You are on page 1of 282

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 Hahn–Banach Theorem ................................. 30
§ 3.6. The Kre–ın–Milman Theorem ................................. 33
§ 3.7. The Balanced Hahn–Banach 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. Banach’s 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 Riesz–Schauder 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. Reflexive Spaces ............................................ 189
§ 10.8. The Space C(Q, R) ........................................ 189
§ 10.9. Radon Measures ............................................ 195
§ 10.10. The Spaces D(Š) and D 0 (Š) ............................... 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 Gelfand–Na–ımark 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 Riesz–Dunford holomorphic functional calculus,
the Fredholm index theory, convex analysis and duality theory for locally convex
spaces.
With standard provisos the presentation is self-sufficient, whereas exposing
about a hundred of celebrated “named” theorems furnished with complete proofs
and culminating in the Gelfand–Naı̆mark–Segal construction for C ∗ -algebras.
The first 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 AMS-TEX, the American Mathematical Soci-
ety’s TEX system.
To clear my conscience completely, I also confess that := stands for the definor,
the assignment operator, / marks the beginning of a (possibly empty) proof, and .
signifies 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
century 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 se-
ries. 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
developed by V. Volterra, C. Arzelà, S. Pincherle, P. Levy, and other representa-
tives of the French and Italian mathematical schools. J. Hadamard’s contribution
to the recent discipline should not be reduced to the invention of the word “func-
tional” (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. Fréchet, 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 Hausdorff’s famous research
heralded the birth of general topology.
For the sequel it is essential to emphasize that one of the most attractive,
difficult, and important sections of classical analysis, the calculus of variations,
became the first 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. Poincaré, et al. The efforts of these mathematicians fertilized soil
for D. Hilbert’s celebrated research into quadratic forms in infinitely 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 influenced theoretical physics, first 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 Minkowski’s geo-
metric ideas. His invention, the apparatus for the finite-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. Carathéodory, 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. Bernoulli’s super-
position principle, the line of set functions and integration theory, the line of opera-
tional calculus, the line of finite differences and fractional derivation, the line of gen-
eral analysis, and many others). These demerits notwithstanding, the three sources
listed above reflect the main, and most principal, regularity: functional analysis has
synthesized and promoted ideas, concepts, and methods from the classical sections
of mathematics: algebra, geometry, and analysis. Therefore, although functional
analysis verbatim means analysis of functions and functionals, even a superficial
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,
unification, 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 finds
diverse applications in modern physics, especially in quantum mechanics.”
The S. Banach treatise Theórie des Operationes Linéares, printed half a cen-
tury ago, inaugurated functional analysis as an essential activity in mathematics.
Its influence on the development of mathematics is seminal: Ubiquitous and om-
nipresent, Banach’s 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. Kreı̆n, 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 Kantorovich’s 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 reflected 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 financial 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 ⊂ A × B the set

dom F := D(F ) := {a ∈ A : (∃ b ∈ B) (a, b) ∈ F }

is the domain (of definition) of F and the set

im F := 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 definition a subset of A2 , i.e. a correspon-
dence from A to A (in words: “F acts in A”).
(3) Let F ⊂ A × B. Then F is a single-valued correspondence if for all
a ∈ A the containments (a, b1 ) ∈ F and (a, b2 ) ∈ F imply b1 = b2 . In particular,
if U ⊂ A and IU := {(a, a) ∈ A2 : a ∈ U }, then IU 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 dom F = 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 ⊂ A × B is denoted by F : A → B. Observe that here dom F always
coincides with A whereas im F may differ from B. The identity relation IU on A is
a mapping if and only if A = U and in this case IU is called the identity mapping
in U or the diagonal of U 2 . The set IU may be treated as a subset of U × A. The
resulting mapping is usually denoted by ι : U 0 = 0 O and is called the identical
embedding of U into A. It is said that “F is a correspondence from A onto B”
if im F = 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 : A0 = 0 O is a family in B (indexed in A by F ), also denoted
by (ba )a∈A or a 7→ ba (a ∈ A) or even (ba ). Specifically, (a, b) ∈ F if and only
if b = ba . 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 identification 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 ) := im F |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 ⊂ A × B, assume that A is the product
of A1 and A2 , i.e. A = A1 × A2 . Fixing a1 in A1 and a2 in A2 , consider the sets

F (a1 , · ) := {(a2 , b) ∈ A2 × B : ((a1 , a2 ), b) ∈ F };


F ( · , a2 ) := {(a1 , b) ∈ A1 × B : ((a1 , a2 ), 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
beneficial agreement is effective in similar events.
1.1.4. Definition. The composite correspondence or composition of corre-
spondences F ⊂ A × B and G ⊂ C × D is the set
G ◦ F := {(a, d) ∈ A × D : (∃ 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 |dom F is a mapping. CB
1.1.7. Let F ⊂ A × B and G ⊂ B × C. Further assume U ⊂ A. Then the
correspondence G ◦ F ⊂ A × C satisfies the equality G ◦ F (U ) = G(F (U )). CB
1.1.8. Let F ⊂ A×B, G ⊂ B ×C, and H ⊂ C ×D. Then the correspondences
H ◦ (G ◦ F ) ⊂ A × D and (H ◦ G) ◦ F ⊂ A × D coincide. CB
1.1.9. Remark. By virtue of 1.1.8, the symbol H ◦ G ◦ F and the like are
defined soundly.
1.1.10. Let F, G, and H be correspondences. Then
[
H ◦G◦F = F −1 (b) × H(c).
(b,c)∈G

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) B
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 Definition 1.1.1!). Experience
justifies 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 = F −1 (b) × G(b).
b∈ imF

C Insert H := G, G := I imF and F := F into 1.1.10. B

1.2. Ordered Sets


1.2.1. Definition. Let σ be a relation on a set X, i.e. σ ⊂ X 2 . Reflexivity
for σ means the inclusion σ ⊃ IX ; transitivity, the inclusion σ◦σ ⊂ σ; antisymmetry,
the inclusion σ ∩ σ −1 ⊂ IX ; and, finally, symmetry, the equality σ = σ −1 .
4 Chapter 1

1.2.2. Definition. A preorder is a reflexive 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, σ), with σ an order on X, is
an ordered set or rarely a poset. The notation x ≤σ y is used instead of y ∈ σ(x).
The terminology and notation are often simplified 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 σ to y,” or “x and y belong to σ,” etc. Anal-
ogous 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 X0 of a set X bearing a relation
σ is endowed with the (induced ) relation σ0 := σ ∩ X0 × X0 .
(2) If σ is an order (preorder) in X, then σ −1 is also an order (preorder)
which is called reverse to σ.
(3) Let f : X0 = 0 O be a mapping and let τ be a relation on Y .
Consider the relation f −1 ◦ τ ◦ f appearing on X. By 1.1.10,
[
f −1 ◦ τ ◦ f = f −1 (y1 ) × f −1 (y2 ).
(y1 ,y2 )∈τ

Hence it follows that (x1 , x2 ) ∈ f −1 ◦ τ ◦ f ⇔ (f (x1 ), f (x2 )) ∈ τ. Thus, if τ


is a preorder then f −1 ◦ τ ◦ f itself is a preorder called the preimage or inverse
image of τ 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 ◦ IY ◦ f ).
(4) Let X be an arbitrary set and let ω be an equivalence on X. Define
a mapping ϕ : X0 = 0 O P(X) by ϕ(x) := ω(x). Recall that P(X) stands for
the powerset of X comprising all subsets of X and also denoted by 2X . Let X :=
X/ω := im ϕ be the quotient set or factor set of X by ω or modulo ω. A member
of X/ω is usually referred to as a coset or equivalence class. The mapping ϕ is
the coset mapping (canonical projection, quotient mapping, etc.). Note that ϕ is
treated as acting onto X. Observe that
[
ω = ϕ−1 ◦ ϕ = ϕ−1 (x) × ϕ−1 (x).
x∈X

Now let f : X0 = 0 O be a mapping. Then f admits factorization through X; i.e.,


there is a mapping f : X0 = 0 O (called a quotient of f by ω) such that f ◦ ϕ = f
if and only if ω ⊂ f −1 ◦ f . CB
An Excursus into Set Theory 5

(5) Let (X, σ) and (Y, τ ) be two preordered sets. A mapping f : X0 =


0 O is increasing or isotone (i.e., x ≤σ y ⇒ f (x) ≤τ f (y)) whenever σ ⊂ f −1 ◦ τ ◦ f .
That f decreases or is antitone means σ ⊂ f −1 ◦ τ −1 ◦ f . CB
1.2.4. Definition. Let (X, σ) be an ordered set and let U ⊂ X. An element x
of X is an upper bound of U (write x ≥ U ) if U ⊂ σ −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 σ −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 definitions 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 first element is defined by reversal.
1.2.7. Let U be a subset of an ordered set (X, σ) and let πσ (U ) be the
collection of all upper bounds of U . Suppose that a member x of X is the greatest
element of U . Then, first, x is the least element of πσ (U ); second, σ(x) ∩ U =
{x}. CB
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 supX 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 infimum, a greatest lower bound inf U or inf X U is defined by reversal.
1.2.10. Definition. Let U be a subset of an ordered set (X, σ). A member x
of X is a maximal element of U if σ(x) ∩ U = {x}. A minimal element is again
defined 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 (x1 , x2 ) of elements of X has a least upper bound, x1 ∨x2 := sup{x1 , x2 },
the join of x1 and x2 , and a greatest lower bound, x1 ∧ x2 := inf{x1 , x2 }, the meet
of x1 and x2 .
6 Chapter 1

1.2.13. Definition. A lattice X is complete if each subset of X has a supre-


mum and an infimum 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. CB
1.2.15. Definition. An ordered set (X, σ) is filtered upward provided that
X = σ −1 ◦ σ. A downward-filtered set is defined by reversal. A nonempty poset is
2

a directed set or simply a direction if it is filtered 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-filtered subset of X
has a least upper bound. CB
1.2.18. Remark. The claim of 1.2.17 implies that for calculating a supremum
of each subset it suffices to find suprema of pairs and increasing nets.
1.2.19. Definition. Let (X, σ) be an ordered set. It is said that X is ordered
linearly whenever X 2 = σ ∪ σ −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. Kuratowski–Zorn Lemma. Each inductive set contains a maximal
element.
1.2.21. Remark. The Kuratowski–Zorn 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 filterbase (on X) if B is filtered
downward. Recall that P(X) is ordered by inclusion. It means that a greater subset
includes a smaller subset by definition; this order is always presumed in P(X).
1.3.2. A subset B of P(X) is a filterbase if and only if
1 B 6= ∅ and ∅ 6∈ B;
(2) B1 , B2 ∈ B ⇒ (∃ B ∈ B) B ⊂ B1 ∩ B2 .
1.3.3. Definition. A subset F of P(X) is a filter (on X) if there is a filter-
base B such that F is the set of all supersets of B; i.e.,
F = fil B := {C ∈ P(X) : (∃ B ∈ B) B ⊂ C}.
In this case B is said to be a base for the filter F (so each filterbase is a filter base).
An Excursus into Set Theory 7

1.3.4. A subset F in P(X) is a filter if and only if


(1) F 6= ∅ and ∅ 6∈ F ;
(2) (A ∈ F & A ⊂ B ⊂ X) ⇒ B ∈ F ;
(3) A1 , A2 ∈ F ⇒ A1 ∩ A2 ∈ F . CB
1.3.5. Examples.
(1) Let F ⊂ X × Y be a correspondence and let B be a downward-
filtered subset of P(X). Put F (B) := {F (B) : B ∈ B}. It is easy to see that
F (B) is filtered downward. The notation is alleviated by putting F (B) := fil F (B).
If F is a filter on X and B ∩ dom F 6= ∅ for all B ∈ F , then F (F ) is a filter
(on Y ) called the image of F under F . In particular, if F is a mapping then the
image of a filter on X is a filter on Y .
(2) Let (X, σ) be a direction. Clearly, B := {σ(x) : x ∈ X} is
a filterbase. For a net F : X0 = 0 O , the filter fil F (B) is the tail filter of F . Let
(X, σ) and F : X → Y be another direction and another net in Y . If the tail filter
of F includes the tail filter 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)x∈X
in the direction (X, σ)) 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 filters on X. Take
F1 , F2 ∈ F (X). Say that F1 is finer than F2 or F2 refines F1 (in other words,
F1 is coarser than F2 or F1 coarsens F2 ) whenever F1 ⊃ F2 .
1.3.7. The set F (X) with the relation “to be finer” is a poset. CB
1.3.8. Let N be a direction in F (X). Then N has a supremum F0 :=
sup N . Moreover, F0 = ∪{F : F ∈ N }.
C To prove this, it is necessary to show that F0 is a filter. Since N is not
empty it is clear that F0 6= ∅ and ∅ ∈ / F0 . If A ∈ F0 and B ⊃ A then, choosing
F in N for which A ∈ F , conclude that B ∈ F ⊂ F0 . Given A1 , A2 ∈ F0 ,
find an element F of N satisfying A1 , A2 ∈ F , which is possible because N is
a direction. By 1.3.4, A1 ∩ A2 ∈ F ⊂ F0 . B
1.3.9. Definition. An ultrafilter is a maximal element of the ordered set
F (X) of all filters on X.
1.3.10. Each filter is coarser than an ultrafilter.
C By 1.3.8, the set of filters finer than a given filter is inductive. Recalling the
Kuratowski–Zorn Lemma completes the proof. B
8 Chapter 1

1.3.11. A filter F is an ultrafilter if and only if for all A ⊂ X either A ∈ F


or X \ A ∈ F .
C ⇒: Suppose that A 6∈ F and B := X \ A 6∈ F . Note that A 6= ∅ and
B 6= ∅. Put F1 := {C ∈ P(X) : A ∪ C ∈ F }. Then A 6∈ F ⇒ ∅ 6∈ F1 and
B ∈ F1 ⇒ F1 6= ∅. The checking of 1.3.4 (2) and 1.3.4 (3) is similar. Hence, F1
is an ultrafilter. By definition, F1 ⊃ F . In addition, F is an ultrafilter and so
F1 = F . Observe that B 6∈ F and B ∈ F , a contradiction.
⇐: Take F1 ∈ F (X) and let F1 ⊃ F . If A ∈ F1 and A 6∈ F then X \ A ∈ F
by hypothesis. Hence, X \A ∈ F1 ; i.e., ∅ = A∩(X \A) ∈ F1 , which is impossible. B
1.3.12. If f is a mapping from X into Y and F is an ultrafilter on X then
f (F ) is an ultrafilter on Y . CB
1.3.13. Let X := XF0 := {F ∈ F (X) : F ⊂ F0 } for F0 ∈ F (X). Then X
is a complete lattice.
C It is obvious that F0 is the greatest element of X and {X} is the least
element of X . Therefore, the empty set has a supremum and an infimum in X :
in fact, sup ∅ = inf X = {X} and inf ∅ = sup X = F0 . By 1.2.17 and 1.3.8,
it suffices to show that the join F1 ∨ F2 is available for all F1 , F2 ∈ X . Consider
F := {A1 ∩ A2 : A1 ∈ F1 , A2 ∈ F2 }. Clearly, F ⊂ F0 while F ⊃ F1 and
F ⊃ F2 . Thus to verify the equality F = F1 ∨ F2 , it is necessary to establish
that F is a filter.
Plainly, F 6= ∅ and ∅ 6∈ F . It is also immediate that (B1 , B2 ∈ F ⇒
B1 ∩ B2 ∈ F ). Moreover, if C ⊃ A1 ∩ A2 where A1 ∈ F1 and A2 ∈ F2 , then
C = {A1 ∩ A2 } ∪ C = (A1 ∪ C) ∩ (A2 ∪ C). Since A1 ∪ C ∈ F1 and A2 ∪ C ∈ F2 ,
conclude that C ∈ F . Appealing to 1.3.4, complete the proof. B

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 RM × RN 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 finite 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 finite set?
1.11. Let F be an increasing idempotent mapping of a set X into itself. Assume that F
dominates the identity mapping: F ≥ IX . Such an F is an abstract closure operator or, briefly,
an (upper) envelope. Study fixed points of a closure operator. (Recall that an element x is a fixed
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 filters are there on a finite set?
1.15. How do the least upper and greatest lower bounds of a set of filters look like?
1.16. Let f be a mapping from X onto Y . Prove that each ultrafilter on Y is the image
of some ultrafilter on X under f .
1.17. Prove that an ultrafilter refining the intersection of two filters is finer than either
of them.
1.18. Prove that each filter is the intersection of all ultrafilters finer than it.
1.19. Let A be an ultrafilter on N containing cofinite subsets (a cofinite subset is a subset
with finite complement). Given x, y ∈ s := RN , put x ∼A y := (∃ A ∈ A ) x|A = y|A . Denote
∗R := RN /∼A . For t ∈ R the notation ∗t symbolizes the coset with the constant sequence t defined
as t(n) := t (n ∈ N). Prove that ∗R \ {∗t : t ∈ R} 6= ∅. 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 defined 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 × X0 = 0 O 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 field is the field R of real numbers or the field C of
complex numbers. The symbol F stands for a basic field. 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 field. A module X over F is a vector space
(over F). An element of the ground field 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 field F.
The operation + : X × X0 = 0 O is addition in X and · : F × X0 = 0 O 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 field F is a vector space over F.
(2) Let (X, F, +, · ) be a vector space. Consider (X, F, +, ·∗ ), where
·∗ : (λ, x) 7→ λ∗ x for λ ∈ F and x ∈ X, the symbol λ∗ standing for the conventional
complex conjugate of λ. The so-defined 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 (X0 , F, +, · ) is called a subspace of a vector space


(X, F, +, · ), if X0 is a subgroup of X and scalar multiplication in X0 is the re-
striction of that in X to F × X0 . Such a set X0 is a linear set in X, whereas X
is referred to as an ambient space (for X0 ). It is convenient although not perfectly
puristic to treat X0 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 fields may seem to have a point
in common, zero.
Q (4) Take (Xξ )ξ∈„ , a family of vector spaces over F, and let X :=
X
ξ∈„ ξ be the product of the underlying sets, i.e. the collection of mappings
O
x : „0 = 0 ξ∈„ Xξ such that xξ := x(ξ) ∈ Xξ as ξ ∈ „ (of course, here „ is not
empty). Endow X with the coordinatewise or pointwise operations of addition and
scalar multiplication:

(x1 + x2 )(ξ) := x1 (ξ) + x2 (ξ) (x1 , x2 ∈ X , ξ ∈ „);


(λ · x)(ξ) := λ · x(ξ) (x ∈ X , λ ∈ F, ξ ∈ „)

(below, as a rule, we write λx and sometimes xλ rather than λ · x). The so-
constructed vector space X over F is the product of (Xξ )ξ∈„ . If „ := {1, 2, . . . , N }
then X1 × X2 × . . . × XN := X . In the case Xξ = X for all ξ ∈ „, the designation
X „ := X is used. Given „ := {1, 2, . . . , N }, put X N := X .
(5) LetP (Xξ )ξ∈„ be a family of vector spaces over F. Consider Q their
direct sum X0 := ξ∈„ Xξ . By definition, X0 is the subset of X := ξ∈„ Xξ
which comprises all x0 such that x0 („\„0 ) ⊂ 0 for a finite subset „0 ⊂ „ (routinely
speaking, „0 is dependent on x0 ). It is easily seen that X0 is a linear set in X .
The vector space associated with X0 presents a subspace of the product of (Xξ )ξ∈„
and is the direct sum of (Xξ )ξ∈„ .
(6) Given a subspace (X, F, +, · ) of a vector space (X0 , F, +, · ),
introduce
∼X0 := {(x1 , x2 ) ∈ X 2 : x1 − x2 ∈ X0 }.
Then ∼X0 is an equivalence on X. Denote X := X/∼X0 and let ϕ : X0 = 0 O X be
the coset mapping. Define operations on X by letting

x1 + x2 := ϕ(ϕ−1 (x1 ) + ϕ−1 (x2 )) (x1 , x2 ∈ X );


λx := ϕ(λϕ−1 (x)) (x ∈ X , λ ∈ F).
12 Chapter 2

Here, as usual, for subsets S1 and S2 of X, a subset ƒ of F, and a scalar λ, a member


of F, it is assumed that

S1 + S2 := +{S1 × S2 };
ƒS1 := · (ƒ × S1 ); λS1 := {λ}S1 .

Thus X is furnished with the structure of a vector space. This space, denoted
by X/X0 , is the quotient (space) of X by X0 or the factor space of X modulo X0 .

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.
C 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. B

2.1.6. Remark. With X1 , X2 ∈ Lat(X), the equality X1 ∨ X2 = X1 + X2


holds. It is evident that inf E = ∩{X0 : X0 ∈ E } for a nonempty subset E
of Lat(X). Provided that E is filtered upward, sup E = ∪ {X0 : X0 ∈ E }. CB

2.1.7. Definition. Subspaces X1 and X2 of a vector space X split X into


(algebraic) direct sum decomposition (in symbols, X = X1 ⊕ X2 ), if X1 ∧ X2 = 0
and X1 ∨ X2 = X. In this case X2 is an (algebraic) complement of X1 to X, and
X1 is an (algebraic) complement of X2 to X. It is also said that X1 and X2 are
(algebraically) complementary to one another.

2.1.8. Each subspace of a vector space has an algebraic complement.


C Take a subspace X1 of X. Put

E := {X0 ∈ Lat(X) : X0 ∧ X1 = 0}.

Obviously, 0 ∈ E . Given a chain E i0 in E , from 2.1.6 infer that X1 ∧ 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, X2 . If x ∈ X \ (X1 + X2 ) then

(X2 + {λx : λ ∈ F}) ∧ X1 = 0.

Indeed, if x2 + λx = x1 with x1 ∈ X1 , x2 ∈ X2 and λ ∈ F, then λx ∈ X1 + X2 and


so λ = 0. Hence, x1 = x2 = 0 as X1 ∧ X2 = 0. Therefore, X2 + {λx : λ ∈ F} = X2
because X2 is maximal. It follows that x = 0. At the same time, it is clear that
x 6= 0. Finally, X1 ∨ X2 = X1 + X2 = X. B
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 dom S 6= X and say that T is given
on X or T is defined everywhere or even T is a total operator, whereas S is referred
to as not-everywhere-defined or partially-defined 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 dom T = X and
T (λ1 x1 + λ2 x2 ) = λ1 T x1 + λ2 T x2 (λ1 , λ2 ∈ F; x1 , x2 ∈ X). CB
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 . CB
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 = IX and T ◦ S = IY (in this
event S = T −1 and T = S −1 ). CB
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 ) 7→ ST is usually treated as the mapping
◦ : L (Y, Z) × L (X, Y )0 = 0 O 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 X1 is a subspace of a vector space X and X2 is an algebraic
complement of X1 then X2 is isomorphic with X/X1 . Indeed, if ϕ : X0 = 0 O /X1
is the coset mapping then its restriction to X2 , i.e. the operator x2 7→ ϕ(x2 ) with
x2 ∈ X2 , implements a desired isomorphism.CB
(3) Consider X := ξ∈„ Xξ , the product of vector spaces (Xξ )ξ∈„ .
Q
Take a coordinate projection, i.e. a mapping Prξ : X 0 = 0 O ξ defined by Prξ x := xξ .
Clearly, Prξ is a linear operator, Prξ ∈ L (X , Xξ ). Such an operator is often
treated as an endomorphism of X , a memberQof L (X ), on implying a natural
isomorphism between Xξ and Xξ , where Xξ := η∈„ X η with X η := 0 if η 6= ξ and
X ξ := Xξ .
(4) Let X := X1 ⊕ X2 . Since +−1 is an isomorphism between X and
X1 × X2 , we may define P1 , P2 ∈ L (X) as P1 := PX1 ||X2 := Pr1 ◦(+−1 ) and
P2 := PX2 ||X1 := Pr2 ◦(+−1 ). The operator P1 is the projection of X onto X1 along
X2 and P2 is the complementary projection to P1 or the complement of P1 (in
symbols, P2 = P1d ). In turn, P1 is complementary to P2 , and P2 projects X onto
X2 along X1 . Observe also that P1 + P2 = IX . Moreover, P12 := P1 P1 = P1 , 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 P T P = T P holds if and only
if T (X0 ) ⊂ X0 with X0 = im P (read: X0 is invariant under T ). CB
The equality T PX1 ||X2 = PX1 ||X2 T holds whenever both X1 and X2 are invari-
ant under T , and in this case the direct sum decomposition X = X1 ⊕ X2 reduces T .
The restriction of T to X1 is acknowledged as an element T1 of L (X1 ) which is
called the part of T in X1 . If T2 ∈ L (X2 ) is the part of T in X2 , then T is
expressible in matrix form  
T1 0
T ∼ .
0 T2
Namely, an element x of X1 ⊕ X2 is regarded as a “column vector” with compo-
nents x1 and x2 , where x1 = PX1 ||X2 x and x2 = PX1 ||X2 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 T1 x1 and T2 x2 , is certainly considered
as T x (in this case, we also write T x1 and T x2 ). In other words, T is identified
with the mapping from X1 × X2 to X1 × X2 acting as
    
x1 T1 0 x1
7→ .
x2 0 T2 x2
Vector Spaces 15

In a similar way we can introduce matrix presentation for general operators con-
tained in L (X1 ⊕ X2 , Y1 ⊕ Y2 ). CB
(5) A finite subset E of X is linearly independent (in X) provided that
e∈E λe e = 0, with λe ∈ F (e ∈ E ), implies λe = 0 for all e ∈ E . An arbitrary
P
subset E of X is linearly independent if every finite 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 dim X. Every vector space is isomorphic to the direct sum of a family
(F )ξ∈„ with „ of cardinality dim X. Suppose that X1 is a subspace of X. The codi-
mension of X1 is the dimension of X/X1 , with codim X1 standing for the former.
If X = X1 ⊕ X2 then codim X1 = dim X2 and dim X = dim X1 + codim X1 . CB

2.3. Equations in Operators


2.3.1. Definition. Given T ∈ L (X, Y ), define the kernel of T as ker T :=
T −1 (0), the cokernel of T as coker T := Y / im T , and the coimage of T as coim T :=
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 im T = Y .
2.3.2. An operator is an isomorphism if and only if it is a monomorphism and
an epimorphism simultaneously. CB
2.3.3. Remark. Below use is made of the concept of commutative diagram.
So the phrase, “The following diagram commutes,”
α1
X - Y 
@ α2 α4
α3@
R ?
@
V α- W
5

encodes the containments α1 ∈ L (X, Y ), α2 ∈ L (Y, W ), α3 ∈ L (X, W ),


α4 ∈ L (V, Y ) and α5 ∈ L (V, W ), as well as the equalities α2 α1 = α3 and
α5 = α2 α4 .
T S
2.3.4. Definition. A diagram X −→ Y −→ Z is an exact sequence (at the
term Y ), if ker S = im T . A sequence . . . 0 = 0 O k−1 0 = 0 O k 0 = 0 O k+1 0 = 0 O is exact
at Xk , if for all k the subsequence Xk−1 0 = 0 O k 0 = 0 O k+1 (symbols of operators
are omitted), and is exact if it is exact at every term (except the first and the last,
if any).
16 Chapter 2

2.3.5. Examples.
T S
(1) An exact sequence X −→ Y −→ Z is semi-exact, i.e. ST = 0. The
converse is not true.
(2) A sequence 00 = 0 O −→ Y is exact if and only if T is a monomor-
T

phism. (Throughout the book 00 = 0 O certainly denotes the sole element of L (0, X),
zero (cf. 2.1.4 (3)).)
(3) A sequence X −→ Y 0 = 0 O is exact if and only if T is an epi-
T

morphism. (Plainly, the symbol Y 0 = 0 O 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 00 = 0 O −→ Y 0 = 0 O is exact.
T

(5) Suppose that X0 is a subspace of X. Let ι : X0 0 = 0 O stand for


the identical embedding of X0 into X. Consider the quotient space X/X0 and let
ϕ : X0 = 0 O /X0 be the corresponding coset mapping. Then the sequence

00 = 0 O 0 −→ X −→ X/X0 0 = 0 O
ι ϕ

is exact. (On similar occasions the letters ι and ϕ are usually omitted.) In a sense,
this sequence is unique. Namely, consider a so-called short sequence

00 = 0 O −→ Y −→ Z0 = 0 O
T S

and assume that it is exact. Putting Y0 := im T , arrange the following commutative


diagram
T S
0 - X - Y - Z - 0
α β γ
? ? ?
0 - Y0 - Y - Y /Y0 - 0

where α, β, and γ are some isomorphisms. In other words, a short exact sequence
actually presents a subspace and the corresponding quotient space. CB
(6) Every T in L (X, Y ) generates the exact sequence

00 = 0 O T 0 = 0 O −→ Y 0 = 0 O T 0 = 0 O
T

which is called the canonical exact sequence for T .


2.3.6. Definition. Given T0 ∈ L (X0 , Y ), with X0 a subspace of X, call
an operator T from X to Y an extension of T0 (onto X, in symbols, T ⊃ T0 ),
provided that T0 = T ι, where ι : X0 0 = 0 O is the identical embedding of X0
into X.
Vector Spaces 17

2.3.7. Let X and Y be vector spaces and let X0 be a subspace of X. Then


each T0 in L (X0 , Y ) has an extension T in L (X, Y ).
C Putting T := T0 PX0 , where PX0 is a projection onto X0 , settles the claim. B
2.3.8. Theorem. Let X, Y , and Z be vector spaces. Take A ∈ L (X, Y ) and
B ∈ L (X, Z). The diagram
A
X - Y

X
@
B@@
R ?
Z
is commutative for some X in L (Y, Z) if and only if ker A ⊂ ker B.
C ⇒: It is evident that ker A ⊂ ker B in case B = X A.
⇐: Set X := B◦A−1 . Clearly, X ◦A(x) = B◦(A−1 ◦A)x = B(x+ker A) = Bx.
Show that X0 := X |im A is a linear operator. It suffices to check that X is single-
valued. Suppose that y ∈ im A and z1 , z2 ∈ X (y). Then z1 = Bx1 , z2 = Bx2 and
Ax1 = Ax2 = y. By hypothesis B(x1 − x2 ) = 0; therefore, z1 = z2 . Applying 2.3.7,
find an extension X of X0 to Y . B
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 ∼codim T .
C Immediate from 2.3.8 and 2.3.9. B
2.3.11. Remark. The operator T is sometimes called the monoquotient of T
and treated as acting onto im T . In this connection, observe that T is expressible
as the composition T = ιT ϕ, with ϕ an epimorphism, T an isomorphism, and ι
a monomorphism; i.e., the following diagram commutes:
T -
coim T im T
ϕ 6 ι
T ?
X - Y

2.3.12. Let X be a vector space and let f0 , f1 , . . . , fNP belong to X # . The


N
functional f0 is a linear combination of f1 , . . . , fN (i.e., f0 = j=1 λj fj with λj in
F ) if and only if ker f0 ⊃ ∩N
j=1 ker fj .
18 Chapter 2

C Define the linear operator (f1 , . . . , fN ) : X0 = 0 O F N by (f1 , . . . , fN )x :=


(f1 (x), . . . , fN (x)). Obviously, ker(f1 , . . . , fN ) = ∩N
j=1 ker fj . Now apply 2.3.8 to
the problem

(f1 ,... ,fN )


X - FN
@
@
f0@
@
R ?
@
F

on recalling what F N # is. B

2.3.13. Theorem. Let X, Y , and Z be vector spaces. Take A ∈ L (Y, X)


and B ∈ L (Z, X). The diagram
A
X  Y

X
@
I 6
@
B @
Z

is commutative for some X in L (Z, Y ) if and only if im A ⊃ im B.

C ⇒: im B = B(Z) = A(X (Z)) ⊂ A(Y ) = im A.


⇐: Let Y0 be an algebraic complement of ker A to Y and A0 := A|Y0 . Then A0
is an isomorphism between Y0 and im A. The operator X := ιA−1 0 B is obviously
a sought solution, with ι the identical embedding of Y0 into Y . B

2.3.14. Remark. Provided that the operator A is a monomorphism, there is


a unique operator X in 2.3.13. CB

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. Snowflake Lemma. Let S ∈ L (Y, Z) and T ∈ L (X, Y ). There


are unique operators α1 , . . . , α6 such that the following diagram commutes:
Vector Spaces 19
0 0
S
w α2 /

ker ST - ker S
α1  7 S  S α3
 S  S
 w
S T /
 w
S
0 - ker T - X - Y - coker T - 0
S S  
S ST SSw  /
 S 
S 
α6 S Z  α4
S  S 
S  S 
w 
S / w 
S /
coker S  coker ST
α5
/
 w
S
0 0
Moreover, the sequence

00 = 0 O T −→ coker S0 = 0 O
α1 α2 α3 α4 α5
ker ST −→ ker S −→ coker T −→ coker ST −→

is exact. CB

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 field Z2 .
2.3. Describe vector spaces with a countable Hamel basis.
2.4. Prove that there is a discontinuous solution f : R0 = 0 O 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 (Xξ ) is presentable as the product
of the algebraic duals of (Xξ ).
2.6. Let X ⊃ X0 ⊃ X00 . Prove that the spaces X/X00 and (X/X0 )/(X00 /X0 ) are isomor-
phic.
2.7. Define the “double sharp” mapping by the rule

x## : x# 7→ hx | x# i (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 finite-dimensional spaces and only such spaces are algebraically reflexive;
i.e.,
##
(X) = X ## ⇔ dim X < +∞.
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 satisfies the conditions T n−1 6=
0 and T n = 0 for a natural n. Prove that the operators T 0 , T, . . . , T n−1 are linearly independent.
2.12. Describe the structure of a linear operator defined 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 (xe )e∈E , find a family of functionals
(x#
e e∈E satisfying the following conditions:
)

hxe | x#
e i=1 (e ∈ E );
hxe | x#
e i =0 (e, e0 ∈ E, e 6= e0 ).

2.17. Given a family of linearly independent functionals (x#


e )e∈E , find a family of vectors
(xe )e∈E satisfying the following conditions:

hxe | x#
e i=1 (e ∈ E );
hxe | x#
e i=0 (e, e0 ∈ E, e 6= e0 ).

2.18. Find compatibility conditions for simultaneous linear equations and linear inequalities
in real vector spaces.
2.19. Consider the commutative diagram

T
W −→ X −→ Y −→ Z
α↓ β↓ γ↓ δ↓
T
W −→ X −→ Y −→ Z

with exact rows and such that α is an epimorphism, and δ is a monomorphism. Prove that
ker γ = T (ker β) and T −1 (im γ) = im β.
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 (λ1 , λ2 ) ∈ € ⇒ λ1 U + λ2 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 € := R2 then a nonempty € -set in a vector space X is a real
subspace of X.
2
(4) By definition, a cone is a nonempty € -set with € := R+ . In other
words, a nonempty set K is a cone if and only if K + K ⊂ K and αK ⊂ K for all
2
α ∈ R+ . A nonempty R+ \ 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 € := {(λ1 , λ2 ) ∈ F 2 : λ1 + λ2 = 1}, is
an affine variety or a flat. If X0 is a subspace of X and x ∈ X then x+X0 := {x}+X0
is an affine variety in X. Conversely, if L is an affine variety in X and x ∈ L then
L − x := L + {−x} is a linear set in X. CB
(6) If € := {(λ1 , λ2 ) ∈ F 2 : |λ1 | + |λ2 | ≤ 1} then a nonempty € -set is
absolutely convex.
(7) If € := {(λ, 0) ∈ F 2 : |λ| ≤ 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 € := {(λ1 , λ2 ) ∈ R2 : λ1 ≥ 0, λ2 ≥
0, λ1 + λ2 = 1}.
22 Chapter 3

(9) A conical segment or conical slice is by definition a nonempty € -set


2
with € := {(λ1 , λ2 ) ∈ R+ : λ1 + λ2 ≤ 1}. A set is a conical segment if and only
if it is a convex set containing zero. CB
(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 ∈ im E } ∈ (€ ). Provided that im E is filtered upward (by inclusion),
∪{U : U ∈ im E } ∈ (€ ). CB
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 ∈ (€ ).
C If U or V is nonempty then U × V = ∅, and there is nothing to prove.
Now take u1 , u2 ∈ U , v1 , v2 ∈ V , and (λ1 , λ2 ) ∈ € . Find λ1 u1 + λ2 u2 ∈ U and
λ1 v2 + λ2 v1 ∈ V . Hence, (λ1 u1 + λ2 u2 , λ1 v1 + λ2 v2 ) ∈ U × V . B
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 specific attribute, the latter is pre-
served for naming a € -correspondence. With this in mind, we speak about linear
and convex correspondences, affine 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 ).
C Take y1 , y2 ∈ T (U ). Then (x1 , y1 ) ∈ T and (x2 , y2 ) ∈ T with some
x1 , x2 ∈ U . Given (λ1 , λ2 ) ∈ €2 , observe that λ1 (x1 , y1 ) + λ2 (x2 , y2 ) ∈ T ,
because by hypothesis (λ1 , λ2 ) ∈ €1 . Finally, infer that λ1 y1 + λ2 y2 ∈ T (U ). B
3.1.9. The composition of € -correspondences is also a € -correspondence.
C Let F ⊂ X × V and G ⊂ W × Y , with F, G ∈ (€ ). Note that

(x1 , y1 ) ∈ G ◦ F ⇔ (∃ v1 ) (x1 , v1 ) ∈ F & (v1 , y1 ) ∈ G;


(x2 , y2 ) ∈ G ◦ F ⇔ (∃ v2 ) (x2 , v2 ) ∈ F & (v2 , y2 ) ∈ G.

To complete the proof, “multiply the first row by λ1 ; the second, by λ2 , where
(λ1 , λ2 ) ∈ € ; and sum the results.” B
Convex Analysis 23

3.1.10. If subsets U and V of a vector space are € -sets for some € ⊂ F 2 then
αU + βV ∈ (€ ) for all α, β ∈ F.
C The claim is immediate from 3.1.5, 3.1.8, and 3.1.9. B
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) U1 ⊂ U2 ⇒ H€ (U1 ) ⊂ H€ (U2 );
(4) U ∈ (€ ) ⇔ U = H€ (U );
(5) H€ (H€ (U )) = H€ (U ). CB
3.1.13. The Motzkin formula holds:

H€ (U ) = ∪{H€ (U0 ) : U0 is a finite subset of U }.

C Denote the right side of the Motzkin formula by V . Since U0 ⊂ U ; applying


3.1.12 (3), deduce that H€ (U0 ) ⊂ H€ (U ); and, hence, H€ (U ) ⊃ V . By 3.1.12 (2),
it is necessary (and, surely, sufficient) to verify that V ∈ (€ ). The last follows from
3.1.3 and the inclusion H€ (U0 ) ∪ H€ (U1 ) ⊂ H€ (U0 ∪ U1 ). B
3.1.14. Remark. The Motzkin formula reduces the problem of describing
arbitrary € -hulls to calculating € -hulls of finite sets. Observe also that € -hulls for
concrete € s have special (but natural) designations. For instance, if € := {(λ1 , λ2 ) ∈
2
R+ : λ1 + λ2 = 1}, then the term “convex hull” is used and co(U ) stands for H€ (U ).
If U 6= ∅, the notation for HF 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 affine hull, conical hull, etc. are introduced in a similar fashion. Note also that
the convex hull of a finite set of points comprises their convex combinations; i.e.,
(N )
X
co({x1 , . . . , xN }) = λk xk : λk ≥ 0, λ1 + . . . + λN = 1 . /.
k=1

3.2. Ordered Vector Spaces


3.2.1. Definition. Let (X, R, +, · ) be a vector space. A preorder σ on X
is compatible with vector structure if σ is a cone in X 2 ; in this case X is an ordered
vector space. (It is more precise to call (X, R, +, · , σ) a preordered vector space,
reserving the term “ordered vector space” for the case in which σ is an order.)
24 Chapter 3

3.2.2. If X is an ordered vector space and σ is the corresponding preorder


then σ(0) is a cone and σ(x) = x + σ(0) for all x ∈ X.
C By 3.1.3, σ(0) is a cone. The equality (x, y) = (x, x) + (0, y − x) yields
the equivalence (x, y) ∈ σ ⇔ y − x ∈ σ(0). B
3.2.3. Let K be a cone in a vector space X. Denote

σ := {(x, y) ∈ X 2 : y − x ∈ K}.

Then σ is a preorder compatible with vector structure and K coincides with the
cone σ(0) of positive elements of X. The relation σ is an order if and only
if K ∩ (−K) = 0.
C It is clear that 0 ∈ K ⇒ IX ⊂ σ and K + K ⊂ K ⇒ σ ◦ σ ⊂ σ. Furthermore,
σ −1 = {(x, y) ∈ X 2 : x − y ∈ K}. Therefore, σ ∩ σ −1 ⊂ IX ⇔ K ∩ (−K) = 0.
To show that σ is a cone, take (x1 , y1 ), (x2 , y2 ) ∈ σ and α1 , α2 ∈ R+ . Find
α1 y1 + α2 y2 − (α1 x1 + α2 x2 ) = α1 (y1 − x1 ) + α2 (y2 − x2 ) ∈ α1 K + α2 K ⊂ K. B
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 X0
is a subspace of X then the order induced in X0 is defined by the cone X0 ∩ X+ .
In this way X0 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 Lr (X, Y ),
and a member of Lr (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 briefly
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.
C Let U be bounded below: x ≤ U for some x. So, −x ≥ −U . Applying 3.2.8,
find sup(−U ). Obviously, − sup(−U ) = inf U . B
3.2.10. If U and V are nonempty bounded above subsets of a K-space then
sup(U + V ) = sup U + sup V.
C 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 ) = sup{sup(u + V ) : u ∈ U }
= sup{u + sup V : u ∈ U } = sup V + sup{u : 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 λU = λ sup U for λ ∈ 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.
C x + y − x ∧ y = x + y + (−x) ∨ (−y) = y ∨ x B
3.2.14. x = x+ − x− ; |x| = x+ + x− .
C The first equality follows from 3.2.13 on letting y := 0. Furthermore, |x| =
x ∨ (−x) = −x + (2x) ∨ 0 = −x + 2x+ = (x+ − x− ) + 2x+ = x+ + x− . B
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] := σ(u) ∩ σ −1 (v) is the (order) interval with endpoints u and v.)
C The inclusion [0, x]+[0, y] ⊂ [0, x+y] is obvious. Assume that 0 ≤ z ≤ x+y
and put z1 := z ∧ x. It is easy that z1 ∈ [0, x]. Now if z2 := z − z1 then z2 ≥ 0 and
z2 = z − z ∧ x = z + (−z) ∨ (−x) = 0 ∨ (z − x) ≤ 0 ∨ (x + y − x) = 0 ∨ y = y. B
3.2.16. Remark. The conclusion of the Interval Addition Lemma is often
referred to as the Riesz Decomposition Property.
3.2.17. Riesz–Kantorovich Theorem. Let X be a vector space and let Y
be a K-space. The space of regular operators Lr (X, Y ), ordered by the cone
of positive operators L+ (X, Y ), is a K-space. CB
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 X0 be C([0, 1], R), the subspace of X comprising all
continuous functions. Put Y := X0 and equip X0 , X, and Y with the natural order
(cf. 3.2.6 (1) and 3.2.6 (2)). Consider the problem of extending the identical em-
bedding T0 : X0 0 = 0 O to a positive operator T in L+ (X, Y ). If the problem had
a solution T , then each nonempty bounded set E in X0 would have a least upper
bound supX0 E calculated in X0 . Namely, supX0 E = T supX E , where supX E is
the least upper bound of E in X; whereas, undoubtedly, Y fails to be a K-space.
(2) Denote by s := RN the sequence space and furnish s with the natural
order. Let c be the subspace of s comprising all convergent sequences, the convergent
sequence space. Demonstrate that the positive functional f0 : c0 = 0 O R defined
by f0 (x) := lim x(n) has no positive extension to s. Indeed, assume that f ∈
s# , f ≥ 0 and f ⊃ f0 . Put x0 (n) := n and xk (n) := k ∧ n for k, n ∈ N. Plainly,
f0 (xk ) = k; moreover, f (x0 ) ≥ f (xk ) ≥ 0, since x0 ≥ xk ≥ 0, a contradiction.
3.3.2. Definition. A subspace X0 of an ordered vector space X with positive
cone X+ is massive (in X) if X0 + X+ = X. The terms “coinitial” or “minorizing”
are also in common parlance.
3.3.3. A subspace X0 is massive in X if and only if for all x ∈ X there are
elements x0 and x0 in X0 such that x0 ≤ x ≤ x0 . CB
3.3.4. Kantorovich Theorem. If X is an ordered vector space, X0 is massive
in X, and Y is a K-space; then each positive operator T0 , a member of L+ (X0 , Y ),
has a positive extension T , a member of L+ (X, Y ).
C Step I. First, let X := X0 ⊕ X1 , where X1 is a one-dimensional subspace,
X1 := {αx : α ∈ R}. Since X0 is massive and T0 is positive, the set U := {T0 x0 :
x0 ∈ X0 , x0 ≥ x} is nonempty and bounded below. Consequently, there is some y
such that y := inf U . Assign T x := {T0 x0 +αy : x = x0 +αx, x0 ∈ X0 , α ∈ R}. It is
clear that T is a single-valued correspondence. Further, T ⊃ T0 and dom T = X.
So, only the positivity of T needs checking. If x = x0 + αx and x ≥ 0, then
the case of α equal to 0 is trivial. If α > 0 then x ≥ −x0 /α. This implies that
−T0 x0 /α ≤ y, i.e., T x ∈ Y+ . In a similar way for α < 0 observe that x ≤ −x0 /α.
Thus, y ≤ −T0 x0 /α and, finally, T x = T0 x0 + αy ∈ Y+ .
Step II. Now let E be the collection of single-valued correspondences S ⊂
X × Y such that S ⊃ T0 and S(X+ ) ⊂ Y+ . By 3.1.3, E is inductive in order
by inclusion and so, by the Kuratowski–Zorn Lemma, E has a maximal element T .
If x ∈ X \ dom T , apply the result of Step I with X := dom T ⊕ X1 , X0 :=
dom T, T0 := T and X1 := {αx : α ∈ R} to obtain an extension of T . But T is
maximal, a contradiction. Thus, T is a sought operator. B
Convex Analysis 27

3.3.5. Remark. When Y := R, Theorem 3.3.4 is sometimes referred to as


the Kreı̆n–Rutman 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+ .
C Let T be such functional and := X+ − X+ . Take f ∈ X # . It suffices to
check that ker f ⊃⇒ = . Evidently, T + f ∈ [0, T ]; i.e., T + f = αT for some
α ∈ [0, 1]. If T | = 0 then 2T ∈ [0, T ]. Hence, T = 0 and f = 0. Now if T (x0 ) 6= 0
for some x0 ∈, then α = 1 and f = 0. B

3.3.8. Discrete Kreı̆n–Rutman Let X be a massive subspace of an ordered


vector space X and let T0 be a discrete functional on X0 . Then there is a discrete
functional T on X extending T0 .
C Adjust the proof of 3.3.4.
Step I. Observe that the exhibited functional T is discrete. For, if T 0 ∈ [0, T ]
then there is some α ∈ [0, 1] such that T 0 (x0 ) = αT (x0 ) for all x0 ∈ X0 and so
(T − T 0 )(x0 ) = (1 − α)T (x0 ). Estimating, find

T 0 (x) ≤ inf{T 0 (x0 ) : x0 ≥ x, x0 ∈ X0 } = αT (x);


(T − T 0 )(x) ≤ inf{(T − T 0 )(x0 ) : x0 ≥ x, x0 ∈ X0 } = (1 − α)T (x).

Therefore, T 0 = αT 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 Ed , the set
comprising all S in E such that the restriction S|dom S is a discrete functional
on dom S. Show that Ed is inductive. To this end, take a chain E0 in Ed and put
S := ∪{S0 : S0 ∈ E0 }; obviously, S ∈ E . Verifying the discreteness of S will
complete the proof.
Suppose that 0 ≤ S 0 (x0 ) ≤ S(x0 ) for all x0 ∈ (dom S)+ and S 0 ∈ (dom S)# .
If S(x0 ) = 0 for all x0 then S 0 = 0S, as was needed. In the case S(x0 ) 6= 0 for
some x0 ∈ (dom S)+ choose S0 ∈ E0 such that S0 (x0 ) = S(x0 ). Since S0 is discrete,
deduce that S 0 (x0 ) = αS(x0 ) for all x0 ∈ dom S0 . Furthermore, α = S 0 (x0 )/S(x0 );
i.e., α does not depend on the choice of S0 . Since E0 is a chain, infer that S 0 = αS. B

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 effective:
α(+∞) := +∞ (α ∈ R+ ) and +∞ + x := x + (+∞) := +∞ (x ∈ R· ).
28 Chapter 3

3.4.2. Definition. Let f : X0 = 0 O R· be a mapping (also called an extended


function). The epigraph of f is the set

0 = 0 OO := {(x, t) ∈ X × R : t ≥ f (x)}.

The effective domain of definition or simply the domain of f is the set

dom f := {x ∈ X : f (x) < +∞}.

3.4.3. Remark. Inconsistency in overusing the symbol dom f is illusory. Na-


mely, the effective domain of definition of f : X0 = 0 O R· coincides with the domain
of definition of the single-valued correspondence f ∩ X × R. We thus continue to
write f : X0 = 0 O R, omitting the dot in the symbol R· whenever dom f = X.
3.4.4. Definition. If X is a real vector space then a mapping f : X0 = 0 O R·
is a convex function provided that the epigraph 0 = 0 OO is convex.
3.4.5. A function f : X0 = 0 O R· is convex if and only if the Jensen inequality
holds:
f (α1 x1 + α2 x2 ) ≤ α1 f (x1 ) + α2 f (x2 )
for all α1 , α2 ≥ 0, α1 + α2 = 1 and x1 , x2 ∈ X.
C ⇒: Take α1 , α2 ≥ 0, α1 + α2 = 1. If either of x1 and x2 fails to be-
long to dom f then the Jensen inequality is evident. Let x1 , x2 ∈ dom f . Then
(x1 , f (x1 )) ∈ 0 = 0 OO and (x2 , f (x2 )) ∈ 0 = 0 OO . Appealing to 3.1.2 (8), find
α1 (x1 , f (x1 )) + α2 (x2 , f (x2 )) ∈ 0 = 0 OO .
⇐: Take (x1 , t1 ) ∈ 0 = 0 OO and (x2 , t2 ) ∈ 0 = 0 OO , i.e. t1 ≥ f (x1 ) and
t2 ≥ f (x2 ) (if dom f = ∅ then f (x) = +∞ (x ∈ X) and 0 = 0 OO = ∅). Applying
the Jensen inequality, observe the containment (α1 x1 +α2 x2 , α1 t1 +α2 t2 ) ∈ 0 = 0 OO
for all α1 , α2 ≥ 0, α1 + α2 = 1. B
3.4.6. Definition. A mapping p : X0 = 0 O R· is a sublinear functional pro-
vided that 0 = 0 OO is a cone.
3.4.7. If dom p 6= 0 then the following statements are equivalent:
(1) p is a sublinear functional;
(2) p is a convex function that is positively homogeneous:
p(αx) = αp(x) for all α ≥ 0 and x ∈ dom p;
(3) if α1 , α2 ∈ R+ and x1 , x2 ∈ X, then p(α1 x1 + α2 x2 ) ≤
α1 p(x1 ) + α2 p(x2 );
(4) p is a positively homogeneous functional that is subadditive:
p(x1 + x2 ) ≤ p(x1 ) + p(x2 ) (x1 , x2 ∈ X). CB
Convex Analysis 29

3.4.8. Examples.
(1) A linear functional is sublinear; an affine functional is a convex
function.
(2) Let U be a convex set in X. Define the indicator function δ(U ) :
X0 = 0 O R· of U as the mapping

if x ∈ U

0,
δ(U )(x) :=
+∞, if x 6∈ U .

It is clear that δ(U ) is a convex function. If U is a cone then δ(U ) is a sublinear


function. If U is an affine set then δ(U ) is an affine function.
(3) The sum of finitely 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 affine operator, i.e.
an everywhere-defined single-valued affine 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 = ∪n∈N nU .
3.4.10. Let T ⊂ X × Y be a linear correspondence with im T = Y . If U is
absorbing (in X) then T (U ) is absorbing (in Y ).
/ Y = T (X) = T (∪n∈N nU ) = ∪n∈N T (nU ) = ∪n∈N 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 : X0 = 0 O R· is a convex function and x ∈ core dom f then for all
h ∈ X there is a limit

f (x + αh) − f (x) f (x + αh) − f (x)


f 0 (x)(h) := lim = inf .
α↓0 α α>0 α

Moreover, the mapping f 0 (x) : h 7→ f 0 (x)h is a sublinear functional f 0 (x) : X0 =


0 O R, the directional derivative of f at x.
C Set ϕ(α) := f (x + αh). By 3.4.8 (4), ϕ : R0 = 0 O R· is a convex function
and 0 ∈ core dom ϕ. The mapping α 7→ (ϕ(α) − ϕ(0))/α (α > 0) is increasing and
bounded above; i.e., ϕ0 (0)(1) is available. By definition, f 0 (x)(h) = ϕ0 (0)(1).
30 Chapter 3

Given β > 0 and h ∈ H, successively find

f (x + αβh) − f (x) f (x + αβh) − f (x)


f 0 (x)(βh) = inf = β inf = βf 0 (x)(h).
α αβ

Moreover, using the above result, for h1 , h2 ∈ X infer that

1

f x + / 2 α(h 1 + h2 ) − f (x)
f 0 (x)(h1 + h2 ) = 2 lim
α↓0 α
1 1

f /2 (x + αh1 ) + /2 (x + αh2 ) − f (x)
= 2 lim
α↓0 α
f (x + αh1 ) − f (x) f (x + αh2 ) − f (x)
≤ lim + lim
α↓0 α α↓0 α
0 0
= f (x)(h1 ) + f (x)(h2 ).

Appealing to 3.4.7 ends the proof. B

3.5. The Hahn–Banach Theorem


3.5.1. Definition. Let X be a real vector space and let f : X0 = 0 O R· be
a convex function. The subdifferential of f at a point x in dom f is the set

∂x (f ) := {l ∈ X # : (∀ y ∈ X) l(y) − l(x) ≤ f (y) − f (x)}.

3.5.2. Examples.
(1) Let p : X0 = 0 O 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 X0 be a subspace of X. Then

∂ (δ(X0 )) = {l ∈ X # : ker l ⊃ X0 }.

(4) If f : X0 = 0 O R· is a convex function then

∂x (f ) = ∂ (f 0 (x))

whenever x ∈ core dom f . CB


Convex Analysis 31

3.5.3. Hahn–Banach Theorem. Let T ∈ L (X, Y ) be a linear operator and


let f : Y 0 = 0 O R· be a convex function. If x ∈ X and T x ∈ core dom f then
∂x (f ◦ T ) = ∂T x (f ) ◦ T.
C By 3.4.10 it follows that x ∈ core dom f . From 3.5.2 (4) obtain ∂x (f ◦ T ) =
∂ ((f ◦ T )0 (x)). Moreover,

(f ◦ T )(x + αh) − (f ◦ T )(x)


(f ◦ T )0 (x)(h) = lim
α↓0 α
f (T x + αT h) − f (T x)
= lim = f 0 (T x)(T h)
α↓0 α
for h ∈ X. Put p := f 0 (T x). By 3.4.12, p is a sublinear functional. Whence, on
appealing again to 3.5.2 (4), infer that

∂ (p) = ∂ (f 0 (T x)) = ∂ T x (f );
∂ (p ◦ T ) = ∂ ((f ◦ T )0 (x)) = ∂x (f ◦ T ).
Thereby, the only claim left unproven is the equality
∂ (p ◦ T ) = ∂ (p) ◦ T.
If l ∈ ∂ (p) ◦ T (i.e., l = l1 ◦ T , where l1 ∈ ∂ (p)) then l1 (y) ≤ p(y) for all y ∈ Y .
In particular, l(x) ∈ l1 (T x) ≤ p(T x) = 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 T x = 0 then l(x) ≤ p(T x) = 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 = l1 ◦ T for some l1 ∈ Y # . Putting Y0 := T (X), observe that the
functional l1 ◦ ι belongs to ∂ (p ◦ ι), with ι the identical embedding of Y0 into Y .
With ∂ (p ◦ ι) ⊂ ∂ (p) ◦ ι proven, observe that l1 ◦ ι = l2 ◦ ι for some l2 ∈ ∂ (p), which
consequently yields l = l1 ◦ T = l1 ◦ ι ◦ T = l2 ◦ ι ◦ T = l2 ◦ T ; i.e., l ∈ ∂ (p) ◦ T .
Thus, to complete the proof of the Hahn–Banach Theorem, showing that ∂ (p ◦
ι) ⊂ ∂ (p) ◦ ι is in order.
Take an element l0 in ∂ (p◦ι) and consider the functional T0 : (y0 , t) 7→ t−l0 (y0 )
given on the subspace Y0 := Y0 × R of Y := Y × R. Equip Y with the cone
Y+ := 0 = 0 OO . Note that, first, Y0 is massive since
(y, t) = (0, t − p(y)) + (y, p(y)) (y ∈ Y, t ∈ R).
Second, by 3.4.2, t ≥ p(y0 ) for (y0 , t) ∈ Y0 ∩ Y+ , and so T0 (y0 , t) = t − l0 (y0 ) ≥ 0;
i.e., T0 is a positive functional on Y0 . By 3.3.4, there is a positive functional T
defined on Y which is an extension of T0 . Put l(y) := T (−y, 0) for y ∈ Y . It is
clear that l ◦ ι = l0 . Furthermore, T (0, t) = T0 (0, t) = t. Hence, 0 ≤ T (y, p(y)) =
p(y) − l(y), i.e., l ∈ ∂ (p). B
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 subdifferential sign or the Hahn–Banach
Theorem in subdifferential form. Observe that the inclusion ∂ (p ◦ ι) ⊂ ∂ (p) ◦ ι is
often referred to as the Hahn–Banach 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 X0 is a subspace of a vector space X and p : X0 = 0 O R
is a sublinear functional then the (asymmetric) Hahn–Banach formula holds:

∂ (p + δ(X0 )) = ∂ (p) + ∂ (δ(X0 )).

C It is obvious that the left side includes the right side. To prove the reverse
inclusion, take l ∈ ∂ (p + δ(X0 )). Then l ◦ ι ∈ ∂ (p ◦ ι), where ι is the identical
embedding of X0 into X. By 3.5.3, l ◦ ι ∈ ∂ (p) ◦ ι, i.e., l ◦ ι = l1 ◦ ι for some
l1 ∈ ∂ (p). Put l2 := l − l1 . By definition, find l2 ◦ ι = (l − l1 ) ◦ ι = l ◦ ι − l1 ◦ ι = 0,
i.e., ker l2 ⊃ X0 . As was mentioned in 3.5.2 (3), this means l2 ∈ ∂ (δ(X0 )). B
3.5.6. Corollary. If f : X0 = 0 O R· is a linear functional and x ∈ core dom f
then ∂x (f ) 6= ∅.
C Let p := f 0 (x) and let ι : 00 = 0 O be the identical embedding of 0 to X.
Plainly, 0 ∈ ∂ (p ◦ ι), i.e., ∂ (p ◦ ι) 6= ∅. By the Hahn–Banach Theorem, ∂ (p) 6= ∅
(otherwise, ∅ = ∂ (p) ◦ ι = ∂ (p ◦ ι)). To complete the proof, apply 3.5.2 (4). B
3.5.7. Corollary. Let f1 , f2 : X0 = 0 O R· be convex functions. Assume
further that x ∈ core dom f1 ∩ core dom f2 . Then

∂x (f1 + f2 ) = ∂x (f1 ) + ∂x (f2 ).

C Let p1 := f10 (x) and p2 := f20 (x). Given x1 , x2 ∈ X, define p(x1 , x2 ) :=


p1 (x1 ) + p2 (x2 ) and ι(x1 ) := (x1 , x1 ). Using 3.5.2 (4) and 3.5.3, infer that

∂x (f1 + f2 ) = ∂ (p1 + p2 ) = ∂ (p ◦ ι)
= ∂ (p) ◦ ι = ∂ (p1 ) + ∂ (p2 ) = ∂x (f1 ) + ∂x (f2 ). .

3.5.8. Remark. Corollary 3.5.6 is sometimes called the Nonempty Subdiffer-


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: Define p T (y) := inf{p(y + T x) − l(x) : x ∈ X},
where l ∈ ∂ (p) and the notation of 3.5.3 is employed. Check that pT is sublinear
and every l1 in ∂ (p T ) satisfies the equality l = l1 ◦ T . Thus, the Nonempty Subdif-
ferential Theorem and the Hahn–Banach Theorem in subdifferential form constitute
a precious (rather than vicious) circle.
Convex Analysis 33

3.6. The Kreı̆n–Milman Theorem


3.6.1. Definition. Let X be a real vector space. Putting

seg(x1 , x2 ) := {α1 x1 + α2 x2 : α1 , α2 > 0, α1 + α2 = 1},

define 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 segU is the restriction of seg
to U 2 , then a convex subset V of U is extreme in U if seg−1 2
U (V ) ⊂ V . 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 u1 , u2 ∈ U and
α1 , α2 > 0, α1 + α2 = 1, the containment α1 u1 + α2 u2 ∈ V implies that u1 ∈ V
and u2 ∈ V . CB
3.6.3. Examples.
(1) Let p : X0 = 0 O R· be a sublinear functional and let a point x of X
belong to dom p. Then ∂x (p) is an extreme subset of ∂ (p).
C For, if α1 l1 + α2 l2 ∈ ∂x (p) and l1 , l2 ∈ ∂ (p), where α1 , α2 > 0, α1 + α2 = 1;
then 0 = p(x) − (α1 l1 (x) + α2 l2 (x)) = α1 (p(x) − l1 (x)) + α2 (p(x) − l2 (x)) ≥ 0.
Moreover, p(x)−l1 (x) ≥ 0 and p(x)−l2 (x) ≥ 0. Hence, l1 ∈ ∂x (p) and l2 ∈ ∂x (p). B
(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 . CB
(3) If X is an ordered vector space then a positive element x of X is
discrete if and only if the cone {αx : α ∈ R+ } is an extreme subset of X+ .
C ⇐: If 0 ≤ y ≤ x then x = 1 /2 (2y) + 1 /2 (2(x − y)). Therefore, by 3.6.2,
2y = αx and 2(x − y) = βx for some α, β ∈ R+ . Thus, 2x = (α + β)x. The case
x = 0 is trivial. Now if x 6= 0 then α /2 ∈ [0, 1] and, consequently, [0, x] ⊂ [0, 1]x.
The reverse inclusion is evident.
⇒: Let [0, x] = [0, 1]x, and suppose that αx = α1 y1 +α2 y2 for α ≥ 0; α1 , α2 >
0, α1 + α2 = 1 and y1 , y2 ∈ X+ . If α = 0 then α1 y1 ∈ [0, x] and α2 y2 ∈ [0, x];
hence, y1 is a positive multiple of x; the same is valid for y2 . In the case α > 0
observe that (α1 /α )y1 = tx for some t ∈ [0, 1]. Finally, (α2 /α )y2 = (1 − t)x. B
(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 . CB
3.6.4. Extreme and Discrete Lemma. Let p : X0 = 0 O R be a sublinear
functional and l ∈ ∂ (p). Assign := X × R, + := = OO , and Tl : (x, t) 7→ t − l(x) (x ∈
X, t ∈ R). Then the functional l is an extreme point of ∂ (p) if and only if Tl is
a discrete functional.
34 Chapter 3

C ⇒: Consider T 0 ∈# such that T 0 ∈ [0, Tl ]. Put

t1 := T 0 (0, 1), l1 (x) := T 0 (−x, 0);


t2 := (Tl − T 0 )(0, 1), l2 (x) := (Tl − T 0 )(−x, 0).

It is clear that t1 ≥ 0, t2 ≥ 0, t1 + t2 = 1; l1 ∈ ∂ (t1 p), l2 ∈ ∂ (t2 p), and l1 + l2 = l.


If t1 = 0 then l1 = 0; i.e., T 0 = 0 and T 0 ∈ [0, 1]Tl . Now if t2 = 0 then t1 = 1; i.e.,
T 0 = Tl , and so T 0 ∈ [0, 1]Tl . Assume that t1 , t2 > 0. In 1
 this case /t1 l1 ∈ ∂ (p)
1 1 1
and /t2 l2 ∈ ∂ (p); moreover, l = t1 /t1 l1 + t2 /t2 l2 . Since, by hypothesis,
l ∈ ext ∂ (p), from 3.6.2 it follows that l1 = t1 l; i.e., T 0 = t1 Tl . B
⇐: Let l = α1 l1 + α2 l2 , where l1 , l2 ∈ ∂ (p) and α1 , α2 > 0, α1 + α2 = 1. The
functionals T 0 := α1 Tl1 and T 00 := α2 Tl2 are positive; moreover, T 0 ∈ [0, Tl ], since
T 0 + T 00 = Tl . Therefore, there is some β ∈ [0, 1] such that T 0 = βTl . At the point
(0, 1), find α1 = β. Hence, l1 = l. By analogy, l2 = l. CB
3.6.5. Kreı̆n–Milman Let T ∈ L (X, Y ) be a linear operator and let f :
Y 0 = 0 O R· be a convex function. If x ∈ X and T x ∈ core dom f then

ext ∂x (f ◦ T ) ⊂ (ext ∂ T x (f )) ◦ T.

C Start arguing as in the proof of the Hahn–Banach Theorem: Put p := f 0 (T x)


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 f0 the restriction of f to
Y0 := im T and notice that f0 ∈ ext ∂ (p ◦ ι), with ι the identical imbedding of Y0
into Y .
Now in Y := Y ×R consider Y+ := 0 = 0 OO and introduce the space Y0 := Y0 ×R.
Note that Y+ ∩ Y0 = 0 = 0 OO p ◦ ι). Applying 3.6.4 with X := Y0 , l := f0 , and
p := p ◦ ι, observe that Tf0 is a discrete functional on Y0 ; moreover, Y0 is massive
in Y (cf. the proof of 3.5.3). By 3.3.8, find a discrete extension S ∈ Y# of Tf0 .
Evidently, S = Tg , 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 (T x) = f0 (T x) = g(T x) for all
x ∈ X. Finally, l ∈ (ext ∂ (p)) ◦ T . B
3.6.6. Corollary. If p : X0 = 0 O 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).
C From 3.6.5 it is easy that ext ∂ (p) 6= ∅ for every p (cf. 3.5.6). Using
this and 3.4.12, choose l in ext ∂x (p0 (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). B
Convex Analysis 35

3.7. The Balanced Hahn–Banach Theorem


3.7.1. Definition. Let (X, F, +, · ) be a vector space over a basic field F.
The vector space (X, R, +, · |R ×X ) is called the real carrier or realification
of (X, F, +, · ) and is denoted by XR .
3.7.2. Definition. Given a linear functional f on a vector space X, define
the mapping Re f : x 7→ Re f (x) (x ∈ X). The real part map or realifier is the
mapping Re : (X # )R 0 = 0 O XR )# .
3.7.3. The real part map Re is a linear operator and, moreover, an isomor-
phism of (X # )R onto (XR )# .
C 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 Re f = 0 then 0 = Re f (ix) = Re(if (x)) = Re(i(Re f (x) + i Im f (x))) =
− Im f (x). Hence, f = 0 and so Re is a monomorphism.
Now take g ∈ (XR )# and put f (x) := g(x)−ig(ix). Evidently, f ∈ L (XR , C R )
and Re f (x) = g(x) for x ∈ X. It is sufficient 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. B
3.7.4. Definition. The lear trap map or complexifier is the inverse Re−1 :
(XR )# 0 = 0 O X # )R of the real part map.
3.7.5. Remark. By 3.7.3, for the complex scalar field

Re−1 g : x 7→ g(x) − ig(ix) (g ∈ (XR )# , x ∈ X).

In the case of the reals, F := R, the complexifier 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 : X0 = 0 O R· such that dom p 6= ∅ and

p(λ1 x1 + λ2 x2 ) ≤ |λ1 |p(x1 ) + |λ2 |p(x2 )

for x1 , x2 ∈ X and λ1 , λ2 ∈ 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 : X0 = 0 O R· is a seminorm then the balanced subdif-


ferential of p is the set
|∂ |(p) := {l ∈ X # : |l(x)| ≤ p(x) for all x ∈ X}.
3.7.9. Balanced Subdifferential Lemma. Let p : X0 = 0 O R· be a semi-
norm. Then
|∂ |(p) = Re−1 (∂ (p)); Re (|∂ |(p)) = ∂ (p)
for the subdifferentials |∂ |(p) and ∂ (p) of p.
C 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 (Re l)(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) 6= 0 set θ := |f (x)|/f (x). Thereby |f (x)| = θf (x) = f (θx) =
Re f (θx) = g(θx) ≤ p(θx) = |θ|p(x) = p(x), since |θ| = 1. Finally, f ∈ |∂ |(p). B
3.7.10. Let X be a vector space, let p : X0 = 0 O R be a seminorm, and let X0
be a subspace of X. The asymmetric balanced Hahn–Banach formula holds:
|∂ |(p + δ(X0 )) = |∂ |(p) + |∂ |(δ(X0 )).
C From 3.7.9, 3.5.5, and the results of Section 3.1 obtain

|∂ |(p + δ(X0 )) = Re−1 (∂ (p + δ(X0 ))) = Re−1 (∂ (p) + ∂ (δ(X0 )))


= Re−1 (∂ (p)) + Re−1 (∂ (δ(X0 ))) = |∂ |(p) + |∂ |(δ(X0 )). .
3.7.11. If X and Y are vector spaces, T ∈ L (X, Y ) is a linear operator, and
p : Y 0 = 0 O R is a seminorm; then p ◦ T is also a seminorm and, moreover,
|∂ |(p ◦ T ) = |∂ |(p) ◦ T.
C Applying 2.3.8 and 3.7.10, successively infer that
|∂ |(p ◦ T ) = |∂ |(p + δ(im T )) ◦ T = (|∂ |(p) + |∂ |(δ(im T ))) ◦ T
= |∂ |(p) ◦ T + |∂ |(δ(im T )) ◦ T = |∂ |(p) ◦ T. .
3.7.12. Remark. If T is the identical embedding of a subspace and the ground
field is C then 3.7.11 is referred to as the Sukhomlinov–Bohnenblust–Sobczyk Theo-
rem.
3.7.13. Balanced Hahn–Banach Let X be a vector space. Assume further
that p : X0 = 0 O R is a seminorm and X0 is a subspace of X. If l0 is a linear
functional given on X0 such that |l0 (x0 )| ≤ p(x0 ) for x0 ∈ X0 , then there is a linear
functional l on X such that l(x0 ) = l0 (x0 ) whenever x0 ∈ X0 and |l(x)| ≤ p(x) for
all x ∈ X. CB
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 : X0 = 0 O 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 7→ Ut (t ∈ T ) be
a family of subsets of X. There is a function f : X0 = 0 O R such that
{f < t} ⊂ Ut ⊂ {f ≤ t} (t ∈ T )
if and only if the mapping t 7→ Ut increases.
C ⇒: Suppose that T contains at least two elements s and t. (Otherwise there
is nothing left to proof.) If s < t then
Us ⊂ {f ≤ s} ⊂ {f < t} ⊂ Ut .
⇐: Define a mapping f : X0 = 0 O R by setting f (x) := inf{t ∈ T : x ∈ Ut }.
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 ∈ Us and s < t. Thus, {f < t} ⊂ Us ⊂ Ut .
Continuing, for x ∈ Ut find f (x) ≤ t; i.e., Ut ⊂ {f ≤ t}. B
3.8.3. Function Comparison Lemma. Let f, g : X0 = 0 O R be functions
defined by (Ut )t∈T and (Vt )t∈T as follows:
{f < t} ⊂ Ut ⊂ {f ≤ t};
{g < t} ⊂ Vt ⊂ {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
X
f ≤ g (in R ; i.e., f (x) ≤ g(x) for x ∈ X) holds if and only if
(t1 , t2 ∈ T & t1 < t2 ) ⇒ Vt1 ⊂ Ut2 .
C ⇒: This is immediate from the inclusions
Vt1 ⊂ {g ≤ t1 } ⊂ {f ≤ t1 } ⊂ {f < t2 } ⊂ Ut2 .
⇐: Assume that g(x) 6= +∞ (if not, f (x) ≤ g(x) for obvious reasons). Given
t ∈ R such that g(x) < t < +∞, choose t1 , t2 ∈ T so as to satisfy the conditions
g(x) < t1 < t2 < t. Now
x ∈ {g < t1 } ⊂ Vt1 ⊂ Ut2 ⊂ {f ≤ t2 } ⊂ {f < t}.
Thus, f (x) < t. Since t is arbitrary, obtain f (x) ≤ g(x). B
38 Chapter 3

3.8.4. Corollary. If T is dense in R and the mapping t 7→ Ut (t ∈ T ) increases


then there is a unique function f : X0 = 0 O R such that

{f < t} ⊂ Ut ⊂ {f ≤ t} (t ∈ T ).

Moreover, the level sets of f may be presented as follows:

{f < t} = ∪ {Us : s < t, s ∈ T };

{f ≤ t} = ∩{Ur : t < r, r ∈ T } (t ∈ R).


C It is immediate from 3.8.2 and 3.8.3 that f exists and is unique. If s < t and
s ∈ T then Us ⊂ {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} ⊂ Us , which
proves the formula for {f < t}.
Suppose that r > t, r ∈ T . Then {f ≤ t} ⊂ {f < r} ⊂ Ur . In turn, if x ∈ Ur
for r ∈ T, r > t; then f (x) ≤ r for all r > t. Hence, f (x) ≤ t. B
3.8.5. Let X be a vector space and let S be a conical segment in X. Given
t ∈ R, put Ut := ∅ if t < 0 and Ut := tS if t ≥ 0. The mapping t 7→ Ut (t ∈ R)
increases.
C If 0 ≤ t1 < t2 and x ∈ t1 S then x ∈ (t1 /t2 ) t2 S. Hence, x ∈ t2 S. B
3.8.6. Definition. The Minkowski functional or the gauge function or simply
the gauge of a conical segment S is a functional pS : X0 = 0 O R such that

{pS < t} ⊂ tS ⊂ {pS ≤ 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,
pS (x) = inf{t > 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}.
C Consider a conical segment S and its Minkowski functional pS . Let x ∈ X.
The inequality pS (x) ≥ 0 is evident. Take α > 0. Then

pS (αx) = inf{t > 0 : αx ∈ tS} = inf t > 0 : x ∈ t /α S




= inf{αβ > 0 : x ∈ βS, β > 0}


= α inf{β > 0 : x ∈ βS} = αpS (x).
Convex Analysis 39

To start checking that pS is subadditive, take x1 , x2 ∈ X. Noting the inclusion


t1 S + t2 S ⊂ (t1 + t2 )S for t1 , t2 > 0, in view of the identity
 
t1 t2
t1 x1 + t2 x2 = (t1 + t2 ) x1 + x2 ,
t1 + t2 t1 + t2

successively infer that

pS (x1 + x2 ) = inf{t > 0 : x1 + x2 ∈ tS}


≤ inf{t : t = t1 + t2 ; t1 , t2 > 0, x1 ∈ t1 S, x2 ∈ t2 S}
= inf{t1 > 0 : x1 ∈ t1 S} + inf{t2 > 0 : x2 ∈ t2 S} = pS (x1 ) + pS (x2 ).

Let p : X0 = 0 O R· be an arbitrary sublinear functional with positive values


and let {p < 1} ⊂ S ⊂ {p ≤ 1}. Given t ∈ R+ , put Vt := {p < t} and Ut := tS;
given t < 0, put Vt := Ut := ∅. Plainly,

{pS < t} ⊂ Ut ⊂ {pS ≤ t}; {p < t} ⊂ Vt ⊂ {p ≤ t}

for t ∈ R. If 0 ≤ t1 < t2 then Vt1 = {p < t1 } = t1 {p < 1} ⊂ t1 S = Ut1 ⊂ Ut2 .


Moreover, Ut1 ⊂ t1 {p ≤ 1} ⊂ {p ≤ t1 } ⊂ {p < t2 } ⊂ Vt2 . Therefore, by 3.8.3 and
3.8.4, p = pS . B
3.8.8. Remark. A conical segment S in X is absorbing in X if and only
if dom pS = X. Also, pS is a seminorm whenever S is absolutely convex. Conversely,
for every seminorm p the sets {p < 1} and {p ≤ 1} are absolutely convex. CB
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 field of X. An element of X/H is called
a hyperplane in X parallel to H. By a hyperplane in X we mean an affine variety
parallel to some hypersubspace of X. An affine variety H is a hyperplane if H −h 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 # . CB
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 affine variety in X. If L ∩ U = ∅
then there is a hyperplane H in X such that H ⊃ L and H ∩ core U = ∅.
C Without loss of generality, it may be supposed that core U 6= ∅ (otherwise
there is nothing left unproven) and, moreover, 0 ∈ core U . Take x ∈ L and put X0 :=
L − x. Consider the quotient space X/X0 and the corresponding coset mapping
ϕ : X0 = 0 O /X0 . Applying 3.1.8 and 3.4.10, observe that ϕ(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 := pϕ(U ) is the quotient X/X0 ; moreover,

ϕ(core U ) ⊂ core ϕ(U ) ⊂ {p < 1} ⊂ ϕ(U ).

In particular, this entails the inequality p(ϕ(x)) ≥ 1, since ϕ(x) 6∈ ϕ(U ).


Using 3.5.6, find a functional f in ∂x (p ◦ ϕ); now the Hahn–Banach Theorem
implies
f ∈ ∂x (p ◦ ϕ) = ∂ ϕ(x) (p) ◦ ϕ.

Put H := {f = p ◦ ϕ(x)}. It is clear that H is a real hyperplane and, undoubtedly,


H ⊃ L. Appealing to 3.5.2 (1), conclude that H ∩ core 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. B
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 Hahn–Banach Theorem in geometric form or as the Minkowski–Ascoli–Mazur
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 different halfspaces defined by H; i.e.,
if there is a presentation H = {f = t}, where f ∈ (XR )# 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 6= ∅ and U ∩ core V = ∅, then there is a hyperplane separating U
and V and disjoint from core V . CB

Exercises

3.1. Show that a hyperplane is precisely an affine set maximal by inclusion and other than
the whole space.
3.2. Prove that every affine 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 finite set. What use
can be made of finite dimensions?
S
3.5. Given sets S1 and S2 , let S := 0≤λ≤1
λS1 ∩(1−λ)S2 . Prove that S is convex whenever
so are S1 and S2 .
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 Sp and Sq . Express S via Sp and Sq .

3.8. Describe sublinear functionals with domain RN .


3.9. Calculate the subdifferential of the upper envelope of a finite set of linear or sublinear
functionals.
3.10. Let p and q be sublinear functionals in general position, i.e. such that dom p−dom q =
dom q − dom p. Prove the symmetric Hahn–Banach formula (cf. 3.5.7): ∂ (p + q) = ∂ (p) + ∂ (q).
3.11. Let p, q : X0 = 0 O R be total (= everywhere-defined) 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 define the structure of a Kantorovich space?
3.14. Each ordered vector space admits a nonzero positive linear functional, doesn’t it?
3.15. What are the means for transforming RN into a Kantorovich space?
3.16. Under what conditions does the claim of the Hahn–Banach Theorem in analytical
form hold for a partial (= not-everywhere-defined) sublinear functional?
3.17. Find the extreme points of the subdifferential of the conventional norm on l∞ .
3.18. Find possible generalizations of the Hahn–Banach Theorem for a mapping acting into
a Kantorovich space.
3.19. Given a set C in a space X, define the Hörmander transform H(C) of C as

H(C) = {(x, t) ∈ X × R : x ∈ tC}.

Study the properties of the Hörmander 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 0 = 0 O 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 0 = 0 O R+ is a metric if and only if


(1) {d ≤ 0} = IX ;
(2) {d ≤ t} = {d ≤ t}−1 (t ∈ R+ );
(3) {d ≤ t1 } ◦ {d ≤ t2 } ⊂ {d ≤ t1 + t2 } (t1 , t2 ∈ R+ ).
C Items 4.1.2 (1)–4.1.2 (3) reformulate 4.1.1 (1)–4.1.1 (3). B

4.1.3. Definition. Let (X, d) be a metric space and take ε ∈ R+ \0, a strictly
positive real. The relation Bε := Bd,ε := {d ≤ ε} is the closed cylinder of size ε. The
◦ ◦
set B ε := B d,ε := {d < ε} is the open cylinder of size ε. The image Bε (x) of a point x
under the relation Bε is called the closed ball with center x and radius ε. By analogy,

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

4.1.4. In a nonempty metric space open cylinders as well as closed cylinders


form bases for the same filter. CB
An Excursus into Metric Spaces 43

4.1.5. Definition. The filter on X 2 with the filterbase of all cylinders of


a nonempty metric space (X, d) is the metric uniformity on X. It is denoted
by UX , Ud , or even U , if the space under consideration is implied. Given X := ∅,
put UX := {∅}. An element of UX is an entourage on X.
4.1.6. If U is a metric uniformity then
(1) U ⊂ fil {IX };
(2) U ∈ U ⇒ U −1 ∈ U ;
(3) (∀ U ∈ U ) (∃ V ∈ U ) V ◦ V ⊂ U ;
(4) ∩{U : U ∈ U } = IX . CB
4.1.7. Remark. Property 4.1.6 (4) reflecting 4.1.1 (1) is often expressed as
follows: “U is a Hausdorff or separated uniformity.”
4.1.8. Given a space X with uniformity UX , put

τ (x) := {U (x) : U ∈ U }.

Then τ (x) is a filter for every x in X. Moreover,


(1) τ (x) ⊂ fil {x};
(2) (∀ U ∈ τ (x)) (∃ V ∈ τ (x) & V ⊂ U ) (∀ y ∈ V ) V ∈ τ (y). CB
4.1.9. Definition. The mapping τ : x 7→ τ (x) is the metric topology on X.
An element of τ (x) is a neighborhood of x or about x. More complete designations
for the topology are also in use: τX , τ (U ), etc.
4.1.10. Remark. All closed balls centered at x form a base for the neighbor-
hood filter of x. The same is true of open balls. Note also that there are disjoint
(= nonintersecting) neighborhoods of different points in X. This property, ciphered
in 4.1.6 (4), reads: “τX is a Hausdorff 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(τ ) ⇔ ((∀ x ∈ G) G ∈ τ (x))).
A subset F of X is a closed set in X whenever its complement to X is open
(in symbols, F ∈ Cl(τ ) ⇔ (X \ F ∈ Op(τ ))).
4.1.12. The union of a family of open sets and the intersection of a finite
family of open sets are open. The intersection of a family of closed sets and the
union of a finite family of closed sets are closed. CB
4.1.13. Definition. Given a subset U of X, put

int U := U := ∪{G ∈ Op(τX ) : G ⊂ U };
cl U := U := ∩{F ∈ Cl(τX ) : 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 . CB
An Excursus into Metric Spaces 45

4.1.15. Remark. In connection with 4.1.14, the set Op(τX ) is also referred
to as the topology of the space X under study, since τX is uniquely determined
from Op(τX ). The same relates to Cl(τX ), the collection of all closed sets in X.

4.1.16. Definition. A filterbase B on X converges to x in X or x is a limit


of B (in symbols, B0 = 0 O ) if fil B is finer than the neighborhood filter of x; i.e.,
fil B ⊃ τ (x).

4.1.17. Definition. A net or (generalized) sequence (xξ )ξ∈„ converges to x


(in symbols, xξ 0 = 0 O ) if the tail filter of (xξ ) converges to x. Other familiar terms
and designations are freely employed. For instance, x = limξ xξ and x is a limit
of (xξ ) as ξ ranges over „.

4.1.18. Remark. A limit of a filter, 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. CB

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 filter F such that F 0 = 0 O and U ∈ F ;
(3) there is a sequence (xξ )ξ∈„ whose entries are in U and which con-
verges to x.
C (1) ⇒ (2): Since x is not exterior to U , the join F := τ (x)∨fil {U } is available
of the pair of the filters τ (x) and fil {U }.
(2) ⇒ (3): Let F 0 = 0 O and U ∈ F . Direct F by reverse inclusion. Take
xV ∈ V ∩ U for V ∈ F . It is clear that xV 0 = 0 O .
(3) ⇒ (1): Let V be a closed set. Take a sequence (xξ )ξ∈„ in V such that
xξ 0 = 0 O . In this case it suffices to show that x ∈ V , which is happily evident.
Indeed, were x in X \ V we would find at least one ξ ∈ „ such that xξ ∈ X \ V . B

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 first axiom of countability is fulfilled.”

4.2. Continuity and Uniform Continuity

4.2.1. If f : X0 = 0 O and τX and τY are topologies on X and Y then the


following conditions are equivalent:
46 Chapter 4

(1) G ∈ Op(τY ) ⇒ f −1 (G) ∈ Op(τX );


(2) F ∈ Cl(τY ) ⇒ f −1 (F ) ∈ Cl(τX );
(3) f (τX (x)) ⊃ τY (f (x)) for all x ∈ X;
(4) (x ∈ X & F 0 = 0 O ) ⇒ (f (F )0 = 0 O (x)) for a filter F ;
(5) f (xξ )0 = 0 O (x) for every point x and every sequence (xξ ) conver-
gent to x.
C The equivalence (1) ⇔ (2) follows from 4.1.11. It remains to demonstrate
that (1) ⇒ (3) ⇒ (4) ⇒ (5) ⇒ (2).
(1) ⇒ (3): If V ∈ τY (f (x)) then W := int V ∈ Op(τY ) and f (x) ∈ W . Hence,
f (W ) ∈ Op(τX ) and x ∈ f −1 (W ). In other words, f −1 (W ) ∈ τX (x) (see 4.1.14).
−1

Moreover, f −1 (V ) ⊃ f −1 (W ) and, consequently, f −1 (V ) ∈ τX (x). Finally, V ⊃


f (f −1 (V )).
(3) ⇒ (4): Given F 0 = 0 O , by Definition 4.1.16 fil F ⊃ τX (x). From the
hypothesis infer that f (F ) ⊃ f (τX (x)) ⊃ τY (f (x)). Appealing to 4.1.16, again
reveal the sought instance of convergence, f (F )0 = 0 O (x).
(4) ⇒ (5): The image of the tail filter of (xξ )ξ∈„ under f is coarser than the
tail filter of (f (xξ ))ξ∈„ .
(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 (xξ )ξ∈„ converging to x and consisting of points in f −1 (F ) (the
claim of 4.1.19 yields existence). Then f (xξ ) ∈ F and f (xξ )0 = 0 O (x). Another
citation of 4.1.19 guarantees that f (x) ∈ F and, consequently, x ∈ f −1 (F ). B
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 fixed 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.
C Apply 4.2.1 (5) thrice. B
4.2.4. Let f : X0 = 0 O and let UX and UY be uniformities on X and Y . The
following statements are equivalent:
(1) (∀ V ∈ UY ) (∃ U ∈ UX ) ((∀ x, y)(x, y) ∈ U ⇒ (f (x), f (y)) ∈ V );
(2) (∀ V ∈ UY ) f −1 ◦ V ◦ f ∈ UX ;
(3) f × (UX ) ⊃ UY , with f × : X 2 0 = 0 O defined as f × : (x, y) 7→
(f (x), f (y));
(4) (∀ V ∈ UY ) f ×−1 (V ) ∈ UX ; i.e., f ×−1 (UY ) ⊂ UX .
An Excursus into Metric Spaces 47

C By 1.1.10, given U ⊂ X 2 and V ⊂ Y 2 , observe that


[
f −1 ◦ V ◦ f = f −1 (v1 ) × f −1 (v2 )
(v1 ,v2 )∈V

= {(x, y) ∈ X 2 : (f (x), f (y)) ∈ V } = f ×−1 (V );


[
f ◦ U ◦ f −1 = f (u1 ) × f (u2 )
(u1 ,u2 )∈U

= {(f (u1 ), f (u2 )) : (u1 , u2 ) ∈ U } = f × (U ). .


4.2.5. Definition. A mapping f : X0 = 0 O 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.
C Consider f : X0 = 0 O , g : Y 0 = 0 O and h := g ◦ f : X0 = 0 O . Plainly,
h× (x, y) = (h(x), h(y)) = (g(f (x)), g(f (y))) = g × (f (x), f (y)) = g × ◦ f × (x, y)
for all x, y ∈ X. Hence, by 4.2.4 (3) h× (UX ) = g × (f × (UX )) ⊃ g × (UY ) ⊃ UZ ,
meaning that h is uniformly continuous. B
4.2.7. Every uniformly continuous mapping is continuous. CB
4.2.8. Definition. Let E be a set of mappings from X into Y and let UX
and UY be the uniformities in X and Y . The set E is equicontinuous if
\
(∀ V ∈ UY ) f −1 ◦ V ◦ f ∈ UX .
f ∈E

4.2.9. Every equicontinuous set consists of uniformly continuous mappings.


Every finite set of uniformly continuous mappings is equicontinuous. CB

4.3. Semicontinuity
4.3.1. Let (X1 , d1 ) and (X2 , d2 ) be metric spaces, and := X × X. Given
x := (x1 , x2 ) and y := (y1 , y2 ), put
d(x, y) := d1 (x1 , y1 ) + d2 (x2 , y2 ).
Then d is a metric on . Moreover, for every x := (x1 , x2 ) in the presentation holds:
τ(x) = fil{U1 × U2 : U1 ∈ τX1 (x1 ), U2 ∈ τX2 (x2 )}. /.
4.3.2. Definition. The topology τ is called the product of τX1 and τX2 or the
product topology of X1 × X2 . This topology on X1 × X2 is denoted by τX1 × τX2 .
48 Chapter 4

4.3.3. Definition. A function f : X0 = 0 O R· is lower semicontinuous if its


epigraph 0 = 0 OO is closed in the product topology of X × R.
4.3.4. Examples.
(1) A continuous real-valued function f : X0 = 0 O R is lower semicon-
tinuous.
(2) If fξ : X0 = 0 O R· is lower semicontinuous for all ξ ∈ „, then the
upper envelope f (x) := sup{fξ (x) : ξ ∈ „} (x ∈ X) is also lower semicontinuous.
A simple reason behind this is the equality 0 = 0 OO = ∩ξ∈„ 0 = 0 OO ξ .
4.3.5. A function f : X0 = 0 O R· is lower semicontinuous if and only if

x ∈ X ⇒ f (x) = lim O inf f (y).


y0=0

Here, as usual,

lim O inf f (y) := lim f (y) := sup inf f (U )


y0=0 y0=0 O U ∈τ (x)

is the lower limit of f at x (with respect to τ (x)).


C ⇒: If x 6∈ dom f then (x, t) 6∈ 0 = 0 OO for all t ∈ R. Hence, there is a neigh-
borhood Ut of x such that inf f (Ut ) > t. This implies that limy0=0 O inf f (y) =
+∞ = f (x). Suppose that x ∈ dom f . Then inf f (V ) > −∞ for some neighbor-
hood V of x. Choose ε > 0 and for an arbitrary U in τ (x) included in V find
xU in U so that inf f (U ) ≥ f (xU ) − ε. By construction, xU ∈ dom f and, more-
over, xU 0 = 0 O (by implication, the set of neighborhoods of x is endowed with
the conventional order by inclusion, cf. 1.3.1). Put tU := inf f (U ) + ε. It is clear
that tU 0 = 0 O := limy0=0 O inf f (y) + ε. Since (xU , tU ) ∈ 0 = 0 OO , from the closure
property of 0 = 0 OO obtain (x, t) ∈ 0 = 0 OO . Thus,

lim O inf f (y) + ε ≥ f (x) ≥ lim O inf f (y).


y0=0 y0=0

⇐: If (x, t) 6∈ 0 = 0 OO then

t < lim O inf f (y) = sup inf f (U ).


y0=0 U ∈τ (x)

Therefore, inf f (U ) > t for some neighborhood U of x. It follows that the comple-
ment of 0 = 0 OO to X × R is open. B
4.3.6. Remark. The property, stated in 4.3.5, may be accepted as an initial
definition of lower semicontinuity at a point.
An Excursus into Metric Spaces 49

4.3.7. A function f : X0 = 0 O R is continuous if and only if both f and −f


are lower semicontinuous. CB
4.3.8. A function f : X0 = 0 O R· is lower semicontinuous if and only if for
every t ∈ R the level set {f ≤ t} is closed.
C ⇒: If x 6∈ {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 limy0=0 O inf f (y) ≤ t < f (x) for some x ∈ X and t ∈ R, choose
ε > 0 such that t + ε < f (x). Repeating the argument of 4.3.5, given U ∈ τ (x) take
a point xU in U ∩ {f ≤ inf f (U ) + ε}. Undoubtedly xU ∈ {f ≤ t + ε} and xU 0 = 0 O ,
a contradiction. B

4.4. Compactness
4.4.1. Definition. A subset C of X is called a compact set whenever for
every subset E of Op(τX ) with the property C ⊂ ∪{G : G ∈ E } there is a finite
subset E0 in E such that still C ⊂ ∪{G : G ∈ E0 }.
4.4.2. Remark. Definition 4.4.1 is verbalized as follows: “A set is compact
if its every open cover has a finite subcover.” The terms “covering” and “subcovering”
are also in current usage.
4.4.3. Every closed subset of a compact set is also compact. Every compact
set is closed. CB
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.
C The inverse images of sets in an open cover of the image compose an open
cover of the original set. B
4.4.6. Each lower semicontinuous real-valued function, defined on a nonempty
compact set, assumes its least value; i.e., the image of a compact domain has a least
element.
C Suppose that f : X0 = 0 O R· and X is compact. Let t0 := inf f (X). In the
case t0 = +∞ there is nothing left to proof. If t0 < +∞ then put T := {t ∈ R :
t > t0 }. The set Ut := {f ≤ t} with t ∈ T is nonempty and closed. Check that
∩{Ut : 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 \ Ut : t ∈ T }
compose an open cover of X. Refining a finite subcover, deduce ∩{Ut : t ∈ T0 } = ∅.
However, this equality is false, since Ut1 ∩ Ut2 = Ut1 ∧t2 for t1 , t2 ∈ T . B
50 Chapter 4

4.4.7. Bourbaki Criterion. A space is compact if and only if every ultrafilter


on it converges (cf. 9.4.4).
4.4.8. The product of compact sets is compact.
C It suffices to apply the Bourbaki Criterion twice. B
4.4.9. Cantor Theorem. Every continuous mapping on a compact set is
uniformly continuous. CB

4.5. Completeness Completeness


4.5.1. If B is a filterbase on X then {B 2 : B ∈ B} is a filterbase (and a base
for the filter B × ) on X 2 .
C (B1 × B1 ) ∩ (B2 × B2 ) ⊃ (B1 ∩ B2 ) × (B1 ∩ B2 ) B
4.5.2. Definition. Let UX be a uniformity on X. A filter F is called
a Cauchy filter or even Cauchy (the latter might seem preposterous) if F × ⊃ UX .
A net in X is a Cauchy net or a fundamental net if its tail filter is a Cauchy filter.
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 Bε -small (or simply ε-small) if and
only if the diameter of U , diam U := sup(U 2 ), is less than or equal to ε. In this
connection, the definition of a Cauchy filter is expressed as follows: “A filter is
a Cauchy filter if and only if it contains arbitrarily small sets.”
4.5.4. In a metric space the following conditions are equivalent:
(1) every Cauchy filter converges;
(2) each Cauchy net has a limit;
(3) every fundamental sequence converges.
C The implications (1) ⇒ (2) ⇒ (3) are obvious; therefore, we are left with
establishing (3) ⇒ (1).
Given a Cauchy filter F , let Un in F be a B1/n -small set. Put Vn := U1 ∩. . .∩Un
and take xn ∈ Vn . Observe that V1 ⊃ V2 ⊃ . . . and diam Vn ≤ 1 /n . Hence, (xn )
is a fundamental sequence. By hypothesis it has a limit, x := lim xn . Check that
F 0 = 0 O . To this end, choose n0 ∈ N such that d(xm , x) ≤ 1 /2n as m ≥ n0 . Then
for all n ∈ N deduce that d(xp , y) ≤ diam Vp ≤ 1 /2n and d(xp , x) ≤ 1 /2n whenever
p := n0 ∨ 2n and y ∈ Vp . It follows that y ∈ Vp ⇒ d(x, y) ≤ 1 /n ; i.e., Vp ⊂ B1/n (x).
In conclusion, F ⊃ τ (x). B
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.
An Excursus into Metric Spaces 51

4.5.6. Cantor Criterion. A metric space X is complete if and only if ev-


ery nonempty downward-filtered family of nonempty closed subsets of X whose
diameters tend to zero has a point of intersection.
C ⇒: If B is such family then by Definition 1.3.1 B is a filterbase. By hy-
pothesis, B is a base for a Cauchy filter. Therefore, there is a limit: B0 = 0 O . The
point x meets the claim.
⇐: Let F be a Cauchy filter. 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 0 = 0 O . Indeed, let V be an ε /2 -small member of F and y ∈ V . Some
y 0 in V is such that d(x, y 0 ) ≤ ε /2 . Therefore, d(x, y) ≤ d(x, y 0 ) + d(y 0 , y) ≤ ε.
Consequently, V ⊂ Bε (x) and so Bε (x) ∈ F . B
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. CB
4.5.8. The image of a Cauchy filter under a uniformly continuous mapping is
a Cauchy filter.
C Let a mapping f act from a space X with uniformity UX into a space Y with
uniformity UY and let F be a Cauchy filter on X. If V ∈ UY then, by Definition
4.2.5, f −1 ◦ V ◦ f ∈ UX (cf. 4.2.4 (2)). Since F is a Cauchy filter, U 2 ⊂ f −1 ◦ V ◦ f
for some U ∈ F . It turns out that f (U ) is V -small. Indeed,
[
f (U )2 = f (u1 ) × f (u2 )
(u1 ,u2 )∈U 2
−1 −1
=f ◦U ◦f2
⊂ f ◦ (f ◦ V ◦ f ) ◦ f −1 = (f ◦ f −1 ) ◦ V ◦ (f ◦ f −1 ) ⊂ V,

because, by 1.1.6, f ◦ f −1 = Iim f ⊂ IY . B


4.5.9. The product of complete spaces is complete.
C The claim is immediate from 4.5.8 and 4.5.4. B
4.5.10. Let X0 be dense in X (i.e., cl X0 = X). Assume further that f0 :
X0 0 = 0 O is a uniformly continuous mapping from X0 into some complete space Y .
Then there is a unique uniformly continuous mapping f : X0 = 0 O extending f0 ;
i.e., f |X0 = f0 .
C For x ∈ X, the filter Fx := {U ∩ X0 : U ∈ τX (x)} is a Cauchy filter on X0 .
Therefore, 4.5.8 implies that f0 (FX ) is a Cauchy filter on Y . By the completeness
of Y , there is a limit y ∈ Y ; that is, f0 (Fx )0 = 0 O . Moreover, this limit is unique
(cf. 4.1.18). Define f (x) := y. Checking uniform continuity for f readily completes
the proof. B
52 Chapter 4

4.5.11. Definition. An isometry or isometric embedding or isometric map-


ping of X into Xe is a mapping f : (X, d)0 = 0 O X, e de) such that d = de ◦ f × .
A mapping f is an isometry onto X e if f is an isometry of X into X
e and, moreover,
im f = X. The expressions, “an isometry of X and X” or “an isometry between X
e e
and X”
e and the like, are also in common parlance.

4.5.12. Hausdorff Completion Theorem. If (X, d) is a metric space then


there are a complete metric space (X, e de) and an isometry ι : (X, d)0 = 0 O X,e de)
onto a dense subspace of (X, e de). The space (X, e de) is unique to within isometry
in the following sense: The diagram
ι
(X, d) - (X,
e de)

ι1
@
‰
@
R
@ ?
(Xe1 , de1 )
commutes for some isometry ‰ : (X, e de)0 = 0 O Xe1 , de1 ), where ι1 : (X, d)0 =
0OXe1 , de1 ) is an isometry of X onto a dense subspace of a complete space (X e1 , de1 ).
C Uniqueness up to isometry follows from 4.5.10. For, if ‰0 := ι1 ◦ ι−1 then
‰0 is an isometry of the dense subspace ι(X) of X e onto the dense subspace ι1 (X)
of Xe1 . Let ‰ be the unique extension of ‰0 to the whole of X. e It is sufficient
to show that ‰ acts onto X1 . Take x
e e1 in X1 . This element is the limit of some
e
sequence (ι1 (xn )), with xn ∈ X. Clearly, the sequence (xn ) is fundamental. Thus,
(ι(xn )) is a fundamental sequence in X. e Let x e := lim ι(xn ), x e ∈ X.
e Proceed as
−1
follows: ‰(e x) = lim ‰0 (ι(xn )) = lim ι1 ◦ ι (ι(xn )) = lim ι1 (xn ) = xe1 .
We now sketch out the proof that X exists. Consider the set of all fundamental
e
sequences in X. Define some equivalence relation in X by putting x1 ∼ x2 ⇔
d(x1 (n), x2 (n))0 = 0 O . Assign Xe := /∼ and d(ϕ(x
e 1 ), ϕ(x2 )) := lim d(x1 (n), x2 (n)),
where ϕ : = O X is the coset mapping. An isometry ι : (X, d)0 = 0 O X, e de) is
immediate: ι(x) := ϕ(n 7→ x (n ∈ N)). B
4.5.13. Definition. The space (X, e de) 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 X0 in a metric space (X, d) is said to be complete
if the metric space (X0 , d|X02 ), a subspace of (X, d), is complete.
4.5.15. Every closed subset of a complete space is complete. Every complete
set is closed. CB
4.5.16. If X0 is a subspace of a complete metric space X then a completion
of X0 is isometric to the closure of X0 in X.
An Excursus into Metric Spaces 53

C Let X e := cl X0 and let ι : X0 0 = 0 O X be the identical embedding. It is


evident that ι is an isometry onto a dense subspace. Moreover, by 4.5.15 X e is
complete. Appealing to 4.5.12 ends the proof. B

4.6. Compactness and Completeness


4.6.1. A compact space is complete. CB
4.6.2. Definition. Let U be a subset of X and V ∈ UX . 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 UX there is a finite 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.
C Let V ∈ UX and W ∈ UX be such that W ◦ W ⊂ V . Take a totally bounded
W -net F for U ; i.e., U ⊂ W (F ). Since F is totally bounded, there is a finite 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 finite V -net for U . B


4.6.5. A subset U of X is totally bounded if and only if for every V in UX
there is a family U1 , . . . , Un of subsets of U such that U = U1 ∪ . . . ∪ Un and each
of the sets U1 , . . . , Un is V -small. CB
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 finite covers consisting of arbitrarily small sets.”
4.6.7. Hausdorff Criterion. A set is compact if and only if it is complete
and totally bounded. CB
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 Chebyshëv metric

d(f, g) := sup dF (f (x), g(x)) := sup |f (x) − g(x)| (f, g ∈ C(Q, F));
x∈Q x∈Q

and, given θ ∈ UF , put

Uθ := (f, g) ∈ C(Q, F)2 : g ◦ f −1 ⊂ θ .




Then Ud = fil {Uθ : θ ∈ UF }. CB


54 Chapter 4

4.6.9. The space C(Q, F) is complete. CB


4.6.10. Ascoli–Arzelà 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.
C ⇒: It is beyond a doubt that ∪{g(Q) : g ∈ E } is totally bounded. To
show equicontinuity for E take θ ∈ UF and choose a symmetric entourage θ0 such
that θ0 ◦ θ0 ◦ θ0 ⊂ θ. By the Hausdorff Criterion, there is a finite Uθ0 -net E 0 for E .
Consider the entourage U ∈ UQ defined as
\
U := f −1 ◦ θ0 ◦ f
f ∈E 0

(cf. 4.2.9). Given g ∈ E and f ∈ E 0 such that g ◦ f −1 ⊂ θ0 , observe that


θ0 = θ0−1 ⊃ (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 ◦ θ0 ◦ f ) ◦ g −1
⊂ (g ◦ f −1 ) ◦ θ0 ◦ (f ◦ g −1 ) ⊂ θ0 ◦ θ0 ◦ θ0 ⊂ θ.
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 sufficient to construct a finite Uθ -net
for E given θ ∈ UF . Choose θ0 ∈ UF such that θ0 ◦ θ0 ◦ θ0 ⊂ θ and find an open
symmetric entourage U ∈ UQ from the condition
\
U⊂ g −1 ◦ θ0 ◦ g
g∈E

(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, refine a finite
subcover {U (x0 ) : x0 ∈ Q0 }. In particular, from 1.1.10 derive
[
IQ ⊂ U (x0 ) × U (x0 )
x0 ∈Q0
[
= U −1 (x0 ) × U (x0 ) = U ◦ IQ0 ◦ U.
(x0 ,x0 )∈IQ0

The set {g|Q0 : g ∈ E } is totally bounded in F Q0 . Consequently, there is a finite


θ0 -net for this set. Speaking more precisely, there is a finite subset E 0 of E with the
following property:
g ◦ IQ0 ◦ f −1 ⊂ θ0
An Excursus into Metric Spaces 55

for every g in E and some f in E 0 . Using the estimates, successively infer that

g ◦ f −1 = g ◦ IQ ◦ f −1 ⊂ g ◦ (U ◦ IQ0 ◦ U ) ◦ f −1
⊂ g ◦ (g −1 ◦ θ0 ◦ g) ◦ IQ0 ◦ (f −1 ◦ θ0 ◦ f ) ◦ f −1
= (g ◦ g −1 ) ◦ θ0 ◦ (g ◦ IQ0 ◦ f −1 ) ◦ θ0 ◦ (f ◦ f −1 )
= Iimg ◦ θ0 ◦ (g ◦ IQ0 ◦ f −1 ) ◦ θ0 ◦ Iimf
⊂ θ0 ◦ θ0 ◦ θ0 ⊂ θ.

Thus, by 4.6.8, E 0 is a finite Uθ -net for E . B


4.6.11. Remark. It is an enlightening exercise to translate the proof of the
Ascoli–Arzelà Theorem into the “ε-δ” language. The necessary vocabulary is in hand:
“θ and Uθ stand for ε,” “θ0 is ε /3 ,” and “δ is U .” It is also rewarding and instructive
to find generalizations of the Ascoli–Arzelà 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 first
category if U is included into a countable union of rare sets; i.e., U ⊂ ∪n∈N Un with
int cl Un = ∅. A nonmeager set (which is not of first 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.
C (1) ⇒ (2): If U := ∪n∈N Un , Un = cl Un , and int Un = ∅ 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 := ∩n∈N Gn , where Gn ’s are open and cl Gn = X. Then
X \U = X \∩n∈N Gn = ∪n∈N (X \Gn ). Moreover, X \Gn is closed and int(X \Gn ) =
∅ (since cl Gn = 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 ⊂ ∪n∈N Un and int cl Un = ∅.


It may be assumed that Un = cl Un . Then Gn := X \ Un is open and everywhere
dense. By hypothesis, ∩n∈N Gn = X \ ∪n∈N Un 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 X \ U is not everywhere dense.
Consequently, U is nonmeager. B
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 (fξ : X0 = 0 O R)ξ∈„
be a family of lower semicontinuous functions such that sup{fξ (x) : ξ ∈ „} < +∞
for all x ∈ X. Then each nonempty open set G in X includes a nonempty open
subset G0 on which (fξ )ξ∈„ is uniformly bounded above; i.e., supx∈G0 sup{fξ (x) :
ξ ∈ „} ≤ +∞. CB
4.7.6. Baire Category Theorem. A complete metric space is a Baire space.
C Let G be a nonempty open set and x0 ∈ G. Suppose by way of contradiction
that G is meager; i.e., G ⊂ ∪n∈N Un , where int Un = ∅ and Un = cl Un . Take
ε0 > 0 satisfying Bε0 (x0 ) ⊂ G. It is obvious that U1 is not included into Bε0 /2 (x0 ).
Consequently, there is an element x1 in Bε0 /2 (x0 ) \ U1 . Since U1 is closed, find ε1
satisfying 0 < ε1 ≤ ε0 /2 and Bε1 (x1 ) ∩ U1 = ∅. Check that Bε1 (x1 ) ⊂ Bε0 (x0 ).
For, if d(x1 , y1 ) ≤ ε1 , then d(y1 , x0 ) ≤ d(y1 , x1 ) + d(x1 , x0 ) ≤ ε1 + ε0 /2 , because
d(x1 , x0 ) ≤ ε0 /2 . The ball Bε1 /2 (x1 ) does not lie in U2 entirely. It is thus possible
to find x2 ∈ Bε1 /2 (x1 ) \ U2 and 0 < ε2 ≤ ε1 /2 satisfying Bε2 (x2 ) ∩ U2 = ∅. It is
easy that again Bε2 (x2 ) ⊂ Bε1 (x1 ). Proceeding by induction, obtain the sequence
of nested balls Bε0 (x0 ) ⊃ Bε1 (x1 ) ⊃ Bε2 (x2 ) ⊃ . . . ; moreover, εn+1 ≤ εn /2 and
Bεn (xn ) ∩ Un = ∅. By the Nested Ball Theorem, the balls have a common point,
x := lim xn . Further, x 6= ∪n∈N Un ; and, hence, x 6∈ G. On the other hand,
x ∈ Bε0 (x0 ) ⊂ G, a contradiction. B
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 differentiable functions.
Given f : [0, 1]0 = 0 O R and x ∈ [0, 1), put

f (x + h) − f (x)
D+ f (x) := lim inf ;
h↓0 h
f (x + h) − f (x)
D+ f (x) := lim sup .
h↓0 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 finite. 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 find and
grasp. Below a few of the most popular are listed:

X hh4n xii
van der Waerden’s function: ,
n=0
4n

with hhxii := (x − [x]) ∧ (1 + [x] − x), the distance from x to the whole number
nearest to x;
+∞
X 1
Riemann’s function: 2
sin (n2 πx);
n=0
n
and, finally, the historically first

X
Weierstrass’s function: bn cos (an πx),
n=0

with a an odd positive integer, 0 < b < 1 and ab > 1 + 3π /2 .

4.8. The Jordan Curve Theorem and Rough Drafts


4.8.1. Remark. In topology, in particular, many significant and curious facts
of the metric space R2 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 definition 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 γ is a simple loop in R2 then there are
open sets G1 and G2 such that
G1 ∪ G2 = R2 \ γ; γ = ∂G1 = ∂G2 . /.
4.8.4. Remark. Either G1 or G2 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 D1 , . . . , Dn and D be closed disks (= closed balls)


in the plane which satisfy D1 , . . . , Dn ⊂ int D and Dm ∩ Dk = ∅ as m 6= k. The
set
n
[
D\ int Dk
k=1

is a holey disk or, more formally, a perforated disk. A subset of the plane which is
diffeomorphic (= “smoothly homeomorphic”) to a holey disk is called a connected
elementary compactum. The union of a finite family of pairwise disjoint connected
elementary compacta is an elementary compactum.
4.8.6. Remark. The boundary ∂F of an elementary compactum F comprises
finitely many disjoint smooth loops. Furthermore, the embedding of F into the
(oriented) plane R2 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 filter 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

Z1
|f (t) − g(t)|
d(f, g) := dt (f, g ∈ S)
1 + |f (t) − g(t)|
0

with some natural identification implied (specify it!). Find the meaning of convergence in the
space.
4.4. Given α, β ∈ NN , put

d(α, β) := 1/ min {k ∈ N : αk 6= βk }.

Check that d is a metric and the space NN 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 definitions of continuity and
uniform continuity.
4.8. Given nonempty compact subsets A and B of the spaces RN, define
   
d(A, B) := sup inf |x − y| ∨ sup inf |x − y| .
x∈A y∈B y∈B x∈A

Show that d is a metric. The metric is called the Hausdorff metric. What is the meaning
of convergence in this metric?
4.9. Prove that nonempty compact convex subsets of a compact convex set in RN constitute
a compact set with respect to the Hausdorff metric. How does this claim relate to the Ascoli–Arzelà
Theorem?
4.10. Prove that each lower semicontinuous function on RN 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 0 = 0 O is called expanding
whenever d(F (x), F (y)) ≥ βd(x, y) for some β > 1 and all x, y ∈ Y . Assume that an expanding
mapping F : Y 0 = 0 O acts onto Y . Prove that F is one-to-one and possesses a sole fixed 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 infimum 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 field F and let p : X0 = 0 O R· be
a seminorm. Then
(1) dom p 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 Bp := {p ≤ 1} are absolutely convex;

moreover, p is the Minkowski functional of every set B such that B p ⊂ B ⊂ Bp ;

(5) X = dom p if and only if B p is absorbing.
C If x1 , x2 ∈ dom p and α1 , α2 ∈ F then by 3.7.6

p(α1 x1 + α2 x2 ) ≤ |α1 |p(x1 ) + |α2 |p(x2 ) < +∞ + (+∞) = +∞.

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. B
5.1.2. If p, q : X0 = 0 O R· are two seminorms then the inequality p ≤ q
(in (R· )X ) holds if and only if Bp ⊃ Bq .
C ⇒: Evidently, {q ≤ 1} ⊂ {p ≤ 1}.
⇐: In view of 5.1.1 (4), observe that p = pBp and q = pBq . Take t1 , t2 ∈ R
such that t1 < t2 . If t1 < 0 then {q ≤ t1 } = ∅, and so {q ≤ t1 } ⊂ {p ≤ t2 }.
If t1 ≥ 0 then t1 Bq ⊂ t1 Bp ⊂ t2 Bp . Thus, by 3.8.3, p ≤ q. B
Multinormed and Banach Spaces 61

5.1.3. Let X and Y be vector spaces, let T ⊂ X×Y be a linear correspondence,


and let p : Y 0 = 0 O R· be a seminorm. If pT (x) := inf p ◦ T (x) for x ∈ X then
pT : X0 = 0 O R· is a seminorm and the set BT := T −1 (Bp ) is absolutely convex.
In addition, pT = pBT .
C Given x1 , x2 ∈ X and α1 , α2 ∈ F, infer that

pT (α1 x1 + α2 x2 ) = inf p(T (α1 x1 + α2 x2 ))

≤ inf p(α1 T (x1 ) + α2 T (x2 )) ≤ inf(|α1 |p(T (x1 )) + |α2 |p(T (x2 )))
= |α1 |pT (x1 ) + |α2 |pT (x2 );
i.e., pT is a seminorm.
The absolute convexity of BT is a consequence of 5.1.1 (4) and 3.1.8. If x ∈ BT
then (x, y) ∈ T for some y ∈ Bp . Hence, pT (x) ≤ p(y) ≤ 1; that is, BT ⊂ BpT .

If in turn x ∈ B pT then pT (x) = inf{p(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 (Bp ) = BT . Finally,

B pT ⊂ BT ⊂ BpT . Referring to 5.1.1 (4), conclude that pBT = pT . B
5.1.4. Definition. The seminorm pT , constructed in 5.1.3, is the inverse
image or preimage of p under T .
5.1.5. Definition. Let p : X0 = 0 O R be a seminorm (by 3.4.3, this implies
that dom p = 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 RX )
of everywhere-defined seminorms. Such a multinorm is denoted by MX or simply
by M, if the underlying vector space is clear from the context. A pair (X, MX ),
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. CB
5.1.8. Definition. A multinorm MX is a Hausdorff or separated multinorm
whenever for all x ∈ X, x 6= 0, there is a seminorm p ∈ MX such that p(x) 6= 0.
In this case X is called a Hausdorff or separated multinormed space.
5.1.9. Definition. A norm is a Hausdorff 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 k · k or (rarely) by k · kX or even k · | Xk if it is necessary
to indicate the space X. A pair (X, k · k) 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-defined) seminorms on a space X
then M is a Hausdorff multinorm called the finest multinorm on X.
(3) Let (Y, N) be a multinormed space, and let T ⊂ X × Y be a linear
correspondence such that dom T = X. By 3.4.10 and 5.1.1 (5), for every p in N the
seminorm pT is defined everywhere, and hence M := {pT : 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 NT . 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 Y0 of Y and T is the identical
embedding ι : Y0 0 = 0 O . At this juncture Y0 is treated as a multinormed space
with multinorm N ◦ ι. Moreover, the abuse of the phrase “N is a multinorm on Y0 ”
is very convenient.
(4) Each basic field F is endowed, as is well known, with the standard
norm | · | : F0 = 0 O R, the modulus of a scalar. Consider a vector space X and
f ∈ X # . Since f : X0 = 0 O F , it is possible to define the inverse image of the norm
on F as pf (x) := |f (x)| (x ∈ X). If is some subspace of X # then the multinorm
σ(X, ) := { : ∈} is the weak multinorm on X induced by .
(5) Let (X, p) be a seminormed space. Assume further that X0 is a sub-
space of X and ϕ : X0 = 0 O /X0 is the coset mapping. The linear correspondence
ϕ−1 is defined on the whole of X/X0 . Hence, the seminorm pϕ−1 appears, called the
quotient seminorm of p by X0 and denoted by pX/X0 . The space (X/X0 , pX/X0 )
is called the quotient space of (X, p) by X0 . The definition 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 X0 := ∩{dom p :
p ∈ M}. More precisely, thinking of the multinormed space (X0 , {pι : p ∈ M}),
where ι is the identical embedding of X0 into X, we say: “M is a multinorm,” or
“Consider the (multinormed) space specified by M.” The next example is typical:
The family of seminorms
 
α β
pα,β (f ) := sup |x ∂ f (x)| : α and β are multi-indices
x∈RN

specifies the (multinormed) space of infinitely differentiable functions on RN de-


creasing rapidly at infinity (such functions are often called tempered, cf. 10.11.6).
Multinormed and Banach Spaces 63

(7) Let (X, k·k) and (Y, k· k) be normed spaces (over the same ground
field F). Given T ∈ L (X, Y ), consider the operator norm of T , i.e. the quantity

kT xk
kT k := sup{kT xk : x ∈ X, kxk ≤ 1} = sup .
x∈X kxk

(From now on, in analogous situations we presume 0 /0 := 0.)


It is easily seen that k · k : L (X, Y )0 = 0 O R· is a seminorm. Indeed, putting
BX := {k · kX ≤ 1} for T1 , T2 ∈ L (X, Y ) and α1 , α2 ∈ F, deduce that

kα1 T1 + α2 T2 k = sup k · kα1 T1 +α2 T2 (BX )


= sup k · k((α1 T1 + α2 T2 )(BX )) ≤ sup kα1 T1 (BX ) + α2 T2 (BX )k
≤ |α1 | sup k · kT1 (BX ) + |α2 | sup k · kT2 (BX ) = |α1 | kT1 k + |α2 | kT2 k.

The subspace B(X, Y ), the effective domain of definition 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

kT xkY ≤ K kxkX (x ∈ X).

Moreover, kT k is the greatest lower bound of the set of Ks appearing in the nor-
mative inequality. CB
(8) Let X be a vector space over F and let k·k be a norm on X. Assume
further that X 0 := B(X, F) is the (normed) dual of X, i.e. the space of bounded
functionals f s with the dual norm

|f (x)|
kf k = sup{|f (x)| : kxk ≤ 1} = sup .
x∈X kxk

Consider X 00 := (X 0 )0 := B(X 0 , F), the second dual of X. Given x ∈ X and


f ∈ X 0 , put x00 := ι(x) : f 7→ f (x). Undoubtedly, ι(x) ∈ (X 0 )# = L (X 0 , F ).
In addition,

kx00 k = kι(x)k = sup{|ι(x)(f )| : kf | X 0 ≤ 1}


= sup{|f (x)| : (∀x ∈ X) |f (x)| ≤ kxkX } = sup{|f (x)| : f ∈ |∂ |(k · kX )} = kxkX .

The final equality follows for instance from Theorem 3.6.5 and Lemma 3.7.9.
Thus, ι(x) ∈ X 00 for every x in X. It is plain that the operator ι : X0 = 0 O 00 ,
64 Chapter 5

defined as ι : x 7→ ι(x), is linear and bounded; moreover, ι is a monomorphism and


kιxk = kxk for all x ∈ X. The operator ι 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 x00 := ιx as the same element; i.e., to consider X as
a subspace of X 00 . A normed space X is reflexive if X and X 00 coincide (under the
indicated embedding!). Reflexive spaces possess many advantages. Evidently, not
all spaces are reflexive. Unfortunately, such is C([0, 1], F) which is irreflexive (the
term “nonreflexive” is also is common parlance). CB
5.1.11. Remark. The constructions, carried out in 5.1.10 (8), show some
symmetry or duality between X and X 0 . In this regard, the notation (x, f ) :=
hx | f i := f (x) symbolizes the action of x ∈ X on f ∈ X 0 (or the action of f on x).
To achieve and ensure the utmost conformity, it is a common practice to denote
elements of X 0 by symbols like x0 ; for instance, hx | x0 i = (x, x0 ) = x0 (x).

5.2. The Uniformity and Topology of a Multinormed


Space
5.2.1. Let (X, p) be a seminormed space. Given x1 , x2 ∈ X, put dp (x1 , x2 )
:= p(x1 − x2 ). Then
(1) dp (X 2 ) ⊂ R+ and {d ≤ 0} ⊃ IX ;
(2) {dp ≤ t} = {dp ≤ t}−1 and {dp ≤ t} = t{dp ≤ 1} (t ∈ R+ \ 0);
(3) {dp ≤ t1 } ◦ {dp ≤ t2 } ⊂ {dp ≤ t1 + t2 } (t1 , t2 ∈ R+ );
(4) {dp ≤ t1 } ∩ {dp ≤ t2 } ⊃ {dp ≤ t1 ∧ t2 } (t1 , t2 ∈ R+ );
(5) p is a norm ⇔ dp is a metric. CB
5.2.2. Definition. The uniformity of a seminormed space (X, p) is the filter
Up := fil {{dp ≤ t} : t ∈ R+ \ 0}.
5.2.3. If Up is the uniformity of a seminormed space (X, p) then
(1) Up ⊂ fil {IX };
(2) U ∈ Up ⇒ U −1 ∈ Up ;
(3) (∀ U ∈ Up ) (∃ V ∈ Up ) V ◦ V ⊂ U . CB
5.2.4. Definition. Let (X, M) be a multinormed space. The filter U :=
sup{Up : p ∈ M} is called the uniformity of X (the other designations are UM , UX ,
etc.). (By virtue of 5.2.3 (1) and 1.3.13, the definition is sound.)
5.2.5. If (X, M) is a multinormed space and U is the uniformity of X then
(1) U ⊂ fil {IX };
(2) U ∈ U ⇒ U −1 ∈ U ;
(3) (∀ U ∈ U ) (∃ V ∈ U ) V ◦ V ⊂ U .
Multinormed and Banach Spaces 65

C Examine (3). Given U ∈ U , by 1.2.18 and 1.3.8 there are seminorms


p1 , . . . , pn ∈ M such that U = U{p1 ,... ,pn } = Up1 ∨ . . . ∨ Upn . Using 1.3.13, find
sets Uk ∈ Upk satisfying U ⊃ U1 ∩ . . . ∩ Un . Applying 5.2.3 (3), choose Vk ∈ Upk so
as to have Vk ◦ Vk ⊂ Uk . It is clear that

(V1 ∩ . . . ∩ Vn ) ◦ (V1 ∩ . . . ∩ Vn ) ⊂ V1 ◦ V1 ∩ . . . ∩ Vn ◦ Vn
⊂ U1 ∩ . . . ∩ Un .
Moreover, V1 ∩ . . . ∩ Vn ∈ Up1 ∨ . . . ∨ Upn ⊂ U . B
5.2.6. A multinorm M on X is separated if and only if so is the uniformity
UM ; i.e., ∩{V : V ∈ UM } = IX .
C ⇒: Let (x, y) 6∈ IX ; i.e., x 6= y. Then p(x − y) > 0 for some seminorm p
in M. Hence, (x, y) 6∈ {dp ≤ 1 /2 p(x − y)}. But the last set is included in Up ,
and thus in UM . Consequently, X 2 \ IX ⊂ X 2 \ ∩{V : V ∈ UM }. Furthermore,
IX ⊂ ∩{V : V ∈ UM }.
⇐: If p(x) = 0 for all p ∈ M then (x, 0) ∈ V for every V in UM . Hence,
(x, 0) ∈ IX by hypothesis. Therefore, x = 0. B
5.2.7. Given a space X with uniformity UX , define
τ (x) := {U (x) : U ∈ UX } (x ∈ X).
Then τ (x) is a filter for every x ∈ X. Moreover,
(1) τ (x) ⊂ fil {x};
(2) (∀ U ∈ τ (x)) (∃ V ∈ τ (x) & V ⊂ U ) (∀ y ∈ V ) V ∈ τ (y).
C All is evident (cf. 4.1.8). B
5.2.8. Definition. The mapping τ : x 7→ τ (x) is called the topology of a multi-
normed space (X, M); a member of τ (x) is a neighborhood about x. The designation
for the topology can be more detailed: τX , τM , τ (UM ), etc.
5.2.9. The following presentation holds:
τX (x) = sup{τp (x) : p ∈ MX }
for all x ∈ X. CB
5.2.10. If X is a multinormed space then
U ∈ τ (x) ⇔ U − x ∈ τX (0)
for all x ∈ X.
C By 5.2.9 and 1.3.13, it suffices to consider a seminormed space (X, p). In this
case for every ε > 0 the equality holds: {dp ≤ ε}(x) = εBp +x, where Bp := {p ≤ 1}.
Indeed, if p(y − x) ≤ ε then y = ε(ε−1 (y − x)) + x and ε−1 (y − x) ∈ Bp . In turn,
if y ∈ εBp + x, then p(y − x) = inf{t > 0 : y − x ∈ tBp } ≤ ε. B
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 Bp , BX , etc.
5.2.12. A multinorm MX is separated if and only if so is the topology τX ;
i.e., given distinct x1 and x2 in X, there are neighborhoods U1 in τX (x1 ) and U2
in τX (x2 ) such that U1 ∩ U2 = ∅.
C ⇒: Let x1 6= x2 and let ε := p(x1 −x2 ) > 0 for p ∈ MX . Put U1 := x1 + ε /3 Bp
and U2 := x2 + ε /3 Bp . By 5.2.10, Uk ∈ τX (xk ). Verify that U1 ∩ U2 = ∅. Indeed,
if y ∈ U1 ∩ U2 then p(x1 − y) ≤ ε /3 and p(x2 − y) ≤ ε /3 . Therefore, p(x1 − x2 ) ≤
2
/3 ε < ε = p(x1 − x2 ), which is impossible.
⇐: If (x1 , x2 ) ∈ ∩{V : V ∈ UX } then x2 ∈ ∩{V (x1 ) : V ∈ UX }. Thus,
x1 = x2 and, consequently, MX is separated by 5.2.6. B
5.2.13. Remark. The presence of a uniformity and the corresponding topol-
ogy in a multinormed space readily justifies 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 X0 be a subspace of X.
The quotient space (X/X0 , pX/X0 ) is separated if and only if X0 is closed.
C ⇒: If x 6∈ X0 then ϕ(x) 6= 0 where, as usual, ϕ : X0 = 0 O /X0 is the coset
mapping. By hypothesis, 0 6= ε := pX/X0 (ϕ(x)) = pϕ−1 (ϕ(x)) = inf{p(x + x0 ) :
x0 ∈ X0 }. Hence, the ball x + ε /2 Bp does not meet X0 and x is an exterior point
of X0 . Thus, X0 is closed.
⇐: Suppose that x is a nonzero point of X/X0 and x = ϕ(x) for some x in X.
If pX/X0 (x) = 0 then 0 = inf{p(x − x0 ) : x0 ∈ X0 }. In other words, there is
a sequence (xn ) in X0 converging to x. Consequently, by 4.1.19, x ∈ X0 and x = 0,
a contradiction. B
5.2.15. The closure of a € -set is a € -set.
C Given U ∈ (€ ), suppose that U 6= ∅ (otherwise there is nothing worthy
of proving). By 4.1.9, for x, y ∈ cl U there are nets (xγ ) and (yγ ) in U such that
xγ 0 = 0 O and yγ 0 = 0 O . If (α, β) ∈ € then αxγ + βyγ ∈ U . Appealing to 4.1.19
again, infer αx + βy = lim(αxγ + βyγ ) ∈ cl U . B

5.3. Comparison Between Topologies


5.3.1. Definition. If M and N are two multinorms on a vector space then M
is said to be finer or stronger than N (in symbols, M  N) if UM ⊃ UN . 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 M and N on a vector space X the following statements


are equivalent:
(1) M  N;
(2) the inclusion τM (x) ⊃ τN (x) holds for all x ∈ X;
(3) τM (0) ⊃ τN (0);
(4) (∀ q ∈ N) (∃ p1 , . . . , pn ∈ M) (∃ ε1 , . . . , εn ∈ R+ \ 0)
Bq ⊃ ε1 Bp1 ∩ . . . ∩ εn Bpn ;
(5) (∀ q ∈ N) (∃ p1 , . . . , pn ∈ M) (∃ t > 0) q ≤ t(p1 ∨ . . . ∨ pn ) (with
respect to the order of the K-space RX ).
C The implications (1) ⇒ (2) ⇒ (3) ⇒ (4) are evident.
(4) ⇒ (5): Applying the Gauge Theorem (cf. 5.1.2), infer that
q ≤ pBp1 /ε1 ∨ . . . ∨ pBpn /εn = 1 /ε1 p1 ∨ . . . ∨ 1 /εn pn
 

≤ 1 /ε1 ∨ . . . ∨ 1 /εn p1 ∨ . . . ∨ pn .
 

(5) ⇒ (1): It is sufficient to check that M  {q} for a seminorm q in N.


If V ∈ Uq then V ⊃ {dq ≤ ε} for some ε > 0. By hypothesis
{dq ≤ ε} ⊃ {dp1 ≤ ε /t } ∩ . . . ∩ {dpn ≤ ε /t }
with suitable p1 , . . . , pn ∈ M and t > 0. The right side of this inclusion is an element
of Up1 ∨ . . . ∨ Upn = U{p1 ,... ,pn } ⊂ UM . Hence, V is also a member of UM . B
5.3.3. Definition. Let p, q : X0 = 0 O R be two seminorms on X. Say that p
is finer 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) Bq ⊃ tBp ;
p ∼ q ⇔ (∃ t1 , t2 > 0) t2 p ≤ q ≤ t1 p ⇔ (∃ t1 , t2 > 0) t1 Bp ⊂ Bq ⊂ t2 Bp .
C Everything follows from 5.3.2 and 5.1.2. B
5.3.5. Riesz Theorem. If p, q : F N 0 = 0 O R are seminorms on the finite-
dimensional space F N then p  q ⇔ ker p ⊂ ker q. CB
5.3.6. Corollary. All norms in finite dimensions are equivalent. CB
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 × (UX ) ⊃ UY and T ×−1 (UY ) ⊂ UX ;
(3) x ∈ X ⇒ T (τX (x)) ⊃ τY (T x);
(4) T (τX (0)) ⊃ τY (0) and τX (0) ⊃ T −1 (τY (0));
(5) (∀ q ∈ N) (∃ p1 , . . . , pn ∈ M) q ◦ T ≺ p1 ∨ . . . ∨ pn . CB
68 Chapter 5

5.3.8. Let (X, k · kX ) and (Y, k · kY ) 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) k · kX  k · kY ◦ T ;
(3) T is uniformly continuous;
(4) T is continuous;
(5) T is continuous at zero.
C Each of the claims is a particular case of 5.3.7. B
5.3.9. Remark. The message of 5.3.7 shows that it is sometimes convenient
to substitute for M a multinorm equivalent to M but filtered upward (with respect
to the relation ≥ or ). For example, we may take the multinorm M := {sup M0 :
M0 is a nonempty finite subset of M}. Observe that unfiltered multinorms should
be treated with due precaution.
5.3.10. Counterexample. Let X := F „ ; and let X0 comprise all constant
functions; i.e., X0 := F1, where 1 : ξ 7→ 1 (ξ ∈ „). Set M := {pξ : ξ ∈ „},
with pξ (x) := |x(ξ)| (x ∈ F „ ). It is clear that M is a multinorm on X. Now let
ϕ : X0 = 0 O /X0 stand for the coset mapping. Undoubtedly, Mϕ−1 consists of the
sole element, zero. At the same time Mϕ−1 is separated.
5.3.11. Definition. Let (X, M) be a multinormed space and let X0 be a sub-
space of X. The multinorm Mϕ−1 , with ϕ : X0 = 0 O /X0 the coset mapping,
is referred to as the quotient multinorm and is denoted by MX/X0 . The space
(X/X0 , MX/X0 ) is called the quotient space of X by X0 .
5.3.12. The quotient space X/X0 is separated if and only if X0 is closed. CB

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 UM = Ud . 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.
C ⇒: Let UM = Ud . Passing if necessary to the multinorm M, assume that
for every n in N it is possible to find a seminorm pn in M and tn > 0 such that
{d ≤ 1 /n } ⊃ {dpn ≤ tn }. Put N := {pn : n ∈ N}. Clearly, M  N. If V ∈ UM
then V ⊃ {d ≤ 1 /n } for some n ∈ N by the definition of metric uniformity. Hence,
by construction, V ∈ Upn ⊂ UM , i.e., M ≺ N. Thus, M ∼ N. The uniformity Ud
Multinormed and Banach Spaces 69

is separated, as indicated in 4.1.7. Applying 5.2.6, observe that UM and UN 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 := {pn : n ∈ N} and M
is a Hausdorff multinorm on X. Given x1 , x2 ∈ X, define

X 1 pk (x1 − x2 )
d(x1 , x2 ) :=
2k 1 + pk (x1 − x2 )
k=1

(the series
P∞ on the right side of the above formula is dominated by the convergent
series k=1 1 /2k , and so d is defined soundly).
Check that d is a metric. It suffices to validate the triangle inequality. For
a start, put α(t) := t(1 + t)−1 (t ∈ R+ ). It is evident that α0 (t) = (1 + t)−2 > 0.
Therefore, α increases. Furthermore, α is subadditive:

α(t1 + t2 ) = (t1 + t2 )(1 + t1 + t2 )−1

= t1 (1 + t1 + t2 )−1 + t2 (1 + t1 + t2 )−1 ≤ t1 (1 + t1 )−1 + t2 (1 + t2 )−1


= α(t1 ) + α(t2 ).
Thus, given x, y, z ∈ X, infer that

∞ ∞
X 1 X 1
d(x, y) = k
α(pk (x − y)) ≤ α(pk (x − z) + pk (z − y))
2 2k
k=1 k=1

X 1
≤ (α(pk (x − z)) + α(pk (z − y))) = d(x, z) + d(z, y).
2k
k=1

It remains to established that Ud and UM coincide.


First, show that Ud ⊂ UM . Take a cylinder, say, {d ≤ ε}; and let (x, y) ∈
{dp1 ≤ t} ∩ . . . ∩ {dpn ≤ t}. Since α is an increasing function, deduce that
n ∞
X 1 pk (x − y) X 1 pk (x − y)
d(x, y) = k
+
2 1 + pk (x − y) 2k 1 + pk (x − y)
k=1 k=n+1

n ∞
t X 1 X 1 t 1
≤ k
+ k
≤ + n.
1+t 2 2 1+t 2
k=1 k=n+1

Since t(1 + t)−1 + 2−n tends to zero as n0 = 0 O and t0 = 0 O , for appropriate t


and n observe that (x, y) ∈ {d ≤ ε}. Hence, {d ≤ ε} ∈ UM , which is required.
70 Chapter 5

Now establish that UM ⊂ Ud . To demonstrate the inclusion, given pn ∈ M


and t > 0, find ε > 0 such that {dpn ≤ t} ⊃ {d ≤ ε}. For this purpose, take

1 t
ε := ,
2n 1 + t

which suffices, since from the relations

1 pn (x − y) 1 t
n
≤ d(x, y) ≤ ε = n
2 1 + pn (x − y) 2 1+t

holding for all x, y ∈ X it follows that pn (x − y) ≤ t. B


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 ∈ M which 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 (xn )n∈N in V and every sequence (λn )n∈N in F
such that λn 0 = 0 O , the sequence (λn xn ) vanishes: λn xn 0 = 0 O
O
(i.e., p(λn xn )0 = 0 for each seminorm p in M);
(3) every neighborhood of zero absorbs V .
C (1) ⇒ (2): p(λn xn ) ≤ |λn |p(xn ) ≤ |λn | sup p(V )0 = 0 O .
(2) ⇒ (3): Let U ∈ τX (0) and suppose that U fails to absorb V . From Definition
3.4.9, it follows that (∀ n ∈ N) (∃ xn ∈ V ) xn 6∈ nU . Thus, 1 /n xn 6∈ U for all n ∈ N;
i.e., (1 /n xn ) does not converge to zero.
(3) ⇒ (1): Given p ∈ M, find n ∈ N satisfying V ⊂ nBp . Obviously, sup p(V ) ≤
sup p(nBp ) = n < +∞. B
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.
C ⇒: It is clear.
⇐: Let V be a bounded neighborhood of zero. Without loss of generality,
it may be assumed that V = Bp for some seminorm p in the given multinorm M.
Undoubtedly, p ≺ M. Now if U ∈ τM (0) then nU ⊃ V for some n ∈ N. Hence,
U ∈ τp (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. B
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 Fréchet space, complete metrizable multi-
normed space, serves as a natural abstraction of Banach space. It may be shown
that the class of Fréchet spaces is the least among those containing all Banach
spaces and closed under the taking of countable products. CB
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.
P∞
C ⇒: Let n=1 kxn k < +∞ for some (countable) sequence (xn ). Then the
sequence of partial sums sn := x1 + · · · + xn is fundamental because
m m
kxn k0 = 0 O
X X
ksm − sk k = k xn k ≤
n=k+1 n=k+1

for m > k.
⇐: Given a fundamental sequence (xn ), choose an increasing sequence (nk )k∈N
such that kxn − xm k≤ 2−k as n, m ≥ nk . Then the series xn1 + (xn2 − xn1 ) + (xn3 −
xn2 ) + · · · converges in norm to some x; i.e., xnk 0 = 0 O . Observe simultaneously
that xn 0 = 0 O . B
5.5.4. If X is a Banach space and X0 is a closed subspace of X then the
quotient space X/X0 is also a Banach space.
C Let ϕ : X0 = 0 O := X/X0 be the coset mapping. Undoubtedly, P for every
−1 ∞
x ∈ there is some x ∈ ϕ (x) such that 2kxk ≥ kxk ≥ kxk. Hence, given n=1 xn ,
a norm convergent
P∞ series in , it is possible to choose xn ∈ ϕ−1 (xn ) so Pthat the

norm series n=1 kxn k be convergent. According to 5.5.3, the sum x := n=1 xn
is available. If x := ϕ(x) then

n n
xk k0 = 0 O .
X X
kx − xk k ≤ kx −
k=1 k=1

Appealing to 5.5.3 again, conclude that is a Banach space. B


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. CB
5.5.6. Theorem. If X and Y are normed spaces and X = 6 0 then B(X, Y ),
the space of bounded operators, is a Banach space if and only if so is Y .
72 Chapter 5

C ⇐: Consider a Cauchy sequence (Tn ) in B(X, Y ). By the normative inequal-


ity, kTm x − Tk xk ≤ kTm − Tk k kxk0 = 0 O for all x ∈ X; i.e., (Tn x) is fundamental
in Y . Thus, there is a limit T x := lim Tn x. Plainly, the so-defined operator T is
linear. By virtue of the estimate |kTm k − kTk k| ≤ kTm − Tk k the sequence (kTn k) is
fundamental in R and, hence, bounded; that is, supn kTn k < +∞. Therefore, pass-
ing to the limit in kTn xk ≤ supn kTn k kxk, obtain kT k < +∞. It remains to check
that kTn − T k0 = 0 O . Given ε > 0, choose a number n0 such that kTm − Tn k ≤ ε /2
as m, n ≥ n0 . Further, for x ∈ BX find m ≥ n0 satisfying kTm x − T xk ≤ ε /2 .
Then kTn x − T xk ≤ kTn x − Tm xk + kTm x − T xk ≤ kTn − Tm k + kTm x − T xk ≤ ε
as n ≥ n0 . In other words, kTn − T k = sup{kTn x − T xk : x ∈ BX } ≤ ε for all
sufficiently large n.
⇒: Let (yn ) be a Cauchy sequence in Y . By hypothesis, there is a norm-one
element x in X; i.e., x has norm 1: kxk = 1. Applying 3.5.6 and 3.5.2 (1), find
an element x0 ∈ |∂ |(k · k) satisfying (x, x0 ) = kxk = 1. Obviously, the rank-
one operator (with range of dimension 1) Tn := x0 ⊗ yn : x 7→ (x, x0 )yn belongs
to B(X, Y ), since kTn k = kx0 k kyn k. Hence, kTm − Tk k = kx0 ⊗ (ym − yk )k=
kx0 k kym − yk k = kym − yk k, i.e., (Tn ) is fundamental in B(X, Y ). Assign T :=
lim Tn . Then kT x − Tn xk = kT x − yn k ≤ kT − Tn k kxk0 = 0 O . In other words, T x
is the limit of (yn ) in Y . B
5.5.7. Corollary. The dual of a normed space (furnished with the dual norm)
is a Banach space. CB
5.5.8. Corollary. Let X be a normed space; and let ι : X0 = 0 O 00 , the double
prime mapping, be the canonical embedding of X into the second dual X 00 . Then
the closure cl ι(X) is a completion of X.
C By virtue of 5.5.7, X 00 is a Banach space. By 5.1.10 (8), ι is an isometry
from X into X 00 . Appealing to 4.5.16 ends the proof. B
5.5.9. Examples.
(1) “Abstract” examples: a basic field, a closed subspace of a Banach
space, the product of Banach spaces, and 5.5.4–5.5.8.
(2) Let E be a nonempty set. Given x ∈ F E , put kxk∞ := sup |x(E )|.
The space l∞ (E ) := l∞ (E , F) := dom k · k∞ 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 infinite, i.e. not finite, and let F stand for a filter
on E . By definition, x ∈ c(E , F ) ⇔ (x ∈ l∞ (E ) and x(F ) is a Cauchy filter on F).
In the case E := N and F is the finite complement filter (comprising all cofinite sets
each of which is the complement of a finite 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 c0 (E , F ) := {x ∈ c(E , F ) : x(F )0 = 0 O } is distinguished. If F is


the finite complement filter then the shorter notation c0 (E ) is used and we speak
of the space of functions vanishing at infinity. Given E := N, write c0 := c0 (E ).
The space c0 is referred to as the space of vanishing sequences. It is worth keeping
in mind that each of these spaces without further specification is endowed with the
norm taken from the corresponding space l∞ (E , F ).
R
(4) Let S := (E , X, ) be a system with integration. This means that
X is a vector sublattice of RE , with the lattice operationsR in X coincident
R with
E O #
R
those in R , and : X0 = 0 R is a (pre)integral; i.e., ∈ X+ and xn ↓ 0
whenever xn ∈ X and xn (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 finite
and almost everywhere defined measurable functions).
1
Denote Np (f ) := ( |f |p ) /p Rfor p ≥ 1, where is the corresponding Lebesgue
R R

extension of the initial integral . (The traditional liberty is taken of using the
same symbol for the original and its successor.)
An element of dom N1 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
Im f , both members of RE . For the sake of completeness, recall the definition
 Z 
N (g) := inf sup xn : (xn ) ⊂ X, xn ≤ xn+1 , (∀ e ∈ E ) |g(e)| = lim xn (e)
n

for an arbitrary g in F E . If F = R then dom N1 obviously presents the closure of X


in the normed space (dom N, N ).
The Hölder inequality is valid:

N1 (f g) ≤ Np (f )Np0 (g),

with p0 the conjugate exponent of p, i.e. 1 /p + 1 /p0 = 1 for p > 1.


p p0
C This is a consequence of the Young inequality, xy − x /p ≤ y /p0 for all
x, y ∈ R+ , applied to |f |/Np (f ) and |g|/Np0 (g) when Np (f ) and Np0 (g) are both
nonzero. If Np (f )Np0 (g) = 0 then the Hölder inequality is beyond a doubt. B
The set Lp := dom Np is a vector space.

C |f + g|p ≤ (|f | + |g|)p ≤ 2p (|f | ∨ |g|)p = 2p (|f |p ∨ |g|p ) ≤ 2p (|f |p + |g|p ) .

The function Np is a seminorm, satisfying the Minkowski inequality:

Np (f + g) ≤ Np (f ) + Np (g).
74 Chapter 5

C For p = 1, this is trivial. For p > 1 the Minkowski inequality follows from
the presentation
Np (f ) = sup{N1 (f g)/Np0 (g) : 0 < Np0 (g) < +∞} (f ∈ Lp )
whose right side is the upper envelope of a family of seminorms.p To prove the above
presentation, using the Hölder inequality, note that g := |f | /p0 lies Rin Lpq when
Np (f ) > 0; furthermore, Np (f ) = N1 (f g)/Np0 (g). Indeed, N1 (f g) = |f | /p0 +1 =
0
Np (f )p , because p /p0 +1 = p 1 − 1 /p +1 = p. Continue arguing to find Np0 (g)p =

R p0 R p
|g| = |f |p = Np (f )p , and so Np0 (g) = Np (f ) /p0. Finally,
p p−p / 0 1
N1 (f g)/N (g) = Np (f ) /Np (f ) = Np (f ) = Np (f )p(1− /p 0 )
= Np (f ). .
p /p0 p
p0

The quotient space Lp / ker Np , together with the corresponding quotient norm
k · kp , is called the space of p-summable
R functions or Lp space with more complete
designations Lp (S), Lp (E , X, ), etc.
Finally, if a system with integration S arises from inspection of measurable step
functions on a measure space (Š, A , µ) then it is customary to write Lp (Š, A , µ),
Lp (Š, µ) and even Lp (µ), with the unspecified parameters clear from the context.
Riesz–Fisher Completeness Theorem. Each Lp space is a Banach space.
P∞
C We P sketch out the proof. Consider t := k=1 Np (fk ), where fk ∈ Lp .
n Pn
Put σn := k=1 fk and sn := k=1 |fk |. It is seen thatR (sn ) has positive entries
and increases. The same is true of (spn ). Furthermore, spn ≤ tp < +∞. Hence,
by the Levy Monotone Convergence Theorem, for almost all e ∈ E there exists
a limit g(e) := lim spn (e), with the resulting function g a member of L1 . Putting
h(e) := g /p (e), observe that h ∈ Lp and sn (e)0 = 0 O (e) for almost all P
1
e ∈ E . The

inequalities |σn | ≤ sn ≤ h imply that for almost all e ∈ E the series k=1 fk (e)
converges. For the sum f0 (e) the estimate holds: |f0 (e)| ≤ h(e). Hence, it may
be assumed that f0 ∈ Lp . Appealing to the Lebesgue Dominated Convergence
1 / p
Theorem, conclude that Np (σn − f0 ) = |σn − f0 |p 0 = 0 O . Thus, in the
R
seminormed space under consideration each seminorm convergent series converges.
To complete the proof, apply 5.5.3–5.5.5. B
If S is the system of conventional summation on E ; i.e.,
P
R X :=P e∈E R is the
direct sum of suitably many copies of the ground field R and x := e∈E x(e), then
Lp comprises all p-summable families. This space is denoted by lp (E ). Further,
1
p /p
. In the case E := N the notation lp is used and lp is
P
kxkp := e∈E |x(e)|
referred to as the space of p-summable sequences.
(5) Define L∞ as follows: Let X be an ordered vector space and let
e ∈ X+ be a positive element. The seminorm pe associated with e is the Minkowski
functional of the order interval [−e, e], i.e.,
pe (x) := inf{t > 0 : −te ≤ x ≤ te}.
Multinormed and Banach Spaces 75

The effective domain of definition of pe is the space of bounded elements (with


respect to e); the element e itself is referred to as the strong order-unit in Xe .
An element of ker pe is said to be nonarchimedean (with respect to e). The quotient
space Xe / ker pe 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 := 1Q : q 7→ 1 (q ∈ Q)
(in itself). In RE the same element 1 generates theR 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 ) := inf{t > 0 : |f | ≤ t1} < +∞,
where ≤ means “less almost everywhere than.” This space is called the space of es-
sentially bounded functions and is labelled with L∞ . To denote the quotient space
L∞ / ker N∞ and its norm the symbols L∞ and k · k∞ 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. CB
The space L∞ , as well as the spaces C(Q, F ), lp (E ), c0 (E ), c, lp , and Lp (p ≥
1), also bears the unifying title “classical Banach space.” Nowadays a Lindenstrauss
space which is a space whose dual is isometric to L1 (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 0 is isomorphic to one of the Lp spaces
with p ≥ 1.
(6) Consider a system with integration S := E , X,
R
and let p ≥ 1.
Suppose that for everyQ e in E there is a Banach space (Ye , k · kYe ). Given an arbi-
trary element f in e∈E Ye , define |||f ||| : e 7→ kf (e)kYe . Put Np (f ) := inf{Np (g) :
g ∈ Lp , g ≥ |||f |||}. It is clear that dom Np is a vector space equipped with the
seminorm Np . The sum of the family in the sense of Lp or simply the p-sum
of (Ye )e∈E (with respect to the system with integration S) is the quotient space
dom Np / ker Np under the corresponding (quotient) norm |||·|||p .
The p-sum
P∞ of a family of Banach spaces is a BanachPspace. n
C If k=1 Np (fk ) < +∞ then the sequence (sn := k=1 |||fk |||) tends to some
almost everywhere finite positive function g and Np (g)P < +∞. It follows that for

almost all e ∈ E the sequence (sn (e)) P(i.e., the series k=1 kfk (e)kYe ) converges.

By the completeness of Ye , the series k=1 fk (e) converges to some sum f0 (e) in Ye
with kf0 (e)kYe ≤ g(e) for almost every e ∈ E P. ∞Therefore, it may be assumed that
f0 ∈ dom Np . Finally, Np ( k=1 fk − f0 ) ≤ k=n+1 Np (fk )0 = 0 O . .
Pn
In the case when E := N with conventional summation, for the sum Y of a se-
quence of Banach spaces (Yn )n∈N (in the sense of Lp ) the following notation is often
employed:
Y := (Y1 ⊕ Y2 ⊕ · · · )p ,
76 Chapter 5

with p the type of summation. An element y in Y presents a sequence (yn )n∈N such
that yn ∈ Yn and

!1 /p
X
|||y|||p := kyn kpYn < +∞.
k=1
In the case Ye := X for all e ∈ E , where X is some Banach space over F, put Fp :=
dom Np and Fp := Fp / ker Np . An element of the so-constructed space is a vector
field or a X-valued function on E (having a p-summable norm). Undoubtedly, Fp
is a Banach space. At the same time, if the initial system with integration contains
a nonmeasurable set then extraordinary elements are plentiful in Fp (in particular,
for the usual Lebesgue system with integration Fp 6= Lp ). In this connection the
functions with finite range, assuming each value on a measurable set, are selected
in Fp . Such a function, as well as the corresponding coset in Fp , is a simple, finite-
valued or step function. The closure in Fp of the set of simple functions is denoted
by Lp (or more completely Lp (X), Lp (S, X), Lp (Š, A , µ), Lp (Š, µ), etc.) and is
the space of X-valued p-summable functions. Evidently, Lp (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 finite combination of character-
istic functions: X
f= χf −1 (x) x,
x∈imf
−1
where f (x) is a measurable set as x ∈ im f , with χE (e) = 1 for e ∈ E and
χE (e) = 0 otherwise. Moreover,
Z Z X
|||f ||| = kχf −1 (x) xk
x∈imf
Z X X Z
= χf −1 (x) kxk = kxk χf −1 (x) < +∞.
x∈imf x∈imf

Next, associate with each simple function f some element in X by the rule
Z X Z
f := χf −1 (x) x.
x∈imf
R
Straightforward calculation shows that the integral defined on the subspace of sim-
ple functions is linear. Furthermore, it is bounded because
Z X Z X Z
k fk = k χf −1 (x) xk ≤ χf −1 (x) kxk
x∈imf x∈imf
Z X Z
= kxkχf −1 (x) = |||f |||.
x∈imf
Multinormed and Banach Spaces 77
R
By virtue of 4.5.10 and 5.3.8, the operator has a unique extension
R to
R an element
of B(L1 (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,
infinite dimensions bring about significant complications.
Let (xn ) 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 (kxn k),
i.e. the norm convergence of (xn ) as a series. Consequently, (xn ) has at most
countably many nonzero P∞ elements and may thus be treated as a (countable) se-
quence. Moreover, n=1 kxn k < +∞; i.e. P∞the series x1 + x2 + · · · converges in
norm. P By 5.5.3, for the series sum x = n=1 xn observe that x = limθ sθ , where
sθ := n∈θ xn is a partial sum and θ ranges over the direction of all finite sub-
sets of N. In this case, the resulting x is sometimes called the unordered sum of
(xn ), whereas the sequence
P (xn ) is called unconditionally or unorderly summable
to x (in symbols, x = n∈N xn ). Using these terms, observe that summability in
norm implies unconditional summability (to the same sum). If dim X < +∞ then
the converse holds which is the Riemann Theorem on Series. The general case is
explained by the following deep and profound assertion:
Dvoretzky–Rogers Theorem. In an arbitrary infinite-dimensionalP∞ Banach
space X, for every sequence of positive numbers (tn ) such that n=1 t2n < +∞
there is an unconditionally summable sequence of elements (xn ) with kxn k = tn for
all n ∈ N.
In this regard for a family of elements (xe )e∈E of an arbitrary multinormed
space (X, M) the following terminology is accepted: Say that P (xe )e∈E is summable
or unconditionally summable (to a sum x) and write x := e∈E xe whenever x is
the limit in (X, M) of the corresponding net of partial sums sθ , with Pθ a finite
subset of E ; i.e., sθ 0 = 0 O in (X, M). If for every p there is a sum e∈E p(xe )
then the family (xe )e∈E 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 L1 (X) into L1 (Y) by putting T f :
e 7→ T f (e) (e ∈ E ) for an arbitrary simpleR X-valued R function f . Then, given
f ∈ L1 (X), observe that T f ∈ L1 (Y) and E T f = T E f . This fact is verbalized
as follows: “The Bochner integral commutes with every bounded operator.” CB

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 kST k ≤ kSk kT k; i.e., the operator norm is submultiplica-
tive.
C Given x ∈ X and using the normative inequality twice, infer that

kST xk ≤ kSk kT xk ≤ kSk kT k kxk. .

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) 7→ 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: λ(ab) =
(λa)b = a(λb) for all a, b ∈ A and λ ∈ 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 field) is an associative
algebra (over this field) 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 6= 0 then B(X) has a neutral
element (with respect to multiplication), the identity operator IX ; i.e., B(X) is
an algebra with unity. Moreover, kIX k = 1. The algebra B(X) is a Banach algebra
if and only if X is a Banach space.
C If X = 0 then there is nothing left to proof. Given X 6= 0, apply 5.5.6. B
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 λ ∈ F, it is convenient to retain the same symbol λ for λIX . (In partic-
ular, 1 = I0 = 0!) For X 6= 0 this procedure may be thought as identification of F
with FIX .
5.6.6. Definition. If X is a normed n space and T ∈
o B(X) then the spectral
n 1 /n
radius of T is the number r(T ) := inf kT k : 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 .
C Indeed, by 5.6.1, the inequality kT n k ≤ kT kn is valid. B
5.6.8. The Gelfand formula holds:
p
n
r(T ) = lim kT n k.
Multinormed and Banach Spaces 79
1
C Take ε > 0 and let s ∈ N satisfy kT s k /s ≤ r(T )+ε. 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) ≤ s − 1.
Hence,

kT n k = kT k(n)s T l(n) k ≤ kT s kk(n) kT l(n) k


≤ 1 ∨ kT k ∨ . . . ∨ kT s−1 k kT s kk(n) = M kT s kk(n) .


Consequently,

1 1 k(n)
r(T ) ≤ kT n k /n
≤M /n
kT s k /n

1 k(n)s 1 (n−l(n))
/n /n /n /n
≤M (r(T ) + ε) =M (r(T ) + ε) .

Since M /n 0 = 0 O and (n−l(n)) /n 0 = 0 O , find r(T ) ≤ lim supkT n k /n ≤ r(T ) + ε.


1 1

1
The inequality lim inf kT n k /n ≥ r(T ) is evident. Recall that ε is arbitrary, thus
completing the proof. B
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) kT k k < 1 for some k and N;
(3) r(T ) < 1.
P∞
If one of the conditions (1)–(3) holds then k=0 T k = (1 − T )−1 .
C (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 ε > 0 and a sufficiently large
1
k ∈ N, observe
P∞ that r(T ) ≤ kT k k /k ≤ r(T ) + ε < 1. In other words, some tail
of the series k=0 kT k k is dominated
P∞ by a convergent series. The completeness
k
of B(X) and 5.5.3 imply that k=0 T converges in B(X).
P∞ Pn
Now let S := k=0 T k and Sn := k=0 T k . Then

S(1 − T ) = lim Sn (1 − T ) = lim (1 + T + · · · + T n ) (1 − T ) = lim(1 − T n+1 ) = 1;


(1 − T )S = lim(1 − T )Sn = lim(1 − T )(1 + T + · · · + T n ) = lim 1 − T n+1 = 1,


because T n 0 = 0 O . Thus, by 2.2.7 S = (1 − T )−1 . B


80 Chapter 5

5.6.10. Corollary. If kT k < 1 then (1 − T ) is invertible (= has a bounded


inverse); i.e., the inverse correspondence (1 − T )−1 is a bounded linear operator.
Moreover, k(1 − T )−1 k ≤ (1 − kT k)−1 .
C The Neumann series converges and

X X∞
−1
k(1 − T ) k ≤ k
kT k ≤ kT kk = (1 − kT k)−1 . .
k=0 k=0
5.6.11. Corollary. If k1 − T k < 1 then T is invertible and
k1 − T k
k1 − T −1 k ≤ .
1 − k1 − T k
C By Theorem 5.6.9,

X X∞
1+ k
(1 − T ) = (1 − T )k = (1 − (1 − T ))−1 = T −1 .
k=1 k=0
Hence,

X ∞
X ∞
X
−1 k k
kT − 1k = k (1 − T ) k ≤ k(1 − T ) k ≤ k1 − T kk . .
k=1 k=1 k=1
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 7→ T −1 acting from Inv(X, Y ) to Inv(Y, X) is continuous.
C Take S, T ∈ B(X, Y ) such that T −1 ∈ B(Y, X) and kT −1 k kS − T k ≤ 1 /2 .
Consider the operator T −1 S in B(X). Observe that
k1 − T −1 Sk = kT −1 T − T −1 Sk ≤ kT −1 k kT − Sk ≤ 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,
kS −1 k − kT −1 k ≤ kS −1 − T −1 k
= kS −1 (T − S)T −1 k ≤ kS −1 k kT − Sk kT −1 k ≤ 1 /2 kS −1 k.
This implies kS −1 k ≤ 2kT −1 k, yielding the inequalities
kS −1 − T −1 k ≤ kS −1 k kT − Sk kT −1 k ≤ 2kT −1 k2 kT − Sk. .
5.6.13. Definition. If X is a Banach space over F and T ∈ B(X) then a scalar
λ ∈ F is a regular or resolvent value of T whenever (λ − T )−1 ∈ B(X). In this case
put R (T, λ) := (λ − T )−1 and say that R (T, λ) is the resolvent of T at λ. The set
of the resolvent values of T is denoted by res(T ) and called the resolvent set of T .
The mapping λ 7→ R (T, λ) from res(T ) into B(X) is naturally called the resolvent
of T . The set F \ res(T ) is referred to as the spectrum of T and is denoted by Sp(T )
or σ(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 6= 0. In the case X 6= 0 for F := R the spectra of some operators can also
be void, whereas for F := C it is impossible (cf. 8.1.11). CB
5.6.15. The set res(T ) is open. If λ0 ∈ res(T ) then

X
R (T, λ) = (−1)k (λ − λ0 )k R (T, λ0 )k+1
k=0

in some neighborhood of λ0 . If |λ| > kT k then λ ∈ res(T ) and the expansion



1 X Tk
R (T, λ) =
λ λk
k=0

holds. Moreover, kR (T, λ)k0 = 0 O as |λ|0 = 0 O ∞.


C Since k(λ−T )−(λ0 −T )k = |λ−λ0 |, the openness property of res(T ) follows
from 5.6.12. Proceed along the lines

λ − T = (λ − λ0 ) + (λ0 − T ) = (λ0 − T )R (T, λ0 )(λ − λ0 ) + (λ0 − T )


= (λ0 − T )((λ − λ0 )R (T, λ0 ) + 1) = (λ0 − T )(1 − ((−1)(λ − λ0 )R (T, λ0 ))).

In a suitable neighborhood of λ0 , from 5.6.9 derive

R (T, λ) = (λ − T )−1
X∞
−1 −1
= (1 − ((−1)(λ − λ0 )R (T, λ0 ))) (λ0 − T ) = (−1)k (λ − λ0 )k R (T, λ0 )k+1 .
k=0

−1
According to 5.6.9, for |λ| > kT k there is an operator 1 − T /λ presenting
the sum of the Neumann series; i.e.,

1 T
−1 1 X Tk
R (T, λ) = 1 − /λ = .
λ λ λk
k=0

It is clear that
1 1
kR (T, λ)k ≤ · ..
|λ| 1 − kT k /|λ|
5.6.16. The spectrum of every operator is compact. CB
82 Chapter 5

5.6.17. Remark. It is worth keeping in mind that the inequality |λ| > r(T )
is a necessary
P∞and ksufficient condition for the convergence of the Laurent series,
k+1
R (T, λ) = k=0 T /λ , which expands the resolvent of T at infinity.
5.6.18. An operator S commutes with an operator T if and only if S commutes
with the resolvent of T .
C ⇒: ST = T S ⇒ S(λ − T ) = λS − ST = λS − T S = (λ − T )S ⇒
R (T, λ)S(λ − T ) = S ⇒ R (T, λ)S = S R (T, λ) (λ ∈ res(T )).
⇐: SR (T, λ0 ) = R (T, λ0 )S ⇒ S = R (T, λ0 )S(λ0 − T ) ⇒ (λ0 − T )S =
S(λ0 − T ) ⇒ T S = ST . B
5.6.19. If λ, µ ∈ res(T ) then the first resolvent equation, the Hilbert identity,
holds:
R (T, λ) − R (T, µ) = (µ − λ)R (T, µ)R (T, λ).
C “Multiplying the equality µ − λ = (µ − T ) − (λ − T ), first, by R (T, λ)
from the right and, second, by R (T, µ) from the left,” successively infer the sought
identity. B
5.6.20. If λ, µ ∈ res(T ) then R (T, λ)R (T, µ) = R (T, µ)R (T, λ). CB
5.6.21. For λ ∈ res(T ) the equality holds:

dk
R (T, λ) = (−1)k k! R (T, λ)k+1 . /.
dλk

5.6.22. Composition Spectrum Theorem. The spectra Sp(ST ) and


Sp(T S) may differ only by zero.
C It suffices to establish that 1 6∈ Sp(ST ) ⇒ 1 6∈ Sp(T S). Indeed, from
λ 6∈ Sp(ST ) and λ 6= 0 it will follow that

1 6∈ 1 /λ Sp(ST ) ⇒ 1 6∈ Sp 1 1
 
/λ ST ⇒ 1 6∈ Sp /λ T S ⇒ λ 6∈ Sp(T S).

Therefore, consider the case 1 6∈ Sp(ST ). The formal Neumann series expan-
sions

(1 − ST )−1 ∼ 1 + ST + (ST )(ST ) + (ST )(ST )(ST ) + · · · ,


T (1 − ST )−1 S ∼ T S + T ST S + T ST ST S + · · · ∼ (1 − T S)−1 − 1

lead to conjecturing that the presentation is valid:

(1 − T S)−1 = 1 + T (1 − ST )−1 S
Multinormed and Banach Spaces 83

(which in turn means 1 6∈ Sp(T S)). Straightforward calculation demonstrates the


above presentation, thus completing the entire proof:

(1 + T (1 − ST )−1 S)(1 − T S) = 1 + T (1 − ST )−1 S − T S + T (1 − ST )−1 (−ST )S


= 1 + T (1 − ST )−1 S − T S + T (1 − ST )−1 (1 − ST − 1)S
= 1 + T (1 − ST )−1 S − T S + T S − T (1 − ST )−1 S = 1;
(1 − T S)(1 + T (1 − ST )−1 S) = 1 − T S + T (1 − ST )−1 S + T (−ST )(1 − ST )−1 S
= 1 − T S + T (1 − ST )−1 S + T (1 − ST − 1)(1 − ST )−1 S
= 1 − T S + T (1 − ST )−1 S + T S − T (1 − ST )−1 S = 1. .

Exercises

5.1. Prove that a normed space is finite-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 finite-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 finite-
dimensional space.

5.5. Each norm on RN is appropriate for norming the product of finitely many normed
spaces, isn’t it?

5.6. Find conditions for continuity of an operator acting between multinormed spaces and
having finite-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 reflexivity for classical Banach spaces.
5.11. Find the mutual disposition of the spaces lp and lp0 as well as Lp and Lp0 . 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 define 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 Lp (X)0 = Lp0 (X 0 ), with X a Banach space.


5.18. Let (Xn ) be a sequence of normed spaces and let
 
O
Y
X0 = x∈ Xn : kxn k0 = 0
n∈N

be their c0 -sum (with the norm kxk = sup{ kxn k : n ∈ N} induced from the l∞ -sum). Prove that
X0 is separable if and only if so is each of the spaces Xn .
5.19. Prove that the space C (p) [0, 1] presents the sum of a finite-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 field F. A mapping
f : H 2 0 = 0 O F is a hermitian form on H provided that
(1) the mapping f ( · , y) : x 7→ f (x, y) belongs to H # for every y in Y ;
(2) f (x, y) = f (y, x)∗ for all x, y ∈ H, where λ 7→ λ∗ 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 7→ (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 first argument and
∗-linear in the second.
3. Every hermitian form f satisfies the polarization identity:

f (x + y, x + y) − f (x − y, x − y) = 4 Re f (x, y) (x, y ∈ H).

C 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)) B
4. Definition. A hermitian form f is positive or positive semidefinite provided
that f (x, x) ≥ 0 for all x ∈ H. In this event, write (x, y) := hx | yi := 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 definite) scalar
product whenever (x, x) = 0 ⇒ x = 0 with x ∈ H.
5. The Cauchy–Bunyakovskiı̆–Schwarz inequality holds:

|(x, y)|2 ≤ (x, x)(y, y) (x, y ∈ H).


86 Chapter 6

C If (x, x) = (y, y) = 0 then 0 ≤ (x+ty, x +ty) = t(x, y)∗ +t∗ (x, y). Letting
t := −(x, y), find that −2|(x, y)|2 ≥ 0; i.e., in this case the claim is established.
If, for definiteness, (y, y) 6= 0; then in view of the estimate

0 ≤ (x + ty, x + ty) = (x, x) + 2t Re(x, y) + t2 (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) 6= 0 then let θ :=
|(x, y)| (x, y)−1 and x := θx. Now |θ| = 1 and, furthermore,

(x, x) = (θx, θx) = θθ∗ (x, x) = |θ|2 (x, x) = (x, x);


|(x, y)| = θ(x, y) = (θx, y) = (x, y) = Re(x, y).

Consequently, |(x, y)|2 = Re(x, y)2 ≤ (x, x)(y, y). .


1
/2
6. If ( · , · ) is a semi-inner product on H then the mapping k·k : x 7→ (x, x)
is a seminorm on H.
C It suffices to prove the triangle inequality. Applying the Cauchy–Bunyakov-
skiı̆–Schwarz inequality, observe that

kx + yk2 = (x, x) + (y, y) + 2 Re(x, y)


≤ (x, x) + (y, y) + 2kxk kyk = (kxk + kyk)2 . .

7. Definition. A space H endowed with a semi-inner product ( · , · ) and the


associate seminorm k · k is a pre-Hilbert space. A pre-Hilbert space H is a Hilbert
space provided that the seminormed space (H, k · k) is a Banach space.
8. In a pre-Hilbert space H, the Parallelogram Law is effective

kx + yk2 + kx − yk2 = 2(kxk2 + kyk2 ) (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.

/ kx + yk2 = (x + y, x + y) = kxk2 + 2 Re(x, y) + kyk2 ;


kx − yk2 = (x − y, x − y) = kxk2 − 2 Re(x, y) + kyk2 .

9. Von Neumann–Jordan Theorem. If a seminormed space (H, k · k) obeys


the Parallelogram Law then H is a pre-Hilbert space; i.e., there is a unique semi-
1
inner product ( · , · ) on H such that kxk = (x, x) /2 for all x ∈ H.
Hilbert Spaces 87

C Considering the real carrier HR of H and x, y ∈ HR , put

(x, y)R := 1 /4 kx + yk2 − kx − yk2 .




Given the mapping ( · , y)R , from the Parallelogram Law successively derive

(x1 , y)R + (x2 , y)R


= 1 /4 kx1 + yk2 − kx1 − yk2 + kx2 + yk2 − kx2 − yk2


= 1 /4 kx1 + yk2 + kx2 + yk2 − kx1 − yk2 + kx2 − yk2


 

= 1 /4 1 /2 (k(x1 + y) + (x2 + y)k2 + kx1 − x2 k2




−1 /2 k(x1 − y) + (x2 − y)k2 + kx1 − x2 k2 )




= 1 /4 1 /2 kx1 + x2 + 2yk2 − 1 /2 kx1 + x2 − 2yk2



 2 2 
1 (x1 +x2 ) (x1 −x2 )
= /2 /2 + y − /2 − y

 
= 2 (x1 +x2 ) /2 , y .
R

In particular, (x2 , y)R = 0 in case x2 := 0, i.e. 1 /2 (x1 , y)R = 1
/2 x1 , y R
. Analo-
gously, given x1 := 2x1 and x2 := 2x2 , infer that

(x1 + x2 , y)R = (x1 , y)R + (x2 , y)R .

Since the mapping ( · , y)R is continuous for obvious reasons, conclude that
( · , y)R ∈ (HR )# . Put
(x, y) := (Re−1 ( · , y)R )(x),
where Re−1 is the complexifier (see 3.7.5).
In case F := R it is clear that (x, y) = (x, y)R = (y, x) and (x, x) = kxk2 ;
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 kx + iyk2 − kx − iyk2




= 1 /4 |i| ky − ixk2 − | − i| kix + yk2 = −(ix, y)R .



88 Chapter 6

Furthermore,

(x, x) = (x, x)R − i(ix, x)R


= kx2 k − i /4 kix + xk2 − kix − xk2


= kxk2 1 − i /4 |1 + i|2 − |1 − i|2 = kxk2 .




The claim of uniqueness follows from 6.1.3. B


10. Examples.
(1) A Hilbert space is exemplified by the L2 space (over R some system
with integration), the inner P product introduced as follows (f, g) := f g ∗ for f, g ∈
L2 . In particular, (x, y) := e∈E xe ye∗ for x, y ∈ l2 (E ).
(2) Assume that H is a pre-Hilbert space and ( · , · ) : H 2 0 = 0 O F
is a semi-inner product on H. It is clear that the real carrier HR with the semi-
inner product ( · , · )R : (x, y) 7→ 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 HR .
The pre-Hilbert space (HR , ( · , · )R ) is the realification or decomplexification of
(H, ( · , · )). In turn, if the real carrier of a seminormed space is a pre-Hilbert space
then the process of complexification 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 H0 := ker k · k be the kernel
of the seminorm k · k on H. Using the Cauchy–Bunyakovskiı̆–Schwarz inequality,
Theorem 2.3.8 and 6.1.10 (3), observe that there is a natural inner product on
the quotient space H/H0 : If x1 := ϕ(x1 ) and x2 := ϕ(x2 ), with x1 , x2 ∈ H and
ϕ : H0 = 0 O /H0 the coset mapping, then (x1 , x2 ) := (x1 , x2 ). Moreover, the pre-
Hilbert space H/H0 may be considered as the quotient space of the seminormed
space (H, k·k) by the kernel of the seminorm k·k. Thus, H/H0 is a Hausdorff space
referred to as the Hausdorff pre-Hilbert space associated with H. Completing the
normed space H/H0 , 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 (He )e∈E is a family of Hilbert spaces and H is the
2-sum of the family; i.e., h ∈ H if and only if h := (he )e∈E , where he ∈ He for e ∈ E
and
!1 /2
X
khk := khe k2 < +∞.
e∈E
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 kf + gk2 + kf − gk2

!
X X
1 2 2
= /2 kfe + ge k + + kfe − ge k
e∈E e∈E
X  X
1
/2 kfe + ge k2 + kfe − ge k2 = kfe k2 + kge k2 = kf k2 + kgk2 .

=
e∈E e∈E

Consequently, H is a Hilbert space by the von Neumann–Jordan Theorem. The


space H, the Hilbert sum of the family (He )e∈E , is denoted by ⊕e∈E He . With
E := N, it is customary to use the symbol H1 ⊕ H2 ⊕ . . . for H.
(6) Let H be a Hilbert space and let S be a system with integration.
The space L2 (S, H) comprising all H-valued square-integrable functions is also
a Hilbert space. CB

6.2. Orthoprojections
6.2.1. Let Uε be a convex subset of the spherical layer (r + ε)BH \ rBH with
r, ε > 0 in a Hilbert space H. Then the diameter of Uε vanishes as ε tends to 0.
C Given x, y ∈ Uε , on considering that 1 /2 (x + y) ∈ Uε and applying the
Parallelogram Law, for ε ≤ r infer that

(x+y) 2

2 2 2

kx − yk = 2 kxk + kyk − 4 /2
≤ 4(r + ε)2 − 4r2 = 8rε + 4ε2 ≤ 12rε. .

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 u0 of U such
that
kx − u0 k = inf{kx − uk : u ∈ U }.
C Put Uε := {u ∈ U : kx − uk ≤ inf kU − xk + ε}. By 6.2.1, the family (Uε )ε>0
constitutes a base for a Cauchy filter in U . B
6.2.3. Definition. The element u0 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 H0 be a closed subspace of a Hilbert space H and x ∈ H \ H0 .
An element x0 of H0 is the projection of x to H0 if and only if (x − x0 , h0 ) = 0 for
every h0 in H0 .
90 Chapter 6

C It suffices to consider the real carrier (H0 )R of H0 . The convex function


f (h0 ) := (h0 − x, h0 − x) is defined on (H0 )R . Further, x0 in H0 serves as the
projection of x to H0 if and only if 0 ∈ ∂x0 (f ). In view of 3.5.2 (4) this containment
means that (x − x0 , h0 ) = 0 for every h0 in H0 , because f 0 (x0 ) = 2(x0 − x, · ). B
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 H0 be a closed subspace of a Hilbert space H. The orthogonal
complement of H0 , the set H0⊥ , is a closed subspace and H = H0 ⊕ H0⊥ .
C The closure property of H0⊥ in H is evident. It is also clear that H0 ∧ H0⊥ =
H0 ∩ H0⊥ = 0. We are left with showing only that H0 ∨ H0⊥ = H0 + H0⊥ = H. Take
an element h of H \ H0 . In virtue of 6.2.2 the projection h0 of H to H0 is available
and, by 6.2.4, h − h0 ∈ H0⊥ . Finally, h = h0 + (h − h0 ) ∈ H0 + H0⊥ . .
6.2.7. Definition. The projection onto a (closed) subspace H0 along H0⊥ is
the orthoprojection onto H0 , denoted by PH0 .
6.2.8. Pythagoras Lemma. x ⊥ y ⇒ kx + yk2 = kxk2 + kyk2 . /.
6.2.9. Corollary. The norm of an orthoprojection is at most one: (H 6= 0 &
H0 6= 0) ⇒ kPH0 k = 1. /.
6.2.10. Orthoprojection Theorem. For an operator P in L (H) such that
2
P = P , the following statements are equivalent:
(1) P is the orthoprojection onto H0 := im P ;
(2) khk ≤ 1 ⇒ kP hk ≤ 1;
(3) (P x, P d y) = 0 for all x, y ∈ H, with P d the complement of P , i.e.
P d = IH − P ;
(4) (P x, y) = (x, P y) for x, y ∈ H.
C (1) ⇒ (2): This is observed in 6.2.9.
(2) ⇒ (3): Let H1 := ker P = im P d . Take x ∈ H1⊥ . Since x = P x + P d x and
x ⊥ P d x; therefore, kxk2 ≥ kP xk2 = (x − P d x, x − P d x) = (x, x) − 2 Re(x, P d x) +
(P d x, P d x) = kxk2 + kP d xk2 . Whence P d x = 0; i.e., x ∈ im P . Considering 6.2.6,
from H1 = ker P and H1⊥ ⊂ im P deduce the equalities H1⊥ = im P = H0 . Thus
(P x, P d y) = 0 for all x, y ∈ H, since P x ∈ H0 and P d y ∈ H1 .
(3) ⇒ (4): (P x, y) = (P x, P y + P d y) = (P x, P y) = (P x, P y) + (P d x, P y) =
(x, P y).
Hilbert Spaces 91

(4) ⇒ (1): Show first that H0 is a closed subspace. Let h0 := lim hn with
hn ∈ H0 , i.e. P hn = hn . For every x in H, from the continuity property of the
functionals ( · , x) and ( · , P x) successively derive

(h0 , x) = lim (hn , x) = lim (P hn , x) = lim (hn , P x) = (P h0 , x).

Whence (h0 − P h0 , h0 − P h0 ) = 0; i.e., h0 ∈ im P .


Given x ∈ H and h0 ∈ H0 , now infer that (x − P x, h0 ) = (x − P x, P h0 ) =
(P (x − P x), h0 ) = (P x − P 2 x, h0 ) = (P x − P x, h0 ) = 0. Therefore, from 6.2.4
obtain P x = PH0 x. .
6.2.11. Let P1 , and P2 be orthoprojections with P1 P2 = 0. Then P2 P1 = 0.
/ P1 P2 = 0 ⇒ im P2 ⊂ ker P1 ⇒ im P1 = (ker P1 )⊥ ⊂ (im P2 )⊥ = ker P2 ⇒
P2 P1 = 0 .
6.2.12. Definition. Orthoprojections P1 and P2 are orthogonal, in symbols,
P1 ⊥ P2 or P2 ⊥ P1 , provided that P1 P2 = 0.
6.2.13. Theorem. Let P1 , . . . , Pn be orthoprojections. The operator P :=
P1 + . . . + Pn is an orthoprojection if and only if Pl ⊥ Pm for l 6= m.
C ⇒: First, given an orthoprojection P0 , observe that kP0 xk2 = (P0 x, P0 x) =
(P02 x, x) = (P0 x, x) by Theorem 6.2.10. Consequently,

kPl xk2 + kPm xk2


n
X n
X
2
≤ kPk xk = (Pk x, x) = (P x, x) = kP xk2 ≤ kxk2
k=1 k=1

for x ∈ H and l 6= m.
In particular, putting x := Pl x, observe that

kPl xk2 + kPm Pl xk2 kPl xk2 ⇒ kPm Pl k = 0.

⇐: Straightforward calculation shows that P is an idempotent operator. Indeed,

n
!2 n n n
X X X X
2
P = Pk = Pl Pm = Pk2 = P.
k=1 l=1 m=1 k=1

Furthermore, in virtue of 6.2.10 (4), (Pk x, y) = (x, Pk y) and so (P x, y) = (x, P y).


It suffices to appeal to 6.2.10 (4) once again. B
6.2.14. Remark. Theorem 6.2.13 is usually referred to as the pairwise orthog-
onality criterion for finitely many orthoprojections.
92 Chapter 6

6.3. A Hilbert Basis


6.3.1. Definition. A family (xe )e∈E of elements of a Hilbert space H is or-
thogonal, if e1 6= e2 ⇒ xe1 ⊥ xe2 . By specification, a subset E of a Hilbert space H
is orthogonal if so is the family (e)e∈E .
6.3.2. Pythagoras Theorem. An orthogonal family (xe )e∈E of elements of
a Hilbert space is (unconditionally) summable if and only if the numeric family
(kxe k2 )e∈E is summable. Moreover,
X 2 X
xe = kxe k2 .


e∈E e∈E

e∈E xe , where θ is a finite subset of E . By 6.2.8, ksθ k =


2
P
P C Let2 sθ :=
e∈θ kxe k . Given a finite subset θ of E which includes θ, thus observe that
0

X
ksθ0 − sθ k2 = ksθ0 \θ k2 = kxe k2 .
e∈θ 0 \θ

In other words, the fundamentalness of (sθ ) amounts to the fundamentalness of the


net of partial sums of the family (kxe k2 )e∈E . On using 4.5.4, complete the proof. B
6.3.3. Orthoprojection Summation Theorem. Let (Pe )e∈E be a family
of pairwise orthogonal orthoprojections in a Hilbert space H. Then for every x
in H the
P family (Pe x)e∈E is (unconditionally) summable. Moreover, the operator
P x := e∈E Pe x is the orthoprojection onto the subspace
( )
X X
H := xe : xe ∈ He := im Pe , kxe k2 < +∞ .
e∈E e∈E

C Given a finite subset θ of E , put sθ := e∈θ Pe . ByPTheorem 6.2.13, sθ is


P
an orthoprojection. Hence, in view of 6.2.8, ksθ xk2 = 2
e∈θ kPe xk ≤ kxk
2

for every x in H. Consequently, the family (kPe xk2 )e∈E is summable (the net
of partial sums isPincreasing and bounded). By the Pythagoras Theorem, there
2
is a sum P x := e∈E Pe x; i.e., P x = limθ sθ x. Whence P x = limθ sθ P x =
limθ sθ limθ0 sθ0 x = limθ limθ0 sθ sθ0 x = limθ limθ0 sθ∩θ0 x = limθ sθ x = P x. Finally,
kP xk = k limθ sθ xk = limθ ksθ xk ≤ kxk and, moreover, P 2 = P . Appealing to
6.2.10, conclude that P is the orthoprojection
P onto im P .
If x ∈ im P , i.e., P x = x; then x = e∈E e x and by the Pythagoras Theorem
P
PHe (e ∈ E ); therefore, x ∈ H .
2 2 2
P
e∈E kPe xk = P kxk = kP xk < +∞. Since Pe x ∈
2
If xe ∈ He and e∈E kxe k < +∞ then for x := P e∈E xe (existence
P follows from
the same Pythagoras Theorem) observe that x = e∈E xe = e∈E Pe xe = P x;
i.e., x ∈ im P . Thus, im P = 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 (He )e∈E are isomorphic. The identification 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., khk = 1. Assume further that H0 := Fh is a one-dimensional subspace of H
spanned over h0 . For every element x of H and every scalar λ, a member of F,
observe that

(x − (x, h)h, λh) = λ∗ ((x, h) − (x, h))(h, h) = 0.

Therefore, by 6.2.4, PH0 = ( · , h) ⊗ h. To denote this orthoprojection, it is conve-


nient to use the symbol hhi. Thus, hhi : x 7→ (x, h)h (x ∈ H).
6.3.6. Definition. A family of elements of a Hilbert space is called orthonor-
mal (or orthonormalized) if, first, the family is orthogonal and, second, the norm of
each member of the family equals one. Orthonormal sets are defined by specifica-
tion.
6.3.7. For every orthonormal subset E of H and every element x of H, the
family (heix)e∈E is (unconditionally) summable. Moreover, the Bessel inequality
holds: X
kxk2 ≥ |(x, e)|2 .
e∈E

C It suffices to refer to the Orthoprojection Summation Theorem, for


2 2
X X X 2
kxk2 ≥ heix = (x, e)e = k(x, e)ek . .


e∈E e∈E e∈E

6.3.8. Definition.
P An orthonormal set E in a Hilbert space H is a Hilbert
basis (for H) if x = e∈E heix 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. CB
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.
C ⇒: Let h ∈ E ⊥ . Then h = e∈E heih = e∈E (h, e)e = e∈E 0 = 0.
P P P

⇐: For x ∈ H, in virtue of 6.3.3 and 6.2.4, x − e∈E heix ∈ E ⊥ . .


P
94 Chapter 6

6.3.12. Theorem. Each Hilbert space has a Hilbert basis.


C By the Kuratowski–Zorn Lemma, each Hilbert space H has an orthonormal
set E maximal by inclusion. If there were some h in H \ H0 , with H0 := cl lin(E );
then the element h1 := h − PH0 h would be orthogonal to every element in E . Thus,
for H 6= 0 we would have E ∪ {kh1 k−1 h1 } = E . A contradiction. In case H = 0
there is nothing left to proof. B
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 (xn )n∈N be a countable sequence of linearly independent
elements of a Hilbert space H. Put x0 := 0, e0 := 0 and
n−1
X yn
yn := xn − hek ixn , en := (n ∈ N).
kyn k
k=0

Evidently, (yn , ek ) = 0 for 0 ≤ k ≤ n − 1 (for instance, by 6.2.13). Also, yn 6= 0,


since H is infinite-dimensional. Say that the orthonormal sequence (en )n∈N results
from the sequence (xn )n∈N by the Gram–Schmidt 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. CB
6.3.15. Definition. Let E be a Hilbert basis for a space H and x ∈ H. The
numeric family x xe )e∈E in F E , given by the identity x
b := (b be := (x, e), is the Fourier
coefficient family of x with respect to E or the Fourier transform of X (relative
to E ).
6.3.16. Riesz–Fisher Isomorphism Theorem. Let E be a Hilbert basis
for H. The Fourier transform F : x 7→ x b (relative to E ) is an isometric isomorphism
of H onto l2 (E ). The inverse Fourier transform, the Fourier summation F −1 :
l2 (E )0 = 0 O , acts by the rule F (x) :=
−1
P
e∈E xe e for x := (xe )e∈E ∈ l2 (E ).
Moreover, for all x, y ∈ H the Parseval identity holds:
X
(x, y) = be ybe∗ .
x
e∈E

C By the Pythagoras Theorem, the Fourier transform acts in l2 (E ). By The-


orem 6.3.3, b is an epimorphism. By the Steklov Theorem, b is a monomorphism.
It is beyond a doubt that F −1 x
b = x for x ∈ H and F\ −1 (x) = x for x ∈ l (E ).
2
The equality X
kxk2 = xe k2 = kb
kb xk22 (x ∈ H)
e∈E
Hilbert Spaces 95

follows from the Pythagoras Theorem. At the same time


!
X X X X
(x, y) = x
be e, ybe e = xbe ybe∗0 (e, e0 ) = be ybe∗ . .
x
e∈E e∈E e,e0 ∈E e∈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 Riesz–Fisher 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
x0 := ( · , x). Then the prime mapping x 7→ x0 presents an isometric isomorphism
of H∗ onto H 0 .
C It is clear that x = 0 ⇒ x0 = 0. If x 6= 0 then

kx0 kH 0 = sup |(y, x)| ≤ sup kyk kxk ≤ kxk;


kyk≤1 kyk≤1

kx0 kH 0 = sup |(y, x)| ≥ |(x/kxk, x)| = kxk.


kyk≤1

Therefore, x 7→ x0 is an isometry of H∗ into H 0 . Check that this mapping is


an epimorphism.
Let l ∈ H 0 and H0 := ker l 6= H (if there no such l then nothing is to be
proven). Choose a norm-one element e in H0⊥ and put grad l := l(e)∗ e. If x ∈ H0
then
(grad l)0 (x) = (x, grad l) = (x, l(e)∗ e) = l(e)∗∗ (x, e) = 0.
Consequently, for some α in F and all x ∈ H in virtue of 2.3.12 (grad l)0 (x) = αl(x).
In particular, letting x := e, find

(grad l)0 (e) = (e, grad l) = l(e)(e, e) = αl(e);

i.e., α = 1. B
6.4.2. Remark. From the Riesz Prime Theorem it follows that the dual
space H 0 possesses a natural structure of a Hilbert space and the prime mapping
x 7→ x0 implements a Hilbert space isomorphism from H∗ onto H 0 . The inverse
mapping now coincides with the gradient mapping l 7→ 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 reflexive.


C Let ι : H0 = 0 O 00 be the double prime mapping; i.e. the canonical embedding
of H into the second dual H 00 which is determined by the rule x00 (l) = ι(x)(l) = l(x),
where x ∈ H and l ∈ H 0 (see 5.1.10 (8)). Check that ι is an epimorphism. Let
f ∈ H 00 . Consider the mapping y 7→ f (y 0 ) 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 0 ) for every y in H. Observe that
ι(x)(y 0 ) = y 0 (x) = (x, y) = f (y 0 ) for all y ∈ H. Since by the Riesz Prime Theorem
y 7→ y 0 is a mapping onto H 0 , conclude that ι(x) = f. .
6.4.4. Let H1 and H2 be Hilbert spaces and T ∈ B(H1 , H2 ). Then there is
a unique mapping T ∗ : H2 0 = 0 O 1 such that

(T x, y) = (x, T ∗ y)

for all x ∈ H1 and y ∈ H2 . Moreover, T ∗ ∈ B(H2 , H1 ) and kT ∗ k = kT k.


C Let y ∈ H2 . The mapping x 7→ (T x, y) is the composition y 0 ◦ T ; i.e.,
it presents a continuous linear functional over H1 . By the Riesz Prime Theorem
there is precisely one element x of H1 for which x0 = y 0 ◦ T . Put T ∗ y := x.
It is clear that T ∗ ∈ L (H2 , H1 ). Furthermore, using the Cauchy–Bunyakovskiı̆–
Schwarz inequality and the normative inequality, infer that

|(T ∗ y, T ∗ y)| = |(T T ∗ y, y)| ≤ kT T ∗ yk kyk ≤ kT k kT ∗ yk kyk.

Hence, kT ∗ yk ≤ kT k kyk for all y ∈ H2 ; i.e., kT ∗ k ≤ kT k. At the same time


T = T ∗∗ := (T ∗ )∗ ; i.e., kT k = kT ∗∗ k ≤ kT ∗ k. .
6.4.5. Definition. For T ∈ B(H1 , H2 ), the operator T ∗ , the member of
B(H2 , H1 ) 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 H1 and H2 be Hilbert spaces. Assume further that S, T ∈ B(H1 , H2 )
and λ ∈ F. Then
(1) T ∗∗ = T ;
(2) (S + T )∗ = S ∗ + T ∗ ;
(3) (λT )∗ = λ∗ T ∗ ;
(4) kT ∗ T k = kT k2 .
C (1)–(3) are obvious properties. If kxk ≤ 1 then

kT xk2 = (T x, T x) = |(T x, T x)| = |(T ∗ T x, x)| ≤ kT ∗ T xk kxk ≤ kT ∗ T k.

Furthermore, using the submultiplicativity of the operator norm and 6.4.4, infer
kT ∗ T k ≤ kT ∗ k kT k = kT k2 , which proves (4). B
Hilbert Spaces 97

6.4.7. Let H1 , H2 , and H3 be three Hilbert spaces. Assume further that


T ∈ B(H1 , H2 ) and S ∈ B(H2 , H3 ). Then (ST )∗ = T ∗ S ∗ .
C (ST x, z) = (T x, S ∗ z) = (x, T ∗ S ∗ z) (x ∈ H1 , z ∈ H3 ) .
T
6.4.8. Definition. Consider an elementary diagram H1 −→ H2 . The diagram
T∗
H1 ←− H2 is the adjoint of the initial elementary diagram. Given an arbitrary
diagram composed of bounded linear mappings between Hilbert spaces, assume
that each elementary subdiagram is replaced with its adjoint. Then the resulting
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.
C Follows from 6.4.7 and 6.4.6 (1). B
6.4.10. Corollary. An operator T is invertible if and only if T ∗ is invertible.
Moreover, T ∗−1 = T −1∗ . CB
6.4.11. Corollary. If T ∈ B(H) then λ ∈ Sp(T ) ⇔ λ∗ ∈ Sp(T ∗ ). CB
6.4.12. Sequence Star Principle (cf. 7.6.13). A sequence

...0 = 0 O −−−k−→ Hk −−−−→ Hk+1 0 = 0 O


T Tk+1
k−1

is exact if and only if so is the sequence star



Tk∗ Tk+1
. . . ← Hk−1 ←−−−− Hk ←−−−− Hk+1 ← . . . . /.

6.4.13. Definition. An involutive algebra or ∗-algebra A (over a ground


field F) is an algebra with an involution ∗, i.e. with a mapping a 7→ a∗ from A
to A such that
(1) a∗∗ = a (a ∈ A);
(2) (a + b)∗ = a∗ + b∗ (a, b ∈ A);
(3) (λa)∗ = λ∗ a∗ (λ ∈ F, a ∈ A);
(4) (ab)∗ = b∗ a∗ (a, b ∈ A).
A Banach algebra A with involution ∗ satisfying ka∗ ak = kak2 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). CB
98 Chapter 6

6.5. Hermitian Operators


6.5.1. Definition. Let H be a Hilbert space over a ground field 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:

kT k = sup |(T x, x)|.


kxk≤1

C Put t := sup{|(T x, x)| : kxk ≤ 1}. It is clear that |(T x, x)| ≤ kT xk kxk ≤
kT k provided kxk ≤ 1. Thus, t ≤ kT k.
Since T = T ∗ ; therefore, (T x, y) = (x, T y) = (T y, x)∗ = (y, T x)∗ ; i.e.,
(x, y) 7→ (T x, y) is a hermitian form. Consequently, in virtue of 6.1.3 and 6.1.8

4 Re(T x, y) = (T (x + y), x + y) − (T (x − y), x − y)


≤ t(kx + yk2 + kx − yk2 ) = 2t(kxk2 + kyk2 ).

If T x = 0 then it is plain that kT xk ≤ t. Assume T x 6= 0. Given kxk ≤ 1 and


putting y := kT xk−1 T x, infer that
 
Tx Tx
kT xk = kT xk ,
kT xk kT xk
 2 
= (T x, y) = Re(T x, y) ≤ 1 /2 t kxk2 + T x /kT xk ≤ t;

i.e., kT k = sup{kT xk : kxk ≤ 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 fT (x, y) := (T x, 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 T y of H such that
f ( · , y) = (T y)0 . Evidently, T ∈ L (H) and (x, T y) = f (x, y) = f (y, x)∗ =
(y, T x)∗ = (T x, y). It is possible to show that in this case T ∈ B(H) and T = T ∗ .
In addition, f = fT . Therefore, the condition T ∈ B(H) in Definition 6.5.1 can be
replaced with the condition T ∈ L (H) (the Hellinger–Toeplitz Theorem).
6.5.4. Weyl Criterion. A scalar λ belongs to the spectrum of a hermitian
operator T if and only if
inf kλx − T xk = 0.
kxk=1

C ⇒: Put t := inf{kλx − T xk : x ∈ H, kxk = 1} > 0. Demonstrate that


λ∈
/ Sp(T ). Given an x in H, observe that kλx − T xk ≥ tkxk. Thus, first, (λ − T )
Hilbert Spaces 99

is a monomorphism; second, H0 := im(λ − T ) is a closed subspace (because k(λ −


T )xm − (λ − T )xk k ≥ tkxm − xk k; i.e., the inverse image of a Cauchy sequence
is a Cauchy sequence); and, third, which is final, (λ − T )−1 ∈ B(H) whenever
H = H0 (in such situation kR (T, λ)k ≤ t−1 ). Suppose to the contrary that
H 6= H0 . Then there is some y in H0⊥ satisfying kyk = 1. For all x ∈ H, note that
0 = (λx − T x, y) = (x, λ∗ y − T y); i.e., λ∗ y = T y. Further, λ∗ = (T y, y)/(y, y)
and the hermiticity of T guarantees λ∗ ∈ R. Whence λ∗ = λ and y ∈ ker(λ − T ).
We arrive at a contradiction: 1 = kyk = k0k = 0.
⇐: If λ ∈/ Sp(T ) then the resolvent of T at λ, the member R (T, λ) of B(H),
is available. Hence, inf{kλx − T xk : kxk = 1} ≥ kR (T, λ)k−1 . .
6.5.5. Spectral Endpoint Theorem. Let T be a hermitian operator in
a Hilbert space. Put

mT := inf (T x, x), MT := sup (T x, x).


kxk=1 kxk=1

Then Sp(T ) ⊂ [mT , MT ] and mT , MT ∈ Sp(T ).


C Considering that the operator T − Re λ is hermitian in the space H under
study, from the identity

kλx − T xk2 = | Im λ|2 kxk2 + kT x − Re λxk2

infer the inclusion Sp(T ) ⊂ R by 6.5.4. Given a norm-one element x of H and


invoking the Cauchy–Bunyakovskiı̆–Schwarz inequality, in case λ < mT deduce
that

kλx − T xk = kλx − T xk kxk ≥ |(λx − T x, x)|


= |λ − (T x, x)| = (T x, x) − λ ≥ mT − λ > 0.

On appealing to 6.5.4, find λ ∈ res(T ). In case λ > MT , similarly infer that

kλx − T xk ≥ |(λx − T x, x)| = |λ − (T x, x)| = λ − (T x, x) ≥ λ − MT > 0.

Once again λ ∈ res(T ). Finally, Sp(T ) ⊂ [mT , MT ].


Since (T x, x) ∈ R for x ∈ H; therefore, in virtue of 6.5.2

kT k = sup{|(T x, x)| : kxk ≤ 1}


= sup{(T x, x) ∨ (−(T x, x)) : kxk ≤ 1} = MT ∨ (−mT ).

Assume first that λ := kT k = MT . If kxk = 1 then

kλx − T xk2 = λ2 − 2λ(T x, x) + kT xk2 ≤ 2kT k2 − 2kT k(T x, x).


100 Chapter 6

In other words, the next estimate holds:


inf kλx − T xk2 ≤ 2kT k inf (kT k − (T x, x)) = 0.
kxk=1 kxk=1

Using 6.5.4, conclude that λ ∈ Sp(T ).


Now consider the operator S := T −mT . It is clear that MS = MT −mT ≥ 0 and
mS = mT − mT = 0. Therefore, kSk = MS and in view of the above MS ∈ Sp(S).
Whence it follows that MT belongs to Sp(T ), since T = S+mT and MT = MS +mT .
It suffices to observe that mT = −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). CB
6.5.7. Corollary. A hermitian operator is zero if and only if its spectrum
consists of zero. CB

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 (BX ) of
the unit ball BX of X is relatively compact in Y .
6.6.2. Remark. Detailed study of compact operators in Banach spaces is the
purpose of the Riesz–Schauder theory to be exposed in Chapter 8.
6.6.3. Let T be a compact hermitian operator. If 0 6= λ ∈ Sp(T ) then λ is
an eigenvalue of T ; i.e., ker(λ − T ) 6= 0.
C By the Weyl Criterion, λxn − T xn 0 = 0 O for some sequence (xn ) such that
kxn k = 1. Without loss of generality, assume that the sequence (T xn ) converges to
y := lim T xn . Then from the identity λxn = (λxn − T xn ) + T xn obtain that there is
a limit (λxn ) and y = lim λxn . Consequently, T y = T (lim λxn ) = λ lim T xn = λy.
Since kyk = |λ|, conclude that y is an eigenvector of T , i.e. y ∈ ker(λ − T ). B
6.6.4. Let λ1 and λ2 be distinct eigenvalues of a hermitian operator T . Assume
further that x1 and x2 are eigenvectors with eigenvalues λ1 and λ2 (i.e., xs ∈
ker(λs − T ), s := 1, 2). Then x1 and x2 are orthogonal.
/ (x1 , x2 ) = λ11 (T x1 , x2 ) = λ11 (x1 , T x2 ) = λλ21 (x1 , x2 ) .
6.6.5. For whatever strictly positive ε, there are only finitely many eigenvalues
of a compact hermitian operator beyond the interval [−ε, ε].
C Let (λn )n∈N be a sequence of pairwise distinct eigenvalues of T satisfying
|λn | > ε. Further, let xn be an eigenvector corresponding to λn and such that
kxn k = 1. By virtue of 6.6.4 (xk , xm ) = 0 for m 6= k. Consequently,
kT xm − T xk k2 = kT xm k2 + kT xk k2 = λ2m + λ2k ≥ 2ε2 ;
i.e., the sequence (T xn )n∈N 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 6= λ ∈ Sp(T ). Put Hλ := ker(λ − T ). Then Hλ
is finite-dimensional and the decomposition H = Hλ ⊕ Hλ⊥ reduces T . Moreover,
the matrix presentation holds
 
λ 0
T ∼ ,
0 Tλ

where the operator Tλ , the part of T in Hλ⊥ , is hermitian and compact, with
Sp(Tλ ) = Sp(T ) \ {λ}.
C The subspace Hλ is finite-dimensional in view of the compactness of T .
Furthermore, Hλ is invariant under T . Consequently, the orthogonal complement
Hλ⊥ of Hλ is an invariant subspace of T ∗ (coincident with T ), since h ∈ Hλ⊥ ⇒
(∀ x ∈ Hλ ) x ⊥ h ⇒ (∀ x ∈ Hλ ) 0 = (h, T x) = (T ∗ h, x) ⇒ T ∗ h ∈ Hλ⊥ .
The part of T in Hλ is clearly λ. The part Tλ of T in Hλ⊥ is undoubtedly
compact and hermitian. Obviously, for µ 6= λ, the operator
 
µ−λ 0
µ−T ∼
0 µ − Tλ

is invertible if and only if so is µ − Tλ . It is also clear that λ is not an eigenvalue


of Tλ . .
6.6.7. Hilbert–Schmidt Theorem. Let H be a Hilbert space and let T be
a compact hermitian operator in H. Assume further that Pλ is the orthoprojection
onto ker(λ − T ) for λ ∈ Sp(T ). Then
X
T = λPλ .
λ∈Sp(T )

C Using 6.5.6 and 6.6.6 as many times as need be, for every finite subset θ
of Sp(T ) obtain the equality

X
T − λPλ = sup{|λ| : λ ∈ (Sp(T ) ∪ 0) \ θ}.


λ∈θ

It suffices to refer to 6.6.5. B


6.6.8. Remark. The Hilbert–Schmidt Theorem provides essentially new in-
formation, as compared with the case of finite dimensions, only if the operator T
has infinite-rank, that is, the dimension of its range is infinite or, which is the same,
H0⊥ is an infinite-dimensional space, where H0 := ker T . In fact, if the operator T
102 Chapter 6

has finite rank (i.e., its range is finite-dimensional) then, since the subspace H0⊥ is
isomorphic with the range of T , observe that
n
X n
X
T = λk hek i = λk e0k ⊗ ek ,
k=1 k=1

where λ1 , . . . , λn are nonzero points of Sp(T ) counted with multiplicity, and {e1 , . . . , en }
is a properly-chosen orthonormal basis for H0⊥ .
The Hilbert–Schmidt Theorem shows that, to within substitution of series for
sum, an infinite-rank compact hermitian operator looks like a finite-rank operator.
Indeed, for λ 6= µ, where λ and µ are nonzero points of Sp(T ), the eigenspaces
Hλ and Hµ are finite-dimensional and orthogonal. Moreover, the Hilbert sum
⊕λ∈Sp(T )\0 Hλ equals H0⊥ = cl im T , because H0 = (im T )⊥ . Successively se-
lecting a basis for each finite-dimensional space Hλ by enumerating the eigen-
values in decreasing order of magnitudes with multiplicity counted; i.e., putting
λ1 := λ2 := . . . := λdim Hλ1 := λ1 , λdim Hλ1 +1 := . . . := λdim Hλ1 +dim Hλ2 := λ2 , etc.,
obtain the decomposition H = H0 ⊕ Hλ1 ⊕ Hλ2 ⊕ . . . and the presentation

X ∞
X
T = λk hek i = λk e0k ⊗ ek ,
k=1 k=1

where the series is summed in operator norm. CB


6.6.9. Theorem. Let T in K (H1 , H2 ) be an infinite-rank compact operator
from a Hilbert space H1 to a Hilbert space H2 . There are orthonormal families
(ek )k∈N in H1 , (fk )k∈N in H2 , and a numeric family (µk )k∈N in R+ \ 0, µk ↓ 0, such
that the following presentation holds:

X
T = µk e0k ⊗ fk .
k=1

C Put S := T ∗ T . It is clear that S ∈ B(H1 ) and S is compact. Furthermore,


(Sx, x) = (T ∗ T x, x) = (T x, T x) = kT xk2 . Consequently, in virtue of 6.4.6, S
is hermitian and H0 := ker S = ker T . Observe also that Sp(S) ⊂ R+ by Theorem
6.5.5.
Let (ek )k∈N be an orthonormal basis for H0⊥ comprising all eigenvalues of S and
let (λk )k∈N be a corresponding decreasing sequence of strictly positive eigenvalues
λk > 0, k ∈ N (cf. 6.6.8). Then every element x ∈ H1 expands into the Fourier
series

X
x − PH 0 x = (x, ek )ek .
k=1
Hilbert Spaces 103

Therefore, considering that T PH0 = 0 and assigning µk := λk and fk := µ−1
k T ek ,
find
∞ ∞ ∞
X X µk X
Tx = (x, ek )T ek = (x, ek ) T ek = µk (x, ek )fk .
µk
k=1 k=1 k=1

The family (fk )k∈N is orthonormal because


 
T en T em 1
(fn , fm ) = , = (T en , T em )
µn µm µn µm
1 1
= (T ∗ T en , em ) = (Sen , em )
µn µm µn µm
1 µn
= (λn en , em ) = (en , em ).
µn , µm µm

Successively using the Pythagoras Theorem and the Bessel inequality, argue as
follows:
n
! 2 ∞ 2
X X
T− µk e0k ⊗ fk x = µk (x, ek )fk


k=1 k=n+1

X ∞
X
= µ2k |(x, ek )|2 ≤ λn+1 |(x, ek )|2 ≤ λn+1 kxk2 .
k=n+1 k=n+1

Since λk ↓ 0, finally deduce that


n

µk e0k ⊗ fk ≤ µn+1 0 = 0 O . .
X
T −


k=1

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 finite-rank op-
erators. 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 infinite-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 L2 space on an interval.
6.9. Prove that x ⊥ y if and only if kx + yk2 = kxk2 + kyk2 and kx + iyk2 = kxk2 + kyk2 .
6.10. Given a bounded operator T , prove that

(ker T )⊥ = cl im T ∗ , (im T )⊥ = 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 finite-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 (akl )k,l∈N be an infinite matrix such that akl ≥ 0 for all k and l. Assume further
that there are also pk and β, γ > 0 satisfying

∞ ∞
X X
akl pk ≤ βpl ; akl pl ≤ γpk (k, l ∈ N).
k=1 l=1

p
Then there is some T in B(l2 ) such that (ek , el ) = akl and kT k = βγ, where ek is the
characteristic function of k, a member of N.
Chapter 7
Principles of Banach Spaces

7.1. Banach’s Fundamental Principle


7.1.1. Lemma. Let U be a convex set with nonempty interior in a multi-
normed space: int U 6= ∅. Then
(1) 0 ≤ α < 1 ⇒ α cl U + (1 − α) int U ⊂ int U ;
(2) core U = int U ;
(3) cl U = cl int U ;
(4) int cl U = int U .
C (1): For u0 ∈ int U , the set int U − u0 is an open neighborhood of zero
in virtue of 5.2.10. Whence, given 0 ≤ α < 1, obtain

α cl U = cl αU ⊂ αU + (1 − α)(int U − u0 )
= αU + (1 − α) int U − (1 − α)u0
⊂ αU + (1 − α)U − (1 − α)u0 ⊂ U − (1 − α)u0 .

Thus, (1 − α)u0 + α cl U ⊂ U and so U includes (1 − α) int U + α cl U . The last set


is open as presenting the sum of α cl U and (1 − α) int U , an open set.
(2): Undoubtedly, int U ⊂ core U . If u0 ∈ int U and u ∈ core U then u =
αu0 + (1 − α)u1 for some u1 in U and 0 < α < 1. Since u1 ∈ 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
u0 in the set int U and putting uα := αu0 +(1−α)u, infer that uα 0 = 0 O as α0 = 0 O
and uα ∈ int U when 0 < α < 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 u0
in the set int U again, find u1 ∈ cl U and 0 < α < 1 satisfying u = αu0 + (1 − α)u1 .
Using (1), finally infer that u ∈ int U . B
106 Chapter 7

7.1.2. Remark. In the case of finite dimensions, the condition int U 6= ∅ 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 := Bc0 ∩ X, where c0 is the space of vanishing sequences and X is the
subspace of terminating sequences in c0 , the direct sum of countably many copies
of a basic field. Evidently, core U = ∅ and at the same time cl U = Bc0 . CB
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,
P∞ U is ideally convex if for all sequences (αn )n∈N
P∞and (un )n∈N , with
αn ∈ R+ , n=1 αn = 1 and un ∈ U such that the series n=1 αn un converges
in X to some u, the containment holds: u ∈ U .
7.1.4. Examples.
(1) Translation (by a vector) preserves ideal convexity. CB
(2) Every closed convex set is ideally convex. CB
(3) Every open convex set is ideally convex.
C Take an open and convex U . If U = ∅ then nothing is left to proof. If U 6= ∅
then by 7.1.4 (1) it may be assumed that 0 ∈ U and, consequently, U = {pU < 1},
where pU is the MinkowskiP functional of U . Let (un )n∈N and (αnP)n∈N be sequences
∞ ∞
in U and in R+ such that n=1 αn = 1, with the element u := n=1 αn un failing
PU∞ = {pU ≤ 1} and so pUP
to lie in U . By virtue of 7.1.4 (2), u belongs to cl (u) = 1.

On the other hand, it P is clear that pU (u) ≤ α p
P∞ n U n
n=1 (u ) ≤ 1 = n=1 αn

(cf. 7.2.1). Thus, 0 = n=1 (αn − αn pU (un )) = n=1 αn (1 − pU (un )). Whence
αn = 0 for all n ∈ N. We arrive at a contradiction. B
(4) The intersection of a family of ideally convex sets is ideally con-
vex. CB
(5) Every convex subset of a finite-dimensional space is ideally con-
vex. CB
7.1.5. Banach’s Fundamental Principle. In a Banach space, each ideally
convex set with absorbing closure is a neighborhood of zero.
C Let U be such set in a Banach space X. By hypothesis X = ∪n∈N n cl U . By
the Baire Category Theorem, X is nonmeager and so there is some n in N such that
int n cl U 6= ∅. Therefore, int cl U = 1 /n int n cl U 6= ∅. By hypothesis 0 ∈ core cl U .
Consequently, from 7.1.1 it follows that 0 ∈ int cl U . In other words, there is a δ > 0
such that cl U ⊃ δBX . Consequently,

ε > 0 ⇒ cl 1 /ε U ⊃ δ /ε BX .

Using the above implication, show that U ⊃ δ /2 BX .


Principles of Banach Spaces 107

Let x0 ∈ δ /2 BX . Putting ε := 2, choose y1 ∈ 1 /ε U from


the condition
ky1 −
x0 k ≤ 1 /2ε δ. Thus obtain an element u1 of U such that 1 /2 u1 − x0 ≤ 1 /2ε δ =
1
/4 δ.
Now putting x0 := −1 /2 u1 + x0 and ε := 4 and applying 1 the argument of the
1
preceding paragraph, find an element u2 of U satisfying /4 u2 + /2 u1 − x0 ≤

1
/2ε δ = 1 /8 δ. Proceeding by induction, construct a sequence
P∞ 1 (un )n∈N of the points
of U which
P∞ 1 possesses the property that the series n=1 /2 un converges to x0 .
n

Since n=1 /2 = 1 and the set U is ideally convex, deduce x0 ∈ U. .


n

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 .
C 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, U −u is a neighborhood of zero by Banach’s
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.
C A closed convex set is ideally convex. B
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 Fréchet spaces.
7.1.9. Counterexample. Each infinite-dimensional Banach space contains
absolutely convex, absorbing and not ideally convex sets.
C Using, for instance, a Hamel basis, take a discontinuous linear functional f .
Then the set {|f | ≤ 1} is what was sought. B

7.2. Boundedness Principles


7.2.1. Let p : X0 = 0 O R be a sublinear functional on a normed space (X, k·k).
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) kpk := sup{|p(x)| : kxk ≤ 1} < +∞; i.e., p is bounded.
108 Chapter 7

C The implications (1) ⇒ (2) ⇒ (3) ⇒ (4) are immediate.


(4) ⇒ (5): There is some t > 0 such that t−1 BX ⊂ {p ≤ 1}. Given kxk ≤ 1,
thus find p(x) ≤ t. In addition, the inequality −p(−x) ≤ p(x) implies that −p(x) ≤
t for x ∈ BX . Finally, kpk ≤ 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) ≤ kpk kx − yk. .


7.2.2. Gelfand Theorem. Every lower semicontinuous sublinear functional
with domain a Banach space is continuous.
C Let p be such functional. Then the set {p ≤ 1} is closed (cf. 4.3.8). Since
dom p is the whole space; therefore, by 3.8.8, {p ≤ 1} is an absorbing set. By
Banach’s Fundamental Principle {p ≤ 1} is a neighborhood of zero. Application
to 7.2.1 completes the proof. B
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 dom p = X may be slightly relaxed by assuming dom p 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 {T x : T ∈ E } is
bounded in Y ;
(2) E is equicontinuous.
C (1) ⇒ (2): Put q(x) := sup{p(T x) : T ∈ E }, with p the (semi)norm of Y .
Evidently, q is a lower semicontinuous sublinear functional and so by the Gelfand
Theorem kqk < +∞; i.e., p(T (x − y)) ≤ kqk kx − yk for all T ∈ E . Consequently,
T ×−1 ({dp ≤ ε}) ⊃ {dk·k ≤ ε/kqk} for every T in E , where ε > 0 is taken arbitrarily.
This means the equicontinuity property of E .
(2) ⇒ (1): Straightforward. B
7.2.5. Uniform Boundedness Pronciple. Let X be a Banach space and
let Y be a normed space. For every nonempty family (Tξ )ξ∈„ of bounded operators
the following statements are equivalent:
(1) x ∈ X ⇒ supξ∈„ kTξ xk < +∞;
(2) supξ∈„ kTξ k < +∞.
C It suffices to observe that 7.2.5 (2) is another expression for 7.2.4 (2). B
Principles of Banach Spaces 109

7.2.6. Let X be a Banach space and let U be a subset of X 0 . Then the


following statements are equivalent:
(1) U is bounded in X 0 ;
(2) for every x in X the numeric set {hx | x0 i : x0 ∈ U } is bounded
in F.
C This is a particular case of 7.2.5. B
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 x0 in X 0 the numeric set {hx | x0 i : x ∈ U } is bounded
in F.
C Only (2) ⇒ (1) needs examining. Observe that X 0 is a Banach space
(cf. 5.5.7) and X is isometrically embedded into X 00 by the double prime map-
ping (cf. 5.1.10 (8)). So, the claim follows from 7.2.6. B
7.2.8. Remark. The message of 7.2.7 (2) may be reformulated as “U is bounded
in the space (X, σ(X, X 0 ))” 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. Banach–Steinhaus Theorem. Let X and Y be Banach spaces. As-
sume further that (Tn )n∈N , Tn ∈ B(X, Y ), is a sequence of bounded operators.
Put E := {x ∈ X : ∃ lim Tn x}. The following conditions are equivalent:
(1) E = X;
(2) supn∈N kTn k < +∞ and E is dense in X.
Under either (and, hence, both) of the conditions (1) and (2) the mapping T0 :
X0 = 0 O , defined as T0 x := lim Tn x, presents a bounded linear operator and kT0 k ≤
lim inf kTn k.
C If E = X then, of course, cl E = X. In addition, for every x in X the se-
quence (Tn x)n∈N is bounded in Y (for, it converges). Consequently, by the Uniform
Boundedness Principle supn∈N kTn k < +∞ and (1) ⇒ (2) is proven.
If (2) holds and x ∈ X then, given x ∈ E and m, k ∈ N, infer that

kTm x − Tk xk = kTm x − Tm x + Tm x − Tk x + Tk x − Tk xk
≤ kTm x − Tm xk + kTm x − Tk xk + kTk x − Tk xk
≤ kTm k kx − xk + kTm x − Tk xk + kTk k kx − xk
≤ 2 sup kTn k kx − xk + kTm x − Tk xk.
n∈N

Take ε > 0 and choose, first, x ∈ E such that 2 supn kTn k kx − xk ≤ ε /2 , and,
second, n ∈ N such that kTm x − Tk xk ≤ ε /2 for m, k ≥ n. By virtue of what was
110 Chapter 7

proven kTm x − Tk xk ≤ ε; i.e., (Tn x)n∈N 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 suffices to observe that
kT0 xk = lim kTn xk ≤ lim inf kTn k kxk
for all x ∈ X, because every norm is a continuous function. B
7.2.10. Remark. Under the hypotheses of the Banach–Steinhaus Theorem,
the validity of either of the equivalent items 7.2.9 (1) and 7.2.9 (2) implies that
(Tn ) converges to T0 compactly on X (= uniformly on every compact subset of X).
In other words,
sup kTn x − T0 xk0 = 0 O
x∈Q

for every (nonempty) compact set Q in X.


C Indeed, it follows from the Gelfand Theorem that the sublinear functional
pn (x) := sup{kTm x − T0 xk : m ≥ n} is continuous. Moreover, pn (x) ≥ pn+1 (x) and
pn (x)0 = 0 O 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.” B
7.2.11. Singularity Fixation Principle. Let X be a Banach space and let
Y be a normed space. If (Tn )n∈N is a sequence of operators, Tn ∈ B(X, Y )
and supn kTn k = +∞ then there is a point x of X satisfying supn kTn xk = +∞.
The set of such points “fixing a singularity” is residual.
C The first part of the assertion is contained in the Uniform Boundedness
Principle. The second requires referring to 7.2.3 and 4.7.4. B
7.2.12. Singularity Condensation Principle. Let X be a Banach space
and let Y be a normed space. If (Tn,m )n,m∈N is a family of operators, Tn ∈ B(X, Y ),
such that supn kTn,m k = +∞ for every m ∈ N then there is a point x of X satisfying
supn kTn,m xk = +∞ for all m ∈ N. CB

7.3. The Ideal Correspondence Principle


7.3.1. Let X and Y be vector spaces. A correspondence F ⊂ X × Y is convex
if and only if for x1 , x2 ∈ X and α1 , α2 ∈ R+ such that α1 + α2 = 1, the inclusion
holds:
F (α1 x1 + α2 x2 ) ⊃ α1 F (x1 ) + α2 F (x2 ).
C ⇐: If (x1 , y1 ), (x2 , y2 ) ∈ F and α1 , α2 ≥ 0, α1 +α2 = 1, then α1 y +α2 y2 ∈
F (α1 x1 + α2 x2 ) since y1 ∈ F (x1 ) and y2 ∈ F (x2 ).
⇒: If either x1 or x2 fails to enter in dom F then there is nothing to prove. If
y1 ∈ F (x1 ) and y2 ∈ F (x2 ) with x1 , x2 ∈ dom F then α1 (x1 , y1 ) + α2 (x2 , y2 ) ∈ F
for α1 , α2 ≥ 0, α1 + α2 = 1 (cf. 3.1.2 (8)). B
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 τX and τY . For instance, it is
possible to put k(x, y)k := kxkX + kykY ; i.e., to define 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 definition is sound.

7.3.3. Definition. A correspondence F ⊂ X × Y , with X and Y Banach


spaces, is called ideally convex or, briefly, 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.
C Let F ⊂ X × Y be an ideal correspondence and let U be a bounded ideally
convex set in X. If U ∩ dom F = ∅ then F (U ) = ∅ and nothing is left unproven.
Let now (yn )n∈N ⊂ F (U ); i.e., yn ∈ F (xn ), where xnP∈ U and n ∈ N. Let, finally,

(αn ) be a sequence of positive P numbers such that n=1 αn = 1 and, moreover,

there is a sum of the series y := n=1 αn yn in Y . It is beyond a doubt that


X ∞
X ∞
X
kαn xn k = αn kxn k ≤ αn sup kU k = sup kU k < +∞
n=1 n=1 n=1

in view of the boundedness property of U .P Since X is complete, from 5.5.3 it



follows
P∞ that X contains the element x := n=1 αn xn . Consequently, (x, y) =
α
n=1 n n(x , 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).
C ⇒: 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 εBX for some ε > 0, it suffices to settle the case
U := BX . Since U is bounded; by 7.3.4, F (U ) is ideally convex. To complete the
proof, it suffices 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 α in R+ such that αy ∈ F (X). In other words, αy ∈ F (X) for some x in X.
112 Chapter 7

If kxk ≤ 1 then there is nothing to prove. If kxk > 1 then λ := kxk−1 < 1. Whence,
using 7.3.1, infer that

αλy = (1 − λ)0 + λαy ∈ (1 − λ)F (0) + λF (x)


⊂ F ((1 − λ)0 + λx) = F (λx) ⊂ F (BX ) = F (U ).

Here use was made of the fact that kλxk = 1; i.e., λx ∈ BX . B


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
Banach’s Fundamental Principle. Nevertheless, the gap is tiny and can be easily
filled 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 modification, 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 Banach’s
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 (xn ) in X
and (yn ) in Y such that xn ∈ dom F, yn ∈ F (xn ) and xn 0 = 0 O , yn 0 = 0 O , it
follows that x ∈ dom F and y ∈ F (x). CB
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 im F . Then F carries
each neighborhood of x onto a neighborhood about y.
C A closed convex set is ideally convex and so all follows from 7.3.5. B
7.3.12. Definition. A correspondence F ⊂ X × Y 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 im F an open set. Then F is
an open correspondence.
C 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 im F . 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. B
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 im T = Y and Y is a Banach space.
C It is obvious that im T = 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 coim T := X/ ker T endowed with the quotient norm. In virtue of 5.5.4,
coim T is a Banach space. In addition, by 2.3.11 there is a unique quotient T of T
by coim T , the monoquotient of T . Taking account of the definition of quotient norm
and 5.1.3, conclude that T is a homomorphism. Furthermore, T is a monomorphism
by definition. It remains to observe that im T = im T = Y. .
7.4.3. Remark. As regards the monoquotient T : coim T 0 = 0 O of T , it may
be asserted that kT k = kT k. CB
7.4.4. Banach Homomorphism Theorem. Every bounded epimorphism
from one Banach space onto the other is a homomorphism.
C Let T ∈ B(X, Y ) and im T = Y . On applying the Open Correspondence
Principle to T , complete the proof. B
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 im T = Y ; then T −1 ∈ B(Y, X).
C A particular case of 7.4.4. B
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 briefly
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
T x = y, with T ∈ B(X, Y ) and X and Y Banach spaces, is uniquely 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.
C The correspondence T −1 is ideal and T −1 (Y ) = X. B
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 (xn )n∈N in X, together with some x in X and y
in Y satisfying xn 0 = 0 O and T xn 0 = 0 O , it happens that y = T x.
C (2): This is a reformulation of the closure property of T . B
7.4.9. Definition. A subspace X1 of a Banach space X is complemented
(rarely, a topologically complemented), if X1 is closed and, moreover, there is a closed
subspace X2 such that X = X1 ⊕ X2 (i.e., X1 ∧ X2 = 0 and X1 ∨ X2 = X). These
X1 and X2 are called complementary to one another.
7.4.10. Complementation Principle. For a subspace X1 of some Banach
space X one of the following conditions amounts to the other:
(1) X1 is complemented;
(2) X1 is the range of a bounded projection; i.e., there is a member P
of B(X) such that P 2 = P and im P = X1 .
C (1) ⇒ (2): Let P be the projection of X onto X1 along X2 (cf. 2.2.9 (4)).
Let (xn )n∈N be a sequence in X with xn 0 = 0 O and P xn 0 = 0 O . It is clear that
P xn ∈ X1 for n ∈ N. Since X1 is closed, by 4.1.19 y ∈ X1 . Similarly, the condition
(xn − P xn ∈ X2 for n ∈ N) implies that x − y ∈ X2 . Consequently, P (x − y) = 0.
Furthermore, y = P y; i.e., y = P x. It remains to refer to 7.4.8.
(2) ⇒ (1): It needs showing only that X1 equal to im P is closed. Take a se-
quence (xn )n∈N in X1 such that xn 0 = 0 O in X. Then P xn 0 = 0 O x in view of the
boundedness of P . Obtain P xn = xn , because xn ∈ im P and P is an idempotent.
Finally, x = P x, i.e. x ∈ X1 , what was required. B
7.4.11. Examples.
(1) Every finite-dimensional subspace is complemented. CB
(2) The space c0 is not complemented in l∞ .
C It turns out more convenient to work with X := l∞ (Q) and Y := c0 (Q),
where Q is the set of rational numbers. Given t ∈ R, choose a sequence (tn )
of pairwise distinct rational numbers other than t and such that tn 0 = 0 O . Let
Qt := {tn : n ∈ N}. Observe that Qt0 ∩ Qt00 is a finite set if t0 6= t00 .
Let χt be the coset containing the characteristic function of Qt in the quotient
space X/Y and V := {χt : t ∈ R}. Since χt0 6= χt00 for t0 6= t00 , the set V is
uncountable. Take f ∈ (X/Y )0 and put Vf := {v ∈ V : f (v) 6= 0}. It is evident
that Vf = ∪n∈N Vf (n), where Vf (n) := {v ∈ V : |f (v)| ≥ 1 /n }. Given P m ∈ N and
n
pairwise distinct v1 , . . . , vm in Vf (n), putPαk := |f (vk )|/f (vkP
) and x := k=1 αk vk
m m
to find kxk ≤ 1 and kf k ≥ |f (x)| = | k=1 αk f (vk )| = | k=1 |f (vk )|| ≥ m /n .
Principles of Banach Spaces 115

Hence, Vf (n) is a finite set. Consequently, Vf is countable. Whence it follows


that for every countable subset F of (X/Y )0 there is an element v of V satisfying
(∀ f ∈ F ) f (v) = 0. At the same time the countable set of the coordinate projections
δq : x 7→ x(q) (q ∈ Q) is total over l∞ (Q); i.e., (∀ q ∈ Q) δq (x) = 0 ⇒ x = 0 for
x ∈ l∞ (Q). It remains to compare the above observations. B
(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 dim X ≥ 3, each closed
subspace is the range of some projection P with kP k ≤ 1; then X is isometrically
isomorphic to a Hilbert space (this is the Kakutani Theorem). The next fact is
much deeper:
Lindenstrauss–TzafririTheorem. 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 im A is a complemented
subspace in Y. The diagram
A
X - Y

X
@
B@@
R ?
Z
is commutative for some X in B(Y, Z) if and only if ker A ⊂ ker B.
C Only ⇐ needs examining. Moreover, in the case im A = Y the sole member
X0 of L (Y, Z) such that X0 A = B is continuous. Indeed, X0−1 (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 X0 ∈ B(im A, Z) and take as X the operator X0 P ,
where P is some continuous projection of Y onto im A. (Such a projection is
available by the Complementation Principle.) B
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
A
X  Y

X
@
I 6
@
B @
Z
is commutative for some X in B(Z, Y ) if and only if im A ⊃ im B.
116 Chapter 7

C Once again only ⇐ needs examining. Using the definition of complemented


subspace, express Y as the direct sum of ker A and Y0 , where Y0 is a closed subspace.
By 5.5.9 (1), Y0 is a Banach space. Consider the part A0 of A in Y0 . Undoubtedly,
im A0 = im A ⊃ im B. Consequently, by 2.3.13 and 2.3.14 the equation A0 X0 = B
has a unique solution X0 := A−10 . It suffices to prove that the operator X0 , treated
as a member of L (Z, Y0 ), is bounded.
The operator X0 is closed. Indeed (cf. 7.4.8), if zn 0 = 0 O and A−10 Bzn 0 = 0
O
O
then Bzn 0 = 0 z, since B is bounded. In addition, by the continuity of A0 ,
the correspondence A−1 0 ⊂ X × Y0 is closed; and so 7.3.10 yields the equality
y = A−1
0 Bz. B
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.
C For definiteness, assume that k · k2  k · k1 in X. Consider the diagram
IX
(X, k · k1 )  (X, k · k2 )

X
@
I 6
@
IX @
(X, k · k1 )
By the Phillips Theorem some continuous operator X makes the diagram commu-
tative. Such an operator is unique: it is IX . B
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, define the graph norm of x as
kxkgr T := kxkX + kT xkY . Then k · kgr T ∼ k · kX .
C Observe that the space (X, k · kgr T ) is complete. Further, k · kgr T ≥ k · kX .
It remains to refer to the Two Norm Principle. B
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 = X1 ⊕ X2 , where
X1 , X2 ∈ Lat(X). The subspaces X1 and X2 are closed if and only if each of them
is a Banach range.
Principles of Banach Spaces 117

C ⇒: 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 kzk0 := kT −1 zkY . It is clear that (Z, k · k0 ) is
a Banach space and kzk = kT T −1 zk ≤ kT kkT −1 zk = kT kkzk0 ; i.e., k · k0  k · kZ .
Applying this construction to X1 and X2 , obtain Banach spaces (X1 , k · k1 ) and
(X2 , k · k2 ). Now k · kk  k · kX on Xk for k := 1, 2.
Given x1 ∈ X1 and x2 ∈ X2 , put kx1 + x2 k0 := kx1 k1 + kx2 k2 . Thereby we
introduce in X some norm k · k that is stronger than the initial norm k · kX . By con-
struction (X, k · k0 ) is a Banach space. It remains to refer to 7.4.16. B

7.5. The Automatic Continuity Principle


7.5.1. Lemma. Let f : X0 = 0 O R· be a convex function on a (multi)normed
space X. The following statements are equivalent:
(1) U := int dom f 6= ∅ and f |U is a continuous function;
(2) there is a nonempty open set V such that sup f (V ) < +∞.
C (1) ⇒ (2): This is obvious.
(2) ⇒ (1): It is clear that U 6= ∅. 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 α ∈ R+ such that
α ≤ 1 and for an arbitrary v in W , obtain

f (αv) = f (αv + (1 − α)0) ≤ αf (v) + (1 − α)f (0) = αf (v);


f (αv) + αf (−v) ≥ f (αv) + f (α(−v))
= 2(1 /2 f αv) + 1 /2 f (−αv) ≥ 2f (0) = 0.


Therefore, |f (αW )| ≤ αt, which implies that f is continuous at zero. B


7.5.2. Corollary. If x ∈ int dom f and f is continuous at x then the sub-
differential ∂x (f ) contains only continuous functionals.
C 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. B
7.5.3. Corollary. Every convex function on a finite-dimensional space is con-
tinuous on the interior of its domain. CB
7.5.4. Definition. A function f : X0 = 0 O R· is called ideally convex if 0 = 0 OO
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.
C Let f be such function. If core dom f = ∅ then there is nothing to prove.
If x ∈ core dom f then put t := f (x) and F := (0 = 0 OO )−1 ⊂ R × X. Applying
the Ideal Correspondence Principle, find a δ > 0 from the condition F (t + BR ) ⊃
x + δBX . Whence, in particular, infer the estimate f (x + δBX ) ≤ t + 1. In virtue
of 7.5.1, f is continuous on int dom f . Since x ∈ int dom f ; therefore, by Lemma
7.1.1, core dom f = int dom f . B
7.5.6. Remark. Using 7.3.6, it is possible to prove that an ideally convex
function f , defined in a Banach space on a subset with nonempty core, is locally
Lipschitz on int dom f . In other words, given x0 ∈ int dom f , there are a positive
number L and a neighborhood U about x0 such that kf (x) − f (x0 )k ≤ Lkx − x0 k
whenever x ∈ U . CB
7.5.7. Corollary. Let f : X0 = 0 O R· be an ideally convex function on a Ba-
nach space X and x ∈ core dom f . Then the directional derivative f 0 (x) is a con-
tinuous sublinear functional and ∂x (f ) ⊂ X 0 .
C Apply the Automatic Continuity Principle twice. B
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 subdifferential of a function
f : X0 = 0 O R· at a point x; i.e., we agree to define
∂x (f ) := ∂x (f ) ∩ X 0 .
Proceed likewise in (multi)normed spaces. If a need is felt to distinguish the “old”
(wider) subdifferential, a subset of X # , from the “new” (narrower) subdifferential,
a subset of X 0 ; then the first 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 subdifferentials. Observe finally
that if f := p is a seminorm on X then, for similar reasons, it is customary to put
|∂|(p) := |∂|(p) ∩ X 0 .
7.5.9. Ideal Hahn–Banach Theorem. Let f : Y 0 = 0 O 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 T x ∈ core dom f then
∂x (f ◦ T ) = ∂ T x (f ) ◦ T.
C The right side of the sought formula is included into its left side for obvious
reasons. If l in X 0 belongs to ∂x (f ◦ T ), then by the Hahn–Banach Theorem there
is an element l1 of the algebraic subdifferential of f at T x for which l = l1 ◦ T .
It suffices to observe that, in virtue of 7.5.7, l1 is an element of Y 0 and so it is
a member of the topological subdifferential ∂ T x (f ). B
Principles of Banach Spaces 119

7.5.10. Balanced Hahn–Banach Theorem. Suppose that X and Y are


normed spaces. Given T ∈ B(X, Y ), let p : Y 0 = 0 O R be a continuous seminorm.
Then
|∂|(p ◦ T ) = |∂|(p) ◦ T.
C If l ∈ |∂|(p ◦ T ) then l = l1 ◦ T for some l1 in the algebraic balanced
subdifferential of p (cf. 3.7.11). From 7.5.2 it follows that l1 is continuous. Thus,
|∂|(p ◦ T ) ⊂ |∂|(p) ◦ T . The reverse inclusion raises no doubts. B
7.5.11. Continuous Extension Principle. Let X0 be a subspace of X and
let l0 be a continuous linear functional on X0 . Then there is a continuous linear
functional l on X extending l0 and such that klk = kl0 k.
C Take p := kl0 k k · k, and consider the identical embedding ι : X0 0 = 0 O .
On account of 7.5.10, l0 ∈ |∂|(p ◦ ι) = |∂|(p) ◦ ι = kl0 k |∂|(k · k) ◦ ι. It suffices to
observe that |∂|(k · kX ) = BX 0 . B
7.5.12. Topological Separation Theorem. Let U be a convex set with
nonempty interior in a space X. If L is an affine variety in X and L ∩ int U = ∅
then there is a closed hyperplane H in X such that H ⊃ L and H ∩ int U = ∅. CB
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. CB
7.5.14. Corollary. Let X0 be a subspace of X. Then

cl X0 = ∩ {ker f : f ∈ X 0 , ker f ⊃ X0 }.

C It is clear that (f ∈ X 0 & ker f ⊃ X0 ) ⇒ ker f ⊃ cl X0 . If x0 ∈ / cl X0


then there is an open convex neighborhood about x0 disjoint from cl X0 . In virtue
of 7.5.12 and 7.5.13 there is a functional f0 , a member of (XR )0 , such that ker f0 ⊃
cl X0 and f0 (x0 ) = 1. From the properties of the complexifier infer that the func-
tional Re−1 f0 vanishes on X0 and differs from zero at the point x0 . It is also beyond
a doubt that the functional is continuous. B

7.6. Prime Principles


7.6.1. Let X and Y be (multi)normed vector spaces (over the same ground
field F). Assume further that X 0 and Y 0 are the duals of X and Y respectively.
Take a continuous linear operator T from X to Y . Then y 0 ◦ T ∈ X 0 for y 0 ∈ Y 0
and the mapping y 0 7→ y 0 ◦ T is a linear operator. CB
7.6.2. Definition. The operator T 0 : Y 0 0 = 0 O 0 , constructed in 7.6.1, is the
dual or transpose of T : X0 = 0 O .
120 Chapter 7

7.6.3. Theorem. The prime mapping T 7→ T 0 implements a linear isometry


of the space B(X, Y ) into the space B(Y 0 , X 0 ).
C The prime mapping is clearly a linear operator from B(X, Y ) to L (Y 0 , X 0 ).
Furthermore, since kyk = sup{|l(y) : l ∈ |∂|(k · k)}; therefore,

kT 0 k = sup{kT 0 y 0 k : ky 0 k ≤ 1}
= sup{|y 0 (T x)| : ky 0 k ≤ 1, kxk ≤ 1} = sup{kT xk : kxk ≤ 1} = kT k,

what was required. B


7.6.4. Examples.
(1) Let X and Y be Hilbert spaces. Take T ∈ B(X, Y ). Observe first
that, in a plain sense, T ∈ B(X, Y ) ⇔ T ∈ B(X∗ , Y∗ ). Denote the prime mapping
of X by (·)0X : X∗ 0 = 0 O 0 , i.e., x 7→ x0 := ( · , x); and denote the prime mapping
of Y by (·)0Y : Y∗ 0 = 0 O 0 , i.e., y 7→ y 0 := ( · , y).
The adjoint of T , the member T ∗ of B(Y, X), and the dual of T , the mem-
ber T 0 of B(Y 0 , X 0 ), are related by the commutative diagram:

T∗
X∗ ←−Y∗
0
(·)X ↓ ↓ (·)0Y
T0
X 0 ←−Y 0

C Indeed, it is necessary to show the equality T 0 y 0 = (T ∗ y)0 for y ∈ Y . Given


x ∈ X, by definition observe that

T 0 y 0 (x) = y 0 (T x) = (T x, y) = (x, T ∗ y) = (T ∗ y)0 (x).

Since x is arbitrary, the proof is complete. B


(2) Let ι : X0 0 = 0 O be the identical embedding of X0 into X. Then
ι : X0 = 0 O 00 . Moreover, ι0 (x0 )(x0 ) = x0 (x0 ) for all x0 ∈ X0 and x0 ∈ X 0 and ι0 is
0
ι0
an epimorphism; i.e., X 0 −→ X 0 0 = 0 O is an exact sequence. CB
0
T
7.6.5. Definition. Let an elementary diagram X −→ Y be given. The dia-
T0
gram Y 0 −→ X 0 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
T 00
7.6.6. Double Prime Lemma. Let X 00 −→ Y 00 be the diagram that results
T
from setting primes in the diagram X −→ Y twice. Then the following diagram
commutes:
T
X −→Y
00
↓ ↓00
T 00
X 00 −→Y 00
Here 00 : X0 = 0 O 00 and 00 : Y 0 = 0 O 00 are the respective double prime mappings;
i.e. the canonical embeddings of X into X 00 and of Y into Y 00 (cf. 5.1.10 (8)).
C Let x ∈ X. We have to prove that T 00 x00 = (T x)00 . Take y 0 ∈ Y 0 . Then

T 00 x00 (y 0 ) = x00 (T 0 y 0 ) = T 0 y 0 (x) = y 0 (T x) = (T x)00 (y 0 ).

Since y 0 ∈ Y 0 is arbitrary, the proof is complete. B


7.6.7. Diagram Prime Principle. A diagram is commutative if and only
if so is its diagram prime.
C It suffices to convince oneself that the triangles

T T0
X −→ Y X 0 ←− Y 0
R& .S 0 0
R - %S
Z 0
Z

are commutative or not simultaneously. Since R = ST ⇒ R0 = (ST )0 = T 0 S 0 ;


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 R00 = S 00 T 00 . Using 7.6.6, argue as follows: (Rx)00 = R00 x00 = S 00 T 00 x00 =
S 00 (T 00 x00 ) = S 00 (T x)00 = (ST x)00 for all x ∈ X. Consequently, R = ST . B
7.6.8. Definition. Let X0 be a subspace of X and let X0 be a subspace
of X 0 . Put

X0⊥ := {f ∈ X 0 : ker f ⊃ X0 } = |∂|(δ(X0 ));



X0 := {x ∈ X : f ∈ X0 ⇒ f (x) = 0} = ∩ {ker f : f ∈ X0 }.

The subspace X0⊥ is the (direct) polar of X0 , and the subspace ⊥


X0 is the reverse
polar of X0 . 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 im T 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 im T .
C ⇒: The Banach Homomorphism Theorem.
⇐: Refer to 7.4.2. B
7.6.11. Polar Lemma. Let T ∈ B(X, Y ). Then
(1) (im T )⊥ = ker(T 0 );
(2) if T is normally solvable then

im T = ⊥ ker(T 0 ), (ker T )⊥ = im(T 0 ).

C (1): y 0 ∈ ker(T 0 ) ⇔ T 0 y 0 = 0
⇔ (∀ x ∈ X) T 0 y 0 (x) = 0 ⇔ (∀ x ∈ X) y 0 (T x) = 0 ⇔ y 0 ∈ (im T )⊥ .
(2): The equality cl im T = ⊥ ker(T 0 ) follows from 7.5.13. Furthermore, by hy-
pothesis im T is closed.
If x0 = T 0 y 0 and T x = 0 then x0 (x) = T 0 y 0 (x) = y 0 (T x) = 0, which means
that x0 ∈ (ker T )⊥ . Consequently, im(T 0 ) ⊂ (ker T )⊥ . Now take x0 ∈ (ker T )⊥ .
Considering the operator T acting onto im T , apply the Sard Theorem to the left
side of the diagram
T
X −→ im T −→ Y
x0 & ↓ y00 . y 0
F
As a result of this, obtain y 00 in (im T )0 such that y 00 ◦T = x0 . By the Continuous
Extension Principle there is an element y 0 of Y 0 satisfying y 0 ⊃ y 00 . Thus, x0 = T 0 y 0 ,
i.e. x0 ∈ im(T 0 ). B
7.6.12. Hausdorff 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 0 is normally
solvable.
C ⇒: In virtue of 7.6.11 (2), im(T 0 ) = (ker T )⊥ . Evidently, the subspace
(ker T )⊥ is closed.
⇐: To begin with, suppose cl im T = Y . It is clear that 0 = Y ⊥ = (cl im T )⊥
= (im T )⊥ = ker(T 0 ) in virtue of 7.6.11. By the Banach Isomorphism Theorem
there is some S ∈ B(im(T 0 ), Y 0 ) such that ST 0 = IY 0 . The case r := kSk = 0 is
trivial. Therefore, it may be assumed that kT 0 y 0 k ≥ 1 /r ky 0 k for all y 0 ∈ Y 0 .
Show now that cl T (BX ) ⊃ 1 /2r BY . With this at hand, from the ideal con-
vexity of T (BX ) it is possible to infer that T (BX ) ⊃ 1 /4r BY . The last inclusion
implies that T is a homomorphism.
Let y ∈/ cl T (BX ). Then y does not belong to some open convex set including
T (BX ). 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, find some nonzero y 0 in Y 0 satisfying


ky 0 k kyk ≥ y 0 (y) ≥ sup y 0 (T x) = kT 0 y 0 k ≥ 1 /r ky 0 k.
kxk≤1

Whence kyk ≥ 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 Y0 := cl im T and let ι : Y0 0 = 0 O be
the identical embedding. Then T = ιT , where T : X0 = 0 O 0 is the operator acting
by the rule T x = T x for x ∈ X. In addition, im(T 0 ) = im(T 0 ι0 ) = T 0 (im(ι0 )) =
T 0 (Y00 ), because ι0 (Y 0 ) = Y00 (cf. 7.6.4 (2)). Thus, T 0 is a normally solvable operator.
By what was proven, T is normally solvable. It remains to observe that im T =
im T . B
7.6.13. Sequence Prime Principle. A sequence
. . . 0 = 0 O k−1 −−−k−→ Xk −−−−→ Xk+1 0 = 0 O
T Tk+1

is exact if and only if so is the sequence prime


0
0 Tk0 Tk+1
... ← Xk−1 ←−−−− Xk0 0
←−−−− Xk+1 ← ... .
C ⇒: Since im Tk+1 = ker Tk+2 ; therefore, Tk+1 is normally solvable. Using
the Polar Lemma, conclude that
ker(Tk0 ) = (im Tk )⊥ = (ker Tk+1 )⊥ = im(Tk+1
0
).
⇒: By the Hausdorff Theorem Tk+1 is normally solvable. Once again appealing
to 7.6.11 (2), infer that
(im Tk )⊥ = ker(Tk0 ) = im(Tk+1
0
) = (ker Tk+1 )⊥ .
Since Tk is normally solvable by Theorem 7.6.12; therefore, im Tk is a closed sub-
space. Using 7.5.14, observe that
im T k = ⊥ ((im Tk )⊥ ) = ⊥ ((ker Tk+1 )⊥ ) = ker Tk+1 .
Here account was taken of the fact that ker Tk+1 itself is a closed subspace. B
7.6.14. Corollary. For a normally solvable operator T , the following isomor-
phisms hold: (ker T )0 ' coker(T 0 ) and (coker T )0 ' ker(T 0 ).
C By virtue of 2.3.5 (6) the sequence
00 = 0 O T 0 = 0 O −→ Y 0 = 0 O T 0 = 0 O
T

is exact. From 7.6.13 obtain that the sequence


0
00 = 0 O coker T )0 0 = 0 O 0 −→ X 0 0 = 0 O ker T )0 0 = 0 O
T

is exact. B
7.6.15. Corollary. T is an isomorphism ⇔ T 0 is an isomorphism. CB
7.6.16. Corollary. Sp(T ) = Sp(T 0 ). CB
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 : X0 = 0 O . Prove that kT xkY ≥ tkxkX for some
strictly positive t and all x ∈ X if and only if ker T = 0 and im T 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 l1 (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 finite-dimensional.
7.8. Let X and Y be different Banach spaces, with X continuously embedded into Y . Es-
tablish that X is a meager subset of Y .
7.9. Let X1 and X2 be nonzero closed subspaces of a Banach space and X1 ∩ X2 = 0. Prove
that the sum X1 + X2 is closed if and only if the next quantity

inf{kx1 − x2 k/kx1 k : x1 6= 0, x1 ∈ X1 , x2 ∈ X2 }

is strictly positive.
7.10. Let (amn ) be a double countable sequence such that there is a sequence (x(m) ) of
P∞ (m)
elements of l1 for which all the series
P∞amn xn fail to converge in norm. Prove that there
n=1
is a sequence x in l1 such that the series a x fail to converge in norm for all m ∈ N.
n=1 mn n
7.11. Let T be an endomorphism of a Hilbert space H which satisfies hT x | yi = hx | T yi 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 = x1 − x2 with
x1 , x2 ∈ X + , the estimates hold: kx1 k ≤ tkxk and kx2 k ≤ tkxk.
7.13. Let lower semicontinuous sublinear functionals p and q on a Banach space X be such
that the cones dom p and dom q are closed and the subspace dom p − dom q = dom q − dom p is
complemented in X. Prove that
∂ (p + q) = ∂ (p) + ∂ (q)
in the case of topological subdifferentials (cf. Exercise 3.10).
7.14. Let p be a continuous sublinear functional defined on a normed space X and let T be
a continuous endomorphism of X. Assume further that the dual T 0 of T takes the subdifferential
∂ (p) into itself. Establish that ∂ (p) contains a fixed point of T 0 .
7.15. Given a function f : X0 = 0 O R· on a (multi)normed space X, put

f ∗ (x0 ) := sup{hx | x0 i − f (x) : x ∈ dom f } (x0 ∈ X 0 );


f ∗∗ (x) := sup{hx | x0 i − f ∗ (x0 ) : x0 ∈ 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 infinite-dimensional complemented
subspaces is isomorphic to X. Verify that c0 and lp (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 im T is nonmeager. Prove that T is normally solvable.
7.19. Let X0 be a closed subspace of a normed space X. Assume further that X0 and X/X0
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 ƒ of the ball BX 0
in the dual space X 0 is called norming (for X) if kxk = sup{|l(x)| : l ∈ ƒ}
for all x ∈ X. If each subset U of X satisfies sup kU k < +∞ on condition that
sup{|l(u)| : u ∈ U } < +∞ for all l ∈ ƒ, then ƒ is a fully norming set.
8.1.2. Examples.
(1) The ball BX 0 is a fully norming set in virtue of 5.1.10 (8) and 7.2.7.
(2) If ƒ0 is a (fully) norming set and ƒ0 ⊂ ƒ1 ⊂ BX 0 then ƒ1 itself is
a (fully) norming set.
(3) The set ext BX 0 of the extreme points of BX 0 is norming in virtue
of the Kreı̆n–Milman Theorem in subdifferential form and the obvious equality
BX 0 = |∂|(k · kX ) which has already been used many times. However, ext BX 0
can fail to be fully norming (in particular, the possibility is realized in the space
C([0, 1], R)). CB
(4) Let X and Y be Banach spaces (over the same ground field F) and
let ƒY be a norming set for Y . Put

ƒB := {δ(y,0 x) : y 0 ∈ ƒY , x ∈ BX },

where δ(y,0 x) (T ) := y 0 (T x) for y 0 ∈ Y, x ∈ X and T ∈ B(X, Y ). It is clear that

|δ(y,0 x) (T )| = |y 0 (T x)| ≤ ky 0 k kT xk ≤ ky 0 k kT k kxk;

i.e., δ(y,0 x) ∈ B(X, Y )0 . Furthermore, given T ∈ B(X, Y ), infer that

kT k = sup{kT xk : kxk ≤ 1} = sup{|y 0 (T x)| : y 0 ∈ ƒY , kxk ≤ 1}


= sup{|δ(y,0 x) (T )| : δ(y,0 x) ∈ ƒB }.
Operators in Banach Spaces 127

Therefore, ƒB is a norming set for B(X, Y ). If ƒY is a fully norming set then ƒB


is also a fully norming set. Indeed, if U is such that the numeric set {|y 0 (T x)| :
T ∈ U } is bounded in R for all x ∈ BX and y 0 ∈ ƒY , then by hypothesis the
set {T x : T ∈ U } is bounded in Y for every x in X. By virtue of the Uniform
Boundedness Principle it means that sup kU k < +∞.
8.1.3. Dunford–Hille Theorem. Let X be a complex Banach space and let
ƒ be a fully norming set for X. Assume further that f : D0 = 0 O is an X-va-
lued function with domain D an open (in C R ' R2 ) subset of C. The following
statements are equivalent:
(1) for every z0 in D there is a limit
f (z) − f (z0 )
limO ;
z0=0 0 z − z0
(2) for all z0 ∈ D and l ∈ ƒ there is a limit
l ◦ f (z) − l ◦ f (z0 )
limO ;
z0=0 0 z − z0
i.e., the function l ◦ f : D0 = 0 O C is holomorphic for l ∈ ƒ.
C (1) ⇒ (2): This is obvious.
(2) ⇒ (1): For the sake of simplicity assume that z0 = 0 and f (z0 ) = 0. Con-
sider the disk of radius 2ε centered at zero and included in D; i.e., 2εD ⊂ D, where
D := BC := {z ∈ C : |z| ≤ 1} is the unit disk. As is customary in complex analysis,
treat the disk εD as an (oriented) compact manifold with boundary εT, where T is
the (properly oriented) unit circle T := {z ∈ C : |z| = 1} in the complex plane C R .
Take z1 , z2 ∈ εD \ 0 and the holomorphic function l ◦ f (the functional l lies in ƒ).
Specifying the Cauchy Integral Formula, observe that
l ◦ f (zk ) l ◦ f (z)
Z
1
= dz (k := 1, 2).
zk 2πi z(z − zk )
2εT

If now z1 6= z2 then, using the condition |z − zk | ≥ ε (k := 1, 2) for z ∈ 2εT and


the continuity property of the function l ◦ f on D, find
  
1 f (z1 ) f (z2 )
l
z1 − z2 −
z1 z2

Z  
1 1 1 1
= · l ◦ f (z) − dz
z1 − z2 2πi z(z − z1 ) z(z − z2 )
2εT

Z
1 1
= l ◦ f (z) dz ≤ M sup |l ◦ f (z)| < +∞
2π z(z − z1 )(z − z2 ) z∈2εT
2εT
128 Chapter 8

for a suitable M > 0. Since ƒ is a fully norming set, conclude that



1 f (z1 ) f (z2 )
sup z1 − z2 < +∞.

z 6=z ;z ,z 6=0 |z1 − z2 |
1 2 1 2
|z1 |≤ε,|z2 |≤ε

This final inequality ensures that the sought limit exists. B


8.1.4. Definition. A mapping f : D0 = 0 O satisfying 8.1.3 (1) (or, which is
the same, 8.1.3 (2) for some fully norming set ƒ) is called holomorphic.
8.1.5. Remark. A meticulous terminology is used sometimes. Namely, if f
satisfies 8.1.3 (1) then f is called a strongly holomorphic function. If f satisfies
8.1.3 (2) with ƒ := BX 0 then f is called weakly holomorphic. Under the hypotheses
of 8.1.3 (2) and 8.1.2 (4), i.e. for f : D0 = 0 O (X, Y ), ƒY := BY 0 and the corre-
sponding ƒ := ƒB , the expression, “f is weakly operator holomorphic,” is employed.
With regard to this terminology, the Dunford–Hille 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 finitely
many disjoint simple loops. An obvious meaning is ascribed to the integrals:
Namely, given a loop γ, choose an appropriate (smooth) parametrization ‰ : T0 =
0 O (with orientation accounted for) and put
Z Z
f (z)dz := f ◦ ‰d‰,
γ T

with the integral treated for instance as a suitable Bochner integral (cf. 5.5.9 (6)).
The soundness of the definition is beyond a doubt, since the needed Bochner integral
exists independently of the choice of the parametrization ‰.
8.1.7. Cauchy–Wiener Integral Theorem. Let D be an nonempty open
subset of the complex plane and let f : D0 = 0 O be a holomorphic X-valued
function, with X a Banach space. Assume further that F is a rough draft for the
pair (∅, D). Then Z
f (z)dz = 0.
∂F
Moreover, Z
1 f (z)
f (z0 ) = dz
2πi z − z0
∂F
for z0 ∈ int F .
Operators in Banach Spaces 129

C 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). B
8.1.8. Remark. The Cauchy–Wiener 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 : D0 = 0 O be a holomor-
phic X-valued function, with X a Banach space, and take z0 ∈ D. In every open
disk U := {z ∈ C : |z − z0 | < ε} such that cl U lies in D, the Taylor series expansion
holds (in a compactly convergent power series, cf. 7.2.10):


X
f (z) = cn (z − z0 )n ,
n=0

where the coefficients cn , members of X, are calculated by the formulas:

1 dn f
Z
1 f (z)
cn = dz = (z0 ).
2πi (z − z0 )n+1 n! dz n
∂U

C The proof results from a standard argument: Expand the kernel u 7→


(u − z)−1 of the formula
Z
1 f (u)
f (z) = du (z ∈ cl U )
2πi u−z
∂U 0

in the powers of z − z0 ; i.e.,


1 1 X (z − z0 )n
=  = n+1
.
u−z (u − z0 ) 1 − z−z0
n=0
(u − z0 )
u−z0

The last series converges uniformly in u ∈ ∂U 0 . (Here U 0 = U + qD for some


q > 0 such that cl U 0 ⊂ D.) Taking it into account that sup kf (∂U 0 )k < +∞
and integrating, arrive at the sought presentation of f (z) for z ∈ cl U . Applying
this to U 0 and using 8.1.7, observe that the power series under study converges
in norm at every point of U 0 . This yields uniform convergence on every compact
subset of U 0 , and so on U . B
130 Chapter 8

8.1.10. Liouville Theorem. If an X-valued function f : C0 = 0 O , with X


a Banach space, is holomorphic and sup kf (C)k < +∞ then f is a constant map-
ping.
C Considering the disk εD with ε > 0 and taking note of 8.1.9, infer that

kcn k ≤ sup kf (z)k · ε−n ≤ sup kf (C )k · ε−n


z∈εT

for all n ∈ N and ε > 0. Therefore, cn = 0 for n ∈ N. B


8.1.11. Each bounded endomorphism of a nonzero complex Banach space has
a nonempty spectrum.
C 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. Furthermore,
by 5.6.15, kR (T, λ)k0 = 0 O as |λ|0 = 0 O ∞. By virtue of 8.1.10 conclude that
R (T, · ) = 0. At the same time, using 5.6.15, observe that R (T, λ)(λ − T ) = 1 for
|λ| > kT k. A contradiction. B
8.1.12. The Beurling–Gelfand formula holds:

r(T ) = sup{|λ| : λ ∈ 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 .
C 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 λ ∈ C so that |λ| > sup{|µ| : µ ∈ Sp(T )}. Then the disk of radius
|λ|−1 lies entirely in the domain of the holomorphic function (cf. 5.6.15)

R (T, z −1 ), for z 6= 0 and z −1 ∈ res(T )



f (z) :=
0, for z = 0.

Using 8.1.9 and 5.6.17, conclude that |λ|−1 < r(T )−1 . Consequently, |λ| > r(T ). B
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 : dom f 0 =
0 O C is a holomorphic function with dom f ⊃ K). Given f1 , f2 ∈ H(K), let the
notation f1 ∼ f2 mean that it is possible to find an open subset D of dom f1 ∩dom f2
satisfying K ⊂ D and f1 |D = f2 |D . Then ∼ is an equivalence in H(K). CB
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 f1 , f2 ∈ f and g1 , g2 ∈ g. Put

x ∈ dom f1 ∩ dom g1 ⇒ ϕ1 (x) := f1 (x) + g1 (x),


x ∈ dom f2 ∩ dom g2 ⇒ ϕ2 (x) := f2 (x) + g2 (x).

Then ϕ1 , ϕ2 ∈ H(K) and ϕ1 = ϕ2 .


C Choose open sets D1 and D2 such that K ⊂ D1 ⊂ dom f1 ∩ dom f2 and
K ⊂ D2 ⊂ dom g1 ∩ dom g2 , with f1 |D1 = f2 |D1 and g1 |D2 = g2 |D2 . Observe now
that ϕ1 and ϕ2 agree on D1 ∩ D2 . B
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 defined in 8.1.16 is an algebra. CB
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 \ K0 = 0 O be an
X-valued holomorphic function, with X a Banach space. Further, take f ∈ H (K)
and f1 , f2 ∈ f . If F1 is a rough draft for the pair (K, dom f1 ) and F2 is a rough
draft for the pair (K, dom f2 ) then
Z Z
f1 (z)R (z)dz = f2 (z)R (z)dz.
∂F1 ∂F2

C Let K ⊂ D ⊂ int F1 ∩ int F2 , with D open and f1 |D = f2 |D . Choose a rough


draft F for the pair (K, D). Since f1 R is holomorphic on dom f1 \ K and f2 R is
holomorphic on dom f2 \ K, infer the equalities

Z Z
f1 (z)R (z)dz = f1 (z)R (z)dz,
∂F ∂F1
Z Z
f2 (z)R (z)dz = f2 (z)R (z)dz
∂F ∂F2

(from the nontrivial fact of the validity of their scalar analogs). Since f1 and f2
agree on D, the proof is complete. B
132 Chapter 8

8.1.20. Definition. Under the hypotheses of 8.1.19, given an element h


in H (K), define the contour integral of h with kernel R as the element
I Z
h(z)R (z)dz := f (z)R (z)dz,
∂F

where h = f and F is a rough draft for the pair (K, dom f ).


8.1.21. Remark. The notation h(z) in 8.1.20 is far from being ad hoc. It is
well justified by the fact that w := f1 (z) = f2 (z) for every point z in K and every
two members f1 and f2 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
I
1
RT h := h(z)R (T, z)dz
2πi

and called the Riesz–Dunford integral (of the germ h). If f is a function holomorphic
in a neighborhood about Sp(T ) then put f (T ) := RT f := RT f . We also use more
suggestive designations like
I
1 f (z)
f (T ) = dz.
2πi z−T

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 A1 and A2 be two algebras (over the same field). A morphism from A1
to A2 or a representation of A1 in A2 (rarely, over A2 ) is a multiplicative linear
operator R, i.e. a member R of L (A1 , A2 ) such that R(ab) = R(a)R(b) for all
a, b ∈ A1 . The expression, “T represents A1 in A2 ,” is also in common parlance.
A representation R is called faithful if ker R = 0. The presence of a faithful repre-
sentation R : A1 0 = 0 O 2 makes it possible to treat A1 as a subalgebra of A2 .
Operators in Banach Spaces 133

If A2 is a (sub)algebra of the endomorphism algebra L (X) of some vector


space X (over the same field), then a morphism of A1 in A2 is referred to as
a (linear) representation of A1 on X or as an operator representation of A1 . The
space X is then called the representation space for the algebra A1 .

Given a representation R, suppose that the representation space X for A has


a subspace X1 invariant under all operators R(a), a ∈ A. Then the representa-
tion R1 : A0 = 0 O L(X1 ) arises naturally, acting by the rule R1 (a)x1 = R(a)x1
for x1 ∈ X1 and a ∈ A and called a subrepresentation of R (induced in X1 ).
If X = X1 ⊕ X2 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
R1 and R2 (induced in X1 and X2 ). We mention in passing the significance of
studying arbitrary irreducible representations, each of which has only trivial sub-
representations by definition.

8.2.3. Gelfand–Dunford Theorem. Let T be a bounded endomorphism


of a Banach space X. The Riesz–Dunford integral RT represents the algebra of
germs of holomorphic functions on the spectrum of T on the space X. Moreover,
P∞ P∞
if f (z) = n=0 cn z n (in a neighborhood about Sp(T )) then f (T ) = n=0 cn T n
(summation is understood in the operator norm of B(X)).

C There is no doubt that RT is a linear operator. Check the multiplicativity


property of RT . To this end, take f 1 , f 2 ∈ H (Sp(T )) and choose rough drafts F1
and F2 such that Sp(T ) ⊂ int F1 ⊂ F1 ⊂ int F2 ⊂ F2 ⊂ D, with the functions f1 in
f 1 and f2 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
Z Z
1 1 f1 (z1 ) f2 (z2 )
RT f 1 ◦ RT f 2 = f1 (T )f2 (T ) = dz1 ◦ dz2
2πi 2πi z1 − T z2 − T
∂F1 ∂F2
 
Z Z
1 1
=  f1 (z1 )R (T, z1 )dz1  f2 (z2 )R (T, z2 )dz2
2πi 2πi
∂F2 ∂F
Z Z1
1 1
= f1 (z1 )f2 (z2 )R (T, z1 )R (T, z2 )dz2 dz1
2πi 2πi
∂F1 ∂F2
R (T, z1 ) − R (T, z2 )
Z Z
1 1
= f1 (z1 )f2 (z2 ) dz2 dz1
2πi 2πi z2 − z1
∂F1 ∂F2
 
Z Z
1 1 f2 (z2 )
= f1 (z1 )  dz2  R (T, z1 )dz1
2πi 2πi z2 − z1
∂F1 ∂F2
 
Z Z
1 1 f1 (z1 )
− f2 (z2 )  dz1  R (T, z2 )dz2
2πi 2πi z2 − z1
∂F2 ∂F1
Z
1
= f1 (z1 )f2 (z1 )R (T, z1 )dz1 − 0 = f1 f2 (T ) = RT (f 1 f 2 ).
2πi
γ

Choose a circle γ := εT that lies in res(T ) as well as in the (open) disk of


P∞
convergence of the series f (z) = n=0 cn z n . From 5.6.16 and 5.5.9 (6) derive
Z ∞
1 X
f (T ) = f (z) z −n−1 T n dz
2πi n=0
γ
 
∞ Z ∞ Z ∞
1 X
−n−1 n
X 1 f (z)  n X
= f (z)z T dz = 
n+1
dz T = cn T n
2πi n=0 n=0
2πi z n=0
γ γ

in virtue of 8.1.9. B
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

C Assume first that λ ∈ Sp(f (T )) and f −1 (λ) ∩ Sp(T ) = ∅. Given z ∈


(C \ f −1 (λ)) ∩ dom f , put g(z) := (λ − f (z))−1 . Then g is a holomorphic function
on a neighborhood of Sp(T ), satisfying g(λ − f ) = (λ − f )g = 1C . Using 8.2.3,
observe that λ ∈ res(f (T )), a contradiction. Consequently, f −1 (λ) ∩ Sp(T ) 6= ∅;
i.e., Sp(f (T )) ⊂ f (Sp(T )).
Now take λ ∈ Sp(T ). Put
f (λ) − f (z)
λ 6= z ⇒ g(z) := ; g(λ) := f 0 (λ).
λ−z
Clearly, g is a holomorphic function (the singularity is “removed”). From 8.2.3
obtain
g(T )(λ − T ) = (λ − T )g(T ) = f (λ) − f (T ).
Consequently, if f (λ) ∈ res(f (T )) then the operator R (f (T ), f (λ))g(T ) is inverse
to λ − T . In other words, λ ∈ res(T ), which is a contradiction. Thus,
f (λ) ∈ C \ res(f (T )) = Sp(f (T ));
i.e., f (Sp(T )) ⊂ Sp(f (T )). B
8.2.6. Let K be a nonempty compact subset of C and let g : dom g0 = 0 O C

be a holomorphic function with dom g ⊃ 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). CB
8.2.7. Dunford Theorem. For every function g : dom g0 = 0 O C holomorphic
in the neighborhood dom g about the spectrum Sp(T ) of an endomorphism T of
a Banach space X, the following diagram of representations commutes:


g
H (Sp(T ))  H (Sp(g(T )))
@
RT @ Rg(T )
@

@
R ?
@
B(X)

C Let f ∈ H (g(Sp(T ))) with f : D0 = 0 O C such that f ∈ f and D ⊃


g(Sp(T )) = Sp(g(T )). Let F1 be a rough draft for the pair (Sp(g(T )), D) and
let F2 be a rough draft for the pair (Sp(T ), g −1 (int F1 )). It is clear that now
g(∂F2 ) ⊂ int F1 and, moreover, the function z2 7→ (z1 − g(z2 ))−1 is defined and
holomorphic on int F2 for z1 ∈ ∂F1 . Therefore, by 8.2.3
Z
1 R (T, z2 )
R (g(T ), z1 ) = dz2 (z1 ∈ ∂F1 ).
2πi z1 − g(z2 )
∂F2
136 Chapter 8

From this equality, successively derive


 
Z Z Z
1 f (z1 ) 1 1 R (T, z2 )
Rg(T ) f = dz1 = f (z1 )  dz2  dz1
2πi z1 − g(T ) 2πi 2πi z1 − g(z2 )
∂F1 ∂F1 ∂F2
 
Z Z
1 1 f (z1 )
=  dz1  R (T, z2 )dz2 .
2πi 2πi z1 − g(z2 )
∂F2 ∂F1

Since g(z2 ) ∈ int F1 for z2 ∈ ∂F2 by construction; therefore, the Cauchy Integral
Formula yields the equality
Z
1 f (z1 )
f (g(z2 )) = dz1 .
2πi z1 − g(z2 )
∂F1

Consequently,
Z
1 ◦
Rg(T ) f = f (g(z2 ))R (T, z2 )dz2 = RT g(f ).
2πi
∂F2

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 σ of Sp(T ) is a clopen or isolated part or rarely
an exclave of Sp(T ), if σ and its complement σ 0 := Sp(T ) \ σ are closed.
8.2.10. Let σ be a clopen part of Sp(T ) and let κσ be some function that
equals 1 in an open neighborhood of σ and 0 in an open neighborhood of σ 0 . Further,
assign I
1 κσ (z)
Pσ := κσ (T ) := dz.
2πi z−T
Then Pσ is a projection in X and the (closed) subspace Xσ := im Pσ is invariant
under T .
C Since κσ2 = κσ , it follows from 8.2.3 that κσ (T )2 = κσ (T ). Furthermore,
T = RT IC , where IC : z 7→ z. Hence T Pσ = Pσ T (because I C κ σ = κ σ I C ).
Consequently, in virtue of 2.2.9 (4), Xσ is invariant under T . B
8.2.11. Definition. The projection Pσ is the Riesz projection or the Riesz
idempotent corresponding to σ.
Operators in Banach Spaces 137

8.2.12. Spectral Decomposition Theorem. Let σ 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 = Xσ ⊕ Xσ0 which reduces T to matrix form
 
Tσ 0
T ∼ ,
0 Tσ0
with the part Tσ of T in Xσ and the part Tσ0 of T in Xσ0 satisfying
Sp(Tσ ) = σ, Sp(Tσ0 ) = σ 0 .
C Since κ σ + κ σ0 = κSp(T ) = 1C , in view of 8.2.3 and 8.2.10 it suffices to es-
tablish the claim about the spectrum of Tσ .
From 8.2.5 and 8.2.3 obtain
σ ∪ 0 = κσ IC (Sp(T )) = Sp(κσ IC (T )) = Sp(RT (κσ IC ))
= Sp(RT κσ ◦ RT IC ) = Sp(Pσ T ).
Moreover, in matrix form
 
Tσ 0
Pσ T ∼ .
0 0
Let λ be a nonzero complex number. Then
 
λ − Tσ 0
λ − Pσ T ∼ ;
0 λ
i.e., the operator λ − Pσ T is not invertible if and only if the same is true of the
operator λ − Tσ . Thus,
Sp(Tσ ) \ 0 ⊂ Sp(Pσ T ) \ 0 = (σ ∪ 0) \ 0 ⊂ σ.
Suppose that 0 ∈ Sp(Tσ ) and 0 ∈/ σ. Choose disjoint open sets Dσ and Dσ0 so
that σ ⊂ Dσ , 0 ∈
/ Dσ and σ 0 ⊂ Dσ0 , and put
1
z ∈ Dσ ⇒ h(z) := ;
z
z ∈ Dσ 0 ⇒ h(z) := 0.
By 8.2.3, h(T )T = T h(T ) = Pσ . Moreover, since hκ σ = κ σ h, the decomposition
X = Xσ ⊕Xσ0 reduces h(T ) and h(T )σ Tσ = Tσ h(T )σ = 1 for the part h(T )σ of h(T )
in Xσ . So, Tσ is invertible; i.e., 0 ∈
/ Sp(Tσ ). We thus arrive at a contradiction which
implies that 0 ∈ σ. In other words, Sp(Tσ ) ⊂ σ.
Observe now that res(T ) = res(Tσ ) ∩ res(Tσ0 ). Consequently, by the above
Sp(T ) = C \ res(T ) = C \ (res(Tσ ) ∩ res(Tσ0 ))
= (C \ res(Tσ )) ∪ (C \ res(Tσ0 )) = Sp(Tσ ) ∪ Sp(Tσ0 ) ⊂ σ ∪ σ 0 = Sp(T ).
Considering that σ ∩ σ 0 = ∅, complete the proof. B
138 Chapter 8

8.2.13. Riesz–Dunford Integral Decomposition Let σ be


a clopen part of Sp(T ) for an endomorphism T of a Banach space X. The direct
sum decomposition X = Xσ ⊕ Xσ0 reduces the representation RT of the algebra
H (Sp(T )) in X to the direct sum of the representations Rσ and Rσ0 . Moreover,
the following diagrams of representations commute:

κ σ- κ σ0-
H (Sp(T )) H (σ) H (Sp(T )) H (σ 0 )
@ @
RTσ RTσ0
@ @
Rσ @ Rσ0 @
@ @
R ?
@ R ?
@
B(Xσ ) B(Xσ0 )

Here κ σ (f ) := κσ f and κ σ0 (f ) := κσ0 f for f ∈ H(Sp(T )) are the representations


induced by restricting f onto σ and σ 0 . CB

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 (xn )n∈N of points of the unit ball BX , the se-
quence (Kxn )n∈N has a Cauchy subsequence. CB
8.3.2. Theorem. Let X and Y be Banach spaces over a basic field 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 T KS ∈ K (W, Z);


(3) IF ∈ K (F) := K (F, F ).
Operators in Banach Spaces 139

C That K (X, Y ) is a subspace of B(X, Y ) follows from 8.3.1. If Kn ∈


K (X, Y ) and Kn 0 = 0 O ; then, given ε > 0, for n sufficiently large observe that
kKx − Kn xk ≤ kK − Kn k kxk ≤ ε whenever x ∈ BX . Therefore, Kn (BX ) serves
as an ε-net (= Bε -net) for K(BX ). 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. B
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 (l2 ), and B(l2 ) exhaust the list of
closed bilateral ideals in the endomorphism algebra B(l2 ) of the Hilbert space l2 .
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 finite-
rank operator provided that T ∈ B(X, Y ) and im T is a finite-dimensional subspace
of Y . In notation: T ∈ F (X, Y ). A hasty term “finite-dimensional operator” would
abuse consistency since T as a subspace of X × Y is usually infinite-dimensional.
8.3.7. The linear span of the set of (bounded) rank-one operators comprises
all finite-rank operators:

T ∈ F (X, Y )
n
X
⇔ (∃ x01 , . . . , x0n 0
∈ X ) (∃ y1 , . . . , yn ∈ Y ) T = x0k ⊗ yk . /.
k=1

8.3.8. Definition. Let Q be a (nonempty) compact set in X. Given T ∈


B(X, Y ), put
kT kQ := sup kT (Q)k.
The collection of all seminorms B(X, Y ) of type k · kQ is the Arens multinorm
in B(X, Y ), denoted by κB(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 ε > 0 and every compact set Q in X there is a finite-rank
endomorphism T of X, a member of F (X) := F (X, X), such that kT x − xk ≤ ε
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 κB(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 κB(X,Y ) .
C 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 ε > 0, by hypothesis there is a member T0 of F (X) such that kT0 − IX kT (Q) =
kT0 T − T kQ ≤ ε. Undoubtedly, T0 T ∈ F (W, X).
(1) ⇒ (3): Let T ∈ B(X, Y ). If T = 0 then there is nothing to be proven.
Let T 6= 0, ε > 0 and Q be a nonempty compact set in X. By hypothesis there
is a member T0 of F (X) such that kT0 − IX kQ ≤ εkT k−1 . Then kT T0 − T kQ ≤
kT k kT0 − IX kQ ≤ ε. Furthermore, T T0 ∈ F (X, Y ). B
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. Enflo’s 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(l2 ) lacks the approxi-
mation property.
8.3.14. Davis–Figiel–Szankowski Couterexample. The spaces c0 and lp
with p 6= 2 have closed subspaces lacking the approximation property.

8.4. The Riesz–Schauder Theory


8.4.1. ε-Perpendicular Lemma. Let X0 be a closed subspace of a Banach
space X and X 6= X0 . Given an ε > 0, there is an ε-perpendicular to X0 in X; i.e.,
an element xε in X such that kxε k = 1 and d(xε , X0 ) := inf dk·k ({xε }×X0 ) ≥ 1−ε.
C Take 1 > ε and x ∈ X \ X0 . It is clear that d := d(x, X0 ) > 0. Find x0 in the
subspace X0 satisfying kx−x0 k ≤ d/(1−ε), which is possible because d/(1−ε) > d.
Operators in Banach Spaces 141

Put xε := (x − x0 )kx − x0 k−1 . Then kxε k = 1. Finally, for x0 ∈ X0 , observe that

x − x0


kx0 − xε k = x0 −

kx − x0 k
1 d(x, X0 )
= 0 k(kx − x0 kx0 + x0 ) − xk ≥ ≥ 1 − ε. .
kx − xk kx0 − xk

8.4.2. Riesz Criterion. Let X be a Banach space. The identity operator


in X is compact if and only if X is finite-dimensional.
C Only the implication ⇒ needs proving. If X fails to be finite-dimensional,
then select a sequence of finite-dimensional subspaces X1 ⊂ X2 ⊂ . . . in X such
that Xn+1 6= Xn for all n ∈ N. In virtue of 8.4.1 there is a sequence (xn ) satisfying
xn+1 ∈ Xn+1 , kxn+1 k = 1 and d(xn+1 , Xn ) ≥ 1 /2 , namely, some sequence of 1 /2 -
perpendiculars to Xn in Xn+1 . It is clear that d(xm , xk ) ≥ 1 /2 for m 6= k. In other
words, the sequence (xn ) lacks Cauchy subsequences. Consequently, by 8.3.1 the
operator IX is not compact. B
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 finite rank.
C Only the implication ⇒ needs examining.
Let Y0 := im T be a closed subspace in Y . By the Banach Homomorphism
Theorem, the image T (BX ) of the unit ball of X is a neighborhood of zero in Y0 .
Furthermore, in virtue of the compactness property of T , the set T (BX ) is relatively
compact in Y0 . It remains to apply 8.4.2 to Y0 . B
8.4.4. Let X be a Banach space and K ∈ K (X). Then the operator 1 − K is
normally solvable.
C Put T := 1 − K and X1 := ker T . Undoubtedly, X1 is finite-dimensional
by 8.4.2. In accordance with 7.4.11 (1) a finite-dimensional subspace is comple-
mented. Denote a topological complement of X1 to X by X2 . Considering that X2
is a Banach space and T (X) = T (X2 ), it suffices to verify that kT xk ≥ tkxk for
some t > 0 and all x ∈ X2 . In the opposite case, there is a sequence (xn ) such that
kxn k = 1, xn ∈ X2 and T xn 0 = 0 O . Using the compactness property of K, we may
assume that (Kxn ) converges. Put y := lim Kxn . Then the sequence (xn ) converges
to y, because y = lim(T xn + Kxn ) = lim xn . Moreover, T y = lim T xn = 0; i.e.,
y ∈ X1 . It is beyond a doubt that y ∈ X2 . Thus, y ∈ X1 ∩ X2 ; i.e., y = 0. We
arrive at a contradiction: kyk = lim kxn k = 1. B
8.4.5. For whatever strictly positive ε, there are only finitely many eigenvalues
of a compact operator beyond the disk centered at zero and having radius ε.
142 Chapter 8

C Suppose by way of contradiction that there is a sequence (λn )n∈N of pair-


wise distinct eigenvalues of a compact operator K, with |λn | ≥ ε for all n ∈ N and
ε > 0. Suppose further that xn satisfying 0 6= xn ∈ ker(λn − K) is an eigenvector
with eigenvalue λn . Establish first that the set {xn : n ∈ N} is linearly inde-
pendent.PTo this end, assume the set {x1 , . . . , xn } linearly
Pindependent. In case
n n
xn+1 = k=1 αk xk , we would have 0 = (λn+1 − K)xn+1 = k=1 αk (λn+1 − λk )xk .
Consequently, αk = 0 for k := 1, . . . , n. Whence the false equality xn+1 = 0 would
ensue.
Put Xn := lin({x1 , . . . , xn }). By definition X1 ⊂ X2 ⊂ . . . ; moreover, as was
proven, Xn+1 6= Xn for n ∈ N. By virtue of 8.4.1 there is a sequence (xn ) such that
xn+1 ∈ Xn+1 , kxn+1 k = 1 and d(xn+1 , Xn ) ≥ 1 /2 . For m > k, straightforward
calculation shows that z := (λm+1 − K)xm+1 ∈ Xm and z + Kxk ∈ Xm + Xk ⊂ Xm .
Consequently,

kKxm+1 − Kxk k = k − λm+1 xm+1 + Kxm+1 + λm+1 xm+1 − Kxk k


= kλm+1 xm+1 − (z + Kxk )k ≥ |λm+1 |d(xm+1 , Xm ) ≥ ε /2 .

In other words, the sequence (Kxn ) has no Cauchy subsequences. B


8.4.6. Schauder Theorem. Let X and Y be Banach spaces (over the same
ground field F). Then

K ∈ K (X, Y ) ⇔ K 0 ∈ K (Y 0 , X 0 ).

C ⇒: Observe first of all that the restriction mapping x0 7→ x0 |BX implements


an isometry of X 0 into l∞ (BX ). Therefore, to check that K 0 (BY 0 ) is relatively
compact we are to show the same for the set V := {K 0 y 0 |BX : y 0 ∈ BY 0 }. Since
K 0 y 0 |BX (x) = y 0 ◦ K|BX (x) = y 0 (Kx) for x ∈ BX and y 0 ∈ BY 0 , consider the

compact set Q := cl K(BX ) and the mapping K : C(Q, F)0 = 0 O ∞ (BX ) defined
◦ ◦
by the rule Kg : x 7→ g(Kx). Undoubtedly, the operator K is bounded and, hence,
continuous. Now put S := {y 0 |Q : y 0 ∈ BY 0 }. It is clear that S is simultaneously
an equicontinuous and bounded subset of C(Q, F ). Consequently, by the Ascoli–
Arzelà 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 0 |Q = K 0 y 0 |BX

for y 0 ∈ BY 0 ; i.e., K(S) = V .
⇐: If K 0 ∈ K (Y 0 , X 0 ) then, as is proven, K 00 ∈ K (X 00 , Y 00 ). By the Double
Prime Lemma, K 00 |X = K. Whence it follows that the operator K is compact. B
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

C 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 ε, there are only finitely many points of Sp(K) beyond the disk
centered at zero and having radius ε. B
8.4.8. Riesz–Schauder Theorem. The spectrum of a compact operator K
in an infinite-dimensional space contains zero. Each nonzero point of the spectrum
of K is isolated and presents an eigenvalue of K with the corresponding eigenspace
finite-dimensional.
C Considering K, a compact endomorphism of a Banach space X, we must
only demonstrate the implication
0 6= λ ∈ Sp(K) ⇒ ker(λ − K) 6= 0.
First, settle the case F := C. Note that {λ} is a clopen part of Sp(K). Putting
g(z) := 1 /z in some neighborhood about λ and g(z) := 0 for z in a suitable
neighborhood about {λ}0 , observe that κ{λ} = gI C . Thus, by 8.2.3 and 8.2.10,
P{λ} = g(K)K. By virtue of 8.3.2 (2), P{λ} ∈ K (X). From 8.4.3 it follows that
im P{λ} is a finite-dimensional space. It remains to invoke the Spectral Decompo-
sition Theorem.
In the case of the reals, F := R, implement the process of complexification.
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. Define 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 λ ∈ Sp(K). Consequently, λ is an eigenvalue of K
by what was proven. Whence it follows that λ is an eigenvalue of K. B
8.4.9. Theorem. Let X be a complex Banach space. Given T ∈ B(X),
assume further that f : C0 = 0 O C is a holomorphic function vanishing only at zero
and such that f (T ) ∈ K (X). Then every nonzero point λ of the spectrum of T is
isolated and the Riesz projection P{λ} is compact.
C Suppose the contrary; i.e., find a sequence (λn )n∈N of distinct points of Sp(T )
such that λn 0 = 0 O 6= 0 (in particular, X is infinite-dimensional). Then f (λn )0 =
0 O (λ) and f (λ) 6= 0 by hypothesis. By the Spectral Mapping Theorem, Sp(f (T )) =
f (Sp(T )). Thus, by 8.4.8, f (λn ) = f (λ) for all sufficiently large n. Whence it
follows that f (z) = f (λ) for all z ∈ C and so f (T ) = f (λ). By the Riesz Criterion
in this case X is finite-dimensional. We come to a contradiction meaning that λ
is an isolated point of Sp(T ). Letting g(z) := f (z)−1 in some neighborhood about
λ disjoint from zero, infer that gf = κ{λ} . Consequently, by the Gelfand–Dunford
Theorem, P{λ} = g(T )f (T ); i.e., in virtue of 8.3.2 (2) the Riesz projection P{λ} is
compact. B
144 Chapter 8

8.4.10. Remark. Theorem 8.4.9 is sometimes referred to as the Generalized


Riesz–Schauder Theorem.

8.5. Fredholm Operators


8.5.1. Definition. Let X and Y be Banach spaces (over the same ground
field F). An operator T , a member of B(X, Y ), is a Fredholm operator (in symbols,
T ∈ F r(X, Y )) if ker T := T −1 (0) and coker T := Y / im T are finite-dimensional;
i.e., if the following quantities, called the nullity and the deficiency of T , are finite:

α(T ) := nul T := dim ker T ; β(T ) := def T := dim coker T.

The integer ind T := α(T )−β(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.
C Immediate from the Kato Criterion. B
8.5.4. For T ∈ B(X, Y ), the equivalence holds:

T ∈ F r(X, Y ) ⇔ T 0 ∈ F r(Y 0 , X 0 ).

Moreover, ind T = − ind T 0 .


C 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 0 = 0 O −→ Y → coker T → 0;
T

T0
0 ← (ker T )0 ← X 0 ←− Y 0 ← (coker T )0 ← 0;
T0
0 → ker(T 0 ) → Y 0 −→ X 0 → coker(T 0 ) → 0;
T
0 ← (ker(T 0 ))0 ← Y ←− X ← (coker(T 0 ))0 ← 0.

Moreover, α(T ) = β(T 0 ) and β(T ) = α(T 0 ) (cf. 7.6.14). B


8.5.5. An operator T is an index-zero Fredholm operator if and only if so is
the dual of T .
C This is a particular case of 8.5.4. B
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 T x = 0 has a sole solution, zero. The
homogeneous conjugate equation T 0 y 0 = 0 has a sole solution, zero. The equation
T x = y is solvable and has a unique solution given an arbitrary right side. The
conjugate equation T 0 y 0 = x0 is solvable and has a unique solution given an arbitrary
right side.
(2) The homogeneous equation T x = 0 has a nonzero solution. The ho-
mogeneous conjugate equation T 0 y 0 = 0 has a nonzero solution. The homogeneous
equation T x = 0 has finitely many linearly independent solutions x1 , . . . , xn . The
homogeneous conjugate equation T 0 y 0 = 0 has finitely many linearly independent
solutions y 01 , . . . , y 0n .
The equation T x = y is solvable if and only if y 01 (y) = . . . = y 0n (y) = 0.
Moreover, the general solution x is the sum of a particular solution x0 and the
general solution of the homogeneous equation; i.e., it has the form
n
X
x = x0 + λk xk (λk ∈ F ).
k=1

The conjugate equation T 0 y 0 = x0 is solvable if and only if x0 (x1 ) = . . . =


x (xn ) = 0. Moreover, the general solution y 0 is the sum of a particular solution y 00
0

and the general solution of the homogeneous equation; i.e., it has the form
n
X
y 0 = y 00 + µk y 0k (µk ∈ F ).
k=1

C This is a reformulation of 8.5.5 with account taken of the Polar Lemma. B


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 im T be the rank of T .
Then α(T ) = n − rank T and β(T ) = m − rank T . Consequently, T ∈ F r(F n , F m )
and ind T = n − m.
(3) Let T ∈ B(X) and X = X1 ⊕ X2 . Assume that this direct sum
decomposition of X reduces T to matrix form
 
T1 0
T ∼ .
0 T2
Undoubtedly, T is a Fredholm operator if and only if its parts are Fredholm
operators. Moreover, α(T ) = α(T1 )+α(T2 ) and β(T ) = β(T1 )+β(T2 ); i.e., ind T =
ind T1 + ind T2 . CB
146 Chapter 8

8.5.8. Fredholm Theorem. Let K ∈ K (X). Then 1 − K is an index-zero


Fredholm operator.
C 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 Riesz–Schauder Theorem
and the Spectral Decomposition Theorem there is a decomposition X = X1 ⊕ X2
such that X1 is finite-dimensional and 1 ∈/ Sp(K2 ), with K2 the part of K in X2 .
Furthermore,  
1 − K1 0
1−K ∼ .
0 1 − K2
By 8.5.7 (2), ind (1 − K1 ) = 0 and, by 8.5.7 (3), ind (1 − K) = ind (1 − K1 ) + ind (1 −
K2 ) = 0.
In the case of the reals, F := R, proceed by way of complexification 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 difference between R
and C, observe that α(1 − K) = α(1 − K) and β(1 − K) = β(1 − K). Finally,
ind (1 − K) = 0. B
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 T R − 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 + KX (KX ∈ K (X)) ⇒ LT R = R + KX R;
T R = 1 + KY (KY ∈ K (Y )) ⇒ LT R = L + LKY .

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.
C 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 := L − R ∈
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 . B
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 ϕ(S)ϕ(T ) = ϕ(T )ϕ(S) = 1, where
ϕ : B(X)0 = 0 O (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.
C ⇒: Let T ∈ F r(X, Y ). Using the Complementation Principle, consider the
decompositions X = ker T ⊕ X1 and Y = im T ⊕ Y1 and the respective finite-rank
projections P which carries X onto ker T along X1 and Q which carries Y onto
Y1 along im T . It is clear that the restriction T1 := T |X1 is an invertible operator
T1 : X1 0 = 0 O T . Put S := T1−1 (1 − Q). The operator S may be viewed as a member
of B(Y, X). Moreover, it is beyond a doubt that ST + P = 1 and T S + Q = 1.
⇐: Let S be an approximate inverse of T ; i.e., ST = 1 + KX and T S = 1 + KY
for appropriate compact operators KX and KY . Consequently, ker T ⊂ ker(1+KX );
i.e., ker T is finite-dimensional since so is ker(1+KX ) in virtue of 8.5.8. Furthermore,
im T ⊃ im(1 + KY ); and so the range of T is of finite codimension because 1 + KY
is an index-zero Fredholm operator. B
8.5.15. Corollary. If T ∈ F r(X, Y ) and S is an approximate inverse of T
then S ∈ F r(Y, X). CB
8.5.16. Corollary. The product of Fredholm operators is itself a Fredholm
operator.
C The composition of approximate inverses (taken in due succession) is an ap-
proximate inverse to the composition of the originals. B
8.5.17. Consider an exact sequence

00 = 0 O 1 0 = 0 O 2 0 = 0 O 0 = 0 O n−1 0 = 0 O n0 = 0 O

of finite-dimensional vector spaces. Then the Euler identity holds:


n
X
(−1)k dim Xk = 0.
k=1

C For n = 1 the exactness of the sequence 00 = 0 O 1 0 = 0 O means that X1 = 0,


and for n = 2 the exactness of 00 = 0 O 1 0 = 0 O 2 0 = 0 O amounts to isomorphy
between X1 and X2 (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 → X1 0 = 0 O 2 0 = 0 O → Xn−2 −−−−→ Xn−1 −−−−→ Xn → 0


Tn−2 Tn−1

reduces to the exact sequence


Tn−2
0 → X1 → X2 → . . . → Xn−2 −−−→ ker Tn−1 → 0.

By hypothesis,
n−2
X
(−1)k dim Xk + (−1)n−1 dim ker Tn−1 = 0.
k=1

Furthermore, since Tn−1 is an epimorphism,

dim Xn−1 = dim ker Tn−1 + dim Xn .

Finally,

n−2
X
0= (−1)k dim Xk + (−1)n−1 (dim Xn−1 − dim Xn )
k=1
n
X
= (−1)k dim Xk . .
k=1

8.5.18. Atkinson Theorem. The index of the product of Fredholm operators


equals the sum of the indices of the factors.
C Let T ∈ F r(X, Y ) and S ∈ F r(Y, Z). By virtue of 8.5.16, ST ∈ F r(X, Z).
Using the Snowflake Lemma, obtain the exact sequence of finite-dimensional spaces

00 = 0 O T 0 = 0 O ST 0 = 0 O S0 = 0 O T 0 = 0 O ST 0 = 0 O S0 = 0 O .

Applying 8.5.17, infer that

α(T ) − α(ST ) + α(S) − β(T ) + β(ST ) − β(S) = 0;

whence ind (ST ) = ind S + ind T . B


8.5.19. Corollary. Let T be a Fredholm operator and let S be an approximate
inverse of T . Then ind T = − ind S.
C ind (ST ) = ind (1 + K) for some compact operator K. By Theorem 8.5.8,
1 + K is an index-zero Fredholm operator. B
Operators in Banach Spaces 149

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 ∈ F r(X, Y ) and K ∈ K (X, Y ) then T + K ∈ F r(X, Y )
and ind (T + K) = ind T .
C Let S be an approximate inverse to T ; i.e.,

ST = 1 + KX ; T S = 1 + KY

with some KX ∈ K (X) and KY ∈ K (Y ) (that S exists is ensured by 8.5.14). It is


clear that

S(T + K) = ST + SK = 1 + KX + SK ∈ 1 + K (X);
(T + K)S = T S + KS = 1 + KY + KS ∈ 1 + K (Y );

i.e., S is an approximate inverse of T + K. By virtue of 8.5.14, T + K ∈ F r(X, Y ).


Finally, from 8.5.19 infer the equalities ind (T +K) = − ind S and ind T = − ind S. B
8.5.21. Bounded Index Stability Theorem. The property of being
a Fredholm operator and the index of a Fredholm operator are preserved under
sufficiently small bounded perturbations: the set F r(X, Y ) is open in the space of
bounded operators, and the index of a Fredholm operator ind : F r(X, Y )0 = 0 O Z
is a continuous function.
C Let T ∈ F r(X, Y ). By 8.5.14 there are operators S ∈ B(Y, X), KX ∈
K (X) and KY ∈ K (Y ) such that

ST = 1 + KX ; T S = 1 + KY .

If S = 0 then the spaces X and Y are finite-dimensional by the Riesz Criterion,


i.e., nothing is left to proof: it suffices to refer to 8.5.7 (2). If S 6= 0 then for all
V ∈ B(X, Y ) with kV k < 1/kSk, from the inequality of 5.6.1 it follows: kSV k < 1
and kV Sk < 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 + KX + SV )


= 1 + (1 + SV )−1 KX ∈ 1 + K (X);

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 + KY + V S)(1 + V S)−1


= 1 + KY (1 + V S)−1 ∈ 1 + K (Y ).
150 Chapter 8

By 8.5.12, T +V is approximately invertible. In virtue of 8.5.14, T +V ∈ F r(X, Y ).


This proves the openness property of F r(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. B
8.5.22. Nikol 0skiı̆ 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.
C ⇒: Let T ∈ F r(X, Y ) and ind T = 0. Consider the direct sum decom-
positions X = X1 ⊕ ker T and Y = im T ⊕ Y1 . It is beyond a doubt that the
operator T1 , the restriction of T to X1 , implements an isomorphism between X1
and im T . Furthermore, in virtue of 8.5.5, dim Y1 = β(T ) = α(T ); i.e., there is
a natural isomorphism Id : ker T 0 = 0 O 1 . Therefore, T admits the matrix presen-
tation      
T1 0 T1 0 0 0
T ∼ = + .
0 0 0 Id 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. B
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
Nikol0 skiı̆ 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 finite-rank
operators. CB

Exercises
8.1. Study the Riesz–Dunford integral in finite dimensions.
8.2. Describe the kernel of the Riesz–Dunford integral.
8.3. Given n ∈ N, let fn be a function holomorphic in a neighborhood U about the spectrum
of an operator T . Prove that the uniform convergence of (fn ) to zero on U follows from the
convergence of (fn (T )) to zero in operator norm.
Operators in Banach Spaces 151

8.4. Let σ be an isolated part of the spectrum of an operator T . Assume that the part
σ 0 := Sp(T ) \ σ is separated from σ by some open disk with center a and radius r so that
σ ⊂ {z ∈ C : |z − a| < r}. Considering the Riesz projection Pσ , prove that

Pσ = lim (1 − z −n (T − a)n )−1 ;


n
1
x ∈ im(Pσ ) ⇔ lim sup k(a − T )n xk /n
< r.
n

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 finite-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 finite-dimensional or finite-codimensional
(the latter means that the kernel has a finite-dimensional algebraic complement).
8.8. Let 1 ≤ p < r < +∞. Prove that every bounded operator from lr to lp or from c0 to lp
is compact.
8.9. Let H be a separable Hilbert space. Given an operator T in B(H) and a Hilbert basis
(en ) for H, define the Hilbert–Schmidt norm as


!1 /2
X
kT k2 := kT en k2 .
n=1

(Examine soundness!) An operator with finite Hilbert–Schmidt norm is a Hilbert–Schmidt oper-


ator. Demonstrate that an operator T is a Hilbert–Schmidt operator if and only if T is compact
P∞ 2 1
and n=1 n
λ < +∞, where (λn ) ranges over the eigenvalues of some operator (T ∗ T ) /2 (define
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 finite 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 finite descent and finite ascent, the two quantities, a(T ) and d(T ), coincide.
8.11. An operator T is a Riesz–Schauder operator, if T is a Fredholm operator and has
finite descent and finite ascent. Prove that an operator T is a Riesz–Schauder operator if and only
if T is of the form T = U + V , where U is invertible and V is of finite rank (or compact) and
commutes with U .
8.12. Let T be a bounded endomorphism of a Banach space X which has finite descent and
finite 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.
152 Chapter 8

8.13. Let T be a normally solvable operator. If either of the next quantities is finite

α(T ) := dim ker T, β(T ) := dim coker T,

then T is called semi-Fredholm. Put

Ф+ (X) := {T ∈ B(X) : im T ∈ Cl(X), α(T ) < +∞};


Ф− (X) := {T ∈ B(X) : im T ∈ Cl(X), β(T ) < +∞}.

Prove that

T ∈ Ф+ (X) ⇔ T 0 ∈ Ф− (X 0 );
T ∈ Ф− (X) ⇔ T 0 ∈ Ф+ (X 0 ).

8.14. Let T be a bounded endomorphism. Prove that T belongs to Ф+ (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 λ the operator (λ − T ) is a Fredholm operator. Prove that T is a Riesz operator if and
only if for all λ ∈ C, λ 6= 0, the following conditions are fulfilled:
(1) the operator (λ − T ) has finite descent and finite ascent;
(2) the kernel of (λ − T )k is finite-dimensional for every k ∈ N;
(3) the range of (λ − T )k has finite deficiency 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 ε, there are only finitely many
points of Sp(T ) beyond the disk centered at zero with radius ε).
8.16. Establish the isometric isomorphisms: (X/Y )0 ' Y ⊥ and X 0 /Y ⊥ ' Y 0 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 τ : X0 = 0 O P (P(X)) is
a pretopology on X if
(1) x ∈ X ⇒ τ (x) is a filter on X;
(2) x ∈ X ⇒ τ (x) ⊂ fil{x}.
A member of τ (x) is a (pre)neighborhood about x or of x. The pair (X, τ ), 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 τ1 , τ2 ∈ T (X) then τ1 is said to be stronger than τ2 or finer than τ2 (in symbols,
τ1 ≥ τ2 ) provided that x ∈ X ⇒ τ1 (x) ⊃ τ2 (x). Of course, τ2 is weaker or coarser
than τ1 .
9.1.3. The set T (X) with the relation “to be stronger” presents a complete
lattice.
C If X = ∅ then T (X) = {∅} and there is nothing to be proven. If X 6= ∅
then refer to 1.3.13. B
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(τ ) ⇔ (∀ x ∈ G) (G ∈ τ (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(τ ) ⇔ X \ F ∈ Op(τ ).
9.1.5. The union of a family of open sets and the intersection of a finite family
of open sets are open. The intersection of a family of closed sets and the union
of a finite family of closed sets are closed. CB
9.1.6. Let (X, τ ) be a pretopological space. Given x ∈ X, put
U ∈ t(τ )(x) ⇔ (∃ V ∈ Op(τ )) x ∈ V & U ⊃ V.
The mapping t(τ ) : x 7→ t(τ )(x) is a pretopology on X. CB
154 Chapter 9

9.1.7. Definition. A pretopology τ on X is a topology if τ = t(τ ). The pair


(X, τ ), 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 τ◦ := inf T (X). It is clear that τ◦ (x) = {X} for x ∈ X.
Consequently, Op(τ◦ ) = {∅, X} and so τ◦ = t(τ◦ ); i.e., τ◦ is a topology. This
topology is called trivial, or antidiscrete, or even indiscrete.
(4) Let τ ◦ := sup T (X). It is clear that τ ◦ (x) = fil {x} for x ∈ X.
Consequently, Op(τ ◦ ) = 2X and so τ ◦ = t(τ ◦ ); i.e., τ ◦ 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 finite
subfamilies. Then there is a unique topology τ on X such that Op(τ ) = Op.
C Put τ (x) := fil {V ∈ Op : x ∈ V } for x ∈ X (in case X = ∅ there is nothing
to prove). Observe that τ (x) 6= ∅ since the intersection of the empty family equals
X (cf. inf ∅ = +∞). From the construction derive that t(τ ) = τ and Op ⊂ Op(τ ).
If G ∈ Op(τ ) then G = ∪{V : V ∈ Op, V ⊂ G} and so G ∈ Op by hypothesis.
The claim of uniqueness raises no doubts. B
9.1.9. Let the mapping t : T (X)0 = 0 O T (X) act by the rule t : τ 7→ t(τ ).
Then
(1) im t = T(X); i.e., τ ∈ T (X) ⇒ t(τ ) ∈ T(X);
(2) τ1 ≤ τ2 ⇒ t(τ1 ) ≤ t(τ2 ) (τ1 , τ2 ∈ T (X));
(3) t ◦ t = t;
(4) τ ∈ T (X) ⇒ t(τ ) ≤ τ ;
(5) Op(τ ) = Op(t(τ )) (τ ∈ T (X)).
C The inclusion Op(τ ) ⊃ Op(t(τ )) holds because it is easier to be open in τ .
The reverse inclusion Op(τ ) ⊂ Op(t(τ )) follows from the definition of t(τ ). The
equality Op(τ ) = Op(t(τ )) makes everything evident. B
9.1.10. A pretopology τ on X is a topology if and only if

(∀ U ∈ τ (x))(∃ V ∈ τ (x) & V ⊂ U )(∀ y) (y ∈ V ⇒ V ∈ τ (y))

for x ∈ X.
C Straightforward from 9.1.9 (5). B
An Excursus into General Topology 155

9.1.11. Let τ1 , τ2 ∈ T(X). The following statements are equivalent:


(1) τ1 ≥ τ2 ;
(2) Op(τ1 ) ⊃ Op(τ2 );
(3) Cl(τ1 ) ⊃ Cl(τ2 ). CB
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, τ ) makes X into the topological space
(X, t(τ )) with the same open sets in stock. Therefore, given a pretopology τ , the
topology t(τ ) is usually called the topology associated with τ .
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:
supT(X) E = supT (X) E .
C Evidently, t(supT (X) E ) ≥ supT (X) t(E ) ≥ supT (X) E ≥ t(supT (X) E ).
Thus, τ := supT (X) E belongs to T(X). It is clear that τ ≥ E . Furthermore,
if τ0 ≥ E and τ0 ∈ T(X) then τ0 ≥ τ and so τ = supT(X) E . It remains to refer
to 1.2.14. B
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 ) ⇔ (∀ τ ∈ E ) U ∈ Op(τ ).


In other words, \
Op(inf T(X) E ) = Op(τ ).
τ ∈E

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 ). CB

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 filters 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. Birkhoff 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 filter containing U and converging to x;
(3) there is a net of elements of U which converges to x. CB
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 filter of an arbitrary point
x is coarser than the neighborhood filter of f (x);
(4) for all x, the mapping f transforms each filter convergent to x into
a filter 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). CB
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, τX ) → (Y, τY ) 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
difference is immaterial between the definitions of the continuity property at a point
of X and the general continuity property (on X). Indeed, if we let τx (x) := τX (x)
and τx (x) := fil {x} for x ∈ X, x 6= x, then the continuity property of f at x (with
respect to the topology τX in X) amounts to that of f : (X, τx ) → (Y, τY ) (at every
point of the space X with topology τx ).
9.2.6. Let τ1 , τ2 ∈ T(X). Then τ1 ≥ τ2 if and only if IX : (X, τ1 ) → (X, τ2 )
is continuous. CB
9.2.7. Let f : (X, τ ) → (Y, ω) be a continuous mapping and let τ1 ∈ T(X)
and ω1 ∈ T(Y ) be such that τ1 ≥ τ and ω ≥ ω1 . Then f : (X, τ1 ) → (Y, ω1 ) is
continuous.
C By hypothesis the following diagram commutes:
f
(X, τ ) −→ (Y, ω)
IX ↑ ↓ IY
f
(X, τ1 ) −→ (Y, ω1 )
An Excursus into General Topology 157

It suffices to observe that every composition of continuous mappings is continu-


ous. B
9.2.8. Inverse Image Topology Theorem. Let f : X → (Y, ω). Put

T0 := {τ ∈ T(X) : f : (X, τ ) → (Y, ω) is continuous}.

Then the topology f −1 (ω) := inf T0 belongs to T0 .


C From 9.2.3 (1) it follows that

τ ∈ T0 ⇔ (x ∈ X ⇒ f −1 (ω(f (x))) ⊂ τ (x)).

Let τ (x) := f −1 (ω(f (x))). Undoubtedly, t(τ ) = τ . Furthermore, f (τ (x)) =


f (f −1 (ω(f (x)))) ⊃ ω(f (x)); i.e., τ ∈ T0 by 9.2.3 (3). Thus, f −1 (ω) = τ . B
9.2.9. Definition. The topology f −1 (ω) is the inverse image of ω 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, (xξ → x in f −1 (ω)) ⇔ (f (xξ ) → f (x) in ω); likewise,
(F → x in f −1 (ω)) ⇔ (f (F ) → f (x) in ω) for a filter F . CB
9.2.11. Image Topology Let f : (X, τ ) → Y . Put

Š0 := {ω ∈ T(Y ) : f : (X, τ ) → (Y, ω) is continuous}.

Then the topology f (τ ) := sup Š0 belongs to Š0 .


C Appealing to 9.1.13, observe that

f (τ )(y) = (supT(Y ) Š0 )(y) = (supT (Y ) Š0 )(y) = sup{ω(y) : ω ∈ Š0 }

for y ∈ Y . By virtue of 9.2.3 (3),

ω ∈ Š0 ⇔ (x ∈ X ⇒ f (τ (x)) ⊃ ω(f (x))).

Comparing the formulas, infer that f (τ ) ∈ Š0 . B


9.2.12. Definition. The topology f (τ ) is the image of τ 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 (fξ : X → (Yξ , ωξ ))ξ∈„ be a family of mappings.
Further, put τ := supξ∈„ fξ−1 (ωξ ). Then τ is the weakest (= least) topology on X
making all the mappings fξ (ξ ∈ „) continuous.
C Using 9.2.8, note that

(fξ : (X, τ ) → (Yξ , ωξ ) is continuous) ⇔ τ ≥ fξ−1 (ωξ ). .

9.2.15. Theorem. Let (fξ : (Xξ , τξ ) → Y )ξ∈„ be a family of mappings.


Further, assign ω := inf ξ∈„ fξ (τξ ). Then ω is the strongest (= greatest) topology
on Y making all the mappings fξ (ξ ∈ „) continuous.
C Appealing to 9.2.11, conclude that

(fξ : (Xξ , τξ ) → (Y, ω) is continuous) ⇔ ω ≤ fξ (τξ ). .

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, τ ) be a topological space and let X0 be a subset of X.
Denote the identical embedding of X0 into X by ι : X0 → X. Put τ0 := ι−1 (τ ).
The topology τ0 is the induced topology (by τ in X0 ), or the relative or subspace
topology; and the space (X0 , τ0 ) is a subspace of (X, τ ).
(2) Let (Xξ , τξ )ξ∈„ be a family of topological spaces and let X :=
the product of (Xξ )ξ∈„ . Put τ := supξ∈„ Pr−1
Q
X
ξ∈„ ξ be ξ (τξ ), where Prξ : X → Xξ
is the coordinate projection (onto Xξ ); i.e., Prξ x = xξ (ξ ∈ „). The topology τ
is the product topology or the product of the topologies (τξ )ξ∈„ , or the Tychonoff
topology of X. The space (X, τ ) is the Tychonoff product of the topological spaces
under study. In particular, if Xξ := [0, 1] for all ξ ∈ „ then X := [0, 1]„ (with the
Tychonoff topology) is a Tychonoff 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.
C To prove, it suffices to observe that

y ∈ cl{x} ⇔ (∀ V ∈ τ (y)) x ∈ V ⇔ x ∈ ∩{V : V ∈ τ (y)},

where x and y are points of a space with topology τ . B


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 T1 -space.
The topology of a T1 -space is called a separated topology or (rarely) a T1 -topology.
9.3.3. Remark. By way of expressiveness, one often says: “A T1 -space is
a space with closed points.”
9.3.4. For a topological space the following conditions are equivalent:
(1) each filter 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.
C (1) ⇒ (2): If y ∈ ∩U ∈τ (x) cl U then U ∩ V 6= ∅ for all V ∈ τ (y), provided
that U ∈ τ (x). Therefore, the join F := τ (x) ∨ τ (y) is available. Clearly, F → x
and F → y. By hypothesis, x = y.
(2) ⇒ (3): Let x, y ∈ X, x 6= 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 τ (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 filter on X. If F → x and F → y then F ⊃ τ (x)
and F ⊃ τ (y). Thus, U ∩ V 6= ∅ for U ∈ τ (x) and V ∈ τ (y), which means that
x = y. B
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 Hausdorff space or a T2 -space. A nat-
ural meaning is ascribed to the term “Hausdorff topology.”
9.3.6. Remark. By way of expressiveness, one often says: “A T2 -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 filter that is the neighborhood filter
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 filter 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.
C (1) ⇒ (2): If x ∈ X and U ∈ τ (x) then V := X \ int U is closed and x ∈ / V.
By hypothesis there is a set F in Cl(τ ) such that x ∈ / F and int F ⊃ V . Put
G := X \ F . Clearly, G ∈ τ (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(τ ) with x ∈ / F then X \ F ∈ τ (x). Thus, there
is a closed neighborhood U in τ (x) lying in X \ F . Thus, X \ U is a neighborhood
of F disjoint from U .
(3) ⇒ (1): If F ∈ Cl(τ ) and int G ⊃ F ⇒ y ∈ cl G, then U ∩ G 6= ∅ for every
U in τ (y) and every neighborhood G of F . This means that y ∈ F . B
9.3.9. Definition. A T3 -space is a topological space satisfying one (and hence
all) of the equivalent conditions 9.3.8 (1)–9.3.8 (3). A separated T3 -space is called
regular.
9.3.10. Urysohn Little Lemma. For a topological space the following con-
ditions are equivalent:
(1) the neighborhood filter of each nonempty closed set has a base of
closed sets;
(2) if two closed sets are disjoint then they have disjoint neighborhoods.
C (1) ⇒ (2): Let F1 and F2 be closed sets in some space X with F1 ∩ F2 = ∅.
Put G := X \ F1 . Obviously, G is open and G ⊃ F2 . If F2 = ∅, then there is
nothing to be proven. It may be assumed consequently that F2 6= ∅. Then there is
a closed set V2 such that G ⊃ V2 ⊃ int V2 ⊃ F2 . Put V1 := X \ V2 . It is clear that
V1 is open, and V1 ∩ V2 = ∅. Moreover, V1 ⊃ X \ G = X \ (X \ F1 ) = F1 .
(2) ⇒ (1): Let F = cl F , G = int G and G ⊃ F . Put F1 := X \ G. Then
F1 = cl F1 and so there are open sets U and U1 satisfying U ∩ U1 = ∅, with F ⊂ U
and F1 ⊂ U1 . Finally, cl U ⊂ X \ U1 ⊂ X \ F1 = G. B
9.3.11. Definition. A T4 -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 T4 -space
is called normal.
9.3.12. Continuous Function Recovery Lemma. Let a subset T be dense
in R and let t 7→ Ut (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} ⊂ Ut ⊂ {f ≤ t} (t ∈ T )
An Excursus into General Topology 161

if and only if
(t, s ∈ T & t < s) ⇒ cl Ut ⊂ int Us .
C ⇒: Take t < s. Since {f ≤ t} is closed and {f < s} is open, the inclusions
hold:
cl Ut ⊂ {f ≤ t} ⊂ {f < s} ⊂ int Us .
⇐: Since Ut ⊂ cl Ut ⊂ int Us ⊂ Us for t < s, the family t 7→ Ut (t ∈ T ) increases
by inclusion. Therefore, f exists by 3.8.2 and is unique by 3.8.4. Consider the
families t 7→ Vt := cl Ut and t 7→ Wt := int Ut . These families increase by inclusion.
Consequently, on applying 3.8.2 once again, find functions g, h : X → R satisfying

{g < t} ⊂ Vt ⊂ {g ≤ t}, {h < t} ⊂ Wt ⊂ {h ≤ t}

for all t ∈ T . If t, s ∈ T and t < s, then in view of 3.8.3

Wt = int Ut ⊂ Ut ⊂ Us ⇒ f ≤ h;
Vt = cl Ut ⊂ int Us = Ws ⇒ h ≤ g;
Ut ⊂ Us ⊂ cl Us = Vs ⇒ g ≤ f.

So, f = g = h. Taking account of 3.8.4 and 9.1.5 and given t ∈ R, find

{f < t} = {h < t} = ∪{Ws : s < t, s ∈ T } ∈ Op(τX );


{f ≤ t} = {g ≤ t} = ∩{Vs : t < s, s ∈ T } ∈ Cl(τX ).

These inclusions readily provide continuity for f . B


9.3.13. Urysohn Great Lemma. Let X be a T4 -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.
C Put Ut := ∅ for t < 0 and Ut := X for t > 1. Consider the set T of the
dyadic-rational points of the interval [0, 1]; i.e., T := ∪n∈N Tn with Tn := {k2−n+1 :
k := 0, 1, . . . , 2n−1 }. It suffices to define Ut for all t in T so that the family t 7→ Ut
(t ∈ T := T ∪ (R \ [0, 1])) satisfy the criterion of 9.3.12. This is done by way
of induction.
If t ∈ T1 , i.e., t ∈ {0, 1}; then put U0 := F and U1 := G. Assume now
that, for t ∈ Tn and n ≥ 1, some set Ut has already been constructed, satisfying
cl Ut ⊂ int Us whenever t, s ∈ Tn and t < s. Take t ∈ Tn+1 and find the two points
tl and tr in Tn nearest to t:

tl := sup{s ∈ Tn : s ≤ t};
tr := inf{s ∈ Tn : t ≤ s}.
162 Chapter 9

If t = tl or t = tr , then Ut exists by the induction hypothesis. If t 6= tl and


t 6= tr , then tl < t < tr and again by the induction hypothesis cl Utl ⊂ int Utr .
By virtue of 9.3.11 there is a closed set Ut such that

cl Utl ⊂ int Ut ⊂ Ut = cl Ut ⊂ int Utr .

It remains to show that the resulting family satisfies the criterion of 9.3.12.
To this end, take t, s ∈ Tn+1 with t < s. If tr = sl , then for s > sl by con-
struction
cl Ut ⊂ cl Utr = cl Usl ⊂ int Us .
For t < tr = sl similarly deduce the following:

cl Ut ⊂ int Utr = inf Usl ⊂ int Us .

If tr < sl then, on using the induction hypothesis, infer that

cl Ut ⊂ cl Utr ⊂ int Usl ⊂ int Us ,

what was required. B


9.3.14. Urysohn Theorem. A topological space X is a T4 -space if and only
if to every pair of disjoint closed sets F1 and F2 in X there corresponds a continuous
function f : X → [0, 1] such that f (x) = 0 for x ∈ F1 and f (x) = 1 for x ∈ F2 .
C ⇒: It suffices to apply 9.3.13 with F := F1 and G := X \ F2 .
⇐: If F1 ∩ F2 = ∅ and F1 and F2 are closed sets, then for a corresponding
function f the sets G1 := {f < 1 /2 } and G2 := {f > 1 /2 } are open and disjoint.
Moreover, G1 ⊃ F1 and G2 ⊃ F2 . B
9.3.15. Definition. A topological space X is a T31 /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 T31 /2 -space is a Tychonoff
space or a completely regular space.
9.3.16. Every normal space is a Tychonoff space.
C Straightforward from 9.3.1 and 9.3.14. B

9.4. Compactness
9.4.1. Let B be a filterbase 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 fil B;
(2) B → x ⇒ x ∈ cl B;
(3) (B is an ultrafilter and x ∈ cl B) ⇒ B → x.
C Only (3) needs demonstrating, since (1) and (2) are evident. Given U ∈ τ (x)
and B ∈ B, observe that U ∩ B 6= ∅. In other words, the join F := τ (x) ∨
B is available. It is clear that F → x. Furthermore, F = B, because B is
an ultrafilter. B
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 finite 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 filterbase B lacks adherent points in C then there is a member
B of B such that B ∩ C = ∅;
(3) each filterbase containing C has an adherent point in C;
(4) each ultrafilter containing C has a limit in C.
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 finite
subset B0 of B such that C ⊂ ∪{X \ cl B0 : B0 ∈ B0 } = X \ ∩{cl B0 : B0 ∈ B0 }.
Let B ∈ B satisfy B ⊂ ∩{B0 : B0 ∈ B0 } ⊂ ∩{cl B0 : B0 ∈ B0 }. Then
C ∩ B ⊂ C ∩ (∩{cl B0 : B0 ∈ B0 }) = ∅.
(2) ⇒ (3): In case C = ∅, there is nothing to be proven. If C 6= ∅, then for
B ∈ B by hypothesis B ∩ C 6= ∅ because C ∈ B. Thus, cl B ∩ C 6= ∅.
(3) ⇒ (4): It suffices to appeal to 9.4.1.
(4) ⇒ (1): Assume that C 6= ∅ (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 finite subset E0 of E , the
inclusion C ⊂ ∪{G : G ∈ E0 } fails. Put
( )
\
B := X \ G : E0 is a finite subset of E .
G∈E0
It is clear that B is a filterbase. Furthermore,
cl B = ∩{cl B : B ∈ B} = ∩{X \ G : G ∈ E }
= X \ ∪{G : G ∈ E } ⊂ X \ C.
Now choose an ultrafilter 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 6= ∅. At the same time cl F ⊂ cl B. We arrive at a contradiction. B
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 ultrafilter on it converges” (cf. 4.4.7). An ultranet is a net whose
tail filter is an ultrafilter. 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). CB
9.4.6. Let X0 be a subspace of a topological space X and let C be a subset
of X0 . Then C is compact in X0 if and only if C is compact in X.
C ⇒: Immediate from 9.4.5 and 9.2.17 (1).
⇐: Let B be a filterbase on X0 . Further, let V := clX0 B stand for the
adherence of B relative to X0 . Suppose that V ∩C = ∅. Since B is also a filterbase
on X, it makes sense to speak of the adherence W := clX B of B relative to X.
It is clear that V = W ∩ X0 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 X0 . B
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 confine study to compact spaces,
i.e. to sets “compact in themselves.” A topology τ on a set C, making C into
a compact space, is usually called a compact topology on C. Also, such C is referred
to as “compact with respect to τ .”
9.4.8. Tychonoff Theorem. The Tychonoff product of compact spaces is
compact.
Q
C Let X := ξ∈„ Xξ be the product of such spaces. If at least one of the
spaces Xξ is nonempty then X = ∅ and nothing is left to proof. Let X 6= ∅ and
let F be an ultrafilter on X. By 1.3.12, given ξ ∈ „ and considering the coordinate
projection Prξ : X → Xξ , observe that Prξ (F ) is an ultrafilter on Xξ . Consequently,
in virtue of 9.4.3 there is some xξ in Xξ such that Prξ (F ) → xξ . Let x : ξ 7→ xξ .
It is clear that F → x (cf. 9.2.10). Appealing to 9.4.3 once more, infer that X is
compact. B
9.4.9. Every closed subset of a compact space is compact.
C Let X be compact and C ∈ Cl(X). Assume further that F is an ultrafilter
on X and C ∈ F . By Theorem 9.4.3, F has a limit x in X: that is, F → x. By the
Birkhoff Theorem, x ∈ cl C = C. Using 9.4.3 again, conclude that C is compact. B
An Excursus into General Topology 165

9.4.10. Every compact subset of a Hausdorff space is closed.


C Let C be compact in a Hausdorff space X. If C = ∅ then there is nothing
to prove. Let C 6= ∅ and x ∈ cl C. By virtue of 9.2.2 there is a filter F0 on X such
that C ∈ F0 and F0 → x. Let F be an ultrafilter finer than F0 . 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. B
9.4.11. Let f : (X, τ ) → (Y, ω) be a continuous one-to-one mapping with
f (X) = Y . If τ is a compact topology and ω is a Hausdorff topology, then f is
a homomorphism.
C It suffices to establish that f −1 is continuous. To this end, we are to demon-
strate that F ∈ Cl(τ ) ⇒ f (F ) ∈ Cl(ω). Take F ∈ Cl(τ ). Then F is compact
by 9.4.9. Successively applying 9.4.5 and 9.4.10, infer that f (F ) is closed. B
9.4.12. Let τ1 and τ2 be two topologies on a set X. If (X, τ1 ) is a compact
space and (X, τ2 ) is a Hausdorff space with τ1 ≥ τ2 , then τ1 = τ2 . CB
9.4.13. Remark. The message of 9.4.12 is customarily verbalized as follows:
“A compact topology is minimal among Hausdorff topologies.”
9.4.14. Theorem. Every Hausdorff compact space is normal.
C Let X be the space under study and let B be some filterbase 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∩(X\int 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 := τ (x). By virtue of 9.3.4, cl B = {x}
and, consequently, the filter τ (x) has a base of closed sets. Thus, X is regular.
Now take nonempty closed subset F of X. Take as B the neighborhood filter
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. B
9.4.15. Corollary. Each Hausdorff compact space is (to within a homeomor-
phism) a closed subset of a Tychonoff cube.
C The compactness property of a closed subset of a Tychonoff cube follows
from 9.4.8 and 9.4.9. Moreover, every cube is a Hausdorff space and so such is each
of its subspaces.
Now take some Hausdorff compact space X. Let Q be the collection of all
continuous function from X to [0, 1]. Define 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. B
166 Chapter 9

9.4.16. Remark. Corollary 9.4.15 presents a part of a more general assertion.


Namely, a Tychonoff space is (to within a homeomorphism) a subspace of a Ty-
chonoff cube. CB
9.4.17. Remark. Sometimes a Hausdorff compact space is also called a com-
pactum (cf. 4.5 and 4.6).
9.4.18. Dieudonnè Lemma. Let F be a closed set and let G1 , . . . , Gn be
open sets in a normal topological space, with F ⊂ G1 ∪ . . . ∪ Gn . There are closed
sets F1 , . . . , Fn such that F = F1 ∪ . . . ∪ Fn and Fk ⊂ Gk (k := 1, . . . , n).
C It suffices to settle the case n := 2. For k := 1, 2 the set Uk := F \ Gk
is closed. Moreover, U1 ∩ U2 = ∅. By 9.3.10 there are open V1 and V2 such that
U1 ⊂ V1 , U2 ⊂ V2 and V1 ∩V2 = ∅. Put Fk := F \Vk . It is clear that Fk is closed and
Fk ⊂ F \Uk = F \(F \Gk ) ⊂ Gk for k := 1, 2. Finally, F1 ∪F2 = F \(V1 ∩V2 ) = F . B
9.4.19. Remark. From 9.3.14 we deduce that under the hypotheses of 9.4.18
there are continuous functions h1 , . . . , hn : X → [0, 1] such that hk |G0k = 0 and
Pn
k=1 hk (x) = 1 for every point x in some neighborhood about F . (As usual,
0
Gk := X \ Gk .)
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 Hausdorff 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.
C ⇐: In virtue of the Weierstrass Theorem it suffices 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 filter 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 ultrafilter 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 ∞. B
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 Alexandroff compactification of X.

9.5. Uniform and Multimetric Spaces


9.5.1. Definition. Let X be a nonempty set and let UX be a filter on X 2 .
The filter UX is a uniformity on X if
An Excursus into General Topology 167

(1) UX ⊂ fil {IX };


(2) U ∈ UX ⇒ U −1 ∈ UX ;
(3) (∀ U ∈ UX )(∃ V ∈ UX ) V ◦ V ⊂ U .
The uniformity of the empty set is by definition {∅}. The pair (X, UX ), as well
as the underlying set X, is called a uniform space.
9.5.2. Given a uniform space (X, UX ), put

x ∈ X ⇒ τ (x) := {U (x) : U ∈ UX }.

The mapping τ : x 7→ τ (x) is a topology on X.


C Clearly, τ is a pretopology. If W ∈ τ (x) then W = U (x) for some U
in UX . Choose a member V of UX 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. B
9.5.3. Definition. The topology τ appearing in 9.5.2 is the topology of the uni-
form space (X, UX ) under consideration or the uniform topology on X. It is also
denoted by τ (UX ), τX , etc.
9.5.4. Definition. A topological space (X, τ ) is called uniformizable pro-
vided that there is a uniformity U on X such that τ coincides with the uniform
topology τ (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, UY ) and f −1 (UY ) := f ×−1 (UY ), where as usual
f × (x1 , x2 ) := (f (x1 ), f (x2 )) for (x1 , x2 ) ∈ X 2 . Evidently, f −1 (UY ) is a uniformity
on X. Moreover,
τ (f −1 (UY )) = f −1 (τ (UY )).
The uniformity f −1 (UY ) is the inverse image of UY under f . Therefore, the inverse
image of a uniform topology is uniformizable.
Q(4) Let (Xξ , Uξ )ξ∈„ be a family of uniform spaces. Assume further
that X := ξ∈„ Xξ is the product of the family. Put UX := supξ∈„ Pr−1 ξ (Uξ ). The
uniformity UX is the Tychonoff uniformity. It is beyond a doubt that the uniform
topology τ (UX ) is the Tychonoff topology of the product of (Xξ , τ (Uξ ))ξ∈„ . CB
168 Chapter 9

(5) Each Hausdorff compact space is uniformizable in a unique fashion.


C By virtue of 9.4.15 such space X may be treated as a subspace of a Ty-
chonoff 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 IX of X 2 implies that every neighborhood of IX
belongs to UX . On the other hand, each entourage is always a neighborhood of the
diagonal. B
(6) Assume that X and Y are nonempty sets, UY is a uniformity on Y
and B is an upward-filtered subset of P(X). Given B ∈ B and θ ∈ UY , put
UB,θ := {(f, g) ∈ Y X × Y X : g ◦ IB ◦ f −1 ⊂ θ}.
Then U := fil {UB,θ : B ∈ B, θ ∈ UY } 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 finite subsets of X, then U coincides with the
Tychonoff 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 τ (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 4.2.4–4.2.9, 4.5.8, 4.5.9,
and 4.6.1–4.6.7 are preserved. It is a rewarding practice to ponder over a possibility
of completing a uniform space, to validate a uniform version of the Hausdorff Cri-
terion, to inspect the proof of the Ascoli–Arzelá 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 Ud := fil {{d ≤ ε} : ε > 0}. Then
Ud is a uniformity. CB
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 defined as UM := sup{Ud : d ∈ M}.
An Excursus into General Topology 169

9.5.10. Definition. A uniform space is called multimetrizable, if its unifor-


mity coincides with some multimetric uniformity. A multimetrizable topological
space is defined by analogy.
9.5.11. Assume that X, Y , and Z are sets, T is a dense subset of R, and
(Ut )t∈T and (Vt )t∈T 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} ⊂ Ut ⊂ {f ≤ t}, {g < t} ⊂ Vt ⊂ {g ≤ t},


{h < t} ⊂ Vt ◦ Ut ⊂ {h ≤ t} (t ∈ T ).

Moreover, the presentation holds:

h(x, y) = inf{f (x, z) ∨ g(z, y) : z ∈ Z}.

C 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. B
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

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) ultra-
metric inequality. In virtue of 9.5.12 this inequality may be rewritten as K −1 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 .
C Let f1 := f and fn+1 := fn
We are left with proving that 1 /2 f ≤ d. To this end, show that fn ≥ 1 /2 f for
n ∈ N. Proceed by way of induction.
The desired inequalities are obvious when n := 1, 2. Assume now that f ≥ f1 ≥
. . . ≥ fn ≥ 1 /2 f and at the same time fn+1 (x, y) < 1 /2 f (x, y) for some (x, y)
in X 2 and n ≥ 2.
170 Chapter 9

For suitable z1 , . . . , zn in X by construction

t := f (x, z1 ) + f (z1 , z2 ) + . . . + f (zn−1 , zn ) + f (zn , y)


< 1 /2 f (x, y).

If f (x, z1 ) ≥ t /2 then t /2 ≥ f (z1 , z2 ) + . . . + f (zn , y) ≥ 1 /2 f (z1 , y). It follows


that t ≥ f (x, z1 ) and t ≥ f (z1 , y). On account of 9.5.13 (3), 1 /2 f (x, y) ≤
f (x, z1 ) ∨ f (z1 , y) ≤ t. Whence we come to 1 /2 f (x, y) > t ≥ 1 /2 f (x, y), which is
false.
Thus, f (x, z1 ) < t /2 . Find m ∈ N, m < n, satisfying

f (x, z1 ) + . . . + f (zm−1 , zm ) < t /2 ;

f (x, z1 ) + . . . + f (zm , zm+1 ) ≥ t /2 .


This is possible, since assuming m = n would entail the false inequality f (zn , y) ≥
t
/2 . (We would have t /2 ≥ f (x, z1 ) + . . . + f (zn−1 , zn ) ≥ 1 /2 f (x, zn ) and so
1
/2 f (x, y) > t ≥ f (x, z n ) ∨ f (zn , y) ≥ 1 /2 f (x, y).)
We obtain the inequality

f (zm+1 , zm+2 ) + . . . + f (zn−1 , zn ) + f (zn , y) < t /2 .

Using the induction hypothesis, conclude that

f (x, zm ) ≤ 2(f (x, z1 ) + . . . + f (zm−1 , zm )) ≤ t;


f (zm , zm+1 ) ≤ t;
f (zm+1 , y) ≤ 2(f (zm+1 , zm+2 ) + . . . + f (zn , y)) ≤ t.

Consequently, by the definition of 2-ultrametric

1
/2 f (x, y) ≤ f (x, zm ) ∨ f (zm , zm+1 ) ∨ f (zm+1 , y) ≤ t
< 1 /2 f (x, y).

We arrive at a contradiction, completing the proof. B


9.5.16. Theorem. Every uniform space is multimetrizable.
C Let (X, UX ) be a uniform space. Take V ∈ UX . Put V1 := V ∩ V −1 . If now
−1
Vn ∈ UX then find a symmetric entourage V = V , a member of UX , satisfying
V ◦ V ◦ V ⊂ Vn . Define Vn+1 := V . Since by construction

Vn ⊃ Vn+1 ◦ Vn+1 ◦ Vn+1 ⊃ Vn+1 ◦ IX ◦ IX ⊃ Vn+1 ;


An Excursus into General Topology 171

therefore, (Vn )n∈N is a decreasing family.


Given t ∈ R, define a set Ut by the rule

t<0

 ∅,

 IX , t=0



Ut := Vinf{n∈N : t≥2−n } , 0<t<1

V1 , t=1




 2
X , t > 1.

By definition, the family t 7→ Ut (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} ⊂ Ut ⊂ {f ≤ t} (t ∈ R).

If Wt := U2t for t ∈ R then

Us ◦ Us ◦ Us ⊂ Wt

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, find a semimetric dV such that 1 /2 f ≤ dV ≤ f . Clearly, UdV =
fil {Vn : n ∈ N}. It is also beyond a doubt that UM = UX for the multimetric
M := {dV : V ∈ UX }. B
9.5.17. Corollary. A topological space is uniformizable if and only if it is
a T31 /2 -space. CB
9.5.18. Corollary. A Tychonoff space is the same as a separated multimetric
space. CB

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 refines 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 refinement of F . Observe that if F is a subcover of E
(i.e., F ⊂ E ) then E refines F .
9.6.2. Definition. A cover E of a set X is called locally finite (with respect
to a topology τ on X), if each point in X possesses a neighborhood (in the sense
of τ ) meeting only finite many members of E . In the case of the discrete topology on
X, such cover is called point finite. If X is regarded as furnished with a prescribed
topology τ then, speaking of a locally finite cover of X, we imply the topology τ .
172 Chapter 9

9.6.3. Lefschetz Lemma. Let E be a point finite open cover of a normal


space X. Then there is an open cover {GE : E ∈ E } such that cl GE ⊂ E for all
E ∈ E.
C 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 s1 and s2 , put
s1 ≤ s2 := (∀ E ∈ E ) (s1 (E) 6= E ⇒ s2 (E) = s1 (E)). It is evident that (S, ≤) is
an ordered set and IE ∈ S. Show that S is inductive.
Given a chain S0 in S, for all E ∈ E put s0 (E) := ∩{s(E) : s ∈ S0 }. If s0 (E) =
E then s(E) = E for all s ∈ S0 . If the case s0 (E) 6= E observe that s0 (E) =
∩{s(E) : s(E) 6= E, s ∈ S0 }.
Since the order of S0 is linear, infer that s0 (E) = s(E) for s ∈ S0 with s(E) 6=
E. Hence, s0 (E ) ⊂ Op(X) and s0 ≥ S0 . It remains to verify that s0 is a cover of X
(and so s0 ∈ S). By the hypothesis of point finiteness, given x ∈ X, there are some
E1 , . . . , En in E such that x ∈ E1 ∩ . . . ∩ En and x ∈ / E for the other members E
of E . If s0 (Ek ) = Ek for some k, then there is nothing to prove, for x ∈ ∪s0 (E ).
In the case when s0 (Ek ) 6= Ek for every k, there are s1 , . . . , sn ∈ S0 meeting the
conditions sk (Ek ) 6= Ek (k := 1, 2 . . . , n). Since S0 is a chain, it may be assumed
that sn ≥ {s1 , . . . , sn−1 }. Moreover, x ∈ sn (E) ⊂ E for an appropriate E in E .
It is clear that E ∈ {E1 , . . . , En } (because x ∈/ E for the other members E of E ).
Since s0 (E) = sn (E), it follows that x ∈ s0 (E).
By the Kuratowski–Zorn 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 6= 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) 6= E,
then cl s(E) ⊂ E by definition. Thus, s is a sought cover. B
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 ) := cl{x ∈ X : f (x) 6= 0}
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 (fe )e∈E be a family of numeric functions on X and let E := {supp(fe ) :
e ∈ E } be the family of their supports. If E is a point finite cover of X then the
family (fe )e∈E is summable pointwise. If in addition
P E is locally finite and every
member of (fe )e∈E is continuous, then the sum e∈E fe is also continuous.
C It suffices to observe that in a suitable neighborhood about a point in X
only finitely many members of the family (fe )e∈E are distinct from zero. B
9.6.6. Definition. It is said that a family of functions (f : X → [0, 1])f ∈F is
a partition of unity on a subset U of X, if the supports of the members of the family
An Excursus into General Topology 173
P
composes a point finite cover of X, and f ∈F 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 } refines E then F is a partition of unity subordinate to E .
A possibility of finding such an F for E is also verbalized as follows: “E admits
a partition of unity.”
9.6.8. Each locally finite open cover of a normal space admits a partition
of unity.
C By the Lefschetz Lemma, such cover {Uξ : ξ ∈ „} has an open refinement
{Vξ : ξ ∈ „} satisfying the condition cl Vξ ⊂ Uξ for all ξ ∈ „. By the Urysohn
Theorem, there is a continuous function gξ : X → [0, 1] such that gξ (x) = 1 for
x ∈ Vξ and gξ (x) = 0 for x ∈ X \Uξ . Consequently, supp(gξ ) ⊂ Uξ . In virtue of 9.6.5
the family (gξ )ξ∈„ is summable pointwise to a continuous function g. Moreover,
g(x) > 0 for all x ∈ X by construction. Put fξ := gξ /g (ξ ∈ „). The family (fξ )ξ∈„
is what we need. B
9.6.9. Definition. A topological space X is called paracompact, if each cover
of X has a locally finite open refinement.
9.6.10. Remark. The theory of paracompactness contains deep and surprising
facts.
9.6.11. Theorem. Every metric space is paracompact.
9.6.12. Theorem. A Hausdorff topological space is paracompact if and only
if its every open cover admits a partition of unity.
9.6.13. Remark. The metric space RN possesses a number of additional struc-
tures providing a stock of well-behaved, smooth (= infinitely differentiable) func-
tions (cf. 4.8.1).
9.6.14. Definition. A mollifier or a mollifying kernel on RN 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 ∈ RN : |x| ≤ 1} is
the (unit Euclidean) ball B := BR N .
9.6.15. Definition. A delta-like sequence is a family of real-valued (smooth)
R ε )ε>0 such that, first, limε→0 (sup | supp(bε )|) = 0 and, second, the equal-
functions (b
ity holds R N bε (x) dx = 1 for all ε > 0. The terms “δ-sequence” and “δ-like se-
quence” are also in current usage. Such a sequence is often assumed countable
without further specification.
174 Chapter 9

9.6.16. Example. The function a(x) := t exp(−(|x|2 −1)−1 ) is taken as a most


popular mollifier when extended by zeroR beyond the open ball int B, with the con-
stant t determined from the condition R N a(x) dx = 1. Each mollifier generates
the delta-like sequence aε (x) := ε−N a(x/ε) (x ∈ RN ).
9.6.17. Definition. Let f ∈ L1,loc (RN ); i.e., let f be a locally integrable
function, that is, a function whose restriction to each compact subset of RN is
integrable. For a compactly-supported integrable function g the convolution f ∗ g
is defined as Z
f ∗ g(x) := f (x − y)g(y) dy (x ∈ RN ).
RN
9.6.18. Remark. The role of a mollifying kernel and the corresponding delta-
like sequence (aε )ε>0 becomes clear from inspecting the aftermath of applying the
smoothing process f 7→ (f ∗aε )ε>0 to a function f belonging to L1,loc (RN ) (cf. 10.10.7
(5)).
9.6.19. The following statements are valid:
(1) to every compact set K in the space RN and every neighborhood U
of K there corresponds a truncator (= a bump function) ψ := ψK,U , i.e. a smooth
mapping ψ : RN → [0, 1] such that K ⊂ int{ψ = 1} and supp(ψ) ⊂ U ;
(2) assume that U1 , . . . , Un ∈ Op(RN ) and U1 ∪ . . . ∪ Un is a neigh-
N
Pn are smooth functions ψ1 , . . . , ψn : R → [0, 1]
borhood of a compact set K; there
such that supp(ψk ) ⊂ Uk and k=1 ψk (x) = 1 for x in some neighborhood of K.
C (1) Put ε := d(K, RN \ U ) := inf{|x − y| : x ∈ K, y ∈ / U }. It is clear that
ε > 0. Given β > 0, denote the characteristic function of K + εB by χβ . Take
a delta-like sequence (bγ )γ>0 of positive functions and put ψ := χβ ∗ bγ . When
γ ≤ β and β + γ ≤ ε with γ := sup | supp(bγ )|, observe that ψ is a sought function.
(2) By the Diedonnè Lemma there are closed sets Fk , with Fk ⊂ Uk , composing
a cover of K. P Put Kk := Fk ∩ K and choose some P truncators ψk := ψKk ,Uk . The
n n
functions ψk / k=1 ψk (k := 1, P. . . , n), defined on { k=1 ψk > 0}, meet the claim
n
after extension by zero onto { k=1 ψk = 0} and multiplication by a truncator
corresponding to an appropriate neighborhood of K. B
9.6.20. Countable Partition Theorem. Let E be a family of open sets
in RN and Š := ∪E . There is a countable partition of unity which is composed
of smooth compactly-supported functions on RN and subordinate to the cover E
of Š.
C Refine from E a countable locally finite cover A of Š with compact sets so
that the family (α := int α)α∈A be also an open cover of Š. Choose an open cover
(Vα )α∈A of Š from the condition cl Vα ⊂ α for α ∈ A. In Pvirtue of 9.6.19 (1) there
are truncators ψ α := ψcl Vα ,α . Putting ψα (x) := ψ α (x)/ α∈A ψ α (x) for x ∈ Š and
ψα (x) := 0 for x ∈ RN \ Š, arrive to a sought partition. B
An Excursus into General Topology 175

9.6.21. Remark. It is worth observing that the so-constructed partition of unity


(ψα )α∈A possesses the property that to each compact subset K ofPŠ there corre-
spond a finite subset A0 of A and a neighborhood U of K such that α∈A0 ψα (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 filters or sequences?
9.3. Establish relations between topologies and preorders on a finite 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 (fξ : X0 = 0 O Yξ , τξ ))ξ∈„ be a family of mappings. A topology σ on X is called
admissible (in the case under study), if for every topological space (Z, ω) and every mapping
g : Z0 = 0 O the following statement holds: g : (Z, ω)0 = 0 O X, σ) is continuous if and only
if so is each mapping fξ ◦ g (ξ ∈ „). Demonstrate that the weakest topology on X making every
fξ (ξ ∈ „) continuous is the strongest admissible topology (in the case under study).
9.6. Let (fξ : (Xξ , σξ )0 = 0 O )ξ∈„ be a family of mappings. A topology τ on Y is called
admissible (in the case under study), if for every topological space (Z, ω) and every mapping
g : Y 0 = 0 O the following statement is true: g : (Y, τ )0 = 0 O Z, ω) is continuous if and only
if each mapping g ◦ fξ (ξ ∈ „) is continuous. Demonstrate that the strongest topology on Y
making every fξ (ξ ∈ „) continuous is the weakest admissible topology (in the case under study).
9.7. Prove that in the Tychonoff product of topological spaces, the closure of the product
of subsets of the factors is the product of closures:
!
Y Y
cl Aξ = cl Aξ .
ξ∈„ ξ∈„

9.8. Show that a Tychonoff product is a Hausdorff space if and only if so is every factor.
9.9. Establish compactness criteria for subsets of classical Banach spaces.
9.10. A Hausdorff space X is called H-closed, if X is closed in every ambient Hausdorff
space. Prove that a regular H-closed space is compact.
9.11. Study possibilities of compactifying a topological space.
9.12. Prove that the Tychonoff 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 field F.
A topology τ on X is a topology compatible with vector structure or, briefly, a vector
topology, if the following mappings are continuous:

+ : (X × X, τ × τ ) → (X, τ ),
· : (F × X, τF × τ ) → (X, τ ).

The space (X, τ ) is then referred to as a topological vector space.


10.1.2. Let τX be a vector topology. The mappings

x 7→ x + x0 , x 7→ αx (x0 ∈ X, α ∈ F \ 0)

are topological isomorphisms in (X, τX ). CB


10.1.3. Remark. It is beyond a doubt that a vector topology τ on a space X
possesses the next “linearity” property:

τ (αx + βy) = ατ (x) + βτ (y) (α, β ∈ F \ 0; x, y ∈ X),

where in accordance with the general agreements (cf. 1.3.5 (1))

Uαx+βy ∈ ατ (x) + βτ (y)


⇔ (∃ Ux ∈ τ (x) & Uy ∈ τ (y)) αUx + βUy ⊂ Uαx+βy .

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.
Duality and Its Applications 177

10.1.4. Theorem. Let X be a vector space and let N be a filter on X.


There is a vector topology τ on X such that N = τ (0) if and only if the following
conditions are fulfilled:
(1) N + N = N ;
(2) N consists of absorbing sets;
(3) N has a base of balanced sets.
Moreover, τ (x) = x + N for all x ∈ X.
C ⇒: Let τ be a vector topology and N = τ (0). From 10.1.2 infer that
τ (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 τF (0)x ⊃ N for every x in X, which is the continuity property of the mapping
α 7→ αx 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 τF (0)N = N , which is the continuity
property of scalar multiplication at zero (of the space F × X).
⇐: Let N be a filter satisfying (1)–(3). It is evident that N ⊂ fil {0}. Put
τ (x) := x + N . Then τ is a pretopology. From the definition of τ and (1) it follows
that τ 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 (λ, x) 7→ λx is jointly continuous at zero
and continuous at zero in the first argument given the second argument. By virtue
of the identity

λx − λ0 x0 = λ0 (x − x0 ) + (λ − λ0 )x0 + (λ − λ0 )(x − x0 ),

we are left with examining the continuity property of scalar multiplication at zero
in the second argument given the first argument. In other words, it is necessary
to show that λN ⊃ N for λ ∈ F. With this in mind, find n ∈ N such that |λ| ≤ n.
Let V in N and W in N be such that  W is balanced and W1 + . . . + Wn ⊂ V ,
λ
where Wk := W . Then λW = n /n W ⊂ nW ⊂ W1 + . . . + Wn ⊂ V . B
10.1.5. Theorem. The set VT(X) of all vector topologies on X presents
a complete lattice. Moreover,

supVT(X) E = supT(X) E

for every subset E of VT(X).


C Let τ := supT(X) E . Since for τ ∈ E each translation by a vector is a topolog-
ical isomorphism in (X, τ ); therefore, this mapping is a topological isomorphism in
(X, τ ). Using 9.1.13, observe that the filter τ (0) meets conditions 10.1.4 (1)–10.1.4
(3), since these conditions are fulfilled for every filter τ (0) with τ ∈ E . It remains
to refer to 1.2.14. B
178 Chapter 10

10.1.6. Theorem. The inverse image of a vector topology under a linear op-
erator is a vector topology.
C Take T ∈ L (X, Y ) and ω ∈ VT(Y ). Put τ := T −1 (ω). If xγ → x and
yγ → y in (X, τ ) then by 9.2.8 T xγ → T x and T yγ → T y. So T (xγ +yγ ) → T (x+y).
This means in virtue of 9.2.10 that xγ +yγ → x+y in (X, τ ). Thus, τ (x) = x+τ (0)
for all x ∈ X and, moreover, τ (0) + τ (0) = τ (0). Successively applying 3.4.10
and 3.1.8 to the linear correspondence T −1 , observe that the filter τ (0) = T −1 (ω(0))
consists of absorbing sets and has a base of balanced sets. The reason is as follows:
by 10.1.4 the filter ω(0) possesses these two properties. Once again using 10.1.4,
conclude that τ ∈ VT(X). B
10.1.7. The product of vector topologies is a vector topology.
C Immediate from 10.1.5 and 10.1.6. B
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 τ on X there corresponds a unique uniformity
Uτ having a base of IX -stable sets and such that τ = τ (Uτ ).
C Given U ∈ τ (0), put VU := {(x, y) ∈ X 2 : y − x ∈ U }. Observe the obvious
properties:

IX ⊂ VU ; VU + IX = VU ; (VU )−1 = V−U ;


VU1 ∩U2 ⊂ VU1 ∩ VU2 ; VU1 ◦ VU2 ⊂ VU1 +U2

for all U , U1 , U2 ∈ τ (0). Using 10.1.4, infer that Uτ := fil {VU : U ∈ τ (0)} is
a uniformity and τ = τ (Uτ ). It is also beyond a doubt that Uτ has a base of
IX -stable sets.
If now U is another uniformity such that τ (U ) = τ , and W is some IX -stable
entourage in U ; then W = VW (0) . Whence the sought uniqueness follows. B
10.1.10. Definition. Let (X, τ ) be a topological vector space. The unifor-
mity Uτ , 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 specification.

10.2. Locally Convex Topologies


10.2.1. Definition. A vector topology is locally convex if the neighborhood
filter of each point has a base of convex sets.
Duality and Its Applications 179

10.2.2. Theorem. Let X be a vector space and let N be a filter on X. There


is a locally convex topology τ on X such that N = τ (0) if and only if
(1) 1 /2 N = N ;
(2) N has a base of absorbing absolutely convex sets.
C ⇒: By virtue of 10.1.2 the mapping x 7→ 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, find 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 satisfies 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. B
10.2.3. Corollary. The set LCT (X) of all locally convex topologies on X is
a complete lattice. Moreover,
supLCT (X) E = supT(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. CB
10.2.5. Corollary. The product of locally convex topologies is a locally con-
vex topology. CB
10.2.6. The topology of a multinormed space is locally convex. CB
10.2.7. Definition. Let τ be a locally convex topology on X. The set of
all everywhere-defined continuous seminorms on X is called the mirror (rarely, the
spectrum) of τ and is denoted by Mτ . The multinormed space (X, Mτ ) is called
associated with (X, τ ).
10.2.8. Theorem. Each locally convex topology coincides with the topology
of the associated multinormed space.
C Let τ be a locally convex topology on X and let ω := τ (Mτ ) be the topology
of the associated space (X, Mτ ). Take V ∈ τ (0). By 10.2.2 there is an absolutely
convex neighborhood B of zero, B ∈ τ (0), such that B ⊂ V . In virtue of 3.8.7
{pB < 1} ⊂ B ⊂ {pB ≤ 1}.
It is obvious that pB is continuous (cf. 7.5.1); i.e., pB ∈ Mτ , and so {pB < 1} ∈ ω(0).
Consequently, V ∈ ω(0). Using 5.2.10, infer that ω(x) = x+ω(0) ⊃ x+τ (0) = τ (x);
i.e., ω ≥ τ . Furthermore, τ ≥ ω by definition. B
180 Chapter 10

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 sepa-
rated multinormed space have the same scope.” For that reason, the terminology
connected with the associated multinormed space is lavishly applied to studying
a locally convex space (cf. 5.2.13).
10.2.11. Definition. Let τ be a locally convex topology on X. The symbol
(X, τ )0 (or, in short, X 0 ) denotes the subspace of X # that comprises all continuous
linear functionals. The space (X, τ )0 is the dual (or τ -dual) of (X, τ ).
10.2.12. (X, τ )0 = ∪ {|∂|(p) : p ∈ Mτ }. CB
10.2.13. Prime Theorem. The prime mapping τ 7→ (X, τ )0 from LCT (X)
to Lat(X # ) preserves suprema; i.e.,
(X, sup E )0 = sup{(X, τ )0 : τ ∈ E }
for every subset E of LCT (X).
C If E = ∅ then sup E is the trivial topology τ◦ of X and, consequently,
(X, τ◦ )0 = 0 = inf Lat(X # ) = supLat(X # ) ∅. By virtue of 9.2.7 the prime mapping
increases. Given a nonempty E , from 2.1.5 infer that
(X, sup E )0 ≥ sup{(X, τ )0 : τ ∈ E }.
If f ∈ (X, sup E )0 , then in view of 10.2.12 and 9.1.13 there are topologies
τ1 , . . . , τn ∈ E such that f ∈ (X, τ1 ∨ . . . ∨ τn )0 . Using 10.2.12 and 5.3.7, find
p1 ∈ Mτ1 , . . . , pn ∈ Mτn satisfying f ∈ |∂|(p1 ∨ . . . ∨ pn ). Recalling 3.5.7 and 3.7.9,
observe that |∂|(p1 + . . . + pn ) = |∂|(p1 ) + . . . + |∂|(pn ). Finally,
f ∈ (X, τ1 )0 + . . . + (X, τn )0 = (X, τ1 )0 ∨ . . . ∨ (X, τn )0 . .
10.3. Duality Between Vector Spaces
10.3.1. Definition. Let X and Y be vector spaces over the same ground field
F. Also, Assume given a bilinear form (or, as it is called sometimes, a bracketing)
h· | ·i acting from X × Y to F, i.e. a mapping linear in each of its arguments. Given
x ∈ X and y ∈ Y , put

hx | : y 7→ hx | yi, h· | : X → F Y , hX | ⊂ Y # ;
| yi : x 7→ hx | yi, | ·i : Y → F X, | Y i ⊂ X #.
The mappings h· | and | ·i are the bra-mapping and the ket-mapping of the initial
bilinear form. By analogy, a member of hX | is a bra-functional on X and a member
of | Y i is a ket-functional on Y .
Duality and Its Applications 181

10.3.2. The bra-mapping and the ket-mapping are linear operators. CB


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 h· | ·i. Given (y, x) ∈ Y × X,
put hy | xi := hx | yi. 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 unified 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 hx | yiR := Rehx | yi sets in duality the real
carriers XR and YR . By way of taking liberties, the previous notation is sometimes
reserved for the arising duality XR ↔ YR ; i.e., it is assumed that hx | yi := hx | yiR ,
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, τ ) be a locally convex space and let X 0 be the dual of X.
The natural evaluation mapping (x, x0 ) 7→ x0 (x) sets X and X 0 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# ) 7→ x# (x) sets the spaces
in duality.
10.3.5. Definition. Let X ↔ Y . The inverse image in X of the Tychonoff
topology on F Y under the bra-mapping, further denoted by σ(X, Y ), is the bra-
topology or the weak topology on X induced by Y . The bra-topology σ(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.
C xγ → x (in σ(X, Y )) ⇔ hxγ | → hx | (in F Y ) ⇔ (∀ y ∈ Y ) hxγ | (y) →
hx | (y) ⇔ (∀ y ∈ Y ) hxγ | yi → hx | yi ⇔ (∀ y ∈ Y ) | yi(xγ ) → | yi(x) ⇔ (∀ y ∈ Y )
xγ → x (in | yi−1 (τF )) B
182 Chapter 10

10.3.7. Remark. The notation σ(X, Y ) agrees perfectly with that of the
weak multinorm in 5.1.10 (4). Namely, σ(X, Y ) is the topology of the multinorm
{|h· | yi| : y ∈ Y }. Likewise, σ(Y, X) is the topology of the multinorm {|hx | ·i| :
x ∈ X}. CB

10.3.8. The spaces (X, σ(X, Y )) and (Y, σ(Y, X)) are locally convex.
C Immediate from 10.2.4 and 10.2.5. B

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.
C Consider a duality pair X ↔ Y . We are to prove exactness for the sequences

h·| |·i
0 → X −→(Y, σ(Y, X))0 → 0; 0 → Y −→(X, σ(X, Y ))0 → 0.

Since the ket-mapping of X ↔ Y is the bra-mapping of Y ↔ X, it suffices to show


that the first sequence is exact. The bra-mapping is a monomorphism by definition.
Furthermore, from 10.2.13 and 10.3.6 it follows that

(Y, σ(Y, X))0 = (Y, sup{hx |−1 (τF ) : x ∈ X})0


= sup{(Y, hx |−1 (τF ))0 : x ∈ X} = lin({(Y, f −1 (τF ))0 : f ∈ hX |}) = hX |,

since in view of 5.3.7 and 2.3.12 (Y, f −1 (τF ))0 = {λf : λ ∈ 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) := hx | yi, 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, σ(Y, X))0 and Y with (X, σ(X, Y ))0 , which justifies writing
X 0 = Y and Y 0 = X.

10.3.11. Remark. A somewhat obsolete convention relates to X ↔ X 0 with


X a normed space. The ket-topology σ(X 0 , X) is customarily called the weak∗
topology (read: weak-star topology) in X 0 , which reflects the concurrent notation
X ∗ for X 0 . The term “weak∗ ” proliferates in a routine fashion elsewhere.
Duality and Its Applications 183

10.4. Topologies Compatible with Duality


10.4.1. Definition. Take a duality pair X ↔ Y and let τ be a locally convex
topology on X. It is said that τ is compatible with duality (between X and Y
by pairing X ↔ Y ), provided that (X, τ )0 = | Y i. A locally convex topology ω
on Y is compatible with duality (by pairing X ↔ Y ), if ω is compatible with duality
(by pairing Y ↔ X); i.e., if the equality holds: (Y, ω)0 = hX |. A unified concise
term “compatible topology” is also current in each of the above cases.
10.4.2. Weak topologies are compatible.
C Follows from 10.3.9. B
10.4.3. Let τ (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 τ (X, Y ) is also compatible.
C Denote the set of all compatible topologies on X by E . Theorem 10.2.13
readily yields the equalities

(X, τ (X, Y ))0 = sup{(X, τ )0 : τ ∈ E } = sup{| Y i : τ ∈ E } = | Y i,

because E is nonempty by 10.4.2. B


10.4.4. Definition. The topology τ (X, Y ), constructed in 10.4.3 (i.e., the
finest 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. Mackey–Arens Theorem. A locally convex topology τ on X is com-
patible with duality between X and Y if and only if

σ(X, Y ) ≤ τ ≤ τ (X, Y ).

C By 10.2.13 the prime mapping τ 7→ (X, τ )0 preserves suprema and, in par-


ticular, increases. Therefore, given τ in the interval of topologies, from 10.4.2
and 10.4.3 obtain

| Y i = (X, σ(X, Y )) ⊂ (X, τ )0 ⊂ (X, τ (X, Y ))0 = | Y i.

The remaining claim is obvious. B


10.4.6. Mackey Theorem. All compatible topologies have the same
bounded sets in stock.
184 Chapter 10

C A stronger topology has fewer bounded sets. So, to prove the theorem
it suffices 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 X0 := X/ ker p and p0 := pX/ ker p . By 5.2.14, p0 is
clearly a norm. Let ϕ : X → X0 be the coset mapping. It is beyond a doubt that
ϕ(U ) is weakly bounded in (X0 , p0 ). From 7.2.7 it follows that ϕ(U ) is bounded
in the norm p0 . Since p0 ◦ ϕ = p; therefore, U is bounded in (X, p). B
10.4.7. Corollary. Let X be a normed space. Then the Mackey topology
τ (X, X 0 ) coincides with the initial norm topology on X.
C It suffices to refer to the Kolmogorov Normability Criterion implying that the
space X with the topology τ (X, X 0 ) finer than the original topology is normable.
Appealing to 5.3.4 completes the proof. B
10.4.8. Strict Separation Theorem. Let (X, τ ) 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, τ )0 , such
that
sup Re f (K) < inf Re f (V ).

C 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 suffices to consider the filterbases 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 (XR )# , such that the
hyperplane {l = 1} in XR separates V from U and does not meet the core of U .
Obviously, l is bounded above on W and so l ∈ (XR , τ )0 by 7.5.1. If f := Re−1 l
then, of view of 3.7.5, f ∈ (X, τ )0 . It is clear that f is a sought functional. B
10.4.9. Mazur Theorem. All compatible topologies have the same closed
convex sets in stock.
C A stronger topology has more closed sets. So, to prove the theorem it suffices
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. B

10.5. Polars
10.5.1. Definition. Let X and Y be sets and let F ⊂ X × Y be a correspon-
Duality and Its Applications 185

dence. Given a subset U of X and a subset V of Y, put

π(U ) := πF (U ) := {y ∈ Y : F −1 (y) ⊃ U };
π −1 (V ) := πF−1 (V ) := {x ∈ X : F (u) ⊃ V }.

The set π(U ) is the (direct) polar of U (under F ), and the set π −1 (V ) is the (reverse)
polar of V (under F ).
10.5.2. The following statements are valid:
(1) π(u) := π({u}) = F (u) and π(U ) = ∩u∈U π(u);
(2) π(∪ξ∈„ Uξ ) = ∩ξ∈„ π(Uξ );
(3) πF−1 (V ) = πF −1 (V );
(4) U1 ⊂ U2 ⇒ π(U1 ) ⊃ π(U2 );
(5) U × V ⊂ F ⇒ (V ⊂ π(U ) & U ⊂ π −1 (V );
(6) U ⊂ π −1 (π(U )). CB
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 ⊂ π −1 (y), / π −1 (y).
x∈

C ⇒: If U = π −1 (V ) then U = ∩v∈V π −1 (v) by 10.5.2 (1).


⇐: The inclusion U ⊂ π −1 (y) means that y ∈ π(U ). Thus, by hypothesis
U = ∩y∈π(U ) π −1 (y) = π −1 (π(U )). B
10.5.4. Corollary. The set π −1 (π(U )) is the (inclusion) least polar greater
than U . CB
10.5.5. Definition. The set πF−1 (πF (U )) is the bipolar of a subset U (under
the correspondence F ).
10.5.6. Examples.
(1) Let (X, σ) be an ordered set and let U be a subset of X. Then
πσ (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 π(U ) = π −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 0 be the dual of X. Consider
F := {(x, x0 ) : x0 (x) = 0}. Then π(X0 ) = X0⊥ and π −1 (X0 ) = ⊥X0 for a subspace
X0 of X and a subspace X0 of X 0 (cf. 7.6.8). Moreover, π −1 (π(X0 )) = cl X0
by 7.5.14.
186 Chapter 10

10.5.7. Definition. Let X ↔ Y . Put


pol := {(x, y) ∈ X × Y : Rehx | yi ≤ 1};
abs pol := {(x, y) ∈ X × Y : |hx | yi| ≤ 1}.
To refer to direct or inverse polars under pol , we use the unified term “polar” (with
respect to X ↔ Y ) and the unified designations π(U ) and π(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 π 2 (U ) := π(π(U )) is the (inclusion)
least weakly closed conical segment greater than U .
C Straightforward from 10.4.8 and the Akilov Criterion. B
10.5.9. Absolute Bipolar Theorem. The absolute bipolar U ◦◦ := (U ◦ )◦ is
the (inclusion) least weakly closed absolutely convex set greater than U .
C It suffices, first, to observe that the polar of a balanced set U coincides with
the absolute polar of U and, second, to apply 10.5.8. B

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) subdifferential ∂(p)
is compact in the topology σ(X 0 , X).
Q
C Put Q := x∈X [−p(−x), p(x)] and endow Q with the Tychonoff topology.
Evidently ∂(p) ⊂ Q, with the Tychonoff topology on Q and σ(X 0 , 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 Tychonoff Theorem and 9.4.9, conclude that
∂(p) is a σ(X 0 , X)-compact set. B
10.6.2. The balanced subdifferential of each continuous seminorm is weakly
compact. CB
10.6.3. Theorem. Let X be a real vector space. A subset U of X # is the
subdifferential of a (unique total) sublinear functional sU : X → R if and only if U
is nonempty, convex and σ(X # , X)-compact.
C ⇒: Let U = ∂(sU ) for some sU . The uniqueness of sU is ensured by 3.6.6.
In view of 10.2.12 it is easy that the mirror of the Mackey topology τ (X, X # ) is
the strongest multinorm on X (cf. 5.1.10 (2)). Whence we infer that the functional
sU is continuous with respect to τ (X, X # ). In virtue of 10.6.1 the set U is compact
in σ(X # , X). The convexity and nonemptiness of U are obvious.
⇐: Put sU (x) := sup{l(x) : l ∈ U }. Undoubtedly, sU is a sublinear functional
and dom sU = X. By definition, U ⊂ ∂(sU ). If l ∈ ∂(sU ) and l ∈ / U , then by the
Strict Separation Theorem and the Dualization Theorem sU (x) < l(x) for some x
in X. This is a contradiction. B
Duality and Its Applications 187

10.6.4. Definition. The sublinear functional sU , constructed in 10.6.3, is the


supporting function of U . The term “support function” is also in current usage.
10.6.5. Kreı̆n–Milman 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.
C Let U be such subset of a space X. It may be assumed that the space
X is real and U 6= ∅. By virtue of 9.4.12, U is compact with respect to the
#
topology σ(X, X 0 ). Since σ(X, X 0 ) is induced in X by the topology σ(X 0 , X 0 )
#
on X 0 ; therefore, U = ∂(sU ). Here (cf. 10.6.3) sU : X 0 → R acts by the rule
sU (x0 ) := sup x0 (U ). By the Kreı̆n–Milman Theorem in subdifferential 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 subdifferential by Theorem 10.6.3. Moreover, this set has sU as its
supporting function, thus coinciding with U (cf. 3.6.6). B
10.6.6. Let X ↔ Y and let S be a conical segment in X. Assume further that
pS is the Minkowski functional of S. The polar π(S) is the inverse image of the
(algebraic) subdifferential ∂(pS ) under the ket-mapping; i.e.,
π(S) = | ∂(pS ) i−1
R .
If S is absolutely convex, then the absolute polar S ◦ is the inverse image of the
(algebraic) balanced subdifferential |∂|(pS ) under the ket-mapping; i.e.,
S ◦ = | |∂|(pS ) i−1 .
C If y ∈ YR and y ∈ | ∂(pS ) i−1
R then | yiR belongs to ∂(pS ). Hence, Rehx | yi =
hx | yiR = | yiR (x) ≤ pS (x) ≤ 1 for x ∈ S, because S ⊂ {pS ≤ 1} by the Gauge
Theorem. Consequently, y ∈ π(S).
If, in turn, y ∈ π(S) then | yiR belongs to ∂(pS ). Indeed, 1 > pS (α−1 x) for
all x in XR and α > pS (x); i.e., α−1 x ∈ {pS < 1} ⊂ S. Whence hα−1 x | yiR =
Rehα−1 x | yi = α−1 Rehx | yi ≤ 1. Finally, observe that | yiR (x) ≤ α. Since α
is arbitrary, this inequality means that | yiR (x) ≤ pS (x). In other words, y ∈
| ∂(ps )i−1 −1
R , which implies that π(S) = | ∂(pS )iR . The remaining claim follows
from the properties of the complexifier (cf. 3.7.3 and 3.7.9). B
10.6.7. Alaoglu–Bourbaki Theorem. The polar of a neighborhood of zero
of each compatible topology is a weakly compact convex set.
C Let U be a neighborhood of zero in a space X and let π(U ) be the polar of U
(with respect to X ↔ X 0 ). Since U ⊃ {p ≤ 1} for some continuous seminorm p,
by 10.5.2 (4), π(U ) ⊂ π({p ≤ 1}) = π(Bp ) = Bp◦ . Using 10.6.6 and recalling that
p is the Minkowski functional of Bp , obtain the inclusion π(U ) ⊂ |∂|(p). By virtue
of 10.6.2 the topological balanced subdifferential |∂|(p) is σ(X 0 , X)-compact. By def-
inition π(U ) is weakly closed. To infer the σ(X 0 , X)-compactness of π(U ), it re-
mains to appeal to 9.4.9. The convexity property of π(U ) is beyond a doubt. B
188 Chapter 10

10.7. Reflexive Spaces


10.7.1. Kakutani Criterion. A normed space is reflexive if and only if its
unit ball is weakly compact.
C ⇒: Let X be reflexive, i.e. 00 (X) = X 00 . In other words, the image of X
under the double prime mapping coincides with X 00 . Since the ball BX 00 is the
polar of the ball BX 0 with respect to X 00 ↔ X 0 ; therefore, BX 00 is a σ(X 00 , X 0 )-
compact set by the Alaoglu–Bourbaki Theorem. It remains to observe that BX 00 is
(the image under the double prime mapping of) BX , and σ(X, X 0 ) is (the inverse
image under the double prime mapping of) σ(X 00 , X 0 ).
⇐: Consider the duality pair X 00 ↔ X 0 . By definition, the ball BX 00 presents
the bipolar of BX (more precisely, the bipolar of (BX )00 ). Using the Absolute
Bipolar Theorem and observing that the weak topology σ(X, X 0 ) is induced in X
by the topology σ(X 00 , X 0 ), conclude that BX 00 = BX (because of the obvious
convexity and closure properties of BX , the latter following from compactness since
X is separated). Thus, X is reflexive. B
10.7.2. Corollary. A space X is reflexive if and only if every bounded closed
convex set in X is weakly compact. CB
10.7.3. Corollary. Every closed subspace of a reflexive space is reflexive.
C By the Mazur Theorem, such subspace and, hence, the unit ball of it are
weakly closed. It thus suffices to apply the Kakutani Criterion twice. B
10.7.4. Pettis Theorem. A Banach space and its dual are (or are not) re-
flexive simultaneously.
C If X is reflexive then σ(X 0 , X) coincides with σ(X 0 , X 00 ). Therefore, by the
Alaoglu–Bourbaki Theorem, BX 0 is σ(X 0 , X 00 )-compact. Consequently, X 0 is re-
flexive. If, in turn, X 0 is reflexive, then so is X 00 by what was proven. However, X,
as a Banach space, is a closed subspace of X 00 . Thus, X is reflexive by 10.7.3. B
10.7.5. James Theorem. A Banach space is reflexive 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-
dorff compact space, denoting by C(Q, R) the set of continuous real-valued func-
tions on Q. Unless specified otherwise, C(Q, R) is furnished with the natural point-
wise algebraic operations and order and equipped with the sup-norm k · k := k · k∞
related to the Chebyshëv 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 specified deliberately.
Duality and Its Applications 189

10.8.2. Definition. A subset L of C(Q, R) is a sublattice (in C(Q, R)) if


f1 ∨ f2 ∈ L and f1 ∧ f2 ∈ L for f1 , f2 ∈ L, where, as usual,

f1 ∨ f2 (q) := f1 (q) ∨ f2 (q),


f1 ∧ f2 (q) := f1 (q) ∧ f2 (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 Q0 be a subset of Q. Put

LQ0 := {f ∈ C(Q, R) : (∃ g ∈ L) g(q) = f (q) (q ∈ Q0 )}.

Then LQ0 is a sublattice. Moreover, L ⊂ LQ0 .


(4) Let Q0 be a compact subset of Q. Given a sublattice L in C(Q, R),
put 
L Q := f Q : f ∈ L .
0 0

Therefore, 
LQ0 = f ∈ C(Q, R) : f Q0 ∈ L Q0 .

It is clear that L Q0 is a sublattice of C(Q0 , 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 is a vector sublattice of C(Q0 , R) (certainly, if Q0 6= ∅).
0

(5) Let Q := {1, 2}. Then C(Q, R) ' R2 . Each nonzero vector sublat-
2
tice of R is given as
{(x1 , x2 ) ∈ R2 : α1 x1 = α2 x2 }
for some α1 , α2 ∈ R+ .
(6) Let L be a vector sublattice of C(Q, R). For q ∈ Q, the alternative
is offered: either L{q} = C(Q, R) or L{q} = {f ∈ C(Q, R) : f (q) = 0}. If q1 and
q2 are distinct points of Q and L {q1 ,q2 } 6= 0, then by 10.8.4 (5) there are some α1 ,
α2 ∈ R+ such that

L{q1 ,q2 } = {f ∈ C(Q, R) : α1 f (q1 ) = α2 f (q2 )}.

Moreover, if L contains a constant function other than zero (i.e. a nonzero multiple
of the constantly-one function 1) then as α1 and α2 in the above presentation
of L{q1 ,q2 } the unity, 1, may be taken. CB
190 Chapter 10

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 ε > 0 and (x, y) ∈ Q2 there is a function
f := fx,y,ε in L satisfying the conditions

f (x) − f (x) < ε, f (y) − f (y) > −ε.

C ⇒: This is obvious.
⇐: Basing it on 3.2.10 and 3.2.11, assume that f = 0. Take ε > 0. Fix x ∈ Q
and consider the function gy := fx,y,ε ∈ L. Let Vy := {q ∈ Q : gy (q) > −ε}. Then
Vy is an open set and y ∈ Vy . By a standard compactness argument, there are
y1 , . . . , yn ∈ Q such that Q = Vy1 ∪ . . . ∪ Vyn . Put fx := gy1 ∨ . . . ∨ gyn . It is
clear that fx ∈ L. Furthermore, fx (x) < ε and fx (y) > −ε for all y ∈ Q. Now let
Ux := {q ∈ Q : fx (q) < ε}. Then Ux is open and x ∈ Ux . Using the compactness
of Q once again, find x1 , . . . , xm ∈ Q such that Q = Ux1 ∪ . . . ∪ Uxm . Finally, put
l := fx1 ∧ . . . ∧ fxm . It is beyond a doubt that l ∈ L and klk < ε. B
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 = cl L{q1 ,q2 } .
(q1 ,q2 )∈Q2

C The inclusion of cl L into cl L{q1 ,q2 } for all (q1 , q2 ) ∈ Q2 raises no doubts.
If f ∈ cl L{q1 ,q2 } for all such q1 and q2 , then by 10.8.5, f ∈ cl L. B
10.8.8. Corollary. Every vector sublattice L of C(Q, R) is expressible as
\
cl L = L{q1 ,q2 } .
(q1 ,q2 )∈Q2

C Observe that every set of the form L{q1 ,q2 } is closed. B


10.8.9. Definition. A subset U of F Q separates the points of Q, if for all q1 ,
q2 ∈ Q such that q1 6= q2 there is a function u ∈ U assuming different values at q1
and q2 , i.e. u(q1 ) 6= u(q2 ).
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).
C Given such sublattice L, observe that

L{q1 ,q2 } = C(Q, R){q1 ,q2 }

for every pair (q1 , q2 ) in Q2 (cf. 10.8.4 (6)). It remains to appeal to 10.8.8. B
Duality and Its Applications 191

10.8.11. Let µ ∈ C(Q, R)0 . Put

N (µ) := {f ∈ C(Q, R) : [0, |f |] ⊂ ker µ}.

Then there is a unique closed subset supp(µ) of Q such that

f ∈ N (µ) ⇔ f supp(µ) = 0.

C By the Interval Addition Lemma

[0, |f |] + [0, |g|] = [0, |f | + |g|].

Thus, f, g ∈ N (µ) ⇒ |f | + |g| ∈ N (µ). Since N (µ) is an order ideal; i.e.,


(f ∈ N (µ) & 0 ≤ |g| ≤ |f | ⇒ g ∈ N (µ)), conclude that N (µ) is a linear set.
Moreover, N (µ) is closed. Indeed, assuming fn ≥ 0, fn → f and fn ∈ N (µ), for
g ∈ [0, f ] find g ∧ fn → g and g ∧ fn ∈ [0, fn ]. Whence it follows that µ(g) = 0;
i.e., f ∈ N (µ).
Since N (µ) is an order ideal, from 10.8.8 deduce that
\
N (µ) = N (µ){q} .
q∈Q

Define the set supp(µ) as

q ∈ supp(µ) ⇔ N (µ){q} 6= C(Q, R) ⇔ (f ∈ N (µ) ⇒ f (q) = 0).

It is beyond a doubt that supp(µ) is closed. Moreover, the equalities hold:


\
N (µ) = N (µ){q} = f ∈ C(Q, R) : f supp(µ) = 0 .


q∈supp(µ)

The claim of uniqueness follows from the normality of Q (cf. 9.4.14) and the Urysohn
Theorem. B
10.8.12. Definition. The set supp(µ) under discussion in 10.8.11 is the sup-
port of µ (cf. 10.9.4 (5)).
10.8.13. Remark. If µ is positive then

N (µ) = {f ∈ C(Q, R) : µ(|f |) = 0}.



Consequently, when µ(f g) = 0 for all g ∈ C(Q, R), observe that f supp(µ) = 0.
By analogy supp(µ) = ∅ ⇔ N (µ) = C(Q, R) ⇔ µ = 0. Therefore, it is quite
convenient to work with the support of a positive functional.
192 Chapter 10

Let F be a closed subset of Q. It is said that F supports or carries µ or that


X \ F lacks µ or is void of µ if µ(|f |) = 0 for every continuous function f with
supp(f ) ⊂ Q\F . The support supp(µ) of µ carries µ; moreover, supp(µ) is included
in every closed subset of Q supporting µ. In other words, the support of µ is the
complement of the greatest open set void of µ (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 µ there correspond some positive (and hence bounded) func-
tionals µ+ , µ− , and |µ| defined as

µ+ (f ) = sup µ[0, f ]; µ− (f ) = − inf µ[0, f ]; |µ| = µ+ + µ− ,

given f ∈ C(Q, R)+ .


Moreover, C(Q, R)0 is a Kantorovich space (cf. 3.2.16). CB
10.8.14. The supports of µ and |µ| coincide.
C By definition N (µ) = N (|µ|). B
10.8.15. Considering a ∈ C(Q, R) with 0 ≤ a ≤ 1, define aµ : f 7→ µ(af ) for
f ∈ C(Q, R) and µ ∈ C(Q, R)0 . Then |aµ| = a|µ|.
C Given f ∈ C(Q, R)+ , infer that

(aµ)+ (f ) = sup{µ(ag) : 0 ≤ g ≤ f } ≤ sup µ[0, af ]


= µ+ (af ) = aµ+ (f ).

Furthermore,

µ+ = (aµ + (1 − a)µ)+ ≤ (aµ)+ + ((1 − a)µ)+ ≤ aµ+ + (1 − a)µ+ = µ+ .

Consequently, (aµ)+ = aµ+ , whence the claim follows. B


10.8.16. De Branges Lemma. Let A be a subalgebra of C(Q, R) containing
constant functions. Take µ ∈ ext(A⊥ ∩ BC(Q, R)0 ). Then the restriction of each
member of A to the support of µ is a constant function.
C If µ = 0 then supp(µ) = ∅, and there is nothing to be proven. If µ 6= 0
then, certainly, kµk = 1. Take a ∈ A. Since the subalgebra A contains constant
functions, it suffices to settle the case in which 0 ≤ a ≤ 1 and

q ∈ supp(µ) ⇒ 0 < a(q) < 1.

Put µ1 := aµ and µ2 := (1 − a)µ. It is clear that µ1 + µ2 = µ and the functionals


µ1 and µ2 are both nonzero. Moreover,

kµk ≤ kµ1 k + kµ2 k


Duality and Its Applications 193

= sup µ(af ) + sup µ((1 − a)g) = sup µ(af + (1 − a)g) ≤ kµk,


kf k≤1 kgk≤1 kf k≤1,kgk≤1

because it is obvious that

aBC(Q, R) + (1 − a)BC(Q, R) ⊂ BC(Q, R) .

Thus, kµk = kµ1 k + kµ2 k. Since


µ1 µ2
µ = kµ1 k + kµ2 k ,
kµ1 k kµ2 k

and µ1 , µ2 ∈ A⊥ , conclude that µ1 = kµ1 kµ. By 10.8.15, a|µ| = |aµ| = |µ1 | =


kµ1 k |µ|. Consequently, |µ|((a − kµ1 k1)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 µ. B
10.8.17. Stone–Weierstrass 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).
C Proceeding by way of contradiction, assume the contrary. By the Absolute
Bipolar Theorem, the subspace A⊥ (coincident with A◦ ) of C(Q, R)0 is nonzero.
Using the Alaoglu–Bourbaki Theorem, observe that A⊥ ∩ BC(Q,R)0 is a nonempty
absolutely convex weakly compact set. Thus, by the Kreı̆n–Milman Theorem, the
set has an extreme point, say, µ.
Undoubtedly, µ is a nonzero functional. By the de Branges Lemma the support
of µ fails to contain two distinct points, since A separates the points of Q. The
support of µ is not a singleton, since µ annihilates constant functions. Thus, supp(µ)
is empty. But then µ is zero (cf. 10.8.13). We arrive at a contradiction completing
the proof. B
10.8.18. Corollary. The closure of a subalgebra of C(Q, R) is a vector sub-
lattice of C(Q, R).
C Using the Stone–Weierstrass Theorem, find a polynomial pn satisfying

|pn (t) − |t| | ≤ 1 /2n

for all t ∈ [−1, 1]. Then |pn (0)| ≤ 1 /2n . Therefore, the polynomial

pn (t) := pn (t) − pn (0)

maintains the inequality |pn (t) − |t| | ≤ 1 /n when −1 ≤ t ≤ 1. By construction, pn


lacks the constant term. Now, if a function a lies in a subalgebra A of C(Q, R)
and kak ≤ 1, then
|pn (a(q)) − |a(q)| | ≤ 1 /n (q ∈ Q).
Moreover, the function q 7→ pn (a(q)) is clearly a member of A. B
194 Chapter 10

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 7→ |t| on the interval [−1, 1]. It takes no
pains to demonstrate 10.8.17, with 10.8.18 available.
10.8.20. Tietze–Urysohn Theorem. Let Q0 be a compact subset of a com-
set Q and f0 ∈ C(Q0 , R). Then there is a function f in C(Q, R) such that
pact
f Q0 = f0 .
C Suppose that Q0 6= ∅ (otherwise, there is nothing to prove). Consider the

identical embedding ι : Q0 → Q and the bounded linear operator ι : C(Q, R) →
◦ ◦
C(Q0 , R) acting by the rule ιf := f ◦ ι. We have to show that ι is an epimorphism.

It is beyond a doubt that im ι is a subalgebra of C(Q0 , R) separating the points
of Q0 and containing constant functions. In virtue of 10.8.17 it is thus sufficient

(and, clearly, necessary) to examine the closure of im ι.

Consider the monoquotient ι of the operator ι and the coset mapping ϕ :
C(Q, R)0 = 0 O ι. Given f ∈ C(Q, R), put

g := (f ∧ sup |f (Q0 )|1) ∨ (− sup |f (Q0 )|1).



By definition f Q0 = g Q0 ; i.e., f := ϕ(f ) = ϕ(g). Consequently, kgk ≥ kf k.
Furthermore,
 ◦ 
kf k = inf khkC(Q,R) : ι(h − f ) = 0 = inf khkC(Q,R) : h Q = f Q
0 0

≥ inf kh Q0 kC(Q,R) : h Q0 = f Q0 = sup |f (Q0 )| = kgk ≥ kf k.

Therefore,
◦ ◦
kιf k = kιgk = kιgkC(Q0 ,R)
= kg ◦ ιkC(Q0 ,R) = sup |g(Q0 )| = kgk = kf k;

i.e., ι is an isometry. Successively applying 5.5.4 and 4.5.15, infer first that coim ι is
a Banach space and second that im ι is closed in C(Q0 , R). It suffices to observe

that im ι = im ι. B

10.9. Radon Measures


10.9.1. Definition. Let Š be a locally compact topological space. Put K(Š)
:= K(Š, F ) := {f ∈ C(Š, F) : supp(f ) is compact}. If Q is compact in Š then
let K(Q) := KŠ (Q) := {f ∈ K(Š) : supp(f ) ⊂ Q}. The space K(Q) is furnished
with the norm k · k∞ . Given E ∈ Op (Š), put K(E) := ∪ {K(Q) : Q b E}. (The
notation Q b E for a subset E of Š means that Q is compact and Q lies in the
interior of E relative to Š.)
Duality and Its Applications 195

10.9.2. The following statements are valid:


(1) if Q b Š 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, Q1 , and Q2 be compact sets and Q b Q1 ×Q2 ; the linear span
in the space C(Q, F ) of the restrictions to Q of the functions like u1 · u2 (q1 , q2 ) :=
u1 ⊗ u2 (q1 , q2 ) := u1 (q1 )u2 (q2 ) with us ∈ K(Qs ) is dense in C(Q, F );
(3) if Š is compact then K(Š) = C(Š, F); if Š fails to be compact and
is embedded naturally into C(Š· , F), with Š· := Š ∪ {∞} the Alexandroff com-
pactification of Š, then the space K(Š) is dense in the hyperplane {f ∈ C(Š· , F ) :
f (∞) = 0};
(4) the mapping E ∈ Op (Š) 7→ K(E) ∈ Lat (K(Š)) preserves suprema;
(5) for E 0 , E 00 ∈ Op (Š) the following sequence is exact:
ι σ
00 = 0 O (E 0 ∩ E 00 ) −−−−−→ K(E 0 ) × K(E 00 ) −−−−−→ K(E 0 ∪ E 00 )0 = 0 O ,
0 00
(E ,E ) (E ,E ) 0 00

with ι(E 0 ,E 00 ) f := (f, −f ), and σ(E 0 ,E 00 ) (f, g) := f + g.


C (1): The boundary ∂Q of Q is at the same time the boundary of the exterior
of Q, the set int(Š \ 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(Š) is an order ideal separating the points of Š· (cf. 10.8.11).
(4): Clearly, K(sup ∅) = K(∅) = 0. If E ⊂ Op (Š) and E is filtered 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 E1 , . . . , En ∈
Op (Š) and f ∈ K(E1P∪ . . . ∪ En ). In accordance P with 9.4.18 there are some
n n
ψk ∈ K(Ek ) such that k=1 ψk = 1. Moreover, f = k=1 ψk f and supp(f ψk ) ⊂
Ek (k := 1, . . . , n).
(5): This is straightforward from (4). B
10.9.3. Definition. A functional µ, a member of K(Š, F)# , is a measure
(or,
amply, a Radon F-measure) on Š, in symbols, µ ∈ M (Š) := M (Š, F); if
0
µ K(Q) ∈ K(Q) whenever Q b Š. The following notation is current:

Z Z Z
f dµ := f dµ := f (x) dµ(x) := µ(f ) (f ∈ K(Š)).
Š

The scalar µ(f ) is the integral of f with respect to µ. In this connection µ is also
called an integral.
196 Chapter 10

10.9.4. Examples.
(1) For q ∈ Š the Dirac measure f 7→ f (q) (f ∈ K(Š)) presents
a Radon measure. It is usually denoted by the symbol δq and called the delta-
function at q.
Suppose also that Š is furnished with group structure so that the taking
of an inverse q ∈ Š 7→ q −1 ∈ Š and the group operation (s, t) ∈ Š × Š 7→ st ∈ Š
are continuous; i.e., Š is a locally compact group. By δ we denote δe where e is the
unity of Š (recall the concurrent terms: “identity,” “unit,” and “neutral element,”
all meaning the same: es = se = s for all s ∈ Š). The nomenclature pertinent
to addition is routinely involved in abelian (commutative) groups.
Given a ∈ Š, acknowledge the operation of (left or right) translation or shift
by a in K(Š) (in fact, every function is shifted in Š × F):

(a τ f )(q) := a f (q) := f (a−1 q), (τa f )(q) := fa (q) := f (qa−1 )

with f ∈ K(Š) and q ∈ Š. Clearly, a τ , τa ∈ L (K(Š)).


A propitious circumstance of paramount importance is the presence of a non-
trivial measure on Š, a member of M (Š, 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 RN is Haar measure on the abelian group RN . 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
Z Z
−1
f (a x) dx = f (x) dx (f ∈ K(Š), a ∈ Š).
Š Š

(2) Let M (Š) := (K(Š), k · k∞ )0 . A member of M (Š) is a bounded


Radon measure. It is clear that a bounded measure belongs to the space C(Š· , F )0
(cf. 10.9.2 (2)).
(3) Given µ ∈ M (Š), put µ∗ (f ) = µ(f ∗ )∗ , where f ∗ (q) := f (q)∗ for
q ∈ Š and f ∈ K(Š). The measure µ∗ is the conjugate of µ. Distinction be-
tween µ∗ and µ is perceptible only if F = C. In case µ = µ∗ , we speak of a real
C-measure. It is clear that µ = µ1 + iµ2 , where µ1 and µ2 are uniquely-determined
real C-measures. In turn, each real C-measure is generates by two R-measures (also
called real measures), members of M (Š, R), because K(Š, C) coincides with the
complexification K(Š, R) ⊕ iK(Š, R) of K(Š, R). The real R-measures obvi-
ously constitute a Kantorovich space. Moreover, the integral with respect to such
Duality and Its Applications 197

measure serves as (pre)integral. So, an opportunity is offered 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-
ification.
To every Radon measure µ we assign the positive measure |µ| that is defined
for f ∈ K(Š, R), f ≥ 0, by the rule

|µ|(f ) := sup{|µ(g)| : g ∈ K(Š, 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 µ and ν are measures and |µ| ∧ |ν| = 0 then µ and ν are called disjoint or
independent (of one another). A measure ν is absolutely continuous with respect
to µ on condition that ν is independent of every measure independent of µ. Such
a measure ν may be given as ν = f µ, where f ∈ L1,loc (µ) and the measure f µ
having density f with respect to µ acts by the rule (f µ)(g) := µ(f g) (g ∈ K(Š))
(this is the Radon–Nikodým Theorem).
(4) Given Š0 ∈ Op (Š) and µ ∈ M (Š), consider the restriction µŠ0 :=
µ K(Š0 ) of µ to K(Š0 ). The restriction operator µ 7→ µŠ0 from M (Š) to M (Š0 ),

viewed as depending on a subset of Š, meets the agreement condition: if Š00 ⊂


Š0 ⊂ Š and µ ∈ M (Š) then µŠ00 = (µŠ0 )Š00 . This situation is verbalized as
follows: “The mapping M : E ∈ Op (Š) 7→ M (E) and the restriction operator
(referred to jointly as the functor M ) defines a presheaf (of vector spaces over Š).”
It stands to reason to convince oneself that the (values of the) restriction operator
need not be an epimorphism.
(5) Let E ∈ Op (Š) and µ ∈ M (Š). It is said that E lacks µ or is
void of µ or that Š \ E supports or carries µ if µE = 0. By 10.9.2 (4) there is
a least closed set supp(µ) supporting µ, the support of µ. It may be shown that
supp(µ) = supp(|µ|). The above definition agrees with 10.8.12. The Dirac measure
δq is a unique Radon measure supported at {q} to within scaling.
(6) Let Šk be a locally compact space and µk ∈ M (Šk ) (k := 1, 2).
There is a unique measure µ on the product Š1 × Š2 such that
Z Z Z
u1 (x)u2 (y) dµ(x, y) = u1 (x) dµ1 (x) u2 (y) dµ2 (y)
Š1 ׊2 Š1 Š2

with uk ∈ K(Šk ). The next designations are popular: µ1 × µ2 := µ1 ⊗ µ2 := µ.


Using 10.9.2 (4), infer that for f ∈ K(Š1 × Š2 ) the value µ1 × µ2 (f ) may be
calculated by repeated integration (this is the Fubini Theorem).
198 Chapter 10

(7) Let G be a locally compact group, and µ, ν ∈ M (G). For f ∈ K(G)


the function f_(s, t) := f (st) is continuous and |(µ × ν)(f_)| ≤ kµk kνk kf k∞ . This
defines the Radon measure µ ∗ ν(f ) := (µ × ν)(f_) (f ∈ K(G)), the convolution of µ
and ν. Using vector integrals, obtain the presentations:
Z Z Z
µ∗ν = δs ∗ δt dµ(s)dν(t) = δs ∗ ν dµ(s) = µ ∗ δt dν(t).
G×G G G

The space M (G) of bounded measures with convolution as multiplication, ex-


emplifies a Banach convolution algebra.
This algebra is commutative if and only if G is an abelian group. In that event
the space L1 (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 (L1 (G), ∗) is the group algebra of G. Thus, for f, g ∈ L1 (G), the defini-
tions of convolution for functions and measures agree with one another (cf. 9.6.17):
(f ∗ g)dm = f dm ∗ gdm. By analogy, the convolution of µ ∈ M (G) and f ∈ L1 (G)
is defined as (µ ∗ f )dm := µ ∗ (f dm); i.e., as the density of the convolution of µ
and f dm with respect to Haar measure dm. In particular,
Z Z
f ∗ g = (δx ∗ g)f (x) dm(x) = τx (g)f (x) dm(x).
G G

Wendel Theorem. Let T ∈ B(L1 (G)). Then the following statements are
equivalent:
(i) there is a measure µ ∈ M (G) such that T f = µ ∗ f for f ∈ L1 (G);
(ii) T commutes with translations: T τa = τa T for a ∈ G, where τa is
a unique bounded extension to L1 (G) of translation by the element
a in K(G);
(iii) T (f ∗ g) = (T f ) ∗ g for f, g ∈ L1 (G);
(iv) T (f ∗ ν) = (T f ) ∗ ν for ν ∈ M (G) and f ∈ L1 (G).
10.9.5. Definition. The spaces K(Š) and M (Š) are set in duality (induced
by the duality bracket K(Š) ↔ K(Š)# ). In this case, the space M (Š) is fur-
nished with the topology σ(M (Š), K(Š)), usually called vague. The space K(Š)
is furnished with the Mackey topology τK(Š) := τ (K(Š), M (Š)) (thereby, in par-
ticular, (K(Š), τK(Š) )0 = M (Š)). The space of bounded measures M (Š) is usually
considered with the dual norm:
kµk := sup{|µ(f )| : kf k∞ ≤ 1, f ∈ K(Š)} (µ ∈ M (Š)).
10.9.6. The topology τK(Š) is strongest among all locally convex topologies
making the identical embedding of K(Q) to K(Š) continuous for every Q with
Q b Š (i.e., τK(Š) is the so-called inductive limit topology (cf. 9.2.15)).
Duality and Its Applications 199

C If τ is the inductive limit topology and µ ∈ (K(Š), τ )0 then by definition


µ ∈ M (Š), because µ ◦ ιK(Q) is continuous for Q b Š. In turn, for µ ∈ M (Š)
the set VQ := {f ∈ K(Q) : |µ(f )| ≤ 1} is a neighborhood of zero in K(Q). From
the definition of τ , infer that ∪ {VQ : Q b Š} = {f ∈ K(Š) : |µ(f )| ≤ 1} is
a neighborhood of zero in τ . Thus, µ ∈ (K(Š), τ )0 and τ is compatible with
duality. Therefore, τ ≤ τK(Š) .
On the other hand, if p is a seminorm in the mirror of the Mackey topology
then p is the supporting function of a subdifferential lying in M (Š). Consequently,
the restriction q := p ◦ ιK(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 ιK(Q) : K(Q) → (K(Š), τK(Š) ) is
continuous and τ ≥ τK(Š) by the definition of inductive limit topology. B
10.9.7. A subset A of K(RN ) is bounded (in the inductive limit topology),
if sup kAk∞ < +∞ and, moreover, the supports of the members of A lie in a com-
mon compact set.
C Suppose to the contrary that, for some Q such that Q b RN , the inclusion
A ⊂ K(Q) fails. In other words, for n ∈ N there are some qn ∈ RN and an ∈ A
satisfying an (qn ) 6= 0 and |qn | > n. Put B := {n|an (qn )|−1 δqn : n ∈ N} and observe
that this set of Radon measures is vaguely bounded and so the seminorm p(f ) :=
sup{|µ|(|f |) : µ ∈ B} is continuous. Moreover, p(an ) ≥ n|an (qn )|−1 δqn (|an |) = n,
which contradicts the boundedness property of A. B
10.9.8. Remark. Let (fn ) ⊂ K(RN ). The notation fn  0 symbolizes the
K
proposition (∃ Q b RN )(∀ n) supp(fn ) ⊂ Q & kfn k∞ → 0. From 10.9.7 it is
immediate that µ ∈ K(RN )# is a Radon measure if µ(fn ) → 0 whenever fn  0.
K
Observe also that the same holds for every locally compact Š countable at infinity
or σ-compact, i.e. for Š presenting the union of countably many compact sets.
10.9.9. Remark. There is a sequence (pn ) of real-valued positive polynomials
on R such that the measures pn dx converge vaguely to δ as n → +∞. Considering
products of the measures, arrive at some polynomials Pn on the RN , with (Pn dx)
converging vaguely to δ (here, as usual, dx := dx1 × . . . × dxN is the Lebesgue
measure on RN ). Now, take f ∈ K(RN ) 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 ∗Pn ), 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 RN (cf. 10.10.2 (4)).
10.9.10. Measure Localization Pronciple. Let E be an open cover of Š
and let (µE )E∈E be a family of Radon measures, µE ∈ M (E). Assume further
200 Chapter 10

that, for every pair (E 0 , E 00 ) of the members of E , the restrictions of µE 0 and µE 00


to E 0 ∩ E 00 coincide. Then there is a unique measure µ on Š whose restriction to E
equals µE for all E ∈ E .
C Using 10.9.2 (5), arrange the sequence
ι σ
X X
K(E 0 ∩ E 00 ) −→ K(E) −→ K(Š) → 0,
0 00 E∈E
{E ,E }
0 00
E ,E ∈E, E 0 6=E 00

where ι is generated by summation of the coordinate identical embeddings ι(E 0,E 00 )


and σ 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(Š) = ∪QbŠ K(Q) and in view
of 10.9.2 (4), it suffices to settle the case of a finite cover by checking exactness at the
second term. Thus, proceeding by induction, suppose that for every n-element cover
{E1 , . . . , En } (n ≥ 2) the following sequence is exact:
n
ιn σ
Y n
Kn −→ K(Ek ) −→ K(E1 ∪ . . . ∪ En ) → 0,
k=1

where ιn is the “restriction” of ι to Kn , the symbol σn stands for addition and


Y
Kn := K(Ek ∩ El ).
k<l
k,l∈{1,... ,n}

By hypothesis, im ιn = ker σn . If σn+1 (fe, fn+1 ) = 0 where fe:= (f1 , . . . , fn ), then


σn fe = −fn+1 and fn+1 ∈ K((E1 ∪ . . . ∪ En ) ∩ En+1 ). Since σn is an epimorphism
by 10.9.2 (5), there are some θk ∈ K(Ek ∩ En+1 ) such that σn θ = −fn+1 for
θ := (θ1 , . . . , θn ). Whence (fe − θ) ∈ ker σn , and by hypothesis there is a member κ
of Kn such that ιn κ = fe − θ. Clearly,
n
Y
Kn+1 = Kn × K(Ek ∩ En+1 )
k=1

(to within isomorphism), κ := (κ, θ1 , . . . , θn ) ∈ Kn+1 , and ιn+1 κ = (fe, fn+1 ).


Passing to the diagram prime (cf. 7.6.13), arrive at the exact sequence
σ0 ι0
Y Y
0 → M (Š) −→ M (E) −→ M (E 0 ∩ E 00 ).
E∈E {E 0 ,E 00 }
E 0,E 00 ∈E ,E 0 6=E 00

The proof is complete. B


Duality and Its Applications 201

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 Š 7→ M (Š) of Radon measures is a sheaf or, putting it more
categorically, the functor M is a sheaf (of vector spaces over Š, cf. 10.9.4 (4)).

10.10. The Spaces D() and D 0 ()


10.10.1. Definition. A compactly-supported smooth function f : RN → F is
a test function; in symbols, f ∈ D(RN ) := D(RN , F).
Given Q b RN and Š ∈
Op (R ), designate D(Q) := f ∈ D(R ) : supp(f ) ⊂ Q and D(Š) := ∪ {D(Q) :
N N

Q b Š}.
10.10.2. The following statements are valid:
(1) D(Q) = 0 ⇔ int Q = ∅;
(2) given Q b RN , put
X X
kf kn,Q := k∂ α f kC(Q) := sup |(∂ α1 . . . ∂ αn f )(Q)|
|α|≤n α∈(Z+)N
α1 +...+αN ≤n

for a function f smooth in a neighborhood of Q (as usual, Z+ := N ∪ {0}); the


multinorm MQ := {k · kn,Q : n ∈ N} makes D(Q) into a Fréchet space;
(3) the space of smooth functions C∞ (Š) := E (Š) on Š in Op (RN )
with the multinorm MŠ := {k · kn,Q : n ∈ N, Q b Š} is a Fréchet space; moreover,
D(Š) is dense in C∞ (Š);
(4) if Q1 b RN , Q2 b RM and Q b Q1 × Q2 ; then the linear span
in D(Q) of the restrictions to Q of the functions like f1 f2 (q1 , q2 ) := f1 ⊗f2 (q1 , q2 ) :=
f1 (q1 )f2 (q2 ), with qk ∈ Qk and fk ∈ D(Qk ), is dense in D(Q);
(5) the mapping E ∈ Op (Š) 7→ D(E) ∈ Lat (D(Š)) preserves suprema:

D(E 0 ∩ E 00 ) = D(E 0 ) ∩ D(E 00 ), D(E 0 ∪ E 00 ) = D(E 0 ) + D(E 00 );


D(∪E ) = L (∪ {D(E) : E ∈ E }) (E ⊂ Op (Š)).

Moreover, the next sequence is exact (cf. 10.9.2 (5)):


ι σ
00 = 0 O D(E 0 ∩ E 00 ) −−−−−→ D(E 0 ) × D(E 00 ) −−−−−→ D(E 0 ∪ E 00 )0 = 0 O .
0
(E ,E ) 00 (E ,E ) 0 00

C (1) and (2) are obvious.


(3): Choose a sequence (Qm )m∈N such that Qm b Š, Qm b Qm+1 and
∪m∈N Qm = Š. The multinorm {k · kn,Qm : n ∈ N, m ∈ N} is countable and
202 Chapter 10

equivalent to MŠ . A reference to 5.4.2 yields the metrizability of C∞ (Š). The


completeness of C∞ (Š) raises no doubts.
To show that D(Š) is dense in C∞ (Š), consider the truncator set Tr(Š) := {ψ ∈
D(Š) : 0 ≤ ψ ≤ 1}. Make Tr (Š) a direction on letting ψ1 ≤ ψ2 ⇔ supp(ψ1 ) ⊂
int{ψ2 = 1}. It is clear that, for f ∈ C∞ (Š), the net (ψf )ψ∈Tr (Š) approximates f
as is needed.
(4): Take q 0 ∈ RN and q 00 ∈ RM . Let a(q 0 , q 00 ) := a0 (q 0 )a00 (q 00 ), where a0
and a00 are mollifiers on RN and RM , respectively. Given f ∈ D(Q), n ∈ N, and
ε > 0, choose χ from the condition kf − f ∗ aχ kn,Q ≤ ε /2 . Using the equicontinuity
property of the family F := {∂ α f (q)τq (aχ ) : |α| ≤ n, q ∈ Q1 × Q2 }, find finite sets
0 and 00 , with 0 ⊂ Q1 and 00 ⊂ Q2 , so that the integral of each function in F
be approximated to within 1 /2 (N + 1)−n ε by a Riemann sum of it using the points
of 0 ׁ00 . This yields a function f , a member of the linear span of D(Q1 )×D(Q2 ),
which was sought; i.e., kf − f kn,Q ≤ ε.
(5): Check this as in 10.9.2 (4), on replacing 9.4.18 with 9.6.19 (2). B
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(Š, F )# , is a distribution
or a generalized function whenever u |D(Q) ∈ D 0 (Q) := D(Q)0 for all Q b Š. This is
expressed in writing as u ∈ D 0 (Š) := D 0 (Š, F ). Sometimes a reference is appended
to the nature of the ground field F.
The usual designations are as follows: hu, f i := hf | ui := u(f ). Often we use
the most suggestive and ubiquitous symbol
Z
f (x)u(x) dx := u(f ) (f ∈ D(Š)).

10.10.5. Examples.
(1) Let g ∈ L1,loc (RN ) be a locally integrable function. The mapping
Z
ug (f ) := f (x)g(x) dx (f ∈ D(Š))

determines a distribution. A distribution of this type is regular. To denote a regular


distribution ug , a more convenient symbol g is also employed. In this connection,
we write D(Š) ⊂ D 0 (Š) and ug = | gi.
(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.
Duality and Its Applications 203

(3) A distribution u is said to has order at most m, if to every Q such


that Q b RN there corresponds a number tQ satisfying

|u(f )| ≤ tQ kf km,Q (f ∈ D(Q)).

The notions of the order of a distribution and of a distribution of finite order are
understood in a matter-of-fact fashion. Evidently, it is false that every distribution
has finite order.
(4) Let α be a multi-index, α ∈ (Z+ )N ; and let u be a distribution,
u ∈ D 0 (Š). Given f ∈ D(Š), put (∂ α u)(f ) := (−1)|α| u(∂ α f ). The distribution
∂ α u is the derivative of u (of order α). We also speak of generalized differentiation,
of derivatives in the distribution sense, etc. and use the conventional symbols
of differential calculus.
A derivative (of nonzero order) of a Dirac measure is not a measure. At the
same time δ ∈ D 0 (R) is the derivative of the Heaviside function δ (−1) := H, where
H : R → R is the characteristic function of R+ . If a derivative of a (regular)
distribution u is some regular distribution ug , 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 0 (Š), put u∗ (f ) := u(f ∗ )∗ . The distribution u∗ is the
conjugate of u. The presence of the involution ∗ routinely justifies speaking of real
distributions and complex distributions (cf. 10.9.3 (3)).
(6) Let E ∈ Op (Š) and u ∈ D 0 (Š). For f ∈ D(E), the scalar u(f )
is easily determined. This gives rise to the distribution uE , a member of D 0 (E),
called the restriction of u to E. The functor D 0 is clearly a presheaf.
Given u ∈ D 0 (Š) and E ∈ Op (Š), say that E lacks or is void of u, if uE = 0.
By 10.10.4 (5), if the members of a family of open subsets of Š are void of u then
so is the union of the family. The complement (to RN ) of the greatest open set
void of u is the support of u, denoted by supp(u). Observe that supp(∂ α u) ⊂
supp(u). Moreover, a distribution with compact support (= compactly-supported
distribution) has finite order.
(7) Let u ∈ D 0 (Š) and f ∈ C∞ (Š). If g ∈ D(Š) then f g ∈ D(Š). Put
(f u)(g) := u(f g). The resulting distribution f u is the product of f and u. Consider
the truncator direction Tr (Š). If there is a limit limψ∈Tr (Š) u(f ψ) then say that u
applies to f . It is clear that a compactly-supported distribution u applies to every
function in C∞ (Š). Moreover, u ∈ E 0 (Š) := C∞ (Š)0 . In turn, each element u
of E 0 (Š) (cf. 10.10.2 (3)) uniquely determines a distribution with compact support,
which is implicit in the notation
u ∈ D 0 (Š).
α
If f ∈ C∞ (Š) and ∂ f supp(u) = 0 for all α, |α| ≤ m, where u is a compactly-
supported distribution of order at most m, then it is easy that u(f ) = 0. In con-
204 Chapter 10

sequence, only linear combinations of a Dirac measure and its derivatives are sup-
ported at a singleton. CB
(8) Let Š1 , Š2 ∈ Op (RN ) and uk ∈ D 0 (Šk ). There is a unique distri-
bution u on Š1 × Š2 such that u(f1 f2 ) = u1 (f1 )u2 (f2 ) for all fk ∈ D(Šk ). The
distribution is denoted by u1 × u2 or u1 ⊗ u2 . Using 10.10.2 (4), infer that for
f ∈ D(Š1 × Š2 ) the value u(f ) of u at f appears from successive application of u1
and u2 . Strictly speaking,

u(f ) = u2 (y ∈ Š2 7→ u1 (f ( · , y))) = u1 (x ∈ Š1 7→ u2 (f (x, · ))).

More suggestive designations prompt the Fubini Theorem:


ZZ
f (x, y)(u1 × u2 )(x, y) dxdy
Š1 ׊2

   
Z Z Z Z
=  f (x, y)u1 (x) dx u2 (y) dy =  f (x, y)u2 (y) dy  u1 (x) dx.
Š2 Š1 Š1 Š2

It is worth noting that

supp(u1 × u2 ) = supp(u1 ) × supp(u2 ).

+
(9) Let u, v ∈ D 0 (RN ). Given f ∈ D(RN ), put f := f ◦ +. It is
+
clear that f ∈ C∞ (RN × RN ). 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∞ (RN × RN ) for each f in D(RN ). It is easy (cf. 10.10.10)
+
that the resulting linear functional f 7→ (u × v)( f ) (f ∈ D(RN )) 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 RN (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 0 (RN ) of compactly-supported
distributions with convolution as multiplication presents an (associative and com-
mutative) algebra with unity the delta-function δ. Further, ∂ α u = ∂ α δ ∗ u and
∂ α (u ∗ v) = ∂ α u ∗ v = u ∗ ∂ α v. Moreover, the following remarkable equality, the
Lions Theorem of Supports, holds:

co (supp(u ∗ v)) = co (supp(u)) + co (supp(v)).


Duality and Its Applications 205

It is worth emphasizing that the pairwise convolutivity of distributions fails in gen-


eral to guarantee the associativity of convolution (for instance, (1 ∗ δ 0 ) ∗ δ (−1) = 0
whereas 1 ∗ (δ 0 ∗ δ (−1) ) = 1, with 1 := 1R ).
Each distribution u convolutes with a test function f , yielding some regular
distribution (u ∗ f )(x) = u(τx (f e )), where f e:= fe is the reflection of f ; i.e.,
f e(x) := f (−x) (x ∈ RN ). The operator u∗ : f 7→ u ∗ f from D(RN ) to C∞ (RN )
is continuous and commutes with translations: (u∗)τx = τx u∗ for x ∈ RN . It is
easily seen that the above properties are characteristic of u∗; i.e., if an operator T ,
a member of L (D(RN ), C∞ (RN )), is continuous and commutes with translations,
then there is a unique distribution u such that T = u∗; namely, u(f ) := (T 0 δ)(fe)
for f ∈ D(RN ) (cf. the Wendel Theorem).
10.10.6. Definition. The spaces D(Š) and D 0 (Š) are set in duality (in-
duced by the duality bracket D(Š) ↔ D(Š)# ). Moreover, D 0 (Š) is furnished
with the topology of the distribution space σ(D 0 (Š), D(Š)), and D(Š) is furnished
with the topology of the test function space, the Mackey topology τD := τD(Š) :=
τ (D(Š), D 0 (Š)).
10.10.7. Let Š ∈ Op (RN ). Then
(1) τD is the strongest of the locally convex topologies making the
identical embedding of D(Q) into D(Š) continuous for all Q b Š (i.e., τD is the
inductive limit topology);
(2) a subset A of D(Š) is bounded if and only if A lies in D(Q) for
some Q such that Q b Š and is bounded in D(Q);
(3) a sequence (fn ) converges to f in (D(Š), τD ) if and only if there
is a compact set Q such that Q b Š, supp(fn ) ⊂ Q, supp(f ) ⊂ Q and (∂ α fn )
converges to ∂ α f uniformly on Q for every multi-index α (in symbols, fn  f );
(4) an operator T , a member of L (D(Š), Y ) with Y a locally convex
space, is continuous if and only if T fn → 0 provided that fn  0;
(5) a delta-like sequence (bn ) serves as a (convolution) approximate
unity in D(RN ) as well as in D 0 (RN ); i.e., bn ∗ f  f (in D(RN )) and bn ∗ u → u
(in D 0 (RN )) for f ∈ D(RN ) and u ∈ D 0 (RN ).
C (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 Š as the union Š = ∪n∈N Qn , where Qn b Qn+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 b Š. By 10.10.2 (2) the space D(Q) is metrizable. It remains
to refer to 10.10.7 (3).
206 Chapter 10

(5): It is clear that all of the supports supp(bn ∗ f ) lie in some compact neigh-
borhood about supp(f ). Furthermore, for g ∈ C(RN ), it is evident that bn ∗ g → g
uniformly on compact subsets of RN . Applying this to ∂ α f and considering (3),
infer that bn ∗ f  f .
On account of 10.10.5 (9), observe the following:

u(fe) = (u ∗ f )(0) = lim(u ∗ (bn ∗ f ))(0)


n

= lim((u ∗ bn ) ∗ f )(0) = lim(bn ∗ u)(fe)


n n

for f ∈ D(R ). B
N

10.10.8. Remark. In view of 10.10.7 (3), for Š ∈ Op (RN ) and m ∈ Z+ it is of-


(m)
ten convenient to consider the space D (m) (Š) := C0 (Š) comprising all compactly-
supported functions f whose derivatives ∂ α f are continuous for all |α| ≤ m. The
space D (m) (Q) := {f ∈ D (m) (Š) : supp(f ) ⊂ Q} for Q b Š is furnished with the
norm k·km,Q making it into a Banach space. In that event, D (m) (Š) is endowed with
the inductive limit topology. Thus, D (0) (Š) = K(Š) and D(Š) = ∩m∈N D (m) (Š).
For a sequence (fn ) to converge in D (m) (Š) means to converge uniformly with all
derivatives up to order m on some Q such that Q b Š and supp(fn ) ⊂ Q for all suf-
ficiently large n. Note that D (m) (Š)0 comprises all distributions of order at most m.
Correspondingly, [
D 0F (Š) := D (m) (Š)0
m∈N
is the space of finite-order distributions.
10.10.9. Let Š ∈ Op (RN ). Then
(1) the space D(Š) is barreled; i.e., each barrel, a closed absorbing
absolutely convex subset, is a neighborhood of zero;
(2) every bounded closed set in D(Š) is compact; i.e., D(Š) is a Montel
space;
(3) every absolutely convex subset of D(Š), absorbing each bounded
set, is a neighborhood of zero; i.e., D(Š) is a bornological space;
(4) the test functions are dense in the distribution space.
C (1): Given a barrel V in D(Š), observe that VQ := V ∩ D(Q) is a barrel
in D(Q) for all Q b Š. Thus, VQ is a neighborhood of zero in D(Q) (cf. 7.1.8).
(2): Such a set lies in D(Q) for some Q b Š 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 b Š.
(4): Let g ∈ | D(Š)i◦ , with the polar takenR with respect to D(Š) ↔ D 0 (Š).
It is clear that uf (g) = 0 for f ∈ D(Š); i.e., g(x)f (x) dx = 0. Thus, g = 0.
It remains to refer to 10.5.9. B
Duality and Its Applications 207

10.10.10. Schwartz Theorem. Let (uk )k∈N be a sequence of distributions.


Assume that for every f in D(Š) there is a sum

X
u(f ) := uk (f ).
k=1

Then u is a distribution and



X
∂αu = ∂ α uk
k=1

for every multi-index α.


C The continuity of u is guaranteed by 10.10.9 (1). Furthermore, for f ∈ D(Š)
by definition (cf. 10.10.5 (4))
  ∞
X   ∞
X
α |α| α |α| α
∂ u(f ) = u (−1) ∂ f = uk (−1) ∂ f = ∂ α uk (f ). .
k=1 k=1

10.10.11. Theorem. The functor D 0 is a sheaf.


C It is immediate (cf. 10.9.10 and 10.9.11). B
10.10.12. Remark. The possibility of recovering a distribution from local
data, the Distribution Localization Principle stated in 10.10.11, admits clarification
in view of the paracompactness of RN . Namely, consider an open cover E of Š and
a distribution u ∈ D 0 (Š) with local data (uE )E∈E . Take a countable
P∞ (locally finite)
partition of unity (ψk )k∈N subordinate to E . Evidently, u = k=1 ψk uk , where
uk := uEk and supp(ψk uk ) ⊂ Ek (k ∈ N).
10.10.13. Theorem. Each distribution u on Š of order at most m may be
expressed as sum of derivatives of Radon measures:
X
u= ∂ α µα ,
|α|≤m

where µα ∈ M (Š).
C To begin with, assume that u has compact support. Let Q with Q b Š be
a compact neighborhood of supp(u). By hypothesis (cf. 10.10.5 (7) and 10.10.8)
X
|u(f )| ≤ t k∂ α f k∞ (f ∈ D(Q))
|α|≤m

for some t ≥ 0.
208 Chapter 10

Using 3.5.7 and 3.5.3, from 10.9.4 (2) obtain


X X
u=t να ◦ ∂ α = t (−1)|α| ∂ α να
|α|≤m |α|≤m

for a suitable family (να )|α|≤m , where να ∈ |∂|(k · k∞ ).


Passing to the general case and invoking the Countable Partition Theorem, find
a partition of unity (ψk )k∈N with ψk ∈ D(Š) such that some neighborhoods Qk
of supp(ψk ) compose a locally finite cover of Š (cf. 10.10.12). For each of the
distributions (ψk u)k∈N it is already proven that
X
ψk u = ∂ α µk,α ,
|α|≤m

where every µk,α is a Radon measure on Š and supp(µk,α ) ⊂ Qk .


From the Schwartz Theorem it is immediate that the sum

X
µα (f ) := µk,α (f )
k=1

exists for all f ∈ K(Š). Moreover, the resulting distribution µα is a Radon measure.
Once again appealing to 10.10.10, infer that
∞ ∞ ∞
!
X X X X X X
u= ψk u = ∂ α µk,α = ∂α µk,α = ∂ α µα ,
k=1 k=1 |α|≤m |α|≤m k=1 |α|≤m

which was required. B


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 clarification are avail-
able. For instance, it may be verified 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
differentiation of a conventional function.

10.11. The Fourier Transform of a Distribution


10.11.1. Let χ be a nonzero functional over the space L1 (RN ) := L1 (RN , C ).
The following statements are equivalent:
(1) χ is a character of the group algebra (L1 (RN ), ∗); i.e., χ 6= 0,
χ ∈ L1 (RN )0 and
χ(f ∗ g) = χ(f )χ(g) (f, g ∈ L1 (RN ))
(in symbols, χ ∈ X(L1 (RN )), cf. 11.6.4);
Duality and Its Applications 209

(2) there is a unique vector t in RN such that


Z
χ(f ) = fb(t) := (f ∗ et )(0) := f (x)ei(x,t) dx
RN

for all f ∈ L1 (RN ).


C (1) ⇒ (2): Suppose that χ(f )χ(g) 6= 0. If x ∈ RN then

χ(δx ∗ f ∗ g) = χ(δx ∗ f )χ(g) = χ(δx ∗ g)χ(f ).

Put ψ(x) := χ(f )−1 χ(δx ∗ f ). This soundly defines some continuous mapping ψ :
RN → C. Moreover,

ψ(x + y)
−1
= χ(f ∗ g) χ(δx+y ∗ (f ∗ g)) = χ(f )−1 χ(g)−1 χ(δx ∗ f ∗ δy ∗ g)
= χ(f )−1 χ(δx ∗ f )χ(g)−1 χ(δy ∗ g) = ψ(x)ψ(y)

for x, y ∈ RN ; i.e., ψ is a (unitary) group character: ψ ∈ X(RN ). Calculus shows


that ψ = et for some (obviously, unique) t ∈ RN . Further, by the properties of the
Bochner integral
 
Z
χ(f )χ(g) = χ(f ∗ g) = χ  (δx ∗ g)f (x) dx
RN
Z Z Z
= χ(δx ∗ g)f (x) dx = f (x)χ(g)ψ(x) dx = χ(g) f (x)ψ(x) dx.
RN RN RN

Thus, Z
χ(f ) = f (x)ψ(x) dx (f ∈ L1 (RN )).
RN

(2) ⇒ (1): Given t ∈ RN and treating f , g and f ∗ g as distributions, infer that

f[∗ g(t) = uf ∗g (et )


Z Z Z Z
= f (x)g(y)et (x + y) dxdy = f (x)et (x) dx g(y)et (y) dy
RN RN RN RN
210 Chapter 10

= uf (et )ug (et ) = fb(t)b


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(L1 (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 ψ : G → C satisfying

|ψ(x)| = 1, ψ(x + y) = ψ(x)ψ(y) (x, y ∈ G).

Endowed with pointwise multiplication, the set G b := X(G) becomes a commutative


group. In virtue of the Alaoglu–Bourbaki Theorem, X(L1 (G)) is locally compact
in the weak topology σ((L1 (G))0 , L1 (G)). So, G b may 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 defines the character bqb : qb ∈ G
b 7→ qb(q) ∈ C of the dual group G.
b The resulting
embedding of G into G is surprisingly an isomorphism of the locally compact abelian
bb

groups G and G (the Pontryagin–van Kampen Duality Theorem).


bb

10.11.3. Definition. For a function f in L1 (RN ), the mapping fb : RN → C,


defined by the rule
fb(t) := fb(t) := (f ∗ et )(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 oper-
N
ator F : L1 (RN ) → C R acting by the rule F f := fb but also for its modifications
(cf. 10.11.13). Second, the Fourier transform F is identified with each operator
Fθ f := fb ◦ θ, where θ is an automorphism (= isomorphism with itself) of RN .
Among the most popular are the functions: θ(x) := ∼ (x) := −x, θ(x) := 2π (x) := 2πx
and θ(x) := −2π (x) := −2πx (x ∈ RN ). In other words, the Fourier transform is
usually defined by one the following formulas:
Z
F∼ f (t) = f (x)e−i(x,t) dx,
RN
Z
F2π f (t) = f (x)e2πi(x,t) dx,
RN
Z
F−2π f (t) = f (x)e−2πi(x,t) dx.
RN
Duality and Its Applications 211

Since the character groups of isomorphic groups are also isomorphic, there are
grounds to using the same notation fb for generally distinct functions F f , F∼ f ,
and F±2π f . The choice of the symbol b for F2π (F−2π ) dictates the denotation ∨
for F−2π (F2π ) (cf. 10.11.12).

10.11.5. Examples.
(1) Let f : R0 = 0 O R be the characteristic function of the interval
[−1, 1]. Clearly, fb(t) = 2t−1 sin t. Observe that if kπ ≥ t0 > 0 then

Z Z ∞
X Z
|fb(t)| dt ≥ |fb(t)| dt = |fb(t)| dt
[t0 ,+∞) [kπ,+∞) n=k [nπ,(n+1)π]

∞ Z ∞
X 2| sin t| X 1
≥ dt = 4 = +∞.
(n + 1)π (n + 1)π
n=k [nπ,(n+1)π] n=k

Thus, fb ∈
/ L1 (R).
(2) For f ∈ L1 (RN ) the function fb is continuous, with the inequality
kfbk∞ ≤ kf k1 holding.
C The continuity of fb follows from the Lebesgue Dominated Convergence The-
orem; and the boundedness of fb, from the obvious estimate
Z
|fb(t)| ≤ |f (x)| dx = kf k1 (t ∈ RN ). .
RN

(3) If f ∈ L1 (RN ), then |fb(t)| → 0 as |t| → +∞ (= the Riemann–


Lebesgue Lemma).
/ The claim is obvious for compactly-supported step functions. It suffices
to refer to 5.5.9 (6) and the containment F ∈ B(L1 (RN ), l∞ (RN )). .
(4) Let f ∈ L1 (RN ), ε > 0, and fε (x) := f (εx) (x ∈ RN ). Then
fbε (t) = ε−N fb(t /ε ) (t ∈ RN ).
f (εx)et (x) dx = ε−N f (εx)et/ε (εx) dεx = ε−N fb(t /ε ) .
R R
/ fbε (t) =
RN RN

(5) F (f ∗ ) = (F∼ f )∗ , τd b d b N
x f = ex f , and ex f = τx f for all f ∈ L1 (R )
and x ∈ RN .
212 Chapter 10

C We will only check the first equality. Since a∗ b = (ab∗ )∗ for a, b ∈ C;


on using the properties of conjugation and integration, given t ∈ RN , infer that
 ∗
Z Z

F (f ∗ )(t) = f (x)ei(x,t) dx =  f (x) ei(x,t) dx


RN RN
 ∗
Z
= f (x)e−i(x,t) dx = (F∼ f )∗ (t). .
RN

(6) If f, g ∈ L1 (RN ) then


Z Z
∗ g = fbgb;
f[ fbg = f gb.
RN RN

C The first equality is straightforward from 10.11.1. The second, the multipli-
cation formula, follows on applying the Fubini Theorem:
Z Z Z
fg =
b f (x)et (x) dxg(t) dt
RN RN RN
 
Z Z Z
=  g(t)et (x) dt f (x) dx = f gb. .
RN RN RN

(7) If fb, f and g belong to L1 (RN ) then (fbg)b = f e∗ gb.


C Given x ∈ RN , observe that
Z Z Z
(fbg)b(x) = g(t)fb(t)et (x) dt = g(t)f (y)et (y)et (x) dydt
RN RN RN
Z Z
= f (y)g(t)et (x + y) dtdy
RN RN
Z Z
= f (y)b
g (x + y) dy = f (y − x)b
g (y) dy = f e∗ g(x). .
RN RN

(8) If f ∈ D(RN ) and α ∈ (Z+ )N then

F (∂ α f ) = i|α| tα F f, ∂ α (F f ) = i|α| F (xα f );


F2π (∂ α f ) = (2πi)|α| tα F2π f, ∂ α (F2π f ) = (2πi)|α| F2π (xα f )
Duality and Its Applications 213

(these equalities take the rather common liberty of designating xα := tα := (·)α :


αN
y ∈ RN 7→ y1α1 · . . . · yN ).
C It suffices (cf. 10.11.4) to examine only the first row. Since ∂ α et = i|α| tα et ;
therefore,

F (∂ α f )(t) = et ∗ ∂ α f (0)


= ∂ α et ∗ f (0) = i|α| tα (et ∗ f )(0) = i|α| tα fb(t).




By analogy, on differentiating under the integral sign, infer that


Z Z
∂ ∂
(F f )(t) = f (x)e i(x,t)
dx = f (x)ix1 ei(x,t) dx = F (ix1 f )(t). .
∂t1 ∂t1
RN RN

 N
(9) If fN (x) := exp −1 /2 |x|2 for x ∈ RN , then fbN = (2π) /2 fN .
C It is clear that
N Z
1
/2 |x|2k
Y
fbN (t) = eitk xk e− dxk (t ∈ RN ).
k=1 R

Consequently, the matter reduces to the case N = 1. Now, given y ∈ R, observe


that
Z Z
−1 /2 x2 ixy 1
/2 (x−iy)2 −1 /2 y 2
fb1 (y) = e e dx = e− dx
R R
Z
1
/2 (x−iy)2
= f1 (y) e− dx.
R

To calculate the integral A that we are interested in, consider (concurrently ori-
ented) straight lines λ1 and λ2 parallel to the real axis R in the complex plane
C R ' R2 . Applying the Cauchy Integral Theorem to the holomorphic function
2
f (z) := exp − z /2 (z ∈ C ) and a rectangular
R with vertices
R on λ1 and λ2 and
properly passing to the limit, conclude that λ1 f (z) dz = λ2 f (z) dz. Whence it fol-
lows that
Z
−1 /2 (x−iy)2
Z
1
/2 x 2

A= e dx = e− dx = 2π. .
R R
214 Chapter 10

10.11.6. Definition. The Schwartz space is the set of tempered (or rapidly
decreasing) functions, cf. 10.11.17 (2),

S (RN ) := f ∈ C∞ (RN ) : (∀α, β ∈ (Z+ )N ) |x| → +∞ ⇒ xα ∂ β f (x) → 0




with the multinorm {pα,β : α, β ∈ (Z+ )N }, where pα,β (f ) := kxα ∂ β f k∞ .


This space is treated as a subspace of the space of all functions from RN to C.

10.11.7. The following statements are valid:


(1) S (RN ) is a Fréchet space;
(2) the operators of multiplication by a polynomial and differentiation
are continuous endomorphisms of S (RN );
(3) the topology of S (RN ) may equivalently be given by the multinorm
{pn : n ∈ N}, where
X
pn (f ) := k(1 + | · |2 )n ∂ α f k∞ (f ∈ S (RN ))
|α|≤n

(as usual, |x| stands for the Euclidean norm of a vector x in RN );


(4) D(RN ) is dense in S (RN ); furthermore, the identical embedding
of D(R ) into S (RN ) is continuous and S (RN )0 ⊂ D 0 (RN );
N

(5) S (RN ) ⊂ L1 (RN ).


C We will check (4), because the other claims are easier.
Take f ∈ S (RN ) and let ψ be a truncator in D(RN ) such that B ⊂ {ψ = 1}.
Given x ∈ RN and ξ > 0, put ψξ (x) := ψ(ξx), and fξ = ψξ f. Evidently, fξ ∈ D(RN ).
Let ε > 0 and α, β ∈ (Z+ )N . It is an easy matter to show that sup{k∂ γ (ψξ −1)k∞ :
γ ≤ β, γ ∈ (Z+ )N } < +∞ for 0 < ξ ≤ 1. Considering that xα ∂ β f (x) → 0 as
|x| → +∞, find r > 1 satisfying |xα ∂ β ((ψξ (x) − 1)f (x))| < ε whenever |x| > r.
Moreover, fξ (x) − f (x) = (ψ(ξx) − 1)f (x) = 0 for |x| ≤ ξ −1 . Therefore,

pα,β (fξ − f ) = sup |xα ∂ β ((ψξ (x) − 1)f (x))|


|x|>ξ −1

≤ sup |xα ∂ β ((ψξ (x) − 1)f (x))| < ε


|x|>r

for ξ ≤ r−1 . Thus, pα,β (fξ − f ) → 0 as ξ → 0; i.e., fξ → f in S (RN ). The required


continuity of the identical embedding raises no doubt. B
Duality and Its Applications 215

10.11.8. The Fourier transform is a continuous endomorphism of S (RN ).


C Given f ∈ D(RN ), from 10.11.5 (8), 10.11.5 (2) and the Hölder inequality
obtain
ktα fbk∞ = k(∂ α f )bk∞ ≤ k∂ α f k1 ≤ Kk∂ α f k∞ .
Thus,
ktα ∂ β fbk∞ = ktα (xβ f )bk∞ ≤ K 0 k∂ α (xβ f )k∞ .

Whence it is easily seen that fb ∈ S (RN ) and the restriction of F to D(Q) with
Q b RN is continuous. It remains to refer to 10.10.7 (4) and 10.11.7 (4). B
10.11.9. Theorem. The repeated Fourier transform in the space S (RN ) is
proportional to the reflection.
C Let f ∈ S (RN ) and g(x) := fN (x) = exp −1 /2 |x|2 . From 10.11.8 and


10.11.7 derive that fb, f , g ∈ L1 (RN ) and so, by 10.11.5 (7), (fbg)b = f e∗ gb. Put
gε (x) := g(εx) for x ∈ RN and ε > 0. Then for the same x, in view of 10.11.5 (4)
find
Z Z Z
1 y
g(εt)f (t)et (x) dt = N
b f (y − x)b
g dy = f (εy − x)b
g (y) dy.
ε ε
RN RN RN

Using 10.11.5 (9) and the Lebesgue Dominated Convergence Theorem as ε → 0,


infer that
Z Z
g(0) fb(t)et (x) dt = f (−x) gb(y) dy
RN RN
Z
N 1
/2 |x|2
= (2π) /2
f (x) e− dx = (2π)N f (−x).
RN

Finally, F 2 f = (2π)N f e. B
10.11.10. Corollary. F2π2
is the reflection and (F2π )−1 = F−2π .
C Given f ∈ S (RN ) and t ∈ RN , deduce that
Z Z
N i(x,t)
f (−t) = (2π) e fb(x) dx = e2πi(x,t) fb(2πx) dx
RN RN

= (F2π (F2π f ))(t).


Since F2π f e= F−2π f , the proof is complete. B
216 Chapter 10

10.11.11. Corollary. S (RN ) is a convolution algebra (= an algebra with


convolution as multiplication).
C For f , g ∈ S (RN ) the product f g is an element of S (RN ) and so fbgb ∈
S (RN ). From 10.11.5 (6) infer that F2π (f ∗ g) ∈ S (RN ) and, consequently,
by 10.11.10, f ∗ g = F−2π (F2π (f ∗ g)) ∈ S (RN ). B
10.11.12. Inversion Theorem. The Fourier transform F := F2π is a topo-
logical automorphism of the Schwartz space S (RN ), with convolution carried into
pointwise multiplication. The inverse transform F−1 equals F−2π , with pointwise
multiplication carried into convolution. Moreover, the Parseval identity holds:
Z Z
fg = ∗
fbgb ∗ (f, g ∈ S (RN )).
RN RN

C 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 (fbg)(0) = (fe ∗
gb)(0). Using 10.11.5 (5), conclude that
Z
f g ∗ = (F(F−1 f )g ∗ )b(0) = ((F−1 f )e∗ Fg ∗ )(0)
RN
Z Z Z
∗ ∗
= Ff (Fg )edx = Ff F∼ (g ) dx = Ff (Fg)∗ . .
RN RN RN

10.11.13. Remark. In view of 10.11.9, the theorem on the repeated Fourier


transform, the following mutually inverse operators
Z
1
Ff (t) = f (x)ei(x,t) dx;
(2π)N /2
RN
Z
−1 1
F f (x) = f (t)e−i(x,t) dt
(2π)N /2
RN

are considered alongside F. In this case, an analog of 10.11.12 is valid on condition


N
that convolution is redefined as f ∗ g := (2π)− /2 f ∗g (f, g ∈ L1 (RN )). The merits
−1
of F and F are connected with some simplification of 10.11.5 (8). In the case of F,
a similar goal is achieved by introducing the differential operator Dα := (2πi)−|α| ∂ α
with α ∈ (Z+ )N .
Duality and Its Applications 217

10.11.14. Plancherel Theorem. The Fourier transform in the Schwartz


space S (RN ) is uniquely extendible to an isometric automorphism of L2 (RN ).
C Immediate from 10.11.12 and 4.5.10 since S (RN ) is dense in L2 (RN ). B
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 Fourier–Plancherel transform or the L2 -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 ∈ L2 (RN ) which are treated as the results of appropriate
passage to the limit in L2 (RN ).
10.11.16. Definition. Let u ∈ S 0 (RN ) := S (RN )0 . 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 0 (RN ) of tempered distributions
is furnished with the weak topology σ(S 0 (RN ), S (RN )) and is sometimes called
the Schwartz space (as well as S (RN )).
10.11.17. Examples.
(1) Lp (RN ) ⊂ S 0 (RN ) for 1 ≤ p ≤ +∞.
C Let f ∈ Lp (RN ), ψ ∈ S (RN ), p < +∞ and 1 /p0 + 1 /p = 1. Using Hölder
inequality, for suitable positive K, K 0, and K 00 successively infer that

 1 /p  1 /p
Z Z
|ψ|p  (1 + |x|2 )N (1 + |x|2 )−N ψ(x) p dx

kψkp0 ≤  +


B R N\ B
  1 /p
Z
dx
≤ K 0 kψk∞ + k(1 + | · |2 )N ψk∞  ≤ K 00 p1 (ψ).
 
(1 + |x|2 )N p

R N\ B

Once again using the Hölder inequality, observe that



Z

|uf (ψ)| = |hψ | f i| =
f ψ ≤ kf kp kψkp0 ≤ Kp1 (ψ).
N
R

The case p = +∞ raises no doubts. B


(2) S (RN ) is dense in S 0 (RN ).
C Follows from 10.11.7 (4), 10.11.17 (1), 10.11.7 (5), and 10.10.9 (4). B
218 Chapter 10

(3) Let µ ∈ M (RN ) be a tempered Radon measure; i.e.,


Z
d|µ|(x)
< +∞
(1 + |x|2 )n
RN

for some n ∈ N. Evidently, µ is a tempered distribution.


(4) If u ∈ S 0 (RN ), f ∈ S (RN ) and α ∈ (Z+ )N then f u ∈ S 0 (RN )
and ∂ α u ∈ S 0 (RN ) in virtue of 10.11.7 (2). By a similar argument, putting
Dα u(f ) := (−1)|α| uDα f for f ∈ S (RN ), infer that Dα u ∈ S 0 (RN ) and Dα u =
(2πi)−|α| ∂ α u.
(5) Every compactly-supported distribution is tempered.
C In accordance with 10.10.5 (7) such u in D 0 (RN ) may be identified with
a member of E 0 (RN ). Since the topology of S (RN ) is stronger than that induced
by the identical embedding in C∞ (RN ), conclude that u ∈ S 0 (RN ). B
(6) Let u ∈ S 0 (RN ). If f ∈ S (RN ) then u convolutes with f and
u ∗ f ∈ S (RN ). In may be shown that u also convolutes with every distribution v,
a member of E 0 (RN ), and u ∗ v ∈ S 0 (RN ).
(7) Take u ∈ D 0 (RN ) and x ∈ RN . Let τx u := (τ−x )0 u = u ◦ τ−x be
the corresponding translation of u. A distribution u is periodic (with period x) if
τx u = u. Every periodic distribution is tempered. Periodicity is preserved under
differentiation and convolution.
P∞(8) If un ∈ S (R ) (u f ∈ S (RN ) there is a sum
0 N
∈ N) and for every

u(f ) := n=1 un (f ), then u ∈ S 0 (RN ) and ∂ α u = n=1 ∂ α un (cf. 10.10.10).
P

10.11.18. Theorem. Each tempered distribution is the sum of derivatives


of tempered measures.
C Let u ∈ S 0 (RN ). On account of 10.11.7 (3) and 5.3.7, for some n ∈ N
and K > 0, observe that
X
(1 + | · |2 )n ∂ α f (f ∈ S (RN )).

|u(f )| ≤ K ∞
|α|≤n

From 3.5.3 and 3.5.7, for some µα ∈ M (RN ), obtain


X
µα (1 + | · |2 )n ∂ α f (f ∈ S (RN )).

u(f ) =
|α|≤n

Let να := (−1)|α| (1 + | · |2 )n µα . Then να is tempered and u = ∂ α να . .


P
|α|≤n
Duality and Its Applications 219

10.11.19. Definition. For u ∈ S 0 (RN ), the distribution Fu acting as

hf | Fui = hFf | ui (f ∈ S (RN ))

is the Fourier transform or, amply, the Fourier–Schwartz transform of u.


10.11.20. Theorem. The Fourier–Schwartz transform F is a unique extension
of the Fourier transform in S (RN ) to a topological automorphism of S 0 (RN ). The
inverse F−1 of F is a unique continuous extension of the inverse Fourier transform
in S (RN ).
C The Fourier–Schwartz transform in S 0 (RN ) is the dual of the Fourier trans-
form in S (RN ). It remains only to appeal to 10.11.7 (5), 10.11.12, 10.11.17 (2),
and 4.5.10. B

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 Hausdorff topological vector space is finite-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 finite-dimensional.
10.5. Describe weak convergence in classical Banach spaces.
10.6. Prove that a normed space is finite-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 finite 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, σ(X, X 0 )) to (Y, σ(Y, Y 0 ))).
10.9. Let k · k1 and k · k2 be two norms making X into a Banach space. Assume further
that (X, k · k1 )0 ∩ (X, k · k2 )0 separates the points of X. Prove that these norms are equivalent.
10.10. Let S act from Y 0 to X 0 . When does S serve as the dual of some operator from X
to Y ?
10.11. What is the Mackey topology τ (X, X # )? Q
10.12. Let (Xξ )ξ∈„ be a family of locally convex spaces, and let X := ξ∈„ Xξ be the prod-
uct of the family. Validate the presentations:
Y Y
σ(X, X 0 ) = σ(Xξ , Xξ0 ); τ (X, X 0 ) = τ (Xξ , Xξ0 ).
ξ∈„ ξ∈„

10.13. Let X and Y be Banach spaces and let T be a element of B(X, Y ) satisfying
im T = Y . Demonstrate that Y is reflexive provided so is X.
220 Chapter 10

10.14. Show that the spaces (X 0 )00 and (X 00 )0 coincide.


10.15. Prove that the space c0 has no infinite-dimensional reflexive 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 0 (∂p) ⊂
T 0 (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 6= 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 specification, we only consider unital alge-


bras over a basic field 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 A1 and A2 ,
by a representation of A1 in A2 we henceforth mean a morphism, a multiplicative
linear operator, from A1 to A2 which sends the unity element of A1 to the unity
element of A2 .
For an algebra A without unity, the process of unitization or adjunction of unity
is in order. Namely, the vector space Ae := A × C is transformed into an algebra
by putting (a, λ)(b, µ) := (ab + µa + λb, λµ), where a, b ∈ A and λ, µ ∈ C. In the
normed case, it is also taken for granted that k(a, λ)kAe := kakA + |λ|.

11.1.3. Definition. An element ar in A is a right inverse of a if aar = e.


An element al of A is a left inverse of a if al a = e.

11.1.4. If an element has left and right inverses then the latter coincide.
/ ar = (al a)ar = al (aar ) = al e = al .

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 := ar = al . 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.
222 Chapter 11

11.1.6. Theorem. Let A be a Banach algebra. Given a ∈ A, put La : x 7→ ax


(x ∈ A). Then the mapping

LA := L : a 7→ La (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.
C Clearly,

L(ab) : x 7→ Lab (x) = abx = a(bx) = La (Lb 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 Ar coinciding with A
as a vector space and equipped with the reversed multiplication ab := ba (a, b ∈ A).
Let R := LAr , i.e. Ra := Ra : x 7→ xa for a ∈ A. Check that L(A) is in fact
the centralizer of the range R (A), i.e. the closed subalgebra

Z(im R) := {T ∈ B(A) : T Ra = Ra T (a ∈ A)}.

Indeed, if T ∈ L(A), i.e. T = La for some a ∈ A, then La Rb (x) = axb =


Rb (La (x)) = Rb La (x) for all b ∈ A. Hence, T ∈ Z(R (A)). If, in turn, T ∈ Z(R (A))
then, putting a := T e, find

La x = ax = (T e)x = Rx (T e) = (Rx T )e = (T Rx )e = T (Rx e) = T x

for all x ∈ A. Consequently, T = La ∈ L(A). Thus, L(A) is a Banach subalgebra


of B(A).
For T = La there is a T −1 in B(A). Put b := T −1 e and observe that ab =
La b = T b = T T −1 e = e. Moreover, ab = e ⇒ aba = a ⇒ T (ba) = La ba = aba =
a = La e = T e. Whence ba = e, because T is a monomorphism. Thus, L(A) is
a subalgebra of A.
By the definition of a Banach algebra, the norm is submultiplicative, providing

kLk = sup{kLak : kak ≤ 1} = sup{kabk : kak ≤ 1, kbk ≤ 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). B
11.1.7. Definition. The representation LA , constructed in 11.1.6, is the
canonical operator representation of A.
Banach Algebras 223

11.1.8. Remark. The presence of the canonical operator representation allows


us to confine the subsequent exposition to studying Banach algebras with norm-one
unity.
For such an algebra A the canonical operator representation LA 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, LA 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, justifies the
use of the symbol λ in place of λe for λ ∈ C, where e is the unity of A (cf. 5.6.5).
In other words, henceforth the isometric representation λ 7→ λe is considered as the
identification 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 λ in C
is a resolvent value of a, in writing λ ∈ res(a)), if (λ − a) ∈ Inv(A). The resolvent
R (a, λ) of a at λ is R (a, λ) := λ−a1
:= (λ − 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 SpA (a) are in order.
11.2.2. For a ∈ A the equalities hold:

SpA (a) = SpL(A) (La ) = Sp(La );


LR (a, λ) = R (La , λ) (λ ∈ res(a) = res(La )). /.

11.2.3. Gelfand–Mazur Theorem. The field of complex numbers is up to


isometric isomorphism the sole Banach division algebra (or skew field); i.e., each
complex Banach algebra with norm-one unity and invertible nonzero elements has
an isometric representation on C.
C Consider ‰ : λ 7→ λe, with e the unity of A and λ ∈ C. It is clear that
‰ represents C in A. Take a ∈ A. By virtue of 11.2.2 and 8.1.11, Sp(a) 6= ∅.
Consequently, there is λ ∈ C such that the element (λ − a) is not invertible; i.e.,
a = λe by hypothesis. Hence, ‰ is an epimorphism. Moreover, k‰(λ)k = kλek =
|λ| kek = |λ| so that ‰ is an isometry. B
11.2.4. Shilov Theorem. Let A be a Banach algebra and let B be a closed
(unital) subalgebra of A. Then

SpB (b) ⊃ SpA (b), ∂ SpA (b) ⊃ ∂ SpB (b)

for all b ∈ B.
224 Chapter 11

C If b := λ − b ∈ Inv(B) then surely b ∈ Inv(A). Whence resB (b) ⊂ resA (b);


i.e.,
SpB (b) = C \ resB (b) ⊃ C \ resA (b) = SpA (b).
If λ ∈ ∂ SpB (b) then b ∈ ∂ Inv(B). Therefore,
−1 there is a sequence (bn ), bn ∈
Inv(B), convergent to b. Putting t := supn∈N bn , deduce that
−1
bn − b−1
−1
= bn (1 − bn b−1
−1 −1
≤ t2 bn − bm .

m ) = bn (bm − bn )bm
m

In other words, if t < +∞ then there is a limit a := lim b−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
−1 −1 −1
bn → +∞. Put an := b−1 n
bn . Then

ban = (b − bn )an + bn an
−1 −1
≤ b − bn kan k + b−1

n
bn bn → 0.
−1
Whence it follows that b is not invertible. Indeed, in the opposite case for a := b
it would hold that

1 = kan k = aban ≤ kak ban → 0.
Finally, conclude that λ − b does not belong to Inv(A); i.e., λ ∈ SpA (b). Since
λ is a boundary point of a larger set SpB (b); undoubtedly, λ ∈ ∂ SpA (b). B
11.2.5. Corollary. If SpB (b) lacks interior points then SpB (b) = SpA (b).
/ SpB (b) = ∂ SpB (b) ⊂ ∂ SpB (b) ⊂ ∂ SpA (b) ⊂ SpA (b) ⊂ SpB (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
I
1 h(z)
Ra h := dz.
2πi z−a
The element Ra h of A is the Riesz–Dunford 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) := Ra f .
Banach Algebras 225

11.3.2. Gelfand–Dunford The Riesz–Dunford integral Ra represents the al-


gebra of germs of holomorphic functions
P∞on the spectrum of an element a of a Banach
algebra PA in A. Moreover, if f (z) := n=0 cn z n (in a neighborhood of Sp(a)) then

f (a) := n=0 cn an .
From 11.2.3, 8.2.1, and 11.2.2 obtain

(LRa h)(b) = LRa h b = (Ra h)b


I I
1 1
= h(z)R (a, z) dzb = h(z)R (a, z) bdz
2πi 2πi
I I
1 1
= h(z)R (La , z) bdz = h(z)R (La , z) dzb
2πi 2πi
= RLa h(b)
for all b ∈ A. In particular, im L includes the range of RLa (H (Sp(a))). Therefore,
the already-proven commutativity of the diagram

H (Sp(a))
@
@ R
RLa @ a
? @
L R
@
B(A)  A

implies that the following diagram also commutes:

H (Sp(a))
@
@ R
RLa @ a
@
? −1 @
L(A) L
R
-A

It remains to appeal to 11.1.6 and the Gelfand–Dunford Theorem in an operator


setting. B
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.
226 Chapter 11

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 C 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,

supJ(A) E = supLat(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 infima of arbitrary subsets.
C Clearly, 0 is the least ideal, whereas A is the greatest ideal. Furthermore,
the intersection of ideals and the sum of finitely many ideals are ideals. It remains
to refer to 2.1.5 and 2.1.6. B
11.4.3. Let J0 C A. Assume further that ϕ : A → A/J0 is the coset mapping
of A onto the quotient algebra A := A/J0 . Then

J C A ⇒ ϕ(J) C A; J C A ⇒ ϕ−1 (J) C A.

C Since by definition ab := ϕ(ϕ−1 (a)ϕ−1 (b)) for a, b ∈ A, the operator ϕ is


multiplicative: ϕ(ab) = ϕ(a)ϕ(b) for a, b ∈ A. Whence successively derive

ϕ(J) ⊂ Aϕ(J) = ϕ(A)ϕ(J) ⊂ ϕ(AJ) ⊂ ϕ(J);


−1
ϕ (J) ⊂ Aϕ−1 (J) ⊂ ϕ−1 (ϕ(A)J) = ϕ−1 (AJ) ⊂ ϕ−1 (J). .

11.4.4. Let J C A and J 6= 0. The following conditions are equivalent:


(1) A 6= J;
(2) 1 ∈
/ J;
(3) no element of J has a left inverse. CB
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. CB
11.4.7. Let J be a proper ideal of A. Then (J is maximal) ⇔ (A/J is a field).
Banach Algebras 227

C ⇒: Let J C A/J. Then, by 11.4.3, ϕ−1 (J) C A. Since, beyond a doubt,


J ⊂ ϕ−1 (J); therefore, either J = ϕ−1 (J) and 0 = ϕ(J) = ϕ(ϕ−1 (J)) = J, or
A = ϕ−1 (J) and J = ϕ(ϕ−1 (J)) = ϕ(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 J0 C A and J0 ⊂ J. Then, by 11.4.3, ϕ(J0 ) C A/J. In virtue of 11.4.6,
either ϕ(J0 ) = 0 or ϕ(J0 ) = A/J. In the first case, J0 ⊂ ϕ−1 ◦ ϕ(J0 ) ⊂ ϕ−1 (0) = J
and J = J0 . In the second case, ϕ(J0 ) = ϕ(A); i.e., A = J0 + J ⊂ J0 + J0 = J0 ⊂ A.
Thus, J is a maximal ideal. B
11.4.8. Krull Theorem. Each proper ideal is included in some maximal ideal.
C Let J0 be a proper ideal of an algebra A. Assume further that E is the set
comprising all proper ideals J of A such that J0 ⊂ J. In virtue of 11.4.2 each chain
E0 in E has a least upper bound: sup E = ∪ {J : J ∈ E0 }. By 11.4.4 the ideal sup E0
is proper. Thus, E is inductive and the claim follows from the Kuratowski–Zorn
Lemma. B

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

Q0 := ∩ {f −1 (0) : f ∈ J};
J0 := {f ∈ C(Q, C) : int f −1 (0) ⊃ Q0 }.

Then J0 C C(Q, C) and J0 ⊂ J.


C Let Q1 := cl(Q \ f −1 (0)) for a function f , a member of J0 . By hypothesis,
Q1 ∩ Q0 = ∅. To prove the containment f ∈ J it is necessary (and, certainly,
sufficient) to find u ∈ J satisfying u(q) = 1 for all q ∈ Q1 . Indeed, in that event
uf = f .
With this in mind, observe first that for q ∈ Q1 there is a function fq in J such
that fq (q) 6= 0. Putting gq := fq∗ fq , where as usual fq∗ : x 7→ fq (x)∗ is the conjugate
of fq , observe that gq ≥ 0 and, moreover, gq (q) > 0. It is also clear that gq ∈ J for
q ∈ Q1 . The family (Uq )q∈Q1 , with Uq := {x ∈ Q1 : gq (x) > 0}, is an open cover
of Q1 . Using a standard compactness argument, choose a finite subset {q1 , . . . , qn }
of Q1 such that Q1 ⊂ Uq1 ∪ . . . ∪ Uqn . Put g := gq1 + . . . + gqn . Undoubtedly, g ∈ J
and g(q) > 0 for q ∈ Q1 . Let h0 (q) := g(q)−1 for q ∈ Q1 . By the Tietze–Urysohn
Theorem, there is a function h in C(Q, R) satisfying h Q1 = h0 . Finally, let u := hg.
This u is a sought function.
We have thus demonstrated that J0 ⊂ J. Moreover, J0 is an ideal of C(Q, C)
for obvious reasons. B
228 Chapter 11

11.5.2. For every closed ideal J of the algebra C(Q, C ) there is a unique
compact subset Q0 of Q such that

J = J(Q0 ) := {f ∈ C(Q, C) : q ∈ Q0 ⇒ f (q) = 0}.

C Uniqueness follows from the Urysohn Theorem. Define Q0 as in 11.5.1.


Then, surely, J ⊂ J(Q0 ). Take f ∈ J(Q0 ) and, given n ∈ N, put

Un := |f | ≤ 1 /2n , Vn := |f | ≥ 1 /n .
 

Once again
using the Urysohn
Theorem, find hn ∈ C(Q, R) satisfying 0 ≤ hn ≤ 1
with hn U = 0 and hn V = 1. Consider fn := f hn . Since
n n

int fn−1 (0) ⊃ int Un ⊃ Q0 ,

therefore, from 11.5.1 derive fn ∈ J. It suffices to observe that fn → f by con-


struction. B
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.
C Follows from 11.5.2, because the closure of an ideal is also an ideal. B

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 ϕ : A → A/J is the coset mapping then ϕ(1) is the unity
of A/J, the operator ϕ is multiplicative and kϕk = 1.
C Given a, b ∈ A, from 5.1.10 (5) derive

kϕ(a)ϕ(b)kA/J = inf{ka0 b0 kA : ϕ(a0 ) = ϕ(a), ϕ(b0 ) = ϕ(b)}


≤ inf{ka0 kA kb0 kA : ϕ(a0 ) = ϕ(a), ϕ(b0 ) = ϕ(b)}
= kϕ(a)kA/J kϕ(b)kA/J .

In other words, the norm of A/J is submultiplicative. Consequently, kϕ(1)k ≥ 1.


Furthermore,

kϕ(1)kA/J = inf{kakA : ϕ(a) = ϕ(1)} ≤ k1kA = 1;

i.e., kϕ(1)k = 1. Whence, in particular, the equality kϕk = 1 follows. The remaining
claims are evident. B
Banach Algebras 229

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 χ : A → C be a nonzero multiplicative linear functional on A.
Then χ is continuous and kχk = χ(1) = 1 (in particular, χ is a representation of A
in C).
C Since χ 6= 0; therefore, 0 6= χ(a) = χ(a1) = χ(a)χ(1) for some a in A.
Consequently, χ(1) = 1. If now a in A and λ in C are such that |λ| > kak, then
λ−a ∈ Inv(A) (cf. 5.6.15). So, 1 = χ(1)χ(λ−a)χ((λ−a)−1 ). Whence χ(λ−a) 6= 0;
i.e., χ(a) 6= λ. Thus, |χ(a)| ≤ kak and kχk ≤ 1. Since kχk = kχk k1k ≥ |χ(1)| = 1,
conclude that kχk = 1. B
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 σ(A0 , A)) and called the character space of A.
11.6.5. The character space is a compactum.
C It is beyond a doubt that X(A) is a Hausdorff space. By virtue of 11.6.3,
X(A) is a σ(A0 , A)-closed subset of the ball BA0 . The latter is σ(A0 , A)-compact
by the Alaoglu–Bourbaki Theorem. It remains to refer to 9.4.9. B
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
χ 7→ ker χ from the character space X(A) onto the set M(A) of all maximal ideals
of A is one-to-one.
C Let χ ∈ X(A) be a character of A. Clearly, ker χ C A. From 2.3.11 it follows
that the monoquotient χ : A/ ker χ → C of χ is a monomorphism. In view of 11.6.1,
χ(1) = χ(1) = 1; i.e., χ is an isomorphism of A/ ker χ and C. Consequently,
A/ ker χ is a field. Using 11.4.7, infer that ker χ is a maximal ideal; i.e., ker χ ∈
M(A). Now, let m ∈ M(A) be some maximal ideal of A. It is clear that m ⊂ cl m,
cl m C 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 ϕ : A → A/m. In view
of 11.4.7 and 11.6.1, A/m is a Banach field. By the Gelfand–Mazur Theorem, there
is an isometric representation ψ : A/m → C. Put χ := ψ ◦ ϕ. It is evident that
χ ∈ X(A) and ker χ = χ−1 (0) = ϕ−1 (ψ −1 (0)) = ϕ−1 (0) = m.
To complete the proof, it suffices to show that the mapping χ 7→ ker χ is one-
to-one. Indeed, let ker χ1 = ker χ2 for χ1 , χ2 ∈ X(A). By 2.3.11 χ1 = λχ2 for some
λ ∈ C. Furthermore, by 11.6.3, 1 = χ1 (1) = λχ2 (1) = λ. Finally, χ1 = χ2 . B
230 Chapter 11

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 χ 7→ ker χ.
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 identified 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 χ ∈ X(A), put b a(χ) := χ(a). The
a : χ 7→ b
resulting function b a(χ), defined on X(A), is the Gelfand transform of a.
The mapping a 7→ b a, with a ∈ A is the Gelfand transform of A, denoted by GA
(or b ).
11.6.9. Gelfand Transform Theorem. The Gelfand transform GA : a 7→ b a
is a representation of a commutative Banach algebra A in the algebra C(X(A), C ).
Moreover,

Sp(a) = Sp(b
a) = b
a(X(A)),
kb
ak = r(a),

with a ∈ A and r(a) standing for the spectral radius of a.


C The implications a ∈ A ⇒ b a ∈ C(X(A), C ), b 1 = 1 and a, b ∈ A ⇒
ab = b
b ab follow from definitions and 11.6.3. The linearity of GA raises no doubts.
b
Consequently, the mapping GA is a representation.
To begin with, take λ ∈ Sp(a). Then λ − a is not invertible, and so the ideal
Jλ−a := A(λ − 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 ⊃ Jλ−a . By Theorem 11.6.6,
m = ker χ for a suitable character χ. In particular, χ(λ − a) = 0; i.e., λ = λχ(1) =
χ(λ) = χ(a) = b a(χ). Consequently, λ ∈ Sp(ba).
If, in turn, λ ∈ Sp(b
a) then (λ − b
a) is not invertible in C(X(A), C); i.e., there
is a character χ ∈ X(A) such that λ = b a(χ). In other words, χ(λ − a) = 0. Thus,
the assumption λ − a ∈ Inv(A) leads to the following contradiction:

1 = χ(1) = χ((λ − a)−1 (λ − a)) = χ((λ − a)−1 )χ(λ − a) = 0.

Hence, λ ∈ Sp(a). Finally, Sp(a) = Sp(b


a).
Using the Beurling–Gelfand formula (cf. 11.3.3 and 8.1.12), infer that

r(a) = sup{|λ| : λ ∈ Sp(a)} = sup{|λ| : λ ∈ Sp(b


a)}
= sup{|λ| : λ ∈ b
a(X(A))} = sup{|ba(χ)| : χ ∈ X(A)} = kbak,

what was required. B


Banach Algebras 231

11.6.10. The Gelfand transform of a commutative Banach algebra A is an iso-


metric embedding if and only if ka2 k = kak2 for all a ∈ A.
C ⇒: The mapping t 7→ t2 , viewed as acting on R+ , and the inverse of the
mapping on R+ are both increasing. Therefore, from 10.6.9 obtain

ka2 k = kb
a2 kC(X(A),C ) = sup |b
a2 (χ)| = sup |χ(a2 )|
χ∈X(A) χ∈X(A)

= sup |χ(a)χ(a)| = sup |χ(a)|2


χ∈X(A) χ∈X(A)
!2
= sup |χ(a)| ak2 = kak2 .
= kb
χ∈X(A)

2n ⇐:
By the Gelfand formula, r(a) = lim kan k /n . In particular, observe that
n
a = kak2 ; i.e., r(a) = kak. By 10.6.9, r(a) = kb
ak, completing the proof. B
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 GA is the intersection of all maximal ideals, called the radi-
cal of A. Therefore, the condition for GA 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(b
a))
@
@ R
Ra @ ^a
@
GA -@
? R
A C(X(A), C)

a) = f ◦ b
Moreover, f (b a) for f ∈ H(Sp(a)).
a = f (b
C Take χ ∈ X(A). Given z ∈ res(a), observe that
   
1 1 1 1
χ (z − a) = 1 ⇒ χ = = .
z−a z−a χ(z − a) z − χ(a)
In other words,

1
d 1 1
R\
(a, z)(χ) = (χ) = = (χ) = R (b
a, z)(χ).
z−a z−b
a(χ) z−ba
232 Chapter 11

Therefore, appealing to the properties of the Bochner integral (cf. 5.5.9 (6)) and
given f ∈ H(Sp(a)), infer that

 I 
1
fd(a) = GA ◦ Ra f = GA f (z)R (a, z) dz
2πi
I I
1 1
= f (z)GA (R (a, z)) dz = f (z)R\ (a, z)) dz
2πi 2πi
I
1
= a, z) dz = R^a (f ) = f (b
f (z)R (b a).
2πi

Furthermore, given χ ∈ X(A) and using the Cauchy Integral Formula, derive the
following chain of equalities

f ◦ba(χ) = f (b
a(χ)) = f (χ(a))
I I  
1 f (z) 1 f (z)
= dz = χ dz
2πi z − χ(a) 2πi z−a
I 
1 f (z)
= χ dz = fd (a)(χ) = f (b
a)(χ). .
2πi z−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
Definitions 11.6.4 and 11.6.8 verbatim. A character χ in X(A) generates some
character χe in X(Ae ) by the rule χe (a, λ) := χ(a) + λ (a ∈ A, λ ∈ C). The set
X(Ae ) \ {χe : χ ∈ X(A)} is a singleton consisting of the sole element χ∞ (a, λ) := λ
(a ∈ A, λ ∈ C ). The space X(A) is locally compact (cf. 9.4.19), because the
mapping χ ∈ X(A) 7→ χe ∈ X(Ae )\{χ∞ } is a homeomorphism. Moreover, ker χ∞ =
A × 0. Consequently, the Gelfand transform of a commutative Banach algebra
without unity represents it in the algebra of continuous complex functions defined
on a locally compact space and vanishing at infinity. Given the group algebra
(L1 (RN ), ∗), observe that by 10.11.1 and 10.11.3 the Fourier transform coincides
with the Gelfand transform, which in turn entails the Riemann–Lebesgue 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).
Banach Algebras 233

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,
1 1
x = (a + a∗ ), y = (a − a∗ ).
2 2i

Moreover, a = x − iy.
C Only the claim of uniqueness needs examining. If a = x1 + iy1 then, using
the properties of involution (cf. 6.4.13), proceed as follows: a∗ = x∗1 + (iy1 )∗ =
x∗1 − iy1∗ = x1 − iy1 . Thus, x1 = x and y1 = y. B
11.7.3. The unity of A is a hermitian element of A.
/ 1∗ = 1∗ 1 = 1∗ 1∗∗ = (1∗ 1)∗ = 1∗∗ = 1 .
11.7.4. a ∈ Inv(A) ⇔ a∗ ∈ Inv(A). Moreover, involution and inversion com-
mute.
C For a ∈ Inv(A) by definition 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. B
11.7.5. Sp(a∗ ) = Sp(a)∗ . CB
11.7.6. The spectrum of a unitary element of a C ∗ -algebra is a subset of the
unit circle.
C By Definition 6.4.13, ka2 k = ka∗ ak ≤ ka∗ k kak for an arbitrary a. In other
words, kak ≤ ka∗ k. Therefore, from a = a∗∗ infer that kak = ka∗ k. If now a is
a unitary element, a∗ = a−1 ; then kak2 = ka∗ ak = ka−1 ak = 1. Consequently,
kak = ka∗ k = ka−1 k = 1. Whence it follows that Sp(a) and Sp(a−1 ) both lie
within the unit disk. Furthermore, Sp(a−1 ) = Sp(a)−1 . B
11.7.7. The spectrum of a hermitian element of a C ∗ -algebra is real.
C Take a in A arbitrarily. From the Gelfand–Dunford Theorem in an algebraic
setting derive

!∗ ∞ ∞
X a n X (an )∗ X (a∗ )n
exp(a)∗ = = = = exp(a∗ ).
n=0
n! n=0
n! n=0
n!
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 λ ∈ Sp(h) then by the Spectral Mapping Theorem
(also cf. 11.3.3) exp(iλ) ∈ Sp(a) ⊂ T. Thus, 1 = | exp(iλ)| = | exp(i Re λ − Im λ)| =
exp(− Im λ). Finally, Im λ = 0; i.e., λ ∈ R. B
234 Chapter 11

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.


C 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 suffices to show only that (b∗ b)−1 belongs to B. Since b∗ b
is hermitian in B, the inclusion holds: SpB (b∗ b) ⊂ R (cf. 11.7.7). Using 11.2.5, infer
that SpA (b∗ b) = SpB (b∗ b). Since 0 ∈
/ SpA (b∗ b); therefore, b∗ b ∈ Inv(B). Finally,
b ∈ Inv(B). 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

SpB (b) = SpA (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 Gelfand–Naı̆mark Theorem

11.8.1. The Banach algebra C(Q, C ) with the natural involution f 7→ f ∗ ,


where f ∗ (q) := f (q)∗ for q ∈ Q, is a C ∗ -algebra.

/ kf ∗ f k = sup{|f (q)∗ f (q)| : q ∈ Q} = sup{|f (q)|2 : q ∈ Q}


= (sup |f (Q)|)2= kf k2 .

11.8.2. Stone–Weierstrass Every unital C ∗ -subalgebra of the C ∗ -algebra


C(Q, C ), which separates the points of Q, is dense in C(Q, C).
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 Stone–Weierstrass Theorem for C(Q, R), the subalgebra Re A is dense
in C(Q, R). It remains to refer to 11.7.2. B
Banach Algebras 235

11.8.3. Definition. A representation of a ∗-algebra agreeing with involution ∗


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:

R
A−→B
∗↓ ↓∗
R
A−→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 Gelfand–Naı̆mark Theorem. The Gelfand trans-


form of a commutative C ∗ -algebra A implements an isometric ∗-isomorphism of A
and C(X(A), C ).
C Given a ∈ A, observe that

1 1
ka2 k = k(a2 )∗ a2 k /2
= ka∗ aa∗ ak /2
= ka∗ ak = kak2 .

In virtue of 11.6.10 the Gelfand transform GA is thus an isometry of A and a closed


subalgebra A b of C(X(A), C). Undoubtedly, A b separates the points of X(A) and
contains constant functions.
By virtue of 11.6.9 and 11.7.7, b 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 χ(x) ∈ R and
χ(y) ∈ R hold for every character χ, a member of X(A).
With this in mind, successively infer that

GA (a)∗ (χ) = b
a∗ (χ) = b
a(χ)∗ = χ(a)∗ = χ(x + iy)∗
= (χ(x) + iχ(y))∗ = χ(x) − iχ(y) = χ(x − iy) = χ(a∗ )
= ab∗ (χ) = GA (a∗ )(χ) (χ ∈ X(A)).

Consequently, the Gelfand transform GA is a ∗-representation and, in particular,


Ab is a C ∗ -subalgebra of C(X(A), C ). It remains to appeal to 11.8.2 and conclude
that A b = C(X(A), C ). B

11.8.5. Assume that R : A → B is a ∗-representation of a C ∗ -algebra A


in a C ∗ -algebra B. Then kRak ≤ kak for a ∈ A.
236 Chapter 11

C Since R(1) = 1; therefore, R(Inv(A)) ⊂ Inv(B). Hence, SpB (R(a)) ⊂


SpA (a) for a ∈ A. Whence it follows from the Beurling–Gelfand formula that the
inequality rA (a) ≥ rB (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
A0 is the least closed C ∗ -subalgebra containing a and B0 is an analogous subalgebra
containing R(a), then A0 and B0 are commutative C ∗ -algebras. Therefore, from
Theorems 11.8.4 and 11.6.9 obtain

kR(a)k = kR(a)kB0 = kGB0 (R(a))k = rB0 (R(a))


= rB (R(a)) ≤ rA (a) = rA0 (a) = kGA0 (a)k = kak.

Given a ∈ A, it is easy to observe that a∗ a is a hermitian element. Thus,

kR(a)k2 = kR(a)∗ R(a)k = kR(a∗ a)k ≤ ka∗ ak = kak2 . .

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 Ra
of C(Sp(a), C) in A such that a = Ra (ISp(a) ).
C 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) = SpA (a) = SpB (a). The Gelfand transform b a := GB (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 b
a is a homeomorphism. Whence it is immediate

that the mapping R : f 7→ f ◦ b 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 Riesz–Dunford integral which is revealed in 11.6.12, for the identity mapping
observe that

a = R^a IC = IC ◦ b

b a = IC ^a(X(B)) ◦ b
a = IC Sp(a) ◦ b
a = ISp(a) ◦ b
a = R(ISp(a) ).

Now, put

Ra := GB−1 ◦ R.
Clearly, Ra is an isometric embedding and a ∗-representation. Moreover,

Ra (ISp(a) ) = GB−1 ◦ R(ISp(a) ) = GB−1 (b
a) = a.

Uniqueness for such representation Ra 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 ISp(a) . B
Banach Algebras 237

11.8.7. Definition. The representation Ra : C(Sp(a), C) → A of 11.8.6 is


the continuous functional calculus (for a normal element a of A). The element
Ra (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 defined 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, define the element Ra (f ) := Ra 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 different symbols one and the same element. This
circumstance may be expressed in visual form.
Namely, let · Sp(a) stand for the mapping sending a germ h, 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

·|Sp(a)-
H (Sp(a)) C(Sp(a), C )
@
Ra
@
Ra @
@
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 A0
is a state of A, in writing s ∈ S(A), if ksk = 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+ .
C Let a ≥ 0 and ksk = s(1) = 1. We have to prove that s(a) ≥ 0. Take z ∈ C
and ε > 0 such that the disk Bε (z) := z + εD includes a(Q). Then ka − zk ≤ ε and,
consequently, |s(a − z)| ≤ ε. Hence, |s(a) − z| = |s(a) − s(z)| ≤ ε; i.e., s(a) ∈ Bε (z).
Observe that

∩ {Bε (z) : Bε (z) ⊃ a(Q)} = cl co(a(Q)) ⊂ R+ .


238 Chapter 11

Thus, s(a) ∈ R+ . B
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+ .
C 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 b a := GB (a) provides b a(X(B)) = SpB (a). In view of 11.7.10,
SpB (a) = Sp(a). In other words, for λ ∈ Sp(a) there is a character χ of B satisfying
the condition χ(a) = λ. By 11.6.3, kχk = χ(1) = 1. Using 7.5.11, find a norm-
preserving extension s of χ onto A. Then s is a state of A and s(a) = λ. 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 b 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 b
a ≥ 0. Thus, Sp(a) ⊂ R+ ⇒ N (a) ⊂ R+ , which ends the
proof. B
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.
C It is clear that N (a+b) ⊂ N (a)+N (b) and N (αa) = αN (a) for a, b ∈ A and
α ∈ R+ . Hence, 11.9.3 ensures the inclusion α1 A+ + α2 A+ ⊂ A+ for α1 , α2 ∈ 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 kak = 0. B
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.

C Everything is immediate from the continuous functional calculus. B


11.9.7. Kaplansky–Fukamija Lemma. An element a of a C ∗ -algebra A is
positive if and only if a = b∗ b for some b in A.
C ⇒: 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 = u − v, where uv = vu = 0 with u ≥ 0 and v ≥ 0 (in the ordered
vector space (AR , A+ )). Straightforward calculation yields the equalities

(bv)∗ bv = v ∗ b∗ bv = vb∗ bv = v(u − v)v = (vu − v 2 )v = −v 3 .


Banach Algebras 239

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 differ only by zero. Therefore, bv(bv)∗ ≤ 0.
In virtue of 11.7.2, bv = a1 + ia2 for suitable hermitian elements a1 and a2 .
It is evident that a21 , a22 ∈ A+ and (bv)∗ = a1 − ia2 . Using 11.9.5 twice, arrive
at the estimates
0 ≥ (bv)∗ bv + bv(bv)∗ = 2 a21 + a22 ≥ 0.


By 11.9.5, a1 = a2 = 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+ . B
11.9.8. Every state s of a C ∗ -algebra A is hermitian; i.e.,

s(a∗ ) = s(a)∗ (a ∈ A).

C 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,

Im s(a) + Im s(a∗ ) = 0; Re(−s(a)) + Re s(a∗ ) = 0.

Whence it follows that

s(a∗ ) = Re s(a∗ ) + i Im s(a∗ ) = Re s(a) − i Im s(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.
C 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-definite hermitian
form. B
11.9.10. GNS-Construction Theorem. To every state s of an arbitrary
C ∗ -algebra A there correspond a Hilbert space (Hs , ( · , · )s ), an element xs in Hs
and a ∗-representation Rs : A → B(Hs ) such that s(a) = (Rs (a)xs , xs )s for all
a ∈ A and the set {Rs (a)xs : a ∈ A} is dense in Hs .
240 Chapter 11

C In virtue of 11.9.9, putting (a, b)p s := s(b a) for a, b ∈ A, obtain a pre-
Hilbert space (A, ( · , · )s ). Let ps (a) := (a, a)s stand for the seminorm of the
space. Assume that ϕs : A → A/ ker ps is the coset mapping of A onto the
Hausdorff pre-Hilbert space A/ ker ps associated with A. Assume further that
ιs : A/ ker ps → Hs is an embedding (for instance, by the double prime map-
ping) of A/ ker ps onto a dense subspace of the Hilbert space Hs associated with
(A, ( · , · )s ) (cf. 6.1.10 (4)). The inner product in Hs retains the previous notation
( · , · )s . Therefore, in particular,
(ιs ϕs a, ιs ϕs b)s = (a, b)s = s(b∗ a) (a, b ∈ A).
Given a ∈ A, consider (the image under the canonical operator representation)
La : b 7→ ab (b ∈ A). Demonstrate first that there are unique bounded operators
La and Rs (a) making the following diagram commutative:
ϕs s ι
A−→A/ ker ps −→H s
La ↓ ↓ La ↓ Rs (a)
ϕs ιs
A−→A/ ker ps −→H s

A sought operator La is a solution to the equation Xϕs = ϕs La . Using 2.3.8,


observe next that the necessary and sufficient condition for solvability of the equa-
tion in linear operators consists in invariance of the subspace ker ps under La . Thus,
examine the inclusion La (ker ps ) ⊂ ker ps . To this end, take an element b of ker ps ,
i.e. ps (b) = 0. By definition and the Cauchy–Bunyakovskiı̆–Schwarz inequality,
deduce that

0 ≤ (La b, La b)s = (ab, ab)s = s((ab)∗ ab)


= s(b∗ a∗ ab) = (a∗ ab, b)s ≤ ps (b)ps (a∗ ab) = 0;
i.e., La b ∈ ker ps . Uniqueness for La is provided by 2.3.9, since ϕs is an epi-
morphism. Observe also that ϕs is an open mapping (cf. 5.1.3). Whence the
continuity of La is immediate. Therefore, in virtue of 5.3.8 the correspondence
ιs ◦ La ◦ (ιs )−1 may be considered as a bounded linear operator from ιs (A/ ker ps )
to the Banach space Hs . By 4.5.10 such an operator extends uniquely to an operator
Rs (a) in B(Hs ).
Demonstrate now that Rs : a 7→ Rs (a) is a sought representation. By 11.1.6,
Lab = La Lb for a, b ∈ A. Consequently,
ϕs Lab = ϕs La Lb = La ϕs Lb = La Lb ϕs .
Since Lab is a unique solution to the equation Xϕs = ϕs Lab , infer the equality
Lab = La Lb , which guarantees multiplicativity for Rs . The linearity of Rs may be
verified likewise. Furthermore,
L1 ϕs = ϕs L1 = ϕs IA = ϕs = IA/ ker ps ϕs = 1ϕs ;
Banach Algebras 241

i.e., Rs (1) = 1.
For simplicity, put ψs := ιs ϕs . Then, on account taken of the definition
of the inner product on Hs (cf. 6.1.10 (4)) and the involution in B(Hs ) (cf. 6.4.14
and 6.4.5), given elements a, b, and y in A, infer that

(Rs (a∗ )ψs x, ψs y)s = (ψs La∗ x, ψs y)s


= (La∗ x, y)s = (a∗ x, y)s = s(y ∗ a∗ x) = s((ay)∗ x) = (x, ay)s
= (x, La y)s = (ψs x, ψs La y)s = (ψs x, Rs (a)ψs y)s = (Rs (a)∗ ψs x, ψs y)s .

Now, since im ψs is dense in Hs it follows that Rs (a∗ ) = Rs (a)∗ for all a in A; i.e.,
Rs is a ∗-representation.
Let xs := ψs 1. Then

Rs (a)xs = Rs (a)ψs 1 = ψs La 1 = ψs a (a ∈ A).

Consequently, the set {Rs (a)xs : a ∈ A} is dense in Hs . Furthermore,

(Rs (a)xs , xs )s = (ψs a, ψs 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 Gelfand–Naı̆mark–Segal con-
struction, which is reflected in the name of 11.9.10.
11.9.12. Gelfand–Naı̆mark Each C ∗ -algebra has an isometric ∗-representation
in the endomorphism algebra of a suitable Hilbert space.
C Let A be a C ∗ -algebra. We have to find 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
(Hs )s∈S(A) which exists in virtue of the GNS-Construction Theorem; i.e.,
 
 Y X 
H := ⊕ Hs = h := (hs )s∈S(A) ∈ Hs : khs k2Hs < +∞ .
s∈S(A)  
s∈S(A) s∈S(A)

Observe that the inner product of h := (hs )s∈S(A) and g := (gs )s∈S(A) is calcu-
lated by the rule (cf. 6.1.10 (5) and 6.1.9):
X
(h, g) = (hs , gs )s .
s∈S(A)
242 Chapter 11

Assume further that Rs is a ∗-representation of A on the space Hs correspond-


ing to s in S(A). Since in view of 11.8.5 there is an estimate kRs (a)kB(Hs ) ≤ kak
for a ∈ A; therefore, given h ∈ H, infer that
X X X
kRs (a)hs k2Hs ≤ kRs (a)k2B(Hs ) khs k2Hs ≤ kak2 khs k2Hs .
s∈S(A) s∈S(A) s∈S(A)

Whence it follows that the expression R(a)h : s 7→ Rs (a)hs defines an element


R(a)h of H. The resulting operator R(a) : h 7→ R(a)h is a member of B(H).
Moreover, the mapping R : a 7→ R(a) (a ∈ A) is a sought isometric ∗-representation
of A.
Indeed, from the definition of R and the properties of Rs 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
X
(R(a∗ )h, g) = (Rs (a∗ )hs , gs )s
s∈S(A)
X X
= (Rs (a)∗ hs , gs )s = (hs , Rs (a)gs )s
s∈S(A) s∈S(A)

= (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 kR(a)k = kak for all a ∈ A. First, assume a positive. From the Spectral
Theorem and the Weierstrass Theorem it follows that kak ∈ Sp(a). In virtue
of 11.9.3 (1) there is a state s ∈ S(A) such that s(a) = kak. Using the properties
of the vector xs corresponding to the ∗-representation Rs (cf. 11.9.10) and applying
the Cauchy–Bunyakovskiı̆–Schwarz inequality, infer that

kak = s(a) = (Rs (a)xs , xs )s ≤ kRs (a)xs kHs kxs kHs


≤ kRs (a)kB(Hs ) kxs k2Hs = kRs (a)kB(Hs ) (xs , xs )s
= kRs (a)kB(Hs ) (Rs (1)xs , xs )s = kRs (a)kB(Hs ) s(1) = kRs (a)kB(Hs ) .
From the estimates kR(a)k ≥ kRs (a)kB(Hs ) and kak ≥ kR(a)k, the former
obvious and the latter indicated in 11.8.5, derive
kak ≥ kR(a)k ≥ kRs (a)kB(Hs ) ≥ kak.
Finally, take a ∈ A. By the Kaplansky–Fukamija Lemma, a∗ a is positive. So,
kR(a)k2 = kR(a)∗ R(a)k = kR(a∗ )R(a)k = kR(a∗ a)k = ka∗ ak = kak2 .
No further explanation is needed. B
Banach Algebras 243

Exercises
11.1. Give examples of Banach algebras and non-Banach algebras.
11.2. Let A be a Banach algebra. Take χ ∈ A# such that χ(1) = 1 and χ(Inv(A)) ⊂ Inv(C).
Prove that χ 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 ε > 0 such that Sp(a + b) ⊂ U for all b ∈ A satisfying kbk ≤ ε.
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 infinite summable sequences l1 (Z) with
multiplication

X
(a ∗ b)(n) := an−k bk .
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 kb ak∞ = sup |b
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 1B = 1A .
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 SpA (a). Demonstrate
that SpB (T a) ⊂ SpA (a) and T f (a) = f (T a).
11.11. Let f ∈ A0 , 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 kf k = 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(Q1 , C) and C(Q2 , C), with Q1 and Q2 compact, are
isomorphic if and only if Q1 and Q2 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 kT k ≤ 1. Then T (a∗ ) =
(T a)∗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)0 = 0 O 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 ρ be a ∗-monomorphism from A to B. Prove
that ρ 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, Khar0 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. Alexandroff P. S., Introduction to Set Theory and General Topology [in Russian], Nauka,
Moscow (1977).
10. Aliprantis Ch. D. and Border K. C., Infinite-Dimensional Analysis. A Hitchhiker’s 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, Birkhäuser, Basel etc. (1986).
14. Antonevich A. B., Knyazev P. N., and Radyno Ya. B., Problems and Exercises in Functional
Analysis [in Russian], Vysheı̆shaya 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., Mikusiński 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. Arkhangel0 skiı̆ A. V., Topological Function Spaces [in Russian], Moscow University Publ.
House, Moscow (1989).
19. Arkhangel0 skiı̆ A. V. and Ponomarëv 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).
References 245

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, Birkhäuser,
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., Theórie des Operationes Linéares, 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. Berezanskiı̆ Yu. M. and Kondrat0 ev Yu. G., Spectral Methods in Infinite-Dimensional Anal-
ysis [in Russian], Naukova Dumka, Kiev (1988).
39. Berezanskiı̆ Yu. M., Us G. F., and Sheftel0 Z. G., Functional Analysis. A Lecture Course
[in Russian], Vishcha Shkola, Kiev (1990).
40. Berg J. and Löfström 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., Il0 in V. P., and Nikol0 skiı̆ S. M., Integral Representations and Embedding
Theorems [in Russian], Nauka, Moscow (1975).
43. Bessaga Cz. and Pe␣lczyński A., Selected Topics in Infinite-Dimensional Topology, Polish
Scientific Publishers, Warsaw (1975).
44. Birkhoff G., Lattice Theory, Amer. Math. Soc., Providence (1967).
45. Birkhoff G. and Kreyszig E., “The establishment of functional analysis,” Historia Math., 11,
No. 3, 258–321 (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. Bollobás 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 Atiyah–Singer 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. 1–8 [in French], Hermann, Paris (1967).
246 References

56. Bourbaki N., Spectral Theory [in French], Hermann, Paris (1967).
57. Bourbaki N., General Topology. Ch. 5–10 [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 Radon–Nikodým 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., Differentiation 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 Hahn–Banach Theorem Surveyed, Diss. Math., 327, Warsaw (1993).
71. Caradus S., Plaffenberger 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., Tsirelson’s 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. 1–3, 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 Riesz–Dunford 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,
129–183 (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 Stone–Weierstrass theorem,” Proc. Amer. Math. Soc., 10, No. 5,
822–824 (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).
References 247

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. Dieudonné J., Foundations of Modern Analysis, Academic Press, New York (1969).
95. Dieudonné J., A Panorama of Pure Mathematics. As Seen by N. Bourbaki, Academic Press,
New York etc. (1982).
96. Dieudonné 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 Belfi V., Characterizations of C ∗ -Algebras. The Gelfand–Naı̆mark 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 Differential 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. Efimov A. V., Zolotarëv 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. Enflo P., “A counterexample to the approximation property in Banach
spaces,” Acta Math., 130, No. 3–4, 309–317 (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, Birkhäuser, Basel
etc., 1983, pp. 120–130.
118. Floret K., Weakly Compact Sets, Springer-Verlag, Berlin etc. (1980).
119. Folland G. B., Fourier Analysis and Its Applications, Wadsworth and Brooks, Pacific 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 Cliffs (1969).
122. Gelbaum B. and Olmsted G., Counterexamples in Analysis, Holden-Day, San Francisco
(1964).
248 References

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., Raı̆kov 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. Goffman C. and Pedrick G., First Course in Functional Analysis, Prentice-Hall, Englewood
Cliffs (1965).
132. Gohberg I. C. and Kreı̆n 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, Birkhäuser, Boston (1981).
134. Goldberg S., Unbounded Linear Operators. Theory and Applications, Dover, New York
(1985).
135. Gol0 dshteı̆n V. M. and Reshetnyak Yu. G., Introduction to the Theory of Functions with
Generalized Derivatives and Quasiconformal Mappings [in Russian], Nauka, Moscow (1983).
136. Griffel 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. Gurariı̆ 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,
561–588 (1990).
147. Halmos P. and Sunder V., Bounded Integral Operators on L2 Spaces, Springer-Verlag, New
York (1978).
148. Halperin I., Introduction to the Theory of Distributions, University of Toronto Press, Toronto
(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. 6–133.
151. Havin V. P. and Nikol0 skiı̆ 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. Hervé 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).
References 249

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 Hahn–Banach theorem,” Math. Intelligencer, 2,
No. 3, 123–125 (1980).
161. Hoffman K., Banach Spaces of Analytic Functions, Prentice-Hall, Englewood Cliffs (1962).
162. Hoffman 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. Hörmander L., Introduction to Complex Analysis in Several Variables, D. Van Nostrand
Company Inc., Princeton (1966).
166. Hörmander L., The Analysis of Linear Differential 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. Ioffe 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. 911–
937 (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. Jörgens 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).
250 References

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. Khelemskiı̆ 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, 795–798 (1989).
197. Knyazev P. N., Functional Analysis [in Russian], Vysheı̆shaya 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. Körner T. W., Fourier Analysis, Cambridge University Press, Cambridge (1988).
202. Korotkov V. B., Integral Operators [in Russian], Nauka, Novosibirsk (1983).
203. Köthe 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. Krasnosel0 skiı̆ M. A., Positive Solutions of Operator Equations, P.Noordhoff Ltd., Groningen
(1964).
206. Krasnosel0 skiı̆ M. A., Lifshits E. A., and Sobolev A. V., Positive Linear Systems. The
Method of Positive Operators [in Russian], Nauka, Moscow (1985).
207. Krasnosel0 skiı̆ M. A. and Rutitskiı̆ Ya. B., Convex Functions and Orlicz Spaces, Noordhoff,
Groningen (1961).
208. Krasnosel0 skiı̆ M. A. and Zabreı̆ko P. P., Geometric Methods of Nonlinear Analysis, Springer-
Verlag, Berlin (1984).
209. Krasnosel0 skiı̆ M. A. et al., Integral Operators in the Spaces of Summable Functions, No-
ordhoff International Publishing, Leyden (1976).
210. Kreı̆n S. G., Linear Differential Equations in Banach Space [in Russian], Nauka, Moscow
(1967).
211. Kreı̆n S. G., Linear Equations in Banach Space [in Russian], Nauka, Moscow (1971).
212. Kreı̆n S. G., Petunin Yu. I., and Semënov 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., Subdifferentials: 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).
References 251

220. Ladyzhenskaya O. A., Boundary Value Problems of Mathematical Physics [in Russian],
Nauka, Moscow (1973).
221. Lang S., Introduction to the Theory of Differentiable 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. Lévy 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 Differentiable 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 Russian],
Vishcha Shkola, Khar0 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 Differentiable 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 Zitný 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, 3–47 (1989).
245. Maslov V. P., Operator Methods [in Russian], Nauka, Moscow (1973).
246. Maurin K., Methods of Hilbert Space, Polish Scientific Publishers, Warsaw (1967).
247. Maurin K., Analysis. Vol. 2: Integration, Distributions, Holomorphic Functions, Tensor and
Harmonic Analysis, Polish Scientific Publishers, Warsaw (1980).
248. Maz0 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 Differential Equations [in Russian], Vysshaya Shkola, Moscow
(1977).
252 References

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 Differential 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. 361–387.
259. Naı̆mark M. A., Normed Rings, Noordhoff, 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. Nikol0 skiı̆ N. K., Lectures on the Shift Operator [in Russian], Nauka, Moscow (1980).
267. Nikol0 skiı̆ S. M., Approximation to Functions in Several Variables and Embedding Theorems
[in Russian], Nauka, Moscow (1977).
268. Nirenberg L., Topics in Nonlinear Functional Analysis. 1973–1974 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 Cliffs (1979).
270. Palais R. (with contributions by M. F. Atiyah et al.), Seminar on the Atiyah–Singer Index
Theorem, Princeton University Press, Princeton (1965).
271. Palamodov V. P., Linear Differential Operators with Constant Coefficients, Springer-Verlag,
Berlin etc. (1970).
272. Pedersen G. K., Analysis Now, Springer-Verlag, New York etc. (1989).
273. Pe␣lczyński 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 Differentiability,
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. Prössdorf 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. Raı̆kov D. A., Vector Spaces [in Russian], Fizmatgiz, Moscow (1962).
References 253

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 Real 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 Szökefalvi-Nagy B., Lectures on Functional Analysis [in French], Akadémi
Kiadó, 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], Pánstwowe 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. Sadovnichiı̆ 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. Samuélidés M. and Touzillier L., Functional Analysis [in French], Toulouse, Cepadues Éditions
(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. 135–141 (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 Scientific 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).
254 References

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.,
359–381 (1973).
325. Steen L. A. and Seebach J. A., Counterexamples in Topology, Springer-Verlag, Berlin etc.
(1978).
326. Stein E. M., Singular Integrals and Differentiability 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, 89–108 (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. 5–101.
340. Trenogin V. A., Functional Analysis [in Russian], Nauka, Moscow (1980).
341. Trenogin V. A., Pisarevskiı̆ B. M., and Soboleva T. S., Problems and Exercises in Functional
Analysis [in Russian], Nauka, Moscow (1984).
342. Tréves F., Locally Convex Spaces and Linear Partial Differential Equations, Springer-Verlag,
Berlin etc. (1967).
343. Triebel H., Theory of Function Spaces, Birkhäuser, Basel (1983).
344. Uspenskiı̆ S. V., Demidenko G. V., and Perepëlkin V. G., Embedding Theorems and Appli-
cations to Differential Equations [in Russian], Nauka, Novosibirsk (1984).
345. Vaı̆nberg 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-Differential 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, Noordhoff, Groningen
(1967).
References 255

353. Waelbroeck L., Topological Vector Spaces and Algebras, Springer-Verlag, Berlin etc. (1971).
354. Wagon S., The Banach–Tarski 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 Differentiable 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 Differential 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 J C A, 11.4.1
/, ., ix K(E), 10.9.1
Ar , 11.1.6 K(Q), 10.9.1
A × B, 1.1.1 K(Š), 10.9.1
Bp , 5.1.1, 5.2.11 LA , 11.1.6
◦ Lp , 5.5.9 (4), 5.5.9 (6)
B p , 5.1.1
Lp (X), 5.5.9
BT , 5.1.3
LQ0 , 10.8.4 (3)
BX , 5.1.10, 5.2.11
B(X), 5.6.4, L , 10.8.4 (4)
Q0
B(E , F ), 5.5.9 (2) L∞ , 5.5.9 (5)
B(X, Y ), 5.1.10 (7) M (Š), 10.9.4 (2)
C(Q, F ), 4.6.8 N (a), 11.9.1
C (m) , 10.9.9 Np , 5.5.9 (6)
C∞ (Š), 10.10.2 (3) PH0 , 6.2.7
Dα , 10.11.13 Pσ , 8.2.10
F −1 , 1.1.3 (1) PX1 ||X2 , 2.2.9 (4)
F (B ), 1.3.5 (1) P1 ⊥ P2 , 6.2.12
Fp , 5.5.9 (6) R (a, λ), 11.2.1
F |U , 1.1.3 (5) R (T, λ), 5.6.13
F (U ), 1.1.3 (5) S(A), 11.9.1
F (a1 , · ), 1.1.3 (6) T 0 , 7.6.2
F ( · , a2 ), 1.1.3 (6) T ∗ , 6.4.4
F ( · , · ), 1.1.3 (6) kT k, 5.1.10 (7)
F (X, Y ), 8.3.6 U ◦ , 10.5.7
U ⊥ , 6.2.5
G
b , 10.11.2
U ∈ (€ ), 3.1.1
G
b , 10.11.2 hX |, 10.3.1
b
G ◦ F , 1.1.4 X 0 , 5.1.10 (8), 10.2.11
H∗ , 6.1.10 (3) X 00 , 5.1.10 (8)
H(K), 8.1.13 X∗ , 2.1.4 (2)
H€ (U ), 3.1.11 X+ , 3.2.5
IC , 8.2.10 X0⊥ , 7.6.8
IU , 1.1.3 (3) Xσ , 8.2.10
J(q), 11.5.3 X N , 2.1.4 (4)
J(Q0 ), 11.5.2 X # , 2.2.4
Notation Index 257

XR , 3.7.1 M (Š), 10.9.3


X „ , 2.1.4 (4) N (µ), 10.8.11
X = X1 ⊕ X2 , 2.1.7 Np (f ), 5.5.9 (4)
X ⊕ iX, 8.4.8 N∞ , 5.5.9 (5)
X1 × X2 × . . . × XN , 2.1.4 (4) P (X), 1.2.3 (4)
(X, τ )0 , 10.2.11 RT , 8.2.1
X/X0 , 2.1.4 (6) Ra h, 11.3.1
(X/X0 , pX/X0 ), 5.1.10 (5) S (RN ), 10.11.6
X ↔ Y , 10.3.3 S 0 (RN ), 10.11.16
X ' Y , 2.2.6 T (X), 9.1.2
| Y i, 10.3.1 Up , 5.2.2
B, 9.6.14 UM , 5.2.4
C, 2.1.2 UX , 4.1.5, 5.2.4
⊥ X , 7.6.8
D, 8.1.3 0
F, 2.1.2 F u, 10.11.19
N, 1.2.16 M , 5.3.9
Q, 7.4.11 M ∼ N , 5.3.1
R, 2.1.2 M  N , 5.3.1
R· , 3.4.1 MX , 5.1.6
R+ , 3.1.2 (4) Mτ , 10.2.7
N ◦ T , 5.1.10 (3)
R, 3.8.1
NT , 5.1.10 (3)
Re, 3.7.3 Ra , 11.8.7
Re−1 , 3.7.4 Cl(τ ), 4.1.15, 9.1.4
T, 8.1.3 Im f , 5.5.9 (4)
Z, 8.5.1 Inv(A), 11.1.5
Z+ , 10.10.2 (2) Inv(X, Y ), 5.6.12
Ae , 11.1.2 ƒB , 8.1.2 (4)
D (Q), 10.10.1 Lat(X), 2.1.5
D (Š), 10.10.1 LCT (X), 10.2.3
D 0 (Š), 10.10.4 M(A), 11.6.6
D 0F (Š), 10.10.8 Op(τ ), 4.1.11, 9.1.4
D (m) (Q), 10.10.8 Re, 2.1.2
D (m) (Š), 10.10.8 Re f , 5.5.9 (4)
D (m) (Š)0 , 10.10.8 Sp(a), 11.2.1
E (Š), 10.10.2 (3) SpA (a), 11.2.1
E 0 (RN ), 10.10.5 (9) Sp(T ), 5.6.13
E ◦ T , 2.2.8 T1 , 9.3.2
F , 10.11.4 T2 , 9.3.5
Fp , 5.5.9 (6) T3 , 9.3.9
F r(X, Y ), 8.5.1 T31 /2 , 9.3.15
F (X), 1.3.6 T4 , 9.3.11
GA , 11.6.8 T(X), 9.1.7
H (K), 8.1.14 Tr (Š), 10.10.2
K (X), 8.3.3 VT(X), 10.1.5
K (X, Y ), 6.6.1 X(A), 11.6.4
L (X), 2.2.8 δ, 10.9.4 (1)
L (X, Y ), 2.2.3 δ (−1) , 10.10.5 (4)
Lr (X, Y ), 3.2.6 (3) δq , 10.9.4 (1)
L∞ , 5.5.9 (5) µ∗ , 10.9.4 (3)
258 Notation Index

µ+ , 10.8.13 aµ, 10.8.15


µ− , 10.8.13 a τ f , 10.9.4 (1)
|µ|, 10.8.13, 10.9.4 (3) (a, b)s , 11.9.9
kµk, 10.9.5 c, 3.3.1 (2), 5.5.9 (3)
µŠ0 , 10.9.4 (4) c (E , F ), 5.5.9 (3)
µ1 ⊗ µ2 , 10.9.4 (6) c0 , 5.5.9 (3)
µ1 × µ2 , 10.9.4 (6) c0 (E ), 5.5.9
µ ∗ f , 10.9.4 (7) c0 (E , F ), 5.5.9 (3)
µ ∗ ν, 10.9.4 (7) ∂ α u, 10.10.5 (4)
π(U ), 10.5.1, 10.5.7 ∂ (p), 3.5.2 (1)
π −1 (V ), 10.5.1 |∂ |(p), 3.7.8
−1
πF (πF (U )), 10.5.5 ∂U , 4.1.13
2π , 10.11.4 ∂ x (f ), 3.5.1
σ 0 , 8.2.9 dp , 5.2.1
σ(T ), 5.6.13 dx, 10.9.9
σ(X, Y ), 10.3.5 e, 10.9.4 (1), 11.1.1
τ (X, Y ), 10.4.4
f , 10.11.3
b
τa f , 10.9.4 (1)
f (a), 11.3.1
τM , 5.2.8
{f < t}, 3.8.1
κσ , 8.2.10 {f = t}, 3.8.1
abs pol , 10.5.7
{f ≤ t}, 3.8.1
cl U , 4.1.13
f (T ), 8.2.1
co(U ), 3.1.14
f e, 10.10.5 (9)
codim X, 2.2.9 (5)
f µ, 10.9.4 (3)
coim T , 2.3.1
f ∗ , 10.9.4 (3)
coker T , 2.3.1
f u, 10.10.5 (7)
core U , 3.4.11
fn  f , 10.10.7 (3)
diam U , 4.5.3
dim X, 2.2.9 (5) fn  0, 10.9.8
K
dom f , 3.4.2
bg, 10.11.2
b
dom F , 1.1.2 ◦
0 = 0 OO , 3.4.2 g(f ), 8.2.6
ext V , 3.6.1 hhi, 6.3.5
fil B , 1.3.3 lp , lp (E ), 5.5.9 (4)
fr U , 4.1.13 l∞ , l∞ (E ), 5.5.9 (2)
im F , 1.1.2 m, 5.5.9 (2)
inf U , 1.2.9 p  q, 5.3.3
int U , 4.1.13 pe , 5.5.9 (5)
ker T , 2.3.1 pS , 3.8.6
lin(U ), 3.1.14 p T , 5.1.3
pol , 10.5.7 pX/X0 , 5.1.10 (5)
rank T , 8.5.7 (2) r(T ), 5.6.6
res(a), 11.2.1 s, Ex. 1.19
res(T ), 5.6.13 tα , 10.11.5 (8)
seg, 3.6.1 ug , 10.10.5 (1)
sup U , 1.2.9 u∗ , 10.10.5 (5)
supp(f ), 9.6.4 u ∗ f , 10.10.5 (9)
supp(µ), 10.8.11, 10.9.4 (5) u1 ⊗ u2 , 10.10.5 (8)
supp(u), 10.10.5 (6) u1 × u2 , 10.10.5 (8)
a, 11.6.8
b hx |, 10.3.1
Notation Index 259

x0 , 6.4.1 |||y|||p , 5.5.9 (6)


x00 , 5.1.10 (8) k · k, 5.1.9
xα , 10.11.5 (8) k · kn,Q , 10.10.2 (2)
x+ , 3.2.12 k · k∞ , 5.5.9 (5)
x− , 3.2.12 k · kX , 5.1.9
|x|, 3.2.12 k · | Xk, 5.1.9
kxkp , 5.5.9 (4) | ·i, 10.3.1
kxk∞ , 5.5.9 (2) 1, 5.3.10, 10.8.4 (6)
2X , 1.2.3 (4)
∼ (x), 10.11.4
∗, 6.4.13
∼X0 , 2.1.4
P (6)
x := x , 5.5.9 (7)
b
R , 10.9.1
e∈E e , 5.5.9 (6)
0
x 7→ x , 6.4.1 E
h· | ·i, 10.3.1
x1 ∨ x2 , x1 ∧ x2 , 1.2.12
h· |, 10.3.1
(x1 , x2 ), 1.2.12 ∼,
hx | f i, 5.1.11 P 1.2.2
ξ∈„ ξ
X , 2.1.4 (5)
x ≤σ y, 1.2.2 Q
X , 2.1.4 (4)
x0 ⊗ y, 5.5.6 H ξ∈„ ξ
x ⊥ y, 6.2.5 h(z)R(z)dz, 8.1.20
| yi, 10.3.1 b, 11.6.8
Subject Index

Absolute Bipolar Theorem, 10.5.9, 178 algebraically interior point, 3.4.11, 28


absolute concept, 9.4.7, 157 algebraically isomorphic spaces,
absolute polar, 10.5.7, 178 2.2.6, 13
absolutely continuous measure, algebraically reflexive space, Ex. 2.8, 19
10.9.4 (3), 190 ambient space, 2.1.4 (3), 11
absolutely convex set, 3.1.2 (6), 20 annihilator, 7.6.8, 116
absolutely fundamental family antidiscrete topology, 9.1.8 (3), 147
of vectors, 5.5.9 (7), 73 antisymmetric relation, 1.2.1, 3
absorbing set, 3.4.9, 28 antitone mapping, 1.2.3, 4
addition in a vector space, 2.1.3, 10 approximate inverse, 8.5.9, 139
adherence of a filterbase, 9.4.1, 155 approximately invertible operator,
adherent point, 4.1.13, 41 8.5.9, 139
adherent point of a filterbase, 9.4.1, 155 approximation property, 8.3.10, 133
adjoint diagram, 6.4.8, 92
approximation property in Hilbert
adjoint of an operator, 6.4.5, 91 space, 6.6.10, 98
adjunction of unity, 11.1.2, 213
arc, 4.8.2, 54
affine hull, 3.1.14, 22
affine mapping, 3.1.7, 21 Arens multinorm, 8.3.8, 133
affine operator, 3.4.8 (4), 28 ascent, Ex. 8.10, 144
affine variety, 3.1.2 (5), 20 Ascoli–ArzelΓ a Theorem, 4.6.10, 50
agreement condition, 10.9.4 (4), 190 assignment operator, ix
Akilov Criterion, 10.5.3, 178 associate seminorm, 6.1.7, 81
Alaoglu–Bourbaki Theorem, 10.6.7, 180 associated Hausdorff pre-Hilbert
Alexandroff compactification, space, 6.1.10 (4), 83
9.4.22, 159 associated Hilbert space,
algebra, 5.6.2, 73 6.1.10 (4), 83
algebra of bounded operators, 5.6.5, 74 associated multinormed space,
algebra of germs of holomorphic 10.2.7, 172
functions, 8.1.18, 125 associated topology, 9.1.12, 148
algebraic basis, 2.2.9 (5), 14 associativity of least upper bounds,
algebraic complement, 2.1.7, 12 3.2.10, 24
algebraic dual, 2.2.4, 13 asymmetric balanced Hahn–Banach
algebraic isomorphism, 2.2.5, 13 formula, 3.7.10, 34
algebraic subdifferential, 7.5.8, 113 asymmetric Hahn–Banach formula,
algebraically complementary subspace, 3.5.5, 30
2.1.7, 12 Atkinson Theorem, 8.5.18, 141
Subject Index 261

Automatic Continuity Principle, bounded below, 3.2.9, 23


7.5.5, 112 bounded endomorphism algebra,
automorphism, 10.11.4, 203 5.6.5, 74
Bounded Index Stability Theorem,
Baire Category Theorem, 4.7.6, 52 8.5.21, 142
Baire space, 4.7.2, 52 bounded operator, 5.1.10 (7), 59
Balanced Hahn–Banach Theorem, bounded Radon measure,
3.7.13, 35 10.9.4 (2), 189
Balanced Hahn–Banach Theorem bounded set, 5.4.3, 66
in a topological setting, 7.5.10, 113 boundedly order complete lattice,
balanced set, 3.1.2 (7), 20 3.2.8, 23
balanced subdifferential, 3.7.8, 34 Bourbaki Criterion, 4.4.7, 46; 9.4.4, 156
Balanced Subdifferential Lemma, bracketing of vector spaces, 10.3.1, 173
3.7.9, 34 bra-functional, 10.3.1, 173
ball, 9.6.14, 166 bra-mapping, 10.3.1, 173
Banach algebra, 5.6.3, 74 bra-topology, 10.3.5, 174
Banach Closed Graph Theorem, B-stable, 10.1.8, 171
7.4.7, 108 bump function, 9.6.19, 167
Banach Homomorphism Theorem,
7.4.4, 108 Calkin algebra, 8.3.5, 132
Banach Inversion Stability Theorem, Calkin Theorem, 8.3.4, 132
5.6.12, 76 canonical embedding, 5.1.10 (8), 59
Banach Isomorphism Theorem, canonical exact sequence,
7.4.5, 108 2.3.5 (6), 16
Banach range, 7.4.18, 111 canonical operator representation,
Banach space, 5.5.1, 66 11.1.7, 214
Banach’s Fundamental Principle, canonical projection, 1.2.3 (4), 4
7.1.5, 101 Cantor Criterion, 4.5.6, 47
Banach’s Fundamental Principle for Cantor Theorem, 4.4.9, 46
a Correspondence, 7.3.7, 107 cap, 3.6.3 (4), 32
Banach–Steinhaus Theorem, 7.2.9, 104 Cauchy–BunyakovskiΦ ß–Schwarz
barrel, 10.10.9 (1), 199 inequality, 6.1.5, 80
barreled normed space, 7.1.8, 102 Cauchy filter, 4.5.2, 47
barreled space, 10.10.9 (1), 199 Cauchy net, 4.5.2, 47
base for a filter, 1.3.3, 6 Cauchy–Wiener Integral Theorem,
basic field, 2.1.2, 10 8.1.7, 122
Bessel inequality, 6.3.7, 88 centralizer, 11.1.6, 214
best approximation, 6.2.3, 84 chain, 1.2.19, 6
Beurling–Gelfand formula, character group, 10.11.2, 203
8.1.12 (2), 124 character of a group algebra,
bilateral ideal, 8.3.3, 132; 11.6.2, 220 10.11.1 (1), 201
bilinear form, 6.1.2, 80 character of an algebra, 11.6.4, 221
bipolar, 10.5.5, 178 character space of an algebra,
Bipolar Theorem, 10.5.8, 178 11.6.4, 221
Birkhoff Theorem, 9.2.2, 148 characteristic function, 5.5.9 (6), 72
Bochner integral, 5.5.9 (6), 72 charge, 10.9.4 (3), 190
bornological space, 10.10.9 (3), 199 ChebyshΞ ev metric, 4.6.8, 50
boundary of an algebra, Ex. 11.8, 234 classical Banach space, 5.5.9 (5), 71
boundary of a set, 4.1.13, 41 clopen part of a spectrum, 8.2.9, 130
boundary point, 4.1.13, 41 closed ball, 4.1.3, 40
bounded above, 1.2.19, 6 closed convex hull, 10.6.5, 179
262 Subject Index

closed correspondence, 7.3.8, 107 completely regular space, 9.3.15, 155


closed cylinder, 4.1.3, 40 completion, 4.5.13, 49
closed-graph correspondence, 7.3.9, 107 complex conjugate, 2.1.4 (2), 10
closed halfspace, Ex. 3.3, 39 complex distribution, 10.10.5 (5), 196
closed linear span, 10.5.6, 178 complex plane, 8.1.3, 121
closed set, 9.1.4, 146 complex vector space, 2.1.3, 10
closed set in a metric space, 4.1.11, 41 complexification, 8.4.8, 136
closure of a set, 4.1.13, 41 complexifier, 3.7.4, 34
closure operator, Ex. 1.11, 8 composite correspondence, 1.1.4, 2
coarser cover, 9.6.1, 164 Composite Function Theorem,
coarser filter, 1.3.6, 7 8.2.8, 129
coarser pretopology, 9.1.2, 146 composition, 1.1.4, 2
codimension, 2.2.9 (5), 14 Composition Spectrum Theorem,
codomain, 1.1.2, 1 5.6.22, 78
cofinite set, Ex. 1.19, 9 cone, 3.1.2 (4), 20
coimage of an operator, 2.3.1, 15 conical hull, 3.1.14, 22
coincidence of the algebraic and conical segment, 3.1.2 (9), 20
topological subdifferentials, conical slice, 3.1.2 (9), 20
7.5.8, 113 conjugate distribution, 10.10.5 (5), 196
coinitial set, 3.3.2, 25 conjugate exponent, 5.5.9 (4), 69
cokernel of an operator, 2.3.1, 15 conjugate-linear functional, 2.2.4, 13
comeager set, 4.7.4, 52 conjugate measure, 10.9.4 (3), 189
commutative diagram, 2.3.3, 15 connected elementary compactum,
Commutative Gelfand–NaΦ ßmark 4.8.5, 54
Theorem, 11.8.4, 227 connected set, 4.8.4, 54
compact convergence, 7.2.10, 105 constant function, 5.3.10, 64;
Compact Index Stability Theorem, 10.8.4 (6), 182
8.5.20, 142 Continuous Extension Principle,
compact-open topology, 8.3.8, 133 7.5.11, 113
compact operator, 6.6.1, 95 continuous function at a point, 4.2.2,
compact set, 9.4.2, 155 43; 9.2.5, 149
compact set in a metric space, 4.4.1, 46 Continuous Function Recovery
compact space, 9.4.7, 157 Lemma, 9.3.12, 153
compact topology, 9.4.7, 157 continuous functional calculus,
compactly-supported distribution, 11.8.7, 228
10.10.5 (6), 196 continuous mapping of a metric
compactly-supported function, space, 4.2.2, 43
9.6.4, 165 continuous mapping of a topological
compactum, 9.4.17, 158 space, 9.2.4, 149
compatible topology, 10.4.1, 175 continuous partition of unity, 9.6.6, 166
complementary projection, 2.2.9 (4), 14 contour integral, 8.1.20, 125
complementary subspace, 7.4.9, 109 conventional summation, 5.5.9 (4), 70
Complementation Principle, 7.4.10, 109 convergent filterbase, 4.1.16, 42
complemented subspace, 7.4.9, 108 convergent net, 4.1.17, 42
complement of an orthoprojection, convergent sequence space, 3.3.1 (2), 25
6.2.10, 85 convex combination, 3.1.14, 22
complement of a projection, convex correspondence, 3.1.7, 21
2.2.9 (4), 14 convex function, 3.4.4, 27
complete lattice, 1.2.13, 5 convex hull, 3.1.14, 22
complete metric space, 4.5.5, 47 convex set, 3.1.2 (8), 20
complete set, 4.5.14, 49 convolution algebra, 10.9.4 (7), 190
Subject Index 263

convolution of a measure and descent, Ex. 8.10, 144


a function, 10.9.4 (7), 191 diagonal, 1.1.3 (3), 2
convolution of distributions, diagram prime, 7.6.5, 115
10.10.5 (9), 197 Diagram Prime Principle, 7.6.7, 115
convolution of functions, 9.6.17, 167 diagram star, 6.4.8, 92
convolution of measures, Diagram Star Principle, 6.4.9, 92
10.9.4 (7), 190 diameter, 4.5.3, 47
convolutive distributions, DieudonnΓ e Lemma, 9.4.18, 158
10.10.5 (9), 197 dimension, 2.2.9 (5), 14
coordinate projection, 2.2.9 (3), 13 Dini Theorem, 7.2.10, 105
coordinatewise operation, Dirac measure, 10.9.4 (1), 188
2.1.4 (4), 11 direct polar, 7.6.8, 116; 10.5.1, 177
core, 3.4.11, 28 direct sum decomposition, 2.1.7, 12
correspondence, 1.1.1, 1 direct sum of vector spaces, 2.1.4 (5), 11
correspondence in two arguments, directed set, 1.2.15, 6
1.1.3 (6), 2 direction, 1.2.15, 6
correspondence onto, 1.1.3 (3), 2 directional derivative, 3.4.12, 28
coset, 1.2.3 (4), 4 discrete element, 3.3.6, 26
coset mapping, 1.2.3 (4), 4 Discrete KreΦ ßn–Rutman Theorem,
countable convex combination, 3.3.8, 26
7.1.3, 101 discrete topology, 9.1.8 (4), 147
Countable Partition Theorem, disjoint measures, 10.9.4 (3), 190
9.6.20, 167 disjoint sets, 4.1.10, 41
countable sequence, 1.2.16, 6 distance, 4.1.1, 40
countably normable space, 5.4.1, 64 distribution, 10.10.4, 195
cover of a set, 9.6.1, 164 distribution applies to a function,
C ∗ -algebra, 6.4.13, 92 10.10.5 (7), 196
C ∗ -subalgebra, 11.7.8, 225 Distribution Localization Principle,
10.10.12, 200
Davis–Figiel–Szankowski distribution of finite order, 10.10.5 (3),
Counterexample, 8.3.14, 134 195
de Branges Lemma, 10.8.16, 185 distribution size at most m,
decomplexification, 6.1.10 (2), 83 10.10.5 (3), 195
decomposition reduces an operator, distribution of slow growth,
2.2.9 (4), 14 10.11.16, 209
decreasing mapping, 1.2.3, 4 distributions admitting convolution,
Dedekind complete vector lattice, 10.10.5 (9), 197
3.2.8, 23 distributions convolute, 10.10.5 (9), 197
deficiency, 8.5.1, 137 division algebra, 11.2.3, 215
definor, ix domain, 1.1.2, 1
delta-function, 10.9.4 (1), 188 Dominated Extension Theorem,
delta-like sequence, 9.6.15, 166 3.5.4, 30
δ-like sequence, 9.6.15, 166 Double Prime Lemma, 7.6.6, 115
δ-sequence, 9.6.15, 166 double prime mapping, 5.1.10 (8), 59
dense set, 4.5.10, 48 double sharp, Ex. 2.7, 19
denseness, 4.5.10, 48 downward-filtered set, 1.2.15, 6
density of a measure, 10.9.4 (3), 190 dual diagram, 7.6.5, 115
derivative in the distribution sense, dual group, 10.11.2, 203
10.10.5 (4), 196 dual norm of a functional, 5.1.10 (8), 59
derivative of a distribution, 10.10.5 (4), dual of a locally convex space,
196 10.2.11, 173
264 Subject Index

dual of an operator, 7.6.2, 114 exterior point, 4.1.13, 41


duality bracket, 10.3.3, 174 Extreme and Discrete Lemma, 3.6.4, 32
duality pair, 10.3.3, 174 extreme point, 3.6.1, 31
dualization, 10.3.3, 174 extreme set, 3.6.1, 31
Dualization Theorem, 10.3.9, 175
Dunford–Hille Theorem, 8.1.3, 121 face, 3.6.1, 31
Dunford Theorem, 8.2.7 (2), 129 factor set, 1.2.3 (4), 4
Dvoretzky–Rogers Theorem, faithful representation, 8.2.2, 126
5.5.9 (7), 73 family, 1.1.3 (4), 2
dyadic-rational point, 9.3.13, 154 filter, 1.3.3, 6
filterbase, 1.3.1, 6
effective domain of definition, 3.4.2, 27 finer cover, 9.6.1, 164
Eidelheit Separation Theorem, finer filter, 1.3.6, 7
3.8.14, 39 finer multinorm, 5.3.1, 62
eigenvalue, 6.6.3 (4), 95 finer pretopology, 9.1.2, 146
eigenvector, 6.6.3, 95 finer seminorm, 5.3.3, 63
element of a set, 1.1.3 (4), 2 finest multinorm, 5.1.10 (2), 58
elementary compactum, 4.8.5, 54 finite complement filter, 5.5.9 (3), 68
endomorphism, 2.2.1, 12; 8.2.1, 126 finite descent, Ex. 8.10, 144
endomorphism algebra, 2.2.8, 13; finite-rank operator, 6.6.8, 97; 8.3.6, 132
5.6.5, 74 finite-valued function, 5.5.9 (6), 72
endomorphism space, 2.2.8, 13 first category set, 4.7.1, 52
Enflo counterexample, 8.3.12, 134 first element, 1.2.6, 5
entourage, 4.1.5, 40 fixed point, Ex. 1.11, 8
envelope, Ex. 1.11, 8 flat, 3.1.2 (5), 20
epigraph, 3.4.2, 26 formal duality, 2.3.15, 18
epimorphism, 2.3.1, 15 Fourier coefficient family, 6.3.15, 89
ε-net, 8.3.2, 132 Fourier–Plancherel transform,
ε-perpendicular, 8.4.1, 134 10.11.15, 209
ε-Perpendicular Lemma, 8.4.1, 134 Fourier–Schwartz transform,
Equicontinuity Principle, 7.2.4, 103 10.11.19, 211
equicontinuous set, 4.2.8, 44 Fourier series, 6.3.16, 89
equivalence, 1.2.2, 3 Fourier transform of a distribution,
equivalence class, 1.2.3 (4), 4 10.11.19, 211
equivalent multinorms, 5.3.1, 62 Fourier transform of a function,
equivalent seminorms, 5.3.3, 63 10.11.3, 203
estimate for the diameter of a spherical Fourier transform relative to a basis,
layer, 6.2.1, 84 6.3.15, 89
Euler identity, 8.5.17, 141 Fr∆echet space, 5.5.2, 66
evaluation mapping, 10.3.4 (3), 174 Fredholm Alternative, 8.5.6, 138
everywhere-defined operator, 2.2.1, 12 Fredholm index, 8.5.1, 137
everywhere dense set, 4.7.3 (3), 52 Fredholm operator, 8.5.1, 137
exact sequence, 2.3.4, 15 Fredholm Theorem, 8.5.8, 139
exact sequence at a term, 2.3.4, 15 frontier of a set, 4.1.13, 41
exclave, 8.2.9, 130 from A into/to B, 1.1.1, 1
expanding mapping, Ex. 4.14, 55 Fubini Theorem for distributions,
extended function, 3.4.2, 26 10.10.5 (8), 197
extended real axis, 3.8.1, 35 Fubini Theorem for measures,
extended reals, 3.8.1, 35 10.9.4 (6), 190
extension of an operator, 2.3.6, 16 full subalgebra, 11.1.5, 213
exterior of a set, 4.1.13, 41 fully norming set, 8.1.1, 120
Subject Index 265

Function Comparison Lemma, 3.8.3, 36 gradient mapping, 6.4.2, 90


function of class C (m) , 10.9.9, 192 Gram–Schmidt orthogonalization
function of compact support, 9.6.4, 165 process, 6.3.14, 89
Function Recovery Lemma, 3.8.2, 35 graph norm, 7.4.17, 111
functor, 10.9.4 (4), 190 Graph Norm Principle, 7.4.17, 111
fundamental net, 4.5.2, 47 greatest element, 1.2.6, 5
fundamental sequence, 4.5.2, 47 greatest lower bound, 1.2.9, 5
fundamentally summable family Grothendieck Criterion, 8.3.11, 133
of vectors, 5.5.9 (7), 73 Grothendieck Theorem, 8.3.9, 133
ground field, 2.1.3, 10
gauge, 3.8.6, 37 ground ring, 2.1.1, 10
gauge function, 3.8.6, 37 group algebra, 10.9.4 (7), 191
Gauge Theorem, 3.8.7, 37 group character, 10.11.1, 202
-correspondence, 3.1.6, 21
-hull, 3.1.11, 21 Haar integral, 10.9.4 (1), 189
-set, 3.1.1, 20 Hahn–Banach Theorem, 3.5.3, 29
Gelfand–Dunford Theorem in Hahn–Banach Theorem in analytical
an operator setting, 8.2.3, 127 form, 3.5.4, 30
Gelfand–Dunford Theorem Hahn–Banach Theorem in geometric
in an algebraic setting, 11.3.2, 216 form, 3.8.12, 38
Gelfand formula, 5.6.8, 74 Hahn–Banach Theorem in subdifferential
Gelfand–Mazur Theorem, 11.2.3, 215 form, 3.5.4, 30
Gelfand–NaΦ ßmark–Segal construction, Hamel basis, 2.2.9 (5), 14
11.9.11, 232 Hausdorff Completion Theorem,
Gelfand Theorem, 7.2.2, 103 4.5.12, 48
Gelfand transform of an algebra, Hausdorff Criterion, 4.6.7, 50
11.6.8, 222 Hausdorff metric, Ex. 4.8, 55
Gelfand transform of an element, Hausdorff multinorm, 5.1.8, 57
11.6.8, 221 Hausdorff multinormed space, 5.1.8, 57
Gelfand Transform Theorem, Hausdorff space, 9.3.5, 152
11.6.9, 222 Hausdorff Theorem, 7.6.12, 117
general form of a compact operator Hausdorff topology, 9.3.5, 152
in Hilbert space, 6.6.9, 97 H-closed space, Ex. 9.10, 168
general form of a linear functional Heaviside function, 10.10.5 (4), 196
in Hilbert space, 6.4.2, 90 Hellinger–Toeplitz Theorem, 6.5.3, 93
general form of a weakly continuous hermitian element, 11.7.1, 224
functional, 10.3.10, 175 hermitian form, 6.1.1, 80
general position, Ex. 3.10, 39 hermitian operator, 6.5.1, 93
generalized derivative in the Sobolev hermitian state, 11.9.8, 230
sense, 10.10.5 (4), 196 Hilbert basis, 6.3.8, 88
Generalized Dini Theorem, 10.8.6, 183 Hilbert cube, 9.2.17 (2), 151
generalized function, 10.10.4, 195 Hilbert dimension, 6.3.13, 89
Generalized Riesz–Schauder Theorem, Hilbert identity, 5.6.19, 78
8.4.10, 137 Hilbert isomorphy, 6.3.17, 90
generalized sequence, 1.2.16, 6 Hilbert–Schmidt norm, Ex. 8.9, 144
Generalized Weierstrass Theorem, Hilbert–Schmidt operator,
10.9.9, 192 Ex. 8.9, 144
germ, 8.1.14, 124 Hilbert–Schmidt Theorem, 6.6.7, 96
GNS-construction, 11.9.11, 232 Hilbert space, 6.1.7, 81
GNS-Construction Theorem, Hilbert-space isomorphism, 6.3.17, 90
11.9.10, 231 Hilbert sum, 6.1.10 (5), 84
266 Subject Index


older inequality, 5.5.9 (4), 69 integral with respect to a measure,
holey disk, 4.8.5, 54 10.9.3, 188
holomorphic function, 8.1.4, 122 interior of a set, 4.1.13, 41
Holomorphy Theorem, 8.1.5, 122 interior point, 4.1.13, 41
homeomorphism, 9.2.4, 149 intersection of topologies, 9.1.14, 148
homomorphism, 7.4.1, 107 interval, 3.2.15, 24

ormander transform, Ex. 3.19, 39 Interval Addition Lemma, 3.2.15, 24
hyperplane, 3.8.9, 38 invariant subspace, 2.2.9 (4), 14
hypersubspace, 3.8.9, 38 inverse-closed subalgebra, 11.1.5, 213
inverse image of a multinorm,
ideal, 11.4.1, 217
5.1.10 (3), 58
Ideal and Character Theorem,
inverse image of a preorder, 1.2.3 (3), 4
11.6.6, 221
ideal correspondence, 7.3.3, 106 inverse image of a seminorm, 5.1.4, 57
Ideal Correspondence Lemma, inverse image of a set, 1.1.3 (5), 2
7.3.4, 106 inverse image of a topology, 9.2.9, 150
Ideal Correspondence Principle, inverse image of a uniformity,
7.3.5, 106 9.5.5 (3), 160
Ideal Hahn–Banach Theorem, 7.5.9, 113 inverse image topology, 9.2.9, 150
ideally convex function, 7.5.4, 112 Inverse Image Topology Theorem,
ideally convex set, 7.1.3, 101 9.2.8, 149
idempotent operator, 2.2.9 (4), 14 inverse of a correspondence, 1.1.3 (1), 1
identical embedding, 1.1.3 (3), 2 inverse of an element in an algebra,
identity, 10.9.4, 188 11.1.5, 213
identity element, 11.1.1, 213 Inversion Theorem, 10.11.12, 208
identity mapping, 1.1.3 (3), 2 invertible element, 11.1.5, 213
identity relation, 1.1.3 (3), 1 invertible operator, 5.6.10, 75
image, 1.1.2, 1 involution, 6.4.13, 92
image of a filterbase, 1.3.5 (1), 7 involutive algebra, 6.4.13, 92
image of a set, 1.1.3 (5), 2 irreducible representation, 8.2.2, 127
image of a topology, 9.2.12, 150 irreflexive space, 5.1.10 (8), 59
image topology, 9.2.12, 150 isolated part of a spectrum, 8.2.9, 130
Image Topology Theorem, isolated point, 8.4.7, 136
9.2.11, 150 isometric embedding, 4.5.11, 48
imaginary part of a function, isometric isomorphism of algebras,
5.5.9 (4), 69 11.1.8, 215
increasing mapping, 1.2.3 (5), 4 isometric mapping, 4.5.11, 48
independent measure, 10.9.4 (3), 190 isometric representation, 11.1.8, 214
index, 8.5.1, 137 isometric ∗-isomorphism, 11.8.3, 226
indicator function, 3.4.8 (2), 27 isometric ∗-representation, 11.8.3, 226
indiscrete topology, 9.1.8 (3), 147 isometry into, 4.5.11, 48
induced relation, 1.2.3 (1), 4 isometry onto, 4.5.11, 48
induced topology, 9.2.17 (1), 151 isomorphism, 2.2.5, 13
inductive limit topology, 10.9.6, 191 isotone mapping, 1.2.3, 4
inductive set, 1.2.19, 6
infimum, 1.2.9, 5 James Theorem, 10.7.5, 181
infinite-rank operator, 6.6.8, 97 Jensen inequality, 3.4.5, 27
infinite set, 5.5.9 (3), 68 join, 1.2.12, 5
inner product, 6.1.4, 80 Jordan arc, 4.8.2, 54
integrable function, 5.5.9 (4), 69 Jordan Curve Theorem, 4.8.3, 54
integral, 5.5.9 (4), 68 juxtaposition, 2.2.8, 13
Subject Index 267

Kakutani Criterion, 10.7.1, 180 linear set, 2.1.4 (3), 11


Kakutani Lemma, 10.8.7, 183 linear space, 2.1.4 (3), 11
Kakutani Theorem, 7.4.11 (3), 109 linear span, 3.1.14, 22
Kantorovich space, 3.2.8, 23 linear topological space, 10.1.3, 169
Kantorovich Theorem, 3.3.4, 25 linear topology, 10.1.3, 169
Kaplansky–Fukamija Lemma, linearly independent set, 2.2.9 (5), 14
11.9.7, 230 linearly-ordered set, 1.2.19, 6
Kato Criterion, 7.4.19, 111 Lions Theorem of Supports,
kernel of an operator, 2.3.1, 15 10.10.5 (9), 197
ket-mapping, 10.3.1, 173 Liouville Theorem, 8.1.10, 123
ket-topology, 10.3.5, 174 local data, 10.9.11, 193
Kolmogorov Normability Criterion, locally compact group, 10.9.4 (1), 188
5.4.5, 66 locally compact space, 9.4.20, 159
KreΦßn–Milman Theorem, 10.6.5, 179 locally compact topology, 9.4.20, 159
KreΦßn–Milman Theorem in subdifferential locally convex space, 10.2.9, 172
form, 3.6.5, 33 locally convex topology, 10.2.1, 171
KreΦßn–Rutman Theorem, 3.3.5, 26 locally finite cover, 9.6.2, 164
Krull Theorem, 11.4.8, 219 locally integrable function, 9.6.17, 167
Kuratowski–Zorn Lemma, 1.2.20, 6 locally Lipschitz function, 7.5.6, 112
K-space, 3.2.8, 23 loop, 4.8.2, 54
K-ultrametric, 9.5.13, 162 lower bound, 1.2.4, 5
lower limit, 4.3.5, 45
last element, 1.2.6, 5 lower right Dini derivative, 4.7.7, 53
lattice, 1.2.12, 5 lower semicontinuous, 4.3.3, 44
lear trap map, 3.7.4, 34 L2 -Fourier transform, 10.11.15, 209
least element, 1.2.6, 5
Lebesgue measure, 10.9.4 (1), 189 Mackey–Arens Theorem, 10.4.5, 176
Lebesgue set, 3.8.1, 35 Mackey Theorem, 10.4.6, 176
Lefschetz Lemma, 9.6.3, 165 Mackey topology, 10.4.4, 176
left approximate inverse, 8.5.9, 139 mapping, 1.1.3 (3), 1
left Haar measure, 10.9.4 (1), 189 massive subspace, 3.3.2, 25
left inverse of an element in an algebra, matrix form, 2.2.9 (4), 14
11.1.3, 213 maximal element, 1.2.10, 5
lemma on continuity of a convex maximal ideal, 11.4.5, 218
function, 7.5.1, 112 maximal ideal space, 11.6.7, 221
lemma on the numeric range Maximal Ideal Theorem, 11.5.3, 220
of a hermitian element, 11.9.3, 229 Mazur Theorem, 10.4.9, 177
level set, 3.8.1, 35 meager set, 4.7.1, 52
Levy Projection Theorem, 6.2.2, 84 measure, 10.9.3, 188
limit of a filterbase, 4.1.16, 42 Measure Localization Principle,
Lindenstrauss space, 5.5.9 (5), 71 10.9.10, 192
Lindenstrauss–Tzafriri Theorem, measure space, 5.5.9 (4), 69
7.4.11 (3), 110 meet, 1.2.12, 5
linear change of a variable under the member of a set, 1.1.3 (4), 2
subdifferential sign, 3.5.4, 30 metric, 4.1.1, 40
linear combination, 2.3.12, 17 metric space, 4.1.1, 40
linear correspondence, 2.2.1, 12; metric topology, 4.1.9, 41
3.1.7, 21 metric uniformity, 4.1.5, 40
linear functional, 2.2.4, 13 Metrizability Criterion, 5.4.2, 64
linear operator, 2.2.1, 12 metrizable multinormed space, 5.4.1, 64
linear representation, 8.2.2, 126 minimal element, 1.2.10, 5
268 Subject Index

Minimal Ideal Theorem, 11.5.1, 219 net having a subnet, 1.3.5 (2), 7
Minkowski–Ascoli–Mazur Theorem, net lacking a subnet, 1.3.5 (2), 7
3.8.12, 38 Neumann series, 5.6.9, 75
Minkowski functional, 3.8.6, 37 Neumann Series Expansion Theorem,
Minkowski inequality, 5.5.9 (4), 69 5.6.9, 75
minorizing set, 3.3.2, 25 neutral element, 2.1.4 (3), 11;
mirror, 10.2.7, 172 10.9.4, 188
module, 2.1.1, 10 Nikol 0skiΦ
ß Criterion, 8.5.22, 143
modulus of a scalar, 5.1.10 (4), 58 Noether Criterion, 8.5.14, 140
modulus of a vector, 3.2.12, 24 nonarchimedean element,
mollifier, 9.6.14, 166 5.5.9 (5), 70
mollifying kernel, 9.6.14, 166 nonconvex cone, 3.1.2 (4), 20
monomorphism, 2.3.1, 15 Nonempty Subdifferential Theorem,
monoquotient, 2.3.11, 17 3.5.8, 31
Montel space, 10.10.9 (2), 199 non-everywhere-defined operator,
Moore subnet, 1.3.5 (2), 7 2.2.1, 12
morphism, 8.2.2, 126; 11.1.2, 213 nonmeager set, 4.7.1, 52
morphism representing an algebra, nonpointed cone, 3.1.2 (4), 20
8.2.2, 126 nonreflexive space, 5.1.10 (8), 59
Motzkin formula, 3.1.13 (5), 22 norm, 5.1.9, 57
multimetric, 9.5.9, 161 norm convergence, 5.5.9 (7), 73
multimetric space, 9.5.9, 161 normable multinormed space, 5.4.1, 64
multimetric uniformity, 9.5.9, 161 normal element, 11.7.1, 224
multimetrizable topological space, normal operator, Ex. 8.17, 145
9.5.10, 161 normal space, 9.3.11, 153
multimetrizable uniform space, normalized element, 6.3.5, 88
9.5.10, 161 normally solvable operator, 7.6.9, 116
multinorm, 5.1.6, 57 normative inequality, 5.1.10 (7), 59
Multinorm Comparison Theorem, normed algebra, 5.6.3, 74
5.3.2, 62 normed dual, 5.1.10 (8), 59
multinorm summable family of vectors, normed space, 5.1.9, 57
5.5.9 (7), 73 normed space of bounded elements,
multinormed space, 5.1.6, 57 5.5.9 (5), 70
multiplication formula, 10.11.5, 205 norming set, 8.1.1, 120
multiplication of a germ by a complex norm-one element, 5.5.6, 68
number, 8.1.16, 125 nowhere dense set, 4.7.1, 52
multiplicative linear operator, 8.2.2, 126 nullity, 8.5.1, 137
numeric family, 1.1.3 (4), 2
natural order, 3.2.6 (1), 23 numeric function, 9.6.4, 165
negative part, 3.2.12, 24 numeric range, 11.9.1, 229
neighborhood about a point, numeric set, 1.1.3 (4), 2
9.1.1 (2), 146
neighborhood about a point in a metric one-point compactification, 9.4.22, 159
space, 4.1.9, 41 one-to-one correspondence, 1.1.3 (3), 2
neighborhood filter, 4.1.10, 41 open ball, 4.1.3, 40
neighborhood filter of a set, 9.3.7, 152 open ball of RN , 9.6.16, 166
neighborhood of a set, 8.1.13 (2), open correspondence, 7.3.12, 107
124; 9.3.7, 152 Open Correspondence Principle,
Nested Ball Theorem, 4.5.7, 47 7.3.13, 107
nested sequence, 4.5.7, 47 open cylinder, 4.1.3, 40
net, 1.2.16, 6 open halfspace, Ex. 3.3, 39
Subject Index 269

Open Mapping Theorem, 7.4.6, 108 patch, 10.9.11, 193


open segment, 3.6.1, 31 perforated disk, 4.8.5, 54
open set, 9.1.4, 146 periodic distribution, 10.11.17 (7), 211
open set in a metric space, 4.1.11, 41 Pettis Theorem, 10.7.4, 181
openness at a point, 7.3.6, 107 Phillips Theorem, 7.4.13, 110
operator, 2.2.1, 12 Plancherel Theorem, 10.11.14, 209
operator ideal, 8.3.3, 132 point finite cover, 9.6.2, 164
operator norm, 5.1.10 (7), 58 point in a metric space, 4.1.1, 40
operator representation, 8.2.2, 126 point in a space, 2.1.4 (3), 11
order, 1.2.2, 3 point in a vector space, 2.1.3, 10
order by inclusion, 1.3.1, 6 pointwise convergence, 9.5.5 (6), 161
order compatible with vector structure, pointwise operation, 2.1.4 (4), 11
3.2.1, 22 polar, 7.6.8, 116; 10.5.1, 177
order ideal, 10.8.11, 183 Polar Lemma, 7.6.11, 116
order of a distribution, 10.10.5 (3), 195 polarization identity, 6.1.3, 80
ordered set, 1.2.2, 3 Pontryagin–van Kampen Duality
ordered vector space, 3.2.1, 22 Theorem, 10.11.2, 203
ordering, 1.2.2, 3 poset, 1.2.2, 3
ordering cone, 3.2.4, 23 positive cone, 3.2.5, 23
oriented envelope, 4.8.8, 54 positive definite inner product, 6.1.4, 80
orthocomplement, 6.2.5, 85 positive distribution, 10.10.5 (2), 195
orthogonal complement, 6.2.5, 85 positive element of a C ∗ -algebra,
orthogonal family, 6.3.1, 87 11.9.4, 230
orthogonal orthoprojections, 6.2.12, 86 positive form on a C ∗ -algebra,
orthogonal set, 6.3.1, 87 Ex. 11.11, 235
orthogonal vectors, 6.2.5, 85 positive hermitian form, 6.1.4, 80
orthonormal family, 6.3.6, 88 positive matrix, Ex. 3.13, 39
orthonormal set, 6.3.6, 88 positive operator, 3.2.6 (3), 23
orthonormalized family, 6.3.6, 88 positive part, 3.2.12, 24
orthoprojection, 6.2.7, 85 positive semidefinite hermitian
Orthoprojection Summation Theorem, form, 6.1.4, 80
6.3.3, 87 positively homogeneous functional,
Orthoprojection Theorem, 6.2.10, 85 3.4.7 (2), 27
Osgood Theorem, 4.7.5, 52 powerset, 1.2.3 (4), 4
precompact set, Ex. 9.16, 168
pair-dual space, 10.3.3, 174 pre-Hilbert space, 6.1.7, 81
pairing, 10.3.3, 174 preimage of a multinorm,
pairwise orthogonality of finitely many 5.1.10 (3), 58
orthoprojections, 6.2.14, 86 preimage of a seminorm, 5.1.4, 57
paracompact space, 9.6.9, 166 preimage of a set, 1.1.3 (5), 2
Parallelogram Law, 6.1.8, 81 preintegral, 5.5.9 (4), 68
Parseval identity, 6.3.16, 89; preneighborhood, 9.1.1 (2), 146
10.11.12, 208 preorder, 1.2.2, 3
part of an operator, 2.2.9 (4), 14 preordered set, 1.2.2, 4
partial correspondence, 1.1.3 (6), 2 preordered vector space, 3.2.1, 22
partial operator, 2.2.1, 12 presheaf, 10.9.4 (4), 190
partial order, 1.2.2, 3 pretopological space, 9.1.1 (2), 146
partial sum, 5.5.9 (7), 73 pretopology, 9.1.1, 146
partition of unity, 9.6.6, 165 primary Banach space, Ex. 7.17, 119
partition of unity subordinate prime mapping, 6.4.1, 90
to a cover, 9.6.7, 166 Prime Theorem, 10.2.13, 173
270 Subject Index

Principal Theorem of the Holomorphic real subspace, 3.1.2 (3), 20


Functional Calculus, 8.2.4, 128 real vector space, 2.1.3, 10
product, 4.3.2, 44 realification, 3.7.1, 33
product of a distribution and realification of a pre-Hilbert space,
a function, 10.10.5 (7), 196 6.1.10 (2), 83
product of germs, 8.1.16, 125 realifier, 3.7.2, 33
product of sets, 1.1.1, 1; 2.1.4 (4), 11 reducible representation, 8.2.2, 127
product of topologies, 9.2.17 (2), 151 refinement, 9.6.1, 164
product of vector spaces, 2.1.4 (4), 11 reflection of a function, 10.10.5, 197
product topology, 4.3.2, 44; 9.2.17 (2), reflexive relation, 1.2.1, 3
151 reflexive space, 5.1.10 (8), 59
projection onto X1 along X2 , regular distribution, 10.10.5 (1), 195
2.2.9 (4), 14 regular operator, 3.2.6 (3), 23
projection to a set, 6.2.3, 84 regular space, 9.3.9, 153
proper ideal, 11.4.5, 218 regular value of an operator, 5.6.13, 76
pseudometric, 9.5.7, 161
relation, 1.1.3 (2), 1
p-sum, 5.5.9 (6), 71
relative topology, 9.2.17 (1), 151
p-summable family, 5.5.9 (4), 70
punctured compactum, 9.4.21, 159 relatively compact set, 4.4.4, 46
pure subalgebra, 11.1.5, 213 removable singularity, 8.2.5 (2), 128
Pythagoras Lemma, 6.2.8, 85 representation, 8.2.2, 126
Pythagoras Theorem, 6.3.2, 87 representation space, 8.2.2, 126
reproducing cone, Ex. 7.12, 119
quasinilpotent, Ex. 8.18, 145 residual set, 4.7.4, 52
quotient mapping, 1.2.3 (4), 4 resolvent of an element of an algebra,
quotient multinorm, 5.3.11, 64 11.2.1, 215
quotient of a mapping, 1.2.3 (4), 4 resolvent of an operator, 5.6.13, 76
quotient of a seminormed space, resolvent set of an operator, 5.6.13, 76
5.1.10 (5), 58 resolvent value of an element
quotient seminorm, 5.1.10 (5), 58 of an algebra, 11.2.1, 215
quotient set, 1.2.3 (4), 4 resolvent value of an operator,
quotient space of a multinormed 5.6.13, 76
space, 5.3.11, 64 restriction, 1.1.3 (5), 2
quotient vector space, 2.1.4 (6), 12 restriction of a distribution,
radical, 11.6.11, 223 10.10.5 (6), 196
Radon F-measure, 10.9.3, 188 restriction of a measure,
Radon–Nikod∆ ym Theorem, 10.9.4 (3), 10.9.4 (4), 190
190 restriction operator, 10.9.4 (4), 190
range of a correspondence, 1.1.2, 1 reversal, 1.2.5, 5
rank, 8.5.7 (2), 139 reverse order, 1.2.3 (2), 4
rare set, 4.7.1, 52 reverse polar, 7.6.8, 116; 10.5.1, 177
Rayleigh Theorem, 6.5.2, 93 reversed multiplication, 11.1.6, 214
real axis, 2.1.2, 10 Riemann function, 4.7.7, 53
real carrier, 3.7.1, 33 Riemann–Lebesgue Lemma,
real C-measure, 10.9.4 (3), 189 10.11.5 (3), 204
real distribution, 10.10.5 (5), 196 Riemann Theorem on Series,
real hyperplane, 3.8.9, 38 5.5.9 (7), 73
real measure, 10.9.4, 189 Riesz Criterion, 8.4.2, 134
real part map, 3.7.2, 33 Riesz Decomposition Property,
real part of a function, 5.5.9 (4), 69 3.2.16, 24
real part of a number, 2.1.2, 10 Riesz–Dunford integral, 8.2.1, 126
Subject Index 271

Riesz–Dunford Integral Decomposition seminormed space, 5.1.5, 57


Theorem, 8.2.13, 131 semisimple algebra, 11.6.11, 223
Riesz–Dunford integral in an algebraic separable space, 6.3.14, 89
setting, 11.3.1, 216 separated multinorm, 5.1.8, 57
Riesz–Fisher Completeness Theorem, separated multinormed space, 5.1.8, 57
5.5.9 (4), 70 separated topological space, 9.3.2, 151
Riesz–Fisher Isomorphism Theorem, separated topology, 9.3.2, 151
6.3.16, 89 separating hyperplane, 3.8.13, 39
Riesz idempotent, 8.2.11, 130 Separation Theorem, 3.8.11, 38
Riesz–Kantorovich Theorem, 3.2.17, 24 Sequence Prime Principle, 7.6.13, 117
Riesz operator, Ex. 8.15, 145 sequence space, 3.3.1 (2), 25
Riesz Prime Theorem, 6.4.1, 90 Sequence Star Principle, 6.4.12, 92
Riesz projection, 8.2.11, 130 series sum, 5.5.9 (7), 73
Riesz–Schauder operator, sesquilinear form, 6.1.2, 80
Ex. 8.11, 144 set absorbing another set, 3.4.9, 28
Riesz–Schauder Theorem, 8.4.8, 136 set in a space, 2.1.4 (3), 11
Riesz space, 3.2.7, 23 set lacking a distribution,
Riesz Theorem, 5.3.5, 63 10.10.5 (6), 196
right approximate inverse, 8.5.9, 139 set lacking a functional, 10.8.13, 184
right Haar measure, 10.9.4 (1), 189 set lacking a measure, 10.9.4 (5), 190
right inverse of an element set of arrival, 1.1.1, 1
in an algebra, 11.1.3, 213 set of departure, 1.1.1, 1
R-measure, 10.9.4 (3), 189 set of second category, 4.7.1, 52
rough draft, 4.8.8, 54 set supporting a measure,
row-by-column rule, 2.2.9 (4), 14 10.9.4 (5), 190
salient cone, 3.2.4, 23 set that separates the points of another
Sard Theorem, 7.4.12, 110 set, 10.8.9, 183
scalar, 2.1.3, 10 set void of a distribution,
scalar field, 2.1.3, 10 10.10.5 (6), 196
scalar multiplication, 2.1.3, 10 set void of a functional, 10.8.13, 184
scalar product, 6.1.4, 80 set void of a measure, 10.9.4 (5), 190
scalar-valued function, 9.6.4, 165 setting in duality, 10.3.3, 174
Schauder Theorem, 8.4.6, 135 setting primes, 7.6.5, 115
Schwartz space of distributions, sheaf, 10.9.11, 193
10.11.16, 209 shift, 10.9.4 (1), 189
Schwartz space of functions, Shilov boundary, Ex. 11.8, 234
10.11.6, 206 Shilov Theorem, 11.2.4, 215
Schwartz Theorem, 10.10.10, 199 short sequence, 2.3.5, 16
second dual, 5.1.10 (8), 59 σ-compact, 10.9.8, 192
selfadjoint operator, 6.5.1, 93 signed measure, 10.9.4 (3), 190
semi-extended real axis, 3.4.1, 26 simple convergence, 9.5.5 (6), 161
semi-Fredholm operator, simple function, 5.5.9 (6), 72
Ex. 8.13, 145 simple Jordan loop, 4.8.2, 54
semi-inner product, 6.1.4, 80 single-valued correspondence,
semimetric, 9.5.7, 161 1.1.3 (3), 1
semimetric space, 9.5.7, 161 Singularity Condensation Principle,
seminorm, 3.7.6, 34 7.2.12, 105
seminorm associated with a positive Singularity Fixation Principle,
element, 5.5.9 (5), 70 7.2.11, 105
seminormable space, 5.4.6, 66 skew field, 11.2.3, 215
272 Subject Index

slowly increasing distribution, ∗-isomorphism, 11.8.3, 226


10.11.16, 209 ∗-linear functional, 2.2.4, 13
smooth function, 9.6.13, 166 ∗-representation, 11.8.3, 226
smoothing process, 9.6.18, 167 star-shaped set, 3.1.2 (7), 20
Snowflake Lemma, 2.3.16, 18 state, 11.9.1, 229
space countable at infinity, 10.9.8, 192 Steklov condition, 6.3.10, 88
space of bounded elements, 5.5.9 (5), 70 Steklov Theorem, 6.3.11, 88
space of bounded functions, 5.5.9 (2), 68 step function, 5.5.9 (6), 72
space of bounded operators, Stone Theorem, 10.8.10, 183
5.1.10 (7), 59 Stone–Weierstrass Theorem for
space of compactly-supported C(Q, C), 11.8.2, 226
distributions, 10.10.5 (9), 197 Stone–Weierstrass Theorem for
space of convergent sequences, C(Q, R), 10.8.17, 186
5.5.9 (3), 68 Strict Separation Theorem, 10.4.8, 177
space of distributions of order at most strict subnet, 1.3.5 (2), 7
m, 10.10.8, 199 strictly positive real, 4.1.3, 40
space of essentially bounded functions, strong order-unit, 5.5.9 (5), 70
5.5.9 (5), 70 strong uniformity, 9.5.5 (6), 161
space of finite-order distributions, stronger multinorm, 5.3.1, 62
10.10.8, 199 stronger pretopology, 9.1.2, 146
space of functions vanishing at infinity, stronger seminorm, 5.3.3, 63
5.5.9 (3), 68 strongly holomorphic function,
space of X-valued p-summable 8.1.5, 122
functions, 5.5.9 (6), 72 structure of a subdifferential,
space of p-summable functions, 10.6.3, 179
5.5.9 (4), 69 subadditive functional, 3.4.7 (4), 27
space of p-summable sequences, subcover, 9.6.1, 164
5.5.9 (4), 70 subdifferential, 3.5.1, 29
space of tempered distributions, sublattice, 10.8.2, 181
10.11.16, 209 sublinear functional, 3.4.6, 27
space of vanishing sequences, submultiplicative norm, 5.6.1, 73
5.5.9 (3), 68 subnet, 1.3.5 (2), 7
Spectral Decomposition Lemma, subnet in a broad sense, 1.3.5 (2), 7
6.6.6, 96 subrepresentation, 8.2.2, 126
Spectral Decomposition Theorem, subspace of a metric space, 4.5.14, 49
8.2.12, 130 subspace of a topological space,
Spectral Endpoint Theorem, 6.5.5, 94 9.2.17 (1), 151
Spectral Mapping Theorem, 8.2.5, 128 subspace of an ordered vector space,
Spectral Purity Theorem, 3.2.6 (2), 23
11.7.11, 226 subspace topology, 9.2.17 (1), 151
spectral radius of an operator, 5.6.6, 74 Sukhomlinov–Bohnenblust–Sobczyk
Spectral Theorem, 11.8.6, 227 Theorem, 3.7.12, 35
spectral value of an element sum of a family in the sense of Lp ,
of an algebra, 11.2.1, 215 5.5.9 (6), 71
spectral value of an operator, 5.6.13, 76 sum of germs, 8.1.16, 125
spectrum, 10.2.7, 172 summable family of vectors,
spectrum of an element of an algebra, 5.5.9 (7), 73
11.2.1, 215 summable function, 5.5.9 (4), 69
spectrum of an operator, 5.6.13, 76 superset, 1.3.3, 6
spherical layer, 6.2.1, 84 sup-norm, 10.8.1, 181
∗-algebra, 6.4.13, 92 support function, 10.6.4, 179
Subject Index 273

support of a distribution, topological space, 9.1.7, 147


10.10.5 (6), 196 topological structure of a convex
support of a functional, 10.8.12, 184 set, 7.1.1, 100
support of a measure, 10.9.4 (5), 190 topological subdifferential, 7.5.8, 113
supporting function, 10.6.4, 179 topological vector space, 10.1.1, 169
supremum, 1.2.9, 5 topologically complemented subspace,
symmetric Hahn–Banach formula, 7.4.9, 108
Ex. 3.10, 39 topology, 9.1.7, 147
symmetric relation, 1.2.1, 3 topology compatible with duality,
symmetric set, 3.1.2 (7), 20 10.4.1, 175
system with integration, 5.5.9 (4), 68 topology compatible with vector
Szankowski Counterexample, structure, 10.1.1, 169
8.3.13, 134 topology given by open sets, 9.1.12, 148
topology of a multinormed space,
tail filter, 1.3.5 (2), 7 5.2.8, 61
τ -dual of a locally convex space, topology of a uniform space, 9.5.3, 160
10.2.11, 173 topology of the distribution space,
Taylor Series Expansion Theorem, 10.10.6, 198
8.1.9, 123 topology of the test function space,
tempered distribution, 10.11.16, 209 10.10.6, 198
tempered function, 5.1.10 (6), 58; total operator, 2.2.1, 12
10.11.6, 206 total set of functionals, 7.4.11 (2), 109
tempered Radon measure, 10.11.17 (3), totally bounded, 4.6.3, 49
210 transitive relation, 1.2.1, 3
test function, 10.10.1, 194 translation, 10.9.4 (1), 189
test function space, 10.10.1, 194 translation of a distribution,
theorem on Hilbert isomorphy, 10.11.17 (7), 211
X
6.3.17, 90 transpose of an operator, 7.6.2, 114
theorem on the equation A = B, trivial topology, 9.1.8 (3), 147
X
2.3.13, 17 truncator, 9.6.19 (1), 167
theorem on the equation A = B, truncator direction, 10.10.2 (5), 194
2.3.8, 16 truncator set, 10.10.2, 194
theorem on the general form twin of a Hilbert space, 6.1.10 (3), 83
of a distribution, 10.10.14, 201 twin of a vector space, 2.1.4 (2), 10
theorem on the inverse image Two Norm Principle, 7.4.16, 111
of a vector topology, 10.1.6, 171 two-sided ideal, 8.3.3, 132; 11.6.2, 220
theorem on the repeated Fourier Tychonoff cube, 9.2.17 (2), 151
transform, 10.11.13, 209 Tychonoff product, 9.2.17 (2), 151
theorem on the structure of a locally Tychonoff space, 9.3.15, 155
convex topology, 10.2.2, 171 Tychonoff Theorem, 9.4.8, 157
theorem on the structure of a vector Tychonoff topology, 9.2.17 (2), 151
topology, 10.1.4, 170 Tychonoff uniformity, 9.5.5 (4), 160
theorem on topologizing by a family T1 -space, 9.3.2, 151
of mappings, 9.2.16, 151 T1 -topology, 9.3.2, 151
there is a unique x, 2.3.9, 17 T2 -space, 9.3.5, 152
Tietze–Urysohn Theorem, T3 -space, 9.3.9, 153
10.8.20, 186 T31 /2 -space, 9.3.15, 155
topological isomorphism, 9.2.4, 149 T4 -space, 9.3.11, 153
topological mapping, 9.2.4, 149
Topological Separation Theorem, ultrafilter, 1.3.9, 7
7.5.12, 114 ultrametric inequality, 9.5.14, 162
274 Subject Index

ultranet, 9.4.4, 156 upper right Dini derivative, 4.7.7, 53


unconditionally summable family upward-filtered set, 1.2.15, 6
of vectors, 5.5.9 (7), 73 Urysohn Great Lemma, 9.3.13, 154
unconditionally summable sequence, Urysohn Little Lemma, 9.3.10, 153
5.5.9 (7), 73 Urysohn Theorem, 9.3.14, 155
underlying set, 2.1.3, 10 2-Ultrametric Lemma, 9.5.15, 162
Uniform Boundedness Principle,
7.2.5, 103 vague topology, 10.9.5, 191
uniform convergence, 7.2.10, 105; value of a germ at a point, 8.1.21, 126
9.5.5 (6), 161 van der Waerden function, 4.7.7, 53
uniform space, 9.5.1, 159 vector, 2.1.3, 10
uniformity, 9.5.1, 159 vector addition, 2.1.3, 10
uniformity of a multinormed space, vector field, 5.5.9 (6), 71
5.2.4, 60 vector lattice, 3.2.7, 23
uniformity of a seminormed space, vector space, 2.1.3, 10
5.2.2, 60 vector sublattice, 10.8.4 (4), 182
uniformity of a topological vector vector topology, 10.1.1, 169
space, 10.1.10, 171 Volterra operator, Ex. 5.12, 79
uniformity of the empty set, 9.5.1, 159 von Neumann–Jordan Theorem,
uniformizable space, 9.5.4, 160 6.1.9, 81
uniformly continuous mapping, 4.2.5, 44 V -net, 4.6.2, 49
unit, 10.9.4, 188 V -small, 4.5.3, 47
unit ball, 5.2.11, 61 weak derivative, 10.10.5 (4), 196
unit circle, 8.1.3, 121 weak multinorm, 5.1.10 (4), 58
unit disk, 8.1.3, 121 weak topology, 10.3.5, 174
unit element, 11.1.1, 213 weak∗ topology, 10.3.11, 175
unit sphere, Ex. 10.6, 212 weak uniformity, 9.5.5 (6), 161
unit vector, 6.3.5, 88 weaker pretopology, 9.1.2, 146
unital algebra, 11.1.1, 213 weakly holomorphic function, 8.1.5, 122
unitary element, 11.7.1, 224 weakly operator holomorphic function,
unitary operator, 6.3.17, 90 8.1.5, 122
unitization, 11.1.2, 213 Weierstrass function, 4.7.7, 53
unity, 11.1.1, 213 Weierstrass Theorem, 4.4.5, 46;
unity of a group, 10.9.4 (1), 188 9.4.5, 157
unity of an algebra, 11.1.1, 213 Well-Posedness Principle, 7.4.6, 108
unordered sum, 5.5.9 (7), 73 Wendel Theorem, 10.9.4 (7), 191
unorderly summable sequence, Weyl Criterion, 6.5.4, 93
5.5.9 (7), 73
Unremovable Spectral Boundary X-valued function, 5.5.9 (6), 71
Theorem, 11.2.6, 216 Young inequality, 5.5.9 (4), 69
upper bound, 1.2.4, 4
upper envelope, 3.4.8 (3), 28 zero of a vector space, 2.1.4 (3), 11

You might also like