
|f|
m,Q
(f D(Q)).
The notions of the order of a distribution and of a distribution of nite order are
understood in a matter-of-fact fashion. Evidently, it is false that every distribution
has nite order.
(4) Let be a multi-index, (Z
+
)
N
; and let u be a distribution,
u D

(). Given f D(), put (

u)(f) := (1)
]]
u(

f). The distribution

u is the derivative of u (of order ). We also speak of generalized dierentiation,


of derivatives in the distribution sense, etc. and use the conventional symbols
of dierential calculus.
A derivative (of nonzero order) of a Dirac measure is not a measure. At the
same time D

(R) is the derivative of the Heaviside function


(1)
:= H, where
H : R R is the characteristic function of R
+
. If a derivative of a (regular)
distribution u is some regular distribution u
g
, then g is a weak derivative of u
or a generalized derivative of u in the Sobolev sense. For a test function, such
derivative coincides with its ordinary counterpart.
(5) Given u D

(), put u

(f) := u(f

. The distribution u

is the
conjugate of u. The presence of the involution routinely justies speaking of real
distributions and complex distributions (cf. 10.9.3 (3)).
(6) Let E Op () and u D

(). For f D(E), the scalar u(f)


is easily determined. This gives rise to the distribution u
E
, a member of D

(E),
called the restriction of u to E. The functor D

is clearly a presheaf.
Given u D

() and E Op (), say that E lacks or is void of u, if u


E
= 0.
By 10.10.4 (5), if the members of a family of open subsets of are void of u then
so is the union of the family. The complement (to R
N
) of the greatest open set
void of u is the support of u, denoted by supp(u). Observe that supp(

u)
supp(u). Moreover, a distribution with compact support (= compactly-supported
distribution) has nite order.
(7) Let u D

() and f C

(). If g D() then fg D(). Put


(fu)(g):= u(fg). The resulting distribution fu is the product of f and u. Consider
the truncator direction Tr(). If there is a limit lim
Tr()
u(f) then say that u
applies to f. It is clear that a compactly-supported distribution u applies to every
function in C

(). Moreover, u E

() := C

()

. In turn, each element u


of E

() (cf. 10.10.2 (3)) uniquely determines a distribution with compact support,


which is implicit in the notation u D

().
If f C

() and

supp(u)
= 0 for all , [[ m, where u is a compactly-
supported distribution of order at most m, then it is easy that u(f) = 0. In con-
Duality and Its Applications 205
sequence, only linear combinations of a Dirac measure and its derivatives are sup-
ported at a singleton.
(8) Let
1
,
2
Op (R
N
) and u
k
D

(
k
). There is a unique distri-
bution u on
1

2
such that u(f
1
f
2
) = u
1
(f
1
)u
2
(f
2
) for all f
k
D(
k
). The
distribution is denoted by u
1
u
2
or u
1
u
2
. Using 10.10.2 (4), infer that for
f D(
1

2
) the value u(f) of u at f appears from successive application of u
1
and u
2
. Strictly speaking,
u(f) = u
2
(y
2
u
1
(f( , y))) = u
1
(x
1
u
2
(f(x, ))).
More suggestive designations prompt the Fubini Theorem:
__

2
f(x, y)(u
1
u
2
)(x, y) dxdy
=
_

2
_
_
_

1
f(x, y)u
1
(x) dx
_
_
u
2
(y) dy =
_

1
_
_
_

2
f(x, y)u
2
(y) dy
_
_
u
1
(x) dx.
It is worth noting that
supp(u
1
u
2
) = supp(u
1
) supp(u
2
).
(9) Let u, v D

(R
N
). Given f D(R
N
), put
+
f := f +. It is
clear that
+
f C

(R
N
R
N
). Say that the distributions u and v convolute or
admit convolution or are convolutive provide that the product u v applies to
the function
+
f C

(R
N
R
N
) for each f in D(R
N
). It is easy (cf. 10.10.10)
that the resulting linear functional f (u v)(
+
f ) (f D(R
N
)) is a distribu-
tion called the convolution of u and v and denoted by u v. It is beyond a doubt
that the convolutions of functions (cf. 9.6.17) and measures on R
N
(cf. 10.9.4 (7))
are particular cases of the convolution of distributions. In some classes, every two
distributions convolute. For instance, the space E

(R
N
) of compactly-supported
distributions with convolution as multiplication presents an (associative and com-
mutative) algebra with unity the delta-function . Further,

u =

u and

(u v) =

u v = u

v. Moreover, the following remarkable equality, the


Lions Theorem of Supports, holds:
co (supp(u v)) = co (supp(u)) + co (supp(v)).
206 Chapter 10
It is worth emphasizing that the pairwise convolutivity of distributions fails in gen-
eral to guarantee the associativity of convolution (for instance, (1

)
(1)
= 0
whereas 1 (


(1)
) = 1, with 1:= 1
R
).
Each distribution u convolutes with a test function f, yielding some regular
distribution (u f)(x) = u(
x
(f )), where f :=

f is the reection of f; i.e.,
f (x) := f(x) (x R
N
). The operator u : f u f from D(R
N
) to C

(R
N
)
is continuous and commutes with translations: (u)
x
=
x
u for x R
N
. It is
easily seen that the above properties are characteristic of u; i.e., if an operator T,
a member of L(D(R
N
), C

(R
N
)), is continuous and commutes with translations,
then there is a unique distribution u such that T = u; namely, u(f) := (T

)(

f)
for f D(R
N
) (cf. the Wendel Theorem).
10.10.6. Definition. The spaces D() and D

() are set in duality (in-


duced by the duality bracket D() D()
#
). Moreover, D

() is furnished
with the topology of the distribution space (D

(), D()), and D() is furnished


with the topology of the test function space, the Mackey topology
D
:=
D()
:=
(D(), D

()).
10.10.7. Let Op (R
N
). Then
(1)
D
is the strongest of the locally convex topologies making the
identical embedding of D(Q) into D() continuous for all Q (i.e.,
D
is the
inductive limit topology);
(2) a subset A of D() is bounded if and only if A lies in D(Q) for
some Q such that Q and is bounded in D(Q);
(3) a sequence (f
n
) converges to f in (D(),
D
) if and only if there
is a compact set Q such that Q , supp(f
n
) Q, supp(f) Q and (

f
n
)
converges to

f uniformly on Q for every multi-index (in symbols, f


n
f);
(4) an operator T, a member of L(D(), Y ) with Y a locally convex
space, is continuous if and only if Tf
n
0 provided that f
n
0;
(5) a delta-like sequence (b
n
) serves as a (convolution) approximate
unity in D(R
N
) as well as in D

(R
N
); i.e., b
n
f f (in D(R
N
)) and b
n
u u
(in D

(R
N
)) for f D(R
N
) and u D

(R
N
).
(1): This is established in much the same way as 10.9.6; and (2), by analogy
with 10.9.7 using the presentation of as the union =
nN
Q
n
, where Q
n
Q
n+1
for n N.
(3): Note that a convergent countable sequence is bounded, and apply 10.10.7
(2) (cf. 10.9.8).
(4): In virtue of 10.10.7 (1) the continuity of T amounts to that of the restriction
T

D(Q)
for all Q . By 10.10.2 (2) the space D(Q) is metrizable. It remains
to refer to 10.10.7 (3).
Duality and Its Applications 207
(5): It is clear that all of the supports supp(b
n
f) lie in some compact neigh-
borhood about supp(f). Furthermore, for g C(R
N
), it is evident that b
n
g g
uniformly on compact subsets of R
N
. Applying this to

f and considering (3),


infer that b
n
f f.
On account of 10.10.5 (9), observe the following:
u(

f) = (u f)(0) = lim
n
(u (b
n
f))(0)
= lim
n
((u b
n
) f)(0) = lim
n
(b
n
u)(

f)
for f D(R
N
).
10.10.8. Remark. In view of 10.10.7 (3), for Op (R
N
) and m Z
+
it is of-
ten convenient to consider the space D
(m)
():= C
(m)
0
() comprising all compactly-
supported functions f whose derivatives

f are continuous for all [[ m. The


space D
(m)
(Q) := f D
(m)
() : supp(f) Q for Q is furnished with the
norm||
m,Q
making it into a Banach space. In that event, D
(m)
() is endowed with
the inductive limit topology. Thus, D
(0)
() = K() and D() =
mN
D
(m)
().
For a sequence (f
n
) to converge in D
(m)
() means to converge uniformly with all
derivatives up to order m on some Q such that Q and supp(f
n
) Q for all suf-
ciently large n. Note that D
(m)
()

comprises all distributions of order at most m.


Correspondingly,
D

F
():=
_
mN
D
(m)
()

is the space of nite-order distributions.


10.10.9. Let Op (R
N
). Then
(1) the space D() is barreled; i.e., each barrel, a closed absorbing
absolutely convex subset, is a neighborhood of zero;
(2) every bounded closed set in D() is compact; i.e., D() is a Montel
space;
(3) every absolutely convex subset of D(), absorbing each bounded
set, is a neighborhood of zero; i.e., D() is a bornological space;
(4) the test functions are dense in the distribution space.
(1): Given a barrel V in D(), observe that V
Q
:= V D(Q) is a barrel
in D(Q) for all Q . Thus, V
Q
is a neighborhood of zero in D(Q) (cf. 7.1.8).
(2): Such a set lies in D(Q) for some Q by 10.10.7 (2). In virtue of 10.10.2
(2), D(Q) is metrizable. Applying 4.6.10 and 4.6.11 proves the claim.
(3): This follows from the barreledness of D(Q) for Q .
(4): Let g [ D()

, with the polar taken with respect to D() D

().
It is clear that u
f
(g) = 0 for f D(); i.e.,
_
g(x)f(x) dx = 0. Thus, g = 0.
It remains to refer to 10.5.9.
208 Chapter 10
10.10.10. Schwartz Theorem. Let (u
k
)
kN
be a sequence of distributions.
Assume that for every f in D() there is a sum
u(f):=

k=1
u
k
(f).
Then u is a distribution and

u =

k=1

u
k
for every multi-index .
The continuity of u is guaranteed by 10.10.9 (1). Furthermore, for f D()
by denition (cf. 10.10.5 (4))

u(f) = u
_
(1)
]]

f
_
=

k=1
u
k
_
(1)
]]

f
_
=

k=1

u
k
(f).
10.10.11. Theorem. The functor D

is a sheaf.
It is immediate (cf. 10.9.10 and 10.9.11).
10.10.12. Remark. The possibility of recovering a distribution from local
data, the Distribution Localization Principle stated in 10.10.11, admits clarication
in view of the paracompactness of R
N
. Namely, consider an open cover E of and
a distribution u D

() with local data (u


E
)
EE
. Take a countable (locally nite)
partition of unity (
k
)
kN
subordinate to E . Evidently, u =

k=1

k
u
k
, where
u
k
:= u
E
k
and supp(
k
u
k
) E
k
(k N).
10.10.13. Theorem. Each distribution u on of order at most m may be
expressed as sum of derivatives of Radon measures:
u =

]]m

,
where

M().
To begin with, assume that u has compact support. Let Q with Q be
a compact neighborhood of supp(u). By hypothesis (cf. 10.10.5 (7) and 10.10.8)
[u(f)[ t

]]m
|

f|

(f D(Q))
for some t 0.
Duality and Its Applications 209
Using 3.5.7 and 3.5.3, from 10.9.4 (2) obtain
u = t

]]m

= t

]]m
(1)
]]

for a suitable family (

)
]]m
, where

[[(| |

).
Passing to the general case and invoking the Countable Partition Theorem, nd
a partition of unity (
k
)
kN
with
k
D() such that some neighborhoods Q
k
of supp(
k
) compose a locally nite cover of (cf. 10.10.12). For each of the
distributions (
k
u)
kN
it is already proven that

k
u =

]]m

k,
,
where every
k,
is a Radon measure on and supp(
k,
) Q
k
.
From the Schwartz Theorem it is immediate that the sum

(f) :=

k=1

k,
(f)
exists for all f K(). Moreover, the resulting distribution

is a Radon measure.
Once again appealing to 10.10.10, infer that
u =

k=1

k
u =

k=1

]]m

k,
=

]]m

k=1

k,
_
=

]]m

,
which was required.
10.10.14. Remark. The claim of 10.10.13 is often referred to as the theorem
on the general form of a distribution. Further abstraction and clarication are avail-
able. For instance, it may be veried that a compactly-supported Radon measure
serves as a derivative (in the distribution sense) of suitable order of some contin-
uous function. This enables us to view each distribution as a result of generalized
dierentiation of a conventional function.
10.11. The Fourier Transform of a Distribution
10.11.1. Let be a nonzero functional over the space L
1
(R
N
):= L
1
(R
N
, C).
The following statements are equivalent:
(1) is a character of the group algebra (L
1
(R
N
), ); i.e., ,= 0,
L
1
(R
N
)

and
(f g) = (f)(g) (f, g L
1
(R
N
))
(in symbols, X(L
1
(R
N
)), cf. 11.6.4);
210 Chapter 10
(2) there is a unique vector t in R
N
such that
(f) =

f(t):= (f e
t
)(0):=
_
R
N
f(x)e
i(x,t)
dx
for all f L
1
(R
N
).
(1) (2): Suppose that (f)(g) ,= 0. If x R
N
then
(
x
f g) = (
x
f)(g) = (
x
g)(f).
Put (x) := (f)
1
(
x
f). This soundly denes some continuous mapping :
R
N
C. Moreover,
(x +y)
= (f g)
1
(
x+y
(f g)) = (f)
1
(g)
1
(
x
f
y
g)
= (f)
1
(
x
f)(g)
1
(
y
g) = (x)(y)
for x, y R
N
; i.e., is a (unitary) group character: X(R
N
). Calculus shows
that = e
t
for some (obviously, unique) t R
N
. Further, by the properties of the
Bochner integral
(f)(g) = (f g) =
_
_
_
R
N
(
x
g)f(x) dx
_
_
=
_
R
N
(
x
g)f(x) dx =
_
R
N
f(x)(g)(x) dx = (g)
_
R
N
f(x)(x) dx.
Thus,
(f) =
_
R
N
f(x)(x) dx (f L
1
(R
N
)).
(2) (1): Given t R
N
and treating f, g and f g as distributions, infer that

f g(t) = u
fg
(e
t
)
=
_
R
N
_
R
N
f(x)g(y)e
t
(x +y) dxdy =
_
R
N
f(x)e
t
(x) dx
_
R
N
g(y)e
t
(y) dy
Duality and Its Applications 211
= u
f
(e
t
)u
g
(e
t
) =

f(t) g(t).
10.11.2. Remark. The essential steps of the above argument remain valid
for every locally compact abelian group G, and so there is a one-to-one correspon-
dence between the character space X(L
1
(G)) of the group algebra and the set X(G)
comprising all (unitary) group characters of G. Recall that such character is a con-
tinuous mapping : G C satisfying
[(x)[ = 1, (x +y) = (x)(y) (x, y G).
Endowed with pointwise multiplication, the set

G:= X(G) becomes a commutative
group. In virtue of the AlaogluBourbaki Theorem, X(L
1
(G)) is locally compact
in the weak topology ((L
1
(G))

, L
1
(G)). So,

Gmay be treated as a locally compact
abelian group called the character group of G or the dual group of G. Each element
q in G denes the character

q : q

G q(q) C of the dual group

G. The resulting
embedding of G into

G is surprisingly an isomorphism of the locally compact abelian


groups G and

G (the Pontryaginvan Kampen Duality Theorem).


10.11.3. Definition. For a function f in L
1
(R
N
), the mapping

f : R
N
C,
dened by the rule

f(t):= f(t):= (f e
t
)(0),
is the Fourier transform of f.
10.11.4. Remark. By way of taking convenient liberties, we customarily use
the term Fourier transform expansively. First, it is retained not only for the
operator F : L
1
(R
N
) C
R
N
acting by the rule Ff :=

f but also for its modi-
cations (cf. 10.11.13). Second, the Fourier transform F is identied with each
operator F

f :=

f , where is an automorphism (= isomorphism with it-
self) of R
N
. Among the most popular are the functions: (x) :=

(x) := x,
(x) :=
2
(x) := 2x and (x) :=
2
(x) := 2x (x R
N
). In other words, the
Fourier transform is usually dened by one the following formulas:
F

f(t) =
_
R
N
f(x)e
i(x,t)
dx,
F
2
f(t) =
_
R
N
f(x)e
2i(x,t)
dx,
F
2
f(t) =
_
R
N
f(x)e
2i(x,t)
dx.
212 Chapter 10
Since the character groups of isomorphic groups are also isomorphic, there are
grounds to using the same notation

f for generally distinct functions Ff, F

f,
and F
2
f. The choice of the symbol for F
2
(F
2
) dictates the denotation

for F
2
(F
2
) (cf. 10.11.12).
10.11.5. Examples.
(1) Let f : R R be the characteristic function of the interval [1, 1].
Clearly,

f(t) = 2t
1
sin t. Observe that if k t
0
> 0 then
_
[t
0
,+)
[

f(t)[ dt
_
[k,+)
[

f(t)[ dt =

n=k
_
[n,(n+1)]
[

f(t)[ dt

n=k
_
[n,(n+1)]
2[ sint[
(n + 1)
dt = 4

n=k
1
(n + 1)
= +.
Thus,

f / L
1
(R).
(2) For f L
1
(R
N
) the function

f is continuous, with the inequality
|

f|

|f|
1
holding.
The continuity of

f follows from the Lebesgue Dominated Convergence The-
orem; and the boundedness of

f, from the obvious estimate
[

f(t)[
_
R
N
[f(x)[ dx = |f|
1
(t R
N
).
(3) If f L
1
(R
N
), then [

f(t)[ 0 as [t[ + (= the Riemann


Lebesgue Lemma).
The claim is obvious for compactly-supported step functions. It suces
to refer to 5.5.9 (6) and the containment F B(L
1
(R
N
), l

(R
N
)).
(4) Let f L
1
(R
N
), > 0, and f

(x) := f(x) (x R
N
). Then

(t) =
N

f (
t
/

) (t R
N
).

(t) =
_
R
N
f(x)e
t
(x) dx =
N
_
R
N
f(x)e
t/
(x) dx =
N

f (
t
/

)
(5) F(f

) = (F

f)

x
f = e
x

f, and

e
x
f =
x

f for all f L
1
(R
N
)
and x R
N
.
Duality and Its Applications 213
We will only check the rst equality. Since a

b = (ab

for a, b C;
on using the properties of conjugation and integration, given t R
N
, infer that
F(f

)(t) =
_
R
N
f(x)e
i(x,t)
dx =
_
_
_
R
N
f(x)
_
e
i(x,t)
_

dx
_
_

=
_
_
_
R
N
f(x)e
i(x,t)
dx
_
_

= (F

f)

(t).
(6) If f, g L
1
(R
N
) then

f g =

f g;
_
R
N

fg =
_
R
N
fg.
The rst equality is straightforward from 10.11.1. The second, the multipli-
cation formula, follows on applying the Fubini Theorem:
_
R
N

fg =
_
R
N
_
R
N
f(x)e
t
(x) dxg(t) dt
=
_
R
N
_
_
_
R
N
g(t)e
t
(x) dt
_
_
f(x) dx =
_
R
N
fg.
(7) If

f, f and g belong to L
1
(R
N
) then (

fg)

= f g.
Given x R
N
, observe that
(

fg)

(x) =
_
R
N
g(t)

f(t)e
t
(x) dt =
_
R
N
_
R
N
g(t)f(y)e
t
(y)e
t
(x) dydt
=
_
R
N
_
R
N
f(y)g(t)e
t
(x +y) dtdy
=
_
R
N
f(y) g(x +y) dy =
_
R
N
f(y x) g(y) dy = f g(x).
(8) If f D(R
N
) and (Z
+
)
N
then
F(

f) = i
]]
t

Ff,

(Ff) = i
]]
F(x

f);
F
2
(

f) = (2i)
]]
t

F
2
f,

(F
2
f) = (2i)
]]
F
2
(x

f)
214 Chapter 10
(these equalities take the rather common liberty of designating x

:= t

:= ()

:
y R
N
y

1
1
. . . y

N
N
).
It suces (cf. 10.11.4) to examine only the rst row. Since

e
t
= i
]]
t

e
t
;
therefore,
F(

f)(t) =
_
e
t

f
_
(0)
=
_

e
t
f
_
(0) = i
]]
t

(e
t
f)(0) = i
]]
t

f(t).
By analogy, on dierentiating under the integral sign, infer that

t
1
(Ff)(t) =

t
1
_
R
N
f(x)e
i(x,t)
dx =
_
R
N
f(x)ix
1
e
i(x,t)
dx = F(ix
1
f)(t).
(9) If f
N
(x):= exp
_

1
/
2
[x[
2
_
for x R
N
, then

f
N
= (2)
N
/
2
f
N
.
It is clear that

f
N
(t) =
N

k=1
_
R
e
it
k
x
k
e

1
/
2
]x]
2
k
dx
k
(t R
N
).
Consequently, the matter reduces to the case N = 1. Now, given y R, observe
that

f
1
(y) =
_
R
e

1
/
2
x
2
e
ixy
dx =
_
R
e

1
/
2
(xiy)
2

1
/
2
y
2
dx
= f
1
(y)
_
R
e

1
/
2
(xiy)
2
dx.
To calculate the integral A that we are interested in, consider (concurrently ori-
ented) straight lines
1
and
2
parallel to the real axis R in the complex plane
C
R
R
2
. Applying the Cauchy Integral Theorem to the holomorphic function
f(z) := exp
_

z
2
/
2
_
(z C) and a rectangular with vertices on
1
and
2
and
properly passing to the limit, conclude that
_

1
f(z) dz =
_

2
f(z) dz. Whence it fol-
lows that
A =
_
R
e

1
/
2
(xiy)
2
dx =
_
R
e

1
/
2
x
2
dx =

2.
Duality and Its Applications 215
10.11.6. Definition. The Schwartz space is the set of tempered (or rapidly
decreasing) functions, cf. 10.11.17 (2),
S(R
N
):=
_
f C

(R
N
) : (, (Z
+
)
N
) [x[ +x

f(x) 0
_
with the multinorm p
,
: , (Z
+
)
N
, where p
,
(f):= |x

f|

.
This space is treated as a subspace of the space of all functions from R
N
to C.
10.11.7. The following statements are valid:
(1) S(R
N
) is a Frechet space;
(2) the operators of multiplication by a polynomial and dierentiation
are continuous endomorphisms of S(R
N
);
(3) the topology of S(R
N
) may equivalently be given by the multinorm
p
n
: n N, where
p
n
(f):=

]]n
|(1 +[ [
2
)
n

f|

(f S(R
N
))
(as usual, [x[ stands for the Euclidean norm of a vector x in R
N
);
(4) D(R
N
) is dense in S(R
N
); furthermore, the identical embedding
of D(R
N
) into S(R
N
) is continuous and S(R
N
)

(R
N
);
(5) S(R
N
) L
1
(R
N
).
We will check (4), because the other claims are easier.
Take f S(R
N
) and let be a truncator in D(R
N
) such that B = 1.
Given x R
N
and > 0, put

(x):= (x), and f

f. Evidently, f

D(R
N
).
Let > 0 and , (Z
+
)
N
. It is an easy matter to show that sup|

1)|

:
, (Z
+
)
N
< + for 0 < 1. Considering that x

f(x) 0 as
[x[ +, nd r > 1 satisfying [x

((

(x) 1)f(x))[ < whenever [x[ > r.


Moreover, f

(x) f(x) = ((x) 1)f(x) = 0 for [x[


1
. Therefore,
p
,
(f

f) = sup
]x]>
1
[x

((

(x) 1)f(x))[
sup
]x]>r
[x

((

(x) 1)f(x))[ <


for r
1
. Thus, p
,
(f

f) 0 as 0; i.e., f

f in S(R
N
). The required
continuity of the identical embedding raises no doubt.
216 Chapter 10
10.11.8. The Fourier transform is a continuous endomorphism of S(R
N
).
Given f D(R
N
), from 10.11.5 (8), 10.11.5 (2) and the Holder inequality
obtain
|t

f|

= |(

f)

f|
1
K|

f|

.
Thus,
|t

f|

= |t

(x

f)

(x

f)|

.
Whence it is easily seen that

f S(R
N
) and the restriction of F to D(Q) with
Q R
N
is continuous. It remains to refer to 10.10.7 (4) and 10.11.7 (4).
10.11.9. Theorem. The repeated Fourier transform in the space S(R
N
) is
proportional to the reection.
Let f S(R
N
) and g(x) := f
N
(x) = exp
_

1
/
2
[x[
2
_
. From 10.11.8 and
10.11.7 derive that

f, f, g L
1
(R
N
) and so, by 10.11.5 (7), (

fg)

= f g. Put
g

(x) := g(x) for x R


N
and > 0. Then for the same x, in view of 10.11.5 (4)
nd
_
R
N
g(t)

f(t)e
t
(x) dt =
1

N
_
R
N
f(y x) g
_
y

_
dy =
_
R
N
f(y x) g(y) dy.
Using 10.11.5 (9) and the Lebesgue Dominated Convergence Theorem as 0,
infer that
g(0)
_
R
N

f(t)e
t
(x) dt = f(x)
_
R
N
g(y) dy
= (2)
N
/
2
f(x)
_
R
N
e

1
/
2
]x]
2
dx = (2)
N
f(x).
Finally, F
2
f = (2)
N
f .
10.11.10. Corollary. F
2
2
is the reection and (F
2
)
1
= F
2
.
Given f S(R
N
) and t R
N
, deduce that
f(t) = (2)
N
_
R
N
e
i(x,t)

f(x) dx =
_
R
N
e
2i(x,t)

f(2x) dx
= (F
2
(F
2
f))(t).
Since F
2
f = F
2
f, the proof is complete.
Duality and Its Applications 217
10.11.11. Corollary. S(R
N
) is a convolution algebra (= an algebra with
convolution as multiplication).
For f, g S(R
N
) the product fg is an element of S(R
N
) and so

f g
S(R
N
). From 10.11.5 (6) infer that F
2
(f g) S(R
N
) and, consequently,
by 10.11.10, f g = F
2
(F
2
(f g)) S(R
N
).
10.11.12. Inversion Theorem. The Fourier transform F:= F
2
is a topo-
logical automorphism of the Schwartz space S(R
N
), with convolution carried into
pointwise multiplication. The inverse transform F
1
equals F
2
, with pointwise
multiplication carried into convolution. Moreover, the Parseval identity holds:
_
R
N
f g

=
_
R
N

f g

(f, g S(R
N
)).
In view of 10.11.10 and 10.11.5 (5), only the sought identity needs examining.
Moreover, given f and g, from 10.11.5 (7) and 10.11.7 (4), obtain (

fg)(0) = (

f
g)(0). Using 10.11.5 (5), conclude that
_
R
N
fg

= (F(F
1
f)g

(0) = ((F
1
f) Fg

)(0)
=
_
R
N
Ff(Fg

)dx =
_
R
N
F
f
F

(g

) dx =
_
R
N
Ff(Fg)

.
10.11.13. Remark. In view of 10.11.9, the theorem on the repeated Fourier
transform, the following mutually inverse operators
Ff(t) =
1
(2)
N
/
2
_
R
N
f(x)e
i(x,t)
dx;
F
1
f(x) =
1
(2)
N
/
2
_
R
N
f(t)e
i(x,t)
dt
are considered alongside F. In this case, an analog of 10.11.12 is valid on condition
that convolution is redened as f g := (2)

N
/
2
fg (f, g L
1
(R
N
)). The merits
of F and F
1
are connected with some simplication of 10.11.5 (8). In the case of F,
a similar goal is achieved by introducing the dierential operator D

:= (2i)
]]

with (Z
+
)
N
.
218 Chapter 10
10.11.14. Plancherel Theorem. The Fourier transform in the Schwartz
space S(R
N
) is uniquely extendible to an isometric automorphism of L
2
(R
N
).
Immediate from 10.11.12 and 4.5.10 since S(R
N
) is dense in L
2
(R
N
).
10.11.15. Remark. The extension, guaranteed by 10.11.14, retains the pre-
vious name and notation. Rarely (in search of emphasizing distinctions and sub-
tleties) one speaks of the FourierPlancherel transform or the L
2
-Fourier transform.
It that event it stands to reason to specify the understanding of the integral formulas
for Ff and F
1
f with f L
2
(R
N
) which are treated as the results of appropriate
passage to the limit in L
2
(R
N
).
10.11.16. Definition. Let u S

(R
N
) := S(R
N
)

. Such a distribution
u is referred to as a tempered distribution (variants: a distribution of slow growth,
a slowly increasing distribution, etc.). The space S

(R
N
) of tempered distributions
is furnished with the weak topology (S

(R
N
), S(R
N
)) and is sometimes called
the Schwartz space (as well as S(R
N
)).
10.11.17. Examples.
(1) L
p
(R
N
) S

(R
N
) for 1 p +.
Let f L
p
(R
N
), S(R
N
), p < + and
1
/
p
+
1
/
p
= 1. Using Holder
inequality, for suitable positive K, K

, and K

successively infer that


||
p

_
_
_
B
[[
p
_
_
1
/
p
+
_
_
_
_
R
N
\ B

(1 +[x[
2
)
N
(1 +[x[
2
)
N
(x)

p
dx
_
_
_
1
/
p
K

||

+|(1 +[ [
2
)
N
|

_
_
_
_
R
N
\ B
dx
(1 +[x[
2
)
Np
_
_
_
1
/
p
K

p
1
().
Once again using the Holder inequality, observe that
[u
f
()[ = [ [ f[ =

_
R
N
f

|f|
p
||
p
Kp
1
().
The case p = + raises no doubts.
(2) S(R
N
) is dense in S

(R
N
).
Follows from 10.11.7 (4), 10.11.17 (1), 10.11.7 (5), and 10.10.9 (4).
Duality and Its Applications 219
(3) Let M(R
N
) be a tempered Radon measure; i.e.,
_
R
N
d[[(x)
(1 +[x[
2
)
n
< +
for some n N. Evidently, is a tempered distribution.
(4) If u S

(R
N
), f S(R
N
) and (Z
+
)
N
then fu S

(R
N
)
and

u S

(R
N
) in virtue of 10.11.7 (2). By a similar argument, putting
D

u(f) := (1)
]]
uD

f for f S(R
N
), infer that D

u S

(R
N
) and D

u =
(2i)
]]

u.
(5) Every compactly-supported distribution is tempered.
In accordance with 10.10.5 (7) such u in D

(R
N
) may be identied with
a member of E

(R
N
). Since the topology of S(R
N
) is stronger than that induced
by the identical embedding in C

(R
N
), conclude that u S

(R
N
).
(6) Let u S

(R
N
). If f S(R
N
) then u convolutes with f and
u f S(R
N
). In may be shown that u also convolutes with every distribution v,
a member of E

(R
N
), and u v S

(R
N
).
(7) Take u D

(R
N
) and x R
N
. Let
x
u := (
x
)

u = u
x
be
the corresponding translation of u. A distribution u is periodic (with period x) if

x
u = u. Every periodic distribution is tempered. Periodicity is preserved under
dierentiation and convolution.
(8) If u
n
S

(R
N
) (u N) and for every f S(R
N
) there is a sum
u(f):=

n=1
u
n
(f), then u S

(R
N
) and

u =

n=1

u
n
(cf. 10.10.10).
10.11.18. Theorem. Each tempered distribution is the sum of derivatives
of tempered measures.
Let u S

(R
N
). On account of 10.11.7 (3) and 5.3.7, for some n N
and K > 0, observe that
[u(f)[ K

]]n
_
_
(1 +[ [
2
)
n

f
_
_

(f S(R
N
)).
From 3.5.3 and 3.5.7, for some

M(R
N
), obtain
u(f) =

]]n

_
(1 +[ [
2
)
n

f
_
(f S(R
N
)).
Let

:= (1)
]]
(1 +[ [
2
)
n

. Then

is tempered and u =

]]n

.
220 Chapter 10
10.11.19. Definition. For u S

(R
N
), the distribution Fu acting as
f [ Fu = Ff [ u (f S(R
N
))
is the Fourier transform or, amply, the FourierSchwartz transform of u.
10.11.20. Theorem. The FourierSchwartz transformF is a unique extension
of the Fourier transform in S(R
N
) to a topological automorphism of S

(R
N
). The
inverse F
1
of F is a unique continuous extension of the inverse Fourier transform
in S(R
N
).
The FourierSchwartz transform in S

(R
N
) is the dual of the Fourier trans-
form in S(R
N
). It remains only to appeal to 10.11.7 (5), 10.11.12, 10.11.17 (2),
and 4.5.10.
Exercises
10.1. Give examples of linear topological spaces and locally convex spaces as well as con-
structions leading to them.
10.2. Prove that a Hausdor topological vector space is nite-dimensional if and only if it is
locally compact.
10.3. Characterize weakly continuous sublinear functionals.
10.4. Prove that the weak topology of a locally convex space is normable or metrizable
if and only if the space is nite-dimensional.
10.5. Describe weak convergence in classical Banach spaces.
10.6. Prove that a normed space is nite-dimensional if and only if its unit sphere, compris-
ing all norm-one vectors, is weakly closed.
10.7. Assume that an operator T carries each weakly convergent net into a norm convergent
net. Prove that T has nite rank.
10.8. Let X and Y be Banach spaces and let T be a linear operator from X to Y . Prove
that T is bounded if and only if T is weakly continuous (i.e., continuous as a mapping from
(X, (X, X

)) to (Y, (Y, Y

))).
10.9. Let | |
1
and | |
2
be two norms making X into a Banach space. Assume further
that (X, | |
1
)

(X, | |
2
)

separates the points of X. Prove that these norms are equivalent.


10.10. Let S act from Y

to X

. When does S serve as the dual of some operator from X


to Y ?
10.11. What is the Mackey topology (X, X
#
)?
10.12. Let (X

be a family of locally convex spaces, and let X:=

be the prod-
uct of the family. Validate the presentations:
(X, X

) =

(X

, X

); (X, X

) =

(X

, X

).
10.13. Let X and Y be Banach spaces and let T be a element of B(X, Y ) satisfying
imT = Y . Demonstrate that Y is reexive provided so is X.
Duality and Its Applications 221
10.14. Show that the spaces (X

and (X

coincide.
10.15. Prove that the space c
0
has no innite-dimensional reexive subspaces.
10.16. Let p be a continuous sublinear functional on Y , and let T L (X, Y ) be a contin-
uous linear operator. Establish the following inclusion of the sets of extreme points: ext T

(p)
T

(ext p).
10.17. Let p be a continuous seminorm on X and let X be a subspace of X. Prove that
f ext(X

p) if and only if the next equality holds:


X = cl X+p f 1 p f 1.
10.18. Prove that the absolutely convex hull of a totally bounded subset of a locally convex
space is also totally bounded.
10.19. Establish that barreledness is preserved under passage to the inductive limit. What
happens with other linear topological properties?
Chapter 11
Banach Algebras
11.1. The Canonical Operator Representation
11.1.1. Definition. An element e of an algebra A is called a unity element
if e ,= 0 and ea = ae = a for all a A. Such an element is obviously unique and is
also referred to as the unity or the identity or the unit of A. An algebra A is unital
provided that A has unity.
11.1.2. Remark. Without further specication, we only consider unital alge-
bras over a basic eld F. Moreover, for simplicity, it is assumed that F:= C, unless
stated otherwise. In studying a representation of unital algebras, we naturally
presume that it preserves unity. In other words, given some algebras A
1
and A
2
,
by a representation of A
1
in A
2
we henceforth mean a morphism, a multiplicative
linear operator, from A
1
to A
2
which sends the unity element of A
1
to the unity
element of A
2
.
For an algebra Awithout unity, the process of unitization or adjunction of unity
is in order. Namely, the vector space A
e
:= A C is transformed into an algebra
by putting (a, )(b, ):= (ab +a +b, ), where a, b A and , C. In the
normed case, it is also taken for granted that |(a, )|
A
e
:= |a|
A
+[[.
11.1.3. Definition. An element a
r
in A is a right inverse of a if aa
r
= e.
An element a
l
of A is a left inverse of a if a
l
a = e.
11.1.4. If an element has left and right inverses then the latter coincide.
a
r
= (a
l
a)a
r
= a
l
(aa
r
) = a
l
e = a
l

11.1.5. Definition. An element a of an algebra A is called invertible, in
writing a Inv(A), if a has a left and right inverse. Denote a
1
:= a
r
= a
l
. The
element a
1
is the inverse of a. A subalgebra (with unity) B of an algebra A is
called pure or full or inverse-closed in A if Inv(B) = Inv(A) B.
Banach Algebras 223
11.1.6. Theorem. Let A be a Banach algebra. Given a A, put L
a
: x ax
(x A). Then the mapping
L
A
:= L : a L
a
(a A)
is a faithful operator representation. Moreover, L(A) is a pure closed subalgebra
of B(A) and L : A L(A) is a topological isomorphism.
Clearly,
L(ab) : x L
ab
(x) = abx = a(bx) = L
a
(L
b
x) = (La)(Lb)x
for x, a, b A; i.e., L is a representation (because the linearity of L is obvious).
If La = 0 then 0 = La(e) = ae = a, so that L is a faithful representation. To prove
the closure property of the range L(A), consider the algebra A
r
coinciding with A
as a vector space and equipped with the reversed multiplication ab:= ba (a, b A).
Let R:= L
A
r
, i.e. R
a
:= Ra : x xa for a A. Check that L(A) is in fact
the centralizer of the range R(A), i.e. the closed subalgebra
Z(imR):= T B(A) : TR
a
= R
a
T (a A).
Indeed, if T L(A), i.e. T = L
a
for some a A, then L
a
R
b
(x) = axb =
R
b
(L
a
(x)) = R
b
L
a
(x) for all b A. Hence, T Z(R(A)). If, in turn, T Z(R(A))
then, putting a:= Te, nd
L
a
x = ax = (Te)x = R
x
(Te) = (R
x
T)e = (TR
x
)e = T(R
x
e) = Tx
for all x A. Consequently, T = L
a
L(A). Thus, L(A) is a Banach subalgebra
of B(A).
For T = L
a
there is a T
1
in B(A). Put b := T
1
e and observe that ab =
L
a
b = Tb = TT
1
e = e. Moreover, ab = e aba = a T(ba) = L
a
ba = aba =
a = L
a
e = Te. Whence ba = e, because T is a monomorphism. Thus, L(A) is
a subalgebra of A.
By the denition of a Banach algebra, the norm is submultiplicative, providing
|L| = sup|La| : |a| 1 = sup|ab| : |a| 1, |b| 1 1.
Using the Banach Isomorphism Theorem, conclude that L is a topological isomor-
phism (i.e., L
1
is a continuous operator from L(A) onto A).
11.1.7. Definition. The representation L
A
, constructed in 11.1.6, is the
canonical operator representation of A.
224 Chapter 11
11.1.8. Remark. The presence of the canonical operator representation allows
us to conne the subsequent exposition to studying Banach algebras with norm-one
unity.
For such an algebra A the canonical operator representation L
A
implements
an isometric embedding of A into the endomorphism algebra B(A) or, in short, an
isometric representation of A in B(A). In this case, L
A
implements an isometric
isomorphism between the algebras A and L(A). The same natural terminology is
used for studying representations of arbitrary Banach algebras. Observe immedi-
ately that the canonical operator representation of A, in particular, justies the
use of the symbol in place of e for C, where e is the unity of A (cf. 5.6.5).
In other words, henceforth the isometric representation e is considered as the
identication of C with the subalgebra Ce of A.
11.2. The Spectrum of an Element of an Algebra
11.2.1. Definition. Let A be a Banach algebra and a A. A scalar in C
is a resolvent value of a, in writing res(a)), if ( a) Inv(A). The resolvent
R(a, ) of a at is R(a, ) :=
1
a
:= ( a)
1
. The set Sp(a) := C res(a) is
the spectrum of a, with a point of Sp(a) a spectral value of a. When it is necessary,
more detailed designations like Sp
A
(a) are in order.
11.2.2. For a A the equalities hold:
Sp
A
(a) = Sp
L(A)
(L
a
) = Sp(L
a
);
LR(a, ) = R(L
a
, ) ( res(a) = res(L
a
)).
11.2.3. GelfandMazur Theorem. The eld of complex numbers is up to
isometric isomorphism the sole Banach division algebra (or skew eld); i.e., each
complex Banach algebra with norm-one unity and invertible nonzero elements has
an isometric representation on C.
Consider : e, with e the unity of A and C. It is clear that
represents C in A. Take a A. By virtue of 11.2.2 and 8.1.11, Sp(a) ,= .
Consequently, there is C such that the element ( a) is not invertible; i.e.,
a = e by hypothesis. Hence, is an epimorphism. Moreover, |()| = |e| =
[[ |e| = [[ so that is an isometry.
11.2.4. Shilov Theorem. Let A be a Banach algebra and let B be a closed
(unital) subalgebra of A. Then
Sp
B
(b) Sp
A
(b), Sp
A
(b) Sp
B
(b)
for all b B.
Banach Algebras 225
If b := b Inv(B) then surely b Inv(A). Whence res
B
(b) res
A
(b);
i.e.,
Sp
B
(b) = C res
B
(b) C res
A
(b) = Sp
A
(b).
If Sp
B
(b) then b Inv(B). Therefore, there is a sequence (b
n
), b
n

Inv(B), convergent to b. Putting t := sup
nN
_
_
b
1
n
_
_
, deduce that
_
_
b
1
n
b
1
m
_
_
=
_
_
b
1
n
(1 b
n
b
1
m
)
_
_
=
_
_
b
1
n
(b
m
b
n
)b
1
m
_
_
t
2
_
_
b
n
b
m
_
_
.
In other words, if t < + then there is a limit a := limb
1
n
in B. Multiplication
is jointly continuous, and so ab = ba = 1; i.e., b Inv(B). Since Inv(B) is open
by the Banach Inversion Stability Theorem and 11.1.6, arrive at a contradiction
to the containment b Inv(B).
Therefore, it may be assumed (on dropping to a subsequence, if need be) that
_
_
b
1
n
_
_
+. Put a
n
:=
_
_
b
1
n
_
_
1
b
1
n
. Then
_
_
ba
n
_
_
=
_
_
(b b
n
)a
n
+b
n
a
n
_
_

_
_
b b
n
_
_
|a
n
| +
_
_
b
1
n
_
_
1
_
_
b
n
b
1
n
_
_
0.
Whence it follows that b is not invertible. Indeed, in the opposite case for a:= b
1
it would hold that
1 = |a
n
| =
_
_
aba
n
_
_
|a|
_
_
ba
n
_
_
0.
Finally, conclude that b does not belong to Inv(A); i.e., Sp
A
(b). Since
is a boundary point of a larger set Sp
B
(b); undoubtedly, Sp
A
(b).
11.2.5. Corollary. If Sp
B
(b) lacks interior points then Sp
B
(b) = Sp
A
(b).
Sp
B
(b) = Sp
B
(b) Sp
B
(b) Sp
A
(b) Sp
A
(b) Sp
B
(b)
11.2.6. Remark. The Shilov Theorem is often referred to as the Unremovable
Spectral Boundary Theorem or Spectral Permanence Theorem and verbalized as
follows: A boundary spectral value is unremovable.
11.3. The Holomorphic Functional Calculus in Algebras
11.3.1. Definition. Let a be an element of a Banach algebra A, and let
h H (Sp(a)) be a germ of a holomorphic function on the spectrum of a. Put
R
a
h:=
1
2i
_
h(z)
z a
dz.
The element R
a
h of A is the RieszDunford integral of h. If, in particular, f
H(Sp(a)) is a function holomorphic in a neighborhood about the spectrum of a,
then f(a):= R
a
f.
226 Chapter 11
11.3.2. GelfandDunford The RieszDunford integral R
a
represents the al-
gebra of germs of holomorphic functions on the spectrum of an element a of a Ba-
nach algebra A in A. Moreover, if f(z):=

n=0
c
n
z
n
(in a neighborhood of Sp(a))
then f(a):=

n=0
c
n
a
n
.
From 11.2.3, 8.2.1, and 11.2.2 obtain
(LR
a
h)(b) = L
R
a
h
b = (R
a
h)b
=
1
2i
_
h(z)R(a, z) dzb =
1
2i
_
h(z)R(a, z) bdz
=
1
2i
_
h(z)R(L
a
, z) bdz =
1
2i
_
h(z)R(L
a
, z) dzb
= R
L
a
h(b)
for all b A. In particular, imL includes the range of R
L
a
(H (Sp(a))). Therefore,
the already-proven commutativity of the diagram
H (Sp(a))
B(A)
A
R
a
L
R
L
a


implies that the following diagram also commutes:
H (Sp(a))
L(A)
A
R
a
L
1
R
L
a


It remains to appeal to 11.1.6 and the GelfandDunford Theorem in an operator
setting.
11.3.3. Remark. All that we have established enables us to use in the sequel
the rules of the holomorphic functional calculus which were exposed in 8.2 for the
endomorphism algebra B(X), with X a Banach space.
Banach Algebras 227
11.4. Ideals of Commutative Algebras
11.4.1. Definition. Let A be a commutative algebra. A subspace J of A is
an ideal of A, in writing J A, provided that AJ J.
11.4.2. The set J(A) of all ideals of A, ordered by inclusion, is a complete
lattice. Moreover,
sup
J(A)
E = sup
Lat(A)
E , inf
J(A)
E = inf
Lat(A)
E ,
for every subset E of J(A); i.e., J(A) is embedded into the complete lattice Lat(A)
of all subspaces of A with preservation of suprema and inma of arbitrary subsets.
Clearly, 0 is the least ideal, whereas A is the greatest ideal. Furthermore,
the intersection of ideals and the sum of nitely many ideals are ideals. It remains
to refer to 2.1.5 and 2.1.6.
11.4.3. Let J
0
A. Assume further that : A A/J
0
is the coset mapping
of A onto the quotient algebra A:= A/J
0
. Then
J A (J) A; J A
1
(J) A.
Since by denition ab := (
1
(a)
1
(b)) for a, b A, the operator is
multiplicative: (ab) = (a)(b) for a, b A. Whence successively derive
(J) A(J) = (A)(J) (AJ) (J);

1
(J) A
1
(J)
1
((A)J) =
1
(AJ)
1
(J).
11.4.4. Let J A and J ,= 0. The following conditions are equivalent:
(1) A ,= J;
(2) 1 / J;
(3) no element of J has a left inverse.
11.4.5. Definition. An ideal J of A is called proper if J is other than A.
A maximal element of the set of proper ideals (ordered by inclusion) is a maximal
ideal.
11.4.6. A commutative algebra is a division algebra if and only if it has no
proper ideals other than zero.
11.4.7. Let J be a proper ideal of A. Then (J is maximal) (A/J is a eld).
228 Chapter 11
: Let J A/J. Then, by 11.4.3,
1
(J) A. Since, beyond a doubt,
J
1
(J); therefore, either J =
1
(J) and 0 = (J) = (
1
(J)) = J, or
A =
1
(J) and J = (
1
(J)) = (A) = A/J in virtue of 1.1.6. Consequently,
A/J has no proper ideals other than zero. It remains to refer to 11.4.6.
: Let J
0
A and J
0
J. Then, by 11.4.3, (J
0
) A/J. In virtue of 11.4.6,
either (J
0
) = 0 or (J
0
) = A/J. In the rst case, J
0

1
(J
0
)
1
(0) = J
and J = J
0
. In the second case, (J
0
) = (A); i.e., A = J
0
+J J
0
+J
0
= J
0
A.
Thus, J is a maximal ideal.
11.4.8. Krull Theorem. Each proper ideal is included in some maximal ideal.
Let J
0
be a proper ideal of an algebra A. Assume further that E is the set
comprising all proper ideals J of A such that J
0
J. In virtue of 11.4.2 each chain
E
0
in E has a least upper bound: supE = J : J E
0
. By 11.4.4 the ideal sup E
0
is proper. Thus, E is inductive and the claim follows from the KuratowskiZorn
Lemma.
11.5. Ideals of the Algebra C(Q, C)
11.5.1. Minimal Ideal Theorem. Let J be an ideal of the algebra C(Q, C)
of complex continuous functions on a compactum Q. Assume further that
Q
0
:= f
1
(0) : f J;
J
0
:= f C(Q, C) : int f
1
(0) Q
0
.
Then J
0
C(Q, C) and J
0
J.
Let Q
1
:= cl(Q f
1
(0)) for a function f, a member of J
0
. By hypothesis,
Q
1
Q
0
= . To prove the containment f J it is necessary (and, certainly,
sucient) to nd u J satisfying u(q) = 1 for all q Q
1
. Indeed, in that event
uf = f.
With this in mind, observe rst that for q Q
1
there is a function f
q
in J such
that f
q
(q) ,= 0. Putting g
q
:= f

q
f
q
, where as usual f

q
: x f
q
(x)

is the conjugate
of f
q
, observe that g
q
0 and, moreover, g
q
(q) > 0. It is also clear that g
q
J for
q Q
1
. The family (U
q
)
qQ
1
, with U
q
:= x Q
1
: g
q
(x) > 0, is an open cover
of Q
1
. Using a standard compactness argument, choose a nite subset q
1
, . . . , q
n

of Q
1
such that Q
1
U
q
1
. . . U
q
n
. Put g := g
q
1
+. . . +g
q
n
. Undoubtedly, g J
and g(q) > 0 for q Q
1
. Let h
0
(q) := g(q)
1
for q Q
1
. By the TietzeUrysohn
Theorem, there is a function h in C(Q, R) satisfying h

Q
1
= h
0
. Finally, let u:= hg.
This u is a sought function.
We have thus demonstrated that J
0
J. Moreover, J
0
is an ideal of C(Q, C)
for obvious reasons.
Banach Algebras 229
11.5.2. For every closed ideal J of the algebra C(Q, C) there is a unique
compact subset Q
0
of Q such that
J = J(Q
0
):= f C(Q, C) : q Q
0
f(q) = 0.
Uniqueness follows from the Urysohn Theorem. Dene Q
0
as in 11.5.1.
Then, surely, J J(Q
0
). Take f J(Q
0
) and, given n N, put
U
n
:=
_
[f[
1
/
2n
_
, V
n
:=
_
[f[
1
/
n
_
.
Once again using the Urysohn Theorem, nd h
n
C(Q, R) satisfying 0 h
n
1
with h
n

U
n
= 0 and h
n

V
n
= 1. Consider f
n
:= fh
n
. Since
int f
1
n
(0) int U
n
Q
0
,
therefore, from 11.5.1 derive f
n
J. It suces to observe that f
n
f by con-
struction.
11.5.3. Maximal Ideal Theorem. A maximal ideal of C(Q, C) has the
form
J(q):= J(q) = f C(Q, C) : f(q) = 0,
with q a point of Q.
Follows from 11.5.2, because the closure of an ideal is also an ideal.
11.6. The Gelfand Transform
11.6.1. Let A be a commutative Banach algebra, and let J be a closed ideal of
A other than A. Then the quotient algebra A/J, endowed with the quotient norm,
is a Banach algebra. If : A A/J is the coset mapping then (1) is the unity
of A/J, the operator is multiplicative and || = 1.
Given a, b A, from 5.1.10 (5) derive
|(a)(b)|
A/J
= inf|a

|
A
: (a

) = (a), (b

) = (b)
inf|a

|
A
|b

|
A
: (a

) = (a), (b

) = (b)
= |(a)|
A/J
|(b)|
A/J
.
In other words, the norm of A/J is submultiplicative. Consequently, |(1)| 1.
Furthermore,
|(1)|
A/J
= inf|a|
A
: (a) = (1) |1|
A
= 1;
i.e., |(1)| = 1. Whence, in particular, the equality || = 1 follows. The remaining
claims are evident.
230 Chapter 11
11.6.2. Remark. The message of 11.6.1 remains valid for a noncommutative
Banach algebra A under the additional assumption that J is a bilateral ideal of A;
i.e., J is a subspace of A satisfying the condition AJA J.
11.6.3. Let : A C be a nonzero multiplicative linear functional on A.
Then is continuous and || = (1) = 1 (in particular, is a representation of A
in C).
Since ,= 0; therefore, 0 ,= (a) = (a1) = (a)(1) for some a in A.
Consequently, (1) = 1. If now a in A and in C are such that [[ > |a|, then
a Inv(A) (cf. 5.6.15). So, 1 = (1)(a)((a)
1
). Whence (a) ,= 0;
i.e., (a) ,= . Thus, [(a)[ |a| and || 1. Since || = || |1| [(1)[ = 1,
conclude that || = 1.
11.6.4. Definition. A nonzero multiplicative linear functional on an alge-
bra A is a character of A. The set of all characters of A is denoted by X(A), fur-
nished with the topology of pointwise convergence (induced in X(A) by the weak
topology (A

, A)) and called the character space of A.


11.6.5. The character space is a compactum.
It is beyond a doubt that X(A) is a Hausdor space. By virtue of 11.6.3,
X(A) is a (A

, A)-closed subset of the ball B


A
. The latter is (A

, A)-compact
by the AlaogluBourbaki Theorem. It remains to refer to 9.4.9.
11.6.6. Ideal and Character Theorem. Each maximal ideal of a commu-
tative Banach algebra A is the kernel of a character of A. Moreover, the mapping
ker from the character space X(A) onto the set M(A) of all maximal ideals
of A is one-to-one.
Let X(A) be a character of A. Clearly, ker A. From 2.3.11 it follows
that the monoquotient : A/ ker C of is a monomorphism. In view of 11.6.1,
(1) = (1) = 1; i.e., is an isomorphism of A/ ker and C. Consequently,
A/ ker is a eld. Using 11.4.7, infer that ker is a maximal ideal; i.e., ker
M(A). Now, let m M(A) be some maximal ideal of A. It is clear that m cl m,
cl m A, and 1 / cl m (because 1 Inv(A), and the last set is open by the
Banach Inversion Stability Theorem and 11.1.6). Therefore, the ideal m is closed.
Consider the quotient algebra A/m and the coset mapping : A A/m. In view
of 11.4.7 and 11.6.1, A/m is a Banach eld. By the GelfandMazur Theorem, there
is an isometric representation : A/m C. Put := . It is evident that
X(A) and ker =
1
(0) =
1
(
1
(0)) =
1
(0) = m.
To complete the proof, it suces to show that the mapping ker is one-
to-one. Indeed, let ker
1
= ker
2
for
1
,
2
X(A). By 2.3.11
1
=
2
for some
C. Furthermore, by 11.6.3, 1 =
1
(1) =
2
(1) = . Finally,
1
=
2
.
Banach Algebras 231
11.6.7. Remark. Theorem 11.6.6 makes it natural to furnish M(A) with the
inverse image topology translated from X(A) to M(A) by the mapping ker .
In this regard, M(A) is referred to as the compact maximal ideal space of A. In other
words, the character space and the maximal ideal space are often identied along
the lines of 11.6.6.
11.6.8. Definition. Let A be a commutative Banach algebra and let X(A)
be the character space of A. Given a A and X(A), put a() := (a). The
resulting function a : a(), dened on X(A), is the Gelfand transform of a.
The mapping a a, with a A is the Gelfand transform of A, denoted by G
A
(or

).
11.6.9. Gelfand Transform Theorem. The Gelfand transform G
A
: a a
is a representation of a commutative Banach algebra A in the algebra C(X(A), C).
Moreover,
Sp(a) = Sp(a) = a(X(A)),
|a| = r(a),
with a A and r(a) standing for the spectral radius of a.
The implications a A a C(X(A), C),

1 = 1 and a, b A

ab = a

b follow from denitions and 11.6.3. The linearity of G


A
raises no doubts.
Consequently, the mapping G
A
is a representation.
To begin with, take Sp(a). Then a is not invertible, and so the ideal
J
a
:= A( a) is proper in virtue of 11.4.4. By the Krull Theorem, there is
a maximal ideal m of A satisfying the condition m J
a
. By Theorem 11.6.6,
m = ker for a suitable character . In particular, ( a) = 0; i.e., = (1) =
() = (a) = a(). Consequently, Sp(a).
If, in turn, Sp(a) then ( a) is not invertible in C(X(A), C); i.e., there
is a character X(A) such that = a(). In other words, ( a) = 0. Thus,
the assumption a Inv(A) leads to the following contradiction:
1 = (1) = (( a)
1
( a)) = (( a)
1
)( a) = 0.
Hence, Sp(a). Finally, Sp(a) = Sp(a).
Using the BeurlingGelfand formula (cf. 11.3.3 and 8.1.12), infer that
r(a) = sup[[ : Sp(a) = sup[[ : Sp(a)
= sup[[ : a(X(A)) = sup[a()[ : X(A) = |a|,
what was required.
232 Chapter 11
11.6.10. The Gelfand transform of a commutative Banach algebra A is an iso-
metric embedding if and only if |a
2
| = |a|
2
for all a A.
: The mapping t t
2
, viewed as acting on R
+
, and the inverse of the
mapping on R
+
are both increasing. Therefore, from 10.6.9 obtain
|a
2
| = |a
2
|
C(X(A),C)
= sup
X(A)
[a
2
()[ = sup
X(A)
[(a
2
)[
= sup
X(A)
[(a)(a)[ = sup
X(A)
[(a)[
2
=
_
sup
X(A)
[(a)[
_
2
= |a|
2
= |a|
2
.
: By the Gelfand formula, r(a) = lim|a
n
|
1
/
n
. In particular, observe that
_
_
a
2
n
_
_
= |a|
2
n
; i.e., r(a) = |a|. By 10.6.9, r(a) = |a|, completing the proof.
11.6.11. Remark. It is interesting sometimes to grasp situations in which
the Gelfand transform of A is faithful but possibly not isometric. The kernel of the
Gelfand transform G
A
is the intersection of all maximal ideals, called the radi-
cal of A. Therefore, the condition for G
A
to be a faithful representation of A
in C(X(A), C) reads: A is semisimple or, which is the same, The radical of A
is trivial.
11.6.12. Theorem. For an element a of a commutative Banach algebra A the
following diagram of representations commutes:
H (Sp(a)) = H (Sp(a))
A
C(X(A), C)
R
a
G
A
R
a


Moreover, f(a) = f a = f(a) for f H(Sp(a)).
Take X(A). Given z res(a), observe that

_
1
z a
(z a)
_
= 1
_
1
z a
_
=
1
(z a)
=
1
z (a)
.
In other words,

R(a, z)() =

1
z a
() =
1
z a()
=
1
z a
() = R(a, z)().
Banach Algebras 233
Therefore, appealing to the properties of the Bochner integral (cf. 5.5.9 (6)) and
given f H(Sp(a)), infer that

f(a) = G
A
R
a
f = G
A
_
1
2i
_
f(z)R(a, z) dz
_
=
1
2i
_
f(z)G
A
(R(a, z)) dz =
1
2i
_
f(z)

R(a, z)) dz
=
1
2i
_
f(z)R(a, z) dz = R
a
(f) = f(a).
Furthermore, given X(A) and using the Cauchy Integral Formula, derive the
following chain of equalities
f a() = f(a()) = f((a))
=
1
2i
_
f(z)
z (a)
dz =
1
2i
_

_
f(z)
z a
_
dz
=
1
2i

__
f(z)
z a
dz
_
=

f(a)() = f(a)().
11.6.13. Remark. The theory of the Gelfand transform may be naturally
generalized to the case of a commutative Banach algebra A without unity. Retain
Denitions 11.6.4 and 11.6.8 verbatim. A character in X(A) generates some
character
e
in X(A
e
) by the rule
e
(a, ) := (a) + (a A, C). The set
X(A
e
)
e
: X(A) is a singleton consisting of the sole element

(a, ):=
(a A, C). The space X(A) is locally compact (cf. 9.4.19), because the
mapping X(A)
e
X(A
e
)

is a homeomorphism. Moreover, ker

=
A 0. Consequently, the Gelfand transform of a commutative Banach algebra
without unity represents it in the algebra of continuous complex functions dened
on a locally compact space and vanishing at innity. Given the group algebra
(L
1
(R
N
), ), observe that by 10.11.1 and 10.11.3 the Fourier transform coincides
with the Gelfand transform, which in turn entails the RiemannLebesgue Lemma
as well as the multiplication formula 10.11.5 (6).
11.7. The Spectrum of an Element of a C

-Algebra
11.7.1. Definition. An element a of an involutive algebra A is called hermit-
ian if a

= a. An element a of A is called normal if a

a = aa

. Finally, an element
a is called unitary if aa

= a

a = 1 (i.e. a, a

Inv(A) with a
1
= a

and
a
1
= a).
234 Chapter 11
11.7.2. Hermitian elements of an involutive algebra A compose a real subspace
of A. Moreover, for every a in A there are unique hermitian elements x, y A
such that a = x +iy. Namely,
x =
1
2
(a +a

), y =
1
2i
(a a

).
Moreover, a

= x iy.
Only the claim of uniqueness needs examining. If a = x
1
+ iy
1
then, using
the properties of involution (cf. 6.4.13), proceed as follows: a

= x

1
+ (iy
1
)

=
x

1
iy

1
= x
1
iy
1
. Thus, x
1
= x and y
1
= y.
11.7.3. The unity of A is a hermitian element of A.
1

= 1

1 = 1

= (1

1)

= 1

= 1
11.7.4. a Inv(A) a

Inv(A). Moreover, involution and inversion com-


mute.
For a Inv(A) by denition aa
1
= a
1
a = 1. Consequently, a
1
a

=
a

a
1
= 1

. Using 11.7.3, infer that a

Inv(A) and a
1
= a
1
. Repeating this
argument for a:= a

, complete the proof.


11.7.5. Sp(a

) = Sp(a)

.
11.7.6. The spectrum of a unitary element of a C

-algebra is a subset of the


unit circle.
By Denition 6.4.13, |a
2
| = |a

a| |a

| |a| for an arbitrary a. In other


words, |a| |a

|. Therefore, from a = a

infer that |a| = |a

|. If now a is
a unitary element, a

= a
1
; then |a|
2
= |a

a| = |a
1
a| = 1. Consequently,
|a| = |a

| = |a
1
| = 1. Whence it follows that Sp(a) and Sp(a
1
) both lie
within the unit disk. Furthermore, Sp(a
1
) = Sp(a)
1
.
11.7.7. The spectrum of a hermitian element of a C

-algebra is real.
Take a in A arbitrarily. From the GelfandDunford Theorem in an algebraic
setting derive
exp(a)

=
_

n=0
a
n
n!
_

n=0
(a
n
)

n!
=

n=0
(a

)
n
n!
= exp(a

).
If now h = h

is a hermitian element of A then, applying the holomorphic


functional calculus to a:= exp(ih), deduce that
a

= exp(ih)

= exp((ih)

) = exp(ih

) = exp(ih) = a
1
.
Consequently, a is a unitary element of A, and by 11.7.6 the spectrum Sp(a) of A is
a subset of the unit circle T. If Sp(h) then by the Spectral Mapping Theorem
(also cf. 11.3.3) exp(i) Sp(a) T. Thus, 1 = [ exp(i)[ = [ exp(i Re Im)[ =
exp(Im). Finally, Im = 0; i.e., R.
Banach Algebras 235
11.7.8. Definition. Let A be a C

-algebra. A subalgebra B of A is called


a C

-subalgebra of A if b B b

B. In this event, B is considered with the


norm induced from A.
11.7.9. Theorem. Every closed C

-subalgebra of a C

-algebra is pure.
Let B be a closed (unital) C

-subalgebra of a C

-algebra A and b B.
If b Inv(B) then it is easy that b Inv(A). Let now b Inv(A). From 11.7.4
derive b

Inv(A). Consequently, b

b Inv(A) and the element (b

b)
1
b

is a left
inverse of b. By virtue of 11.1.4, it means that b
1
= (b

b)
1
b

. Consequently,
to complete the proof it suces to show only that (b

b)
1
belongs to B. Since b

b
is hermitian in B, the inclusion holds: Sp
B
(b

b) R (cf. 11.7.7). Using 11.2.5, infer


that Sp
A
(b

b) = Sp
B
(b

b). Since 0 / Sp
A
(b

b); therefore, b

b Inv(B). Finally,
b Inv(B).
11.7.10. Corollary. Let b be an element of a C

-algebra A and let B be some


closed C

-subalgebra of A with b B. Then


Sp
B
(b) = Sp
A
(b).
11.7.11. Remark. In view of 11.7.10, Theorem 11.7.9 is often referred to as
the Spectral Purity Theorem for a C

-algebra. It asserts that the concept of the


spectrum of an element a of a C

-algebra is absolute, i.e. independent of the choice


of a C

-subalgebra containing a.
11.8. The Commutative GelfandNamark Theorem
11.8.1. The Banach algebra C(Q, C) with the natural involution f f

,
where f

(q):= f(q)

for q Q, is a C

-algebra.
|f

f| = sup[f(q)

f(q)[ : q Q = sup[f(q)[
2
: q Q
= (sup [f(Q)[)
2
= |f|
2

11.8.2. StoneWeierstrass Every unital C

-subalgebra of the C

-algebra
C(Q, C), which separates the points of Q, is dense in C(Q, C).
Let A be such subalgebra. Since f A f

A; therefore, f A
Re f A and so the set Re A:= Re f : f A is a real subalgebra of C(Q, R).
It is beyond a doubt that Re A contains constant functions and separates the points
of Q. By the StoneWeierstrass Theorem for C(Q, R), the subalgebra Re A is dense
in C(Q, R). It remains to refer to 11.7.2.
236 Chapter 11
11.8.3. Definition. A representation of a -algebra agreeing with involu-
tion is a -representation. In other words, if (A, ) and (B, ) are involutive
algebras and R : A B is a multiplicative linear operator, then R is called
a -representation of A in B whenever the following diagram commutes:
A
R
B

A
R
B
If R is also an isomorphism then R is a -isomorphism of A and B. In the pres-
ence of norms in the algebras, the naturally understood terms isometric -representation
and isometric -isomorphism are in common parlance.
11.8.4. Commutative GelfandNamark Theorem. The Gelfand trans-
form of a commutative C

-algebra A implements an isometric -isomorphism of A


and C(X(A), C).
Given a A, observe that
|a
2
| = |(a
2
)

a
2
|
1
/
2
= |a

aa

a|
1
/
2
= |a

a| = |a|
2
.
In virtue of 11.6.10 the Gelfand transform G
A
is thus an isometry of A and a closed
subalgebra

A of C(X(A), C). Undoubtedly,

A separates the points of X(A) and
contains constant functions.
By virtue of 11.6.9 and 11.7.7,

h(X(A)) = Sp(h) R for every hermitian
element h = h

of A. Now, take an arbitrary element a of A. Using 11.7.2, write


a = x+iy, where x and y are hermitian elements. The containments (x) R and
(y) R hold for every character , a member of X(A).
With this in mind, successively infer that
G
A
(a)

() = a

() = a()

= (a)

= (x +iy)

= ((x) +i(y))

= (x) i(y) = (x iy) = (a

)
=

a

() = G
A
(a

)() ( X(A)).
Consequently, the Gelfand transform G
A
is a -representation and, in particular,

A is a C

-subalgebra of C(X(A), C). It remains to appeal to 11.8.2 and conclude


that

A = C(X(A), C).
11.8.5. Assume that R : A B is a -representation of a C

-algebra A
in a C

-algebra B. Then |Ra| |a| for a A.


Banach Algebras 237
Since R(1) = 1; therefore, R(Inv(A)) Inv(B). Hence, Sp
B
(R(a))
Sp
A
(a) for a A. Whence it follows from the BeurlingGelfand formula that the
inequality r
A
(a) r
B
(R(a)) holds for the spectral radii. If a is a hermitian element
of A then R(a) is a hermitian element of B, because R(a)

= R(a

) = R(a). If now
A
0
is the least closed C

-subalgebra containing a and B


0
is an analogous subalgebra
containing R(a), then A
0
and B
0
are commutative C

-algebras. Therefore, from


Theorems 11.8.4 and 11.6.9 obtain
|R(a)| = |R(a)|
B
0
= |G
B
0
(R(a))| = r
B
0
(R(a))
= r
B
(R(a)) r
A
(a) = r
A
0
(a) = |G
A
0
(a)| = |a|.
Given a A, it is easy to observe that a

a is a hermitian element. Thus,


|R(a)|
2
= |R(a)

R(a)| = |R(a

a)| |a

a| = |a|
2
.
11.8.6. Spectral Theorem. Let a be a normal element of a C

-algebra A,
with Sp(a) the spectrum of a. There is a unique isometric -representation R
a
of C(Sp(a), C) in A such that a = R
a
(I
Sp(a)
).
Let B be the least closed C

-subalgebra of A containing a. It is clear


that the algebra B is commutative by the normality of a (this algebra presents
the closure of the algebra of all polynomials in a and a

). Moreover, by 11.7.10,
Sp(a) = Sp
A
(a) = Sp
B
(a). The Gelfand transform a := G
B
(a) of a acts from
X(B) onto Sp(a) by 11.6.9 and is evidently one-to-one. Since X(B) and Sp(a) are
compact sets; on using 9.4.11, conclude that a is a homeomorphism. Whence it is
immediate that the mapping

R : f f a implements an isometric -isomorphism


between C(Sp(a), C) and C(X(B), C).
Using Theorem 11.3.2 and the connection between the Gelfand transform and
the RieszDunford integral which is revealed in 11.6.12, for the identity mapping
observe that
a = R
a
I
C
= I
C
a = I
C

a(X(B))
a = I
C

Sp(a)
a = I
Sp(a)
a =

R(I
Sp(a)
).
Now, put
R
a
:= G
1
B

R.
Clearly, R
a
is an isometric embedding and a -representation. Moreover,
R
a
(I
Sp(a)
) = G
1
B

R(I
Sp(a)
) = G
1
B
(a) = a.
Uniqueness for such representation R
a
is guaranteed by 11.8.5 and the Stone
Weierstrass Theorem implies that the C

-algebra C(Sp(a), C) is its least closed


(unital) C

-subalgebra containing I
Sp(a)
.
238 Chapter 11
11.8.7. Definition. The representation R
a
: C(Sp(a), C) A of 11.8.6 is
the continuous functional calculus (for a normal element a of A). The element
R
a
(f) with f C(Sp(a), C) is usually denoted by f(a).
11.8.8. Remark. Let f be a holomorphic function in a neighborhood about
the spectrum of a normal element a of some C

-algebra A; i.e., f H(Sp(a)). Then


the element f(a) of A was dened by the holomorphic functional calculus. Retain
the symbol f for the restriction of f to the set Sp(a). Then, using the continuous
functional calculus, dene the element R
a
(f):= R
a
_
f

Sp(a)
_
of A. This element, as
mentioned in 11.8.7, is also denoted by f(a). The use of the designation is by far
not incidental (and sound in virtue of 11.6.12 and 11.8.6). Indeed, it would be
weird to deliberately denote by dierent symbols one and the same element. This
circumstance may be expressed in visual form.
Namely, let

Sp(a)
stand for the mapping sending a germh, a member of H (Sp(a)),
to its restriction to Sp(a); i.e., let h

Sp(a)
at a point z stand for the value of h at z
(cf. 8.1.21). It is clear that

Sp(a)
: H (Sp(a)) C(Sp(a), C). The above con-
nection between the continuous and holomorphic functional calculuses for a normal
element a of the C

-algebra A may be expressed as follows: The next diagram


commutes
A
H (Sp(a)) C(Sp(a), C)
R
a
[
Sp(a)
R
a

11.9. Operator -Representations of a C

-Algebra
11.9.1. Definition. Let A be a (unital) Banach algebra. An element s in A

is a state of A, in writing s S(A), if |s| = s(1) = 1. For a A, the set


N(a):= s(a) : s S(A) is the numeric range of a.
11.9.2. The numeric range of a positive function, a member of C(Q, C), lies
in R
+
.
Let a 0 and |s| = s(1) = 1. We have to prove that s(a) 0. Take z C
and > 0 such that the disk B

(z):= z +D includes a(Q). Then |a z| and,


consequently, [s(az)[ . Hence, [s(a) z[ = [s(a) s(z)[ ; i.e., s(a) B

(z).
Observe that
B

(z) : B

(z) a(Q) = cl co(a(Q)) R


+
.
Banach Algebras 239
Thus, s(a) R
+
.
11.9.3. Lemma. Let a be a hermitian element of a C

-algebra. Then
(1) Sp(a) N(a);
(2) Sp(a) R
+
N(a) R
+
.
Let B be the least closed C

-subalgebra, of the algebra A under study, which


contains a. It is evident that B is a commutative algebra. By virtue of 11.6.9,
the Gelfand transform a:= G
B
(a) provides a(X(B)) = Sp
B
(a). In view of 11.7.10,
Sp
B
(a) = Sp(a). In other words, for Sp(a) there is a character of B satisfying
the condition (a) = . By 11.6.3, || = (1) = 1. Using 7.5.11, nd a norm-
preserving extension s of onto A. Then s is a state of A and s(a) = . Finally,
Sp(a) N(a) (in particular, if N(a) R
+
then Sp(a) R
+
). Now, let s stand
for an arbitrary state of A. It is clear that the restriction s

B
is a state of B.
It is an easy matter to show that a maps X(B) onto Sp(a) in a one-to-one fashion.
Consequently, B may be treated as the algebra C(Sp(a), C). From 11.9.2 derive
s(a) = s

B
(a) 0 for a 0. Thus, Sp(a) R
+
N(a) R
+
, which ends the
proof.
11.9.4. Definition. An element a of a C

-algebra A is called positive if a is


hermitian and Sp(a) R
+
. The set of all positive elements of A is denoted by A
+
.
11.9.5. In each C

-algebra A the set A


+
is an ordering cone.
It is clear that N(a+b) N(a)+N(b) and N(a) = N(a) for a, b A and
R
+
. Hence, 11.9.3 ensures the inclusion
1
A
+
+
2
A
+
A
+
for
1
,
2
R
+
.
Thus, A
+
is a cone. If a A
+
(A
+
) then Sp(a) = 0. Since a is a hermitian
element, from Theorem 11.8.6 deduce that |a| = 0.
11.9.6. To every hermitian element a of a C

-algebra A there correspond some


elements a
+
and a

of A
+
such that
a = a
+
a

; a
+
a

= a

a
+
= 0.
Everything is immediate from the continuous functional calculus.
11.9.7. KaplanskyFukamija Lemma. An element a of a C

-algebra A is
positive if and only if a = b

b for some b in A.
: Let a A
+
; i.e., a = a

and Sp(a) R
+
. Then (cf. 11.8.6) there is
a square root b:=

a. Moreover, b = b

and b

b = a.
: If a = b

b then a is hermitian. Therefore, in view of 11.9.6, it may be


assumed that b

b = uv, where uv = vu = 0 with u 0 and v 0 (in the ordered


vector space (A
R
, A
+
)). Straightforward calculation yields the equalities
(bv)

bv = v

bv = vb

bv = v(u v)v = (vu v


2
)v = v
3
.
240 Chapter 11
Since v 0, it follows that v
3
0; i.e., (bv)

bv 0. By Theorem 5.6.22, Sp((bv)

bv)
and Sp(bv(bv)

) may dier only by zero. Therefore, bv(bv)

0.
In virtue of 11.7.2, bv = a
1
+ ia
2
for suitable hermitian elements a
1
and a
2
.
It is evident that a
2
1
, a
2
2
A
+
and (bv)

= a
1
ia
2
. Using 11.9.5 twice, arrive
at the estimates
0 (bv)

bv +bv(bv)

= 2
_
a
2
1
+a
2
2
_
0.
By 11.9.5, a
1
= a
2
= 0; i.e., bv = 0. Hence, v
3
= (bv)

bv = 0. The second appeal


to 11.9.5 shows v = 0. Finally, a = b

b = u v = u 0; i.e., a A
+
.
11.9.8. Every state s of a C

-algebra A is hermitian; i.e.,


s(a

) = s(a)

(a A).
By Lemmas 11.9.7 and 11.9.3, s(a

a) 0 for all a A. Putting a:= a + 1


and a:= a +i, successively infer that
0 s((a + 1)

(a + 1)) = s(a

a +a +a

+ 1) s(a) +s(a

) R;
0 s((a +i)

(a +i)) = s(a

a ia +ia

+ 1) i(s(a) +s(a

)) R.
In other words,
Ims(a) + Ims(a

) = 0; Re(s(a)) + Re s(a

) = 0.
Whence it follows that
s(a

) = Re s(a

) +i Ims(a

) = Re s(a) i Ims(a) = s(a)

.
11.9.9. Let s be a state of a C

-algebra A. Given a, b A, denote (a, b)


s
:=
s(b

a). Then ( , )
s
is a semi-inner product on A.
From 11.9.8 derive
(a, b)
s
= s(b

a) = s((a

b)

) = s(a

b)

= (b, a)

s
.
Hence, ( , )
s
is a hermitian form. Since a

a 0 for a A in virtue of 11.9.7,


(a, a)
s
= s(a

a) 0 by 11.9.3. Consequently, ( , )
s
is a positive-denite hermit-
ian form.
11.9.10. GNS-Construction Theorem. To every state s of an arbitrary
C

-algebra A there correspond a Hilbert space (H


s
, ( , )
s
), an element x
s
in H
s
and a -representation R
s
: A B(H
s
) such that s(a) = (R
s
(a)x
s
, x
s
)
s
for all
a A and the set R
s
(a)x
s
: a A is dense in H
s
.
Banach Algebras 241
In virtue of 11.9.9, putting (a, b)
s
:= s(b

a) for a, b A, obtain a pre-


Hilbert space (A, ( , )
s
). Let p
s
(a) :=
_
(a, a)
s
stand for the seminorm of the
space. Assume that
s
: A A/ ker p
s
is the coset mapping of A onto the
Hausdor pre-Hilbert space A/ ker p
s
associated with A. Assume further that

s
: A/ ker p
s
H
s
is an embedding (for instance, by the double prime map-
ping) of A/ ker p
s
onto a dense subspace of the Hilbert space H
s
associated with
(A, ( , )
s
) (cf. 6.1.10 (4)). The inner product in H
s
retains the previous notation
( , )
s
. Therefore, in particular,
(
s

s
a,
s

s
b)
s
= (a, b)
s
= s(b

a) (a, b A).
Given a A, consider (the image under the canonical operator representation)
L
a
: b ab (b A). Demonstrate rst that there are unique bounded operators
L
a
and R
s
(a) making the following diagram commutative:
A

s
A/ ker p
s

s
H
s
L
a
L
a
R
s
(a)
A

s
A/ ker p
s

s
H
s
A sought operator L
a
is a solution to the equation X
s
=
s
L
a
. Using 2.3.8,
observe next that the necessary and sucient condition for solvability of the equa-
tion in linear operators consists in invariance of the subspace ker p
s
under L
a
. Thus,
examine the inclusion L
a
(ker p
s
) ker p
s
. To this end, take an element b of ker p
s
,
i.e. p
s
(b) = 0. By denition and the CauchyBunyakovskiSchwarz inequality,
deduce that
0 (L
a
b, L
a
b)
s
= (ab, ab)
s
= s((ab)

ab)
= s(b

ab) = (a

ab, b)
s
p
s
(b)p
s
(a

ab) = 0;
i.e., L
a
b ker p
s
. Uniqueness for L
a
is provided by 2.3.9, since
s
is an epi-
morphism. Observe also that
s
is an open mapping (cf. 5.1.3). Whence the
continuity of L
a
is immediate. Therefore, in virtue of 5.3.8 the correspondence

s
L
a
(
s
)
1
may be considered as a bounded linear operator from
s
(A/ ker p
s
)
to the Banach space H
s
. By 4.5.10 such an operator extends uniquely to an operator
R
s
(a) in B(H
s
).
Demonstrate now that R
s
: a R
s
(a) is a sought representation. By 11.1.6,
L
ab
= L
a
L
b
for a, b A. Consequently,

s
L
ab
=
s
L
a
L
b
= L
a

s
L
b
= L
a
L
b

s
.
Since L
ab
is a unique solution to the equation X
s
=
s
L
ab
, infer the equality
L
ab
= L
a
L
b
, which guarantees multiplicativity for R
s
. The linearity of R
s
may be
veried likewise. Furthermore,
L
1

s
=
s
L
1
=
s
I
A
=
s
= I
A/ ker p
s

s
= 1
s
;
242 Chapter 11
i.e., R
s
(1) = 1.
For simplicity, put
s
:=
s

s
. Then, on account taken of the denition
of the inner product on H
s
(cf. 6.1.10 (4)) and the involution in B(H
s
) (cf. 6.4.14
and 6.4.5), given elements a, b, and y in A, infer that
(R
s
(a

)
s
x,
s
y)
s
= (
s
L
a
x,
s
y)
s
= (L
a
x, y)
s
= (a

x, y)
s
= s(y

x) = s((ay)

x) = (x, ay)
s
= (x, L
a
y)
s
= (
s
x,
s
L
a
y)
s
= (
s
x, R
s
(a)
s
y)
s
= (R
s
(a)

s
x,
s
y)
s
.
Now, since im
s
is dense in H
s
it follows that R
s
(a

) = R
s
(a)

for all a in A; i.e.,


R
s
is a -representation.
Let x
s
:=
s
1. Then
R
s
(a)x
s
= R
s
(a)
s
1 =
s
L
a
1 =
s
a (a A).
Consequently, the set R
s
(a)x
s
: a A is dense in H
s
. Furthermore,
(R
s
(a)x
s
, x
s
)
s
= (
s
a,
s
1)
s
= (a, 1)
s
= s(1

a) = s(a).
11.9.11. Remark. The construction, presented in the proof of 11.9.10, is
called the GNS-construction or, in expanded form, the GelfandNamarkSegal
construction, which is reected in the name of 11.9.10.
11.9.12. GelfandNamark Each C

-algebra has an isometric -representation


in the endomorphism algebra of a suitable Hilbert space.
Let A be a C

-algebra. We have to nd a Hilbert space H and an isomet-


ric -representation R of A in the C

-algebra B(H) of bounded endomorphisms


of H. For this purpose, consider the Hilbert sum H of the family of Hilbert spaces
(H
s
)
sS(A)
which exists in virtue of the GNS-Construction Theorem; i.e.,
H:=
sS(A)
H
s
=
_
_
_
h:= (h
s
)
sS(A)

sS(A)
H
s
:

sS(A)
|h
s
|
2
H
s
< +
_
_
_
.
Observe that the inner product of h:= (h
s
)
sS(A)
and g := (g
s
)
sS(A)
is calcu-
lated by the rule (cf. 6.1.10 (5) and 6.1.9):
(h, g) =

sS(A)
(h
s
, g
s
)
s
.
Banach Algebras 243
Assume further that R
s
is a -representation of A on the space H
s
correspond-
ing to s in S(A). Since in view of 11.8.5 there is an estimate |R
s
(a)|
B(H
s
)
|a|
for a A; therefore, given h H, infer that

sS(A)
|R
s
(a)h
s
|
2
H
s

sS(A)
|R
s
(a)|
2
B(H
s
)
|h
s
|
2
H
s
|a|
2

sS(A)
|h
s
|
2
H
s
.
Whence it follows that the expression R(a)h : s R
s
(a)h
s
denes an element
R(a)h of H. The resulting operator R(a) : h R(a)h is a member of B(H).
Moreover, the mapping R : a R(a) (a A) is a sought isometric -representation
of A.
Indeed, from the denition of Rand the properties of R
s
for s S(A), it follows
easily that R is a -representation of A in B(H). Check for instance that R agrees
with involution. To this end, take a A and h, g H. Then
(R(a

)h, g) =

sS(A)
(R
s
(a

)h
s
, g
s
)
s
=

sS(A)
(R
s
(a)

h
s
, g
s
)
s
=

sS(A)
(h
s
, R
s
(a)g
s
)
s
= (h, R(a)g) = (R(a)

h, g).
Since h and g in H are arbitrary, conclude that R(a

) = R(a)

.
It remains to establish only that the -representation R is an isometry, i.e. the
equalities |R(a)| = |a| for all a A. First, assume a positive. From the Spectral
Theorem and the Weierstrass Theorem it follows that |a| Sp(a). In virtue
of 11.9.3 (1) there is a state s S(A) such that s(a) = |a|. Using the properties
of the vector x
s
corresponding to the -representation R
s
(cf. 11.9.10) and applying
the CauchyBunyakovskiSchwarz inequality, infer that
|a| = s(a) = (R
s
(a)x
s
, x
s
)
s
|R
s
(a)x
s
|
H
s
|x
s
|
H
s
|R
s
(a)|
B(H
s
)
|x
s
|
2
H
s
= |R
s
(a)|
B(H
s
)
(x
s
, x
s
)
s
= |R
s
(a)|
B(H
s
)
(R
s
(1)x
s
, x
s
)
s
= |R
s
(a)|
B(H
s
)
s(1) = |R
s
(a)|
B(H
s
)
.
From the estimates |R(a)| |R
s
(a)|
B(H
s
)
and |a| |R(a)|, the former
obvious and the latter indicated in 11.8.5, derive
|a| |R(a)| |R
s
(a)|
B(H
s
)
|a|.
Finally, take a A. By the KaplanskyFukamija Lemma, a

a is positive. So,
|R(a)|
2
= |R(a)

R(a)| = |R(a

)R(a)| = |R(a

a)| = |a

a| = |a|
2
.
No further explanation is needed.
244 Chapter 11
Exercises
11.1. Give examples of Banach algebras and non-Banach algebras.
11.2. Let A be a Banach algebra. Take A
#
such that (1) = 1 and (Inv(A)) Inv(C).
Prove that is multiplicative and continuous.
11.3. Let the spectrum Sp(a) of an element a of a Banach algebra A lie in an open set U.
Prove that there is a number > 0 such that Sp(a +b) U for all b A satisfying |b| .
11.4. Describe the maximal ideal spaces of the algebras C(Q, C) and C
(1)
([0, 1], C) with
pointwise multiplication, and of the algebra of two-way innite summable sequences l
1
(Z) with
multiplication
(a b)(n):=

k=
a
nk
b
k
.
11.5. Show that a member T of the endomorphism algebra B(X) of a Banach space X has
a left inverse if and only if T is a monomorphism and the range of T is complemented in X.
11.6. Show that a member T of the endomorphism algebra B(X) of a Banach space X has
a right inverse if and only if T is an epimorphism and the kernel of T is complemented in X.
11.7. Assume that a Banach algebra A has an element with disconnected spectrum (having
a proper clopen part). Prove that A has a nontrivial idempotent.
11.8. Let A be a unital commutative Banach algebra and let E be some set of maximal
ideals of A. Such a set E is a boundary of A if |a|

= sup [a(E)[ for all a A. Prove that the


intersection of all closed boundaries of A is also a boundary of A. This is the Shilov boundary
of A.
11.9. Let A and B be unital commutative Banach algebras, with B A and 1
B
= 1
A
.
Prove that each maximal ideal of the Shilov boundary of B lies in some maximal ideal of A.
11.10. Let A and B be unital C

-algebras and let T be a morphism from A to B. Assume


further that a is a normal element of A and f is a continuous function on Sp
A
(a). Demonstrate
that Sp
B
(Ta) Sp
A
(a) and Tf(a) = f(Ta).
11.11. Let f A

, with A a commutative C

-algebra. Show that f is a positive form (i.e.,


f(a

a) 0 for a A) if and only if |f| = f(1).


11.12. Describe extreme rays of the set of positive forms on a commutative C

-algebra.
11.13. Prove that the algebras C(Q
1
, C) and C(Q
2
, C), with Q
1
and Q
2
compact, are
isomorphic if and only if Q
1
and Q
2
are homeomorphic.
11.14. Let a normal element a of a C

-algebra has real spectrum. Prove that a is hermitian.


11.15. Using the continuous functional calculus, develop a spectral theory for normal oper-
ators in a Hilbert space. Describe compact normal operators.
11.16. Let T be an algebraic morphism between C

-algebras, and |T| 1. Then T(a

) =
(Ta)

for all a.
11.17. Let T be a normal operator in a Hilbert space H. Show that there are a hermitian
operator S in H and a continuous function f : Sp(S) C such that T = f(S). Is an analogous
assertion valid in C

-algebras?
11.18. Let A and B be C

-algebras and let be a -monomorphism from A to B. Prove


that is an isometric embedding of A into B.
11.19. Let a and b be hermitian elements of a C

-algebra A. Assume that ab = ba and,


moreover, a b. Prove that f(a) f(b) for (suitable restrictions of) every increasing continuous
scalar function f over R.
References
1. Adams R., Sobolev Spaces, Academic Press, New York (1975).
2. Adash N., Ernst B., and Keim D., Topological Vector Spaces. The Theory Without Con-
vexity Conditions, Springer-Verlag, Berlin etc. (1978).
3. Akhiezer N. I., Lectures on Approximation Theory [in Russian], Nauka, Moscow (1965).
4. Akhiezer N. I., Lectures on Integral Transforms [in Russian], Vishcha Shkola, Khar

kov
(1984).
5. Akhiezer N. I. and Glazman I. M., Theory of Linear Operators in Hilbert Spaces. Vol. 1
and 2, Pitman, Boston etc. (1981).
6. Akilov G. P. and Dyatlov V. N., Fundamentals of Mathematical Analysis [in Russian],
Nauka, Novosibirsk (1980).
7. Akilov G. P. and Kutateladze S. S., Ordered Vector Spaces [in Russian], Nauka, Novosibirsk
(1978).
8. Alekseev V. M., Tikhomirov V. M., and Fomin S. V., Optimal Control [in Russian], Nauka,
Moscow (1979).
9. Alexandro P. S., Introduction to Set Theory and General Topology [in Russian], Nauka,
Moscow (1977).
10. Aliprantis Ch. D. and Border K. C., Innite-Dimensional Analysis. A Hitchhikers Guide,
Springer-Verlag, New York etc. (1994).
11. Aliprantis Ch. D. and Burkinshaw O., Locally Solid Riesz Spaces, Academic Press, New
York (1978).
12. Aliprantis Ch. D. and Burkinshaw O., Positive Operators, Academic Press, Orlando etc.
(1985).
13. Amir D., Characterizations of Inner Product Spaces, Birkhauser, Basel etc. (1986).
14. Antonevich A. B., Knyazev P. N., and Radyno Ya. B., Problems and Exercises in Functional
Analysis [in Russian], Vysheshaya Shkola, Minsk (1978).
15. Antonevich A. B. and Radyno Ya. B., Functional Analysis and Integral Equations [in Rus-
sian], Universitetskoe Publ. House, Minsk (1984).
16. Antosik P., Mikusi nski Ya., and Sikorski R., Theory of Distributions. A Sequential Ap-
proach, Elsevier, Amsterdam (1973).
17. Approximation of Hilbert Space Operators, Pitman, Boston etc. Vol. 1: Herrero D. A.
(1982); Vol. 2: Apostol C. et al. (1984).
18. Arkhangel

ski A. V., Topological Function Spaces [in Russian], Moscow University Publ.
House, Moscow (1989).
19. Arkhangel

ski A. V. and Ponomarev V. I., Fundamentals of General Topology in Problems


and Exercises [in Russian], Nauka, Moscow (1974).
20. Arveson W., An Invitation to C

-Algebra, Springer-Verlag, Berlin etc. (1976).


21. Aubin J.-P., Applied Abstract Analysis, Wiley-Interscience, New York (1977).
22. Aubin J.-P., Applied Functional Analysis, Wiley-Interscience, New York
(1979).
246 References
23. Aubin J.-P., Mathematical Methods of Game and Economic Theory, North-Holland, Ams-
terdam (1979).
24. Aubin J.-P., Nonlinear Analysis and Motivations from Economics [in French], Masson, Paris
(1984).
25. Aubin J.-P., Optima and Equilibria. An Introduction to Nonlinear Analysis, Springer-
Verlag, Berlin etc. (1993).
26. Aubin J.-P. and Ekeland I., Applied Nonlinear Analysis, Wiley-Interscience, New York etc.
(1984).
27. Aubin J.-P. and Frankowska H., Set-Valued Analysis. Systems and Control, Birkhauser,
Boston (1990).
28. Baggett L. W., Functional Analysis. A Primer, Dekker, New York etc. (1991).
29. Baggett L. W., Functional Analysis, Dekker, New York etc. (1992).
30. Baiocchi C. and Capelo A., Variational and Quasivariational Inequalities. Application to
Free Boundary Problems [in Italian], Pitagora Editrice, Bologna (1978).
31. Balakrishnan A. V., Applied Functional Analysis, Springer-Verlag, New York etc. (1981).
32. Banach S., Theorie des Operationes Lineares, Monografje Mat., Warsaw (1932).
33. Banach S., Theory of Linear Operations, North-Holland, Amsterdam (1987).
34. Bauer H., Probability Theory and Elements of Measure Theory, Academic Press, New York
etc. (1981).
35. Beals R., Advanced Mathematical Analysis, Springer-Verlag, New York etc. (1973).
36. Beauzamy B., Introduction to Banach Spaces and Their Geometry, North-Holland, Amster-
dam etc. (1985).
37. Berberian St., Lectures in Functional Analysis and Operator Theory, Springer-Verlag, Berlin
etc. (1974).
38. Berezanski Yu. M. and Kondrat

ev Yu. G., Spectral Methods in Innite-Dimensional Anal-


ysis [in Russian], Naukova Dumka, Kiev (1988).
39. Berezanski Yu. M., Us G. F., and Sheftel

Z. G., Functional Analysis. A Lecture Course


[in Russian], Vishcha Shkola, Kiev (1990).
40. Berg J. and Lofstrom J., Interpolation Spaces. An Introduction, Springer-Verlag, Berlin etc.
(1976).
41. Berger M., Nonlinearity and Functional Analysis, Academic Press, New York (1977).
42. Besov O. V., Il

in V. P., and Nikol

ski S. M., Integral Representations and Embedding


Theorems [in Russian], Nauka, Moscow (1975).
43. Bessaga Cz. and Pelczy nski A., Selected Topics in Innite-Dimensional Topology, Polish
Scientic Publishers, Warsaw (1975).
44. Birkho G., Lattice Theory, Amer. Math. Soc., Providence (1967).
45. Birkho G. and Kreyszig E., The establishment of functional analysis, Historia Math.,
11, No. 3, 258321 (1984).
46. Birman M. Sh. et al., Functional Analysis [in Russian], Nauka, Moscow (1972).
47. Birman M. Sh. and Solomyak M. Z., Spectral Theory of Selfadjoint Operators in Hilbert
Space [in Russian], Leningrad University Publ. House, Leningrad (1980).
48. Boccara N., Functional Analysis. An Introduction for Physicists, Academic Press, New York
etc. (1990).
49. Bogolyubov N. N., Logunov A. A., Oksak A. I., and Todorov I. T., General Principles
of Quantum Field Theory [in Russian], Nauka, Moscow (1987).
50. Bollobas B., Linear Analysis. An Introductory Course, Cambridge University Press, Cam-
bridge (1990).
51. Bonsall F. F. and Duncan J., Complete Normed Algebras, Springer-Verlag, Berlin etc.
(1973).
52. Boos B. and Bleecker D., Topology and Analysis. The AtiyahSinger Index Formula and
Gauge-Theoretic Physics, Springer-Verlag, Berlin etc. (1985).
53. Bourbaki N., Set Theory [in French], Hermann, Paris (1958).
54. Bourbaki N., General Topology. Parts 1 and 2, Addison-Wesley, Reading (1966).
55. Bourbaki N., Integration. Ch. 18 [in French], Hermann, Paris (1967).
References 247
56. Bourbaki N., Spectral Theory [in French], Hermann, Paris (1967).
57. Bourbaki N., General Topology. Ch. 510 [in French], Hermann, Paris (1974).
58. Bourbaki N., Topological Vector Spaces, Springer-Verlag, Berlin etc. (1981).
59. Bourbaki N., Commutative Algebra, Springer-Verlag, Berlin etc. (1989).
60. Bourgain J., New Classes of L
p
-Spaces, Springer-Verlag, Berlin etc. (1981).
61. Bourgin R. D., Geometric Aspects of Convex Sets with the RadonNikod ym Property,
Springer-Verlag, Berlin etc. (1983).
62. Bratteli O. and Robertson D., Operator Algebras and Quantum Statistical Mechanics.
Vol. 1, Springer-Verlag, New York etc. (1982).
63. Bremermann H., Distributions, Complex Variables and Fourier Transforms, Addison-Wesley,
Reading (1965).
64. Brezis H., Functional Analysis. Theory and Applications [in French], Masson, Paris etc.
(1983).
65. Brown A. and Pearcy C., Introduction to Operator Theory. Vol. 1: Elements of Functional
Analysis, Springer-Verlag, Berlin etc. (1977).
66. Bruckner A., Dierentiation of Real Functions, Amer. Math. Soc., Providence (1991).
67. Bukhvalov A. V. et al., Vector Lattices and Integral Operators [in Russian], Nauka, Novo-
sibirsk (1991).
68. Buldyrev V. S. and Pavlov P. S., Linear Algebra and Functions in Many Variables [in Rus-
sian], Leningrad University Publ. House, Leningrad (1985).
69. Burckel R., Characterization of C(X) Among Its Subalgebras, Dekker, New York etc.
(1972).
70. Buskes G., The HahnBanach Theorem Surveyed, Diss. Math., 327, Warsaw (1993).
71. Caradus S., Plaenberger W., and Yood B., Calkin Algebras of Operators on Banach Spaces,
Dekker, New York etc. (1974).
72. Carreras P. P. and Bonet J., Barrelled Locally Convex Spaces, North-Holland, Amsterdam
etc. (1987).
73. Cartan A., Elementary Theory of Analytic Functions in One and Several Complex Variables
[in French], Hermann, Paris (1961).
74. Casazza P. G. and Shura Th., Tsirelsons Spaces, Springer-Verlag, Berlin etc. (1989).
75. Chandrasekharan P. S., Classical Fourier Transform, Springer-Verlag, Berlin etc. (1980).
76. Choquet G., Lectures on Analysis. Vol. 13, Benjamin, New York and Amsterdam (1969).
77. Colombeau J.-F., Elementary Introduction to New Generalized Functions, North-Holland,
Amsterdam etc. (1985).
78. Constantinescu C., Weber K., and Sontag A., Integration Theory. Vol. 1: Measure and
Integration, Wiley, New York etc. (1985).
79. Conway J. B., A Course in Functional Analysis, Springer-Verlag, Berlin etc. (1985).
80. Conway J. B., Herrero D., and Morrel B., Completing the RieszDunford Functional Cal-
culus, Amer. Math. Soc., Providence (1989).
81. Courant R. and Hilbert D., Methods of Mathematical Physics. Vol. 1 and 2, Wiley-
Interscience, New York (1953, 1962).
82. Cryer C., Numerical Functional Analysis, Clarendon Press, New York (1982).
83. Dales H. G., Automatic continuity: a survey, Bull. London Math. Soc., 10, No. 29,
129183 (1978).
84. Dautray R. and Lions J.-L., Mathematical Analysis and Numerical Methods for Science and
Technology. Vol. 2 and 3, Springer-Verlag, Berlin etc. (1988, 1990).
85. Day M., Normed Linear Spaces, Springer-Verlag, Berlin etc. (1973).
86. De Branges L., The StoneWeierstrass theorem, Proc. Amer. Math. Soc., 10, No. 5,
822824 (1959).
87. De Branges L. and Rovnyak J., Square Summable Power Series, Holt, Rinehart and Winston,
New York (1966).
88. Deimling K., Nonlinear Functional Analysis, Springer-Verlag, Berlin etc.
(1985).
89. DeVito C. L., Functional Analysis, Academic Press, New York etc. (1978).
248 References
90. De Wilde M., Closed Graph Theorems and Webbed Spaces, Pitman, London (1978).
91. Diestel J., Geometry of Banach Spaces Selected Topics, Springer-Verlag, Berlin etc.
(1975).
92. Diestel J., Sequences and Series in Banach Spaces, Springer-Verlag, Berlin etc. (1984).
93. Diestel J. and Uhl J. J., Vector Measures, Amer. Math. Soc., Providence (1977).
94. Dieudonne J., Foundations of Modern Analysis, Academic Press, New York (1969).
95. Dieudonne J., A Panorama of Pure Mathematics. As Seen by N. Bourbaki, Academic Press,
New York etc. (1982).
96. Dieudonne J., History of Functional Analysis, North-Holland, Amsterdam etc. (1983).
97. Dinculeanu N., Vector Measures, Verlag der Wissenschaften, Berlin (1966).
98. Dixmier J., C

-Algebras and Their Representations [in French], Gauthier-Villars, Paris


(1964).
99. Dixmier J., Algebras of Operators in Hilbert Space (Algebras of Von Neumann) [in French],
Gauthier-Villars, Paris (1969).
100. Donoghue W. F. Jr., Distributions and Fourier Transforms, Academic Press, New York etc.
(1969).
101. Doob J., Measure Theory, Springer-Verlag, Berlin etc. (1993).
102. Doran R. and Bel V., Characterizations of C

-Algebras. The GelfandNamark Theorem,


Dekker, New York and Basel (1986).
103. Dowson H. R., Spectral Theory of Linear Operators, Academic Press, London etc. (1978).
104. Dugundji J., Topology, Allyn and Bacon, Boston (1966).
105. Dunford N. and Schwartz G. (with the assistance of W. G. Bade and R. G. Bartle), Linear
Operators. Vol. 1: General Theory, Interscience, New York (1958).
106. Edmunds D. E. and Evans W. D., Spectral Theory and Dierential Operators, Clarendon
Press, Oxford (1987).
107. Edwards R., Functional Analysis. Theory and Applications, Holt, Rinehart and Winston,
New York etc. (1965).
108. Edwards R., Fourier Series. A Modern Introduction. Vol. 1 and 2, Springer-Verlag, New
York etc. (1979).
109. Emov A. V., Zolotarev Yu. G., and Terpigorev V. M., Mathematical Analysis (Special
Sections). Vol. 2: Application of Some Methods of Mathematical and Functional Analysis
[in Russian], Vysshaya Shkola, Moscow (1980).
110. Ekeland I. and Temam R., Convex Analysis and Variational Problems, North-Holland, Am-
sterdam (1976).
111. Emch G., Algebraic Methods in Statistic Mechanics and Quantum Field Theory, Wiley-
Interscience, New York etc. (1972).
112. Eno P., A counterexample to the approximation property in Banach
spaces, Acta Math., 130, No. 34, 309317 (1979).
113. Engelkind R., General Topology, Springer-Verlag, Berlin etc. (1985).
114. Erdelyi I. and Shengwang W., A Local Spectral Theory for Closed Operators, Cambridge
University Press, Cambridge (1985).
115. Faris W. G., Selfadjoint Operators, Springer-Verlag, Berlin etc. (1975).
116. Fenchel W., Convex Cones, Sets and Functions, Princeton University Press, Princeton
(1953).
117. Fenchel W., Convexity through ages, in: Convexity and Its Applications, Birkhauser,
Basel etc., 1983, pp. 120130.
118. Floret K., Weakly Compact Sets, Springer-Verlag, Berlin etc. (1980).
119. Folland G. B., Fourier Analysis and Its Applications, Wadsworth and Brooks, Pacic Grove
(1992).
120. Friedrichs K. O., Spectral Theory of Operators in Hilbert Space, Springer-Verlag, New York
etc. (1980).
121. Gamelin T., Uniform Algebras, Prentice-Hall, Englewood Clis (1969).
122. Gelbaum B. and Olmsted G., Counterexamples in Analysis, Holden-Day, San Francisco
(1964).
References 249
123. Gelfand I. M., Lectures on Linear Algebra [in Russian], Nauka, Moscow (1966).
124. Gelfand I. M. and Shilov G. E., Generalized Functions and Operations over Them [in Rus-
sian], Fizmatgiz, Moscow (1958).
125. Gelfand I. M. and Shilov G. E., Spaces of Test and Generalized Functions [in Russian],
Fizmatgiz, Moscow (1958).
126. Gelfand I. M., Rakov D. A., and Shilov G. E., Commutative Normed Rings, Chelsea Pub-
lishing Company, New York (1964).
127. Gelfand I. M. and Vilenkin N. Ya., Certain Applications of Harmonic Analysis. Rigged
Hilbert Spaces, Academic Press, New York (1964).
128. Gillman L. and Jerison M., Rings of Continuous Functions, Springer-Verlag, Berlin etc.
(1976).
129. Glazman I. M. and Lyubich Yu. I., Finite-Dimensional Linear Analysis, M.I.T. Press, Cam-
bridge (1974).
130. Godement R., Algebraic Topology and Sheaf Theory [in French], Hermann, Paris (1958).
131. Goman C. and Pedrick G., First Course in Functional Analysis, Prentice-Hall, Englewood
Clis (1965).
132. Gohberg I. C. and Kren M. G., Introduction to the Theory of Nonselfadjoint Linear Oper-
ators, Amer. Math. Soc., Providence (1969).
133. Gohberg I. and Goldberg S., Basic Operator Theory, Birkhauser, Boston (1981).
134. Goldberg S., Unbounded Linear Operators. Theory and Applications, Dover, New York
(1985).
135. Gol

dshten V. M. and Reshetnyak Yu. G., Introduction to the Theory of Functions with
Generalized Derivatives and Quasiconformal Mappings [in Russian], Nauka, Moscow (1983).
136. Griel P. H., Applied Functional Analysis, Wiley, New York (1981).
137. Grothendieck A., Topological Vector Spaces, Gordon and Breach, New York etc. (1973).
138. Guerre-Delabriere S., Classical Sequences in Banach Spaces, Dekker, New York etc. (1992).
139. Gurari V. P., Group Methods in Commutative Harmonic Analysis [in Russian], VINITI,
Moscow (1988).
140. Halmos P., Finite Dimensional Vector Spaces, D. Van Nostrand Company Inc., Princeton
(1958).
141. Halmos P., Naive Set Theory, D. Van Nostrand Company Inc., New York (1960).
142. Halmos P., Introduction to Hilbert Space, Chelsea, New York (1964).
143. Halmos P., Measure Theory, Springer-Verlag, New York (1974).
144. Halmos P., A Hilbert Space Problem Book, Springer-Verlag, New York (1982).
145. Halmos P., Selecta: Expository Writing, Springer-Verlag, Berlin etc. (1983).
146. Halmos P., Has progress in mathematics slowed down? Amer. Math. Monthly, 97, No. 7,
561588 (1990).
147. Halmos P. and Sunder V., Bounded Integral Operators on L
2
Spaces, Springer-Verlag, New
York (1978).
148. Halperin I., Introduction to the Theory of Distributions, University of Toronto Press, Toron-
to (1952).
149. Harte R., Invertibility and Singularity for Bounded Linear Operators, Dekker, New York
and Basel (1988).
150. Havin V. P., Methods and structure of commutative harmonic analysis, in: Current Prob-
lems of Mathematics. Fundamental Trends. Vol. 15 [in Russian], VINITI, Moscow, 1987,
pp. 6133.
151. Havin V. P. and Nikol

ski N. K. (eds.), Linear and Complex Analysis Problem Book 3.


Parts 1 and 2, Springer-Verlag, Berlin etc. (1994).
152. Helmberg G., Introduction to Spectral Theory in Hilbert Space, North-Holland, Amsterdam
etc. (1969).
153. Herve M., The Fourier Transform and Distributions [in French], Presses Universitaires
de France, Paris (1986).
154. Heuser H., Functional Analysis, Wiley, New York (1982).
155. Heuser H., Functional Analysis [in German], Teubner, Stuttgart (1986).
250 References
156. Hewitt E. and Ross K. A., Abstract Harmonic Analysis. Vol. 1 and 2, Springer-Verlag, New
York (1994).
157. Hewitt E. and Stromberg K., Real and Abstract Analysis, Springer-Verlag, Berlin etc.
(1975).
158. Heyer H., Probability Measures on Locally Compact Groups, Springer-Verlag, Berlin etc.
(1977).
159. Hille E. and Phillips R., Functional Analysis and Semigroups, Amer. Math. Soc., Providence
(1957).
160. Hochstadt H., Edward Helly, father of the HahnBanach theorem, Math. Intelligencer,
2, No. 3, 123125 (1980).
161. Homan K., Banach Spaces of Analytic Functions, Prentice-Hall, Englewood Clis (1962).
162. Homan K., Fundamentals of Banach Algebras, University do Parana, Curitaba (1962).
163. Hog-Nlend H., Bornologies and Functional Analysis, North-Holland, Amsterdam etc. (1977).
164. Holmes R. B., Geometric Functional Analysis and Its Applications, Springer-Verlag, Berlin
etc. (1975).
165. Hormander L., Introduction to Complex Analysis in Several Variables, D. Van Nostrand
Company Inc., Princeton (1966).
166. Hormander L., The Analysis of Linear Dierential Equations. Vol. 1, Springer-Verlag, New
York etc. (1983).
167. Horn R. and Johnson Ch., Matrix Analysis, Cambridge University Press, Cambridge etc.
(1986).
168. Horvath J., Topological Vector Spaces and Distributions. Vol. 1, Addison-Wesley, Reading
(1966).
169. Husain T., The Open Mapping and Closed Graphs Theorems in Topological Vector Spaces,
Clarendon Press, Oxford (1965).
170. Husain T. and Khaleelulla S. M., Barrelledness in Topological and Ordered Vector Spaces,
Springer-Verlag, Berlin etc. (1978).
171. Hutson W. and Pym G., Applications of Functional Analysis and Operator Theory, Aca-
demic Press, London etc. (1980).
172. Ioe A. D. and Tikhomirov V. M., Theory of Extremal Problems, North-Holland, Amster-
dam (1979).
173. Istratescu V. I., Inner Product Structures, Reidel, Dordrecht and Boston (1987).
174. James R. C., Some interesting Banach spaces, Rocky Mountain J. Math., 23, No. 2.
911937 (1993).
175. Jameson G. J. O., Ordered Linear Spaces, Springer-Verlag, Berlin etc. (1970).
176. Jarchow H., Locally Convex Spaces, Teubner, Stuttgart (1981).
177. Jones D. S., Generalized Functions, McGraw-Hill Book Co., New York etc. (1966).
178. Jonge De and Van Rooij A. C. M., Introduction to Riesz Spaces, Mathematisch Centrum,
Amsterdam (1977).
179. Jorgens K., Linear Integral Operators, Pitman, Boston etc. (1982).
180. Kadison R. V. and Ringrose J. R., Fundamentals of the Theory of Operator Algebras. Vol. 1
and 2, Academic Press, New York (1983, 1986).
181. Kahane J.-P., Absolutely Convergent Fourier Series [in French], Springer-Verlag, Berlin etc.
(1970).
182. Kamthan P. K. and Gupta M., Sequence Spaces and Series, Dekker, New York and Basel
(1981).
183. Kantorovich L. V. and Akilov G. P., Functional Analysis in Normed Spaces, Pergamon
Press, Oxford etc. (1964).
184. Kantorovich L. V. and Akilov G. P., Functional Analysis, Pergamon Press, Oxford and New
York (1982).
185. Kantorovich L. V., Vulikh B. Z., and Pinsker A. G., Functional Analysis in Semiordered
Spaces [in Russian], Gostekhizdat, Moscow and Leningrad (1950).
186. Kashin B. S. and Saakyan A. A., Orthogonal Series [in Russian], Nauka, Moscow (1984).
187. Kato T., Perturbation Theory for Linear Operators, Springer-Verlag, Berlin etc. (1995).
References 251
188. Kelly J. L., General Topology, Springer-Verlag, New York etc. (1975).
189. Kelly J. L. and Namioka I., Linear Topological Spaces, Springer-Verlag, Berlin etc. (1976).
190. Kelly J. L. and Srinivasan T. P., Measure and Integral. Vol. 1, Springer-Verlag, New York
etc. (1988).
191. Kesavan S., Topics in Functional Analysis and Applications, Wiley, New York etc. (1989).
192. Khaleelulla S. M., Counterexamples in Topological Vector Spaces, Springer-Verlag, Berlin
etc. (1982).
193. Khelemski A. Ya., Banach and Polynormed Algebras: General Theory, Representation and
Homotopy [in Russian], Nauka, Moscow (1989).
194. Kirillov A. A., Elements of Representation Theory [in Russian], Nauka, Moscow (1978).
195. Kirillov A. A. and Gvishiani A. D., Theorems and Problems of Functional Analysis [in Rus-
sian], Nauka, Moscow (1988).
196. Kislyakov S. V., Regular uniform algebras are not complemented, Dokl. Akad. Nauk
SSSR, 304, No. 1, 795798 (1989).
197. Knyazev P. N., Functional Analysis [in Russian], Vysheshaya Shkola, Minsk (1985).
198. Kollatz L., Functional Analysis and Numeric Mathematics [in German],
Springer-Verlag, Berlin etc. (1964).
199. Kolmogorov A. N., Selected Works. Mathematics and Mechanics [in Russian], Nauka,
Moscow (1985).
200. Kolmogorov A. N. and Fomin S. V., Elements of Function Theory and Functional Analysis
[in Russian], Nauka, Moscow (1989).
201. Korner T. W., Fourier Analysis, Cambridge University Press, Cambridge (1988).
202. Korotkov V. B., Integral Operators [in Russian], Nauka, Novosibirsk (1983).
203. Kothe G., Topological Vector Spaces. Vol. 1 and 2, Springer-Verlag, Berlin etc. (1969,
1980).
204. Kostrikin A. I. and Manin Yu. I., Linear Algebra and Geometry [in Russian], Moscow
University Publ. House, Moscow (1986).
205. Krasnosel

ski M. A., Positive Solutions of Operator Equations, P.Noordho Ltd., Groningen


(1964).
206. Krasnosel

ski M. A., Lifshits E. A., and Sobolev A. V., Positive Linear Systems. The
Method of Positive Operators [in Russian], Nauka, Moscow (1985).
207. Krasnosel

ski M. A. and Rutitski Ya. B., Convex Functions and Orlicz Spaces, Noordho,
Groningen (1961).
208. Krasnosel

ski M. A. and Zabreko P. P., Geometric Methods of Nonlinear Analysis, Springer-


Verlag, Berlin (1984).
209. Krasnosel

ski M. A. et al., Integral Operators in the Spaces of Summable Functions, No-


ordho International Publishing, Leyden (1976).
210. Kren S. G., Linear Dierential Equations in Banach Space [in Russian], Nauka, Moscow
(1967).
211. Kren S. G., Linear Equations in Banach Space [in Russian], Nauka, Moscow (1971).
212. Kren S. G., Petunin Yu. I., and Semenov E. M., Interpolation of Linear Operators [in Rus-
sian], Nauka, Moscow (1978).
213. Kreyszig E., Introductory Functional Analysis with Applications, Wiley, New York (1989).
214. Kudryavtsev L. D., A Course in Mathematical Analysis. Vol. 2 [in Russian], Vysshaya
Shkola, Moscow (1981).
215. Kuratowski K., Topology. Vol. 1 and 2, Academic Press, New York and London (1966,
1968).
216. Kusraev A. G., Vector Duality and Its Applications [in Russian], Nauka, Novosibirsk (1985).
217. Kusraev A. G. and Kutateladze S. S., Subdierentials: Theory and Applications, Kluwer,
Dordrecht (1995).
218. Kutateladze S. S. and Rubinov A. M., Minkowski Duality and Its Applications [in Russian],
Nauka, Novosibirsk (1976).
219. Lacey H., The Isometric Theory of Classical Banach Spaces, Springer-Verlag, Berlin etc.
(1973).
252 References
220. Ladyzhenskaya O. A., Boundary Value Problems of Mathematical Physics [in Russian],
Nauka, Moscow (1973).
221. Lang S., Introduction to the Theory of Dierentiable Manifolds, Columbia University, New
York (1962).
222. Lang S., Algebra, Addison-Wesley, Reading (1965).
223. Lang S., SL(2, R), Addison-Wesley, Reading (1975).
224. Lang S., Real and Functional Analysis, Springer-Verlag, New York etc. (1993).
225. Larsen R., Banach Algebras, an Introduction, Dekker, New York etc. (1973).
226. Larsen R., Functional Analysis, an Introduction, Dekker, New York etc. (1973).
227. Leifman L. (ed.), Functional Analysis, Optimization and Mathematical Economics, Oxford
University Press, New York and Oxford (1990).
228. Levin V. L., Convex Analysis in Spaces of Measurable Functions and Its Application in Math-
ematics and Economics [in Russian], Nauka, Moscow (1985).
229. Levy A., Basic Set Theory, Springer-Verlag, Berlin etc. (1979).
230. Levy P., Concrete Problems of Functional Analysis (with a supplement by F. Pellegrino on
analytic functionals) [in French], Gauthier-Villars, Paris (1951).
231. Lindenstrauss J. and Tzafriri L., Classical Banach Spaces, Springer-Verlag, Berlin etc.
Vol. 1: Sequence Spaces (1977). Vol. 2: Function Spaces (1979).
232. Llavona J. G., Approximation of Continuously Dierentiable Functions,
North-Holland, Amsterdam etc. (1988).
233. Loomis L. H., An Introduction to Abstract Harmonic Analysis, D. Van Nostrand Company
Inc., Princeton (1953).
234. Luecking P. H. and Rubel L. A., Complex Analysis. A Functional Analysis Approach,
Springer-Verlag, Berlin etc. (1984).
235. Luxemburg W. A. J. and Zaanen A. C., Riesz Spaces. Vol. 1, North-Holland, Amsterdam
etc. (1971).
236. Lyubich Yu. I., Introduction to the Theory of Banach Representations of Groups [in Rus-
sian], Vishcha Shkola, Khar

kov (1985).
237. Lyubich Yu. I., Linear Functional Analysis, Springer-Verlag, Berlin etc.
(1992).
238. Lyusternik L. A. and Sobolev V. I., A Concise Course in Functional Analysis [in Russian],
Vysshaya Shkola, Moscow (1982).
239. Mackey G. W., The Mathematical Foundations of Quantum Mechanics, Benjamin, New
York (1964).
240. Maddox I. J., Elements of Functional Analysis, Cambridge University Press, Cambridge
(1988).
241. Malgrange B., Ideals of Dierentiable Functions, Oxford University Press, New York (1967).
242. Malliavin P., Integration of Probabilities. Fourier Analysis and Spectral Analysis [in French],
Masson, Paris (1982).
243. Marek I. and Zitn y K., Matrix Analysis for Applied Sciences. Vol. 1, Teubner, Leipzig
(1983).
244. Mascioni V., Topics in the theory of complemented subspaces in Banach spaces, Exposi-
tiones Math., 7, No. 1, 347 (1989).
245. Maslov V. P., Operator Methods [in Russian], Nauka, Moscow (1973).
246. Maurin K., Methods of Hilbert Space, Polish Scientic Publishers, Warsaw (1967).
247. Maurin K., Analysis. Vol. 2: Integration, Distributions, Holomorphic Functions, Tensor and
Harmonic Analysis, Polish Scientic Publishers, Warsaw (1980).
248. Maz

ya V. G., The Spaces of S. L. Sobolev [in Russian], Leningrad University Publ. House,
Leningrad (1985).
249. Meyer-Nieberg P., Banach Lattices, Springer-Verlag, Berlin etc. (1991).
250. Michor P. W., Functors and Categories of Banach Spaces, Springer-Verlag, Berlin etc.
(1978).
251. Mikhlin S. G., Linear Partial Dierential Equations [in Russian], Vysshaya Shkola, Moscow
(1977).
References 253
252. Milman V. D. and Schechtman G., Asymptotic Theory of Finite Dimensional Normed
Spaces, Springer-Verlag, Berlin etc. (1986).
253. Miranda C., Fundamentals of Linear Functional Analysis. Vol. 1 and 2 [in Italian], Pitagora
Editrice, Bologna (1978, 1979).
254. Misra O. P. and Lavoine J. L., Transform Analysis of Generalized Functions, North-Holland,
Amsterdam etc. (1986).
255. Mizohata S., Theory of Partial Dierential Equations [in Russian], Mir, Moscow (1977).
256. Moore R., Computational Functional Analysis, Wiley, New York (1985).
257. Morris S., Pontryagin Duality and the Structure of Locally Compact Abelian Groups, Cam-
bridge University Press, Cambridge (1977).
258. Motzkin T. S., Endovectors in convexity, in: Proc. Sympos. Pure Math., 7, Amer. Math.
Soc., Providence, 1963, pp. 361387.
259. Namark M. A., Normed Rings, Noordho, Groningen (1959).
260. Napalkov V. V., Convolution Equations in Multidimensional Spaces [in Russian], Nauka,
Moscow (1982).
261. Narici L. and Beckenstein E., Topological Vector Spaces, Dekker, New York (1985).
262. Naylor A. and Sell G., Linear Operator Theory in Engineering and Science, Springer-Verlag,
Berlin etc. (1982).
263. Neumann J. von, Mathematical Foundations of Quantum Mechanics [in German], Springer-
Verlag, Berlin (1932).
264. Neumann J. von, Collected Works, Pergamon Press, Oxford (1961).
265. Neveu J., Mathematical Foundations of Probability Theory [in French], Masson et Cie, Paris
(1964).
266. Nikol

ski N. K., Lectures on the Shift Operator [in Russian], Nauka, Moscow (1980).
267. Nikol

ski S. M., Approximation to Functions in Several Variables and Embedding Theorems


[in Russian], Nauka, Moscow (1977).
268. Nirenberg L., Topics in Nonlinear Functional Analysis. 19731974 Notes by R. A Artino,
Courant Institute, New York (1974).
269. Oden J. T., Applied Functional Analysis. A First Course for Students of Mechanics and
Engineering Science, Prentice-Hall, Englewood Clis (1979).
270. Palais R. (with contributions by M. F. Atiyah et al.), Seminar on the AtiyahSinger Index
Theorem, Princeton University Press, Princeton (1965).
271. Palamodov V. P., Linear Dierential Operators with Constant Coecients, Springer-Verlag,
Berlin etc. (1970).
272. Pedersen G. K., Analysis Now, Springer-Verlag, New York etc. (1989).
273. Pelczy nski A., Banach Spaces of Analytic Functions and Absolutely Summing Operators,
Amer. Math. Soc., Providence (1977).
274. Petunin Yu. I. and Plichko A. N., Theory of Characteristics of Subspaces and Its Applications
[in Russian], Vishcha Shkola, Kiev (1980).
275. Phelps R., Convex Functions, Monotone Operators and Dierentiability,
Springer-Verlag, Berlin etc. (1989).
276. Pietsch A., Nuclear Locally Convex Spaces [in German], Akademie-Verlag, Berlin (1967).
277. Pietsch A., Operator Ideals, VEB Deutschen Verlag der Wissenschaften, Berlin (1978).
278. Pietsch A., Eigenvalues and S-Numbers, Akademie-Verlag, Leipzig (1987).
279. Pisier G., Factorization of Linear Operators and Geometry of Banach Spaces, Amer. Math.
Soc., Providence (1986).
280. Plesner A. I., Spectral Theory of Linear Operators [in Russian], Nauka, Moscow (1965).
281. Pr ossdorf S., Some Classes of Singular Equations [in German], Akademie-Verlag, Berlin
(1974).
282. Radjavi H. and Rosenthal P., Invariant Subspaces, Springer-Verlag, Berlin etc. (1973).
283. Radyno Ya. V., Linear Equations and Bornology [in Russian], The Belorussian State Uni-
versity Publ. House, Minsk (1982).
284. Rakov D. A., Vector Spaces [in Russian], Fizmatgiz, Moscow (1962).
254 References
285. Reed M. and Simon B., Methods of Modern Mathematical Physics, Academic Press, New
York and London (1972).
286. Reshetnyak Yu. G., Vector Measures and Some Questions of the Theory of Functions of a Re-
al Variable [in Russian], Novosibirsk University Publ. House, Novosibirsk (1982).
287. Richards J. Ian and Joun H. K., Theory of Distributions: a Nontechnical Introduction,
Cambridge University Press, Cambridge (1990).
288. Richtmyer R., Principles of Advanced Mathematical Physics. Vol. 1, Springer-Verlag, New
York etc. (1978).
289. Rickart Ch., General Theory of Banach Algebras, D. Van Nostrand Company Inc., Princeton
(1960).
290. Riesz F. and Szokefalvi-Nagy B., Lectures on Functional Analysis [in French], Akademi
Kiado, Budapest (1972).
291. Robertson A. and Robertson V., Topological Vector Spaces, Cambridge University Press,
Cambridge (1964).
292. Rockafellar R., Convex Analysis, Princeton University Press, Princeton
(1970).
293. Rolewicz S., Functional Analysis and Control Theory [in Polish], Panstwowe Wydawnictwo
Naukowe, Warsaw (1977).
294. Rolewicz S., Metric Linear Spaces, Reidel, Dordrecht etc. (1984).
295. Roman St., Advanced Linear Algebra, Springer-Verlag, Berlin etc. (1992).
296. Rudin W., Fourier Analysis on Groups, Interscience, New York (1962).
297. Rudin W., Functional Analysis, McGraw-Hill Book Co., New York (1973).
298. Sadovnichi V. A., Operator Theory [in Russian], Moscow University Publ. House, Moscow
(1986).
299. Sakai S., C

-Algebras and W

-Algebras, Springer-Verlag, Berlin etc. (1971).


300. Samuelides M. and Touzillier L., Functional Analysis [in French], Toulouse, Cepadues

Editions
(1983).
301. Sard A., Linear Approximation, Amer. Math. Soc., Providence (1963).
302. Schaefer H. H., Topological Vector Spaces, Springer-Verlag, New York etc. (1971).
303. Schaefer H. H., Banach Lattices and Positive Operators, Springer-Verlag, Berlin etc. (1974).
304. Schapira P., Theory of Hyperfunctions [in French], Springer-Verlag, Berlin etc. (1970).
305. Schechter M., Principles of Functional Analysis, Academic Press, New York etc. (1971).
306. Schwartz L. (with participation of Denise Huet), Mathematical Methods for Physical Sci-
ences [in French], Hermann, Paris (1961).
307. Schwartz L., Theory of Distributions [in French], Hermann, Paris (1966).
308. Schwartz L., Analysis. Vol. 1 [in French], Hermann, Paris (1967).
309. Schwartz L., Analysis: General Topology and Functional Analysis [in French], Hermann,
Paris (1970).
310. Schwartz L., Hilbertian Analysis [in French], Hermann, Paris (1979).
311. Schwartz L., Geometry and probability in Banach spaces, Bull. Amer. Math. Soc., 4,
No. 2. 135141 (1981).
312. Schwarz H.-U., Banach Lattices and Operators, Teubner, Leipzig (1984).
313. Segal I. and Kunze R., Integrals and Operators, Springer-Verlag, Berlin etc. (1978).
314. Semadeni Zb., Banach Spaces of Continuous Functions, Warsaw, Polish Scientic Publishers
(1971).
315. Shafarevich I. R., Basic Concepts of Algebra [in Russian], VINITI, Moscow (1986).
316. Shilov G. E., Mathematical Analysis. Second Optional Course [in Russian], Nauka, Moscow
(1965).
317. Shilov G. E. and Gurevich B. L., Integral, Measure, and Derivative [in Russian], Nauka,
Moscow (1967).
318. Sinclair A., Automatic Continuity of Linear Operators, Cambridge University Press, Cam-
bridge (1976).
319. Singer I., Bases in Banach Spaces. Vol. 1 and 2, Springer-Verlag, Berlin etc. (1970, 1981).
320. Smirnov V. I., A Course of Higher Mathematics. Vol. 5, Pergamon Press, New York (1964).
References 255
321. Sobolev S. L., Selected Topics of the Theory of Function Spaces and Generalized Functions
[in Russian], Nauka, Moscow (1989).
322. Sobolev S. L., Applications of Functional Analysis in Mathematical Physics, Amer. Math.
Soc., Providence (1991).
323. Sobolev S. L., Cubature Formulas and Modern Analysis, Gordon and Breach, Montreux
(1992).
324. Steen L. A., Highlights in the history of spectral theory, Amer. Math. Monthly, 80,
No. 4., 359381 (1973).
325. Steen L. A. and Seebach J. A., Counterexamples in Topology, Springer-Verlag, Berlin etc.
(1978).
326. Stein E. M., Singular Integrals and Dierentiability Properties of Functions,
Princeton University Press, Princeton (1971).
327. Stein E. M., Harmonic Analysis, Real-Variable Methods, Orthogonality, and Oscillatory
Integrals, Princeton University Press, Princeton (1993).
328. Stein E. M. and Weiss G., Introduction to Harmonic Analysis on Euclidean
Spaces, Princeton University Press, Princeton (1970).
329. Stone M., Linear Transformations in Hilbert Space and Their Application to Analysis, Amer.
Math. Soc., New York (1932).
330. Sundaresan K. and Swaminathan Sz., Geometry and Nonlinear Analysis in Banach Spaces,
Berlin etc. (1985).
331. Sunder V. S., An Invitation to Von Neumann Algebras, Springer-Verlag, New York etc.
(1987).
332. Swartz Ch., An Introduction to Functional Analysis, Dekker, New York etc. (1992).
333. Szankowski A., B(H) does not have the approximation property, Acta Math., 147, No. 1
2, 89108 (1979).
334. Takeuti G. and Zaring W., Introduction to Axiomatic Set Theory, Springer-Verlag, New
York etc. (1982).
335. Taldykin A. T., Elements of Applied Functional Analysis [in Russian], Vysshaya Shkola,
Moscow (1982).
336. Taylor A. E. and Lay D. C., Introduction to Functional Analysis, Wiley, New York (1980).
337. Taylor J. L., Measure Algebras, Amer. Math. Soc., Providence (1973).
338. Tiel J. van, Convex Analysis. An Introductory Theory, Wiley, Chichester (1984).
339. Tikhomirov V. M., Convex analysis, in: Current Problems of Mathematics. Fundamental
Trends. Vol. 14 [in Russian], VINITI, Moscow, 1987, pp. 5101.
340. Trenogin V. A., Functional Analysis [in Russian], Nauka, Moscow (1980).
341. Trenogin V. A., Pisarevski B. M., and Soboleva T. S., Problems and Exercises in Functional
Analysis [in Russian], Nauka, Moscow (1984).
342. Treves F., Locally Convex Spaces and Linear Partial Dierential Equations, Springer-Verlag,
Berlin etc. (1967).
343. Triebel H., Theory of Function Spaces, Birkhauser, Basel (1983).
344. Uspenski S. V., Demidenko G. V., and Perepelkin V. G., Embedding Theorems and Appli-
cations to Dierential Equations [in Russian], Nauka, Novosibirsk (1984).
345. Vanberg M. M., Functional Analysis [in Russian], Prosveshchenie, Moscow (1979).
346. Vladimirov V. S., Generalized Functions in Mathematical Physics [in Russian], Nauka,
Moscow (1976).
347. Vladimirov V. S., Equations of Mathematical Physics [in Russian], Nauka, Moscow (1988).
348. Vladimirov V. S. et al., A Problem Book on Equations of Mathematical Physics [in Russian],
Nauka, Moscow (1982).
349. Voevodin V. V., Linear Algebra [in Russian], Nauka, Moscow (1980).
350. Volterra V., Theory of Functionals, and Integral and Integro-Dierential Equations, Dover,
New York (1959).
351. Vulikh B. Z., Introduction to Functional Analysis, Pergamon Press, Oxford (1963).
352. Vulikh B. Z., Introduction to the Theory of Partially Ordered Spaces, Noordho, Groningen
(1967).
256 References
353. Waelbroeck L., Topological Vector Spaces and Algebras, Springer-Verlag, Berlin etc. (1971).
354. Wagon S., The BanachTarski Paradox, Cambridge University Press, Cambridge (1985).
355. Weidmann J., Linear Operators in Hilbert Spaces, Springer-Verlag, New York etc. (1980).
356. Wells J. H. and Williams L. R., Embeddings and Extensions in Analysis, Springer-Verlag,
Berlin etc. (1975).
357. Wermer J., Banach Algebras and Several Convex Variables, Springer-Verlag, Berlin etc.
(1976).
358. Wiener N., The Fourier Integral and Certain of Its Applications, Cambridge University
Press, New York (1933).
359. Wilanski A., Functional Analysis, Blaisdell, New York (1964).
360. Wilanski A., Topology for Analysis, John Wiley, New York (1970).
361. Wilanski A., Modern Methods in Topological Vector Spaces, McGraw-Hill Book Co., New
York (1980).
362. Wojtaszczyk P., Banach Spaces for Analysis, Cambridge University Press, Cambridge (1991).
363. Wong Yau-Chuen, Introductory Theory of Topological Vector Spaces, Dekker, New York
etc. (1992).
364. Yood B., Banach Algebras, an Introduction, Carleton University, Ottawa (1988).
365. Yosida K., Operational Calculus, Theory of Hyperfunctions, Springer-Verlag, Berlin etc.
(1982).
366. Yosida K., Functional Analysis, Springer-Verlag, New York etc. (1995).
367. Young L., Lectures on the Calculus of Variations and Optimal Control Theory, W. B. Saun-
ders Company, Philadelphia etc. (1969).
368. Zaanen A. C., Riesz Spaces. Vol. 2, North-Holland, Amsterdam etc. (1983).
369. Zemanian A. H., Distribution Theory and Transform Analysis, Dover, New York (1987).
370. Ziemer W. P., Weakly Dierentiable Functions. Sobolev Spaces and Functions of Bounded
Variation, Springer-Verlag, Berlin etc. (1989).
371. Zimmer R. J., Essential Results of Functional Analysis, Chicago University Press, Chicago
and London (1990).
372. Zorich V. A., Mathematical Analysis. Part 2 [in Russian], Nauka, Moscow (1984).
373. Zuily C., Problems in Distributions and Partial Dierential Equations, North-Holland, Am-
sterdam etc. (1988).
374. Zygmund A., Trigonometric Series. Vol. 1 and 2, Cambridge University Press, New York
(1959).
Notation Index
:=, ix
, , ix
A
r
, 11.1.6
A B, 1.1.1
B
p
, 5.1.1, 5.2.11

B
p
, 5.1.1
B
T
, 5.1.3
B
X
, 5.1.10, 5.2.11
B(X), 5.6.4,
B(E , F), 5.5.9 (2)
B(X, Y ), 5.1.10 (7)
C(Q, F), 4.6.8
C
(m)
, 10.9.9
C

(), 10.10.2 (3)


D

, 10.11.13
F
1
, 1.1.3 (1)
F(B), 1.3.5 (1)
F
p
, 5.5.9 (6)
F[
U
, 1.1.3 (5)
F(U), 1.1.3 (5)
F(a
1
, ), 1.1.3 (6)
F( , a
2
), 1.1.3 (6)
F( , ), 1.1.3 (6)
F(X, Y ), 8.3.6

G, 10.11.2

G, 10.11.2
G F, 1.1.4
H

, 6.1.10 (3)
H(K), 8.1.13
H

(U), 3.1.11
I
C
, 8.2.10
I
U
, 1.1.3 (3)
J(q), 11.5.3
J(Q
0
), 11.5.2
J A, 11.4.1
K(E), 10.9.1
K(Q), 10.9.1
K(), 10.9.1
L
A
, 11.1.6
L
p
, 5.5.9 (4), 5.5.9 (6)
L
p
(X), 5.5.9
L
Q
0
, 10.8.4 (3)
L

Q
0
, 10.8.4 (4)
L

, 5.5.9 (5)
M(), 10.9.4 (2)
N(a), 11.9.1
N
p
, 5.5.9 (6)
P
H
0
, 6.2.7
P

, 8.2.10
P
X
1
||X
2
, 2.2.9 (4)
P
1
P
2
, 6.2.12
R(a, ), 11.2.1
R(T, ), 5.6.13
S(A), 11.9.1
T

, 7.6.2
T

, 6.4.4
|T|, 5.1.10 (7)
U

, 10.5.7
U

, 6.2.5
U (), 3.1.1
X[, 10.3.1
X

, 5.1.10 (8), 10.2.11


X

, 5.1.10 (8)
X

, 2.1.4 (2)
X
+
, 3.2.5
X

0
, 7.6.8
X

, 8.2.10
X
N
, 2.1.4 (4)
X
#
, 2.2.4
258 Notation Index
X
R
, 3.7.1
X

, 2.1.4 (4)
X = X
1
X
2
, 2.1.7
X iX, 8.4.8
X
1
X
2
. . . X
N
, 2.1.4 (4)
(X, )

, 10.2.11
X/X
0
, 2.1.4 (6)
(X/X
0
, p
X/X
0
), 5.1.10 (5)
X Y , 10.3.3
X Y , 2.2.6
[ Y ), 10.3.1
B, 9.6.14
C, 2.1.2
D, 8.1.3
F, 2.1.2
N, 1.2.16
Q, 7.4.11
R, 2.1.2
R

, 3.4.1
R
+
, 3.1.2 (4)
R, 3.8.1
Re, 3.7.3
Re
1
, 3.7.4
T, 8.1.3
Z, 8.5.1
Z
+
, 10.10.2 (2)
A
e
, 11.1.2
D(Q), 10.10.1
D(), 10.10.1
D

(), 10.10.4
D

F
(), 10.10.8
D
(m)
(Q), 10.10.8
D
(m)
(), 10.10.8
D
(m)
()

, 10.10.8
E (), 10.10.2 (3)
E

(R
N
), 10.10.5 (9)
E T, 2.2.8
F, 10.11.4
F
p
, 5.5.9 (6)
Fr(X, Y ), 8.5.1
F(X), 1.3.6
G
A
, 11.6.8
H (K), 8.1.14
K (X), 8.3.3
K (X, Y ), 6.6.1
L(X), 2.2.8
L(X, Y ), 2.2.3
L
r
(X, Y ), 3.2.6 (3)
L

, 5.5.9 (5)
M(), 10.9.3
N (), 10.8.11
N
p
(f), 5.5.9 (4)
N

, 5.5.9 (5)
P(X), 1.2.3 (4)
R
T
, 8.2.1
R
a
h, 11.3.1
S(R
N
), 10.11.6
S

(R
N
), 10.11.16
T (X), 9.1.2
U
p
, 5.2.2
U
M
, 5.2.4
U
X
, 4.1.5, 5.2.4

X
0
, 7.6.8
Fu, 10.11.19
M, 5.3.9
M N, 5.3.1
M N, 5.3.1
M
X
, 5.1.6
M

, 10.2.7
N T, 5.1.10 (3)
N
T
, 5.1.10 (3)
R
a
, 11.8.7
Cl(), 4.1.15, 9.1.4
Imf, 5.5.9 (4)
Inv(A), 11.1.5
Inv(X, Y ), 5.6.12

B
, 8.1.2 (4)
Lat(X), 2.1.5
LCT(X), 10.2.3
M(A), 11.6.6
Op(), 4.1.11, 9.1.4
Re, 2.1.2
Re f, 5.5.9 (4)
Sp(a), 11.2.1
Sp
A
(a), 11.2.1
Sp(T), 5.6.13
T
1
, 9.3.2
T
2
, 9.3.5
T
3
, 9.3.9
T
3
1
/
2
, 9.3.15
T
4
, 9.3.11
T(X), 9.1.7
Tr (), 10.10.2
VT(X), 10.1.5
X(A), 11.6.4
, 10.9.4 (1)

(1)
, 10.10.5 (4)

q
, 10.9.4 (1)

, 10.9.4 (3)
Notation Index 259

+
, 10.8.13

, 10.8.13
[[, 10.8.13, 10.9.4 (3)
||, 10.9.5

, 10.9.4 (4)

1

2
, 10.9.4 (6)

1

2
, 10.9.4 (6)
f, 10.9.4 (7)
, 10.9.4 (7)
(U), 10.5.1, 10.5.7

1
(V ), 10.5.1

1
F
(
F
(U)), 10.5.5
2
, 10.11.4

, 8.2.9
(T), 5.6.13
(X, Y ), 10.3.5
(X, Y ), 10.4.4

a
f, 10.9.4 (1)

M
, 5.2.8

, 8.2.10
abs pol , 10.5.7
cl U, 4.1.13
co(U), 3.1.14
codimX, 2.2.9 (5)
coimT, 2.3.1
coker T, 2.3.1
core U, 3.4.11
diamU, 4.5.3
dimX, 2.2.9 (5)
domf, 3.4.2
domF, 1.1.2
epi f, 3.4.2
ext V , 3.6.1
l B, 1.3.3
fr U, 4.1.13
imF, 1.1.2
inf U, 1.2.9
int U, 4.1.13
ker T, 2.3.1
lin(U), 3.1.14
pol , 10.5.7
rank T, 8.5.7 (2)
res(a), 11.2.1
res(T), 5.6.13
seg, 3.6.1
sup U, 1.2.9
supp(f), 9.6.4
supp(), 10.8.11, 10.9.4 (5)
supp(u), 10.10.5 (6)
a, 11.6.8
a, 10.8.15
a
f, 10.9.4 (1)
(a, b)
s
, 11.9.9
c, 3.3.1 (2), 5.5.9 (3)
c(E , F), 5.5.9 (3)
c
0
, 5.5.9 (3)
c
0
(E ), 5.5.9
c
0
(E , F), 5.5.9 (3)

u, 10.10.5 (4)
(p), 3.5.2 (1)
[[(p), 3.7.8
U, 4.1.13

x
(f), 3.5.1
d
p
, 5.2.1
dx, 10.9.9
e, 10.9.4 (1), 11.1.1

f, 10.11.3
f(a), 11.3.1
f < t, 3.8.1
f = t, 3.8.1
f t, 3.8.1
f(T), 8.2.1
f, 10.10.5 (9)
f, 10.9.4 (3)
f

, 10.9.4 (3)
fu, 10.10.5 (7)
f
n
f, 10.10.7 (3)
f
n

K
0, 10.9.8

g, 10.11.2

g(f), 8.2.6
h), 6.3.5
l
p
, l
p
(E ), 5.5.9 (4)
l

, l

(E ), 5.5.9 (2)
m, 5.5.9 (2)
p q, 5.3.3
p
e
, 5.5.9 (5)
p
S
, 3.8.6
p
T
, 5.1.3
p
X/X
0
, 5.1.10 (5)
r(T), 5.6.6
s, Ex. 1.19
t

, 10.11.5 (8)
u
g
, 10.10.5 (1)
u

, 10.10.5 (5)
u f, 10.10.5 (9)
u
1
u
2
, 10.10.5 (8)
u
1
u
2
, 10.10.5 (8)
x [, 10.3.1
260 Notation Index
x

, 6.4.1
x

, 5.1.10 (8)
x

, 10.11.5 (8)
x
+
, 3.2.12
x

, 3.2.12
[x[, 3.2.12
|x|
p
, 5.5.9 (4)
|x|

, 5.5.9 (2)

(x), 10.11.4

X
0
, 2.1.4 (6)
x :=

eE
x
e
, 5.5.9 (7)
x x

, 6.4.1
x
1
x
2
, x
1
x
2
, 1.2.12
(x
1
, x
2
), 1.2.12
x [ f), 5.1.11
x

y, 1.2.2
x

y, 5.5.6
x y, 6.2.5
[ y), 10.3.1
[[[y[[[
p
, 5.5.9 (6)
| |, 5.1.9
| |
n,Q
, 10.10.2 (2)
| |

, 5.5.9 (5)
| |
X
, 5.1.9
| [ X|, 5.1.9
[ ), 10.3.1
1, 5.3.10, 10.8.4 (6)
2
X
, 1.2.3 (4)
, 6.4.13
, 10.9.1
_
E
, 5.5.9 (6)
[ ), 10.3.1
[, 10.3.1
, 1.2.2

, 2.1.4 (5)

, 2.1.4 (4)
_
h(z)R(z)dz, 8.1.20
, 11.6.8
Subject Index
Absolute Bipolar Theorem, 10.5.9, 178
absolute concept, 9.4.7, 157
absolute polar, 10.5.7, 178
absolutely continuous measure,
10.9.4 (3), 190
absolutely convex set, 3.1.2 (6), 20
absolutely fundamental family
of vectors, 5.5.9 (7), 73
absorbing set, 3.4.9, 28
addition in a vector space, 2.1.3, 10
adherence of a lterbase, 9.4.1, 155
adherent point, 4.1.13, 41
adherent point of a lterbase, 9.4.1, 155
adjoint diagram, 6.4.8, 92
adjoint of an operator, 6.4.5, 91
adjunction of unity, 11.1.2, 213
ane hull, 3.1.14, 22
ane mapping, 3.1.7, 21
ane operator, 3.4.8 (4), 28
ane variety, 3.1.2 (5), 20
agreement condition, 10.9.4 (4), 190
Akilov Criterion, 10.5.3, 178
AlaogluBourbaki Theorem, 10.6.7, 180
Alexandro compactication,
9.4.22, 159
algebra, 5.6.2, 73
algebra of bounded operators, 5.6.5, 74
algebra of germs of holomorphic
functions, 8.1.18, 125
algebraic basis, 2.2.9 (5), 14
algebraic complement, 2.1.7, 12
algebraic dual, 2.2.4, 13
algebraic isomorphism, 2.2.5, 13
algebraic subdierential, 7.5.8, 113
algebraically complementary subspace,
2.1.7, 12
algebraically interior point, 3.4.11, 28
algebraically isomorphic spaces,
2.2.6, 13
algebraically reexive space, Ex. 2.8, 19
ambient space, 2.1.4 (3), 11
annihilator, 7.6.8, 116
antidiscrete topology, 9.1.8 (3), 147
antisymmetric relation, 1.2.1, 3
antitone mapping, 1.2.3, 4
approximate inverse, 8.5.9, 139
approximately invertible operator,
8.5.9, 139
approximation property, 8.3.10, 133
approximation property in Hilbert
space, 6.6.10, 98
arc, 4.8.2, 54
Arens multinorm, 8.3.8, 133
ascent, Ex. 8.10, 144
AscoliArzel`a Theorem, 4.6.10, 50
assignment operator, ix
associate seminorm, 6.1.7, 81
associated Hausdor pre-Hilbert
space, 6.1.10 (4), 83
associated Hilbert space,
6.1.10 (4), 83
associated multinormed space,
10.2.7, 172
associated topology, 9.1.12, 148
associativity of least upper bounds,
3.2.10, 24
asymmetric balanced HahnBanach
formula, 3.7.10, 34
asymmetric HahnBanach formula,
3.5.5, 30
Atkinson Theorem, 8.5.18, 141
262 Subject Index
Automatic Continuity Principle,
7.5.5, 112
automorphism, 10.11.4, 203
Baire Category Theorem, 4.7.6, 52
Baire space, 4.7.2, 52
Balanced HahnBanach Theorem,
3.7.13, 35
Balanced HahnBanach Theorem
in a topological setting, 7.5.10, 113
balanced set, 3.1.2 (7), 20
balanced subdierential, 3.7.8, 34
Balanced Subdierential Lemma,
3.7.9, 34
ball, 9.6.14, 166
Banach algebra, 5.6.3, 74
Banach Closed Graph Theorem,
7.4.7, 108
Banach Homomorphism Theorem,
7.4.4, 108
Banach Inversion Stability Theorem,
5.6.12, 76
Banach Isomorphism Theorem,
7.4.5, 108
Banach range, 7.4.18, 111
Banach space, 5.5.1, 66
Banachs Fundamental Principle,
7.1.5, 101
Banachs Fundamental Principle for
a Correspondence, 7.3.7, 107
BanachSteinhaus Theorem, 7.2.9, 104
barrel, 10.10.9 (1), 199
barreled normed space, 7.1.8, 102
barreled space, 10.10.9 (1), 199
base for a lter, 1.3.3, 6
basic eld, 2.1.2, 10
Bessel inequality, 6.3.7, 88
best approximation, 6.2.3, 84
BeurlingGelfand formula,
8.1.12 (2), 124
bilateral ideal, 8.3.3, 132; 11.6.2, 220
bilinear form, 6.1.2, 80
bipolar, 10.5.5, 178
Bipolar Theorem, 10.5.8, 178
Birkho Theorem, 9.2.2, 148
Bochner integral, 5.5.9 (6), 72
bornological space, 10.10.9 (3), 199
boundary of an algebra, Ex. 11.8, 234
boundary of a set, 4.1.13, 41
boundary point, 4.1.13, 41
bounded above, 1.2.19, 6
bounded below, 3.2.9, 23
bounded endomorphism algebra,
5.6.5, 74
Bounded Index Stability Theorem,
8.5.21, 142
bounded operator, 5.1.10 (7), 59
bounded Radon measure,
10.9.4 (2), 189
bounded set, 5.4.3, 66
boundedly order complete lattice,
3.2.8, 23
Bourbaki Criterion, 4.4.7, 46; 9.4.4, 156
bracketing of vector spaces, 10.3.1, 173
bra-functional, 10.3.1, 173
bra-mapping, 10.3.1, 173
bra-topology, 10.3.5, 174
B-stable, 10.1.8, 171
bump function, 9.6.19, 167
Calkin algebra, 8.3.5, 132
Calkin Theorem, 8.3.4, 132
canonical embedding, 5.1.10 (8), 59
canonical exact sequence,
2.3.5 (6), 16
canonical operator representation,
11.1.7, 214
canonical projection, 1.2.3 (4), 4
Cantor Criterion, 4.5.6, 47
Cantor Theorem, 4.4.9, 46
cap, 3.6.3 (4), 32
CauchyBunyakovskiSchwarz
inequality, 6.1.5, 80
Cauchy lter, 4.5.2, 47
Cauchy net, 4.5.2, 47
CauchyWiener Integral Theorem,
8.1.7, 122
centralizer, 11.1.6, 214
chain, 1.2.19, 6
character group, 10.11.2, 203
character of a group algebra,
10.11.1 (1), 201
character of an algebra, 11.6.4, 221
character space of an algebra,
11.6.4, 221
characteristic function, 5.5.9 (6), 72
charge, 10.9.4 (3), 190
Chebyshev metric, 4.6.8, 50
classical Banach space, 5.5.9 (5), 71
clopen part of a spectrum, 8.2.9, 130
closed ball, 4.1.3, 40
closed convex hull, 10.6.5, 179
Subject Index 263
closed correspondence, 7.3.8, 107
closed cylinder, 4.1.3, 40
closed-graph correspondence, 7.3.9, 107
closed halfspace, Ex. 3.3, 39
closed linear span, 10.5.6, 178
closed set, 9.1.4, 146
closed set in a metric space, 4.1.11, 41
closure of a set, 4.1.13, 41
closure operator, Ex. 1.11, 8
coarser cover, 9.6.1, 164
coarser lter, 1.3.6, 7
coarser pretopology, 9.1.2, 146
codimension, 2.2.9 (5), 14
codomain, 1.1.2, 1
conite set, Ex. 1.19, 9
coimage of an operator, 2.3.1, 15
coincidence of the algebraic and
topological subdierentials,
7.5.8, 113
coinitial set, 3.3.2, 25
cokernel of an operator, 2.3.1, 15
comeager set, 4.7.4, 52
commutative diagram, 2.3.3, 15
Commutative GelfandNamark
Theorem, 11.8.4, 227
compact convergence, 7.2.10, 105
Compact Index Stability Theorem,
8.5.20, 142
compact-open topology, 8.3.8, 133
compact operator, 6.6.1, 95
compact set, 9.4.2, 155
compact set in a metric space, 4.4.1, 46
compact space, 9.4.7, 157
compact topology, 9.4.7, 157
compactly-supported distribution,
10.10.5 (6), 196
compactly-supported function,
9.6.4, 165
compactum, 9.4.17, 158
compatible topology, 10.4.1, 175
complementary projection, 2.2.9 (4), 14
complementary subspace, 7.4.9, 109
Complementation Principle, 7.4.10, 109
complemented subspace, 7.4.9, 108
complement of an orthoprojection,
6.2.10, 85
complement of a projection,
2.2.9 (4), 14
complete lattice, 1.2.13, 5
complete metric space, 4.5.5, 47
complete set, 4.5.14, 49
completely regular space, 9.3.15, 155
completion, 4.5.13, 49
complex conjugate, 2.1.4 (2), 10
complex distribution, 10.10.5 (5), 196
complex plane, 8.1.3, 121
complex vector space, 2.1.3, 10
complexication, 8.4.8, 136
complexier, 3.7.4, 34
composite correspondence, 1.1.4, 2
Composite Function Theorem,
8.2.8, 129
composition, 1.1.4, 2
Composition Spectrum Theorem,
5.6.22, 78
cone, 3.1.2 (4), 20
conical hull, 3.1.14, 22
conical segment, 3.1.2 (9), 20
conical slice, 3.1.2 (9), 20
conjugate distribution, 10.10.5 (5), 196
conjugate exponent, 5.5.9 (4), 69
conjugate-linear functional, 2.2.4, 13
conjugate measure, 10.9.4 (3), 189
connected elementary compactum,
4.8.5, 54
connected set, 4.8.4, 54
constant function, 5.3.10, 64;
10.8.4 (6), 182
Continuous Extension Principle,
7.5.11, 113
continuous function at a point, 4.2.2,
43; 9.2.5, 149
Continuous Function Recovery
Lemma, 9.3.12, 153
continuous functional calculus,
11.8.7, 228
continuous mapping of a metric
space, 4.2.2, 43
continuous mapping of a topological
space, 9.2.4, 149
continuous partition of unity, 9.6.6, 166
contour integral, 8.1.20, 125
conventional summation, 5.5.9 (4), 70
convergent lterbase, 4.1.16, 42
convergent net, 4.1.17, 42
convergent sequence space, 3.3.1 (2), 25
convex combination, 3.1.14, 22
convex correspondence, 3.1.7, 21
convex function, 3.4.4, 27
convex hull, 3.1.14, 22
convex set, 3.1.2 (8), 20
convolution algebra, 10.9.4 (7), 190
264 Subject Index
convolution of a measure and
a function, 10.9.4 (7), 191
convolution of distributions,
10.10.5 (9), 197
convolution of functions, 9.6.17, 167
convolution of measures,
10.9.4 (7), 190
convolutive distributions,
10.10.5 (9), 197
coordinate projection, 2.2.9 (3), 13
coordinatewise operation,
2.1.4 (4), 11
core, 3.4.11, 28
correspondence, 1.1.1, 1
correspondence in two arguments,
1.1.3 (6), 2
correspondence onto, 1.1.3 (3), 2
coset, 1.2.3 (4), 4
coset mapping, 1.2.3 (4), 4
countable convex combination,
7.1.3, 101
Countable Partition Theorem,
9.6.20, 167
countable sequence, 1.2.16, 6
countably normable space, 5.4.1, 64
cover of a set, 9.6.1, 164
C

-algebra, 6.4.13, 92
C

-subalgebra, 11.7.8, 225


DavisFigielSzankowski
Counterexample, 8.3.14, 134
de Branges Lemma, 10.8.16, 185
decomplexication, 6.1.10 (2), 83
decomposition reduces an operator,
2.2.9 (4), 14
decreasing mapping, 1.2.3, 4
Dedekind complete vector lattice,
3.2.8, 23
deciency, 8.5.1, 137
denor, ix
delta-function, 10.9.4 (1), 188
delta-like sequence, 9.6.15, 166
-like sequence, 9.6.15, 166
-sequence, 9.6.15, 166
dense set, 4.5.10, 48
denseness, 4.5.10, 48
density of a measure, 10.9.4 (3), 190
derivative in the distribution sense,
10.10.5 (4), 196
derivative of a distribution, 10.10.5 (4),
196
descent, Ex. 8.10, 144
diagonal, 1.1.3 (3), 2
diagram prime, 7.6.5, 115
Diagram Prime Principle, 7.6.7, 115
diagram star, 6.4.8, 92
Diagram Star Principle, 6.4.9, 92
diameter, 4.5.3, 47
Dieudonn`e Lemma, 9.4.18, 158
dimension, 2.2.9 (5), 14
Dini Theorem, 7.2.10, 105
Dirac measure, 10.9.4 (1), 188
direct polar, 7.6.8, 116; 10.5.1, 177
direct sum decomposition, 2.1.7, 12
direct sum of vector spaces, 2.1.4 (5), 11
directed set, 1.2.15, 6
direction, 1.2.15, 6
directional derivative, 3.4.12, 28
discrete element, 3.3.6, 26
Discrete KrenRutman Theorem,
3.3.8, 26
discrete topology, 9.1.8 (4), 147
disjoint measures, 10.9.4 (3), 190
disjoint sets, 4.1.10, 41
distance, 4.1.1, 40
distribution, 10.10.4, 195
distribution applies to a function,
10.10.5 (7), 196
Distribution Localization Principle,
10.10.12, 200
distribution of nite order, 10.10.5 (3),
195
distribution size at most m,
10.10.5 (3), 195
distribution of slow growth,
10.11.16, 209
distributions admitting convolution,
10.10.5 (9), 197
distributions convolute, 10.10.5 (9), 197
division algebra, 11.2.3, 215
domain, 1.1.2, 1
Dominated Extension Theorem,
3.5.4, 30
Double Prime Lemma, 7.6.6, 115
double prime mapping, 5.1.10 (8), 59
double sharp, Ex. 2.7, 19
downward-ltered set, 1.2.15, 6
dual diagram, 7.6.5, 115
dual group, 10.11.2, 203
dual norm of a functional, 5.1.10 (8), 59
dual of a locally convex space,
10.2.11, 173
Subject Index 265
dual of an operator, 7.6.2, 114
duality bracket, 10.3.3, 174
duality pair, 10.3.3, 174
dualization, 10.3.3, 174
Dualization Theorem, 10.3.9, 175
DunfordHille Theorem, 8.1.3, 121
Dunford Theorem, 8.2.7 (2), 129
DvoretzkyRogers Theorem,
5.5.9 (7), 73
dyadic-rational point, 9.3.13, 154
eective domain of denition, 3.4.2, 27
Eidelheit Separation Theorem,
3.8.14, 39
eigenvalue, 6.6.3 (4), 95
eigenvector, 6.6.3, 95
element of a set, 1.1.3 (4), 2
elementary compactum, 4.8.5, 54
endomorphism, 2.2.1, 12; 8.2.1, 126
endomorphism algebra, 2.2.8, 13;
5.6.5, 74
endomorphism space, 2.2.8, 13
Eno counterexample, 8.3.12, 134
entourage, 4.1.5, 40
envelope, Ex. 1.11, 8
epigraph, 3.4.2, 26
epimorphism, 2.3.1, 15
-net, 8.3.2, 132
-perpendicular, 8.4.1, 134
-Perpendicular Lemma, 8.4.1, 134
Equicontinuity Principle, 7.2.4, 103
equicontinuous set, 4.2.8, 44
equivalence, 1.2.2, 3
equivalence class, 1.2.3 (4), 4
equivalent multinorms, 5.3.1, 62
equivalent seminorms, 5.3.3, 63
estimate for the diameter of a spherical
layer, 6.2.1, 84
Euler identity, 8.5.17, 141
evaluation mapping, 10.3.4 (3), 174
everywhere-dened operator, 2.2.1, 12
everywhere dense set, 4.7.3 (3), 52
exact sequence, 2.3.4, 15
exact sequence at a term, 2.3.4, 15
exclave, 8.2.9, 130
expanding mapping, Ex. 4.14, 55
extended function, 3.4.2, 26
extended real axis, 3.8.1, 35
extended reals, 3.8.1, 35
extension of an operator, 2.3.6, 16
exterior of a set, 4.1.13, 41
exterior point, 4.1.13, 41
Extreme and Discrete Lemma, 3.6.4, 32
extreme point, 3.6.1, 31
extreme set, 3.6.1, 31
face, 3.6.1, 31
factor set, 1.2.3 (4), 4
faithful representation, 8.2.2, 126
family, 1.1.3 (4), 2
lter, 1.3.3, 6
lterbase, 1.3.1, 6
ner cover, 9.6.1, 164
ner lter, 1.3.6, 7
ner multinorm, 5.3.1, 62
ner pretopology, 9.1.2, 146
ner seminorm, 5.3.3, 63
nest multinorm, 5.1.10 (2), 58
nite complement lter, 5.5.9 (3), 68
nite descent, Ex. 8.10, 144
nite-rank operator, 6.6.8, 97; 8.3.6, 132
nite-valued function, 5.5.9 (6), 72
rst category set, 4.7.1, 52
rst element, 1.2.6, 5
xed point, Ex. 1.11, 8
at, 3.1.2 (5), 20
formal duality, 2.3.15, 18
Fourier coecient family, 6.3.15, 89
FourierPlancherel transform,
10.11.15, 209
FourierSchwartz transform,
10.11.19, 211
Fourier series, 6.3.16, 89
Fourier transform of a distribution,
10.11.19, 211
Fourier transform of a function,
10.11.3, 203
Fourier transform relative to a basis,
6.3.15, 89
Frechet space, 5.5.2, 66
Fredholm Alternative, 8.5.6, 138
Fredholm index, 8.5.1, 137
Fredholm operator, 8.5.1, 137
Fredholm Theorem, 8.5.8, 139
frontier of a set, 4.1.13, 41
from A into/to B, 1.1.1, 1
Fubini Theorem for distributions,
10.10.5 (8), 197
Fubini Theorem for measures,
10.9.4 (6), 190
full subalgebra, 11.1.5, 213
fully norming set, 8.1.1, 120
266 Subject Index
Function Comparison Lemma, 3.8.3, 36
function of class C
(m)
, 10.9.9, 192
function of compact support, 9.6.4, 165
Function Recovery Lemma, 3.8.2, 35
functor, 10.9.4 (4), 190
fundamental net, 4.5.2, 47
fundamental sequence, 4.5.2, 47
fundamentally summable family
of vectors, 5.5.9 (7), 73
gauge, 3.8.6, 37
gauge function, 3.8.6, 37
Gauge Theorem, 3.8.7, 37
-correspondence, 3.1.6, 21
-hull, 3.1.11, 21
-set, 3.1.1, 20
GelfandDunford Theorem in
an operator setting, 8.2.3, 127
GelfandDunford Theorem
in an algebraic setting, 11.3.2, 216
Gelfand formula, 5.6.8, 74
GelfandMazur Theorem, 11.2.3, 215
GelfandNamarkSegal construction,
11.9.11, 232
Gelfand Theorem, 7.2.2, 103
Gelfand transform of an algebra,
11.6.8, 222
Gelfand transform of an element,
11.6.8, 221
Gelfand Transform Theorem,
11.6.9, 222
general form of a compact operator
in Hilbert space, 6.6.9, 97
general form of a linear functional
in Hilbert space, 6.4.2, 90
general form of a weakly continuous
functional, 10.3.10, 175
general position, Ex. 3.10, 39
generalized derivative in the Sobolev
sense, 10.10.5 (4), 196
Generalized Dini Theorem, 10.8.6, 183
generalized function, 10.10.4, 195
Generalized RieszSchauder Theorem,
8.4.10, 137
generalized sequence, 1.2.16, 6
Generalized Weierstrass Theorem,
10.9.9, 192
germ, 8.1.14, 124
GNS-construction, 11.9.11, 232
GNS-Construction Theorem,
11.9.10, 231
gradient mapping, 6.4.2, 90
GramSchmidt orthogonalization
process, 6.3.14, 89
graph norm, 7.4.17, 111
Graph Norm Principle, 7.4.17, 111
greatest element, 1.2.6, 5
greatest lower bound, 1.2.9, 5
Grothendieck Criterion, 8.3.11, 133
Grothendieck Theorem, 8.3.9, 133
ground eld, 2.1.3, 10
ground ring, 2.1.1, 10
group algebra, 10.9.4 (7), 191
group character, 10.11.1, 202
Haar integral, 10.9.4 (1), 189
HahnBanach Theorem, 3.5.3, 29
HahnBanach Theorem in analytical
form, 3.5.4, 30
HahnBanach Theorem in geometric
form, 3.8.12, 38
HahnBanach Theorem in subdierential
form, 3.5.4, 30
Hamel basis, 2.2.9 (5), 14
Hausdor Completion Theorem,
4.5.12, 48
Hausdor Criterion, 4.6.7, 50
Hausdor metric, Ex. 4.8, 55
Hausdor multinorm, 5.1.8, 57
Hausdor multinormed space, 5.1.8, 57
Hausdor space, 9.3.5, 152
Hausdor Theorem, 7.6.12, 117
Hausdor topology, 9.3.5, 152
H-closed space, Ex. 9.10, 168
Heaviside function, 10.10.5 (4), 196
HellingerToeplitz Theorem, 6.5.3, 93
hermitian element, 11.7.1, 224
hermitian form, 6.1.1, 80
hermitian operator, 6.5.1, 93
hermitian state, 11.9.8, 230
Hilbert basis, 6.3.8, 88
Hilbert cube, 9.2.17 (2), 151
Hilbert dimension, 6.3.13, 89
Hilbert identity, 5.6.19, 78
Hilbert isomorphy, 6.3.17, 90
HilbertSchmidt norm, Ex. 8.9, 144
HilbertSchmidt operator,
Ex. 8.9, 144
HilbertSchmidt Theorem, 6.6.7, 96
Hilbert space, 6.1.7, 81
Hilbert-space isomorphism, 6.3.17, 90
Hilbert sum, 6.1.10 (5), 84
Subject Index 267
Holder inequality, 5.5.9 (4), 69
holey disk, 4.8.5, 54
holomorphic function, 8.1.4, 122
Holomorphy Theorem, 8.1.5, 122
homeomorphism, 9.2.4, 149
homomorphism, 7.4.1, 107
Hormander transform, Ex. 3.19, 39
hyperplane, 3.8.9, 38
hypersubspace, 3.8.9, 38
ideal, 11.4.1, 217
Ideal and Character Theorem,
11.6.6, 221
ideal correspondence, 7.3.3, 106
Ideal Correspondence Lemma,
7.3.4, 106
Ideal Correspondence Principle,
7.3.5, 106
Ideal HahnBanach Theorem, 7.5.9, 113
ideally convex function, 7.5.4, 112
ideally convex set, 7.1.3, 101
idempotent operator, 2.2.9 (4), 14
identical embedding, 1.1.3 (3), 2
identity, 10.9.4, 188
identity element, 11.1.1, 213
identity mapping, 1.1.3 (3), 2
identity relation, 1.1.3 (3), 1
image, 1.1.2, 1
image of a lterbase, 1.3.5 (1), 7
image of a set, 1.1.3 (5), 2
image of a topology, 9.2.12, 150
image topology, 9.2.12, 150
Image Topology Theorem,
9.2.11, 150
imaginary part of a function,
5.5.9 (4), 69
increasing mapping, 1.2.3 (5), 4
independent measure, 10.9.4 (3), 190
index, 8.5.1, 137
indicator function, 3.4.8 (2), 27
indiscrete topology, 9.1.8 (3), 147
induced relation, 1.2.3 (1), 4
induced topology, 9.2.17 (1), 151
inductive limit topology, 10.9.6, 191
inductive set, 1.2.19, 6
inmum, 1.2.9, 5
innite-rank operator, 6.6.8, 97
innite set, 5.5.9 (3), 68
inner product, 6.1.4, 80
integrable function, 5.5.9 (4), 69
integral, 5.5.9 (4), 68
integral with respect to a measure,
10.9.3, 188
interior of a set, 4.1.13, 41
interior point, 4.1.13, 41
intersection of topologies, 9.1.14, 148
interval, 3.2.15, 24
Interval Addition Lemma, 3.2.15, 24
invariant subspace, 2.2.9 (4), 14
inverse-closed subalgebra, 11.1.5, 213
inverse image of a multinorm,
5.1.10 (3), 58
inverse image of a preorder, 1.2.3 (3), 4
inverse image of a seminorm, 5.1.4, 57
inverse image of a set, 1.1.3 (5), 2
inverse image of a topology, 9.2.9, 150
inverse image of a uniformity,
9.5.5 (3), 160
inverse image topology, 9.2.9, 150
Inverse Image Topology Theorem,
9.2.8, 149
inverse of a correspondence, 1.1.3 (1), 1
inverse of an element in an algebra,
11.1.5, 213
Inversion Theorem, 10.11.12, 208
invertible element, 11.1.5, 213
invertible operator, 5.6.10, 75
involution, 6.4.13, 92
involutive algebra, 6.4.13, 92
irreducible representation, 8.2.2, 127
irreexive space, 5.1.10 (8), 59
isolated part of a spectrum, 8.2.9, 130
isolated point, 8.4.7, 136
isometric embedding, 4.5.11, 48
isometric isomorphism of algebras,
11.1.8, 215
isometric mapping, 4.5.11, 48
isometric representation, 11.1.8, 214
isometric -isomorphism, 11.8.3, 226
isometric -representation, 11.8.3, 226
isometry into, 4.5.11, 48
isometry onto, 4.5.11, 48
isomorphism, 2.2.5, 13
isotone mapping, 1.2.3, 4
James Theorem, 10.7.5, 181
Jensen inequality, 3.4.5, 27
join, 1.2.12, 5
Jordan arc, 4.8.2, 54
Jordan Curve Theorem, 4.8.3, 54
juxtaposition, 2.2.8, 13
268 Subject Index
Kakutani Criterion, 10.7.1, 180
Kakutani Lemma, 10.8.7, 183
Kakutani Theorem, 7.4.11 (3), 109
Kantorovich space, 3.2.8, 23
Kantorovich Theorem, 3.3.4, 25
KaplanskyFukamija Lemma,
11.9.7, 230
Kato Criterion, 7.4.19, 111
kernel of an operator, 2.3.1, 15
ket-mapping, 10.3.1, 173
ket-topology, 10.3.5, 174
Kolmogorov Normability Criterion,
5.4.5, 66
KrenMilman Theorem, 10.6.5, 179
KrenMilman Theorem in subdierential
form, 3.6.5, 33
KrenRutman Theorem, 3.3.5, 26
Krull Theorem, 11.4.8, 219
KuratowskiZorn Lemma, 1.2.20, 6
K-space, 3.2.8, 23
K-ultrametric, 9.5.13, 162
last element, 1.2.6, 5
lattice, 1.2.12, 5
lear trap map, 3.7.4, 34
least element, 1.2.6, 5
Lebesgue measure, 10.9.4 (1), 189
Lebesgue set, 3.8.1, 35
Lefschetz Lemma, 9.6.3, 165
left approximate inverse, 8.5.9, 139
left Haar measure, 10.9.4 (1), 189
left inverse of an element in an algebra,
11.1.3, 213
lemma on continuity of a convex
function, 7.5.1, 112
lemma on the numeric range
of a hermitian element, 11.9.3, 229
level set, 3.8.1, 35
Levy Projection Theorem, 6.2.2, 84
limit of a lterbase, 4.1.16, 42
Lindenstrauss space, 5.5.9 (5), 71
LindenstraussTzafriri Theorem,
7.4.11 (3), 110
linear change of a variable under the
subdierential sign, 3.5.4, 30
linear combination, 2.3.12, 17
linear correspondence, 2.2.1, 12;
3.1.7, 21
linear functional, 2.2.4, 13
linear operator, 2.2.1, 12
linear representation, 8.2.2, 126
linear set, 2.1.4 (3), 11
linear space, 2.1.4 (3), 11
linear span, 3.1.14, 22
linear topological space, 10.1.3, 169
linear topology, 10.1.3, 169
linearly independent set, 2.2.9 (5), 14
linearly-ordered set, 1.2.19, 6
Lions Theorem of Supports,
10.10.5 (9), 197
Liouville Theorem, 8.1.10, 123
local data, 10.9.11, 193
locally compact group, 10.9.4 (1), 188
locally compact space, 9.4.20, 159
locally compact topology, 9.4.20, 159
locally convex space, 10.2.9, 172
locally convex topology, 10.2.1, 171
locally nite cover, 9.6.2, 164
locally integrable function, 9.6.17, 167
locally Lipschitz function, 7.5.6, 112
loop, 4.8.2, 54
lower bound, 1.2.4, 5
lower limit, 4.3.5, 45
lower right Dini derivative, 4.7.7, 53
lower semicontinuous, 4.3.3, 44
L
2
-Fourier transform, 10.11.15, 209
MackeyArens Theorem, 10.4.5, 176
Mackey Theorem, 10.4.6, 176
Mackey topology, 10.4.4, 176
mapping, 1.1.3 (3), 1
massive subspace, 3.3.2, 25
matrix form, 2.2.9 (4), 14
maximal element, 1.2.10, 5
maximal ideal, 11.4.5, 218
maximal ideal space, 11.6.7, 221
Maximal Ideal Theorem, 11.5.3, 220
Mazur Theorem, 10.4.9, 177
meager set, 4.7.1, 52
measure, 10.9.3, 188
Measure Localization Principle,
10.9.10, 192
measure space, 5.5.9 (4), 69
meet, 1.2.12, 5
member of a set, 1.1.3 (4), 2
metric, 4.1.1, 40
metric space, 4.1.1, 40
metric topology, 4.1.9, 41
metric uniformity, 4.1.5, 40
Metrizability Criterion, 5.4.2, 64
metrizable multinormed space, 5.4.1, 64
minimal element, 1.2.10, 5
Subject Index 269
Minimal Ideal Theorem, 11.5.1, 219
MinkowskiAscoliMazur Theorem,
3.8.12, 38
Minkowski functional, 3.8.6, 37
Minkowski inequality, 5.5.9 (4), 69
minorizing set, 3.3.2, 25
mirror, 10.2.7, 172
module, 2.1.1, 10
modulus of a scalar, 5.1.10 (4), 58
modulus of a vector, 3.2.12, 24
mollier, 9.6.14, 166
mollifying kernel, 9.6.14, 166
monomorphism, 2.3.1, 15
monoquotient, 2.3.11, 17
Montel space, 10.10.9 (2), 199
Moore subnet, 1.3.5 (2), 7
morphism, 8.2.2, 126; 11.1.2, 213
morphism representing an algebra,
8.2.2, 126
Motzkin formula, 3.1.13 (5), 22
multimetric, 9.5.9, 161
multimetric space, 9.5.9, 161
multimetric uniformity, 9.5.9, 161
multimetrizable topological space,
9.5.10, 161
multimetrizable uniform space,
9.5.10, 161
multinorm, 5.1.6, 57
Multinorm Comparison Theorem,
5.3.2, 62
multinorm summable family of vectors,
5.5.9 (7), 73
multinormed space, 5.1.6, 57
multiplication formula, 10.11.5, 205
multiplication of a germ by a complex
number, 8.1.16, 125
multiplicative linear operator, 8.2.2, 126
natural order, 3.2.6 (1), 23
negative part, 3.2.12, 24
neighborhood about a point,
9.1.1 (2), 146
neighborhood about a point in a metric
space, 4.1.9, 41
neighborhood lter, 4.1.10, 41
neighborhood lter of a set, 9.3.7, 152
neighborhood of a set, 8.1.13 (2),
124; 9.3.7, 152
Nested Ball Theorem, 4.5.7, 47
nested sequence, 4.5.7, 47
net, 1.2.16, 6
net having a subnet, 1.3.5 (2), 7
net lacking a subnet, 1.3.5 (2), 7
Neumann series, 5.6.9, 75
Neumann Series Expansion Theorem,
5.6.9, 75
neutral element, 2.1.4 (3), 11;
10.9.4, 188
Nikol

ski Criterion, 8.5.22, 143


Noether Criterion, 8.5.14, 140
nonarchimedean element,
5.5.9 (5), 70
nonconvex cone, 3.1.2 (4), 20
Nonempty Subdierential Theorem,
3.5.8, 31
non-everywhere-dened operator,
2.2.1, 12
nonmeager set, 4.7.1, 52
nonpointed cone, 3.1.2 (4), 20
nonreexive space, 5.1.10 (8), 59
norm, 5.1.9, 57
norm convergence, 5.5.9 (7), 73
normable multinormed space, 5.4.1, 64
normal element, 11.7.1, 224
normal operator, Ex. 8.17, 145
normal space, 9.3.11, 153
normalized element, 6.3.5, 88
normally solvable operator, 7.6.9, 116
normative inequality, 5.1.10 (7), 59
normed algebra, 5.6.3, 74
normed dual, 5.1.10 (8), 59
normed space, 5.1.9, 57
normed space of bounded elements,
5.5.9 (5), 70
norming set, 8.1.1, 120
norm-one element, 5.5.6, 68
nowhere dense set, 4.7.1, 52
nullity, 8.5.1, 137
numeric family, 1.1.3 (4), 2
numeric function, 9.6.4, 165
numeric range, 11.9.1, 229
numeric set, 1.1.3 (4), 2
one-point compactication, 9.4.22, 159
one-to-one correspondence, 1.1.3 (3), 2
open ball, 4.1.3, 40
open ball of R
N
, 9.6.16, 166
open correspondence, 7.3.12, 107
Open Correspondence Principle,
7.3.13, 107
open cylinder, 4.1.3, 40
open halfspace, Ex. 3.3, 39
270 Subject Index
Open Mapping Theorem, 7.4.6, 108
open segment, 3.6.1, 31
open set, 9.1.4, 146
open set in a metric space, 4.1.11, 41
openness at a point, 7.3.6, 107
operator, 2.2.1, 12
operator ideal, 8.3.3, 132
operator norm, 5.1.10 (7), 58
operator representation, 8.2.2, 126
order, 1.2.2, 3
order by inclusion, 1.3.1, 6
order compatible with vector structure,
3.2.1, 22
order ideal, 10.8.11, 183
order of a distribution, 10.10.5 (3), 195
ordered set, 1.2.2, 3
ordered vector space, 3.2.1, 22
ordering, 1.2.2, 3
ordering cone, 3.2.4, 23
oriented envelope, 4.8.8, 54
orthocomplement, 6.2.5, 85
orthogonal complement, 6.2.5, 85
orthogonal family, 6.3.1, 87
orthogonal orthoprojections, 6.2.12, 86
orthogonal set, 6.3.1, 87
orthogonal vectors, 6.2.5, 85
orthonormal family, 6.3.6, 88
orthonormal set, 6.3.6, 88
orthonormalized family, 6.3.6, 88
orthoprojection, 6.2.7, 85
Orthoprojection Summation Theorem,
6.3.3, 87
Orthoprojection Theorem, 6.2.10, 85
Osgood Theorem, 4.7.5, 52
pair-dual space, 10.3.3, 174
pairing, 10.3.3, 174
pairwise orthogonality of nitely many
orthoprojections, 6.2.14, 86
paracompact space, 9.6.9, 166
Parallelogram Law, 6.1.8, 81
Parseval identity, 6.3.16, 89;
10.11.12, 208
part of an operator, 2.2.9 (4), 14
partial correspondence, 1.1.3 (6), 2
partial operator, 2.2.1, 12
partial order, 1.2.2, 3
partial sum, 5.5.9 (7), 73
partition of unity, 9.6.6, 165
partition of unity subordinate
to a cover, 9.6.7, 166
patch, 10.9.11, 193
perforated disk, 4.8.5, 54
periodic distribution, 10.11.17 (7), 211
Pettis Theorem, 10.7.4, 181
Phillips Theorem, 7.4.13, 110
Plancherel Theorem, 10.11.14, 209
point nite cover, 9.6.2, 164
point in a metric space, 4.1.1, 40
point in a space, 2.1.4 (3), 11
point in a vector space, 2.1.3, 10
pointwise convergence, 9.5.5 (6), 161
pointwise operation, 2.1.4 (4), 11
polar, 7.6.8, 116; 10.5.1, 177
Polar Lemma, 7.6.11, 116
polarization identity, 6.1.3, 80
Pontryaginvan Kampen Duality
Theorem, 10.11.2, 203
poset, 1.2.2, 3
positive cone, 3.2.5, 23
positive denite inner product, 6.1.4, 80
positive distribution, 10.10.5 (2), 195
positive element of a C

-algebra,
11.9.4, 230
positive form on a C

-algebra,
Ex. 11.11, 235
positive hermitian form, 6.1.4, 80
positive matrix, Ex. 3.13, 39
positive operator, 3.2.6 (3), 23
positive part, 3.2.12, 24
positive semidenite hermitian
form, 6.1.4, 80
positively homogeneous functional,
3.4.7 (2), 27
powerset, 1.2.3 (4), 4
precompact set, Ex. 9.16, 168
pre-Hilbert space, 6.1.7, 81
preimage of a multinorm,
5.1.10 (3), 58
preimage of a seminorm, 5.1.4, 57
preimage of a set, 1.1.3 (5), 2
preintegral, 5.5.9 (4), 68
preneighborhood, 9.1.1 (2), 146
preorder, 1.2.2, 3
preordered set, 1.2.2, 4
preordered vector space, 3.2.1, 22
presheaf, 10.9.4 (4), 190
pretopological space, 9.1.1 (2), 146
pretopology, 9.1.1, 146
primary Banach space, Ex. 7.17, 119
prime mapping, 6.4.1, 90
Prime Theorem, 10.2.13, 173
Subject Index 271
Principal Theorem of the Holomorphic
Functional Calculus, 8.2.4, 128
product, 4.3.2, 44
product of a distribution and
a function, 10.10.5 (7), 196
product of germs, 8.1.16, 125
product of sets, 1.1.1, 1; 2.1.4 (4), 11
product of topologies, 9.2.17 (2), 151
product of vector spaces, 2.1.4 (4), 11
product topology, 4.3.2, 44; 9.2.17 (2),
151
projection onto X
1
along X
2
,
2.2.9 (4), 14
projection to a set, 6.2.3, 84
proper ideal, 11.4.5, 218
pseudometric, 9.5.7, 161
p-sum, 5.5.9 (6), 71
p-summable family, 5.5.9 (4), 70
punctured compactum, 9.4.21, 159
pure subalgebra, 11.1.5, 213
Pythagoras Lemma, 6.2.8, 85
Pythagoras Theorem, 6.3.2, 87
quasinilpotent, Ex. 8.18, 145
quotient mapping, 1.2.3 (4), 4
quotient multinorm, 5.3.11, 64
quotient of a mapping, 1.2.3 (4), 4
quotient of a seminormed space,
5.1.10 (5), 58
quotient seminorm, 5.1.10 (5), 58
quotient set, 1.2.3 (4), 4
quotient space of a multinormed
space, 5.3.11, 64
quotient vector space, 2.1.4 (6), 12
radical, 11.6.11, 223
Radon F-measure, 10.9.3, 188
RadonNikod ym Theorem, 10.9.4 (3),
190
range of a correspondence, 1.1.2, 1
rank, 8.5.7 (2), 139
rare set, 4.7.1, 52
Rayleigh Theorem, 6.5.2, 93
real axis, 2.1.2, 10
real carrier, 3.7.1, 33
real C-measure, 10.9.4 (3), 189
real distribution, 10.10.5 (5), 196
real hyperplane, 3.8.9, 38
real measure, 10.9.4, 189
real part map, 3.7.2, 33
real part of a function, 5.5.9 (4), 69
real part of a number, 2.1.2, 10
real subspace, 3.1.2 (3), 20
real vector space, 2.1.3, 10
realication, 3.7.1, 33
realication of a pre-Hilbert space,
6.1.10 (2), 83
realier, 3.7.2, 33
reducible representation, 8.2.2, 127
renement, 9.6.1, 164
reection of a function, 10.10.5, 197
reexive relation, 1.2.1, 3
reexive space, 5.1.10 (8), 59
regular distribution, 10.10.5 (1), 195
regular operator, 3.2.6 (3), 23
regular space, 9.3.9, 153
regular value of an operator, 5.6.13, 76
relation, 1.1.3 (2), 1
relative topology, 9.2.17 (1), 151
relatively compact set, 4.4.4, 46
removable singularity, 8.2.5 (2), 128
representation, 8.2.2, 126
representation space, 8.2.2, 126
reproducing cone, Ex. 7.12, 119
residual set, 4.7.4, 52
resolvent of an element of an algebra,
11.2.1, 215
resolvent of an operator, 5.6.13, 76
resolvent set of an operator, 5.6.13, 76
resolvent value of an element
of an algebra, 11.2.1, 215
resolvent value of an operator,
5.6.13, 76
restriction, 1.1.3 (5), 2
restriction of a distribution,
10.10.5 (6), 196
restriction of a measure,
10.9.4 (4), 190
restriction operator, 10.9.4 (4), 190
reversal, 1.2.5, 5
reverse order, 1.2.3 (2), 4
reverse polar, 7.6.8, 116; 10.5.1, 177
reversed multiplication, 11.1.6, 214
Riemann function, 4.7.7, 53
RiemannLebesgue Lemma,
10.11.5 (3), 204
Riemann Theorem on Series,
5.5.9 (7), 73
Riesz Criterion, 8.4.2, 134
Riesz Decomposition Property,
3.2.16, 24
RieszDunford integral, 8.2.1, 126
272 Subject Index
RieszDunford Integral Decomposition
Theorem, 8.2.13, 131
RieszDunford integral in an algebraic
setting, 11.3.1, 216
RieszFisher Completeness Theorem,
5.5.9 (4), 70
RieszFisher Isomorphism Theorem,
6.3.16, 89
Riesz idempotent, 8.2.11, 130
RieszKantorovich Theorem, 3.2.17, 24
Riesz operator, Ex. 8.15, 145
Riesz Prime Theorem, 6.4.1, 90
Riesz projection, 8.2.11, 130
RieszSchauder operator,
Ex. 8.11, 144
RieszSchauder Theorem, 8.4.8, 136
Riesz space, 3.2.7, 23
Riesz Theorem, 5.3.5, 63
right approximate inverse, 8.5.9, 139
right Haar measure, 10.9.4 (1), 189
right inverse of an element
in an algebra, 11.1.3, 213
R-measure, 10.9.4 (3), 189
rough draft, 4.8.8, 54
row-by-column rule, 2.2.9 (4), 14
salient cone, 3.2.4, 23
Sard Theorem, 7.4.12, 110
scalar, 2.1.3, 10
scalar eld, 2.1.3, 10
scalar multiplication, 2.1.3, 10
scalar product, 6.1.4, 80
scalar-valued function, 9.6.4, 165
Schauder Theorem, 8.4.6, 135
Schwartz space of distributions,
10.11.16, 209
Schwartz space of functions,
10.11.6, 206
Schwartz Theorem, 10.10.10, 199
second dual, 5.1.10 (8), 59
selfadjoint operator, 6.5.1, 93
semi-extended real axis, 3.4.1, 26
semi-Fredholm operator,
Ex. 8.13, 145
semi-inner product, 6.1.4, 80
semimetric, 9.5.7, 161
semimetric space, 9.5.7, 161
seminorm, 3.7.6, 34
seminorm associated with a positive
element, 5.5.9 (5), 70
seminormable space, 5.4.6, 66
seminormed space, 5.1.5, 57
semisimple algebra, 11.6.11, 223
separable space, 6.3.14, 89
separated multinorm, 5.1.8, 57
separated multinormed space, 5.1.8, 57
separated topological space, 9.3.2, 151
separated topology, 9.3.2, 151
separating hyperplane, 3.8.13, 39
Separation Theorem, 3.8.11, 38
Sequence Prime Principle, 7.6.13, 117
sequence space, 3.3.1 (2), 25
Sequence Star Principle, 6.4.12, 92
series sum, 5.5.9 (7), 73
sesquilinear form, 6.1.2, 80
set absorbing another set, 3.4.9, 28
set in a space, 2.1.4 (3), 11
set lacking a distribution,
10.10.5 (6), 196
set lacking a functional, 10.8.13, 184
set lacking a measure, 10.9.4 (5), 190
set of arrival, 1.1.1, 1
set of departure, 1.1.1, 1
set of second category, 4.7.1, 52
set supporting a measure,
10.9.4 (5), 190
set that separates the points of another
set, 10.8.9, 183
set void of a distribution,
10.10.5 (6), 196
set void of a functional, 10.8.13, 184
set void of a measure, 10.9.4 (5), 190
setting in duality, 10.3.3, 174
setting primes, 7.6.5, 115
sheaf, 10.9.11, 193
shift, 10.9.4 (1), 189
Shilov boundary, Ex. 11.8, 234
Shilov Theorem, 11.2.4, 215
short sequence, 2.3.5, 16
-compact, 10.9.8, 192
signed measure, 10.9.4 (3), 190
simple convergence, 9.5.5 (6), 161
simple function, 5.5.9 (6), 72
simple Jordan loop, 4.8.2, 54
single-valued correspondence,
1.1.3 (3), 1
Singularity Condensation Principle,
7.2.12, 105
Singularity Fixation Principle,
7.2.11, 105
skew eld, 11.2.3, 215
Subject Index 273
slowly increasing distribution,
10.11.16, 209
smooth function, 9.6.13, 166
smoothing process, 9.6.18, 167
Snowake Lemma, 2.3.16, 18
space countable at innity, 10.9.8, 192
space of bounded elements, 5.5.9 (5), 70
space of bounded functions, 5.5.9 (2), 68
space of bounded operators,
5.1.10 (7), 59
space of compactly-supported
distributions, 10.10.5 (9), 197
space of convergent sequences,
5.5.9 (3), 68
space of distributions of order at most
m, 10.10.8, 199
space of essentially bounded functions,
5.5.9 (5), 70
space of nite-order distributions,
10.10.8, 199
space of functions vanishing at innity,
5.5.9 (3), 68
space of X-valued p-summable
functions, 5.5.9 (6), 72
space of p-summable functions,
5.5.9 (4), 69
space of p-summable sequences,
5.5.9 (4), 70
space of tempered distributions,
10.11.16, 209
space of vanishing sequences,
5.5.9 (3), 68
Spectral Decomposition Lemma,
6.6.6, 96
Spectral Decomposition Theorem,
8.2.12, 130
Spectral Endpoint Theorem, 6.5.5, 94
Spectral Mapping Theorem, 8.2.5, 128
Spectral Purity Theorem,
11.7.11, 226
spectral radius of an operator, 5.6.6, 74
Spectral Theorem, 11.8.6, 227
spectral value of an element
of an algebra, 11.2.1, 215
spectral value of an operator, 5.6.13, 76
spectrum, 10.2.7, 172
spectrum of an element of an algebra,
11.2.1, 215
spectrum of an operator, 5.6.13, 76
spherical layer, 6.2.1, 84
-algebra, 6.4.13, 92
-isomorphism, 11.8.3, 226
-linear functional, 2.2.4, 13
-representation, 11.8.3, 226
star-shaped set, 3.1.2 (7), 20
state, 11.9.1, 229
Steklov condition, 6.3.10, 88
Steklov Theorem, 6.3.11, 88
step function, 5.5.9 (6), 72
Stone Theorem, 10.8.10, 183
StoneWeierstrass Theorem for
C(Q, C), 11.8.2, 226
StoneWeierstrass Theorem for
C(Q, R), 10.8.17, 186
Strict Separation Theorem, 10.4.8, 177
strict subnet, 1.3.5 (2), 7
strictly positive real, 4.1.3, 40
strong order-unit, 5.5.9 (5), 70
strong uniformity, 9.5.5 (6), 161
stronger multinorm, 5.3.1, 62
stronger pretopology, 9.1.2, 146
stronger seminorm, 5.3.3, 63
strongly holomorphic function,
8.1.5, 122
structure of a subdierential,
10.6.3, 179
subadditive functional, 3.4.7 (4), 27
subcover, 9.6.1, 164
subdierential, 3.5.1, 29
sublattice, 10.8.2, 181
sublinear functional, 3.4.6, 27
submultiplicative norm, 5.6.1, 73
subnet, 1.3.5 (2), 7
subnet in a broad sense, 1.3.5 (2), 7
subrepresentation, 8.2.2, 126
subspace of a metric space, 4.5.14, 49
subspace of a topological space,
9.2.17 (1), 151
subspace of an ordered vector space,
3.2.6 (2), 23
subspace topology, 9.2.17 (1), 151
SukhomlinovBohnenblustSobczyk
Theorem, 3.7.12, 35
sum of a family in the sense of L
p
,
5.5.9 (6), 71
sum of germs, 8.1.16, 125
summable family of vectors,
5.5.9 (7), 73
summable function, 5.5.9 (4), 69
superset, 1.3.3, 6
sup-norm, 10.8.1, 181
support function, 10.6.4, 179
274 Subject Index
support of a distribution,
10.10.5 (6), 196
support of a functional, 10.8.12, 184
support of a measure, 10.9.4 (5), 190
supporting function, 10.6.4, 179
supremum, 1.2.9, 5
symmetric HahnBanach formula,
Ex. 3.10, 39
symmetric relation, 1.2.1, 3
symmetric set, 3.1.2 (7), 20
system with integration, 5.5.9 (4), 68
Szankowski Counterexample,
8.3.13, 134
tail lter, 1.3.5 (2), 7
-dual of a locally convex space,
10.2.11, 173
Taylor Series Expansion Theorem,
8.1.9, 123
tempered distribution, 10.11.16, 209
tempered function, 5.1.10 (6), 58;
10.11.6, 206
tempered Radon measure, 10.11.17 (3),
210
test function, 10.10.1, 194
test function space, 10.10.1, 194
theorem on Hilbert isomorphy,
6.3.17, 90
theorem on the equation AX = B,
2.3.13, 17
theorem on the equation XA = B,
2.3.8, 16
theorem on the general form
of a distribution, 10.10.14, 201
theorem on the inverse image
of a vector topology, 10.1.6, 171
theorem on the repeated Fourier
transform, 10.11.13, 209
theorem on the structure of a locally
convex topology, 10.2.2, 171
theorem on the structure of a vector
topology, 10.1.4, 170
theorem on topologizing by a family
of mappings, 9.2.16, 151
there is a unique x, 2.3.9, 17
TietzeUrysohn Theorem,
10.8.20, 186
topological isomorphism, 9.2.4, 149
topological mapping, 9.2.4, 149
Topological Separation Theorem,
7.5.12, 114
topological space, 9.1.7, 147
topological structure of a convex
set, 7.1.1, 100
topological subdierential, 7.5.8, 113
topological vector space, 10.1.1, 169
topologically complemented subspace,
7.4.9, 108
topology, 9.1.7, 147
topology compatible with duality,
10.4.1, 175
topology compatible with vector
structure, 10.1.1, 169
topology given by open sets, 9.1.12, 148
topology of a multinormed space,
5.2.8, 61
topology of a uniform space, 9.5.3, 160
topology of the distribution space,
10.10.6, 198
topology of the test function space,
10.10.6, 198
total operator, 2.2.1, 12
total set of functionals, 7.4.11 (2), 109
totally bounded, 4.6.3, 49
transitive relation, 1.2.1, 3
translation, 10.9.4 (1), 189
translation of a distribution,
10.11.17 (7), 211
transpose of an operator, 7.6.2, 114
trivial topology, 9.1.8 (3), 147
truncator, 9.6.19 (1), 167
truncator direction, 10.10.2 (5), 194
truncator set, 10.10.2, 194
twin of a Hilbert space, 6.1.10 (3), 83
twin of a vector space, 2.1.4 (2), 10
Two Norm Principle, 7.4.16, 111
two-sided ideal, 8.3.3, 132; 11.6.2, 220
Tychono cube, 9.2.17 (2), 151
Tychono product, 9.2.17 (2), 151
Tychono space, 9.3.15, 155
Tychono Theorem, 9.4.8, 157
Tychono topology, 9.2.17 (2), 151
Tychono uniformity, 9.5.5 (4), 160
T
1
-space, 9.3.2, 151
T
1
-topology, 9.3.2, 151
T
2
-space, 9.3.5, 152
T
3
-space, 9.3.9, 153
T
3
1
/
2
-space, 9.3.15, 155
T
4
-space, 9.3.11, 153
ultralter, 1.3.9, 7
ultrametric inequality, 9.5.14, 162
Subject Index 275
ultranet, 9.4.4, 156
unconditionally summable family
of vectors, 5.5.9 (7), 73
unconditionally summable sequence,
5.5.9 (7), 73
underlying set, 2.1.3, 10
Uniform Boundedness Principle,
7.2.5, 103
uniform convergence, 7.2.10, 105;
9.5.5 (6), 161
uniform space, 9.5.1, 159
uniformity, 9.5.1, 159
uniformity of a multinormed space,
5.2.4, 60
uniformity of a seminormed space,
5.2.2, 60
uniformity of a topological vector
space, 10.1.10, 171
uniformity of the empty set, 9.5.1, 159
uniformizable space, 9.5.4, 160
uniformly continuous mapping, 4.2.5, 44
unit, 10.9.4, 188
unit ball, 5.2.11, 61
unit circle, 8.1.3, 121
unit disk, 8.1.3, 121
unit element, 11.1.1, 213
unit sphere, Ex. 10.6, 212
unit vector, 6.3.5, 88
unital algebra, 11.1.1, 213
unitary element, 11.7.1, 224
unitary operator, 6.3.17, 90
unitization, 11.1.2, 213
unity, 11.1.1, 213
unity of a group, 10.9.4 (1), 188
unity of an algebra, 11.1.1, 213
unordered sum, 5.5.9 (7), 73
unorderly summable sequence,
5.5.9 (7), 73
Unremovable Spectral Boundary
Theorem, 11.2.6, 216
upper bound, 1.2.4, 4
upper envelope, 3.4.8 (3), 28
upper right Dini derivative, 4.7.7, 53
upward-ltered set, 1.2.15, 6
Urysohn Great Lemma, 9.3.13, 154
Urysohn Little Lemma, 9.3.10, 153
Urysohn Theorem, 9.3.14, 155
2-Ultrametric Lemma, 9.5.15, 162
vague topology, 10.9.5, 191
value of a germ at a point, 8.1.21, 126
van der Waerden function, 4.7.7, 53
vector, 2.1.3, 10
vector addition, 2.1.3, 10
vector eld, 5.5.9 (6), 71
vector lattice, 3.2.7, 23
vector space, 2.1.3, 10
vector sublattice, 10.8.4 (4), 182
vector topology, 10.1.1, 169
Volterra operator, Ex. 5.12, 79
von NeumannJordan Theorem,
6.1.9, 81
V -net, 4.6.2, 49
V -small, 4.5.3, 47
weak derivative, 10.10.5 (4), 196
weak multinorm, 5.1.10 (4), 58
weak topology, 10.3.5, 174
weak

topology, 10.3.11, 175


weak uniformity, 9.5.5 (6), 161
weaker pretopology, 9.1.2, 146
weakly holomorphic function, 8.1.5, 122
weakly operator holomorphic function,
8.1.5, 122
Weierstrass function, 4.7.7, 53
Weierstrass Theorem, 4.4.5, 46;
9.4.5, 157
Well-Posedness Principle, 7.4.6, 108
Wendel Theorem, 10.9.4 (7), 191
Weyl Criterion, 6.5.4, 93
X-valued function, 5.5.9 (6), 71
Young inequality, 5.5.9 (4), 69
zero of a vector space, 2.1.4 (3), 11

S-ar putea să vă placă și